0807.4646/aqo.tex
1: \documentclass[12pt]{article}
2: \usepackage{amssymb,bbm,amsthm,amscd,amsmath}
3: \usepackage{epsfig,graphics}
4: \usepackage{setspace}
5: 
6: 
7: 
8: \newcommand{\II}{{\mathbbm{1}}}
9: \parindent0pt
10: 
11: 
12: %----------------------------------------------------------------------------
13: %------------------------- Beginn des Vorspanns -----------------------------
14: %----------------------------------------------------------------------------
15: 
16: 
17: \topmargin-0.5cm
18: \oddsidemargin-0.05cm
19: \evensidemargin-0.05cm
20: \textheight22cm
21: \textwidth16cm
22: \pagenumbering{arabic}
23: \pagestyle{plain}
24: \footskip2cm
25: \parindent0pt
26: \setcounter{secnumdepth}{2}
27: \setcounter{tocdepth}{2}
28: %
29: % Umgebungen
30: %
31: \theoremstyle{theorem}
32: \newtheorem{Theorem}{Theorem}
33: \newtheorem{Proposition}{Proposition}
34: \newtheorem{Lemma}{Lemma}
35: \newtheorem{Corollary}{Corollary}
36: \newtheorem{Definition}{Definition}
37: \newtheorem{Punkt}{}[subsection]
38: 
39: \theoremstyle{remark}
40: \newtheorem{Remark}{Remark}
41: 
42: \newcommand{\btm}{\begin{Theorem}}
43: \newcommand{\etm}{\end{Theorem}}
44: 
45: \newcommand{\ben}{\begin{enumerate}}
46: \newcommand{\een}{\end{enumerate}}
47: 
48: \newcommand{\bpu}{\begin{Punkt}\rm}
49: \newcommand{\epu}{\end{Punkt}}
50: 
51: \newcommand{\bre}{\begin{Remark}\rm}
52: \newcommand{\ere}{\end{Remark}}
53: 
54: \newcommand{\ble}{\begin{Lemma}}
55: \newcommand{\ele}{\end{Lemma}}
56: 
57: \newcommand{\bpr}{\begin{Proposition}}
58: \newcommand{\epr}{\end{Proposition}}
59: 
60: \newcommand{\beq}{\begin{equation}}
61: \newcommand{\eeq}{\end{equation}}
62: 
63: %
64: %
65: % mathbb-Zeichen
66: %
67: \renewcommand{\AA}{{\mathbb{A}}}
68: \newcommand{\BB}{{\mathbb{B}}}
69: \newcommand{\CC}{{\mathbb{C}}}
70: \newcommand{\DD}{{\mathbb{D}}}
71: \newcommand{\EE}{{\mathbb{E}}}
72: \newcommand{\FF}{{\mathbb{F}}}
73: \newcommand{\GG}{{\mathbb{G}}}
74: \newcommand{\HH}{{\mathbb{H}}}
75: \newcommand{\JJ}{{\mathbb{J}}}
76: \newcommand{\KK}{{\mathbb{K}}}
77: \newcommand{\LL}{{\mathbb{L}}}
78: \newcommand{\MM}{{\mathbb{M}}}
79: \newcommand{\NN}{{\mathbb{N}}}
80: \newcommand{\OO}{{\mathbb{O}}}
81: \newcommand{\PP}{{\mathbb{P}}}
82: \newcommand{\QQ}{{\mathbb{Q}}}
83: \newcommand{\RR}{{\mathbb{R}}}
84: \renewcommand{\SS}{{\mathbb{S}}}
85: \newcommand{\TT}{{\mathbb{T}}}
86: \newcommand{\UU}{{\mathbb{U}}}
87: \newcommand{\VV}{{\mathbb{V}}}
88: \newcommand{\WW}{{\mathbb{W}}}
89: \newcommand{\XX}{{\mathbb{X}}}
90: \newcommand{\YY}{{\mathbb{Y}}}
91: \newcommand{\ZZ}{{\mathbb{Z}}}
92: %
93: %
94: % zusaetzlich fuer Clf-Teil:
95: %
96: %
97: \DeclareMathOperator{\Ad}{Ad}
98: \DeclareMathOperator{\diag}{diag}
99: \DeclareMathOperator{\Imag}{Im}
100: \DeclareMathOperator{\Pol}{Pol}
101: \DeclareMathOperator{\tr}{tr}
102: \newcommand{\sinfn}[1]{\sin(#1)}
103: \newcommand{\cosfn}[1]{\cos(#1)}
104: \newcommand{\arcsinfn}[1]{\arcsin(#1)}
105: \newcommand{\sinfnpot}[2]{\sin^{#1}(#2)}
106: \newcommand{\cosfnpot}[2]{\cos^{#1}(#2)}
107: \newcommand{\tanfn}[1]{\tan(#1)}
108: \newcommand{\AC}{\mr{AC}}
109: \newcommand{\cdic}{C^\infty_\mr{ev}[0,\pi]}
110: \newcommand{\mc}[1]{\mathcal{#1}}
111: \newcommand{\mf}[1]{\mathfrak{#1}}
112: \newcommand{\mr}[1]{\mathrm{#1}}
113: \newcommand{\OPg}{{\mr P}}
114: \newcommand{\Hi}{{\mc H}}
115: \newcommand{\scale}{\beta}
116: \newcommand{\ipro}{\lrcorner}
117: \newcommand{\comment}[1]{}
118: \newcommand{\verweis}[1]{}
119: \newcommand{\todo}[1]{}
120: \newcommand{\ddx}{\frac{\mr d}{\mr d x}}
121: \newcommand{\ddxx}{\frac{\mr d^2}{\mr d x^2}}
122: \newcommand{\ddt}{\frac{\mr d}{\mr d t}}
123: \newcommand{\ddtn}{\left.\frac{\mr d}{\mr d t}\right|_{t=0}}
124: \newcommand{\ddsn}{\left.\frac{\mr d}{\mr d s}\right|_{s=0}}
125: \newcommand{\SU}{\mr{SU}}
126: \newcommand{\su}{\mr{su}}
127: \newcommand{\SL}{\mr{SL}}
128: \newcommand{\inco}{\nu}
129: \newcommand{\group}{G}
130: \newcommand{\lieal}{{\mf g}}
131: \newcommand{\cfg}{{\mc X}}
132: \newcommand{\pha}{{\mc P}}
133: \newcommand{\vol}{\mr{vol}}
134: \newcommand{\Ann}{\mr{Ann}}
135: \newcommand{\ket}[1]{|#1\rangle}
136: \newcommand{\bra}[1]{\langle#1|}
137: \newcommand{\braket}[2]{\langle#1|#2\rangle}
138: \newcommand{\ve}{\varepsilon}
139: \newcommand{\vf}{{\mathfrak X}}
140: \newcommand{\vp}{\varphi}
141: \newcommand{\vr}{\varrho}
142: \newcommand{\wh}{\widehat}
143: \newcommand{\ctg}{\mr T^\ast}
144: \newcommand{\tg}{\mr T}
145: \newcommand{\rref}[1]{{\rm \ref{#1}}}
146: \newcommand{\cl}{\mr{c}}
147: \newcommand{\ol}[1]{\overline{#1}}
148: 
149: 
150: 
151: \newcommand{\linie}[3]{\put(#1){\line(#2){#3}}}
152: \newcommand{\marke}[3]{\put(#1){\put(0.05,0.1){\makebox(-0.1,-0.2)[#2]{$#3$}}}}
153: 
154: 
155: 
156: 
157: 
158: %----------------------------------------------------------------------------
159: %------------------------- Ende des Vorspanns -------------------------------
160: %----------------------------------------------------------------------------
161: %
162: 
163: 
164: 
165: \begin{document}
166: 
167: 
168: 
169: \title{\bf On the algebra of quantum observables for a certain
170: gauge model}
171: 
172: \author{
173:     G.~Rudolph and M.~Schmidt\\
174:     Institut f\"ur Theoretische Physik, Universit\"at Leipzig\\
175:     Augustusplatz 10/11, 04109 Leipzig, Germany\\
176:     }
177: 
178: \maketitle
179: 
180: 
181: \comment{
182: \vspace{2cm}
183: 
184: {\bf Keywords:}~
185: \\
186: 
187: {\bf MSC:}~ 70G65, 70S15
188: }
189: \vspace{2cm}
190: 
191: 
192: \begin{abstract}
193: 
194: \noindent
195: We prove that the algebra of observables of a certain gauge model is generated
196: by unbounded elements in the sense of  Woronowicz. The generators are
197: constructed from the classical generators of invariant polynomials by means of
198: geometric quantization.
199: 
200: \end{abstract}
201: 
202: \newpage
203: 
204: 
205: 
206: %\doublespacing
207: 
208: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
209: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
210: 
211: \section{Introduction}
212: \label{S-intro}
213: 
214: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
215: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
216: 
217: 
218: 
219: One of the fundamental structures in nonperturbative quantum field theory is the algebra of
220: observables and its representations. To construct the observable algebra and to find its
221: irreducible representations for a gauge theory is a complicated task, see
222: \cite{StrocchiWightman}, \cite{Fredenhagen} and \cite{Froehlich} for attempts
223: made in the seventies and eighties. Roughly speaking, one has to start with a
224: model of the field algebra carrying the action of the gauge group by
225: automorphisms, next one has to pass to the algebra of gauge invariant elements
226: and, finally, one has to factorize this algebra by an ideal generated by the
227: Gauss law. Unfortunately, standard charge superselection theory
228: \cite{DHR71,DHR74,DR90} does
229: not apply to genuine local gauge theories, see \cite{Buchholz82,Buchholz86}.
230: 
231: In order to separate functional analytical problems related to the mathematical
232: nature of quantum fields on continuous space time from those related to the
233: gauge structure, one is tempted to consider, in a first step, models
234: approximated on a finite lattice. In this context, we have constructed the
235: observable algebras and classified their irreducible representations both for
236: quantum electrodynamics \cite{KiRuTh,KiRuSl} and for quantum chromodynamics
237: \cite{KiRu02,KiRu05}. An additional challenge comes from the fact that on the
238: classical level there are nongeneric gauge orbit strata, see \cite{rsv:review}
239: for a review, which should have an impact on quantum level as well. In
240: \cite{hrs} we have shown that one can include these singularities by using the
241: concept of a costratified Hilbert space \cite{Hue:qr}. In the case of
242: chromodynamics, a full understanding of the observable algebra in terms of
243: generators and defining relations is still lacking, see \cite{JaKiRu} for
244: preliminary results. Generally speaking, gauge invariant generators are
245: polynomial invariants built from gauge and matter fields, corresponding to
246: classical generators of the algebra of polynomial invariants. Since typical
247: quantum observables are unbounded operators, one cannot hope to incorporate all
248: observables in a na\"ive sense into the observable algebra. Fortunately, there is
249: a suitable approach developed by Woronowicz in the nineties \cite{Woro}, which
250: makes it possible to say that a  given number of unbounded elements generates a
251: certain $C^*$-algebra, with the generators being affiliated with the algebra
252: under consideration in the $C^*$-sense. We remark that recently another
253: construction of a $C^\ast$-algebra of observables from unbounded physical
254: quantities was invented, see \cite{BuchholzGrundling}. In \cite{KiRu05} we have
255: shown that the field algebra of quantum chromodynamics is a $C^\ast$-algebra of
256: this type. In the present paper we prove that the algebra of observables of the
257: model studied in \cite{hrs} is also generated by unbounded operators in the
258: sense of Woronowicz. It is a challenge to extend this result to full chromodynamics on
259: a finite lattice in the future. In the case at hand, the generating operators
260: are the quantum counterparts of the generators of the algebra of real invariant
261: polynomials on the reduced phase space. This is an interesting fact in itself,
262: because in the Woronowicz theory there does not exist a general method to find a set of
263: generators of a given  $C^\ast$-algebra, nor does there exist a general method to find
264: the $C^\ast$-algebra generated by a given set of unbounded operators.
265: 
266: 
267: The paper is organized as follows: In Section \ref{S-model} we briefly present
268: the underlying classical model. In Section \ref{S-cobs} we present the algebra
269: of classical observables and its generators. Section \ref{S-qobs} is devoted to quantum observables. First
270: we quantize the classical generators using geometric quantization. Next, we discuss
271: the spectral properties of the quantized generators and the quantum
272: counterpart of the relation amongst the classical generators. Then, 
273: we construct the algebra of quantum observables and discuss the relations between our generators and the generators 
274: defined in \cite{KiRu05}. Finally, we comment on quantum dynamics and give an outlook.
275: 
276: 
277: 
278: 
279: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
280: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
281: 
282: \section{The model}
283: \label{S-model}
284: 
285: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
286: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
287: 
288: 
289: 
290: 
291: The model was explained in detail in \cite{hrs}. We recall the main facts.
292: The configuration space is the group manifold $G = \SU(2)$, acted upon by $G$
293: itself by inner automorphisms,
294: $$
295: g\cdot a = g a g^{-1}\,.
296: $$
297: The phase space is given by the cotangent bundle $\ctg G$, acted upon by the
298: lifted action. This action is symplectic and it possesses a natural equivariant
299: momentum mapping $\mu : \ctg G\to\mf g^\ast$, where $\mf g$ denotes the Lie
300: algebra of $G$. Thus, the phase space carries the structure of a Hamiltonian
301: $G$-manifold. We trivialize $\ctg G\cong G\times\mf g$ by means of an invariant
302: scalar product $\langle\cdot,\cdot\rangle$ on $\mf g$ and left translation. In
303: these coordinates, the lifted
304: action is given by
305: $$
306: g\cdot(a,X)
307:  =
308: (gag^{-1},\Ad(g)X)
309:  \,,~~~~~~
310: a\in G, X\in \mf g, g\in G\,,
311: $$
312: and the natural momentum mapping is given by
313:  \beq\label{G-momap}
314: \mu(a,X) = aXa^{-1} - X\,.
315:  \eeq
316: W.r.t.\ the natural decomposition
317:  \beq\label{G-decotg}
318: \tg_{(a,X)} (G\times\mf g)
319:  =
320: \tg_a G \oplus \tg_X\mf g\,,
321:  \eeq
322: tangent vectors at $(a,X)\in G\times \mf g$ can be written in the form
323:  \beq\label{G-vf}
324: (\mr L_a' A, B)\,,
325:  \eeq
326: where $A,B\in\mf g$ and $\mr L_a$ means left multiplication by $a$. In this
327: notation, the symplectic potential reads
328:  \beq\label{G-splpot}
329: \theta_{(a,X)}\big((\mr L_a' A,B)\big) = \langle X,A \rangle
330:  \,,~~~~~~
331: A,B\in\mf g\,,
332:  \eeq
333: and the symplectic form $\omega = -\mr d\theta$ is given by
334:  \beq\label{G-omega}
335: \omega_{(a,X)}\big((\mr L_a' A_1,B_1)\,,\,(\mr L_a' A_2,B_2)\big)
336:  =
337: \langle A_1,B_2 \rangle
338:  -
339: \langle A_2,B_1 \rangle
340:  +
341: \langle X,[A_1,A_2]\rangle\,.
342:  \eeq
343: The model can be interpreted as an $\SU(2)$-lattice gauge theory on a single
344: spatial plaquette in the Hamiltonian approach in the tree gauge,
345: or as $\SU(2)$-gauge theory on a space-time cylinder in the temporal gauge and
346: after reduction by the group of based gauge transformations, see
347: \cite{hrs}. In both cases, the classical Hamiltonian is given by
348:  \beq\label{G-Ham}
349:  \textstyle
350: H(a,X)
351:  =
352:  -
353: \frac{1}{2} |X|^2
354:  +
355: \frac{\inco}{2}
356:  \left(3 - \Re\,\tr(a)\right), \ a\in \group,\, X \in \lieal\,.
357:  \eeq
358: Let $T$ denote the subgroup of $G$ of diagonal matrices and $\mf t$ the
359: subalgebra of $\mf g$ of diagonal matrices. Let $W$ denote the Weyl group. It
360: acts on $T$ and $\mf t$ by permutation of entries. The reduced configuration
361: space $\cfg$ is given by the adjoint quotient
362: $$
363: \cfg = G/\Ad(G) \cong T/W\,.
364: $$
365: For general $\SU(n)$, this is an $(n-1)$-simplex.
366: For $\SU(2)$, the parameterization
367:  \beq\label{G-paramT}
368: \phi:\RR\to T
369:  \,,~~~~~~
370: x\mapsto\diag(\mr e^{\mr i x},\mr e^{-\mr i x})
371:  \eeq
372: induces a homeomorphism $[0,\pi]\cong\cfg$.
373: The reduced phase space is the zero level singular symplectic quotient
374: $$
375: \pha = \mu^{-1}(0)/G\,.
376: $$
377: Since, according to \eqref{G-momap}, $\mu(a,X)=0$ means that $a$
378: and $X$ commute and hence can be simultaneously diagonalized, $\pha$ may be
379: identified with the quotient $(T\times\mf t)/W$. For $\SU(2)$, this amounts to
380: the cylinder $\mr U(1)\times\RR$, factorized by reflection about the (virtual)
381: line connecting the points $(1,0)$ and $(-1,0)$, see Figure
382: \rref{Fig-canoe}. The space arising this way is known as the canoe. It coincides
383: with the phase space of a spherical pendulum, reduced at zero angular momentum
384: by the rotations about the vertical axis.
385: 
386: The reduced configuration space and the reduced phase space are
387: stratified by connected components of orbit type subsets,
388: $$
389: \cfg = \cfg_0 \cup \cfg_+ \cup \cfg_-
390:  \,,~~~~~~
391: \pha = \pha_0 \cup \pha_+ \cup \pha_-\,,
392: $$
393: where $\cfg_\pm$ consists of the class of $\pm\II$ and $\pha_\pm$ consists of
394: the class of the zero covector over $\pm\II$, see Figure \rref{Fig-canoe}.
395: 
396:  \begin{figure}
397: 
398: \unitlength1cm
399: 
400:  \begin{picture}(6,7.8)
401: 
402: \put(0.5,4.25){
403:  \put(1,0.27){
404:  \marke{4,1.82}{cl}{~-\!\II}
405:  \marke{0,1.82}{cr}{+\II~~}
406:  \marke{2,0}{tc}{T}
407:  \put(-0.135,1.82){\circle*{0.2}}
408:  \put(3.985,1.82){\circle*{0.2}}
409:  }
410:  \put(-1,0){
411:  \put(0,0.5){\epsfig{file=circle.eps,width=6cm,height=2cm}}
412:  }
413: } \put(0.5,1){
414:  \put(1,0.27){
415:  \marke{0,0}{tr}{\cfg_+~}
416:  \marke{2,0}{bc}{\cfg_0}
417:  \marke{4,0}{tl}{~\cfg_-}
418:  \marke{2,-0.5}{tc}{\cfg \cong T/W}
419:  \put(0,-0.06){\circle*{0.2}}
420:  \put(4,-0.06){\circle*{0.2}}
421:  \linie{0,-0.06}{1,0}{4}
422:  }
423: } \put(8.5,4.25){
424:  \put(1,0.27){
425:  \marke{4,1.82}{cl}{~(-\II,0)}
426:  \marke{0,1.82}{cr}{(+\II,0)~~}
427:  \marke{2,0}{tc}{T\times\mf t}
428:  \put(-0.135,1.82){\circle*{0.2}}
429:  \put(3.985,1.82){\circle*{0.2}}
430:  }
431:  \put(-1,0){
432:  \put(1.85,0.55){\epsfig{file=cylinder.eps,width=4.15cm,height=3cm}}
433:  }
434: } \put(8.5,1){
435:  \put(1,0.27){
436:  \marke{0,0}{tr}{\pha_+}
437:  \marke{4,0}{tl}{~\pha_-}
438:  \marke{4,1.75}{tl}{~\pha_0}
439:  \marke{2,-0.5}{tc}{\pha = \big(T\times\mf t\big)/W}
440:  \put(0,-0.06){\circle*{0.2}}
441:  \put(4,-0.06){\circle*{0.2}}
442:  }
443:  \put(-1,0){
444:  \put(0,0.02){\epsfig{file=canoe.eps,width=6cm,height=2cm}}
445:  }
446: }
447:  \end{picture}
448: \caption{\label{Fig-canoe}
449: Reduced configuration space $\cfg$ and reduced phase space
450: $\pha$ of the model together with their stratification by connected components
451: of orbit types}
452: 
453:  \end{figure}
454: 
455: 
456: 
457: 
458: 
459: 
460: 
461: 
462: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
463: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
464: 
465: \section{Classical observables}
466: \label{S-cobs}
467: 
468: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
469: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
470: 
471: 
472: 
473: The algebra of classical observables, as provided by standard singular
474: symplectic reduction at level $0$, is given by the quotient Poisson algebra
475: $$
476: \mc O_\cl = C^\infty(\ctg G)^G/V^G\,,
477: $$
478: where $V$ denotes the vanishing ideal of the closed subset $\mu^{-1}(0)$
479: \cite{ACG}. This algebra contains as a Poisson subalgebra the quotient
480: $\Pol(\ctg G)^G/V_{\mr{Pol}}^G$, where $\Pol(\ctg G)$
481: denotes the algebra of real polynomials on $\ctg G$ and $V_{\mr{Pol}}$
482: is the vanishing ideal of $\mu^{-1}(0)$ in this algebra. By definition, a
483: function on $\ctg G$ is polynomial if via the diffeomorphism $\ctg G\cong
484: G\times \mf g$ it corresponds to a function that is polynomial in the matrix
485: entries.
486: 
487: 
488: \bre\label{Rem-pnmalg}
489: 
490: One could also define polynomial functions on $\ctg G$ to be functions which via
491: the diffeomorphism $\ctg G \cong G^\CC$ correspond to elements of
492: $\Pol(G^\CC)$, i.e., to the functions on $G^\CC$ that are polynomial in the
493: matrix entries. This type of polynomial functions was used in
494: \cite{Hue:bedlewo}. Since polar decomposition is non-polynomial,
495: the two types of polynomial functions lead to completely different subalgebras
496: of $\mc O_\cl$ which intersect only in the constants.
497: 
498: \ere
499: 
500: 
501: The generators of $\Pol(\ctg G)^G$ are provided by invariant theory. For
502: $G=\SU(n)$ it is known that, via the diffeomorphism $\ctg G\cong G\times\mf g$,
503: a set of generators is provided by the real and imaginary parts of arbitrary
504: trace monomials of order $2^n-1$ in $a,a^\dagger \in G$ and $X\in\mf t$
505: \cite{Weyl}. By means of the fundamental trace identity and the Cayley-Hamilton
506: theorem this set of generators can be reduced considerably. For $\SU(2)$ there
507: remain 3 generators,
508:  \beq\label{G-generators}
509:  \textstyle
510: f_0(a,X) = \tr(a)
511:  \,,~~~~~~
512: f_1(a,X) = \frac{1}{2\beta^2} \tr(aX)
513:  \,,~~~~~~
514: f_2(a,X) = - \frac{1}{2\beta^2} \tr(X^2)
515:  \,.
516:  \eeq
517: Here $\beta$ is a scaling factor, defined by
518: $$
519:  \textstyle
520: \langle X,Y\rangle = -\frac{1}{2\beta^2} \tr(XY)
521:  \,,~~~~~~
522: X,Y\in\mf g\,.
523: $$
524: The functions $f_0$, $f_1$, $f_2$ are already real. For convenience, the
525: generators $f_1$ and $f_2$ have been rescaled by the scaling factor of the
526: invariant scalar product on $\mf g$. This way, $f_2$ is twice the kinetic
527: energy. In terms of the generators, the Hamiltonian \eqref{G-Ham} reads
528:  \beq\label{G-Ham-ivr}
529:  \textstyle
530: H = \frac{1}{2} f_2 + \frac{1}{2 g^2} (3 - f_0)\,.
531:  \eeq
532: I.e., up to a shift and up to a coupling parameter, $f_0$ is the potential
533: energy of the system.
534: 
535: 
536: \bre
537: 
538: For $G=\SU(n)$, $n\geq 3$, to cut to size the set of generators one
539: also has to make use of the fact that in the level set $\mu^{-1}(0)$, $a$ and
540: $X$ commute. This is not necessary for $\SU(2)$ though. I.e., here the set of
541: generators of invariant polynomials for the reduced phase space $\pha$ and for
542: the full quotient $\ctg G/G$ coincide.
543: 
544: \ere
545: 
546: 
547: The generators $f_0$, $f_1$, $f_2$ define a map $\pha\to\RR^3$, known as the
548: Hilbert map associated with this set of generators. It is common knowledge, see
549: e.g.\ \cite{Schwarz}, that the Hilbert map is a homeomorphism onto its image and
550: that the image is a semialgebraic subset, i.e., a subset defined by equalities
551: and inequalities. The defining equalities and inequalities for our case are
552: obtained as follows. Up to diagonal conjugation, an arbitrary element $(a,X)\in
553: G\times\mf g$ can be written
554: $$
555: a = \left[\begin{array}{cc} \alpha & 0 \\ 0 & \ol{\alpha} \end{array}\right]
556:  \,,~~~~~~
557: X = \left[\begin{array}{cc} \mr i x & z \\ -\ol z & -\mr i x \end{array}\right]
558:  \,,~~~~~~
559: \alpha\in\mr U(1)\,,~~x\in\RR\,,~~z\in\CC\,.
560: $$
561: Then
562: $$
563:  \textstyle
564: f_0(a,X) = 2\Re(\alpha)
565:  \,,~~~~~~
566: f_1(a,X) = - \frac{1}{\beta^2} x \Imag(\alpha)
567:  \,,~~~~~~
568: f_2(a,X) = \frac{1}{\beta^2}(x^2 + |z|^2)\,.
569: $$
570: Eliminating $x$ and $\alpha$ we obtain the relations
571:  \beq\label{G-fullquotient}
572: (\scale^2 f_2 - |z|^2)(4 - f_0^2) - 4 \beta^4 f_1^2 = 0
573:  \,,~~~~~~
574: \scale^2 f_2 - |z|^2 \geq 0\,,
575:  \eeq
576: where now $f_0$, $f_1$ and $f_2$ are interpreted as standard coordinates in
577: $\RR^3$. Hence, the image of the full quotient $\ctg G/G$ under the Hilbert map
578: coincides with the set of points $(f_0,f_1,f_2)\in\RR^3$ satisfying
579: \eqref{G-fullquotient} for some $z$. Since \eqref{G-fullquotient} implies
580: $4-f_1^2 \geq 0$ and since $|z|$ can take any nonnegative value, this subset is
581: given by the two inequalities
582: $$
583: f_2 (4 - f_0^2) - 4 \scale^2 f_1^2 \geq 0
584:  \,,~~~~~~
585: 4 - f_0^2 \geq 0\,.
586: $$
587: If $a$ and $X$ commute then, up to diagonal conjugation, $z=0$.
588: Hence, the reduced phase space $\pha\subseteq\ctg G/G$ corresponds to the subset
589: of \eqref{G-fullquotient} defined by $z=0$. Thus, this subset is given by the
590: relation
591:  \beq\label{G-rel}
592: f_2 (4 - f_0^2) - 4 \scale^2 f_1^2 = 0
593:  \eeq
594: and the inequality
595:  \beq\label{G-ineq}
596: f_2 \geq 0\,.
597:  \eeq
598: This subset is shown in Figure \rref{Fig-Himap}. It is of course a
599: concrete realization of the canoe, see Figure \rref{Fig-canoe}.
600: The image of the full quotient $\ctg G/G$ corresponds to this
601: subset together with the interior.
602:  \begin{figure}
603: 
604: \centering
605: 
606: \unitlength1cm
607: 
608:  \comment{
609:  \begin{picture}(12,4)
610:  \put(5.4,0.3){
611: \put(0,0){\vector(1,0){4}}
612: \put(0,0){\vector(0,1){4}}
613: \put(0,0){\vector(2,1){4}}
614:  \marke{4,0}{cl}{f_0}
615:  \marke{0,4}{bc}{f_2}
616:  \marke{4,2}{cl}{f_1}
617:  }
618:  }
619: 
620:  \begin{picture}(12,4.5)
621: \put(2.5,0.5){\epsfig{file=canoe2.eps,width=6cm,height=4cm}}
622:  \put(11,1.5){
623: \put(0,0){\vector(0,1){0.5}}
624: \put(0,0){\vector(4,-1){0.5}}
625: \put(0,0){\vector(2,1){0.5}}
626:  \marke{0,0.5}{br}{f_2}
627:  \marke{0.5,-0.125}{tl}{f_0}
628:  \marke{0.5,0.25}{cl}{f_1}
629:  }
630: 
631:  \end{picture}
632: 
633: \caption{\label{Fig-Himap}Image of $\pha$ under the Hilbert map defined by the generators
634: $f_0,f_1,f_2$. Equivalently, the semialgebraic subset of $\RR^3$ defined by
635: \eqref{G-rel} and \eqref{G-ineq}}
636: 
637:  \end{figure}
638: 
639: 
640: \bre
641: 
642: From Figure \rref{Fig-Himap} it is obvious that, topologically, $\pha$ is just a
643: copy of $\RR^2$. In fact, the Hilbert map defined by the natural generators of
644: the polynomial algebra $\Pol(G^\CC)^G$, see Remark \rref{Rem-pnmalg}, identifies
645: $\pha$ with the complex plane. However, as Poisson spaces, $\pha$ and $\CC$ are
646: distinct. This generalizes to $\SU(n)$. See \cite{Hue:bedlewo} for details.
647: 
648: \ere
649: 
650: 
651: Next, we compute the Hamiltonian vector fields $X_{f_i}$ associated with the
652: generators $f_i$ and the Poisson brackets between the generators. Let
653: $\OPg:\mr M_2(\CC)\to \mf g$ denote the orthogonal projection, i.e.,
654:  \beq\label{G-defP}
655:  \textstyle
656: \OPg(A) = \frac12(A-A^\dagger) - \frac{\mr i}{2} (\Imag\tr A) \II\,.
657:  \eeq
658: The defining equation for $X_{f_i}$ is
659: $
660: \omega(X_{f_i},Y) = - Y(f_i)
661: $
662: for all vector fields $Y$ on $G\times\mf g$. Writing
663:  \beq\label{G-Havf-AB}
664: (X_{f_i})_{(a,X)} = (\mr L_a' A_i,B_i)
665:  \eeq
666: and $Y_{(a,X)} = (\mr L_a' C,D)$ and using \eqref{G-omega} we obtain
667: $$
668:  \textstyle
669: \langle A_i,D \rangle - \langle B_i + [A_i,X],C \rangle
670:  =
671: - \ddtn f_i\left( a \mr e^{Ct} , X + t D \right)
672:  ~~~~~~
673: \forall~C,D\in\mf g\,.
674: $$
675: Evaluating the r.h.s.\ and solving for $A_i$ and $B_i$ we arrive at
676:  \begin{align}\label{G-Havf0}
677: A_0 & = 0 \,, & B_0 & = -2\scale^2 \OPg(a)\,,
678: \\ \label{G-Havf1}
679: A_1 & = \OPg(a)\,, & B_1 & = -\OPg(aX)\,,
680: \\ \label{G-Havf2}
681: A_2 & = -2X \,, & B_2 & = 0 \, .
682:  \end{align}
683: (Calculations are simplified by observing that $\tr(a)$ and $\tr(aX)$ are real,
684: hence the trace term in \eqref{G-defP} is absent in both cases.)
685: 
686: 
687: \ble\label{L-Havf}
688: 
689: The Hamiltonian vector fields $X_{f_0}$, $X_{f_1}$ and $X_{f_2}$ are
690: complete.
691: 
692: \ele
693: 
694: 
695: {\it Proof.}~ The flows of $X_{f_0}$ and $X_{f_2}$ are immediate:
696: $$
697: \Phi^{X_{f_0}}_t (a,X) = (a,X - 2\beta^2 \OPg(a) t)
698:  \,,~~~~~~
699: \Phi^{X_{f_2}}_t (a,X) = (a \mr e^{-2Xt},X)\,.
700: $$
701: They are defined for all $t\in\RR$. The flow equations for $X_{f_1}$ are
702: $$
703: \dot a = \mr L_a'\OPg(a)
704:  \,,~~~~~~
705: \dot X = -\OPg(aX)\,.
706: $$
707: Let $(a(0),X(0))$ be arbitrary but fixed initial values. Since $\dot a$ does
708: not depend on $X$ and since $G$ is compact, the solution $a(t)$ exists for all
709: $t\in\RR$. Hence, for $X_{f_1}$ to be complete it suffices that $|X(t)|$ be
710: finite for any $t\in\RR$. Consider the function $f(t) = |X(t)|^2$.
711: A brief computation using the Cayley-Hamilton theorem for $X$ yields $\ddt f(t)=
712: - \tr(a(t)) f(t)$. Then $\ddt f(t) \leq |\ddt f(t)| \leq 2 f(t)$. Thus, $f(t)$
713: is a nonnegative function whose derivative at $t$ is bounded by $2f(t)$. It
714: follows $f(t) \leq f(0) \mr e^{2t}$, hence the assertion.
715:  \qed
716:  \\
717: 
718: Finally, we calculate the Poisson brackets between the generators,
719: $$
720:  \textstyle
721: \{f_i,f_j\} = X_{f_i} f_j = \ddtn f_j(a\mr e^{A_it},X + tB_i)\,.
722: $$
723: Using the Cayley-Hamilton theorem to reduce powers of $a$ and $X$ we obtain
724:  \beq\label{G-Poibra}
725:  \textstyle
726: \{f_0,f_1\} = 2 - \frac12 f_0^2 \verweis{(SOA-16)}
727:  \,,~~~~~~
728: \{f_0,f_2\} = 4 \scale^2 f_1  \verweis{(SOA-16)}
729:  \,,~~~~~~
730: \{f_1,f_2\} = - f_0 f_2\,.  \verweis{(SOA-16)}
731:  \eeq
732: Since restriction of $f_i$ to the singular strata $\pha_\pm$ yields
733: $f_0|_{\pha_\pm} = \pm 2$ and $f_1|_{\pha_\pm} = f_2|_{\pha_\pm} = 0$,
734: $$
735: \{f_i,f_j\} |_{\pha_\pm} = 0\,.
736: $$
737: Thus, the Poisson structure of the reduced phase space $\pha$, given by
738: \eqref{G-Poibra}, reduces consistently to the (necessarily trivial) Poisson
739: structure on the singular strata $\pha_\pm$.
740: 
741: 
742: 
743: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
744: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
745: 
746: \section{Quantum observables}
747: \label{S-qobs}
748: 
749: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
750: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
751: 
752: 
753: 
754: 
755: 
756: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
757: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
758: 
759: \subsection{Quantization of classical generators}
760: 
761: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
762: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
763: 
764: 
765: 
766: 
767: The algebra of quantum observables will be constructed as follows. We quantize
768: the generators $f_i$ of the algebra of classical observables by means of
769: geometric quantization in the vertical polarization ('Schr\"odinger
770: quantization') on the unreduced phase space and subsequent reduction. As the
771: algebra
772: of quantum observables we will then take the $C^\ast$-algebra generated by these
773: operators in the sense of Woronowicz. The latter notion will be explained below.
774: We will loosely speak of the quantized generators as quantum observables as
775: well, although they do not belong to the algebra of quantum observables so
776: constructed.
777: 
778: The Hilbert space of Schr\"odinger quantization on $\ctg G$ can be identified
779: canonically with $L^2(G)$ with scalar product
780: $$
781: \braket{\psi_1}{\psi_2} = \frac{1}{\vol(G)} \int_G \ol{\psi_1}\psi_2 \mr
782: d a\,.
783: $$
784: Here $\mr d a$ stands for the volume form associated with the
785: bi-invariant Riemannian metric on $G$ defined by the invariant
786: scalar product on $\mf g$. By virtue of the isomorphism $G\times\mf g
787: \cong \tg G$, $f_2$ corresponds to the bi-invariant Riemannian metric defined
788: by the invariant scalar product on $\mf g$. Let $\Delta_G$ denote the Laplacian
789: associated with this metric. Since $G$ is closed, $\Delta_G$ is essentially
790: self-adjoint on the domain $C^\infty(G)$. The quantum observable $\hat f_2$
791: associated with $f_2$ is the unique self-adjoint extension of $- \hbar^2
792: \Delta_G$ see \cite[\S 9.7]{Woodhouse}. Thus, on the core $C^\infty(G)$,
793:  \beq\label{G-qobs2}
794: \hat f_2\psi = - \hbar^2 \Delta_G \psi
795:  \,,~~~~~~\psi\in C^\infty(G)\,.
796:  \eeq
797: In order to determine the quantum observables $\hat
798: f_0$ and $\hat f_1$ associated with the generators $f_0$ and $f_1$,
799: respectively, we have to recall the main steps in the construction of the
800: Hilbert space and the quantum observables in the Schr\"odinger quantization
801: \cite{Hall:Compact,Woodhouse}. Prequantization renders the complex line bundle
802: $L=\ctg G \times\CC$ with Hermitian form $h\big((a,z_1),(a,z_2)\big) = \ol{z_1}
803: z_2$ and connection $\nabla = \mr d + \theta \, , $ where $\theta$ denotes the symplectic potential 
804: of $\ctg G\, .$  Let $\pi:\ctg G \to G$ denote the canonical projection.
805: The vertical polarization is given by the vertical distribution $D\subseteq
806: \tg(\ctg G)$ induced by the fibres of $\ctg G$.
807:  \bigskip
808: 
809: {\it Hilbert space:}~ Consider the tautological complex line bundle $\kappa :=
810: \Lambda^n\Ann(D^\CC)$, where $n = \dim(G)$ and $\Ann$ denotes the annihilator of
811: $D^\CC$ in $(\tg^\CC)^\ast(\ctg G)$. The pull-back $\pi^\ast v$ of the volume
812: form $v$ associated with the Riemannian metric on $G$ defines a global section
813: in $\kappa$. Hence $\kappa$ is trivial and there exists a real line
814: bundle $\delta$ over $\ctg G$ such that $\kappa := (\delta\otimes\delta)^\CC$.
815: The bundle $\delta$ is called the half-form bundle associated with $D$. By
816: choosing a square root $\sqrt{\pi^\ast v}$ of $\pi^\ast v$ one obtains a global
817: nowhere vanishing section in $\delta$, hence $\delta$ is trivial, too.
818: Let $\Gamma_\mr{pol}(L\otimes\delta)$ denote the space of polarized sections in
819: $L\otimes\delta$. By definition, a section $\vp\otimes\nu$ in $L\otimes \delta$
820: is polarized if so are $\vp$ and $\nu$. A section $\vp\in \Gamma(L)$, viewed as
821: a function on $\ctg G$, is polarized if it is constant along the fibres, i.e.,
822: if $\vp = \pi^\ast \psi$ for some $\psi\in C^\infty(G)$. A section $\nu$ in
823: $\delta$ is polarized if $\nu\otimes\nu = \pi^\ast\alpha$ for some $n$-form
824: $\alpha$ on $G$. The Hilbert space $L^2_\mr{pol}(L\otimes\delta)$
825: is defined as the completion of $\Gamma_\mr{pol}(L\otimes\delta)$ w.r.t.\ the
826: norm defined by the following intrinsic scalar product:~ if $\vp_1\otimes\nu_1$
827: and $\vp_2\otimes\nu_2$ are polarized, $h(\vp_1,\vp_2)\nu_1\otimes\nu_2=
828: \pi^\ast\beta$ for some $n$-form $\beta$ on $G$. Then
829: $$
830: \braket{\vp_1\otimes\nu_1}{\vp_2\otimes\nu_2}
831:  :=
832: \frac{1}{\vol(G)} \int_G \beta\,.
833: $$
834: Finally, one can pass from half forms to functions on $G$ by observing that any
835: element of $\Gamma_\mr{pol}(L\otimes\delta)$ can be written in the form
836: $\vp\otimes\sqrt{\pi^\ast v}$ with $\vp = \pi^\ast \psi$ for some
837: $\psi\in C^\infty(G)$. By construction,
838: $$
839: \braket{\pi^\ast\psi_1\otimes\sqrt{\pi^\ast v}
840:  }{
841: \pi^\ast\psi_2\otimes\sqrt{\pi^\ast v}\rangle
842:  }
843:  =
844: \braket{\psi_1}{\psi_2}\,.
845: $$
846: Hence, the assignment
847:  \beq\label{G-ismSrqiz}
848: \psi\mapsto \pi^\ast\psi\otimes\sqrt{\pi^\ast v}
849:  \,,~~~~~~
850: \psi\in C^\infty(G)\,,
851:  \eeq
852: defines a unitary isomorphism from $L^2(G)$ onto $L^2_\mr{pol}(L\otimes\delta)$.
853:  \relax
854:  \bigskip
855: 
856: {\it Quantization of polarized classical observables:}~ A
857: classical observable $f\in C^\infty(\ctg G)$ is polarized if the
858: Hamiltonian vector field $X_f$ associated with $f$ satisfies
859: $[X_f,\Gamma(D)]\subseteq\Gamma(D)$. The operator $\hat f$
860: associated with $f$ is then defined by
861:  \beq\label{G-qobs-vert0}
862: \hat f(\vp \otimes \nu)
863:  =
864:  \big(
865: (\mr i \hbar X_f + \theta(X_f) + f)\vp
866:  \big)
867: \otimes \nu
868:  +
869: \vp\otimes(\mr i \hbar \mc L_{X_f} \nu)
870:  \,,~~~~~~
871: \vp\otimes\nu \in\Gamma_\mr{pol}(L\otimes\delta)\,.
872:  \eeq
873: Here, $\mc L_X$ denotes the Lie derivative w.r.t.\ the vector
874: field $X$ on $\ctg G$, which is defined on sections of $\delta$ by
875: virtue of the Leibniz rule
876:  \beq\label{G-defLie}
877:  \textstyle
878: (\mc L_{X} \nu) \otimes\nu := \frac12 \mc L_{X} (\nu\otimes\nu)\,.
879:  \eeq
880: The first term in \eqref{G-qobs-vert0} contains the ordinary
881: quantization formula of Kostant and Souriau, whereas the second
882: term represents the half-form correction. If $X_f$ is complete,
883: $\hat f$ is essentially self-adjoint \cite{Woodhouse}. The
884: argument is as follows. For any polarized $f$, the flow of $X_f$
885: lifts to a flow on $L\otimes\delta$, where the lift to $\delta$ is
886: natural and the lift to $L$ is defined by the connection $\nabla$.
887: If $X_f$ is complete, the lifted flow induces a strongly continuous
888: 1-parameter group of unitary transformations on
889: $L^2_\mr{pol}(L\otimes\delta)$. The self-adjoint generator of this
890: group, which exists due to Stone's theorem, has the subspace
891: $\Gamma_\mr{pol}(L\otimes\delta)$ as a core and on this
892: core it is given by \cite{Woodhouse}:
893: $$
894: \hat f(\vp \otimes \nu)
895:  =
896:  \big(
897: (\mr i \hbar X_f + \theta(X_f))\vp
898:  \big)
899: \otimes \nu
900:  +
901: \vp\otimes(\mr i \hbar \mc L_{X_f} \nu)
902:  \,,~~~~~~
903: \vp\otimes\nu \in\Gamma_\mr{pol}(L\otimes\delta)\,.
904: $$
905: By adding the multiplication operator  by the real function $f$ we obtain  
906: $\hat f$ as an essentially self-adjoint operator. 
907: 
908: By virtue of the isomorphism between $L^2(G)$ and
909: $L^2_\mr{pol}(L\otimes\delta)$ defined by \eqref{G-ismSrqiz},
910: $\hat f$ is mapped to an essentially self-adjoint operator on 
911: $L^2(G)$ which will be denoted by $\hat f$ as well. According to 
912: \eqref{G-qobs-vert0}, the defining equation for this operator is
913:  \beq\label{G-qobs-vert}
914: \pi^\ast(\hat f \psi) \otimes \sqrt{\pi^\ast v}
915:  =
916:  \big(
917: (\mr i \hbar X_f + \theta(X_f) + f) \pi^\ast\psi
918:  \big)
919:  \otimes\sqrt{\pi^\ast v}
920:  +
921: \pi^\ast\psi\otimes(\mr i \hbar \mc L_{X_f} \sqrt{\pi^\ast v})\,,
922:  \eeq
923: where $\psi\in C^\infty(G)$.
924:  \bigskip
925: 
926: 
927: 
928: {\it Quantization of $f_0$ and $f_1$:}~ We check that $f_0$ and $f_1$ are
929: polarized. Since in the decomposition \eqref{G-decotg}, elements
930: of $\Gamma(D)$ are characterized by having zero first component,
931: it suffices to take the commutator of $X_{f_i}$ with the constant
932: vector fields $(0,B)$ where $B\in\mf g$. Since, according to
933: \eqref{G-Havf0} and \eqref{G-Havf1}, the first components of
934: $X_{f_0}$ and $X_{f_1}$ do not depend on the momentum variable
935: $X$, only their second components can contribute to the commutator
936: $[X_{f_i},(0,B)]$. Since $\Gamma(D)$ is integrable, then
937: $[X_{f_i},(0,B)]\in\Gamma(D)$, $i=0,1$.
938: 
939: Next, we determine $\hat f_0$ and $\hat f_1$ using \eqref{G-qobs-vert}.
940: For $f = f_0$, \eqref{G-splpot} and \eqref{G-Havf0}
941: yield $\theta(X_{f_0}) = 0$ as well as $X_{f_0} \pi^\ast \psi= 0$ and $\mc
942: L_{X_{f_0}} \pi^\ast v = 0$, hence $\mc L_{X_{f_0}}\sqrt{\pi^\ast v} = 0$. Thus,
943: \eqref{G-qobs-vert} yields
944:  \beq\label{G-qobs0}
945:  \textstyle
946: \hat f_0 \psi = f_0\psi
947:  \,,~~~~~~
948: \psi\in C^\infty(G)\,,
949:  \eeq
950: where on the r.h.s., $f_0$ is viewed as a function on $G$ rather than on $\ctg
951: G$. As $G$ is compact, $f_0$ is bounded, hence $\hat f_0$ extends to a bounded
952: self-adjoint operator on $L^2(G)$, which will be denoted by the same symbol.
953: 
954: For $f = f_1$, \eqref{G-splpot} and \eqref{G-Havf1} yield
955: $
956: \theta_{(a,X)}(X_{f_1})
957:  =
958: \langle X , \OPg(a) \rangle
959:  =
960: - \frac{1}{2\scale^2} \tr(Xa)
961:  =
962: - f_1(a,X)\,.
963: $
964: Hence,
965:  \beq\label{G-f1-1}
966: \theta\big(X_{f_1}\big) + f_1 = 0\,.
967:  \eeq
968: Furthermore, we observed before that the first component of $X_{f_1}$ in
969: the decomposition \eqref{G-decotg} does not depend on the momentum variable $X$.
970: Hence, this component defines a vector field $Y_{f_1}$ on $G$. According to
971: \eqref{G-Havf1},
972:  \beq\label{G-Yi}
973: \big(Y_{f_1}\big)_a = \mr L_a' \OPg(a)
974:  \,,~~~~~~
975: a\in G\,.
976:  \eeq
977: By construction,
978:  \begin{align} \label{G-f1-2}
979: X_{f_1}\pi^\ast\psi & = \pi^\ast\big(Y_{f_1}\psi\big)\,,
980: \\ \label{G-f1-3}
981: \mc L_{X_{f_1}}\pi^\ast v & = \pi^\ast \big(\mc L_{Y_{f_1}} v\big)\,.
982:  \end{align}
983: A straightforward computation yields
984:  \beq\label{G-f1-4}
985: \mc L_{Y_{f_1}} v = \frac{3}{2} f_0 \, v\,,
986:  \eeq
987: see the appendix. Then \eqref{G-defLie} and \eqref{G-f1-3} yield
988: $\mc L_{X_{f_1}} \sqrt{\pi^\ast v} = \frac{3}{4} f_0 \, \sqrt{\pi^\ast v}$.
989: Plugging in this as well as \eqref{G-f1-1} and \eqref{G-f1-2} into
990: \eqref{G-qobs-vert} we arrive at
991:  \beq\label{G-qobs1}
992:  \textstyle
993: \hat f_1\psi = \mr i \hbar \left(Y_{f_1} + \frac34 \hat f_0\right)\psi
994:  \,,~~~~~~\psi\in C^\infty(G)\,.
995:  \eeq
996: Since according to Lemma \rref{L-Havf}, the Hamiltonian vector field $X_{f_1}$ 
997: is complete, $\hat f_1$ is essentially self-adjoint. From now on, $\hat f_1$
998: will denote the self-adjoint extension. Thus, all the operators $\hat f_0$,
999: $\hat f_1$ and $\hat f_2$ are self-adjoint and have $C^\infty(G)$ as a common
1000: invariant core. 
1001: \bre
1002: 
1003: Consider the operator $\mr i\hbar Y_{f_1}$ on $C^\infty(G)$. Since $\hat f_1$
1004: and $\hat f_0$ are symmetric, $\mr i\hbar Y_{f_1} = \hat f_1  - \mr
1005: i\hbar\frac34\hat f_0$ is not. Thus, the term $\mr i\hbar \frac34 \hat f_0$,
1006: playing the role of the half-form correction in the quantization of $f_1$, can
1007: be characterized as the unique purely imaginary multiplication operator which
1008: has to be added to the 'na\"ive quantization' $\mr i\hbar Y_{f_1}$ of $f_1$ in
1009: order to obtain a symmetric operator.
1010: 
1011: \ere
1012: 
1013: Finally, by reduction after quantization we arrive at the
1014: Hilbert space $L^2(G)^G$ of $G$-invariant elements. Since the functions
1015: $f_i$ are $G$-invariant, by restriction, the operators $\hat f_i$
1016: define self-adjoint operators on $L^2(G)^G$ which will be denoted by
1017: the same symbols. The subspace $C^\infty(G)^G$ is a common invariant core for
1018: these operators.
1019: 
1020: 
1021: 
1022: 
1023: An orthonormal basis in $L^2(G)^G$ is
1024: provided by the real characters $\chi_n$, where $n=0,1,2,\dots$ is twice the
1025: spin and labels the irreducible representations of $G$. To have the
1026: formulae in the following proposition valid for all $n$, let $\chi_{-1} = 0$.
1027: 
1028: 
1029:  \bpr\label{P-qobs-bs}
1030: 
1031: In the basis of characters, the quantum observables $\hat f_i$ are
1032: given by
1033:  \begin{eqnarray}\label{G-qobs-bs-0}
1034:  \textstyle
1035: \hat f_0 \chi_n
1036:  & = &
1037:  \textstyle
1038: \chi_{n+1} + \chi_{n-1}\,,
1039: \\ \label{G-qobs-bs-1}
1040:  \textstyle
1041: \hat f_1 \chi_n
1042:  & = &
1043:  \textstyle
1044: \mr i \hbar \left(\frac{2n+3}{4} \chi_{n+1}
1045: - \frac{2n+1}{4}\chi_{n-1}\right)\,,
1046:  \textstyle
1047: \\ \label{G-qobs-bs-2}
1048: \hat f_2 \chi_n
1049:  & = &
1050:  \textstyle
1051: \hbar^2 \scale^2 n(n+2) \chi_n\,.
1052:  \end{eqnarray}
1053: Accordingly, their matrix elements are
1054:  \begin{eqnarray*}
1055: (\hat f_0)_{nm}
1056:  & = &
1057:  \textstyle
1058: \delta_{n\,m+1} + \delta_{n\,m-1}\,,
1059: \\
1060: (\hat f_1)_{nm}
1061:  & = &
1062:  \textstyle
1063: \mr i\hbar \left(\frac{2m+3}{4}
1064: \delta_{n\,m+1} - \frac{2m+1}{4} \delta_{n\,m-1}\right)\,,
1065: \\
1066: (\hat f_2)_{nm}
1067:  & = &
1068:  \textstyle
1069: \hbar^2 \beta^2 \frac{n(n+2)}{2} \delta_{nm}\,.
1070:  \end{eqnarray*}
1071:  \verweis{(SQMG-244)}%
1072: 
1073: \epr
1074: 
1075: 
1076: {\it Proof.}~ As $\hat f_2$ is the negative of the Laplacian on
1077: $G$, the formula for $\hat f_2$ is standard. As $\hat f_0$ is
1078: multiplication by $f_0$ and $f_0$ is the character of the
1079: fundamental representation, the formula for $\hat f_0$ reflects
1080: the ordinary reduction formula for tensor products. For $\hat
1081: f_1$, it suffices to determine $(\hat f_1\chi_n)(a)$ for $a\in T$.
1082: Write $a = \diag(\alpha,\ol\alpha)$ with $\alpha\in\mr U(1)$. Then
1083: $$
1084: \chi_n(a)
1085:  =
1086: \alpha^n + \alpha^{n-2} + \cdots + \alpha^{-n}\,.
1087: $$
1088: We compute
1089:  $
1090:  \textstyle
1091: \left(Y_{f_1}\right)_a \chi_n
1092:  =
1093: \ddtn \chi_n\left( a \mr e^{\OPg(a)t} \right)
1094:  $.
1095: Since
1096:  $
1097: a \mr e^{\OPg(a)t} = \diag\left(\alpha \mr e^{\frac12(\alpha
1098: - \ol\alpha)t} , \ol\alpha \mr e^{-\frac12(\alpha -
1099: \ol\alpha)t}\right)
1100:  $,
1101: we have
1102: $$
1103: \chi_n\left( a \mr e^{\OPg(a)t} \right)
1104:  =
1105: \alpha^n \mr e^{\frac n 2(\alpha - \ol\alpha)t}
1106:  +
1107: \alpha^{n-2} \mr e^{\frac {n-2} 2(\alpha - \ol\alpha)t}
1108:  + \cdots +
1109: \alpha^{-n} \mr e^{-\frac n 2 (\alpha - \ol\alpha)t}\,.
1110: $$
1111: Taking the derivative and sorting by powers of $\alpha$ we obtain
1112: $$
1113:  \textstyle
1114: \left(Y_{f_1}\right)_a \chi_n
1115:  =
1116: \frac n 2 \alpha^{n+1} - \alpha^{n-1} - \alpha^{n-3}
1117:  - \cdots -
1118: \alpha^{-n+1} + \frac n 2 \alpha^{-n-1}
1119:  =
1120: \frac n 2 \chi_{n+1}(a) - \frac {n+2} 2 \chi_{n-1}(a)\,.
1121: $$
1122: Combining this with \eqref{G-qobs-bs-0} we arrive at \eqref{G-qobs-bs-1}.
1123:  \qed
1124:  \bigskip
1125: 
1126: 
1127: 
1128: 
1129: \bre
1130: 
1131: Composition of the trivialization $\ctg G\cong G\times \mf g$ with
1132: the inverse of the polar decomposition on $G^\CC$ yields a natural
1133: diffeomorphism $\ctg G\cong G^\CC$. In our situation, $G^\CC = \SL(2,\CC)$. By
1134: virtue of this diffeomorphism, the complex structure of $G^\CC$ and the
1135: symplectic structure of $\ctg G$ combine to a K\"ahler structure. Therefore, in
1136: addition to the vertical polarization defined by the fibres, $\ctg G$ carries a
1137: canonical K\"ahler polarization defined by the K\"ahler structure. For
1138: quantization in this polarization (K\"ahler quantization) on a general compact
1139: Lie group see \cite{Hall:Compact}. The Hilbert space $\mc H_{\text{K\"ahler}}$
1140: of K\"ahler quantization consists of holomorphic function on $G^\CC$ which are
1141: square-integrable w.r.t.\ a certain measure. Since the elements of $\mc
1142: H_{\text{K\"ahler}}$ are true functions rather than classes of functions, it is
1143: this space on which one constructs the costratied Hilbert space structure that
1144: implements the stratification of the reduced phase space on the level of the
1145: quantum theory, see \cite{Hue:qr} and \cite{hrs}.
1146: There exists a natural unitary isomorphism between the Hilbert spaces of the
1147: Schr\"odinger and the K\"ahler quantization ('generalized Bargmann-Segal
1148: transformation'). For general compact $G$, this isomorphism was first given in
1149: \cite{Hall:Compact} in terms of the heat kernel on $G^\CC$. Later on, in
1150: \cite{Hue:holopewe} a Peter-Weyl theorem for the Hilbert space of K\"ahler
1151: quantization was proved and it was used to show that the isomorphism between the
1152: two Hilbert spaces can also be obtained by matching irreducible components of
1153: the standard $G\times G$-representations on these two Hilbert spaces. For the
1154: subspaces of invariants this implies that here the unitary isomorphism is given
1155: by mapping each $\chi_n$ to the corresponding character on $G^\CC$, normalized
1156: w.r.t.\ the specific scalar product on $\mc H_{\text{K\"ahler}}$.
1157: 
1158: \ere
1159: 
1160: 
1161: 
1162: 
1163: 
1164: 
1165: 
1166: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1167: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1168: 
1169: \subsection{Domains, eigenvalues and spectra of the quantized generators}
1170: 
1171: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1172: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1173: 
1174: 
1175: 
1176: To investigate the operators $\hat f_i$ we pass from $L^2(G)^G$ to $L^2[0,\pi]$
1177: as follows. Let $C^\infty[0,\pi]$ denote the Whitney
1178: smooth functions on the closed interval $[0,\pi]$ (i.e., smooth functions on the
1179: open interval $]0,\pi[$ that can be smoothly extended outside $[0,\pi]$). Take
1180: the parameterization $\phi$ of the subgroup $T\subseteq G$ of diagonal matrices,
1181: see \eqref{G-paramT}, and define a map $\Gamma : C^\infty(G) \to
1182: C^\infty[0,\pi]$ by
1183: $$
1184: (\Gamma \psi)(x) = \sqrt2 ~ \sinfn x ~ \psi\big(\phi(x)\big)
1185:  \,,~~~~~~
1186: x\in[0,\pi]\,.
1187: $$
1188: 
1189: 
1190: \ble\label{L-L2}
1191: 
1192: $\Gamma$ extends to a unitary Hilbert space isomorphism $L^2(G)^G\to
1193: L^2[0,\pi]$.
1194: 
1195: \ele
1196: 
1197: 
1198: {\it Proof.}~ We have to check that $\Gamma$ is isometric and that its image is
1199: dense in $L^2[0,\pi]$. Let $\psi,\vp\in L^2(G)^G$. From the Weyl integration
1200: formula we know that
1201: $$
1202:  \textstyle
1203: \int_G \overline\psi \vp \mr da
1204:  =
1205: \int_T \overline\psi \vp v\mr dt\,,
1206: $$
1207: where $\mr da$ and $\mr dt$ denote the Haar measures on $G$ and $T$,
1208: respectively, and $v$ is a density function that accounts
1209: for the volume of the orbits under inner automorphisms of $G$. For $G = \SU(2)$,
1210: $$
1211:  \textstyle
1212: \phi^\ast(v\mr d t) = \frac{\vol(G)}{\pi} \sinfnpot 2 x\mr dx\,.
1213: $$
1214: Hence,
1215:  \begin{align*}
1216: \braket{\psi}{\vp}
1217:  & =
1218:  \textstyle
1219: \frac{1}{\vol(G)} \int_T \overline{\psi} \, \vp v\, \mr dt
1220: \\
1221:  & =
1222:  \textstyle
1223: \frac{1}{\vol(G)} \int_{-\pi}^\pi \phi^\ast
1224:  \left(\overline{\psi} \, \vp \, v \, \mr dt\right)
1225: \\
1226:  & =
1227:  \textstyle
1228: \frac{1}{\pi} \int_{-\pi}^\pi
1229:  ~
1230: \overline{\psi\big(\phi(x)\big)}
1231:  ~
1232: \vp\big(\phi(x)\big)
1233:  ~
1234: \sinfnpot 2 x \, \mr dx\,.
1235: \\
1236:  & =
1237:  \textstyle
1238: \frac{1}{2\pi} \int_{-\pi}^\pi
1239:  ~
1240: \overline{(\Gamma\psi)(x)}
1241:  ~
1242: (\Gamma\vp)(x) ~ \mr dx\,.
1243:  \end{align*}
1244: Since $\psi$ and $\vp$ are invariant under inner automorphisms, $\Gamma\psi$ and
1245: $\Gamma\vp$ are invariant under reflection $x\mapsto -x$. Hence, the integral
1246: over $[-\pi,\pi]$ gives twice the integral over $[0,\pi]$. This shows that
1247: $\Gamma$ is isometric. Since the image of $\Gamma$ contains the smooth
1248: functions with compact support inside the open interval $]0,\pi[$, it is dense
1249: in $L^2[0,\pi]$.
1250:  \qed
1251: \\
1252: 
1253: 
1254: We will need the image of $C^\infty(G)^G$ under $\Gamma$. Let $\cdic$ denote the
1255: subspace of $C^\infty[0,\pi]$ of functions whose even order derivatives
1256: $\psi^{(2n)}$, $n=0,1,2,,\dots$, vanish in $0$ and $\pi$:
1257: $$
1258: \cdic = \{\psi\in C^\infty[0,\pi] : \psi^{(2n)} (0) = \psi^{(2n)}
1259: (\pi) = 0\,,~n=0,1,2,\dots\}\,.
1260: $$
1261: 
1262: 
1263: \ble\label{L-core}
1264: 
1265: $\Gamma\big(C^\infty(G)^G\big) = \cdic$.
1266: 
1267: \ele
1268: 
1269: 
1270: {\it Proof.}~ First, let $\vp\in C^\infty(G)^G$. Define $\tilde\vp\in
1271: C^\infty(\RR)$ by $\tilde\vp(x) := \vp\big(\phi(x)\big)$. Then $\Gamma(\vp)(x) =
1272: \sqrt 2 \sinfn x \tilde\vp(x)$ and the iterated Leibniz rule yields for the
1273: derivative of order $2n$
1274:  \begin{align*}
1275:  \textstyle
1276: \Gamma(\vp)^{(2n)}(x)
1277:  & =
1278:  \textstyle
1279: \sqrt2
1280:  \Big\{
1281: \sum_{k=0}^n (-1)^{n-k} {2n \choose 2k} \sinfn x \tilde\vp^{(2k)}(x)
1282: \\
1283:  & \hspace{5cm} +
1284:  \textstyle
1285: \sum_{k=0}^{n-1} (-1)^{n-k-1} {2n \choose 2k+1} \cosfn x
1286: \tilde\vp^{(2k+1)}(x)\,.
1287:  \Big\}
1288:  \end{align*}
1289: By construction, the function $\tilde\vp$ is $2\pi$-periodic and has even
1290: parity, i.e., $\tilde\vp(-x) = \tilde\vp(x)$. Hence, the derivative $\vp^{(k)}$
1291: is $2\pi$-periodic and has even parity for even $k$ and odd parity for odd $k$.
1292: It follows $\vp^{(2n+1)} (0) = \vp^{(2n+1)} (\pi) = 0$ and hence
1293: $\Gamma(\vp)^{(2n)}(0) = \Gamma(\vp)^{(2n)}(\pi) = 0$, for any $n$.
1294: 
1295: Conversely, let $\psi\in\cdic$. Since $\psi(0) = \psi(\pi) = 0$, we can extend
1296: $\psi$ to a well-defined function on the whole of $\RR$ by setting $\psi(-x) =
1297: -\psi(x)$ and $\psi(x+2\pi m) = \psi(x)$, $x\in[0,\pi]$, $m$ an integer. Then
1298: for any $x\in\RR\setminus2\pi\ZZ$, any $k=0,1,2,\dots$ and any $m\in\ZZ$ there
1299: holds $\psi^{(k)}(-x) = -(-1)^k \psi^{(k)}(x)$ and $\psi^{(k)}(x+2\pi m) =
1300: \psi^{(k)}(x)$. In addition, $\lim_{x\to2\pi m} \psi^{(2k)}(x) = 0$. This
1301: implies that derivatives of $\psi$ of arbitrary order are continuous in $x =
1302: 2\pi m$, hence $\psi$ is smooth. Now define a function $\tilde\vp$ on
1303: $\RR\setminus2\pi\ZZ$ by $\tilde\vp(x) = \frac{1}{\sqrt
1304: 2}\,\frac{\psi(x)}{\sinfn x}$. We
1305: claim that $\tilde\vp$ extends to a smooth function on the whole of $\RR$. To
1306: see this, it suffices to show smoothness in $x=0$ and $x=\pi$. We give the
1307: argument for $x=0$ only; the case $x=\pi$ is analogous. Since the sine function
1308: is a local diffeomorphism in a neighbourhood of $x=0$, $\tilde\vp$ is smooth in
1309: $0$ iff so is $\tilde\vp\circ\arcsin$. Denote $f(x) = \psi\big(\arcsinfn
1310: x\big)$. Then $f$ is smooth in a neighbourhood of $x=0$ and
1311: $\tilde\vp\circ\arcsin(x) = \frac{1}{\sqrt 2}\,\frac{f(x)}{x}$. Hence,
1312: $$
1313:  \textstyle
1314: (\tilde\vp\circ\arcsin)^{(k)}(x)
1315:  =
1316: \frac{1}{\sqrt 2}\,\frac{1}{x^{n+1}}
1317:  ~
1318: \sum_{l=0}^k ~ (-1)^{n-k} \, {k \choose l} \, (k-l)! \, x^k \, f^{(k)}(x)\,.
1319: $$
1320: Since $f(0) = 0$, the r.h.s.\ yields an indefinite expression for $x\to 0$.
1321: The derivative of the enumerator is $x^k f^{(k+1)}(x)$. Hence, the rule
1322: of de l'Hospital yields
1323: $$
1324:  \textstyle
1325: \lim_{x\to 0}(\tilde\vp\circ\arcsin)^{(k)}(x)
1326:  =
1327: \frac{1}{\sqrt2} \, \frac{f^{(k+1)}(0)}{k+1}\,.
1328: $$
1329: This proves that $\tilde\vp$ extends to a smooth function on $\RR$. Since it is
1330: $2\pi$-periodic and has even parity by construction, there exists $\vp\in
1331: C^\infty(G)^G$ such that $\tilde\vp = \vp\circ\phi$. Then $\psi = \Gamma(\vp)$.
1332:  \qed
1333:  \\
1334: 
1335: 
1336: 
1337: 
1338: The operators $\hat f_i$ on $L^2(G)^G$ induce operators $\Gamma \hat f_i
1339: \Gamma^{-1}$ on $L^2[0,\pi]$. These induced operators will also be denoted by
1340: $\hat f_i$. We derive explicit expressions. Since $\Gamma \hat f_0 \Gamma^{-1}$
1341: is multiplication by the function $f_0\circ\phi$,
1342:  \beq\label{G-f0expr}
1343: \hat f_0\psi(x) = 2\cosfn{x} \, \psi(x)\,.
1344:  \eeq
1345: Let $\AC[0,\pi]$ denote the space of absolutely
1346: continuous functions and let
1347:  \begin{align*}
1348: \AC^1[0,\pi] & = \{\psi\in\AC[0,\pi] : \psi'\in L^2[0,\pi]\}\,,
1349:  \\
1350: \AC^2[0,\pi] & = \{\psi\in\AC^1[0,\pi] : \psi'\in\AC^1[0,\pi]\}\,.
1351:  \end{align*}
1352: 
1353: 
1354: \bpr\label{P-dom}~
1355: 
1356: The operator $\hat f_1$ has domain
1357: $
1358: \mr D(\hat f_1)
1359:  =
1360: \{\psi\in L^2[0,\pi] : \sinfn{x}\psi(x) \in \AC^1[0,\pi]\}
1361: $
1362: and is given by the expression
1363:  \beq\label{G-f1expr}
1364:  \textstyle
1365: \hat f_1 = \mr i \hbar \left(\ddx \sinfn{x} \,-\, \frac12\cosfn x\right)\,.
1366:  \eeq
1367: The operator $\hat f_2$ has domain
1368: $
1369: \mr D(\hat f_2)
1370:  =
1371: \{\psi\in L^2[0,\pi] : \psi \in \AC^2[0,\pi], \psi(0) = \psi(\pi) = 0\}
1372: $
1373: and is given by the expression
1374:  \beq\label{G-f2expr}
1375:  \textstyle
1376: \hat f_2 = - \hbar^2\beta^2\left(\ddxx + 1\right)\,.
1377:  \eeq
1378: The subspace $\cdic$ is a common invariant core for $\hat f_0$, $\hat
1379: f_1$, $\hat f_2$.
1380: 
1381: \epr
1382: 
1383: 
1384: \bre
1385: 
1386: For $\psi\in C^\infty[0,\pi]$, one has
1387:  \begin{align}\nonumber
1388:  \textstyle
1389: \hat f_1 \psi
1390:  & =
1391:  \textstyle
1392: \mr i \hbar \left(\ddx \sinfn{x} \,-\, \frac12\cosfn x\right) \psi(x)
1393: \\ \nonumber
1394:  & =
1395:  \textstyle
1396: \mr i \hbar \left(\sqrt{\sinfn{x}}\ddx\sqrt{\sinfn{x}}\right) \psi(x)
1397: \\ \label{G-f1variants}
1398:  & =
1399:  \textstyle
1400: \mr i \hbar \left(\sinfn{x}\ddx \,+\, \frac12\cosfn x\right) \psi(x)\,,
1401:  \end{align}
1402: whereas it is only the first of these three expressions that extends to the
1403: whole of $\mr D(\hat f_1)$.
1404: 
1405:  \ere
1406: 
1407: 
1408:  \bre
1409: 
1410: For general $\SU(n)$ the Hilbert space $L^2(G)^G$ can be realized as
1411: $L^2(\sigma^{n-1},v\rm dt)$, where $\sigma^{n-1}$ is the $(n-1)$-simplex (more
1412: concretely, a Weyl alcove in the Lie algebra $\mf g$) and $v\mr dt$ is an
1413: appropriate measure on $\sigma^{n-1}$. In \cite{Wren} it is proved
1414: that in this  realization a core for the group Laplacian $\Delta_G$
1415: is given by Neumann boundary conditions at the boundary of $\sigma^{n-1}$. In
1416: our situation, $L^2(\sigma^{n-1},v\mr dt)$ corresponds to $L^2([0,\pi],\sinfnpot
1417: 2 x \mr dx)$ and the core isolated in \cite{Wren} amounts to $\{\psi\in C^\infty[0,\pi]
1418: : \psi'(0) = \psi'(\pi) = 0\}$. By means of the isomorphism $\Gamma$, this core
1419: is mapped into $\{\psi\in C^\infty[0,\pi] : \psi(0) = \psi(\pi) = 0\}$. Thus, the
1420: assertion about the core for $\hat f_2$ in Proposition \rref{P-dom} is
1421: consistent with the result of \cite{Wren}.
1422: 
1423:  \ere
1424: 
1425: 
1426:  \bre
1427: 
1428: The domain of $\hat f_1$ is unusually large for a differential operator,
1429: it contains e.g.\ all smooth functions. This is due to the fact that, in $\hat
1430: f_1$, the derivative is combined with the sine function which destroys any
1431: information about the boundary values of the function whose derivative is taken.
1432: In particular, $C^\infty[0,\pi]$ may also be taken as a core for $\hat f_1$.
1433: 
1434:  \ere
1435: 
1436: 
1437:  \bre
1438: 
1439: Occasionally we will have to deal with the operator $\hat f_1^2$
1440: below. For further use we note that the domain of $\hat f_1^2$
1441: contains $\AC^2[0,\pi]$ as a proper subspace and that on
1442: $\AC^2[0,\pi]$,
1443:  \beq\label{G-f12expr}
1444:  \textstyle
1445: \hat f_1^2
1446:  =
1447: -\hbar^2
1448:  \left(
1449: \sinfn{x}\ddxx\sinfn{x} - \frac14\cosfnpot 2 x + \frac12
1450:  \right)\,.
1451:  \eeq
1452: 
1453:  \ere
1454: 
1455: 
1456: 
1457: {\it Proof.}~ The last statement follows from the fact that $C^\infty(G)^G$ is a
1458: common invariant core for $\hat f_0$, $\hat f_1$, $\hat f_2$ and Lemma
1459: \rref{L-core}.
1460: 
1461: First, consider $\hat f_2$. According to the general formula for the radial part
1462: of the Laplacian on a compact group, see \cite[\S II.3.4]{Helgason}, the
1463: restriction of $\hat f_2$ to $\cdic$ is given by the r.h.s.\ of
1464: \eqref{G-f2expr}. The assertion about the domain then follows by standard
1465: extension theory for the operator of second derivative. Since the r.h.s.\ of \eqref{G-f2expr} is well defined on
1466: $\AC^2[0,\pi]$, $\hat f_2$ is given by this expression on the whole of its
1467: domain.
1468: 
1469: Next, consider $\hat f_1$. According to \eqref{G-qobs1}, for $\psi\in \cdic$,
1470: $$
1471:  \textstyle
1472: \hat f_1\psi
1473:  \equiv
1474: \Gamma \hat f_1 \Gamma^{-1}\psi
1475:  =
1476: \mr i \hbar \Gamma \big(Y_{f_1} + \frac34 \hat f_0\big) \Gamma^{-1}\psi\,.
1477: $$
1478: According to \eqref{G-Yi} and \eqref{G-defP},
1479: $$
1480:  \textstyle
1481: \big(Y_{f_1}\psi\big)(x)
1482:  =
1483: \sqrt 2 \sinfn x \, \ddtn (\Gamma^{-1}\psi)
1484:  \left(
1485: \phi(x) \mr e^{\frac12\left(\phi(x) - \phi(x)^\dagger\right)t}
1486:  \right)\,.
1487: $$
1488: A brief computation shows
1489: $
1490: \phi(x) \mr e^{\frac12\left(\phi(x) - \phi(x)^\dagger\right)t}
1491:  =
1492: \phi(x + t \sinfn x)\,.
1493: $
1494: Hence,
1495: $$
1496:  \textstyle
1497: \big(Y_{f_1}\psi\big)(x)
1498:  =
1499: \sqrt 2 \sinfn x \, \ddtn \frac{\psi(x + t \sinfn x)}{\sqrt2\sin(x + t \sinfn x)}
1500:  =
1501: \psi'(x) \sinfn x - \psi(x) \cosfn x\,.
1502: $$
1503: Together with \eqref{G-f0expr} and \eqref{G-f1variants} this yields
1504: $
1505: \big(\hat f_1\psi\big)(x)
1506:  =
1507: \mr i \hbar \left(\ddx \sinfn{x} \,-\, \frac12\cosfn x\right) \psi(x)\,,
1508: $
1509: hence on $\cdic$, $\hat f_1$ is given by the r.h.s.\ of \eqref{G-f1expr}. Denote $D = \{\psi\in L^2[0,\pi] : \sinfn{x}\psi(x)
1510: \in \AC^1[0,\pi]\}$. Since the r.h.s.\ of \eqref{G-f1expr} is well-defined for
1511: all $\psi\in D$, $\hat f_1$ is given by this expression on the whole of $D$. It
1512: remains to show $\mr D(\hat f_1) = D$.
1513: 
1514: Let $A$ be defined by restriction of $\hat f_1$ to the core $\cdic$. Since $f_1$ is self-adjoint, 
1515: $$
1516: A^\dagger = \ol A{}^\dagger = \hat f_1^\dagger = \hat f_1 \, .
1517: $$
1518: Hence, it suffices to
1519: show $\mr D(A^\dagger) = D$. Let $\psi\in D$. Then $\sinfn{x}\psi(x)\in
1520: \AC^1[0,\pi]$, hence it has a derivative $(\sinfn{x}\psi(x))'\in L^1[0,\pi]$
1521: and $(\sinfn{x}\psi(x))'\in L^2[0,\pi]$ . Then
1522: $$
1523:  \textstyle
1524: \tilde\psi(x)
1525:  :=
1526: \mr i\hbar \big((\sinfn{x}\psi(x))' - \frac12\cosfn x\psi(x)\big) \in
1527: L^2[0,\pi]\,.
1528: $$
1529: For any $\vp\in \cdic$, integration by parts yields
1530:  \begin{align*}
1531:  \textstyle
1532: \braket{\tilde\psi}{\vp}
1533:  & =
1534:  \textstyle
1535: -\frac{\mr i \hbar}{\pi} \int_0^\pi
1536:  \ol{
1537: \big((\sinfn{x}\psi)' - \frac12\cosfn x\psi(x)\big)
1538:  }
1539: \vp(x) \, \mr dx
1540: \\
1541:  & =
1542:  \textstyle
1543: \frac{\mr i \hbar}{\pi}
1544:  \int_0^\pi
1545: \ol{\psi(x)}\big(\sinfn{x}\vp' + \frac12\cosfn x\vp(x)\big)
1546:   \, \mr dx
1547: \\
1548:  & =
1549:  \textstyle
1550: \braket{\psi}{A\vp}\,,
1551:  \end{align*}
1552: hence $\psi\in\mr D(A^\dagger)$. Conversely, let $\psi\in\mr D(A^\dagger)$. Then
1553: there exists $\tilde\psi\in L^2[0,\pi]$ such that $\braket{\psi}{A\vp} =
1554: \braket{\tilde\psi}{\vp}$ for all $\vp\in \cdic$. Write this equation in the
1555: form
1556: $$
1557:  \textstyle
1558: \int_0^\pi \ol{\sinfn{x}\psi(x)}
1559:  ~
1560: \mr i\ddx \vp(x)
1561:  ~
1562: \mr dx
1563:  =
1564: \int_0^\pi
1565:  \ol{\left(
1566: \frac1\hbar \tilde\psi
1567:  +
1568: \frac{\mr i}{2} \cosfn x\psi(x)
1569:  \right)}
1570:  ~
1571: \vp(x)
1572:  ~
1573: \mr dx
1574:  \,,~~~~~~
1575: \forall~~\vp\in \cdic\,.
1576: $$
1577: We conclude that $\sinfn{x}\psi(x)$ belongs to the domain of the adjoint of the restriction of $\mr i\ddx$ to the 
1578: subspace $\cdic \,. $ Since  $\cdic $ is a core for $\mr i\ddx$ and since the domain of the self-adjoint operator $\mr i\ddx$ is $\AC^1[0,\pi]$ it follows that  $\sinfn{x}\psi(x)\in\AC^1[0,\pi]$, i.e., $\psi\in D \, .$ 
1579: \qed
1580:  \\
1581:  
1582: 
1583: 
1584: Next, we discuss the eigenvalues and the spectra of the operators $\hat f_i$.
1585: According to \eqref{G-qobs-bs-2}, $\hat f_2$ has pure point spectrum,
1586: $$
1587: \sigma(\hat f_2) = \{\hbar^2n(n+2) : n=0,1,2,\dots\}
1588: $$
1589: and the characters form an orthonormal basis of eigenvectors.
1590: 
1591: 
1592: \bpr\label{P-spec}
1593: 
1594: The operators $\hat f_0$, $\hat f_1$ and $\hat f_1^2$ do not possess
1595: eigenvalues. Their spectra are
1596: $$
1597: \sigma(\hat f_0) = [-2,2]
1598:  \,,~~~~~~
1599: \sigma(\hat f_1) = \RR
1600:  \,,~~~~~~
1601: \sigma(\hat f_1^2) = [0,\infty[
1602:  \,.
1603: $$
1604: 
1605: \epr
1606: 
1607: 
1608: {\it Proof.}~ First, consider $\hat f_0$. The eigenvalue equation $(\hat f_0
1609: -\lambda)\psi = 0$ reads $(2\cosfn x-\lambda)\psi(x) = 0$, hence $\psi=0$ a.e.\ for any
1610: $\lambda\in\RR$. Thus, there are no eigenvalues. The assertion about the
1611: spectrum follows from the spectral mapping theorem.
1612: 
1613: Next, consider $\hat f_1$. According to \eqref{G-f1expr}, the eigenvalue
1614: equation amounts to the differential equation
1615:  \beq\label{G-evaleqf1}
1616:  \textstyle
1617: \big(\hat f_1 - \lambda\big) \psi(x)
1618:  =
1619: \big\{\mr i\hbar\big(\ddx\sinfn x - \frac12 \cosfn x\big) - \lambda\big\} \psi(x)
1620:  =
1621: 0
1622:  \eeq
1623: which on the open interval $]0,\pi[$ can be written in the form
1624: $$
1625:  \textstyle
1626: \mr i \hbar
1627:  \left\{
1628: \ddx
1629:  +
1630: \left(\frac{\mr i\lambda}{\hbar\sinfn x} - \frac12\cot(x)\right)
1631:  \right\}
1632:  ~
1633: (\sinfn{x}\psi(x))
1634:  =
1635: 0\,.
1636: $$
1637: For any $\lambda\in\RR$ the
1638: solution is
1639:  \beq\label{G-psil}
1640:  \textstyle
1641: \psi_\lambda(x)
1642:  =
1643: \frac{1}{\sqrt{2\hbar}}
1644:  ~
1645: \frac{\mr e^{-\frac{\mr i}{\hbar} \lambda \ln \tanfn{\frac x2}}
1646:  }{
1647: \sqrt{\sinfn x}
1648:  }\,.
1649: \eeq
1650: The particular choice of normalization will be justified below.
1651: Since neither of the functions $\psi_\lambda$ is square
1652: integrable, $\hat f_1$ does not have eigenvalues.
1653: 
1654: To determine the spectrum of $\hat f_1$, let $\lambda\in\RR$. If $\hat f_1-\lambda$ had a
1655: bounded inverse, there would exist $C>0$ such that $\|\psi\| = \|(\hat
1656: f_1-\lambda)^{-1}(\hat f_1-\lambda)\psi\|
1657: \leq C \|(\hat f_1-\lambda)\psi\|$ for any $\psi\in\mr D(\hat f_1)$. Thus, in
1658: order to show that $\lambda\in\sigma(\hat f_1)$ it
1659: suffices to construct a sequence $\psi_n$ in $\mr D(\hat f_1)$ such that
1660: $\frac{\|\psi_n\|}{\|(\hat f_1-\lambda)\psi_n\|} \to \infty$.
1661: Choose a smooth function $j$ on $\RR$ with support in the open interval
1662: $]-1,1[$ such that $0\leq j(x)\leq 1$ and
1663: $\int_{-\infty}^\infty j(x) \, \mr d x
1664: = 1$. Define $g_n(x) = n\int_{-\infty}^x \big\{ j(nx'-2) - j(n(x'-\pi)+2)
1665: \big\}\,\mr d x'$ and $\psi_n := g_n\psi_\lambda$. Since $g_n$ has
1666: support in $]0,\pi[$, $\psi_n\in L^2[0,\pi]$ and hence $\psi_n\in\mr D(\hat
1667: f_1)$. On the open interval $]0,\pi[$ we have
1668: $$
1669:  \textstyle
1670: \big((\hat f_1-\lambda)\psi_n\big)(x)
1671:  =
1672: \mr i\hbar \big(\ddx g_n\big)(x) \sinfn{x}\psi_\lambda(x)
1673:  +
1674: g_n(x) \left(\mr i\hbar \big(\ddx\sinfn{x} - \frac12\cosfn x\big)-\lambda\right)
1675: \psi_\lambda(x)\,.
1676: $$
1677: The second term vanishes because $\psi_\lambda$ solves \eqref{G-evaleqf1} on
1678: $]0,\pi[$. Hence
1679: $$
1680:  \textstyle
1681: \big((\hat f_1-\lambda)\psi_n\big)(x)
1682:  =
1683: \mr i \sqrt{\frac{\hbar}{2}}
1684:  ~
1685: \big(nj(nx-2) - nj(n(x-\pi)+2)\big)
1686:  ~
1687: \sqrt{\sinfn x}
1688:  ~
1689: \mr e^{-\frac{\mr i}{\hbar}\lambda \ln\tanfn{\frac x2}}
1690: $$
1691: and therefore
1692: $$
1693:  \textstyle
1694: \|(\hat f_1-\lambda)\psi_n\|^2
1695:  =
1696: \frac{\hbar}{2\pi} n^2
1697:  \left\{
1698: \int_0^\pi j(nx-2)^2 \, \sinfn x \, \mr d x
1699:  +
1700: \int_0^\pi j\big(n(x-\pi)+2\big)^2 \, \sinfn x \, \mr d x
1701:  \right\}\,.
1702: $$
1703: The mixed term vanishes for $n$ large enough because $j(nx-2)$ has support
1704: in $]\frac1n,\frac3n[$ and $j(n(x-\pi)+2)$ has support in
1705: $]\pi-\frac3n,\pi-\frac1n[$. For the same reason,
1706: $j(nx-2)\sinfn x\leq\frac3n$ and $j(n(x-\pi)+2)\sinfn x\leq\frac3n$. Hence
1707: $$
1708:  \textstyle
1709: \|(\hat f_1-\lambda)\psi_n\|^2
1710:  \leq
1711: \frac{3\hbar}{2\pi} n
1712:  \left\{
1713: \int_0^\pi j(nx-2) \, \mr d x
1714:  +
1715: \int_0^\pi j\big(n(x-\pi)+2\big) \, \mr d x
1716:  \right\}
1717:  =
1718: \frac{3\hbar}{\pi}\,.
1719: $$
1720: It follows
1721: $$
1722:  \textstyle
1723: \frac{\|\psi_n\|^2}{\|(\hat f_1-\lambda)\psi_n\|^2}
1724:  \geq
1725: \frac{\pi}{3\hbar} ~ \|\psi_n\|^2
1726:  =
1727: \frac{1}{6\hbar^2} \int_0^\pi \frac{g_n^2(x)}{\sinfn x} \, \mr d x
1728:  \to
1729: \infty
1730: $$
1731: and hence $\lambda\in\sigma(\hat f_1)$.
1732: 
1733: Finally, consider $\hat f_1^2$. According to \eqref{G-f12expr},
1734: the eigenvalue equation $\big(\hat f_1^2 - \lambda^2\big) \psi =
1735: 0$ can be written in the form
1736: $$
1737:  \textstyle
1738: -\hbar^2\sinfn x
1739:  \left(
1740: \ddxx
1741:  -
1742: \frac14\cot^2(x)
1743:  +
1744: \left(\frac12 + \frac{\lambda^2}{\hbar^2}\right) \,
1745: \frac{1}{\sinfnpot 2 x}
1746:  \right) ~ (\sinfn{x}\psi(x))
1747:   =
1748: 0\,,~~~~~~ x\in]0,\pi[\,,
1749: $$
1750: where it is manifest that for any $\lambda \geq 0$ the solution
1751: space has dimension $2$. Hence, in case $\lambda^2\neq 0$, any
1752: solution is a linear combination of $\psi_\lambda$ and
1753: $\psi_{-\lambda}$ and, therefore, is not square-integrable. Thus,
1754: $\lambda^2$ is not an eigenvalue. In case $\lambda^2 =0$ we
1755: observe that, in addition to $\psi_0(x) =
1756: \frac{1}{\sqrt{\sinfn x}}$, a further solution is given by
1757: $\tilde\psi_0(x) = \frac{\ln\tanfn{\frac x2}}{\sqrt{\sinfn x}}$. Since
1758: neither $\psi_0$ nor $\tilde\psi_0$ is square-integrable,
1759: $\lambda^2 = 0$ is not an eigenvalue, too.
1760: 
1761: To prove the assertion about the spectrum we choose $\lambda\geq0$
1762: and consider the sequence $\psi_n$ defined above. Obviously,
1763: $\psi_n\in\mr D(\hat f_1^2)$ for all $n$. On $]0,\pi[$ we find, using
1764: \eqref{G-f1expr},
1765:  \begin{align*}
1766:  \textstyle
1767: \big(\hat f_1^2-\lambda^2\big)\psi_n(x)
1768:  & =
1769:  \textstyle
1770: -\hbar^2
1771:  \left\{
1772: g_n'(x) \sinfn{x}
1773:  \left(
1774: \frac12\cosfn{x} - \mr i \frac{\lambda}{\hbar}
1775:  \right)
1776:  +
1777: g_n''(x)\sinfnpot 2 x
1778:  \right\}
1779: \psi_\lambda(x)
1780: \\
1781:  & \hspace{4cm} +
1782:  \textstyle
1783: g_n(x)
1784:  \left(
1785: - \hbar^2
1786:  \left(
1787: \ddx\sinfn{x} - \frac12 \cosfn{x}
1788:  \right)^2
1789: -\lambda^2
1790:  \right)
1791: \psi_\lambda(x)\,.
1792:  \end{align*}
1793: The last term vanishes. Moreover, $g_n''(x) = n^2 \big(j'(nx - 2)
1794: - j'(n(x-\pi)+2)\big)$. Hence,
1795:  \begin{align*}
1796:  \textstyle
1797: \big(\big(\hat f_1^2-\lambda^2\big)\psi_n\big) (x)
1798:  & =
1799:  \textstyle
1800: -\sqrt{\frac{\hbar^3}{2}}
1801:  \left\{
1802: n\big(j(nx-2) - j(n(x-\pi)+2)\big) \sqrt{\sinfn{x}}
1803:  \left(
1804: \frac12\cosfn{x} - \mr i \frac{\lambda}{\hbar}
1805:  \right)\right.
1806: \\
1807:  & \hspace{2cm}+ \left.
1808:  \textstyle
1809: n^2 \big(j'(nx-2)-j'(n(x-\pi)+2)\big) \sinfnpot{\frac32}{x}
1810:  \right\}
1811: \mr e^{-\frac{\mr i}{\hbar} \ln\tanfn{\frac x2}}\,.
1812:  \end{align*}
1813: Consequently, there are two contributions to $\big\|\big(\hat
1814: f_1^2-\lambda^2\big)\psi_n\big\|^2$. One is centered near $x=0$ and is given by
1815:  \beq\label{G-specf12-1}
1816:  \textstyle
1817: \frac{\hbar^3}{2\pi}
1818:  \int_0^\pi
1819:  \Big\{
1820: \frac{\lambda^2}{\hbar^2} n^2 j(nx-2)^2 \sinfn{x}
1821:  +
1822:  \Big(
1823: \frac12 n j(nx-2) \cosfn x + n^2 j'(nx-2) \sinfn{x}
1824:  \Big)^2
1825: \sinfn{x}
1826:  \Big\}
1827:  \,
1828: \mr dx\,,
1829:  \eeq
1830: the other one is centered near $x=\pi$ and is obtained by replacing
1831: $j(nx-2)$ by $j(n(x-\pi)+2)$ and $j'(nx-2)$ by $j'(n(x-\pi)+2)$ in
1832: \eqref{G-specf12-1}. The first term
1833: in \eqref{G-specf12-1} already appeared in the discussion of the
1834: spectrum of $\hat f_1$, hence we know that it is bounded for all
1835: $n$. The integrand of the second term has support in $[\frac1n,\frac3n]$.
1836: There exists $C>0$ such that $|j'(y)|\leq C$ for all $y\in\RR$. Then
1837: $$
1838:  \textstyle
1839:  \big(
1840: \frac12 n j(nx-2) \cosfn x + n^2 j'(nx-2) \sinfn{x}
1841:  \big)^2
1842: \sinfn{x}
1843:  \leq
1844:  \big(
1845: \frac12 n + n^2 C \frac3n
1846:  \big)^2
1847:  \,
1848: \frac3n
1849:  \leq
1850: \left(\frac12+3C\right)^2 n\,,
1851: $$
1852: hence upon integration, this term is bounded by $2\left(\frac12+3C\right)^2$. An
1853: analogous argument applies to the contribution centered near $x=\pi$. Thus,
1854: $\|\big(\hat f_1^2-\lambda^2\big)\psi_n\|$ is bounded for all $n$.
1855: Then $\frac{\|\psi_n\|}{\|(\hat
1856: f_1^2-\lambda^2)\psi_n\|}\to\infty$ and hence $\lambda^2$ belongs
1857: to the spectrum of $\hat f_1^2$.
1858:  \qed
1859: \bigskip
1860: 
1861: For any $\lambda\in\RR$, the function $\psi_\lambda$ defines a linear functional
1862: on the subspace $C^\infty_0]0,\pi[$ of $L^2[0,\pi]$ by
1863: $$
1864: \braket{\psi_\lambda}{\vp}
1865:  :=
1866: \frac1\pi\int_0^\pi \ol{\psi_\lambda(x)} \, \vp(x) \, \mr d x
1867:  \,,~~~~~~
1868: \vp\in C^\infty_0]0,\pi[\,.
1869: $$
1870: Integration by parts yields $\braket{\psi_\lambda}{(\hat f_1-\lambda)\vp} =
1871: 0$ for all $\vp\in C^\infty_0]0,\pi[$, so that $\psi_\lambda$ can be viewed as
1872: a generalized eigenvector of $\hat f_1$.
1873: 
1874: 
1875: \bpr\label{P-genevecf1}
1876: 
1877: The set of generalized eigenvectors $\{\psi_\lambda : \lambda\in\RR\}$ of $\hat
1878: f_1$ is complete and orthogonal in the distributional sense, i.e.,
1879:  \begin{align} \label{G-complete}
1880: \int_{-\infty}^\infty \ol{\psi_\lambda(x)} \, \psi_\lambda(y)
1881:  \,
1882: \mr d\lambda
1883:  & =
1884: \pi \, \delta(x-y)\,,
1885:  &&
1886: x,y\in]0,\pi[\,,
1887: \\ \label{G-ogon}
1888: \frac1\pi\int_0^\pi
1889: \ol{\psi_{\lambda}(x)} \, \psi_{\mu}(x) \, \mr dx
1890:  & =
1891: \delta(\lambda - \mu)\,,
1892:  &&
1893: \lambda,\mu\in\RR\,.
1894:  \end{align}
1895: The assignment of $\tilde\vp(\lambda) =
1896: \braket{\psi_\lambda}{\vp}$ to $\vp\in C^\infty_0]0,\pi[$ extends to a unitary
1897: isomorphism of $L^2[0,\pi]$ onto $L^2(\RR)$ with inverse $\vp(x) =
1898: \int_{-\infty}^\infty \psi_\lambda(x) \, \tilde\vp(\lambda) \, \mr d \lambda$.
1899: 
1900: \epr
1901: 
1902: 
1903: {\it Proof.}~ For $x,y\in]0,\pi[$,
1904: $$
1905:  \textstyle
1906: \int_{-\infty}^\infty \ol{\psi_\lambda(x)}
1907:  \,
1908: \psi_\lambda(y)
1909:  \,
1910: \mr d\lambda
1911:  =
1912: \frac{1}{2\hbar} ~ \int_{-\infty}^\infty  \frac{\mr e^{\frac{\mr
1913: i}{\hbar}\lambda (\ln\tanfn{\frac x2}-\ln\tanfn{\frac y2})}}{\sqrt{\sinfn x\sinfn y}}
1914:  \,
1915: \mr d\lambda
1916:  =
1917: \pi ~
1918:  \frac{
1919: \delta\big(\ln\tanfn{\frac x2}-\ln\tanfn{\frac y2}\big)
1920:  }{
1921: \sqrt{\sinfn x\sinfn y}
1922:  }\,.
1923: $$
1924: The argument of the $\delta$-distribution, viewed as a function of
1925: $x$ with parameter $y$, has derivative $\frac{1}{\sinfn x}$ and a
1926: single zero at $x=y$. Hence, $ \delta\big(\ln\tan{\frac x2}
1927:  -
1928: \ln\tan{\frac y2}\big)
1929:  =
1930: \sinfn{y} \, \delta(x-y)\,.
1931: $
1932: This proves \eqref{G-complete}. For $\lambda,\mu\in\RR$, the substitution $y =
1933: \frac1\hbar \ln\tanfn{\frac x2}$ yields
1934: $$
1935:  \textstyle
1936: \frac1\pi \int_0^\pi \ol{\psi_{\lambda}(x)} \, \psi_{\mu}(x) \, \mr dx
1937:  =
1938: \frac{1}{2\pi\hbar} \int_0^\pi
1939:  \frac{
1940: \mr e^{\frac{\mr i}{\hbar}(\lambda - \mu)\ln\tanfn{\frac x2}}
1941:  }{
1942: \sinfn x
1943:  } \, \mr d x
1944:  =
1945: \frac{1}{2\pi} \int_{-\infty}^\infty
1946: \mr e^{\mr i (\lambda - \mu) y} \, \mr dy
1947:  =
1948: \delta(\lambda-\mu)\,,
1949: $$
1950: hence \eqref{G-ogon}. To prove the last assertion, we observe that
1951: \eqref{G-complete} implies
1952: $
1953: \int_{-\infty}^\infty \ol{\tilde\vp_1(\lambda)} \, \tilde\vp_2(\lambda)
1954:  \,
1955: \mr d\lambda
1956:  =
1957: \braket{\vp_1}{\vp_2}
1958: $
1959: for all $\vp_1,\vp_2\in C^\infty_0]0,\pi[$.
1960: Hence, $\tilde\vp(\lambda)\in L^2(\RR)$ and the assignment $\vp\mapsto\tilde\vp$
1961: extends to an isometric map $L^2[0,\pi]\to L^2(\RR)$. It remains to check that
1962: this map has dense image. Let $\vp\in C^\infty_0(\RR)$. Since the integrals
1963: $\int_{-\infty}^\infty \psi_\lambda(x)\vp(\lambda)\,\mr d\lambda$ exist for any
1964: $x\in]0,\pi[$, they define a function $\vp_0$ on $]0,\pi[$. Due to
1965: \eqref{G-ogon},
1966: $
1967: \frac1\pi \int_0^\pi |\vp_0(x)|^2 \, \mr dx
1968:  =
1969: \int_{-\infty}^\infty |\vp(x)|^2 \, \mr d\lambda\,,
1970: $
1971: hence $\vp_0\in L^2[0,\pi]$. This proves that the extended map is a unitary
1972: isomorphism.
1973:  \qed
1974:  \\
1975: 
1976: 
1977: 
1978: 
1979: 
1980: 
1981: 
1982: 
1983: 
1984: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1985: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1986: 
1987: \subsection{Quantum analogue of the classical relation between generators}
1988: \label{S-q-SS-rel}
1989: 
1990: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1991: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1992: 
1993: 
1994: 
1995: 
1996: 
1997: Consider the relation \eqref{G-rel} and the inequality
1998: \eqref{G-ineq} satisfied by the classical generators $f_i$. Since
1999: $\hat f_2 \geq 0$, the inequality \eqref{G-ineq} has an obvious
2000: quantum counterpart. Concerning the relation \eqref{G-rel}, we
2001: recall that on the domain of $\hat f_2$, $\hat f_1^2$ is given by
2002: \eqref{G-f12expr}. Expressing the r.h.s.\ of this equation in terms of the
2003: $\hat f_i$ we obtain
2004: $$
2005:  \textstyle
2006: \hat f_1^2
2007:  =
2008: \frac{1}{4\beta^2} \sqrt{4-\hat f_0^2} \, \hat f_2 \, \sqrt{4-\hat f_0^2}
2009:  -
2010: \frac{3\hbar^2}{16} \hat f_0^2 + \frac{\hbar^2}{2}\,.
2011: $$
2012: This can be written in the form
2013:  \beq\label{G-Asqrt-f1}
2014:  \textstyle
2015: \sqrt{4-\hat f_0^2} \, \hat f_2 \, \sqrt{4-\hat f_0^2} - 4\beta^2 \hat
2016: f_1^2
2017:  =
2018: \hbar^2\beta^2\left(\frac34\hat f_0^2 - 2\right)\,.
2019:  \eeq
2020: We observe that when replacing $\hat f_i$ by $f_i$,
2021: \eqref{G-Asqrt-f1} reproduces the relation \eqref{G-rel} in the limit $\hbar \to
2022: 0$. We will now derive a relation between the quantum observables
2023: $\hat f_0$, $\hat f_1$, $\hat f_2$ which {\em exactly} reproduces
2024: the classical relation \eqref{G-rel} under the na\"ive replacement
2025: of the operators $\hat f_i$ by the phase space functions $f_i$,
2026: $i=0,1,2$. The attribute 'na\"ive' shall remind us that this
2027: operation is well-defined on the level of formal expressions in
2028: the variables $\hat f_0$, $\hat f_1$, $\hat f_2$ but not
2029: necessarily on the level of the operators defined by these expressions.
2030: To begin with, let $A_1$ be the operator defined on $\mr D(\hat f_2)$ by
2031: the l.h.s.\ of
2032: \eqref{G-Asqrt-f1}. Since the domain of $\hat f_2$ is invariant under $\hat f_0$
2033: and contained in the domain of $\hat f_1^2$, on $\mr D(\hat f_2)$ we can define, additionally,  
2034: the following operators:
2035:  \begin{align*}
2036:  \textstyle
2037: A_2 & := 4 \hat f_2- \hat f_0^2 \hat f_2 - 4\beta^2 \hat f_1^2\,,
2038: \\
2039: A_3 & := 4 \hat f_2 - \hat f_0 \hat f_2 \hat f_0 - 4\beta^2 \hat f_1^2\,,
2040: \\
2041: A_4 & := 4 \hat f_2 - \hat f_2 \hat f_0^2 - 4\beta^2 \hat f_1^2\,.
2042:  \end{align*}
2043: Like $A_1$, these operators correspond to the classical phase space function
2044: $(4-f_0^2)f_2 - 4\beta^2 f_1^2$, but contrary to $A_1$ they are polynomial in
2045: the $\hat f_i$. A straightforward computation using
2046: \eqref{G-f0expr}--\eqref{G-f2expr} yields that on $\mr D(\hat f_2)$ there
2047: holds
2048:  \begin{eqnarray}\label{G-A002-f1}
2049:  \textstyle
2050: \frac12(A_2 + A_4)
2051:  & = &
2052:  \textstyle
2053: \hbar^2\scale^2\left(-\frac14 \hat f_0^2 - 2\cdot\II\right)\,,
2054:  \\ \label{G-A020-f1}
2055:  \textstyle
2056: A_3
2057:  & = &
2058:  \textstyle
2059: \hbar^2\scale^2 \left( \frac34 \hat f_0^2 - 6\cdot\II\right)\,,
2060:  \end{eqnarray}
2061: First, we observe that,
2062: similar to \eqref{G-Asqrt-f1}, when replacing $\hat f_i$
2063: by $f_i$, both \eqref{G-A002-f1} and \eqref{G-A020-f1} reproduce the relation
2064: \eqref{G-rel} in the limit $\hbar \to 0$. In addition, we observe that the three
2065: operators $A_1$, $\frac12(A_2 + A_4)$ and $A_3$ are contained in the real
2066: vector space spanned by $\II$ and $\hat f_0^2$. A brief computation
2067: reveals that the sum of the coefficients of a vanishing linear
2068: combination is nonzero. Hence, these coefficients can be chosen so
2069: that they add up to $1$. The corresponding linear combination is
2070: $$
2071:  \textstyle
2072: \frac34 A_1 + \frac38 (A_2 + A_4) - \frac12 A_4 = 0\,.
2073: $$
2074: This yields the relation
2075:  \beq\label{G-qrel}
2076:  \textstyle
2077: \hat f_2
2078:  -
2079:  \left(
2080: \frac38 \hat f_0^2 \hat f_2
2081:  -
2082: \frac12 \hat f_0 \hat f_2 \hat f_0
2083:  +
2084: \frac38 \hat f_2 \hat f_0^2
2085:  \right)
2086:  +
2087: \frac34 \sqrt{4-\hat f_0^2} \, \hat f_2 \, \sqrt{4-\hat f_0^2}
2088:  -
2089: 4 \scale^2 \hat f_1^2
2090:  =
2091: 0\,.
2092:  \eeq
2093: It holds exactly on the domain of $\hat f_2$ and reproduces the classical
2094: relation \eqref{G-rel} under the na\"ive replacement of the operators $\hat f_i$
2095: by the phase space functions $f_i$, $i=0,1,2$.
2096: 
2097: 
2098: From the above relations we can derive
2099: relations, valid on the whole of $\mr D(\hat f_2)$, expressing the operators
2100: $(4-\hat f_0^2) \hat f_2$, $(4-\hat f_0^2)^{\frac12} \, \hat f_2 \, (4-\hat
2101: f_0^2)^{\frac12}$ or $\hat f_2(4-\hat f_0^2)$ as  polynomials in $\hat
2102: f_0$, $\hat f_1$ and the identity. In any of these expressions, the contribution
2103: of the identity is nonzero. Since $(4-\hat f_0^2)$ is given by multiplication by
2104: $4\sinfnpot 2 x$, any subspace on which such a relation can be resolved for
2105: $f_2$ must be contained in
2106:  \beq\label{G-nocore}
2107:  \textstyle
2108: \{\psi\in\mr D(\hat f_2) : \frac{\psi(x)}{\sinfnpot 2 x} \in L^2[0,\pi]\}\,.
2109:  \eeq
2110: 
2111: 
2112: \bpr
2113: \label{P-nocore}
2114: The subspace \eqref{G-nocore} is not a core for $\hat f_2$.
2115: 
2116: \epr
2117: 
2118: 
2119: Thus, none of the above relations determines
2120: $\hat f_2$ completely in terms of $\hat f_0$ and $\hat f_1$. 
2121: \\
2122: 
2123: 
2124: {\it Proof.}~ Let $D_0$ denote the subspace \eqref{G-nocore}. Let $A$ be defined
2125: as the restriction of $\hat f_2$ to the domain $D_0$. First, we show
2126: that any $\psi\in D_0$ satisfies $\psi'(0) = \psi'(\pi) = 0$. Indeed, since
2127: $\psi'\in\AC^1[0,\pi]$, $\psi'$ is continuous and can be extended
2128: continuously outside $[0,\pi]$. Hence, $\psi'\in C^1[0,\pi]$. Consider the
2129: function $\frac{\psi(x)}{\sinfn x}$ on $]0,\pi[$. Since for $x\to +0$,
2130: $\frac{\psi'(x)}{\cosfn x} \to \psi'(0)$, by the rule of de l'Hospital,
2131: $\frac{\psi(x)}{\sinfn x} \to \psi'(0)$. Hence there exists $x_0>0$ such that
2132: for any $x\in[0,x_0]$ there holds
2133: $\left|\frac{\psi(x)}{\sinfn x}\right| \geq \left|\frac{\psi'(0)}{2}\right|$.
2134: Since $\frac{\psi(x)}{\sinfnpot 2 x}$ is square-integrable then $\psi'(0) = 0$.
2135: A similar argument shows the assertion for $\psi'(\pi)$. Now, integration by
2136: parts yields that for any $\vp\in\AC^2[0,\pi]$ and any $\psi\in D_0$ there holds
2137: $\braket{\vp}{\hat f_2\psi} = \braket{-\hbar^2\beta^2(\vp''+\vp)}{\psi}$. It
2138: follows that $\AC^2[0,\pi] \subseteq \mr D(A^\dagger)$, hence $D_0$ is not a
2139: core for $\hat f_2$, as asserted.
2140:  \qed
2141: 
2142: 
2143: 
2144: 
2145: 
2146: 
2147: 
2148: 
2149: 
2150: 
2151: 
2152:  \todo{
2153: 
2154: -- Gerds Korrekturen
2155: 
2156: -- Alexanders Korrekturen
2157: 
2158: 
2159: -- Erzeugung von $\hat f_2$
2160: 
2161: 
2162: 
2163: 
2164: Problem: Schreibt man dieselbe Operatorrelation anders (d. h. mit anderen
2165: Polynomen in den $\hat f_i$) auf, dann liefert sie mglw eine andere klassische
2166: Relation, die gar nicht gelten muss. (Pruefe, ob diese Gefahr wirklich besteht.)
2167: 
2168: 
2169: Evtl.\ Beispiel angeben, dass naive Ersetzung inkonsistent ist.
2170:  }%
2171: 
2172: 
2173: 
2174: 
2175: 
2176: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2177: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2178: 
2179: \subsection{Algebra of quantum observables}
2180: 
2181: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2182: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2183: 
2184: 
2185: 
2186: As the algebra of quantum observables we would like to take an
2187: algebra which is generated, in some natural way, by the quantized
2188: generators of the algebra of classical observables (to be precise,
2189: the subalgebra of observables polynomial in the position
2190: and momentum variables). There is a natural choice for that
2191: algebra. It relies on the notion of a $C^\ast$-algebra generated
2192: by unbounded operators in the sense of Woronowicz.
2193: 
2194: 
2195: \begin{Definition}{\rm \cite{Woro}}\label{D-Woro}
2196: 
2197: Let $\Hi$ be a separable Hilbert space. Let $\mc A$ be a $C^\ast$-subalgebra of
2198: $\mr B(\Hi)$ and let $T_1,\dots,T_N$ be closed, densely defined operators on
2199: $\Hi$ affiliated with $\mc A$. Then $\mc A$ is generated by $T_1,\dots,T_N$ in
2200: the sense of Woronowicz if for all nondegenerate representations $\pi$ of $\mc
2201: A$ on $\Hi$ and all nondegenerate $C^\ast$-subalgebras $\mc B\subseteq\mr
2202: B(\Hi)$ there holds: if $\pi(T_1),\dots,\pi(T_N)$ are affiliated with $\mc B$,
2203: then $\pi(\mc A)\subseteq \mr M(\mc B)$ and $\pi(\mc A)\mc B$ is dense in $\mc
2204: B$.
2205: 
2206: \end{Definition}
2207: 
2208: 
2209: Here,
2210: $
2211: \mr M(\mc B) = \{ b\in\mr B(\Hi) ~:~ b\mc B, \mc Bb \subseteq \mc
2212: B \}
2213: $
2214: is the multiplier algebra of $\mc B$.
2215: Let us recall the notions entering this definition, see \cite{Woro}.
2216: For a closed, densely defined operator $T$ on $\Hi$, the $z$-transform is
2217: defined by
2218: $$
2219: z_T = T(\II + T^\dagger T)^{-\frac12}\,.
2220: $$
2221: This is a bounded operator on $\Hi$. $T$ can be recovered from $z_T$ by
2222:  \beq\label{G-TfromzT}
2223: T = z_T(\II-z_T^\dagger z_T)^{-\frac12}\,.
2224:  \eeq
2225: A closed, densely defined operator $T$ on $\Hi$ is said to be affiliated with
2226: $\mc A\subseteq\mr B(\Hi)$ (in the sense of Baaj and Julg \cite{BaajJulg})
2227: if $z_T\in\mr M(\mc A)$,
2228: $\II - z_T^\dagger z_T\geq 0$ and $(\II - z_T^\dagger z_T)\mc A$ is dense
2229: in $\mc A$. A nondegenerate representation of $\mc A$ is a $\ast$-morphism
2230: $\pi:\mc A\to\mr B(\Hi)$ such that $\pi(\mc A)\Hi$ is dense in $\Hi$. (Note that
2231: the assumption in Definition \rref{D-Woro} that $\pi(\mc A)$ is dense in
2232: $\mc B$ may not follow from the assumption that $\vp(\mc A)\Hi$ is dense in
2233: $\Hi$ made here.) The representation $\pi:\mc A\to\mr B(\Hi)$ can be extended to
2234: affiliated operators $T$ by extending $\pi$ to $\mr M(\mc A)$ through
2235: $$
2236: \pi(b)\pi(a)\psi = \pi(ba)\psi
2237:  \,,~~~~~~
2238: b\in \mr M(\mc A) \,,~ a\in\mc A\,,~ \psi\in\Hi\,,
2239: $$
2240: and defining $\pi(T)$ by $z_{\pi(T)} = \pi(z_T)$. This definition makes sense,
2241: because $\pi(\mc A)\Hi$ is dense in $\Hi$.
2242: 
2243: The fundamental criterion to test whether $\mc A$ is generated by a given set of
2244: affiliated operators is
2245: 
2246: 
2247: \btm{\rm\cite[Thm.\ 3.3]{Woro}} \label{T-Woro-1}
2248: 
2249: Let $\mc A$ be a $C^\ast$-subalgebra of $\mr B(\Hi)$ and let
2250: $T_1,\dots,T_N$ be closed, densely defined operators on $\Hi$
2251: affiliated with $\mc A$. Then $\mc A$ is generated by
2252: $T_1,\dots,T_N$ if
2253: 
2254: {\rm 1.}~ some product built from $(\II + T_i^\dagger T_i)^{-1}$,
2255: $(\II + T_i T_i^\dagger)^{-1}$ belongs to $\mc A$,
2256: 
2257: {\rm 2.}~ $T_1,\dots,T_N$ separate the representations of $\mc A$
2258: on $\Hi$.
2259:  \qed
2260: 
2261: \etm
2262: 
2263: Separation of representations means that for any two distinct
2264: representations $\pi_1,\pi_2 : \mc A \to \mr B(\Hi)$ there exists an 
2265: $i\in\{1,\dots,N\}$ such that $\pi_1(T_i)\neq \pi_2(T_i)$.
2266: Prominent examples of $C^\ast$-algebras generated by unbounded
2267: operators are:
2268: \medskip
2269: 
2270: 
2271: 1.~ Let $\mc A$ be a unital $C^\ast$-subalgebra of $\mr B(\Hi)$
2272: and let $T_1,\dots,T_N$ be elements of $\mc A$ such that the
2273: algebra generated by $T_1,\dots,T_N$ and the identity is dense in $\mc A$. Then
2274: $T_1,\dots,T_N$ generate $\mc A$ in the sense of Woronowicz. Thus, the concept
2275: is a generalization of the ordinary notion of generation of an
2276: algebra by a subset.
2277: \medskip
2278: 
2279: 
2280: 2.~ Let $G$ be a connected Lie group, let $\Hi = L^2(G)$ and let
2281: $\mc A$ be the group $C^\ast$-algebra $C^\ast(G)$. Then $\mc A$ is
2282: generated in the sense of Woronowicz by any basis in the Lie
2283: algebra $\mf g$, where the basis elements are viewed as first
2284: order differential operators on $G$.
2285:  \medskip
2286: 
2287: 3.~ Let $\Hi = L^2(\RR)$, let $\mc A$ be the $C^\ast$-algebra $\mr
2288: K(\Hi)$ of compact operators on $\Hi$. Then $\mc A$ is generated
2289: in the sense of Woronowicz by the position operator $T_1 = x$ and
2290: the momentum operator $T_2 = \frac\hbar{\mr i} \frac{\mr d}{\mr
2291: dx}$.
2292:  \medskip
2293: 
2294: 
2295:  \bre
2296: 
2297: At the present stage this theory has the disadvantage of not providing a general method
2298: how to construct the algebra for a given set of generators.
2299: 
2300:  \ere
2301: 
2302: 
2303: \btm\label{T-Woro}
2304: 
2305: Each of the sets $\{\hat f_0, \hat f_2\}$ and $\{\hat f_1, \hat
2306: f_2\}$ generates the $C^\ast$-algebra $\mr K\big(L^2(G)^G\big)$ in
2307: the sense of Woronowicz.
2308: 
2309: \etm
2310: 
2311: 
2312: Thus, it is natural to define the algebra of quantum observables
2313: to be
2314: $$
2315: \mc O_q := \mr K\big(L^2(G)^G\big)\,.
2316: $$
2317: 
2318: {\it Proof.}~ In the proof, denote $\Hi = L^2(G)^G$. The $\hat
2319: f_i$ are affiliated with $\mr K(\Hi)$, because this holds for any
2320: closed and densely defined operator. To prove Theorem
2321: \rref{T-Woro}, we use the criterion given in Theorem
2322: \rref{T-Woro-1}. The first condition holds, because $\hat f_2^2$
2323: has eigenbasis $\chi_n$ with eigenvalues $\hbar^4 \beta^4
2324: n^2(n+2)^2$ and hence has compact resolvent. To check the second
2325: condition, let $k=0$ or $1$. Assume that we are given
2326: representations $\pi_1$, $\pi_2$ with $\pi_1(\hat f_i) =
2327: \pi_2(\hat f_i)$, $i=k,2$. By definition of the operators
2328: $\pi_j(\hat f_i)$, then $\pi_1(z_{\hat f_i}) = \pi_2(z_{\hat
2329: f_i})$, $i=k,2$. Hence, $\pi_1$ and $\pi_2$ coincide on the
2330: subalgebra $\tilde A$ of the multiplier algebra of $\mr K(\Hi)$
2331: generated by $\II$, $z_{\hat f_k}$, $z_{\hat f_2}$. To prove
2332: $\pi_1 = \pi_2$ it suffices to show $\mr K(\Hi) \subseteq
2333: \tilde{\mc A}$. As the multiplier algebra of $\mr K(\Hi)$ is $\mr
2334: B(\Hi)$, we can apply the following criterion.
2335: 
2336: 
2337: \ble\cite[Prop.\ 10.4.1]{KaRi}\label{L-Woro-2}
2338: 
2339: Let $\tilde{\mc A}$ be a $C^\ast$-subalgebra of $\mr B(\Hi)$. If
2340: 
2341: {\rm 1.}~ $\tilde{\mc A} \cap \mr K(\Hi) \neq 0$,
2342: 
2343: {\rm 2.}~ $\tilde{\mc A}$ is irreducibly represented on $\Hi$,
2344: 
2345: then $\mr K(\Hi) \subseteq \tilde{\mc A}$.
2346:  \qed
2347: 
2348: \ele
2349: 
2350: 
2351: We check these two conditions. For the first one, we use that
2352: for a closed, densely defined operator $T$ there holds the
2353: identity
2354:  \begin{align*}
2355: \II - z_T^\dagger z_T
2356:  & =
2357: (\II + T^\dagger T)^{-1}\,,
2358:  \end{align*}
2359: see e.g.\ \cite{Lance}. Plugging in $\hat f_2$ for $T$ we observe: the
2360: l.h.s.\ belongs to $\tilde{\mc A}$ and the r.h.s.\ was shown above
2361: to belong to $\mr K(\Hi)$. Thus, the first condition holds,
2362: indeed. To prove the second condition, we apply the lemma of
2363: Schur. Assume that we are given a bounded operator $S$ on $\Hi$
2364: that commutes with all elements of $\tilde{\mc A}$. Then $S$
2365: commutes with $z_{\hat f_k}$ and $z_{\hat f_2}$. In particular, $S$ leaves
2366: invariant the eigenspaces of $z_{\hat f_2}$. According to \eqref{G-qobs-bs-2},
2367: $\chi_n$ is a basis of eigenvectors of $z_{\hat f_2}$ with eigenvalues
2368: $\frac{\hbar^2\beta^2 n(n+2)}{\sqrt{1+\hbar^4\beta^4 n^2(n+2)^2}}$. Since this
2369: is a strictly monotonous function of $n$, the eigenspaces have dimension $1$.
2370: Hence, $S\chi_n = \lambda_n \chi_n$, $n=0,1,2,\dots$ with
2371: $\lambda_n\in\CC$. According to \eqref{G-TfromzT}, $[S,z_{\hat f_k}] = 0$
2372: implies $S\hat f_k \chi_n - \hat f_k S \chi_n = 0$ for all $n$. This yields,
2373: respectively,
2374:  \begin{align*}
2375:  \textstyle
2376: (\lambda_{n+1} - \lambda_n) \chi_{n+1}
2377:  +
2378: (\lambda_{n-1} - \lambda_n) \chi_{n-1}
2379:  & = 0 & & (k=0)
2380: \\
2381:  \textstyle
2382: \frac{2n+3}{4} \left(\lambda_{n+1} - \lambda_n\right) \chi_{n+1}
2383:  -
2384: \frac{2n+1}{4} \left(\lambda_{n-1} - \lambda_n\right) \chi_{n-1}
2385:  & = 0 & & (k=1)
2386:  \end{align*}
2387: for all $n$. In both cases, it follows $\lambda_{n+1} = \lambda_n$ for all $n$,
2388: hence $S = \lambda \II$. Then $\tilde{\mc A}$ is irreducibly represented on
2389: $\Hi$ by the lemma of Schur and hence the second condition of Lemma
2390: \rref{L-Woro-2} is satisfied. This shows that condition 2 of Theorem
2391: \rref{T-Woro-1} holds and, therefore, completes the proof of Theorem
2392: \rref{T-Woro}.
2393:  \qed
2394: 
2395: 
2396: \bre
2397: 
2398: There remains the question whether the set $\{\hat f_0,\hat f_1\}$ generates
2399: $\mr K\big(L^2(G)^G\big)$  as well. The crucial point
2400: is Condition 1 of Theorem \rref{T-Woro-1}. While according to Proposition
2401: \rref{P-spec}, the operators $(\II+\hat f_0^2)^{-1}$ and $(\II+\hat
2402: f_1^2)^{-1}$ do not belong to $\mr K(L^2(G)^G)$, it would be sufficient to show that some product of these 
2403: operators is compact. We did not succeed to clarify this point.
2404:  \todo{
2405: 
2406: \RA~ pruefe, ob nicht doch ein Produkt von $(\II + \hat
2407: f_0^2)^{-1}$ und $(\II + \hat f_1^2)^{-1}$ kompakt ist.
2408: 
2409:  }
2410: 
2411: \ere
2412: 
2413: 
2414: 
2415: 
2416: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2417: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2418: 
2419: \subsection{Relation with the algebra of bosonic quantum
2420: observables in lattice gauge theory}
2421: 
2422: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2423: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2424: 
2425: 
2426: 
2427: We discuss the relation between the algebra of quantum observables
2428: of our model and the bosonic part of the algebra of observables of
2429: a quantum lattice gauge theory of \cite{KiRu05}. This paper
2430: concentrates on the case of lattice quantum chromodynamics, i.e.,
2431: gauge group $\SU(3)$ rather than $\SU(2)$ as in our model.
2432: However, the results are valid for general $\SU(n)$. We recall the
2433: construction of the bosonic observable algebra for the case of a
2434: single plaquette without external links, after having implemented
2435: the tree gauge. The bosonic field algebra is the crossed product
2436: algebra $\mc F = C(G)\otimes_\alpha G$ associated with the
2437: $C^\ast$-dynamical system $(C(G),G,(\mr L_{g^{-1}})^\ast)$. For
2438: the notions of $C^\ast$-dynamical system and crossed-product
2439: algebra, see \cite{Pedersen}.
2440: $\mc F$ carries a natural $G$-action. The bosonic observable
2441: algebra $\mc O$ is defined as the quotient of the subalgebra of
2442: $G$-invariant elements of $\mc F$ by the ideal defined by the
2443: generators of the $G$-action. This factorization corresponds to
2444: imposing the Gauss law, which is the quantum analogue of the
2445: restriction of the phase space to the zero level set of the
2446: momentum mapping. If there are external links, this definition of
2447: $\mc O$ yields the subalgebra of internal observables. The natural
2448: covariant representation of $(C(G),G,\alpha)$ on $L^2(G)$
2449: naturally induces a representation of $\mc F$ on $L^2(G)$, mapping
2450: $\mc F$ to $\mr K(L^2(G))$.
2451:  \todo{
2452: 
2453: Evtl. Definition der kovar Dst und der induzierten von $\mc F$
2454: angeben. Erstere: $\pi(f)\psi = f\psi$, $u(g)\psi = \mr
2455: (L_{g^{-1}})^\ast\psi$.
2456: 
2457:  }%
2458: It is shown in \cite{KiRu05} that this representation is the
2459: unique irreducible representation of $\mc F$. It is therefore
2460: called the generalized Schr\"odinger representation. Using this
2461: representation, it is then shown that $\mc O$ can be identified
2462: with the compact operators on the closed subspace $L^2(G)^G$.
2463: Thus, through this identification, the algebra of quantum
2464: observables $\mc O_q$ of our model coincides with the bosonic
2465: observable algebra $\mc O$ of \cite{KiRu05}, specified to the case
2466: of a single plaquette without external links.
2467: 
2468: Next, we compare generators. Let $U^A{}_B : G\to\CC$ denote the matrix entry
2469: functions. Choose a basis $T_i$ in $\mf g$ orthonormal
2470: w.r.t.\ the trace form and define vector fields $E^A{}_B$ on $G$
2471: by
2472: $$
2473:  \textstyle
2474: E^A{}_B = \sum_i (T_i)^A{}_B T_i\,,
2475: $$
2476: where $(T_i)^A{}_B$ are the entries of the basis element $T_i$
2477: when viewed as a matrix, whereas the second $T_i$ is viewed as a
2478: vector field. It is stated in \cite{KiRu05} that $\mc F$, when
2479: realized as $\mr K(L^2(G))$, is generated in the sense of
2480: Woronowicz by the multiplication operators $U^A{}_B$ and the first
2481: order differential operators $E^A{}_B$. It was not clarified in \cite{KiRu05} whether gauge invariant 
2482: combinations of $U^A{}_B$ and $E^A{}_B$ generate the observable algebra. 
2483: 
2484: Our quantum observables $\hat f_0$, $\hat f_1$ and $\hat f_2$ can be expressed
2485: in terms of the gauge invariant combinations $U^A{}_A\, , $ $U^A{}_B E^B{}_A$ and $E^A{}_B E^B{}_A$ as follows. 
2486: Since $U^A{}_A = f_0$ as functions
2487: on $G$, for the multiplication operators we have
2488: $$
2489: \hat f_0 = U^A{}_A\,.
2490: $$
2491: For the value of the vector field $U^A{}_B E^B{}_A$ at
2492: $a\in G$ we find
2493: $\left(U^A{}_B E^B{}_A\right)_a
2494:  =
2495: \mr L_a' \sum_i \tr(a T_i) T_i$. Since $T_i$ is orthonormal w.r.t.\ the trace
2496: form, $\sum_i \tr(a T_i) T_i = -\OPg(a)$. According to \eqref{G-Yi}, then
2497: $U^A{}_B E^B{}_A = - Y_{f_1}$ and hence
2498: $$
2499:  \textstyle
2500: \hat f_1
2501:  =
2502: \mr i\hbar\big( - U^A{}_B E^B{}_A + \frac34 U^A{}_A\big)\,.
2503: $$
2504: Finally, a similar computation shows
2505: $$
2506: \hat f_2 = 2\hbar^2\beta^2 E^A{}_B E^B{}_A\,.
2507: $$
2508: Thus, Theorem \rref{T-Woro} implies that, in the case of a single plaquette
2509: without external links, the algebra of quantum observables of \cite{KiRu05} is
2510: generated, in the sense of Woronowicz, by $U^A{}_A$ and $E^A{}_B E^B{}_A$ or by
2511: $U^A{}_B E^B{}_A$ and $E^A{}_B E^B{}_A$. This
2512: extends the result of \cite{KiRu05} on the generation of the field algebra by
2513: unbounded operators  to the algebra of observables, at
2514: least in the simple case at hand.
2515: 
2516: In \cite{KiRu05} it was argued that on a purely algebraic level the observable algebra is 
2517: generated by $U^A{}_A$ and $U^A{}_B E^B{}_A$ and that all other invariants can be 
2518: expressed in terms of these generators. This is, however, the pair of operators for 
2519: which we could not prove that they generate the algebra in the sense of Woronowicz. Moreover, from Proposition \ref{P-nocore} 
2520: we conclude that e.g. the quadratic Casimir operator $E^A{}_B E^B{}_A$ cannot be expressed in 
2521: terms of $U^A{}_A$ and $U^A{}_B E^B{}_A$ on a core. These observations show that any na\"ive algebraic procedure of 
2522: reducing the number of independent generators has to be handled with care.
2523: 
2524: 
2525: 
2526: 
2527: 
2528: 
2529: 
2530: 
2531: 
2532: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2533: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2534: 
2535: \subsection{Towards quantum dynamics}
2536: 
2537: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2538: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2539: 
2540: 
2541: 
2542: Quantization of the classical Hamiltonian \eqref{G-Ham-ivr} yields the quantum
2543: Hamiltonian
2544: $$
2545:  \textstyle
2546: \hat H = \frac12 \hat f_2 + \frac{1}{2g^2}(3-\hat f_0)
2547: $$
2548: which is a time-independent self-adjoint operator with domain $\mr D(\hat H) =
2549: \mr D(\hat f_2)$. On the level of pure states (Schr\"odinger picture),
2550: dynamics is given by the 1-parameter group of unitary transformations of
2551: $L^2(G)^G$ generated by $\hat H$,
2552: $$
2553: U_t = \mr e^{-\frac{\mr i}{\hbar} \hat H t}\,.
2554: $$
2555: Since the algebra of compact operators is invariant under unitary
2556: tranformations, $U_t$ induces a 1-parameter automorphisms group $\alpha_t$ of
2557: the algebra of quantum observables by
2558: $$
2559: \alpha_t(A) = U_t A U_t^\dagger\,.
2560: $$
2561: On the level of observables (Heisenberg picture), dynamics is given by the
2562: 1-parameter automorphism 
2563: group $\alpha_t$. It is interesting as well to study the dynamics of the
2564: generators $\hat f_k$. On the common invariant core $C^\infty(G)^G$ it is given
2565: by 
2566: \beq
2567: \label{G-Hbeq0}
2568: \hat f_k(t) = U_t \hat f_k U_t^\dagger \, .
2569: \eeq
2570: The corresponding equation of motion, on this core, reads
2571:  \beq
2572:  \label{G-Hbeq}
2573: \ddt \hat f_k(t) = \frac{\mr i}{\hbar} [\hat H,\hat f_k(t)]
2574:  \,,~~~~~~
2575: \hat f_k(0) = \hat f_k
2576:  \,,~~~~~~
2577: k=0,1,2\,.
2578:  \eeq
2579: The automorphism group $\alpha_t$ and the operators $\hat f_k(t)$ will be
2580: studied elsewhere. 
2581: 
2582: We conclude with a discussion of the commutators between the
2583: generators $\hat f_k \, . $ These commutators are relevant for the evaluation of the right-hand side of \eqref{G-Hbeq0} 
2584: and for the iterative solution of \eqref{G-Hbeq}, respectively. Since $\hat f_0$ leaves invariant the domains of
2585: $\hat f_1$ and 
2586: $\hat f_2$, the commutators $[\hat f_0,\hat f_1]$ and $[\hat f_0,\hat f_2]$ are
2587: defined on these domains. A straightforward computation using
2588: \eqref{G-f0expr}--\eqref{G-f2expr} yields 
2589:  \beq\label{G-ctr-f0}
2590: [\hat f_0,\hat f_1] = \mr i \hbar \left(2 - \frac12\hat
2591: f_0^2\right)
2592:  \,,~~~~~~
2593: [\hat f_0,\hat f_2] = 4 \scale^2 \, \mr i \hbar \, \hat f_1\,.
2594:  \eeq
2595: We claim that the commutator of $\hat f_1$ and $\hat f_2$ is
2596: defined on $\mr D(\hat f_2)$ and is given by
2597:  \beq\label{G-ctr-f1f2}
2598:  \textstyle
2599: [\hat f_1,\hat f_2]
2600:  =
2601: - \frac12 \mr i \hbar
2602:  \left(
2603: \hat f_0 \hat f_2 + \hat f_2 \hat f_0 + 3 \hbar^2 \hat f_0
2604:  \right)\,.
2605:  \eeq
2606: To see this, write
2607:  \begin{align*}
2608: \hat f_1 \hat f_2 - \hat f_2 \hat f_1
2609:  & =
2610:  \textstyle
2611: -\mr i \hbar^3 \beta^2
2612:  \left\{
2613:  \left(
2614: \ddx\sinfn x - \frac12\cosfn x
2615:  \right)
2616:  \big(
2617: \ddxx + 1
2618:  \big)
2619:  -
2620:  \big(
2621: \ddxx + 1
2622:  \big)
2623:  \left(
2624: \ddx\sinfn x - \frac12\cosfn x
2625:  \right)
2626:  \right\}
2627: \\
2628:  & =
2629:  \textstyle
2630: \mr i \hbar^3 \beta^2
2631:  \left\{
2632: \ddx
2633:  \big(
2634: \ddxx\sinfn x - \sinfn x\ddxx
2635:  \big)
2636:  +
2637: \frac12
2638:  \big(
2639: \cosfn x \ddxx - \ddxx \cosfn x
2640:  \big)
2641:  \right\}
2642:  \end{align*}
2643: and observe that for $\psi\in\AC^2[0,\pi]$,
2644: $$
2645:  \textstyle
2646: \left(\ddxx\sinfn x - \sinfn x\ddxx\right) \psi(x)
2647:  =
2648: \left(-\sinfn x + 2\cosfn x \ddx\right) \psi(x)\,.
2649: $$
2650: Hence $\hat f_1\hat f_2 - \hat f_2 \hat f_1$ contains derivatives
2651: up to second order only and is therefore defined on $\mr D(\hat f_2) \subseteq
2652: \AC^2[0,\pi]$, indeed. Then a straighforward calculation yields
2653: \eqref{G-ctr-f1f2}. Thus, all the commutators between the quantum observables
2654: $\hat f_k$ are defined on $\mr D(\hat f_2)$.
2655: 
2656: 
2657: \bre
2658: 
2659: Comparing the commutators \eqref{G-ctr-f0} and \eqref{G-ctr-f1f2} with the
2660: corresponding Poisson brackets \eqref{G-Poibra} we observe
2661: that for the combinations of $\hat f_0$ with $\hat f_1$ and $\hat f_2$ the
2662: na\"ive relation between the commutator and the Poisson bracket (replacing, in
2663: the commutator, the operators by their classical counterparts,
2664: provided the latter are well-defined) holds exactly. For the
2665: combination of $\hat f_1$ and $\hat f_2$, this relation holds in the limit
2666: $\hbar \to 0$.
2667: 
2668: \ere
2669: 
2670: 
2671: 
2672: 
2673: 
2674: 
2675: 
2676: 
2677: 
2678: 
2679: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2680: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2681: 
2682: \section{Outlook}
2683: 
2684: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2685: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2686: 
2687: 
2688: 
2689: An obvious future task is to generalize the results of this paper
2690: to a general compact Lie group. Another task is to study how the
2691: algebra of quantum observables depends on the choice of what phase
2692: space functions should be considered polynomial, cf.\ Remark
2693: \rref{Rem-pnmalg}. E.g., one should carry out a similar
2694: construction for the generators of the algebra of real invariant
2695: polynomials on $G^\CC$ and compare the resulting algebra of
2696: quantum observables with the one obtained above. The concept used in this paper
2697: should be also compared with an alternative approach proposed by Buchholz and Grundling 
2698: \cite{BuchholzGrundling}, who define the notion
2699: of resolvent algebra associated with a symplectic space and
2700: propose to take this algebra as the algebra of observables in a
2701: bosonic field theory. The resolvent algebra is a unital
2702: $C^\ast$-algebra defined abstractly in terms of generators and
2703: relations. Equivalently, it can be viewed as  generated by
2704: the resolvents $(\mr i\lambda - \phi(f))^{-1}$ of the field
2705: operators $\phi(f)$ of some quantum field $\phi$.
2706: Following these ideas, in our model one may take the unital
2707: $C^\ast$-algebra generated by the resolvents $(\mr i\lambda - \hat
2708: f_k)^{-1}$ of the quantized generators $\hat f_0$, $\hat f_1$,
2709: $\hat f_2$ as the algebra of observables. This algebra will be
2710: studied elsewhere. It is definitely distinct from the algebra of
2711: quantum observables $\mc O_q$ constructed above. The trivial
2712: reason for that is that this algebra is unital by definition;
2713: another reason is that, according to Proposition \rref{P-spec},
2714: the resolvents $(\mr i\lambda - \hat f_0)^{-1}$ and $(\mr i\lambda
2715: - \hat f_1)^{-1}$ are not compact.
2716: 
2717: Furthermore, we address the problem of studying the influence of the
2718: stratification of the classical configuration and phase spaces on the quantum
2719: theory. For that purpose, one has to find a quantum structure that implements
2720: this stratification. On the level of pure states, such a quantum structure is
2721: given by a costratification of the Hilbert space \cite{Hue:qr}. One may think of a
2722: costratification as a family of closed subspaces, indexed by the
2723: strata, and a family of orthoprojectors, indexed by the inclusion relations
2724: between the closures of the strata. For the model under consideration, the
2725: costratified Hilbert space was studied in \cite{hrs}. On the other hand, it is
2726: not clear how to implement the stratification on the level of observables.
2727: For the concrete algebra of observables at hand, this problem will be studied in
2728: detail in a future work.
2729: 
2730: 
2731: \section{Acknowledgements}
2732: 
2733: The authors are grateful to A.\ Hertsch, J.\ Huebschmann, J.\ Kijowski and
2734: K.\ Schm\"udgen for helpful discussions and to K.\ Schm\"udgen for reading part
2735: of the manuscript.
2736: 
2737: 
2738: 
2739: 
2740: 
2741: \begin{appendix}
2742: 
2743: 
2744: \section*{Appendix}
2745: 
2746: 
2747: We prove Formula \eqref{G-f1-4}. Choose an orthonormal basis $B_i$ in $\mf g$.
2748: Let $\beta_i$ denote the
2749: elements of the dual basis in $\mf g^\ast$. Then $\beta_i(A) =
2750: \langle B_i,A \rangle$ for any $A\in\mf g$. We have $v =
2751: \beta_1\wedge\beta_2\wedge\beta_3$, where the $\beta_i$ are viewed
2752: as left-invariant forms. Using the derivation property of the Lie
2753: derivative, expanding
2754: $$
2755:  \textstyle
2756: \mc L_{Y_{f_1}} \beta_i
2757:  =
2758: \sum\nolimits_{j=1}^3 ~ (\mc L_{Y_{f_1}} \beta_i) (B_j) ~ \beta_j
2759: $$
2760: and rewriting $(\mc L_{Y_{f_1}} \beta_i) (B_i) = - \beta_i (\mc
2761: L_{Y_{f_1}} B_i) = - \langle B_i,\mc L_{Y_{f_1}} B_i \rangle$ we
2762: obtain
2763:  \beq\label{G-gf1}
2764:  \textstyle
2765: \mc L_{Y_{f_1}} v
2766:  =
2767: -
2768:  \left(
2769: \sum\nolimits_{i=1}^3 ~ \langle B_i,\mc L_{Y_{f_1}} B_i\rangle
2770:  \right)
2771: v\,.
2772:  \eeq
2773: We calculate $\mc L_{Y_{f_1}} B_i$ by taking derivatives in the
2774: ambient vector space $\mr M_2(\CC)$. According to \eqref{G-Yi}, for $a\in G$,
2775: $$
2776:  \textstyle
2777: (\mc L_{Y_{f_1}} B_i)_a
2778:  =
2779: [Y_{f_1},B_i]_a
2780:  =
2781: \ddtn \ddsn a \mr e^{\OPg(a)t} \mr e^{B_i s}
2782:  -
2783: \ddtn \ddsn a \mr e^{B_i t} \mr e^{\OPg(a\mr e^{B_it})s}\,.
2784: $$
2785: This yields $a \OPg(a) B_i - a B_i \OPg(a) - a \OPg(aB_i)$, which
2786: can be rewritten as $- a \OPg(B_i a)$. Hence,
2787: $$
2788: (\mc L_{Y_{f_1}} B_i)_a = - \mr L_a' \OPg(B_i a)\,.
2789: $$
2790: Then
2791: $$
2792:  \textstyle
2793: \langle B_i,\mc L_{Y_{f_1}} B_i\rangle(a)
2794:  =
2795: -\langle B_i,\OPg(B_ia) \rangle
2796:  =
2797: \frac{1}{2\scale^2}
2798:  ~\frac12~
2799: \tr\left(B_i^2 (a + a^\dagger) \right)
2800:  =
2801: \frac{1}{2\scale^2} ~\frac12~ \tr(B_i^2) \tr(a)
2802:  =
2803: - \frac12\tr(a)\,,
2804: $$
2805: where we have used $B_i^2 -\frac12\tr(B_i^2)\II = 0$, due to the
2806: Cayley-Hamilton theorem. Then \eqref{G-gf1} yields $\mc L_{Y_{f_1}} v =
2807: \frac{3}{2} f_0 \, v$, i.e., Formula \eqref{G-f1-4}.
2808: 
2809: \end{appendix}
2810: 
2811: 
2812: 
2813: 
2814: \begin{thebibliography}{99}
2815: 
2816: \onehalfspacing
2817: 
2818: \bibitem{ACG}
2819:  J.M.\ Arms, R.H.\ Cushman, M.J.\ Gotay,
2820:  in {\it The geometry of Hamiltonian systems},
2821:  Math.\ Sci.\ Res.\ Inst.\ Publ.\ 22, pp.\ 33--51 (Springer, New York, 1991).
2822: 
2823: 
2824: \bibitem{BaajJulg}
2825:  S.\ Baaj, P.\ Julg,
2826:  C.\ R.\ Acad.\ Sci.\ Paris, Sér. I Math., {\bf 296}, 875--878 (1983).
2827: 
2828: \bibitem{Buchholz82}
2829:  D.\ Buchholz,
2830:  Commun.\ Math.\ Phys.\ {\bf 85}, 49--71 (1982).
2831: 
2832: 
2833: \bibitem{Buchholz86}
2834:  D.\ Buchholz,
2835:  Phys.\ Lett. B {\bf 174}, 331--334 (1986).
2836: 
2837: 
2838: \bibitem{BuchholzGrundling}
2839:  D.\ Buchholz, H.\ Grundling,
2840:  arXiv:0705.1988v3; to appear in J.\ Funct.\ Anal.
2841: 
2842: \bibitem{DHR71}
2843:  S.\ Doplicher, R.\ Haag, J.\ Roberts,
2844:  Commun.\ Math.\ Phys.\ {\bf 23},  199--230 (1971).
2845: 
2846: 
2847: \bibitem{DHR74}
2848:  S.\ Doplicher, R.\ Haag, J.\ Roberts,
2849:  Commun.\ Math.\ Phys.\ {\bf 35},  49--85 (1974).
2850: 
2851: 
2852: \bibitem{DR90}
2853:  S.\ Doplicher, J.\ Roberts,
2854:  Commun.\ Math.\ Phys.\ {\bf 131}, 51--107 (1990).
2855: 
2856: 
2857: \bibitem{Fredenhagen}
2858:  K.\ Fredenhagen, M.\ Marcu,
2859:  Commun.\ Math.\ Phys.\ {\bf 92}, 81--119 (1983).
2860: 
2861: 
2862: \bibitem{Froehlich}
2863:  J.\ Fr\"ohlich,
2864:  Commun.\ Math.\ Phys.\ {\bf 66}, 223--265 (1979).
2865: 
2866: 
2867: \bibitem{Hall:Compact}
2868:  B.C.\ Hall,
2869:  Commun.\ Math.\ Phys.\ {\bf 226}, 233--268 (2002).
2870: 
2871: 
2872: \bibitem{Helgason}
2873:  S.\ Helgason,
2874:  {\em Groups and geometric analysis. Integral geometry, invariant differential
2875:  operators, and spherical functions} (Academic Press 1984).
2876: 
2877: 
2878: \bibitem{Hue:qr}
2879:  J.\ Huebschmann,
2880:  J.\ reine angew.\ Math.\ \textbf{591}, 75--109 (2006).
2881: 
2882: 
2883: \bibitem{Hue:bedlewo}
2884:  J.\ Huebschmann,
2885:  in {\em The mathematical legacy of C.\ Ehresmann}, pp.\  325--347 
2886:  (Banach Center Publications 76, Warsaw 2007).
2887: 
2888: 
2889: \bibitem{Hue:holopewe}
2890:  J.\ Huebschmann,
2891:  J.\ Geom.\ Phys.\ 58, 833--848 (2008).
2892: 
2893: 
2894: \bibitem{hrs}
2895:  J.\ Huebschmann, G.\ Rudolph, M.\ Schmidt,
2896:  hep-th/0702017.
2897: \\
2898:  J.\ Huebschmann, G.\ Rudolph, M.\ Schmidt,
2899:  in {\it Lie Theory and its Applications in
2900:  Physics}, pp.\ 190--210 (Heron Press, Sofia, 2008).
2901: 
2902: \bibitem{JaKiRu}
2903:  P.D.\ Jarvis, J.\ Kijowski, G.\ Rudolph,
2904:  J.\ Phys.\ A {\bf 38}, 5359--5377 (2005).
2905: 
2906: 
2907: \bibitem{KaRi} 
2908:  R.V.\ Kadison, J.R.\ Ringrose,
2909:  {\em Fundamentals of the Theory of Operator Algebras}
2910:  (Academic Press 1986).
2911: 
2912: 
2913: \bibitem{KiRu02}
2914:  J.\ Kijowski, G.\ Rudolph,
2915:  J.\ Math.\ Phys.\ {\bf 43}, 1796--1808 (2002).
2916: 
2917: 
2918: \bibitem{KiRu05}
2919:  J.\ Kijowski, G.\ Rudolph,
2920:  J.\ Math.\ Phys.\ {\bf 46}, 032303 (2005).
2921: 
2922: 
2923: \bibitem{KiRuSl}
2924:  J.\ Kijowski, G.\ Rudolph, C.\ \'Sliwa,
2925:  Lett.\ Math.\ Phys.\ {\bf 43}, 299--308 (1998).
2926: 
2927: 
2928: \bibitem{KiRuTh}
2929:  J.\ Kijowski, G.\ Rudolph, A.\ Thielmann,
2930:  Commun.\ Math.\ Phys.\ {\bf 188}, 535--564 (1997).
2931: 
2932: 
2933: \bibitem{Lance}
2934:  E.\ Lance,
2935:  {\it Hilbert $C^\ast$-modules}
2936:  (Cambridge University Press, Cambridge, 1995).
2937: 
2938: 
2939: \bibitem{Pedersen}
2940:  G.K.\ Pedersen,
2941:  {\em $C\sp{*} $-algebras and their automorphism groups}
2942:  (Academic Press, Inc., London-New York, 1979).
2943: 
2944: 
2945: \bibitem{rsv:review}
2946:  G.\ Rudolph, M.\ Schmidt, I.P.\ Volobuev,
2947:  J.\ Phys.\ A: Math.\ Gen.\ {\bf 35}, R1--R50 (2002).
2948: 
2949: \bibitem{Schwarz}
2950:  G.W.\ Schwarz,
2951:  Topology {\bf 14}, 63--68 (1975).
2952: 
2953: 
2954: \bibitem{Woodhouse}
2955:  J.\ \'Sniatycki,
2956:  {\em Geometric quantization and Quantum Mechanics}
2957:  (Springer, 1980).
2958: \\
2959:  N.M.J.\ Woodhouse,
2960:  {\em Geometric quantization}
2961:  (Clarendon Press, Oxford 1991).
2962: 
2963: 
2964: \bibitem{StrocchiWightman}
2965:  F.\ Strocchi, A.\ Wightman,
2966:  J.\ Math.\ Phys.\ {\bf 15}, 2198--2224 (1974).
2967:  \\
2968:  F.\ Strocchi,
2969:  Phys.\ Rev.\ D {\bf 17}, 2010--2021 (1978).
2970: 
2971: 
2972: \bibitem{Weyl}
2973:  H.\ Weyl,
2974:  {\it The classical groups}
2975:  (Princeton University Press, Princeton, N.J., 1946).
2976: 
2977: 
2978: \bibitem{Woro}
2979:  S.L.\ Woronowicz,
2980:  Rev.\ Math.\ Phys.\ {\bf 7}, 481--521 (1995).
2981: 
2982: 
2983: \bibitem{Wren}
2984:  K.K.\ Wren,
2985:  J.\ Geom.\ Phys.\ {\bf 24}, 173--202 (1998).
2986:  \\
2987:  K.K.\ Wren,
2988:  Nucl.\ Phys.\ {\bf B 521}, 471--502 (1998).
2989: 
2990: 
2991: 
2992: 
2993: \end{thebibliography}
2994: 
2995: 
2996: 
2997: 
2998: \end{document}
2999: 
3000: 
3001: 
3002: 
3003: 
3004: 
3005: VOLLSTAENDIGES LITVERZ
3006: 
3007: \begin{thebibliography}{99}
3008: 
3009: 
3010: \bibitem{ACG}
3011:  J.M.\ Arms, R.H.\ Cushman, M.J.\ Gotay:
3012:  A universal reduction procedure for Hamiltonian group actions.
3013:  In: {\it The geometry of Hamiltonian systems (Berkeley, CA, 1989)},
3014:  Math.\ Sci.\ Res.\ Inst.\ Publ.\ 22, Springer, New York, 1991, pp.\ 33--51
3015: 
3016: 
3017: \bibitem{BaajJulg}
3018:  S.\ Baaj, P.\ Julg:
3019:  Th\'eorie bivariante de Kasparow et op\'erateurs non born\'es dans les
3020:  $C^\ast$-modules hilbertien.
3021:  C.\ R.\ Acad.\ Sci.\ Paris, Sér. I Math., 296 (1983), no.\ 21, 875--878
3022: 
3023: \bibitem{Buchholz82}
3024:  D.\ Buchholz:
3025:  The physical state space of quantum electrodynamics.
3026:  Commun.\ Math.\ Phys.\ {\bf 85} (1982) 49--71
3027: 
3028: 
3029: \bibitem{Buchholz86}
3030:  D.\ Buchholz:
3031:  Gauss' law and the infraparticle problem.
3032:  Phys.\ Lett. B {\bf 174} (1986) 331--334
3033: 
3034: 
3035: \bibitem{BuchholzGrundling}
3036:  D.\ Buchholz, H.\ Grundling:
3037:  The Resolvent Algebra: A New Approach to Canonical Quantum Systems.
3038:  arXiv:0705.1988v3; to appear in J.\ Funct.\ Anal.
3039: 
3040: \bibitem{DHR71}
3041:  S.\ Doplicher, R.\ Haag, J.\ Roberts:
3042:  Local observables and particle statistics. I.
3043:  Commun.\ Math.\ Phys.\ {\bf 23}(1971) 199--230
3044: 
3045: 
3046: \bibitem{DHR74}
3047:  S.\ Doplicher, R.\ Haag, J.\ Roberts:
3048:  Local observables and particle statistics. II.
3049:  Commun.\ Math.\ Phys.\ {\bf 35} (1974) 49--85
3050: 
3051: 
3052: \bibitem{DR90}
3053:  S.\ Doplicher, J.\ Roberts:
3054:  Why there is a field algebra with a compact gauge group describing the
3055:  superselection structure in particle physics.
3056:  Commun.\ Math.\ Phys.\ {\bf 131} (1990) 51--107
3057: 
3058: 
3059: \bibitem{Fredenhagen}
3060:  K.\ Fredenhagen, M.\ Marcu:
3061:  Charged states in $\ZZ_2$ gauge theories.
3062:  Commun.\ Math.\ Phys.\ {\bf 92} (1983) 81--119
3063: 
3064: 
3065: \bibitem{Froehlich}
3066:  J.\ Fr\"ohlich:
3067:  The charged sectors of quantum electrodynamics in a framework of local
3068:  observables.
3069:  Commun.\ Math.\ Phys.\ {\bf 66} (1979) 223--265
3070: 
3071: 
3072: \bibitem{Hall:Compact}
3073:  B.C.\ Hall:
3074:  Geometric quantization and the generalized Segal-Bargmann transform
3075:  for Lie groups of compact type.
3076:  Commun.\ Math.\ Phys.\ {\bf 226} (2002) 233--268
3077: 
3078: 
3079: \bibitem{Helgason}
3080:  S.\ Helgason:
3081:  {\em Groups and geometric analysis. Integral geometry, invariant differential
3082:  operators, and spherical functions.}
3083:  Academic Press 1984
3084: 
3085: 
3086: \bibitem{Hue:qr}
3087:  J.\ Huebschmann:
3088:  K\"ahler quantization and reduction.
3089:  J.\ reine angew.\ Math.\ \textbf{591} (2006) 75--109
3090: 
3091: 
3092: \bibitem{Hue:bedlewo}
3093:  J.\ Huebschmann:
3094:  Singular Poisson-K\"ahler geometry of certain adjoint quotients.
3095:  In: {\em The mathematical legacy of C.\ Ehresmann}, Proceedings, Bedlewo,
3096:  2005, Banach Center Publications (to appear),
3097:  {\tt math.SG/0610614}
3098: 
3099: 
3100: \bibitem{Hue:holopewe}
3101:  J.\ Huebschmann:
3102:  Kirillov's character formula, the holomorphic Peter-Weyl theorem, and the
3103:  Blattner-Kostant-Sternberg pairing.
3104:  J.\ Geom.\ Phys.\ 58 (2008), 833-848; {\tt math.DG/0610613}
3105: 
3106: 
3107: \bibitem{hrs}
3108:  J.\ Huebschmann, G.\ Rudolph, M.\ Schmidt:
3109:  A gauge model for quantum mechanics on a stratified space.
3110:  hep-th/0702017
3111: 
3112: \bibitem{hrs2}
3113:  J.\ Huebschmann, G.\ Rudolph, M.\ Schmidt:
3114:  Quantum mechanical tunneling on a space with singularities.
3115:  H.D.\ Doebner, V.K.\ Dobrev (eds.): {\it Lie Theory and its Applications in
3116:  Physics.} Proc.\ of the VIIth International Workshop, Varna, Bulgaria, 18-24
3117:  June, 2007. Heron Press, Sofia, 2008, pp.\ 190--210
3118: 
3119: \bibitem{JaKiRu}
3120:  P.D.\ Jarvis, J.\ Kijowski, G.\ Rudolph:
3121:  On the structure of the observable algebra of QCD on the lattice.
3122:  J.\ Phys.\ A {\bf 38}, 5359--5377 (2005)
3123: 
3124: 
3125: \bibitem{KaRi} Kadison, Ringrose
3126:  R.V.\ Kadison, J.R.\ Ringrose:
3127:  {\em Fundamentals of the Theory of Operator Algebras.}
3128:  Academic Press 1986
3129: 
3130: 
3131: \bibitem{KiRu02}
3132:  J.\ Kijowski, G.\ Rudolph:
3133:  On the Gauss law and global charge for quantum chromodynamics.
3134:  J.\ Math.\ Phys.\ {\bf 43} (2002) 1796--1808
3135: 
3136: 
3137: \bibitem{KiRu05}
3138:  J.\ Kijowski, G.\ Rudolph:
3139:  Charge superselection sectors for qcd on the lattice.
3140:  J.\ Math.\ Phys.\ {\bf 46} (2005) 032303
3141: 
3142: 
3143: \bibitem{KiRuSl}
3144:  J.\ Kijowski, G.\ Rudolph, C.\ \'Sliwa:
3145:  On the structure of the observable algebra for QED on the lattice.
3146:  Lett.\ Math.\ Phys.\ {\bf 43}, 299--308 (1998);~
3147: 
3148: 
3149: \bibitem{KiRuTh}
3150:  J.\ Kijowski, G.\ Rudolph, A.\ Thielmann:
3151:  Algebra of observables and charge superselection sectors for QED on the
3152:  lattice.
3153:  Commun.\ Math.\ Phys.\ {\bf 188}, 535--564 (1997)
3154: 
3155: 
3156: \bibitem{Lance}
3157:  E.\ Lance: {\it Hilbert $C^\ast$-modules.}
3158:  London Mathematical Society Lecture Note Series, 210.
3159:  Cambridge University Press, Cambridge, 1995
3160: 
3161: 
3162: \bibitem{Pedersen}
3163:  Pedersen, G.K.:
3164:  {\em $C\sp{*} $-algebras and their automorphism groups.}
3165:  London Mathematical Society Monographs, 14.
3166:  Academic Press, Inc., London-New York, 1979
3167: 
3168: 
3169: \bibitem{rsv:review}
3170:  G.\ Rudolph, M.\ Schmidt, I.P.\ Volobuev:
3171:  On the Gauge Orbit Space Stratification. A Review.
3172:  J.\ Phys.\ A: Math.\ Gen.\ {\bf 35} (2002) R1--R50
3173: 
3174: \bibitem{Schwarz}
3175: Schwarz, G.W.:
3176: Smooth functions invariant under the action of a compact Lie group.
3177: Topology {\bf 14} (1975) 63--68
3178: 
3179: 
3180: \bibitem{StrocchiWightman}
3181:  F.\ Strocchi, A.\ Wightman:
3182:  Proof of the charge superselection rule in local relativistic quantum field
3183:  theory.
3184:  J.\ Math.\ Phys.\ {\bf 15} (1974) 2198--2224
3185:  \\
3186:  F.\ Strocchi:
3187:  Local and covariant gauge quantum field theories. Cluster property,
3188:  superselection rules, and the infrared problem.
3189:  Phys.\ Rev.\ D {\bf 17} (1978) 2010--2021
3190: 
3191: 
3192: \bibitem{Weyl}
3193:  Weyl, H.:
3194:  {\it The classical groups.}
3195:  Princeton University Press, Princeton, N.J., 1946
3196: 
3197: 
3198: \bibitem{Woro}
3199:  S.L.\ Woronowicz:
3200:  $C^\ast$-algebras generated by unbounded elements.
3201:  Rev.\ Math.\ Phys.\ {\bf 7} (1995) 481--521
3202: 
3203: 
3204: \bibitem{Woodhouse}
3205:  J.\ \'Sniatycki:
3206:  {\em Geometric quantization and Quantum Mechanics.}
3207:  Springer, 1980
3208: \\
3209:  N.M.J.\ Woodhouse:
3210:  {\em Geometric quantization.}
3211:  Clarendon Press, Oxford 1991
3212: 
3213: \bibitem{Wren}
3214:  K.K.\ Wren:
3215:  Quantization of constrained systems with singularities using Rieffel
3216:  induction. J.\ Geom.\ Phys.\ {\bf 24} (1998) 173--202;
3217:  Constrained quantisation and $\theta$-angles. II.
3218:  Nucl.\ Phys.\ {\bf B 521} (1998) 471--502
3219: 
3220: 
3221: 
3222: 
3223: \end{thebibliography}
3224: 
3225: