1 Introduction
The classical CCR algebra (see [5]) provides a standard 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 on in the standard Schrödinger representation gives rise to -automorphisms of the CCR algebra over the standard symplectic space
only in the trivial case of an identically vanishing potential function . In order to resolve such problems, D. Buchholz and H. Grundling have proposed a new approach to CCR by introducing the resolvent algebra associated to a symplectic space 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 -space equipped with the standard symplectic structure. As is well-known, in this setting the CCR algebra is identified with the algebra generated by Toeplitz operators with symbols in the space TP of trigonometric polynomials. It is an interesting observation that this Toeplitz 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 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 , which are the symbols of Toeplitz operators linearly generating the resolvent algebra.
In dimension we show that the space corresponding to the resolvent algebra is the uniform closure of the classical resolvent functions. However, in higher dimensions we only can prove a weaker result (see Proposition 4.26). Nevertheless, we show that coincides with a 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 and in the sense of Theorem 4.5.
In the second part of the paper, we replace by a separable infinite dimensional complex symplectic Hilbert spaces . 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 algebras generated by Toeplitz operators over . Due to the non-nulcearity of (as well as of the space of entire functions over 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 with Hilbert-Schmidt embedding inside the full Toeplitz algebra over .
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 with the standard symplectic structure as a algebra generated by Toeplitz operators.
Section 3 provides some basic facts on Fock-Bargmann spaces over and Toeplitz operators
acting on . 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 be a symplectic vector space. Consider a Hilbert space and a real linear map from into the space of (unbounded) self-adjoint operators
on . Let us assume that all operators are essentially self-adjoint on a common domain which forms a core for each . Recall that the canonical
commutation relation (CCR) have the form:
|
|
|
(2.1) |
We call and 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 algebraic models which suitably encode (CCR). One standard idea consists in replacing above by suitable bounded functions of .
A classical approach due to H. Weyl amounts in considering the CCR algebra generated by the unitary operators , . The modified CCR are:
|
|
|
and self-adjointness of implies that .
More abstractly, one is interested in the algebra generated by the relations
|
|
|
|
|
|
|
|
Such algebras are usually called CCR algebras, see [5].
With put and consider the standard symplectic space
with symplectic form
|
|
|
One obtains a representation of as operators acting on the Fock-Bargmann space (see [8] and Section 3 for precise definitions). In fact, with our previous notation we put:
|
|
|
where is a Toeplitz operator with complex valued symbol on .
The corresponding CCR algebra is generated by the unitary Weyl operators (3.3) below which satisfy
|
|
|
i.e. is the generator of the unitary one-parameter group . Hence,
|
|
|
(2.2) |
The algebra is a classical object and in a more general setup it is discussed in [5].
In particular, the Fock-Bargmann representation of maps into the full Toeplitz algebra , i.e. the algebra generated by
Toeplitz operators with arbitrary bounded measurable symbols.
Theorem 2.1 (L. A. Coburn [8]).
It holds
|
|
|
where TP is the space of trigonometric polynomials:
|
|
|
Here we write for the operator norm closure of a given set .
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 algebra generated by the resolvents of the (in general unbounded) operators .
Definition 2.2 (D. Buchholz, H. Grundling [7]).
Given a symplectic space we define as the universal unital -algebra generated by the set
together with the relations:
-
(1)
,
-
(2)
,
-
(3)
,
-
(4)
,
-
(5)
,
-
(6)
,
for and .
The resolvent algebra is defined as the closure of with respect to a certain semi-norm
obtained through the GNS construction (cf. [7]).
Remark 2.3.
-
1.
Note that Equations (1) and (2) encode and the self-adjointness of , respectively. The third and sixth equation reflect
the -linearity of , whereas Equation (4) is the resolvent identity. Finally, the fifth equation is the substitute for the (CCR).
-
2.
When one considers a representation of as a concrete -algebra generated by resolvents of self-adjoint operators on a
Hilbert space, then coincides with the operator norm closure of .
In the present paper we study the Fock-Bargmann representation of the resolvent algebra over the standard symplectic space
as well as on an infinite dimensional Hilbert space with a symplectic structure. In the finite dimensional setting of it is well-known
(cf. [3, 8] and Section 3) that the generators are resolvents of self-adjoint Toeplitz operators , defined below.
Therefore, the resolvent algebra
is the algebra generated by the resolvents of Toeplitz operators:
|
|
|
(2.3) |
In fact, this will be the starting point of our analysis.
3 Fock-Bargmann space and Toeplitz operators
Let and consider positive real numbers where . We write and on we define the
probability measure by
|
|
|
Here, denotes the Lebesgue measure on . The Fock-Bargmann space is defined as
|
|
|
(3.1) |
where denotes the space of entire functions on . The reader experienced with the analysis on such spaces may wonder why we introduce the parameters in the above definition, as the standard approach corresponds to the choice . We will use the notation 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 and of by
|
|
|
and we write for the induced norm. As is well-known, is a reproducing kernel Hilbert space with kernel function
|
|
|
where In order to simplify notations, we write
|
|
|
(3.2) |
such that
.
We express the normalized reproducing kernels as
|
|
|
Let be a bounded linear operator on . We define the Berezin transform of
by
|
|
|
Note that is a bounded real-analytic function. Moreover, the map is known to be injective, [12].
As is a closed subspace of , there is an orthogonal projection acting on as
|
|
|
For a measurable function , we let denote the Toeplitz operator with symbol , which is defined by
|
|
|
on the natural domain
.
If , then is a bounded operator. For a given set , we will denote by the algebra generated by all Toeplitz operators with symbols in and set
|
|
|
for the full Toeplitz algebra. We also define to be the closed linear span of Toeplitz operators with symbols in .
Let us introduce Weyl operators on the Fock-Bargmann space : if then we define by
|
|
|
(3.3) |
Weyl operators are well-known to be unitary. Moreover, they fulfill the relation
|
|
|
Here, the symplectic form on with parameter is defined by
|
|
|
As a matter of fact, Weyl operators are Toeplitz operators themselves. If we define the family of bounded functions on by
|
|
|
then it holds
|
|
|
(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 in its Fock-Bargmann representation. Of course, it is not hard to reduce the analysis
of to that of . Again, we emphazise that the extra flexibility coming from the parameter set
will be useful when passing to the infinite dimensional limit . Therefore, we take the (mostly notational) burden upon us to carry all the way through this section. For readability, we will denote the Fock-Bargmann representation of the resolvent algebra also by .
Since the Toeplitz operators fulfill the canonical commutation relation (equivalently, the Weyl operators fulfill the relation of the CCR algebra), the resolvent algebra is the algebra generated by resolvents of Toeplitz operators. The following integral representations allow us to study more in detail:
Lemma 4.1.
For the map defines a strongly continuous unitary one-parameter group with generator . In particular, for the following integral representations of the resolvents hold in the strong sense:
|
|
|
(4.1) |
and
|
|
|
(4.2) |
Proof.
It is easy to check that is indeed a strongly continuous unitary one-parameter group. For determining its generator we compute
|
|
|
for all such that the limit exists.
According to Stone’s Theorem there is a unique self-adjoint operator with domain such that generates the group, i.e. for it holds
|
|
|
(4.3) |
in the norm sense. Since norm convergence in the reproducing kernel Hilbert space implies pointwise convergence, it suffices to determine the pointwise limit of
the right hand side of (4.3). Let and note that:
|
|
|
The integrand converges as :
|
|
|
An easy application of the dominated convergence theorem yields
|
|
|
for those such that .
Hence,
|
|
|
Therefore, is the generator of the one-parameter group . Since all operators
are unitary, the group has growth bound . For it follows (cf. [10, Theorem I.1.10])
|
|
|
strongly, i.e.
|
|
|
The integral representation (4.2) follows from:
|
|
|
|
which completes the proof.
∎
Remark 4.2.
An inspection of the above arguments shows that the limit in Equation (4.3) exists for all . Moreover, one easily verifies that
the space is invariant under the action of . Then, [17, Theorem VIII.10] implies that is essentially self-adjoint on .
Define for :
|
|
|
For simplicity, we will occasionally suppress in the notation and shortly write . Before we continue with our investigation, we need to recall the following well-known expansion of the resolvent.
Lemma 4.3.
Let such that . Then, it holds
|
|
|
(4.4) |
where the series converges in operator norm. In particular:
|
|
|
(4.5) |
The previous lemma has the following important consequence:
Corollary 4.4.
We further note that Equations (4.1) and (4.2) extend to with the same proof as in the case :
|
|
|
(4.6) |
and
|
|
|
(4.7) |
We mention two ways of analyzing the connection between the algebra 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 , the algebra of bounded, uniformly continuous functions on ,
with corresponding subspaces of .
We start with some notation. Let be a function and , then put:
|
|
|
Given an operator we define its shift by through
|
|
|
where is a Weyl operator. Clearly, is the subalgebra of consisting of functions
for which is continuous with respect to the -norm. The analogous space on the operator side is
|
|
|
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 . The case
of positive weight parameters in the definition of the Gaussian measure follows by identical proofs and obvious modifications:
Theorem 4.5 (Correspondence Theorem, [20, 13]).
Let be a closed, -invariant subspace (meaning for every ). Then, there is a unique -invariant closed subspace of such that
|
|
|
Further, for an operator the following statements are equivalent:
|
|
|
Finally, can be computed as follows:
In the following we call and corresponding spaces.
Here, we used the notation
|
|
|
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 algebras coincide: .
We mention that the previous two results extend to the case of the -Fock space where , 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 .
Let denote the ideal of compact operators on a Hilbert space .
Theorem 4.7 ([1]).
Let . Then, the following are equivalent:
|
|
|
Here, 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 and . Then:
|
|
|
Proof.
Applying the CCR of Weyl operators, we get for :
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
The case follows analogously.
∎
Proposition 4.9.
is a closed, -invariant subspace of .
Proof.
The -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 . It suffices to show that for and . This is an easy consequence of Lemma 4.3 and Lemma 4.8. In fact, for sufficiently small such that simultaneously
|
|
|
it follows:
|
|
|
|
|
|
|
|
|
|
|
|
This last expression tends to as .
∎
Proposition 4.9 and the correspondence theorem (Theorem 4.5) imply that
there is a closed, -invariant subspace of (possibly depending on the parameter tuple ) such that
|
|
|
Since itself is a algebra, we get:
Theorem 4.10.
is a Toeplitz algebra. More precisely,
|
|
|
for a suitable -invariant subspace .
Our next task is to determine the space explicitly.
4.2 Computing
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, corresponds to the compact operators and the almost periodic functions
correspond to the CCR algebra. In this respect, one might hope that the space corresponding to is:
|
|
|
Note that does not depend on the choice of the parameter set and we can replace by in its definition
(the “contribution” of the parameters can be absorbed into ).
Since is indeed a subalgebra of , which is further -invariant as well as invariant under the parity operation
Theorem 3.13 in
[13] implies that:
|
|
|
Therefore, if and only if .
Next, we collect some useful facts:
Proposition 4.11.
The following inclusions hold:
-
1.
.
-
2.
.
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 and . Given a multi-index and
we have:
|
|
|
|
|
|
where
|
|
|
Here, we have used the standard multi-index notation together with:
|
|
|
|
|
|
|
|
|
|
|
|
Proof.
Using the relation , we can reduce the proof to the case where for . From Lemma
4.3 we obtain that
|
|
|
where we use the short notation:
|
|
|
According to Lemma 4.8 we have:
|
|
|
|
|
|
|
|
|
|
|
|
By using analyticity of the resolvent maps it follows that the difference quotients converge in operator norm. Hence, differentiation can be interchanged with the inner product, which yields:
|
|
|
|
|
|
|
|
|
Now, we insert the integral expression of the resolvent in (4.6):
|
|
|
|
The inner product in the integrand can be calculated explicitly:
|
|
|
Finally, the assertion follows by inserting this expression into the last integral and performing the -derivatives.
∎
Similarly, one computes the Berezin transform of the classical resolvent functions:
Lemma 4.13.
Let , and . For any set of complex numbers
consider the function
|
|
|
With the notation in Proposition 4.12 the Berezin transform of is given by:
|
|
|
|
|
|
|
|
Proof.
The lemma follows by a direct calculation. First, note that
|
|
|
(4.8) |
Without loss of generality may assume that for such that has an integral representation (Laplace transform):
|
|
|
Inserting the last expression into (4.8) and interchanging the
order of integrations shows:
|
|
|
The inner integration can be evaluated explicitly:
|
|
|
which implies the statement of the lemma.
∎
We now prove a first inclusion of algebras:
Lemma 4.14.
It holds for every and . In particular,
|
|
|
Proof.
Since we already know that , it suffices to show that by the
correspondence theorem. Using we may assume without loss of generality that .
By transformation of the integral and according to Lemma 4.13:
|
|
|
|
|
|
|
|
|
|
|
|
where is a classical resolvent, namely:
|
|
|
Note that the Berezin transform behaves under dilations as follows:
|
|
|
where , and . Hence, we obtain:
|
|
|
Since again is a resolvent function and is invariant under the Berezin transform (which is simply the convolution by an appropriate Gaussian function), the inclusion
follows.
∎
Recall that an isotropic subspace is a (real) subspace such that for all . Every isotropic subspace
is of real dimension , and if , then is called Lagrangian. To every Lagrangian subspace
there exists a complementary Lagrangian subspace , i.e. . Indeed, one can choose and we will make this choice in the following for convenience.
If we now fix a Lagrangian subspace , then Proposition
4.12 shows that the unital algebra
|
|
|
is commutative. Note that is also -invariant according to Lemma 4.8.
It is our next aim to show that , in the sense of Theorem 4.5, corresponds to the space
|
|
|
i.e.
|
|
|
Before we approach this goal, note the following facts:
First, by Theorem 3.13 of [14], we have , i.e. it suffices to prove that . Secondly, since is also invariant under dilations
where , we obtain as in the proof of Lemma 4.14. Therefore, we only need to prove that where is a product of classical resolvent functions with . Since the operator algebra 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 .
Using the explicit formulas for the Berezin transform, one observes that point evaluations of the Berezin transforms are multiplicative linear functionals on .
Since for every and and , there is no loss of generality in considering only point evaluations of the Berezin transform at . Since the Berezin transform is injective, we can expect the Gelfand spectrum of to be a suitable compactification of . We will describe this compactification now and start by recalling a compactification of a real inner product space first constructed in [19]. Therein the author described the maximal ideal space of as such
a compactification of .
Let be a finite-dimensional real inner product space. We denote by the orthogonal projection onto a given subspace . By we denote the
affine Grassmannian of , i.e. the set of all affine subspaces of . As a set, this can be written as
|
|
|
The precise topology with which is endowed can be found in [19] and will not be described here. We will denote by the set endowed with this particular topology. We collect some facts about in the following lemma:
Lemma 4.15 ([19]).
Let be a finite-dimensional real inner product space.
-
1.
is a compact Hausdorff space.
-
2.
Together with the embedding , is a compactification of .
-
3.
A net converges to if and only if the following hold:
-
•
-
•
eventually
-
•
There is no affine subspace such that there exists a subnet of with and eventually (along the subnet).
One of the main theorems of [19] is the following:
Theorem 4.16 ([19]).
The Gelfand spectrum of can be identified with the above compactification of , i.e. .
Let us briefly describe the identification of with in more detail:
Given a function and an affine line , , it is not hard to verify
that exists. For an affine subspace and any with ,
the value of this limit is almost everywhere, with respect to the surface measure on , the same. If we denote this value by , then this defines a multiplicative linear functional and every element of can be obtained in this way.
Indeed, the same can be done for , and the following holds true:
Proposition 4.17.
The Gelfand spectrum of can be identified with the above compactification of
, i.e. .
Here, we consider as a real inner product space with the -weighted inner product induced from , i.e. . 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 , every classical resolvent with extends to a function in .
As we will see next, the Gelfand spectrum of is indeed the same as the one of :
Proposition 4.18.
The Gelfand spectrum of can be identified with , where all multiplicative linear functionals on have the form:
|
|
|
Here, we denote by the multiplicative functional of introduced above.
Before proving Proposition 4.18, let us derive our intended result from this:
Theorem 4.19.
It is .
Proof.
The inclusion was already stated above.
Let . Since is translation invariant and closed, we conclude that
. Therefore, is
the Gelfand transform of an operator in . This operator must be , i.e. .
∎
We will present a sequence of lemmas that lead to a proof of Proposition 4.18.
Lemma 4.20.
For any resolvent , where , and it holds:
|
|
|
Proof.
Let and first observe that iff (note that by our choice ). Hence, we have
|
|
|
|
|
|
|
|
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 where and any affine line the
limit exists and is given by
|
|
|
(4.9) |
Proof.
If , then such that and the equality follows from the previous lemma. Assume that and let
(the case follows similarly). According to Proposition 4.12 the Berezin transform of the resolvent at has the value:
|
|
|
We use integration by parts
in order to show that the right hand side converges to as . With the short notation we have:
|
|
|
|
|
|
|
|
|
|
|
|
Since the integral
|
|
|
over the absolute value is finite, we obtain as .
∎
Lemma 4.22.
Let with be a resolvent in . Then, its restriction to extends to a continuous
function on .
Proof.
We first describe the values that (the extension of) takes on and discuss continuity in a second step.
The value of the Berezin transform at is given by . At each affine line the value of the extension of the Berezin transform to is defined by the right hand side of (4.9). For a given affine spaces with , we distinguish
two cases: If , then the value of coincides with . Otherwise,
if , then we set it to be zero.
Note that these choices are in accordance with the definition of the multiplicative linear functional . We explain what we mean by this only in the last example:
If , then is a subspace of of dimension strictly smaller than the dimension of . Hence, the unit sphere of is a zero set with respect to the surface measure of the unit sphere of . Hence,
It remains to prove that the above extension of from to the compactification is in fact continuous. According to
Bourbaki’s Extension Theorem, [4, Theorem 1, p. 82], it suffices to show that
|
|
|
(4.10) |
for any net such that . We distinguish several cases:
-
1.
If , then (4.10) follows from the continuity of the Berezin transform on .
-
2.
If with , then:
|
|
|
|
|
|
|
|
-
3.
If with , then we conclude that
|
|
|
In fact, assuming the opposite there exists a subnet, also denoted by , such that is bounded away from zero.
Take such that and put . By denote the orthogonal complement
of in . We obtain a decomposition of :
|
|
|
(4.11) |
-
(a)
If was unbounded, we could pass to a subnet such that . For this subnet, one could show by using the method of stationary phase
(as in the proof of Lemma 4.21) that . However, this is impossible as we assumed that
is bounded away from zero.
-
(b)
If is bounded, we can pass to a subnet such that converges (say, to some ).
From the orthogonal decomposition (4.11) we see that:
|
|
|
From it follows:
.
However this contradicts the assumption . ∎
Since the Berezin transform is multiplicative on and in implies uniformly, we conclude that
for every extends to a continuous function on . Now we can give the proof of Proposition 4.18:
Proof of Proposition 4.18.
We claim that the map
|
|
|
is a bijective and unital homomorphism of algebras. By what has been said before, defines a continuous and injective unital -homomorphism.
We are left with verifying surjectivity. Since the range of is a -algebra, we only need to show that it separates points of in order to apply the
Stone-Weierstrass Theorem. Showing that the points are separated by 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 .
-
1.
is separated from any other affine subspace in the following way: If is a proper subspace, then let and consider
for . Then, can be chosen such that
|
|
|
However, as long as . Hence the affine spaces and are separated. Separating from
can be done by considering .
-
2.
Affine spaces are separated as follows: We necessarily have and . If for some , then consider for suitable .
Otherwise, we can find such that . Let such that . Hence for :
|
|
|
|
|
|
|
|
Now, one has to choose accordingly such that these values are different (see the formula in Proposition 4.12).
-
3.
If are proper subspaces of , then we consider two cases:
-
(a)
Recall that we always have for an affine subspace . If with are real multiples of each other, then we obtain with :
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Now, arrange such that these are not equal.
-
(b)
If and are not real multiples of each other, then it must be either or . Assume that
. For :
|
|
|
|
|
|
|
|
The value at now depends on whether is orthogonal to or not: If , then
|
|
|
|
|
|
|
|
Since , we can arrange such that the values of
for are different.
If, on the other hand, , then
|
|
|
Hence, we have to arrange such that (e.g. by letting ).∎
Before continuing, we state a consequence of the previous considerations:
Corollary 4.23.
Let and . Consider the function
|
|
|
where and is Lagrangian. Then, .
Proof.
By definition we have . Now Theorem 4.19 implies that
.
∎
Now we consider resolvent functions for which the vectors are either taken from or from .
Corollary 4.24.
Let and be two Lagrangian subspaces of such that . Let and and such that and . Set
|
|
|
where and . Then, .
Proof.
For we have
|
|
|
As , the first factor tends to . As , the second factor tends to . In conclusion, . Hence, . Since according to Proposition 4.11 the statement follows.
∎
Theorem 4.25.
Let , then .
Proof.
If , then any classical resolvent function either fulfills the assumption of Corollary 4.23 or the assumptions of Corollary 4.24. In either case, we obtain
.
∎
For , the situation is more complicated. We will show only a weaker result for this case.
Since every is contained in some Lagrangian subspace , we conclude that each resolvent is contained in some
. Thus, we see that
|
|
|
contains every resolvent . Further, since , we obtain
|
|
|
where is given by
|
|
|
As a weak substitute for Theorem 4.25 in higher dimensions, we get:
Proposition 4.26.
.
It might of course be true that . In this case, the previous result would imply . 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 be a summable sequence of strictly positive real numbers. Then, we set with the usual norm
and define:
|
|
|
|
|
|
|
|
We may think of and as the range of and , respectively, where is the trace class
operator obtained by linearly extending the map with being the standard orthonormal basis of .
We consider as a symplectic Hilbert space equipped with the symplectic form:
|
|
|
Note that is well-defined even for and . We can write , where is the canonical inner product of :
|
|
|
Moreover, denotes the corresponding norm: .
We recall some basic facts about Gaussian measures on infinite dimensional Hilbert spaces, cf. [18, 9, 2]. The infinite product measure of the Gaussian measures
|
|
|
gives a well-defined probability measure on . It is concentrated on .
By , we also denote its restriction to the measurable space , where is the Borel -algebra of . Note that agrees with the -algebra generated by cylindrical Borel sets. The measure is called the (centered) Gaussian measure on
with covariance operator : This is due to the fact that is naturally related to by
|
|
|
(5.1) |
The space introduced above is also known as the Cameron-Martin space of . It is a measurable set of measure zero: .
Hence, so is .
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 for the set of all sequences with values in such that all but finitely many entries are zero. In the following
we use the standard multi-index notation: let and , then we put:
|
|
|
The monomials below form an orthonormal system of complex analytic functions on inside :
|
|
|
We write and for the standard inner product and norm in , respectively.
Elements in the linear span of are referred to as analytic polynomials. In analogy to the finite dimensional setting of we define the Bargmann-Fock space to be the
-closure of analytic polynomials:
|
|
|
Let and be an analytic polynomial. Extending (3.3) we define the Weyl operator on by:
|
|
|
Here, denotes the normalized reproducing kernel defined by:
|
|
|
Let and assume that the analytic polynomial only depends on the variables . We write
, where and . Define the projections
for a sequence . Since we obtain:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Therefore, and acts isometrically (say, on analytic polynomials). In finite dimensions, it would now be clear that , since defines a holomorphic function. In infinite dimensions, this statement is not entirely trivial, as is defined as the closure of the analytic polynomials. Nevertheless, remains true, cf. [15]. Hence, extends to an isometric operator on for every .
Further, note that for , in we clearly have
|
|
|
pointwise almost everywhere on . Hence, Scheffé’s Lemma shows that the assignment is continuous from
to for every analytic polynomial . Since these are dense in , the Weyl operators extend to isometric operators on such is continuous on with respect to the strong operator topology.
An even stronger statement holds, namely the map is even strongly continuous with respect to the coarser topology of , cf. [15]. Therefore, it extends to a map from to the unitary operators on .
By applying the Weyl operators with to analytic polynomials, it is not hard to see that
|
|
|
These relations extend to , as well:
Lemma 5.1 ([15, 16]).
The Weyl operators depend continuously on with respect to the
strong operator topology and
|
|
|
for every .
Fix a bounded operator and . We define the Berezin transform of at by
|
|
|
Lemma 5.2.
The map is injective on .
Proof.
If , then for every . We denote by the closed linear span in
of the set of all analytic polynomials in the finitely many variables . Moreover, we write for the orthogonal
projection onto . Then, for and hence:
|
|
|
Note that can be naturally identified with the -variable Fock-Bargmann space with weight parameter . Moreover,
|
|
|
On the right hand side of the last formula we have use the Berezin transform on .
Since the Berezin transform is injective on (see [12]), we conclude that .
Hence, for any analytic polynomial and sufficiently large. Finally, we have
|
|
|
in , i.e. for analytic polynomials showing that .
∎
We denote by the orthogonal projection from onto . Given we define the Toeplitz operator by:
|
|
|
By applying the injectivity of the Berezin transform in Lemma 5.2, it is not hard to verify that the Weyl operators for
have a representation as Toeplitz operators. More precisely:
|
|
|
(5.2) |
Due to the importance of this fact, we derive an explicit formula for the Berezin transform of both operators:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
It is an important but non-trivial fact that extends to . First, we have to explain in what sense the
map
defines a measurable function on for . Clearly, the expression is pointwise
well-defined on . However, since is a set of measure zero, this is of no big help. We need a preliminary result:
Lemma 5.3.
for .
Proof.
We introduce a complex parameter . Then, it is not hard to verify the following equalities through standard results on differentiation of parameter integrals, where we use that (this is easily established):
|
|
|
On the other hand, using Equality (5.1):
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
We therefore obtain:
|
|
|
Choosing yields the desired equality.
∎
Hence, if is a Cauchy sequence with respect to , then is Cauchy in .
In conclusion, the family of functions continuously extends to . In particular, for each , is a well-defined expression for almost every and, as a function of , measurable. In particular,
|
|
|
is an almost everywhere well-defined function with
|
|
|
We also note that, since for every , the same is true for . Iterating
the procedure from the proof of Lemma 5.3, one easily obtains:
|
|
|
Next, we calculate:
|
|
|
We obtain:
|
|
|
for every . By a standard density argument, we have:
|
|
|
Now, one can compute the Berezin transform of as above. For every and one obtains:
|
|
|
Further, since continuously (in strong operator topology) depends on the parameter , we conclude that the Berezin transform
continuously (in the topology of pointwise convergence) depends on , which gives
|
|
|
for every and . Comparing Berezin transforms, we have shown that extends to .
In an abuse of notation, we will write for the representation of the resolvent algebra on with
respect to the symplectic space . Since the Weyl operators , satisfy the CCR in
Lemma 5.1, we start again by expressing the resolvent in form of a Laplace transform of the one-parameter group
.
For we have:
|
|
|
which exists as an integral in strong operator topology. There is an analogous formula in the case . Consider the space of (classical) symbols:
|
|
|
Moreover, we write for the C∗ algebra generated by Toeplitz operators over with symbols in
.
Theorem 5.4.
The following inclusion holds true: .
Proof.
Without loss of generality we assume . 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:
|
|
|
|
|
|
|
|
|
|
|
|
Since the Weyl operators are unitary, and hence satisfy , it follows that:
|
|
|
Therefore, the dominated convergence theorem gives as :
|
|
|
We have therefore seen that
|
|
|
where the series converges in operator norm. Fix for the moment . Since the Riemann integral exists as a limit of Riemann sums in strong operator topology (this easily follows from the fact that the mapping is strongly continuous), we obtain for the Berezin transform:
|
|
|
Therefore, injectivity of the Berezin transform (Lemma 5.2) shows:
|
|
|
Since
|
|
|
uniformly, this gives
|
|
|
For the symbol, standard facts on the Laplace transform yield
|
|
|
This is now, as a function of , 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 , 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
and (the latter being the space of trace class operators). It is not clear what an appropriate interpretation of “”
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.