Abstract

We apply a special case, the restriction principle (for which we give a definition simpler than the usual one), of a basic result in functional analysis (the polar decomposition of an operator) in order to define , the -version of the Segal-Bargmann transform, associated with a finite Coxeter group acting in and a given value of Planck's constant, where is a multiplicity function on the roots defining the Coxeter group. Then we immediately prove that is a unitary isomorphism. To accomplish this we identify the reproducing kernel function of the appropriate Hilbert space of holomorphic functions. As a consequence we prove that the Segal-Bargmann transforms for Versions , , and are also unitary isomorphisms though not by a direct application of the restriction principle. The point is that the -version is the only version where a restriction principle, in our definition of this method, applies directly. This reinforces the idea that the -version is the most fundamental, most natural version of the Segal-Bargmann transform.

1. Introduction

The basic idea involved in the restriction principle is the use of the polar decomposition of an operator in order to define a unitary transformation. The polar decomposition (e.g., see [1, 2]) is a well-known result in functional analysis that says that one can write , where and is a partial isometry. Here is a closed (possibly unbounded), densely defined linear operator mapping its domain to , where and are complex Hilbert spaces. It turns out that is the positive square root of the densely defined self-adjoint operator and so maps a domain in to . The partial isometry maps to . We are generally interested in the case when the partial isometry is a unitary isomorphism from onto , which is true if and only if is one-to-one and has dense range.

Applying the polar decomposition theorem as a means for constructing unitary operators is a very general method. Also this method has nothing to do with the structures of the complex Hilbert spaces and . And these can be advantages or disadvantages depending on one's particular interest.

But if we assume that is some set of complex-valued functions (and in general not equivalence classes of functions) on a set and is a Hilbert space of complex-valued functions (or possibly equivalence classes of functions) on a subset of , then we define the restriction operator by for all and all . This is at a formal level only, since in general, we do not know that restricted to is an element of . Then we apply the polar decomposition to the adjoint of the restriction operator (provided that is a closed, densely defined operator) to get , where is a positive operator of no further interest and is a partial isometry from to . We say that is defined by the restriction principle. We then have to show that is one-to-one and has dense range in order to prove that the partial isometry is a unitary isomorphism from onto , this being the case of interest for us. In this paper will be a reproducing kernel Hilbert space. This turns out to be quite useful for deriving explicit formulas, but it is not a necessary aspect of this approach.

We note that our definitions here differ from those of other authors. For us a restriction operator is simply restriction to a subset and nothing else. Other authors allow for operators that are the composition of restriction to a subset followed or preceded by another operator, often a multiplication operator. Then these authors apply the polar decomposition to these more general โ€œrestrictionโ€ operators. Now this introduces another operator as a deus ex machina; that is, something that arrives on the stage without rhyme or reason but that saves the day by making everything work out well. We object to such an approach to constructing a mathematical theory on general principles, both aesthetic and logical. Moreover, in the context of generalizations of Segal-Bargmann analysis it seems that the application of the restriction principle (using our definition of this) in the context of the -version of Segal-Bargmann analysis eliminates any need to introduce unmotivated factors. This is clearly seen in this paper as well as in [3, 4]. Also, as we will see, developing the theory first for the -version gives us enough information to dispose easily of the other versions, including an explanation of where the โ€œmysteriousโ€ multiplication factors come from for the , , and versions. See Hall [5] for the original use of this nomenclature of โ€œversionsโ€ and [6] for its use in the context of finite Coxeter groups.

When the above sketch can be filled in rigorously, this is a simple way of defining a unitary isomorphism . Moreover, the simplicity of the definition often allows one to prove results about in a straightforward way. However, the devil lies in the details as the saying goes, and the details can sabotage this approach. For example, the definition of the restriction operator might not make sense on the domain though it always makes sense on the subspace . However, it could happen that is the zero subspace, in which case this method is for naught.

The full history of this method is not our primary interest, but we present what we know about this in the area of mathematical physics and related areas of analysis. In this paragraph, and only in this paragraph, the phrase โ€œrestriction principleโ€ is used in the sense of the authors cited. Peetre and Zhang in 1992 in [7] used polar decomposition to get the Berezin transform. A polar decomposition was used by ร˜rsted and Zhang in [8] in order to define and study the Weyl transform. The article [4] by ร“lafsson and ร˜rsted contains some applications of restriction principles in order to understand the work of Hall in [5] and Hijab in [9]. The approach in [4] was recently followed up by Hilgert and Zhang in [3] in their study of compact Lie groups. Also Davidson et al. used a restriction principle in [10] in order to study Laguerre polynomials. See [10] for more references on this topic and on the Berezin transform. Zhang in [11] used a restriction principle to study the Segal-Bargmann transform of a weighted Bergman space on a bounded symmetric domain. In [12] a restriction principle was used by Ben Saรฏd and ร˜rsted to produce a โ€œgeneralized Segal-Bargmann transformโ€ associated with a finite Coxeter group acting on . We first learned about this method by reading [8] within some six months of its publication. But our recent interest was stimulated by our desire to understand [12].

We should note that the same generalized Segal-Bargmann space as found in [12] together with its associated Segal-Bargmann transform (but called the chaotic transform) can be found for the case , (dimension ) in Sifi and Soltani [13] and for , (arbitrary finite dimension ) in Soltani [14]. However, neither [13] nor [14] used a restriction principle. The case , is discussed by us in [15] and in the references found there, while we studied the arbitrary finite dimensional case , in [6]. Our point of view in [6, 15] was to use the approach of Hall [5], which is directly based on heat kernel analysis, rather than using the restriction principle. While the restriction principle can be considered as an alternative to the approach of Hall, this approach still relies in an essential way, at least in this paper, on the heat kernel of the Dunkl theory as we will see.

The restriction principle approach has various limitations. For example, and need not be manifolds and even if they are, need not be the cotangent bundle of so that the theory can lose contact with physics and symplectic geometry. Also, the Hilbert spaces are not constructed, but they must be known prior to applying this approach. And there is no necessary connection with heat kernel analysis. Of course, these attributes can be viewed as strengths rather than weaknesses, since they could allow for more general application than other approaches.

In this paper we will use the restriction principle to define the version of the Segal-Bargmann transform associated with a finite Coxeter group acting on and with a value of Planck's constant. (We will discuss the multiplicity function later on.) We also show that is a unitary isomorphism. This is a new way to construct and prove that it is a unitary isomorphism. Along the way we have to find an explicit formula for the reproducing kernel function for the Hilbert space that turns out to be the range of the unitary transform .

A major point of this paper is that our original proof of the unitarity of the transform , as given in [6], depends on using the previously established unitarity of , the -version of the Segal-Bargmann transform. Since none of the versions of the Segal-Bargmann transform appears as the most natural version in the analysis given in [6], there is no logical reason to start with the -version. However, using that approach, things in the end do work out quite nicely. But the proof given here seems to us to be more natural, since the starting point, namely the -version, plays a distinguished role, while the remaining versions are obtained as secondary constructs.

Having established these results in the -version, it is then simple for us to prove the corresponding results for Versions , , and . In particular we show as an immediate consequence of our work how the โ€œrestrictionโ€ operator used in [12] (which is actually restriction followed by multiplication by an unmotivated factor) arises in a natural way from our restriction operator, which is simply restriction without multiplication by some fudge factor.

The upshot is that the restriction principle for the -version can be used as a starting point for defining all of the versions of the Segal-Bargmann transform associated to a finite Coxeter group. Therefore the restriction principle is a fundamental principle in Segal-Bargmann analysis. So, this paper complements the approach in our recent paper [6], where we showed by using the Dunkl heat kernel that the versions , , and of the Segal-Bargmann transform associated with a finite Coxeter group are analogous to the versions of the Segal-Bargmann transform as introduced by Hall in [5], where he used the appropriate heat kernel.

Since many authors now take the -version to be the most fundamental version of the Segal-Bargmann transform, we feel that our result has an impact on that approach to this field of research. We also feel that the current approach is better than that in [6], since we now emphasize how the -version is singled out in yet another way as more fundamental than the other versions.

2. Definitions and Other Preliminaries

We follow the definitions and notation of [6]. Consult [6] and the references given there for a more leisurely review of this material. In that paper we studied various versions of the Segal-Bargmann transform associated with a finite Coxeter group acting on the Euclidean space . One of these versions (known as Version or the -version) is, as we will see, a unitary isomorphism of Hilbert spaces where the density function (with respect to Lebesgue measure) for is Throughout this paper we let denote Planck's constant. Here the Macdonald-Mehta-Selberg constant is defined by (In a moment we will discuss , the finite set and .) Since this integral does not depend on the value of (by dilating), we do not include this parameter in the notation on the left side. Clearly, .

We define the Version Segal-Bargmann transform as the integral kernel operator for and and , where the integral kernel is defined for and by where (given that ) is a holomorphic function and is the usual Euclidean norm squared. The function will be introduced momentarily. Using (2.9) below and the Cauchy-Schwarz inequality one shows the absolute convergence of the integral in (2.4). Our paper [6] provides motivation for formula (2.5).

In the above is a certain finite subset of , known as a root system, is a multiplicity function (see [6] for definitions) and It may be possible to weaken the hypothesis that we are imposing here while still having the same results. We work with a fixed root system and a fixed multiplicity function throughout this paper. See [6] for the details about how gives rise to a finite Coxeter group acting as orthogonal transformations of .

The space introduced above ([12, 14]), which is called the Version Segal-Bargmann space, is the reproducing kernel Hilbert space of holomorphic functions whose reproducing kernel is defined for and by where is the Dunkl kernel function associated with the Coxeter group (associated itself to the root system ) and the multiplicity function . For any , we let denote its complex conjugate. The Dunkl kernel (see [16โ€“18]) is a holomorphic function with many properties. We simply note for now that for all and all . In the first equation 0 denotes the zero vector in . Also, in the obvious notation. We will also be using the estimate (see [19]) for all , which holds if . (Here, is the Euclidean norm of . Also recall that is assumed throughout this paper.)

For a Hilbert space we use the notations and for its inner product and norm, respectively. The inner product is antilinear in its first argument, linear in its second. All Hilbert spaces considered are over the field of complex numbers.

We will be using dilations. Our present notation for these operators is , where is a function in some appropriate function space. The proof of the next result is straightforward and so is left to the reader.

Lemma 2.1. For every and , we have that is a unitary isomorphism.

Finally, we want to introduce the Dunkl heat kernel (see [18, 19]) for the heat equation associated with the Dunkl Laplacian ; namely The Dunkl Laplacian is defined and discussed in [19]. In particular, it has a realization in as an unbounded, self-adjoint operator with and spectrum . Specifically, we have for and that solves (2.11) for any initial condition (see [18] for more details), where the Dunkl heat kernelโ€‰โ€‰ is given for all and by This has an analytic extension , which we also denote as . One of the basic results of [6] is that for and we have which, in accordance with the approach of Hall [5], indicates that (2.4) is justifiably called the Version Segal-Bargmann transform associated with a finite Coxeter group. This formula also clarifies the nature of the seemingly arbitrary definition (2.5) of the kernel function of the integral transform .

Notice that the reproducing kernel function for clearly satisfies This identity shows that the reproducing kernel function for the Hilbert space is determined by the Dunkl heat kernel . Or, in other words, we can get the Segal-Bargmann space for Version from the Dunkl heat kernel. Another way to write this reproducing kernel in terms of the Dunkl heat kernel is to consider equation (46) in Hall [5]. In the present context the analogous result says that for all we have as the reader can check. (Hint: one needs an identity involving the Dunkl kernel. See [12, Equationโ€‰], or [18, Propositionโ€‰2.37, equation (2)].) Even though we will not be using these two formulas for , we present them to show how the Dunkl heat kernel determines the reproducing kernel of . As we will show later, the reproducing kernel function for the Version Segal-Bargmann space is also determined by the Dunkl heat kernel .

We gather here some basic results of functional analysis that we will be using. (See [1], especially Chapter III, Sectionโ€‰5 and Chapter V, Section โ€‰3, for more details.) Let and be complex Hilbert spaces with a linear operator which is densely defined (which means is a dense subspace in ). Let denote the adjoint of . If is closable (namely, has a closure), then we denote the closure of by . We denote the kernel and range of by and , respectively. We say that is globally defined if . For any subset in a Hilbert space, is its closure in the norm topology and is its orthogonal complement. The following proposition comes from elementary functional analysis.

Proposition 2.2. Let be densely defined, as above. Then we have the following: (1)if is closable, then is closed, densely defined and , (2), (3)if is closed, then , (4)if is bounded (i.e., there exists such that for all ), then is closable and is globally defined and bounded (with the same bound as ). In particular, if is bounded and closed, then is globally defined, that is, .

As we have already mentioned, we will use a standard result of functional analysis known as the polar decomposition of an operator. For the reader's convenience we state this result. We present a modification of the statement of Theorem VIII.32 in [2]. A very thorough discussion of this topic is also given in [1]. (See Chapter VI, Sectionโ€‰2.7.) We state this theorem for a closed densely defined linear operator (that is, it may be bounded or not).

Theorem 2.3 (Polar Decomposition). Let and be Hilbert spaces and be a closed linear operator, defined in the dense linear domain . Then there exists a positive self-adjoint operator with and there exists a partial isometry with initial space and final space such that on their common domain . Also, and are uniquely determined by and the above properties.
In particular, is one-to-one if and only if , while is onto if and only if is dense.
Consequently, is a unitary isomorphism of onto if and only if and is dense.

Remarks 2.4. Theorem 2.3 is stated in terms of the structures of Hilbert spaces, nothing else. So it is invariant under unitary isomorphisms. To make this more explicit we suppose are unitary isomorphisms for , where and are Hilbert spaces. (We continue using the notation of Theorem 2.3.) Then define , a subset of , and by . Clearly, is a closed, densely defined operator. So, according to Theorem 2.3, we have that , where and is a uniquely determined partial isometry. Then the relation of the polar decomposition of with that of is Moreover, if where is a restriction operator, then according to our definition is defined by a restriction principle. Nonetheless, need not be defined by a restriction principle; that is, need not be the adjoint of a restriction operator even though is. However, is well defined by polar decomposition. While the restriction principle is not a unitary invariant, this discussion shows that there is a straightforward method for transforming a polar decomposition by unitary transformations. There is absolutely no guesswork involved.
It seems to be a rule of thumb in Segal-Bargmann analysis that it is rather straightforward to prove that a Segal-Bargmann transform is injective, while to prove that it is surjective requires a rather detailed argument. However, that is not so for the restriction principle we will consider. On the contrary, as we will see in the next section, proving that the transform is surjective is immediate (using uniqueness of analytic continuation), while proving that it is injective does involve a bit more work (using that the Dunkl transform, to be discussed later, is injective) though is not all that difficult.

3. Version

In this section we will show how Version of the Segal-Bargmann transform associated to a Coxeter group arises from the restriction principle. We feel that using the restriction principle is a more fundamental approach to this theory.

We recall from [6] that the Version (or -version) Segal-Bargmann transform for and is defined by where . This integral converges absolutely by using the estimate (2.9). This definition is the natural analogue in this context of the definition of the -version given in [5]. Then we proved in [6] that this gives a unitary isomorphism The definition of the Hilbert space of holomorphic functions will be given below. Other details may be found in [6]. First, we will identify the reproducing kernel function for this Hilbert space. But we want this space to be the image of the coherent state transform , and it is well known in the theory of coherent states that the reproducing kernel in the codomain Hilbert space (if it exists) is necessarily given by the inner product in the domain Hilbert space of two coherent states. In this context that will be which can be calculated to give the formula for in the next theorem. Essentially, this involves the time evolution for the time interval of the heat kernel at time . This is why the result is (given our conventions, proportional to) the heat kernel at time . An important point here is that there is no liberty in the choice for this reproducing kernel, but rather it results from evaluating an integral.

We again call to the reader's attention that restriction principles do not define the Hilbert spaces, which must be introduced prior to the application of a restriction principle. And so it is in the present case with the Hilbert space .

Theorem 3.1. The reproducing kernel function for the Hilbert space is given by for all .

Remarks 3.2. Note the similarity of formula (3.4) with the reproducing kernel for the Version generalized Segal-Bargmann space for compact, connected Lie groups as given by Hall in [5] (Theoremโ€‰6, page 127): See [5] for the definition of this notation and further details. Also, note that this formula occurs in Segal-Bargmann analysis in the context of Heisenberg groups in [20] and in the context of the compact Heckman-Opdam setting in [21]. Admittedly, the factors of 2 in our formula look strange and are not found in these references. These factors are a consequence of the unusual convention we have introduced in [6] for normalizing the Dunkl heat kernel and the measure .

Proof. We let denote the space of all of the holomorphic functions . We recall three definitions from [6]. For we define by for all . (Note that is never zero.) Then we define which becomes a Hilbert space with its inner product defined by for .
The reproducing kernel of a Hilbert space must satisfy two characteristic properties. The first of these is that must be an element in the Hilbert space . The second is that for all and .
We start with the first property. Now if and only if is an element of as a function of .
So we calculate Here is the reproducing kernel function for the Hilbert space , which implies that for all and so as desired.
Now for the second property we evaluate the right side for (which implies ) and use to get for all . So the second property has also been established, thereby completing the proof without ever using the transform .

The definition of the Hilbert space given in (3.6), (3.7), and (3.8) is what we were naturally led to while preparing [6]. It is the range space of the Version Segal-Bargmann transform introduced there. However, the result of Theorem 3.1 gives us an intrinsic way of defining , namely as the Hilbert space of holomorphic functions with reproducing kernel defined by (3.4). This is arguably a better approach. However, the natural way to do this would be to omit the factors of 2 from (3.4). This would simply give us a different normalization of the Version of the Segal-Bargmann space. But either way the Hilbert space must be defined before applying a restriction principle, as we noted earlier.

Of course, in order to apply the restriction principle, we need to define the restriction operator rigorously.

Definition 3.3. We define the restriction operator by for all in a domain and all . The definition of the domain of in is the obvious one Note that does depend on and , since these parameters appear in both the domain and codomain spaces of this operator.
We will show later on that is a globally defined, bounded operator. Still this is a bit surprising since the following standard estimates do not prove it. Indeed, for any we have that Here we used (2.9) in the second inequality and the usual pointwise estimate for functions in a reproducing kernel Hilbert space in the first inequality.
As far as we know at this point of our exposition it could well be the case that . We now show that this domain is actually dense along with other properties of .

Theorem 3.4. The operator defined on its domain is a closed, densely defined operator that is one-to-one and has dense range in . Also its adjoint is densely defined, closed, one-to-one and has dense range. In particular, we have that for all .

Proof. By the uniqueness of analytic continuation from to , we have immediately that is one-to-one, that is, .
We claim that the functions are all in . This follows from for and , which (using and (2.9)) gives the estimate This clearly implies that is integrable with respect to the measure . And so . Now, by the theory of reproducing kernel Hilbert spaces, the finite linear combinations of the functions with form a dense subspace of and so is dense; that is, is a densely defined operator.
The proof that the graph of is closed is a standard argument, which we leave to the reader. So, is a closed operator.The proof that is a densely defined and closed operator follows by applying Proposition 2.2 to the closed operator .
To prove that is injective, we first find a formula for . So we take and with the intention of calculating in general. Introducing the reproducing kernel in the second equality and using in the third equality we calculate as follows: Notice how the factors of 2 combined with to form , which is the measure we want to use in integrals involving the Dunkl heat kernel .
Now put for in (3.18) to get where is the Dunkl transform. (See [16โ€“19] for information on this transform and [6] for our notation and conventions. For this argument, we only need to know that is injective.) At this point, let us note that implies that so that (3.19) makes sense.
We now assume that . So, for all . Using that is injective on , it follows from (3.19) that almost everywhere with respect to the measure . Hence almost everywhere with respect to . This shows that is injective.
To prove that the ranges are dense, we will again use Proposition 2.2. Since is closed we have that and that . The last equality then implies that . (We use the symbol 0 here to designate ambiguously the zero subspace of the appropriate Hilbert space.)

We have shown that the range of the restriction operator is dense only for the sake of completeness. This will not be used later on.

We continue with our main result.

Theorem 3.5 (Restriction principle: Version ). (i) Suppose that the multiplicity function satisfies . The restriction principle says that the partial isometry produced by writing the adjoint of the restriction operator, namely , in its polar decomposition; that is, actually gives a unitary isomorphism .
(ii) Moreover, we have that where is defined by equation (3.1).
So it follows that , the -version of the Segal-Bargmann transform associated with a finite Coxeter group and the value of Planck's constant, is a unitary isomorphism.

Remark 3.6. Instead of using the definitionโ€‰ from [6], we can use the first part of this theorem to define . It is in this sense that the restriction principle can be said to define the -version of the Segal-Bargmann transform.

Proof. We begin by finding another formula for . So we take and . Continuing the calculation given above in (3.18), we obtain (Parenthetically, we warn the reader that this equation does not say that is equal to . This quite simply can not be true, since the codomains of these two operators are not the same space. The correct statement is that is equal to followed by analytic continuation to . Also, it is clear that , since the domains of and are equal as sets.)
To get the polar decomposition of we have to analyze the operator . But , since is closed. So we consider from now on. By using the definition of we immediately get for and that and so on which is dense in by a theorem of von Neumann. (See [1], Chapter 5, Sectionโ€‰3, Theoremโ€‰3.24, page 275.) But is closed (being self-adjoint by standard functional analysis) and bounded (being a restriction of the bounded operator ) and so is a globally defined, bounded operator by Proposition 2.2. Moreover, on . So, is a globally defined, bounded operator with on , since the operator , is globally defined, bounded and its square is .
Next the polar decomposition theorem tells us that on , where is partial isometry from to . But and so is globally defined and equal by (3.26) to the composition of two bounded operators on . Therefore is also bounded. Since is closed, we have . This displays as the adjoint of the globally defined, bounded operator . We then conclude that is a globally defined, bounded operator as well.
Now by a โ€œone-pageโ€ argument, we have shown that , and so, is one-to-one. And by a โ€œone-lineโ€ proof, we have seen that is dense, and so is onto. The two preceding assertions about follow from the polar decomposition Theorem 2.3. We conclude that is a unitary isomorphism.
We now write (3.26) equivalently as for all and all . Now we apply to both sides, recalling that there is an implicit analytic continuation on the left side which cancels with , to get for all and all . So, we have the operator equation where each side is a bounded operator from to itself. Also all of the operators in this equation are bounded. This then implies that on . Of course, is not a bounded operator. However, its domain is dense in . (Proof: using the Dunkl transform (see [16โ€“19]), one shows that the bounded operator is unitarily equivalent to multiplication by acting on , where is the variable in . But the range of multiplication by clearly contains and so is dense by a standard argument in analysis.) Moreover, we also have on as one sees by applying both sides to an arbitrary element , where , and by using the semigroup property. This in turn gives us on the dense domain . Since both and are globally defined, bounded operators that are equal on a dense domain, it follows that on . So for all and all , we obtain Next, we write out the left side as follows: So for all and all , we find that and so . Using that is injective (i.e., uniqueness of analytic continuation), we finally arrive at the desired identity, , and therefore, is a unitary isomorphism as we wanted to prove.

During the proof of the previous theorem we proved the statement made earlier that is bounded. We now state this result separately and amplify on it.

Theorem 3.7. The operator is bounded and has operator norm . Also the operator is bounded with operator norm .

Proof. In this proof we denote all operator norms by . We already have shown that is a self-adjoint, bounded operator acting on and that and are globally defined, bounded operators. We take in the following, getting We also compute directly since and . The result now follows.

4. Versions , , and

Now we will apply the method indicated after the statement of the polar decomposition Theorem 2.3 in order to show that the -version of the Segal-Bargmann transform can be obtained by a polar decomposition which is related to the polar decomposition (namely, the restriction principle) used to obtain the -version. So, we are looking for two unitary isomorphisms, and , making the following diagram commute: 365085.fig.001(4.1) Then we can use these two unitaries to change the polar decomposition which gave us into a polar decomposition giving . Of course, the very existence of such a pair, and , already would prove that is a unitary isomorphism.

We use a known relation between the and -versions in order to start. The rest of the construction then follows in a systematic, algorithmic manner. The relation between these two versions that we use starts from this identity for the integral kernels for all and all . (See [6, Theoremโ€‰3.5]). Now we translate this relation into a relation between the integral transforms themselves. From the defining equation (2.5) we have the scaling relation for . By taking and replacing with in this, we have Next we replace with to obtain To understand the integral kernel , we take and evaluate as follows: where we used , the scaling property , and the definition of the dilation operator . Next, by multiplying both sides of (4.4) by and then integrating with respect to , we get A crucial point here is that the factor does not depend on the variable of integration and so factors out in front of of the integral. Equivalently, which itself is equivalent to the operator equation where denotes the operator of multiplication by the function . Now we solve the last equation for getting

Next, we want the operator to be sandwiched between two unitary operators, and so it is not initially clear how to divide up the factors of 2 in (4.9) to get multiples of and of that are unitaries. But by Lemma 2.1 we know that is a unitary isomorphism. The desired domain and the desired codomain of this unitary operator are determined by diagram (4.1). So it remains to show what is happening with the operator . According to the diagram (4.1) this should be the unitary isomorphism indicated there.

Therefore we would like to take and calculate the norms and and then show they are equal. But we do not have closed formulas for these norms for general elements in these reproducing kernel Hilbert spaces. However, it suffices to consider the case when , where is arbitrary. See (3.4). In spite of the quantity of details, this does work out in an algorithmic manner.

Nevertheless, purely for the sake of simplicity, we prefer to give a shorter proof by relating with known entities. We note first that is a unitary isomorphism. (See (3.6) and the subsequent discussion.) And second from a result in [6] we have that is also a unitary isomorphism. So the composition is again a unitary isomorphism. For any we use (3.6) to calculate this composition, giving for all that which in turn implies the operator equation It follows that is a unitary isomorphism.

We are now ready to apply the method discussed in the remarks just after the polar decomposition Theorem 2.3. Using the notation established there, we let be defined as Also, we already defined . So we have shown above that and are unitary isomorphisms and that diagram (4.1) commutes.

Of course, we have from (4.9) and the subsequent results that is a unitary isomorphism, since it is the composition of three unitary isomorphisms. We now want to see how arises explicitly from the corresponding polar decomposition (which, according to our definition, will turn out not to be a restriction principle) and how this polar decomposition relates to the unmotivated definition of a โ€œrestrictionโ€ operator in [12]. So, continuing with the notation established earlier, we have that arises in the polar decomposition ; that is, , where . (Recall that we have shown that and are globally defined, bounded operators.) It follows that , and therefore, arises from the restriction principle according to our definition exactly when is the restriction operator ; namely, . We know that and so . But from (3.6), we immediately have So for we have for that Then since is simply restriction, we obtain for all that Finally, applying yields for all and which is not the restriction operator. Consequently, this polar decomposition is not a restriction principle. However, notice that the operator is globally defined and bounded, since is globally defined and bounded. This fact is not so obvious by merely inspecting the right side of (4.18).

The operator in (4.18) does not compare very well at first sight with the โ€œrestriction operatorโ€ defined in [12, page 298]. But this discrepancy is easily understood. In [6, Corollaryโ€‰3.1] we give the unitary equivalence between and the โ€œgeneralized Segal-Bargmann transformโ€ defined in [12]. (N.B. only the case is considered in [12].) Using this we can conjugate the polar decomposition used above in order to obtain to get an operator, say , whose polar decomposition gives us . We note that is globally defined and bounded, since it is unitarily equivalent to . The adjoint of (which should be the restriction operator) for all and turns out to be which is not a restriction operator according to our definition. Except for the positive multiplicative constant , this agrees with the โ€œrestriction operatorโ€ given in [12]. But for any closed, densely defined operator and any , the polar decompositions of and give the same partial isometry. And this explains how the unmotivated โ€œrestriction operatorโ€ used in [12] arises in a natural manner in our presentation.

We wish to note that formula (4.18) was forced on us by our method, once we had established that the unitary operators and change the transform into . (cp. diagram (4.1).) And these two unitaries arose in a natural, motivated way directly from an identity that relates the kernel functions of these transforms. So the -version arises by applying polar decomposition to a particular operator. When one thinks of it this way, this is a rather unimpressive result. Actually, every unitary operator between two Hilbert spaces can be realized via a polar decomposition. And any closed, densely defined operator which satisfies two additional hypotheses (injectivity and dense range) gives us a unitary operator in its polar decomposition.

Moreover, we could have used another pair of unitary isomorphisms, say and in place of and , to change into , using a diagram analogous to (4.1). Then arises from the polar decomposition that comes from the restriction principle used to produce . However, this polar decomposition in general will not be a restriction principle. (e.g., the codomains of and need not even be function spaces.) Actually, any unitary isomorphism between separable, complex Hilbert spaces of infinite dimension can arise this way by an appropriate, but far from unique, choice of the two unitaries and . So in general it would be misleading to dub with a name that indicates that it forms a part of Segal-Bargmann analysis.

However, the transform does arise naturally and uniquely from the heat kernel method as a part of Segal-Bargmann analysis. (See [6].) So it is reasonable to ask (and answer, as we have done in this section) how the restriction principle for gives us a polar decomposition of . On the other hand, we have not been able to find in [12] a satisfactory, explicit justification for considering the transform defined there as a part of Segal-Bargmann analysis. For example, Remarkโ€‰4.3 ([12, page 301]) only indicates what happens when (in our notation). In our opinion this is very far from justifying the terminology โ€œSegal-Bargmannโ€ for the case of general .

One point of this section is to show where the unmotivated exponential factor comes from in the definition of the โ€œrestriction operatorโ€ in [12]. It is truly a deus ex machina in [12]. Here it flows out naturally from an analysis based on the -version. The second point of this section is to provide contrast with the method used to define the -version in the last section. While that was also a polar decomposition, it was a particular, uniquely defined special case, namely the restriction principle. The worst that could happen with an analysis based on the restriction principle is that the technical details do not work out and therefore no unitary isomorphism at all is produced. In short, the result of the method is unique but may not exist.

As for the remaining two versions of the Segal-Bargmann, the Version (resp., ) is defined by a unitary transformation (a change of measure) on the domain space starting with the Version (resp., ). (See [6] for details about Version . Version is related to Version analogously.) So, the restriction principle for the -version implies that these remaining two versions can also be obtained from the polar decomposition of an explicitly defined operator. The details are left to the interested reader. We do wish to comment that these polar decompositions are not restriction principles. The brevity of our discussion in this paragraph is not meant to indicate that these versions are less important than the -version. On the contrary, we think that the three versions , , and have the same relative relation to the truly important and logically central -version.

5. Concluding Remarks

Our confusion over the role in [12] of their โ€œrestriction principleโ€ in the Segal-Bargmann analysis motivated our study of this topic. The upshot is our discovery of the central role of the restriction principle in the -version of the Segal-Bargmann analysis associated to a finite Coxeter group. We wish to underscore that only the -version of the Segal-Bargmann analysis is considered in [12]. This can be clearly seen in the reproducing kernel for the space of holomorphic functions in [12], which is therefore the -version space. Also the โ€œgeneralized Segal-Bargmann transformโ€ in [12] has an integral kernel which is not the analytically continued heat kernel (as in the -version), but rather something that corresponds to our uniquely defined -version (modulo normalization and dilation). There is no mention in [12] of the -version nor even of the existence of other versions of the Segal-Bargmann analysis.

In summary, we think that this paper shows that the restriction principle and the -version (and not any other version) of Segal-Bargmann analysis are naturally and closely related with each other. So this is a new way for understanding how the -version in general is the most fundamental version of the Segal-Bargmann analysis.

As for future endeavors, we note that we have studied only the case and so it might be interesting to understand what happens when we drop or weaken that condition.

Acknowledgments

The author thanks Greg Stavroudis for a most hospitable venue for connecting to the Internet as well as for his excellent coffee, both being critical elements for doing the research for this paper. Also he warmly thanks Jean-Pierre Gazeau for being his academic host at the Universitรฉ Paris Diderot Paris 7 during the spring of 2011 when this paper was finished. This research is partially supported by CONACYT (Mexico) project 49187.