This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The essential spectrum, norm, and spectral radius of
abstract multiplication operators.

Anton R. Schep Department of Mathematics
University of South Carolina
Columbia, SC 29208
[email protected]
Abstract.

Let EE be a complex Banach lattice and TT is an operator in the center Z(E)={T:|T|λI for some λ}Z(E)=\{T:|T|\leq\lambda I\mbox{ for some }\lambda\} of EE. Then the essential norm Te\|T\|_{e} of TT equals the essential spectral radius re(T)r_{e}(T) of TT. We also prove re(T)=max{TAd,re(TA)}r_{e}(T)=\max\{\|T_{A^{d}}\|,r_{e}(T_{A})\}, where TAT_{A} is the atomic part of TT and TAdT_{A^{d}} is the non-atomic part of TT. Moreover re(TA)=lim supλar_{e}(T_{A})=\limsup_{\mathcal{F}}\lambda_{a}, where \mathcal{F} is the Fréchet filter on the set AA of all positive atoms in EE of norm one and λa\lambda_{a} is given by TAa=λaaT_{A}a=\lambda_{a}a for all aAa\in A.

2000 Mathematics Subject Classification:
Primary: 47G10; Secondary: 47B65, 47B34

1. Introduction.

Over the past two decades numerous authors have studied boundedness, invertibility and compactness of multiplication operators on a variety of spaces of measurable functions on a measure space (X,μ)(X,\mu) (or sequence spaces, when the measure is atomic). We refer to a recent preprint of J. Voigt ([7]) for some references. It seems that these authors are not aware that these multiplication operators are examples of order bounded band preserving operators (or orthomorphisms). Zaanen showed in 1977 in [9] already that the collection of all such orthomorphisms is lattice isometric to L(X,μ)L^{\infty}(X,\mu), so that a function defining a multiplication operator has to be essentially bounded and the operator norm of the multiplication operator equals the LL^{\infty}-norm. Around that time it was also established that for an arbitrary Banach lattice EE the collection of band preserving operators was equal to the center Z(E)Z(E), where Z(E)={T(E):|T|λI for some λ}Z(E)=\{T\in\mathcal{L}(E):|T|\leq\lambda I\mbox{ for some }\lambda\} and that T=inf{λ:|T|λI}\|T\|=\inf\{\lambda:|T|\leq\lambda I\} for TZ(E)T\in Z(E). For this and additional basic properties of the center we refer to the books [4] and [10]. In particular, Proposition 3.1.3 of [4] shows that for a Dedekind complete Banach lattice the center as defined here coincides with the collection of regular operators which commute with band projections. It was proved in [5] that Z(E)Z(E) is a full sub-algebra of the algebra (E)\mathcal{L}(E) of norm bounded operators, i.e., TZ(E)T\in Z(E) is invertible in (E)\mathcal{L}(E) if and only if there exists c>0c>0 such that |T|cI|T|\geq cI. Therefore σ(T)\sigma(T) is equal to the essential range of TT, when TT is a multiplication operator on a Banach function space. Moreover the spectral radius r(T)=Tr(T)=\|T\|. The study of the essential spectrum of operators TZ(E)T\in Z(E) was initiated in [5], where it was proved that σe(T)=σ(T)\sigma_{e}(T)=\sigma(T) for TZ(E)T\in Z(E) with EE a non-atomic Banach lattice. Here the essential spectrum σe(T)\sigma_{e}(T) denotes the spectrum of T+𝒦(E)T+\mathcal{K}(E) in the Calkin algebra (E)/𝒦(E)\mathcal{L}(E)/\mathcal{K}(E). A few years later in [1] other spectra and essential spectra of operators TZ(E)T\in Z(E) were considered and an alternative proof of σe(T)=σ(T)\sigma_{e}(T)=\sigma(T) for TZ(E)T\in Z(E) with EE a non-atomic Banach lattice was given. A description of compact TZ(E)T\in Z(E) can be derived from the more general description of compact disjointness preserving operators due to De Pagter (unpublished) and Wickstead [8].

More recently the essential norm Te\|T\|_{e} in the Calkin algebra has been considered for certain concrete Banach sequence spaces and LpL_{p}-spaces (see [7] and [2]) for more references. The current paper was prompted by the preprint [7], which didn’t use our results from [5]. By using the term abstract multiplication operator in our title we hope that future duplication of our results by authors working on these questions for concrete function spaces will be avoided by a web search or MathScinet search. The current paper describes completely the essential spectrum and norm of an operator TZ(E)T\in Z(E) for an arbitrary Banach lattice EE. The paper is organized as follows. In section 2 we describe for non-atomic Banach lattices the essential spectrum and norm for an operator TZ(E)T\in Z(E). To be able to describe the essential spectral radius for the atomic case where we might have uncountably many atoms, we describe in section 3 what we call Fréchet cluster points of a complex valued function and the Fréchet limit superior and inferior of the modulus of such a function. For the countable case this is standard textbook material, but for the uncountable case one can find some of this in General Topology textbooks, but for the convenience of the reader we have included this section. Then in section 4 the atomic case is handled. In the process of proving the results we reprove the description of a compact TZ(E)T\in Z(E). In the final section we combine the atomic and non-atomic case to describe the essential spectrum and norm of TZ(E)T\in Z(E) for an arbitrary Banach lattice EE. As in [7] we associate in that case with TZ(E)T\in Z(E) an atomic part TAT_{A} and non-atomic part TAdT_{A^{d}}, but in general (i.e., when EE is not Dedekind complete) we don’t have a decomposition of TT, as these parts are not defined on all of EE. We give in this section a concrete example to illustrate this. We conclude the paper by applying our results to prove that if TZ(E)T\in Z(E) is essentially quasi-nilpotent, then TT is compact (and thus T+𝒦(E)=0T+\mathcal{K}(E)=0 in the Calkin algebra).

2. The non-atomic case.

Let EE be a complex Banach lattice. Let AEA\subset E. Then the disjoint complement AdA^{d} of AA is defined as Ad={xE:|x||a|=0 for all aA}A^{d}=\{x\in E:|x|\wedge|a|=0\mbox{ for all }a\in A\}. Then AddA^{dd} is defined as (Ad)d(A^{d})^{d}. Recall that aEa\in E is called an atom, if |x||a||x|\leq|a| implies that x=λax=\lambda a for some scalar λ\lambda, i.e., {a}dd={λa:λ}\{a\}^{dd}=\{\lambda a:\lambda\in\mathbb{C}\}. Now EE is called non-atomic if EE doesn’t contain any atoms. Typical examples of non-atomic Banach lattices are Banach function spaces over a non-atomic σ\sigma-finite measure space. The following theorem was proved by the author in 1980 in [5] (Theorem 1.11).

Theorem 2.1.

Let EE be a complex non-atomic Banach lattice and TZ(E)T\in Z(E). Then σ(T)=σe(T)\sigma(T)=\sigma_{e}(T) and thus re(T)=r(T)=Tr_{e}(T)=r(T)=\|T\|.

Using the methods of the above mentioned paper we can now prove the following theorem. The proof of this theorem, as well as the following theorem, depend on Lemma 1.10 of [5]. As one of the referees correctly pointed out, this lemma is incorrect as stated. Fortunately the fix is easy. In Lemma 1.9 of [5] one needs to replace 0unu0\leq u_{n}\leq u by |un|u|u_{n}|\leq u.Then the proof of Lemma 1.9 of [5] is correct. Note that in the final step the Eberlein-Smulyan theorem is used. With this correction, Lemma 1.10 of [5] is true if we replace 0unB0\leq u_{n}\in B by unBu_{n}\in B in the statement.

Theorem 2.2.

Let EE be a complex non-atomic Banach lattice and TZ(E)T\in Z(E). Then Te=T=re(T)\|T\|_{e}=\|T\|=r_{e}(T).

Proof.

Let 0<c<T0<c<\|T\|. Then (|T|cI)+>0(|T|-cI)^{+}>0. Therefore there exists 0<uE0<u\in E such that v=(|T|cI)+u>0v=(|T|-cI)^{+}u>0. Now (|T|cI)v=((|T|cI)+)2u0(|T|-cI)v=((|T|-cI)^{+})^{2}u\geq 0. Therefore |T|vcv|T|v\geq cv. By order continuity of TT this implies that |Tw|c|w||Tw|\geq c|w| for all w{v}ddw\in\{v\}^{dd}. In particular Twcw\|Tw\|\geq c\|w\| for all w{v}ddw\in\{v\}^{dd}. Now by the corrected (as indicated above) Lemma 1.10 of [5] there exist wn{v}ddw_{n}\in\{v\}^{dd} with wn=1\|w_{n}\|=1 such that wn0w_{n}\to 0 in the weak topology σ(E,E)\sigma(E,E^{*}). Let now K𝒦(E)K\in\mathcal{K}(E). Then, by compactness of KK, we have that Kwn0\|Kw_{n}\|\to 0 as nn\to\infty. From this it follows that

TK\displaystyle\|T-K\| lim supTwnKwn\displaystyle\geq\limsup\|Tw_{n}-Kw_{n}\|
lim sup|TwnKwn|=lim supTwnc.\displaystyle\geq\limsup|\|Tw_{n}\|-\|Kw_{n}\||=\limsup\|Tw_{n}\|\geq c.

As this holds for all 0<c<T0<c<\|T\|, it follows that TKT\|T-K\|\geq\|T\|. Hence TeT\|T\|_{e}\geq\|T\| and thus Te=T\|T\|_{e}=\|T\|. The result follows now from the above theorem. ∎

Remark.

In case EE^{*} is non-atomic, it was already proved in [6] that Te=T\|T\|_{e}=\|T\| for the larger class of disjointness preserving operators.

3. Fréchet cluster points and the Fréchet limit superior.

Let AA be a non-empty set and f:Af:A\to\mathbb{C} a function. Denote by \mathcal{F} the Fréchet filter on AA, i.e., a subset BB of AA is in \mathcal{F} if BcB^{c} is a finite set.

Definition 3.1.

A point z𝒞z\in\mathcal{C} is called a Fréchet cluster point of ff if for all ϵ>0\epsilon>0 and all BB\in\mathcal{F} there is an element aBa\in B such that |f(a)z|<ϵ|f(a)-z|<\epsilon.

In this case we will also say that zz is an \mathcal{F}-cluster point. Note that one can equivalently say that zz is a Fréchet cluster point of ff if and only if zz is a cluster point of f()f(\mathcal{F}) in the topological sense if and only if there exist distinct anAa_{n}\in A (n=1,2,n=1,2,\cdots) such that f(an)zf(a_{n})\to z. One can show (assuming the axiom of choice) the following proposition.

Proposition 3.2.

Let AA be a non-empty set and f:Af:A\to\mathbb{C} a function. Then zz is a Fréchet cluster point of ff if and only there exist a free ultrafilter 𝒰\mathcal{U} on AA such that z=lim𝒰fz=\lim_{\mathcal{U}}f.

It is also known that the set of all \mathcal{F}-cluster points of ff is a closed subset of 𝒞\mathcal{C}. Therefore we have the following proposition.

Proposition 3.3.

Let AA be a non-empty set and f:Af:A\to\mathbb{C} a bounded function. Then there exists an \mathcal{F}-cluster point of ff with largest and smallest modulus.

We now describe the modulus this largest and smallest \mathcal{F}-cluster point.

Definition 3.4.

Let AA be a non-empty set and f:Af:A\to\mathbb{R} a bounded function. Then the \mathcal{F}-limit superior of ff is defined as

lim supf=limFsupf[F]=infFsupf[F]\limsup_{\mathcal{F}}f=\lim_{F\in\mathcal{F}}\sup f[F]=\inf_{F\in\mathcal{F}}\sup f[F]

and the \mathcal{F}-limit inferior of ff is defined as

lim inff=limFinff[F]=supFinff[F].\liminf_{\mathcal{F}}f=\lim_{F\in\mathcal{F}}\inf f[F]=\sup_{F\in\mathcal{F}}\inf f[F].

Note that these notions can be extended to unbounded function by allowing ++\infty or -\infty as values. As in the sequential case one can prove now the following theorem.

Theorem 3.5.

Let AA be a non-empty set and f:Af:A\to\mathbb{R} a bounded function. Then the \mathcal{F}-limit superior of ff is the largest \mathcal{F}-cluster point of ff and lim inff\liminf_{\mathcal{F}}f is the smallest \mathcal{F}-cluster point of ff.

Corollary 3.6.

Let AA be a non-empty set and f:Af:A\to\mathbb{C} a bounded function with λ\lambda an \mathcal{F}-cluster point of ff of largest modulus. Then

|λ|=lim sup|f|.|\lambda|=\limsup_{\mathcal{F}}|f|.
Proof.

The inequality ||z1||z2|||z1z2|||z_{1}|-|z_{2}||\leq|z_{1}-z_{2}| implies immediately that if zz is an \mathcal{F}-cluster point of ff, then |z||z| is an \mathcal{F}-cluster point of |f||f|. Conversely, if xx is an \mathcal{F}-cluster point of |f||f|, then there exist distinct ana_{n} in AA such that |f(an)|x|f(a_{n})|\to x. By passing to a subsequence we can assume that there exists zz such that f(ai)zf(a_{i})\to z. Then |z|=x|z|=x. This shows that xx\in\mathbb{R} is an \mathcal{F}-cluster point of ff if and only there exist a an \mathcal{F}-cluster point zz of ff with |z|=x|z|=x. From this we the corollary follows immediately.

We indicate in the next section a case, where lim sup|f|\limsup_{\mathcal{F}}|f| naturally occurs as the quotient norm of (A)\ell_{\infty}(A) by c0(A)c_{0}(A).

4. The atomic case.

Let EE be an infinite dimensional (complex) Banach lattice throughout this section. Recall that aEa\in E is called an atom, if the band generated by aa is one dimensional, i.e., {a}dd={λa:λ}\{a\}^{dd}=\{\lambda a:\lambda\in\mathbb{C}\}. Denote by AA the set of all positive atoms in EE of norm one. Then EE is called an atomic Banach lattice if the band AddA^{dd} generated by AA equals EE. If we denote by PaP_{a} the band projection from EE onto {a}dd\{a\}^{dd}, then every xEx\in E can be written as an order convergent series

x=aAPax.x=\sum_{a\in A}P_{a}x.

Note that if AA is countable, then we have the usual Banach sequence spaces. Now every TZ(E)T\in Z(E) can be represented as a multiplication operator, where for each aAa\in A there exists λa\lambda_{a}\in\mathbb{C} such that Ta=λaaTa=\lambda_{a}a. Then, if x=aAPaxx=\sum_{a\in A}P_{a}x, we have Tx=aAλaPaxTx=\sum_{a\in A}\lambda_{a}P_{a}x. The following lemma is well-known in the countable case. We include for convenience of the reader the similar proof for the uncountable case.

Lemma 4.1.

Let EE be an atomic Banach lattice with, as above, the set AA the set of all positive atoms of norm one. Let TZ(E)T\in Z(E) as above. Then T𝒦(E)T\in\mathcal{K}(E) if and only if

lim|λa|=lim sup|λa|=0.\lim_{\mathcal{F}}|\lambda_{a}|=\limsup_{\mathcal{F}}|\lambda_{a}|=0.
Proof.

Assume first that T𝒦(E)T\in\mathcal{K}(E). Then there exists ϵ>0\epsilon>0 such that lim sup|λa|>ϵ>0\limsup_{\mathcal{F}}|\lambda_{a}|>\epsilon>0. Then there exist distinct anAa_{n}\in A, for n=1,2,n=1,2,\cdots, such that |λan|ϵ|\lambda_{a_{n}}|\geq\epsilon for all nn. As TT is a compact operator we can assume, by passing to a subsequence, that Tan=λananxETa_{n}=\lambda_{a_{n}}a_{n}\to x\in E. As |λan|T|\lambda_{a_{n}}|\leq\|T\| we can, by passing to a further subsequence, assume that λanλ0\lambda_{a_{n}}\to\lambda_{0}, where λ00\lambda_{0}\neq 0. Now λ0an=(λ0λan)an+λananx\lambda_{0}a_{n}=(\lambda_{0}-\lambda_{a_{n}})a_{n}+\lambda_{a_{n}}a_{n}\to x, which implies that an1λ0xa_{n}\to\frac{1}{\lambda_{0}}x. This contradicts that anam=an+aman=1\|a_{n}-a_{m}\|=\|a_{n}+a_{m}\|\geq\|a_{n}\|=1 for all nmn\neq m and thus lim sup|λa|=0\limsup_{\mathcal{F}}|\lambda_{a}|=0. This implies that lim inf|λa|=lim sup|λa|=0\liminf_{\mathcal{F}}|\lambda_{a}|=\limsup_{\mathcal{F}}|\lambda_{a}|=0 and thus lim|λa|=0\lim_{\mathcal{F}}|\lambda_{a}|=0. Conversely if lim sup|λa|=0\limsup_{\mathcal{F}}|\lambda_{a}|=0. Then for all ϵ>0\epsilon>0 we have |λa|ϵ|\lambda_{a}|\geq\epsilon for at most finitely many aAa\in A. This implies that {aA:λa0}\{a\in A:\lambda_{a}\neq 0\} is countable. Let {an:n=1,2,}={aA:λa0}\{a_{n}:n=1,2,\cdots\}=\{a\in A:\lambda_{a}\neq 0\} (with the obvious modification if the set on the right is finite). We have then that limnλan=0\lim_{n\to\infty}\lambda_{a_{n}}=0. Now the estimate

Tn=1NλanPansupnN+1|λan|\|T-\sum_{n=1}^{N}\lambda_{a_{n}}P_{a_{n}}\|\leq\sup_{n\geq N+1}|\lambda_{a_{n}}|

implies that TT is a norm limit of finite rank operators and thus compact. ∎

We now show that the expression lim sup|λa|\limsup_{\mathcal{F}}|\lambda_{a}| is actually a quotient norm. Let (A)\ell_{\infty}(A) denote the Banach lattice of all bounded functions f:Af:A\to\mathbb{C} with the supremum norm and c0(A)c_{0}(A) the closed ideal of all f(A)f\in\ell_{\infty}(A) with lim sup|λa|=0\limsup_{\mathcal{F}}|\lambda_{a}|=0. Then

Theorem 4.2.

Let (A)\ell_{\infty}(A) and c0(A)c_{0}(A) be as above. Then we have f+c0(A)=lim sup|f|\|f+c_{0}(A)\|=\limsup_{\mathcal{F}}|f| for all f(A)f\in\ell_{\infty}(A).

Proof.

Note first that

f+c0(A)=inf{supa|f(a)g(a)|:limg(a)=0}.\|f+c_{0}(A)\|=\inf\{\sup_{a}|f(a)-g(a)|:\lim_{\mathcal{F}}g(a)=0\}.

Let ϵ>0\epsilon>0. Then f+c0(A)<f+c0(A)+ϵ\|f+c_{0}(A)\|<\|f+c_{0}(A)\|+\epsilon implies that there exists gc0(A)g\in c_{0}(A) such that supa|f(a)g(a)|<f+c0(A)+ϵ\sup_{a}|f(a)-g(a)|<\|f+c_{0}(A)\|+\epsilon. Now gc0(A)g\in c_{0}(A) implies that there is a finite subset BB of AA such that |g(a)|<ϵ|g(a)|<\epsilon for all aABa\in A\setminus B. This implies that supa|f(a)|<f+c0(A)+2ϵ\sup_{a}|f(a)|<\|f+c_{0}(A)\|+2\epsilon for all aABa\in A\setminus B. Hence

lim sup|f|f+c0(A).\limsup_{\mathcal{F}}|f|\leq\|f+c_{0}(A)\|.

Similarly for ϵ>0\epsilon>0 there exists a finite subset BB of AA such that |f(a)|<lim sup|f|+ϵ|f(a)|<\limsup_{\mathcal{F}}|f|+\epsilon for all aABa\in A\setminus B. Now define gc0(A)g\in c_{0}(A) by g(a)=f(a)g(a)=f(a) for aBa\in B and g(a)=0g(a)=0 otherwise. Then supa|f(a)g(a)|<lim sup|f|+ϵ\sup_{a}|f(a)-g(a)|<\limsup_{\mathcal{F}}|f|+\epsilon. Hence

f+c0(A)lim sup|f|,\|f+c_{0}(A)\|\leq\limsup_{\mathcal{F}}|f|,

which completes the proof. ∎

Corollary 4.3.

Let EE be an atomic Banach lattice with, as above, the set AA the set of all positive atoms of norm one. Let TZ(E)T\in Z(E). Then

inf{TK:KZ(E)𝒦(E)}=lim sup|λa|.\inf\{\|T-K\|:K\in Z(E)\cap\mathcal{K}(E)\}=\limsup_{\mathcal{F}}|\lambda_{a}|.
Theorem 4.4.

Let EE be an atomic Banach lattice with, as above, the set AA the set of all positive atoms of norm one. Let TZ(E)T\in Z(E). Then the essential spectrum σe(T)\sigma_{e}(T) is given by

σe(T)={λ:λ is a Fréchet cluster point of the function aλa}.\sigma_{e}(T)=\{\lambda\in\mathbb{C}:\lambda\mbox{ is a Fr\'{e}chet cluster point of the function }a\mapsto\lambda_{a}\}.

Moreover, the essential spectral radius re(T)r_{e}(T) of TT is given by

re(T)=lim sup|λa|,r_{e}(T)=\limsup_{\mathcal{F}}|\lambda_{a}|,

and the essential norm Te\|T\|_{e} of TT satisfies Te=re(T)\|T\|_{e}=r_{e}(T).

Proof.

We note that it is sufficient to prove 0σe(T)0\in\sigma_{e}(T) if and only if 0 is a Fréchet cluster point of the function aλaa\mapsto\lambda_{a}, as TZ(E)T\in Z(E) if and only TλIZ(E)T-\lambda I\in Z(E). Assume first that 0σe(T)0\in\sigma_{e}(T) and that 0 is not a Fréchet cluster point of the function aλaa\mapsto\lambda_{a}. Then by Theorem 3.5 we have lim inf|λa|>0\liminf_{\mathcal{F}}|\lambda_{a}|>0. Let 0<ϵ<lim inf|λa|0<\epsilon<\liminf_{\mathcal{F}}|\lambda_{a}|. Then there exists a finite subset BB of AA such that |λa|ϵ|\lambda_{a}|\geq\epsilon for all aABa\in A\setminus B. Now aBλaPa\sum_{a\in B}\lambda_{a}P_{a} is a finite rank operator, so that σe(T)=σe(TaBλaPa)\sigma_{e}(T)=\sigma_{e}(T-\sum_{a\in B}\lambda_{a}P_{a}). Now |λa|ϵ|\lambda_{a}|\geq\epsilon for all aABa\in A\setminus B implies that |TaBλaPa|ϵaABPa|T-\sum_{a\in B}\lambda_{a}P_{a}|\geq\epsilon\sum_{a\in A\setminus B}P_{a}. As I=aBPa+aABPaI=\sum_{a\in B}P_{a}+\sum_{a\in A\setminus B}P_{a} and as aBPa\sum_{a\in B}P_{a} is a band projection, it follows that also aABPa\sum_{a\in A\setminus B}P_{a} is a band projection. Combined with the lower estimate we obtained above, this implies that TaABλaPaT-\sum_{a\in A\setminus B}\lambda_{a}P_{a} is invertible on aABPa(E)\sum_{a\in A\setminus B}P_{a}(E). This implies that TaABλaPaT-\sum_{a\in A\setminus B}\lambda_{a}P_{a} is a Fredholm operator on EE and thus 0σe(TaBλaPa)=σe(T)0\notin\sigma_{e}(T-\sum_{a\in B}\lambda_{a}P_{a})=\sigma_{e}(T), which contradicts our assumption. Now assume that 0 is a Fréchet cluster point of the function aλaa\mapsto\lambda_{a}. Then there exist infinitely many distinct ana_{n} with n=1,2,n=1,2,\cdots such that λan0\lambda_{a_{n}}\to 0. This implies that 0σ(T)0\in\sigma(T) and that 0 is not isolated in σ(T)\sigma(T). It is well-known that this implies that 0σe(E)0\in\sigma_{e}(E), which completes the proof of the first part of the theorem. The formula for the essential spectral radius follows now immediately from Corollary 3.6. The formula for the essential norm follows from the above and the previous corollary as follows

Te=inf{TK:K𝒦}inf{TK:TZ(E)𝒦(E)}=lim sup|λa|=re(T).\|T\|_{e}=\inf\{\|T-K\|:K\in\mathcal{K}\}\leq\inf\{\|T-K\|:T\in Z(E)\cap\mathcal{K}(E)\}=\limsup_{\mathcal{F}}|\lambda_{a}|=r_{e}(T).

5. The general case.

Let EE be a (complex) Banach lattice. Denote by AA the set of all positive atoms in EE of norm one. Then the atomic part of EE is the band AddA^{dd} generated by AA and will be denoted by EAE_{A}. The disjoint complement AdA^{d} of AA will be called the non-atomic part (or continuous part) of EE and will be denoted by EAdE_{A^{d}}. Below we shall see by means of an example that these bands in general don’t need to be projection bands (but are so of course when EE is Dedekind complete). Therefore we have in general that EAEAdE_{A}\oplus E_{A^{d}} is only a norm closed, order dense ideal in EE. We denote the restriction of TZ(E)T\in Z(E) to EAE_{A} by TAT_{A} (the atomic part of TT) and to EAdE_{A^{d}} by TAdT_{A^{d}} (the non-atomic or continuous part of TT). As TT is band preserving we have TAZ(EA)T_{A}\in Z(E_{A}) and TAdZ(EAd)T_{A^{d}}\in Z(E_{A^{d}}), but in general TT is not the direct sum of these two operators, as TATAdT_{A}\oplus T_{A^{d}} is only defined on EAEAdE_{A}\oplus E_{A^{d}}. To overcome this technical difficulty, we use the following lemma.

Proposition 5.1.

Let EE be a Banach lattice and FEF\subset E a norm closed, order dense ideal in EE. Let TZ(E)T\in Z(E). Then TF=T\|T_{F}\|=\|T\|.

Proof.

As TFT_{F} is a restriction of TT we always have TFT\|T_{F}\|\leq\|T\|. To prove the reverse inequality, observe that

TF=inf{λ:|TF|λIF}.\|T_{F}\|=\inf\{\lambda:|T_{F}|\leq\lambda I_{F}\}.

Now |TF|λIF|T_{F}|\leq\lambda I_{F} implies by order continuity of TT that also |T|λI|T|\leq\lambda I. This implies that TTF\|T\|\leq\|T_{F}\| and equality of norms follows. ∎

Theorem 5.2.

Let EE be a (complex) Banach lattice and TZ(E)T\in Z(E). Then

σe(T)=σe(TA)σ(TAd)\sigma_{e}(T)=\sigma_{e}(T_{A})\cup\sigma(T_{A^{d}})

and thus

re(T)=max{re(TA),TAd}.r_{e}(T)=\max\{r_{e}(T_{A}),\|T_{A^{d}}\|\}.

Moreover Te=re(T)\|T\|_{e}=r_{e}(T).

Proof.

It is easy to see that σe(TA)σe(T)\sigma_{e}(T_{A})\subset\sigma_{e}(T) and σe(TAd)σe(T)\sigma_{e}(T_{A^{d}})\subset\sigma_{e}(T), so σe(TA)σe(TAd)σe(T)\sigma_{e}(T_{A})\cup\sigma_{e}(T_{A^{d}})\subset\sigma_{e}(T). Therefore assume that λσe(TA)σe(TAd)\lambda\notin\sigma_{e}(T_{A})\cup\sigma_{e}(T_{A^{d}}). Then TAdλIAdT_{A^{d}}-\lambda I_{A^{d}} is Fredholm on EAdE_{A^{d}}. As σ(TAdλIAd)=σe(TAdλIAd)\sigma(T_{A^{d}}-\lambda I_{A^{d}})=\sigma_{e}(T_{A^{d}}-\lambda I_{A^{d}}) by the main result in the non-atomic case, we see that there exists c1>0c_{1}>0 such that |TAdλIAd|c1IAd|T_{A^{d}}-\lambda I_{A^{d}}|\geq c_{1}I_{A^{d}}. Now TAλIAT_{A}-\lambda I_{A} is Fredholm on EAE_{A} implies that there exist a finite subset BB of AA such that the kernel of TAλIAT_{A}-\lambda I_{A} is PB(EA)P_{B}(E_{A}). Then TABλIABT_{A\setminus B}-\lambda I_{A\setminus B} is invertible on PABP_{A\setminus B} , so there exists c2>0c_{2}>0 such that |TABλIAB|c2IAB|T_{A\setminus B}-\lambda I_{A\setminus B}|\geq c_{2}I_{A\setminus B}. This implies that for c=min{c1,c2}c=\min\{c_{1},c_{2}\} that |TAdABλIAdAB|cIAdAB|T_{A^{d}\cup A\setminus B}-\lambda I_{A^{d}\cup A\setminus B}|\geq cI_{A^{d}\cup A\setminus B}. This implies that TAdABλIAdT_{A^{d}\cup A\setminus B}-\lambda I_{A^{d}} is invertible on EAdABE_{A^{d}\cup A\setminus B}. As EBE_{B} is finite dimensional this implies that λσe(T)\lambda\notin\sigma_{e}(T), which shows that σe(T)=σe(TA)σ(TAd)\sigma_{e}(T)=\sigma_{e}(T_{A})\cup\sigma(T_{A^{d}}). The formula for the essential spectral radius follows now from the previous results. To prove the result about the essential norm, observe that re(T)Ter_{e}(T)\leq\|T\|_{e} is always true, so it remains to show that Tere(T)\|T\|_{e}\leq r_{e}(T). Let F=EAEAdF=E_{A}\oplus E_{A^{d}}. Then FF is a closed order dense ideal in EE. Therefore, using the preceding proposition, we have

Te\displaystyle\|T\|_{e} =inf{TK:K𝒦(E)}\displaystyle=\inf\{\|T-K\|:K\in\mathcal{K}(E)\}
inf{TK:K𝒦(E)Z(E)}\displaystyle\leq\inf\{\|T-K\|:K\in\mathcal{K}(E)\cap Z(E)\}
=inf{(TK)F:K𝒦(E)Z(E)}\displaystyle=\inf\{\|(T-K)_{F}\|:K\in\mathcal{K}(E)\cap Z(E)\}
inf{max{TAd,TAKA}:K𝒦(E)Z(E)}\displaystyle\leq\inf\{\max\{\|T_{A^{d}}\|,\|T_{A}-K_{A}\|\}:K\in\mathcal{K}(E)\cap Z(E)\}
=max{TAd,inf{TAKA:K𝒦(E)Z(E)}}=re(T)\displaystyle=\max\{\|T_{A^{d}}\|,\inf\{\|T_{A}-K_{A}\|:K\in\mathcal{K}(E)\cap Z(E)\}\}=r_{e}(T)

by the above. Hence Te=re(T)\|T\|_{e}=r_{e}(T). ∎

We conclude with an example of a Banach lattice EE for which the atomic part EAE_{A} is not a projection band.

Example 5.3.

Let xn=12n+1x_{n}=\frac{1}{2^{n+1}} and In=[12n+1+2,12n+1+1]I_{n}=[\frac{1}{2^{n+1}+2},\frac{1}{2^{n+1}+1}] for n1n\geq 1. Then it is easy to verify that xnImx_{n}\notin I_{m} for n,m1n,m\geq 1. Now put K=n1In{xn:n1}{0}K=\cup_{n\geq 1}I_{n}\cup\{x_{n}:n\geq 1\}\cup\{0\}. Then KK is compact in the induced Euclidean topology. Let E=C(K)E=C(K). Let δx\delta_{x} denote the Dirac delta function supported by xx. Then it is clear that the set AA of positive atoms of norm one equals {δxn:n1}\{\delta_{x_{n}}:n\geq 1\}. Note that δ0E\delta_{0}\notin E, as δ0\delta_{0} is not continuous at x=0x=0. Now the continuous part EAdE_{A^{d}} of EE equals Ad={fE:f(xn)=0 for all n1}A^{d}=\{f\in E:f(x_{n})=0\mbox{ for all }n\geq 1\}. Note fAdf\in A^{d} implies by continuity that f(0)=0f(0)=0. Now the atomic part EAE_{A} is equal to Add={fE:f(x)=0 on n1In}A^{dd}=\{f\in E:f(x)=0\mbox{ on }\cup_{n\geq 1}I_{n}\}. Again by continuity f(0)=0f(0)=0 for fEAf\in E_{A}. Hence EAEAdEE_{A}\oplus E_{A^{d}}\neq E, as the function identical one is in EE. Therefore EAE_{A} is not a projection band. It is well-known that for a C(K)C(K) space EE each TZ(E)T\in Z(E) is given by a multiplication by a function pC(K)p\in C(K), as Z(E)Z(E) is lattice isometric to C(K)C(K). Denote by TpT_{p} the operator Tpf=pfT_{p}f=pf.Then we have σ(Tp)={p(x):xK}\sigma(T_{p})=\{p(x):x\in K\}. Thus we have r(Tp)=T=pr(T_{p})=\|T\|=\|p\|_{\infty}. Now Tp(δxn)=p(xn)δxnT_{p}(\delta_{x_{n}})=p(x_{n})\delta_{x_{n}} for all n1n\geq 1. As 0 is the only cluster point of the sequence (p(xn))(p(x_{n})), it follows that σe(TA)={0}\sigma_{e}(T_{A})=\{0\}. The essential spectrum of the non-atomic part of TpT_{p} is

σe((Tp)Ad)={p(x):xn=1In{0}}.\sigma_{e}((T_{p})_{A^{d}})=\{p(x):x\in\bigcup_{n=1}^{\infty}I_{n}\cup\{0\}\}.

Therefore

Tpe=(Tp)Ad=max{|p(x)|:xn=1In{0}}.\|T_{p}\|_{e}=\|(T_{p})_{A^{d}}\|=\max\{|p(x)|:x\in\bigcup_{n=1}^{\infty}I_{n}\cup\{0\}\}.

We also observe that (Tp)A(T_{p})_{A} is compact if and only if we have p(xn)0p(x_{n})\to 0, i.e., if and only if p(0)=0p(0)=0. Moreover we can write Tp=Tp1+Tp2T_{p}=T_{p_{1}}+T_{p_{2}} with (Tp1)A=(Tp)A(T_{p_{1}})_{A}=(T_{p})_{A} if and only if p(0)=0p(0)=0, i.e., if and only if (Tp)A(T_{p})_{A} is compact.

As the above example shows that in general, if TZ(E)T\in Z(E), then TT is not the sum of T1T_{1} and T2T_{2}, where TA=(T1)AT_{A}=(T_{1})_{A} and TAd=(T2)AdT_{A^{d}}=(T_{2})_{A^{d}}.However that is so if only if (Tp)A(T_{p})_{A} is compact. We now show that this is always the case for the if part of the statement.

Proposition 5.4.

Let EE be a (complex) Banach lattice and TZ(E)T\in Z(E) such that TAT_{A} is compact. Then there exist T1,T2Z(E)T_{1},T_{2}\in Z(E) with T1T_{1} compact such that T=T1+T2T=T_{1}+T_{2} where TA=(T1)AT_{A}=(T_{1})_{A} and TAd=(T2)AdT_{A^{d}}=(T_{2})_{A^{d}}.

Proof.

Let TAea=λaeaT_{A}e_{a}=\lambda_{a}e_{a}. Then TAT_{A} is compact implies that lim sup|λa|=0\limsup_{\mathcal{F}}|\lambda_{a}|=0. This implies that the set B={aA:λa0}B=\{a\in A:\lambda_{a}\neq 0\} is countable and TAT_{A} is a norm limit of the series aBλaPa\sum_{a\in B}\lambda_{a}P_{a} in Z(EA)Z(E_{A}). As this series also converges in Z(E)Z(E), it follows that TAT_{A} extends to a compact operator T1T_{1} in Z(E)Z(E) such that TA=(T1)AT_{A}=(T_{1})_{A}. Let T2=TT1T_{2}=T-T_{1}. Then it clear from (T1)Ad=0(T_{1})_{A^{d}}=0 that TAd=(T2)AdT_{A^{d}}=(T_{2})_{A^{d}}. ∎

In [3] it was observed that if σe(A+B)=σe(A)\sigma_{e}(A+B)=\sigma_{e}(A) for all bounded self-adjoint operators AA on a Hilbert space, for a given self-adjoint BB, then BB is compact. The following theorem is an order analogue of this.

Theorem 5.5.

Let EE be a (complex) Banach lattice and TZ(E)T\in Z(E) such that σe(I+T)=σe(I)\sigma_{e}(I+T)=\sigma_{e}(I). Then TT is compact.

Proof.

From σe(I+T)=σe(I)\sigma_{e}(I+T)=\sigma_{e}(I) we conclude that σe(T)={0}\sigma_{e}(T)=\{0\}. From Theorem 5.2 we see that σe(TA)={0}\sigma_{e}(T_{A})=\{0\} and σe(TAd)={0}\sigma_{e}(T_{A^{d}})=\{0\}. From Theorem 4.4 we see that ra(TA)=lim supλa=0r_{a}(T_{A})=\limsup_{\mathcal{F}}\lambda_{a}=0. This implies that TAT_{A} is compact on EAE_{A}. By the above proposition there exist T1,T2Z(E)T_{1},T_{2}\in Z(E) with T1T_{1} compact such that T=T1+T2T=T_{1}+T_{2} where TA=(T1)AT_{A}=(T_{1})_{A} and TAd=(T2)AdT_{A^{d}}=(T_{2})_{A^{d}}. Now σ(TAd)=σe(TAd)={0}\sigma(T_{A^{d}})=\sigma_{e}(T_{A^{d}})=\{0\} by Theorem 2.1 and thus r(TAd)=||TAd=0r(T_{A^{d}})=||T_{A^{d}}\|=0. This implies that (T2)AdAdd=0(T_{2})_{A^{d}\oplus A^{dd}}=0, which by order density of AdAddA^{d}\oplus A^{dd} and order continuity of T2T_{2} implies that T2=0T_{2}=0. Hence T=T1T=T_{1} is compact. ∎

Remark.

We could have phrased the above result alternatively as: If TZ(E)T\in Z(E) is essentially quasi-nilpotent, then TT is compact.

References

  • [1] W. Arendt and A. R. Sourour, Perturbation of regular operators and the order essential spectrum, Nederl. Akad. Wetensch. Indag. Math. 48 (1986), no. 2, 109–122.
  • [2] René E. Castillo, Yesid A. Lemus-Abril, and Julio C. Ramos-Fernández, Essential norm estimates for multiplication operators on Lp(μ)L_{p}(\mu) spaces, Afr. Mat. 32 (2021), no. 7-8, 1595–1603.
  • [3] K. Gustafson and J. Weidmann, On the essential spectrum, J. Math. Anal. Appl. 25 (1969), 121–127.
  • [4] P. Meyer-Nieberg, Banach lattices, Springer-Verlag, 1991.
  • [5] A. R. Schep, Positive diagonal and triangular operators, J. Operator Theory 3 (1980), no. 2, 165–178.
  • [6] A. R. Schep, The measure of noncompactness of a disjointness preserving operator, J. Operator Theory 21 (1989), no. 2, 397–402.
  • [7] Jürgen Voigt, The essential norm of multiplication operators on Lp(μ){L_{p}(\mu)}, arXiv:2203.13096 [math.FA], 2022.
  • [8] Anthony W. Wickstead, Extremal structure of cones of operators, Quart. J. Math. Oxford Ser. (2) 32 (1981), no. 126, 239–253.
  • [9] A. C. Zaanen, Examples of orthomorphisms, J. Approximation Theory 13 (1975), 192–204.
  • [10] by same author, Riesz spaces. II, North-Holland Mathematical Library, vol. 30, North-Holland Publishing Co., Amsterdam, 1983.