1 Introduction and preliminaries

Let \(T\in L(X)\) be a bounded linear operator on an infinite-dimensional complex Banach spaces X,  and denote by \(\alpha (T)\) and \(\beta (T)\) the dimension of the kernel N(T) and the codimension of the range \(R(T)=T(X),\) respectively. Let \(\Phi _+(X):=\{ T\in L(X:\alpha (T)<\infty , T(X) \ \text{ is } \text{ closed } \}\) be the class of all upper semi-Fredholm operators and \(\Phi _-(X):=\{ T\in L(X): \beta (T)<\infty \} \) the class of all lower semi-Fredholm operators. If \(T\in \Phi _-^+(X):=\Phi _+(X) \cup \Phi _-(X),\) the index of T is defined by \(ind(T):=\alpha (T)-\beta (T).\) If \(\Phi (X):=\Phi _+(X) \cap \Phi _-(X),\) denotes the set of all Fredholm operators, the class of Weyl operators is defined by

$$\begin{aligned} W(X)=\{ T\in \Phi (X): ind(T)=0\}, \end{aligned}$$

the class of upper semi-Weyl operators is defined by

$$\begin{aligned} W_+(X)=\{ T\in \Phi _+(X): ind(T)\leqslant 0\}, \end{aligned}$$

while the class lower semi-Weyl operators is defined by

$$\begin{aligned} W_-(X)=\{ T\in \Phi _-(X): ind(T)\geqslant 0\}. \end{aligned}$$

Evidently, \(W(X)= W_-(X)\cap W_+(X)\). If \(T^*\) denotes the dual of \(T\in L(X)\), it is well known that \(T\in W_+(X)\) (respectively, \(T\in W_-(X)\)) if and only if \(T^*\in W_-(X^*)\) (respectively, \(T^*\in W_+(X^*)\)). The classes of operators above defined generate the following spectra: the Weyl spectrum, defined by

$$\begin{aligned} \sigma _{w}(T):= \{ \lambda \in {\mathbb {C}}: \lambda I -T \not \in W(X)\}, \end{aligned}$$

the upper semi-Weyl spectrum, defined by

$$\begin{aligned} \sigma _{uw}(T):= \{ \lambda \in {\mathbb {C}}: \lambda I-T \not \in W_+(X)\}, \end{aligned}$$

and the lower semi-Weyl spectrum, defined by

$$\begin{aligned} \sigma _{lw}(T):= \{ \lambda \in {\mathbb {C}}: \lambda I-T \not \in W_-(X)\}. \end{aligned}$$

Two classical quantities in operator theory are defined as follows. The ascent of an operator T,  is the smallest non-negative integer \(p: = p(T)\) such that \(N( T^p)=N( T^{p+1}).\) If such integer does not exist we put \(p(T)=\infty .\) Analogously, the descent of T,  is the smallest non-negative integer \(q:=q(T)\) such that \(R(T^q)=R(T^{q+1}),\) and if such integer q does not exist we put \(q(T)=\infty .\) It is well known that if p(T) and q(T) are both finite then \(p(T)=q(T),\) see [1, Theorem 1.20]. Moreover, \(0<p(\lambda I-T) = q(\lambda I-T)< \infty \) exactly when \(\lambda \) is a pole of the resolvent of T, see [14, Proposition 50.2]. The class of all Browder operators is defined

$$\begin{aligned} B(X):=\{ T\in \Phi (X): p(T)=q(T)<\infty ,\}, \end{aligned}$$

while the class of all upper semi-Browder operators is defined

$$\begin{aligned} B_+(X):=\{ T\in \Phi _+(X): p(T)<\infty ,\}. \end{aligned}$$

Obviously, \(B(X)\subseteq W(X)\) and \(B_+(X)\subseteq W_+(X),\) see [1, Chapter 3]. If \(\sigma _{b}(T)\) and \(\sigma _{ub}(T)\) denote the Browder spectrum and the upper semi-Browder spectrum, respectively, then \(\sigma _{w}(T)\subseteq \sigma _{b}(T)\) and \(\sigma _{uw}(T)\subseteq \sigma _{ub}(T)\).

In [6, 8, 10] Berkani et al. generalize semi-Fredholm operators 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\(=T_{[0]}\)). \(T\in L(X)\) is said to be semi B-Fredholm,(respectively, B-Fredholm, upper semi B-Fredholm, lower semi B-Fredholm) if, for some integer n, the range \(T^n(X)\) is closed and \(T_{[n]}\) is a semi-Fredholm operator (resp. Fredholm, upper semi-Fredholm, lower semi-Fredholm). If \(T_{[n]}\) is a semi-Fredholm operator, then \(T_{[n]}\) is a semi-Fredholm operator for all \(m\geqslant n\) ( [10]), with the same index of \(T_{[n]}.\) This enables one to define the index of a semi-B-Fredholm as \(ind(T)=ind(T_{[n]}).\) Analogously, 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 \(n\geqslant 0\) the range \(T^n(X)\) is closed and \(T_{[n]}\) is Weyl (respectively, upper semi-Weyl, lower semi-Weyl ). Analogous definitions are given for semi B-Browder operators. The B-Weyl spectrum is defined as

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

and analogously, the upper semi B-Weyl spectrum of T is defined by

$$\begin{aligned} \sigma _{ubw}(T):= \{ \lambda \in {\mathbb {C}}: \lambda I -T\ is \ not \ upper \ semi \ B-Weyl \}. \end{aligned}$$

The concept of Drazin invertibility has been introduced in a more abstract setting than operator theory. In the case the Banach algebra L(X),  an operator \(T\in L(X)\) is said to be Drazin if \(p(T)=q(T)<\infty \). Evidently, if T is Drazin invertible then either \(\lambda I-T\) is invertible or \(\lambda \) is a pole of the resolvent of T. An operator \(T\in L(X)\) is said to be left Drazin invertible if \(p=p(T)<\infty \) and \(R(T^{p+1})\) is closed. The Drazin spectrum is then defined as

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

while the left Drazin spectrum is defined as

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

In the sequel we denote by \(\sigma _a(T)\) the approximate point spectrum, defined by \(\sigma _a(T)=\{ \lambda \in {\mathbb {C}}:\lambda I-T\) is not bounded below}, where an operator is said to be bounded below if it is injective and has closed range. The classical surjective spectrum of T is denoted by \(\sigma _s(T).\)

The following property has a fundamental role in local spectral theory. An operator \(T\in L(X)\) 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 \(T\in L(X)\) is said to have the SVEP if T has this property at every \(\lambda \in {\mathbb {C}}\). (See [1] and [15] for more details about this concept). Evidently, an operator \(T\in L(X)\) has SVEP at every point of the resolvent \(\rho (T)={\mathbb {C}}\setminus \sigma (T),\) and both T and \(T^*\) have SVEP at the isolated points of the spectrum. Note that

$$\begin{aligned} p(\lambda I- T)<\infty \Longrightarrow T \ \text{ has } \text{ SVEP } \text{ at } \ \lambda , \end{aligned}$$
(1)

and dually,

$$\begin{aligned} q(\lambda I- T)<\infty \Longrightarrow T^*\ \text{ has } \text{ SVEP } \text{ at } \ \lambda , \end{aligned}$$
(2)

see [1, Chapter 2]. Furthermore, if \(\text{ acc } \, F\) denote the set of all cluster points of \(F\subseteq {\mathbb {C}}\), we have:

$$\begin{aligned} \lambda \notin \text{ acc }\,\sigma _a(T)\Longrightarrow T \ \text{ has } \text{ SVEP } \text{ at } \ \lambda , \end{aligned}$$
(3)

and dually,

$$\begin{aligned} \lambda \notin \text{ acc }\,\sigma _s(T)\Longrightarrow T^*\ \text{ has } \text{ SVEP } \text{ at } \ \lambda , \end{aligned}$$
(4)

see [1, Chapter 2].

Remark 1.1

In [1, Chapter 2] it is shown that the implications above are equivalences if \(\lambda I-T\) semi B-Fredholm, in particular semi-Fredholm.

2 Property \(( UW {\scriptstyle \Pi })\)

Denote by \(p_{00}(T):=\sigma (T)\setminus \sigma _b(T)\) the set of all poles of T having finite rank, by \(\Pi (T)=\sigma (T)\setminus \sigma _d(T)\) the set of all poles of T, and by \(p_{00}^a(T)=\sigma _a(T)\setminus \sigma _{ub}(T)\) the set of all left poles of T having finite rank. It is easy to check that \(p_{00}(T)\subseteq p_{00}^a(T)\) for all \(T\in L(X),\) and obviously every point of \(p_{00}(T)\) is an isolated point of \(\sigma (T)\), and hence an isolated point of \(\sigma _a(T)\) (since every isolated point of the spectrum belongs to \(\sigma _a (T)\)). Set

$$\begin{aligned} \Delta (T): = \sigma (T)\setminus \sigma _{w}(T) \quad \text{ and }\quad \Delta _a (T): = \sigma _a(T)\setminus \sigma _{uw}(T). \end{aligned}$$

Let \(\Pi _a(T): = \sigma _a(T)\setminus \sigma _{ld}(T)\) be the set of all left poles of the resolvent of T. Obviously, \(p_{00}(T) \subseteq \Delta _a (T)\), since each point of \(p_{00}(T)\) is an eigenvalue of T and every Browder operator is upper semi-Weyl.

Hereafter, the symbol \(\bigsqcup \) stands for disjoint union.

Definition 2.1

A bounded operator \(T\in L(X)\) is said to satisfy:

  1. (i)

    Browder’s theorem if \(\sigma _w(T)=\sigma _b(T)\) or equivalently \(\sigma (T)\setminus \sigma _w(T)= p_{00}(T)\), ([13]).

  2. (ii)

    a-Browder’s theorem if \(\sigma _{uw}(T)=\sigma _{ub}(T)\) or equivalently \(\Delta _a(T)= p_{00}^a(T)\), ([13, 1, Chapter 5]).

  3. (iii)

    Property \((UW {\scriptstyle \Pi })\) if \(\Delta _a(T)= \Pi (T),\) or equivalently \(\sigma _a (T)= \sigma _{uw}(T) \bigsqcup \Pi (T)\), ( [2, 9]).

Remark 2.2

a-Browder’s theorem entails Browder’s theorem. Furthermore, a-Browder’s theorem is equivalent to generalized a-Browder’s theorem, i.e., a-Browder’s theorem is equivalent to saying that the equality \(\sigma _{ubw}(T)= \sigma _{ld}(T)\) holds, where \(\sigma _{ubw}(T)\) denotes the upper semi B-Weyl spectrum, see [5].

Property \((UW {\scriptstyle \Pi })\) may be characterized as follows ([2]):

Theorem 2.3

Let \(T\in L(X).\) Then the following assertions are equivalent:

  1. (i)

    T satisfies property \((UW {\scriptstyle \Pi });\)

  2. (ii)

    T satisfies a-Browder’s theorem, and \(p_{00}^a(T)=\Pi (T)\);

  3. (iii)

    \(T^*\) has SVEP at every \(\lambda \in \Delta _a(T)\) and \(\sigma _b (T)= \sigma _d (T)\).

Property \((UW {\scriptstyle \Pi })\) entails some relevant equalities between parts of the spectrum:

Theorem 2.4

Suppose that T satisfies property \(( UW {\scriptstyle \Pi })\). Then we have:

  1. (i)

    \(\sigma _{w}(T) \setminus \sigma _{uw}(T)= \sigma (T)\setminus \sigma _a(T)\).

  2. (ii)

    \(\sigma _{uw}(T) \setminus \sigma _{ubw}= \Pi _a(T)\setminus \Pi (T)\).

  3. (iii)

    \(\sigma (T)\setminus \sigma _w(T)= \Pi (T)\).

Proof

If \(\lambda \in \sigma _{w}(T) \setminus \sigma _{uw}(T)\) then \(\alpha (\lambda I-T)< \infty \) and \((\lambda I-T)(X)\) is closed. Suppose that \(0< \alpha (\lambda I-T)\). Then \(\lambda \in \sigma _{a}(T) \setminus \sigma _{uw}(T) = \Pi (T)\). Hence, \(p(\lambda I-T)= q(\lambda I-T)< \infty \) and consequently, by [1, Theorem 1.22], \(\lambda I -T\) is Browder, in particular \(\lambda \notin \sigma _{w}(T)\), a contradiction. Hence \(\alpha (\lambda I-T)= 0\), so \(\lambda \notin \sigma _a (T)\), and consequently \(\lambda \in \sigma (T)\setminus \sigma _a(T)\).

Conversely, if \(\lambda \in \sigma (T)\setminus \sigma _a(T)\) then \(\lambda \notin \sigma _{uw}(T)\), since \(\sigma _{uw}(T)\subseteq \sigma _a (T)\). Suppose that \(\lambda \notin \sigma _w(T)\). Then \(\alpha (\lambda I-T)= \beta (\lambda I-T)\), and since \(\alpha (\lambda I-T)=0\) we have \(\lambda \notin \sigma (T)\), a contradiction. Hence \(\lambda \in \sigma _{w}(T) \setminus \sigma _{uw}(T)\).

(ii) Suppose that T has property \(( UW {\scriptstyle \Pi })\). Then a-Browder’s theorem holds for T, or equivalently generalized a-Browder’s theorem holds for T. If \(\lambda \in \sigma _{uw}(T) \setminus \sigma _{ubw}\) then \(\lambda \notin \sigma _{ubw} (T) =\sigma _{ld}(T)\), and \(\lambda \in \sigma _a (T)\), since \(\sigma _{uw}(T)\subseteq \sigma _a(T)\). Hence, \(\lambda \in \Pi _a (T)\). Since a-Browder’s theorem holds for T we then have \(\lambda \in \sigma _{ub}(T)\), hence \(\lambda \notin \sigma _a (T)\setminus \sigma _{ub}(T)= \pi _{00}^a\), and hence \(\lambda \notin \Pi (T)\), by Theorem 2.3.

Conversely, suppose that \(\lambda \in \Pi _a(T)\setminus \Pi (T)\). Then \(\lambda \in \Pi _a(T)\) and hence \(\lambda \notin \sigma _{ld}(T)= \sigma _{ubw}(T)\), since generalized a-Browder’s theorem holds for T. On the other hand, \(\lambda \notin \Pi (T)= p_{00}^a (T)\), by Theorem 2.3, and since \(\lambda \in \sigma _a(T)\), then \(\lambda \in \sigma _{ub}(T)= \sigma _{uw}(T)\).

(iii) Write \(\sigma (T)= \sigma (T)\setminus \sigma _a(T) \bigsqcup \sigma _a (T)\). Since \(\sigma _a (T)= \sigma _{uw}(T) \bigsqcup \Pi (T)\), by part (i) we then have

$$\begin{aligned} \sigma (T)= & {} \sigma _{w}(T) \setminus \sigma _{uw}(T) \bigsqcup \sigma _a (T) \\= & {} \sigma _{w}(T) \setminus \sigma _{uw}(T) \bigsqcup \sigma _{uw}(T) \bigsqcup \Pi (T) = \sigma _{w}(T)\bigsqcup \Pi (T), \end{aligned}$$

so \(\sigma (T)\setminus \sigma _w(T)= \Pi (T)\). \(\square \)

The following variant of Weyl’s theorem and Browder’s theorem has been introduced recently by Zariouh [19]. An operator \(T\in L(X)\) is said to satisfies property\((Z_{\Pi _a})\) if \(\Delta (T): = \sigma (T)\setminus \sigma _w(T)\) coincides with the set \(\Pi _a(T)\) of left poles of T.

The next theorem improves a result of [2], which was proved by assuming that \(T^*\) has SVEP.

Theorem 2.5

If \(T\in L(X)\) satisfies property \((UW {\scriptstyle \Pi }),\) then T satisfies property \((Z{\scriptstyle \Pi _a}).\)

Proof

Suppose that T satisfies property \((UW {\scriptstyle \Pi })\), i.e. \(\sigma _a (T)= \sigma _{uw}(T) \bigsqcup \Pi (T)\). By part (i) and part (iii) of Theorem 2.4 we have

$$\begin{aligned} \sigma (T)= & {} \sigma _w (T)\bigsqcup \Pi (T)= (\sigma _w (T)\setminus \sigma _{uw}(T)) \bigsqcup \sigma _{uw}(T) \bigsqcup \Pi (T)\\= & {} (\sigma (T)\setminus \sigma _{a}(T)) \bigsqcup \sigma _{uw}(T) \bigsqcup \Pi (T)\\= & {} (\sigma (T)\setminus \sigma _{a}(T)) \bigsqcup [(\sigma _{uw}(T)\setminus \sigma _{ubw}(T))\bigsqcup \sigma _{ubw}(T)] \bigsqcup \Pi (T). \end{aligned}$$

From part (ii) of Theorem 2.4 we then obtain

$$\begin{aligned} \sigma (T)= & {} (\sigma _w (T)\setminus \sigma _{uw}(T)) \bigsqcup (\Pi _a (T)\setminus \Pi (T)) \bigsqcup \sigma _{ubw}(T) \bigsqcup \Pi (T)\\= & {} (\sigma _w (T)\setminus \sigma _{uw}(T)) \bigsqcup \Pi _a (T)\bigsqcup \sigma _{ubw}(T). \end{aligned}$$

Since \(\sigma _{ubw}(T)\subseteq \sigma _{uw}(T)\) then

$$\begin{aligned} \sigma (T) \subseteq (\sigma _w (T)\setminus \sigma _{uw}(T)) \bigsqcup \Pi _a (T) \bigsqcup \sigma _{uw}(T) = \sigma _w (T) \bigsqcup \Pi _a (T). \end{aligned}$$

Trivially, \(\sigma _w (T) \bigsqcup \Pi _a (T) \subseteq \sigma (T)\) Consequently, \(\sigma (T)= \sigma _w (T) \bigsqcup \Pi _a (T)\) and hence property \((Z{\scriptstyle \Pi _a})\) holds for T. \(\square \)

The converse of Theorem 2.5 holds if \(T^*\) has SVEP, see [2]. The precise relationship between property \((UW {\scriptstyle \Pi })\) and property \((Z{\scriptstyle \Pi _a})\) is described by the following theorem.

Theorem 2.6

Let \(T \in L(X)\). Then the following statements are equivalent:

  1. (i)

    T satisfies property \((UW {\scriptstyle \Pi });\)

  2. (ii)

    T satisfies property \((Z{\scriptstyle \Pi _a})\) and \(\sigma _{w}(T)\setminus \sigma _{uw}(T)=\sigma (T)\setminus \sigma _{a}(T);\)

  3. (iii)

    T satisfies property \((Z {\scriptstyle \Pi _a})\) and \(\Delta _a (T)\cap \sigma _w (T)= \emptyset \).

Proof

(i) \(\Longleftrightarrow \) (ii) Suppose that T satisfies property \((UW {\scriptstyle \Pi }).\) Then T has property \((Z{\scriptstyle \Pi _a})\) and, by Theorem 2.4, we have \(\sigma _{w}(T)\setminus \sigma _{uw}(T)=\sigma (T)\setminus \sigma _{a}(T).\) Conversely, if T has property \((Z{\scriptstyle \Pi _a})\) and \(\sigma _{w}(T)\setminus \sigma _{uw}(T)=\sigma (T)\setminus \sigma _{a}(T)\), we have \(\sigma (T)=\sigma _w(T)\bigsqcup \Pi _a(T).\) Our assumption \(\sigma _{w}(T)\setminus \sigma _{uw}(T)=\sigma (T)\setminus \sigma _{a}(T)\) entails that \(\sigma (T) =[\sigma (T)\setminus \sigma _a(T)] \bigsqcup [\sigma _{uw}(T) \bigsqcup \Pi _a(T)],\) hence \(\sigma _a(T)=\sigma _{uw}(T) \bigsqcup \Pi _a(T).\) From [19, Lemma 2.9] we have that \(\Pi _a(T)=\Pi (T),\) so \(\sigma _a(T)= \sigma _{uw}(T) \bigsqcup \Pi (T),\) hence T satisfies property \((UW {\scriptstyle \Pi }).\)

(i) \(\Longleftrightarrow \) (iii) Suppose that T satisfies property \((UW {\scriptstyle \Pi }).\) Then T has property \((Z{\scriptstyle \Pi _a})\). Let \(\lambda \in \Delta _a (T)=\Pi (T)\). Then \(\lambda I-T\in W_+(X)\) and since \(p(\lambda I-T)= q(\lambda I-T)< \infty \) it follows, from [1, Theorem 1.22] that \(\lambda I-T \in B(X)\) and in particular \(\lambda I-T\in W(X)\), so \(\lambda \notin \sigma _w(T)\).

Conversely, suppose that T satisfies property \((Z {\scriptstyle \Pi _a})\) and \(\Delta _a (T)\cap \sigma _w (T)= \emptyset \). If \(\lambda \in \Delta _a (T)\) then \(\lambda \notin \sigma _w(T)\) and hence \(\lambda \notin \sigma _{uw} (T)\). This implies that \(\lambda \in \Delta _a (T)= \Pi _a (T)= \Pi (T)\), always by [19, Lemma 2.9]. Therefore T has property \((UW {\scriptstyle \Pi }).\)\(\square \)

3 Property \((UW {\scriptstyle \Pi })\) and perturbations

An operator \(R\in L(X)\) is called a Riesz operator if \(\lambda I-R\) is Fredholm for all nonzero \(\lambda \in \mathbb {C}\). Evidently, quasi-nilpotent operators and compact operators are Riesz operators. Also every operator K for which, for some \(n\in {\mathbb {N}}\), \(K^n\) is finite-dimensional is a Riesz operator. Indeed, \(K^n\) is a Riesz operator and hence K is a Riesz operator, see [14]. The spectrum of a Riesz operator is either a finite set or a sequence of eigenvalues which clusters at 0. It is known that if \(T\in L(X)\) then \(\sigma _b (T+R)= \sigma _b (T)\) for every Riesz operator R commuting with T, see [1, Chapter 2], and \(\sigma _{uw}(T+R)= \sigma _{uw}(T)\) see [17].

  • Every Riesz operator T having infinite spectrum has property \((UW {\scriptstyle \Pi })\). To see this, observe first that \(\sigma _a (T)= \sigma (T)\), since \(T^*\) has SVEP. Further, \(\sigma _{uw} (T)= \sigma _d (T)= \{0\}\), so \(\Delta _a(T)= \Pi (T)\).

  • An operator \(T\in L(X)\) is said to be algebraic if there exists a nontrivial complex polynomial h such that h(T)=0. Examples of algebraic operators are idempotent operators, nilpotent operators and every operator K such that \(K^n\) is finite-dimensional for some \(n\in {\mathbb {N}}\).If T is algebraic then \(\sigma (T)\) is a finite set of poles [1, Chapter 3], say \(\{\lambda _1,\dots , \lambda _n\}\). Therefore, \(\sigma _d (T)= \emptyset \) and hence \(\Pi (T)= \{\lambda _1,\dots , \lambda _n\}\). Moreover, \(\sigma _a (T) = \sigma (T)\), since a pole is always an eigenvalue and \(\sigma _{uw}(T) \ne \emptyset \) for every operator \(T\in L(X)\). Therefore, for an algebraic operator we have \(\Delta _a (T) \ne \Pi (T)\), i.e., for every algebraic operator property \((UW {\scriptstyle \Pi })\) fails.

  • A Riesz operator which has finite spectrum may fail property \((UW {\scriptstyle \Pi })\). Indeed, every finite-dimensional operator K does not satisfy this property, since K is algebraic.

  • If Q be a quasi-nilpotent operator, then

    $$\begin{aligned} Q \ \text{ has } \text{ property } \ (UW {\scriptstyle \Pi }) \Longleftrightarrow \ 0\notin \sigma _d(Q). \end{aligned}$$

    Indeed, if 0 is not a pole then \(\{0\}=\sigma (Q)= \sigma _{ uw}(Q)= \sigma _{ a}(Q)= \sigma _d (Q)\), since both \(\sigma _{ uw}(Q)\) and \(\sigma _{ a}(Q)\) are non-empty. Therefore \(\Delta _a (Q) = \Pi (Q)= \emptyset \), so Q has property \((UW {\scriptstyle \Pi })\). Conversely, if Q has property \((UW {\scriptstyle \Pi })\) then, we have \(\emptyset =\sigma _a (Q)\setminus \sigma _{uw}(Q)= \Pi (T)\) and since \(\sigma (Q)= \{0\}\) it then follows that \(\sigma _{d}(Q)=\{0\}\).

  • It should be noted that if \(K^n\) is finite-dimensional for some \(n\in {\mathbb {N}}\) and \(KT=TK\) then

    $$\begin{aligned} \sigma _d (T)= \sigma _d(T+K), \end{aligned}$$

    see [21], or [1, Chapter 3] for an alternative proof. Moreover, if \(\text{ acc }\, F\) denotes the set of all cluster points of \(F\subseteq {\mathbb {C}}\), then

    $$\begin{aligned} \text{ acc }\,\sigma (T)= \text{ acc }\, \sigma (T+K) \quad \text{ and } \quad \text{ acc }\,\sigma _a (T)= \text{ acc } \, \sigma _a (T+K) \end{aligned}$$

    see [20].

Theorem 3.1

Suppose that \(T\in L(X)\) has property \((UW {\scriptstyle \Pi })\) and \(T^*\) has SVEP. If \(K\in L(X)\) commutes with T and \(K^n\) is finite-dimensional for some \(n\in {\mathbb {N}}\), then \(T+K\) has property \((UW {\scriptstyle \Pi }).\)

Proof

Since T has property \((UW {\scriptstyle \Pi })\), by Theorem 2.3 we have \(\sigma _b (T) = \sigma _d(T)\). The spectrum \(\sigma _b (T)\) is invariant under Riesz commuting perturbations, and being K a Riesz operator, we then have \(\sigma _b (T+K)= \sigma _b(T)\), while the equality \(\sigma _d (T+K) = \sigma _d (T)\), has been already observed. Therefore, \(\sigma _b (T+K)= \sigma _d (T+K)\) . Now, K is a Riesz operator and hence also \(K^*\) is a Riesz operator. Consequently, \(T^*+K^*\) has SVEP, see [1, Chapter 2]. By Theorem 2.3 it then follows that \(T+K\) has property \((UW {\scriptstyle \Pi }).\)

\(\square \)

Corollary 3.2

Suppose that \(Q \in L(X)\) is a quasi-nilpotent such that \(\alpha (Q)< \infty \). If \(K \in L(X)\) commutes with Q and \(K^n\) is finite-dimensional for some \(n\in {\mathbb {N}}\), then \(Q+K\) satisfies property \((UW {\scriptstyle \Pi })\).

Proof

If \(\alpha (Q) < \infty \) then \(0\in \sigma _d (Q)\), otherwise, by [1, Theorem 1.22], we would have \(0 \notin \sigma _b(Q)\), hence \(\sigma _b (Q)= \emptyset \), and this is impossible. Therefore, 0 is not a pole of the resolvent of T and, as noted before, Q has property \((UW {\scriptstyle \Pi })\). Theorem 3.1 then applies, since \(Q^*\) has SVEP. \(\square \)

Property \((UW {\scriptstyle \Pi })\) is also transmitted to \(T+K\) if we assume that the approximate-point spectrum \(\sigma _a(T)\) has no isolated points:

Theorem 3.3

Suppose that \(T, K\in L(X)\) commute, and \(K^n\) is finite-dimensional for some \(n\in {\mathbb {N}}\). If \(\text{ iso }\, \sigma _a (T)= \emptyset \) and T has property \((UW {\scriptstyle \Pi })\), then \(T+K\) has property \((UW {\scriptstyle \Pi }).\)

Proof

Observe that if \(\text{ iso }\, \sigma _a (T)= \emptyset \) then \(\text{ iso }\, \sigma (T)= \emptyset \), since \( \text{ iso } \, \sigma (T)\subseteq \text{ iso }\,\sigma _a(T)\). Therefore,

$$\begin{aligned} \sigma (T)= \text{ iso }\, \sigma (T) \cup \text{ acc }\, \sigma (T)=\text{ acc }\, \sigma (T), \end{aligned}$$

and, similarly, \(\sigma _a (T)= \text{ acc }\, \sigma _a (T)\). Since \(\text{ acc }\, \sigma _{a}(T)=\text{ acc }\ \sigma _{a}(T+K) \), and \(\text{ acc }\, \sigma _{}(T)=\text{ acc }\, \sigma _{}(T+K) \) see [20], we then have \(\sigma _a (T+K)= \sigma _a(T)\) and \(\sigma (T+K)= \sigma (T)\). Also the spectrum \(\sigma _{uw}(T)\) is invariant under Riesz commuting perturbations, so \(\sigma _{uw}(T+K)= \sigma _{uw}(T)\). Hence,

$$\begin{aligned} \Delta _a (T+K)= \sigma _a (T+K)\setminus \sigma _{uw}(T+K)= \sigma _a (T)\setminus \sigma _{uw}(T)=\Delta _a (T). \end{aligned}$$

As observed before, \(\sigma _d (T+K) = \sigma _d (T)\), so

$$\begin{aligned} \Pi (T+K)= \sigma (T+K)\setminus \sigma _d(T+K)= \sigma (T)\setminus \sigma _d(T)= \Pi (T). \end{aligned}$$

Therefore, \(\Delta _a (T+K)= \Pi (T+K) = \Pi (T)= \Delta _a(T)\). \(\square \)

Theorem 3.3 applies to nilpotent commuting perturbations. However, in this case no assumption on the spectrum is required. It is well-known that \(\sigma (T)\) and \(\sigma _a(T)\) are invariant under commuting quasi-nilpotent perturbations.

Theorem 3.4

Let \(T\in L(X)\) and let N be a nilpotent operator which commutes with T. If T has property \((UW {\scriptstyle \Pi })\), then \(T+N\) has property \((UW {\scriptstyle \Pi }).\)

Proof

We have that \(\sigma (T+N)= \sigma (T)\), \(\sigma _a (T+N)= \sigma _a (T)\) and \(\sigma _d(T+N)= \sigma _d(T)\), so \(\Delta _a (T+N) = \Delta _a (T)= \Pi (T)= \Pi (T+N)\). \(\square \)

The following example shows that the result of Theorem 3.4 cannot be extended to quasi-nilpotent commuting perturbations:

Example 3.5

Let \(Q\in L(X)\) denote an injective quasi-nilpotent operator (for instance the Volterra integral operator defined on the space of all continuous function C[ab]). Then \(\Delta _a (Q)= \sigma _a (Q)\setminus \sigma _{uw}(Q)= \emptyset \). Since \(\alpha (Q)= 0\), the point 0 cannot be a pole of the resolvent, i.e. \(\sigma _d(Q) = \{0 \}\), otherwise the condition \(p(Q)= q(Q)< \infty \) would implies \(\alpha (Q) = \beta (Q)= 0\) and hence \(0\notin \sigma (Q)\)). Therefore, Q has property \((UW {\scriptstyle \Pi }).\) On the other hand, \(0=Q-Q\) does not have property \((UW {\scriptstyle \Pi })\) since \(\Delta _a (0)= \emptyset \), while \(\sigma _d(0) = \emptyset \) and \(\sigma (0)= \{0\}\), hence \(\Pi (0) = \{0\}\).

Theorem 3.6

Suppose that \(T\in L(X)\) has property \((UW {\scriptstyle \Pi })\). If \(\text{ iso }\, \sigma _b(T)= \emptyset \), or \(\text{ iso }\, \sigma _{ub}(T)= \emptyset \), then \(T+Q\) has property \((UW {\scriptstyle \Pi })\) for every commuting quasi-nilpotent operator Q.

Proof

The spectra \(\sigma (T)\), \(\sigma _a (T)\) and \(\sigma _{uw}(T)\) are invariant under a commuting quasi-nilpotent perturbation Q, so, if T has property \((UW {\scriptstyle \Pi })\), then \(T+Q\) has property \((UW {\scriptstyle \Pi })\) exactly when \(\sigma _d(T+Q) = \sigma _d(T)\). Each one of the assumptions \(\text{ iso }\, \sigma _b(T)= \emptyset \), or \(\text{ iso }\, \sigma _{ub}(T)= \emptyset \), entails the equality \(\sigma _d(T+Q) = \sigma _d(T)\), see [11, Proposition 2.7]. \(\square \)

The result of Corollary 3.2 may be strongly improved if we consider injective quasi-nilpotent perturbations. To see this we need a preliminary lemma.

Lemma 3.7

Let \(T\in L(X)\) be such that \(\alpha (T)< \infty \). Suppose that there exists an injective quasi-nilpotent operator \(Q\in L(X)\) such that \(TQ=QT\). Then T is injective.

Proof

Set \(Y: = \ker \, T\). Clearly, Y is invariant under Q and the restriction \((\lambda I-Q)|Y\) is injective for all \(\lambda \ne 0\). Since Y is finite-dimensional then \((\lambda I-Q)|Y\) is also surjective for all \(\lambda \ne 0\), thus \(\sigma (Q|Y)\subseteq \{0\}\). On the other hand, from assumption we know that Q|Y is injective and hence Q|Y is surjective, so \(\sigma (Q|Y)= \emptyset \), from which we conclude that \(Y= \{0\}\). \(\square \)

Recall that \(T\in L(X)\) is said to be finite-isoloid (respectively, finite a-isoloid) if every isolated point of \(\sigma (T)\) (respectively, \(\sigma _a (T)\)) is an eigenvalue having finite rank. It is easily seen that if T is finite-isoloid then \(\sigma _b (T)= \sigma _d (T)\). Indeed, if \(\lambda \notin \sigma _d(T)\) then \(\lambda \in \text{ iso } \, \sigma (T)\), so \(\alpha (\lambda I-T)< \infty \). Since \(p(\lambda I-T)= q(\lambda I-T)< \infty \), then \(\lambda I-T\) is Browder, by [1, Chapter 1], so \(\lambda \notin \sigma _b(T)\). Therefore, \(\sigma _b(T)\subseteq \sigma _d (T)\), and being the opposite inclusion true for every operator, we have \(\sigma _b (T)= \sigma _d (T)\).

Theorem 3.8

Suppose that \(T \in L(X)\) is finite-isoloid. If Q is an injective quasi-nilpotent operator commuting with T, then both T and \(T+Q\) have property \((UW {\scriptstyle \Pi })\).

Proof

Let \(\lambda \in \Delta _a (T)= \sigma _a (T)\setminus \sigma _{uw}(T)\). Since \(\lambda I-T \in W_+(X)\) then \(\alpha (\lambda I-T)< \infty \) and \(\lambda I-T\) has closed range. Since \(\lambda I-T\) commutes with Q it then follows, by Lemma 3.7, that \(\lambda I-T\) is injective, so \(\lambda \notin \sigma _{ a} (T)\), a contradiction. Therefore, \(\Delta _a (T)\) is empty. Also \(\Pi (T)\) is empty. Indeed, suppose that \(\lambda \in \Pi (T)=\sigma (T)\setminus \sigma _d(T)\). Then \(\lambda \) is an isolated point of \(\sigma (T)\), and since T is finite-isoloid then \(0<\alpha (\lambda I-T)< \infty \). But by Lemma 3.7 we have \(\alpha (\lambda I-T)=0\), a contradiction. Therefore \(\Pi (T) = \emptyset \), and hence T has property \((UW {\scriptstyle \Pi })\).

To show that \(T+Q\) has property \((UW {\scriptstyle \Pi })\), observe first that

$$\begin{aligned} \Delta _a (T+Q)= \sigma _a (T+Q)\setminus \sigma _{uw} (T+Q)= \sigma _a (T)\setminus \sigma _{uw}(T)=\Delta _a (T), \end{aligned}$$

so, from the first part, \(\Delta _a (T+Q) = \emptyset \). Also \(\Pi (T+Q)\) is empty. Indeed, suppose that there exists \(\lambda \in \Pi (T+Q)\). Then \(\lambda \in \text{ iso }\, \sigma (T+Q)= \text{ iso }\, \sigma (T)\), and hence, since T is finite-isoloid, \(0<\alpha (\lambda I-T)< \infty \). Again by Lemma 3.7 we have \(\alpha (\lambda I-T)=0\), a contradiction. \(\square \)

The assumption that T is finite-isoloid in Theorem 3.8 is crucial. For instance, if T is an injective quasi-nilpotent operator Q then T is not finite-isoloid, and, as observed before, T has property \((UW {\scriptstyle \Pi })\), while \(T-Q=0\) does not have property \((UW {\scriptstyle \Pi })\).

In the next result, we give a necessary and sufficient condition under which property \((UW {\scriptstyle \Pi })\) is preserved under commuting Riesz perturbations.

Theorem 3.9

Let \(T\in L(X)\) and let R be a Riesz operator commuting with T. If T satisfies property \((UW {\scriptstyle \Pi })\) then the following assertions are equivalent:

  1. (i)

    \(T+R\) satisfies property \((UW {\scriptstyle \Pi });\)

  2. (ii)

    \(p_{00}^a(T+R)=\Pi (T+R)\).

Proof

(i)\( \Rightarrow \) (ii) Suppose that \(T+R\) satisfies property \((UW {\scriptstyle \Pi }).\) By Theorem 2.3 we then have \(\ p_{00}^a(T+R)=\Pi (T+R).\)

(ii) \(\Rightarrow \) (i) Assume hat \(\ p_{00}^a(T+R)=\Pi (T+R)\). Since T satisfies property \((UW {\scriptstyle \Pi })\) then T satisfies a-Browder’s theorem. Hence \(\sigma _{uw}(T)=\sigma _{ub}(T)\). Since \(\sigma _{uw}(T+R)=\sigma _{uw}(T)\) and \(\sigma _{ub}(T+R)=\sigma _{ub}(T),\) then so \(\sigma _{ub}(T+R)=\sigma _{uw}(T+R)\), and \(T+R\) satisfies a-Browder’s theorem i.e \(\Delta _a(T+R)= p_{00}^a(T+R)\). Since \( p_{00}^a(T+R)=\Pi (T+R)\), then \(\Delta _a(T+R)=\Pi (T+R).\) Therefore \(T+R\) satisfies property \((UW {\scriptstyle \Pi })\).\(\square \)

Property \((UW {\scriptstyle \Pi }))\) is transmitted under commuting Riesz perturbations in a very special case. Recall that an operator \(T\in L(X)\) is said to be finitely a-polaroid if every isolated point of \(\sigma _a (T)\) is a pole of T having finite rank. Recall that, in general, the equality \(\sigma _a (T+R) = \sigma _a (T)\), for a Riesz commuting perturbation, does not hold, even if T is a finite-rank operator.

Theorem 3.10

Suppose that \(T\in L(X)\) is finitely a-polaroid and let \(R\in L(X)\) be a commuting Riesz operator such that \(\sigma _a (T+R)= \sigma _a (T)\). If T has property \((UW {\scriptstyle \Pi }))\) then \(T+R\) has property \((UW {\scriptstyle \Pi })\).

Proof

By Theorem 3.9 it suffices to show the equality \(\Pi (T+R)= p_{00}^a (T+R)\). Let \(\lambda \in \Pi (T+R)\). Then \(\lambda \in \text{ iso } \,\sigma _a (T+R)= \text{ iso }\,\sigma _a (T)\). Since T is finitely a-polaroid then \(\lambda I-T\) is Drazin invertible and \(\alpha (\lambda I-T)< \infty \). Then \(\lambda I-T\in B(X)\), see [1, Chapter 1], in particular is upper semi-Browder. This implies that \(\lambda I- (T+R) \in B_+(X)\), so \(\lambda \in \sigma _a (T+R)\setminus \sigma _{ub}(T+R) = p_{00}^a (T+R)\).

Conversely, suppose that \(\lambda \in p_{00}^a (T+R)\). Then

$$\begin{aligned} \lambda\in & {} \sigma _a (T+R)\setminus \sigma _{ub}(T+R) \subseteq \sigma _a (T+R)\setminus \sigma _{uw}(T+R)\\= & {} \sigma _a (T)\setminus \sigma _{uw}(T) = \Delta _a (T)=\Pi (T). \end{aligned}$$

In particular, \(\lambda \) is an isolated point of \(\sigma _a (T)\), and consequently, a pole of T having finite rank. This implies that \(\lambda I-T \in B(X)\), see [1, Chapter 3], and hence \(\lambda I- (T+R) \in B(X)\), so \(\lambda \in \Pi (T+R)\). \(\square \)

4 Property \((UW {\scriptstyle \Pi })\) under functional calculus

In this section we study the preservation of property \((UW {\scriptstyle \Pi })\) under functional calculus. Furthermore, we show that this property is transferred from a Drazin invertible operator to its Drazin inverse. Let \({\mathcal {H}}(\sigma (T))\) be the set of all analytic functions defined on a neighborhood of \(\sigma (T)\), and for every \(f\in {\mathcal {H}}(\sigma (T))\) let f(T) be defined by means of the classical functional calculus. The spectral theorem for the spectrum asserts that \(\sigma (f(T))= f (\sigma (T))\) for every \(f\in {\mathcal {H}}(\sigma (T))\) and a similar equality holds for the approximate point spectrum.

Lemma 4.1

Let \(T \in L(X)\) and let \(\{\lambda _1, \cdots , \lambda _k \}\) be a finite subset of \({\mathbb {C}}\) such that \(\lambda _i \ne \lambda _j \) for \(i\ne j\). Assume that \(\{\nu _1, \cdots , \nu _n\}\subset {\mathbb {N}} \) and set \( h(\lambda ) : = \prod _{i=1}^{n}(\lambda _i - \lambda )^{\nu _i}\). Then, for the operator \( h(T) : = \prod _{i=1}^{n}(\lambda _i I - T)^{\nu _i}\) we have

$$\begin{aligned} \mathrm{ker}\ h(T) = \bigoplus _{i=1}^n \mathrm{ker} \ (\lambda _i I - T)^{\nu _i} \end{aligned}$$
(5)

and

$$\begin{aligned} h(T)(X) = \bigcap _{i=1}^n (\lambda _i I - T)^{\nu _i}(X). \end{aligned}$$
(6)

Furthermore,

$$\begin{aligned} 0\in \Pi (h(T)) \Leftrightarrow \lambda _i \in \Pi (T) \ \text{ for } \text{ all } \ i=1, \dots , n. \end{aligned}$$

Proof

A proof of the equalities (5) and (6) may be found in [1]. Suppose that \(\lambda _i \in \Pi (T)\) for all \(i=1, \dots , n\), i.e., \(\lambda _i \in \sigma (T)\) and \(\lambda _i I-T\) Drazin invertible, for all \(i=1, \dots , n\). Then, for k large enough, we have \(\ker (\lambda _i I-T)^{\nu _i k}= \ker (\lambda _i I-T)^{\nu _i (k+1)}\) and \((\lambda _i -T)^{\nu _i k} (X)= (\lambda _i -T)^{\nu _i (k+1)} (X)\). This implies, by (5) and (6), that \(\ker \, (h(T)^k)= \ker \, (h(T)^{k+1}) \) and \(h(T)^k (X)= h(T)^{k+1}(X)\), so h(T) is Drazin invertible. Since \(0 =h(\lambda _i)\in h(\sigma (T)) = \sigma (h(T))\) it then follows that \(0\in \Pi (h(T))\).

Conversely, assume that \(\lambda _j \notin \Pi (T)\) for some \(j\in \{1,\dots , n\}\). Let \(p_j\) and \(q_j\) denote the ascent and the descent of \(\lambda _j I-T\), respectively. We have either \(\lambda _j \notin \sigma (T)\) or \(\lambda _j \in \sigma (T)\). In the case where \(\lambda _j \notin \sigma (T)\), then \(0= h(\lambda _j)\notin h(\sigma (T))= \sigma (h(T))\), so \(0\notin \Pi (T)\). Consider the other case \(\lambda _j \in \sigma (T)\). Since \(\lambda _j \notin \Pi (T)\) then \(\lambda _j I-T\) is not Drazin invertible, so either \(p_j=\infty \), i.e. \(\ker \, (\lambda _j I-T)^k\) is properly contained in \(\ker \, (\lambda _j I-T)^{k+1}\) for every \(k\in {\mathbb {N}}\), or \(q_j=\infty \), i.e., \((\lambda _j I-T)^{k+1} (X)\) is properly contained in \((\lambda _j I-T)^{k} (X)\) for every \(k\in {\mathbb {N}}\). If \(p_j=\infty \), being \(\ker \, (\lambda _i I -T)^{\nu _i} \cap \ker \, (\lambda _j I -T)^{\nu _j}= \emptyset \) for all \(i\ne j\), it then follows that \(p(h(T))= \infty \). Analogously, if \(q_j= \infty \) then \(q(h(T))= \infty \). Therefore, also in this case \(0\notin \Pi (T)\). \(\square \)

Remark 4.2

If \(T\in L(X)\) is invertible and \(S\in L(X)\) commutes with T then \(N(S^n)= N((TS))^n\) and \((TS)^n (X)= S^n (X)\) for all \(n\in {\mathbb {N}}\). Consequently, TS is Drazin invertible if and only if S is Drazin invertible, while \(0\in \Pi (TS)\) if and only if \(0\in \Pi (S)\).

In the sequel we need the following lemma.

Lemma 4.3

For every \(T\in L(X)\), X a Banach space, and \(f\in {\mathcal {H}} (\sigma (T))\) we have

$$\begin{aligned} \sigma _a (f(T))\setminus \Pi (f(T)) \subseteq f \left( \sigma _a (T)\setminus \Pi (T) \right) . \end{aligned}$$
(7)

Furthermore, if T is a-isoloid then

$$\begin{aligned} \sigma _a (f(T))\setminus \Pi (f(T)) = f\left( \sigma _a (T) \setminus \Pi (T)\right) . \end{aligned}$$
(8)

Proof

Suppose that \(\lambda _0 \in \sigma _a (f(T))\setminus \Pi (f(T))\). We consider two cases.

Case (I). Suppose first that \(\lambda _0 \notin \text{ iso } \, f(\sigma _a (T))= \text{ iso } \, \sigma _a (f(T))\). In this case \(\lambda _0 \notin \Pi (f(T))\), since each point of \(\Pi (f(T))\) is an isolated point of \(\sigma (f(T))\) and hence an isolated point of \(\sigma _a (f(T))\). Let \((\lambda _n)\subset f(\sigma _a (T))= \sigma _a (f(T))\) be a sequence such that \(\lambda _n \rightarrow \lambda _0\) and let \((\mu _n)\in \sigma _a (T)\) be such that \(f(\mu _n) =\lambda _n\). Then \((\mu _n)\) admits a subsequence which converge to some \(\mu _0 \in \sigma _a (T))\). There is no harm if we assume that \(\mu _n \rightarrow \mu _0\). Then \(f(\mu _n)\rightarrow f(\mu _0)= \lambda _0\). Evidently, \(\mu _0 \notin \Pi (T)\), since \(\mu _0\) is a cluster point of \(\sigma _a (T)\). Hence

$$\begin{aligned} \lambda _0 =f(\mu _0) \in f\left( \sigma _a (T) \setminus \Pi (T)\right) . \end{aligned}$$

Case (II). Suppose that \(\lambda \in \text{ iso } \, f(\sigma _a (T)) =\text{ iso } \, \sigma _a (f(T))\). Define \(g(\lambda ): = \lambda _0 -f(\lambda )\), Since \(g(\lambda )\) is analytic then g has only a finite number of zeros in \(\sigma (T)\), say \(\{\lambda _1, \dots , \lambda _n\}\). Write

$$\begin{aligned} g(\lambda ) = h(\lambda ) k(\lambda ) \quad \text{ and } \quad h(\lambda ) = \prod _{i=1}^{k}(\lambda _i - \lambda )^{\nu _i}, \end{aligned}$$
(9)

where \(k(\lambda )\) has no zero in \(\sigma (T)\). Then

$$\begin{aligned} g(T)= \lambda _0 I- f(T)= h(T) k(T), \end{aligned}$$

where k(T) invertible, and \(h(T): = \prod _{i=1}^{k}(\lambda _i I - T)^{\nu _i}\).

Now, \(\lambda _0 \notin \Pi (f(T)\) implies that \(\lambda _0 I-f(T)\) is invertible, or \(\lambda _0 I-f(T)\) is not Drazin invertible. In the first case g(T) is invertible, so \(0\notin \Pi (g(T))\). If \(\lambda _0 I-f(T)\) is not Drazin invertible, we have either \(p(g(T))= p(\lambda _0 I-f(T))= \infty \) or \(q(g(T))= q(\lambda _0 I-f(T))= \infty \), so also in this case \(0\notin \Pi (g(T))\). According Remark 4.2 then \(0\notin \Pi (h(T))\). By Lemma 4.1 it then follows that \(\lambda _j \notin \Pi (T)\) for some j. Thus, \(\lambda _j \in \sigma _a (T)\setminus \Pi (T)\), and hence \(\lambda _0 = f(\lambda _j)\in f\left( \sigma _a (T) \setminus \Pi (T)\right) .\)

To show the equality (8), assume that T is a-isoloid.

To show the equality (8) we need only to prove the inclusion \(\supseteq \). Let \(\lambda _0 \in f\left( \sigma _a(T) \setminus \Pi (T)\right) \) be arbitrary given. Since \(\lambda _0 \in f(\sigma _a (T))= \sigma _a (f(T)) \subseteq \sigma (f (T))\) it then suffices to prove that \(\lambda _0 \notin \Pi (f(T))\). To do this, suppose that \(\lambda _0 \in \Pi (f(T))\). Then \(0 \in \Pi (h(T)\), and by Lemma 4.1 it then follows that \(\lambda _i \in \Pi (T)\) for all \(i=1, \dots , n\). Thus, \(\lambda _i \notin \sigma _a (T)\setminus \Pi (T)\) for all i. Since \(0= g(\lambda _i)= \lambda _0-f(\lambda _i)\), we have \(\lambda _0 = f(\lambda _i) \notin f\left( \sigma _a(T) \setminus \Pi (T)\right) \), a contradiction. Therefore, \(\lambda _0 \notin \Pi (f(T))\), and hence

$$\begin{aligned} \lambda _0 \in \sigma _a (f(T))\setminus \Pi (f(T)), \end{aligned}$$

as desired. \(\square \)

Recall that, in general, the spectral theorem does not hold for \(\sigma _{uw} (T)\), see [1, Chapter 3].

Theorem 4.4

Let \(T\in L(X)\) be a-isoloid and \(f\in {\mathcal {H}} (\sigma (T))\). If T satisfies property \((UW {\scriptstyle \Pi }))\) then the following are equivalent:

  1. (i)

    f(T) satisfies property \((UW {\scriptstyle \Pi })\);

  2. (ii)

    \(f(\sigma _{uw}(T))= \sigma _{uw} (f(T))\).

Proof

(i) \(\rightarrow \) (ii) Observe first that the spectral mapping theorem holds for \(\sigma _{\mathrm{ub}}(T)\), i.e., \(f(\sigma _{\mathrm{ub}}(T))= \sigma _{\mathrm{ub}} (f(T))\) for every \(f\in {\mathcal {H}} (\sigma (T))\), see [1, Chapter 3]. Since f(T) satisfies property \((UW {\scriptstyle \Pi })\), then both T and f(T) satisfies a-Browder’s theorem, by Theorem 2.3, so \(\sigma _{\mathrm{uw}}(T)= \sigma _{\mathrm{ub}} (T)\) and \(\sigma _{\mathrm{uw}}(f(T))= \sigma _{\mathrm{ub}} (f(T))\). Then we have

$$\begin{aligned} f(\sigma _{\mathrm{uw}}(T))= f(\sigma _{\mathrm{ub}}(T))= \sigma _{\mathrm{ub}} (f(T))= \sigma _{\mathrm{uw}}(f(T)). \end{aligned}$$

(ii) \(\rightarrow \) (i) Since T satisfies property \((UW {\scriptstyle \Pi }))\) we have \(\sigma _{uw}(T) = \sigma _a (T)\setminus \Pi (T)\). By Lemma 4.3 it then follows that

$$\begin{aligned} \sigma _{uw} (f(T))= f(\sigma _{uw}(T))= f ( \sigma _a (T)\setminus \Pi (T))= \sigma _a (f(T))\setminus \Pi (f(T)), \end{aligned}$$

so f(T) satisfies property \((UW {\scriptstyle \Pi })\). \(\square \)

It is known that the spectral theorem for \(\sigma _{uw}(T)\) holds if T or \(T^*\) has SVEP ( [1, Chapter 3], or if f is injective ( [18]. Consequently, property \((UW {\scriptstyle \Pi }))\) is transmitted from T to f(T) if T is a-isoloid and T or \(T^*\) has SVEP. In the case, that \(T^*\) has SVEP, then we can require that T is isoloid, since in this case \(\sigma _a(T)= \sigma (T)\) and hence the properties of being a-isoloid and isoloid coincide.

Lemma 4.5

Suppose that for a bounded operator \(T\in L(X)\) there exists \(\lambda _0 \in {\mathbb {C}}\) such that \( K(\lambda _0 I-T) = \{0\}\) and \(\ker \, (\lambda _0 I-T)= \{0\}\). Then \(\sigma _{p}(T)= \emptyset \).

Proof

For all complex \(\lambda \ne \lambda _0\) we have \(\ker \, (\lambda I-T) \subseteq K(\lambda _0 I-T)\), so that \(\ker \, (\lambda I-T)= \{0\}\), for \(\lambda \ne \lambda _0\). Since \(\ker \, (\lambda _0 I-T)= \{0\}\) we then conclude that \(\ker \, (\lambda I-T)= \{0\}\) for all \(\lambda \in {\mathbb {C}}\). \(\square \)

Theorem 4.6

Let \(T\in L(X)\) be such that there exists \(\lambda _0 \in {\mathbb {C}}\) such that

$$\begin{aligned} K(\lambda _0 I-T) = \{0\} \quad \text{ and } \quad \ker \, (\lambda _0 I-T)= \{0\}. \end{aligned}$$
(10)

Then property \((UW {\scriptstyle \Pi })\) holds for f(T) for all \(f\in {\mathcal {H}}(\sigma (f(T))\).

Proof

We know from Lemma 4.5 that \(\sigma _{p}(T)= \emptyset \), so T has SVEP. We show that also \(\sigma _{p}(f(T))= \emptyset \). Let \(\mu \in \sigma (f(T))\) and write \(\mu -f(\lambda ) = p (\lambda ) g(\lambda )\), where g is analytic on an open neighborhood \({\mathcal {U}}\) containing \(\sigma (T)\) and without zeros in \(\sigma (T)\), p a polynomial of the form \( p(\lambda ) = \Pi _{k=1}^n (\lambda _k- \lambda )^{\nu _k}, \) with distinct roots \(\lambda _1, \dots , \lambda _n\) lying in \(\sigma (T)\). Then

$$\begin{aligned} \mu I- f(T) = \Pi _{k=1}^n (\lambda _k I - T)^{\nu _k} g(T). \end{aligned}$$

Since g(T) is invertible, \(\sigma _{\mathrm{p}}(T)= \emptyset \) implies that \(\ker \, (\mu I - f(T))= \{0\}\) for all \(\mu \in {\mathbb {C}}\), so \(\sigma _{p}(f(T))= \emptyset \). Since T has SVEP then f(T) has SVEP, see [1, Chapter 2], so that a-Browder’s theorem holds for f(T) ( [1, Chapter 5]). To prove that property \((UW {\scriptstyle \Pi })\) holds for f(T), by Theorem 2.3 it then suffices to prove that

$$\begin{aligned} p_{00}^a (f(T))= \Pi (f(T)). \end{aligned}$$

Obviously, the condition \(\sigma _{p}(f(T))= \emptyset \) entails that \(\Pi (f(T))= \emptyset \), since every point of \(\Pi (f(T))\) is an eigenvalue of f(T). On the other hand, if \(\lambda \in p_{00}^a (f(T))\) then \(\lambda \in \sigma _a (f(T))\) and \(\lambda \notin \sigma _{ub}(f(T)\), so \(\lambda I-f(T)\) has closed range, since it is a upper semi-Browder operator. This implies that \(\alpha (\lambda I-f(T)) >0\) and we get a contradiction, since \(\sigma _{p}(f(T))= \emptyset \). \(\square \)

The conditions of Theorem 4.6 are satisfied by any injective operator for which the hyperrange\(T^\infty (X): = \bigcap T^n (X)\) is \(\{0\}\). In fact, \( K(T)\subseteq T^\infty (X)\) for all \(T\in L(X)\), so that \(K(T)= \{0\}\). In particular, the conditions of Theorem 4.6 are satisfied by a semi-shift T, i.e. T is an isometry for which \(T^\infty (X)= \{0\},\) see [15] for details on this class of operators. Clearly, a semi-shift T on a non-trivial Banach space is a non-invertible isometry.

The following result applies to several operator, in particular to shift operators.

Lemma 4.7

Suppose that for \(T\in L(X)\) we have \(\text{ iso }\, \sigma _{\mathrm{a}}\, (T)= \emptyset \). If \(K\in L(X)\) commutes with T and \(K^n\) is finite-dimensional for some \(n\in {\mathbb {N}}\), then \(\text{ iso }\, \sigma _{\mathrm{a}} (T+K)= \emptyset \). Consequently, \( \sigma _{\mathrm{a}} (T+K)= \sigma _{\mathrm{a}} (T)\).

Proof

By Lemma 4.7 we have

$$\begin{aligned} \sigma _{\mathrm{a}} (T)= & {} \text{ iso }\, \sigma _{\mathrm{a}}\, (T) \cup \text{ acc }\, \sigma _{\mathrm{a}}\, (T)= \text{ acc }\, \sigma _{\mathrm{a}}\, (T)\\= & {} \text{ acc }\, \sigma _{\mathrm{a}}\, (T+K )\subseteq \sigma _{\mathrm{a}} (T+K). \end{aligned}$$

On the other hand, if \(\sigma _ \mathrm{a} (K)= \{\lambda _1, \lambda _2, \dots \lambda _n\}\) we have

$$\begin{aligned} \text{ iso }\, \sigma _{\mathrm{a}} (T+K)\subseteq \text{ iso }\,(\sigma _\mathrm{a} (T) + \sigma _ \mathrm{a} (K)) = \text{ iso }\,\bigcup _{k=1}^n (\lambda _k + \sigma _{\mathrm{a}} (T))= \emptyset , \end{aligned}$$

hence,

$$\begin{aligned} \sigma _{\mathrm{a}} (T+K)= & {} \text{ iso }\, \sigma _{\mathrm{a}} (T+K) \cup \text{ acc }\, \sigma _{\mathrm{a}} (T+K)= \text{ acc }\, \sigma _{\mathrm{a}} (T+K)\\= & {} \text{ acc }\, \sigma _{\mathrm{a}} (T) = \sigma _{\mathrm{a}} (T), \end{aligned}$$

so \(\sigma _{\mathrm{a}} (T+K)= \sigma _{\mathrm{a}} (T)\) holds.

\(\square \)

Theorem 4.8

Let \(T\in L(X)\) be a semi-shift and suppose that \(K \in L(X)\) is such that \(K^n\) is finite-dimensional for some \(n\in {\mathbb {N}}\) and that \(TK=KT\). Then property \((UW {\scriptstyle \Pi })\) holds for \(f(T)+K\) for all \(f\in {\mathcal {H}}(\sigma (f(T))\).

Proof

Since T is a non-invertible isometry, the aproximate point spectrum \(\sigma _{\mathrm{a}}(T)\) is the closed unit circle of \({\mathbb {C}}\), see [15, Proposition 1.6.2]. Hence \(\text{ iso }\, \sigma _{\mathrm{a}}(T)= \emptyset \). From Lemma 4.7 and Theorem 3.3 then property \((UW {\scriptstyle \Pi })\) holds for \(f(T)+K\). \(\square \)

Recall that an operator \(T\in L(X)\) is Drazin invertible if and only if there exists an operator \(S\in L(X)\) and \(n\in {\mathbb {N}}\) such that

$$\begin{aligned} TS = ST, \quad STS=S, \quad T^n ST=T^n , \end{aligned}$$
(11)

see [16, Chap.3, Theorem 10]. The operator S is called the Drazin inverse of T. From [16, Chap. 3, Theorem 10] we also know that if \(T\in L(X)\) is Drazin invertible if and only if there exist two closed invariant subspaces Y and Z such that \(X= Y\oplus Z\) and, with respect to this decomposition,

$$\begin{aligned} T= T_1 \oplus T_2, \quad \text{ with } \ T_1: = T|Y \ \text{ nilpotent } \ \text{ and } \ T_2: =T|Z \ \text{ invertible }. \end{aligned}$$
(12)

Note that the Drazin inverse S of an operator, if it exists, is uniquely determined ( [12]), and with respect to the decomposition \(X=Y\oplus Z\), the Drazin inverse S may be represented as the directed sum

$$\begin{aligned} S: = 0\oplus S_2 \quad \text{ with } \ S_2: = {T_2}^{-1}. \end{aligned}$$
(13)

From the decomposition (13) it is obvious that the Drazin inverse S is also Drazin invertible. Evidently, if T is invertible then \(S= T^{-1}\), while \(0 \in \sigma (T)\) if and only if \(0\in \sigma (S)\).

The decompositions (12) and (13) are very useful in order to study the spectral properties of a Drazin invertible operator, see [3] and [4], in particular the decomposition (13) shows that the Drazin inverse S is itself Drazin invertible, since is the direct sum of the nilpotent operator 0 and the invertible operator \(S_2\). It should be noted that if \(0\in \sigma (T)\) then 0 is a pole of the first order of the resolvent of S, see [18]. Furthermore, the following relationship of reciprocity holds for the spectra of S and T:

$$\begin{aligned} \sigma (S)\setminus \{0\}= \left\{ \frac{1}{\lambda }: \lambda \in \sigma (T)\setminus \{0\} \right\} , \end{aligned}$$
(14)

see [1, Chapter 1].

We also have,

$$\begin{aligned} \sigma _{ a}(S)\setminus \{0\}= \left\{ \frac{1}{\lambda }: \lambda \in \sigma _{ a}(T)\setminus \{0\}\right\} , \end{aligned}$$

and

$$\begin{aligned} \sigma _{ uw}(S)\setminus \{0\}= \left\{ \frac{1}{\lambda }: \lambda \in \sigma _{uw}(T)\setminus \{0\}\right\} , \end{aligned}$$

see [4]. By [4, Theorem 2.8] we also have

$$\begin{aligned} \Pi (S) \setminus \{0\} = \left\{ \frac{1}{\lambda }: \lambda \in \Pi (T) \setminus \{0\}\right\} . \end{aligned}$$

Lemma 4.9

Suppose that \(T\in L(X)\) is Drazin invertible with Drazin inverse S. Then \(T \in W_+(X)\) if and only if \(S\in W_+(X)\).

Proof

(i) If \(0\notin \sigma (T)\) then T is invertible and the Drazin inverse is \(S=T^{-1}\) so the assertion is trivial in this case. Suppose that \(0\in \sigma (T)\) and that T is upper semi-Weyl. Since \(p(\lambda I-T)= q(\lambda I-T) < \infty \) then, see [1, Chapter 1], T is Browder. Then 0 is a pole of the resolvent of T and is also a pole (of the first order) of the resolvent of S. Let \(X=Y\oplus Z\) such that \(T= T_1\oplus T_2\), \(T_1=T|Y\) nilpotent and \(T_2=T|Z\) invertible. Observe that

$$\begin{aligned} \ker \, T = \ker \, T_1 \oplus \ker T_2= \ker T_1 \oplus \{0\}, \end{aligned}$$
(15)

and, analogously, since \(S= 0\oplus S_2\) with \(S_2= {T_2}^{-1}\), we have

$$\begin{aligned} \ker \, S = \ker \,0 \oplus \ker S_2= Y \oplus \{0\}. \end{aligned}$$
(16)

Since T is upper semi-Weyl we have \(\alpha (T)= \text{ dim }\, \ker \, T< \infty \), and from the inclusion \(\ker \, T_1 \subseteq \ker \, T\) we obtain \(\alpha (T_1)< \infty \). Consequently, \(\alpha (T_1^n)< \infty \) for all \(n\in {\mathbb {N}}\). Let \(T_1^\nu =0\). Since \(Y = \ker \, T_1^\nu \) the subspace Y is finite-dimensional, and hence \(\ker \, S = Y \oplus \{0\}\) is finite-dimensional, i.e. \(\alpha (S)< \infty \). Now, S is Drazin invertible, so \(p(S)=q(S)< \infty \) and hence, see [1, Chapter 1], \(\alpha (S)= \beta (S) < \infty \). Hence S is Browder, and in particular upper semi-Weyl.

Conversely, suppose that S is upper semi-Weyl. Then \(\alpha (S) <\infty \) and hence by (16) the subspace Y is finite-dimensional, from which it follows that also \(\ker \, T_1= \ker \, T|Y\) is finite-dimensional. From (15) we then have that \(\alpha (T)< \infty \) and since \(p(T)= q(T) < \infty \) we then conclude that \(\alpha (T)= \beta (T) \), see [1, Chapter 1]. Therefore, T is a Browder operator, in particular upper semi-Weyl. \(\square \)

In [4] it is shown that several Browder type theorems and Weyl type theorems are transferred from a Drazin invertible operator to its Drazin inverse. We show now that the same happens for property \((UW {\scriptstyle \Pi })\).

Theorem 4.10

Suppose that \(T\in L(X)\) is Drazin invertible with Drazin inverse S. Then T satisfies property \((UW {\scriptstyle \Pi })\) if and only if S satisfies property \((UW {\scriptstyle \Pi })\).

Proof

Suppose that T satisfies property \((UW {\scriptstyle \Pi })\). Consider first the case that T is invertible. Then \(S= T^{-1}\). Suppose that \(\lambda \in \sigma _a (T^{-1})\). Then \(\lambda \ne 0\), and hence \(\frac{1}{\lambda } \in \sigma _a (T)= \sigma _{uw}(T) \bigsqcup \Pi (T)\). If \(\frac{1}{\lambda } \in \sigma _{uw}(T)\) then \(\lambda \in \sigma _{uw}(T^{-1}\), while if \(\frac{1}{\lambda } \in \Pi (T)\) then \(\lambda \in \sigma _{uw}(T^{-1}\). This shows that \(\sigma _a (T^{-1})\subseteq \sigma _{uw}(T^{-1}) \bigsqcup \Pi (T^{-1})\), and since the opposite inclusion always holds we then have

$$\begin{aligned} \sigma _a (T^{-1})= \sigma _{uw}(T^{-1}) \bigsqcup \Pi (T^{-1}), \end{aligned}$$

i.e., property \((UW {\scriptstyle \Pi })\) holds for \(T^{-1}\). The proof that property \((UW {\scriptstyle \Pi })\) for \(T^{-1}\) implies property \((UW {\scriptstyle \Pi })\) for T is similar.

Consider the case that T is not invertible. Then \(0\in \sigma (T)\), as well as \(0\in \sigma (S)\), and in this case 0 is a pole of both T and S. Let \(\lambda \in \sigma _a (S)\). If \(\lambda =0\) we have already observed that \(0 \in \Pi (S)\) and \(0\in \Pi (T)\). Since T has property \((UW {\scriptstyle \Pi })\) then \(\sigma _a (T)= \sigma _{uw}(T) \bigsqcup \Pi (T)\), so \(0\notin \sigma _{uw} (T)\). But \(p(T)= q(T) < \infty \), so, by [1, Chapter 1], T is Browder, and in particular is upper semi-Weyl. By Lemma 4.9, S is upper semi-Weyl, so \(0\notin \sigma _{uw} (S)\). Therefore, \(0\in \sigma _{uw}(T) \bigsqcup \Pi (T)\). Suppose that \(\lambda \ne 0\). Then \(\frac{1}{\lambda } \in \sigma _{a}(T)\), and hence \(\frac{1}{\lambda } \in \sigma _{uw}(T) \bigsqcup \Pi (T)\). This implies that either \(\lambda \in \sigma _{uw}(S)\) or \(\lambda \in \Pi (S)\), so \(\lambda \in \sigma _{uw}(T) \bigsqcup \Pi (T)\). We have shown that \( \sigma _a (S)\subseteq \sigma _{uw}(T) \bigsqcup \Pi (T)\) and since the reverse inclusion holds we then have \( \sigma _a (S)\subseteq \sigma _{uw}(T) \bigsqcup \Pi (T)\).

Conversely, suppose that S has property \((UW {\scriptstyle \Pi })\). The Drazin inverse of S is \(U: = T^2 S=TST\) and the Drazin inverse of U is T, see [1, Chapter 1]. Therefore, from the first part, U inherits property \((UW {\scriptstyle \Pi })\) from S and property \((UW {\scriptstyle \Pi })\) is then transferred from U to T. \(\square \)