1 Introduction

Let \(\mathscr {X}\) be a complex Banach space. Denote by \(\mathcal {B}(\mathscr {X})\) the algebra of all bounded linear operators on \(\mathscr {X}\). A closed subspace \(\mathscr {M}\subseteq \mathscr {X}\) is said to be invariant for an operator \(T\in \mathcal {B}(\mathscr {X})\) if \(T\mathscr {M}\subseteq \mathscr {M}\). Let \(\mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) be a non-empty set of operators. Then, \(\mathscr {M}\) is an invariant subspace of \(\mathcal {S}\) if it is invariant for every operator in \(\mathcal {S}\). If \(\mathscr {M}\) is invariant for every operator in \(\mathcal {S}\) and for every operator in the commutant \(\mathcal {S}'=\{ T\in \mathcal {B}(\mathscr {X});\; TS=ST\,\text {for every}\, S\in \mathcal {S}\}\), then it is a hyperinvariant subspace of \(\mathcal {S}\). Of course, the trivial subspaces \(\{ 0\}\) and \(\mathscr {X}\) are (hyper)invariant for any set of operators. We are interested in the existence of non-trivial invariant and hyperinvariant subspaces. The problem of existence of invariant and hyperinvariant subspaces for a given operator or a non-empty set of operators is an extensively studied topic in operator theory. The problem is solved in the finite-dimensional setting by Burnside’s theorem (see [14, Theorem 1.2.2]). In the context of infinite-dimensional Banach spaces, the problem is open for reflexive Banach spaces, in particular, for the infinite-dimensional separable Hilbert space. However, there are some Banach spaces for which we know either that every operator has a non-trivial invariant subspace or that there exist operators without it. For instance, Argyros and Haydon [1] have proved the existence of an infinite-dimensional Banach spaces \(\mathscr {X}\) such that every operator in \(\mathcal {B}(\mathscr {X})\) is of the form \(\lambda I+K\), where I is the identity operator and K is compact. It follows, by the celebrated von Neumann-Aronszajn-Smith theorem [2] and Lomonosov’s theorem [12], that any operator in \(\mathcal {B}(\mathscr {X})\) has a non-trivial invariant subspace and any non-scalar operator in \(\mathcal {B}(\mathscr {X})\) has a non-trivial hyperinvariant subspace. On the other hand, several examples of Banach spaces (including \(\ell _1\)) with operators without a non-trivial invariant subspace are known (see [3, 15] for the first examples and [6] for a general approach to Read’s type constructions of operators without non-trivial invariant closed subspaces).

With the von Neumann–Aronszajn–Smith theorem and Lomonosov’s theorem in mind, it is not a surprise that suitable compactness conditions imply existence of non-trivial invariant and hyperinvariant subspaces for different classes of operators and sets of operators. For instance, Shulman [16, Theorem 2] proved that an algebra of operators whose radical contains a non-zero compact operator has a non-trivial hyperinvariant subspace. Turovskii [17, Corollary 5] extended this result to semigroups of quasinilpotent operators. Another type of results are those related to triangularizability of a set of operators. Recall that a non-empty set \(\mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) is triangularizable if there exists a chain \(\mathscr {C}\) which is maximal as a chain of subspaces of \(\mathscr {X}\) and every subspace in \(\mathscr {C}\) is invariant for all operators in \(\mathcal {S}\). Every commutative set of compact operators is triangularizable (see [14, Theorem 7.2.1]). Konvalinka [10, Corollary 2.6] has extended this result by showing that a commuting family of polynomially compact operators is triangularizable. Another result in this direction, obtained by the second author [9], says that for a norm closed subalgebra \(\mathcal {A}\subseteq \mathcal {B}(\mathscr {X})\) of power compact operators the following assertions are equivalent: (a) \(\mathcal {A}\) is triangularizable; (b) the Jacobson radical \(\mathcal {R}(\mathcal {A})\) consists precisely of quasinilpotent operators in \(\mathcal {A}\); (c) the quotient algebra \(\mathcal {A}/\mathcal {R}(\mathcal {A})\) is commutative.

The aim of this paper is to consider the problem of existence of a non-trivial hyperinvariant subspace for sets of polynomially compact operators. For instance, we prove (Theorem 3.3) that a non-trivial norm closed algebra \(\mathcal {A}\subseteq \mathcal {B}(\mathscr {X})\) which consists of polynomially compact quasinilpotent operators has a non-trivial hyperinvariant subspace. Another result (Theorem 4.3) which we mention here is related to operator bands, that is, to semigroups of idempotent operators. It says that an operator band \(\mathcal {S}\) has a non-trivial hyperinvariant subspace if there exists a non-zero compact operator in the norm closure of the algebra generated by \(\mathcal {S}\).

2 Preliminaries

2.1 Notation

Let \(\mathscr {X}\) be a non-trivial complex Banach space. Since the results proved in this paper are either trivial or well known when \(\mathscr {X}\) is finite-dimensional, we always assume that \(\dim (\mathscr {X})=\infty\). Let \(\mathcal {B}(\mathscr {X})\) denote the Banach algebra of all bounded linear operators on \(\mathscr {X}\) and let \(\mathcal {K}(\mathscr {X})\subseteq \mathcal {B}(\mathscr {X})\) be the ideal of compact operators. The identity operator is denoted by I and an operator is said to be scalar if it is a scalar multiple of I. The norm closure of a set \(\mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) is denoted by \(\overline{\mathcal {S}}\).

For two operators \(S_1, S_2\in \mathcal {B}(\mathscr {X})\), we denote their commutator \(S_1S_2-S_2S_1\) by \([S_1,S_2]\). A non-empty set \(\mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) is commutative if any two operators from \(\mathcal {S}\) commute, that is, \([S_1,S_2]=0\) for all \(S_1,S_2\in \mathcal {S}\). Similarly, \(\mathcal {S}\) is essentially commutative if \([S_1,S_2]\in \mathcal {K}(\mathscr {X})\). For an arbitrary non-empty subset \(\mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) the commutant of \(\mathcal {S}\) is \(\mathcal {S}'=\{ T\in \mathcal {B}(\mathscr {X});\; [T,S]=0,\, \forall \, S\in \mathcal {S}\}\). It is clear that \(\mathcal {S}'\) is a closed subalgebra of \(\mathcal {B}(\mathscr {X})\). If \(\mathcal {S}\) is commutative, then \(\mathcal {S}\subseteq \mathcal {S}'\).

2.2 Invariant subspaces

A non-empty subset \(\mathscr {M}\subseteq \mathscr {X}\) is a subspace if it is a closed linear manifold. It is said that a subspace \(\mathscr {M}\) of \(\mathscr {X}\) is invariant for the operator T if \(T\mathscr {M}\subseteq \mathscr {M}\). An invariant subspace \(\mathscr {M}\) is non-trivial if \(\{ 0\}\ne \mathcal {M}\ne \mathscr {X}\). A non-empty set \(\mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) is reducible if there exists a non-trivial subspace \(\mathscr {M}\subseteq \mathscr {X}\) which is invariant for every \(T\in \mathcal {S}\). If there exists a chain \(\mathscr {C}\) which is maximal as a chain of subspaces of \(\mathscr {X}\) and every subspace in \(\mathscr {C}\) is invariant for all operators in \(\mathcal {S}\), then \(\mathcal {S}\) is said to be triangularizable.

If a subspace \(\mathscr {M}\) is invariant for every operator T in a set \(\mathcal {S}\) and in its commutant \(\mathcal {S}'\), then \(\mathscr {M}\) is said to be a hyperinvariant subspace for \(\mathcal {S}\).

Next theorem (see Assertion in the end of [12]) is one of the deepest results in the theory of invariant subspaces.

Theorem 2.1

(Lomonosov) Every non-scalar operator which commutes with a non-zero compact operator has a non-trivial hyperinvariant subspace.

The proof of Theorem 2.1 relies on the following useful lemma.

Lemma 2.2

(Lomonosov’s Lemma) Let \(\mathcal {A}\subseteq \mathcal {B}(\mathscr {X})\) be an algebra. If \(\mathcal {A}\) is not reducible, then for every non-zero compact operator \(K\in \mathcal {B}(\mathscr {X})\) there exist an operator \(A\in \mathcal {A}\) and a non-zero vector \(x\in \mathscr {X}\) such that \(AKx=x\).

2.3 Spectral radius

The spectrum of an operator \(T\in \mathcal {B}(\mathscr {X})\) is denoted by \(\sigma (T)\) and the spectral radius of T is \(\rho (T)=\max \{ |z|;\; z\in \sigma (T)\}\). By the spectral radius formula (Gelfand’s formula), \(\rho (T)=\lim \limits _{n\rightarrow \infty } \Vert T^n\Vert ^{\frac{1}{n}}\). An operator \(T\in \mathcal {B}(\mathscr {X})\) is quasinilpotent if \(\rho (T)=0\). A compact quasinilpotent operator is called a Volterra operator.

Let \(\mathcal {F}\) be a non-empty set of operators in \(\mathcal {B}(\mathscr {X})\). For each \(n\in \mathbb {N}\), let \(\mathcal {F}^{(n)}=\{T_1\cdots T_n;\; T_1,\ldots , T_n\in \mathcal {F}\}\). By \(\Vert \mathcal {F}\Vert =\sup \{\Vert T\Vert ;\; T\in \mathcal {F}\}\) we denote the joint norm of \(\mathcal {F}\) and by \(\rho (\mathcal {F})\), we denote the joint spectral radius of \(\mathcal {F}\) defined as

$$\begin{aligned} \rho (\mathcal {F})=\limsup _{n\rightarrow \infty } \Vert \mathcal {F}^{(n)}\Vert ^{\frac{1}{n}}. \end{aligned}$$

A subalgebra \(\mathcal {A}\subseteq \mathcal {B}(\mathscr {X})\) is said to be finitely quasinilpotent if \(\rho (\mathcal {F})=0\) for every finite subset \(\mathcal {F}\) of \(\mathcal {A}\). By [16, Theorem 1], every subalgebra \(\mathcal {A}\subseteq \mathcal {B}(\mathscr {X})\) of Volterra operators is finitely quasinilpotent.

2.4 Polynomially compact operators

An operator \(T\in \mathcal {B}(\mathscr {X})\) is polynomially compact if there exists a non-zero complex polynomial p such that p(T) is a compact operator. In particular, algebraic operators (nilpotents, idempotents, etc.) are polynomially compact. Hence, T is polynomially compact if and only if \(\pi (T)\), where \(\pi :\mathcal {B}(\mathscr {X})\rightarrow \mathcal {B}(\mathscr {X})/\mathcal {K}(\mathscr {X})\) is the quotient projection, is an algebraic element in the Calkin algebra \(\mathcal {B}(\mathscr {X})/\mathcal {K}(\mathscr {X})\). If \(T^n\) is compact for some \(n\in \mathbb {N}\), then T is said to be power compact. For a polynomially compact operator T, there exists a unique monic polynomial \(m_T\) of the smallest degree such that \(m_T(T)\) is a compact operator. The polynomial \(m_T\) is called the minimal polynomial of T. The following is the structure theorem for polynomially compact operators proved by Gilfeather [4, Theorem 1].

Theorem 2.3

Let \(T\in \mathcal {B}(\mathscr {X})\) be a polynomially compact operator with minimal polynomial \(m_T(z)=(z-\lambda _1)^{n_1}\cdots (z-\lambda _k)^{n_k}\). Then there exist invariant subspaces \(\mathscr {X}_1,\ldots , \mathscr {X}_k\) for T such that \(\mathscr {X}=\mathscr {X}_1 \oplus \cdots \oplus \mathscr {X}_k\) and \(T=T_1\oplus \cdots \oplus T_k\), where \(T_i\) is the restriction of T to \(\mathscr {X}_i\). The operators \((T_j-\lambda _j I_j)^{n_j}\) are all compact.

The spectrum of T consists of countably many points with \(\{ \lambda _1,\ldots ,\lambda _k\}\) as the only possible limit points and such that all but possibly \(\{ \lambda _1,\ldots ,\lambda _k\}\) are eigenvalues with finite-dimensional generalized eigenspaces. Each point \(\lambda _j\) \((j=1, \ldots ,k)\) is either the limit of eigenvalues of T or else \(\mathscr {X}_j\) is infinite-dimensional and \(T_j-\lambda _j I_j\) is a quasinilpotent operator on \(\mathscr {X}_j\).

Corollary 2.4

A quasinilpotent operator \(T\in \mathcal {B}(\mathscr {X})\) is polynomially compact if and only if it is power compact.

Proof

It is clear that every power compact operator is polynomially compact. On the other hand, if T is a polynomially compact and quasinilpotent, then \(\sigma (T)=\{0\}\) and therefore \(m_T(z)=z^n\) for a positive integer n, by Theorem 2.3, that is, T is power compact. \(\square\)

2.5 Algebras and ideals

For a non-empty set of operators \(\mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\), let \(\mathcal {A}(\mathcal {S})\) be the subalgebra of \(\mathcal {B}(\mathscr {X})\) generated by \(\mathcal {S}\) and let \(\mathcal {A}_1(\mathcal {S})\) be the subalgebra of \(\mathcal {B}(\mathscr {X})\) generated by \(\mathcal {S}\) and I. By \(\mathcal {H}(\mathcal {S})\) we denote the algebra which is generated by \(\mathcal {S}\) and its commutant \(\mathcal {S}'\). We will call it the hyperalgebra of \(\mathcal {S}\). If \(\mathcal {S}\) is a semigroup, then an operator \(T\in \mathcal {B}(\mathscr {X})\) is in \(\mathcal {H}(\mathcal {S})\) if and only if there exist \(n\in \mathbb {N}\) and operators \(S_1, \ldots , S_n\in \mathcal {S}\) and \(T_0, T_1, \ldots , T_n\in \mathcal {S}'\) such that

$$\begin{aligned} T=T_0+S_1 T_1+\cdots + S_n T_n=T_0+T_1 S_1+\cdots + T_n S_n . \end{aligned}$$

Here, we used the fact that \(I\in \mathcal {S}'\). Since \(\mathcal {S}\subseteq \mathcal {H}(\mathcal {S})\) and \(\mathcal {A}(\mathcal {S})\) is the smallest algebra which contains \(\mathcal {S}\), we have \(\mathcal {A}(\mathcal {S})\subseteq \mathcal {H}(\mathcal {S})\). On the other hand, it is obvious that \(\mathcal {S}'=\mathcal {A}(\mathcal {S})'\) and therefore \(\mathcal {A}(\mathcal {S})'\subseteq \mathcal {H}(\mathcal {S})\). We conclude that the hyperalgebra of \(\mathcal {S}\) is generated by \(\mathcal {A}(\mathcal {S})\) and \(\mathcal {A}(\mathcal {S})'\), that is, \(\mathcal {H}(\mathcal {S})=\mathcal {H}\bigl (\mathcal {A}(\mathcal {S})\bigr )\).

If \(\mathcal {M}\) and \(\mathcal {N}\) are non-empty subsets of \(\mathcal {B}(\mathscr {X})\), then let \(\mathcal {M}+\mathcal {N}=\{ M+N;\; M\in \mathcal {M}, N\in \mathcal {N}\}\) and let \(\mathcal {M}\mathcal {N}\) be the set of all finite sums \(M_1 N_1+\cdots +M_k N_k\), where \(M_i\in \mathcal {M}\) and \(N_i\in \mathcal {N}\) for each \(i=1, \ldots , k\). Hence, if \(\mathcal {A}\) is a subalgebra of \(\mathcal {B}(\mathscr {X})\), then its hyperalgebra is \(\mathcal {H}(\mathcal {A})=\mathcal {A}'+\mathcal {A}\mathcal {A}'=\mathcal {A}'+\mathcal {A}'\mathcal {A}\).

If \(\mathcal {J}\) is an ideal in an algebra \(\mathcal {A}\subseteq \mathcal {B}(\mathscr {X})\), then we write \(\mathcal {J}\triangleleft \mathcal {A}\). We will denote by \(\mathcal {J}_{_{\mathcal {H}}}\) the ideal in \(\mathcal {H}(\mathcal {A})\) generated by \(\mathcal {J}\).

Lemma 2.5

Let \(\mathcal {A}\subseteq \mathcal {B}(\mathscr {X})\) be an algebra. If \(\mathcal {J}\triangleleft \mathcal {A}\), then \(\mathcal {J}_{_{\mathcal {H}}}=\mathcal {A}'\mathcal {J}=\mathcal {J}\mathcal {A}'\).

Proof

From equalities \(\mathcal {J}_{_{\mathcal {H}}}=\mathcal {J}+\mathcal {H}(\mathcal {A})\mathcal {J}+\mathcal {J}\mathcal {H}(\mathcal {A})+\mathcal {H}(\mathcal {A})\mathcal {J}\mathcal {H}(\mathcal {A})\) and \(\mathcal {H}(\mathcal {A})=\mathcal {A}'+\mathcal {A}'\mathcal {A}\), we conclude

$$\begin{aligned} \mathcal {J}_{_{\mathcal {H}}}&=\mathcal {J}+\mathcal {H}(\mathcal {A})\mathcal {J}+\mathcal {J}\mathcal {H}(\mathcal {A})+\mathcal {H}(\mathcal {A})\mathcal {J}\mathcal {H}(\mathcal {A})\\&=\mathcal {J}+\mathcal (\mathcal {A}'+\mathcal {A}'\mathcal {A})\mathcal {J}+\mathcal {J}\mathcal (\mathcal {A}'+\mathcal {A}'\mathcal {A})+\mathcal (\mathcal {A}'+\mathcal {A}'\mathcal {A})\mathcal {J}\mathcal (\mathcal {A}'+\mathcal {A}'\mathcal {A})\\&\subseteq \mathcal {J}+\mathcal {A}'\mathcal {J}\subseteq \mathcal {A}' \mathcal {J}\end{aligned}$$

as \(\mathcal {A}'\) contains the identity operator. On the other hand, since we also have \(\mathcal {A}'\mathcal {J}\subseteq \mathcal {J}_{_{\mathcal {H}}}\), we obtain the equality \(\mathcal {J}_{_{\mathcal {H}}}=\mathcal {A}'\mathcal {J}\). \(\square\)

An algebra \(\mathcal {A}\) over an arbitrary field is said to be a nil-algebra if every element of \(\mathcal {A}\) is nilpotent. A nil-algebra \(\mathcal {A}\) is of bounded nil-index if there exists a positive integer n such that \(x^n=0\) for each \(x\in \mathcal {A}\). If there exists \(n\in \mathbb {N}\) such that \(a_1\cdots a_n=0\) for all \(a_1,\ldots ,a_n\in \mathcal {A}\), then \(\mathcal {A}\) is said to be a nilpotent algebra. By the celebrated Nagata-Higman theorem (see [8, 13]), every nil-algebra of bounded nil-index is nilpotent.

Lemma 2.6

Let \(\mathcal {A}\subseteq \mathcal {B}(\mathscr {X})\) be an algebra. An ideal \(\mathcal {J}\triangleleft \mathcal {A}\) is nilpotent if and only if \(\mathcal {J}_{_{\mathcal {H}}}\triangleleft \mathcal {H}(\mathcal {A})\) is nilpotent. The nilpotency indices of \(\mathcal {J}\) and \(\mathcal {J}_{_{\mathcal {H}}}\) are equal.

Proof

Since \(\mathcal {A}'\) and \(\mathcal {J}\) commute, an easy induction shows that for each \(n\in \mathbb {N}\) we have \((\mathcal {A}'\mathcal {J})^n=\mathcal {A}' \mathcal {J}^n\). Hence, if \(\mathcal {J}^n=\{0\}\), then \((\mathcal {A}'\mathcal {J})^n=\{0\}\), as well. If \((\mathcal {A}'\mathcal {J})^n=\{0\}\), then \(\mathcal {J}^n \subseteq \mathcal {A}'\mathcal {J}^n=(\mathcal {A}'\mathcal {J})^n=\{0\}\) yields that \(\mathcal {J}^n=\{0\}\). \(\square\)

If \(\mathcal {J}\) is a nil-ideal of bounded nil-index, then \(\mathcal {J}\) is nilpotent by the Nagata-Higman theorem. This immediately implies that \(\mathcal {J}_{_{\mathcal {H}}}\) is nilpotent. In particular, if \(\mathcal {A}\) is a nilpotent algebra, then \(\mathcal {A}\mathcal {A}'\) is a nilpotent ideal in the hyperalgebra \(\mathcal {H}(\mathcal {A})\).

3 Hyperinvariant subspaces of algebras of polynomially compact operators

The simplest polynomially compact operators which are not necessary compact are algebraic operators, in particular nilpotent operators. Hadwin et al. [7, Corollary 4.2] proved that a norm closed algebra of nilpotent operators on the separable infinite-dimensional complex Hilbert space is triangularizable. The following proposition shows that a norm closed algebra of nilpotent operators on an arbitrary complex Banach space has a non-trivial hyperinvariant subspace.

Proposition 3.1

If a subalgebra \(\{ 0\}\ne \mathcal {A}\subseteq \mathcal {B}(\mathscr {X})\) consists of nilpotent operators, then its hyperalgebra \(\mathcal {H}(\mathcal {A})\) is reducible in either of the following cases.

  1. (a)

    The algebra \(\mathcal {A}\) is nilpotent.

  2. (b)

    The algebra \(\mathcal {A}\) is norm closed.

Proof

(a) There exists \(n\in \mathbb {N}\) such that an arbitrary product of at least n operators from \(\mathcal {A}\) is the zero operator. Let \(n_0\) be the smallest positive integer with this property. Since \(\mathcal {A}\ne \{ 0\}\) we have \(n_0>1.\) There exist operators \(A_1,\ldots ,A_{n_0-1}\in \mathcal {A}\) such that \(A_0:=A_1,\ldots ,A_{n_0-1}\ne 0\). Note that \(A_0T=TA_0=0\) for every operator \(T\in \mathcal {A}\), that is, \(\mathcal {A}\subseteq (A_0)'\), where \((A_0)'\) is the commutant of \(A_0\). It is clear that \(\mathcal {A}'\subseteq (A_0)'\). Hence, \(\mathcal {H}(\mathcal {A})\subseteq (A_0)'\). Since \(A_0\ne 0\) the kernel \(\ker (A_0)\) is a non-trivial subspace of \(\mathscr {X}\) and it is hyperinvariant for \(A_0\). It follows that \(\ker (A_0)\) is a non-trivial hyperinvariant subspace for \(\mathcal {A}\), that is, \(\mathcal {H}(\mathcal {A})\) is reducible.

(b) Since \(\mathcal {A}\) is a closed subalgebra of \(\mathcal {B}(\mathscr {X})\), it is a Banach algebra. Therefore, \(\mathcal {A}\) is a Banach algebra which is also a nil-algebra, so that by Grabiner’s theorem [5], \(\mathcal {A}\) is a nilpotent algebra. Now, we apply (a). \(\square\)

Hadwin et al. have constructed a semi-simple algebra of nilpotent operators on a separable Hilbert space such that for each pair (xy) of vectors, where \(x\ne 0\), there exists an operator \(A\in \mathcal {A}\) such that \(Ax=y\) (see [7, Section 4]). Hence, Proposition 3.1(b) (and consequently Theorem 3.3 below) does not hold, in general, for non-closed algebras. However, as the following example shows there are simple special cases when a not necessarily closed algebra generated by a set of nilpotents has a non-trivial hyperinvariant subspace.

Example

The idea for this example is from [7, Theorem 4.3] where quadratic nilpotent operators are considered. Choose and fix \(\lambda \in \mathbb {C}\). Let \(\mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) be a non-empty set of operators such that \(A^2=\lambda A\) for every \(A\in \mathcal {S}\). If \(\mathcal {S}\) is a linear manifold and contains a non-scalar operator, then \(\mathcal {A}(\mathcal {S})\) has a non-trivial hyperinvariant subspace. To see this, choose a non-scalar operator \(A \in \mathcal {S}\). Then its kernel \(\ker (A)\) is a non-trivial hyperinvariant subspace for A. Since \(\mathcal {A}(\mathcal {S})'\subseteq (A)'\), we see that \(\ker (A)\) is invariant for every operator from \(\mathcal {A}(\mathcal {S})'\). Let \(B\in \mathcal {S}\) be arbitrary. It follows from \((A+B)^2=\lambda (A+B)\) that \(AB=-BA\). Hence, \(ABx=-BAx=0\) for every \(x\in \ker (A)\), that is, \(\ker (A)\) is invariant for every \(B\in \mathcal {S}\). We conclude that \(\ker (A)\) is invariant for every operator from the hyperalgebra \(\mathcal {H}(\mathcal {S})\).

Proposition 3.2

Let \(\mathcal {A}\subseteq \mathcal {B}(\mathscr {X})\) be a non-trivial subalgebra. If there exists a non-trivial nilpotent ideal \(\mathcal {J}\triangleleft \mathcal {A}\), then \(\mathcal {A}\) has a non-trivial hyperinvariant subspace.

Proof

Let \(\mathcal {J}\) be a non-trivial nilpotent ideal in \(\mathcal {A}(\mathcal {S})\). Then, by Lemma 2.6, the ideal \(\mathcal {J}_{_{\mathcal {H}}}\) which is generated by \(\mathcal {J}\) in the hyperalgebra \(\mathcal {H}(\mathcal {A})\) is nilpotent, as well. By Proposition 3.1(a) the ideal \(\mathcal {J}_{_{\mathcal {H}}}\) has a non-trivial hyperinvariant subspace, in particular, it is reducible. By [14, Lemma 7.4.6], \(\mathcal {H}(\mathcal {A})\) is reducible, as well. \(\square\)

The following two theorems show that an algebra \(\mathcal {A}\) of polynomially compact operators has a non-trivial hyperinvariant subspace if \(\mathcal {A}\) satisfies some additional condition. For instance, as we already mentioned, the key assumption in Theorem 3.3 is that the involved algebra is norm closed.

Theorem 3.3

Let \(\{ 0\}\ne \mathcal {A}\subseteq \mathcal {B}(\mathscr {X})\) be a norm closed algebra. If every operator in \(\mathcal {A}\) is quasinilpotent and polynomially compact, then \(\mathcal {A}\) has a non-trivial hyperinvariant subspace.

Proof

Note first that each operator in \(\mathcal {A}\) is power compact, by Corollary 2.4. If each operator in \(\mathcal {A}\) is nilpotent, then \(\mathcal {H}(\mathcal {A})\) is reducible, by Proposition 3.1(b). Assume therefore that there exists an operator \(T\in \mathcal {A}\) which is not nilpotent. Since T is power compact there exists \(m\in \mathbb {N}\) such that \(K=T^m\ne 0\) is compact. Let \(\mathcal {J}\) be the two-sided ideal in \(\mathcal {A}_1\) generated by K. It is clear that \(K\in \mathcal {J}\subseteq \mathcal {A}\). Hence, \(\mathcal {J}\) is an algebra of Volterra operators. By [16, Theorem 1], \(\mathcal {J}\) is finitely quasinilpotent. Let \(A=B_0+\sum _{i=1}^{n} A_iB_i\), where \(A_i\in \mathcal {A}\) and \(B_i\in \mathcal {A}'\), be an arbitrary operator in \(\mathcal {H}(\mathcal {A})\) and let \(\mathcal {M}=\{K,A_1K,\ldots ,A_nK\}\). Since \(\mathcal {M}\) is a finite subset of \(\mathcal {J}\) it is a quasinilpotent set, that is, \(\rho (\mathcal {M})=0\). Let \(\mathcal {N}=\{B_0, B_1,\ldots ,B_n\}\subseteq \mathcal {A}'\). Since \(B_0, B_1,\ldots ,B_n\) commute with operators from \(\mathcal {M}\) we have \(\rho (AK)\le (n+1)\rho (\mathcal {M}) \rho (\mathcal {N})=0\), by [16, Lemma 1]. This shows that the operator AK is quasinilpotent for each \(A\in \mathcal {H}(\mathcal {A})\). It follows, by Lemma 2.2, that the hyperalgebra \(\mathcal {H}(\mathcal {A})\) is reducible.

Theorem 3.4

Let \(\mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) be an essentially commutative set of polynomially compact operators which contains at least one non-scalar operator. If \(\mathcal {S}\) is triangularizable, then \(\mathcal {A}(\mathcal {S})\) has a non-trivial hyperinvariant subspace.

Proof

It is obvious that every subspace of \(\mathscr {X}\) which is invariant for \(\mathcal {S}\) is invariant for \(\mathcal {A}(\mathcal {S})\), as well. Hence, \(\mathcal {A}(\mathcal {S})\) is triangularizable. Since \(\mathcal {S}\) consists of polynomially compact operators, the same holds for the algebra \(\mathcal {A}(\mathcal {S})\), by [10, Theorem 1.5]. Denote by \(\pi \bigl (\mathcal {A}(\mathcal {S})\bigr )\) and \(\pi (\mathcal {S})\) the image of \(\mathcal {A}(\mathcal {S})\) and \(\mathcal {S}\), respectively, in the Calkin algebra \(\mathcal {B}(\mathscr {X})/\mathcal {K}(\mathscr {X})\). Since \(\pi \bigl (\mathcal {A}(\mathcal {S})\bigr )\) is the subalgebra of the Calkin algebra generated by the commutative set \(\pi (\mathcal {S})\) we see that \(\pi \bigl (\mathcal {A}(\mathcal {S})\bigr )\) is commutative, as well. It follows that \(\mathcal {A}(\mathcal {S})\) is an essentially commutative subalgebra of \(\mathcal {B}(\mathscr {X})\). By [9, Theorem 3.5], for an algebra of essentially commuting polynomially compact operators triangularizability of \(\mathcal {A}(\mathcal {S})\) is equivalent to the fact that each commutator [ST], where \(S, T\in \mathcal {A}(\mathcal {S})\), is a quasinilpotent operator.

Suppose that there exist operators \(S, T\in \mathcal {A}(\mathcal {S})\) such that the commutator \(K:=[S,T]\) is non-zero. Let \(\mathcal {J}\) be the ideal in \(\mathcal {A}(\mathcal {S})\) generated by K. Of course, \(\mathcal {J}\) consists of compact operators and, by Ringrose’s Theorem (see [14, Theorem 7.2.3]), every operator in \(\mathcal {J}\) is quasinilpotent. Hence, \(\mathcal {J}\) is an algebra of Volterra operators. Let \(\mathcal {J}_{_{\mathcal {H}}}\) be the ideal in the hyperalgebra \(\mathcal {H}(\mathcal {S})\) generated by \(\mathcal {J}\). It is clear that \(\mathcal {J}_{_{\mathcal {H}}}\subseteq \mathcal {K}(\mathscr {X})\). We claim that \(\mathcal {J}_{_{\mathcal {H}}}\) consists of quasinilpotent operators, as well. Choose an arbitrary operator \(A\in \mathcal {J}_{_{\mathcal {H}}}\). By Lemma 2.5, there exist operators \(J_1,\ldots ,J_n\in \mathcal {J}\) and operators \(B_1,\ldots ,B_n\in \mathcal {A}(\mathcal {S})'\) such that \(A=\sum _{i=1}^{n} J_i B_i\). By [16, Lemma 1], \(\rho (A)\le n \rho (\{J_1,\ldots ,J_n\}) \rho (\{B_1,\ldots , B_n\})\). Since each finite subset of an algebra of Volterra operators is quasinilpotent, by [16, Theorem 1], we have \(\rho (\{J_1,\ldots ,J_n\})=0\) and consequently \(\rho (A)=0\). We have proved that \(\mathcal {J}_{_{\mathcal {H}}}\) is a non-trivial ideal of Volterra operators. By [16, Theorem 2], the hyperalgebra \(\mathcal {H}(\mathcal {S})\) is reducible.

Suppose now that \(\mathcal {A}(\mathcal {S})\) is commutative. Hence, \(\mathcal {A}(\mathcal {S})\subseteq \mathcal {A}(\mathcal {S})'\). Let T be any non-scalar polynomially compact operator in \(\mathcal {A}(\mathcal {S})\) and let \(m_T\) be its minimal polynomial. Hence, \(m_T(T)\) is a compact operator. If m(T) is a non-zero compact operator, then it has a non-trivial hyperinvariant subspace \(\mathscr {Y}\), by Lomonosov’s Theorem. Since \(\mathcal {A}(\mathcal {S})\subseteq \mathcal {A}(\mathcal {S})'\subseteq (m_T(T))'\) subspace \(\mathscr {Y}\) is invariant for every operator in \(\mathcal {A}(\mathcal {S})\) and \(\mathcal {A}(\mathcal {S})'\). Thus, \(\mathcal {H}(\mathcal {S})\) is reducible. On the other hand, if \(m_T(T)=0\), then T is a non-scalar algebraic operator. Hence, for every \(\lambda \in \sigma (T)\), the kernel \(\ker (T-\lambda I)\) is a non-trivial hyperinvariant subspace for T, and consequently, for \(\mathcal {A}(\mathcal {S})\) as \(\mathcal {A}(\mathcal {S})\subseteq \mathcal {A}(\mathcal {S})'\subseteq (T)'\). \(\square\)

4 Hyperinvariant subspaces for operator bands

An operator band on a Banach space \(\mathscr {X}\) is a (multiplicative) semigroup \(\mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) of idempotents, that is, \(S^2=S\) for each \(S\in \mathcal {S}\). The linear span of an operator band \(\mathcal {S}\) is the algebra \(\mathcal {A}(\mathcal {S})\) called a band algebra. We will denote by \(\mathcal {N}(\mathcal {S})\), the set of all nilpotent operators in the band algebra \(\mathcal {A}(\mathcal {S})\). By [11, Theorem 5.2], \(\mathcal {N}(\mathcal {S})\) is the linear span of \([\mathcal {S},\mathcal {S}]\), that is, \(\mathcal {N}(\mathcal {S})=\{ \sum _{i=1}^{n}[A_i,B_i];\; n\in \mathbb {N},\, A_i,B_i\in \mathcal {A}(\mathcal {S})\}\). By [11, Corollary 5.5], the set \(\mathcal {N}(\mathcal {S})\) of nilpotent operators in \(\mathcal {A}(\mathcal {S})\) coincides with the Jacobson radical \({\mathcal {R}}\bigl (\mathcal {A}(\mathcal {S})\bigr )\) of \(\mathcal {A}(\mathcal {S})\).

Proposition 4.1

Let \(\{ 0\}\ne \mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) be an operator band. If \(\mathcal {N}\subseteq \mathcal {N}(\mathcal {S})\) is a non-empty set, then the ideal \(\mathcal {J}_{_{\mathcal {H}}}\) in the hyperalgebra \(\mathcal {H}(\mathcal {S})\) generated by \(\mathcal {N}\) is a nil-ideal.

Proof

Denote by \(\mathcal {J}\) the ideal in \(\mathcal {A}(\mathcal {S})\) generated by \(\mathcal {N}\). We have already mentioned that \({\mathcal {R}}\bigl (\mathcal {A}(\mathcal {S})\bigr )=\mathcal {N}(\mathcal {S})\). Since the radical is an ideal, we have that \(\mathcal {J}\subseteq {\mathcal {R}}\bigl (\mathcal {A}(\mathcal {S})\bigr )\). Hence, \(\mathcal {J}\) is a nil-ideal in \(\mathcal {A}(\mathcal {S})\).

It is clear that \(\mathcal {N}\) and ideal \(\mathcal {J}\) generate the same ideal \(\mathcal {J}_{_{\mathcal {H}}}\) in the hyperalgebra \(\mathcal {H}(\mathcal {S})\). Let \(A\in \mathcal {J}_{_{\mathcal {H}}}\) be arbitrary. By Lemma 2.5, there exist \(n\in \mathbb {N}\), \(J_i\in \mathcal {J}\) and \(B_i\in \mathcal {A}(\mathcal {S})'\) such that \(A=\sum _{i=1}^n J_i B_i\). Let \(\mathcal {F}\) be a finite subset of \(\mathcal {S}\) such that \(J_1,\ldots ,J_n\) are contained in the algebra \(\mathcal {A}(\mathcal {F})\) generated by \(\mathcal {F}\). Note that such finite sets exist, because each \(J_i\) is of the form \(p_i(S_{1}^{(i)}, \ldots , S_{k_i}^{(i)})\), where \(p_i\) is a polynomial of \(k_i\) non-commuting variables and \(S_{1}^{(i)}, \ldots , S_{k_i}^{(i)}\) \((i=1, \ldots , n)\) are idempotents from \(\mathcal {S}\). Denote by \(\mathcal {S}_0\) the operator band generated by \(\mathcal {F}\). Thus, \(\mathcal {A}(\mathcal {F})\subseteq \mathcal {A}(\mathcal {S}_0)\). Since \(\mathcal {F}\) is finite, the operator band \(\mathcal {S}_0\) is finite, as well, by the Green-Rees theorem (see [14, Theorem 9.3.11]). Now, we apply [14, Theorem 9.3.15] which says that there exists a finite chain \(\{0\}=\mathscr {X}_0\subseteq \mathscr {X}_1\subseteq \cdots \subseteq \mathscr {X}_{m}=\mathscr {X}\) of invariant subspaces for \(\mathcal {S}_0\) such that for each \(E\in \mathcal {S}_0\) the operator induced by E on \(\mathscr {X}_i/\mathscr {X}_{i-1}\) is either zero or the identity operator. This implies that operator induced by \(J_i\) on \(\mathscr {X}_j/\mathscr {X}_{j-1}\) is a scalar multiple of the identity operator. Since each \(J_i\) is nilpotent, every operator \(J_i\) induces the zero operator on \(\mathscr {X}_j/\mathscr {X}_{j-1}\). Hence, for each i and each j, we have \(J_i(\mathscr {X}_j)\subseteq \mathscr {X}_{j-1}\). This implies that an arbitrary product of length at least m with letters from \(\{J_1,\ldots ,J_n\}\) is zero. Now it is obvious that \(A^{m}=\left( \sum _{i=1}^n J_i B_i\right) ^m=0\) as \(J_i\) and \(B_j\) commute. \(\square\)

Proposition 4.2

Let \(\{ 0\}\ne \mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) be an operator band. If \(K\in \overline{\mathcal {A}(\mathcal {S})}\) is a compact operator, then for each operator \(A\in \overline{\mathcal {H}(\mathcal {S})}\) the commutator [KA] is in the Jacobson radical of \(\overline{\mathcal {H}(\mathcal {S})}\), that is, \([K,\overline{\mathcal {H}(\mathcal {S})}]\subseteq {\mathcal {R}}\bigl (\overline{\mathcal {H}(\mathcal {S})}\bigr )\).

Proof

Let \(K\in \overline{\mathcal {A}(\mathcal {S})}\) be a compact operator and let \(A\in \overline{\mathcal {H}(\mathcal {S})}\) be an arbitrary operator. To prove that the commutator [KA] is in \({\mathcal {R}}\bigl (\overline{\mathcal {H}(\mathcal {S})}\bigr )\), we need to show that [KA]C is quasinilpotent for each \(C\in \overline{\mathcal {H}(\mathcal {S})}\). Since [KA]C is compact, the continuity of the spectral radius at compact operators yields that it suffices to prove that [KA]C is quasinilpotent for all \(A, C\in \mathcal {H}(\mathcal {S})\).

Let \(A,C\in \mathcal {H}(\mathcal {S})\) be arbitrary. Since K belongs to the closure of \(\mathcal {A}(\mathcal {S})\), there is a sequence \((K_n)_{n\in \mathbb {N}}\subseteq \mathcal {A}(\mathcal {S})\) which converges to K. We claim that for each \(n\in \mathbb {N}\) the commutator \([K_n,A]\) is nilpotent and belongs to the ideal \(\mathcal {N}(\mathcal {S})_{_{\mathcal {H}}}\) in \(\mathcal {H}(\mathcal {S})\) generated by the Jacobson radical \(\mathcal {N}(\mathcal {S})\) of the band algebra \(\mathcal {A}(\mathcal {S})\). Since \(A\in \mathcal {H}(\mathcal {S})\), there exist operators \(A_1,\ldots , A_k\in \mathcal {A}(\mathcal {S})\) and \(B_0,\ldots ,B_k\in \mathcal {A}(\mathcal {S})'\) such that \(A=B_0+\sum _{i=1}^kA_iB_i\). It follows that \([K_n,A]=\sum _{i=1}^k[K_n,A_i]B_i.\) Since every commutator \([K_n,A_i]\) is in \(\mathcal {N}(\mathcal {S})\) and \(\mathcal {N}(\mathcal {S})\subseteq \mathcal {N}(\mathcal {S})_{_{\mathcal {H}}}\), it follows \([K_n,A_i] \in \mathcal {N}(\mathcal {S})_{_{\mathcal {H}}}\). Since \(\mathcal {N}(\mathcal {S})_{_{\mathcal {H}}}\) is an ideal in the hyperalgebra \(\mathcal {H}(\mathcal {S})\) we have \([K_n,A]C\in \mathcal {N}(\mathcal {S})_{_{\mathcal {H}}}\). By Proposition 4.1, the operator \([K_n,A]C\) is nilpotent. To finish the proof, we apply the fact that the spectral radius is continuous at compact operators and that the compact operator [KA]C is the limit of the sequence \(([K_n,A]C)_{n\in \mathbb {N}}\) of nilpotent operators. \(\square\)

Theorem 4.3

Let \(\mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) be an operator band. If there exists a non-zero compact operator in \(\overline{\mathcal {A}(\mathcal {S})}\), then \(\mathcal {S}\) has a non-trivial hyperinvariant subspace.

In this proof we added the closures (\overline) eight times around \mathcal {H}(\mathcal {S})

Let K be a non-zero compact operator in \(\overline{\mathcal {A}(\mathcal {S})}\). Then, by Proposition 4.2, [KA] is a compact operator contained in the radical \({\mathcal {R}}\bigl (\overline{\mathcal {H}(\mathcal {S})}\bigr )\) for each \(A\in \overline{\mathcal {H}(\mathcal {S})}\). If \([K,A]=0\) for every \(A\in \overline{\mathcal {H}(\mathcal {S})}\), then \(\overline{\mathcal {H}(\mathcal {S})}\subseteq (K)'\) and, therefore, every non-trivial hyperinvariant subspace of K is a non-trivial hyperinvariant subspace for \(\mathcal {A}(\mathcal {S})\). Hence, we may assume that for some \(A\in \overline{\mathcal {H}(\mathcal {S})}\) the commutator [KA] is a non-zero operator. Let \(\mathcal {J}\) be the ideal in \(\overline{\mathcal {H}(\mathcal {S})}\) generated by [KA]. Of course, each operator in \(\mathcal {J}\) is compact. Since, by Proposition 4.2, [KA] is in the Jacobson ideal of \({\mathcal {R}}\bigl (\overline{\mathcal {H}(\mathcal {S})}\bigr )\) we have \(\mathcal {J}=\overline{\mathcal {H}(\mathcal {S})} [K,A] \overline{\mathcal {H}(\mathcal {S})}\subseteq \overline{\mathcal {H}(\mathcal {S})} {\mathcal {R}}\bigl (\overline{\mathcal {H}(\mathcal {S})}\bigr ) \overline{\mathcal {H}(\mathcal {S})}\subseteq {\mathcal {R}}\bigl (\overline{\mathcal {H}(\mathcal {S})}\bigr )\). It follows that operators in \(\mathcal {J}\) are quasinilpotent. Thus, \(\mathcal {J}\) is a Volterra ideal in \(\overline{\mathcal {H}(\mathcal {S})}\). By [16, Theorem 2], \(\mathcal {J}\) is reducible. Now we apply [14, Lemma 7.4.6] and conclude that \(\overline{\mathcal {H}(\mathcal {S})}\) is reducible. \(\square\)

Corollary 4.4

Every essentially commuting non-scalar operator band \(\mathcal {S}\subseteq \mathcal {B}(\mathscr {X})\) has a non-trivial hyperinvariant subspace.

Proof

If \(\mathcal {S}\) is commutative, then \(\mathcal {S}\subseteq \mathcal {S}'\). In this case the kernel of any non-scalar operator from \(\mathcal {S}\) is invariant for each \(S\in \mathcal {S}'\). If \(\mathcal {S}\) is not commutative, then there exist idempotents \(E,F\in \mathcal {S}\) with a non-zero compact commutator \(EF-FE\in \mathcal {A}(\mathcal {S})\). The assertion follows, by Theorem 4.3. \(\square\)

Since every essentially commuting band of operators has an invariant subspace, an application of the Triangularization lemma (see [14, Lemma 7.1.11]) immediately implies the following result.

Corollary 4.5

Every essentially commuting band of operators on a Banach space is triangularizable.