1 Introduction and basic definitions

For T in the Banach algebra \(\mathcal {B}(X)\) of bounded linear operators acting on a Banach space X,  we will denote by \(\sigma (T)\) the spectrum of T,  by \(\sigma _a(T)\) the approximate point spectrum of T,  by \(\mathcal {N}(T)\) the null space of T, by n(T) the nullity of T,  by \(\mathcal {R}(T)\) the range of T and by d(T) its defect. If \(n(T)<\infty \) and \(d(T)<\infty \), then T is called a Fredholm operator and its index is defined by \(\text{ ind }(T)=n(T)-d(T).\) A Weyl operator \(T\in \mathcal {B}(X)\) is a Fredholm operator of index zero and the Weyl spectrum is defined by \(\sigma _{w}(T)=\{\lambda \in \mathbb {C}\, |\,T-\lambda I \text{ is } \text{ not } \text{ a } \text{ Weyl } \text{ operator }\}.\) \(T\in \mathcal {B}(X)\) is called an upper (resp., a lower) semi-Fredholm if \(\mathcal {R}(T)\) is closed and \(n(T)<\infty \) (resp., \(d(T)<\infty ).\) The respective semi-Fredholm spectrum and semi-Weyl spectrum of T are defined respectively, by \(\sigma _{sf}(T)=\{\lambda \in \mathbb {C}\, |\,T-\lambda I \text{ is } \text{ not } \text{ a } \text{ semi-Fedholm } \text{ operator }\},\) and \(\sigma _{sf_+^-}(T)=\{\lambda \in \mathbb {C}\, |\,T-\lambda I \text{ is } \text{ not } \text{ an } \text{ upper } \text{ semi-Fredholm } \text{ operator } \text{ with } \text{ index } \text{ less } \text{ or } \text{ equal } \text{ than } \text{ zero }\}.\)

For a bounded linear operator T and \(n\in \mathbb {N},\) let \(T_{[n]}: \mathcal {R}(T^n)\rightarrow \mathcal {R}(T^n)\) be the restriction of T to \(\mathcal {R}(T^n).\) \(T \in \mathcal {B}(X)\) is said to be a semi B-Fredholm if for some integer \(n \ge 0,\) the range \(\mathcal {R}(T^n)\) is closed and \(T_{[n]}\) is a semi-Fredholm; its index is defined as the index of the semi-Fredholm operator \(T_{[n]}.\) The respective semi B-Fredholm spectrum of T is defined by \(\sigma _{sbf}(T)=\{\lambda \in \mathbb {C}\, |\, T-\lambda I \text{ is } \text{ not } \text{ a } \text{ semi } \text{ B-Fredholm } \text{ operator }\}.\)

The ascent of an operator T is defined by \(a(T)=\text{ inf } \{ n\in \mathbb {N}\, |\,\mathcal {N}(T^n)=\mathcal {N}(T^{n+1})\}\), and the descent of T is defined by \(\delta (T)= \text{ inf } \{ n \in \mathbb {N}\, |\, \mathcal {R}(T^n)= \mathcal {R}(T^{n+1})\},\) with \( \text{ inf }\, \emptyset = \infty .\) According to Heuser [8], a complex number \(\lambda \in \sigma (T)\) is a pole of the resolvent of T if \(T-\lambda I\) has finite ascent and finite descent, and in this case they are equal. An operator \(T\in \mathcal {B}(X)\) is said to be Browder operator if it is a Fredholm with finite ascent and descent, and is said to be an upper Browder operator if it is an upper semi-Fredholm operator with finite ascent. The respective Browder spectrum and upper Browder spectrum of T are defined respectively, by \(\sigma _{b}(T)=\{\lambda \in \mathbb {C}\, |\, T-\lambda I \text{ is } \text{ not } \text{ a } \text{ Browder } \text{ operator }\},\) and \(\sigma _{ub}(T)=\{\lambda \in \mathbb {C}\, |\, T-\lambda I \text{ is } \text{ not } \text{ an } \text{ upper } \text{ Browder } \text{ operator }\}.\)

In the following, we recall the definition of a property which has a relevant role in local spectral theory. For more details about this property see the monographs of Laursen and Neumann [10] and Aiena [1].

Definition 1.1

An operator \(T\in \mathcal {B}(X)\) is said to have the single valued extension property at \(\lambda _{0}\in \mathbb {C}\) (abbreviated SVEP at \(\lambda _{0}\)), if for every open neighborhood \({\mathcal U} \) of \(\lambda _{0}\), the only analytic function \(f:{\mathcal U} \longrightarrow X\) which satisfies the equation \((T-\lambda I)f(\lambda )=0\) for all \(\lambda \in {\mathcal U} \) is the function \(f\equiv 0\). An operator \(T\in \mathcal {B}(X)\) is said to have SVEP if T has SVEP at every point \(\lambda \in \mathbb {C}\).

Evidently, \(T\in \mathcal {B}(X)\) has SVEP at every isolated point of the spectrum. We summarize in the following list the usual notations and symbols needed later.

Notations and symbols

\({\mathcal F}(X) \) :

The ideal of finite rank operators in \(\mathcal {B}(X),\)

\({\mathcal K}(X) \) :

The ideal of compact operators in \(\mathcal {B}(X),\)

\({\mathcal N}(X) \) :

The class of nilpotent operators on X

\({\mathcal Q}(X) \) :

The class of quasi-nilpotent operators on X

\(\text{ iso }\,A\) :

Isolated points of a subset \(A\subset \mathbb {C},\)

\(\text{ acc }\,A\) :

Accumulations points of a subset \(A\subset \mathbb {C},\)

D(0, 1):

The closed unit disc in \(\mathbb {C},\)

C(0, 1):

The unit circle of \(\mathbb {C},\)

\(\Pi (T)\) :

poles of T

\(\Pi ^0(T)\) :

Poles of T of finite rank,

\(\Pi _a(T)\) :

Left poles of T

\(\Pi _a^0(T)\) :

Left poles of T of finite rank,

\(\sigma _{p}(T)\) :

Eigenvalues of T

\(\sigma _{p}^f(T)\) :

Eigenvalues of T of finite multiplicity,

\(E^0(T)\) :

\(\text{ iso }\,\sigma (T)\cap \sigma _{p}^f(T),\)

E(T):

\(\text{ iso }\,\sigma (T)\cap \sigma _{p}(T),\)

\(E_a^0(T)\) :

\(\text{ iso }\,\sigma _a(T)\cap \sigma _{p}^f(T),\)

\(\sigma _{b}(T)= \sigma (T){\setminus }\Pi ^0(T)\) :

Browder spectrum of T

\(\sigma _{ub}(T)= \sigma _a(T){\setminus }\Pi _a^0(T)\) :

Upper-Browder spectrum of T

\(\sigma _{w}(T)\) :

Weyl spectrum of T

\(\sigma _{sf_+^-}(T)\) :

Semi-Weyl spectrum of T

\(\sigma _{sbf}(T)\) :

Semi B-Fredholm spectrum of T.

Definition 1.2

[2, 6, 9, 12] Let \(T\in \mathcal {B}(X).\) T is said to satisfy

  1. (i)

    a-Browder’s theorem if \(\sigma (T){\setminus }\sigma _{sf_+^-}(T)=\Pi _a^0(T),\) or equivalently \(\sigma _{ub}(T)=\sigma _{sf_+^-}(T).\)

  2. (ii)

    Browder’s theorem if \(\sigma (T){\setminus }\sigma _{w}(T)=\Pi ^0(T),\) or equivalently \(\sigma _{b}(T)=\sigma _{w}(T).\)

  3. (iii)

    Weyl’s theorem if \(\sigma (T){\setminus }\sigma _{w}(T)=E^0(T).\)

  4. (iv)

    Property (k) if \(\sigma (T){\setminus }\sigma _{w}(T)=E(T). \)

Definition 1.3

Let \(T\in \mathcal {B}(X)\) and \(S\in \mathcal {B}(X).\) We will say that T and S have a shared stable sign index if for each \(\lambda \not \in \sigma _{sbf}(T)\) and \(\mu \not \in \sigma _{sbf}(S)\), \(\text{ ind }(T-\lambda I)\) and \(\text{ ind }(S-\mu I)\) have the same sign.

For examples we have:

  1. 1.

    It is easily verified that if \(T\in \mathcal {B}(X)\) has SVEP then \(\text{ ind }(T-\mu I)\le 0\) for every \(\mu \not \in \sigma _{sbf}(T).\) So if S and T have SVEP, then they have a shared stable sign index.

  2. 2.

    Here and elsewhere, \(\mathcal {H}\) denotes a Hilbert space. It is well known that every hyponormal operator T acting on \(\mathcal {H}\) has property \((H_1)\) (for the definition of \((H_1),\) see the end of the second section) and hence has SVEP. As a consequence of the first point, every two hyponormal operators have a shared stable sign index. Recall that \(T\in \mathcal {B}(\mathcal {H})\) is said to be hyponormal if \(T^*T-TT^*\ge 0\) (or equivalently \(\Vert T^*x\Vert \le \Vert Tx\Vert \) for all \(x\in \mathcal {H}\)). The class of hyponormal operators includes also subnormal operators and quasinormal operators, see Conway [4].

After giving an introduction and some preliminaries in the first section, we study in the second section the preservation of property (k) introduced and studied by Kaushik and Kashyap [9], under several commuting Riesz-type perturbations. We prove in particular that if T is an isoloid operator acting on a Banach space and satisfies property (k),  then \(T+S\) satisfies property (k) for every finite rank power operator S which commutes with T. Moreover, we give generalization of some perturbation results to commuting Riesz operators such as Theorems 2.3 and 2.8 of Berkani and Zariouh [3]. In the end of this paper, we prove that if S and T are isoloid bounded operators acting on Banach spaces and satisfy property (k),  then \(S\oplus T\) satisfies property (k) if and only if \(\sigma _{w}(S\oplus T)=\sigma _{w}(S)\cup \sigma _{w}(T),\) extending a result of Kaushik and Kashyap [9]. Some crucial examples are also given.

2 Property (k) and perturbations

We recall that an operator \(R\in \mathcal {B}(X)\) is said to be Riesz if \(R-\mu I\) is Fredholm for every non-zero complex \(\mu ,\) that is, \(\pi (R)\) is quasi-nilpotent in the Calkin algebra \(C(X)=\mathcal {B}(X)/{\mathcal K}(X) \) where \(\pi \) is the canonical mapping of \( \mathcal {B}(X)\) into C(X). We denote by \(\mathcal {R}(X)\) the class of Riesz operators and by \({\mathcal F^0}(X) ,\) the class of finite rank power operators as follows:

$$\begin{aligned} {\mathcal F^0}(X) =\left\{ S\in \mathcal {B}(X) : S^n \in {\mathcal F}(X) \text{ for } \text{ some } n\in \mathbb {\mathbb {N}}\right\} . \end{aligned}$$

Clearly, \({\mathcal F}(X) \cup {\mathcal N}(X) \subset {\mathcal F^0}(X) \subset {\mathcal R}(X) ,\) and \({\mathcal K}(X) \cup {\mathcal Q}(X) \subset {\mathcal R}(X) .\)

According to Oberai [11], Rakocěvi\(\acute{c}\) [14] and Tylli [15], we know that for every \(T\in \mathcal {B}(X)\) and \(R\in {\mathcal R}(X) \) such that \(TR=RT,\) \(\sigma _{*}(T+R)=\sigma _{*}(T);\) where \(\sigma _{*}\in \{\sigma _{sf_+^-}, \sigma _{w}, \sigma _{b}, \sigma _{ub}\}.\) From this, we give the following known lemma which we need in the proof of the next main results.

Lemma 2.1

Let \(T\in \mathcal {B}(X)\) and let \(R\in {\mathcal R}(X) \) be a commuting operator with T. The following statements hold:

  1. (i)

    T satisfies Browder’s theorem if and only if \(T+R\) satisfies Browder’s theorem.

  2. (ii)

    T satisfies a-Browder’s theorem if and only if \(T+R\) satisfies a-Browder’s theorem.

We start this section by the following nilpotent perturbation result.

Proposition 2.2

Let \(T\in \mathcal {B}(X)\) and let \(N\in {\mathcal N}(X) \) which commutes with T. Then T satisfies property (k) if and only if \(T+N\) satisfies property (k).

Proof

Since N is nilpotent and commutes with T,  we know that \(\sigma (T+N)=\sigma (T),\) and it is easily seen that \(0<n(T+N)\Longleftrightarrow 0<n(T).\) Since \(\sigma _{w}(T)=\sigma _{w}(T+N),\) it follows that T satisfies property (k) if and only if \(T+N\) satisfies property (k). \(\square \)

  • Note that the assumption of commutativity in the Proposition 2.2 is crucial. Let T and N be defined on \(\ell ^2(\mathbb {N})\) by \( T(x_1, x_2, \ldots )=(0, \frac{x_1}{2}, \frac{x_2}{3}, \ldots )\) and \( N(x_1, x_2, \ldots )=(0, \frac{-x_1}{2}, 0, 0, \ldots ).\) Clearly N is nilpotent and does not commute with T. The property (k) is satisfied by T,  since \(\sigma (T)=\{0\}=\sigma _{w}(T)\) and \(E(T)=\emptyset .\) But \(T+N\) does not satisfy property (k),  because \(\sigma (T+N)=\sigma _{w}(T+N)=\{0\}\) and \(\{0\}= E(T+N).\)

  • The stability of property (k) showed in Proposition 2.2 cannot be extended to commuting quasi-nilpotent operators, as we can see in the next example:

Example 2.3

Let T be the operator defined on \(\ell ^2(\mathbb {N})\) by \(T(x_1,x_2, \ldots )=(0, \frac{x_1}{2},\frac{x_2}{3}, \ldots ).\) Put \(R=-T,\) clearly R is quasi-nilpotent, compact and commutes with T. As it is already mentioned, T satisfies property (k). But \(T+R=0\) does not satisfy this property. Indeed, \(\sigma (T+R)=\{0\}=\sigma _{w}(T+R),\) \(E(T+R)=\{0\}.\) Note also that \(\Pi ^ 0(T+R)=\emptyset ,\) \(\Pi ^0(T)=\emptyset .\)

Theorem 2.4

Let \(R\in {\mathcal R}(X) \) and let \(T\in \mathcal {B}(X)\) be a commuting operator with T. If T satisfies property (k),  then the following statements are equivalent:

  1. (i)

    \(T+R\) satisfies property (k); 

  2. (ii)

    \(\Pi ^0(T+R)=E(T+R);\)

  3. (iii)

    \(E(T+R)\cap \sigma (T)\subset \Pi ^0(T)\).

Proof

  1. (i)

    \(\Longleftrightarrow \) (ii) If \(T+R\) satisfies property (k) then from [9, Theorem 2.5], \(\Pi ^0(T+R)=E(T+R).\) Conversely, since T satisfies property (k),  then it satisfies Browder’s theorem, and from Lemma 2.1, \(T+R\) satisfies Browder’s theorem too. So \(\sigma (T+R){\setminus }\sigma _{w}(T+R)=\Pi ^0(T+R).\) Thus \(T+R\) satisfies property (k).

  2. (ii)

    \(\Longleftrightarrow \) (iii) Suppose that \(\Pi ^0(T+R)=E(T+R)\) and let \(\lambda _0\in E(T+R)\cap \sigma (T)\) be arbitrary. Then \(\lambda _0\in \Pi ^0(T+R)\cap \sigma (T)\) and hence \(\lambda _0\in \sigma (T){\setminus }\sigma _{b}(T)=\Pi ^0(T).\) Consequently, \(E(T+R)\cap \sigma (T)\subset \Pi ^0(T).\) Conversely, suppose that \(E(T+R)\cap \sigma (T)\subset \Pi ^0(T)\) and let \(\mu _0\in E(T+R)\) be arbitrary. We distinguish two cases: the first is \(\mu _{0}\in \sigma (T).\) Then \(\mu _{0}\in E(T+R)\cap \sigma (T)\subset \Pi ^0(T).\) It follows that \(\mu _{0}\not \in \sigma _{b}(T+R)\) and since \(\mu _{0}\in \sigma (T+R),\) then \(\mu _{0}\in \Pi ^0(T+R).\) The second case is \(\mu _{0}\not \in \sigma (T).\) This implies that \(\mu _{0}\not \in \sigma _{b}(T+R)\) and then \(\mu _{0}\in \Pi ^0(T+R).\) In the two cases we have \(\Pi ^0(T+R)\supset E(T+R)\) and as the opposite inclusion is always true, then \(\Pi ^0(T+R)=E(T+R).\) Remark that the statements (ii) and (iii) are always equivalent without the assumption that T satisfies property (k). \(\square \)

As an application of Theorem 2.4 to commuting isoloid operators, we give the following corollary. Recall that an operator \(T\in \mathcal {B}(X)\) is said to be isoloid (resp., polaroid) if \(\text{ iso }\,\sigma (T)=E(T)\) (resp., \(\text{ iso }\,\sigma (T)=\Pi (T)\)).

Corollary 2.5

Let \(S\in {\mathcal F^0}(X) \) and let \(T\in \mathcal {B}(X)\) be an isoloid operator commuting with S. If T satisfies property (k) then \(T+S\) satisfies property (k).

Proof

Let \(\lambda _{0}\in E(T+S)\cap \sigma (T)\) be arbitrary. Then \(\lambda _{0}\not \in \text{ acc }\,\sigma (T+S)=\text{ acc }\,\sigma (T),\) see [16, Theorem 2.2]. As \(\lambda _{0}\in \sigma (T),\) then \(\lambda _{0}\in \text{ iso }\,\sigma (T)=E(T).\) The property (k) for T implies that \(E(T)=\Pi ^0(T)\) and hence \( E(T+S)\cap \sigma (T)\subset \Pi ^0(T).\) But this is equivalent from Theorem 2.4, to say that \(T+S\) satisfies property (k). \(\square \)

The next theorem extends [3, Theorem 2.3] to commuting Riesz perturbations which are not necessary nilpotent or compact. According to Rakocěvić [13], we recall that an operator \(T\in \mathcal {B}(X)\) is said to satisfy a-Weyl’s theorem if \(\sigma _a(T){\setminus }\sigma _{sf_+^-}(T)=E_a^0(T).\)

Theorem 2.6

Let \(R\in {\mathcal R}(X) .\) If \(T\in \mathcal {B}(X)\) satisfies a-Weyl’s theorem and commutes with R,  then the following statements are equivalent:

  1. (i)

    \(T+R\) satisfies a-Weyl’s theorem;

  2. (ii)

    \(\Pi _a^0(T+R)=E_a^0(T+R);\)

  3. (iii)

    \(E_a^0(T+R)\cap \sigma _a(T)\subset E_a^0(T)\).

Proof

  1. (i)

    \(\Longleftrightarrow \) (ii) Suppose that \(T+R\) satisfies a-Weyl’s theorem and let \(\mu _0\in E_a^0(T+R)\) be arbitrary. Then \(\mu _0\in E_a^0(T+R)\Longleftrightarrow \mu _0\in \text{ iso }\,\sigma _a(T+R)\cap \,\sigma _{sf_+^-}(T+R)^C \Longleftrightarrow \mu _0\in \Pi _a^0(T+R),\) where \(\sigma _{sf_+^-}(T+R)^C\) is the complement of the semi-Weyl spectrum of \(T+R.\) Thus \(\Pi _a^0(T+R)=E_a^0(T+R).\) For the converse, since T satisfies a-Weyl’s theorem, then it satisfies a-Browder’s theorem and therefore \(T+R\) satisfies a-Browder’s theorem too, see Lemma 2.1. So \(\sigma _a(T+R){\setminus }\sigma _{sf_+^-}(T+R)=\Pi _a^0(T+R)=E_a^0(T+R).\)

  2. (ii)

    \(\Longleftrightarrow \) (iii) Suppose that \(\Pi _a^0(T+R)=E_a^0(T+R)\) and let \(\lambda _0\in E_a^0(T+R)\cap \, \sigma _a(T)\) be arbitrary. Then \(\lambda _0\in \Pi _a^0(T+R)\cap \, \sigma _a(T).\) Hence \(\lambda _0\in \sigma _a(T){\setminus }\sigma _{ub}(T)=\Pi _a^0(T).\) Consequently, \(E_a^0(T+R)\cap \sigma _a(T)\subset E_a^0(T).\) Conversely, suppose that \(E_a^0(T+R)\cap \sigma _a(T)\subset E_a^0(T)\) and let \(\mu _0\in E_a^0(T+R)\) be arbitrary. We distinguish two cases: the first is \(\mu _{0}\in \sigma _a(T).\) Then \(\mu _{0}\in E_a^0(T+R)\cap \sigma _a(T)\subset E_a^0(T).\) It follows that \(\mu _{0}\not \in \sigma _{sf_+^-}(T+R)\) and since \(\mu _{0}\in \text{ iso }\,\sigma _a(T+R),\) then \(\mu _{0}\in \Pi _a^0(T+R).\) The second case is \(\mu _{0}\not \in \sigma _a(T).\) This implies that \(\mu _{0}\not \in \sigma _{ub}(T+R)\) and so \(\mu _{0}\in \Pi _a^0(T+R).\) In the two cases we have \(\Pi _a^0(T+R)\supset E_a^0(T+R)\) and as the opposite inclusion is always true, then \(\Pi _a^0(T+R)=E_a^0(T+R)\). \(\square \)

The next theorem extends [3, Theorem 2.8] to commuting Riesz perturbations which are not necessary nilpotent or compact.

Theorem 2.7

Let \(R\in {\mathcal R}(X) .\) If \(T\in \mathcal {B}(X)\) satisfies Weyl’s theorem and commutes with R,  then the following statements are equivalent:

  1. (i)

    \(T+R\) satisfies Weyl’s theorem;

  2. (ii)

    \(\Pi ^0(T+R)=E^0(T+R);\)

  3. (iii)

    \(E^0(T+R)\cap \sigma (T)\subset E^0(T)\).

Proof

  1. (i)

    \(\Longleftrightarrow \) (ii) Suppose that \(T+R\) satisfies Weyl’s theorem and let \(\mu _0\in E^0(T+R)\) be arbitrary. Then \(\mu _0\in E^0(T+R)\Longleftrightarrow \mu _0\in \text{ iso }\,\sigma (T+R)\cap \,\sigma _{w}(T+R)^C \Longleftrightarrow \mu _0\in \Pi ^0(T+R),\) where \(\sigma _{w}(T+R)^C\) is the complement of the Weyl spectrum of \(T+R.\) Thus \(\Pi ^0(T+R)=E^0(T+R).\) For the converse, since T satisfies Weyl’s theorem, then it satisfies Browder’s theorem and therefore \(T+R\) satisfies Browder’s theorem too. So \(\sigma (T+R){\setminus }\sigma _{w}(T+R)=\Pi ^0(T+R)=E^0(T+R).\)

  2. (ii)

    \(\Longleftrightarrow \) (iii) Suppose that \(\Pi ^0(T+R)=E^0(T+R)\) and let \(\lambda _0\in E^0(T+R)\cap \, \sigma (T)\) be arbitrary. Then \(\lambda _0\in \Pi ^0(T+R)\cap \, \sigma (T)\) and hence \(\lambda _0\in \sigma (T){\setminus }\sigma _{b}(T)=\Pi ^0(T).\) Consequently, \(E^0(T+R)\cap \,\sigma (T)\subset E^0(T).\) Conversely, suppose that \(E^0(T+R)\cap \, \sigma (T)\subset E^0(T)\) and let \(\mu _0\in E^0(T+R)\) be arbitrary. We distinguish two cases: the first is \(\mu _{0}\in \sigma (T).\) Then \(\mu _{0}\in E^0(T+R)\cap \,\sigma (T)\subset E^0(T).\) It follows that \(\mu _{0}\in \sigma _{w}(T+R)^C\cap \,\text{ iso }\,\sigma (T+R) \Longleftrightarrow \mu _{0}\in \Pi ^0(T+R).\) The second case is \(\mu _{0}\not \in \sigma (T).\) This implies that \(\mu _{0}\not \in \sigma _{b}(T+R).\) Thus \(\mu _{0}\in \Pi ^0(T+R).\) In the two cases we have \(\Pi ^0(T+R)\supset E^0(T+R)\) and as the opposite inclusion is always true, then \(\Pi ^0(T+R)=E^0(T+R)\). \(\square \)

Corollary 2.8

Let \(S\in {\mathcal F^0}(X) \) and let \(T\in \mathcal {B}(X)\) be a bounded operator commuting with S. The following assertions hold:

  1. (i)

    If T satisfies a-Weyl’s theorem and \(\text{ iso }\,\sigma _a(T)=E_a^0(T),\) then \(T+S\) satisfies a-Weyl’s theorem.

  2. (ii)

    If T satisfies Weyl’s theorem and \(\text{ iso }\,\sigma (T)=E^0(T),\) then \(T+S\) satisfies Weyl’s theorem.

Proof

  1. (i)

    Let \(\lambda _{0}\in E_a^0(T+S)\cap \sigma _a(T)\) be arbitrary. Then \(\lambda _{0}\not \in \text{ acc }\,\sigma _a(T+S)=\text{ acc }\,\sigma _a(T),\) see [16, Theorem 2.2]. As \(\lambda _{0}\in \sigma _a(T)\) then \(\lambda _{0}\in \text{ iso }\,\sigma _a(T)=E_a^0(T).\) So \( E_a^0(T+S)\cap \sigma _a(T)\subset E_a^0(T).\) But this is equivalent from Theorem 2.6, to say that \(T+S\) satisfies a-Weyl’s theorem.

  2. (ii)

    Goes similarly with the proof of the first assertion, as an application of Theorem 2.7. \(\square \)

In the next, we explore conditions on \(S\in \mathcal {B}(X)\) and \(T\in \mathcal {B}(Y)\) so that \(S\oplus T\) satisfies property (k). The motivation for this work has come from Duggal and Kubrusly [7]. We begin with an example which shows that even if two operators S and T satisfy property (k),  yet there direct sum may fail to satisfy property (k).

Example 2.9

Let R and L be the operators defined on \(\ell ^2(\mathbb {\mathbb {N}})\) by \(R(x_1,x_2, x_3, \ldots )=(0, x_1, x_2, x_3, \ldots )\) and \(L(x_1,x_2, x_3, \ldots )=( x_2, x_3, \ldots ).\) Then property (k) holds for R and L,  since \(\sigma (R)= \sigma _{w}(R)= D(0, 1),\) \(E(R)=\emptyset ,\) \(\sigma (L)=\sigma _{w}(L)=D(0, 1)\) and \(E(L)=\emptyset \). But it does not hold for \(R\oplus L.\) In fact \(\sigma (R\oplus L)=D(0, 1),\) and as \(n(R\oplus L)=d(R\oplus L)=1\) then \(0\not \in \sigma _{w}(R\oplus L)\). So \(\sigma _{w}(R\oplus L)\subsetneq \sigma (R\oplus L).\) We also remark that \(E(R\oplus L)=\emptyset .\) Thus \(\sigma (R\oplus L){\setminus }\sigma _{w}(R\oplus L)\ne E(R\oplus L).\) Note that S and T are isoloid and \(\sigma _{w}(R\oplus L)\subsetneq \sigma _{w}(R)\cup \sigma _{w}(L)=D(0, 1).\)

Nonetheless, and under the assumption that S and T are isoloid, we give in the following result a characterization of the stability of property (k) under direct sum.

Theorem 2.10

Let \(S\in \mathcal {B}(X)\) and let \(T\in \mathcal {B}(Y),\) X and Y are Banach spaces. If S and T satisfy property (k) and are isoloid, then the following assertions are equivalent:

  1. (i)

    \(S\oplus T\) satisfies property (k);

  2. (ii)

    \(\sigma _{w}(S\oplus T)=\sigma _{w}(S)\cup \sigma _{w}(T)\).

Proof

  1. (i)

    \(\Longrightarrow \) (ii) The property (k) for \(S\oplus T\) implies with no other restriction on either S or T that \(\sigma _{w}(S\oplus T)=\sigma _{w}(S)\cup \sigma _{w}(T)\). Indeed, as \(S\oplus T\) satisfies property (k) then it satisfies Browder’s theorem and so \(\sigma _{w}(S\oplus T)=\sigma _{b}(S\oplus T).\) Since \(\sigma _{b}(S\oplus T)=\sigma _{b}(S)\cup \sigma _{b}(T),\) then \(\sigma _{w}(S\oplus T)=\sigma _{b}(S)\cup \sigma _{b}(T),\) and as \(\sigma _{w}(S)\cup \sigma _{w}(T)\subset \sigma _{b}(S)\cup \sigma _{b}(T)\), we then have \(\sigma _{w}(S)\cup \sigma _{w}(T)\subset \sigma _{w}(S\oplus T).\) Hence \(\sigma _{w}(S\oplus T)=\sigma _{w}(S)\cup \sigma _{w}(T).\)

  2. (ii)

    \(\Longrightarrow \) (i) Suppose that \(\sigma _{w}(S\oplus T)=\sigma _{w}(S)\cup \sigma _{w}(T)\). As S and T are isoloid then

    $$\begin{aligned} E(S\oplus T)= & {} \left[ E(S)\cap \rho (T)\right] \cup \left[ E(T)\cap \rho (S)\right] \cup \left[ E(S)\cap E(T)\right] , \end{aligned}$$

    where \(\rho (.)=\mathbb {\mathbb {C}}{\setminus }\sigma (.)\). On the other hand, since S and T satisfy property (k), i.e. \(\sigma (S){\setminus }\sigma _{w}(S)=E(S)\) and \(\sigma (T){\setminus }\sigma _{w}(T)=E(T),\) we then have

    $$\begin{aligned}&\left[ \sigma (S)\cup \sigma (T)\right] {\setminus } \left[ \sigma _{w}(S)\cup \sigma _{w}(T)\right] \\= & {} \left[ (\sigma (S){\setminus }\sigma _{w}(S))\cap \rho (T)\right] \cup \left[ (\sigma (T){\setminus }\sigma _{w}(T))\cap \rho (S)\right] \\&\cup \left[ (\sigma (S){\setminus }\sigma _{w}(S))\cap (\sigma (T){\setminus }\sigma _{w}(T))\right] \\= & {} \left[ E(S)\cap \rho (T)\right] \cup \left[ E(T)\cap \rho (S)\right] \cup \left[ E(S)\cap E(T)\right] . \end{aligned}$$

Hence \(E(S\oplus T)=[\sigma (S)\cup \sigma (T)]{\setminus }[\sigma _{w}(S)\cup \sigma _{w}(T)]= \sigma (S\oplus T){\setminus }\sigma _{w}(S\oplus T),\) and this shows that property (k) is satisfied by \(S\oplus T\). \(\square \)

Remark 2.11

  1. 1.

    Theorem 2.10 extends [9, Theorem 3.2] which proves that if \(T\in \mathcal {B}(\mathcal {H})\) and \(S\in \mathcal {B}({\mathcal K} )\) are isoloid operators acting on Hilbert spaces \(\mathcal {H}\) and \({\mathcal K} \) satisfying property (k) with the supplementary condition \(\sigma _{w}(S\oplus T)=\sigma _{w}(S)\cup \sigma _{w}(T),\) then \(S\oplus T\) satisfies property (k).

  2. 2.

    The assumption “S and T are isoloid” is essential in Theorem 2.10. For this define on \(\mathbb {\mathbb {C}}^n\oplus \ell ^2(\mathbb {N})\) the operator \(U=0\oplus S\) where S is defined by \( S(x_1, x_2, \ldots )=(0, \frac{x_1}{2}, \frac{x_2}{3}, \ldots ).\) It is clear that the null operator satisfies property (k). Also the operator S satisfies property (k),  since \(\sigma (S)=\sigma _{w}(S)=\{0\}\) and \(E(S)=\emptyset .\) But U does not satisfy this property, since \(\sigma (U)=\sigma _{w}(U)=\{0\}\) and \(E(U)=\{0\}.\) Here \(\sigma _{w}(0\oplus S)=\sigma _{w}(0)\cup \sigma _{w}(S)=\{0\},\) the null operator is isoloid, but S is not isoloid.

Before we state our next corollary as an application of Theorem 2.10 to the class of (H)-operators, we recall the definition of this class and definitions of some classes of operators which are contained in the class (H).

According to the monograph of Aiena [1], the quasinilpotent part \(H_0(T)\) of \(T\in \mathcal {B}(X)\) is defined as the set \(H_{0}(T)=\{x\in X: \displaystyle \lim _{n\rightarrow \infty }\Vert T^{n}(x)\Vert ^{\frac{1}{n}}=0\}.\) Note that generally, \(H_0(T)\) is not closed and from [1, Theorem 2.31] we have if \(H_0(T-\lambda I)\) is closed then T has SVEP at \(\lambda \). We also recall that T is said to belong to the class (H) if for all \(\lambda \in \mathbb {C}\) there exists \(p:=p(\lambda )\in \mathbb {N}\) such that \(H_{0}(T-\lambda I)=\mathcal {N}((T-\lambda I)^p),\) see Aiena [1] for more details about this class of (H)-operators. Of course, every operator T which belongs the class (H) has SVEP, since \(H_0(T-\lambda I)\) is closed. Observe also that \(a(T-\lambda I)\le p,\) for every \(\lambda \in \mathbb {C}.\) The class of operators having the property (H) is rather large. Obviously, it contains every operator having the property \((H_1).\) Recall that an operator \(T\in \mathcal {B}(X)\) is said to have the property \((H_1)\) if \(H_{0}(T-\lambda I)=\mathcal {N}(T-\lambda I)\) for all \(\lambda \in \mathbb {C}.\) Although the property \((H_1)\) seems to be rather strong, the class of operators having the property \((H_1)\) is considerably large. In the sequel we give some important classes of operators which satisfy property \((H_1).\) Every totally paranormal operator has property \((H_1),\) and in particular every hyponormal operator has property \((H_1).\) Also every transaloid operator or log-hyponormal has the property \((H_1).\) Some other operators satisfy property (H);  for example M-hyponormal operators, p-hyponormal operators, algebraically p-hyponormal operators, algebraically M-hyponormal operators, subscalar operators and generalized scalar operators. For more details about these definitions and comments which we cited above, we refer the reader to Aiena [1], Curto and Han [5] and, Laursen and Neumann [10].

Corollary 2.12

Let \(S\in \mathcal {B}(X)\) and \(T\in \mathcal {B}(Y)\) be isoloid operators and have a shared stable sign index. If S and T satisfy property (k), then \(S\oplus T\) satisfies property (k). In particular, if S and T are (H)-operators satisfying property (k) then \(S\oplus T\) satisfies property (k).

Proof

Assume that S and T are isoloid and satisfy property (k). Since S and T have a shared stable sign index, then it is easily seen that \(\sigma _{w}(S\oplus T)=\sigma _{w}(S)\cup \sigma _{w}(T).\) But this is equivalent by Theorem 2.10, to say that property (k) holds for \(S\oplus T.\) In particular if S and T are (H)-operators, then they are polaroid and so isoloid. But every (H)-operator has SVEP. Hence \(\text{ ind }(T-\lambda I)\) and \(\text{ ind }(S-\mu I)\) are less or equal than zero, for each \(\lambda \in \rho _{sbf}(T)\) and \(\mu \in \rho _{sbf}(S)\). Hence \(S\oplus T\) satisfies property (k). \(\square \)