1 Introduction

This paper concerns operators \(T\in L(X)\) that belong to the commutant of an injective quasi-nilpotent operator Q. The spectral properties of such operators, from the point of view of the classical Fredholm theory, have been studied in [7]. In this paper, we consider the spectra originating from the B-Fredholm theory in the sense of Berkani et al., and we show, if a such operator Q does exist, that the upper B-Weyl spectrum \(\sigma _{\textrm{ubw}}(T)\) coincides with the left Drazin spectrum \(\sigma _{\textrm{ld}}(T)\), while the B-Weyl spectrum \(\sigma _{\textrm{bw}}(T)\) coincides with the Drazin spectrum \(\sigma _{\textrm{d}}(T)\). These spectral equalities have, as a consequence, that the B-Weyl spectrum, as well as the upper B-Weyl spectrum, obeys to the spectral mapping theorem. Dually, assuming that the dual \(Q^*\) of a quasi-nilpotent operator is injective, then the lower B-Weyl spectrum \(\sigma _{\textrm{lbw}}(T)\) coincides with the right Drazin spectrum \(\sigma _{\textrm{rd}}(T)\). In this case, both the lower B-Weyl spectrum and the B-Weyl spectrum satisfy the spectral mapping theorem.

The second section of this article concerns some generalized version of Weyl type theorems, as the generalized version of a-Weyl’s theorem and the generalized version of property \((\omega )\). If T commutes with an injective quasi-nilpotent operator, then these generalized versions hold for T assuming that T is left polaroid, or that T is a-polaroid. Note that the generalized versions of Weyl type theorems are stronger than the classical versions, see [1, Chapter 5]. Another stronger variant of Weyl’s theorem, the so-called S-Weyl’s theorem, is also discussed in the last part.

Our results find a natural application to the operators which belong to the commutant of the Volterra V operator in \(L^2[a,b]\), since, as it is well known, both V and its adjoint \(V'\) are quasi-nilpotent and injective. Since the Volterra operator is also compact, our results complement each other with the celebrate Lomonosov result [15] that V admits a non-trivial closed invariant subspace.

2 Preliminaries and Definitions

Let L(X) denote the Banach algebra of all bounded linear operators acting on a complex Banach space X. If \(T\in L(X)\), by \(\alpha (T): =\text{ dim }\, \ker \, T\) and \(\beta (T): = \text{ codim }\, T(X)\), we denote the defects of T. The class of all upper semi-Fredholm operators is defined by

$$\begin{aligned} \Phi _+(X): =\{T \in L(X): \alpha (T)< \infty \ \text{ and }\ T(X) \ \text{ is } \text{ closed }\}, \end{aligned}$$

while the class all lower semi-Fredholm operators is defined by

$$\begin{aligned} \Phi _-(X): =\{T \in L(X): \beta (T)< \infty \}. \end{aligned}$$

The class of all semi-Fredholm operators is defined by \( \Phi _\pm (X): = \Phi _+(X) \cup \Phi _-(X)\). For every \(T\in \Phi _{\pm }(X)\), the index of T is defined by \(\text{ ind }\, T =\alpha (T) - \beta (T)\). The upper semi-Fredholm spectrum is defined by

$$\begin{aligned} \sigma _{\textrm{usf}}(T): =\{\lambda \in {\mathbb {C}}: \lambda I -T \notin \Phi _+(X)\}, \end{aligned}$$

and similarly, it is defined the lower semi-Fredholm spectrum \(\sigma _{\textrm{lsf}}(T)\). Recall that an operator \(T\in L(X)\) is said to be bounded below if is injective and has closed range. The classical approximate point spectrum is defined by

$$\begin{aligned} \sigma _{\textrm{ap}}(T): =\{\lambda \in {\mathbb {C}}: \lambda I-T \ \text{ is } \text{ not } \text{ bounded } \text{ below }\}, \end{aligned}$$

while the surjectivity spectrum is defined as

$$\begin{aligned} \sigma _{{\textrm{s}}}(T): =\{\lambda \in {\mathbb {C}}: \lambda I-T \ \text{ is } \text{ not } \text{ onto }\}. \end{aligned}$$

If \(T^*\) denotes the dual of T, it is well known that \(\sigma _{\textrm{ap}}(T)=\sigma _{\textrm{s}}(T^*)\) and \(\sigma _{\textrm{s}}(T)=\sigma _{\textrm{ap}}(T^*)\).

Denote by \(\Phi (X):= \Phi _+(X)\cap \Phi _-(X)\) the class of all Fredholm operators. An operator \(T\in L(X)\) is said to be a Weyl operator if \(T\in \Phi (X)\) and \(\text{ ind }\, T =0\), \(T\in L(X)\) is said to be upper semi-Weyl if \(T\in \Phi _+(X)\) and \(\text{ ind }\, T\le 0\), while \(T\in L(X)\) is said to be lower semi-Weyl if \(T\in \Phi _-(X)\) and \(\text{ ind }\, T\ge 0\). Denote by \(\sigma _{\textrm{w}}(T)\), \(\sigma _{\textrm{uw}}(T)\) and \(\sigma _{\textrm{lw}}(T)\), the Weyl spectrum, the upper semi-Weyl spectrum, and the lower semi-Weyl spectrum, respectively. Obviously, the inclusions

$$\begin{aligned} \sigma _{\textrm{usf}}(T) \subseteq \sigma _{\textrm{uw}}(T)\subseteq \sigma _{\textrm{ap}}(T) \quad \text{ and } \quad \sigma _{\textrm{lsf}}(T)\subseteq \sigma _{\textrm{lw}}(T)\subseteq \sigma _{\textrm{s}}(T) \end{aligned}$$

hold for every \(T\in L(X)\), and there is a duality

$$\begin{aligned}\sigma _{\textrm{uw}}(T)= \sigma _{\textrm{lw}}(T^*) \quad \text{ and }\quad \sigma _{\textrm{lw}}(T)= \sigma _{\textrm{uw}}(T^*). \end{aligned}$$

Recall that the ascent of \(T\in L(X)\) is the smallest positive integer \(p = p(T)\), whenever it exists, such that \(\ker \ T^{p}\) = \(\ker \ T^{p+1}\). If such p does not exist, we set \(p(T)= \infty \). Analogously, the descent of T is defined to be the smallest integer \(q = q(T)\), whenever it exists, such that \(T^{q+1}(X)\) = \(T^{q}(X)\). If such q does not exist, we set \(q(T)= \infty \). Note that if p(T) and q(T) are both finite, then \(p(T) = q(T)\), see Chapter 1 of [1] Moreover, \(\lambda \) is a pole of the resolvent if and only if \(0<p(\lambda I-T)= q(\lambda I-T)< \infty \); see [14, Proposition 50.2].

An operator \(T\in L(X)\) is said to be Browder if \(T\in \Phi (X)\) and \(p(T)=q(T)<\infty \). \(T\in L(X)\) is said to be upper semi-Browder if \(T\in \Phi _+(X)\) and \(p(T)< \infty \), while \(T\in L(X)\) is said to be lower semi-Browder if \(T\in \Phi _-(X)\) and \(q(T)< \infty \). Every Browder operator is Weyl, and every upper semi-Browder (respectively, lower semi-Browder) operator is upper semi-Weyl (respectively, lower semi-Weyl); see [1, Chapter 3].

The Browder spectrum, the upper semi-Browder spectrum, and the lower semi-Browder spectrum are denoted by \(\sigma _{\textrm{b}}(T)\), \(\sigma _{\textrm{ub}}(T)\), and \(\sigma _{\textrm{lb}}(T)\), respectively. Note that if \(\lambda \) is a spectral point for which \(\lambda I-T\) is Browder, then \(\lambda \) is an isolated point of \(\sigma (T)\). Recall that \(R\in L(X)\) is said to be a Riesz operator if \(\lambda I-T\in \Phi (X)\) for each \(\lambda \ne 0\). Quasi-nilpotent operators and compact operators are examples of Riesz operators. By a well-known result of Rakočević [16] (see also [3]), the Browder spectra are invariant under Riesz commuting perturbations R, that is

$$\begin{aligned} \sigma _{\textrm{ub}}(T)= \sigma _{\textrm{ub}}(T+R), \quad \sigma _{\textrm{lb}}(T)= \sigma _{\textrm{lb}}(T+R), \quad \sigma _{\textrm{b}}(T)= \sigma _{\textrm{b}}(T+R). \end{aligned}$$
(1)

Semi-Fredholm operators have been generalized by Berkani [8, 9] in the following way: for every \(T\in L(X)\) and a nonnegative integer n, let us denote by \(T_{[n]}\) the restriction of T to \(T^n(X)\), viewed as a map from the space \(T^n(X)\) into itself (we set \(T_{[0]} = T\)). \(T\in L(X)\) is said to be semi B-Fredholm (resp. B-Fredholm, upper semi B-Fredholm, lower semi B-Fredholm), if, for some integer \(n\ge 0\), the range \(T^n (X)\) is closed and \(T_{[n]}\) is a semi-Fredholm operator (resp. Fredholm, upper semi-Fredholm, lower semi-Fredholm). In this case, \(T_{[m]}\) is a semi-Fredholm operator for all \(m\ge n\) [8] with the same index of \(T_{[n]}\). This enables one to define the index of a semi B-Fredholm as \(\text{ ind } \ T = \text{ ind }\ T_{[n]}\). The upper semi B-Fredholm spectrum is defined

$$\begin{aligned} \sigma _{\textrm{ubf}}(T): = \{\lambda \in {\mathbb {C}}: \lambda I-T \ \text{ is } \text{ not } \text{ upper } \text{ semi } \text{ B-Fredholm }\}, \end{aligned}$$

and analogously, it may be defined the lower semi B-Fredholm spectrum \(\sigma _{\textrm{lbf}}(T)\).

A bounded operator \(T\in L(X)\) is said to be B-Weyl (respectively, upper semi B-Weyl, lower semi B-Weyl), if, for some integer \(n\ge 0\), the range \(T^n (X)\) is closed and \(T_{[n]}\) is Weyl (respectively, upper semi-Weyl, lower semi-Weyl). The B-Weyl spectrum is defined by

$$\begin{aligned} \sigma _{\textrm{bw}}(T): =\{\lambda \in {\mathbb {C}}: \lambda I-T \ \text{ is } \text{ not } \text{ B-Weyl }\}, \end{aligned}$$

and analogously may be defined the upper semi B-Weyl spectrum \(\sigma _{\textrm{ubw}}(T)\) and the lower semi B-Weyl spectrum \(\sigma _{\textrm{lbw}}(T)\).

Recall that an operator \(T\in L(X)\) is said to be Drazin invertible if \(p(T)=q(T)< \infty \), i.e., 0 is a pole of the resolvent or T is invertible. An operator \(T\in L(X)\) is said to be left Drazin invertible if \(p: =p(T)< \infty \) and \(T^{p+1}(X)\) is closed. A scalar \(\lambda \in {\mathbb {C}}\) is said to be a left pole if \(\lambda I-T\) is left Drazin invertible and \(\lambda \in \sigma _{\textrm{ap}}(T)\). A left pole \(\lambda \) for which \(\alpha (\lambda I-T)< \infty \) is said to have finite rank. Dually, \(T\in L(X)\) is said to be right Drazin invertible if \(q: = q(T)< \infty \) and \(T^q(X)\) is closed. A scalar \(\lambda \in {\mathbb {C}}\) is said to be a right pole if \(\lambda I-T\) is right Drazin invertible and \(\lambda \in \sigma _{\textrm{s}}(T)\).

The Drazin spectrum is defined by

$$\begin{aligned} \sigma _{\textrm{d}}(T): = \{\lambda \in {\mathbb {C}}: \lambda I-T \ \text{ is } \text{ not } \text{ Drazin } \text{ invertible }\}, \end{aligned}$$

and analogously are defined the left Drazin spectrum \(\sigma _{\textrm{ld}}(T)\) and the right Drazin spectrum \(\sigma _{\textrm{rd}}(T)\). It should be noted that there is a perfect duality, i.e., \(\sigma _{\textrm{ld}}(T)= \sigma _{\textrm{rd}}(T^*)\) and \(\sigma _{\textrm{ld}}(T^*)= \sigma _{\textrm{rd}}(T)\); see [1, Chapter 1]. The following inclusions hold for every operator \(T\in L(X)\):

$$\begin{aligned} \sigma _{\textrm{ubf}}(T) \subseteq \sigma _{\textrm{ubw}}(T) \subseteq \sigma _{\textrm{ld}}(T)\subseteq \sigma _{\textrm{ap}}(T), \end{aligned}$$
(2)

and

$$\begin{aligned} \sigma _{\textrm{lbf}}(T) \subseteq \sigma _{\textrm{lbw}}(T) \subseteq \sigma _{\textrm{rd}}(T)\subseteq \sigma _{\textrm{s}}(T); \end{aligned}$$
(3)

see [1, Chapter 1]. The following lemma has been proved in [4, Lemma 3.5].

Lemma 2.1

Suppose that \(\rho _{uw}(T)\) is connected. Then, \(\sigma _{\textrm{uw}}(T)= \sigma _{\textrm{w}}(T)\).

A proof of the following theorem can be found in [1, Theorem 1.140].

Theorem 2.2

Let \(T\in L(X)\). Then, T is left Drazin invertible (respectively, right Drazin invertible, Drazin invertible) if and only if there exists \(n\in {\mathbb {N}}\), such that \(T^n(X)\) is closed and the restriction \(T|T^n(X)\) is bounded below (respectively, onto, invertible).

3 Injective Quasi-Nilpotent Operators

Recall that an operator \(Q\in L(X)\) is said to be quasi-nilpotent if \(\sigma (Q)= \{0\}\). In the sequel by \(comm\,(T)\), we denote the commutant of T. A proof of the following result may be found in [7].

Theorem 3.1

Let \(T\in L(X)\) and suppose that \(Q\in \text{ comm }(T)\) is a quasi-nilpotent operator.

  1. (i)

    If Q is injective, then \(\alpha (T) < \infty \) if and only if T is injective.

  2. (ii)

    If the dual \(Q^*\) is injective, then \(\beta (T) < \infty \) if and only if T is onto.

In the sequel, we set

$$\begin{aligned} \textbf{Q}_i (X):=&\{ T\in L(X)&: \text{ there } \text{ exists } \text{ an } \text{ injective } \text{ quasi-nilpotent } \text{ operator }\ \\&Q\in L(X)&\text{ such } \text{ that } \ TQ=QT\}. \end{aligned}$$

Several examples of operators that commute with an injective quasi-nilpotent operator are given in [7]. Theorem 3.1 has some important consequences:

Theorem 3.2

Let \(T\in L (X)\) and \(Q\in L(X)\) a quasi-nilpotent operator that commutes with T.

  1. (i)

    If Q is injective, then

    $$\begin{aligned} \sigma _{\textrm{usf}}(T) =\sigma _{\textrm{uw}}(T) =\sigma _{\textrm{ub}}(T) = \sigma _{\textrm{ap}}(T) \quad \text{ and } \quad \sigma _{\textrm{b}}(T)= \sigma _{\textrm{w}}(T) = \sigma (T). \end{aligned}$$
    (4)
  2. (ii)

    If \(Q^*\) is injective, then

    $$\begin{aligned} \sigma _{\textrm{lsf}}(T) =\sigma _{\textrm{lw}}(T) =\sigma _{\textrm{lb}}(T) = \sigma _{\textrm{s}}(T) \quad \text{ and } \quad \sigma _{\textrm{b}}(T)= \sigma _{\textrm{w}}(T) = \sigma (T). \end{aligned}$$
    (5)

Proof

(i) Part (i) has been proved in [7, Corollary 3.7 and Theorem 3.8].

(ii) We have \(\sigma _{\textrm{lsf}}(T) \subseteq \sigma _{\textrm{lw}}(T) \subseteq \sigma _{\textrm{lb}}(T) \subseteq \sigma _{\textrm{s}}(T)\), and so, to show the first equalities in (5), it suffices to prove the inclusion \(\sigma _{\textrm{s}}(T) \subseteq \sigma _{\textrm{lsf}}(T)\). Let \(\lambda \notin \sigma _{\textrm{lsf}}(T)\) be arbitrary given. Then, \( \beta (\lambda I-T)< \infty \). Since \(T^*Q^*= Q^*T^*\), by Theorem 3.1, we have \( \beta (\lambda I-T)=0\), so \(\lambda \notin \sigma _{\textrm{s}} (T)\), and hence, \(\sigma _{\textrm{s}}(T) \subseteq \sigma _{\textrm{lw}}(T) \).

The equalities \(\sigma _{\textrm{b}}(T)= \sigma _{\textrm{w}}(T) = \sigma (T)\) may be proved in a similar way. \(\square \)

An operator \(T\in L(X)\), X a Banach space, is said to have the single-valued extension property at \(\lambda _0 \in {{\mathbb {C}}}\), in short T has the SVEP at \(\lambda _0\), if for every open disc \({{\textbf{D}}}_{\lambda _0}\) centered at \(\lambda _0\) the only analytic function \(f: {{\textbf{D}}}_{\lambda _0} \rightarrow X \) which satisfies the equation

$$\begin{aligned} (\lambda I - T) f(\lambda ) = 0 \quad \text{ for } \text{ all } \ \lambda \in {{\textbf{D}}}_{\lambda _0} \end{aligned}$$

is the constant function \(f \equiv 0\). T is said to have the SVEP if T has the SVEP for every \(\lambda \in {\mathbb {C}}\). Evidently, both T and \(T^*\) have SVEP at the points \(\lambda \notin \sigma (T)\).

Remark 3.3

Let \(\lambda _0 \in {\mathbb {C}}\) and suppose that T has SVEP at the points \(\lambda \) of a punctured open disc \({\mathbb {D}}(\lambda _0, \varepsilon ){\setminus }\{\lambda _0\}\). Then, T has SVEP at \(\lambda _0\). Indeed, suppose that \(f: {\mathbb {D}}(\lambda _0, \varepsilon ) \rightarrow X \) is an analytic function, such that \((\lambda I - T) f(\lambda ) = 0\) holds for every \(\lambda \in \mathbb D(\lambda _0, \varepsilon )\). Choose \(\mu \in {\mathbb {D}}(\lambda _0, \varepsilon ){\setminus } \{\lambda _0\}\) and let \({\mathbb {D}}(\mu , \delta )\) be an open disc contained in \({\mathbb {D}}(\lambda _0, \varepsilon ){\setminus } \{\lambda _0\}\). The SVEP for T at \(\mu \) entails \(f(\lambda )=0\) on \({\mathbb {D}}(\mu , \delta )\). Since f is continuous at \(\lambda _0\), we then conclude that \(f(\lambda _0) = 0\). Hence, \(f \equiv 0\) on \({\mathbb {D}}(\lambda _0, \varepsilon )\), and thus, T has the SVEP at \(\lambda _0\).

We now consider the Weyl spectra relative to the B-Fredholm theory.

Theorem 3.4

Let \(T\in L(X)\) and suppose that \(Q\in L(X)\) is a quasi-nilpotent operator, such that \(TQ=QT\).

  1. (i)

    If Q is injective, then

    $$\begin{aligned} \sigma _{\textrm{ubf}}(T) = \sigma _{\textrm{ubw}}(T) =\sigma _{\textrm{ld}}(T) \quad \text{ and } \quad \sigma _{\textrm{bw}}(T) = \sigma _{\textrm{d}}(T). \end{aligned}$$
    (6)
  2. (ii)

    If \(Q^*\) is injective, then

    $$\begin{aligned} \sigma _{\textrm{lbf}}(T) = \sigma _{\textrm{lbw}}(T) =\sigma _{\textrm{rd}}(T) \quad \text{ and } \quad \sigma _{\textrm{bw}}(T) = \sigma _{\textrm{d}}(T). \end{aligned}$$
    (7)

Proof

(i) To show the first equality it suffices, by (2), to prove the inclusion \(\sigma _{\textrm{ld}}(T) \subseteq \sigma _{\textrm{ubf}}(T)\). Let \(\lambda \notin \sigma _{\textrm{ubf}}(T)\) be arbitrarily given. There is no harm if we assume that \(\lambda =0\). Then, T is upper semi B-Fredholm, so \(T^n(X)\) is closed for some \(n\in {\mathbb {N}}\) and \(T_{[n]} =T|T^n(X)\) is upper semi-Fredholm, and hence, \(\alpha ( T_{[n]}) < \infty \) and \(T_{[n]}\) has a closed range. Now, from

$$\begin{aligned}Q(T^n(X)) = T^n (Q(X))\subseteq T^n(X),\end{aligned}$$

we see that \(T^n(X)\) is invariant under Q, so we can consider the restriction \(Q_{[n]} = Q|T^n(X)\), and, since T and Q commutes, we have the following:

$$\begin{aligned}T_{[n]} Q_{[n]} = Q_{[n]} T_{[n]}. \end{aligned}$$

Clearly, \(Q_{[n]}\) is injective and quasi-nilpotent. By Theorem 3.1, then \(\alpha (T_{[n]})=0\). Since \(T_{[n]}\) has closed range, then \(T_{[n]}\) is bounded below, so, by Theorem 2.2, T is left Drazin invertible, i.e., \(0\notin \sigma _{\textrm{ld}}(T)\). Therefore, the first equalities in (6) are proved.

The proof of the equality \(\sigma _{\textrm{bw}}(T)= \sigma _{\textrm{d}}(T)\) is analogous: if \(0 \notin \sigma _{\textrm{bw}}(T)\), then T is semi B-Weyl, so \(T^n(X)\) is closed for some \(n\in {\mathbb {N}}\) and the restriction \(T_{[n]}\) is Weyl, in particular \(\alpha ( T_{[n]} )= \beta ( T_{[n]})< \infty \). The restriction \(Q_{[n]}= Q|T^n(X)\) is injective and commutes with \(T_{[n]}\), so, always by Theorem 3.1, \(\alpha (T_{[n]})=\beta (T_{[n]} )=0\), and hence, \(T_{[n]}\) is invertible. By Theorem 2.2, then T is Drazin invertible, so \(0\notin \sigma _{\textrm{d}}(T)\). Therefore, \(\sigma _{\textrm{bw}}(T)= \sigma _{\textrm{d}}(T)\).

(ii) According the inclusions (3), to show the first equalities in (7), we need only to prove that \(\sigma _{\textrm{rd}}(T)\subseteq \sigma _{\textrm{lbf}}(T)\). Let \(\lambda _0 \notin \sigma _{\textrm{lbf}}(T)\). Then, \(\lambda _0 I-T\) is lower semi B-Fredholm, so, by [1, Theorem 1.117], there exists an open disc \({\mathbb {D}}_\varepsilon (\lambda _0)\) centered at \(\lambda _0\) and radius \(\varepsilon >0\), such that \(\lambda I-T\) is lower semi-Fredholm for all \(\lambda \in {\mathbb {D}}_\varepsilon (\lambda _0){\setminus } \{\lambda _0\}\). Since \(\lambda I-T\) commutes with Q, by part (ii) of Theorem 3.2, then \(\lambda I-T\) is lower semi-Browder, and hence, \(q(\lambda I-T)< \infty \) for all \(\lambda \in {\mathbb {D}}_\varepsilon (\lambda _0){\setminus } \{\lambda _0\}\), and this implies, by [1, Theorem 2.65], that \(T^*\) has SVEP at every \(\lambda \in {\mathbb {D}}_\varepsilon (\lambda _0){\setminus } \{\lambda _0\}\). From Remark 3.3, we then conclude that \(T^*\) has SVEP also at \(\lambda _0\). Since \(\lambda _0 I-T\) has topological uniform descent (see Chapter 1 of [1] for details), then, by Theorem 2.98 of [1], \(\lambda _0 I-T\) is right Drazin invertible, and hence, \(\lambda _0 \notin \sigma _{\textrm{rd}}(T)\). Therefore, \(\sigma _{\textrm{rd}}(T) \subseteq \sigma _{\textrm{lbf}}(T)\).

The proof of the second equality in (7) is similar: we have only to prove the inclusion \(\sigma _{\textrm{d}}(T) \subseteq \sigma _{\textrm{bw}}(T)\). Let \(\lambda _0 \notin \sigma _{\textrm{bw}}(T)\). Then, \(\lambda _0 I-T\) is semi-B-Weyl and, by [1, Theorem 1.117], there exists an open disc \({\mathbb {D}}_\varepsilon (\lambda _0)\) centered at \(\lambda _0\) and radius \(\varepsilon >0\), such that \(\lambda I-T\) is Weyl for all \(\lambda \in {\mathbb {D}}_\varepsilon (\lambda _0){\setminus } \{\lambda _0\}\). From part (ii) of Theorem 3.2, it then follows that \(\lambda I-T\) is Browder; hence, \(p(\lambda I-T)=q(\lambda I-T)< \infty \) for all \(\lambda \in \mathbb D_\varepsilon (\lambda _0){\setminus } \{\lambda _0\}\). This implies, see [1, Theorem 2.65], that both T and \(T^*\) have SVEP at every \(\lambda \in {\mathbb {D}}_\varepsilon (\lambda _0){\setminus } \{\lambda _0\}\). From Remark 3.3, we then deduce that both T and \(T^*\) have SVEP at \(\lambda _0\). Since \(\lambda _0 I-T\) has topological uniform descent, from Theorem 2.97 and Theorem 2.98 of [1], we then conclude that \(\lambda _0 I-T\) is both left and right Drazin invertible, i.e., Drazin invertible; hence, \(\lambda _0 \notin \sigma _{\textrm{d}}(T)\). \(\square \)

Example 3.5

An important example of injective quasi-nilpotent operator is provided by the classical Volterra operator V, on the Banach space X, where \(X: =C[0,1]\), the space of all continuous functions on the closed interval [0, 1], or \(X: =L^2[0,1]\), the Hilbert space of all complex-valued square-integrable functions on the interval [0, 1]. The operator V is defined by means

$$\begin{aligned} (Vf)(x): = \int _0^x f(t){\textrm{d}}t \quad \text{ for } \text{ all } \ f \in X \quad \text{ and } \ x\in [0,1]. \end{aligned}$$

The class of operators which commute with the Volterra operators is large; for instance, examples of operators which commute with V have been studied in the framework of supercyclic operators [17]. Note that the adjoint of the Volterra operator V on \(L^2[0,1]\) is given by

$$\begin{aligned} (V'f)(x): = \int _x^1 f(t){\textrm{d}}t. \end{aligned}$$

Evidently, also \(V'\) is injective and quasi-nilpotent; moreover, \(V'\) commutes with the adjoint of every operator which belongs to the commutant of V.

For operators T defined on a Hilbert spaces, H is better to consider the adjoint \(T'\) instead of the dual \(T^*\). We recall now the relationship between the Hilbert adjoint \(T'\) of an operator T defined on a Hilbert space and the dual \(T^*\). By the Frechét–Riesz representation theorem, there exists a conjugated-linear isometry \( U: H\rightarrow H^*\), \(H^*\) the dual of H, that associates to every \(y\in H\) the linear form defined

$$\begin{aligned}f_y(x):= <x, \,y> \quad \text{ for } \text{ every } \ x \in H.\end{aligned}$$

Moreover

$$\begin{aligned} ( {{\bar{\lambda }}} I - T') = U^{-1}(\lambda I -T^*)U \quad \text{ for } \text{ every }\ \lambda \in {\mathbb {C}}. \end{aligned}$$

Hence, for a Hilbert space operator T

$$\begin{aligned}T' \ \text{ is } \text{ injective } \Leftrightarrow \ T^*\ \text{ is } \text{ injective }. \end{aligned}$$

Since the adjoint \(V'\) of the Volterra operator V in \(L^2[0,1]\) is also injective and quasi-nilpotent, from Theorems 3.2 and 3.4, we then conclude the following:

Corollary 3.6

The equalities (4), (5), (6), and (7) hold for every bounded linear operator \(T:L^2[0,1]\rightarrow L^2[0,1]\) that belongs to the commutant of the Volterra operator V.

Let \(f\in {\mathcal {H}}(\sigma (T))\) be an analytic function defined on an open neighborhood U which contains the spectrum, and let f(T) be defined by the classical functional calculus

$$\begin{aligned} f(T): = \frac{1}{2\pi i}\int _\Gamma f(\lambda )(\lambda I-T)^{-1} \, {\textrm{d}} \lambda , \end{aligned}$$

where \(\Gamma \) is a contour that surrounds \(\sigma (T)\) in U. It is known that, in general, the spectral mapping theorem does not hold for \(\sigma _{\textrm{ubw}}(T)\) and \(\sigma _{\textrm{bw}}(T)\); indeed, we have only the inclusions

$$\begin{aligned}\sigma _{\textrm{ubw}}(f(T)) \subseteq f(\sigma _{\textrm{ubw}}(T)) \quad \text{ and }\quad \sigma _{\textrm{bw}}(f(T)) \subseteq f(\sigma _{\textrm{bw}}(T));\end{aligned}$$

see [1, Theorem 3.24], and these inclusions may be strict.

Let \({\mathcal {H}}_{\textrm{nc}}(\sigma (T))\) denote the subset of \({\mathcal {H}}(\sigma (T))\) of all functions nonconstant on each component of its domain of definition.

Theorem 3.7

Let \(T\in L(X)\) and suppose that there exists a quasi-nilpotent operator Q that commutes with T.

  1. (i)

    If Q is injective, then the spectral mapping theorem holds for \(\sigma _{\textrm{ubw}}(T)\) and \(\sigma _{\textrm{bw}}(T)\), i.e.,

    $$\begin{aligned}\sigma _{\textrm{ubw}}(f(T))= f(\sigma _{\textrm{ubw}}(T)) \quad \text{ for } \text{ all }\ f \in {\mathcal {H}}_{\textrm{nc}}(\sigma (T)),\end{aligned}$$

    and

    $$\begin{aligned}\sigma _{\textrm{bw}}(f(T))= f(\sigma _{\textrm{bw}}(T)) \quad \text{ for } \text{ all }\ f \in {\mathcal {H}}_{\textrm{nc}}(\sigma (T)).\end{aligned}$$
  2. (ii)

    If \(Q^*\) is injective, then the spectral mapping theorem holds for \(\sigma _{\textrm{lbw}}(T)\) and \(\sigma _{\textrm{bw}}(T)\), i.e.,

    $$\begin{aligned}\sigma _{\textrm{lbw}}(f(T))= f(\sigma _{\textrm{lbw}}(T)) \quad \text{ for } \text{ all } \ f \in {\mathcal {H}}_{\textrm{nc}}(\sigma (T)),\end{aligned}$$

    and

    $$\begin{aligned}\sigma _{\textrm{bw}}(f(T))= f(\sigma _{\textrm{bw}}(T)) \quad \text{ for } \text{ all } \ f \in {\mathcal {H}}_{\textrm{nc}}(\sigma (T)).\end{aligned}$$

Proof

(i) By Theorem 3.4, we know that \(\sigma _{\textrm{ubw}}(T)=\sigma _{\textrm{ld}}(T)\). The spectral mapping theorem holds for the left Drazin spectrum, since the set of all left Drazin invertible operators is a regularity; see [1, Theorem 3.109], hence

$$\begin{aligned} f(\sigma _{\textrm{ubw}}(T))= f(\sigma _{\textrm{ld}}(T))= \sigma _{\textrm{ld}}(f(T)). \end{aligned}$$

Let Q be an injective quasi-nilpotent operator Q which commutes with T. Evidently, if \(\lambda \in \rho (T):= {\mathbb {C}}{\setminus } \sigma (T)\), then Q commutes also with \((\lambda I-T)^{-1}\), and consequently, Q commutes with f(T). Therefore, always by Theorem 3.4, we have \(\sigma _{\textrm{ld}}(f(T))= \sigma _{\textrm{ubw}}(f(T))\), so the spectral mapping theorem holds for \(\sigma _{\textrm{ubw}}(T)\).

The spectral mapping theorem for the B-Weyl spectrum follows similarly, taking into account that, by Theorem 3.4, we have \(\sigma _{\textrm{bw}}(T)=\sigma _{\textrm{d}}(T)\) and that the spectral mapping theorem holds for \(\sigma _{\textrm{d}}(T)\).

(ii) Evidently, \(Q^*\) commutes with \(T^*\), and hence, with \(f(T^*)=f(T)^*\), so by Theorem 3.4, we have \(\sigma _{\textrm{rd}}(f(T))= \sigma _{\textrm{lbw}}(f(T))\). The spectral mapping theorem also holds for the right Drazin spectrum, since the set of all right Drazin invertible operators is a regularity, see [1, Theorem 3.109], hence

$$\begin{aligned} f(\sigma _{\textrm{lbw}}(T))= f(\sigma _{\textrm{rd}}(T))= \sigma _{\textrm{rd}}(f(T))= \sigma _{\textrm{lbw}}(f(T)). \end{aligned}$$

The spectral mapping theorem for \(\sigma _{\textrm{bw}}(T)\) is proved similarly using the equality \(\sigma _{\textrm{d}}(f(T))= \sigma _{\textrm{bw}}(f(T))\), proved in part (ii) of Theorem 3.4. \(\square \)

Corollary 3.8

If a bounded linear operator \(T:L^2[0,1] \rightarrow L^2[0,1] )\) commutes with the Volterra operator V, then the spectral mapping theorem holds for all B-Weyl spectra.

4 Weyl Type Theorems

An operator \(T\in L(X)\) is said to verify Weyl’s theorem if \(\sigma (T)\setminus \sigma _{\textrm{w}} (T)=\pi _{00}(T)\), where \(\pi _{00} (T): = \{\lambda \in \text{ iso }\, \sigma (T): 0< \alpha (\lambda I-T) < \infty \}\). The operator \(T\in L(X)\) is said to verify a-Weyl’s theorem if \(\sigma _{\textrm{ap}}(T)\setminus \sigma _{\textrm{uw}} (T)=\pi _{00}^a (T)\), where \(\pi _{00}^a (T): = \{\lambda \in \text{ iso }\, \sigma _{\textrm{ap}}(T): 0< \alpha (\lambda I-T) < \infty \}\), while \(T\in L(X)\) is said to verify property \((\omega )\) if \(\sigma _{\textrm{ap}}(T){\setminus } \sigma _{\textrm{uw}} (T)=\pi _{00} (T)\). Note that

$$\begin{aligned} \text{ either } a\text {-Weyl's theorem or property} \ (\omega ) \ \Rightarrow \text {Weyl' s theorem;} \end{aligned}$$

see [1, Chapter 6]. An operator \(T\in L(X)\) is said to verify generalized Weyl’s theorem (shortly, (gWt)), if \(\sigma _\mathrm{}(T){\setminus } \sigma _{\textrm{bw}} (T)=E(T)\), where \(E(T): = \{\lambda \in \text{ iso }\, \sigma (T): 0< \alpha (\lambda I-T) \}\). The operator \(T\in L(X)\) is said to verify generalized a-Weyl’s theorem (shortly, (gaWt)) if \(\sigma _{\textrm{ap}}(T){\setminus } \sigma _{\textrm{ubw}} (T)=E_a (T)\), where \(E_a (T): = \{\lambda \in \text{ iso }\, \sigma _{\textrm{ap}}(T): 0< \alpha (\lambda I-T) \}\), while \(T\in L(X)\) is said to verify generalized property \((\omega )\) (shortly, \((g\omega )\)) if \(\sigma _{\textrm{ap}}(T){\setminus } \sigma _{\textrm{ubw}} (T)=E (T)\). Note that (gaWt) entails a-Weyl’s theorem and generalized property \((\omega )\) entails property \((\omega )\). Furthermore

$$\begin{aligned} \text{ either }\ (gaWt)\ \text{ or } \ (g\omega ) \ \Rightarrow \ (gWt); \end{aligned}$$

see [1, Chapter 6]. In [7], it has been proved that the existence of an injective quasi-nilpotent operator that commutes with T ensures that a-Weyl’s theorem or property \((\omega )\) hold for T. It is a natural question if the generalized versions of Weyl type theorems hold for \(T\in \textbf{Q}_i (X)\). In the sequel, we shall prove that this is true under some additional conditions.

Let \(\text{ iso }\, F \) denote the isolated points of \(F\subseteq {\mathbb {C}}\). Recall that if Q is any quasi-nilpotent operator that commutes with T, then

$$\begin{aligned} \sigma _{\textrm{ap}}(T) = \sigma _{\textrm{ap}}(T+Q) \quad \text{ and } \quad \sigma _{\textrm{s}}(T) = \sigma _{\textrm{s}}(T+Q); \end{aligned}$$
(8)

see [1, Corollary 3.24]. Since \(\sigma (T)= \sigma _{\textrm{ap}}(T) \cup \sigma _{\textrm{s}}(T)\), this easily implies that \(\sigma (T)= \sigma (T+Q)\).

Theorem 4.1

Let \(T\in L(X)\) and suppose that Q is an injective quasi-nilpotent operator that commutes with T.

  1. (i)

    If \(\text{ iso }\, \sigma (T)= \emptyset \), then both T and \(T+Q\) satisfy generalized Weyl’s theorem.

  2. (ii)

    If \(\text{ iso }\, \sigma _{\textrm{ap}} (T)= \emptyset \), then both T and \(T+Q\) satisfy generalized a-Weyl’s theorem and generalized property \((\omega )\).

Proof

 

  1. (i)

    We show first that \(\sigma _{\textrm{d}}(T)=\sigma (T)\). It suffices to prove that \(\sigma (T)\subseteq \sigma _{\textrm{d }}(T)\). Let \(\lambda \notin \sigma _{\textrm{d}}(T)\). Then, \(\lambda I-T\) is Drazin invertible, and hence, \(\lambda I-T\) is either invertible or \(\lambda \) is a pole of T. If \(\lambda \) is a pole, then \(\lambda \) is an isolated point of \(\sigma (T)\), but this is impossible, since by assumption \(\text{ iso }\, \sigma (T)= \emptyset \). Hence, \(\lambda \notin \sigma (T)\), so \(\sigma _{\textrm{d}}(T)=\sigma (T)\). To show that T satisfies generalized Weyl’s theorem, observe that, by Theorem 3.4, \(\sigma (T){\setminus } \sigma _{\textrm{bw}}(T) =\sigma (T){\setminus } \sigma _{\textrm{d}}(T) = \emptyset \), and obviously, \(E(T)= \emptyset \), since there are no isolated points in \(\sigma (T)\). Since \(T+Q\) commutes with Q and \(\text{ iso }\, \sigma (T+Q)= \text{ iso }\,\sigma (T)= \emptyset \), the argument above proves that also \(T+Q\) satisfies generalized Weyl’s theorem.

  2. (ii)

    We show first that \(\sigma _{\textrm{ld}}(T)=\sigma _{\textrm{ap }}(T)\). Evidently, every bounded below operator is left Drazin invertible, so \(\sigma _{\textrm{ld}}(T)\subseteq \sigma _{{\textrm{ap }}}(T)\). To prove the opposite inclusion, let \(\lambda \notin \sigma _{\textrm{ld}}(T)\) be arbitrary given. Then, \(\lambda I-T\) is left Drazin invertible. There are two possibilities: \(\lambda \in \sigma _{\textrm{ap}} (T)\) or \(\lambda \notin \sigma _{\textrm{ap}} (T)\). If \(\lambda \in \sigma _{\textrm{ap}} (T)\), then \(\lambda \) is a left pole of T, and hence, by [1, Theorem 4.3], an isolated point of \(\sigma _{\textrm{ap}} (T)\), but this is impossible since \(\text{ iso }\, \sigma _{\textrm{ap}} (T)= \emptyset \). Hence, \(\lambda \notin \sigma _{\textrm{ap}} (T)\), so the equality \(\sigma _{\textrm{ld}}(T)=\sigma _{\textrm{ap }}(T)\) holds. Now, by Theorem 3.4,

    $$\begin{aligned}\sigma (T)\setminus \sigma _{\textrm{bw}}(T) =\sigma (T)\setminus \sigma _{\textrm{d}}(T) = \emptyset , \end{aligned}$$

    and obviously, \(E_a(T)= \emptyset \), since there is no isolated point in \(\sigma _{\textrm{ap }}(T)\); hence, T satisfies generalized Weyl’s theorem. Since \(T+Q\) commutes with Q and

    $$\begin{aligned}\text{ iso }\,\sigma _{\textrm{ap}} (T+Q)= \text{ iso }\, \sigma _{\textrm{ap}} (T)= \emptyset ,\end{aligned}$$

    the argument above proves that \(T+Q\) satisfies generalized a-Weyl’s theorem.

To prove the generalized property \((\omega )\) for T, observe, as above, that \(\sigma (T){\setminus } \sigma _{\textrm{bw}}(T) = \emptyset \). Furthermore, also \(E(T)= \emptyset \), since every isolated point \(\lambda \) of the spectrum belongs to \(\sigma _{\textrm{ap}} (T)\); see [1, Theorem 1.12], and hence, \(\lambda \in \text{ iso } \, \sigma _{\textrm{ap}} (T)\) and this is impossible. Therefore, also generalized property \((\omega )\) holds for T.

Since \(T+Q\) commutes with Q and \(\text{ iso }\, \sigma _{\textrm{ap}} (T+Q)= \emptyset \), then the previous reasoning shows that both generalized a-Weyl’s theorem and generalized property \((\omega )\) hold for \(T+Q\). \(\square \)

An operator \(T\in L(X)\) is said to be polaroid if every\(\lambda \in \text{ iso }\, \sigma (T)\) is a pole of the resolvent. \(T\in L(X)\) is said to be left polaroid if every \(\lambda \in \text{ iso }\, \sigma _{\textrm{ap}} (T)\) is left pole of the resolvent, while \(T\in L(X)\) is said to be a-polaroid if every \(\lambda \in \text{ iso }\, \sigma _{\textrm{ap}} (T)\) is pole of the resolvent. Note that

$$\begin{aligned} T \ a\text{-polaroid }\ \Rightarrow T \ \text{ left } \text{ polaroid } \Rightarrow T \ \text{ polaroid }; \end{aligned}$$

see [1, Corollary 4.13]. Recall that the spectral theorem holds for \(\sigma _{\textrm{ap}}(T)\).

Lemma 4.2

Suppose that \(T\in L(X)\) is polaroid (respectively, left polaroid, a-polaroid). Then, f(T) is polaroid (respectively, left polaroid, a-polaroid) for every \(f\in {\mathcal {H}}_{\textrm{nc}}(\sigma (T))\).

Proof

The case where T is polaroid is proved in [1, Theorem 4.19], while for the case where T is left polaroid; see [1, Remark 4.20]. Now, let T be a-polaroid and suppose that \(\lambda _0 \in \text{ iso }\, \sigma _{\textrm{ap}}(f(T))= \text{ iso }\, f(\sigma _{\textrm{ap}}(T))\). We show that \(\lambda _0\) is a pole of the resolvent of f(T). Let us first to show that \(\lambda _0 \in f(\text{ iso }\, \sigma _{\textrm{ap}} (T))\). Select \(\mu _0\in \sigma _{\textrm{ap}}(T)\), such that \(f(\mu _0)= \lambda _0\). Let \(\Omega \) be the connected component of the domain of f which contains \(\mu _0\) and suppose that \(\mu _0\) is not isolated in \(\sigma _{\textrm{ap}}(T)\). Then, there exists a sequence \((\mu _n) \subset \sigma _{\textrm{ap}}(T)\cap \Omega \) of distinct scalars, such that \(\mu _n \rightarrow \mu _0\). The set \(\Gamma : =\{\mu _0, \mu _1, \mu _2, \dots \}\) is a compact subset of \(\Omega \), so, by the Principle of isolated zeros of analytic functions, the function f may assume the value \(\lambda _0= f(\mu _0)\) at only a finite number of points of \(\Gamma \). Consequently, for n sufficiently large \(f(\mu _n)\ne f(\mu _0)= \lambda _0\), and since \(f(\mu _n)\rightarrow f(\mu _0)= \lambda _0\), it then follows that \(\lambda _0\) is not an isolated point of \(f(\sigma _{\textrm{ap}} (T))\), a contradiction.

Hence, \(\lambda _0 =f(\mu _0)\), with \(\mu _0 \in \text{ iso }\ \sigma _{\textrm{ap}}(T)\). Since T is a-polaroid, then \(\mu _0\) is a pole of the resolvent of T, and hence, by [1, Theorem 4.16], \(\lambda _0\) is a pole of the resolvent of f(T). Thus, f(T) is a-polaroid. \(\square \)

Theorem 4.3

Let \(T\in \textbf{Q}_i (X)\). Then, we have the following:

  1. (i)

    If T is polaroid, then f(T) satisfies generalized Weyl’s theorem for every \(f\in {\mathcal {H}}_{\textrm{nc}}(\sigma (T))\).

  2. (ii)

    If T is left polaroid, then f(T) satisfies generalized a Weyl’s theorem theorem for every \(f\in \mathcal H_{\textrm{nc}}(\sigma (T))\).

  3. (iii)

    If T is a-polaroid, then f(T) satisfies generalized property \((\omega )\) for every \(f\in {\mathcal {H}}_{\textrm{nc}}(\sigma (T))\).

Proof

 

  1. (i)

    Let T be polaroid. By Theorem 3.4, we have \(\sigma _{\textrm{bw}} (T)= \sigma _{\textrm{d}} (T)\). Hence

    $$\begin{aligned} \sigma (T){\setminus } \sigma _{\textrm{bw}} (T) = \sigma (T){\setminus } \sigma _{\textrm{d}} (T) = \Pi (T),\end{aligned}$$

    so every \(\lambda \in \sigma (T){\setminus } \sigma _{\textrm{bw}} (T)\) is a pole of the resolvent, hence an isolated point of \(\sigma (T)\) and an eigenvalue of T. Therefore, \(\lambda \in E (T)\) and this show that \(\sigma (T){\setminus } \sigma _{\textrm{bw}} (T) \subseteq E(T)\).

    The opposite inclusion is also true, if \(\lambda \in E(T)\) then \(\lambda \) is a pole, since T is polaroid, and so, by Theorem 3.2, \(\lambda \in \sigma (T){\setminus } \sigma _{\textrm{d}} (T)= \sigma (T){\setminus } \sigma _{\textrm{bw}} (T)\), and hence, (gW) holds for T. By Lemma 4.2, also f(T) is polaroid for every \(f\in {\mathcal {H}}_{\textrm{nc}}(\sigma (T))\), and since f(T) commutes with Q, then (gW) holds for f(T), by the first part of the proof.

  2. (ii)

    Let T be left polaroid. If \(\lambda \in E_a(T)\), then \(\lambda \in \text{ iso }\, \sigma _{\textrm{ap}}(T)\), and hence, \(\lambda \) is a left pole, since T is left polaroid; in particular, \(\lambda I-T\) is left Drazin invertible.Taking into account Theorem 3.2

    $$\begin{aligned} \lambda \in \sigma _{\textrm{ap}}(T)\setminus \sigma _{\textrm{ld}}(T)= \sigma _{\textrm{ap}}(T)\setminus \sigma _{\textrm{ubw}}(T),\end{aligned}$$

    so \(E_a(T)\subseteq \sigma _{\textrm{ap}}(T)\setminus \sigma _{\textrm{ubw}}(T)\).

    On the other hand, the opposite inclusion also holds. Indeed, if

    $$\begin{aligned}\lambda \in \sigma _{\textrm{ap}}(T){\setminus } \sigma _{\textrm{ubw}}(T)= \sigma _{\textrm{ap}}(T){\setminus } \sigma _{\textrm{ld}}(T)= \Pi _a(T),\end{aligned}$$

    then \(\lambda \in \text{ iso }\, \sigma _{\textrm{ap}}(T)\). Furthermore, \(\alpha (\lambda I-T)>0\); otherwise, if were \(\alpha (\lambda I-T) =0\), we would have \(p(\lambda I-T)= 0\), and hence, since \(\lambda \notin \sigma _{\textrm{ld}}(T)\), \((\lambda I-T)^{0+1}(X)= (\lambda I-T)(X)\) would be closed, contradicting the assumption that \(\lambda \in \sigma _{\textrm{ap}}(T)\). Therefore, \(\lambda \in E_a(T)\), so \(\sigma _{\textrm{ap}}(T)\setminus \sigma _{\textrm{ubw}}(T)=E_a(T)\); thus, generalized a-Weyl theorem holds for T.

    As above, the same argument shows that generalized a-Weyl theorem holds for f(T), since f(T) commutes with Q and f(T) is left polaroid, by Lemma 4.2.

  3. (iii)

    Observe first that E(T) is contained in \(\sigma _{\textrm{ap}}(T){\setminus } \sigma _{\textrm{ubw}}(T)\). Indeed, every isolated point of the spectrum belongs to its boundary and hence to \(\sigma _{\textrm{ap}}(T)\), by [1, Theorem 1.12]. Therefore, every \(\lambda \in E(T)\) is an isolated point of \(\sigma _{\textrm{ap}}(T)\), and hence, a pole of the resolvent, since by assumption T is a-polaroid; in particular, \(\lambda \) is a left pole. By Theorem 3.4, then

    $$\begin{aligned}\lambda \in \sigma _{\textrm{ap}}(T)\setminus \sigma _{\textrm{ld}}(T)= \sigma _{\textrm{ap}}(T)\setminus \sigma _{\textrm{ubw}}(T).\end{aligned}$$

To show the opposite inclusion \(\sigma _{\textrm{ap}}(T){\setminus } \sigma _{\textrm{ubw}}(T)\subseteq E(T)\), observe that if \(\lambda \in \sigma _{\textrm{ap}}(T){\setminus } \sigma _{\textrm{lbw}}(T)= \sigma _{\textrm{ap}}(T){\setminus } \sigma _{\textrm{ld}}(T)\), then \(\lambda \) is a left pole, hence an isolated point of \(\sigma _{\textrm{ap}}(T)\). The assumption that T is a-polaroid then entails that \(\lambda \) is a pole of T, hence an isolated point of \(\sigma (T)\). Since \(\lambda \) is also an eigenvalue of T, we then conclude that \(\lambda \in E(T)\), so property \((g\omega )\) holds for T. Since f(T) is a-polaroid then, as above, we conclude that property \((g\omega )\) holds for f(T). \(\square \)

Corollary 4.4

Let \(T\in L(L^2[0,1])\) commute with the Volterra operator and \(f\in {\mathcal {H}}_{\textrm{nc}} (\sigma (T))\). If T is polaroid, then f(T) satisfies generalized Weyl’s theorem. If T is left polaroid, then generalized a-Weyl’s theorem holds for f(T). If T is a-polaroid, then generalized property (w) holds for f(T).

Denote by \(\Pi (T)\) the set of all poles of T, i.e., \(\Pi (T):=\sigma (T){\setminus } \sigma _{\textrm{d}}(T)\) and by \(\Pi _a(T)\) the set of all left poles. i.e., \(\Pi _a(T):=\sigma _{\textrm{ap}} (T){\setminus } \sigma _{\textrm{ld}}(T)\). Note that \(\Pi _a(T) \subseteq \text{ iso }\, \sigma _{\textrm{ap}}(T)\) ( [1, Theorem 4.3]. According to [10] and [11], an operator \(T\in L(X)\) is said to satisfy property (gb) if \(\sigma _{\textrm{ap}}(T){\setminus } \sigma _{ubw}(T)= \Pi (T)\).

Theorem 4.5

Le \(T\in \textbf{Q}_i (X)\) be a-polaroid. Then, T satisfies property (gb).

Proof

From Theorem 3.4, we know that

$$\begin{aligned}\sigma _{\textrm{ap}}(T){\setminus } \sigma _{\textrm{ubw}}(T)= \sigma _{\textrm{ap}}(T) {\setminus } \sigma _{\textrm{ld}}(T)= \Pi _a(T).\end{aligned}$$

Obviously, the inclusion \(\Pi (T) \subseteq \Pi _a(T)\) holds for every operator. If \(\lambda \in \Pi _a(T)\), then \(\lambda \in \text{ iso } \, \sigma _{\textrm{ap}}(T)\), and since T is a-polaroid, then \(\lambda \in \Pi (T)\); hence, \(\Pi _a(T)\subseteq \Pi (T)\), and thus, \(\Pi _a(T)=\Pi (T)\). \(\square \)

To introduce another stronger variant of Weyl’s theorem, we introduce the following property.

Definition 4.6

An operator \(T\in L(X)\) is said to verify property (gaz) if \(\sigma (T)\setminus \sigma _{\textrm{ubw}}(T)= \Pi _a (T)\).

Theorem 4.7

Let \(T\in {\textbf{Q}}_i (X)\). Then, property (gaz) holds for T if and only if \(\sigma _{\textrm{ap}} (T)= \sigma (T)\).

Proof

Suppose that property (gaz) holds for T. Then, \(\sigma _{\textrm{ap}} (T)= \sigma (T)\), by Theorem 3.2 of [5]. Conversely, suppose that \(\sigma _{\textrm{ap}} (T)= \sigma (T)\). Since \(T\in \textbf{Q}_i (X)\), then \(\sigma _{\textrm{ubw}} (T) = \sigma _{\textrm{ld}} (T)\), by Theorem 3.2, so we have

$$\begin{aligned} \sigma (T){\setminus } \sigma _{\textrm{ubw}}(T)= \sigma _{\textrm{ap }} (T){\setminus } \sigma _{\textrm{ld}}(T)= \Pi _a(T), \end{aligned}$$

i.e., (gaz) holds for T. \(\square \)

It should be noted that the equality \(\sigma _{\textrm{ap}} (T)= \sigma (T)\) holds if \(T^*\) has SVEP, and hence, if \(T^\star \) has SVEP, then property (gaz) holds for T.

The following variant of Weyl type theorems has been introduced in [6]. An operator \(T\in L(X)\) is said to satisfy S-Weyl’s theorem, if \(\sigma (T)\setminus \sigma _{\textrm{ubw}}(T)= \pi _{00}(T)\). Note that S-Weyl’s theorem holds for T if and only if T satisfies property (gaz) and the equality \(\Pi _a (T)= \pi _{00}(T)\) holds; see [6, Theorem 3.6]. Furthermore, S-Weyl’s theorem entails both a-Weyl’s theorem and property \((\omega )\); see [6, Theorem 3.8].

Observe that for operators \(T \in \textbf{Q}_i (X)\), we have

$$\begin{aligned}\pi _{00}(T)= \pi _{00}^a(T)= \emptyset , \end{aligned}$$

since, if \(\lambda \) belongs to one of these sets, then \(0<\alpha (\lambda I-T)< \infty \) and this is impossible by Theorem 3.1.

Corollary 4.8

Let \(T\in \textbf{Q}_i (X)\) be such that \(\sigma _{\textrm{ap}} (T)= \sigma (T)\). Then, S-Weyl’s theorem holds for T if and only if T has no left poles. In particular, if \(\text{ iso }\, \sigma _{\textrm{ap}}(T)= \emptyset \), then S-Weyl’s theorem holds for T.

Proof

By Theorem 4.3, if \(\sigma _{\textrm{ap}} (T)= \sigma (T)\), then T has property (gaz). Furthermore, as noted above, \(\pi _{00}(T)= \emptyset \). Hence

$$\begin{aligned} \emptyset = \Pi _a(T)=\pi _{00}(T); \end{aligned}$$

thus, S-Weyl’s theorem holds for T. Conversely, if S-Weyl’s theorem holds for T, then \(\Pi _a(T)=\pi _{00}(T) =\emptyset \), so T has no left poles.

The last assertion is clear: every left pole of the resolvent of an operator is an isolated point of \(\sigma _{\textrm{ap}}(T)\); see [1, Theorem 4.3]. \(\square \)

Theorem 4.9

Suppose that \(\sigma _{\textrm{ap}}(T)\) has no hole and that \(\sigma _{\textrm{ap}}(T)\) has no isolated point. If Q is an injective quasi-nilpotent operator that commutes with T, then S-Weyl’s theorem holds for T and \(T+Q\).

Proof

Since \(\sigma _{\textrm{ap}}(T)= \sigma _{\textrm{uw}}(T)\), by Theorem 3.2, the set \(\rho _{\textrm{uw}} (T):= {\mathbb {C}} {\setminus } \sigma _{\textrm{uw}}(T)\) is connected, so, by Lemma 2.1 and Theorem 3.2, we have

$$\begin{aligned} \sigma _{\textrm{ap}}(T)=\sigma _{\textrm{uw}}(T)= \sigma _{\textrm{w}} (T)= \sigma (T). \end{aligned}$$

Therefore, T has property (gaz), by Theorem 4.3. To prove that S-Weyl’s theorem holds for T, we need to prove that \(\Pi _a (T)= \pi _{00}(T)\). As noted before, \(\pi _{00}(T)\) is empty and also \(\Pi _a(T)= \emptyset \), since \(\Pi _a(T)\subseteq \text{ iso }\, \sigma _{\textrm{ap}}(T)\), and thus, T satisfies S-Weyl’s theorem. To show that S-Weyl’s theorem holds also for \(T+Q\), note that \(T+Q\) commutes with Q, \(\text{ iso }\, \sigma _{\textrm{ap}}(T+Q)=\text{ iso }\, \sigma _{\textrm{ap}}(T) = \emptyset \) and \(\sigma _{\textrm{ap}}(T+Q)= \sigma _{\textrm{ap}}(T)\) has no hole. \(\square \)