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

Multiplication Operators on the Weighted Lipschitz Space of a Tree

Robert F. Allen1, Flavia Colonna2, and Glenn R. Easley3 1Department of Mathematics, University of Wisconsin-La Crosse 2Department of Mathematical Sciences, George Mason University 3System Planning Corporation [email protected], [email protected], [email protected]
Abstract.

We study the multiplication operators on the weighted Lipschitz space w\mathcal{L}_{\textbf{w}} consisting of the complex-valued functions ff on the set of vertices of an infinite tree TT rooted at oo such that supvo|v||f(v)f(v)|<\sup_{v\neq o}|v||f(v)-f(v^{-})|<\infty, where |v||v| denotes the distance between oo and vv and vv^{-} is the neighbor of vv closest to oo. For the multiplication operator, we characterize boundedness, compactness, provide estimates on the operator norm and the essential norm, and determine the spectrum. We prove that there are no isometric multiplication operators or isometric zero divisors on w\mathcal{L}_{\textbf{w}}.

Key words and phrases:
Multiplication operators, Lipschitz space, Trees.
2010 Mathematics Subject Classification:
primary: 47B38; secondary: 05C05.

1. Introduction

Let XX be a Banach space of complex-valued functions on a set Ω\Omega. For a complex-valued function ψ\psi with domain Ω\Omega, we define the multiplication operator with symbol ψ\psi on XX to be Mψf=ψfM_{\psi}f=\psi f for fX.f\in X. The study of such operators with symbol attempts to tie the properties of the operator with the function theoretic properties of the symbol. The operator properties typically considered are boundedness, compactness, and being an isometry. Other aspects of interest are the determination of estimates on the operator norm as well as on the essential norm, and the identification of the spectrum and the essential spectrum.

A setting that has been widely considered in the literature is when Ω\Omega is the open unit disk 𝔻\mathbb{D} and XX is a Banach space of analytic functions on 𝔻\mathbb{D}. Examples of such Banach spaces are the Hardy space HpH^{p}, the Bergman space ApA^{p}, and the Bloch space \mathcal{B} (see [19] for more information on the operator theory on these spaces.)

In recent years, researchers have been developing versions of these spaces where the set Ω\Omega is a discrete space such as a tree or a discrete group. Historically, the function theory on trees has been largely devoted to studying the eigenfunctions of the Laplace operator (and in particular, the harmonic functions), defined as the averaging operator (with respect to a nearest-neighbor transition probability) at the neighbors of a vertex, minus the identity operator.

The study of the harmonic functions on discrete structures can be traced back many years in the literature. It was the harmonic analysis on trees developed by Cartier in [2] that made evident the analogy between trees endowed with the edge-counting metric and the open unit disk in the complex plane under the Poincaré metric.

The Hardy spaces HpH^{p} on trees have been studied by Korányi, Picardello, and Taibleson in [10], and the theory of the HpH^{p} spaces was further developed in [8] by Di Biase and Picardello in the special case when the tree is homogeneous (that is, the vertices have the same number of neighbors).

Operators on discrete structures other than the Laplacian have been studied in a number of papers (e.g., see the works of Pavone [11], [12], [13], Roe [17], and Rabinovich and Roch [14], [15], and [16]). Examples include the composition operators on LpL^{p} spaces associated with homogeneous trees, the Toeplitz operators on discrete groups, and the band-dominated operators defined on p(X)\ell^{p}(X), where XX is a discrete metric space. The band-dominated operators on p(X)\ell^{p}(X) are compositions of shift operators on XX with multiplication operators with symbols in (X)\ell^{\infty}(X) and have a natural connection to Schrödinger operators when XX is a graph.

In [3], Cohen and the second author defined the Bloch space on an isotropic homogeneous tree TT by considering the harmonic functions which are Lipschitz when regarded as function between metric spaces, where the distance on TT counts the edges between pairs of vertices and \mathbb{C} is endowed with the Euclidean distance. However, in [4], where embeddings of homogeneous trees of even degree in the hyperbolic disk were constructed so that the edges are geodesic arcs of the same hyperbolic length, it was shown that the harmonicity condition on a tree from a nearest-neighbor perspective is not related to the classical harmonicity (and hence analyticity) condition on the disk derived through interpolation.

This suggests that for the purpose of the study of certain operators with symbol such as the multiplication, or more generally, the weighted composition operators, these spaces are not natural analogues of their continuous counterparts. In particular, the study of multiplication operators on such spaces is of no interest, since in order for a multiplication operator to preserve harmonicity on a tree, its symbol must be a constant function. So, for the study of the theory of such operators with symbol, the spaces of functions on trees need to be less restrictive.

In [5] (see also [19]) it was shown that the analytic functions f:𝔻f:\mathbb{D}\to\mathbb{C} such that

βf=supz𝔻(1|z|2)|f(z)|<\beta_{f}=\displaystyle\sup_{z\in\mathbb{D}}(1-|z|^{2})|f^{\prime}(z)|<\infty

are precisely the Lipschitz functions with respect to the Poincaré distance ρ\rho on 𝔻\mathbb{D} and the Euclidean distance on \mathbb{C} and βf\beta_{f} is the Lipschitz constant of ff, namely

βf=supzw|f(z)f(w)|ρ(z,w).\beta_{f}=\sup_{z\neq w}\frac{|f(z)-f(w)|}{\rho(z,w)}.

The collection of such functions is called the Bloch space.

In [6], the last two authors defined the Lipschitz space \mathcal{L} on an infinite tree TT rooted at vertex oo to be the collection of all complex-valued functions on the vertices of the tree that are Lipschitz with respect to the edge-counting metric on TT and the Euclidean metric on \mathbb{C}. They showed these are precisely the functions ff for which

supvT|f(v)f(v)|<,\sup_{v\in T^{*}}|f(v)-f(v^{-})|<\infty,

where T=T{o}T^{*}=T\setminus\{o\}. It was shown that \mathcal{L} is a functional Banach space under the norm

f=|f(o)|+supvT|f(v)f(v)|,\|f\|_{\mathcal{L}}=|f(o)|+\sup_{v\in T^{*}}|f(v)-f(v^{-})|,

and the multiplication operator was studied in detail on \mathcal{L} as well as on a closed separable subspace called the little Lipschitz space. The space \mathcal{L} can be viewed as a discrete analogue of the space \mathcal{B}.

In this work, we carry out the study of the multiplication operators on the space w\mathcal{L}_{\textbf{w}} of the complex-valued functions ff on an infinite tree TT rooted at oo satisfying the condition

supvT|v||f(v)f(v)|<,\sup_{v\in T^{*}}|v||f(v)-f(v^{-})|<\infty,

where |v||v| is the number of edges in the unique path from oo to vv and vv^{-} is the neighbor of vv closest to oo. The interest in studying this space is due to the fact that the bounded functions in w\mathcal{L}_{\textbf{w}} are the symbols of the bounded multiplication operators on \mathcal{L} [6]. The space w\mathcal{L}_{\textbf{w}} (where the subscript w stands for weight) can be regarded as a discrete analogue of the weighted Bloch space \mathcal{B}_{\ell} defined as the set of analytic functions ff on 𝔻\mathbb{D} such that

supz𝔻(1|z|2)log21|z|2|f(z)|<,\sup_{z\in\mathbb{D}}(1-|z|^{2})\log\frac{2}{1-|z|^{2}}|f^{\prime}(z)|<\infty,

since the logarithmic weight is closely related to the Poincaré distance

ρ(0,z)=12log1+|z|1|z|.\rho(0,z)=\frac{1}{2}\log\frac{1+|z|}{1-|z|}.

The multiplication operators and cyclic vectors on the weighted Bloch space were studied by Ye in [18]. In this work, we prove the discrete counterparts of several results in [18] and expand the scope of the analysis of such operators.

1.1. Organization of the Paper

After giving some preliminary definitions and notation on trees, in Section 2, we show that w\mathcal{L}_{\textbf{w}} is a Banach space under the norm

fw=|f(o)|+supvT|v||f(v)f(v)|\left\|f\right\|_{\textbf{w}}=|f(o)|+\displaystyle\sup_{v\in T^{*}}|v||f(v)-f(v^{-})|

and define a particular closed subspace w,0\mathcal{L}_{\textbf{w},0} we call the little weighted Lipschitz space. We also give some useful properties that will be needed in the following sections. In Section 3, we define the notion of a cyclic vector for w,0\mathcal{L}_{\textbf{w},0} and determine a class of cyclic vectors.

In Section 4, we characterize the bounded multiplication operators MψM_{\psi} on w\mathcal{L}_{\textbf{w}} and w,0\mathcal{L}_{\textbf{w},0} in terms of the symbol ψ\psi and establish estimates on the operator norm in Section 5. In Section 6, we determine the spectrum, the point spectrum and the approximate spectrum of MψM_{\psi}. We also show that MψM_{\psi} is bounded below if and only if the modulus of ψ\psi is bounded away from 0.

In Section 7, we characterize the compact multiplication operators on w\mathcal{L}_{\textbf{w}} and w,0\mathcal{L}_{\textbf{w},0} in terms of a little-oh condition corresponding to the big-oh condition for boundedness. In Section 8, we determine estimates on the essential norm of MψM_{\psi}.

In Section 9, we characterize the isometric multiplication operators on w\mathcal{L}_{\textbf{w}} and w,0\mathcal{L}_{\textbf{w},0} and show that there are no isometric zero divisors on these spaces.

1.2. Preliminary Definitions and Notation

By a tree TT we mean a locally finite, connected, and simply-connected graph, which, as a set, we identify with the collection of its vertices. By a function on a tree we mean a complex-valued function on the set of its vertices. Two vertices vv and ww are called neighbors if there is an edge [v,w][v,w] connecting them, and we use the notation vwv\sim w. A vertex is called terminal if it has a unique neighbor. A path is a finite or infinite sequence of vertices [v0,v1,][v_{0},v_{1},\dots] such that vkvk+1v_{k}\sim v_{k+1} and vk1vk+1v_{k-1}\neq v_{k+1}, for all kk. Given a tree TT rooted at oo and a vertex vTv\in T, a vertex ww is called a descendant of vv if vv lies in the unique path from oo to ww. The vertex vv is then called an ancestor of ww. The vertex vv is called a child of vv^{-}.

For vTv\in T, the set SvS_{v} consisting of vv and all its descendants is called the sector determined by vv. Define the length of a finite path [v=v0,v1,,w=vn][v=v_{0},v_{1},\dots,w=v_{n}] (with vkvk+1v_{k}\sim v_{k+1} for k=0,,n1k=0,\dots,n-1) to be the number nn of edges connecting vv to ww. The distance, d(v,w)d(v,w), between vertices vv and ww is the length of the unique path connecting vv to ww. Fixing oo as the root of the tree, we define the length of a vertex vv, by |v|=d(o,v)|v|=d(o,v).

In this paper, we shall assume the tree TT to be without terminal vertices (and hence infinite), and rooted at a vertex oo and shall denote by LL^{\infty} the space of the bounded functions ff on the tree equipped with the supremum norm

f=supvT|f(v)|.\|f\|_{\infty}=\sup\limits_{v\in T}|f(v)|.

2. The Weighted Lipschitz Space

Let TT be a tree and let w\mathcal{L}_{\textbf{w}} denote the set of functions ff on TT such that supvT|v|Df(v)<,\sup\limits_{v\in T^{*}}|v|Df(v)<\infty, where Df(v)=|f(v)f(v)|Df(v)=|f(v)-f(v^{-})| for vTv\in T^{*}. For fwf\in\mathcal{L}_{\textbf{w}}, define

fw=|f(o)|+supvT|v|Df(v).\left\|f\right\|_{\textbf{w}}=|f(o)|+\sup_{v\in T^{*}}|v|Df(v).
Proposition 2.1.

If fwf\in\mathcal{L}_{\textbf{w}} and vTv\in T^{*}, then

|f(v)|(1+log|v|)fw.\displaystyle|f(v)|\leq(1+\log|v|)\left\|f\right\|_{\textbf{w}}. (2.1)

For the proof we need the following result.

Lemma 2.2.

For x>1x>1, we have

1xlog(xx1)1x1.\frac{1}{x}\leq\log\left(\frac{x}{x-1}\right)\leq\frac{1}{x-1}.
Proof.

The upper estimate is an immediate consequence of the inequality

log(1+1x1)1x1.\log\left(1+\frac{1}{x-1}\right)\leq\frac{1}{x-1}.

The lower estimate follows from the fact that the function φ(x)=xlog(xx1)\varphi(x)=x\log\left(\displaystyle\frac{x}{x-1}\right) is decreasing and limxφ(x)=1\displaystyle\lim\limits_{x\to\infty}\varphi(x)=1.∎

Proof of Proposition 2.1.

Let us first assume f(o)=0f(o)=0 and argue by induction on |v||v|. For |v|=1|v|=1, we have

|f(v)|=|v|Df(v)fw=(1+log|v|)fw.|f(v)|=|v|Df(v)\leq\left\|f\right\|_{\textbf{w}}=(1+\log|v|)\left\|f\right\|_{\textbf{w}}.

Let nn\in\mathbb{N} and assume |f(w)|(1+log|w|)fw|f(w)|\leq(1+\log|w|)\left\|f\right\|_{\textbf{w}} whenever ww is a vertex such that 1|w|<n1\leq|w|<n. Let vv be a vertex of length nn. Then, by Lemma 2.2 we get

|f(v)|\displaystyle|f(v)| \displaystyle\leq |f(v)|+|f(v)f(v)|(1+log(|v|1)+1|v|)fw\displaystyle|f(v^{-})|+|f(v)-f(v^{-})|\leq\left(1+\log(|v|-1)+\frac{1}{|v|}\right)\left\|f\right\|_{\textbf{w}}
\displaystyle\leq (1+log|v|)fw.\displaystyle(1+\log|v|)\left\|f\right\|_{\textbf{w}}.

On the other hand, if f(o)0f(o)\neq 0, let g(v)=f(v)f(o)g(v)=f(v)-f(o) for vTv\in T. By the previous case, we have |g(v)|(1+log|v|)gw|g(v)|\leq(1+\log|v|)\left\|g\right\|_{\textbf{w}} for vTv\in T^{*}. Since fw=|f(o)|+gw\left\|f\right\|_{\textbf{w}}=|f(o)|+\left\|g\right\|_{\textbf{w}}, we deduce that

|f(v)||f(o)|+|g(v)||f(o)|+(1+log|v|)gw(1+log|v|)fw,|f(v)|\leq|f(o)|+|g(v)|\leq|f(o)|+(1+\log|v|)\left\|g\right\|_{\textbf{w}}\leq(1+\log|v|)\left\|f\right\|_{\textbf{w}},

completing the proof. ∎

Theorem 2.3.

w\rm{\mathcal{L}_{\textbf{w}}} is a Banach space under the norm w\left\|\cdot\right\|_{\textbf{w}}.

Proof.

It is immediate to see that w\mathcal{L}_{\textbf{w}} is a vector space and that ffwf\mapsto\left\|f\right\|_{\textbf{w}} is a semi-norm. It is also evident that the norm of the function identically 0 is 0. Conversely, assume fw=0\left\|f\right\|_{\textbf{w}}=0. Then DfDf is identically 0. Thus, ff is a constant and since f(o)=0f(o)=0, ff is identically 0.

To prove that w\mathcal{L}_{\textbf{w}} is a Banach space, let {fn}\{f_{n}\} be Cauchy in w\mathcal{L}_{\textbf{w}}. For n,mn,m\in\mathbb{N}, since |fn(o)fm(o)|fnfmw|f_{n}(o)-f_{m}(o)|\leq\left\|f_{n}-f_{m}\right\|_{\textbf{w}}, and by Proposition 2.1, for vTv\in T^{*},

|fn(v)fm(v)|(1+log|v|)fnfmw,\displaystyle|f_{n}(v)-f_{m}(v)|\leq(1+\log|v|)\left\|f_{n}-f_{m}\right\|_{\textbf{w}},

the sequence {fn(v)}\{f_{n}(v)\} is Cauchy for each vTv\in T. Hence it converges pointwise to some function ff. We now show that fwf\in\mathcal{L}_{\textbf{w}}.

Let vTv\in T^{*} and fix nn\in\mathbb{N}. Then

|v|Df(v)|v||f(v)fn(v)|+|v|Dfn(v)+|v||fn(v)f(v)|.\displaystyle|v|Df(v)\leq|v||f(v)-f_{n}(v)|+|v|Df_{n}(v)+|v||f_{n}(v^{-})-f(v^{-})|. (2.2)

Since for each vTv\in T^{*}, |v|Dfn(v)fnw|v|Df_{n}(v)\leq\left\|f_{n}\right\|_{\textbf{w}} and {fn}\{f_{n}\} is Cauchy in w\mathcal{L}_{\textbf{w}}, and hence bounded, {|v|Dfn(v)}\{|v|Df_{n}(v)\} is uniformly bounded by some constant CC, and so (2.2) yields

|v|Df(v)lim infn|v|Dfn(v)C.|v|Df(v)\leq\liminf_{n\to\infty}|v|Df_{n}(v)\leq C.

Hence fwf\in\mathcal{L}_{\textbf{w}}.

To conclude the proof of the completeness, we need to show that fnf_{n} converges to ff in norm as nn\to\infty. Since fn(o)f(o)f_{n}(o)\to f(o), it suffices to show that

supvT|v|D(fnf)(v)0\sup\limits_{v\in T^{*}}|v|D(f_{n}-f)(v)\to 0

as nn\to\infty. Arguing by contradiction, suppose there exist ε>0\varepsilon>0 and a subsequence {fnj}j\{f_{n_{j}}\}_{j\in\mathbb{N}} such that supvT|v|D(fnjf)(v)>ε\displaystyle\sup_{v\in T^{*}}|v|D(f_{n_{j}}-f)(v)>\varepsilon for all jj\in\mathbb{N}. Then for each jj\in\mathbb{N}, we may pick two neighboring vertices vnjv_{n_{j}} and wnjw_{n_{j}}, with vnjv_{n_{j}} child of wnjw_{n_{j}}, such that

|vnj||fnj(vnj)f(vnj)(fnj(wnj)f(wnj))|ε.|v_{n_{j}}||f_{n_{j}}(v_{n_{j}})-f(v_{n_{j}})-(f_{n_{j}}(w_{n_{j}})-f(w_{n_{j}}))|\geq\varepsilon.

Since {fnj}\{f_{n_{j}}\} is Cauchy in w\mathcal{L}_{\textbf{w}}, there exists a positive integer j0j_{0} such that for each j,hj0j,h\geq j_{0}, and vTv\in T^{*}, we have

|v||fnj(v)fnh(v)(fnj(v)fnh(v))|fnjfnhw<ε2.|v||f_{n_{j}}(v)-f_{n_{h}}(v)-(f_{n_{j}}(v^{-})-f_{n_{h}}(v^{-}))|\leq\left\|f_{n_{j}}-f_{n_{h}}\right\|_{\textbf{w}}<\frac{\varepsilon}{2}.

In particular, for all hj0h\geq j_{0}, we have

|vnj0||fnj0(vnj0)fnh(vnj0)(fnj0(wnj0)fnh(wnj0))|<ε2.\displaystyle|v_{n_{j_{0}}}||f_{n_{j_{0}}}(v_{n_{j_{0}}})-f_{n_{h}}(v_{n_{j_{0}}})-(f_{n_{j_{0}}}(w_{n_{j_{0}}})-f_{n_{h}}(w_{n_{j_{0}}}))|<\frac{\varepsilon}{2}. (2.3)

On the other hand, by the pointwise convergence of fnhf_{n_{h}} to ff, for all integers hh sufficiently large

|fnh(vnj0)f(vnj0)(fnh(wnj0)f(wnj0))|<ε2|vnj0|.\displaystyle|f_{n_{h}}(v_{n_{j_{0}}})-f(v_{n_{j_{0}}})-(f_{n_{h}}(w_{n_{j_{0}}})-f(w_{n_{j_{0}}}))|<\frac{\varepsilon}{2|v_{n_{j_{0}}}|}. (2.4)

Thus, by the triangle inequality, from (2.3) and (2.4) we deduce that

|vnj0||fnj0(vnj0)f(vnj0)(fnj0(wnj0)f(wnj0))|<ε,|v_{n_{j_{0}}}||f_{n_{j_{0}}}(v_{n_{j_{0}}})-f(v_{n_{j_{0}}})-(f_{n_{j_{0}}}(w_{n_{j_{0}}})-f(w_{n_{j_{0}}}))|<\varepsilon,

contradicting the choice of vnj0v_{n_{j_{0}}} and wnj0w_{n_{j_{0}}}. Therefore w\mathcal{L}_{\textbf{w}} is a Banach space.∎

A Banach space XX of complex-valued functions on a set Ω\Omega is said to be a functional Banach space if for each ωΩ\omega\in\Omega, the point evaluation functional

eω(f)=f(ω),fX,e_{\omega}(f)=f(\omega),\quad f\in X,

is bounded; that is, there exists a constant C>0C>0 such that |f(ω)|Cf|f(\omega)|\leq C\|f\|, for each fXf\in X.

Lemma 2.4 (Lemma 11 of [9]).

Let XX be a functional Banach space on the set Ω\Omega and let ψ\psi be a complex-valued function on Ω\Omega such that MψM_{\psi} maps XX into itself. Then MψM_{\psi} is bounded on XX and |ψ(ω)|Mψ|\psi(\omega)|\leq\|M_{\psi}\| for all ωΩ\omega\in\Omega. In particular, ψ\psi is bounded.

Corollary 2.5.

The set w\mathcal{L}_{\textbf{w}} is a functional Banach space. If MψM_{\psi} is a multiplication operator on w\mathcal{L}_{\textbf{w}}, then MψM_{\psi} is bounded, its symbol ψ\psi is bounded and ψMψ\|\psi\|_{\infty}\leq\|M_{\psi}\|.

Proof.

Let fwf\in\mathcal{L}_{\textbf{w}}. Then |f(o)|fw|f(o)|\leq\left\|f\right\|_{\textbf{w}} and fixing vTv\in T^{*}, inequality (2.1) shows that the point evaluation functional ev(f)=f(v)e_{v}(f)=f(v) is bounded. Thus, w\mathcal{L}_{\textbf{w}} is a functional Banach space. The second statement is an immediate consequence of Lemma 2.4.∎

Define the little weighted Lipschitz space to be the subspace w,0\mathcal{L}_{\textbf{w},0} of w\mathcal{L}_{\textbf{w}} consisting of the functions ff such that

lim|v||v|Df(v)=0.\lim_{|v|\to\infty}|v|Df(v)=0.
Proposition 2.6.

If fw,0f\in\mathcal{L}_{\textbf{w},0}, then lim|v|f(v)log|v|=0.\lim\limits_{|v|\to\infty}\frac{f(v)}{\log|v|}=0.

Proof.

If ff is constant then the result holds trivially. Assume ff is nonconstant, so that βf=supvT|v|Df(v)>0\beta_{f}=\sup\limits_{v\in T^{*}}|v|Df(v)>0, and fix ε(0,βf)\varepsilon\in(0,\beta_{f}). Then, there exists NN\in\mathbb{N} such that |v|Df(v)<ε|v|Df(v)<\varepsilon, for all vTv\in T, with |v|N|v|\geq N. For |w|=N|w|=N and vv a descendant of ww, let u0=w,u1,,u|v|N=vu_{0}=w,u_{1},\dots,u_{|v|-N}=v be the vertices in the path from ww to vv, where uj=uj1{u_{j}}^{-}=u_{j-1}, j=1,,|v|Nj=1,\dots,|v|-N. By the triangle inequality and Proposition 2.1, we have

|f(v)|\displaystyle|f(v)| \displaystyle\leq |f(w)|+j=1|v||w||f(uj)f(uj1)|\displaystyle|f(w)|+\sum_{j=1}^{|v|-|w|}|f(u_{j})-f(u_{j-1})|
\displaystyle\leq (1+logN)fw+εk=N+1|v|1k\displaystyle(1+\log N)\left\|f\right\|_{\textbf{w}}+\varepsilon\sum_{k=N+1}^{|v|}\frac{1}{k}
\displaystyle\leq (1+k=1N11k)fwεk=2N1k+εk=2|v|1k\displaystyle\left(1+\sum_{k=1}^{N-1}\frac{1}{k}\right)\left\|f\right\|_{\textbf{w}}-\varepsilon\sum_{k=2}^{N}\frac{1}{k}+\varepsilon\sum_{k=2}^{|v|}\frac{1}{k}
<\displaystyle< 2fw+(fwε)k=2N1k+εlog|v|.\displaystyle 2\left\|f\right\|_{\textbf{w}}+(\left\|f\right\|_{\textbf{w}}-\varepsilon)\sum_{k=2}^{N}\frac{1}{k}+\varepsilon\log|v|.

Therefore, for all vertices vv of length greater than NN we obtain

|f(v)|log|v|<2fw+(fwε)k=2N1klog|v|+ε.\frac{|f(v)|}{\log|v|}<\frac{2\left\|f\right\|_{\textbf{w}}+(\left\|f\right\|_{\textbf{w}}-\varepsilon)\displaystyle\sum_{k=2}^{N}\frac{1}{k}}{\log|v|}+\varepsilon.

Hence lim|v||f(v)|log|v|ε\lim\limits_{|v|\to\infty}\displaystyle\frac{|f(v)|}{\log|v|}\leq\varepsilon. Letting ε0\varepsilon\to 0, we obtain the result.∎

The following result will be used in Section 8 to derive estimates on the essential norm of the multiplication operators on w\mathcal{L}_{\textbf{w}}.

Proposition 2.7.

Let {fn}\{f_{n}\} be a sequence of functions in w,0\mathcal{L}_{\textbf{w},0} converging to 0 pointwise in TT and such that fnw\left\|f_{n}\right\|_{\textbf{w}} is bounded. Then fn0f_{n}\to 0 weakly in w,0\mathcal{L}_{\textbf{w},0}.

Proof.

First suppose fn(o)=0f_{n}(o)=0 for all nn\in\mathbb{N}, thus fnw=supvT|v|Dfn(v)\left\|f_{n}\right\|_{\textbf{w}}=\sup\limits_{v\in T^{*}}|v|Df_{n}(v). Then, letting μ(v)=|v|\mu(v)=|v| for vTv\in T, the sequence {μDfn}\{\mu Df_{n}\} converges to 0 pointwise. Observe that the subspace of w,0\mathcal{L}_{\textbf{w},0} whose elements send oo to 0 is isomorphic to the space c0c_{0}, consisting of the sequences indexed by TT which vanish at infinity, under the supremum norm via the correspondence fμDff\mapsto\mu Df. The space c0c_{0} has dual isomorphic to the space 1\ell^{1} of absolutely summable sequences (e.g. [7]) via the correspondence g1g~c0g\in\ell^{1}\mapsto\widetilde{g}\in c_{0}^{*}, where for fc0f\in c_{0},

g~(f)=vTf(v)g(v).\widetilde{g}(f)=\sum_{v\in T}f(v)g(v).

Thus, under the identification of w,0\mathcal{L}_{\textbf{w},0} with c0c_{0}, if fnc0f_{n}\in c_{0} converges pointwise to 0 and is bounded in c0c_{0}, then for any g1g\in\ell^{1}, we have

|g~(fn)|=|vTfn(v)g(v)|vT|fn(v)||g(v)|.\displaystyle|\widetilde{g}(f_{n})|=\left|\sum_{v\in T}f_{n}(v)g(v)\right|\leq\sum_{v\in T}|f_{n}(v)||g(v)|. (2.5)

Let c=supn,vT|fn(v)|c=\sup\limits_{n\in\mathbb{N},v\in T}|f_{n}(v)|. Fixing any positive integer NN, we may split the sum on the right-hand side of (2.5) into the two sums

S1(n,N)=|v|N|fn(v)||g(v)| and S2(n,N)=|v|>N|fn(v)||g(v)|.S_{1}(n,N)=\sum_{|v|\leq N}|f_{n}(v)||g(v)|\ \hbox{ and }\ S_{2}(n,N)=\sum_{|v|>N}|f_{n}(v)||g(v)|.

Since fn0f_{n}\to 0 uniformly on the set {vT:|v|N}\{v\in T:|v|\leq N\}, we see that

S1(n,N)max|v|N|fn(v)|g10, as n.S_{1}(n,N)\leq\max_{|v|\leq N}|f_{n}(v)|\|g\|_{1}\to 0,\hbox{ as }n\to\infty.

On the other hand, since g1g\in\ell^{1}, the tail end of the series vT|g(v)|\displaystyle\sum_{v\in T}|g(v)| approaches 0. Therefore, since

limn|g~(fn)|limnS1(n,N)+supnS2(n,N)c|v|>N|g(v)|,\lim_{n\to\infty}|\widetilde{g}(f_{n})|\leq\lim_{n\to\infty}S_{1}(n,N)+\sup_{n\in\mathbb{N}}S_{2}(n,N)\leq c\sum_{|v|>N}|g(v)|,

letting NN\to\infty, we deduce that limng~(fn)=0\displaystyle\lim_{n\to\infty}\widetilde{g}(f_{n})=0.

Hence, if fn(o)=0f_{n}(o)=0, then fnf_{n} converges to 0 weakly. In the general case, define Fn=fnfn(o)F_{n}=f_{n}-f_{n}(o). By the previous part, Fn0F_{n}\to 0 weakly. Since fn(o)0f_{n}(o)\to 0, we conclude that fn0f_{n}\to 0 weakly as well.∎

Denote by χA\chi_{A} the characteristic function of the set AA and use the simpler notation χv\chi_{v} for the function χ{v}\chi_{\{v\}}.

Proposition 2.8.

The set

𝒫={k=1Nakpvk:N,ak,vkT,1kN},\mathcal{P}=\left\{\sum\limits_{k=1}^{N}a_{k}p_{v_{k}}:N\in\mathbb{N},a_{k}\in\mathbb{C},v_{k}\in T,1\leq k\leq N\right\},

is dense in w,0\mathcal{L}_{\textbf{w},0}, where pv=χSvp_{v}=\chi_{S_{v}} for vTv\in T.

Proof.

Fix vTv\in T and observe that Dpv=χvDp_{v}=\chi_{v}, so that for wTw\in T^{*}, we have

|w|Dpv(w)={0 if wv,|v| if w=v.|w|Dp_{v}(w)=\begin{cases}0&\quad\hbox{ if }w\neq v,\\ |v|&\quad\hbox{ if }w=v.\end{cases}

Thus, as |w||w|\to\infty, |w|Dpv(w)0|w|Dp_{v}(w)\to 0, proving that pvw,0p_{v}\in\mathcal{L}_{\textbf{w},0}.

Let fw,0f\in\mathcal{L}_{\textbf{w},0} and for nn\in\mathbb{N}, define

fn(v)={f(v) if |v|n,f(vn) if |v|>n,f_{n}(v)=\begin{cases}f(v)&\quad\hbox{ if }|v|\leq n,\\ f(v_{n})&\quad\hbox{ if }|v|>n,\end{cases}

where vnv_{n} is the ancestor of vv of length nn. Observe that for vTv\in T^{*},

χv=pvwv+pw,\displaystyle\chi_{v}=p_{v}-\sum_{w\in v^{+}}p_{w}, (2.6)

where v+={wT:w=v}v^{+}=\{w\in T:w^{-}=v\}. Therefore, for nn\in\mathbb{N}, we have

fn\displaystyle f_{n} =\displaystyle= |v|<nf(v)χv+|v|=nf(v)pv\displaystyle\sum_{|v|<n}f(v)\chi_{v}+\sum_{|v|=n}f(v)p_{v}
=\displaystyle= |v|<nf(v)(pvwv+pw)+|v|=nf(v)pv\displaystyle\sum_{|v|<n}f(v)\left(p_{v}-\sum_{w\in v^{+}}p_{w}\right)+\sum_{|v|=n}f(v)p_{v}
=\displaystyle= |v|nf(v)pv|v|<nf(v)wv+pw\displaystyle\sum_{|v|\leq n}f(v)p_{v}-\sum_{|v|<n}f(v)\sum_{w\in v^{+}}p_{w}

Thus, fnf_{n} is a finite linear combination of the functions pvp_{v} and

fnfw=sup|v|>n|v|Df(v)0\left\|f_{n}-f\right\|_{\textbf{w}}=\sup_{|v|>n}|v|Df(v)\to 0

as nn\to\infty, proving the result.∎

Remark 2.9.

Since [i]={z:Rez,Imz}\mathbb{Q}[i]=\{z\in\mathbb{C}:{\rm{Re}}\,z,{\rm{Im}}\,z\in\mathbb{Q}\} is dense in \mathbb{C}, and TT is countable, the subset of 𝒫\mathcal{P} consisting of the finite linear combinations of the functions pvp_{v} with coefficients in [i]\mathbb{Z}[i] is countable and dense in w,0\mathcal{L}_{\textbf{w},0}. Therefore, w,0\mathcal{L}_{\textbf{w},0} is a closed separable subspace of w\mathcal{L}_{\textbf{w}}.

3. Cyclic Vectors in the Weighted Little Lipschitz Space

Definition 3.1.

Let XX be a Banach space of functions on TT such that 𝒫\mathcal{P} is dense in XX. A function ff in XX is called a cyclic vector if XX is the closure [f][f] of the functions of the form pvfp_{v}f.

If fw,0f\in\mathcal{L}_{\textbf{w},0} vanishes at some vertex uu, then ff cannot be a cyclic vector since the function χu\chi_{u} cannot be the limit in w,0\mathcal{L}_{\textbf{w},0} of multiples of ff. For the converse, we have the following result.

Theorem 3.2.

Let fw,0f\in\mathcal{L}_{\textbf{w},0} be such that |f(v)|δ>0|f(v)|\geq\delta>0 for all vTv\in T. Then ff is a cyclic vector in w,0\mathcal{L}_{\textbf{w},0}.

Proof.

First observe that to prove the result, it suffices to show that the constant function 1 is a limit in w,0\mathcal{L}_{\textbf{w},0} of functions of the form pvfp_{v}f. Indeed, observe that if v,uTv,u\in T, then

pvpu={0 if SvSu=,pv if u is an ancestor of v,pu if v is an ancestor of u or v=u.p_{v}p_{u}=\begin{cases}0&\quad\hbox{ if }S_{v}\cap S_{u}=\emptyset,\\ p_{v}&\quad\hbox{ if }u\hbox{ is an ancestor of }v,\\ p_{u}&\quad\hbox{ if }v\hbox{ is an ancestor of }u\hbox{ or }v=u.\end{cases}

Thus, 1[f]1\in[f] implies that pv[f]p_{v}\in[f] for all vTv\in T. By Proposition 2.8, it follows that ff is a cyclic vector in w,0\mathcal{L}_{\textbf{w},0}.

For nn\in\mathbb{N}, define fnf_{n} as in the proof of Proposition 2.8. Then

ffn1w=sup|v|>n|v|Df(v)|f(vn)|1δsup|v|>n|v|Df(v)0\displaystyle\left\|\frac{f}{f_{n}}-1\right\|_{\textbf{w}}=\sup_{|v|>n}{|v|}\frac{Df(v)}{|f(v_{n})|}\leq\frac{1}{\delta}\sup_{|v|>n}|v|Df(v)\to 0

as nn\to\infty. On the other hand, letting

gn=|v|n21fn(v)χv,g_{n}=\sum_{|v|\leq n^{2}}\frac{1}{f_{n}(v)}\chi_{v},

we see that

gnfffnw=sup|v|>n2|v|Df(v)f(vn)1δsup|v|>n2|v|Df(v)0\left\|g_{n}f-\frac{f}{f_{n}}\right\|_{\textbf{w}}=\sup_{|v|>n^{2}}\frac{|v|Df(v)}{f(v_{n})}\leq\frac{1}{\delta}\sup_{|v|>n^{2}}|v|Df(v)\to 0

as nn\to\infty. Thus, gnf1w0\left\|g_{n}f-1\right\|_{\textbf{w}}\to 0 as nn\to\infty. For vTv\in T^{*}, recalling (2.6), we see that the functions gnfg_{n}f belong to [f][f]. Therefore, ff is a cyclic vector.∎

It is still an open question as to whether there exist cyclic vectors that are not bounded away from 0.

4. Boundedness of MψM_{\psi}

In this section, we characterize the bounded multiplication operators acting on w\mathcal{L}_{\textbf{w}} and w,0\mathcal{L}_{\textbf{w},0}. This characterization provides a big-oh criterion for boundedness, which corresponds to a little-oh criterion for compactness developed in Section 7.

Theorem 4.1.

For a function ψ\psi on TT the following statements are equivalent.

  1. (a)

    MψM_{\psi} is bounded on w\mathcal{L}_{\textbf{w}}.

  2. (b)

    MψM_{\psi} is bounded on w,0\mathcal{L}_{\textbf{w},0}.

  3. (c)

    ψL\psi\in L^{\infty} and supvT|v|log|v|Dψ(v)<.\displaystyle\sup\limits_{v\in T^{*}}|v|\log|v|D\psi(v)<\infty.

Proof.

We first prove (a)\Longrightarrow(c). Assume MψM_{\psi} is bounded on w\mathcal{L}_{\textbf{w}}. The boundedness of ψ\psi follows immediately from Corollary 2.5.

For vTv\in T, define

f(v)={ 0 if v=o,log|v| if vT.f(v)=\begin{cases}\ \ 0&\quad\hbox{ if }v=o,\\ \log|v|&\quad\hbox{ if }v\in T^{*}.\end{cases}

Then, for |v|=1|v|=1, we have Df(v)=0Df(v)=0, while for |v|2|v|\geq 2, by Lemma 2.2, we obtain

Df(v)=log(|v||v|1)1|v|1.Df(v)=\log\left(\frac{|v|}{|v|-1}\right)\leq\frac{1}{|v|-1}.

Thus, |v|Df(v)|v||v|12.|v|Df(v)\leq\displaystyle\frac{|v|}{|v|-1}\leq 2. Therefore, fwf\in\mathcal{L}_{\textbf{w}}. Furthermore, for vTv\in T^{*} we have

Dψ(v)|f(v)|\displaystyle D\psi(v)|f(v)| \displaystyle\leq |ψ(v)f(v)ψ(v)f(v)|+|ψ(v)f(v)ψ(v)f(v)|\displaystyle|\psi(v)f(v)-\psi(v^{-})f(v^{-})|+|\psi(v^{-})f(v^{-})-\psi(v^{-})f(v)| (4.1)
=\displaystyle= D(ψf)(v)+|ψ(v)|Df(v).\displaystyle D(\psi f)(v)+|\psi(v^{-})|Df(v).

Thus, by the boundedness of MψM_{\psi}, for vTv\in T^{*}, we obtain

|v|Dψ(v)|f(v)|\displaystyle|v|D\psi(v)|f(v)| \displaystyle\leq |v|D(ψf)(v)+|ψ(v)||v|Df(v)\displaystyle|v|D(\psi f)(v)+|\psi(v^{-})||v|Df(v)
\displaystyle\leq Mψfw+ψfw.\displaystyle\left\|M_{\psi}f\right\|_{\textbf{w}}+\|\psi\|_{\infty}\left\|f\right\|_{\textbf{w}}.

Hence

supvT|v|log|v|Dψ(v)<.\displaystyle\sup_{v\in T^{*}}|v|\log|v|D\psi(v)<\infty. (4.2)

Next, we prove (c)\Longrightarrow(a). Assume ψ\psi is bounded and (4.2) holds. Let fwf\in\mathcal{L}_{\textbf{w}} and vTv\in T^{*}. Note that

D(ψf)(v)\displaystyle D(\psi f)(v) \displaystyle\leq |ψ(v)f(v)ψ(v)f(v)|+|ψ(v)f(v)ψ(v)f(v)|\displaystyle|\psi(v)f(v)-\psi(v^{-})f(v)|+|\psi(v^{-})f(v)-\psi(v^{-})f(v^{-})| (4.3)
=\displaystyle= Dψ(v)|f(v)|+|ψ(v)|Df(v).\displaystyle D\psi(v)|f(v)|+|\psi(v^{-})|Df(v).

Thus, by Proposition 2.1, we have

|v|D(ψf)(v)\displaystyle|v|D(\psi f)(v) \displaystyle\leq |v|Dψ(v)|f(v)|+|ψ(v)||v|Df(v)\displaystyle|v|D\psi(v)|f(v)|+|\psi(v^{-})||v|Df(v)
\displaystyle\leq |v|(1+log|v|)Dψ(v)fw+ψfw.\displaystyle|v|(1+\log|v|)D\psi(v)\left\|f\right\|_{\textbf{w}}+\|\psi\|_{\infty}\left\|f\right\|_{\textbf{w}}.

In particular, for |v|3|v|\geq 3, we have

|v|D(ψf)(v)(2|v|log|v|Dψ(v)+ψ)fw,|v|D(\psi f)(v)\leq(2|v|\log|v|D\psi(v)+\|\psi\|_{\infty})\left\|f\right\|_{\textbf{w}},

proving that ψfw\psi f\in\mathcal{L}_{\textbf{w}}. The boundedness of MψM_{\psi} follows from Lemma 2.4.

Now, we prove (b)\Longrightarrow(c). Assume MψM_{\psi} is bounded on w,0\mathcal{L}_{\textbf{w},0}. For 0<α<10<\alpha<1, define

fα(v)={0 if v=o,(log|v|)α if vT.f_{\alpha}(v)=\begin{cases}0&\hbox{ if }v=o,\\ (\log|v|)^{\alpha}&\hbox{ if }v\in T^{*}.\end{cases}

Then |v|Dfα(v)0|v|Df_{\alpha}(v)\to 0 as |v||v|\to\infty, so that fαw,0f_{\alpha}\in\mathcal{L}_{\textbf{w},0}. Since for 0<α<10<\alpha<1, the function xxxαx\mapsto x-x^{\alpha} is increasing for x1x\geq 1, it follows that for |v|2|v|\geq 2, Dfα(v)log(|v|)log(|v|1)Df_{\alpha}(v)\leq\log(|v|)-\log(|v|-1), so by Lemma 2.2, we have

|v|Dfα(v)|v|(log|v|log(|v|1))|v||v|12.|v|Df_{\alpha}(v)\leq|v|(\log|v|-\log(|v|-1))\leq\frac{|v|}{|v|-1}\leq 2.

Furthermore, for |v|=1|v|=1, |v|Dfα(v)=0|v|Df_{\alpha}(v)=0. Thus, fαw2\left\|f_{\alpha}\right\|_{\textbf{w}}\leq 2 for all α(0,1)\alpha\in(0,1). Moreover, by Lemma 2.4, the function ψ\psi is bounded, so by (4.1), for vTv\in T^{*}, we have

|v|Dψ(v)|fα(v)|\displaystyle|v|D\psi(v)|f_{\alpha}(v)| \displaystyle\leq |v|D(ψfα)(v)+|v||ψ(v)|Dfα(v)\displaystyle|v|D(\psi f_{\alpha})(v)+|v||\psi(v^{-})|Df_{\alpha}(v)
\displaystyle\leq Mψfαw+ψfαw.\displaystyle\left\|M_{\psi}f_{\alpha}\right\|_{\textbf{w}}+\|\psi\|_{\infty}\left\|f_{\alpha}\right\|_{\textbf{w}}.

This implies that

|v|Dψ(v)(log|v|)α(Mψ+ψ)fαw.|v|D\psi(v)(\log|v|)^{\alpha}\leq(\|M_{\psi}\|+\|\psi\|_{\infty})\left\|f_{\alpha}\right\|_{\textbf{w}}.

Letting α\alpha approach 1, by the boundedness of fαw\left\|f_{\alpha}\right\|_{\textbf{w}}, we obtain (4.2).

Lastly, we prove (c)\Longrightarrow(b). Assume (c) holds and let fw,0f\in\mathcal{L}_{\textbf{w},0}. By (4.3) and Proposition 2.6, for |v|>1|v|>1 we have

|v|D(ψf)(v)\displaystyle|v|D(\psi f)(v) \displaystyle\leq |v|Dψ(v)|f(v)|+|v||ψ(v)|Df(v)\displaystyle|v|D\psi(v)|f(v)|+|v||\psi(v^{-})|Df(v)
\displaystyle\leq |v|log|v|Dψ(v)|f(v)|log|v|+ψ|v|Df(v)0\displaystyle|v|\log|v|D\psi(v)\frac{|f(v)|}{\log|v|}+\|\psi\|_{\infty}|v|Df(v)\to 0

as |v||v|\to\infty. Therefore, ψfw,0\psi f\in\mathcal{L}_{\textbf{w},0}. The boundedness of MψM_{\psi} on w,0\mathcal{L}_{\textbf{w},0} follows from Lemma 2.4.∎

5. Norm of MψM_{\psi}

In this section, we provide estimates on the norm of the bounded multiplication operators on w\mathcal{L}_{\textbf{w}} and w,0\mathcal{L}_{\textbf{w},0}.

Theorem 5.1.

Let MψM_{\psi} be a bounded multiplication operator on w\mathcal{L}_{\textbf{w}} or w,0\mathcal{L}_{\textbf{w},0}. Then

max{ψw,ψ}Mψψ+supvT|v|(1+log|v|)Dψ(v).\max\{\left\|\psi\right\|_{\textbf{w}},\|\psi\|_{\infty}\}\leq\|M_{\psi}\|\leq\|\psi\|_{\infty}+\displaystyle\sup_{v\in T^{*}}|v|(1+\log|v|)D\psi(v).
Proof.

Let ff be the function identically equal to 1 on TT. Since DfDf is identically 0, fwf\in\mathcal{L}_{\textbf{w}}, and fw=1\left\|f\right\|_{\textbf{w}}=1. Thus, ψ=ψfw\psi=\psi f\in\mathcal{L}_{\textbf{w}} and Mψfw=ψw\left\|M_{\psi}f\right\|_{\textbf{w}}=\left\|\psi\right\|_{\textbf{w}}. Therefore, ψwMψ\left\|\psi\right\|_{\textbf{w}}\leq\|M_{\psi}\|. Moreover, by Lemma 2.4, ψL\psi\in L^{\infty} and ψMψ\|\psi\|_{\infty}\leq\|M_{\psi}\|, proving the lower estimate.

Let fwf\in\mathcal{L}_{\textbf{w}} such that fw=1\left\|f\right\|_{\textbf{w}}=1. Then, using (4.3), Proposition 2.1, and the fact that supvT|v|Df(v)=1|f(o)|\sup\limits_{v\in T^{*}}|v|Df(v)=1-|f(o)|, we obtain

Mψf\displaystyle\|M_{\psi}f\| |ψ(o)||f(o)|+supvT|v||f(v)|Dψ(v)+supvT|v||ψ(v)|Df(v)\displaystyle\leq|\psi(o)||f(o)|+\sup_{v\in T^{*}}|v||f(v)|D\psi(v)+\sup_{v\in T^{*}}|v||\psi(v^{-})|Df(v)
|ψ(o)||f(o)|+supvT|v|(1+log|v|)Dψ(v)+ψsupvT|v|Df(v)\displaystyle\leq|\psi(o)||f(o)|+\sup_{v\in T^{*}}|v|(1+\log|v|)D\psi(v)+\|\psi\|_{\infty}\sup_{v\in T^{*}}|v|Df(v)
=|ψ(o)||f(o)|+supvT|v|(1+log|v|)Dψ(v)+ψ(1|f(o)|)\displaystyle=|\psi(o)||f(o)|+\sup_{v\in T^{*}}|v|(1+\log|v|)D\psi(v)+\|\psi\|_{\infty}(1-|f(o)|)
ψ+supvT|v|(1+log|v|)Dψ(v),\displaystyle\leq\|\psi\|_{\infty}+\sup_{v\in T^{*}}|v|(1+\log|v|)D\psi(v),

proving the upper estimate.∎

6. Spectrum of MψM_{\psi}

In this section, we study the spectra of the bounded multiplication operator MψM_{\psi} on w\mathcal{L}_{\textbf{w}} and w,0\mathcal{L}_{\textbf{w},0}. We show that the point spectrum is nonempty and, in fact, it is a dense subset of the spectrum. We also show that the spectrum and the approximate point spectrum are equal to the closure of the range of the symbol. We deduce a characterization of the bounded multiplications operators that are bounded below.

Recall that for a bounded operator SS on a Banach space XX, the spectrum of SS is defined as

σ(S)={λ:SλI is not invertible},\sigma(S)=\left\{\lambda\in\mathbb{C}:S-\lambda I\text{ is not invertible}\right\},

where II is the identity operator on XX. The point spectrum of SS is defined as

σp(S)={λ:ker(SλI){0}}.\sigma_{p}(S)=\left\{\lambda\in\mathbb{C}:\mathrm{ker}(S-\lambda I)\neq\{0\}\right\}.

The approximate point spectrum of SS is defined as

σap(S)={λ:{xn}X, such that xn=1n, and (SλI)xn0}.\sigma_{ap}(S)=\left\{\lambda\in\mathbb{C}:\exists\{x_{n}\}\subseteq X,\text{ such that }\,\|x_{n}\|=1\ \forall n,\text{ and }\,\|(S-\lambda I)x_{n}\|\to 0\right\}.

The following inclusions hold:

σp(S)σap(S)σ(S).\displaystyle\sigma_{p}(S)\subseteq\sigma_{ap}(S)\subseteq\sigma(S). (6.1)
Theorem 6.1.

Let MψM_{\psi} be a bounded multiplication operator on w\mathcal{L}_{\textbf{w}} or w,0\mathcal{L}_{\textbf{w},0}. Then

  1. (a)

    σp(Mψ)=ψ(T)\sigma_{p}(M_{\psi})=\psi(T).

  2. (a)

    σ(Mψ)=ψ(T)¯\sigma(M_{\psi})=\overline{\psi(T)}.

Proof.

To prove (a), suppose λσp(Mψ)\lambda\in\sigma_{p}(M_{\psi}). Then there exists a non-zero function fw,0f\in\mathcal{L}_{\textbf{w},0} such that MψλfM_{\psi-\lambda}f is identically zero. Since ff is not identically zero, there exists wTw\in T such that f(w)0f(w)\neq 0. Then 0=(Mψλf)(w)=(ψ(w)λ)f(w)0=(M_{\psi-\lambda}f)(w)=(\psi(w)-\lambda)f(w), and so ψ(w)=λ\psi(w)=\lambda, proving that λ\lambda is in the image of ψ\psi.

Conversely, suppose λ\lambda is in the image of ψ\psi. Then there exists wTw\in T such that ψ(w)=λ\psi(w)=\lambda. So we see that MψλχwM_{\psi-\lambda}\chi_{w} is identically zero. Thus λσp(Mψ)\lambda\in\sigma_{p}(M_{\psi}). Therefore σp(Mψ)=ψ(T)\sigma_{p}(M_{\psi})=\psi(T).

To prove (b), observe that since the spectrum is closed, the inclusion ψ(T)¯σ(Mψ)\overline{\psi(T)}\subseteq\sigma(M_{\psi}) follows at once from part (a) by passing to the closure.

Conversely, if λψ(T)¯\lambda\notin\overline{\psi(T)}, then |ψ(v)λ|c|\psi(v)-\lambda|\geq c\, for some positive constant cc and all vTv\in T. Thus, the function g=(ψλ)1g=(\psi-\lambda)^{-1} is bounded on TT. Furthermore,

supvT|v|log|v|Dg(v)\displaystyle\sup_{v\in T^{*}}|v|\log|v|Dg(v) =\displaystyle= supvT|v|log|v||1ψ(v)λ1ψ(v)λ|\displaystyle\sup_{v\in T^{*}}|v|\log|v|\left|\frac{1}{\psi(v)-\lambda}-\frac{1}{\psi(v^{-})-\lambda}\right|
\displaystyle\leq 1c2supvT|v|log|v|Dψ(v)<.\displaystyle\frac{1}{c^{2}}\sup_{v\in T^{*}}|v|\log|v|D\psi(v)<\infty.

By Theorem 4.1, we deduce that Mg=M(ψλ)1M_{g}=M_{(\psi-\lambda)^{-1}} is a bounded operator on w\mathcal{L}_{\textbf{w}} or w,0\mathcal{L}_{\textbf{w},0}, which implies that λσ(Mψ)\lambda\notin\sigma(M_{\psi}). Therefore σ(Mψ)=ψ(T)¯\sigma(M_{\psi})=\overline{\psi(T)}.∎

The following proposition relates the boundary of the spectrum to the approximate point spectrum.

Proposition 6.2 (Proposition 6.7 of [7]).

If SS is a bounded operator on a Banach space, then the boundary of σ(S)\sigma(S) is a subset of σap(S)\sigma_{ap}(S).

Using Theorem 6.1, the inclusions (6.1) and Proposition 6.2, we obtain the following result.

Corollary 6.3.

Let MψM_{\psi} be a bounded multiplication operator on w\mathcal{L}_{\textbf{w}} or w,0\mathcal{L}_{\textbf{w},0}. Then σap(Mψ)=ψ(T)¯\sigma_{ap}(M_{\psi})=\overline{\psi(T)}.

A bounded operator SS on a Banach space XX is said to be bounded below if there exists a positive constant cc such that Sxcx\|Sx\|\geq c\|x\| for all xXx\in X. Note that a bounded operator that is bounded below is necessarily injective.

The following result connects the approximate point spectrum and the operators that are bounded below.

Proposition 6.4 (Proposition 6.4 of [7]).

If SS is a bounded operator on a Banach space, then λσap(S)\lambda\not\in\sigma_{ap}(S) if and only if SλIS-\lambda I is bounded below.

We next characterize the bounded multiplication operators on w\mathcal{L}_{\textbf{w}} or w,0\mathcal{L}_{\textbf{w},0} which are bounded below.

Theorem 6.5.

If MψM_{\psi} is a bounded multiplication operator on w\mathcal{L}_{\textbf{w}} or w,0\mathcal{L}_{\textbf{w},0}, then MψM_{\psi} is bounded below if and only if infvT|ψ(v)|>0\inf\limits_{v\in T}|\psi(v)|>0.

Proof.

By Proposition 6.4, if MψM_{\psi} is a bounded operator on w\mathcal{L}_{\textbf{w}} or w,0\mathcal{L}_{\textbf{w},0}, then MψM_{\psi} is bounded below if and only if 0σap(Mψ)0\notin\sigma_{ap}(M_{\psi}). By Corollary 6.3, this condition is equivalent to 0ψ(T)¯0\notin\overline{\psi(T)}, i.e. infvT|ψ(v)|>0\inf\limits_{v\in T}|\psi(v)|>0.∎

7. Compactness of MψM_{\psi}

In this section, we characterize the compact multiplication operators on w\mathcal{L}_{\textbf{w}} and w,0\mathcal{L}_{\textbf{w},0}.

Lemma 7.1.

A bounded multiplication operator MψM_{\psi} on w\mathcal{L}_{\textbf{w}} (respectively, w,0\mathcal{L}_{\textbf{w},0}) is compact if and only if ψfnw0\left\|\psi f_{n}\right\|_{\textbf{w}}\to 0 as nn\to\infty for every bounded sequence {fn}\{f_{n}\} in w\mathcal{L}_{\textbf{w}} (respectively, w,0\mathcal{L}_{\textbf{w},0}) converging to 0 pointwise.

Proof.

We shall prove the result for the bounded operator MψM_{\psi} acting on w\mathcal{L}_{\textbf{w}}. The proof for the case of w,0\mathcal{L}_{\textbf{w},0} is analogous.

Assume MψM_{\psi} is compact on w\mathcal{L}_{\textbf{w}} and let {fn}\{f_{n}\} be a bounded sequence in w\mathcal{L}_{\textbf{w}} converging to 0 pointwise. By rescaling the sequence, if necessary, we may assume fnw1\left\|f_{n}\right\|_{\textbf{w}}\leq 1 for all nn\in\mathbb{N}. By the compactness of MψM_{\psi}, {fn}\{f_{n}\} has a subsequence {fnk}\{f_{n_{k}}\} such that {ψfnk}\{\psi f_{n_{k}}\} converges in norm to some function fwf\in\mathcal{L}_{\textbf{w}}. Observe that ψ(o)fnk(o)f(o)\psi(o)f_{n_{k}}(o)\to f(o) and for vTv\in T^{*}, by Proposition 2.1 applied to the function ψfnkf\psi f_{n_{k}}-f, we have

|ψ(v)fnk(v)f(v)|(1+log|v|)ψfnkfw.\displaystyle|\psi(v)f_{n_{k}}(v)-f(v)|\leq(1+\log|v|)\left\|\psi f_{n_{k}}-f\right\|_{\textbf{w}}.

Therefore, ψfnkf\psi f_{n_{k}}\to f pointwise. Since by assumption, fn0f_{n}\to 0 pointwise, it follows that ff must be identically 0, whence ψfnw0\left\|\psi f_{n}\right\|_{\textbf{w}}\to 0. Since 0 is the only limit point in w\mathcal{L}_{\textbf{w}} of the sequence {ψfn}\{\psi f_{n}\}, it follows that ψfnw0\left\|\psi f_{n}\right\|_{\textbf{w}}\to 0 as nn\to\infty.

Conversely, suppose that for every bounded sequence {fn}\{f_{n}\} in w\mathcal{L}_{\textbf{w}} converging to 0 pointwise, ψfnw0\left\|\psi f_{n}\right\|_{\textbf{w}}\to 0 as nn\to\infty. Let {gn}\{g_{n}\} be a sequence in w\mathcal{L}_{\textbf{w}} with gnw1\left\|g_{n}\right\|_{\textbf{w}}\leq 1. Then |gn(o)|1|g_{n}(o)|\leq 1 and by Proposition 2.1, for each vTv\in T^{*}, we have |gn(v)|1+log|v||g_{n}(v)|\leq 1+\log|v|. Therefore, gng_{n} is uniformly bounded on finite subsets of TT and so some subsequence, which for notational convenience we reindex as the original sequence, converges to some function gg. Then, for vTv\in T^{*}, we have

Dg(v)|g(v)g(v)(gn(v)gn(v))|+Dgn(v).Dg(v)\leq|g(v)-g(v^{-})-(g_{n}(v)-g_{n}(v^{-}))|+Dg_{n}(v).

Fix ε>0\varepsilon>0 and vTv\in T, |v|2|v|\geq 2. Since gngg_{n}\to g pointwise,

|gn(v)g(v)|<ε/(2|v|)|g_{n}(v)-g(v)|<\varepsilon/(2|v|)

and

|gn(v)g(v)|<ε/(2|v|)|g_{n}(v^{-})-g(v^{-})|<\varepsilon/(2|v^{-}|)

for all nn sufficiently large. Therefore |v|Dg(v)<ε+|v|Dgn(v)|v|Dg(v)<\varepsilon+|v|Dg_{n}(v) for nn sufficiently large, so gwg\in\mathcal{L}_{\textbf{w}}. Therefore, the sequence {fn}\{f_{n}\} defined by fn=gngf_{n}=g_{n}-g is bounded in w\mathcal{L}_{\textbf{w}} and converges to 0 pointwise; hence, by the hypothesis, ψfnw0\left\|\psi f_{n}\right\|_{\textbf{w}}\to 0 as nn\to\infty. We conclude that ψgnψg\psi g_{n}\to\psi g in norm, proving the compactness of MψM_{\psi}.∎

Theorem 7.2.

For MψM_{\psi} a bounded multiplication operator on w\mathcal{L}_{\textbf{w}}, the following are equivalent statements.

  1. (a)

    MψM_{\psi} is compact on w\mathcal{L}_{\textbf{w}}.

  2. (b)

    MψM_{\psi} is compact on w,0\mathcal{L}_{\textbf{w},0}.

  3. (c)

    lim|v|ψ(v)=0\displaystyle\lim\limits_{|v|\to\infty}\psi(v)=0 and lim|v||v|log|v|Dψ(v)=0.\displaystyle\lim\limits_{|v|\to\infty}|v|\log|v|D\psi(v)=0.

Proof.

First, we prove (a)\Longrightarrow(c). Assume MψM_{\psi} is compact on w\mathcal{L}_{\textbf{w}}. Let {vn}\{v_{n}\} be a sequence in TT such that 2<|vn|2<|v_{n}|\to\infty. We are going to show that

limnψ(vn)=0\displaystyle\lim\limits_{n\to\infty}\psi(v_{n})=0 (7.1)
limn|vn|log|vn|Dψ(vn)=0.\displaystyle\lim\limits_{n\to\infty}|v_{n}|\log|v_{n}|D\psi(v_{n})=0. (7.2)

Let fn=1|vn|χvnf_{n}=\displaystyle\frac{1}{|v_{n}|}\chi_{v_{n}}. Then fn0f_{n}\to 0 pointwise and

fnw=supvT|v|Dfn(v)=|vn|+1|vn|<32,\left\|f_{n}\right\|_{\textbf{w}}=\sup_{v\in T^{*}}|v|Df_{n}(v)=\frac{|v_{n}|+1}{|v_{n}|}<\frac{3}{2},

so that by Lemma 7.1, we obtain |ψ(vn)|ψfnw0|\psi(v_{n})|\leq\left\|\psi f_{n}\right\|_{\textbf{w}}\to 0 as nn\to\infty, proving (7.1).

To prove (7.2), for nn\in\mathbb{N} let

gn(v)={0 if |v|<|vn|,2log|v|log|vn| if |vn||v|<|vn|1,log|vn| if |v||vn|1.g_{n}(v)=\begin{cases}0&\quad\hbox{ if }|v|<\sqrt{|v_{n}|},\\ 2\log|v|-\log|v_{n}|&\quad\hbox{ if }\sqrt{|v_{n}|}\leq|v|<|v_{n}|-1,\\ \log|v_{n}|&\quad\hbox{ if }|v|\geq|v_{n}|-1.\end{cases}

Then Dgn(v)=0Dg_{n}(v)=0 if |v||vn||v|\leq\sqrt{|v_{n}|} or |v|>|vn||v|>|v_{n}|, and if |vn|<|v|<|vn|1\sqrt{|v_{n}|}<|v|<|v_{n}|-1 then |v|Dgn(v)4|v|Dg_{n}(v)\leq 4. Thus, {gnw}\{\left\|g_{n}\right\|_{\textbf{w}}\} is bounded and gn0g_{n}\to 0 pointwise as nn\to\infty. By Lemma 7.1, we get |vn|log|vn||ψ(vn)|ψgnw0|v_{n}|\log|v_{n}||\psi(v_{n})|\leq\left\|\psi g_{n}\right\|_{\textbf{w}}\to 0 as nn\to\infty.

Next, we prove (c)\Longrightarrow(a). Assume the conditions in (c) hold and set aside the case when ψ\psi is the constant 0. By Lemma 7.1, to prove that MψM_{\psi} is compact on w\mathcal{L}_{\textbf{w}}, it suffices to show that if {fn}\{f_{n}\} is a sequence in w\mathcal{L}_{\textbf{w}} converging to 0 pointwise and such that s=supnfnw<s=\displaystyle\sup_{n\in\mathbb{N}}\left\|f_{n}\right\|_{\textbf{w}}<\infty, then ψfnw0\left\|\psi f_{n}\right\|_{\textbf{w}}\to 0 as nn\to\infty. Let {fn}\{f_{n}\} be such a sequence and fix a positive number ε\varepsilon. Then |fn(o)|<ε3ψw|f_{n}(o)|<\displaystyle\frac{\varepsilon}{3\left\|\psi\right\|_{\textbf{w}}} for all nn sufficiently large, and there exists MM\in\mathbb{N} such that

|ψ(v)|<ε3s and |v|log|v|Dψ(v)<ε3s for |v|M.|\psi(v)|<\frac{\varepsilon}{3s}\ \hbox{ and }\ |v|\log|v|D\psi(v)<\frac{\varepsilon}{3s}\ \hbox{ for }\ |v|\geq M.

If |v|>M|v|>M, then |v|M|v^{-}|\geq M, and |ψ(v)|<ε3s|\psi(v^{-})|<\displaystyle\frac{\varepsilon}{3s}. Thus, by (4.3) and Proposition 2.1, we obtain

|v|D(ψfn)(v)|v|Dψ(v)(1+log|v|)fnw+ε3.|v|D(\psi f_{n})(v)\leq|v|D\psi(v)(1+\log|v|)\left\|f_{n}\right\|_{\textbf{w}}+\frac{\varepsilon}{3}.

In particular, for |v|>M|v|>M, we get

|v|D(ψfn)(v)2|v|log|v|Dψ(v)s+ε3<ε.|v|D(\psi f_{n})(v)\leq 2|v|\log|v|D\psi(v)s+\frac{\varepsilon}{3}<\varepsilon.

Since fn0f_{n}\to 0 uniformly on {vT:|v|M}\{v\in T:|v|\leq M\} as nn\to\infty, so does the sequence {|w|D(ψfn)(w)}wT\{|w|D(\psi f_{n})(w)\}_{w\in T^{*}} (under an identification of TT with \mathbb{N}). Therefore, for nn sufficiently large and for each vTv\in T^{*}, |v|D(ψfn)(v)<ε|v|D(\psi f_{n})(v)<\varepsilon. On the other hand, fn(o)0f_{n}(o)\to 0, and so ψfnw0\left\|\psi f_{n}\right\|_{\textbf{w}}\to 0, as nn\to\infty.

Note that for nn\in\mathbb{N} the functions fnf_{n} and gng_{n} defined to prove (a)\Longrightarrow(c) are in w,0\mathcal{L}_{\textbf{w},0}. Therefore the proof of (b)\Longrightarrow(c) is analogous. The converse can also be proved similarly. ∎

8. Essential Norm of MψM_{\psi}

In this section, we provide estimates on the essential norm of the bounded multiplication operators on w\mathcal{L}_{\textbf{w}}. We recall that the essential norm Se\|S\|_{e} of an operator SS on a Banach space XX is defined as

Se=inf{SK:K compact operator on X}.\|S\|_{e}=\inf\{\|S-K\|:K\hbox{ compact operator on }X\}.
Definition 8.1.

Given a bounded multiplication operator MψM_{\psi} on w\mathcal{L}_{\textbf{w}} or w,0\mathcal{L}_{\textbf{w},0}, define

A(ψ)\displaystyle A(\psi) =\displaystyle= limnsup|v|n|ψ(v)|\displaystyle\lim_{n\to\infty}\sup_{|v|\geq\,n}|\psi(v)|
B(ψ)\displaystyle B(\psi) =\displaystyle= limnsup|v|n|v|log|v|Dψ(v).\displaystyle\lim_{n\to\infty}\sup_{|v|\geq\,n}|v|\log|v|D\psi(v).
Theorem 8.2.

Let MψM_{\psi} be bounded on w\mathcal{L}_{\textbf{w}} or w,0\mathcal{L}_{\textbf{w},0}. Then

Mψemax{A(ψ),B(ψ)}.\|M_{\psi}\|_{e}\geq\max\left\{A(\psi),B(\psi)\right\}.
Proof.

For each nn\in\mathbb{N} and vTv\in T, define fn=1nχ{v:|v|=n}f_{n}=\displaystyle\frac{1}{n}\chi_{\{v:\,|v|=n\}}. Then fnw,0f_{n}\in\mathcal{L}_{\textbf{w},0}, fnw=n+1n2\left\|f_{n}\right\|_{\textbf{w}}=\displaystyle\frac{n+1}{n}\leq 2, and fn0f_{n}\to 0 pointwise. Therefore, by Proposition 2.7, the sequence {fn}\{f_{n}\} converges weakly to 0 in w,0\mathcal{L}_{\textbf{w},0}. Since compact operators are completely continuous [7], it follows that limnKfnw=0\displaystyle\lim\limits_{n\to\infty}\left\|Kf_{n}\right\|_{\textbf{w}}=0 for any compact operator KK on w,0\mathcal{L}_{\textbf{w},0}. Therefore, if KK is a compact operator on w,0\mathcal{L}_{\textbf{w},0}, then

MψKlim supn(MψK)fnwlim supnMψfnw.\|M_{\psi}-K\|\geq\limsup_{n\to\infty}\left\|(M_{\psi}-K)f_{n}\right\|_{\textbf{w}}\geq\limsup_{n\to\infty}\left\|M_{\psi}f_{n}\right\|_{\textbf{w}}.

Thus,

Mψe\displaystyle\|M_{\psi}\|_{e} \displaystyle\geq inf{MψK:K compact on w,0}\displaystyle\inf\{\|M_{\psi}-K\|:\,K\hbox{ compact on }\mathcal{L}_{\textbf{w},0}\}
\displaystyle\geq lim supnMψfnw\displaystyle\limsup_{n\to\infty}\left\|M_{\psi}f_{n}\right\|_{\textbf{w}}
=\displaystyle= lim supnsupvT|v||ψ(v)fn(v)ψ(v)fn(v)|\displaystyle\limsup_{n\to\infty}\sup_{v\in T^{*}}|v||\psi(v)f_{n}(v)-\psi(v^{-})f_{n}(v^{-})|
=\displaystyle= limnn+1nsup|v|n|ψ(v)|=A(ψ).\displaystyle\lim_{n\to\infty}\frac{n+1}{n}\sup_{|v|\geq n}|\psi(v)|=A(\psi).

Next we show that MψeB(ψ)\|M_{\psi}\|_{e}\geq B(\psi). The result is immediate if B(ψ)=0B(\psi)=0. So assume {vn}\{v_{n}\} is a sequence of vertices of length greater than 1 such that |vn||v_{n}| is increasing unboundedly and

limn|vn|log|vn|Dψ(vn)=B(ψ).\lim_{n\to\infty}|v_{n}|\log|v_{n}|D\psi(v_{n})=B(\psi).

Fix p(0,1)p\in(0,1) and for nn\in\mathbb{N}, let

hp,n(v)={(log(|v|+1))p+1(log|vn|)p if 0|v|<|vn|,log|vn| if |v||vn|.h_{p,n}(v)=\begin{cases}\displaystyle\frac{(\log(|v|+1))^{p+1}}{(\log|v_{n}|)^{p}}&\hbox{ if }\quad 0\leq|v|<|v_{n}|,\\ \log|v_{n}|&\hbox{ if }\quad|v|\geq|v_{n}|.\end{cases}

Then hp,n(o)=0h_{p,n}(o)=0, hp,n(vn)=hp,n(vn)=log|vn|h_{p,n}(v_{n})=h_{p,n}(v_{n}^{-})=\log|v_{n}|, and

|v|Dhp,n(v)={|v|log|vn|p[(log(|v|+1))p+1(log|v|)p+1] if 1|v|<|vn|, 0 if |v||vn|.|v|Dh_{p,n}(v)=\begin{cases}\displaystyle\frac{|v|}{\log|v_{n}|^{p}}\left[(\log(|v|+1))^{p+1}-(\log|v|)^{p+1}\right]&\hbox{ if }1\leq|v|<|v_{n}|,\\ \ \ 0&\hbox{ if }|v|\geq|v_{n}|.\end{cases}

The supremum of v|v|Dhn(v)v\mapsto|v|Dh_{n}(v) is attained at the vertices of length |vn|1|v_{n}|-1 and by a straightforward calculation it can be written as

sp,n=(|vn|1)(log|vn|log(|vn|1))[log(|vn|1)log|vn|1(log(|vn|1)log|vn|)p1log(|vn|1)log|vn|+1].s_{p,n}=(|v_{n}|-1)(\log|v_{n}|-\log(|v_{n}|-1))\left[\frac{\log(|v_{n}|-1)}{\log|v_{n}|}\frac{1-\left(\frac{\log(|v_{n}|-1)}{\log|v_{n}|}\right)^{p}}{1-\frac{\log(|v_{n}|-1)}{\log|v_{n}|}}+1\right].

Since the product of the first two factors approaches 1 as nn\to\infty and the function 1up1u\frac{1-u^{p}}{1-u} approaches pp as u1u\to 1, we see that sp,np+1s_{p,n}\to p+1 as nn\to\infty. In particular, hp,nw=sp,n\left\|h_{p,n}\right\|_{\textbf{w}}=s_{p,n} yields a bounded sequence. Letting gp,n=hp,nsp,ng_{p,n}=\displaystyle\frac{h_{p,n}}{s_{p,n}}, we see that gp,nw,0g_{p,n}\in\mathcal{L}_{\textbf{w},0}, gp,nw=1\left\|g_{p,n}\right\|_{\textbf{w}}=1, and gp,n0g_{p,n}\to 0 pointwise. Consequently, by Proposition 2.7, the sequence {gp,n}\{g_{p,n}\} converges to 0 weakly. This implies that Kgp,nw0\left\|Kg_{p,n}\right\|_{\textbf{w}}\to 0 as nn\to\infty for any compact operator KK on w,0\mathcal{L}_{\textbf{w},0}. We deduce that for any such operator KK

MψKlim supn(MψK)gp,nwlim supnψgp,nw.\|M_{\psi}-K\|\geq\limsup_{n\to\infty}\left\|(M_{\psi}-K)g_{p,n}\right\|_{\textbf{w}}\geq\limsup_{n\to\infty}\left\|\psi g_{p,n}\right\|_{\textbf{w}}.

Therefore

Mψe\displaystyle\|M_{\psi}\|_{e} \displaystyle\geq inf{MψK:K compact on w,0}\displaystyle\inf\{\|M_{\psi}-K\|:\,K\hbox{ compact on }\mathcal{L}_{\textbf{w},0}\} (8.1)
\displaystyle\geq lim supnsupvT|v|D(ψgp,n)(v).\displaystyle\limsup_{n\to\infty}\sup_{v\in T^{*}}|v|D(\psi g_{p,n})(v).

Next, observe that for each nn\in\mathbb{N}, gp,n(vn)=gp,n(vn)=log|vn|sp,ng_{p,n}(v_{n})=g_{p,n}(v_{n}^{-})=\displaystyle\frac{\log|v_{n}|}{s_{p,n}}, so

|vn|D(ψgp,n)(vn)\displaystyle|v_{n}|D(\psi g_{p,n})(v_{n}) =\displaystyle= |vn||ψ(vn)gp,n(vn)ψ(vn)gp,n(vn)|\displaystyle|v_{n}||\psi(v_{n})g_{p,n}(v_{n})-\psi(v_{n}^{-})g_{p,n}(v_{n}^{-})| (8.2)
=\displaystyle= |vn|Dψ(vn)log|vn|sp,n.\displaystyle|v_{n}|D\psi(v_{n})\frac{\log|v_{n}|}{s_{p,n}}.

Therefore, from (8.1) and (8.2), we obtain

Mψe1p+1limn|vn|log|vn|Dψ(vn)=1p+1B(ψ).\displaystyle\|M_{\psi}\|_{e}\geq\frac{1}{p+1}\lim_{n\to\infty}|v_{n}|\log|v_{n}|D\psi(v_{n})=\frac{1}{p+1}B(\psi).

Finally, letting pp approach 0, we deduce MψeB(ψ)\|M_{\psi}\|_{e}\geq B(\psi), completing the proof.∎

We now turn to the upper estimate.

Theorem 8.3.

If MψM_{\psi} is bounded on w\mathcal{L}_{\textbf{w}} (or equivalently, w,0\mathcal{L}_{\textbf{w},0}), then

MψeA(ψ)+B(ψ).\|M_{\psi}\|_{e}\leq A(\psi)+B(\psi).
Proof.

Fix nn\in\mathbb{N}, define the operator KnK_{n} on w\mathcal{L}_{\textbf{w}} by

Knf(v)={f(v) if |v|n,f(vn) if |v|>n,K_{n}f(v)=\begin{cases}f(v)&\hbox{ if }\quad|v|\leq n,\\ f(v_{n})&\hbox{ if }\quad|v|>n,\end{cases}

where fwf\in\mathcal{L}_{\textbf{w}} and vnv_{n} is the ancestor of vv of length nn. In particular, Knfw,0K_{n}f\in\mathcal{L}_{\textbf{w},0} and

Knf(o)=f(o).\displaystyle K_{n}f(o)=f(o). (8.3)

Furthermore, KnfK_{n}f attains finitely many values, whose number does not exceed the number of vertices in the closed ball centered at oo of radius nn. Observe that if {gk}\{g_{k}\} is a sequence in w\mathcal{L}_{\textbf{w}} with gkw1\left\|g_{k}\right\|_{\textbf{w}}\leq 1 for each kk\in\mathbb{N}, then, a=supk|gk(o)|1a=\displaystyle\sup_{k\in\mathbb{N}}|g_{k}(o)|\leq 1 so that |Kngk(o)|a|K_{n}g_{k}(o)|\leq a. Furthermore, as a consequence of Proposition 2.1, for each vTv\in T^{*}, and for each kk\in\mathbb{N}, we have |Kngk(v)|1+logn|K_{n}g_{k}(v)|\leq 1+\log n. Therefore, some subsequence {Kngkj}j\{K_{n}g_{k_{j}}\}_{j\in\mathbb{N}} must converge to a function gg on TT attaining constant values on the sectors determined by the vertices on the sphere centered at oo of radius nn. In particular, gwg\in\mathcal{L}_{\textbf{w}} and since KngkjgK_{n}g_{k_{j}}\to g uniformly on the closed ball centered at oo of radius nn, and DKngkjDK_{n}g_{k_{j}} and DgDg are 0 outside of the ball, we have

Kngkjgw\displaystyle\left\|K_{n}g_{k_{j}}-g\right\|_{\textbf{w}} =\displaystyle= |gkj(o)g(o)|+sup|v|n|v|D(gkjg)(v)\displaystyle|g_{k_{j}}(o)-g(o)|+\sup_{|v|\leq n}|v|D(g_{k_{j}}-g)(v)
\displaystyle\leq |gkj(o)g(o)|\displaystyle|g_{k_{j}}(o)-g(o)|
+nsup|v|n[|gkj(v)g(v)|+|gkj(v)g(v)|],\displaystyle+n\sup_{|v|\leq n}\left[|g_{k_{j}}(v)-g(v)|+\ |g_{k_{j}}(v^{-})-g(v^{-})|\right],

which converges to 0 as jj\to\infty. Thus, KnK_{n} is compact.

Observe that the operator MψKnM_{\psi}K_{n} is also compact. Furthermore, for vTv\in T^{*}, we have

|v|D[(IKn)f](v)|v|Df(v)fw.\displaystyle|v|D[(I-K_{n})f](v)\leq|v|Df(v)\leq\left\|f\right\|_{\textbf{w}}. (8.4)

On the other hand, by Proposition 2.1, we see that

|[(IKn)f](v)|(1+log|v|)fw.\displaystyle\left|[(I-K_{n})f](v)\right|\leq(1+\log|v|)\left\|f\right\|_{\textbf{w}}. (8.5)

We now use (8.5) and (8.4) to estimate ψ(IKn)fw\left\|\psi(I-K_{n})f\right\|_{\textbf{w}}:

ψ(IKn)fw\displaystyle\left\|\psi(I-K_{n})f\right\|_{\textbf{w}} =\displaystyle= sup|v|>n|v||ψ(v)[(IKn)f](v)ψ(v)[(IKn)f](v)|\displaystyle\sup_{|v|>n}|v|\left|\psi(v)[(I-K_{n})f](v)-\psi(v^{-})[(I-K_{n})f](v^{-})\right|
\displaystyle\leq sup|v|>n{|ψ(v)||v|D[(IKn)f](v)+|v|Dψ(v)|[(IKn)f](v)|}\displaystyle\sup_{|v|>n}\left\{|\psi(v^{-})||v|D[(I-K_{n})f](v)+|v|D\psi(v)|[(I-K_{n})f](v)|\right\}
\displaystyle\leq sup|v|>n|ψ(v)||v|D[(IKn)f](v)\displaystyle\sup_{|v|>n}|\psi(v^{-})||v|D[(I-K_{n})f](v)
+sup|v|>n|v|log|v|Dψ(v)|[(IKn)f](v)|1+log|v|1+log|v|log|v|\displaystyle+\ \sup_{|v|>n}|v|\log|v|D\psi(v)\frac{|[(I-K_{n})f](v)|}{1+\log|v|}\frac{1+\log|v|}{\log|v|}
\displaystyle\leq sup|v|>n|ψ(v)|fw+sup|v|>n|v|log|v|Dψ(v)fw(1+lognlogn).\displaystyle\sup_{|v|>n}|\psi(v^{-})|\left\|f\right\|_{\textbf{w}}+\sup_{|v|>n}|v|\log|v|D\psi(v)\left\|f\right\|_{\textbf{w}}\left(\frac{1+\log n}{\log n}\right).

Therefore, using this estimate and taking the limit as nn\to\infty, we obtain

Mψe\displaystyle\|M_{\psi}\|_{e} \displaystyle\leq lim supnMψMψKn\displaystyle\limsup_{n\to\infty}\|M_{\psi}-M_{\psi}K_{n}\|
=\displaystyle= lim supnsupfw=1ψ(IKn)fw\displaystyle\limsup_{n\to\infty}\sup_{\left\|f\right\|_{\textbf{w}}=1}\left\|\psi(I-K_{n})f\right\|_{\textbf{w}}
\displaystyle\leq A(ψ)+B(ψ),\displaystyle A(\psi)+B(\psi),

completing the proof.∎

9. Isometries and Isometric Zero Divisors

In this section we show that, in analogy to the case of the multiplication operators on the Lipschitz space of the tree [6], there are no nontrivial isometric multiplication operators.

Theorem 9.1.

The only isometric multiplication operators on w\mathcal{L}_{\textbf{w}} or w,0\mathcal{L}_{\textbf{w},0} are induced by the constant functions of modulus one.

Proof.

It is clear that the constant functions of modulus one are symbols of isometric multiplication operators on w\mathcal{L}_{\textbf{w}} and w,0\mathcal{L}_{\textbf{w},0}. Thus, assume MψM_{\psi} is an isometry on w\mathcal{L}_{\textbf{w}} or w,0\mathcal{L}_{\textbf{w},0} so that, in particular,

ψw=Mψ1w=1.\displaystyle\left\|\psi\right\|_{\textbf{w}}=\left\|M_{\psi}1\right\|_{\textbf{w}}=1. (9.1)

First we are going to show that ψ\psi has constant modulus 1. Fix vTv\in T^{*} and let fv=1|v|+1χvf_{v}=\displaystyle\frac{1}{|v|+1}\chi_{v}. Then fvw=1\left\|f_{v}\right\|_{\textbf{w}}=1 and so 1=ψfvw=|ψ(v)|1=\left\|\psi f_{v}\right\|_{\textbf{w}}=|\psi(v)|. On the other hand, for g=12χog=\frac{1}{2}\chi_{o}, gw=2|g(o)|=1\left\|g\right\|_{\textbf{w}}=2|g(o)|=1 and thus

1=ψgw=2|ψ(o)g(o)|=|ψ(o)|.1=\left\|\psi g\right\|_{\textbf{w}}=2|\psi(o)g(o)|=|\psi(o)|.

Hence, |ψ(v)|=1|\psi(v)|=1 for all vTv\in T. From (9.1) it follows that DψD\psi must be identically 0. Therefore, ψ\psi is a constant function of modulus one.∎

Inspired by [1], we now define the notion of isometric zero divisor in a tree setting.

Definition 9.2.

Let XX be a Banach space of functions defined on a tree TT and let ZZ be a nonempty subset of TT. A function ψX\psi\in X is called a zero divisor for ZZ if it vanishes precisely at the vertices in ZZ and g/ψXg/\psi\in X for every gXg\in X vanishing on ZZ. The function ψ\psi is said to be an isometric zero divisor if g/ψ=g\|g/\psi\|=\|g\| for each gXg\in X which vanishes on ZZ.

Recalling the set 𝒫\mathcal{P} in Proposition 2.8, we now see that, under certain hypotheses on the space XX, the isometric zero divisors induce isometric multiplication operators on XX.

Theorem 9.3.

Let XX be a functional Banach space on TT containing 𝒫\mathcal{P} and satisfying the following properties:

  1. (a)

    𝒫\mathcal{P} is dense in XX.

  2. (b)

    For each vTv\in T and each fXf\in X, pvfXp_{v}f\in X.

If ψX\psi\in X is an isometric zero divisor, then MψM_{\psi} is an isometry on XX.

Proof.

Let ψ\psi be an isometric zero divisor with zero set ZZ. Then, for each p𝒫p\in\mathcal{P}, pψXp\psi\in X and vanishes at the vertices in ZZ, so

p=pψψ=pψ.\displaystyle\|p\|=\left\|\frac{p\psi}{\psi}\right\|=\|p\psi\|. (9.2)

We wish to show that ψfX\psi f\in X and ψf=f\|\psi f\|=\|f\| for each fXf\in X. Fix fXf\in X and using the density of 𝒫\mathcal{P}, let {pn}\{p_{n}\} be a sequence in 𝒫\mathcal{P} such that pnf0\|p_{n}-f\|\to 0 as nn\to\infty. Since 𝒫\mathcal{P} is closed under addition, using (9.2), for n,mn,m\in\mathbb{N} we have pnψpmψ=pnpm\|p_{n}\psi-p_{m}\psi\|=\|p_{n}-p_{m}\|, so {pnψ}\{p_{n}\psi\} is a Cauchy sequence in XX. By the completeness of XX, there exists gXg\in X such that pnψg0\|p_{n}\psi-g\|\to 0 as nn\to\infty. Since XX is a functional Banach space, the point evaluation functionals are bounded, hence pnψgp_{n}\psi\to g and pnfp_{n}\to f pointwise in TT. Therefore g=ψfg=\psi f, proving that ψfX\psi f\in X. Moreover, by the triangle inequality and (9.2), we have

|ψff|\displaystyle|\|\psi f\|-\|f\|| \displaystyle\leq |ψfpnψ|+|pnψpn|+|pnf|\displaystyle|\|\psi f\|-\|p_{n}\psi\||+|\|p_{n}\psi\|-\|p_{n}\||+|\|p_{n}\|-\|f\||
\displaystyle\leq ψfpnψ+pnf0,\displaystyle\|\psi f-p_{n}\psi\|+\|p_{n}-f\|\to 0,

as nn\to\infty. Therefore, ψf=f\|\psi f\|=\|f\|, as desired.∎

We now turn our attention to the existence of isometric zero divisors on the spaces w\mathcal{L}_{\textbf{w}} and w,0\mathcal{L}_{\textbf{w},0}.

Corollary 9.4.

The space w,0\mathcal{L}_{\textbf{w},0} has no isometric zero divisors.

Proof.

By Theorem 9.1, the only isometric multiplication operators are the constants of modulus one, which do not vanish anywhere. Observe that for each vTv\in T, recalling that pvp_{v} is the characteristic function of the set consisting of vv and all its descendants, the set 𝒫\mathcal{P} is closed under multiplication by pvp_{v}. Thus, by Proposition 2.8, for each vTv\in T and each fw,0f\in\mathcal{L}_{\textbf{w},0}, the function pvfw,0p_{v}f\in\mathcal{L}_{\textbf{w},0}. Therefore, since w,0\mathcal{L}_{\textbf{w},0} is a functional Banach space, the space w,0\mathcal{L}_{\textbf{w},0} satisfies the hypotheses of Theorem 9.3 and thus, no isometric zero divisors can exist.∎

Theorem 9.5.

The space w\mathcal{L}_{\textbf{w}} has no isometric zero-divisors.

Proof.

By Corollary 9.4, it suffices to show that the isometric zero divisors of w\mathcal{L}_{\textbf{w}} are in w,0\mathcal{L}_{\textbf{w},0}.

Assume ψ\psi is an isometric zero divisor of w\mathcal{L}_{\textbf{w}}. We begin by showing that ψ\psi is bounded. Fix wTw\in T^{*} and define fw=1|w|pwf_{w}=\displaystyle\frac{1}{|w|}p_{w}. Then, fw𝒫f_{w}\in\mathcal{P} and for vTv\in T^{*}, we have

|v||fw(v)fw(v)|=|v||w|χw(v)=χw(v),|v||f_{w}(v)-f_{w}(v^{-})|=\frac{|v|}{|w|}\chi_{w}(v)=\chi_{w}(v),

so fww=1\left\|f_{w}\right\|_{\textbf{w}}=1. Therefore, ψfww=1\left\|\psi f_{w}\right\|_{\textbf{w}}=1. Letting DwD_{w} be the set of descendants of ww, we have

ψfww=max{|ψ(w)|,supvDw|v||w||ψ(v)ψ(v)|}|ψ(w)|.\left\|\psi f_{w}\right\|_{\textbf{w}}=\max\left\{|\psi(w)|,\sup_{v\in D_{w}}\frac{|v|}{|w|}|\psi(v)-\psi(v^{-})|\right\}\geq|\psi(w)|.

Hence, |ψ(w)|1|\psi(w)|\leq 1, proving the boundedness of ψ\psi.

For |w|2|w|\geq 2, let us define

gw(v)={0 if v=o,1|w|log|v| if 1|v|<|w|,1|w|log|w| if |v||w|.g_{w}(v)=\begin{cases}0&\quad\hbox{ if }v=o,\\ \frac{1}{|w|}\log|v|&\quad\hbox{ if }1\leq|v|<|w|,\\ \frac{1}{|w|}\log|w|&\quad\hbox{ if }|v|\geq|w|.\end{cases}

Since gwg_{w} has finite support and by (2.6), the function χv\chi_{v} is in 𝒫\mathcal{P}, we deduce that gw𝒫g_{w}\in\mathcal{P}, so by (9.2), we obtain

ψgww=gww=sup2|v||w||v||w|(log|v|log(|v|1))2log2|w|.\displaystyle\ \ \ \ \ \ \left\|\psi g_{w}\right\|_{\textbf{w}}=\left\|g_{w}\right\|_{\textbf{w}}=\sup_{2\leq|v|\leq|w|}\frac{|v|}{|w|}\left(\log|v|-\log(|v|-1)\right)\leq\frac{2\log 2}{|w|}. (9.3)

On the other hand,

ψgww\displaystyle\left\|\psi g_{w}\right\|_{\textbf{w}} \displaystyle\geq sup2|v||w||v||w||(ψ(v)ψ(v))log|v|+ψ(v)log|v||v|1|\displaystyle\sup_{2\leq|v|\leq|w|}\frac{|v|}{|w|}\left|(\psi(v)-\psi(v^{-}))\log|v|+\psi(v^{-})\log\frac{|v|}{|v|-1}\right|
\displaystyle\geq sup2|v||w||v||w||ψ(v)ψ(v)|log|v|sup2|v||w||v||w||ψ(v)|log|v||v|1\displaystyle\sup_{2\leq|v|\leq|w|}\frac{|v|}{|w|}|\psi(v)-\psi(v^{-})|\log|v|-\sup_{2\leq|v|\leq|w|}\frac{|v|}{|w|}|\psi(v^{-})|\log\frac{|v|}{|v|-1}
\displaystyle\geq sup2|v||w||v||w||ψ(v)ψ(v)|log|v|2log2|w|ψ.\displaystyle\sup_{2\leq|v|\leq|w|}\frac{|v|}{|w|}|\psi(v)-\psi(v^{-})|\log|v|-\frac{2\log 2}{|w|}\|\psi\|_{\infty}.

Therefore, using (9.3), we get

sup2|v||w||v||w||ψ(v)ψ(v)|log|v|2log2(1+ψ)|w|.\sup_{2\leq|v|\leq|w|}\frac{|v|}{|w|}|\psi(v)-\psi(v^{-})|\log|v|\leq\frac{2\log 2(1+\|\psi\|_{\infty})}{|w|}.

Multiplying both sides by |w||w| and letting |w||w|\to\infty, we obtain

sup|v|2|v||ψ(v)ψ(v)|log|v|<.\sup_{|v|\geq 2}|v||\psi(v)-\psi(v^{-})|\log|v|<\infty.

Hence lim|v||v||ψ(v)ψ(v)|=0\lim\limits_{|v|\to\infty}|v||\psi(v)-\psi(v^{-})|=0, proving that ψw,0\psi\in\mathcal{L}_{\textbf{w},0}.∎

Acknowledgements

The research of the first author is supported by a grant from the College of Science and Health of the University of Wisconsin-La Crosse.

References

  • [1] Alexandru Aleman, Peter Duren, Maria J. Martin, and Dragan Vukotić, Multiplicative isometries and isometric zero-divisors, Canad. J. Math. 62 (2010), no. 5, 961–974. MR 2730350
  • [2] Pierre Cartier, Fonctions harmoniques sur un arbre, Symposia Mathematica, Vol. IX (Convegno di Calcolo delle Probabilità, INDAM, Rome, 1971), 1972, pp. 203–270. MR 0353467
  • [3] Joel M. Cohen and Flavia Colonna, The Bloch space of a homogeneous tree, vol. 37, 1992, Papers in honor of José Adem (Spanish), pp. 63–82. MR 1317563
  • [4] by same author, Embeddings of trees in the hyperbolic disk, Complex Variables Theory Appl. 24 (1994), no. 3-4, 311–335. MR 1270321
  • [5] Flavia Colonna, Bloch and normal functions and their relation, Rend. Circ. Mat. Palermo (2) 38 (1989), no. 2, 161–180. MR 1029707
  • [6] Flavia Colonna and Glenn R. Easley, Multiplication operators on the Lipschitz space of a tree, Integral Equations Operator Theory 68 (2010), no. 3, 391–411. MR 2735443
  • [7] John B. Conway, A course in functional analysis, second ed., Graduate Texts in Mathematics, vol. 96, Springer-Verlag, New York, 1990. MR 1070713
  • [8] Fausto Di Biase and Massimo A. Picardello, The Green formula and HpH^{p} spaces on trees, Math. Z. 218 (1995), no. 2, 253–272. MR 1318159
  • [9] Peter L. Duren, Bernhard W. Romberg, and Allen L. Shields, Linear functionals on HpH^{p} spaces with 0<p<10<p<1, J. Reine Angew. Math. 238 (1969), 32–60. MR 259579
  • [10] Adam Korányi, Massimo A. Picardello, and Mitchell H. Taibleson, Hardy spaces on nonhomogeneous trees, Symposia Mathematica, Vol. XXIX (Cortona, 1984), Sympos. Math., XXIX, Academic Press, New York, 1987, With an appendix by Picardello and Wolfgang Woess, pp. 205–265. MR 951187
  • [11] Marco Pavone, Chaotic composition operators on trees, Houston J. Math. 18 (1992), no. 1, 47–56. MR 1159439
  • [12] by same author, Toeplitz operators on discrete groups, J. Operator Theory 27 (1992), no. 2, 359–384. MR 1249652
  • [13] by same author, Partially ordered groups, almost invariant sets, and Toeplitz operators, J. Funct. Anal. 113 (1993), no. 1, 1–18. MR 1214895
  • [14] Vladimir S. Rabinovich and Steffen Roch, Essential spectra of difference operators on n\mathbb{Z}^{n}-periodic graphs, J. Phys. A 40 (2007), no. 33, 10109–10128. MR 2371282
  • [15] by same author, Fredholm properties of band-dominated operators on periodic discrete structures, Complex Anal. Oper. Theory 2 (2008), no. 4, 637–668. MR 2465542
  • [16] by same author, Finite sections of band-dominated operators on discrete groups, Recent progress in operator theory and its applications, Oper. Theory Adv. Appl., vol. 220, Birkhäuser/Springer Basel AG, Basel, 2012, pp. 239–253. MR 2953882
  • [17] John Roe, Band-dominated Fredholm operators on discrete groups, Integral Equations Operator Theory 51 (2005), no. 3, 411–416. MR 2126819
  • [18] Shanli Ye, Multipliers and cyclic vectors on the weighted Bloch space, Math. J. Okayama Univ. 48 (2006), 135–143. MR 2291174
  • [19] Kehe Zhu, Operator theory in function spaces, second ed., Mathematical Surveys and Monographs, vol. 138, American Mathematical Society, Providence, RI, 2007. MR 2311536