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

The Hardy-Weyl algebra

Jim Agler and John E. McCarthy Partially supported by National Science Foundation Grant DMS 2054199
Abstract

We study the algebra π’œ{\mathcal{A}} generated by the Hardy operator HH and the operator MxM_{x} of multiplication by xx on L2​[0,1]L^{2}[0,1]. We call π’œ{\mathcal{A}} the Hardy-Weyl algebra. We show that its quotient by the compact operators is isomorphic to the algebra of functions that are continuous on Ξ›\Lambda and analytic on the interior of Ξ›\Lambda for a planar set Ξ›\Lambda = [βˆ’1,0]βˆͺ𝔻​(1,1)Β―[-1,0]\cup\overline{\mathbb{D}(1,1)}, which we call the lollipop. We find a Toeplitz-like short exact sequence for the Cβˆ—C^{*}-algebra generated by π’œ{\mathcal{A}}.

We study the operator Z=Hβˆ’MxZ=H-M_{x}, show that its point spectrum is (βˆ’1,0]βˆͺ𝔻​(1,1)(-1,0]\cup\mathbb{D}(1,1), and that the eigenvalues grow in multiplicity as the points move to 0 from the left.

1 Introduction

The classical Weyl algebra is generated by the operators of multiplication by xx, denoted MxM_{x}, and differentiation, denoted by DD. These operators satisfy the commutator relation

D​Mxβˆ’Mx​D= 1.DM_{x}-M_{x}D\ =\ 1. (1.1)

The algebra generated by these relations has been studied extensively in both algebra and operator theoryβ€”see e.g. the books [14, 12, 19, 9]. In operator theory, the study is complicated by the fact that no bounded operators satisfy (1.1). In this note, we shall study the associated algebras that arise when one replaces the differentiation operator DD by a bounded integration operator.

The Hardy operator HH is defined on L2​[0,1]L^{2}[0,1] by

H​f​(x)=1xβ€‹βˆ«0xf​(t)​𝑑t.Hf(x)\ =\ \frac{1}{x}\int_{0}^{x}f(t)dt.

Let V=Mx​HV=M_{x}H denote the Volterra operator

V:fβ†¦βˆ«0xf​(t)​𝑑t,V:f\mapsto\int_{0}^{x}f(t)dt,

which is a right inverse of DD. These operators give rise to a new set of relations instead of (1.1), namely

V​Mxβˆ’Mx​V\displaystyle VM_{x}-M_{x}V =\displaystyle\ =\ βˆ’V2\displaystyle-V^{2}
H​Mxβˆ’Mx​H\displaystyle HM_{x}-M_{x}H =\displaystyle= βˆ’H​Mx​H=βˆ’H​V.\displaystyle-HM_{x}H\ =\ -HV.

We shall use L2L^{2} to mean L2​[0,1]L^{2}[0,1] throughout.

Definition 1.2.

Let π’œ{\mathcal{A}} denote the closure of the unital algebra generated by HH and MxM_{x} in the norm topology of B​(L2)B(L^{2}). We call π’œ{\mathcal{A}} the Hardy-Weyl algebra.

When is an operator in π’œ{\mathcal{A}}? Can all the elements be described? Since the commutator of HH and MxM_{x} is compact, we first describe the quotient of π’œ{\mathcal{A}} by the compact operators. Define Ξ›\Lambda, a subset of the plane, by

Ξ›=[βˆ’1,0]βˆͺ𝔻​(1,1)Β―.\Lambda\ =\ [-1,0]\cup\overline{\mathbb{D}(1,1)}.

We call Ξ›\Lambda the lollipop. The lollipop algebra A​(Ξ›)A(\Lambda) is the Banach algebra of functions that are continuous on Ξ›\Lambda and analytic on int​(Ξ›){\rm int}(\Lambda), equipped with the maximum modulus norm. Let 𝒦{\mathcal{K}} denote the compact operators on L2L^{2}, and let π’¦π’œ=π’¦βˆ©π’œ{\mathcal{K}}_{{\mathcal{A}}}={\mathcal{K}}\cap{\mathcal{A}},

Theorem 1.3.

There is a Banach algebra isomorphism Ξ³\gamma from A​(Ξ›)A(\Lambda) onto π’œ/π’¦π’œ{\mathcal{A}}/{\mathcal{K}}_{{\mathcal{A}}}. It is given by

γ​(f)=f|[βˆ’1,0]​(βˆ’Mx)+f|𝔻​(1,1)¯​(H)βˆ’f​(0).\gamma(f)\ =\ f|_{[-1,0]}(-M_{x})+f|_{\overline{\mathbb{D}(1,1)}}(H)-f(0).

Let ΞΈ:π’œβ†’A​(Ξ›)\theta:{\mathcal{A}}\to A(\Lambda) be defined by θ​(T)=Ξ³βˆ’1​([T])\theta(T)=\gamma^{-1}([T]), where [T][T] is the projection of TT onto π’œ/π’¦π’œ{\mathcal{A}}/{\mathcal{K}}_{{\mathcal{A}}}. As a corollary to Theorem 1.3 we obtain that every element TT of π’œ{\mathcal{A}} can be written uniquely as

T=MΟ•+g​(H)+K,T\ =\ M_{\phi}+g(H)+K, (1.4)

where Ο•βˆˆC​([0,1])\phi\in C([0,1]), gg is in A​(𝔻​(1,1)Β―)A(\overline{\mathbb{D}(1,1)}), ϕ​(0)=g​(0)\phi(0)=g(0), and Kβˆˆπ’¦π’œK\in{\mathcal{K}}_{{\mathcal{A}}}.

If we look at Cβˆ—β€‹(π’œ)C^{*}({\mathcal{A}}), the Cβˆ—C^{*}-algebra generated by π’œ{\mathcal{A}}, we get something similar to the Toeplitz algebra short exact sequence. See [4, 8, 18] for some recent results on the Toeplitz algebra, and [3, 10, 15, 16] for some applications. The lollipop algebra is replaced by the functions that are continuous on the boundary.

Theorem 1.5.

There is a short exact sequence of Cβˆ—C^{*}-algebras

0→𝒦→Cβˆ—β€‹(π’œ)β†’C​(βˆ‚Ξ›)β†’0.0\to{\mathcal{K}}\to C^{*}({\mathcal{A}})\to C(\partial\Lambda)\to 0.

In addition to studying the algebra generated by HH and MxM_{x}, one may also study the smaller algebra generated by VV and MxM_{x}.

Definition 1.6.

The algebra π’œ0{\mathcal{A}}_{0} is the norm-closed unital algebra generated by VV and MxM_{x}.

One can see that π’œ0{\mathcal{A}}_{0} is a proper subalgebra of π’œ{\mathcal{A}} by noting that its quotient by the compact operators is isomorphic to C​([0,1])C([0,1]).

The lattice of closed invariant subspaces of the Volterra operator VV was shown independently by Brodskii and Donoghue [5, 6] to be

Lat​(V)={{f∈L2:f=0​on​[0,s]}:s∈[0,1]}.{\rm Lat}(V)\ =\ \Big{\{}\{f\in L^{2}:f=0{\rm\ on\ }[0,s]\}:s\in[0,1]\Big{\}}.

We let AlgLat​(V){\rm AlgLat}(V) denote the set of bounded operators on L2L^{2} that leave invariant every element of LatV{\rm LatV}. This is a large algebraβ€”it includes all right translation operators for example. It is described by the following theorem of Radjavi and Rosenthal [17, Example 9.26].

Theorem 1.7.

The algebra AlgLat​(V){\rm AlgLat}(V) is the weak operator topology closure of the algebra generated by VV and MxM_{x}.

It follows from Theorem 1.7 that HH is in the WOT closure of π’œ0{\mathcal{A}}_{0}, and hence π’œ{\mathcal{A}} and π’œ0{\mathcal{A}}_{0} have the same WOT closure. However, no extra compact operators are added in the WOT closure.

Theorem 1.8.
AlgLat​(V)βˆ©π’¦βŠ‚π’œ0.{\rm AlgLat}(V)\cap{\mathcal{K}}\ \subset\ {\mathcal{A}}_{0}.

In Section 6 we consider the operator Z=Hβˆ’MxZ=H-M_{x}. It follows from Theorem 1.3 that [Z][Z] generates π’œ/π’¦π’œ{\mathcal{A}}/{\mathcal{K}}_{{\mathcal{A}}}. We show ZZ has a surprisingly rich collection of eigenvectors.

Theorem 1.9.

Let Z=Hβˆ’MxZ=H-M_{x}. Then

Οƒp​(Z)=(βˆ’1,0]βˆͺ𝔻​(1,1).\sigma_{p}(Z)\ =\ (-1,0]\cup\mathbb{D}(1,1).

The algebraic multiplicity of the eigenvalues of ZZ on the stick (βˆ’1,0](-1,0] increases as Ξ»β†’0βˆ’\lambda\to 0^{-}, and hence operators XX in the closed algebra generated by ZZ have the property that θ​(X)\theta(X) is not just in A​(Ξ›)A(\Lambda), but is smoother. We prove:

Theorem 1.10.

Let XX be in the norm-closed algebra generated by ZZ. Then θ​(X)\theta(X) is CmC^{m} on (βˆ’22​m+1,0)(-\frac{2}{2m+1},0).

2 Preliminaries

Definition 2.1.

A monomial operator is a bounded linear operator T:L2​[0,1]β†’L2​[0,1]T:L^{2}[0,1]\to L^{2}[0,1] with the property that there exist constants cnc_{n} and pnp_{n} so that

T:xn↦cn​xpnβˆ€nβˆˆβ„•.T:x^{n}\mapsto c_{n}\,x^{p_{n}}\qquad\forall n\in{\mathbb{N}}. (2.2)

We call it a flat monomial operator if there exists some Ο„\tau so that pn=n+Ο„p_{n}=n+\tau for all nn.

In [2] we showed that every flat monomial operator is in AlgLat​(V){\rm AlgLat}(V), and hence by Theorem 1.7 in the weak closure of π’œ0{\mathcal{A}}_{0}. The Volterra operator VV is Hilbert-Schmidt, and hence compact. See e.g. [11] for a proof.

Let H2H^{2} denote the Hardy space of the unit disk, and kw​(z)=11βˆ’w¯​zk_{w}(z)=\frac{1}{1-\bar{w}z} the SzegΕ‘ kernel. We shall let S:f​(z)↦z​f​(z)S:f(z)\mapsto zf(z) denote the unilateral shift on H2H^{2}. Let

β​(z)=12βˆ’z,\beta(z)\ =\ \frac{1}{2-z},

and let CΞ²:f↦f∘βC_{\beta}:f\mapsto f\circ\beta denote the composition operator of composing with Ξ²\beta.

There is a unitary U:L2β†’H2U:L^{2}\to H^{2} that is defined on monomials by

U:xs↦1s+1​ksΒ―sΒ―+1,U:x^{s}\mapsto\frac{1}{s+1}k_{\frac{\bar{s}}{\bar{s}+1}}, (2.3)

and extended by linearity and continuity to the whole space. If T∈B​(L2)T\in B(L^{2}), we shall let T^\widehat{T} denote U​T​Uβˆ—UTU^{*}. It is easy to see that UU is unitary, as it preserves inner products. In [1] we prove that UU is given by the formula

U​f​(z)=11βˆ’zβ€‹βˆ«01f​(x)​xz1βˆ’z​𝑑x,Uf(z)\ =\ \frac{1}{1-z}\int_{0}^{1}f(x)x^{\frac{z}{1-z}}dx, (2.4)

and show that

Mx^\displaystyle\widehat{M_{x}} =\displaystyle\ =\ Sβˆ—β€‹CΞ²βˆ—\displaystyle S^{*}C_{\beta}^{*}
V^\displaystyle\widehat{V} =\displaystyle\ =\ (1βˆ’Sβˆ—)​CΞ²βˆ—\displaystyle(1-S^{*})C_{\beta}^{*}
H^\displaystyle\widehat{H} =\displaystyle= 1βˆ’Sβˆ—.\displaystyle 1-S^{*}.

If XX is a compact subset of β„‚\mathbb{C}, we shall let C​(X)C(X) denote the Banach algebra of functions that are continuous on XX, with the maximum modulus norm. We shall let A​(X)A(X) denote the subalgebra of functions that are continuous on XX and analytic on the interior of XX, and P​(X)P(X) denote the closure of the polynomials in C​(X)C(X). A theorem of Mergelyan [13] says that if the complement of XX is connected, then A​(X)=P​(X)A(X)=P(X).

3 The Calkin Hardy-Weyl Algebra

We let 𝒦\mathcal{K} denote the ideal of compact operators acting on L2L^{2} and set

𝒦0=π’œβˆ©π’¦.\mathcal{K}_{0}={\mathcal{A}}\cap\mathcal{K}.

Evidently, 𝒦0\mathcal{K}_{0} is a 2-sided ideal in π’œ{\mathcal{A}}. Consequently, we may define an algebra π’ž\mathcal{C}, the Calkin Hardy-Weyl algebra, by

π’ž=π’œ/𝒦0\mathcal{C}={\mathcal{A}}/\mathcal{K}_{0}

If Tβˆˆπ’œT\in{\mathcal{A}} we let [T][T] denote the coset of TT in π’ž\mathcal{C}, i.e.,

[T]={T+K|Kβˆˆπ’¦0}.[T]=\{T+K\,|\,K\in\mathcal{K}_{0}\}.
Proposition 3.1.

π’ž\mathcal{C} is an abelian Banach algebra.

Proof.

That π’ž\mathcal{C} is a Banach algebra follows from the fact that 𝒦0\mathcal{K}_{0} is closed in π’œ{\mathcal{A}}. To see that π’ž\mathcal{C} is abelian, observe that as MxM_{x} and HH generate π’œ{\mathcal{A}}, [Mx][M_{x}] and [H][H] generate π’ž\mathcal{C}. Furthermore, as Mx​H=Vβˆˆπ’¦0M_{x}H=V\in\mathcal{K}_{0},

[Mx]​[H]=0.[M_{x}][H]=0. (3.2)

Likewise, as H​Mx=(1βˆ’H)​Vβˆˆπ’¦0HM_{x}=(1-H)V\in\mathcal{K}_{0},

[H]​[Mx]=0,[H][M_{x}]=0, (3.3)

so that in particular we have that

[Mx]​[H]=[H]​[Mx].[M_{x}][H]=[H][M_{x}].

As [Mx][M_{x}] and [H][H] commute and generate π’ž\mathcal{C}, π’ž\mathcal{C} is abelian. ∎

3.1 A Uniform Algebra Homeomorphically Isomorphic to π’ž\mathcal{C}

We begin by defining an algebra by gluing together two simpler algebras whose maximal ideal spaces overlap at a single point. Let

𝒫={f=(fβˆ’,f+):fβˆ’βˆˆC​([βˆ’1,0]),f+∈A​(𝔻​(1,1)Β―)​ and ​fβˆ’β€‹(0)=f+​(0)}{\mathcal{P}}\ =\ \{\ f=(f_{-},f_{+})\ :\ f_{-}\in C([-1,0]),\ f_{+}\in A(\overline{\mathbb{D}(1,1)})\text{ and }f_{-}(0)=f_{+}(0)\ \}

where we view 𝒫{\mathcal{P}} as an algebra with the operations

c​f=(c​fβˆ’,c​f+),f+g=(fβˆ’+gβˆ’,f++g+),and ​f​g=(fβˆ’β€‹gβˆ’,f+​g+),cf=(cf_{-},cf_{+}),\ f+g=(f_{-}+g_{-},f_{+}+g_{+}),\text{and }fg=(f_{-}g_{-},f_{+}g_{+}),

and the norm

β€–fβ€–=max⁑{maxt∈[βˆ’1,0]⁑|fβˆ’β€‹(t)|,maxz∈(𝔻​(1,1)¯⁑|f+​(z)|}.\|f\|=\max\big{\{}\max_{t\in[-1,0]}|f_{-}(t)|,\max_{z\in(\overline{\mathbb{D}(1,1)}}|f_{+}(z)|\ \big{\}}.

We abuse notation by letting

f​(0)=fβˆ’β€‹(0)f(0)=f_{-}(0)

when fβˆˆπ’«f\in{\mathcal{P}}.

We note that if f∈C​([βˆ’1,0])f\in C([-1,0]), then as βˆ’Mx-M_{x} is self-adjoint and has spectrum equal to [βˆ’1,0][-1,0], we may form the operator f​(βˆ’Mx)f(-M_{x}). Likewise, as HH is cosubnormal and has spectrum equal to 𝔻​(1,1)Β―\overline{\mathbb{D}(1,1)}, if g∈A​(𝔻​(1,1)Β―)g\in A(\overline{\mathbb{D}(1,1)}), then we may form the operator g​(H)g(H). Concretely,

f​(βˆ’Mx)=Mf​(βˆ’x)​ and ​g​(H)^=Mhβˆ—,f(-M_{x})=M_{f(-x)}\ \text{ and }\ \widehat{g(H)}=M_{h}^{*},

where h​(z)=g​(1βˆ’zΒ―)Β―h(z)=\overline{g(1-\bar{z})}, and MhM_{h} denotes multiplication by hh.

Lemma 3.4.

If f∈C​([βˆ’1,0])f\in C([-1,0]) and g∈A​(𝔻​(1,1)Β―)g\in A(\overline{\mathbb{D}(1,1)}), then

[f​(βˆ’Mx)]​[g​(H)]=g​(0)​[f​(βˆ’Mx)]+f​(0)​[g​(H)]βˆ’f​(0)​g​(0).[f(-M_{x})][g(H)]=g(0)[f(-M_{x})]+f(0)[g(H)]-f(0)g(0).
Proof.

Since [βˆ’1,0][-1,0] is a spectral set for βˆ’Mx-M_{x} and 𝔻​(1,1)Β―\overline{\mathbb{D}(1,1)} is a spectral set for HH it suffices to prove the lemma in the special case when ff and gg are polynomials. Let f​(x)=f​(0)+x​f1​(x)f(x)=f(0)+xf_{1}(x) and g​(x)=g​(0)+x​g1​(x)g(x)=g(0)+xg_{1}(x). Using (3.2) and (3.3) we see that

[f​(βˆ’Mx)]​[g​(H)]\displaystyle[f(-M_{x})][g(H)] =([f​(0)]+[βˆ’Mx]​[f1​(βˆ’Mx)])​([g​(0)]+[H]​[g1​(H)])\displaystyle=\big{(}[f(0)]+[-M_{x}][f_{1}(-M_{x})]\big{)}\big{(}[g(0)]+[H][g_{1}(H)]\big{)}
=f​(0)​g​(0)+g​(0)​[βˆ’Mx]​[f1​(βˆ’Mx)]+f​(0)​[H]​[g1​(H)]\displaystyle=f(0)g(0)+g(0)[-M_{x}][f_{1}(-M_{x})]+f(0)[H][g_{1}(H)]
=f​(0)​g​(0)+g​(0)​[f​(βˆ’Mx)βˆ’f​(0)]+f​(0)​[g​(H)βˆ’g​(0)]\displaystyle=f(0)g(0)+g(0)[f(-M_{x})-f(0)]+f(0)[g(H)-g(0)]
=g​(0)​[f​(βˆ’Mx)]+f​(0)​[g​(H)]βˆ’f​(0)​g​(0).\displaystyle=g(0)[f(-M_{x})]+f(0)[g(H)]-f(0)g(0).

∎

If fβˆˆπ’«f\in{\mathcal{P}} we define γ​(f)βˆˆπ’œ\gamma(f)\in{\mathcal{A}} by the formula

γ​(f)=fβˆ’β€‹(βˆ’Mx)+f+​(H)βˆ’f​(0).\gamma(f)=f_{-}(-M_{x})+f_{+}(H)-f(0).

We also define Ξ“:π’«β†’π’ž\Gamma:{\mathcal{P}}\to\mathcal{C} by the formula

Γ​(f)=[γ​(f)]\Gamma(f)=[\gamma(f)]
Proposition 3.5.

Ξ“\Gamma is a continuous unital homomorphism.

Proof.

Ξ³\gamma is linear and γ​(1)=1\gamma(1)=1. Therefore, Ξ“\Gamma is linear and Γ​(1)=1\Gamma(1)=1. Also,

‖Γ​(f)β€–\displaystyle\|\Gamma(f)\| =β€–[γ​(f)]β€–\displaystyle=\|[\gamma(f)]\|
≀‖γ​(f)β€–\displaystyle\leq\|\gamma(f)\|
=β€–fβˆ’β€‹(βˆ’Mx)+f+​(H)βˆ’f​(0)β€–\displaystyle=\|f_{-}(-M_{x})+f_{+}(H)-f(0)\|
≀‖fβˆ’β€‹(βˆ’Mx)β€–+β€–f+​(H)β€–+|f​(0)|\displaystyle\leq\|f_{-}(-M_{x})\|+\|f_{+}(H)\|+|f(0)|
=maxt∈[βˆ’1,0]⁑|fβˆ’β€‹(t)|+maxzβˆˆπ”»β€‹(1,1)¯⁑|f+​(z)|+|f​(0)|\displaystyle=\max_{t\in[-1,0]}|f_{-}(t)|+\max_{z\in\overline{\mathbb{D}(1,1)}}|f_{+}(z)|+|f(0)|
≀3​‖fβ€–,\displaystyle\leq 3\|f\|,

so Ξ“\Gamma is continuous.

Finally, to see that Ξ“\Gamma preserves products, fix f,gβˆˆπ’«f,g\in{\mathcal{P}}.

Γ​(f)​Γ​(g)\displaystyle\Gamma(f)\Gamma(g) =[γ​(f)]​[γ​(g)]\displaystyle=[\gamma(f)]\ [\gamma(g)]
=[fβˆ’β€‹(βˆ’Mx)+f+​(H)βˆ’f​(0)]​[gβˆ’β€‹(βˆ’Mx)+g+​(H)βˆ’g​(0)]\displaystyle=[f_{-}(-M_{x})+f_{+}(H)-f(0)]\ [g_{-}(-M_{x})+g_{+}(H)-g(0)]
=([fβˆ’β€‹(βˆ’Mx)]​[gβˆ’β€‹(βˆ’Mx)]+[f+​(H)]​[g+​(H)])\displaystyle=\Big{(}[f_{-}(-M_{x})][g_{-}(-M_{x})]+[f_{+}(H)][g_{+}(H)]\Big{)}
+([fβˆ’β€‹(βˆ’Mx)]​[g+​(H)]+[gβˆ’β€‹(βˆ’Mx)]​[f+​(H)])\displaystyle\ \ +\Big{(}[f_{-}(-M_{x})][g_{+}(H)]+[g_{-}(-M_{x})][f_{+}(H)]\Big{)}
βˆ’(f​(0)​[gβˆ’β€‹(βˆ’Mx)+g+​(H)]+g​(0)​[fβˆ’β€‹(βˆ’Mx)+f+​(H)])\displaystyle\ \ -\Big{(}f(0)[g_{-}(-M_{x})+g_{+}(H)]+g(0)[f_{-}(-M_{x})+f_{+}(H)]\Big{)}
+f​(0)​g​(0)\displaystyle\ \ +f(0)g(0)
=A+Bβˆ’C+f​(0)​g​(0).\displaystyle=\ \ A+B-C+f(0)g(0).

But

A\displaystyle A =([fβˆ’β€‹(βˆ’Mx)]​[gβˆ’β€‹(βˆ’Mx)]+[f+​(H)]​[g+​(H)])\displaystyle=\Big{(}[f_{-}(-M_{x})][g_{-}(-M_{x})]+[f_{+}(H)][g_{+}(H)]\Big{)}
=([fβˆ’β€‹gβˆ’β€‹(βˆ’Mx)]+[f+​g+​(H)]βˆ’f​(0)​g​(0))+f​(0)​g​(0)\displaystyle=\Big{(}[f_{-}g_{-}(-M_{x})]+[f_{+}g_{+}(H)]-f(0)g(0)\Big{)}+f(0)g(0)
=[γ​(f​g)]+f​(0)​g​(0)\displaystyle=[\gamma(fg)]+f(0)g(0)
=Γ​(f​g)+f​(0)​g​(0),\displaystyle=\Gamma(fg)+f(0)g(0),

and using Lemma 3.4, we see that

B\displaystyle B =([fβˆ’β€‹(βˆ’Mx)]​[g+​(H)])+([gβˆ’β€‹(βˆ’Mx)]​[f+​(H)])\displaystyle=\Big{(}[f_{-}(-M_{x})][g_{+}(H)]\Big{)}+\Big{(}[g_{-}(-M_{x})][f_{+}(H)]\Big{)}
=(g​(0)​[fβˆ’β€‹(βˆ’Mx)]+[f​(0)​g+​(H)]βˆ’f​(0)​g​(0))+(f​(0)​[gβˆ’β€‹(βˆ’Mx)]+[g​(0)​f+​(H)]βˆ’f​(0)​g​(0))\displaystyle=\Big{(}g(0)[f_{-}(-M_{x})]+[f(0)g_{+}(H)]-f(0)g(0)\Big{)}+\Big{(}f(0)[g_{-}(-M_{x})]+[g(0)f_{+}(H)]-f(0)g(0)\Big{)}
=Cβˆ’2​f​(0)​g​(0).\displaystyle=C-2f(0)g(0).

Therefore,

Γ​(f)​Γ​(g)\displaystyle\Gamma(f)\Gamma(g) =A+Bβˆ’C+f​(0)​g​(0)\displaystyle=A+B-C+f(0)g(0)
=(Γ​(f​g)+f​(0)​g​(0))+(Cβˆ’2​f​(0)​g​(0))βˆ’C+f​(0)​g​(0)\displaystyle=(\Gamma(fg)+f(0)g(0))+(C-2f(0)g(0))-C+f(0)g(0)
=Γ​(f​g).\displaystyle=\Gamma(fg).

∎

Lemma 3.6.

If pp is a polynomial in two variables and we define fβˆˆπ’«f\in{\mathcal{P}} by letting

fβˆ’β€‹(t)=p​(t,0),t∈[βˆ’1,0]andf+​(z)=p​(0,z),zβˆˆπ”»β€‹(1,1)Β―,f_{-}(t)=p(t,0),\ t\in[-1,0]\ \ \ \ \text{and}\ \ \ \ f_{+}(z)=p(0,z),\ z\in\overline{\mathbb{D}(1,1)},

then

p​([βˆ’Mx],[H])=Γ​(f).p([-M_{x}],[H])=\Gamma(f).
Proof.

If p=p​(x,y)p=p(x,y) is a polynomial in two variables and we let

q​(x,y)=p​(x,y)βˆ’p​(x,0)βˆ’p​(0,y)+p​(0,0),q(x,y)=p(x,y)-p(x,0)-p(0,y)+p(0,0),

then

p​(x,y)=p​(x,0)+p​(0,y)βˆ’p​(0,0)+q​(x,y)p(x,y)=p(x,0)+p(0,y)-p(0,0)+q(x,y)

and

q​([βˆ’Mx],[H])=0.q([-M_{x}],[H])=0.

Therefore,

p​([βˆ’Mx],[H])\displaystyle p([-M_{x}],[H]) =p​([βˆ’Mx],0)+p​(0,[H])βˆ’p​(0,0)\displaystyle=p([-M_{x}],0)+p(0,[H])-p(0,0)
=fβˆ’β€‹([Mx])+f+​([h])βˆ’f​(0)\displaystyle=f_{-}([M_{x}])+f_{+}([h])-f(0)
=[γ​(f)]\displaystyle=[\gamma(f)]
=Γ​(f).\displaystyle=\Gamma(f).

∎

Corollary 3.7.

The range of Ξ“\Gamma is dense in π’ž\mathcal{C}.

Proof.

This follows immediately from Lemma 3.6 by recalling that [βˆ’Mx][-M_{x}] and [H][H] generate π’ž\mathcal{C} (cf. proof of Proposition 3.1). ∎

Lemma 3.6 suggests that we consider the subset 𝒫0{\mathcal{P}}_{0} of 𝒫{\mathcal{P}} defined by

𝒫0={fβˆˆπ’«|fβˆ’β€‹Β and ​f+​ are polynomials}.{\mathcal{P}}_{0}=\{f\in{\mathcal{P}}\,|\,f_{-}\text{ and }f_{+}\text{ are polynomials}\}.

We note that it follows from the facts that the polynomials are dense in both C​([βˆ’1,0])C([-1,0]) and A​(𝔻​(1,1)Β―)A(\overline{\mathbb{D}(1,1)}) that 𝒫0{\mathcal{P}}_{0} is dense in 𝒫{\mathcal{P}}.

Lemma 3.8.

If s∈[βˆ’1,0]s\in[-1,0], then

|fβˆ’β€‹(s)|≀‖Γ​(f)β€–|f_{-}(s)|\leq\|\Gamma(f)\| (3.9)

for all fβˆˆπ’«f\in{\mathcal{P}}.

Proof.

As ff is continuous, it suffices to prove the lemma under the assumption that s∈(βˆ’1,0)s\in(-1,0). For nn satisfying 1/n<min⁑{s,1βˆ’s}1/n<\min\{s,1-s\} we define a unit vector Ο‡n∈L2\chi_{n}\in L^{2} by the formula

Ο‡n​(t)={n2if ​|tβˆ’s|≀1/n0if ​|tβˆ’s|>1/n\chi_{n}(t)=\left\{\begin{array}[]{ll}\sqrt{\frac{n}{2}}&\mbox{if }|t-s|\leq 1/n\\ 0&\mbox{if }|t-s|>1/n\end{array}\right.

We observe that the mean value theorem for integrals implies that

limnβ†’βˆžβŸ¨g​χn,Ο‡n⟩=g​(s)\lim_{n\to\infty}\langle g\,\chi_{n},\chi_{n}\rangle=g(s)

whenever g∈C​([0,1])g\in C([0,1]). Also, as Ο‡nβ†’0\chi_{n}\to 0 weakly,

limnβ†’βˆžβ€–K​χnβ€–=0\lim_{n\to\infty}\|K\chi_{n}\|=0

whenever KK is a compact operator acting on L2L^{2}. In particular, as VV is compact and V​χn​(t)=0V\chi_{n}(t)=0 when t∈[0,sβˆ’1/n)t\in[0,s-1/n),

limnβ†’βˆžβ€–H​χnβ€–=limnβ†’βˆžβ€–M1/x​V​χnβ€–=0.\lim_{n\to\infty}\|H\chi_{n}\|=\lim_{n\to\infty}\|M_{1/x}V\chi_{n}\|=0.

More generally, if qq is a polynomial and q​(0)=0q(0)=0, write q​(z)=z​r​(z)q(z)=zr(z), and we get

limnβ†’βˆžβ€–q​(H)​χnβ€–=limnβ†’βˆžβ€–r​(H)​H​χnβ€–=0.\lim_{n\to\infty}\|q(H)\chi_{n}\|=\lim_{n\to\infty}\|r(H)\ H\chi_{n}\|=0.

Now fix fβˆˆπ’«0f\in{\mathcal{P}}_{0} and a compact operator KK acting on L2L^{2}. Using the observations in the previous paragraph we have that

⟨(γ​(f)+K)​χn,Ο‡n⟩\displaystyle\langle(\gamma(f)+K)\ \chi_{n},\chi_{n}\rangle =\displaystyle=\ ⟨(fβˆ’β€‹(βˆ’Mx)+f+​(H)βˆ’f​(0)+K)​χn,Ο‡n⟩\displaystyle\langle(f_{-}(-M_{x})+f_{+}(H)-f(0)+K)\ \chi_{n},\chi_{n}\rangle
=\displaystyle=\ ⟨fβˆ’β€‹(βˆ’x)​χn,Ο‡n⟩+⟨(f+βˆ’f+​(0))​(H)​χn,Ο‡n⟩+⟨K​χn,Ο‡n⟩\displaystyle\langle f_{-}(-x)\chi_{n},\chi_{n}\rangle+\langle(f_{+}-f_{+}(0))(H)\ \chi_{n},\chi_{n}\rangle+\langle K\ \chi_{n},\chi_{n}\rangle
β†’\displaystyle\to\ fβˆ’β€‹(s)+ 0+0\displaystyle\ \ \ \ \ f_{-}(s)\qquad\qquad+\qquad\qquad\ \ 0\qquad\qquad\qquad\ +\qquad 0
=\displaystyle=\ fβˆ’β€‹(s).\displaystyle\ \ f_{-}(s).

Therefore, as β€–Ο‡nβ€–=1\|\chi_{n}\|=1,

|fβˆ’β€‹(s)|≀‖γ​(f)+Kβ€–|f_{-}(s)|\leq\|\gamma(f)+K\|

for all fβˆˆπ’«0f\in{\mathcal{P}}_{0} and KK any compact operator acting on L2L^{2}. Hence,

|fβˆ’β€‹(s)|≀infKβˆˆπ’¦0‖γ​(f)+Kβ€–=‖Γ​(f)β€–|f_{-}(s)|\leq\inf_{K\in{\mathcal{K}}_{0}}\|\gamma(f)+K\|=\|\Gamma(f)\|

for all fβˆˆπ’«0f\in{\mathcal{P}}_{0}. As Ξ“\Gamma is continuous and 𝒫0{\mathcal{P}}_{0} is dense in 𝒫{\mathcal{P}}, it follows that (3.9) holds for all fβˆˆπ’«f\in{\mathcal{P}}. ∎

Lemma 3.10.

If zβˆˆπ”»β€‹(1,1)Β―z\in\overline{\mathbb{D}(1,1)}, then

|f+​(z)|≀‖Γ​(f)β€–|f_{+}(z)|\leq\|\Gamma(f)\| (3.11)

for all fβˆˆπ’«f\in{\mathcal{P}}.

Proof.

We first observe that as f+∈A​(𝔻​(1,1)Β―)f_{+}\in A(\overline{\mathbb{D}(1,1)}), by the Maximum Modulus Theorem it suffices to prove the lemma under the assumption that z=1+Ο„z=1+\tau where Ο„βˆˆπ•‹βˆ–{βˆ’1}\tau\in\mathbb{T}\setminus\{-1\}. For Ξ±βˆˆπ”»\alpha\in\mathbb{D}, let

Ξ₯Ξ±=Uβˆ—β€‹kβˆ’Ξ±Β―β€–kβˆ’Ξ±Β―β€–,\Upsilon_{\alpha}=U^{*}\frac{k_{-\bar{\alpha}}}{\|k_{-\bar{\alpha}}\|},

where UU is as in (2.3). Clearly, as kβˆ’Ξ±Β―/β€–kβˆ’Ξ±Β―β€–k_{-\bar{\alpha}}/\|k_{-\bar{\alpha}}\| is a unit vector and Uβˆ—U^{*} is unitary, Ξ₯Ξ±\Upsilon_{\alpha} is a unit vector. Also, as

(1βˆ’Sβˆ—)​kβˆ’Ξ±Β―β€–kβˆ’Ξ±Β―β€–=(1+Ξ±)​kβˆ’Ξ±Β―β€–kβˆ’Ξ±Β―β€–,(1-S^{*})\frac{k_{-\bar{\alpha}}}{\|k_{-\bar{\alpha}}\|}=(1+\alpha)\frac{k_{-\bar{\alpha}}}{\|k_{-\bar{\alpha}}\|},

it follows that H​Ξ₯Ξ±=(1+Ξ±)​Ξ₯Ξ±H\Upsilon_{\alpha}=(1+\alpha)\Upsilon_{\alpha}, and more generally,

f+​(H)​Ξ₯Ξ±=f+​(1+Ξ±)​Ξ₯Ξ±f_{+}(H)\Upsilon_{\alpha}\ =\ f_{+}(1+\alpha)\Upsilon_{\alpha} (3.12)

for all fβˆˆπ’«f\in{\mathcal{P}}.

Now notice that (2.3) implies that

Ξ₯Ξ±=1βˆ’|Ξ±|21+α​xβˆ’Ξ±1+Ξ±.\Upsilon_{\alpha}\ =\ \frac{\sqrt{1-|\alpha|^{2}}}{1+\alpha}x^{-\frac{\alpha}{1+\alpha}}.
Claim 3.13.

If ρ>0\rho>0 and Ο„βˆˆπ•‹βˆ–{βˆ’1}\tau\in\mathbb{T}\setminus\{-1\}, then

limΞ±β†’Ο„βŸ¨xρ​Ξ₯Ξ±,Ξ₯α⟩=0.\lim_{\alpha\to\tau}\langle x^{\rho}\Upsilon_{\alpha},\Upsilon_{\alpha}\rangle=0. (3.14)
Proof.

First note that

Οβˆ’(Ξ±1+Ξ±+Ξ±Β―1+Ξ±Β―)+1=ρ+1βˆ’|Ξ±|2|1+Ξ±|2,\rho-(\frac{\alpha}{1+\alpha}+\frac{\bar{\alpha}}{1+\bar{\alpha}})+1=\rho+\frac{1-|\alpha|^{2}}{|1+\alpha|^{2}},

so that

∫01xΟβˆ’(Ξ±1+Ξ±+Ξ±Β―1+Ξ±Β―)​𝑑x=(ρ+1βˆ’|Ξ±|2|1+Ξ±|2)βˆ’1.\int_{0}^{1}x^{\rho-(\frac{\alpha}{1+\alpha}+\frac{\bar{\alpha}}{1+\bar{\alpha}})}dx\ =\ \Big{(}\rho+\frac{1-|\alpha|^{2}}{|1+\alpha|^{2}}\Big{)}^{-1}.

Hence,

⟨xρ​Ξ₯Ξ±,Ξ₯α⟩\displaystyle\langle x^{\rho}\Upsilon_{\alpha},\Upsilon_{\alpha}\rangle =1βˆ’|Ξ±|2|1+Ξ±|2β€‹βˆ«01xΟβˆ’(Ξ±1+Ξ±+Ξ±Β―1+Ξ±Β―)​𝑑x\displaystyle=\frac{1-|\alpha|^{2}}{|1+\alpha|^{2}}\int_{0}^{1}x^{\rho-(\frac{\alpha}{1+\alpha}+\frac{\bar{\alpha}}{1+\bar{\alpha}})}dx
=1βˆ’|Ξ±|2|1+Ξ±|2​(ρ+1βˆ’|Ξ±|2|1+Ξ±|2)βˆ’1\displaystyle=\frac{1-|\alpha|^{2}}{|1+\alpha|^{2}}\Big{(}\rho+\frac{1-|\alpha|^{2}}{|1+\alpha|^{2}}\Big{)}^{-1}
=1βˆ’|Ξ±|2|1+Ξ±|2​ρ+1βˆ’|Ξ±|2.\displaystyle=\frac{1-|\alpha|^{2}}{|1+\alpha|^{2}\rho\ +1-|\alpha|^{2}}.

Therefore, if ρ>0\rho>0 and Ο„βˆˆπ•‹βˆ–{βˆ’1}\tau\in\mathbb{T}\setminus\{-1\}, (3.14) holds. ∎

Observe that if qq is a polynomial and Ο„βˆˆπ•‹βˆ–{βˆ’1}\tau\in\mathbb{T}\setminus\{-1\}, then Claim 3.13 implies that ⟨q​Ξ₯Ξ±,Ξ₯Ξ±βŸ©β†’q​(0)\langle q\Upsilon_{\alpha},\Upsilon_{\alpha}\rangle\to q(0) as Ξ±β†’Ο„\alpha\to\tau. In particular,

limΞ±β†’Ο„βŸ¨q​(βˆ’Mx)​Ξ₯Ξ±,Ξ₯α⟩=0\lim_{\alpha\to\tau}\langle q(-M_{x})\Upsilon_{\alpha},\Upsilon_{\alpha}\rangle=0 (3.15)

whenever Ο„βˆˆπ•‹βˆ–{βˆ’1}\tau\in\mathbb{T}\setminus\{-1\} and qq is a polynomial satisfying q​(0)=0q(0)=0.

We now conclude the proof of the lemma. We need to show that if fβˆˆπ’«f\in{\mathcal{P}} and Ο„βˆˆπ•‹βˆ–{βˆ’1}\tau\in\mathbb{T}\setminus\{-1\} then (3.11) holds with z=1+Ο„z=1+\tau. First assume that fβˆˆπ’«0f\in{\mathcal{P}}_{0} and fix Kβˆˆπ’¦0K\in\mathcal{K}_{0}. Since Ξ₯Ξ±β†’0\Upsilon_{\alpha}\to 0 weakly as Ξ±β†’Ο„\alpha\to\tau, using (3.12) and (3.14) we have

⟨(γ​(f)+K)​Ξ₯Ξ±,Ξ₯α⟩\displaystyle\langle(\gamma(f)+K)\Upsilon_{\alpha},\Upsilon_{\alpha}\rangle =\displaystyle\ =\ ⟨(fβˆ’βˆ’fβˆ’β€‹(0))​(βˆ’Mx)​Ξ₯Ξ±,Ξ₯α⟩+⟨f+​(H)​Ξ₯Ξ±,Ξ₯α⟩+⟨K​Ξ₯Ξ±,Ξ₯α⟩\displaystyle\langle(f_{-}-f_{-}(0))(-M_{x})\Upsilon_{\alpha},\Upsilon_{\alpha}\rangle+\langle f_{+}(H)\Upsilon_{\alpha},\Upsilon_{\alpha}\rangle+\langle K\Upsilon_{\alpha},\Upsilon_{\alpha}\rangle
β†’\displaystyle\to 0+f+​(1+Ο„)+ 0\displaystyle\qquad\qquad\qquad 0\qquad\qquad\qquad+\ \ \ \ f_{+}(1+\tau)\ \ \ \ +\ \ 0
=\displaystyle= f+​(1+Ο„).\displaystyle f_{+}(1+\tau).

as Ξ±β†’Ο„\alpha\to\tau. Therefore, if fβˆˆπ’«0f\in{\mathcal{P}}_{0} and Ο„βˆˆπ•‹βˆ–{βˆ’1}\tau\in\mathbb{T}\setminus\{-1\},

|f+​(1+Ο„)|≀‖γ​(f)+Kβ€–.|f_{+}(1+\tau)|\leq\|\gamma(f)+K\|.

Hence, if fβˆˆπ’«0f\in{\mathcal{P}}_{0},

|f+​(1+Ο„)|≀infK∈K0‖γ​(f)+Kβ€–=‖Γ​(f)β€–.|f_{+}(1+\tau)|\leq\inf_{K\in K_{0}}\|\gamma(f)+K\|=\|\Gamma(f)\|.

As Ξ“\Gamma is continuous and 𝒫0{\mathcal{P}}_{0} is dense in 𝒫{\mathcal{P}}, it follows that (3.11) holds with z=1+Ο„z=1+\tau for all fβˆˆπ’«f\in{\mathcal{P}}. ∎

Lemma 3.16.

Ξ“\Gamma is a homeomorphism.

Proof.

In the proof of Proposition 3.5 we showed that

‖Γ​(f)‖≀3​‖fβ€–\|\Gamma(f)\|\leq 3\|f\|

for all fβˆˆπ’«f\in{\mathcal{P}}. On the other hand, Lemma 3.8 implies that

maxt∈[βˆ’1,0]⁑|fβˆ’β€‹(t)|≀‖Γ​(f)β€–\max_{t\in[-1,0]}|f_{-}(t)|\leq\|\Gamma(f)\|

for all fβˆˆπ’«f\in{\mathcal{P}} and Lemma 3.10 implies that

maxzβˆˆπ”»β€‹(1,1)¯⁑|f+​(z)|≀‖Γ​(f)β€–\max_{z\in\overline{\mathbb{D}(1,1)}}|f_{+}(z)|\leq\|\Gamma(f)\|

for all fβˆˆπ’«f\in{\mathcal{P}}. Therefore,

β€–fβ€–=max⁑{maxt∈[βˆ’1,0]⁑|fβˆ’β€‹(t)|,maxzβˆˆπ”»β€‹(1,1)¯⁑|f+​(z)|}≀‖Γ​(f)β€–\|f\|=\max\ \big{\{}\max_{t\in[-1,0]}|f_{-}(t)|,\max_{z\in\overline{\mathbb{D}(1,1)}}|f_{+}(z)|\ \big{\}}\leq\|\Gamma(f)\|

for all fβˆˆπ’«f\in{\mathcal{P}}. ∎

Putting together the results of Subsection 3.1 we get the following theorem.

Theorem 3.17.

The map Ξ“\Gamma is a homeomorphic unital isomorphism from 𝒫{\mathcal{P}} onto π’ž\mathcal{C}.

3.2 Some Observations on the Gelfand Theory of π’ž\mathcal{C}

If Ξ›=[βˆ’1,0]βˆͺ𝔻​(1,1)Β―\Lambda=[-1,0]\cup\overline{\mathbb{D}(1,1)}, then there is an isometric isomorphism from 𝒫{\mathcal{P}} onto the lollipop algebra A​(Ξ›)A(\Lambda) given by

A​(Ξ›)βˆ‹f↦(f|[βˆ’1,0],f|(𝔻​(1,1)Β―)).A(\Lambda)\ni f\mapsto\big{(}\ f|_{[-1,0]}\ ,f|_{(\overline{\mathbb{D}(1,1)})}\ \big{)}.

So one could just as well state Theorem 3.17 with 𝒫{\mathcal{P}} replaced by A​(Ξ›)A(\Lambda) and Ξ“\Gamma replaced with the map Ξ“βˆΌ:A​(Ξ›)β†’π’ž\Gamma^{\sim}:A(\Lambda)\to\mathcal{C} defined by

Ξ“βˆΌβ€‹(f)=[f|[βˆ’1,0]​(βˆ’Mx)+f|𝔻​(1,1)¯​(H)βˆ’f​(0)].\ \Gamma^{\sim}(f)=\Big{[}\ f\big{|}_{[-1,0]}(-M_{x})+f\big{|}_{\overline{\mathbb{D}(1,1)}}(H)-f(0)\ \Big{]}.
Definition 3.18.

Define ΞΈ:π’œβ†’A​(Ξ›)\theta:{\mathcal{A}}\to A(\Lambda) by

θ​(X)=(Ξ“βˆΌ)βˆ’1​([X]).\theta(X)\ =\ (\Gamma^{\sim})^{-1}([X]).

Then Theorem 3.17 says that there is a short exact sequence

0β†’π’¦π’œβ†’π’œβ†’ΞΈA​(Ξ›)β†’0.0\to{\mathcal{K}}_{{\mathcal{A}}}\to{\mathcal{A}}\stackrel{{\scriptstyle\theta}}{{\to}}A(\Lambda)\to 0.
Remark 3.19.

By Mergelyan’s theorem, A​(Ξ›)=P​(Ξ›)A(\Lambda)=P(\Lambda), and since zz generates P​(Ξ›)P(\Lambda), it follows that

Ξ“βˆΌβ€‹(z)=[Hβˆ’Mx]\Gamma^{\sim}(z)=[H-M_{x}]

generates π’ž\mathcal{C}. We shall examine Hβˆ’MxH-M_{x} in Section 6.

4 Cβˆ—β€‹(π’œ)C^{*}({\mathcal{A}})

We shall let ℬ=Cβˆ—β€‹(π’œ){\mathcal{B}}=C^{*}({\mathcal{A}}) denote the Cβˆ—C^{*}-algebra generated by π’œ{\mathcal{A}}. Since it is irreducible, ℬ{\mathcal{B}} contains all the compact operators.

The Toeplitz Cβˆ—C^{*}-algebra 𝒯\mathcal{T} is the Cβˆ—C^{*} algebra generated by the shift SS. There is a short exact sequence

0→𝒦→𝒯→αA​(𝔻¯)β†’0.0\to{\mathcal{K}}\to{\mathcal{T}}\stackrel{{\scriptstyle\alpha}}{{\to}}\ A(\overline{\mathbb{D}})\to 0.

(See e.g. [7, 7.23]). A cross-section of Ξ±\alpha is the map that sends a function mm to the Toeplitz operator TmT_{m} on H2H^{2} with symbol mm.

Since H^=Sβˆ—+1\widehat{H}=S^{*}+1, the Cβˆ—C^{*}-algebra generated by HH is unitarily equivalent to 𝒯\mathcal{T}. We wish to think of it as living on 𝔻​(1,1)Β―\overline{\mathbb{D}(1,1)}, so we must shift things over. Let τ​(z)=z+1\tau(z)=z+1. For any function ff defined on some domain in β„‚\mathbb{C}, let fβˆͺ​(z)=f​(zΒ―)Β―f^{\cup}(z)=\overline{f(\bar{z})} be its reflection in the real axis.

Definition 4.1.

Let ψ∈C​(βˆ‚π”»β€‹(1,1))\psi\in C(\partial\mathbb{D}(1,1)). Let Hψ∈B​(L2)H_{\psi}\in B(L^{2}) be defined by

Hψ=Uβˆ—β€‹T(Οˆβˆ˜Ο„)βˆͺβˆ—β€‹U.H_{\psi}\ =\ U^{*}T^{*}_{(\psi\circ\tau)^{\cup}}U.

The map Οˆβ†¦Hψ\psi\mapsto H_{\psi} is unital and linear, but not multiplicative. One checks that if Οˆβ€‹(z)=zn\psi(z)=z^{n}, then Hzn=HnH_{z^{n}}=H^{n}, and if Οˆβ€‹(z)=zΒ―n\psi(z)=\bar{z}^{n}, then HzΒ―n=(Hβˆ—)nH_{\bar{z}^{n}}=(H^{*})^{n}.

Theorem 4.2.

There is a short exact sequence

0→𝒦→ℬ→πC​(βˆ‚Ξ›)β†’0.0\to{\mathcal{K}}\to{\mathcal{B}}\stackrel{{\scriptstyle\pi}}{{\to}}C(\partial\Lambda)\to 0. (4.3)

For every XX in ℬ{\mathcal{B}}, its coset in ℬ/𝒦{\mathcal{B}}/{\mathcal{K}} can be written uniquely as

[X]=[g​(βˆ’Mx)+HΟˆβˆ’g​(0)][X]\ =\ [g(-M_{x})+H_{\psi}-g(0)] (4.4)

where g∈C​[βˆ’1,0]g\in C[-1,0], ψ∈C​(βˆ‚π”»β€‹(1,1))\psi\in C(\partial\mathbb{D}(1,1)), and g​(0)=Οˆβ€‹(0)g(0)=\psi(0). The essential spectrum of XX as in (4.4) is g​([0,1])βˆͺΟˆβ€‹(βˆ‚π”»β€‹(1,1))g([0,1])\cup\psi(\partial\mathbb{D}(1,1)). If Ξ»βˆ‰Οƒe​(X)\lambda\notin\sigma_{e}(X), then the Fredhom index is given by the winding number of ψ\psi about Ξ»\lambda:

ind​(Xβˆ’Ξ»)=indΟˆβ€‹(Ξ»).{\rm ind}(X-\lambda)\ =\ {\rm ind}_{\psi}(\lambda).

Proof: For Xβˆˆβ„¬X\in{\mathcal{B}}, we shall let [X][X] denote its equivalence class in ℬ/𝒦{\mathcal{B}}/{\mathcal{K}}. We have

[H​Mx]=[Mx​H]=0,[HM_{x}]=[M_{x}H]=0,

and

[H​Hβˆ—]=[Hβˆ—β€‹H]=H+Hβˆ—.[HH^{*}]=[H^{*}H]=H+H^{*}.

So ℬ/𝒦{\mathcal{B}}/{\mathcal{K}} is abelian. Moreover, for any polynomial qq in 3 variables, there are polynomials p1,p2,p3p_{1},p_{2},p_{3} in one variable so that

[q​(Mx,H,Hβˆ—)]=[p1​(Mx)+p2​(H)+p3​(Hβˆ—)].[q(M_{x},H,H^{*})]\ =\ [p_{1}(M_{x})+p_{2}(H)+p_{3}(H^{*})]. (4.5)

Therefore operators of the form (4.5) are dense in ℬ/𝒦{\mathcal{B}}/{\mathcal{K}}.

We wish to prove that ℬ/𝒦{\mathcal{B}}/{\mathcal{K}} is isomorphic to the abelian Cβˆ—C^{*}-algebra C​(βˆ‚Ξ›)C(\partial\Lambda). We will use a similar strategy to the proof of Theorem 3.17. Let

𝒬={f=(fβˆ’,f+):fβˆ’βˆˆC​([βˆ’1,0]),f+∈C​(βˆ‚π”»β€‹(1,1)),fβˆ’β€‹(0)=f+​(0)}.{\mathcal{Q}}\ =\ \{f=(f_{-},f_{+}):f_{-}\in C([-1,0]),f_{+}\in C(\partial\mathbb{D}(1,1)),f_{-}(0)=f_{+}(0)\}.

The algebra 𝒬{\mathcal{Q}} is just C​(βˆ‚Ξ›)C(\partial\Lambda), but it is easier to define the functional calculus on it. Define

Ξ΄:𝒬\displaystyle\delta:{\mathcal{Q}} β†’\displaystyle\ \to\ ℬ\displaystyle{\mathcal{B}}
f\displaystyle f ↦\displaystyle\mapsto fβˆ’β€‹(βˆ’Mx)+Hf+βˆ’f​(0).\displaystyle f_{-}(-M_{x})+H_{f_{+}}-f(0).

β–‘\Box

Let Δ​(f)=[δ​(f)]\Delta(f)=[\delta(f)]. The following lemma is straightforward to prove.

Lemma 4.6.

(i) Let ψ,Ο•βˆˆC​(βˆ‚π”»β€‹(1,1))\psi,\phi\in C(\partial\mathbb{D}(1,1)). Then [HΟˆβ€‹HΟ•]=[HΟˆβ€‹Ο•][H_{\psi}H_{\phi}]=[H_{\psi\phi}].

(ii) Let g∈C​([βˆ’1,0])g\in C([-1,0]) and ψ∈C​(βˆ‚π”»β€‹(1,1))\psi\in C(\partial\mathbb{D}(1,1)). Then

[g​(βˆ’Mx)]​[Hψ]=[Hψ]​[g​(βˆ’Mx)]=[g​(0)​Hψ+Οˆβ€‹(0)​g​(βˆ’Mx)βˆ’g​(0)β€‹Οˆβ€‹(0)].[g(-M_{x})][H_{\psi}]\ =\ [H_{\psi}][g(-M_{x})]\ =\ [g(0)H_{\psi}+\psi(0)g(-M_{x})-g(0)\psi(0)].

Using Lemma 4.6, one can check that Ξ”\Delta is a unital *-homomorphism from 𝒬{\mathcal{Q}} into ℬ/𝒦{\mathcal{B}}/{\mathcal{K}}. Its range is dense, so if we can show it has no kernel, then it is a C*-isomorphism.

Lemma 4.7.

If s∈[βˆ’1,0]s\in[-1,0], then

|fβˆ’β€‹(s)|≀‖Δ​(f)β€–|f_{-}(s)|\leq\|\Delta(f)\| (4.8)

for all fβˆˆπ’¬f\in{\mathcal{Q}}.

Proof.

As ff is continuous, it suffices to prove the lemma under the assumption that s∈(βˆ’1,0)s\in(-1,0), and as Ξ”\Delta is continuous, we can assume that fβˆ’f_{-} is a polynomial, and that f+​(z)=f​(0)+z​p2​(z)+z​p3​(z)Β―f_{+}(z)=f(0)+zp_{2}(z)+\overline{zp_{3}(z)} where p2p_{2} and p3p_{3} are polynomials.

As in Lemma 3.8, for nn satisfying 1/n<min⁑{s,1βˆ’s}1/n<\min\{s,1-s\} we define a unit vector Ο‡n∈L2\chi_{n}\in L^{2} by the formula

Ο‡n​(t)={n2if ​|tβˆ’s|≀1/n0if ​|tβˆ’s|>1/n\chi_{n}(t)=\left\{\begin{array}[]{ll}\sqrt{\frac{n}{2}}&\mbox{if }|t-s|\leq 1/n\\ 0&\mbox{if }|t-s|>1/n\end{array}\right.

Let KK be compact.

⟨(δ​(f)+K)​χn,Ο‡n⟩\displaystyle\langle(\delta(f)+K)\ \chi_{n},\chi_{n}\rangle =\displaystyle=\ ⟨(fβˆ’β€‹(βˆ’Mx)+Hz​p2+Hz¯​pΒ―3+K)​χn,Ο‡n⟩\displaystyle\langle(f_{-}(-M_{x})+H_{zp_{2}}+H_{\bar{z}\bar{p}_{3}}+K)\ \chi_{n},\chi_{n}\rangle
=\displaystyle=\ ⟨fβˆ’β€‹(βˆ’x)​χn,Ο‡n⟩+⟨p2​(H)​H​χn,Ο‡n⟩+βŸ¨Ο‡n,p3​(H)​H​χn⟩+⟨K​χn,Ο‡n⟩\displaystyle\langle f_{-}(-x)\chi_{n},\chi_{n}\rangle+\langle p_{2}(H)H\ \chi_{n},\chi_{n}\rangle+\langle\chi_{n},p_{3}(H)H\ \chi_{n}\rangle+\langle K\ \chi_{n},\chi_{n}\rangle
β†’\displaystyle\to\ fβˆ’β€‹(s)+0+ 0+0\displaystyle\ \ \ f_{-}(s)\qquad\quad\ +\qquad\qquad 0\qquad\ \ \ +\qquad\ 0\qquad\ \ \ \ \ \ \ +\qquad 0
=\displaystyle=\ fβˆ’β€‹(s).\displaystyle\ \ f_{-}(s).

Therefore,

|fβˆ’β€‹(s)|≀infKβˆˆπ’¦β€–Ξ΄β€‹(f)+Kβ€–=‖Δ​(f)β€–.|f_{-}(s)|\leq\inf_{K\in{\mathcal{K}}}\|\delta(f)+K\|=\|\Delta(f)\|.

∎

If Δ​(f)=0\Delta(f)=0, by Lemma 4.7 we must have fβˆ’=0f_{-}=0. So δ​(f)=Hf+\delta(f)=H_{f_{+}} must be compact. But Hf+H_{f_{+}} is unitarily equivalent to a Toeplitz operator, and there are no non-zero compact Toeplitz operators. Therefore Ξ”\Delta has a trivial kernel, and hence is a βˆ—*-isomorphism.

The claim about the spectrum of [x][x] now follows from the fact that the spectrum of a function in C​(βˆ‚Ξ›)C(\partial\Lambda) equals its range. Finally, the claim about the Fredholm index follows from the fact that the Fredholm index at Ξ»\lambda will be unchanged under any homotopy of ff that keeps Ξ»\lambda outside its range. Then ff can be homotoped to (fβˆ’,f+)(f_{-},f_{+}) where f+​(z)=(zβˆ’1βˆ’Ξ»)nf_{+}(z)=(z-1-\lambda)^{n} and fβˆ’β€‹(x)=(βˆ’1βˆ’Ξ»)nf_{-}(x)=(-1-\lambda)^{n} for some integer nn, and the Fredholm index of δ​(f)\delta(f) is nn. β–‘\Box

5 Compact operators in the little algebra π’œ0{\mathcal{A}}_{0}

Recall from Definition 1.6 that π’œ0{\mathcal{A}}_{0} is the norm-closed algebra generated by MxM_{x} and VV. We shall prove that every compact operator in AlgLat​(V){\rm AlgLat}(V) lies not just in π’œ{\mathcal{A}} but in π’œ0{\mathcal{A}}_{0}.

For II an interval in [0,1][0,1], let us write L2​(I)L^{2}(I) for the subspace of L2L^{2} that vanishes a.e. off II, and let PIP_{I} denote projection onto L2​(I)L^{2}(I). For Ο•,ψ∈L2\phi,\psi\in L^{2} we write Ο•βŠ—Οˆ\phi\otimes\psi to denote the rank one operator

Ο•βŠ—Οˆ:fβ†¦βŸ¨f,ΟˆβŸ©β€‹Ο•.\phi\otimes\psi:f\ \mapsto\ \langle f,\psi\rangle\phi.

The key observation is the following:

Lemma 5.1.

Suppose Ο•βˆˆL2​[t,1]\phi\in L^{2}[t,1] and ψ∈L2​[0,t]\psi\in L^{2}[0,t]. Then Ο•βŠ—Οˆ=Mϕ​V​MΟˆβˆ—\phi\otimes\psi=M_{\phi}VM_{\psi}^{*}.

Proof: We have

Mϕ​V​MΟˆβˆ—β€‹f​(x)=ϕ​(x)β€‹βˆ«0xf​(s)β€‹Οˆβ€‹(s)¯​𝑑s.M_{\phi}VM_{\psi}^{*}f(x)\ =\ \phi(x)\int_{0}^{x}f(s)\overline{\psi(s)}ds.

The right-hand side is 0 if x<tx<t, and ϕ​(x)β€‹βŸ¨f,ψ⟩\phi(x)\langle f,\psi\rangle if x>tx>t. β–‘\Box

Lemma 5.2.

Every finite rank operator on L2​[0,1]L^{2}[0,1] can be written as an integral operator whose kernel is in L2​([0,1]Γ—[0,1])L^{2}([0,1]\times[0,1]).

Proof: Let K=βˆ‘j=1nΟ•jβŠ—ΟˆjK=\sum_{j=1}^{n}\phi_{j}\otimes\psi_{j}. Define

k​(x,s)=βˆ‘j=1nΟ•j​(x)β€‹Οˆj​(s)Β―.k(x,s)\ =\ \sum_{j=1}^{n}\phi_{j}(x)\overline{\psi_{j}(s)}.

Then K​f​(x)=∫01k​(x,s)​f​(s)​𝑑sKf(x)=\int_{0}^{1}k(x,s)f(s)ds, and kk is in L2​([0,1]Γ—[0,1])L^{2}([0,1]\times[0,1]). β–‘\Box

Lemma 5.3.

Let kk be in L2​([0,1]Γ—[0,1])L^{2}([0,1]\times[0,1]), and let T​f​(x)=∫01k​(x,s)​f​(s)​𝑑sTf(x)=\int_{0}^{1}k(x,s)f(s)ds. Then TT is in AlgLat​(V){\rm AlgLat}(V) if and only if k​(s,x)=0k(s,x)=0 for s>xs>x.

Proof: Sufficiency is clear. To prove necessity, assume that for some 0<t<10<t<1, the kernel

k​(s,x)​χ[0,t]​(x)​χ[t,1]​(s)k(s,x)\chi_{[0,t](x)}\chi_{[t,1]}(s)

is not 0 a.e. As an integral operator is zero if and only if the kernel is 0 a.e., this means that the corresponding integral operator is non-zero, and hence TT maps a function in L2​(t,1)L^{2}(t,1) to a function that is not 0 a.e. on [0,t][0,t]. β–‘\Box

Lemma 5.4.

Let T=Ο•βŠ—ΟˆT=\phi\otimes\psi be a rank-one operator. Then TT is in AlgLat​(V){\rm AlgLat}(V) if and only if for some 0<t<10<t<1, the support of Ο•\phi is in [t,1][t,1] (i.e. Ο•=0\phi=0 a.e. on [0,t][0,t]) and the support of ψ\psi is in [0,t][0,t]. In this case, Tβˆˆπ’œ0T\in{\mathcal{A}}_{0}.

Proof: The first part follows from Lemma 5.3. For the second part, observe that if the supports of Ο•,ψ\phi,\psi are in [t,1][t,1] and [0,t][0,t] respectively, then Ο•βŠ—Οˆ=Mϕ​V​MΟˆβˆ—\phi\otimes\psi=M_{\phi}VM_{\psi}^{*}. If Ο•\phi and ψ\psi are both in C​([0,1])C([0,1]), this proves that Ο•βŠ—Οˆβˆˆπ’œ0\phi\otimes\psi\in{\mathcal{A}}_{0}.

For the general case, choose continuous functions fnf_{n} and gng_{n} that converge to Ο•\phi and ψ\psi respectively in L2L^{2}. It follows from Lemma 5.1 that Mfn​V​Mgmβˆ—M_{f_{n}}VM_{g_{m}}^{*} converges to Mϕ​V​Mgmβˆ—M_{\phi}VM_{g_{m}}^{*} in norm as nβ†’βˆžn\to\infty, and that Mϕ​V​Mgmβˆ—M_{\phi}VM_{g_{m}}^{*} converges to Mϕ​V​MΟˆβˆ—M_{\phi}VM_{\psi}^{*}. Therefore Ο•βŠ—Οˆβˆˆπ’œ0\phi\otimes\psi\in{\mathcal{A}}_{0}. β–‘\Box

Theorem 5.5.

Let KK be a compact operator in AlgLat​(V){\rm AlgLat}(V). Then Kβˆˆπ’œ0K\in{\mathcal{A}}_{0}, and can be approximated in norm by finite rank operators in π’œ0{\mathcal{A}}_{0}.

Proof: Note that K∈AlgLat​(V)K\in{\rm AlgLat}(V) means that for all 0<s<10<s<1, we have P[0,s]​K​P[s,1]=0P_{[0,s]}KP_{[s,1]}=0.

Let Ξ΅>0\varepsilon>0. First, consider P[1/2,1]​K​P[0,1/2]P_{[1/2,1]}KP_{[0,1/2]}. This can be approximated within Ξ΅/2\varepsilon/2 by a finite rank operator that is a sum of rank one operators that map L2​(0,1/2)L^{2}(0,1/2) to L2​(1/2,1)L^{2}(1/2,1). By Lemma 5.4, this means that this finite rank operator is in π’œ0{\mathcal{A}}_{0}.

A similar argument shows that P[1/4,1/2]​K​P[0,1/4]P_{[1/4,1/2]}KP_{[0,1/4]} and P[3/4,1]​K​P[1/2,3/4]P_{[3/4,1]}KP_{[1/2,3/4]} can both be approximated by finite rank operators in π’œ0{\mathcal{A}}_{0} within Ξ΅/8\varepsilon/8. Iterating, we get that if nn is a power of 22, we can approximate

Kβˆ’βˆ‘j=1nP[(jβˆ’1)/n,j/n]​K​P[(jβˆ’1)/n,j/n]K-\sum_{j=1}^{n}P_{[(j-1)/n,j/n]}KP_{[(j-1)/n,j/n]}

within Ξ΅\varepsilon by a finite rank operator in π’œ0{\mathcal{A}}_{0}.

Finally we observe that

β€–βˆ‘j=1nP[(jβˆ’1)/n,j/n]​K​P[(jβˆ’1)/n,j/n]β€–=max1≀j≀n⁑‖P[(jβˆ’1)/n,j/n]​K​P[(jβˆ’1)/n,j/n]β€–.\|\sum_{j=1}^{n}P_{[(j-1)/n,j/n]}KP_{[(j-1)/n,j/n]}\|\ =\ \max_{1\leq j\leq n}\|P_{[(j-1)/n,j/n]}KP_{[(j-1)/n,j/n]}\|. (5.6)

Since KK is compact,

limnβ†’βˆžsup|I|=1/nβ€–K​PIβ€–=0,\lim_{n\to\infty}\sup_{|I|=1/n}\|KP_{I}\|=0,

so (5.6) tends to 0. β–‘\Box

6 The operator Hβˆ’MxH-M_{x}

Let us write ZZ for the operator Hβˆ’MxH-M_{x}. We know that [Z][Z] generates the Calkin Hardy-Weyl algebra π’ž\mathcal{C}. By Theorem 3.17 we know that the spectrum of [Z][Z] in π’œ/𝒦0{\mathcal{A}}/{\mathcal{K}}_{0} is Ξ›\Lambda. It is not surprising that 𝔻​(1,1)\mathbb{D}(1,1) are eigenvalues of ZZ, since they are eigenvalues of HH. It is perhaps surprising that every point in the stick, except βˆ’1-1, is also an eigenvalue. Moreover as we move up the stick to the bulb of the lollipop, the eigenvalues increase in multiplicity.

Theorem 6.1.

(i) Οƒp​(Z)=𝔻​(1,1)βˆͺ(βˆ’1,0]\sigma_{p}(Z)=\mathbb{D}(1,1)\cup(-1,0].

(ii) The point spectrum of Zβˆ—Z^{*} is empty.

(iii) The spectrum of ZZ is Ξ›\Lambda.

Proof: (i) Suppose (Zβˆ’Ξ»)​f=0(Z-\lambda)f=0. Let F​(x)=V​f​(x)=∫0xf​(t)​𝑑tF(x)=Vf(x)=\int_{0}^{x}f(t)dt. Then we have

1x​F​(x)=(x+Ξ»)​f​(x).\frac{1}{x}F(x)\ =\ (x+\lambda)f(x).

As F′​(x)=f​(x)F^{\prime}(x)=f(x), we get the equation

1x​F​(x)=(x+Ξ»)​F′​(x),\frac{1}{x}F(x)=(x+\lambda)F^{\prime}(x), (6.2)

with the boundary condition

F​(0)= 0.\quad F(0)\ =\ 0. (6.3)

The function FF is continuous. Let Ξ©\Omega denote the relatively open subset of [0,1][0,1] on which it is non-zero.

We get that the solution of (6.2), with Ξ»β‰ 0\lambda\neq 0, is

F​(x)=c​(xx+Ξ»)1/λ​χΩ​(x),F(x)\ =\ c\left(\frac{x}{x+\lambda}\right)^{1/\lambda}\chi_{\Omega}(x)\ , (6.4)

where the constant cc can a priori be different on different components of Ξ©\Omega (though we show below that Ξ©\Omega is actually connected). Hence

f​(x)=c​x1Ξ»βˆ’1​(x+Ξ»)βˆ’1βˆ’1λ​χΩ​(x).f(x)\ =\ cx^{\frac{1}{\lambda}-1}(x+\lambda)^{-1-\frac{1}{\lambda}}\chi_{\Omega}(x). (6.5)

Case: Ξ©=[0,1]\Omega=[0,1].

For ff to be in L2L^{2} with c≠0c\neq 0, we need

β„œβ‘(1Ξ»βˆ’1)>βˆ’12,\Re(\frac{1}{\lambda}-1)\ >\ -\frac{1}{2},

which is the same as Ξ»βˆˆπ”»β€‹(1,1)\lambda\in\mathbb{D}(1,1). For Ξ»βˆˆπ”»β€‹(1,1)\lambda\in\mathbb{D}(1,1), we get the eigenfunctions

f​(x)=x1Ξ»βˆ’1​(x+Ξ»)βˆ’1βˆ’1Ξ».f(x)\ =\ x^{\frac{1}{\lambda}-1}(x+\lambda)^{-1-\frac{1}{\lambda}}. (6.6)

When Ξ»=0\lambda=0, we get

F​(x)=c​eβˆ’1/x,F(x)\ =\ ce^{-1/x},

and

f​(x)=c​1x2​eβˆ’1x.f(x)\ =\ c\frac{1}{x^{2}}e^{-\frac{1}{x}}. (6.7)

Now suppose that Ξ©\Omega is not all of [0,1][0,1]. Decompose Ξ©\Omega as a union of disjoint non-empty intervals. On each interval, we have that ff is given by (6.5), with some constant cc that can depend on the interval. Choose such an interval, II. Then 0 cannot be an end-point of II, or we would have that FF is given by (6.4), and this is discontinuous at the right-hand end-point of II.

So assume that the left-hand end-point of II is t>0t>0. Then the boundary condition (6.3) is replaced by F​(t)=0F(t)=0. On II, we have

F​(x)=c​(xx+Ξ»)1/Ξ»,F(x)\ =\ c\left(\frac{x}{x+\lambda}\right)^{1/\lambda},

so to have F​(t)=0F(t)=0 we need Ξ»=βˆ’t\lambda=-t. By continuity, FF cannot vanish again, so we conclude that I=(t,1]I=(t,1] and that for λ∈(βˆ’1,0)\lambda\in(-1,0) the function

f​(x)=x1Ξ»βˆ’1​(x+Ξ»)βˆ’1βˆ’1λ​χ[βˆ’Ξ»,1]​(x)f(x)\ =\ x^{\frac{1}{\lambda}-1}(x+\lambda)^{-1-\frac{1}{\lambda}}\chi_{[-\lambda,1]}(x) (6.8)

is an eigenvector of ZZ with eigenvalue Ξ»\lambda.

(ii) As Hβˆ—β€‹f​(x)=∫x11t​f​(t)​𝑑tH^{*}f(x)=\int_{x}^{1}\frac{1}{t}f(t)dt, the eigenvalue equation becomes

G​(x)=(x+Ξ»)​f​(x),G(x)\ =\ (x+\lambda)f(x),

where

G​(x)=∫x11t​f​(t)​𝑑tG(x)\ =\ \int_{x}^{1}\frac{1}{t}f(t)dt

As G′​(x)=βˆ’1x​f​(x)G^{\prime}(x)=-\frac{1}{x}f(x), we get the differential equation

G​(x)=βˆ’x​(x+Ξ»)​G′​(x),G​(1)=0.G(x)\ =\ -x(x+\lambda)G^{\prime}(x),\quad G(1)=0. (6.9)

Solving for GG, we get

G​(x)=c​xβˆ’1λ​(x+Ξ»)1λ​χ{Gβ‰ 0}.G(x)\ =\ cx^{-\frac{1}{\lambda}}(x+\lambda)^{\frac{1}{\lambda}}\,\chi_{\{G\neq 0\}}.

The only solution on an interval that vanishes on the right end-point is the zero solution.

(iii) We know

Οƒe​(Z)=βˆ‚Ξ›βŠ†Ξ›=Οƒp​(Z)Β―βŠ†Οƒβ€‹(Z).\sigma_{e}(Z)=\partial\Lambda\subseteq\Lambda=\overline{\sigma_{p}(Z)}\subseteq\sigma(Z).

If Ξ»\lambda is a point in β„‚βˆ–Ξ›\mathbb{C}\setminus\Lambda, it must be a Fredholm point. By (i), it is not an eigenvalue of ZZ, and by (ii) it is not an eigenvalue of Zβˆ—Z^{*}. Therefore Zβˆ’Ξ»Z-\lambda has trivial kernel and cokernel, and closed range. Therefore it is invertible, and Ξ»\lambda is in the resolvent of ZZ. β–‘\Box

Not only are points on the stick of the lollipop eigenvalues, there is some additional smoothness. By a generalized eigenvector of order nn we mean a vector ff that satisfies (Zβˆ’Ξ»)n+1​f=0(Z-\lambda)^{n+1}f=0 but (Zβˆ’Ξ»)n​fβ‰ 0(Z-\lambda)^{n}f\neq 0.

We shall prove the case Ξ»=0\lambda=0 first.

Lemma 6.10.

At 0, the operator ZZ has generalized eigenvectors of all orders.

Proof: We want to show that if we let f0​(z)=1x2​eβˆ’1xf_{0}(z)=\frac{1}{x^{2}}e^{-\frac{1}{x}} from (6.7), then for every nβˆˆβ„•n\in{\mathbb{N}} there exists f∈L2f\in L^{2} so that

Z​fn+1=fn.Zf_{n+1}\ =\ f_{n}. (6.11)
Claim 6.12.

For every nβˆˆβ„•n\in{\mathbb{N}} there exists a polynomial pnp_{n} of degree 2​n+22n+2, with lowest order term of degree n+2n+2, so that the functions

fn​(x)=pn​(1x)​eβˆ’1xf_{n}(x)\ =\ p_{n}(\frac{1}{x})e^{-\frac{1}{x}}

satisfy (6.11).

We prove this by induction on nn. It is true when n=0n=0. Assume we have proved it up to level nn, and we want to prove it for n+1n+1. So we wish to solve the equation

Z​fn+1=fn=pn​(1x)​eβˆ’1xZf_{n+1}\ =\ f_{n}=p_{n}(\frac{1}{x})e^{-\frac{1}{x}} (6.13)

and show that the solution is of the form

fn+1​(x)=pn+1​(1x)​eβˆ’1x.f_{n+1}(x)\ =\ p_{n+1}(\frac{1}{x})e^{-\frac{1}{x}}. (6.14)

Writing Fn+1F_{n+1} for V​fn+1Vf_{n+1}, equation (6.13) is

(Hβˆ’Mx)​fn+1​(x)\displaystyle(H-M_{x})f_{n+1}(x) =\displaystyle\ =\ 1xβ€‹βˆ«0xfn+1​(t)​𝑑tβˆ’x​fn+1​(x)\displaystyle\frac{1}{x}\int_{0}^{x}f_{n+1}(t)dt-xf_{n+1}(x)
=\displaystyle= 1x​Fn+1​(x)βˆ’x​Fn+1′​(x)\displaystyle\frac{1}{x}F_{n+1}(x)-xF_{n+1}^{\prime}(x)
=\displaystyle= fn​(x).\displaystyle f_{n}(x).

This gives us the linear differential equation

Fn+1′​(x)βˆ’1x2​Fn+1​(x)=βˆ’1x​fn​(x).F_{n+1}^{\prime}(x)-\frac{1}{x^{2}}F_{n+1}(x)\ =\ -\frac{1}{x}f_{n}(x).

Multiply by the integrating factor e1xe^{\frac{1}{x}} to get

dd​x​[e1x​Fn+1​(x)]\displaystyle\frac{d}{dx}\left[e^{\frac{1}{x}}F_{n+1}(x)\right] =\displaystyle\ =\ βˆ’1x​e1x​fn​(x)\displaystyle-\frac{1}{x}e^{\frac{1}{x}}f_{n}(x)
=\displaystyle= βˆ’1x​pn​(1x).\displaystyle-\frac{1}{x}p_{n}(\frac{1}{x}).

Therefore

e1x​Fn+1​(x)=qn​(1x),e^{\frac{1}{x}}F_{n+1}(x)\ =\ q_{n}(\frac{1}{x}),

where qnq_{n} is a polynomial of degree 2​n+22n+2 that may have a constant term, and whose next lowest order term is of degree n+2n+2. This gives

fn+1​(x)\displaystyle f_{n+1}(x) =\displaystyle\ =\ dd​x​[eβˆ’1x​qn​(1x)]\displaystyle\frac{d}{dx}\left[e^{-\frac{1}{x}}q_{n}(\frac{1}{x})\right]
=\displaystyle= eβˆ’1x​[1x2​qn​(1x)βˆ’1x2​qn′​(1x)]\displaystyle e^{-\frac{1}{x}}\left[\frac{1}{x^{2}}q_{n}(\frac{1}{x})-\frac{1}{x^{2}}q_{n}^{\prime}(\frac{1}{x})\right]

Let

pn+1​(x)=x2​qn​(x)βˆ’x2​qn′​(x).p_{n+1}(x)\ =\ x^{2}q_{n}(x)-x^{2}q_{n}^{\prime}(x).

The degree of pn+1p_{n+1} is two higher than qnq_{n}, so it is 2​(n+1)+22(n+1)+2. There may be a term of order 22; the next lowest order term is n+3=(n+1)+2n+3=(n+1)+2. But as Z​f0=0Zf_{0}=0, one can subtract a multiple of f0f_{0} from fn+1f_{n+1} without changing (6.13), so we can assume that pn+1p_{n+1} has no term of order 22.

So we have proved Claim 6.12. As any function of the form (6.14) is in L2L^{2}, we are done. β–‘\Box

For points in the stick, a similar method works, but there are restrictions when requiring the generalized eigenvectors to be in L2L^{2}. Here is one result.

Lemma 6.15.

Let λ∈(βˆ’1,0)\lambda\in(-1,0). Then ZZ has a generalized eigenvalue of order 11 at Ξ»\lambda if and only if βˆ’23<Ξ»<0-\frac{2}{3}<\lambda<0.

Proof: Let s=βˆ’Ξ»s=-\lambda.

f0​(x)=(xβˆ’sx)1s​1x​(xβˆ’s)​χ[s,1].f_{0}(x)\ =\ \left(\frac{x-s}{x}\right)^{\frac{1}{s}}\frac{1}{x(x-s)}\chi_{[s,1]}. (6.16)

All the functions below are supported on [s,1][s,1]. We wish to find a function f1f_{1} that satisfies

(Z+s)​f1=f0.(Z+s)f_{1}\ =\ f_{0}. (6.17)

Writing F1F_{1} for V​f1​(x)=∫sxf1​(t)​𝑑tVf_{1}(x)=\int_{s}^{x}f_{1}(t)dt, this becomes

1x​F1βˆ’(xβˆ’s)​F1β€²=f0.\frac{1}{x}F_{1}-(x-s)F_{1}^{\prime}\ =\ f_{0}.

or

F1β€²βˆ’1x​(xβˆ’s)​F1=βˆ’1xβˆ’s​f0.F_{1}^{\prime}-\frac{1}{x(x-s)}F_{1}\ =\ -\frac{1}{x-s}f_{0}. (6.18)

An integrating factor for (6.18) is (xxβˆ’s)1s\left(\frac{x}{x-s}\right)^{\frac{1}{s}}. This yields

dd​x​[(xxβˆ’s)1s​F1]\displaystyle\frac{d}{dx}\left[\left(\frac{x}{x-s}\right)^{\frac{1}{s}}F_{1}\right] =\displaystyle\ =\ βˆ’(xxβˆ’s)1s​1xβˆ’s​f0\displaystyle-\left(\frac{x}{x-s}\right)^{\frac{1}{s}}\frac{1}{x-s}f_{0}
=\displaystyle= βˆ’1x​(xβˆ’s)2.\displaystyle-\frac{1}{x(x-s)^{2}}.

Integrating, we get

(xxβˆ’s)1s​F1=1s2​log⁑xβˆ’sx+1s​1xβˆ’s+c.\left(\frac{x}{x-s}\right)^{\frac{1}{s}}F_{1}\ =\ \frac{1}{s^{2}}\log\frac{x-s}{x}+\frac{1}{s}\frac{1}{x-s}+c.

Dividing through by the integrating factor and differentiating, we get

f1​(x)\displaystyle f_{1}(x) =\displaystyle\ =\ 1s2​[(xβˆ’sx)1sβˆ’1​1x2​log⁑xβˆ’sx+(xβˆ’sx)1s​sx​(xβˆ’s)]\displaystyle\frac{1}{s^{2}}\left[\left(\frac{x-s}{x}\right)^{\frac{1}{s}-1}\frac{1}{x^{2}}\log\frac{x-s}{x}+\left(\frac{x-s}{x}\right)^{\frac{1}{s}}\frac{s}{x(x-s)}\right]
+1s​[βˆ’1(xβˆ’s)2​(xβˆ’sx)1s+1xβˆ’s​(xβˆ’sx)1sβˆ’1​1x2]\displaystyle+\frac{1}{s}\left[-\frac{1}{(x-s)^{2}}\left(\frac{x-s}{x}\right)^{\frac{1}{s}}+\frac{1}{x-s}\left(\frac{x-s}{x}\right)^{\frac{1}{s}-1}\frac{1}{x^{2}}\right]
+c​[(xβˆ’sx)1sβˆ’1​1x2].\displaystyle+c\left[\left(\frac{x-s}{x}\right)^{\frac{1}{s}-1}\frac{1}{x^{2}}\right].

We can choose c=0c=0, since it is the coefficient of f0f_{0}. This gives

f1​(x)=(xβˆ’sx)1s​[1s2​1x​(xβˆ’s)​log⁑xβˆ’sx+1βˆ’ss​1x​(xβˆ’s)2].f_{1}(x)\ =\ \left(\frac{x-s}{x}\right)^{\frac{1}{s}}\left[\frac{1}{s^{2}}\frac{1}{x(x-s)}\log\frac{x-s}{x}+\frac{1-s}{s}\frac{1}{x(x-s)^{2}}\right]. (6.19)

Examining this expression, we see that f1f_{1} is smooth on (s,1](s,1], and the first term

1s2​(xβˆ’s)1sβˆ’1x1s+1​log⁑xβˆ’sx\frac{1}{s^{2}}\frac{(x-s)^{\frac{1}{s}-1}}{x^{\frac{1}{s}+1}}\log\frac{x-s}{x}

vanishes at ss for every s<1s<1. However the second term

1βˆ’ss​(xβˆ’s)1sβˆ’2x1s+1\frac{1-s}{s}\ \frac{(x-s)^{\frac{1}{s}-2}}{x^{\frac{1}{s}+1}}

has a singularity that grows like (xβˆ’s)1sβˆ’2(x-s)^{\frac{1}{s}-2}. This is integrable for every s<1s<1, but it is only in L2​[s,1]L^{2}[s,1] for s<23s<\frac{2}{3}. So we have shown that (6.17) has a solution f1f_{1} in L2L^{2} if and only if Ξ»>βˆ’23\lambda>-\frac{2}{3}. β–‘\Box

One can repeat the argument of Lemma 6.15 to get higher order generalized eigenvectors, as Ξ»\lambda gets closer to 0.

Lemma 6.20.

Let mβ‰₯1m\geq 1. Let Ξ»\lambda lie in the interval (βˆ’22​m+1,0)(-\frac{2}{2m+1},0). Then ZZ has generalized eigenvectors up to order mm at Ξ»\lambda.

Proof: We shall inductively find functions fnf_{n} satisfying (Zβˆ’Ξ»)​fn+1=fn(Z-\lambda)f_{n+1}=f_{n}, with f0f_{0} as in (6.16). Let s=βˆ’Ξ»s=-\lambda, and write

Φ​(x)=(xβˆ’sx)1s​χ[s,1]​(x).\Phi(x)\ =\ \left(\frac{x-s}{x}\right)^{\frac{1}{s}}\chi_{[s,1]}(x).

Then we have

f0​(x)\displaystyle f_{0}(x) =\displaystyle\ =\ Φ​(x)​1x​(xβˆ’s)\displaystyle\Phi(x)\frac{1}{x(x-s)}
f1​(x)\displaystyle f_{1}(x) =\displaystyle\ =\ Φ​(x)​[1s2​1x​(xβˆ’s)​log⁑xβˆ’sx+1βˆ’ss​1x​(xβˆ’s)2].\displaystyle\Phi(x)\left[\frac{1}{s^{2}}\frac{1}{x(x-s)}\log\frac{x-s}{x}+\frac{1-s}{s}\frac{1}{x(x-s)^{2}}\right].

Writing Fn+1F_{n+1} for V​fn+1=∫sxfn+1​(t)​𝑑tVf_{n+1}=\int_{s}^{x}f_{n+1}(t)dt, we want to solve

1x​Fn+1​(x)βˆ’(xβˆ’s)​Fn+1′​(x)=fn​(x).\frac{1}{x}F_{n+1}(x)-(x-s)F_{n+1}^{\prime}(x)\ =\ f_{n}(x). (6.21)

After multiplying by the integrating factor 1/Ξ¦1/\Phi, we have

dd​x​[1Φ​Fn+1]=βˆ’1Φ​1xβˆ’s​fn​(x).\frac{d}{dx}\left[\frac{1}{\Phi}F_{n+1}\right]\ =\ -\frac{1}{\Phi}\frac{1}{x-s}f_{n}(x). (6.22)
Claim 6.23.

There are constants MnM_{n} such that the functions fnf_{n} satisfy

|fn​(x)|≀Mn​(xβˆ’s)1sβˆ’nβˆ’1βˆ€x∈(s,1].|f_{n}(x)|\ \leq\ M_{n}(x-s)^{\frac{1}{s}-n-1}\qquad\forall x\in(s,1]. (6.24)

Proof of Claim 6.23: By induction on nn. It is true when n=0n=0. Assume it is true up to nn. From (6.22) we get for x∈(s,1]x\in(s,1]:

1Φ​(x)​Fn+1​(x)=∫x11Φ​(t)​1tβˆ’s​fn​(t)​𝑑t+cn.\frac{1}{\Phi(x)}F_{n+1}(x)\ =\ \int_{x}^{1}\frac{1}{\Phi(t)}\frac{1}{t-s}f_{n}(t)dt+c_{n}. (6.25)

By the inductive hypothesis, the integrand in (6.25) is O​(tβˆ’s)βˆ’nβˆ’2O(t-s)^{-n-2}, so the integral is O​(xβˆ’s)βˆ’nβˆ’1O(x-s)^{-n-1}. So Fn+1F_{n+1} satisfies

Fn+1​(x)=cn​Φ​(x)+O​(xβˆ’s)1sβˆ’nβˆ’1.F_{n+1}(x)\ =\ c_{n}\Phi(x)+O(x-s)^{\frac{1}{s}-n-1}. (6.26)

From (6.21) we have

fn+1​(x)=1x​(xβˆ’s)​Fn+1​(x)βˆ’1xβˆ’s​fn​(x).f_{n+1}(x)\ =\ \frac{1}{x(x-s)}F_{n+1}(x)-\frac{1}{x-s}f_{n}(x). (6.27)

When we use (6.26) for Fn+1F_{n+1}, we get

fn+1​(x)=cn​f0​(x)βˆ’1xβˆ’s​fn​(x)+O​(xβˆ’s)1sβˆ’nβˆ’2.f_{n+1}(x)\ =\ c_{n}f_{0}(x)-\frac{1}{x-s}f_{n}(x)+O(x-s)^{\frac{1}{s}-n-2}.

Now the claim follows from the inductive hypothesis on fnf_{n}. β–‘\Box

It follows from (6.27) that that fmf_{m} is continuous on (s,1](s,1], and Claim 6.23 shows that its singularity at ss is of order (xβˆ’s)1sβˆ’mβˆ’1(x-s)^{\frac{1}{s}-m-1}. This means fmf_{m} is in L2L^{2} provided 1sβˆ’mβˆ’1>βˆ’12,\frac{1}{s}-m-1\ >\ -\frac{1}{2}, which is the same as s<22​m+1s<\frac{2}{2m+1}. β–‘\Box

For later use, let us note that if you track the constants MnM_{n} in Claim 6.23, you can show:

Lemma 6.28.

In Claim 6.23 one can take M0=1(s1+1/s)M_{0}=\frac{1}{(s^{1+1/s})} and the consants MnM_{n} satisfy

Mn+1≀Mn​(1+M0n+1).M_{n+1}\leq M_{n}\left(1+\frac{M_{0}}{n+1}\right).
Lemma 6.29.

Let mβ‰₯1m\geq 1. On the interval (βˆ’22​m+1,0)(-\frac{2}{2m+1},0) one can choose the generalized eigenvectors of order nn of Zβˆ’Ξ»Z-\lambda continuously in Ξ»\lambda, for every n≀mn\leq m, and satisfying (Zβˆ’Ξ»)​fΞ»,n=fΞ»,nβˆ’1(Z-\lambda)f_{\lambda,n}=f_{\lambda,n-1}, for every 1≀n≀m1\leq n\leq m.

Proof: Let us write fΞ»,nf_{\lambda,n} for the choice of generalized eigenvector of order nn at Ξ»\lambda. Write

Φ​(Ξ»,x)=(x+Ξ»x)βˆ’1λ​χ[βˆ’Ξ»,1]​(x).\Phi(\lambda,x)\ =\ \left(\frac{x+\lambda}{x}\right)^{-\frac{1}{\lambda}}\chi_{[-\lambda,1]}(x).

We have

fΞ»,0​(x)=1x​(x+Ξ»)​Φ​(Ξ»,x).f_{\lambda,0}(x)\ =\ \frac{1}{x(x+\lambda)}\Phi(\lambda,x).

On every compact subset KK, of (βˆ’1,0)(-1,0) the functions {fΞ»,0:λ∈K}\{f_{\lambda,0}:\lambda\in K\} are uniformly bounded, and limΞ»β€²β†’Ξ»fΞ»β€²,0​(x)=fΞ»,0​(x)\lim_{\lambda^{\prime}\to\lambda}f_{\lambda^{\prime},0}(x)=f_{\lambda,0}(x) a.e., so the map λ↦fΞ»,0\lambda\mapsto f_{\lambda,0} is continuous as a map from (βˆ’22​m+1,0)(-\frac{2}{2m+1},0) into L2L^{2}.

For higher nn, we find fn+1f_{n+1} as in Lemma 6.20. At each stage, we take the constant cnc_{n} in (6.26) to be 0. (We can do this because cn​fΞ»,0c_{n}f_{\lambda,0} will be in the kernel of Zβˆ’Ξ»Z-\lambda.) This gives us

fΞ»,n+1​(x)=1x​(xβˆ’s)​Φ​(Ξ»,x)β€‹βˆ«x11Φ​(Ξ»,t)​1tβˆ’s​fΞ»,n​(t)​𝑑tβˆ’1xβˆ’s​fΞ»,n​(x).f_{\lambda,n+1}(x)\ =\ \frac{1}{x(x-s)}\Phi(\lambda,x)\int_{x}^{1}\frac{1}{\Phi(\lambda,t)}\frac{1}{t-s}f_{\lambda,n}(t)dt-\frac{1}{x-s}f_{\lambda,n}(x).

Moreover we have fΞ»,n​(x)=0f_{\lambda,n}(x)=0 if x<βˆ’Ξ»x<-\lambda and

|fΞ»,n​(x)|≀Mn,λ​(x+Ξ»)βˆ’1Ξ»βˆ’nβˆ’1,x>βˆ’Ξ».|f_{\lambda,n}(x)|\ \leq\ M_{n,\lambda}(x+\lambda)^{-\frac{1}{\lambda}-n-1},\qquad x>-\lambda.

By Lemma 6.28, we have that each Mn,Ξ»M_{n,\lambda} can be chosen uniformly in Ξ»\lambda for Ξ»\lambda in (βˆ’22​m+1,0)(-\frac{2}{2m+1},0). For any interval II of length Ξ΄\delta, we get

∫I|fΞ»,n​(x)|2​𝑑x\displaystyle\int_{I}|f_{\lambda,n}(x)|^{2}dx ≀\displaystyle\ \leq\ Mnβ€‹βˆ«0Ξ΄|x+Ξ»|βˆ’2Ξ»βˆ’2​nβˆ’2\displaystyle M_{n}\int_{0}^{\delta}|x+\lambda|^{-\frac{2}{\lambda}-2n-2}
=\displaystyle= Mn​1βˆ’2Ξ»βˆ’2​nβˆ’1β€‹Ξ΄βˆ’2Ξ»βˆ’2​nβˆ’1.\displaystyle M_{n}\frac{1}{-\frac{2}{\lambda}-2n-1}\delta^{-\frac{2}{\lambda}-2n-1}.

Therefore, as Ξ»\lambda ranges over any compact subset of (βˆ’22​m+1,0)(-\frac{2}{2m+1},0), the functions |fΞ»,n​(x)|2|f_{\lambda,n}(x)|^{2} are uniformly integrable in xx. So, by the Vitali convergence theorem, the map

(βˆ’22​m+1,0)\displaystyle(-\frac{2}{2m+1},0) β†’\displaystyle\ \to\ L2\displaystyle L^{2}
Ξ»\displaystyle\lambda ↦\displaystyle\mapsto fΞ»,n\displaystyle f_{\lambda,n}

is continuous. β–‘\Box

These lemmas say that operators in the closed algebra generated by ZZ have certain smoothness properties when mapped by Οƒ\sigma into A​(Ξ›)A(\Lambda). The functions get smoother as we get closer to 0.

Theorem 6.30.

Let XX be in the norm-closed algebra generated by ZZ. Then θ​(X)\theta(X) is CmC^{m} on (βˆ’22​m+1,0)(-\frac{2}{2m+1},0).

Proof: Let θ​(X)=Ο•βˆˆA​(Ξ›)\theta(X)=\phi\in A(\Lambda). Let pjp_{j} be a sequence of polynomials so that β€–pj​(Z)βˆ’Xβ€–β†’0\|p_{j}(Z)-X\|\to 0. It follows from Theorem 3.17 that pjp_{j} converges to Ο•\phi uniformly on Ξ›\Lambda.

Case: m=1m=1. Let fΞ»,nf_{\lambda,n} be as in Lemma 6.29. For any polynomial pp, we have

⟨p​(Z)​fΞ»,0,fΞ»,0⟩=β€–fΞ»,0β€–2​p​(Ξ»).\langle p(Z)f_{\lambda,0},f_{\lambda,0}\rangle\ =\ \|f_{\lambda,0}\|^{2}p(\lambda).

Moreover,

p​(Z)​fΞ»,1=p​(Ξ»)​fΞ»,1+p′​(Ξ»)​fΞ»,0.p(Z)f_{\lambda,1}\ =\ p(\lambda)f_{\lambda,1}+p^{\prime}(\lambda)f_{\lambda,0}.

Let gλ1g^{1}_{\lambda} be the linear combination of fλ,0f_{\lambda,0} and fλ,1f_{\lambda,1} that satisfies ⟨fλ,0,gλ1⟩=1\langle f_{\lambda,0},g^{1}_{\lambda}\rangle=1 and ⟨fλ,1,gλ1⟩=0\langle f_{\lambda,1},g^{1}_{\lambda}\rangle=0. Then

⟨p​(Z)​fΞ»,1,gΞ»1⟩=p′​(Ξ»).\langle p(Z)f_{\lambda,1},g^{1}_{\lambda}\rangle\ =\ p^{\prime}(\lambda).

So as functions on (βˆ’23,0)(-\frac{2}{3},0), we get that pj′​(Ξ»)p_{j}^{\prime}(\lambda) converges to some function Οˆβ€‹(Ξ»)=⟨X​fΞ»,1,gΞ»1⟩\psi(\lambda)=\langle Xf_{\lambda,1},g^{1}_{\lambda}\rangle.

Claim 6.31.

For all xx in (βˆ’23,0)(-\frac{2}{3},0), we have ϕ′​(x)=Οˆβ€‹(x)\phi^{\prime}(x)=\psi(x), and ψ\psi is continuous.

We have gΞ»1=aΞ»,0​fΞ»,0+aΞ»,1​fΞ»,1g^{1}_{\lambda}=a_{\lambda,0}f_{\lambda,0}+a_{\lambda,1}f_{\lambda,1}, where the coefficients aΞ»,0a_{\lambda,0} and aΞ»,1a_{\lambda,1} solve the linear system

(⟨fΞ»,0,fΞ»,0⟩⟨fΞ»,0,fΞ»,1⟩⟨fΞ»,1,fΞ»,0⟩⟨fΞ»,1,fΞ»,1⟩)​(aΞ»,0aΞ»,1)=(10).\begin{pmatrix}\langle f_{\lambda,0},f_{\lambda,0}\rangle&\langle f_{\lambda,0},f_{\lambda,1}\rangle\\ \langle f_{\lambda,1},f_{\lambda,0}\rangle&\langle f_{\lambda,1},f_{\lambda,1}\rangle\end{pmatrix}\begin{pmatrix}a_{\lambda,0}\\ a_{\lambda,1}\end{pmatrix}\ =\ \begin{pmatrix}1\\ 0\end{pmatrix}.

By Lemma 6.29, since fΞ»,0f_{\lambda,0} and fΞ»,1f_{\lambda,1} are continuous in Ξ»\lambda, so as they are linearly independent, we have that gΞ»1g^{1}_{\lambda} is also continuous in Ξ»\lambda. Therefore ψ\psi is continuous, and pjβ€²p_{j}^{\prime} converges to ψ\psi locally uniformly on (βˆ’23,0)(-\frac{2}{3},0). As βˆ«βˆ’13xpj′​(t)​𝑑t=ϕ​(x)βˆ’Ο•β€‹(βˆ’13)\int_{-\frac{1}{3}}^{x}p_{j}^{\prime}(t)dt=\phi(x)-\phi(-\frac{1}{3}) converges to βˆ«βˆ’13xΟˆβ€‹(t)​𝑑t\int_{-\frac{1}{3}}^{x}\psi(t)dt, we get that ϕ′​(x)=Οˆβ€‹(x)\phi^{\prime}(x)=\psi(x). β–‘\Box

We have shown that Ο•\phi is in C1​(βˆ’23,0)C^{1}(-\frac{2}{3},0). A similar argument with higher derivatives proves that Ο•\phi is in Cm​(βˆ’22​m+1,0)C^{m}(-\frac{2}{2m+1},0). β–‘\Box

Of course one can also find generalized eigenvectors for ZZ of all orders at points in 𝔻​(1,1)\mathbb{D}(1,1), but we already know that θ​(X)\theta(X) is analytic on 𝔻​(1,1)\mathbb{D}(1,1) for every Xβˆˆπ’œX\in{\mathcal{A}}.

7 Open Questions

Question 7.1.

Is there a good description of π’¦π’œ{\mathcal{K}}_{{\mathcal{A}}}, the compact operators in π’œ{\mathcal{A}}?

Question 7.2.

Is Cβˆ—β€‹(Z)=Cβˆ—β€‹(π’œ)C^{*}(Z)=C^{*}({\mathcal{A}})? To prove this, it is sufficient to show that Cβˆ—β€‹(Z)C^{*}(Z) is irreducible, since then it would contain all the compacts, and its quotient by the compacts would be all of C​(βˆ‚Ξ›)C(\partial\Lambda).

Question 7.3.

Do the eigenvectors of ZZ span L2L^{2}?

References

  • [1] Jim Agler and JohnΒ E. McCarthy. The asymptotic MΓΌntz-SzΓ‘sz theorem. Preprint, 2022.
  • [2] Jim Agler and JohnΒ E. McCarthy. A generalization of Hardy’s operator and an asymptotic MΓΌntz-SzΓ‘sz theorem. Preprint, 2022.
  • [3] Francesca Arici and Bram Mesland. Toeplitz extensions in noncommutative topology and mathematical physics. In Geometric methods in physics XXXVIII, Trends Math., pages 3–29. BirkhΓ€user/Springer, Cham, [2020] Β©2020.
  • [4] Laura Brake and Wilhelm Winter. The Toeplitz algebra has nuclear dimension one. Bull. Lond. Math. Soc., 51(3):554–562, 2019.
  • [5] M.Β S. Brodskii. On a problem of I. M. Gelfand. Uspehi Mat. Nauk (N.S.), 12(2(74)):129–132, 1957.
  • [6] W.Β F. Donoghue, Jr. The lattice of invariant subspaces of a completely continuous quasi-nilpotent transformation. Pacific J. Math., 7:1031–1035, 1957.
  • [7] R.Β G. Douglas. Banach Algebra Techniques in Operator Theory. Academic Press, New York, 1972.
  • [8] Philip Easo, Esperanza Garijo, Sarunas Kaubrys, David Nkansah, Martin Vrabec, David Watt, Cameron Wilson, Christian BΓΆnicke, Samuel Evington, Marzieh Forough, Sergio GirΓ³nΒ Pacheco, Nicholas Seaton, Stuart White, MichaelΒ F. Whittaker, and Joachim Zacharias. The Cuntz-Toeplitz algebras have nuclear dimension one. J. Funct. Anal., 279(7):108690, 14, 2020.
  • [9] Franco Gallone. Hilbert space and quantum mechanics. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2015.
  • [10] Shin Hayashi. Topological invariants and corner states for Hamiltonians on a three-dimensional lattice. Comm. Math. Phys., 364(1):343–356, 2018.
  • [11] Peter Lax. Functional Analysis. Wiley, 2002.
  • [12] Toufik Mansour and Matthias Schork. Commutation relations, normal ordering, and Stirling numbers. Discrete Mathematics and its Applications (Boca Raton). CRC Press, Boca Raton, FL, 2016.
  • [13] S.Β N. Mergelyan. Uniform approximations of functions of a complex variable. Uspehi Matem. Nauk (N.S.), 7(2(48)):31–122, 1952.
  • [14] Toshio Oshima. Fractional calculus of Weyl algebra and Fuchsian differential equations, volumeΒ 28 of MSJ Memoirs. Mathematical Society of Japan, Tokyo, 2012.
  • [15] Efton Park. Index theory and Toeplitz algebras on certain cones in 𝐙2{\bf Z}^{2}. J. Operator Theory, 23(1):125–146, 1990.
  • [16] Efton Park and Claude Schochet. On the KK-theory of quarter-plane Toeplitz algebras. Internat. J. Math., 2(2):195–204, 1991.
  • [17] Heydar Radjavi and Peter Rosenthal. Invariant subspaces. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 77. Springer-Verlag, New York-Heidelberg, 1973.
  • [18] Adam Rennie, David Robertson, and Aidan Sims. PoincarΓ© duality for Cuntz-Pimsner algebras. Adv. Math., 347:1112–1172, 2019.
  • [19] Konrad SchmΓΌdgen. An invitation to unbounded representations of βˆ—*-algebras on Hilbert space, volume 285 of Graduate Texts in Mathematics. Springer, Cham, [2020].