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

Resolvent algebra in Fock-Bargmann representation

Wolfram Bauer and Robert Fulsche111Institut für Analysis, Welfengarten 1, 30167 Hannover, Germany. Email: [email protected] (W. Bauer) and [email protected] (R. Fulsche)
Abstract

The resolvent algebra (X,σ)\mathcal{R}(X,\sigma) associated to a symplectic space (X,σ)(X,\sigma) was introduced by D. Buchholz and H. Grundling as a convenient model of the canonical commutation relation (CCR) in quantum mechanics. We first study a representation of (n,σ)\mathcal{R}(\mathbb{C}^{n},\sigma) with the standard symplectic form σ\sigma inside the full Toeplitz algebra over the Fock-Bargmann space. We prove that (n,σ)\mathcal{R}(\mathbb{C}^{n},\sigma) itself is a Toeplitz algebra. In the sense of R. Werner’s correspondence theory we determine its corresponding shift-invariant and closed space of symbols. Finally, we discuss a representation of the resolvent algebra (,σ~)\mathcal{R}(\mathcal{H},\tilde{\sigma}) for an infinite dimensional symplectic separable Hilbert space (,σ~)(\mathcal{H},\tilde{\sigma}). More precisely, we find a representation of (,σ~)\mathcal{R}(\mathcal{H},\tilde{\sigma}) inside the full Toeplitz algebra over the Fock-Bargmann space in infinitely many variables.
keywords: canonical commutation relation, Toeplitz CC^{*} algebra, correspondence theorem, Gaussian measure on Hilbert space
Mathematical Subject Classification 2020: Primary: 47L80; Secondary: 46L60

1 Introduction

The classical CCR algebra (see [5]) provides a standard CC^{*} algebraic model for the canonical commutation relation (CCR) in quantum mechanics. In typical applications, however, these algebras have significant disadvantages. In particular, CCR algebras in general do not contain bounded functions of the Hamiltonian (physical observables). Only in rare and physically less relevant cases they are stable under dynamics. More precisely, it was noticed in [11] that time evolution of the Hamiltonian 𝐇:=Δ+V\mathbf{H}:=-\Delta+V on L2(n)L^{2}(\mathbb{R}^{n}) in the standard Schrödinger representation gives rise to \ast-automorphisms of the CCR algebra over the standard symplectic space 2nTn\mathbb{R}^{2n}\cong T^{*}\mathbb{R}^{n} only in the trivial case of an identically vanishing potential function VV. In order to resolve such problems, D. Buchholz and H. Grundling have proposed a new approach to CCR by introducing the resolvent algebra (X,σ)\mathcal{R}(X,\sigma) associated to a symplectic space (X,σ)(X,\sigma) in [7]. This algebra abstractly is defined through certain algebraic relations that mimic properties of the resolvents of the (i.g. unbounded) canonical operators (see (2.1)).

In the present paper, we first consider complex nn-space n2n\mathbb{C}^{n}\cong\mathbb{R}^{2n} equipped with the standard symplectic structure. As is well-known, in this setting the CCR algebra is identified with the CC^{*} algebra generated by Toeplitz operators with symbols in the space TP of trigonometric polynomials. It is an interesting observation that this Toeplitz CC^{*} algebra coincides with the linear closure of Toeplitz operators with symbols in TP (see [8] for a precise statement and further extensions of these results). In a similar manner, our goal is to study the resolvent algebra in the Fock-Bargmann representation and to describe it as a concrete algebra generated by Toeplitz operators with shift-invariant symbol space. Being a shift-invariant CC^{*} algebra, it has turned out that a convenient mathematical framework for the analysis of the resolvent algebra is R. Werner’s quantum harmonic analysis [20] and its extension by R. Fulsche in [13] to Toeplitz operator theory on the Fock-Bargmann space. A particularly useful tool is the correspondence theorem (Theorem 4.5) in [20, 13] which ensures the existence and uniqueness of a closed and shift-invariant space of bounded uniformly continuous functions on n\mathbb{C}^{n}, which are the symbols of Toeplitz operators linearly generating the resolvent algebra.

In dimension n=1n=1 we show that the space corresponding to the resolvent algebra (,σ)\mathcal{R}(\mathbb{C},\sigma) is the uniform closure of the classical resolvent functions. However, in higher dimensions n>1n>1 we only can prove a weaker result (see Proposition 4.26). Nevertheless, we show that (n,σ)\mathcal{R}(\mathbb{C}^{n},\sigma) coincides with a CC^{*} algebra generated by a concrete set of Toeplitz operators with bounded symbols. It remains an open problem whether or not the closure of classical resolvent functions forms the corresponding space to the resolvent algebra in every complex dimension nn\in\mathbb{N} and in the sense of Theorem 4.5.

In the second part of the paper, we replace n\mathbb{C}^{n} by a separable infinite dimensional complex symplectic Hilbert spaces (,σ~)(\mathcal{H},\tilde{\sigma}). In [15, 16] the authors have proposed two definitions of Toeplitz operators on the infinitely many variable Fock-Bargmann space. Here, we follow the measure theoretical approach and consider CC^{*} algebras generated by Toeplitz operators over \mathcal{H}. Due to the non-nulcearity of \mathcal{H} (as well as of the space of entire functions over \mathcal{H} with the compact-open topology) as topological vector spaces and caused by features of the measure theory on infinite dimensional spaces, a variety of new effects can be observed in the theory of Toeplitz operators. In particular, in this setting a correspondence theorem so far is unknown, even though it may exist in a suitable formulation. Our main result of the last section (Theorem 5.4) states that there is a representation of the resolvent algebra corresponding to a symplectic Hilbert space 1/2\mathcal{H}_{1/2} with Hilbert-Schmidt embedding 1/2\mathcal{H}_{1/2}\hookrightarrow\mathcal{H} inside the full Toeplitz algebra over \mathcal{H}.

The structure of the paper is as follows: In Section 2 we recall the definition of the CCR and resolvent algebra [7]. We restate a well-known representation of the CCR algebra associated to n\mathbb{C}^{n} with the standard symplectic structure as a CC^{*} algebra generated by Toeplitz operators.

Section 3 provides some basic facts on Fock-Bargmann spaces F𝐭2F^{2}_{\bf t} over n\mathbb{C}^{n} and Toeplitz operators acting on F𝐭2F^{2}_{\bf t}. We define the Berezin transform and then express Weyl operators in form of Toeplitz operators with bounded symbols. Most of the material is standard and further details on the role of Toeplitz operators in quantum mechanics can be found in [3, 8].

Section 4 describes the Fock-Bargmann representation of the resolvent algebra. We present the correspondence theorem [20] in this setting (see [13]). In passing, we mention some applications to the analysis of Toeplitz operators on Fock-Bargmann spaces. We discuss the unique shift-invariant closed symbol space corresponding to the resolvent algebra. An essential ingredient to the analysis is an integral form of the Berezin transform on products of resolvents (Proposition 4.12). Moreover, in our proofs we essentially use a specific compactification of the affine Grassmannian which was introduced in [19].

Finally, in Section 5, we discuss the resolvent algebra associated to an infinite dimensional symplectic separable Hilbert space in the framework of Toeplitz operators on the Fock-Bargmann space of Gaussian square integrable entire functions in infinitely many variables.
Acknowledgement: We thank Prof. R. Werner for many enlightening discussions on quantum harmonic analysis, the correspondence theorem and the structure of the resolvent algebra.

2 CCR and Resolvent Algebra

Let (X,σ)(X,\sigma) be a symplectic vector space. Consider a Hilbert space \mathcal{H} and a real linear map ϕ\phi from (X,σ)(X,\sigma) into the space of (unbounded) self-adjoint operators on \mathcal{H}. Let us assume that all operators ϕ(f)\phi(f) are essentially self-adjoint on a common domain 𝒟\mathcal{D} which forms a core for each ϕ(f)\phi(f). Recall that the canonical commutation relation (CCR) have the form:

[ϕ(f),ϕ(g)]=iσ(f,g),f,gX.\big{[}\phi(f),\phi(g)\big{]}=i\sigma(f,g),\quad f,g\in X. (2.1)

We call ϕ(f)\phi(f) and ϕ(g)\phi(g) canonical operators. Especially in an algebraic setup the analysis of unbounded operators involves some difficulties (compare e.g. [17, Chapters VIII.5 and VIII.6]). Therefore, one may look for CC^{*} algebraic models which suitably encode (CCR). One standard idea consists in replacing ϕ(f)\phi(f) above by suitable bounded functions of ϕ(f)\phi(f). A classical approach due to H. Weyl amounts in considering the CCR algebra generated by the unitary operators exp(iϕ(f))\exp(i\phi(f)), fXf\in X. The modified CCR are:

exp(iϕ(f))exp(iϕ(g))=eiσ(f,g)exp(iϕ(f+g))\exp\big{(}i\phi(f)\big{)}\exp\big{(}i\phi(g)\big{)}=e^{-i\sigma(f,g)}\exp\big{(}i\phi(f+g)\big{)}

and self-adjointness of ϕ(f)\phi(f) implies that exp(iϕ(f))=exp(iϕ(f))\exp(i\phi(f))^{\ast}=\exp(i\phi(-f)).

More abstractly, one is interested in the CC^{\ast} algebra CCR(X,σ)\textup{CCR}(X,\sigma) generated by the relations

W(f)W(g)\displaystyle W(f)W(g) =eiσ(f,g)W(f+g),f,gX\displaystyle=e^{-i\sigma(f,g)}W(f+g),\hskip 21.52771ptf,g\in X
W(f)\displaystyle W(f)^{\ast} =W(f).\displaystyle=W(-f).

Such CC^{\ast} algebras are usually called CCR algebras, see [5].

With z,wnz,w\in\mathbb{C}^{n} put w,z:=w1z¯1++wnz¯n\langle w,z\rangle:=w_{1}\overline{z}_{1}+\ldots+w_{n}\overline{z}_{n} and consider the standard symplectic space (n,σ)(\mathbb{C}^{n},\sigma) with symplectic form

σ(w,z):=Imw,z.\sigma(w,z):=\operatorname{Im}\langle w,z\rangle.

One obtains a representation of CCR(n,σ)\textup{CCR}(\mathbb{C}^{n},\sigma) as operators acting on the Fock-Bargmann space F12F_{1}^{2} (see [8] and Section 3 for precise definitions). In fact, with our previous notation we put:

ϕ(z):=T2σ(,z),\phi(z):=T_{2\sigma(\cdot,z)},

where TfT_{f} is a Toeplitz operator with complex valued symbol ff on n\mathbb{C}^{n}. The corresponding CCR algebra is generated by the unitary Weyl operators (3.3) below which satisfy

Wz=exp(iT2σ(,z)),W_{z}=\exp\big{(}iT_{2\sigma(\cdot,z)}\big{)},

i.e. iT2σ(,z)iT_{2\sigma(\cdot,z)} is the generator of the unitary one-parameter group (Wtz)t(W_{tz})_{t\in\mathbb{R}}. Hence,

CCR(n,σ)C(Wz:zn).\textup{CCR}(\mathbb{C}^{n},\sigma)\cong C^{\ast}\big{(}W_{z}\>:\>~{}z\in\mathbb{C}^{n}\big{)}. (2.2)

The algebra CCR(n,σ)\textup{CCR}(\mathbb{C}^{n},\sigma) is a classical object and in a more general setup it is discussed in [5]. In particular, the Fock-Bargmann representation of CCR(n,σ)\textup{CCR}(\mathbb{C}^{n},\sigma) maps into the full Toeplitz algebra 𝒯\mathcal{T}, i.e. the CC^{*} algebra generated by Toeplitz operators with arbitrary bounded measurable symbols.

Theorem 2.1 (L. A. Coburn [8]).

It holds

CCR(n,σ){Tϕ:ϕTP}¯,\textup{CCR}(\mathbb{C}^{n},\sigma)\cong\overline{\{T_{\phi}\>:\>\phi\in TP\}},

where TP is the space of trigonometric polynomials:

TP:=span{wexp(iσ(w,z)):zn}L(n).TP:=\operatorname{span}\big{\{}w\mapsto\exp\big{(}i\sigma(w,z)\big{)}\>:\>z\in\mathbb{C}^{n}\big{\}}\subset L^{\infty}(\mathbb{C}^{n}).

Here we write ¯\overline{\mathcal{M}} for the operator norm closure of a given set (F12)\mathcal{M}\subset\mathcal{L}(F_{1}^{2}).

As was mentioned in the introduction, there are certain drawbacks of using CCR algebras as quantum mechanical models. A new algebraic framework was suggested in [7, 6] which partly overcomes the above mentioned obstructions. Therein, the authors consider the resolvent algebra, which is the CC^{\ast} algebra generated by the resolvents of the (in general unbounded) operators ϕ(f)\phi(f).

Definition 2.2 (D. Buchholz, H. Grundling [7]).

Given a symplectic space (X,σ)(X,\sigma) we define 0(X,σ)\mathcal{R}_{0}(X,\sigma) as the universal unital \ast-algebra generated by the set {R(λ,f):λ{0},fX}\{R(\lambda,f)\>:\>~{}\lambda\in\mathbb{R}\setminus\{0\},\>f\in X\} together with the relations:

  • (1)

    R(λ,0)=iλ1R(\lambda,0)=-\frac{i}{\lambda}1,

  • (2)

    R(λ,f)=R(λ,f)R(\lambda,f)^{\ast}=R(-\lambda,f),

  • (3)

    νR(νλ,νf)=R(λ,f)\nu R(\nu\lambda,\nu f)=R(\lambda,f),

  • (4)

    R(λ,f)R(μ,f)=i(μλ)R(λ,f)R(μ,f)R(\lambda,f)-R(\mu,f)=i(\mu-\lambda)R(\lambda,f)R(\mu,f),

  • (5)

    [R(λ,f),R(μ,g)]=iσ(f,g)R(λ,f)R(μ,g)2R(λ,f)[R(\lambda,f),R(\mu,g)]=i\sigma(f,g)R(\lambda,f)R(\mu,g)^{2}R(\lambda,f),

  • (6)

    R(λ,f)R(μ,g)=R(λ+μ,f+g)[R(λ,f)+R(μ,g)+iσ(f,g)R(λ,f)2R(μ,g)]R(\lambda,f)R(\mu,g)=R(\lambda+\mu,f+g)\big{[}R(\lambda,f)+R(\mu,g)+i\sigma(f,g)R(\lambda,f)^{2}R(\mu,g)\big{]},

for λ,μ,ν{0}\lambda,\mu,\nu\in\mathbb{R}\setminus\{0\} and f,gXf,g\in X.

The resolvent algebra (X,σ)\mathcal{R}(X,\sigma) is defined as the closure of 0(X,σ)\mathcal{R}_{0}(X,\sigma) with respect to a certain semi-norm obtained through the GNS construction (cf. [7]).

Remark 2.3.
  1. 1.

    Note that Equations (1) and (2) encode ϕ(0)=1\phi(0)=1 and the self-adjointness of ϕ(f)\phi(f), respectively. The third and sixth equation reflect the \mathbb{R}-linearity of ϕ\phi, whereas Equation (4) is the resolvent identity. Finally, the fifth equation is the substitute for the (CCR).

  2. 2.

    When one considers a representation of 0(X,σ)\mathcal{R}_{0}(X,\sigma) as a concrete \ast-algebra generated by resolvents of self-adjoint operators on a Hilbert space, then (X,σ)\mathcal{R}(X,\sigma) coincides with the operator norm closure of 0(X,σ)\mathcal{R}_{0}(X,\sigma).

In the present paper we study the Fock-Bargmann representation of the resolvent algebra over the standard symplectic space n2n\mathbb{C}^{n}\cong\mathbb{R}^{2n} as well as on an infinite dimensional Hilbert space with a symplectic structure. In the finite dimensional setting of n\mathbb{C}^{n} it is well-known (cf. [3, 8] and Section 3) that the generators are resolvents of self-adjoint Toeplitz operators T2σ(,z)T_{2\sigma(\cdot,z)}, znz\in\mathbb{C}^{n} defined below.

Therefore, the resolvent algebra (n,σ)\mathcal{R}(\mathbb{C}^{n},\sigma) is the CC^{\ast} algebra generated by the resolvents of Toeplitz operators:

(n,σ)C((iλT2σ(,z))1;λ{0},zn).\mathcal{R}(\mathbb{C}^{n},\sigma)\cong C^{\ast}\Big{(}(i\lambda-T_{2\sigma(\cdot,z)})^{-1};~{}\lambda\in\mathbb{R}\setminus\{0\},z\in\mathbb{C}^{n}\Big{)}. (2.3)

In fact, this will be the starting point of our analysis.

3 Fock-Bargmann space and Toeplitz operators

Let nn\in\mathbb{N} and consider positive real numbers tj>0t_{j}>0 where j=1,,nj=1,\ldots,n. We write 𝐭:=(t1,,tn)\mathbf{t}:=(t_{1},\dots,t_{n}) and on n\mathbb{C}^{n} we define the probability measure μ𝐭\mu_{\mathbf{t}} by

dμ𝐭(z)=1πnt1tne(|z1|2t1++|zn|2tn)dV(z),z=(z1,,zn)n.d\mu_{\mathbf{t}}(z)=\frac{1}{\pi^{n}t_{1}\cdot\dots\cdot t_{n}}e^{-\left(\frac{|z_{1}|^{2}}{t_{1}}+\dots+\frac{|z_{n}|^{2}}{t_{n}}\right)}dV(z),\hskip 12.91663ptz=(z_{1},\ldots,z_{n})\in\mathbb{C}^{n}.

Here, VV denotes the Lebesgue measure on n2n\mathbb{C}^{n}\cong\mathbb{R}^{2n}. The Fock-Bargmann space F𝐭2=F𝐭2(n)F_{\mathbf{t}}^{2}=F_{\mathbf{t}}^{2}(\mathbb{C}^{n}) is defined as

F𝐭2:=L2(n,μ𝐭)Hol(n),F_{\mathbf{t}}^{2}:=L^{2}(\mathbb{C}^{n},\mu_{\mathbf{t}})\cap\operatorname{Hol}(\mathbb{C}^{n}), (3.1)

where Hol(n)\operatorname{Hol}(\mathbb{C}^{n}) denotes the space of entire functions on n\mathbb{C}^{n}. The reader experienced with the analysis on such spaces may wonder why we introduce the parameters t1,,tnt_{1},\ldots,t_{n} in the above definition, as the standard approach corresponds to the choice t1=t2==tn=:t>0t_{1}=t_{2}=\dots=t_{n}=:t>0. We will use the notation Ft2=F(t,,t)2F_{t}^{2}=F_{(t,\dots,t)}^{2} when we are in this standard situation. Conceptually, the more general setup does not cause any problems (apart from longer notations), but will be convenient when we pass to infinite dimensional symplectic spaces.

Throughout the paper we denote the standard inner product of L2(n,μ𝐭)L^{2}(\mathbb{C}^{n},\mu_{\mathbf{t}}) and of F𝐭2F_{\mathbf{t}}^{2} by

f,g:=nf(z)g(z)¯𝑑μ𝐭(z)\langle f,g\rangle:=\int_{\mathbb{C}^{n}}f(z)\overline{g(z)}d\mu_{\mathbf{t}}(z)

and we write f:=f,f1/2\|f\|:=\langle f,f\rangle^{1/2} for the induced norm. As is well-known, F𝐭2F_{\mathbf{t}}^{2} is a reproducing kernel Hilbert space with kernel function

Kw𝐭(z):=K𝐭(z,w):=ez1w1¯t1++znwn¯tn,K_{w}^{\mathbf{t}}(z):=K^{\mathbf{t}}(z,w):=e^{\frac{z_{1}\cdot\overline{w_{1}}}{t_{1}}+\dots+\frac{z_{n}\cdot\overline{w_{n}}}{t_{n}}},

where w=(w1,,wn)n.w=(w_{1},\ldots,w_{n})\in\mathbb{C}^{n}. In order to simplify notations, we write

z,w𝐭:=z1w1¯t1++znwn¯tnand z𝐭2:=z,z𝐭\displaystyle\langle z,w\rangle_{\mathbf{t}}:=\frac{z_{1}\cdot\overline{w_{1}}}{t_{1}}+\dots+\frac{z_{n}\cdot\overline{w_{n}}}{t_{n}}\hskip 12.91663pt\mbox{\it and }\hskip 12.91663pt\|z\|_{\mathbf{t}}^{2}:=\langle z,z\rangle_{\mathbf{t}} (3.2)

such that Kw𝐭(z)=ez,w𝐭K_{w}^{\mathbf{t}}(z)=e^{\langle z,w\rangle_{\mathbf{t}}}. We express the normalized reproducing kernels kw𝐭F𝐭2k_{w}^{\mathbf{t}}\in F_{\mathbf{t}}^{2} as

kw𝐭(z):=Kw𝐭(z)Kw𝐭=ez,w𝐭12w𝐭2.k_{w}^{\mathbf{t}}(z):=\frac{K_{w}^{\mathbf{t}}(z)}{\|K_{w}^{\mathbf{t}}\|}=e^{\langle z,w\rangle_{\mathbf{t}}-\frac{1}{2}\|w\|_{\mathbf{t}}^{2}}.

Let A(F𝐭2)A\in\mathcal{L}(F_{\mathbf{t}}^{2}) be a bounded linear operator on F𝐭2F_{\mathbf{t}}^{2}. We define the Berezin transform A~\widetilde{A} of AA by

A~(z):=Akz𝐭,kz𝐭,zn.\widetilde{A}(z):=\big{\langle}Ak_{z}^{\mathbf{t}},k_{z}^{\mathbf{t}}\big{\rangle},\hskip 17.22217ptz\in\mathbb{C}^{n}.

Note that A~(z)\widetilde{A}(z) is a bounded real-analytic function. Moreover, the map AA~A\mapsto\widetilde{A} is known to be injective, [12].

As F𝐭2F_{\mathbf{t}}^{2} is a closed subspace of L2(n,μ𝐭)L^{2}(\mathbb{C}^{n},\mu_{\mathbf{t}}), there is an orthogonal projection P𝐭:L2(n,μ𝐭)F𝐭2P^{\mathbf{t}}:L^{2}(\mathbb{C}^{n},\mu_{\mathbf{t}})\to F_{\mathbf{t}}^{2} acting on fL2(n,μ𝐭)f\in L^{2}(\mathbb{C}^{n},\mu_{\mathbf{t}}) as

P𝐭(f)(w)=f,Kw𝐭,wn.P^{\mathbf{t}}(f)(w)=\big{\langle}f,K_{w}^{\mathbf{t}}\big{\rangle},\quad w\in\mathbb{C}^{n}.

For a measurable function φ:n\varphi:\mathbb{C}^{n}\to\mathbb{C}, we let Tφ𝐭T_{\varphi}^{\mathbf{t}} denote the Toeplitz operator with symbol φ\varphi, which is defined by

Tφ𝐭(g):=P𝐭(φg)T_{\varphi}^{\mathbf{t}}(g):=P^{\mathbf{t}}(\varphi g)

on the natural domain D(Tφ𝐭)={gF𝐭2:φgL2(n,μ𝐭)}D(T_{\varphi}^{\mathbf{t}})=\big{\{}g\in F_{\mathbf{t}}^{2}\>:\>\varphi g\in L^{2}(\mathbb{C}^{n},\mu_{\mathbf{t}})\big{\}}.

If φL(n)\varphi\in L^{\infty}(\mathbb{C}^{n}), then Tφ𝐭T_{\varphi}^{\mathbf{t}} is a bounded operator. For a given set SL(n)S\subset L^{\infty}(\mathbb{C}^{n}), we will denote by 𝒯𝐭(S)\mathcal{T}^{\mathbf{t}}(S) the CC^{\ast} algebra generated by all Toeplitz operators with symbols in SS and set

𝒯𝐭:=𝒯𝐭(L(n))\mathcal{T}^{\mathbf{t}}:=\mathcal{T}^{\mathbf{t}}\big{(}L^{\infty}(\mathbb{C}^{n})\big{)}

for the full Toeplitz algebra. We also define 𝒯lin𝐭(S)\mathcal{T}_{\textup{lin}}^{\mathbf{t}}(S) to be the closed linear span of Toeplitz operators with symbols in SS.

Let us introduce Weyl operators on the Fock-Bargmann space F𝐭2F_{\mathbf{t}}^{2}: if znz\in\mathbb{C}^{n} then we define Wz𝐭(F𝐭2)W_{z}^{\mathbf{t}}\in\mathcal{L}(F_{\mathbf{t}}^{2}) by

Wz𝐭(g)(w)=kz𝐭(w)g(wz).W_{z}^{\mathbf{t}}(g)(w)=k_{z}^{\mathbf{t}}(w)g(w-z). (3.3)

Weyl operators are well-known to be unitary. Moreover, they fulfill the relation

(Wz𝐭)=(Wz𝐭)1=Wz𝐭andWz𝐭Ww𝐭=eiσ𝐭(z,w)Wz+w𝐭.\big{(}W_{z}^{\mathbf{t}}\big{)}^{\ast}=\big{(}W_{z}^{\mathbf{t}}\big{)}^{-1}=W_{-z}^{\mathbf{t}}\quad\mbox{\it and}\quad W_{z}^{\mathbf{t}}W_{w}^{\mathbf{t}}=e^{-i\sigma_{\mathbf{t}}(z,w)}W_{z+w}^{\mathbf{t}}.

Here, the symplectic form σ𝐭\sigma_{\mathbf{t}} on n\mathbb{C}^{n} with parameter 𝐭{\bf t} is defined by

σ𝐭(w,z):=Im(w1z1¯)t1++Im(wnzn¯)tn=Imw,z𝐭.\displaystyle\sigma_{\mathbf{t}}(w,z):=\frac{\operatorname{Im}(w_{1}\cdot\overline{z_{1}})}{t_{1}}+\dots+\frac{\operatorname{Im}(w_{n}\cdot\overline{z_{n}})}{t_{n}}=\operatorname{Im}\langle w,z\rangle_{\mathbf{t}}.

As a matter of fact, Weyl operators are Toeplitz operators themselves. If we define the family (gz𝐭)zn(g_{z}^{\bf t})_{z\in\mathbb{C}^{n}} of bounded functions on n\mathbb{C}^{n} by

gz𝐭(w):=e12z𝐭2+2iσ𝐭(w,z),g_{z}^{\mathbf{t}}(w):=e^{\frac{1}{2}\|z\|_{\mathbf{t}}^{2}+2i\sigma_{\mathbf{t}}(w,z)},

then it holds

Wz𝐭=Tgz𝐭𝐭.W_{z}^{\mathbf{t}}=T_{g_{z}^{\mathbf{t}}}^{\mathbf{t}}. (3.4)

In fact, (3.4) follows by showing that the Berezin transform of both sides coincide (see [3, 8] and the references therein).

4 Resolvent algebra in Fock-Bargmann
representation

In this section, we study the resolvent algebra (n,σ𝐭)\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}) in its Fock-Bargmann representation. Of course, it is not hard to reduce the analysis of (n,σ𝐭)\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}) to that of (n,σ)\mathcal{R}(\mathbb{C}^{n},\sigma). Again, we emphazise that the extra flexibility coming from the parameter set 𝐭=(t1,,tn){\bf t}=(t_{1},\ldots,t_{n}) will be useful when passing to the infinite dimensional limit nn\rightarrow\infty. Therefore, we take the (mostly notational) burden upon us to carry 𝐭\mathbf{t} all the way through this section. For readability, we will denote the Fock-Bargmann representation of the resolvent algebra also by (n,σ𝐭)\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}). Since the Toeplitz operators T2σ𝐭(,z)𝐭T_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}} fulfill the canonical commutation relation (equivalently, the Weyl operators Wz𝐭W_{z}^{\mathbf{t}} fulfill the relation of the CCR algebra), the resolvent algebra is the CC^{\ast} algebra generated by resolvents of Toeplitz operators. The following integral representations allow us to study (n,σ𝐭)\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}) more in detail:

Lemma 4.1.

For zn{0}z\in\mathbb{C}^{n}\setminus\{0\} the map sWsz𝐭\mathbb{R}\ni s\mapsto W_{sz}^{\mathbf{t}} defines a strongly continuous unitary one-parameter group with generator iT2σ𝐭(,z)𝐭iT_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}}. In particular, for λ>0\lambda>0 the following integral representations of the resolvents hold in the strong sense:

(T2σ𝐭(,z)𝐭+iλ)1=(T2σ𝐭(,z)+iλ𝐭)1=i0eλsWsz𝐭𝑑s\big{(}T_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}}+i\lambda\big{)}^{-1}=\big{(}T_{2\sigma_{\mathbf{t}}(\cdot,z)+i\lambda}^{\mathbf{t}}\big{)}^{-1}=-i\int_{0}^{\infty}e^{-\lambda s}W_{sz}^{\mathbf{t}}ds (4.1)

and

(T2σ𝐭(,z)𝐭iλ)1=(T2σ𝐭(,z)iλ𝐭)1=i0eλsWsz𝐭𝑑s.\big{(}T_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}}-i\lambda\big{)}^{-1}=\big{(}T_{2\sigma_{\mathbf{t}}(\cdot,z)-i\lambda}^{\mathbf{t}}\big{)}^{-1}=i\int_{0}^{\infty}e^{-\lambda s}W_{-sz}^{\mathbf{t}}ds. (4.2)
Proof.

It is easy to check that sWsz𝐭\mathbb{R}\ni s\mapsto W_{sz}^{\mathbf{t}} is indeed a strongly continuous unitary one-parameter group. For determining its generator we compute

lims0Wsz𝐭ffs=lims0Tgsz𝐭𝐭ffs=lims0T(gsz𝐭1)/s𝐭f\lim_{s\to 0}\frac{W_{sz}^{\mathbf{t}}f-f}{s}=\lim_{s\to 0}\frac{T_{g_{sz}^{\mathbf{t}}}^{\mathbf{t}}f-f}{s}=\lim_{s\to 0}T_{(g_{sz}^{\mathbf{t}}-1)/s}^{\mathbf{t}}f

for all fF𝐭2f\in F_{\mathbf{t}}^{2} such that the limit exists. According to Stone’s Theorem there is a unique self-adjoint operator AA with domain D(A)D(A) such that iAiA generates the group, i.e. for fD(A)f\in D(A) it holds

lims0T(gsz𝐭1)/s𝐭f=iAf\lim_{s\to 0}T_{(g_{sz}^{\mathbf{t}}-1)/s}^{\mathbf{t}}f=iAf (4.3)

in the norm sense. Since norm convergence in the reproducing kernel Hilbert space F𝐭2F_{\mathbf{t}}^{2} implies pointwise convergence, it suffices to determine the pointwise limit of the right hand side of (4.3). Let unu\in\mathbb{C}^{n} and note that:

Tgsz𝐭1𝐭fs(u)=n1s(es22z𝐭2+2isσ𝐭(w,z)1)eu,w𝐭f(w)𝑑μ𝐭(w).\displaystyle\frac{T_{g_{sz}^{\mathbf{t}}-1}^{\mathbf{t}}f}{s}(u)=\int_{\mathbb{C}^{n}}\frac{1}{s}\left(e^{\frac{s^{2}}{2}\|z\|^{2}_{\mathbf{t}}+2is\sigma_{\mathbf{t}}(w,z)}-1\right)e^{\langle u,w\rangle_{\mathbf{t}}}f(w)d\mu_{\mathbf{t}}(w).

The integrand converges as s0s\rightarrow 0:

1s(es22z𝐭2+2isσ𝐭(w,z)1)s02iσ𝐭(w,z).\displaystyle\frac{1}{s}\left(e^{\frac{s^{2}}{2}\|z\|_{\mathbf{t}}^{2}+2is\sigma_{\mathbf{t}}(w,z)}-1\right)\overset{s\to 0}{\longrightarrow}2i\sigma_{\mathbf{t}}(w,z).

An easy application of the dominated convergence theorem yields

T(gsz𝐭1)/s𝐭f(u)iT2σ𝐭(,z)𝐭f(u)T_{(g_{sz}^{\mathbf{t}}-1)/s}^{\mathbf{t}}f(u)\to iT_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}}f(u)

for those fF𝐭2f\in F_{\mathbf{t}}^{2} such that wσ𝐭(w,z)f(w)L2(n,μ𝐭)w\mapsto\sigma_{\mathbf{t}}(w,z)f(w)\in L^{2}(\mathbb{C}^{n},\mu_{\mathbf{t}}). Hence,

T(gsz𝐭1)/s)𝐭fiT2σ𝐭(,z)𝐭f=iAfforfD(A).T_{(g_{sz}^{\mathbf{t}}-1)/s)}^{\mathbf{t}}f\to iT_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}}f=iAf\hskip 12.91663pt\mbox{\it for}\hskip 17.22217ptf\in D(A).

Therefore, iT2σ𝐭(,z)𝐭iT_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}} is the generator of the one-parameter group sWsz𝐭s\mapsto W_{sz}^{\mathbf{t}}. Since all operators Wsz𝐭W_{sz}^{\mathbf{t}} are unitary, the group (Wsz𝐭)s(W_{sz}^{\mathbf{t}})_{s} has growth bound ω0=0\omega_{0}=0. For λ>0\lambda>0 it follows (cf. [10, Theorem I.1.10])

(λiT2σ𝐭(,z)𝐭)1=0eλsWsz𝐭𝑑s\big{(}\lambda-iT_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}}\big{)}^{-1}=\int_{0}^{\infty}e^{-\lambda s}W_{sz}^{\mathbf{t}}ds

strongly, i.e.

(T2σ𝐭(,z)𝐭+iλ)1=i0eλsWsz𝐭𝑑s.\big{(}T_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}}+i\lambda\big{)}^{-1}=-i\int_{0}^{\infty}e^{-\lambda s}W_{sz}^{\mathbf{t}}ds.

The integral representation (4.2) follows from:

(T2σ𝐭(,z)𝐭iλ)1\displaystyle\big{(}T_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}}-i\lambda\big{)}^{-1} =((T2σ𝐭(,z)𝐭+iλ)1)=i0eλsWsz𝐭𝑑s,\displaystyle=\left((T_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}}+i\lambda)^{-1}\right)^{\ast}=i\int_{0}^{\infty}e^{-\lambda s}W_{-sz}^{\mathbf{t}}ds,

which completes the proof. ∎

Remark 4.2.

An inspection of the above arguments shows that the limit in Equation (4.3) exists for all f=Kw𝐭,wnf=K_{w}^{\mathbf{t}},~{}w\in\mathbb{C}^{n}. Moreover, one easily verifies that the space 𝒳:=span{Kw𝐭:wn}\mathcal{X}:=\operatorname{span}\{K_{w}^{\mathbf{t}}\>:\>w\in\mathbb{C}^{n}\} is invariant under the action of Wsz𝐭W_{sz}^{\mathbf{t}}. Then, [17, Theorem VIII.10] implies that T2σ𝐭(,z)𝐭T_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}} is essentially self-adjoint on 𝒳\mathcal{X}.

Define for zn,λiz\in\mathbb{C}^{n},~{}\lambda\in\mathbb{C}\setminus i\mathbb{R}:

R𝐭(λ,z):=(T2σ𝐭(,z)𝐭iλ)1.R_{\mathbf{t}}(\lambda,z):=\big{(}T_{2\sigma_{\mathbf{t}}(\cdot,z)}^{\mathbf{t}}-i\lambda\big{)}^{-1}.

For simplicity, we will occasionally suppress 𝐭\mathbf{t} in the notation and shortly write R(λ,z)=R𝐭(λ,z)R(\lambda,z)=R_{\mathbf{t}}(\lambda,z). Before we continue with our investigation, we need to recall the following well-known expansion of the resolvent.

Lemma 4.3.

Let λ0,λi\lambda_{0},\lambda\in\mathbb{C}\setminus i\mathbb{R} such that |λ0λ|<|λ0||\lambda_{0}-\lambda|<|\lambda_{0}|. Then, it holds

R(λ,z)=k=0(λλ0)kikR(λ0,z)k+1,R(\lambda,z)=\sum_{k=0}^{\infty}(\lambda-\lambda_{0})^{k}i^{k}R(\lambda_{0},z)^{k+1}, (4.4)

where the series converges in operator norm. In particular:

R(λ0,z)k=ik1(k1)!dk1dλk1|λ=λ0R(λ,z).R(\lambda_{0},z)^{k}=\frac{i^{k-1}}{(k-1)!}\frac{{\rm d}^{k-1}}{{\rm d}\lambda^{k-1}}|_{\lambda=\lambda_{0}}R(\lambda,z). (4.5)

The previous lemma has the following important consequence:

Corollary 4.4.

(n,σ𝐭)=C(R𝐭(λ,z):λi,zn).\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}})=C^{\ast}(R_{\mathbf{t}}(\lambda,z):~{}\lambda\in\mathbb{C}\setminus i\mathbb{R},~{}z\in\mathbb{C}^{n}).

We further note that Equations (4.1) and (4.2) extend to λi\lambda\in\mathbb{C}\setminus i\mathbb{R} with the same proof as in the case λ{0}\lambda\in\mathbb{R}\setminus\{0\}:

R(λ,z)=i0eλsWsz𝐭𝑑s,Re(λ)>0R(\lambda,z)=i\int_{0}^{\infty}e^{-\lambda s}W_{-sz}^{\mathbf{t}}ds,\quad\operatorname{Re}(\lambda)>0 (4.6)

and

R(λ,z)=i0eλsWsz𝐭𝑑s,Re(λ)<0.R(\lambda,z)=-i\int_{0}^{\infty}e^{\lambda s}W_{sz}^{\mathbf{t}}ds,\quad\operatorname{Re}(\lambda)<0. (4.7)

We mention two ways of analyzing the connection between the algebra (n,σ𝐭)\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}) in its Fock-Bargmann space representation and the theory of Toeplitz operators and, more precisely, its realization as a Toeplitz algebra. The computational-heavy method studies the resolvents through their Laplace transform representation, using that the Weyl operators themselves are Toeplitz operators. Alternatively, one can use “soft analysis” arguments from the theory of Toeplitz operators, most prominently those arising in the theory of quantum harmonic analysis [20, 13]. Both approaches have their advantages and so we will try to shed some light on either of them.

4.1 Correspondence Theory in the Fock-Bargmann space

We recall some aspects of quantum harmonic analysis in the setting of the Fock-Bargmann space and explain a particular part of it: the correspondence theory. We refer to [20, 13] for further details. Roughly speaking, correspondence theory relates certain subspaces of BUC(n)\operatorname{BUC}(\mathbb{C}^{n}), the CC^{\ast} algebra of bounded, uniformly continuous functions on n\mathbb{C}^{n}, with corresponding subspaces of (F𝐭2)\mathcal{L}(F_{\mathbf{t}}^{2}). We start with some notation. Let f:nf:\mathbb{C}^{n}\to\mathbb{C} be a function and znz\in\mathbb{C}^{n}, then put:

αz(f)(w):=f(wz).\displaystyle\alpha_{z}(f)(w):=f(w-z).

Given an operator A(F𝐭2)A\in\mathcal{L}(F_{\mathbf{t}}^{2}) we define its shift by znz\in\mathbb{C}^{n} through

αz(A)=Wz𝐭AWz𝐭=Wz𝐭A(Wz𝐭),\displaystyle\alpha_{z}(A)=W_{z}^{\mathbf{t}}AW_{-z}^{\mathbf{t}}=W_{z}^{\mathbf{t}}A(W_{z}^{\mathbf{t}})^{\ast},

where Wz𝐭W_{z}^{\mathbf{t}} is a Weyl operator. Clearly, BUC(n)\operatorname{BUC}(\mathbb{C}^{n}) is the subalgebra of L(n)L^{\infty}(\mathbb{C}^{n}) consisting of functions ff for which zαz(f)z\mapsto\alpha_{z}(f) is continuous with respect to the LL^{\infty}-norm. The analogous space on the operator side is

𝒞1𝐭={A(F𝐭2):zαz(A) is op-continuous}.\displaystyle\mathcal{C}_{1}^{\mathbf{t}}=\big{\{}A\in\mathcal{L}(F_{\mathbf{t}}^{2}):~{}z\mapsto\alpha_{z}(A)\text{ is }\|\cdot\|_{op}\text{-\it continuous}\big{\}}.

The Fock-Bargmann space formulation of the correspondence theorem due to R. Werner in [20] specifically tailored for Toeplitz operators can be found in [13] within the standard situation t1=t2==tn=t>0t_{1}=t_{2}=\dots=t_{n}=t>0. The case of positive weight parameters t1,,tnt_{1},\ldots,t_{n} in the definition of the Gaussian measure μ𝐭\mu_{\mathbf{t}} follows by identical proofs and obvious modifications:

Theorem 4.5 (Correspondence Theorem, [20, 13]).

Let 𝒟1𝒞1𝐭\mathcal{D}_{1}\subset\mathcal{C}_{1}^{\mathbf{t}} be a closed, α\alpha-invariant subspace (meaning αz(A)𝒟1\alpha_{z}(A)\in\mathcal{D}_{1} for every zn,A𝒟1z\in\mathbb{C}^{n},~{}A\in\mathcal{D}_{1}). Then, there is a unique α\alpha-invariant closed subspace 𝒟0\mathcal{D}_{0} of BUC(n)\operatorname{BUC}(\mathbb{C}^{n}) such that

𝒟1=𝒯lin𝐭(𝒟0).\displaystyle\mathcal{D}_{1}=\mathcal{T}_{lin}^{\mathbf{t}}(\mathcal{D}_{0}).

Further, for an operator A𝒞1𝐭A\in\mathcal{C}_{1}^{\mathbf{t}} the following statements are equivalent:

A𝒟1A~𝒟0.\displaystyle A\in\mathcal{D}_{1}\Longleftrightarrow\widetilde{A}\in\mathcal{D}_{0}.

Finally, 𝒟0\mathcal{D}_{0} can be computed as follows: 𝒟0={A~:A𝒟1}¯.\mathcal{D}_{0}=\overline{\{\widetilde{A}:~{}A\in\mathcal{D}_{1}\}}. In the following we call 𝒟0\mathcal{D}_{0} and 𝒟1\mathcal{D}_{1} corresponding spaces.

Here, we used the notation

𝒯lin𝐭(𝒟0)={Tf𝐭:f𝒟0}¯.\displaystyle\mathcal{T}_{lin}^{\mathbf{t}}(\mathcal{D}_{0})=\overline{\big{\{}T_{f}^{\mathbf{t}}:~{}f\in\mathcal{D}_{0}\big{\}}}.

The previous result is a useful tool in the study of Toeplitz operators and Toeplitz algebras when it is combined with the following:

Theorem 4.6 ([13]).

The following CC^{*} algebras coincide: 𝒞1𝐭=𝒯𝐭\mathcal{C}_{1}^{\mathbf{t}}=\mathcal{T}^{\mathbf{t}}.

We mention that the previous two results extend to the case of the pp-Fock space F𝐭pF_{\bf t}^{p} where p2p\neq 2, cf. [13, 14]. They have immediate applications such as simple proofs of a compactness characterization for operators acting on the Fock-Bargmann space. We only present the case p=2p=2.

Let 𝒦()\mathcal{K}(\mathcal{H}) denote the ideal of compact operators on a Hilbert space \mathcal{H}.

Theorem 4.7 ([1]).

Let A(F𝐭2)A\in\mathcal{L}(F_{\mathbf{t}}^{2}). Then, the following are equivalent:

A𝒦(F𝐭2)A𝒯𝐭 and A~C0(n).\displaystyle A\in\mathcal{K}(F_{\mathbf{t}}^{2})\Longleftrightarrow A\in\mathcal{T}^{\mathbf{t}}\text{ and }\widetilde{A}\in C_{0}(\mathbb{C}^{n}).

Here, C0(n)C_{0}(\mathbb{C}^{n}) denotes the space of continuous complex valued functions vanishing at infinity.

The short proofs of the last theorems based on correspondence theory can be found in [13]. We now want to demonstrate how Theorem 4.5 can be applied in order to gain a better understanding of the resolvent algebra. First, we note:

Lemma 4.8.

Let z,wnz,w\in\mathbb{C}^{n} and λi\lambda\in\mathbb{C}\setminus i\mathbb{R}. Then:

αw(R(λ,z))=R(λ+2iσ𝐭(z,w),z).\displaystyle\alpha_{w}\big{(}R(\lambda,z)\big{)}=R\big{(}\lambda+2i\sigma_{\mathbf{t}}(z,w),z\big{)}.
Proof.

Applying the CCR of Weyl operators, we get for Re(λ)>0\operatorname{Re}(\lambda)>0:

αw(R(λ,z))\displaystyle\alpha_{w}\big{(}R(\lambda,z)\big{)} =Ww𝐭i0eλsWsz𝐭𝑑sWw𝐭\displaystyle=W_{w}^{\mathbf{t}}i\int_{0}^{\infty}e^{-\lambda s}W_{-sz}^{\mathbf{t}}~{}dsW_{-w}^{\mathbf{t}}
=i0eλsWw𝐭Wsz𝐭Ww𝐭𝑑s\displaystyle=i\int_{0}^{\infty}e^{-\lambda s}W_{w}^{\mathbf{t}}W_{-sz}^{\mathbf{t}}W_{-w}^{\mathbf{t}}~{}ds
=i0eλte2isσ𝐭(z,w)Wsz𝐭𝑑s\displaystyle=i\int_{0}^{\infty}e^{-\lambda t}e^{-2is\sigma_{\mathbf{t}}(z,w)}W_{-sz}^{\mathbf{t}}~{}ds
=R(λ+2iσ𝐭(z,w),z).\displaystyle=R\big{(}\lambda+2i\sigma_{\mathbf{t}}(z,w),z\big{)}.

The case Re(λ)<0\operatorname{Re}(\lambda)<0 follows analogously. ∎

Proposition 4.9.

(n,σ𝐭)\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}) is a closed, α\alpha-invariant subspace of 𝒞1𝐭\mathcal{C}_{1}^{\mathbf{t}}.

Proof.

The α\alpha-invariance immediately follows from the previous lemma. Further, the resolvent algebra is closed by definition. We only need to prove that it is contained in 𝒞1𝐭\mathcal{C}_{1}^{\mathbf{t}}. It suffices to show that R𝐭(λ,z)𝒞1𝐭R_{\mathbf{t}}(\lambda,z)\in\mathcal{C}_{1}^{\mathbf{t}} for λi\lambda\in\mathbb{C}\setminus i\mathbb{R} and znz\in\mathbb{C}^{n}. This is an easy consequence of Lemma 4.3 and Lemma 4.8. In fact, for |w||w| sufficiently small such that simultaneously

2|σ𝐭(z,w)|<|λ|and |2σ𝐭(z,w)|R(λ,z)<12|\sigma_{\mathbf{t}}(z,w)|<|\lambda|\hskip 12.91663pt\mbox{\it and }\hskip 12.91663pt|2\sigma_{\mathbf{t}}(z,w)|\|R(\lambda,z)\|<1

it follows:

R(λ,z)αw(R(λ,z))\displaystyle\|R(\lambda,z)-\alpha_{w}(R(\lambda,z))\| =R(λ,z)R(λ+2iσ𝐭(z,w),z)\displaystyle=\|R(\lambda,z)-R(\lambda+2i\sigma_{\mathbf{t}}(z,w),z)\|
k=1|2σ𝐭(z,w)|kR(λ,z)k+1\displaystyle\leq\sum_{k=1}^{\infty}|2\sigma_{\mathbf{t}}(z,w)|^{k}\|R(\lambda,z)\|^{k+1}
=2|σ𝐭(z,w)|R(λ,z)212|σ𝐭(z,w)|R(λ,z).\displaystyle=\frac{2|\sigma_{\mathbf{t}}(z,w)|\|R(\lambda,z)\|^{2}}{1-2|\sigma_{\mathbf{t}}(z,w)|\|R(\lambda,z)\|}.

This last expression tends to 0 as |w|0|w|\to 0. ∎

Proposition 4.9 and the correspondence theorem (Theorem 4.5) imply that there is a closed, α\alpha-invariant subspace 𝒟0𝐭\mathcal{D}_{0}^{\mathbf{t}} of BUC(n)\operatorname{BUC}(\mathbb{C}^{n}) (possibly depending on the parameter tuple 𝐭\mathbf{t}) such that

(n,σ𝐭)=𝒯lin𝐭(𝒟0𝐭).\displaystyle\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}})=\mathcal{T}_{\textup{lin}}^{\mathbf{t}}(\mathcal{D}_{0}^{\mathbf{t}}).

Since (n,σ𝐭)\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}) itself is a CC^{\ast} algebra, we get:

Theorem 4.10.

(n,σ𝐭)\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}) is a Toeplitz algebra. More precisely,

(n,σ𝐭)=𝒯𝐭(𝒟0𝐭):=C(Tf𝐭:f𝒟0𝐭)\displaystyle\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}})=\mathcal{T}^{\mathbf{t}}(\mathcal{D}_{0}^{\mathbf{t}}):=C^{\ast}(T_{f}^{\mathbf{t}}:~{}f\in\mathcal{D}_{0}^{\mathbf{t}})

for a suitable α\alpha-invariant subspace 𝒟0𝐭BUC(n)\mathcal{D}_{0}^{\mathbf{t}}\subset\textup{BUC}(\mathbb{C}^{n}).

Our next task is to determine the space 𝒟0𝐭\mathcal{D}_{0}^{\mathbf{t}} explicitly.

4.2 Computing 𝒟0𝐭\mathcal{D}_{0}^{\mathbf{t}}

Correspondence theory has the nice flavour that in several examples the corresponding spaces (in the sense of Theorem 4.5) are what one might naively expect; for example, C0(n)C_{0}(\mathbb{C}^{n}) corresponds to the compact operators 𝒦(F𝐭2)\mathcal{K}(F_{\mathbf{t}}^{2}) and the almost periodic functions correspond to the CCR algebra. In this respect, one might hope that the space 𝒟0𝐭\mathcal{D}_{0}^{\mathbf{t}} corresponding to (n,σ𝐭)\mathcal{R}(\mathbb{C}^{n},\sigma_{{\bf t}}) is:

cl=C({(λ2iσ𝐭(,z))1:λi,zn}).\displaystyle{\mathcal{R}_{cl}}=C^{\ast}\Big{(}\{(\lambda-2i\sigma_{\mathbf{t}}(\cdot,z))^{-1}\>:\>~{}\lambda\in\mathbb{C}\setminus i\mathbb{R},~{}z\in\mathbb{C}^{n}\}\Big{)}.

Note that cl\mathcal{R}_{cl} does not depend on the choice of the parameter set 𝐭=(t1,,tn){\bf t}=(t_{1},\ldots,t_{n}) and we can replace σ𝐭\sigma_{\bf t} by σ=σ1\sigma=\sigma_{1} in its definition (the “contribution” of the parameters tkt_{k} can be absorbed into zz).

Since cl\mathcal{R}_{cl} is indeed a subalgebra of BUC(n)\operatorname{BUC}(\mathbb{C}^{n}), which is further α\alpha-invariant as well as invariant under the parity operation ff(),f\mapsto f(-\cdot), Theorem 3.13 in [13] implies that:

𝒯lin𝐭(cl)=𝒯𝐭(cl).{\mathcal{T}_{lin}^{\mathbf{t}}(\mathcal{R}_{\textup{cl}})=\mathcal{T}^{\mathbf{t}}(\mathcal{R}_{cl})}.

Therefore, 𝒟0𝐭=cl\mathcal{D}_{0}^{\mathbf{t}}=\mathcal{R}_{cl} if and only if (n,σ𝐭)=𝒯𝐭(cl)\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}})=\mathcal{T}^{\mathbf{t}}(\mathcal{R}_{cl}). Next, we collect some useful facts:

Proposition 4.11.

The following inclusions hold:

  1. 1.

    C0(n)clC_{0}(\mathbb{C}^{n})\subset\mathcal{R}_{cl}.

  2. 2.

    𝒦(F𝐭2)(n,σ𝐭)\mathcal{K}(F^{2}_{\mathbf{t}})\subset\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}).

Proof.

The first statement is an easy application of the Stone-Weierstrass Theorem; the second statement is [7, Theorem 5.4].∎

In what follows some computations can no longer be avoided. The Berezin transform of a product of resolvents can be computed explicitly, which turns out to be quite useful. We will not need the formula in full generality, but still provide the complete expression here.

Proposition 4.12.

Let mm\in\mathbb{N} and z1,,zmnz_{1},\dots,z_{m}\in\mathbb{C}^{n}. Given a multi-index 𝐤=(k1,,km)m{\bf k}=(k_{1},\dots,k_{m})\in{\mathbb{N}}^{m} and λ1,,λmi\lambda_{1},\dots,\lambda_{m}\in\mathbb{C}\setminus i\mathbb{R} we have:

(R(λ1,z1)k1R(λm,zm)km)(w)=\displaystyle\Big{(}R(\lambda_{1},z_{1})^{k_{1}}\dots R(\lambda_{m},z_{m})^{k_{m}}\Big{)}^{\sim}(w)=
=Ci|𝐤|(𝐤𝟏)!(0,)m𝐬𝐤𝟏eΛ𝐬2iσ𝐭(w,𝐬𝐳)ij<sjsσ𝐭(zj,z)12𝐬𝐳𝐭2𝑑𝐬,\displaystyle\quad=\frac{Ci^{|\mathbf{k}|}}{(\mathbf{k-1})!}\int_{(0,\infty)^{m}}\mathbf{s}^{\mathbf{k-1}}e^{-\Lambda\cdot\mathbf{s}-2i\sigma_{\mathbf{t}}(w,\mathbf{s}\cdot\mathbf{z})-i\sum_{j<\ell}s_{j}s_{\ell}\sigma_{\mathbf{t}}(z_{j},z_{\ell})-\frac{1}{2}\|\mathbf{s}\cdot\mathbf{z}\|_{\mathbf{t}}^{2}}~{}d\mathbf{s},

where

C=Ck,λ,m:=(1)|𝐤|mj=1msign(Re(λj))kj{1,1}.C=C_{k,\lambda,m}:=(-1)^{|\mathbf{k}|-m}\cdot\prod_{j=1}^{m}\operatorname{sign}\big{(}\operatorname{Re}(\lambda_{j})\big{)}^{k_{j}}\in\{-1,1\}.

Here, we have used the standard multi-index notation together with:

𝟏\displaystyle\mathbf{1} :=(1,,1)m,\displaystyle:=(1,\dots,1)\in\mathbb{N}^{m},
Λ\displaystyle\Lambda :=(sign(Re(λ1))λ1,,sign(Re(λm))λm),\displaystyle:=\big{(}\operatorname{sign}(\operatorname{Re}(\lambda_{1}))\lambda_{1},\dots,\operatorname{sign}(\operatorname{Re}(\lambda_{m}))\lambda_{m}\big{)},
𝐬𝐳\displaystyle\mathbf{s}\cdot\mathbf{z} =s1z1++smzmn.\displaystyle=s_{1}z_{1}+\dots+s_{m}z_{m}\in\mathbb{C}^{n}.
Proof.

Using the relation R(λ,z)=R(λ,z)R(-\lambda,z)=-R(\lambda,-z), we can reduce the proof to the case where Re(λj)>0\operatorname{Re}(\lambda_{j})>0 for j=1,,mj=1,\ldots,m. From Lemma 4.3 we obtain that

(R(λ1,z1)k1R(λm,zm)km)(w)==i|𝐤|m(𝐤𝟏)!𝐤𝟏μ𝐤𝟏|μ=ΛR(μ1,z1)R(μm,zm)kw𝐭,kw𝐭,\Big{(}R(\lambda_{1},z_{1})^{k_{1}}\dots R(\lambda_{m},z_{m})^{k_{m}}\Big{)}^{\sim}(w)=\\ =\frac{i^{|\mathbf{k}|-m}}{(\mathbf{k}-\mathbf{1})!}\left\langle\frac{\partial^{\mathbf{k-1}}}{\partial\mu^{\mathbf{k-1}}}|_{\mu=\Lambda}R(\mu_{1},z_{1})\dots R(\mu_{m},z_{m})k_{w}^{{\bf t}},k_{w}^{\bf t}\right\rangle,

where we use the short notation:

𝐤𝟏μ𝐤𝟏|μ=Λ:=k11μ1k11km1μmkm1|μ1=λ1,,μm=λm.\displaystyle\frac{\partial^{\mathbf{k-1}}}{\partial\mu^{\mathbf{k-1}}}|_{\mu=\Lambda}:=\frac{\partial^{k_{1}-1}}{\partial\mu_{1}^{k_{1}-1}}\dots\frac{\partial^{k_{m}-1}}{\partial\mu_{m}^{k_{m}-1}}|_{\mu_{1}=\lambda_{1},\dots,\mu_{m}=\lambda_{m}}.

According to Lemma 4.8 we have:

(R(μ1,z1)\displaystyle\Big{(}R(\mu_{1},z_{1})\ldots\> R(μm,zm))(w)=αw(R(μ1,z1)R(μm,zm))(0)\displaystyle R(\mu_{m},z_{m})\Big{)}^{\sim}(w)=\alpha_{-w}\Big{(}R(\mu_{1},z_{1})\ldots R(\mu_{m},z_{m})\Big{)}^{\sim}(0)
=(αw[R(μ1,z1)]αw[R(μm,zm)])(0)\displaystyle=\Big{(}\alpha_{-w}\big{[}R(\mu_{1},z_{1})\big{]}\ldots\alpha_{-w}\big{[}R(\mu_{m},z_{m})\big{]}\Big{)}^{\sim}(0)
=(R(μ12iσ𝐭(z1,w),z1)R(μm2iσ𝐭(zm,w),zm))(0).\displaystyle=\Big{(}R\big{(}\mu_{1}-2i\sigma_{\mathbf{t}}(z_{1},w),z_{1}\big{)}\ldots R\big{(}\mu_{m}-2i\sigma_{\mathbf{t}}(z_{m},w),z_{m}\big{)}\Big{)}^{\sim}(0).

By using analyticity of the resolvent maps μjR(μj,zj)(F𝐭2)\mu_{j}\mapsto R(\mu_{j},z_{j})\in\mathcal{L}(F_{\mathbf{t}}^{2}) it follows that the difference quotients converge in operator norm. Hence, differentiation can be interchanged with the inner product, which yields:

(𝐤𝟏)!i|𝐤|m(R(λ1,z1)k1R(λm,zm)km)(w)\displaystyle\frac{(\mathbf{k-1})!}{i^{|\mathbf{k}|-m}}\Big{(}R(\lambda_{1},z_{1})^{k_{1}}\dots R(\lambda_{m},z_{m})^{k_{m}}\Big{)}^{\sim}(w)
=𝐤𝟏μ𝐤𝟏|μ=Λ(R(μ1,z1)R(μm,zm))(w)\displaystyle\quad=\frac{\partial^{\mathbf{k-1}}}{\partial\mu^{\mathbf{k-1}}}|_{\mu=\Lambda}\Big{(}R(\mu_{1},z_{1})\ldots R(\mu_{m},z_{m})\Big{)}^{\sim}(w)
=𝐤𝟏μ𝐤𝟏|μ=Λ(R(μ12iσ𝐭(z1,w),z1)R(μm2iσ𝐭(zm,w)))(0)=().\displaystyle\quad=\frac{\partial^{\mathbf{k-1}}}{\partial\mu^{\mathbf{k-1}}}|_{\mu=\Lambda}\underbrace{\Big{(}R\big{(}\mu_{1}-2i\sigma_{\mathbf{t}}(z_{1},w),z_{1}\big{)}\ldots R\big{(}\mu_{m}-2i\sigma_{\mathbf{t}}(z_{m},w)\big{)}\Big{)}^{\sim}(0)}_{=(*)}.

Now, we insert the integral expression of the resolvent in (4.6):

()\displaystyle(*) =(0,)mej[μj+2iσ𝐭(w,zj))]sjij<sjsσ𝐭(zj,z)W𝐬𝐳𝐭1,1𝑑𝐬.\displaystyle=\int_{(0,\infty)^{m}}e^{-\sum_{j}[\mu_{j}+2i\sigma_{\mathbf{t}}(w,z_{j}))]s_{j}-i\sum_{j<\ell}s_{j}s_{\ell}\sigma_{\mathbf{t}}(z_{j},z_{\ell})}\langle W_{{\bf s\cdot z}}^{\mathbf{t}}1,1\rangle~{}d\mathbf{s}.

The inner product in the integrand can be calculated explicitly:

W𝐬𝐳𝐭1,1=e12𝐬𝐳𝐭2.\langle W_{{\bf s\cdot z}}^{\mathbf{t}}1,1\rangle=e^{-\frac{1}{2}\|{\bf s\cdot z}\|_{\mathbf{t}}^{2}}.

Finally, the assertion follows by inserting this expression into the last integral and performing the μ\mu-derivatives. ∎

Similarly, one computes the Berezin transform of the classical resolvent functions:

Lemma 4.13.

Let mm\in\mathbb{N}, z1,,zmnz_{1},\ldots,z_{m}\in\mathbb{C}^{n} and 𝐤:=(k1,,km)m{\bf k}:=(k_{1},\dots,k_{m})\in{\mathbb{N}}^{m}. For any set of complex numbers λ1,,λmi\lambda_{1},\ldots,\lambda_{m}\in\mathbb{C}\setminus i\mathbb{R} consider the function

g:=(λ12iσ𝐭(,z1))k1(λm2iσ𝐭(,zm))km.g:=\big{(}\lambda_{1}-2i\sigma_{\mathbf{t}}(\cdot,z_{1})\big{)}^{-k_{1}}\dots\big{(}\lambda_{m}-2i\sigma_{\mathbf{t}}(\cdot,z_{m})\big{)}^{-k_{m}}.

With the notation in Proposition 4.12 the Berezin transform of gg is given by:

g~(𝐭)(w):\displaystyle\widetilde{g}^{(\mathbf{t})}(w): =(Tg𝐭)(w)\displaystyle=(T_{g}^{\mathbf{t}})^{\sim}(w)
=1(𝐤𝟏)!(0,)m𝐬𝐤𝟏eΛ𝐬+2iσ𝐭(w,𝐬𝐳)𝐬𝐳𝐭2𝑑𝐬.\displaystyle=\frac{1}{(\mathbf{k-1})!}\int_{(0,\infty)^{m}}\mathbf{s}^{\mathbf{k-1}}e^{-\Lambda\cdot\mathbf{s}+2i\sigma_{\mathbf{t}}(w,\mathbf{s}\cdot\mathbf{z})-\|\mathbf{s}\cdot\mathbf{z}\|_{\mathbf{t}}^{2}}~{}d\mathbf{s}.
Proof.

The lemma follows by a direct calculation. First, note that

g~(𝐭)(w)=αw(g)1,1=ng(v+w)𝑑μ𝐭(v).\widetilde{g}^{(\mathbf{t})}(w)=\big{\langle}\alpha_{-w}(g)1,1\big{\rangle}\\ =\int_{\mathbb{C}^{n}}g(v+w)d\mu_{\mathbf{t}}(v). (4.8)

Without loss of generality may assume that Re(λj)>0\textup{Re}(\lambda_{j})>0 for j=1,,mj=1,\ldots,m such that gg has an integral representation (Laplace transform):

g(w)=1(𝐤𝟏)!(0,)m𝐬𝐤𝟏eΛ𝐬+2iσ𝐭(w,𝐬𝐳)𝑑𝐬.g(w)=\frac{1}{(\mathbf{k-1})!}\int_{(0,\infty)^{m}}\mathbf{s}^{\mathbf{k-1}}e^{-\Lambda\cdot\mathbf{s}+2i\sigma_{\mathbf{t}}(w,\mathbf{s}\cdot\mathbf{z})}d\mathbf{s}.

Inserting the last expression into (4.8) and interchanging the order of integrations shows:

g~(𝐭)(w)=1(𝐤𝟏)!(0,)m𝐬𝐤𝟏eΛ𝐬+2iσ𝐭(w,𝐬𝐳)ne2iσ𝐭(v,𝐬𝐳)𝑑μ𝐭(v)𝑑𝐬.\widetilde{g}^{(\mathbf{t})}(w)=\frac{1}{(\mathbf{k-1})!}\int_{(0,\infty)^{m}}\mathbf{s}^{\mathbf{k-1}}e^{-\Lambda\cdot\mathbf{s}+2i\sigma_{\mathbf{t}}(w,\mathbf{s}\cdot\mathbf{z})}\int_{\mathbb{C}^{n}}e^{2i\sigma_{\mathbf{t}}(v,\mathbf{s}\cdot\mathbf{z})}~{}d\mu_{\mathbf{t}}(v)~{}d\mathbf{s}.

The inner integration can be evaluated explicitly:

ne2iσ𝐭(v,𝐬𝐳)𝑑μ𝐭(v)=K𝐬𝐳𝐭,K𝐬𝐳𝐭=e𝐬𝐳𝐭2,\int_{\mathbb{C}^{n}}e^{2i\sigma_{\mathbf{t}}(v,\mathbf{s}\cdot\mathbf{z})}~{}d\mu_{\mathbf{t}}(v)=\big{\langle}K_{\mathbf{s}\cdot\mathbf{z}}^{\mathbf{t}},K_{-\mathbf{s}\cdot\mathbf{z}}^{\mathbf{t}}\big{\rangle}=e^{-\|{\bf s\cdot z}\|^{2}_{\mathbf{t}}},

which implies the statement of the lemma. ∎

We now prove a first inclusion of algebras:

Lemma 4.14.

It holds R(λ,z)𝒯𝐭(cl)R(\lambda,z)\in\mathcal{T}^{\mathbf{t}}(\mathcal{R}_{cl}) for every λi\lambda\in\mathbb{C}\setminus i\mathbb{R} and znz\in\mathbb{C}^{n}. In particular,

(n,σ𝐭)𝒯𝐭(cl).\mathcal{R}(\mathbb{C}^{n},\sigma_{\bf t})\subset\mathcal{T}^{\mathbf{t}}(\mathcal{R}_{cl}).
Proof.

Since we already know that (n,σ)𝒞1𝐭\mathcal{R}(\mathbb{C}^{n},\sigma)\subset\mathcal{C}_{1}^{\mathbf{t}}, it suffices to show that R(λ,z)~cl\widetilde{R(\lambda,z)}\in\mathcal{R}_{cl} by the correspondence theorem. Using R(λ,z)=R(λ,z)R(\lambda,z)=-R(-\lambda,-z) we may assume without loss of generality that Re(λ)>0\operatorname{Re}(\lambda)>0. By transformation of the integral and according to Lemma 4.13:

R(λ,z)~(w)\displaystyle\widetilde{R(\lambda,z)}(w) =i0eλs2isσ𝐭(w,z)s22z𝐭2𝑑s\displaystyle=-i\int_{0}^{\infty}e^{-\lambda s-2is\sigma_{\mathbf{t}}(w,z)-\frac{s^{2}}{2}\|z\|_{\mathbf{t}}^{2}}~{}ds
=i20eλ2s2isσ𝐭(2w,z)s2z𝐭2𝑑s\displaystyle=i\sqrt{2}\int_{0}^{\infty}e^{-\lambda\sqrt{2}s-2is\sigma_{\mathbf{t}}(\sqrt{2}w,z)-s^{2}\|z\|_{\mathbf{t}}^{2}}ds
=i2g~(𝐭)(2w),\displaystyle=i\sqrt{2}\>\widetilde{g}^{(\mathbf{t})}\big{(}\sqrt{2}w\big{)},

where gg is a classical resolvent, namely:

g(w)=(2λ2iσ𝐭(w,z))1.g(w)=\big{(}\sqrt{2}\lambda-2i\sigma_{\mathbf{t}}(w,-z)\big{)}^{-1}.

Note that the Berezin transform (𝐭)\sim^{(\mathbf{t})} behaves under dilations as follows:

f(𝐭)(ρw)=ρ2fρ(𝐭/ρ2)(w),ρ>0,\overset{\sim}{f}^{(\mathbf{t})}(\rho w)=\rho^{2}\cdot\overset{\sim}{f_{\rho}}^{{(\mathbf{t}/\rho^{2})}}(w),\hskip 17.22217pt\rho>0,

where fL(n)f\in L^{\infty}(\mathbb{C}^{n}), fρ(w):=f(ρw)f_{\rho}(w):=f(\rho w) and 𝐭/ρ2=(t1/ρ2,,tn/ρ2)\mathbf{t}/\rho^{2}=(t_{1}/\rho^{2},\dots,t_{n}/\rho^{2}). Hence, we obtain:

R(λ,z)~(w)=i22g2~(𝐭/2)(w).\displaystyle\widetilde{R(\lambda,z)}(w)=i2\sqrt{2}\cdot\widetilde{g_{\sqrt{2}}}^{(\mathbf{t}/2)}(w).

Since g2g_{\sqrt{2}} again is a resolvent function and cl\mathcal{R}_{cl} is invariant under the Berezin transform (𝐭/2)\sim^{(\mathbf{t}/2)} (which is simply the convolution by an appropriate Gaussian function), the inclusion R(λ,z)~cl\widetilde{R(\lambda,z)}\in\mathcal{R}_{cl} follows. ∎

Recall that an isotropic subspace VnV\subset\mathbb{C}^{n} is a (real) subspace such that σ𝐭(z,w)=0\sigma_{\mathbf{t}}(z,w)=0 for all z,wVz,w\in V. Every isotropic subspace VV is of real dimension n\leq n, and if dim(V)=n\dim_{\mathbb{R}}(V)=n, then VV is called Lagrangian. To every Lagrangian subspace VnV\subset\mathbb{C}^{n} there exists a complementary Lagrangian subspace VnV^{\prime}\subset\mathbb{C}^{n}, i.e. n=VV\mathbb{C}^{n}=V\oplus V^{\prime}. Indeed, one can choose V:={iz:zV}V^{\prime}:=\{iz:z\in V\} and we will make this choice in the following for convenience.

If we now fix a Lagrangian subspace VV, then Proposition 4.12 shows that the unital CC^{\ast} algebra

V:=C(R(λ,z):λi,zV)\displaystyle\mathcal{R}_{V}:=C^{\ast}\big{(}R(\lambda,z):~{}\lambda\in\mathbb{C}\setminus i\mathbb{R},~{}z\in V\big{)}

is commutative. Note that V\mathcal{R}_{V} is also α\alpha-invariant according to Lemma 4.8.

It is our next aim to show that V\mathcal{R}_{V}, in the sense of Theorem 4.5, corresponds to the space

cl,V:=C((λ2iσ𝐭(,z))1:zV,λi),\displaystyle\mathcal{R}_{cl,V}:=C^{\ast}\big{(}(\lambda-2i\sigma_{\mathbf{t}}(\cdot,z))^{-1}:~{}z\in V,~{}\lambda\in\mathbb{C}\setminus i\mathbb{R}\big{)},

i.e.

V=𝒯lin𝐭(cl,V).\displaystyle\mathcal{R}_{V}=\mathcal{T}_{lin}^{\mathbf{t}}(\mathcal{R}_{cl,V}).

Before we approach this goal, note the following facts:

First, by Theorem 3.13 of [14], we have 𝒯lin𝐭(cl,V)=𝒯𝐭(cl,V)\mathcal{T}_{lin}^{\mathbf{t}}(\mathcal{R}_{cl,V})=\mathcal{T}^{\mathbf{t}}(\mathcal{R}_{cl,V}), i.e. it suffices to prove that V=𝒯𝐭(cl,V)\mathcal{R}_{V}=\mathcal{T}^{\mathbf{t}}(\mathcal{R}_{cl,V}). Secondly, since cl,V\mathcal{R}_{cl,V} is also invariant under dilations ff(λ)f\mapsto f(\lambda\>\cdot) where λ>0\lambda>0, we obtain V𝒯𝐭(cl,V)\mathcal{R}_{V}\subseteq\mathcal{T}^{\mathbf{t}}(\mathcal{R}_{cl,V}) as in the proof of Lemma 4.14. Therefore, we only need to prove that Tg𝐭VT_{g}^{\mathbf{t}}\in\mathcal{R}_{V} where gg is a product of classical resolvent functions (λ2iσ𝐭(,z))1(\lambda-2i\sigma_{\mathbf{t}}(\cdot,z))^{-1} with zVz\in V. Since the operator algebra V\mathcal{R}_{V} is commutative, we have more techniques at hand for obtaining this goal. In particular, Gelfand theory turns out to be useful here. The first conceptual goal is therefore describing the Gelfand spectrum of V\mathcal{R}_{V}.

Using the explicit formulas for the Berezin transform, one observes that point evaluations of the Berezin transforms are multiplicative linear functionals on V\mathcal{R}_{V}. Since A~(v+v)=A~(v)\widetilde{A}(v+v^{\prime})=\widetilde{A}(v^{\prime}) for every vVv\in V and vVv^{\prime}\in V^{\prime} and AVA\in\mathcal{R}_{V}, there is no loss of generality in considering only point evaluations of the Berezin transform at VV^{\prime}. Since the Berezin transform is injective, we can expect the Gelfand spectrum of V\mathcal{R}_{V} to be a suitable compactification of VV^{\prime}. We will describe this compactification now and start by recalling a compactification of a real inner product space first constructed in [19]222To be more precise, therein it was only described for n\mathbb{R}^{n}. It is straightforward to generalize the procedure to any finite dimensional real inner product space.. Therein the author described the maximal ideal space of cl\mathcal{R}_{cl} as such a compactification of 2n\mathbb{R}^{2n}.

Let XX be a finite-dimensional real inner product space. We denote by PYP_{Y} the orthogonal projection onto a given subspace YXY\subset X. By Graff(X)\operatorname{Graff}(X) we denote the affine Grassmannian of XX, i.e. the set of all affine subspaces of XX. As a set, this can be written as

Graff(X)={x+Y:Ya linear subspace of X and xY}.\displaystyle\operatorname{Graff}(X)=\big{\{}x+Y:~{}Y\text{\it a linear subspace of }X\text{ \it and }x\perp Y\big{\}}.

The precise topology with which Graff(X)\operatorname{Graff}(X) is endowed can be found in [19] and will not be described here. We will denote by γX\gamma X the set Graff(X)\operatorname{Graff}(X) endowed with this particular topology333In [19], the compactification is denoted by Ω\Omega. We chose to name it differently, as the symbol Ω\Omega is somewhat ambiguous in a symplectic context. We collect some facts about γX\gamma X in the following lemma:

Lemma 4.15 ([19]).

Let XX be a finite-dimensional real inner product space.

  1. 1.

    γX\gamma X is a compact Hausdorff space.

  2. 2.

    Together with the embedding Xxx+{0}γXX\ni x\mapsto x+\{0\}\in\gamma X, γX\gamma X is a compactification of XX.

  3. 3.

    A net (xι+Yι)ιIγX(x_{\iota}+Y_{\iota})_{\iota\in I}\subset\gamma X converges to x+YγXx+Y\in\gamma X if and only if the following hold:

    • PYxιιIxP_{Y^{\perp}}x_{\iota}\overset{\iota\in I}{\longrightarrow}x

    • eventually YιYY_{\iota}\subseteq Y

    • There is no affine subspace x+Yx+Yx^{\prime}+Y^{\prime}\subsetneq x+Y such that there exists a subnet of (xι+Yι)(x_{\iota}+Y_{\iota}) with P(Y)xιxP_{(Y^{\prime})^{\perp}}x_{\iota}\to x^{\prime} and YιYY_{\iota}\subseteq Y^{\prime} eventually (along the subnet).

One of the main theorems of [19] is the following:

Theorem 4.16 ([19]).

The Gelfand spectrum (cl)\mathcal{M}(\mathcal{R}_{cl}) of cl\mathcal{R}_{cl} can be identified with the above compactification of 2n\mathbb{R}^{2n}, i.e. (cl)γ2n\mathcal{M}(\mathcal{R}_{cl})\cong\gamma\mathbb{R}^{2n}.

Let us briefly describe the identification of (cl)\mathcal{M}(\mathcal{R}_{cl}) with γ2n\gamma\mathbb{R}^{2n} in more detail: Given a function fclf\in\mathcal{R}_{cl} and an affine line x+span{y}2nx+\operatorname{span}\{y\}\subset\mathbb{R}^{2n}, y2ny\in\mathbb{R}^{2n}, it is not hard to verify that limαf(x+αy)\lim_{\alpha\to\infty}f(x+\alpha y) exists. For an affine subspace x+Y2nx+Y\subset\mathbb{R}^{2n} and any yYy\in Y with y=1\|y\|=1, the value of this limit is almost everywhere, with respect to the surface measure on {yY:y=1}\{y\in Y:~{}\|y\|=1\}, the same. If we denote this value by φx+Y(f)\varphi_{x+Y}(f), then this defines a multiplicative linear functional and every element of (cl)\mathcal{M}(\mathcal{R}_{cl}) can be obtained in this way.

Indeed, the same can be done for cl,V\mathcal{R}_{cl,V}, and the following holds true:

Proposition 4.17.

The Gelfand spectrum (cl,V)\mathcal{M}(\mathcal{R}_{cl,V}) of cl,V\mathcal{R}_{cl,V} can be identified with the above compactification of VV^{\prime}, i.e. (cl,V)γV\mathcal{M}(\mathcal{R}_{cl,V})\cong\gamma V^{\prime}.

Here, we consider VV^{\prime} as a real inner product space with the 𝐭\mathbf{t}-weighted inner product induced from 2nn\mathbb{R}^{2n}\cong\mathbb{C}^{n}, i.e. z,w=j=1nRe(zw¯)tj\langle z,w\rangle=\sum_{j=1}^{n}\frac{\operatorname{Re}(z\cdot\overline{w})}{t_{j}}. We will not need the result in its fulll strength and we leave it as an exercise to adapt the arguments from [19]. It is sufficient and elementary to verify that via the embedding Vvv+{0}γVV^{\prime}\ni v\mapsto v+\{0\}\in\gamma V^{\prime}, every classical resolvent (λ2iσ𝐭(,z))1(\lambda-2i\sigma_{\mathbf{t}}(\cdot,z))^{-1} with zVz\in V extends to a function in C(γV)C(\gamma V^{\prime}).

As we will see next, the Gelfand spectrum of V\mathcal{R}_{V} is indeed the same as the one of cl,V\mathcal{R}_{cl,V}:

Proposition 4.18.

The Gelfand spectrum (V)\mathcal{M}(\mathcal{R}_{V}) of V\mathcal{R}_{V} can be identified with (V)(cl,V)γV\mathcal{M}(\mathcal{R}_{V})\cong\mathcal{M}(\mathcal{R}_{cl,V})\cong\gamma V^{\prime}, where all multiplicative linear functionals on V\mathcal{R}_{V} have the form:

ψx+Y(A)=φx+Y(A~|V).\displaystyle\psi_{x+Y}(A)=\varphi_{x+Y}\big{(}\widetilde{A}|_{V^{\prime}}\big{)}.

Here, we denote by φx+Y\varphi_{x+Y} the multiplicative functional of cl,V\mathcal{R}_{cl,V} introduced above.

Before proving Proposition 4.18, let us derive our intended result from this:

Theorem 4.19.

It is V=𝒯lin𝐭(cl,V)\mathcal{R}_{V}=\mathcal{T}_{lin}^{\mathbf{t}}(\mathcal{R}_{cl,V}).

Proof.

The inclusion V𝒯lin𝐭(cl,V)\mathcal{R}_{V}\subset\mathcal{T}_{lin}^{\mathbf{t}}(\mathcal{R}_{cl,V}) was already stated above. Let gcl,Vg\in\mathcal{R}_{cl,V}. Since cl,V\mathcal{R}_{cl,V} is translation invariant and closed, we conclude that g~(𝐭)=Tg𝐭~cl,V=C((V))\widetilde{g}^{(\mathbf{t})}=\widetilde{T_{g}^{\mathbf{t}}}\in\mathcal{R}_{cl,V}=C(\mathcal{M}(\mathcal{R}_{V})). Therefore, g~(𝐭)\widetilde{g}^{(\mathbf{t})} is the Gelfand transform of an operator in V\mathcal{R}_{V}. This operator must be Tg𝐭T_{g}^{\mathbf{t}}, i.e. Tg𝐭VT_{g}^{\mathbf{t}}\in\mathcal{R}_{V}. ∎

We will present a sequence of lemmas that lead to a proof of Proposition 4.18.

Lemma 4.20.

For any resolvent R(λ,z)VR(\lambda,z)\in\mathcal{R}_{V}, where zVz\in V, and x,yVx,y\in V^{\prime} it holds:

R(λ,z)~(x+y)=R(λ,z)~(x+Pspan{iz}y).\displaystyle\widetilde{R(\lambda,z)}(x+y)=\widetilde{R(\lambda,z)}\big{(}x+P_{\operatorname{span}\{iz\}}y\big{)}.
Proof.

Let zVz\in V and first observe that σ𝐭(z,y)=0\sigma_{\mathbf{t}}(z,y)=0 iff yizy\perp iz (note that izViz\in V^{\prime} by our choice V={iz:zV}V^{\prime}=\{iz:z\in V\}). Hence, we have

R(λ,z)~(x+y)\displaystyle\widetilde{R(\lambda,z)}(x+y) =R(λ,z)~(x+(IPspan{iz})y+Pspan{iz}y)\displaystyle=\widetilde{R(\lambda,z)}\big{(}x+(I-P_{\operatorname{span}\{iz\}})y+P_{\operatorname{span}\{iz\}}y\big{)}
=R(λ,z)~(x+Pspan{iz}y),\displaystyle=\widetilde{R(\lambda,z)}\big{(}x+P_{\operatorname{span}\{iz\}}y\big{)},

where the last equality follows from the formula for the Berezin transform of a resolvent in Proposition 4.12. ∎

Lemma 4.21.

For any resolvent R(λ,z)VR(\lambda,z)\in\mathcal{R}_{V} where zVz\in V and any affine line x+span{y}γVx+\operatorname{span}\{y\}\in\gamma V^{\prime} the limit limαR(λ,z)~(x+αy)\lim_{\alpha\to\infty}\widetilde{R(\lambda,z)}(x+\alpha y) exists and is given by

limαR(λ,z)~(x+αy)={R(λ,z)~(x),σ𝐭(z,y)=0,0,σ𝐭(z,y)0.\displaystyle\lim_{\alpha\to\infty}\widetilde{R(\lambda,z)}(x+\alpha y)=\begin{cases}\widetilde{R(\lambda,z)}(x),\quad&\sigma_{\mathbf{t}}(z,y)=0,\\ 0,\quad&\sigma_{\mathbf{t}}(z,y)\neq 0.\end{cases} (4.9)
Proof.

If σ𝐭(z,y)=0\sigma_{\mathbf{t}}(z,y)=0, then yizy\perp iz such that Pspan{iz}y=0P_{\operatorname{span}\{iz\}}y=0 and the equality follows from the previous lemma. Assume that σ𝐭(z,y)0\sigma_{\mathbf{t}}(z,y)\neq 0 and let Re(λ)>0\operatorname{Re}(\lambda)>0 (the case Re(λ)>0\operatorname{Re}(\lambda)>0 follows similarly). According to Proposition 4.12 the Berezin transform of the resolvent at x+αyx+\alpha y has the value:

R(λ,z)~(x+αy)=0eλs2isασ𝐭(y,z)2isσ𝐭(x,z)s22z𝐭2𝑑s.\displaystyle\widetilde{R(\lambda,z)}(x+\alpha y)=\int_{0}^{\infty}e^{-\lambda s-2is\alpha\sigma_{\mathbf{t}}(y,z)-2is\sigma_{\mathbf{t}}(x,z)-\frac{s^{2}}{2}\|z\|_{\mathbf{t}}^{2}}~{}ds.

We use integration by parts in order to show that the right hand side converges to 0 as α\alpha\rightarrow\infty. With the short notation σ:=σ𝐭(y,z)0\sigma:=\sigma_{\mathbf{t}}(y,z)\neq 0 we have:

R(λ,z)~(x+αy)\displaystyle\widetilde{R(\lambda,z)}(x+\alpha y)
=12iασ0dds[e2isασ]eλs2isσ𝐭(x,z)s22z𝐭2𝑑s\displaystyle\ =-\frac{1}{2i\alpha\sigma}\int_{0}^{\infty}\frac{\rm d}{{\rm d}s}\left[e^{-2is\alpha\sigma}\right]e^{-\lambda s-2is\sigma_{\mathbf{t}}(x,z)-\frac{s^{2}}{2}\|z\|_{\mathbf{t}}^{2}}~{}ds
=12iασ([e2isασλs2isσ𝐭(x,z)s22z𝐭2]s=0=1\displaystyle\ =-\frac{1}{2i\alpha\sigma}\Big{(}\underbrace{\left[e^{-2is\alpha\sigma-\lambda s-2is\sigma_{\mathbf{t}}(x,z)-\frac{s^{2}}{2}\|z\|_{\mathbf{t}}^{2}}\right]_{s=0}^{\infty}}_{=-1}
0e2isασ(λ2iσ𝐭(x,z)sz𝐭2)eλs2isσ𝐭(x,z)s22z𝐭2ds).\displaystyle\hskip 28.45274pt-\int_{0}^{\infty}e^{-2is\alpha\sigma}\big{(}-\lambda-2i\sigma_{\mathbf{t}}(x,z)-s\|z\|_{\mathbf{t}}^{2}\big{)}e^{-\lambda s-2is\sigma_{\mathbf{t}}(x,z)-\frac{s^{2}}{2}\|z\|_{\mathbf{t}}^{2}}~{}ds\Big{)}.

Since the integral

0|λ+2iσ𝐭(x,z)+sz𝐭2|eλss22z𝐭2ds\displaystyle\int_{0}^{\infty}\big{|}\lambda+2i\sigma_{\mathbf{t}}(x,z)+s\|z\|_{\mathbf{t}}^{2}\big{|}e^{-\lambda s-\frac{s^{2}}{2}\|z\|_{\mathbf{t}}^{2}}~{}ds

over the absolute value is finite, we obtain R(λ,z)~(x+αy)0\widetilde{R(\lambda,z)}(x+\alpha y)\to 0 as α\alpha\to\infty. ∎

Lemma 4.22.

Let R(λ,z)R(\lambda,z) with zVz\in V be a resolvent in V\mathcal{R}_{V}. Then, its restriction R(λ,z)~|V\widetilde{R(\lambda,z)}|_{V^{\prime}} to VV^{\prime} extends to a continuous function on γV\gamma V^{\prime}.

Proof.

We first describe the values that (the extension of) R(λ,z)~|V\widetilde{R(\lambda,z)}|_{V^{\prime}} takes on γV\gamma V^{\prime} and discuss continuity in a second step.

The value of the Berezin transform at xx+{0}x\cong x+\{0\} is given by R(λ,z)~(x)\widetilde{R(\lambda,z)}(x). At each affine line x+span{y}x+\operatorname{span}\{y\} the value of the extension of the Berezin transform to γV\gamma V^{\prime} is defined by the right hand side of (4.9). For a given affine spaces x+Yx+Y with dim(Y)2\dim_{\mathbb{R}}(Y)\geq 2, we distinguish two cases: If Y{wV:σ𝐭(w,z)=0}Y\subseteq\{w\in V^{\prime}:\sigma_{\mathbf{t}}(w,z)=0\}, then the value of R(λ,z)~(x+Y)\widetilde{R(\lambda,z)}(x+Y) coincides with R(λ,z)~(x)\widetilde{R(\lambda,z)}(x). Otherwise, if Y{wV:σ𝐭(w,z)=0}Y\not\subseteq\{w\in V^{\prime}:\sigma_{\mathbf{t}}(w,z)=0\}, then we set it to be zero.

Note that these choices are in accordance with the definition of the multiplicative linear functional φx+Y\varphi_{x+Y}. We explain what we mean by this only in the last example: If Y{wV:σ𝐭(w,z)=0}Y\not\subseteq\{w\in V^{\prime}\>:\>\sigma_{\mathbf{t}}(w,z)=0\}, then {wY:σ𝐭(w,z)=0}\{w\in Y:\sigma_{\mathbf{t}}(w,z)=0\} is a subspace of YY of dimension strictly smaller than the dimension of YY. Hence, the unit sphere of {wY:σ𝐭(w,z)=0}\{w\in Y:\sigma_{\mathbf{t}}(w,z)=0\} is a zero set with respect to the surface measure of the unit sphere of YY. Hence, φx+Y(R(λ,z)~|V)=0.\varphi_{x+Y}(\widetilde{R(\lambda,z)}|_{V^{\prime}})=0.

It remains to prove that the above extension of R(λ,z)~|V\widetilde{R(\lambda,z)}|_{{V^{\prime}}} from VV^{\prime} to the compactification γV\gamma V^{\prime} is in fact continuous. According to Bourbaki’s Extension Theorem, [4, Theorem 1, p. 82], it suffices to show that

R(λ,z)~(xι)φx+Y(R(λ,z)~)\widetilde{R(\lambda,z)}(x_{\iota})\to\varphi_{x+Y}\big{(}\widetilde{R(\lambda,z)}\big{)} (4.10)

for any net (xι)ιIV(x_{\iota})_{\iota\in I}\subset V^{\prime} such that xι+{0}x+Yx_{\iota}+\{0\}\to x+Y. We distinguish several cases:

  1. 1.

    If xι+{0}x+{0}x_{\iota}+\{0\}\to x+\{0\}, then (4.10) follows from the continuity of the Berezin transform R(λ,z)~\widetilde{R(\lambda,z)} on VV^{\prime}.

  2. 2.

    If xι+{0}x+Yx_{\iota}+\{0\}\to x+Y with Y{wV:σ𝐭(w,z)=0}Y\subseteq\{w\in V^{\prime}:~{}\sigma_{\mathbf{t}}(w,z)=0\}, then:

    R(λ,z)~(xι)\displaystyle\widetilde{R(\lambda,z)}(x_{\iota}) =R(λ,z)~(xιPYxι)=R(λ,z)~(PYxι)\displaystyle=\widetilde{R(\lambda,z)}\big{(}x_{\iota}-P_{Y}x_{\iota}\big{)}=\widetilde{R\big{(}\lambda,z)}\big{(}P_{Y^{\perp}}x_{\iota}\big{)}
    𝜄R(λ,z)~(x)=φx+Y(R(λ,z)~).\displaystyle\overset{\iota}{\longrightarrow}\widetilde{R(\lambda,z)}(x)=\varphi_{x+Y}\big{(}\widetilde{R(\lambda,z)}\big{)}.
  3. 3.

    If xι+{0}x+Yx_{\iota}+\{0\}\to x+Y with Y{wV:σ𝐭(w,z)=0}Y\not\subseteq\{w\in V^{\prime}:~{}\sigma_{\mathbf{t}}(w,z)=0\}, then we conclude that

    R(λ,z)~(xι)𝜄0=φx+Y(R(λ,z)~).\widetilde{R(\lambda,z)}(x_{\iota})\overset{\iota}{\longrightarrow}0=\varphi_{x+Y}\big{(}\widetilde{R(\lambda,z)}\big{)}.

    In fact, assuming the opposite there exists a subnet, also denoted by (xι)(x_{\iota}), such that R(λ,z)~(xι)\widetilde{R(\lambda,z)}(x_{\iota}) is bounded away from zero. Take w0Yw_{0}\in Y such that σ𝐭(w0,z)0\sigma_{\mathbf{t}}(w_{0},z)\neq 0 and put Z:=span{w0}YZ:=\textup{span}\{w_{0}\}\subset Y. By YYY^{\prime}\subset Y denote the orthogonal complement of ZZ in YY. We obtain a decomposition of VV^{\prime}:

    V=YY=YYZ.V^{\prime}=Y^{\perp}\oplus Y=Y^{\perp}\oplus Y^{\prime}\oplus Z. (4.11)
    1. (a)

      If PZxιP_{Z}x_{\iota} was unbounded, we could pass to a subnet such that PZxι\|P_{Z}x_{\iota}\|\to\infty. For this subnet, one could show by using the method of stationary phase (as in the proof of Lemma 4.21) that R(λ,z)~(xι)0\widetilde{R(\lambda,z)}(x_{\iota})\to 0. However, this is impossible as we assumed that R(λ,z)~(xι)\widetilde{R(\lambda,z)}(x_{\iota}) is bounded away from zero.

    2. (b)

      If PZxιP_{Z}x_{\iota} is bounded, we can pass to a subnet such that PZxιP_{Z}x_{\iota} converges (say, to some x0Zx_{0}\in Z). From the orthogonal decomposition (4.11) we see that:

      P(Y)xι=PYxι+PZxι𝜄x+x0YZ.\displaystyle P_{(Y^{\prime})^{\perp}}x_{\iota}=P_{Y^{\perp}}x_{\iota}+P_{Z}x_{\iota}\overset{\iota}{\longrightarrow}x+x_{0}\in Y^{\perp}\oplus Z.

      From x0ZYx_{0}\in Z\subset Y it follows: x+x0+Yx+Yx+x_{0}+Y^{\prime}\subset x+Y. However this contradicts the assumption xι+{0}𝜄x+Yx_{\iota}+\{0\}\overset{\iota}{\longrightarrow}x+Y. ∎

Since the Berezin transform is multiplicative on VV^{\prime} and AkAA_{k}\to A in V\mathcal{R}_{V} implies Ak~A~\widetilde{A_{k}}\to\widetilde{A} uniformly, we conclude that A~\widetilde{A} for every AVA\in\mathcal{R}_{V} extends to a continuous function on γV\gamma V^{\prime}. Now we can give the proof of Proposition 4.18:

Proof of Proposition 4.18.

We claim that the map

Φ:VC(γV):Φ(A)=[(x+Y)φx+Y(A~|V)]\Phi:\mathcal{R}_{V}\to C(\gamma V^{\prime}):\>\Phi(A)=\big{[}(x+Y)\mapsto\varphi_{x+Y}\big{(}\widetilde{A}_{|_{V^{\prime}}}\big{)}\big{]}

is a bijective and unital homomorphism of CC^{\ast} algebras. By what has been said before, Φ\Phi defines a continuous and injective unital \ast-homomorphism. We are left with verifying surjectivity. Since the range of Φ\Phi is a \ast-algebra, we only need to show that it separates points of γV\gamma V^{\prime} in order to apply the Stone-Weierstrass Theorem. Showing that the points are separated by Φ(V)\Phi(\mathcal{R}_{V}) is easy, as we can always choose the Berezin transform of a resolvent for separating two given sets. Careful inspection of the proof of the previous lemma indeed gives a guideline on how to choose the resolvent. We briefly explain this here. In what follows, we always assume that Re(λ)>0\operatorname{Re}(\lambda)>0.

  1. 1.

    0+V0+V^{\prime} is separated from any other affine subspace in the following way: If YVY\subsetneq V^{\prime} is a proper subspace, then let 0vY0\neq v\perp Y and consider R(λ,iαv)R(\lambda,i\alpha v) for α\alpha\in\mathbb{R}. Then, α\alpha can be chosen such that

    Φ(R(λ,iαv))(v+Y)=ψv+Y(R(λ,iαv))=R(λ,iαv)~(v)0.\Phi\big{(}R(\lambda,i\alpha v)\big{)}(v+Y)=\psi_{v+Y}\Big{(}R(\lambda,i\alpha v)\Big{)}=\widetilde{R(\lambda,i\alpha v)}(v)\neq 0.

    However, ψ0+V(R(λ,iαv))=0\psi_{0+V^{\prime}}(R(\lambda,i\alpha v))=0 as long as α0\alpha\neq 0. Hence the affine spaces v+Yv+Y and 0+V0+V^{\prime} are separated. Separating 0+V0+V^{\prime} from 0+Y0+Y can be done by considering R(λ,iv)R(\lambda,iv).

  2. 2.

    Affine spaces x1+Yx2+Yx_{1}+Y\neq x_{2}+Y are separated as follows: We necessarily have YVY\not=V^{\prime} and x1x2x_{1}\neq x_{2}. If x1=ρx2x_{1}=\rho x_{2} for some ρ\rho\in\mathbb{R}, then consider R(λ,iαx2)R(\lambda,i\alpha x_{2}) for suitable α\alpha\in\mathbb{R}. Otherwise, we can find z0Yz_{0}\perp Y such that Pspan{z0}x1Pspan{z0}x2P_{\operatorname{span}\{z_{0}\}}x_{1}\neq P_{\operatorname{span}\{z_{0}\}}x_{2}. Let z=iz0Vz=iz_{0}\in V such that Y{wV:σ𝐭(w,z)=0}Y\subset\{w\in V^{\prime}:\sigma_{\mathbf{t}}(w,z)=0\}. Hence for α{0}\alpha\in\mathbb{R}\setminus\{0\}:

    ψx1+Y(R(λ,αz))\displaystyle\psi_{x_{1}+Y}\big{(}R(\lambda,\alpha z)\big{)} =R(λ,αz)~(Pspan{z0}x1),\displaystyle=\widetilde{R(\lambda,\alpha z)}\big{(}P_{\operatorname{span}\{z_{0}\}}x_{1}\big{)},
    ψx2+Y(R(λ,αz))\displaystyle\psi_{x_{2}+Y}\big{(}R(\lambda,\alpha z)\big{)} =R(λ,αz)~(Pspan{z0}x2).\displaystyle=\widetilde{R(\lambda,\alpha z)}\big{(}P_{\operatorname{span}\{z_{0}\}}x_{2}\big{)}.

    Now, one has to choose α\alpha accordingly such that these values are different (see the formula in Proposition 4.12).

  3. 3.

    If Y1Y2Y_{1}\neq Y_{2} are proper subspaces of VV^{\prime}, then we consider two cases:

    1. (a)

      Recall that we always have xYx\perp Y for an affine subspace x+Yx+Y. If x1=ρx2x_{1}=\rho x_{2} with ρ\rho\in\mathbb{R} are real multiples of each other, then we obtain with α\alpha\in\mathbb{R}:

      ψx1+Y1(R(λ,iαx2))\displaystyle\psi_{x_{1}+Y_{1}}\big{(}R(\lambda,i\alpha x_{2})\big{)} =R(λ,iαx2)~(ρx2)\displaystyle=\widetilde{R(\lambda,i\alpha x_{2})}(\rho x_{2})
      =0eλse2isραx2𝐭2s2α22x2𝐭2𝑑s,\displaystyle=\int_{0}^{\infty}e^{-\lambda s}e^{-2is\rho\alpha\|x_{2}\|_{\mathbf{t}}^{2}-\frac{s^{2}\alpha^{2}}{2}\|x_{2}\|_{\mathbf{t}}^{2}}~{}ds,
      ψx2+Y2(R(λ,iαx2))\displaystyle\psi_{x_{2}+Y_{2}}\big{(}R(\lambda,i\alpha x_{2})\big{)} =R(λ,iαx2)~(x2)\displaystyle=\widetilde{R(\lambda,i\alpha x_{2})}(x_{2})
      =0eλse2isαx2𝐭2s2α22x2𝐭2𝑑s.\displaystyle=\int_{0}^{\infty}e^{-\lambda s}e^{-2is\alpha\|x_{2}\|_{\mathbf{t}}^{2}-\frac{s^{2}\alpha^{2}}{2}\|x_{2}\|_{\mathbf{t}}^{2}}~{}ds.

      Now, arrange α\alpha such that these are not equal.

    2. (b)

      If x1x_{1} and x2x_{2} are not real multiples of each other, then it must be either Pspan{x1}x2x1P_{\operatorname{span}\{x_{1}\}}x_{2}\neq x_{1} or Pspan{x2}x1x2P_{\operatorname{span}\{x_{2}\}}x_{1}\neq x_{2}. Assume that Pspan{x1}x2x1P_{\operatorname{span}\{x_{1}\}}x_{2}\neq x_{1}. For α\alpha\in\mathbb{R}:

      ψx1+Y1(R(λ,iαx1))\displaystyle\psi_{x_{1}+Y_{1}}\big{(}R(\lambda,i\alpha x_{1})\big{)} =R(λ,iαx1)~(x1)\displaystyle=\widetilde{R(\lambda,i\alpha x_{1})}(x_{1})
      =0eλse2isαx1𝐭2s2α22x1𝐭2𝑑s.\displaystyle=\int_{0}^{\infty}e^{-\lambda s}e^{-2is\alpha\|x_{1}\|_{\mathbf{t}}^{2}-\frac{s^{2}\alpha^{2}}{2}\|x_{1}\|_{\mathbf{t}}^{2}}~{}ds.

      The value at x2+Y2x_{2}+Y_{2} now depends on whether x1x_{1} is orthogonal to Y2Y_{2} or not: If x1Y2x_{1}\perp Y_{2}, then

      ψx2+Y2(R(λ,iαx1))\displaystyle\psi_{x_{2}+Y_{2}}\big{(}R(\lambda,i\alpha x_{1})\big{)} =R(λ,iαx1)~(x2)\displaystyle=\widetilde{R(\lambda,i\alpha x_{1})}(x_{2})
      =0eλse2isασ𝐭(Pspan{ix1}x2,x1)s2α22x1𝐭2𝑑s.\displaystyle=\int_{0}^{\infty}e^{-\lambda s}e^{-2is\alpha\sigma_{\mathbf{t}}(P_{\operatorname{span}\{ix_{1}\}}x_{2},x_{1})-\frac{s^{2}\alpha^{2}}{2}\|x_{1}\|_{\mathbf{t}}^{2}}~{}ds.

      Since σ𝐭(Pspan{ix1}x2,x1)x1𝐭2\sigma_{\mathbf{t}}(P_{\operatorname{span}\{ix_{1}\}}x_{2},x_{1})\neq\|x_{1}\|_{\mathbf{t}}^{2}, we can arrange α\alpha such that the values of ψxj+Yj(R(λ,iαx1))\psi_{x_{j}+Y_{j}}(R(\lambda,i\alpha x_{1})) for j=1,2j=1,2 are different. If, on the other hand, x1⟂̸Y2x_{1}\not\perp Y_{2}, then

      ψx2+Y2(R(λ,iαx1))=0.\displaystyle\psi_{x_{2}+Y_{2}}\big{(}R(\lambda,i\alpha x_{1})\big{)}=0.

      Hence, we have to arrange α\alpha such that ψx1+Y1(R(λ,iαx1))0\psi_{x_{1}+Y_{1}}(R(\lambda,i\alpha x_{1}))\neq 0 (e.g. by letting α=0\alpha=0).∎

Before continuing, we state a consequence of the previous considerations:

Corollary 4.23.

Let mm\in\mathbb{N} and λ1,,λmi\lambda_{1},\ldots,\lambda_{m}\in\mathbb{C}\setminus i\mathbb{R}. Consider the function

g=(λ12iσ𝐭(,z1))k1(λm2iσ𝐭(,zm))km,g=\big{(}\lambda_{1}-2i\sigma_{\mathbf{t}}(\cdot,z_{1})\big{)}^{-k_{1}}\cdot\dots\cdot\big{(}\lambda_{m}-2i\sigma_{\mathbf{t}}(\cdot,z_{m})\big{)}^{-k_{m}},

where z1,,zmVz_{1},\dots,z_{m}\in V and VnV\subset\mathbb{C}^{n} is Lagrangian. Then, Tg𝐭(n,σ𝐭)T_{g}^{\mathbf{t}}\in\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}).

Proof.

By definition we have gcl,Vg\in\mathcal{R}_{cl,V}. Now Theorem 4.19 implies that Tg𝐭𝒯lin𝐭(cl,V)=V(n,σ𝐭)T_{g}^{\mathbf{t}}\in\mathcal{T}_{lin}^{\mathbf{t}}(\mathcal{R}_{cl,V})=\mathcal{R}_{V}\subset\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}). ∎

Now we consider resolvent functions for which the vectors zjz_{j} are either taken from VV or from VV^{\prime}.

Corollary 4.24.

Let VV and VV^{\prime} be two Lagrangian subspaces of n\mathbb{C}^{n} such that n=VV\mathbb{C}^{n}=V\oplus V^{\prime}. Let m,nm,\ell\geq n and z1,,zmVz_{1},\dots,z_{m}\in V and w1,,wVw_{1},\dots,w_{\ell}\in V^{\prime} such that V=span{z1,,zm}V=\operatorname{span}_{\mathbb{R}}\{z_{1},\dots,z_{m}\} and V=span{w1,,w}V^{\prime}=\operatorname{span}_{\mathbb{R}}\{w_{1},\dots,w_{\ell}\}. Set

g=j=1m(λj2iσ𝐭(,zj))kjj=1(λj2iσ𝐭(,wj))kj,\displaystyle g=\prod_{j=1}^{m}\big{(}\lambda_{j}-2i\sigma_{\mathbf{t}}(\cdot,z_{j})\big{)}^{-k_{j}}\cdot\prod_{j=1}^{\ell}\big{(}\lambda_{j}^{\prime}-2i\sigma_{\mathbf{t}}(\cdot,w_{j})\big{)}^{-k_{j}^{\prime}},

where kj,kjk_{j},k_{j}^{\prime}\in\mathbb{N} and λj,λji\lambda_{j},\lambda_{j}^{\prime}\in\mathbb{C}\setminus i\mathbb{R}. Then, Tg𝐭(n,σ𝐭)T_{g}^{\mathbf{t}}\in\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}).

Proof.

For v=v1+v2VV=nv=v_{1}+v_{2}\in V\oplus V^{\prime}=\mathbb{C}^{n} we have

g(v)=j=1m(λj2iσ𝐭(v2,zj))kjj=1(λj2iσ𝐭(v1,wj))kj.\displaystyle g(v)=\prod_{j=1}^{m}\big{(}\lambda_{j}-2i\sigma_{\mathbf{t}}(v_{2},z_{j})\big{)}^{-k_{j}}\cdot\prod_{j=1}^{\ell}\big{(}\lambda_{j}^{\prime}-2i\sigma_{\mathbf{t}}(v_{1},w_{j})\big{)}^{-k_{j}^{\prime}}.

As v1v_{1}\to\infty, the first factor tends to 0. As v2v_{2}\to\infty, the second factor tends to 0. In conclusion, gC0(n)g\in C_{0}(\mathbb{C}^{n}). Hence, Tg𝐭𝒦(F𝐭2)T_{g}^{\mathbf{t}}\in\mathcal{K}(F_{\mathbf{t}}^{2}). Since 𝒦(F𝐭2)(n,σ𝐭)\mathcal{K}(F_{\mathbf{t}}^{2})\subset\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}) according to Proposition 4.11 the statement follows. ∎

Theorem 4.25.

Let n=1n=1, then 𝒟0t=cl\mathcal{D}_{0}^{t}=\mathcal{R}_{cl}.

Proof.

If n=1n=1, then any classical resolvent function gg either fulfills the assumption of Corollary 4.23 or the assumptions of Corollary 4.24. In either case, we obtain Tg𝐭(n,σ𝐭)T_{g}^{\mathbf{t}}\in\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}}). ∎

For n>1n>1, the situation is more complicated. We will show only a weaker result for this case.

Since every znz\in\mathbb{C}^{n} is contained in some Lagrangian subspace VV, we conclude that each resolvent R(λ,z)R(\lambda,z) is contained in some V\mathcal{R}_{V}. Thus, we see that

0:=V Lagrangian subspaceV\displaystyle\mathcal{R}_{0}:=\sum_{V\text{ Lagrangian subspace}}\mathcal{R}_{V}

contains every resolvent R(λ,z)R(\lambda,z). Further, since V=𝒯lin𝐭(cl,V)\mathcal{R}_{V}=\mathcal{T}_{lin}^{\bf t}(\mathcal{R}_{cl,V}), we obtain

0¯=𝒯lin𝐭(cl,0¯),\displaystyle\overline{\mathcal{R}_{0}}=\mathcal{T}_{lin}^{\mathbf{t}}(\overline{\mathcal{R}_{cl,0}}),

where cl,0\mathcal{R}_{cl,0} is given by

cl,0:=V Lagrangian subspacecl,V.\displaystyle\mathcal{R}_{cl,0}:=\sum_{V\text{ Lagrangian subspace}}\mathcal{R}_{cl,V}.

As a weak substitute for Theorem 4.25 in higher dimensions, we get:

Proposition 4.26.

(n,σ𝐭)=𝒯𝐭(cl,0¯)\mathcal{R}(\mathbb{C}^{n},\sigma_{\mathbf{t}})=\mathcal{T}^{\mathbf{t}}(\overline{\mathcal{R}_{cl,0}}).

It might of course be true that cl,0¯=cl\overline{\mathcal{R}_{cl,0}}=\mathcal{R}_{cl}. In this case, the previous result would imply 𝒟0t=cl\mathcal{D}_{0}^{t}=\mathcal{R}_{cl}. Nevertheless, so far this remains an open question.

5 Infinite dimensional symplectic space

As is well-known there are significant differences between the resolvent algebras of finite and infinite dimensional symplectic spaces: in finite dimensions every two regular irreducible representations are unitarily equivalent. However, in the case of infinite dimensional symplectic spaces, this statement is false. Nevertheless, we can build particular representations of the resolvent algebra by using ideas from the previous sections.

Let 𝐭=(tk)k=11()\mathbf{t}=(t_{k})_{k=1}^{\infty}\in\ell^{1}(\mathbb{N}) be a summable sequence of strictly positive real numbers. Then, we set :=2()\mathcal{H}:=\ell^{2}(\mathbb{N}) with the usual norm 2\|\cdot\|_{\ell^{2}} and define:

1/2𝐭\displaystyle\mathcal{H}_{1/2}^{\mathbf{t}} :={z2():n=1|zn|2tn<},\displaystyle:=\Big{\{}z\in\ell^{2}(\mathbb{N}):~{}\sum_{n=1}^{\infty}\frac{|z_{n}|^{2}}{t_{n}}<\infty\Big{\}},
1𝐭\displaystyle\mathcal{H}_{1}^{\mathbf{t}} :={z2():n=1|zn|2tn2<}.\displaystyle:=\Big{\{}z\in\ell^{2}(\mathbb{N}):~{}\sum_{n=1}^{\infty}\frac{|z_{n}|^{2}}{t_{n}^{2}}<\infty\Big{\}}.

We may think of 1𝐭\mathcal{H}_{1}^{\mathbf{t}} and 1/2𝐭\mathcal{H}_{1/2}^{\mathbf{t}} as the range of BB and B\sqrt{B}, respectively, where BB is the trace class operator obtained by linearly extending the map entnene_{n}\mapsto t_{n}e_{n} with {en}\{e_{n}\} being the standard orthonormal basis of 2()\ell^{2}(\mathbb{N}). We consider 1/2𝐭\mathcal{H}_{1/2}^{\mathbf{t}} as a symplectic Hilbert space equipped with the symplectic form:

σ𝐭(z,w):=n=1Im(znwn¯)tn,z=(zn),w=(wn)1/2𝐭.\displaystyle\sigma_{\mathbf{t}}(z,w):=\sum_{n=1}^{\infty}\frac{\operatorname{Im}(z_{n}\cdot\overline{w_{n}})}{t_{n}},\hskip 12.91663ptz=(z_{n}),w=(w_{n})\in\mathcal{H}_{1/2}^{\mathbf{t}}.

Note that σ𝐭(z,w)\sigma_{\mathbf{t}}(z,w) is well-defined even for z1𝐭z\in\mathcal{H}_{1}^{\mathbf{t}} and ww\in\mathcal{H}. We can write σ𝐭=Im,1/2\sigma_{\mathbf{t}}=\operatorname{Im}\langle\cdot,\cdot\rangle_{1/2}, where ,1/2\langle\cdot,\cdot\rangle_{1/2} is the canonical inner product of 1/2𝐭\mathcal{H}_{1/2}^{\mathbf{t}}:

z,w1/2:=n=1znwn¯tn.\displaystyle\langle z,w\rangle_{1/2}:=\sum_{n=1}^{\infty}\frac{z_{n}\cdot\overline{w_{n}}}{t_{n}}.

Moreover, 1/2\|\cdot\|_{1/2} denotes the corresponding norm: z1/22=z,z1/2\|z\|_{1/2}^{2}=\langle z,z\rangle_{1/2}.

We recall some basic facts about Gaussian measures on infinite dimensional Hilbert spaces, cf. [18, 9, 2]. The infinite product measure μ𝐭:=k=1μtk\mu_{\bf t}:=\prod_{k=1}^{\infty}\mu_{t_{k}} of the Gaussian measures

dμtk(z):=1πtke|z|2tkdV(z),zd\mu_{t_{k}}(z):=\frac{1}{\pi t_{k}}e^{-\frac{|z|^{2}}{t_{k}}}dV(z),\hskip 17.22217ptz\in\mathbb{C}

gives a well-defined probability measure on =k=1\mathbb{C}^{\infty}=\prod_{k=1}^{\infty}\mathbb{C}. It is concentrated on =2()\mathcal{H}=\ell^{2}(\mathbb{N})\subset\mathbb{C}^{\infty}. By μ𝐭\mu_{\mathbf{t}}, we also denote its restriction to the measurable space (,)(\mathcal{H},\mathcal{B}), where \mathcal{B} is the Borel σ\sigma-algebra of \mathcal{H}. Note that \mathcal{B} agrees with the σ\sigma-algebra generated by cylindrical Borel sets. The measure μ𝐭\mu_{\mathbf{t}} is called the (centered) Gaussian measure on \mathcal{H} with covariance operator BB: This is due to the fact that BB is naturally related to μ𝐭\mu_{\mathbf{t}} by

eiRex,y2𝑑μ(x)=e12By,y2,(y).\displaystyle\int_{\mathcal{H}}e^{i\operatorname{Re}\langle x,y\rangle_{\ell^{2}}}~{}d\mu(x)=e^{-\frac{1}{2}\langle By,y\rangle_{\ell^{2}}},\hskip 17.22217pt(y\in\mathcal{H}). (5.1)

The space 1/2𝐭\mathcal{H}_{1/2}^{\mathbf{t}} introduced above is also known as the Cameron-Martin space of μ𝐭\mu_{\mathbf{t}}. It is a measurable set of measure zero: μ𝐭(1/2𝐭)=0\mu_{\mathbf{t}}(\mathcal{H}_{1/2}^{\mathbf{t}})=0. Hence, so is 1\mathcal{H}_{1}.

We will now introduce the Fock-Bargmann space of holomorphic functions in infinitely many variables and Berezin-Toeplitz quantization in this setting. Two different approaches have been proposed in [15, 16]: the first being of measure theoretic nature and the second based on an inductive limit construction. Both approaches yield different theories and we deal with the measure theoretic approach here.

We will write 𝒵0\mathcal{Z}_{0} for the set of all sequences (αn)n=1(\alpha_{n})_{n=1}^{\infty} with values in 0\mathbb{N}_{0} such that all but finitely many entries are zero. In the following we use the standard multi-index notation: let z=(z1,z2,)2()z=(z_{1},z_{2},\ldots)\in\ell^{2}(\mathbb{N}) and α𝒵0\alpha\in\mathcal{Z}_{0}, then we put:

zα=z1α1z2α2,α!=α1!α2!,𝐭α=t1α1t2α2.\displaystyle z^{\alpha}=z_{1}^{\alpha_{1}}z_{2}^{\alpha_{2}}\dots,\quad\alpha!=\alpha_{1}!\alpha_{2}!\dots,\quad\mathbf{t}^{\alpha}=t_{1}^{\alpha_{1}}t_{2}^{\alpha_{2}}\dots.

The monomials \mathcal{M} below form an orthonormal system of complex analytic functions on \mathcal{H} inside L2(,μ𝐭)L^{2}(\mathcal{H},\mu_{\bf t}):

:=[eα(z):=1𝐭αα!zα:α𝒵0].\mathcal{M}:=\Big{[}e_{\alpha}(z):=\frac{1}{\sqrt{\mathbf{t}^{\alpha}\alpha!}}z^{\alpha}:~{}\alpha\in\mathcal{Z}_{0}\Big{]}.

We write ,\langle\cdot,\cdot\rangle and \|\cdot\| for the standard inner product and norm in L2(,μ𝐭)L^{2}(\mathcal{H},\mu_{\bf t}), respectively. Elements in the linear span of \mathcal{M} are referred to as analytic polynomials. In analogy to the finite dimensional setting of n\mathbb{C}^{n} we define the Bargmann-Fock space F𝐭2()F_{\mathbf{t}}^{2}(\mathcal{H}) to be the L2L^{2}-closure of analytic polynomials:

F𝐭2():=span¯.F_{\mathbf{t}}^{2}(\mathcal{H}):=\overline{\operatorname{span}}\>\mathcal{M}.

Let z1z\in\mathcal{H}_{1} and pspanp\in\textup{span}\>\mathcal{M} be an analytic polynomial. Extending (3.3) we define the Weyl operator Wz𝐭W_{z}^{\mathbf{t}} on F𝐭2()F_{\mathbf{t}}^{2}(\mathcal{H}) by:

Wz𝐭p(w):=kz𝐭(w)p(wz),w.\displaystyle W_{z}^{\mathbf{t}}p(w):=k_{z}^{\mathbf{t}}(w)p(w-z),\hskip 17.22217ptw\in\mathcal{H}.

Here, kz𝐭k_{z}^{\mathbf{t}} denotes the normalized reproducing kernel defined by:

kz𝐭(w)=ew,z1/212z1/22.\displaystyle k_{z}^{\mathbf{t}}(w)=e^{\langle w,z\rangle_{1/2}-\frac{1}{2}\|z\|_{1/2}^{2}}.

Let NN\in\mathbb{N} and assume that the analytic polynomial pp only depends on the variables z1,,zNz_{1},\ldots,z_{N}. We write μ𝐭=μ𝐭μ𝐭N+1\mu_{\mathbf{t}}=\mu_{\mathbf{t}^{\prime}}\otimes\mu_{\mathbf{t}_{N+1}}, where 𝐭:=(t1,,tN)\mathbf{t}^{\prime}:=(t_{1},\dots,t_{N}) and 𝐭N+1=(tN+1,tN+2,)\mathbf{t}_{N+1}=(t_{N+1},t_{N+2},\dots). Define the projections πN(z):=(z1,,zN)\pi_{N}(z):=(z_{1},\ldots,z_{N}) for a sequence z=(zj)jz=(z_{j})_{j}\in\mathcal{H}. Since p=pπNp=p\circ\pi_{N} we obtain:

|Wz𝐭p(w)|2𝑑μ𝐭(w)=\displaystyle\int_{\mathcal{H}}|W_{z}^{\bf t}p(w)|^{2}d\mu_{\mathbf{t}}(w)=
=\displaystyle= |kπN(z)𝐭πN(w)pπN(wz)|2|k(IπN)(z)𝐭N+1(IπN)(w)|2𝑑μ𝐭(w)\displaystyle\int_{\mathbb{C}^{\infty}}|k_{\pi_{N}(z)}^{\mathbf{t}^{\prime}}\circ\pi_{N}(w)\>p\circ\pi_{N}(w-z)|^{2}|k_{(I-\pi_{N})(z)}^{\mathbf{t}_{N+1}}\circ(I-\pi_{N})(w)|^{2}~{}d\mu_{\mathbf{t}}(w)
=\displaystyle= N|WπN(z)𝐭p(w)|2𝑑μ𝐭(w)×k=N+1|k(IπN)(z)𝐭N+1(w)|2𝑑μ𝐭N+1(w)=1\displaystyle\int_{\mathbb{C}^{N}}|W_{\pi_{N}(z)}^{\mathbf{t}^{\prime}}p(w)|^{2}~{}d\mu_{\mathbf{t}^{\prime}}(w)\times\underbrace{\int_{\prod_{k=N+1}^{\infty}\mathbb{C}}|k_{(I-\pi_{N})(z)}^{\mathbf{t}_{N+1}}(w)|^{2}~{}d\mu_{\mathbf{t}_{N+1}}(w)}_{=1}
=\displaystyle= |p(w)|2𝑑μ𝐭(w).\displaystyle\int_{\mathcal{H}}|p(w)|^{2}~{}d\mu_{\mathbf{t}}(w).

Therefore, Wz𝐭pL2(,μ𝐭)W_{z}^{\mathbf{t}}p\in L^{2}(\mathcal{H},\mu_{\mathbf{t}}) and Wz𝐭W_{z}^{\mathbf{t}} acts isometrically (say, on analytic polynomials). In finite dimensions, it would now be clear that WztpF𝐭2W_{z}^{t}p\in F_{\mathbf{t}}^{2}, since WztpW_{z}^{t}p defines a holomorphic function. In infinite dimensions, this statement is not entirely trivial, as F𝐭2F_{\mathbf{t}}^{2} is defined as the closure of the analytic polynomials. Nevertheless, WztpF𝐭2W_{z}^{t}p\in F_{\mathbf{t}}^{2} remains true, cf. [15]. Hence, Wz𝐭W_{z}^{\mathbf{t}} extends to an isometric operator on F𝐭2()F_{\mathbf{t}}^{2}(\mathcal{H}) for every z1z\in\mathcal{H}_{1}.

Further, note that for (zk)1𝐭(z_{k})\subset\mathcal{H}_{1}^{\mathbf{t}}, zkzz_{k}\to z in 1𝐭\mathcal{H}_{1}^{\mathbf{t}} we clearly have

kzk𝐭(w)p(wzk)kz𝐭(w)p(wz)\displaystyle k_{z_{k}}^{\mathbf{t}}(w)p(w-z_{k})\to k_{z}^{\mathbf{t}}(w)p(w-z)

pointwise almost everywhere on 2()\ell^{2}(\mathbb{N}). Hence, Scheffé’s Lemma shows that the assignment zWz𝐭pz\mapsto W_{z}^{\mathbf{t}}p is continuous from 1\mathcal{H}_{1} to F𝐭2()F_{\mathbf{t}}^{2}(\mathcal{H}) for every analytic polynomial pp. Since these are dense in F𝐭2()F_{\mathbf{t}}^{2}(\mathcal{H}), the Weyl operators Wz𝐭W_{z}^{\mathbf{t}} extend to isometric operators on F𝐭2()F_{\mathbf{t}}^{2}(\mathcal{H}) such zWz𝐭z\mapsto W_{z}^{\mathbf{t}} is continuous on 1\mathcal{H}_{1} with respect to the strong operator topology.

An even stronger statement holds, namely the map zWz𝐭z\mapsto W_{z}^{\mathbf{t}} is even strongly continuous with respect to the coarser topology of 1/2\mathcal{H}_{1/2}, cf. [15]. Therefore, it extends to a map from 1/2\mathcal{H}_{1/2} to the unitary operators on F𝐭2()F_{\mathbf{t}}^{2}(\mathcal{H}).

By applying the Weyl operators with z,w1z,w\in\mathcal{H}_{1} to analytic polynomials, it is not hard to see that

Wz𝐭Ww𝐭=eiσ𝐭(z,w)Wz+w𝐭and (Wz𝐭)=(Wz𝐭)1=Wz𝐭.W_{z}^{\mathbf{t}}W_{w}^{\mathbf{t}}={e^{-i\sigma_{\mathbf{t}}(z,w)}}W_{z+w}^{\mathbf{t}}\hskip 12.91663pt\mbox{\it and }\hskip 12.91663pt(W_{z}^{\mathbf{t}})^{\ast}=(W_{z}^{\mathbf{t}})^{-1}=W_{-z}^{\mathbf{t}}.

These relations extend to z,w1/2z,w\in\mathcal{H}_{1/2}, as well:

Lemma 5.1 ([15, 16]).

The Weyl operators Wz𝐭(F𝐭2())W_{z}^{\mathbf{t}}\in\mathcal{L}(F_{\mathbf{t}}^{2}(\mathcal{H})) depend continuously on z(1/2,1/2)z\in(\mathcal{H}_{1/2},\|\cdot\|_{1/2}) with respect to the strong operator topology and

Wz𝐭Ww𝐭=eiσ𝐭(z,w)Wz+w𝐭\displaystyle W_{z}^{\mathbf{t}}W_{w}^{\mathbf{t}}={e^{-i\sigma_{\mathbf{t}}(z,w)}}W_{z+w}^{\mathbf{t}}

for every z,w1/2z,w\in\mathcal{H}_{1/2}.

Fix a bounded operator A(F𝐭2())A\in\mathcal{L}(F_{\mathbf{t}}^{2}(\mathcal{H})) and z1z\in\mathcal{H}_{1}. We define the Berezin transform of AA at zz by

A~(z)=Akz𝐭,kz𝐭=Wz𝐭AWz𝐭1,1.\displaystyle\widetilde{A}(z)=\big{\langle}Ak_{z}^{\mathbf{t}},k_{z}^{\mathbf{t}}\big{\rangle}=\big{\langle}W_{-z}^{\mathbf{t}}AW_{z}^{\mathbf{t}}1,1\big{\rangle}.
Lemma 5.2.

The map AA~A\mapsto\widetilde{A} is injective on (F𝐭2())\mathcal{L}(F_{\mathbf{t}}^{2}(\mathcal{H})).

Proof.

If A~=0\widetilde{A}=0, then A~(z)=0\widetilde{A}(z)=0 for every z=(z1,,zN,0,0,)z=(z_{1},\dots,z_{N},0,0,\dots). We denote by FN2F_{N}^{2} the closed linear span in F𝐭2()F_{\mathbf{t}}^{2}(\mathcal{H}) of the set of all analytic polynomials in the finitely many variables z1,,zNz_{1},\dots,z_{N}. Moreover, we write PN(F𝐭2())P_{N}\in\mathcal{L}(F_{\mathbf{t}}^{2}(\mathcal{H})) for the orthogonal projection onto FN2F_{N}^{2}. Then, kz𝐭FN2k_{z}^{\mathbf{t}}\in F_{N}^{2} for z=(z1,,zN,0,0,)z=(z_{1},\dots,z_{N},0,0,\dots) and hence:

A~(z)=Akz𝐭,kz𝐭=PNAkz𝐭,kz𝐭.\displaystyle\widetilde{A}(z)=\big{\langle}Ak_{z}^{\mathbf{t}},k_{z}^{\mathbf{t}}\big{\rangle}=\big{\langle}P_{N}Ak_{z}^{\mathbf{t}},k_{z}^{\mathbf{t}}\big{\rangle}.

Note that FN2F_{N}^{2} can be naturally identified with the NN-variable Fock-Bargmann space F𝐭2=F𝐭2(N)F_{\mathbf{t}^{\prime}}^{2}=F_{\mathbf{t}^{\prime}}^{2}(\mathbb{C}^{N}) with weight parameter 𝐭=(t1,,tN)\mathbf{t}^{\prime}=(t_{1},\dots,t_{N}). Moreover,

PNA|FN2(FN2)(F𝐭2)and 0=A~(z)=(PNA|FN2)(z1,,zN).P_{N}A|_{F_{N}^{2}}\in\mathcal{L}(F_{N}^{2})\cong\mathcal{L}(F_{\mathbf{t}^{\prime}}^{2})\hskip 12.91663pt\mbox{\it and }\hskip 12.91663pt0=\widetilde{A}(z)=(P_{N}A|_{F_{N}^{2}})^{\sim}(z_{1},\dots,z_{N}).

On the right hand side of the last formula we have use the Berezin transform on F𝐭2F_{\mathbf{t}^{\prime}}^{2}. Since the Berezin transform is injective on (F𝐭2)\mathcal{L}(F_{\mathbf{t}^{\prime}}^{2}) (see [12]), we conclude that PNA|FN2=0P_{N}A|_{F_{N}^{2}}=0. Hence, PNAp=0P_{N}Ap=0 for any analytic polynomial pp and NN\in\mathbb{N} sufficiently large. Finally, we have

PNApAp,N,\displaystyle P_{N}Ap\to Ap,\quad N\to\infty,

in F𝐭2()F_{\mathbf{t}}^{2}(\mathcal{H}), i.e. Ap=0Ap=0 for analytic polynomials pp showing that A=0A=0. ∎

We denote by P𝐭P^{\mathbf{t}} the orthogonal projection from L2(,μ𝐭)L^{2}(\mathcal{H},\mu_{\mathbf{t}}) onto F𝐭2()F_{\mathbf{t}}^{2}(\mathcal{H}). Given φL(,μ𝐭)\varphi\in L^{\infty}(\mathcal{H},\mu_{\mathbf{t}}) we define the Toeplitz operator Tφ𝐭(F𝐭2())T_{\varphi}^{\mathbf{t}}\in\mathcal{L}(F_{\mathbf{t}}^{2}(\mathcal{H})) by:

Tφ𝐭(g)=P𝐭(φg).\displaystyle T_{\varphi}^{\mathbf{t}}(g)=P^{\mathbf{t}}(\varphi g).

By applying the injectivity of the Berezin transform in Lemma 5.2, it is not hard to verify that the Weyl operators Wz𝐭W_{z}^{\mathbf{t}} for z1z\in\mathcal{H}_{1} have a representation as Toeplitz operators. More precisely:

Wz𝐭=Tgz𝐭wheregz(w)=e2iσ𝐭(w,z)+12z1/22.W_{z}^{\mathbf{t}}=T_{g_{z}}^{\mathbf{t}}\hskip 12.91663pt\mbox{\it where}\hskip 12.91663ptg_{z}(w)=e^{2i\sigma_{\mathbf{t}}(w,z)+\frac{1}{2}\|z\|_{1/2}^{2}}. (5.2)

Due to the importance of this fact, we derive an explicit formula for the Berezin transform of both operators:

Wz𝐭~(w)\displaystyle\widetilde{W_{z}^{\mathbf{t}}}(w) =ew1/22Wz𝐭Kw𝐭,Kw𝐭\displaystyle=e^{-\|w\|_{1/2}^{2}}\langle W_{z}^{\mathbf{t}}K_{w}^{\mathbf{t}},K_{w}^{\mathbf{t}}\rangle
=ew1/2212z1/22e,z1/2ez,w1/2,e,w1/2\displaystyle=e^{-\|w\|_{1/2}^{2}-\frac{1}{2}\|z\|_{1/2}^{2}}\langle e^{\langle\cdot,z\rangle_{1/2}}e^{\langle\cdot-z,w\rangle_{1/2}},e^{\langle\cdot,w\rangle_{1/2}}\rangle
=ew1/2212z1/22ez,w1/2Kw+z𝐭,Kw𝐭\displaystyle=e^{-\|w\|_{1/2}^{2}-\frac{1}{2}\|z\|_{1/2}^{2}}e^{-\langle z,w\rangle_{1/2}}\langle K_{w+z}^{\mathbf{t}},K_{w}^{\mathbf{t}}\rangle
=ew1/2212z1/22ez,w1/2+w,w+z1/2,\displaystyle=e^{-\|w\|_{1/2}^{2}-\frac{1}{2}\|z\|_{1/2}^{2}}e^{-\langle z,w\rangle_{1/2}+\langle w,w+z\rangle_{1/2}},
gz~(𝐭)(w)\displaystyle\widetilde{g_{z}}^{(\mathbf{t})}(w) =e12z1/22w1/22e2iσ𝐭(,z)Kw𝐭,Kw𝐭\displaystyle=e^{\frac{1}{2}\|z\|_{1/2}^{2}-\|w\|_{1/2}^{2}}\langle e^{2i\sigma_{\mathbf{t}}(\cdot,z)}K_{w}^{\mathbf{t}},K_{w}^{\mathbf{t}}\rangle
=e12z1/22w1/22e,z1/2Kw𝐭,e,z1/2Kw𝐭\displaystyle=e^{\frac{1}{2}\|z\|_{1/2}^{2}-\|w\|_{1/2}^{2}}\langle e^{\langle\cdot,z\rangle_{1/2}}K_{w}^{\mathbf{t}},e^{-\langle\cdot,z\rangle_{1/2}}K_{w}^{\mathbf{t}}\rangle
=e12z1/22w1/22Kw+z𝐭,Kwz𝐭\displaystyle=e^{\frac{1}{2}\|z\|_{1/2}^{2}-\|w\|_{1/2}^{2}}\langle K_{w+z}^{\mathbf{t}},K_{w-z}^{\mathbf{t}}\rangle
=e12z1/22w1/22ewz,w+z1/2=Wz𝐭~(w).\displaystyle=e^{\frac{1}{2}\|z\|_{1/2}^{2}-\|w\|_{1/2}^{2}}e^{\langle w-z,w+z\rangle_{1/2}}=\widetilde{W_{z}^{\mathbf{t}}}(w).

It is an important but non-trivial fact that (5.2)(\ref{Weyl_equals_Toeplitz}) extends to z1/2z\in\mathcal{H}_{1/2}. First, we have to explain in what sense the map ww,z1/2w\mapsto\langle w,z\rangle_{1/2} defines a measurable function on \mathcal{H} for z1/2z\in\mathcal{H}_{1/2}. Clearly, the expression ,z\langle\cdot,z\rangle is pointwise well-defined on 1/2\mathcal{H}_{1/2}. However, since 1/2\mathcal{H}_{1/2} is a set of measure zero, this is of no big help. We need a preliminary result:

Lemma 5.3.

,z122=2z1/22\|\langle\cdot,z\rangle_{\frac{1}{2}}\|^{2}=2\|z\|_{1/2}^{2} for z1z\in\mathcal{H}_{1}.

Proof.

We introduce a complex parameter λ\lambda\in\mathbb{C}. Then, it is not hard to verify the following equalities through standard results on differentiation of parameter integrals, where we use that ,zL2(,μ𝐭)\langle\cdot,z\rangle\in L^{2}(\mathcal{H},\mu_{\mathbf{t}}) (this is easily established):

2λ¯λei2Rew,λz1/2𝑑μ𝐭(w)=2λ¯λeiλ¯w,z1/2+iλz,w1/2𝑑μ𝐭(w)=|z,w1/2|2eiλ¯w,z1/2+iλz,w1/2𝑑μ𝐭(w).\frac{\partial^{2}}{\partial\overline{\lambda}\partial\lambda}\int_{\mathcal{H}}e^{i2\operatorname{Re}\langle w,\lambda z\rangle_{1/2}}~{}d\mu_{\mathbf{t}}(w)=\frac{\partial^{2}}{\partial\overline{\lambda}\partial\lambda}\int_{\mathcal{H}}e^{i\overline{\lambda}\langle w,z\rangle_{1/2}+i\lambda\langle z,w\rangle_{1/2}}~{}d\mu_{\mathbf{t}}(w)\\ =-\int_{\mathcal{H}}|\langle z,w\rangle_{1/2}|^{2}e^{i\overline{\lambda}\langle w,z\rangle_{1/2}+i\lambda\langle z,w\rangle_{1/2}}~{}d\mu_{\mathbf{t}}(w).

On the other hand, using Equality (5.1):

2λ¯λei2Rew,λz1/2𝑑μ𝐭(w)\displaystyle\frac{\partial^{2}}{\partial\overline{\lambda}\partial\lambda}\int_{\mathcal{H}}e^{i2\operatorname{Re}\langle w,\lambda z\rangle_{1/2}}~{}d\mu_{\mathbf{t}}(w) =2λ¯λeiRew,2λB1z2𝑑μ𝐭(w)\displaystyle=\frac{\partial^{2}}{\partial\overline{\lambda}\partial\lambda}\int_{\mathcal{H}}e^{i\operatorname{Re}\langle w,2\lambda B^{-1}z\rangle_{\ell^{2}}}~{}d\mu_{\mathbf{t}}(w)
=2λ¯λe122λBB1z,2λB1z2\displaystyle=\frac{\partial^{2}}{\partial\overline{\lambda}\partial\lambda}e^{-\frac{1}{2}\langle 2\lambda BB^{-1}z,2\lambda B^{-1}z\rangle_{\ell^{2}}}
=2λ¯λe2λλ¯z,B1z2\displaystyle=\frac{\partial^{2}}{\partial\overline{\lambda}\partial\lambda}e^{-2\lambda\overline{\lambda}\langle z,B^{-1}z\rangle_{\ell^{2}}}
=2λ¯λe2λλ¯z,z1/2\displaystyle=\frac{\partial^{2}}{\partial\overline{\lambda}\partial\lambda}e^{-2\lambda\overline{\lambda}\langle z,z\rangle_{1/2}}
=(2z1/22+4|λ|2z1/24)e2λλ¯z1/22.\displaystyle=\Big{(}-2\|z\|_{1/2}^{2}+4|\lambda|^{2}\|z\|_{1/2}^{4}\Big{)}e^{-2\lambda\overline{\lambda}\|z\|_{1/2}^{2}}.

We therefore obtain:

|z,w1/2|2ei2Rew,λz1/2𝑑μ𝐭(w)=(24|λ|2z1/22)z1/22e2|λ|2z1/22.\int_{\mathcal{H}}|\langle z,w\rangle_{1/2}|^{2}e^{i2\operatorname{Re}\langle w,\lambda z\rangle_{1/2}}~{}d\mu_{\mathbf{t}}(w)=\big{(}2-4|\lambda|^{2}\|z\|_{1/2}^{2}\big{)}\|z\|_{1/2}^{2}e^{-2|\lambda|^{2}\|z\|_{1/2}^{2}}.

Choosing λ=0\lambda=0 yields the desired equality. ∎

Hence, if (zk)1(z_{k})\subset\mathcal{H}_{1} is a Cauchy sequence with respect to 1/2\|\cdot\|_{1/2}, then ,zk1/2\langle\cdot,z_{k}\rangle_{1/2} is Cauchy in L2(,μ𝐭)L^{2}(\mathcal{H},\mu_{\mathbf{t}}). In conclusion, the family of functions ,z1/2L2(,μ𝐭)\langle\cdot,z\rangle_{1/2}\in L^{2}(\mathcal{H},\mu_{\mathbf{t}}) continuously extends to z1/2z\in\mathcal{H}_{1/2}. In particular, for each z1/2z\in\mathcal{H}_{1/2}, w,z1/2\langle w,z\rangle_{1/2} is a well-defined expression for almost every ww\in\mathcal{H} and, as a function of ww, measurable. In particular,

gz(w)=e2iσ𝐭(w,z)+12z1/22=ew,z1/2ez,w1/2e12z1/22\displaystyle g_{z}(w)=e^{2i\sigma_{\mathbf{t}}(w,z)+\frac{1}{2}\|z\|_{1/2}^{2}}=e^{\langle w,z\rangle_{1/2}}e^{-\langle z,w\rangle_{1/2}}e^{\frac{1}{2}\|z\|_{1/2}^{2}}

is an almost everywhere well-defined function with

gze12z1/22,z1/2.\|g_{z}\|_{\infty}\leq e^{\frac{1}{2}\|z\|_{1/2}^{2}},\hskip 17.22217ptz\in\mathcal{H}_{1/2}.

We also note that, since ,z1/2F𝐭2()\langle\cdot,z\rangle_{1/2}\in F_{\mathbf{t}}^{2}(\mathcal{H}) for every z1z\in\mathcal{H}_{1}, the same is true for z1/2z\in\mathcal{H}_{1/2}. Iterating the procedure from the proof of Lemma 5.3, one easily obtains:

,z1/2k2=2kz1/22kfor k.\displaystyle\|\langle\cdot,z\rangle_{1/2}^{k}\|^{2}=2^{k}\|z\|_{1/2}^{2k}\hskip 12.91663pt\mbox{\it for }\hskip 12.91663ptk\in\mathbb{N}.

Next, we calculate:

k=0,z1/2kk!=k=02k2z1/2kk!=e2z1/2.\displaystyle\sum_{k=0}^{\infty}\frac{\|\langle\cdot,z\rangle_{1/2}^{k}\|}{k!}=\sum_{k=0}^{\infty}\frac{2^{\frac{k}{2}}\|z\|_{1/2}^{k}}{k!}=e^{\sqrt{2}\|z\|_{1/2}}.

We obtain:

k=0,z1/2kk!=e,z1/2=Kz𝐭F𝐭2()\displaystyle\sum_{k=0}^{\infty}\frac{\langle\cdot,z\rangle_{1/2}^{k}}{k!}=e^{\langle\cdot,z\rangle_{1/2}}=K_{z}^{\mathbf{t}}\in F_{\mathbf{t}}^{2}(\mathcal{H})

for every z1/2z\in\mathcal{H}_{1/2}. By a standard density argument, we have:

Kz,Kw=ew,z1/2forz,w1/2.\displaystyle\langle K_{z},K_{w}\rangle=e^{\langle w,z\rangle_{1/2}}\hskip 12.91663pt\mbox{\it for}\hskip 12.91663ptz,w\in\mathcal{H}_{1/2}.

Now, one can compute the Berezin transform of gzg_{z} as above. For every z1/2z\in\mathcal{H}_{1/2} and w1w\in\mathcal{H}_{1} one obtains:

Tgz𝐭~(w)=e12z1/22w1/22ewz,w+z1/2.\displaystyle\widetilde{T_{g_{z}}^{\mathbf{t}}}(w)=e^{\frac{1}{2}\|z\|_{1/2}^{2}-\|w\|_{1/2}^{2}}e^{\langle w-z,w+z\rangle_{1/2}}.

Further, since Wz𝐭W_{z}^{\mathbf{t}} continuously (in strong operator topology) depends on the parameter z1/2z\in\mathcal{H}_{1/2}, we conclude that the Berezin transform Wz𝐭~\widetilde{W_{z}^{\mathbf{t}}} continuously (in the topology of pointwise convergence) depends on z1/2z\in\mathcal{H}_{1/2}, which gives

Wz𝐭~(w)=e12z1/22w1/22ewz,w+z1/2\displaystyle\widetilde{W_{z}^{\mathbf{t}}}(w)=e^{\frac{1}{2}\|z\|_{1/2}^{2}-\|w\|_{1/2}^{2}}e^{\langle w-z,w+z\rangle_{1/2}}

for every z1/2z\in\mathcal{H}_{1/2} and w1w\in\mathcal{H}_{1}. Comparing Berezin transforms, we have shown that Tgz𝐭=Wz𝐭T_{g_{z}}^{\mathbf{t}}=W_{z}^{\mathbf{t}} extends to z1/2z\in\mathcal{H}_{1/2}.

In an abuse of notation, we will write (1/2,σ𝐭)\mathcal{R}(\mathcal{H}_{1/2},\sigma_{\mathbf{t}}) for the representation of the resolvent algebra on F𝐭2()F_{\mathbf{t}}^{2}(\mathcal{H}) with respect to the symplectic space (1/2,σ𝐭)(\mathcal{H}_{1/2},\sigma_{\bf t}). Since the Weyl operators Wz𝐭W_{z}^{\mathbf{t}}, z1/2z\in\mathcal{H}_{1/2} satisfy the CCR in Lemma 5.1, we start again by expressing the resolvent R(λ,z)R(\lambda,z) in form of a Laplace transform of the one-parameter group (Wtz)t(W_{tz})_{t\in\mathbb{R}}. For Re(λ)>0\operatorname{Re}(\lambda)>0 we have:

R(λ,z)=i0eλsWsz𝐭𝑑s,\displaystyle R(\lambda,z)=i\int_{0}^{\infty}e^{-\lambda s}W_{-sz}^{\mathbf{t}}~{}ds,

which exists as an integral in strong operator topology. There is an analogous formula in the case Re(λ)<0\operatorname{Re}(\lambda)<0. Consider the space of (classical) symbols:

cl𝐭=C({(λ2iσ𝐭(,z))1:λi,z1/2}).\displaystyle\mathcal{R}_{cl}^{\mathbf{t}}=C^{\ast}\Big{(}\{(\lambda-2i\sigma_{\mathbf{t}}(\cdot,z))^{-1}:\lambda\in\mathbb{C}\setminus i\mathbb{R},~{}z\in\mathcal{H}_{1/2}\}\Big{)}.

Moreover, we write 𝒯𝐭(cl𝐭)\mathcal{T}^{\mathbf{t}}(\mathcal{R}_{cl}^{\mathbf{t}}) for the C algebra generated by Toeplitz operators over F𝐭2()F_{\mathbf{t}}^{2}(\mathcal{H}) with symbols in cl𝐭\mathcal{R}_{cl}^{\mathbf{t}}.

Theorem 5.4.

The following inclusion holds true: (1/2,σ𝐭)𝒯𝐭(cl𝐭)\mathcal{R}(\mathcal{H}_{1/2},\sigma_{\mathbf{t}})\subset\mathcal{T}^{\mathbf{t}}(\mathcal{R}_{cl}^{\mathbf{t}}).

Proof.

Without loss of generality we assume Re(λ)>0\operatorname{Re}(\lambda)>0. We will verify that the representation of the resolvent in form of a Laplace transform of Weyl operators defines an element in the Toeplitz algebra. The following integrals are to be understood as improper Riemann integrals in strong operator topology:

0eλsWsz𝐭𝑑s\displaystyle\int_{0}^{\infty}e^{-\lambda s}W_{-sz}^{\mathbf{t}}~{}ds =0eλsTgsz𝐭𝑑s\displaystyle=\int_{0}^{\infty}e^{-\lambda s}T_{g_{-sz}}^{\mathbf{t}}~{}ds
=0eλses22z1/22Texp(2isσ𝐭(,z))𝐭𝑑s\displaystyle=\int_{0}^{\infty}e^{-\lambda s}e^{\frac{s^{2}}{2}\|z\|_{1/2}^{2}}T_{\exp(-2is\sigma_{\mathbf{t}}(\cdot,z))}^{\mathbf{t}}~{}ds
=0eλsk=0s2kz1/22k2kk!Texp(2isσ𝐭(,z))𝐭ds.\displaystyle=\int_{0}^{\infty}e^{-\lambda s}\sum_{k=0}^{\infty}\frac{s^{2k}\|z\|_{1/2}^{2k}}{2^{k}k!}T_{\exp(-2is\sigma_{\mathbf{t}}(\cdot,z))}^{\mathbf{t}}~{}ds.

Since the Weyl operators are unitary, and hence satisfy Wz𝐭=1\|W_{z}^{\mathbf{t}}\|=1, it follows that:

Texp(2isσ𝐭(,z))𝐭=es22z1/22.\displaystyle\|T_{\exp(-2is\sigma_{\mathbf{t}}(\cdot,z))}^{\mathbf{t}}\|=e^{-\frac{s^{2}}{2}\|z\|_{1/2}^{2}}.

Therefore, the dominated convergence theorem gives as mm\rightarrow\infty:

iR(λ,z)k=0mz1/22k2kk!0s2keλsTexp(2isσ𝐭(,z))𝐭𝑑s0eλs|es22z1/22k=0ms2kz1/22kk!2k|es22z1/22𝑑s0,m.\Big{\|}iR(\lambda,z)-\sum_{k=0}^{m}\frac{\|z\|_{1/2}^{2k}}{2^{k}k!}\int_{0}^{\infty}s^{2k}e^{-\lambda s}T_{\exp(-2is\sigma_{\mathbf{t}}(\cdot,z))}^{\mathbf{t}}~{}ds\Big{\|}\\ \leq\int_{0}^{\infty}e^{-\lambda s}\left|e^{\frac{s^{2}}{2}\|z\|_{1/2}^{2}}-\sum_{k=0}^{m}\frac{s^{2k}\|z\|_{1/2}^{2k}}{k!2^{k}}\right|e^{-\frac{s^{2}}{2}\|z\|_{1/2}^{2}}~{}ds\to 0,\quad m\to\infty.

We have therefore seen that

iR(λ,z)=k=01k!z1/22k2k0Ts2kexp(λs+2isσ𝐭(,z))𝐭𝑑s,\displaystyle iR(\lambda,z)=\sum_{k=0}^{\infty}\frac{1}{k!}\frac{\|z\|_{1/2}^{2k}}{2^{k}}\int_{0}^{\infty}T_{s^{2k}\exp(-\lambda s+2is\sigma_{\mathbf{t}}(\cdot,z))}^{\mathbf{t}}~{}ds,

where the series converges in operator norm. Fix for the moment N>0N>0. Since the Riemann integral 0NTs2kexp(λs+2isσ𝐭(,z))𝐭𝑑s\int_{0}^{N}T_{s^{2k}\exp(-\lambda s+2is\sigma_{\mathbf{t}}(\cdot,z))}^{\mathbf{t}}~{}ds exists as a limit of Riemann sums in strong operator topology (this easily follows from the fact that the mapping sWsz𝐭s\mapsto W_{sz}^{\mathbf{t}} is strongly continuous), we obtain for the Berezin transform:

(0NTs2kexp(λs+2isσ𝐭(,z))𝐭𝑑s)(u)=T0Ns2kexp(λs+2isσ𝐭(,z))𝑑s𝐭ku𝐭,ku𝐭.\left(\int_{0}^{N}T_{s^{2k}\exp(-\lambda s+2is\sigma_{\mathbf{t}}(\cdot,z))}^{\mathbf{t}}~{}ds\right)^{\sim}(u)=\Big{\langle}T_{\int_{0}^{N}s^{2k}\exp(-\lambda s+2is\sigma_{\mathbf{t}}(\cdot,z))ds}^{\mathbf{t}}~{}k_{u}^{\mathbf{t}},k_{u}^{\mathbf{t}}\Big{\rangle}.

Therefore, injectivity of the Berezin transform (Lemma 5.2) shows:

0NTs2kexp(λs+2isσ𝐭(,z))𝐭𝑑s=T0Ns2kexp(λs+2isσ𝐭(,z))𝑑s𝐭.\displaystyle\int_{0}^{N}T_{s^{2k}\exp(-\lambda s+2is\sigma_{\mathbf{t}}(\cdot,z))}^{\mathbf{t}}ds=T_{\int_{0}^{N}s^{2k}\exp(-\lambda s+2is\sigma_{\mathbf{t}}(\cdot,z))ds}^{\mathbf{t}}.

Since

0Ns2kexp(λs+2isσ𝐭(,z))𝑑sN0s2kexp(λs+2isσ𝐭(,z))𝑑s,\displaystyle\int_{0}^{N}s^{2k}\exp(-\lambda s+2is\sigma_{\mathbf{t}}(\cdot,z))ds\overset{N\rightarrow\infty}{\longrightarrow}\int_{0}^{\infty}s^{2k}\exp(-\lambda s+2is\sigma_{\mathbf{t}}(\cdot,z))ds,

uniformly, this gives

0Ts2kexp(λs+2isσ𝐭(,z))𝐭𝑑s=T0s2kexp(λs+2isσ𝐭(,z))𝑑s𝐭.\displaystyle\int_{0}^{\infty}T_{s^{2k}\exp(-\lambda s+2is\sigma_{\mathbf{t}}(\cdot,z))}^{\mathbf{t}}~{}ds=T_{\int_{0}^{\infty}{s^{2k}\exp(-\lambda s+2is\sigma_{\mathbf{t}}(\cdot,z))}~{}ds}^{\mathbf{t}}.

For the symbol, standard facts on the Laplace transform yield

0s2keλs+2isσ𝐭(w,z)𝑑s=(2k)!(λ2iσ𝐭(w,z))(2k+1).\displaystyle\int_{0}^{\infty}{s^{2k}e^{-\lambda s+2is\sigma_{\mathbf{t}}(w,z)}}~{}ds=(2k)!(\lambda-2i\sigma_{\mathbf{t}}(w,z))^{-(2k+1)}.

This is now, as a function of ww\in\mathcal{H}, bounded and measurable, which finishes the proof. ∎

The careful reader may have noticed that the proofs of Proposition 4.12 and Lemma 4.13 generalize to the infinite dimensional phase space, i.e. analogous formulas for the Berezin transforms of products of resolvents and the classical resolvent functions, respectively, are valid. Nevertheless, since a correspondence theorem in the infinite dimensional setup (similar to Theorem 4.5) is not available at the moment, this observation is of no further use for proving a refinement of Theorem 5.4. The biggest issue is the lack of a Haar measure on the infinite dimensional symplectic space. In the finite dimensional framework of n\mathbb{C}^{n}, this measure coincides with the Lebesgue measure and is at the heart of quantum harmonic analysis. An important ingredient to the theory is a correspondence between the spaces L1(n)L^{1}(\mathbb{C}^{n}) and 𝒯1(F𝐭2)\mathcal{T}^{1}(F_{\mathbf{t}}^{2}) (the latter being the space of trace class operators). It is not clear what an appropriate interpretation of “L1(1/2)L^{1}(\mathcal{H}_{1/2})” should be in order to develop a quantum harmonic analysis and correspondence theory for infinite dimensional phase spaces. This will be a key issue for our future work.

References

  • [1] W. Bauer and J. Isralowitz, Compactness characterization of operators in the Toeplitz algebra of the Fock space FαpF_{\alpha}^{p}, J. Funct. Anal. 263 (2012), 1323–1355.
  • [2] W. Bauer and M. A. Rodriguez Rodriguez, Commutative Toeplitz algebras and their Gelfand theory: old and new results, Complex Anal. Oper. Theory 16 (2022), 77.
  • [3] C. L. Berger and L. A. Coburn, Toeplitz operators on the Segal-Bargmann space, Trans. Amer. Math. Soc. 301 (1987), 813–829.
  • [4] N. Bourbaki, Elements of mathematics: general topology chapters 1 - 4, Springer, 1989.
  • [5] O. Bratteli and D. W. Robinson, Operator algebras and quantum statistical mechanics 2, 2nd ed., Springer-Verlag, 1997.
  • [6] D. Buchholz, The resolvent algebra: Ideals and dimension, J. Funct. Anal. 266 (2014), 3286–3302.
  • [7] D. Buchholz and H. Grundling, The resolvent algebra: A new approach to canonical quantum systems, J. Funct. Anal. 254 (2008), 2725–2779.
  • [8] L. A. Coburn, The measure algebra of the Heisenberg group, J. Funct. Anal. 161 (1999), 509–525.
  • [9] G. Da Prato, An introduction to infinite-dimensional analysis, Universitext, Springer Verlag, 2006.
  • [10] K.-J. Engel and R. Nagel, One-parameter semigroups for linear evolution equations, Graduate texts in mathematics, vol. 194, Springer-Verlag, 2000.
  • [11] M. Fannes and A. Verbeure, On the time evolution automorphisms of the CCR-algebra for quantum mechanics, Comm. Math. Phys. 35 (1974), 257–264.
  • [12] G. B. Folland, Harmonic analysis in phase space, Princeton University Press, 1989.
  • [13] R. Fulsche, Correspondence theory on pp-Fock spaces with applications to Toeplitz algebras, J. Funct. Anal. 279 (2020), 108661.
  • [14]  , Toeplitz operators on non-reflexive Fock spaces, preprint available under arXiv:2202.11440, 2021.
  • [15] J. Janas and K. Rudol, Toeplitz operators on the Segal-Bargmann space of infinitely many variables, Linear Operators in Function spaces. 12th International Conference on Operator Theory Timişoara (Romania) June 6-16 1988 (H. Helson, B. Sz.-Nagy, and F.-H. Vasilescu, eds.), Operator Theory: Advances and Applications, no. 43, Birkh”auser Verlag, 1990, pp. 217–228.
  • [16]  , Toeplitz operators in infinitely many variables, Topics in operator theory, operator algebras and applications. 15th international conference on operator theory, Timişoara, Romania, June 6–10, 1994., IMAR Bucharest, 1995, pp. 147–160.
  • [17] M. Reed and B. Simon, Methods of modern mathematical physics 1: functional analysis, Academic Press, 1972.
  • [18] A. V. Skorohod, Integration in Hilbert space, Ergebnisse der Mathematik und ihrer Grenzgebiete, Springer Verlag, 1974.
  • [19] T. D. H. van Nuland, Quantization and the resolvent algebra, J. Funct. Anal. 277 (2019), 2815–2838.
  • [20] R. Werner, Quantum harmonic analysis on phase space, J. Math. Phys. 25 (1984), 1404–1411.
  • [21] B. W. Wick and S. Wu, Fock space on \mathbb{C}^{\infty} and Bose-Fock space, J. Math. Anal. Appl. 505 (2022), 125499.