This is an AMSLaTeX document, v1.1. BODY \documentstyle[verbatim,righttag]{amsart} \numberwithin{equation}{section} \newcommand{\ms}{\medskip} \newcommand{\bs}{\bigskip} \newcommand{\noi}{\noindent} \newcommand{\ra}{\rightarrow} \newcommand{\bea}{\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}} \newcommand{\gr}{Groenewold} \newcommand{\vh}{Van Hove} \newcommand{\vn}{von Neumann} \newcommand{\sh}{spherical harmonic} \newcommand{\shs}{spherical harmonics} \newcommand{\cg}{Clebsch-Gordan coefficient} \newcommand{\ci}{S_i} \newcommand{\cii}{S_i\,\!^2} \newcommand{\ciii}{S_i\,\!^3} \newcommand{\cj}{S_\ell} \newcommand{\ck}{S_k} \newcommand{\cij}{S_iS_\ell} \newcommand{\cx}{S_1} \newcommand{\cxx}{S_1\,\!^2} \newcommand{\cxxx}{S_1\,\!^3} \newcommand{\cy}{S_2} \newcommand{\cyy}{S_2\,\!^2} \newcommand{\cyyy}{S_2\,\!^3} \newcommand{\cz}{S_3} \newcommand{\czz}{S_3\,\!^2} \newcommand{\czzz}{S_3\,\!^3} \newcommand{\q}{\cal Q} \newcommand{\qx}{\cal Q(S_1)} \newcommand{\qy}{\cal Q(S_2)} \newcommand{\qz}{\cal Q(S_3)} \newcommand{\qi}{\cal Q(S_i)} \newcommand{\qii}{\cal Q\big(S_i\,\!^2\big)} \newcommand{\qiii}{\cal Q\big(S_i\,\!^3\big)} \newcommand{\qj}{\cal Q(S_\ell)} \newcommand{\qk}{\cal Q(S_k)} \newcommand{\p}{\cal P} \newcommand{\h}{\cal H} \newcommand{\oo}{\cal O} \newcommand{\ld}{\lambda} \newcommand{\qn}{\Bbb Q_{n}^{|n|}} \newcommand{\sm}{\Bbb S_{m}^{|m|}} \newtheorem{thm}{Theorem} \newtheorem{lem}{Lemma} \newtheorem{prop}[thm]{Proposition} \newtheorem{cor}{Lemma}[section] \newtheorem{guess}[thm]{Conjecture} \newtheorem{nnlem}{Lemma} \newtheorem{axiom}{Lemma}[section] \newtheorem{axiomA}{Lemma}[section] \renewcommand{\thennlem}{} \theoremstyle{definition} \newtheorem{defn}{Definition} \newsymbol\ltimes 226E %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{document} \title{A Groenewold\,-Van Hove Theorem for $S^2$} \author{Mark J. Gotay} \address{Department of Mathematics \\University of Hawai`i \\2565 The Mall \\ Honolulu, HI 96822 USA} \email{gotay@@math.hawaii.edu} \author{Hendrik Grundling} \address{Department of Pure Mathematics \\University of New South Wales \\P. O. Box 1 \\ Kensington, NSW 2033 Australia} \email{hendrik@@solution.maths.unsw.edu.au} \author{C. A. Hurst} \address{Department of Physics and Mathematical Physics \\University of Adelaide \\ G. P. O. Box 498 \\ Adelaide, SA 5001 Australia} \email{ahurst@@physics.adelaide.edu.au} \date{February 18, 1995} \maketitle \begin{abstract} We prove that there does not exist a nontrivial quantization of the Poisson algebra of the symplectic manifold $S^2$ which is irreducible on the subalgebra generated by the components $\{S_1,S_2,S_3\}$ of the spin vector. We also show that there does not exist such a quantization of the Poisson subalgebra $\p$ consisting of polynomials in $\{S_1,S_2,S_3\}$. Furthermore, we show that the maximal Poisson subalgebra of $\cal P$ containing $\{1,S_1,S_2,S_3\}$ that can be so quantized is just that generated by $\{1,S_1,S_2,S_3\}$. \end{abstract} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Introduction} In a striking paper, Groenewold \cite{gr} showed that one cannot ``consistently'' quantize all polynomials in the classical positions $q^i$ and momenta $p_i$ on $\Bbb R^{2n}.$ Subsequently Van Hove \cite{vh1,vh2} refined and extended \gr's result, in effect showing that there does not exist a quantization functor which is consistent with the Schr\"odinger quantization of $\Bbb R^{2n}$. (For discussions of \gr's and \vh's work and related results, see \cite{a-m,c,f,go1,g-s,j} and references contained therein.) However, these theorems rely heavily on certain properties of $\Bbb R^{2n}$, and so it is not clear whether they can be generalized. Naturally, one {\em expects} similar ``no\,-go'' theorems to hold in a wide range of situations, but we are not aware of any previous results along these lines. In this paper we prove a Groenewold-Van Hove theorem for the symplectic manifold $S^2$. Our proof is similar to Groenewold's for $\Bbb R^{2n}$, although it differs from his in several important respects and is technically more complicated. On the other hand, as $S^2$ is compact there are no problems with the completeness of the flows generated by the classical observables, and so Van Hove's modification of \gr's theorem is unnecessary in this instance. To set the stage, let $(M,\omega)$ be a symplectic manifold. We are interested in quantizing the Poisson algebra $C^{\infty}(M)$ of smooth real-valued functions on $M$, or at least some subalgebra $\cal C$ of it, in the following sense. \begin{defn} A {\em quantization} of $\cal C$ is a linear map $\cal Q$ from $\cal C$ to an algebra of self-adjoint operators\footnote{Technical difficulties with unbounded operators will be ignored, as they are not important for what follows.} on a Hilbert space such that \ms \begin{enumerate}\begin{enumerate} \item[({\em i\/})] $\cal Q\big(\{f,g\}\big) = -\mbox{i}\big[\cal Q(f),\cal Q(g)\big],$ \end{enumerate}\end{enumerate} \noindent where $\{\,,\,\}$ denotes the Poisson bracket and $[\;,\,]$ the commutator. If $\cal C$ contains the constant function 1, then we also demand \ms \begin{enumerate}\begin{enumerate} \item[({\em ii\/})] $\cal Q(1) = I.$ \end{enumerate}\end{enumerate} \end{defn} As is well known, it is necessary to supplement these conditions for the quantization to be physically meaningful. To this end, one often requires that a certain subalgebra $\cal B$ of observables be represented irreducibly. Exactly which observables should be taken as ``basic'' in this regard depends upon the particular example at hand; one typically uses the components of a momentum map associated to a (transitive) Lie symmetry group. For $\Bbb R^{2n}$ the relevant group is the Heisenberg group \cite{f,g-s} and $\cal B = {\text {span}}\{1,q^i,p_i\,|\, i=1,\ldots,n\}$. In the case of $S^2$ the appropriate group is SU$(2) \times \Bbb R$ whence the basic observables are span$\{1,S_1,S_2,S_3\}$, the $S_i$ being the components of the spin angular momentum. Alternatively, one could require the {\em strong von Neumann rule} \cite{vn} \[\cal Q\big(k(f)\big) = k\big(\cal Q(f)\big)\] \noindent to hold for all polynomials $k$ and all $f \in \cal C$ such that $k(f) \in \cal C$. Usually it is necessary to weaken this condition \cite{f}, insisting only that it hold for a certain subclass $\cal B$ of observables $f$ and certain polynomials $k$. We refer to this simply as a ``von Neumann rule.'' In the case of $\Bbb R^{2n}$, the von Neumann rule as applied to the $q^i$ and $p_i$ with $k(x) = x^2$ is actually implied by the irreducibility of the $\cal Q(q^i)$ and $\cal Q(p_i)$ \cite{c}. But the corresponding statement is not quite true for $S^2$, as we will see. We refer the reader to \cite{f} for further discussion of \vn\ rules. In our view, imposing an irreducibility condition on the quantization map $\cal Q$ seems more compelling physically and pleasing aesthetically than requiring $\cal Q$ to satisfy a \vn\ rule. With this as well as the observations above in mind, we make \begin{defn} An {\em admissible quantization} of the pair $(\cal C,\cal B)$ is a quantization of $\cal C$ which is irreducible on $\cal B$, where $\cal C \subset C^\infty(M)$ is a given Poisson algebra, and $\cal B \subset\cal C$ a given subalgebra. \end{defn} The results of \gr\ and Van Hove may then be interpreted as showing that there does not exist an admissible quantization of the pair $\big(C^{\infty}(\Bbb R^{2n}),\,\text{span}\{1,q^i,p_i\,| \linebreak \, i=1,\ldots,n\} \big)$ nor, for that matter, of the subalgebra of all polynomials in the $q^i$ and $p_i$. We will prove here that there likewise does not exist an admissible quantization of the pair $\big(C^{\infty}(S^2),\,\text{span}\{1,S_1,S_2,S_3\}\big)$ nor, for that matter, of the subalgebra of all polynomials in the components $S_i$ of the spin vector. \ms %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{No\,-Go Theorems} Consider a sphere $S^2$ of radius $s > 0$. We view this sphere as the ``internal'' phase space of a massive particle with spin $s$ and realize it as the subset of $\Bbb R^3$ given by \begin{equation} S_1\,\!^2 + S_2\,\!^2 + S_3\,\!^2 = s^2, \label{cs2} \end{equation} \noindent where $\bold S = (S_1,S_2,S_3)$ is the spin angular momentum. The symplectic form is\footnote{Note that $\omega$ is $1/s$ times the area form $d{\boldsymbol \sigma}$ on $S^2$. It is the symplectic form on $S^2$ viewed as a coadjoint orbit of SU(2).} \[\omega = \frac{1}{2s^2}\sum_{i,j,k=1}^3\epsilon_{ijk}S_i\,dS_j \wedge dS_k\] \noindent with corresponding Poisson bracket \begin{equation} \{f,g\} = \sum_{i,j,k=1}^3\epsilon_{ijk}S_i\frac{\partial f}{\partial S_j} \frac{\partial g}{\partial S_k} \label{pb} \end{equation} \noindent for $f,g \in C^{\infty}(S^2).$ We have the relations $\{S_i,S_j\} = \sum_{k=1}^3 \epsilon_{ijk}S_k.$ The group \,SU(2)\, acts transitively on $S^2$ with momentum map $\bold S= (S_1,S_2,S_3)$, i.e., the pair $\big(S^2,{\text {SU(2)}}\big)$ is an ``elementary system'' in the sense of \cite{w}. Thus it is natural to require that quantization provide an irreducible representation of \,SU(2). In terms of observables, quantization should produce a representation which is irreducible when restricted to the subalgebra generated by $\{\cx,\cy,\cz\}$. However, this subalgebra does not include the constants. To remedy this, we consider instead the central extension ${\text {SU(2)}} \times \Bbb R$ of \,SU(2)\, by $\Bbb R$ with momentum map $(1,S_1,S_2,S_3)$, and take the subalgebra generated by these observables to be the basic set $\cal B$ in the sense of the Introduction. Let $\cal P$ denote the Poisson algebra of polynomials in the components $S_1,S_2,S_3$ of the spin vector $\bold S$ modulo the relation \eqref{cs2}. (This means we are restricting polynomials as functions on $\Bbb R^3$ to $S^2$.) We shall refer to an equivalence class $p \in \cal P$ as a ``polynomial'' and take its degree to be the minimum of the degrees of its polynomial representatives. We denote by $\cal P^k$ the subspace of polynomials of degree at most $k$. In particular, $\cal P^1$ is just the Poisson subalgebra generated by $\{1,S_1,S_2,S_3\}$. When equipped with the $L^2$ inner product given by integration over $S^2$, the vector space $\p^k$ becomes a real Hilbert space which admits the orthogonal direct sum decomposition $\cal P^k = \bigoplus_{l=0}^k \cal H_l$, where $\cal H_l$ is the vector space of spherical harmonics of degree $l$ (i.e., the restrictions to $S^2$ of homogeneous harmonic polynomials of degree $l$ on $\Bbb R^3$ \cite{a-b-r}). Note that $\h_1$ is the Poisson subalgebra generated by $\{S_1,S_2,S_3\}$. The collection of \shs\ $\big\{Y_l^m,\;l = 0,1,\ldots,k,\;m = -l,-l+1,\ldots,l\big\}$ forms the standard (complex) orthogonal basis for the complexification $\p^k_{\Bbb C}$: \begin{equation*} \int_{S^2}Y_{l_1}^{m_1*}Y_{l_2}^{m_2}\,d\sigma = s^2\delta_{l_1l_2}\delta_{m_1m_2}. \end{equation*} \noi Thus if $p \in \cal P^k_{\Bbb C}$, we have the harmonic decomposition \begin{equation} p = p_k + p_{k-1} + \cdots + p_0, \label{hd} \end{equation} \noi where $p_l \in (\h_l)_{\Bbb C}$ is given by \begin{equation} p_l = \frac{1}{s^2}\sum_{m=-l}^l\left(\int_{S^2}Y_l^{m*}p\,d\sigma\right)Y_l^m. \label{hdy} \end{equation} It is well known that O(3) acts orthogonally on $\cal P$, and that this action is irreducible on each $\h_l$ where it is the standard real (orbital) angular momentum $l$ representation. The corresponding infinitesimal generators on $\h_l$ are $L_i = \{S_i,\cdot\}$. If we identify o(3) and $\h_1$ as Lie algebras, it follows that the ``adjoint'' action of $\h_1$ on $\h_l$ given by $S_i \mapsto \{S_i,\cdot\}$ is irreducible as well. \ms Now suppose $\cal Q$ is a quantization of $\cal P$, so that \begin{equation} \big[\cal Q(S_i),\cal Q(S_j)\big] = \mbox{i}\sum_{k=1}^3\epsilon_{ijk}\cal Q(S_k) \label{com} \end{equation} \noi and \vspace{-1ex} \begin{equation} \cal Q\big(\bold S^2\big) = s^2I. \label{s2} \end{equation} \vspace{1ex} \noindent If in addition $\q$ is admissible on $(\cal P,\,\cal P^1)$, then $\p^1$ must be irreducibly represented. Since as a Lie algebra $\cal P^1$ is isomorphic to \,su$(2) \times \Bbb R$, its irreducible representations are all finite-dimensional.\footnote {Every irreducible representation of su$(2) \times \Bbb R$ by (essentially) self-adjoint operators on an invariant dense domain in a Hilbert space can be integrated to a continuous irreducible representation of $\text {SU(2)} \times \Bbb R$ \cite[\S 11.10.7.3]{b-r}. But it is well known that every such representation of this group is finite-dimensional.} These are just the usual (spin angular momentum) representations labeled by $j = 0,\frac{1}{2},1,\ldots$, where\footnote{In what follows we use standard quantum mechanical notation, cf. \cite{m}.} \begin{equation} \sum_{i=1}^3\cal Q(S_i)^2 = j(j+1)I. \label{j2} \end{equation} \noindent The Hilbert space corresponding to the quantum number $j$ has dimension $2j+1$, with the standard orthonormal basis $\big\{|\,j, m\rangle,\: m = -j,-j+1,\ldots,j\big\}$ consisting of eigenvectors of $\cal Q(S_3)$. We regard the representation defined by $j=0$ as trivial, in that it corresponds to quantum spin 0. The following result shows that admissibility implies a weak type of \vn\ rule on $\cal P^1$. \begin{prop} If $\cal Q$ is an admissible quantization of $(\cal P,\,\cal P^1)$, then \begin{equation} \cal Q\big(S_i\,\!^2\big) = a \cal Q(S_i)^2 + cI \label{ii} \end{equation} \noindent for $i = 1,2,3$, where $a$ and $c$ are real constants with $a^2 + c^2 \neq 0$. Furthermore, for $i \neq \ell$, \begin{equation} \cal Q(S_iS_\ell) = \frac{a}{2}\big(\cal Q(S_i)\cal Q(S_\ell)+\cal Q(S_\ell)\cal Q(S_i)\big). \label{pr} \end{equation} \label{qvn} \end{prop} \vspace{-2ex} We have placed the proof, which is rather long and technical, in Appendix A so as not to interrupt the exposition. Observe that on summing \eqref{ii} over $i$, we get $s^2=aj(j+1)+3c$ which fixes the constant $c$ in terms of $s,\:j$ and $a.$ {}From these relations the main result now follows. \begin{thm}[No\,-Go Theorem] There does not exist a nontrivial admissible quantization of $(\cal P,\cal P^1)$. \label{ng2} \end{thm} \begin{pf} Suppose there did exist an admissible quantization of $(\cal P,\cal P^1)$; we shall show that for $j>0$ this leads to a contradiction. First observe that we have the classical equality \[s^2S_3 = \{S_1\,\!^2 - S_2\,\!^2,S_1S_2\} - \{S_2S_3,S_3S_1\}.\] \noindent Quantizing this, a calculation using \eqref{ii}, \eqref{pr}, \eqref{com} and \eqref{j2} gives \[s^2\qz = a^2\bigg (j(j+1) - \frac {3}{4} \bigg ) \cal Q(S_3).\] \noindent Thus either $\cal Q(S_3) = 0$, whence $j=0$, or $j > 0$, in which case \begin{equation} s^2 = a^2\bigg (j(j+1) - \frac {3}{4} \bigg ). \label{qc1} \end{equation} \noindent Observe that when $j = \frac{1}{2}$, this implies that $s = 0$, which is impossible. Henceforth take $j>\frac{1}{2}.$ Next we quantize the relation \[2s^2\cy\cz = \big\{\cy\,\!^2,\{\cx\cy,\cx\cz\}\big\} - \frac{3}{4}\big\{\cx\,\!^2,\{\cx\,\!^2,\cy\cz\}\big\}.\] \noi Using \eqref{pr}, \eqref{ii}, \eqref{com} and \eqref{j2}, the l.h.s. becomes \[as^2\big(\q(S_2)\q(S_3)+\q(S_3)\q(S_2)\big) = as^2 \big (2\qy\qz - \mbox{i}\qx\big )\] \noi while the r.h.s. reduces to \[a^3\bigg (j(j+1) - \frac{9}{4}\bigg) \big(2\qy\qz - \mbox{i}\qx\big ).\] \noindent Since for $j > \frac{1}{2}$ the matrix element \[\big{\langle}\, j, j\,|\,2\qy\qz - \mbox{i}\qx|\,j, j\!-\!1\big{\rangle} = \text {i} \bigg (\frac{1}{2} - j\bigg )\sqrt{2j}\] \noindent is nonzero, it follows that \[as^2 = a^3\bigg (j(j+1) - \frac{9}{4} \bigg ).\] \noindent If $a=0$ \eqref{qc1} yields $s=0$, whereas if $a \neq 0$ this conflicts with \eqref{qc1}. Thus we have derived contradictions provided $j > 0$. Since $j=0$ is the trivial representation, the theorem is proven. \end{pf} This contradiction shows that the quantization goes awry on the level of quadratic polynomials. On the other hand, there are many admissible quantizations of the Poisson subalgebra $\cal P^1$ of all polynomials of degree at most one, viz. the irreducible representations $\q$ of \,su$(2) \times \Bbb R$ with $\q(1) = I$. Thus it is of interest to determine the largest subalgebra of $\cal P$ containing $\{1,S_1,S_2,S_3\}$ that can be so quantized. We will now show that this largest subalgebra is just $\cal P^1$ itself. Unfortunately, this is not entirely straightforward, since $\cal P^1$ is not a maximal Poisson subalgebra of $\cal P$; indeed, if $\cal O$ denotes the Poisson subalgebra of odd polynomials (i.e., polynomials all of whose terms are of odd degree), then $\cal P^1$ is contained in $\tilde{\cal O} = \cal O \oplus \Bbb R$. To prove the result, we proceed in two stages. First we show that $\cal P^1$ is maximal in $\tilde {\cal O}$, and then prove a no\,-go theorem for $\tilde {\cal O}$. \begin{prop} $\cal P^1$ is maximal in $\tilde{\cal O}$. \label{max} \end{prop} \begin{pf} Actually, the constants are unimportant, and it will suffice to prove that the Poisson subalgebra $\h_1$ generated by $\{S_1,S_2,S_3\}$ is maximal in $\cal O$. Set $\cal O^l = \cal O \cap \p^l$, where henceforth $l$ is odd. For $k$ odd it is clear from \eqref{pb} and \eqref{cs2} that $\{\h_k,\h_l\} \subset \oo^{k+l-1}$. Let $\cal R$ be the Poisson algebra generated by a single polynomial $r \in \cal O^l$ of degree $l>1$ together with $\h_1$. Evidently $\cal R \subset \cal O$; we must show that $\cal O \subset \cal R$. We will accomplish this in a series of lemmas. \begin{lem} If in its harmonic decomposition an element of $\cal R$ has a nonzero component in $\h_k$, then $\h_k\subset\cal R$. \label{inhomo} \end{lem} \begin{pf} Let $\cal R' \subset \cal R$ be the span of all elements of the form \[\big\{h_n,\ldots\big\{h_2,\{h_1,r\}\big\}\ldots\big\}\] \noi for $h_i \in \h_1$ and $n \in \Bbb N.$ Then $\cal R'$ is an o(3)-invariant subspace of $\oo^l \subset \p$. Since the representation of o(3) on $\cal P$ is completely reducible, $\cal R'$ must be the direct sum of certain $\h_k$ with $k \leq l.$ Consequently, if when harmonically decomposed an element of $\cal R'$ has a nonzero component in some $\h_k$, then $\h_k \subset \cal R' \subset \cal R$. \renewcommand{\qedsymbol}{$\bigtriangledown$} \end{pf} Now by assumption $r_l \neq 0$ in $\h_l$ and hence $\h_l \subset \cal R.$ Then $\{\h_l,\h_l\} \subset \oo^{2l-1} \cap \cal R$. We will use this fact to show that $\oo^{2l-1} \subset \cal R$. The proof devolves upon an explicit computation of the harmonic decomposition of $\{Y_l^m,Y_l^n\}$. \begin{lem} For each $j$ in the range $0 < j \leq 2l$, we have \begin{equation*} \{Y_l^{l-j},Y_l^l\} = \sum_{k=1}^{l}y_{2k-1}(l-j,l)Y_{2k-1}^{2l-j}. \end{equation*} In particular, when $j = 1$ the top coefficients $y_{2l-1}(l-1,l)$ are nonzero. Furthermore, provided $l\geq 5$, $k \geq \frac{l-1}{2}$ and $k > l - \frac{j+1}{2}$, the coefficients $y_{2k-1}(l-j,l)$ are nonzero. \label{main} \end{lem} Since the proof requires an extended calculation, we defer it until Appendix B. \begin{lem} $\oo^{2l-1} \subset \cal R$. \end{lem} \begin{pf} Decompose $Y_l^m = R_l^m+\text {i}I_l^m$ into real and imaginary parts, with $R_l^m,\;I_l^m\in\cal H_l$. So in $\cal P_{\Bbb C}$ we observe that both $\Re\{Y_l^m,\,Y_l^n\}=\{R_l^m,R_l^n\}- \{I_l^m,I_l^n\}$ and $\Im\{Y_l^m,Y_l^n\}=\{R_l^m,I_l^n\}+\{I_l^m,R_l^n\}$ belong to $\{\cal{H}_l,\,\cal{H}_l\}$. Thus if the harmonic decomposition of $\{Y_l^m,Y_l^n\}$ has a nonzero $k^{\text{th}}$ component, then either its real or imaginary part must be nonzero which allows us to conclude that $\{\cal{H}_l,\cal{H}_l\}$ contains an element with nonzero component in $\cal{H}_k$. If $l=3$, then $\h_3 \subset \cal R$. Now consider the bracket $\{Y_3^{2},\,Y_3^3\}$. By Lemma \ref{main} with $j=1$ it has a nonzero $5^{\text{th}}$ component, so by the preceding and Lemma \ref{inhomo} it follows that $\h_5 \subset \cal R$. Since by definition $\h_1\subset\cal R$, we then have $\oo^5 = \h_1 \oplus \h_3 \oplus \h_5 \subset \cal R.$ If $l\geq 5$, we consider $\{Y_l^{-2},\,Y_l^l\}\in\cal R_{\Bbb C}$. By Lemma \ref{main} with $j=l+2$, the preceding and Lemma \ref{inhomo} we conclude that \[\h_{l-2}\oplus\h_l\oplus\cdots\oplus\h_{2l-1}\subset\cal R.\] \noi Hence $\h_{l-2}\subset\cal R$, so by the same argument applied to $\{Y_{l-2}^{-2},Y_{l-2}^{l-2}\}$ we get that $\h_{l-4}\subset\cal R$. Continuing in this way we obtain $\h_{l-2n}\subset\cal R$ for all $n$ with $l-2n\geq 3$. In particular, taking $n = \frac{l-3}{2}$, we get $\h_3\subset\cal R$. But we have already remarked that $\h_1 \subset \cal R,$ so the lemma is proven. \renewcommand{\qedsymbol}{$\bigtriangledown$} \end{pf} Thus $\cal R$ must contain all odd polynomials of degree at most $2l-1$. To obtain higher degree polynomials, we need only bracket $\h_{2l-1} \subset \cal R$ with itself and apply the argument above to conclude that $\cal O^{4l-3} \subset \cal R$. Continuing in this manner, we have finally that $\cal O \subset \cal R$, and this proves the proposition. \end{pf} Our strategy in proving the no\,-go theorem for $\tilde {\oo}$ is the same as for $\p$. To begin, we use admissibility to obtain a weak version of a cubic \vn\ rule. \begin{prop} If $\q$ is an admissible quantization of $(\tilde{\cal O},\p^1)$, then \begin{equation} \q\big(S_i\,\!^3\big) = a \qi ^3 + c \qi \label{vn3} \end{equation} \noi for $i=1,2,3,$ where $a$ and $c$ are real constants. Furthermore, when $i \neq \ell$, \begin{equation} \q(\ci\cj\ci) = a\qi\qj\qi + \frac{1}{3}(a+c)\qj. \label{iji} \end{equation} \noi Finally, \begin{equation} \q(\cx\cy\cz) = a\qx\qy\qz + \frac{a}{2 \mathrm{i}}\big (\qx^2 - \qy^2 + \qz^2\big). \label{123} \end{equation} \label{cvn} \end{prop} \vspace{-1.5ex} Again the proof is placed in Appendix A. We derive some consequences of these results. Multiplying \eqref{cs2} through by $S_\ell$ and quantizing gives \[\sum_{i=1}^3 \cal Q(S_iS_\ell S_i) = s^2 \cal Q(S_\ell).\] \noindent Applying \eqref{iji} and \eqref{vn3}, this in turn becomes \begin{equation} a\sum_{i=1}^3 \cal Q(S_i)\cal Q(S_\ell)\cal Q(S_i) = \bigg(s^2 - \frac{2a}{3} - \frac{5c}{3}\bigg) \cal Q(S_\ell). \label{s3s} \end{equation} \noi We also find, by rearranging the factors in $\sum_{i=1}^3 \cal Q(S_i)\cal Q(S_\ell)\cal Q(S_i)$, that \begin{equation} \sum_{i=1}^3 \cal Q(S_i)\cal Q(S_\ell)\cal Q(S_i) = \big (j(j+1) -1\big )\qj. \label{s3j} \end{equation} \noi A comparison of \eqref{s3s} and \eqref{s3j} yields \begin{equation} s^2 = a\bigg(j(j+1) -\frac{1}{3}\bigg ) + \frac{5c}{3} \label{js} \end{equation} \noi provided $j>0.$ \begin{thm} There does not exist a nontrivial admissible quantization of $(\tilde{\cal O},\p^1)$. \label{ng3} \end{thm} \vspace{-3ex} \begin{pf} Suppose $\q$ were an admissible quantization of $(\tilde{\cal O},\p^1)$; we will show that then $j=0.$ Consider the Poisson bracket relation \begin{eqnarray*} 3s^4S_3 &\!\!\! =\!\!\! & 4\{S_1\,^3,S_2S_3\,^2\} - 4\{S_2\,^3,S_3\,^2S_1\} + \{S_2\,^2S_1,S_2\,^3\} \\ & & \rule{0ex}{3ex} - \,\{S_2S_1\,^2,S_1\,^3\} - 6\{S_2\,^3,S_1\,^3\} -3\{S_2S_3\,^2,S_3\,^2S_1\}. \end{eqnarray*} \noindent Upon quantizing, an enormous calculation using \eqref{vn3}, \eqref{iji}, \eqref{com} and \eqref{s3j} gives\footnote{This calculation was done using the {\sl Mathematica} package {\sl NCAlgebra} \cite{h-m}.} \begin{eqnarray*} 3s^4 \cal Q(S_3) & = & \bigg(3a^2j^4 + 6a^2j^3 + 14acj^2 + 8a^2j^2 \nonumber \\ & & \mbox{} + 14acj + 5a^2j + \frac{29c^2}{3} - 14 \frac{ac}{3} - \frac{7a^2}{3}\bigg)\cal Q(S_3) \nonumber \\ \rule{0ex}{3ex} & & \mbox{} - (10a^2 + 4ac)\cal Q(S_3)^3 \end{eqnarray*} \noindent which in view of \eqref{js} simplifies to \begin{equation} \hspace{-.25in} \bigg [\frac{4}{3}(2a-c)(a+c) - a(7a+4c)j(j+1)\bigg ]\cal Q(S_3) + a(10a+4c)\qz^3 = 0. \label{cc1} \end{equation} Now suppose $j = \frac{1}{2}$, so that $\qz^2 = \frac{1}{4}I$. Then \eqref{cc1} implies that $a = -4c$ which, when substituted into \eqref{js} yields $s=0$. Similarly, when $j = 1$, $\qz^3 = \qz$. In this case \eqref{cc1} implies that $a = -c$, and again \eqref{js} requires $s=0.$ Thus we have derived contradictions for these two values of $j$. Henceforth take $j>1.$ Next we quantize \[ 6s^2\cx\cy\cz = \{S_1\,\!^3,\cyy\cx\} +\{S_2\,\!^3,\czz\cy\} + \{\czzz,\cxx\cz\}. \] \noi Another computer calculation using \eqref{vn3}, \eqref{iji}, \eqref{123}, \eqref{com} and \eqref{s3j} yields \begin{eqnarray*} \lefteqn{6s^2\Big[a\qx\qy\qz + \frac{a}{2\text{i}}\big(\qx^2 - \qy^2 + \qz^2\big)\Big]} \nonumber \\ \rule{0ex}{1ex} & & \hspace{-.18in} \rule{0ex}{4ex} = -3a\text{i}\big(c-2a+aj(j+1)\big)\big[\qx^2-\qy^2+\qz^2 + 2\text{i}\qx\qy\qz\big]. \label{cc2a} \end{eqnarray*} \noi Since for $j>1$ the matrix element \begin{eqnarray*} \lefteqn{\hspace{-1.5in} \Big {\langle}\,j, j\!-\!2\,\Big|\,\qx\qy\qz + \frac{1}{2\text{i}}\big(\qx^2 - \qy^2 + \qz^2\big)\Big|\,j, j\Big{\rangle}} \\ & & = \rule{0ex}{4ex}\frac{1}{2{\text {i}}}(1-j)\sqrt{j(2j-1)} \nonumber \end{eqnarray*} \noi is nonzero, we conclude that either $a=0$ or \begin{equation} s^2 = c-2a + aj(j+1). \label{cc2} \end{equation} \noi If $a=0$, \eqref{cc1} implies that $c=0$, and then \eqref{js} leads to a contradiction, so \eqref{cc2} must hold. Subtracting \eqref{cc2} from \eqref{js} gives $c = -5a/2$; substituting this into \eqref{cc1} produces \[3a^2(j^2+j-3)\qz = 0.\] \noi But then $j = (-1 \pm \surd{13})/2$, neither of which is permissible. Thus we have derived contradictions for all $j>0$ and the theorem follows. \end{pf} We remark that Theorem \ref{ng3} is actually sharper than Theorem \ref{ng2}; we have included the latter because it is simpler. As Proposition \ref{max} shows, by augmenting $\cal P^1$ with a single odd polynomial, we generate $\tilde{\cal O}$. Similarly we can generate all of $\cal P$ from $\cal P^1$ and a single polynomial not in $\tilde{\cal O}$, which implies that the only Poisson subalgebras of $\cal P$ strictly containing $\cal P^1$ are $\tilde{\cal O}$ and $\cal P$ itself. This can be proven in the same manner as Proposition \ref{max}, but in the interests of economy, we make do with: \begin{lem} Any Poisson subalgebra of $\cal P$ which strictly contains $\cal P^1$ also contains $\tilde{\cal O}$. \end{lem} \begin{pf} By Proposition \ref{max} it suffices to consider the Poisson algebra $\cal T$ generated by $\cal P^1$ and a polynomial $p$ not in $\tilde{\cal O}$. Then $p$ has a component in some $\h_{2k}$ for $k>0$, so by Lemma \ref{inhomo} it follows that $\h_{2k}\subset\cal T$. Now consider the bracket $\{Y_{2k}^{2k-1},\, Y_{2k}^{2k}\}$. According to Lemma \ref{main}, either its real or imaginary part has a nonzero component in $\h_{4k-1}$. Hence $\h_{4k-1}\subset\cal T$, and so by Proposition \ref{max} we have $\tilde{\cal O}\subset\cal T$. \end{pf} Now given any Poisson subalgebra of $\cal P$ strictly containing $\cal P^1$ we only have to apply Theorem \ref{ng3} to the subalgebra $\tilde{\cal O}$ inside it to obtain a contradiction, hence: \begin{thm} No nontrivial quantization of $\cal P^1$ can be admissibly extended beyond $\cal P^1$. \label{dogo} \end{thm} This result stands in marked contrast to the analogous one for $\Bbb R^{2n}.$ There one runs into difficulties with {\em cubic} polynomials in $\{1,q^i,p_i\},$ so that $\cal P^2$ (the Poisson algebra of polynomials of degree at most two) is a maximal polynomial subalgebra containing $\cal P^1$ that can be admissibly quantized \cite{g-s}. This dichotomy seems to be connected with the fact that for $\Bbb R^{2n}$, $\cal P^2$ is the Poisson normalizer of $\cal P^1,$ whereas for $S^2$ the normalizer of $\cal P^1$ is itself. On the other hand, it should be noted that there are other maximal polynomial subalgebras of $C^{\infty}(\Bbb R^{2n})$ containing $\cal P^1$ that can be admissibly quantized, for instance the Schr\"odinger subalgebra \[\left\{\sum_{i=1}^nh^i(q^1,\dots,q^n)p_i + k(q^1,\dots,q^n)\right\}\] \noi where the $h^i$ and $k$ are polynomials. Here one encounters problems when one tries to extend to terms which are quadratic in the momenta. \ms Finally, a word is in order regarding the case $j = 0$ -- the one instance in which we did not derive a contradiction. It happens that the spin 0 representation of $\p^1$ {\em can\/} be extended, in a unique way, to an admissible quantization $\q$ of all of $\p$. Indeed, given $p \in \p$, let $p_0$ denote the constant term in the harmonic decomposition \eqref{hd} of $p$. Then $\q:\p \rightarrow \Bbb C$ defined by $\q(p) = p_0$ is, technically, an admissible quantization of $(\p,\p^1)$. To prove this, it is only necessary to show that $\q$ so defined is a Lie algebra homomorphism, which in this context means $\{p,p'\}_0 = 0$. But from \eqref{hdy} and \eqref{pb}, in vector notation, \[4\pi s^2\{p,p'\}_0 = \int_{S^2}\{p,p'\}\,d\sigma = \int_{S^2}\bold S \cdot (\nabla p \times \nabla p')\,d\sigma = s\int_{S^2} (\nabla p \times \nabla p') \cdot d{\boldsymbol \sigma}, \] \noi which vanishes by the divergence theorem.\footnote{This actually is a consequence of a general fact about momentum maps on compact symplectic manifolds, cf. \cite[p.~ 187]{g-s}.} To show that this quantization is unique, we need \begin{lem} For $l > 0$, $\h_l = \{\h_1,\h_l\}.$ \label{l1l} \end{lem} \begin{pf} For $l>0$ $\{\h_1,\h_l\}$ is a nontrivial invariant subspace of $\h_l$, and hence by the irreducibility of the $\h_1$-action on $\h_l$, $\{\h_1,\h_l\} = \h_l.$ \end{pf} Now suppose $\chi$ is an admissible quantization of $(\p,\p^1)$ with $j=0$ so that by the above, $\chi$ is a linear map $\p \ra \Bbb C$ which must annihilate Poisson brackets. But then $\chi(p) = \chi(p_0)$, since by Lemma \ref{l1l} each term $p_l \in \h_l$ for $l > 0$ in the harmonic decomposition of $p$ must be a sum of terms of the form $\{h_l,r_l\}$ for some $h_l \in \h_1$ and $r_l \in \h_l$. It follows that $\chi$ is uniquely determined by its value on the constants, and as $\chi(1) = I = \q(1)$ we must have $\chi = \q.$ Although the corresponding representation of $\p^1$ is trivial in that $\qi = 0$ for all $i$, it is worth emphasizing that $\q$ is {\em not\/} zero on the remainder of $\p$. For example, $\q(\cii) = \frac{s^2}{3}I$ for all $i$, consistent with Proposition \ref{qvn} and \eqref{s2}. The existence of this trivial yet ``not completely trivial'' representation of $\p$ -- for any nonzero value of the classical spin $s$ -- may be related to a well-known ``anomaly'' in the geometric quantization of spin, cf. \cite[\S7]{t} and \cite[\S11.2]{s}. \ms\ms\ms %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section{Discussion} Theorems \ref{ng2} and \ref{dogo} could have been ``predicted'' on the basis of geometric quantization theory. Here one knows that one can quantize those classical observables $f \in C^{\infty}(M)$ whose flows preserve a given polarization $P$. In our case we take $P$ to be the antiholomorphic polarization on $S^2$ (thought of as $\Bbb CP^1$); then $\p^1$ is exactly the set of polarization-preserving observables. However, in geometric quantization theory one does not expect to be able to consistently quantize observables outside this class \cite{w}. Further corroboration for our results is provided by Rieffel \cite{r}, who showed that there are no strict SU(2)-invariant deformation quantizations of $C^{\infty}(S^2)$. In fact, it seems that only the polynomial algebra $\cal P^1 \subset C^{\infty}(S^2)$ can be rigorously deformation quantized in an SU(2)-invariant way \cite{k}. There are several points we would like to make concerning the \vn\ rules for $S^2$, especially in comparison with those for $\Bbb R^{2n}$. Our no\,-go theorems may be interpreted as stating that the ``Poisson bracket $\ra$ commutator'' rule is {\em totally} incompatible with even the relatively weak \vn\ rules given in Propositions \ref{qvn} and \ref{cvn}. On $\Bbb R^{2n}$, on the other hand, the \vn\ rule $k(x) = x^n$ with $n=2$ does hold for $x\in\p^1 \subset \p^2$ in the metaplectic representation \cite{c,g-s}, and with any $n \geq 0$ for $x=q^i$ in the Schr\"odinger representation. Moreover it is curious that, according to Propositions \ref{qvn} and \ref{cvn}, requiring that the $S_i$ be irreducibly represented does not yield strict \vn\ rules as happens for $\{q^i,p_i\}$ on $\Bbb R^{2n},$ cf. \cite{gr,c}.\footnote{In the case of $\Bbb R^{2n}$, one usually employs the Stone-\vn\ theorem to show that irreducibility implies the \vn\ rules, but in fact one can prove this without recourse to the Stone-\vn\ theorem.} It is not clear why $S^2$ and $\Bbb R^{2n}$ behave differently in these regards. We remark that it is substantially easier to prove the no\,-go theorems \ref{ng2} and \ref{ng3} if one assumed from the start that the strict \vn\ rules $\q(\cii) = \q(\ci)^2$ and $\q(\ciii) = \q(\ci)^3$ hold. We can gain some insight into this as follows. Suppose the strong \vn\ rule applied to one of the $S_i$, say $S_3$, so that $\cal Q(S_3\,\!^n) = \cal Q(S_3)^n$ for all positive integers $n$. Suppose furthermore that $\q$ is injective when restricted to the polynomial algebra $\Bbb R[S_3]$ generated by $S_3$. Provided it is appropriately continuous, $\q$ will then extend to an isomorphism from the real $C^*$--algebra $C^*(S_3)$ (consisting of the closure of $\Bbb R[S_3]$ in the supremum norm on $S^2$ with pointwise operations) to the real $C^*$--algebra generated by $\cal Q(S_3)$. But this implies that the classical spectrum of $S_3$ (i.e., the set of all values it takes) is the same as the operator spectrum of $\cal Q(S_3)$. Since the classical spectrum of $S_3$ is $[-s,s]$ whereas the quantum spectrum is discrete, it is clear -- in retrospect -- why no (strong) \vn\ rule can apply to $S_3$. Since $S^2$ is in a sense the opposite extreme from $\Bbb R^{2n}$ insofar as symplectic manifolds go, our result lends support to the contention that no\,-go theorems should hold in some generality. Nonetheless, these two examples are special in that they are symplectic homogeneous spaces \big($\Bbb R^{2n}$ for the translation group $\Bbb R^{2n}$, and $S^2$ for $\mbox{SU}(2)$\big). Thus in both cases we are quantizing finite-dimensional Lie algebras \big(the Heisenberg algebra and $\mbox{su}(2) \times \Bbb R$, respectively, which are certain central extensions by $\Bbb R$ of $\Bbb R^{2n}$ and su(2)\big). Will a similar analysis work for other symplectic homogeneous spaces, e.g., $\Bbb CP^n$ with group $\mbox{SU}(n+1)$? How does one proceed in the case of symplectic manifolds which do not have such a high degree of symmetry? What set of observables will play the role of the distinguished subalgebras generated by $\{1,q^i,p_i\}$ and $\{1,S_1,S_2,S_3\}$ (which are the components of the momentum mappings for the Hamiltonian actions of the Heisenberg group and $\mbox{SU}(2) \times \Bbb R$, resp.)? For a cotangent bundle $T^*Q$, the obvious counterpart would be the infinite-dimensional abgebra of linear momentum observables $P_X + f,$ where $P_X$ is the momentum in the direction of the vector field $X$ on $Q$ and $f$ is a function on $Q$. (These are the components of the momentum mapping for the transitive action of $\text{Diff}(Q) \ltimes C^{\infty}(Q)$ on $T^*Q$.) In this regard, it is known that geometric quantization (formally) obeys the von Neumann rules $\cal Q\big(P_X\,^2\big) = \cal Q(P_X)^2$ and $\cal Q(f^n) = \cal Q(f)^n$ \cite{go2}. We hope to explore some of these issues in future papers. \ms\ms %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \section*{Acknowledgments} One of us (M.J.G.) would like to thank L. Bos, G. Emch, G. Goldin, M. Karasev and J. Tolar for enlightening conversations, and the University of New South Wales for support while this research was underway. He would also like to express his appreciation to both the organizers of the XIII Workshop on Geometric Methods in Physics and the University of Warsaw, who provided a lively and congenial atmosphere for working on this problem. This research was supported in part by NSF grant DMS-9222241. Another one of us (H.G.) would like to acknowledge the support of an ARC--grant. \ms %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \appendix \section{\vn\ Rules} Here we provide the proofs of Propositions \ref{qvn} and \ref{cvn}. Fix an irreducible representation of $\text {su(2)} \times \Bbb R$ labeled by the quantum number $j$. The proofs depend on the following three facts: \vskip 2pt \begin{enumerate} \item[({\em i\/})] As the representation is irreducible, any endomorphism of the representation space is an element of the enveloping algebra of the generators $\qi$ \cite[Prop. 2.6.5]{d}, and hence can be expressed as a polynomial in the $\qi$. \vskip 6pt \item[({\em ii\/})] Under the su(2)-action, a monomial $\qx\,\!^{n_1}\qy\,\!^{n_2}\qz\,\!^{n_3}$ of degree $|n| = n_1 + n_2 + n_3$ transforms as a tensor operator of rank $|n|$. \vskip 6pt \item[({\em iii\/})] Under the induced action of su(2) on $\p$, a monomial $S_1\,\!^{n_1}S_2\,\!^{n_2}S_3\,\!^{n_3}$ transforms as a symmetric tensor of rank $|n|$. Since the quantization $\q$ is (infinitesimally) equivariant, it follows that $\q(S_1\,\!^{n_1}S_2\,\!^{n_2}S_3\,\!^{n_3})$ also transforms as a symmetric tensor operator of rank $|n|$. \end{enumerate} \noi When $ |n| > 1$, the tensor operators $\qx\,\!^{n_1}\qy\,\!^{n_2}\qz\,\!^{n_3}$ and $\q(S_1\,\!^{n_1}S_2\,\!^{n_2}S_3\,\!^{n_3})$ are reducible. Equation \eqref{ii} then follows from the observation that as $\qii$ is a reducible symmetric tensor operator of rank 2, its irreducible constituents must be of even rank, and hence it must be equal to an even polynomial in the $\qi$ of degree at most 2. Equations \eqref{pr}, \eqref{vn3}, \eqref{iji} and \eqref{123} follow from similar observations. The rest of the argument is the working out of these observations. Letting $n = (n_1,n_2,n_3)$ be a multi-index of length $|n| = n_1+n_2+n_3$, we denote $\qx\,\!^{n_1}\qy\,\!^{n_2}\qz\,\!^{n_3}$ by $\Bbb Q_{\bf n}^{|n|}$ and $S_1\,\!^{n_1}S_2\,\!^{n_2}S_3\,\!^{n_3}$ by $\Bbb S_{n}^{|n|}$. Since the commutation relations \eqref{com} are nonlinear, the difference between each $\Bbb Q_{\bf n}^{|n|}$ and its symmetrization $\Bbb Q_{\bf (n)}^{|n|}$ is a linear combination of tensor operators $\Bbb Q_{\bf m}^{|m|}$ of lower rank $|m|$. Thus we may use the symmetrized tensor operators $\Bbb Q_{\bf (n)}^{|n|}$ as a basis for the enveloping algebra of the operators $\qi$. Then by ({\em i\/}) we can expand \begin{equation} \q\big(\sm\big) = \sum_{|n|=0}^{d}[\,m\,|\,n\,]\,\Bbb Q_{\bf (n)}^{|n|} \label{Expand} \end{equation} \noi as a polynomial of degree $d$, say.\footnote{It can be shown that $d \leq 4j$, but we do not need this fact.} We can further decompose each \begin{equation} \Bbb Q_{\bf (n)}^{|n|} = \sum_{\ld = 0}^{|n|}\sum_{\mu = -\ld}^{\ld}(\,n\,|\,\ld\;\mu\,)\,T^{\ld}_{\mu} \label{t} \end{equation} \noi into a sum of irreducible spherical tensor operators $T^{\ld}_{\mu}$ of rank $\ld$ (with $\mu = -\ld,\dots,\ld$),\footnote {Since they are symmetric the tensor operators $\Bbb Q_{\bf (n)}^{|n|}$ are ``simply reducible,'' i.e., in the decomposition \eqref{t} there is at most one irreducible constituent $T^{\ld}_{\mu}$ for each weight $\ld$. This follows from a consideration of Young tableaux. Likewise, the $\q\big(\sm\big)$, being symmetric, are simply reducible, and hence there is no degeneracy in either \eqref{Irredc} or \eqref{Thirdord} below.} so that \eqref{Expand} becomes \begin{equation} \q\big(\sm\big) = \sum_{|n|=0}^{d}\sum_{\lambda=0}^{|n|}\sum_ {\mu=-\lambda}^{\lambda}[\,m\,|\,n\,]\,(\,n\,|\,\lambda\; \mu\,)\,T_{\mu}^{\lambda}. \label{Exdecom} \end{equation} On the other hand, for $|m|=2$ we may directly decompose \begin{equation} \q\big(\Bbb S^2_m\big) = \sum_{\nu=-2}^2 (\,m\,|\,2\;\nu\,)\,V_{\nu}^2 + (\,m\,|\,0\;0\,)\,V^0_0, \label{Irredc} \end{equation} \noi where the irreducible constituents $V^{\eta}_{\nu}$ of $\q\big(\Bbb S^2_m\big)$ are given by \begin{eqnarray} V_{\nu}^2 & = &\sum_{|m|=2}(\,2\;\nu\,|\,m\,)\,\q(\Bbb S^2_m), \label{Irreda}\\ V_0^0 & = &\sum_{|m|=2}(\,0\;0\,|\,m\,)\,\q(\Bbb S^2_m). \label{Irredb} \end{eqnarray} \noi Here we have used the relation $\sum_{\ld \geq 0}\sum_{\nu = -\ld}^{\ld}(\,m\,|\,\ld\;\nu\,)\,(\,\ld\;\nu\,|\,m'\,) = \delta_{mm'}$. Note that, as it is symmetric, $\q\big(\Bbb S^2_m\big)$ has no irreducible rank 1 constituent. Combining \eqref{Irreda} and \eqref{Irredb} with \eqref{Exdecom}, we obtain \begin{eqnarray} V_{\nu}^2 & = &\sum_{|m|=2}\sum_{|n|=0}^{d}\sum_{\lambda =0}^ {|n|}\sum_{\mu = -\ld}^{\ld}(\,2\;\nu\,|\,m\,)\,[\,m\,|\,n\,]\,(\,n\,|\,\lambda \;\mu\,)\,T_{\mu}^{\lambda}, \label{Exirreda}\\ V^0_0 & = &\sum_{|m|=2}\sum_{|n|=0}^{d}\sum_{\lambda=0}^{|n|} \sum_{\mu=-\ld}^{\ld}(\,0\;0\,|\,m\,)\,[\,m\,|\,n\,]\,(\,n\,|\,\lambda\;\mu\,)\, T_{\mu}^{\lambda}. \label{Exirredb} \end{eqnarray} Now apply a rotation $R \in {\text {SU(2)}}$ to \eqref{Exirreda} to obtain \begin{eqnarray*} \lefteqn{\sum_{\nu'=-2}^2D_{\nu\nu'}^2(R)V_{\nu'}^2 =} \\ & & \sum_{|m|=2}\sum_{|n|=0}^{d}\sum_ {\lambda=0}^{|n|}\sum_{\mu=-\ld}^{\ld}\sum_{\mu'=-\ld}^{\ld}(\,2\;\nu\,|\,m\,)\, [\,m\,|\,n\,]\, (\,n\,|\,\lambda\;\mu\,)\,D_{\mu \mu'}^{\lambda}(R)\,T_{\mu'}^{\lambda}. \label{Rotate} \end{eqnarray*} \noi Multiplying both sides of this by $D_{\nu\rho'}^2(R)^*$ and integrating over the group manifold, the orthogonality theorem for products of representations \cite[\S 7.1]{b-r} yields \begin{equation} V_{\rho'}^2 = \left[\sum_{|m|=2}\sum_{|n|=0}^{d} (\,2\;\nu\,|\,m\,)\,[\,m\,|\,n\,]\,(\,n\,|\,2\;\nu\,)\right]T_{\rho'}^2. \label{FConda} \end{equation} \noi Applying the same procedure to \eqref{Exirredb} but using $D^0_{00}(R)^*$ instead, we similarly obtain \begin{equation} V_0^0 = \left[\sum_{|m|=2}\sum_{|n|=0}^{d}(\,0\;0\,|\,m\,)\,[\,m\,|\,n\,]\, (\,n\,|\,0\;0\,)\right]T_0^0.\label{FCondd} \end{equation} {}From \eqref{FConda} and \eqref{FCondd}, we see that $V_{\rho'}^2$ is proportional to $T_{\rho'}^2$, and $V_0^0$ to $T^0_0$; let the corresponding constants of proportionality be $a$ and $b$. Substituting \eqref{FConda} and \eqref{FCondd} into \eqref{Irredc} and inverting \eqref{t} gives \begin{eqnarray} \q\big(\Bbb S^2_m\big) & = & a\sum_{\nu=-2}^2(\,m\,|\,2\;\nu\,)\,T_{\nu}^2 + b(\,m\,|\,0\;0\,)\,T_0^{0} \nonumber \\ & = & a\sum_{|n|=2}\sum_{\nu=-2}^2(\,m\,|\,2\;\nu\,)\,(\,2\;\nu\,|\,n\,)\,\Bbb Q^2_{(n)} + b(\,m\,|\,0\;0\,)\,(\,0\;0\,|\,0\,)\,\Bbb Q^0_0 \nonumber \\ \rule{0ex}{4ex} & = & a\sum_{|n|=2}^{\,}P_{2}(m,n)\,\Bbb Q^2_{(n)} + bP_0(m,0)I. \label{Secord} \end{eqnarray} \noi In this last expression, $P_{\ld}(m,n) = \sum_{\nu=-\ld}^{\ld}(\,m\,|\,\ld\;\nu)\,(\,\ld\;\nu\,|\,n\,)$ is the projector which picks off the $m^{\text {th}}$-component (with respect to $\Bbb Q^{|m|}_{(m)}$) of the irreducible rank $\ld$ constituent of $\Bbb Q_{(n)}^{|n|}$. Since $m$ has length 2, $m$ must be of the form $1_i+1_\ell$, where $1_i$ is the multi-index with 1 in the $i^{\text{th}}$-slot and zeros elsewhere. Similarly, when $|n| =2$, $n = 1_p+1_q$. Then we have \begin{eqnarray*} P_{2}(1_i+1_\ell,1_p+1_q) & = & \frac{1}{2}\big(\delta_{ip}\delta_{\ell q}+\delta_{iq}\delta_{\ell p}\big) - \frac{1}{3}\delta_{i\ell}\delta_{pq}, \label{Proj2} \\ \rule{0ex}{4ex} P_0(1_i+1_\ell,0) & = & \frac{1}{3}\delta_{i\ell}. \label{Proj0} \end{eqnarray*} Setting $\ell = i$, \eqref{Secord} and \eqref{j2} then yield \begin{equation*} \qii=a\left(\qi^2-\frac{1}{3}j(j+1)I\right)+\frac{1}{3}bI. \label{Prop2} \end{equation*} \noi The constants $a$ and $b$ must both be real \big(as $\qii$ is self-adjoint\big), and both cannot be simultaneously zero \big(for this would contradict \eqref{s2}\big). Thus, upon setting $c = \big(b-aj(j+1)\big)/3$, we obtain \eqref{ii}. Taking $\ell \neq i$, we similarly obtain \eqref{pr}. The same arguments can be applied, {\em mutatis mutandis}, to $\q\big(\Bbb S^3_m\big)$. Then \eqref{Secord} is replaced by \begin{eqnarray} \q\big(\Bbb S^3_m\big) & = & a\sum_{\nu=-3}^3(\,m\,|\,3\;\nu\,)\,T_{\nu}^3 + b\sum_{\nu=-1}^1(\,m\,|\,1\;\nu\,)T_{\nu}^1 \nonumber\\ & = & \rule{0ex}{4ex} a\sum_{|n|=3}P_{3}(m,n)\,\Bbb Q^3_{(n)} + b\sum_{|n|=1}P_{1}(m,n)\,\Bbb Q^1_n. \label{Thirdord} \end{eqnarray} \noi The projectors in this instance are \begin{eqnarray*} \lefteqn{\hspace {-1.17in}P_{3}(1_i+1_k+1_{\ell},1_p+1_q+1_r)} \\ & = & \rule{0ex}{4ex}{1\over 6}\sum\delta_{ip}\delta_{kq}\delta_{\ell r}- {1\over 30}\sum\delta_{ik}(\delta_{pq}\delta_{\ell r}+\delta_{pr}\delta_{\ell q}+\delta_{qr}\delta_{\ell p}), \label{eq:Proj3}\\ \rule{0ex}{4ex} P_{1}(1_i+1_k+1_{\ell},1_p) & = & {1\over 15}(\delta_{ik}\delta_{\ell p}+\delta_{k\ell}\delta_{ip}+\delta_{\ell i}\delta_{kp}), \label{eq:Proj1} \end{eqnarray*} \noi where the sums are over all permutations of $\{i,k,\ell\}$. Setting $i = k = \ell$, \eqref{Thirdord} reduces to \begin{equation*} \qiii=a\left(\qi^3-\frac{3}{5}\Big(j(j+1)-\frac{1}{3}\Big)\qi\right)+\frac{1 }{5}b\qi. \label{Prop3} \end{equation*} \noi Again by self-adjointness, $a$ and $b$ must be real. Upon setting $c = \frac{1}{5}\big(b-3\big[j(j+1) - \frac{1}{3}\big]\big)$ and rearranging the above, we obtain \eqref{vn3}. Finally, taking $i=k \neq \ell$ and $(i,k,\ell) = (1,2,3)$ in \eqref{Thirdord}, we similarly obtain \eqref{iji} and \eqref{123}, respectively. Although it is not apparent from this derivation, it can be shown that \eqref{pr} is actually a consequence of \eqref{ii} and, likewise, that \eqref{iji} and \eqref{123} both follow from \eqref{vn3}. \ms\ms %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \appendix \addtocounter{section}{1} \section{On the Harmonic Decomposition of $\{Y_l^{l-j},Y_l^l\}$} In this appendix we prove Lemma \ref{main}, which computes the harmonic decomposition of $\{Y_l^{l-j},Y_l^l\}$ for $l>0$. Specifically, for each $j,\; 0 < j \leq 2l$, we have \begin{equation*} \{Y_l^{l-j},Y_l^l\} = \sum_{k=1}^{l}y_{2k-1}(l-j,l)Y_{2k-1}^{2l-j}, \end{equation*} \noi where the coefficients $y_{2k-1}(l-j,l)$ have the following properties: \begin{enumerate} \item[({\em i\/})] when $j = 1,$ the top coefficients $y_{2l-1}(l-1,l) \neq 0$, and \ms \item[({\em ii\/})] provided $l\geq 5$, $k \geq \frac{l-1}{2}$ and $k > l - \frac{j+1}{2}$, we have $y_{2k-1}(l-j,l) \neq 0.$ \end{enumerate} \ms The proof will be presented in several steps. We refer the reader to \cite[Chapter XIII and Appendix C]{m} for the relevant background and conventions on spherical harmonics. %%%%%%%%%%%%%%%%%%%%%%%% \ms {\em Step 1.} It is convenient to work in spherical coordinates $(\theta,\phi)$ on $S^2$. The generators $L_3$ and $L_{\pm} = L_1 \pm \text {i}L_2$ of O(3) then take the form \[L_3 = \frac{\partial}{\partial \phi} \;\;\mbox{and}\;\; L_{\pm} = \pm\mbox{i} e^{\pm \text{i}\phi} \bigg (\frac{\partial}{\partial \theta} \pm \mbox{i}\cot \theta \frac{\partial}{\partial \phi}\bigg). \] \noi They satisfy \[L_3Y_l^m = {\text i}mY_l^m, \;\;\;\;L_+Y_l^m = {\text i}\beta_{l,m}Y_l^{m+1}\;\;\mbox{ and }\;\; L_-Y_l^m = {\text i}\beta_{l,m-1}Y_l^{m-1}\] \noi where $\beta_{l,m} = \sqrt{(l+m+1)(l-m)}.$ The Poisson bracket \eqref{pb} becomes \begin{equation*} \{f,g\} = \frac{\csc \theta}{s} \bigg (\frac{\partial f}{\partial \phi}\frac{\partial g}{\partial \theta} - \frac{\partial f}{\partial \theta}\frac{\partial g}{\partial \phi}\bigg ), \end{equation*} \noi which can be rewritten in terms of the $L_i$ as \begin{equation*} \{f,g\} = -\frac{\text{i}}{s}\csc \theta\, e^{-{\text{i}}\phi}\Big (L_3(f)L_+(g)-L_+(f)L_3(g)\Big). \label{pbs} \end{equation*} \ms %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% {\em Step 2.} As a specific instance of this formula, we compute \begin{equation} \{Y_l^m,Y_l^n\}\sin \theta \, e^{\text{i}\phi} = \frac{\mbox i}{s}\Big(m\,\beta_{l,n}Y_l^mY_l^{n+1}-n\,\beta_{l,m}Y_l^{m+1}Y_l^n\Big). \label{mn} \end{equation} \noi Now we have the product decomposition \begin{eqnarray} Y_{l_1}^{m_1}Y_{l_2}^{m_2}&=&\sum_{l = |l_1-l_2|}^{l_1+l_2} \sqrt{\frac{(2l_1+1)(2l_2+1)}{4\pi(2l+1)}}\: \langle\,l_1\,l_2\,0\,0\,|\,l\,0\,\rangle \nonumber\\ & &\times\:\langle\,l_1\,l_2\,m_1\,m_2\,|\,l\;m_1\!+\!m_2\,\rangle Y_l^{m_1+m_2}, \label{pd} \end{eqnarray} \noi where the quantities $\langle\,l_1\,l_2\,m_1\,m_2\,|\,L\,M\,\rangle$ are \cg s. Applying \eqref{pd} to the r.h.s. of \eqref{mn} gives \begin{eqnarray} \{Y_l^m,Y_l^n\}\sin \theta \, e^{\text{i}\phi} & = & \frac{\mbox i}{s} \sum_{j=0}^{2l}\frac{2l+1}{\sqrt{4\pi(2j+1)}}\langle\,l\,l\,0\,0\,|\,j\,0\, \rangle \label{yyr} \\ & & \mbox{}\!\! \times \Big [m\,\beta_{l,n}\langle\,l\,l\,m,\,n\!+\!1\,|\,j,\,m\!+\!n\!+1\,\rangle \nonumber \\ & & \mbox{} \;\;\;\;\; - n\, \beta_{l,m} \langle\,l\,l\,m\!+\!1,\,n\,|\,j,\,m\!+\!n\!+1\,\rangle \Big ]Y_j^{m+n+1}. \nonumber \end{eqnarray} On the other hand, since $\{\h_l,\h_l\} \subset \oo^{2l-1}$ we can expand \begin{equation*} \{Y_l^m,Y_l^n\} = \sum_{k=1}^{l}\sum_{r=-2k+1}^{2k-1}y_{2k-1,r}(m,n)Y_{2k-1}^r. \end{equation*} \noi Multiply this equation through by $\sin \theta\,e^{\text {i}\phi}$. Since $\sin \theta\,e^{\text {i}\phi} = - \sqrt{8\pi/3}\,Y_1^1$, we can apply the product decomposition \eqref{pd} to the r.h.s. of the resulting expression to obtain \begin{eqnarray} \{Y_l^m,Y_l^n\}\sin \theta \, e^{\text{i}\phi} & = & - \sqrt{\frac{8\pi}{3}} \bigg (\sum_{k=1}^{l}\sum_{r=-2k+1}^{2k-1}y_{2k-1,r}(m,n)\big [Y_{2k-1}^rY_1^1\big ]\bigg) \label{yyl} \\ & = & \sum_{k=1}^{l}\sum_{r=-2k+1}^{2k-1}y_{2k-1,r}(m,n)\Big (e_{2k-1,r}Y_{2k-2}^{r+1} - d_{2k-1,r}Y_{2k}^{r+1}\Big), \nonumber \end{eqnarray} \noi where \[d_{t,r} = \sqrt{\frac{(t+r+1)(t+r+2)}{(2t+1)(2t+3)}}\:\:\: \mbox{ and }\:\:e_{t,r} = \sqrt{\frac{(t-r)(t-r-1)}{(2t+1)(2t-1)}}.\] Comparing \eqref{yyr} with \eqref{yyl}, we see that $r = m+n$ and hence \begin{eqnarray*} \lefteqn{\hspace{-.5in} \sum_{k=1}^{l}y_{2k-1,m+n}(m,n) \Big (e_{2k-1,m+n}Y_{2k-2}^{m+n+1} - d_{2k-1,m+n}Y_{2k}^{m+n+1}\Big)} \\ & = & \frac{\mbox i}{s} \sum_{j=0}^{2l}\frac{2l+1}{\sqrt{4\pi(2j+1)}}\langle\,l\,l\,0\,0\,|\,j\,0\, \rangle \\ & & \mbox{} \!\times \Big [m\,\beta_{l,n}\langle\,l\,l\,m,\,n\!+\!1\,|\,j,\,m\!+\!n\!+1\,\rangle \\ & &\mbox{}\!\;\;\;\;\;\; - n\, \beta_{l,m} \langle\,l\,l\,m\!+\!1,\,n\,|\,j,\,m\!+\!n\!+1\,\rangle \Big ]Y_j^{m+n+1}. \end{eqnarray*} \noi Note that the sum on the l.h.s. contains only even degree harmonics, and hence the sum on the r.h.s. must as well. \big(This is reflected by the vanishing of the \cg s $\langle\,l\,l\,0\,0\,|\,j\,0\,\rangle$ for $j$ odd, cf. \cite[eq. (C23.a)]{m}.\big) Thus we may reindex $j=2k,\: k = 0,\ldots,l$ on the r.h.s.. Upon reindexing $k \mapsto k+1$ in the first term on the l.h.s. and equating coefficients of $Y_{2k}^{m+n+1}$ on both sides, we obtain the recursion relation \begin{eqnarray} y_{2k-1}(m,n) & = & \frac{e_{2k+1,m+n}}{d_{2k-1,m+n}}\,y_{2k+1}(m,n) \nonumber \\ & & \mbox{} - \frac{\text{i}}{s} \frac{2l+1}{\sqrt{4\pi(4k+1)}\, d_{2k-1,m+n}}\, \langle \,l\,l\,0\,0\,|\,2k\,0\, \rangle \nonumber \\ & & \mbox{} \times \Big{[} m\,\beta_{l,n}\langle\,l\,l\,m\,n\!+\!1\,|\,2k,\,m\!+\!n\!+1\,\rangle \nonumber \\ & & \mbox{}\;\;\;\;\;\; - n\, \beta_{l,m} \langle\,l\,l\,m\!+\!1,\,n\,|\,2k,\,m\!+\!n\!+1\,\rangle \Big{]} \label{rry} \end{eqnarray} \noi where we have abbreviated $y_{2k-1,m+n}(m,n) =: y_{2k-1}(m,n)$. \ms {\em Step 3.} Now we specialize even further, setting $m = l-j$ and $n=l$ for $j = 1,\cdots,2l$. Then we have \begin{equation} \{Y_l^{l-j},Y_l^l\} = \sum_{k=1}^{l}y_{2k-1}(l-j,l)Y_{2k-1}^{2l-j}, \label{ylyl} \end{equation} \noi as in the statement of the lemma, and \eqref{rry} reduces to \begin{eqnarray} y_{2k-1}(l-j,l) & = & \frac{e_{2k+1,2l-j}}{d_{2k-1,2l-j}}\,y_{2k+1}(l-j,l) \nonumber \\ & & \mbox{} + \frac{\text{i}}{s} \frac{l(2l+1)}{\sqrt{4\pi(4k+1)}\, d_{2k-1,2l-j}}\,\beta_{l,l-j} \langle \,l\,l\,0\,0\,|\,2k\,0\, \rangle \nonumber \\ & & \mbox{} \times \langle\,l\,l\,l\!-\!j\!+\!1,\,l\,|\,2k,\,2l\!-\!j\!+1\,\rangle. \label{rrl} \end{eqnarray} \noi The main reason for this choice of $m$ and $n$ is that it simplifies the last term in \eqref{rry}, since $\beta_{l,l}=0$. Before proceeding, we must evaluate the \cg s in \eqref{rrl}. Using the Racah formula \cite[(C.21), (C.22) and (C.23b)]{m}, we compute \[\langle \,l\,l\,0\,0\,|\,2k\,0 \,\rangle = (-1)^{k+l}\sqrt{4k+1}\, \sqrt{\frac{(2l-2k)!}{(2l+2k+1)!}}\, \frac{(2k)!(l+k)!}{(k!)^2(l-k)!}\] \noi and \begin{eqnarray*} \lefteqn{\hspace{-1in} \langle\,l\,l\,l\!-\!j\!+\!1,\,l\,|\,2k,\,2l\!-\!j\!+1\,\rangle = } \hspace{-1in} \\ & & \\ & & \sqrt{4k+1}\, \sqrt{\frac{(2l)!(j-1)!(2k+2l-j+1)!}{(2l-j+1)!(2l+2k+1)!(2l-2k)!(2k-2l+j-1)!}}. \end{eqnarray*} \noi Substituting these expressions as well as those for $d_{2k-1,2l-j}$, $e_{2k+1,2l-j}$ and $\beta_{l,l-j}$ into \eqref{rrl}, the recursion relation becomes \begin{eqnarray*} \tilde y_{2k-1} & = & \sqrt{\frac{(2k-2l+j+1)(2k-2l+j)(4k-1)} {(2k+2l-j)(2k+2l-j+1)(4k+3)}}\,\tilde y_{2k+1} \\ & & \mbox{} + (-1)^k \sqrt{\frac{(2k+2l-j-1)!}{(2k-2l+j-1)!}}\, \frac{\sqrt{4k-1}(4k+1)(l+k)!(2k)!}{(2l+2k+1)!(l-k)!(k!)^2}, \end{eqnarray*} \noi where $\tilde y_{2k-1}$ is defined according to \[y_{2k-1}(l-j,l) = \frac{\text i}{s}(-1)^l \frac{l(2l+1)}{\sqrt{4\pi}}\,\sqrt{\frac{j!(2l)!}{(2l-j)!}}\,\tilde y_{2k-1}.\] \noi Since $l>0$, $y_{2k-1}(l-j,l) = 0$ iff $\tilde y_{2k-1} = 0$. Finally, we rewrite this in the form \begin{equation} \tilde y_{2k-1} = Z_{2k-1}\big [\tilde y_{2k+1} + (-1)^k W_{2k-1}\big ] \label{rrb} \end{equation} \noi where \[Z_{2k-1}= \sqrt{\frac{(2k-2l+j+1)(2k-2l+j)(4k-1)} {(2k+2l-j)(2k+2l-j+1)(4k+3)}}\] \noi and \[W_{2k-1} = \sqrt{\frac{(2k+2l-j+1)!}{(2k-2l+j+1)!}}\, \frac{\sqrt{4k+3}(4k+1)(l+k)!(2k)!}{(2l+2k+1)!(l-k)!(k!)^2}. \] \noi Notice that for fixed $j$ and $l$, $y_{2k-1}(l-j,l) = 0$ if $2k-1 \leq 2l-j-2.$ Upon setting $k=l$ and $j=1$, \eqref{rrb} gives $\tilde y_{2l-1} = (-1)^l Z_{2l-1}W_{2l-1} \neq 0$ and hence $y_{2l-1}(l-1,l) \neq 0$. This proves ({\em i\/}). %%%%%%%%%%%%%%%%%%%%%%% \ms {\em Step 4.} Fix $l$ and $j$, in which case the coefficients $Z_{2k-1}$ and $W_{2k-1}$ are nonzero whenever $2k-1 > 2l-j-2$. We claim that \begin{equation} Z_{2k-1}W_{2k-1} < W_{2k-3} \label{ratio} \end{equation} \noi provided $l \geq 5$ and $k \geq \frac{l-1}{2}$. Indeed, we compute \[\frac{Z_{2k-1}W_{2k-1}}{W_{2k-3}} = \frac{(4k+1)(l-k+1)(2k-1)}{(4k-3)(2l+2k+1)k}.\] \noi But now one verifies that the maximum of the r.h.s. of this expression is 18/25 on the domain in the $k,l$-plane determined by the above inequalities along with the fact that $k \leq l.$ %%%%%%%%%%%%%%%%%%%%%%% \ms {\em Step 5.} We can now prove statement ({\em ii\/}) of the lemma. Fix $l\geq 5$ and $j\in\{1,\ldots,2l\}$. The solution of the recursion relation \eqref{rrb} with initial condition $\tilde y_{2l+1} = 0$ is \[\tilde y_{2l-2n-1} = (-)^l \sum_{k=0}^n(-1)^{k}U_k\] \noi where $U_k= W_{2l-2k-1}\prod\limits_{t=0}^{n-k}Z_{2l-2n+2t-1}$, and $n=0,1,\ldots,\text{max}\{\frac{l+1}{2},\frac{j-1}{2}\}$. For the alternating sum $\sum_{k=0}^n(-1)^kU_k$ we have by \eqref{ratio} that $U_{k+1}/U_k>1$, so if $n$ is even the sum is \[U_0+(U_2-U_1)+\cdots+(U_n-U_{n-1})>0\] \noi since $U_0$ and all bracketed terms are positive, and if $n$ is odd \[(U_0-U_1)+(U_2-U_3)+\cdots+(U_{n-1}-U_n)<0\] \noi likewise. So the sum is always nonzero in the given ranges of the parameters, hence $\tilde y_{2k-1}\not=0$ as claimed. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \ms\ms \begin{thebibliography}{99} \bibitem[{\bf A-B-R}]{a-b-r} Axler, S., Bourdon P. and Ramey, W. [1992] {\sl Harmonic Function Theory}. Grad. Texts in Math. {\bf 137} (Springer, New York). \bibitem[{\bf A-M}]{a-m} Abraham, R. and Marsden, J.E. [1978] {\sl Foundations of Mechanics}. Second Ed. (Benjamin-Cummings, Reading, MA). \bibitem[{\bf B-R}]{b-r} Barut, A.O. and R\c aczka, R. [1977] {\sl Theory of Group Representations and Applications}. (Polish Scientific Publishers, Warsaw). \bibitem[{\bf C}]{c} Chernoff, P. [1981] Mathematical obstructions to quantization. {\sl Hadronic J.} {\bf 4}, 879-898. \bibitem[{\bf D}]{d} Dixmier, J. [1977] {\sl Enveloping Algebras}. (North Holland, Amsterdam). \bibitem[{\bf F}]{f} Folland, G.B. [1989] {\sl Harmonic Analysis in Phase Space}. Ann. Math. Studies {\bf 122} (Princeton Univ. Press, Princeton). \bibitem[{\bf Go1}]{go1} Gotay, M.J. [1980] Functorial geometric quantization and Van Hove's theorem. {\sl Int. J. Theor. Phys.} {\bf 19}, 139-161. \bibitem[{\bf Go2}]{go2} Gotay, M.J. [1987] Formal quantization of quadratic momentum observables. In: {\sl The Physics of Phase Space}, Y.S. Kim and W.W. Zachary, Eds. Lect. Notes in Physics {\bf 278}, 375-379. \bibitem[{\bf Gr}]{gr} Groenewold, H.J. [1946] On the principles of elementary quantum mechanics. {\sl Physics} {\bf 12}, 405-460. \bibitem[{\bf G-S}]{g-s} Guillemin, V. and Sternberg, S. [1984] {\sl Symplectic Techniques in Physics}. (Cambridge Univ. Press, Cambridge). \bibitem[{\bf H-M}]{h-m} Helton, J.W. and Miller, R.L. [1994] {\sl NCAlgebra: A Mathematica Package for Doing Non Commuting Algebra}. v0.2 (Available from ncalg@@ucsd.edu, La Jolla). \bibitem[{\bf J}]{j} Joseph, A. [1970] Derivations of Lie brackets and canonical quantization. {\sl Comm. Math. Phys.} {\bf 17}, 210-232. \bibitem[{\bf K}]{k} Karasev, M. [1994] Private communication. \bibitem[{\bf M}]{m} Messiah, A. [1962] {\sl Quantum Mechanics II.} (Wiley, New York). \bibitem[{\bf R}]{r} Rieffel, M. A. [1989] Deformation quantization of Heisenberg manifolds. {\sl Commun. Math. Phys.} {\bf 122}, 531-562. \bibitem[{\bf \'S}]{s} \'Sniatycki, J. [1980] {\sl Geometric \cal Quantization and \cal Quantum Mechanics.} Appl. Math. Sci. {\bf 30} (Springer, New York). \bibitem[{\bf T}]{t} Tuynman, G.T. [1987] Generalised Bergman kernels and geometric quantization. {\sl J. Math. Phys.} {\bf 28}, 573-583. \bibitem[{\bf VH1}]{vh1} Van Hove, L. [1951] Sur le probl\`eme des relations entre les transformations unitaires de la m\'ecanique quantique et les transformations canoniques de la m\'ecanique classique. {\sl Acad. Roy. Belgique Bull. Cl. Sci.} (5) {\bf 37}, 610-620. \bibitem[{\bf VH2}]{vh2} Van Hove, L. [1951] Sur certaines repr\'esentations unitaires d'un groupe infini de transformations. {\sl Mem. Acad. Roy. Belgique Cl. Sci} {\bf 26}, 61-102. \bibitem[{\bf vN}]{vn} von Neumann, J. [1955] {\sl Mathematical Foundations of \cal Quantum Mechanics.} (Princeton. Univ. Press, Princeton). \bibitem[{\bf W}]{w} Woodhouse, N.M.J. [1992] {\sl Geometric \cal Quantization.} (Clarendon Press, Oxford). \end{thebibliography} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \end{document}