1: \documentclass[12pt]{article}
2:
3:
4:
5:
6: \usepackage{latexsym}
7: \usepackage{bbm}
8: \usepackage{amsopn,amscd}
9: \usepackage{amsfonts}
10: \usepackage{amssymb}
11: \usepackage{amsthm}
12: \usepackage{amsmath}
13: \usepackage{epsfig,graphics}
14: %\usepackage{t1enc}
15: %\usepackage[cp1250]{inputenc}
16: \usepackage{a4wide}
17: %\usepackage[all]{xy}
18: \usepackage[square,authoryear]{natbib}
19:
20:
21:
22: \topmargin-1.5cm \oddsidemargin-0.55cm \evensidemargin-0.55cm
23: \textheight23cm \textwidth16cm \pagenumbering{arabic}
24: %\parindent0pt
25:
26: \newtheorem{Theorem}{Theorem}[section]
27: \newtheorem{Proposition}[Theorem]{Proposition}
28: \newtheorem{Lemma}[Theorem]{Lemma}
29: \newtheorem{Corollary}[Theorem]{Corollary}
30: \newtheorem{Definition}[Theorem]{Definition}
31: \theoremstyle{remark}
32: \newtheorem{Remark}[Theorem]{Remark}
33:
34:
35: % Beginn Befehle Johannes
36:
37: \newcommand{\nl}{\hfill\newline}
38:
39: \def\zz{z}
40: \def\empha{\em}
41: \def\emphas{}
42: \def\bemphas{}
43: \def\pemphas{}
44: \def\scsc{\sc}
45: \newtheorem{thm}{Theorem}[section]
46: \newtheorem{cor}[thm]{Corollary}
47: \newtheorem{lem}[thm]{Lemma}
48: \newtheorem{prop}[thm]{Proposition}
49: \newtheorem{defi}[thm]{Definition}
50: \newtheorem{examp}[thm]{Example}
51: \newtheorem{rema}[thm]{Remark}
52: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
53:
54: \long
55: \def\MSC#1\EndMSC{\def\arg{#1}\ifx\arg\empty\relax\else
56: {\par\narrower\noindent%
57: 2000 Mathematics Subject Classification: #1\par}\fi}
58:
59: \long
60: \def\KEY#1\EndKEY{\def\arg{#1}\ifx\arg\empty\relax\else
61: {\par\narrower\noindent%
62: Keywords and Phrases: #1\par}\fi}
63:
64: % Ende Befehle Johannes
65:
66:
67: \newcommand{\comment}[1]{}
68: \newcommand{\todo}[1]{}
69: \newcommand{\aufgabe}[1]{{\tt #1}}
70: \newcommand{\acosh}{\mr{acosh}}
71: \newcommand{\meta}{\zeta}
72: \newcommand{\halfsum}{\vr}
73: \newcommand{\weight}{\gamma}
74: \newcommand{\weightc}{\gamma^c}
75: \newcommand{\se}{\mr{se}}
76: \newcommand{\ce}{\mr{ce}}
77: \newcommand{\vanifac}{f}
78: \newcommand{\vanisum}{F}
79: \newcommand{\inco}{\nu}
80:
81: \newcommand{\prin}{1}
82: \newcommand{\sing}{0}
83: \newcommand{\pha}{{\mc P}}
84: \newcommand{\cfg}{{\mc X}}
85: \newcommand{\BS}{\text{BS}}
86: \newcommand{\akonst}{c}
87: \newcommand{\opt}{\mr{opt}}
88: \newcommand{\Hi}{\mc{H}}
89: \newcommand{\group}{K}
90: \newcommand{\lieal}{\mf{k}}
91: \newcommand{\hol}{\mr{hol}}
92: \newcommand{\sr}{\mr{S}}
93: \newcommand{\tinco}{{\tilde{\nu}}}
94: \newcommand{\coco}{g}
95: \newcommand{\vol}{\mr{vol}}
96: \newcommand{\scapro}[2]{\langle #1,#2 \rangle}
97: \newcommand{\scaproC}[2]{\langle #1,#2 \rangle}
98: \newcommand{\scale}{\beta}
99: \newcommand{\ratio}{\tau}
100: \newcommand{\vani}{\mc{V}}
101: \newcommand{\even}{\mr{g}}
102: \newcommand{\odd}{\mr{u}}
103: \newcommand{\ket}[1]{{|#1\rangle}}
104: \newcommand{\abs}[1]{{|#1|}}
105:
106:
107:
108: \newcommand{\mc}[1]{\mathcal{#1}}
109: \newcommand{\mr}[1]{\mathrm{#1}}
110: \newcommand{\SU}{\mr{SU}}
111: \newcommand{\su}{\mr{su}}
112: \newcommand{\Gl}{\mr{GL}}
113: \newcommand{\gl}{\mr{gl}}
114: \newcommand{\SO}{\mr{SO}}
115: \newcommand{\so}{\mr{so}}
116: \newcommand{\SL}{\mr{SL}}
117: \newcommand{\GL}{\mr{GL}}
118: \newcommand{\ddtn}{\left.\frac{\mr d}{\mr d t}\right|_{t=0}}
119: \newcommand{\tree}{{\mc T}}
120: \newcommand{\Pol}{{\mr{Pol}}}
121: \newcommand{\pr}{{\mr{pr}}}
122: \renewcommand{\Re}{{\mr{Re}}}
123: \renewcommand{\Im}{{\mr{Im}}}
124: \newcommand{\ol}[1]{\overline{#1}}
125: \newcommand{\tr}{\mathrm{tr}}
126: \newcommand{\id}{\mathrm{id}}
127: \newcommand{\RR}{\mathbb{R}}
128: \newcommand{\CC}{\mathbb{C}}
129: \newcommand{\ZZ}{\mathbb{Z}}
130: \newcommand{\II}{\mathbbm{1}}
131: \newcommand{\ctg}{\mr T^\ast}
132: \newcommand{\tg}{\mr T}
133: \newcommand{\diag}{{\rm diag}}
134: \newcommand{\mb}{\mathbf}
135: \newcommand{\mf}{\mathfrak}
136: \newcommand{\maxtor}{T}
137: \newcommand{\U}{{\mr U}}
138: \newcommand{\Ad}{{\mr{Ad}}}
139: \newcommand{\stab}{\mr{stab}}
140: \newcommand{\rref}[1]{{\rm \ref{#1}}}
141: \newcommand{\vp}{\varphi}
142: \newcommand{\vr}{\varrho}
143: \newcommand{\ve}{\varepsilon}
144: \newcommand{\ble}{\begin{Lemma}}
145: \newcommand{\ele}{\end{Lemma}}
146: \newcommand{\bpp}{\begin{Proposition}}
147: \newcommand{\epp}{\end{Proposition}}
148: \newcommand{\bpf}{\begin{proof}}
149: \newcommand{\epf}{\end{proof}}
150: \newcommand{\btm}{\begin{Theorem}}
151: \newcommand{\etm}{\end{Theorem}}
152: \newcommand{\bre}{\begin{Remark}\rm}
153: \newcommand{\ere}{\end{Remark}}
154: \newcommand{\beq}{\begin{equation}}
155: \newcommand{\eeq}{\end{equation}}
156: \newcommand{\beqa}{\begin{eqnarray}}
157: \newcommand{\eeqa}{\end{eqnarray}}
158: \newcommand{\beqast}{\begin{eqnarray*}}
159: \newcommand{\eeqast}{\end{eqnarray*}}
160: \newcommand{\linie}[3]{\put(#1){\line(#2){#3}}}
161: \newcommand{\marke}[3]{\put(#1){\put(0.05,0.1){\makebox(-0.1,-0.2)[#2]{$#3$}}}}
162: %\newcommand{\ddtn}{\frac{\mr d}{\mr d t}}
163:
164: \numberwithin{equation}{section}
165:
166: %\renewcommand{\epsfig}[1]{}
167: %\renewcommand{\II}{1}
168:
169:
170: \begin{document}
171:
172: \title{\bf A gauge model for quantum mechanics\\ on a stratified space}
173:
174: \author{
175: J.~Huebschmann$^1$, G.~Rudolph$^2$, M.~Schmidt$^2$
176: \\[0.3cm]
177: $^1$ USTL, UFR de Math\'ematiques\\
178: CNRS-UMR 8524
179: \\
180: 59655 Villeneuve d'Ascq C\'edex, France\\
181: Johannes.Huebschmann@math.univ-lille1.fr
182: \\[0.3cm]
183: $^2$ Institute for Theoretical Physics, University of Leipzig\\
184: Augustusplatz 10/11, 04109 LEIPZIG, Germany
185: \\[0.3cm]
186: }
187:
188:
189: \maketitle
190:
191:
192: \begin{abstract}
193:
194: \noindent
195: In the Hamiltonian approach on a single spatial
196: plaquette, we construct a quantum (lattice) gauge theory which
197: incorporates the classical singularities. The reduced phase space is
198: a stratified K\"ahler space, and we make explicit the requisite
199: singular holomorphic quantization procedure on this space. On the
200: quantum level, this procedure yields a costratified Hilbert
201: space, that is, a Hilbert space together with a system which
202: consists of the
203: subspaces associated with the strata of the reduced phase space and
204: of the corresponding orthoprojectors. The costratified Hilbert space
205: structure reflects the stratification of the reduced phase space.
206: For the special case where the structure group is
207: $\mathrm{SU}(2)$, we discuss the tunneling probabilities between
208: the strata, determine the energy eigenstates and study the
209: corresponding expectation values of the orthoprojectors onto the
210: subspaces associated with the strata in the strong and weak
211: coupling approximations.
212:
213: \end{abstract}
214:
215:
216:
217: \newpage
218:
219: \tableofcontents
220:
221:
222:
223: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
224: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
225:
226: \section{Introduction}
227:
228: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
229: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
230:
231:
232: According to {\scsc Dirac}, the correspondence between a
233: classical theory and its quantum counterpart should be based on an
234: analogy between their mathematical structures. An interesting
235: issue is then that of the role of singularities in quantum
236: problems. Singularities are known to arise in classical phase
237: spaces. For example, in the Hamiltonian picture of a theory,
238: reduction modulo symmetries leads in general to singularities on
239: the classical level. Thus the question arises whether, on the
240: quantum level, there is a suitable structure having the classical
241: singularities as its shadow and whether and how we can uncover
242: it. As far as we know, one of the first papers in this topic is
243: that of {\scsc Emmrich and R\"omer\/} \cite{emmroeme}. This paper
244: indicates that wave functions may \lq\lq congregate\rq\rq\ near a
245: singular point, which goes counter to the sometimes quoted
246: statement that singular points in a quantum problem are a set
247: of measure zero so cannot possibly be important. In a similar
248: vein, {\scsc Asorey et al\/} observed that vacuum nodes correspond
249: to the chiral gauge orbits of reducible gauge fields with
250: non-trivial magnetic monopole components \cite {asfalolu}. It is
251: also noteworthy, cf.\ e.g.\ \cite{armcusgo} and the references
252: there, that in classical mechanics and in classical field theories
253: singularities in the solution spaces are the rule rather than
254: the exception. This is in particular true for Yang-Mills
255: theories and for Einstein's gravitational theory; see for example
256: \cite{armamonc,armamotw}.
257:
258: In \cite{kaehler}, one of us isolated a certain class
259: of K\"ahler spaces with singularities, referred to as stratified K\"ahler
260: spaces. To explore the potential impact of classical
261: phase space singularities on quantum problems, in \cite{qr}, he then developed
262: the notion of costratified Hilbert space. This is the
263: appropriate quantum state space over a stratified space; it
264: consists of a system of Hilbert spaces, one for each stratum which
265: arises from quantization on the closure of that stratum, the
266: stratification provides bounded linear operators between these
267: Hilbert spaces reversing the partial ordering among the strata,
268: and these linear operators are compatible with the quantizations.
269: The notion of costratified Hilbert space is, perhaps, the
270: quantum structure which has the classical singularities as its
271: shadow.
272: In \cite{qr}, the ordinary K\"ahler quantization scheme has been
273: extended to such a scheme over (complex analytic) stratified
274: K\"ahler spaces. The appropriate quantum Hilbert space is, in
275: general, a costratified Hilbert space. Examples abound; one such
276: class of examples, involving holomorphic nilpotent orbits and in
277: particular angular momentum zero spaces, has been treated in
278: \cite{qr}.
279:
280: Gauge theory in the Hamiltonian approach, phrased on a finite
281: spatial lattice, leads to tractable finite-dimensional models for
282: which one can analyze the role of singularities explicitly. Under
283: such circumstances, after a choice of tree gauge has been made,
284: the unreduced classical phase space amounts to the total space
285: $\ctg(\group\times\cdots\times\group)$ of the cotangent bundle
286: on a product of finitely many copies of the manifold underlying the
287: structure group $\group$. Gauge transformations
288: are then given by the lift of the action of $\group$ on
289: $\group\times\cdots\times\group$ by diagonal conjugation. This leads to a
290: finite-dimensional Hamiltonian system with symmetries. For first results on the
291: stratified structure of both the reduced configuration space and the reduced
292: phase space of systems of this type, see
293: \cite{cfg,cfgtop,cdy,locpois,modus}.
294: Within canonical quantization for the unreduced system, the
295: algebra of observables and its representations have been
296: extensively investigated, see \cite{qed1,qed2,qed3} for quantum
297: electrodynamics and \cite{qcd1,qcd2,qcd3} for quantum
298: chromodynamics. However, in this approach, the implementation of
299: singularities is far from being clear.
300:
301: In the present paper we will consider the case of one copy of $\group$. This
302: corresponds to a lattice consisting of a single plaquette.
303: The unreduced phase space $\ctg \group$ carries an invariant
304: complex structure, and the complex and cotangent bundle symplectic structures
305: combine to give an invariant K\"ahler
306: structure. Thus, the stratified K\"ahler quantization scheme of
307: \cite{qr} referred to above can be
308: applied. We construct the costratified Hilbert space on the
309: reduced phase space by reduction after quantization. Ordinary
310: half-form K\"ahler quantization on $\ctg \group$ yields a Hilbert
311: space of holomorphic and, therefore, continuous wave
312: functions on $\ctg \group$, and we take the total Hilbert space of
313: our theory to be the subspace of $\group$-invariants. Given a
314: stratum, we then consider the space of functions in the Hilbert
315: space which vanish on the stratum, and we take the orthogonal
316: complement of this space as the Hilbert space associated with the
317: stratum. Now, in the K\"ahler polarization, among the classical
318: observables, only the constants can be quantized directly.
319: However, the holomorphic Peter-Weyl theorem \cite{holopewe} or,
320: equivalently, a version of the Segal-Bargmann transform
321: \cite{bhallone}, yields an isomorphism between the total Hilbert
322: space arising from K\"ahler quantization and the Hilbert space of
323: the Schr\"odinger representation. Via this isomorphism, the
324: costratified structure passes to the Schr\"odinger picture. On the
325: other hand, observables defined in the Schr\"odinger picture via
326: half-form quantization, for example, the Hamiltonian, can be
327: transferred to the holomorphic picture as well. Our approach
328: includes the quantization of arbitrary conjugation invariant
329: Hamiltonian systems on the total space of the cotangent bundle of
330: a compact Lie group. In this paper we concentrate on the
331: particular case of $\SU(2)$ with a lattice gauge theoretic
332: Hamiltonian.
333:
334: The paper is organized as follows. In Section \rref{Sclassical} we
335: introduce the model and give a brief description of the stratified
336: K\"ahler structure of its reduced classical phase space. Section
337: \rref{Squantum} contains the construction of the costratified
338: Hilbert space structure for general $\SU(n)$. In Section
339: \rref{SSU2}, we then make this construction explicit for $\SU(2)$.
340: In Section \rref{Seigen} we determine the energy eigenvalues and
341: eigenstates of our model for $\SU(2)$. Finally, in Section
342: \rref{Sexpect}, we discuss the corresponding expectation values of
343: the orthoprojectors onto the subspaces associated with the strata
344: and derive approximations for strong and weak coupling.
345:
346:
347: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
348: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
349:
350: \section{The classical picture}
351: \label{Sclassical}
352:
353: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
354: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
355:
356:
357: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
358: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
359:
360: \subsection{The model}
361:
362: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
363: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
364:
365:
366: Let $\group$ be a compact connected Lie group
367: and let $\lieal$ be its Lie algebra. We consider
368: lattice gauge theory with structure group $\group$ in the
369: Hamiltonian approach on a single spatial plaquette. By means of a
370: tree gauge\footnote{For an arbitray lattice $\Lambda$,
371: a tree gauge amounts to a choice of maximal tree in $\Lambda$,
372: the parallel transporters along the on-tree links being set equal to the
373: identity of $\group$; this leaves the
374: parallel transporters along the off-tree links as variables and constant gauge
375: transformations as symmetries. In our simple example, there is only
376: one off-tree
377: link.}, the reduced phase space of the system can be shown to
378: be isomorphic, as a stratified symplectic space, to the reduced
379: phase space of the following simpler system. The unreduced
380: configuration space is the group manifold $\group$ and
381: gauge transformations are given by the action
382: of $\group$ upon itself by inner automorphisms.
383: The unreduced phase space is the
384: cotangent bundle $\ctg\group$, acted upon by the lifted action.
385: This action is well known to be Hamiltonian and the corresponding
386: momentum mapping $\mu\colon\ctg\group\to\lieal^\ast$ is given by a
387: familiar expression \cite{AbrahamMarsden}. We trivialize
388: $\ctg\group$ in the following fashion: Endow $\lieal$ with an
389: invariant positive definite inner product
390: $\langle\cdot,\cdot\rangle$; we could take, for example, the
391: negative of the Killing form, but this is not necessary. By means
392: of the inner product, we identify $\lieal$ with its dual
393: $\lieal^*$ and the total space $\tg \group$ of the tangent bundle
394: of $\group$ with the total space $\ctg \group$ of the cotangent
395: bundle of $\group$. Composing the latter identification with the
396: inverse of left translation we obtain a diffeomorphism
397: \beq\label{Gtrviz}
398: \ctg\group \to \tg\group \to \group\times\lieal\,.
399: \eeq
400: It is $\group$-bi-inivariant w.r.t.\ the action of
401: $\group\times\group$ on $\group\times\lieal$ given by
402: $$
403: (x,Y) \mapsto (axb,\Ad_{b^{-1}} Y)
404: \,,~~~~~~
405: a,b,x\in \group,~~ Y\in \lieal\,.
406: $$
407: In the variables $(x,Y)\in\group\times\lieal$, the lifted action of $K$ reads
408: $$
409: (x,Y) \mapsto (axa^{-1},\Ad_a Y)
410: \,,~~~~~~
411: a\in \group\,,
412: $$
413: and the symplectic potential $\theta \colon \tg\ctg \group\to \RR$ is given by
414: \beq\label{GsplPot}
415: \theta_{(x,Y)}(x V,W) = \langle Y,V \rangle
416: \,,~~~~~~
417: V,W\in\lieal\,,
418: \eeq
419: where the association $(x,V)\mapsto x V$ ($x\in \group,\, V \in
420: \lieal$) refers to left translation in $\mathrm T\group$. Accordingly, the
421: symplectic form $\omega = \mr -d \theta$ has the explicit description
422: $$
423: \omega_{(x,Y)} \big((xV_1,W_1),(xV_2,W_2)\big)
424: =
425: \big\langle V_1,W_2 \big\rangle
426: -
427: \big\langle W_1,V_2 \big\rangle
428: +
429: \big\langle Y,[V_1,V_2] \big\rangle
430: \,,
431: $$
432: where $V_1,V_2,W_1,W_2\in\lieal$. The Poisson bracket of functions
433: $f,g\in C^\infty(\group\times\lieal)$ is given by
434: \beq\label{GPoibra}
435: \{f,g\}(x,Y)
436: =
437: \big\langle f_\group(x,Y), g_\lieal(x,Y) \big\rangle
438: -
439: \big\langle f_\lieal(x,Y), g_\group(x,Y) \big\rangle
440: -
441: \big\langle Y , [f_\lieal(x,Y),g_\lieal(x,Y)] \big\rangle
442: \,,
443: \eeq
444: where $f_\group$ and $f_\lieal$ are $\lieal$-valued functions on
445: $\group\times\lieal$ representing the partial derivatives of $f$
446: along $\group$ and $\lieal$, respectively. They are defined by
447: $$
448: \big\langle f_\group(x,Y) , Z \big\rangle
449: =
450: \left.\frac{\mr d }{\mr d t}\right|_{t=0} f(x\mr e^{t Z},Y)
451: \,,~~~~~~
452: \big\langle f_\lieal(x,Y) , Z \big\rangle
453: =
454: \left.\frac{\mr d }{\mr d t}\right|_{t=0} f(x,Y + tZ)\,,
455: $$
456: for any $Z\in\mf g$. The momentum mapping $\mu$ takes the form
457: \beq\label{Gmomap}
458: \mu(x,Y)=\Ad_x Y - Y, \ x\in \group,\, Y \in \lieal\,.
459: \eeq
460: In \cite{cfg,cfgtop,cdy}, $\ctg\group$ has been trivialized by
461: right translation and the sign conventions necessarily differ. The
462: (classical unreduced) Hamiltonian $H\colon \ctg\group \to \mathbb
463: R$ of our model is given by
464: \beq
465: \label{GHaFn}
466: H(x,Y)
467: =
468: \frac{1}{2} |Y|^2
469: +
470: \frac{\inco}{2}
471: \left(3 - \Re\,\tr(x)\right), \ x\in \group,\, Y \in \lieal\,.
472: \eeq
473: Here $|\cdot|$ denotes the norm defined by the inner product on
474: $\lieal$, the constant $\nu$ is defined by $\inco = 1/\coco^2$, where $\coco$ is
475: the coupling constant, and the trace refers to some representation; below
476: we will suppose $\group$ to be realized as a closed subgroup of some unitary group
477: $\U(n)$. Moreover, we have set the lattice spacing equal
478: to $1$. The Hamiltonian $H$ is manifestly gauge invariant.
479:
480:
481: \bre \label{Remmodel}
482:
483: Ordinary Yang-Mills theory on $S^1$ proceeds by reduction relative
484: to the group of all gauge transformations. As an intermediate
485: step, one can perform reduction relative to the group of based
486: gauge transformations. This procedure provides our unreduced
487: model, i.~e., the Hamiltonian $\group$-space $\ctg \group$. Thus,
488: this model recovers a true continuum theory. Starting at the
489: lattice theory on a single plaquette, we have bypassed the
490: reduction relative to the group of based gauge transformations.
491: Our model therefore includes the continuum theory on $S^1$ and
492: serves as a building block of a lattice gauge theory as well.
493:
494: The quantization of Yang-Mills theory on $S^1$ in the Hamiltonian
495: approach has been worked out in
496: \cite{Dimock, DriverHall, Hall:Coherent, Hetrick, LandsmanWren:Theta,
497: LandsmanWren:Coherent, Wren:Rieffel, Wren:ThetaII}. In
498: \cite{LandsmanWren:Theta,LandsmanWren:Coherent,Wren:ThetaII} the
499: authors proceed through Rieffel induction, starting from the full
500: continuum theory, and arrive at the Hilbert space $L^2(\group,\mr
501: d x)^\group$ of square-integrable functions on $\group$ invariant
502: under inner automorphisms of $\group$. See also \cite[\S\S
503: IV.3.7,8]{Landsman} and the references there. We shall arrive at
504: the same Hilbert space almost immediately, as we start at a later
505: stage in the reduction procedure, but this is only a preliminary
506: stage for what we are aiming at: the construction of a costratified Hilbert
507: space to study the role of singularities in the quantum theory.
508:
509: \ere
510:
511:
512:
513:
514:
515:
516:
517:
518:
519: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
520: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
521:
522: \subsection{The K\"ahler structure on the unreduced phase space}
523: \label{SSKaestr}
524:
525: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
526: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
527:
528:
529: We recall that a K\"ahler manifold is a complex manifold, endowed
530: with a positive definite Hermitian form whose imaginary part,
531: necessarily an ordinary real 2-form, is closed and non-degenerate
532: and hence a symplectic structure. Equivalently, a K\"ahler
533: manifold is a smooth manifold, endowed with a complex and a
534: symplectic structure, and the two structures are required to be
535: compatible. One way of phrasing the compatibility condition is to
536: require that Poisson brackets of holomorphic functions be zero.
537:
538: The unreduced phase space $\ctg\group$ acquires a K\"ahler
539: structure in the following manner:
540: We suppose $\group$ realized as a closed subgroup of some unitary group
541: $\U(n)$; then the complexification $\group^{\mathbb C}$ of $\group$
542: is the complex subgroup of $\GL(n,\CC)$ generated by $K$.
543: By restriction, the polar
544: decomposition map
545: $$
546: \U(n) \times \mr u(n) \longrightarrow \GL(n,\CC)
547: \,,~~~~~~
548: (x,Y) \longmapsto x \mr e^{iY}\,,
549: $$
550: yields a diffeomorphism
551: %
552: \begin{equation}\label{polar}
553: \group\times \lieal \longrightarrow \group^{\mathbb C}
554: \,,~~~~~~
555: (x,Y)\longmapsto x\,\mathrm e^{iY}\,,
556: \end{equation}
557: %
558: commonly referred to as the polar decomposition of
559: $\group^{\mathbb C}$. The polar decomposition is manifestly
560: $\group$-bi-invariant w.r.t.\ the action of $\group\times\group$
561: on $\group\times\lieal$ spelled out above. Thus, the composite of
562: the trivialization \eqref{Gtrviz} of $\ctg\group$ with the polar
563: decomposition map \eqref{polar} is a $\group$-bi-invariant
564: diffeomorphism $\ctg\group\to\group^\CC$. The resulting complex
565: structure on $\ctg\group \cong \group^\CC$ and the
566: cotangent bundle symplectic structure combine to give
567: a $\group$-bi-invariant
568: K\"ahler structure, having as global K\"ahler potential the real
569: analytic function $\kappa$ given by
570: \begin{equation}
571: \kappa(x\,\mathrm e^{iY}) =|Y|^2. \label{kappa}
572: \end{equation}
573: An explicit calculation which justifies this assertion may be
574: found in \cite{bhallone}.
575:
576:
577:
578: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
579: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
580:
581: \subsection{Symmetry reduction}
582: \label{SSsymred}
583:
584: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
585: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
586:
587:
588: Let $\cfg$ denote the adjoint quotient $\group/\Ad$; this is the
589: reduced configuration space of our model. In the standard manner,
590: we decompose $\cfg$ as a disjoint union $\cfg = \bigcup_{\tau,i}
591: \cfg_{\tau,i}$. Here, $\tau$ ranges over the orbit types of the
592: action, $\cfg_{\tau}$ denotes the subset of $\cfg$ which
593: consists of orbits of type $\tau$, and $i$ labels the
594: connected components of this subset. We will refer to this
595: decomposition as the orbit type stratification of $\cfg$.
596: It is a stratification in the sense of e.~g.\ Goresky-MacPherson
597: \cite{Goresky}. For our purposes it suffices to know that it is a
598: manifold decomposition in the ordinary sense, i.\ e., the
599: $\cfg_{\tau,i}$ are manifolds and the frontier condition holds,
600: viz. $\cfg_{\tau_1,i_1}\subseteq\ol{\cfg_{\tau_2,i_2}}$ whenever
601: $\cfg_{\tau_1,i_1}\cap\ol{\cfg_{\tau_2,i_2}}\neq\emptyset$. An
602: explicit description of $\cfg$ arises from a choice of a maximal
603: toral subgroup $T\subseteq\group$. Let $W$ be the Weyl group of
604: $\group$. It is well known that the inclusion $T\hookrightarrow
605: \group$ induces a homeomorphism from the orbit space $T/W$ onto
606: the quotient $\cfg=\group/\Ad$ which identifies orbit type strata.
607:
608: The reduced phase space of our model is the zero momentum reduced
609: space $\mu^{-1}(0)/\group$ obtained by singular Marsden-Weinstein
610: reduction. We denote this space by $\pha$. It acquires a
611: stratified symplectic structure where, similarly to the reduced
612: configuration space $\cfg$, the stratification is given by the
613: connected components of the orbit type subsets, viz.\ $\pha =
614: \bigcup_{\tau,i}\pha_{\tau,i}$. An explicit description of $\pha$
615: is obtained as follows. Let $\mf t\subseteq\lieal$ be the Lie
616: algebra of $T$. Given $(x,Y)\in \group\times\lieal$, according to
617: \eqref{Gmomap}, the vanishing of $\mu(x,Y)$ implies that $x$ and
618: $Y$ commute. Hence, the pair $(x,Y)$ is conjugate to an element of
619: $T\times\mf t$ and the injection $T\times\mf
620: t\hookrightarrow\group\times\lieal$ induces a homeomorphism of
621: $\pha$ onto the quotient $(T\times\mf t)/W$ where $W$ acts
622: simultaneously on $T$ and $\mf t$. This homeomorphism identifies
623: orbit type strata.
624:
625: In the case $\group = \SU(n)$, the torus
626: $T$ can be chosen as the subgroup
627: of diagonal matrices in $\group$. Then $\mf t$ is the subalgebra
628: of diagonal matrices in $\lieal$. The Weyl group $W$ is the
629: symmetric group $S_n$ on $n$ letters, acting on $T$ and $\mf t$ by
630: permutation of entries. The reduced configuration space $\cfg\cong
631: T/W$ amounts to an $(n-1)$-simplex and the orbit type strata
632: correspond to its (open) subsimplices. In particular, the orbit
633: types are labelled by partitions $n = n_1 + \cdots + n_k$ of $n$
634: where the $n_i$'s are positive integers reflecting the
635: multiplicities of the entries of the elements of $T$. Concerning
636: the reduced phase space $\pha$, the orbit types of the action of
637: $W$ on $T\times\mf t$ are given by partitions of $n$ again, where
638: the $n_i$'s now are the dimensions of the common eigenspaces
639: of pairs in $T\times \mf t$.
640:
641: For later use, we shall describe $\cfg$ and $\pha$ for $\group =
642: \SU(2)$ in detail. Here, $T$ amounts to the complex unit circle
643: and $\mf t$ to the imaginary axis. Then the Weyl group $W=S_2$
644: acts on $T$ by complex conjugation and on $\mf t$ by reflection.
645: Hence, the reduced configuration space $\cfg \cong T/W$ is
646: homeomorphic to a closed interval and the reduced phase space
647: $\pha \cong (T\times\mf t)/W$ is homeomorphic to the well-known
648: canoe, see Figure \rref{FKanu}. Corresponding to the partitions
649: $2=2$ and $2=1+1$, there are two orbit types. We denote them by
650: $\sing$ and $\prin$, respectively. The orbit type subset
651: $\cfg_\sing$ consists of the classes of $\pm\II$, i.~e., of the
652: endpoints of the interval; it decomposes into the connected
653: components $\cfg_+$, consisting of the class of $\II$, and
654: $\cfg_-$, consisting of the class of $-\II$. The orbit type subset
655: $\cfg_\prin$ is connected and consists of the remaining classes,
656: i.~e., of the interior of the interval. The orbit type subset
657: $\pha_\sing$ consists of the classes of $(\pm\II,0)$, i.~e., of
658: the vertices of the canoe; it decomposes into the connected
659: components $\pha_+$, consisting of the class of $(\II,0)$, and
660: $\pha_-$, consisting of the class of $(-\II,0)$. The orbit type
661: subset $\pha_\prin$ consists of the remaining classes, has
662: dimension $2$ and is connected.
663:
664: \begin{figure}
665:
666: \begin{center}
667:
668: \unitlength1cm
669:
670: \begin{picture}(6,3)
671: \put(1,0.27){
672: \marke{0,0}{tr}{\pha_+}
673: \marke{4,0}{tl}{~\pha_-}
674: \marke{4,2}{tl}{~\pha_\prin}
675: \put(0,0){\circle*{0.2}}
676: \put(4,0){\circle*{0.2}}
677: }
678: \put(-1,0){
679: \put(0,0){\epsfig{file=canoe.eps,width=6cm,height=3cm}}
680: }
681: \end{picture}
682:
683: \end{center}
684:
685: \caption{\label{FKanu} The reduced phase space $\pha$ for
686: $\group = \SU(2)$.}
687:
688: \end{figure}
689:
690: \bre
691:
692: In the case $\group = \SU(2)$, as a stratified symplectic space,
693: $\pha$ is isomorphic to the reduced phase space of a spherical
694: pendulum, reduced at vertical angular momentum $0$ (whence the
695: pendulum is constrained to move in a plane), see \cite{CuBa}.
696:
697: \ere
698:
699: In \cite{kaehler}, the notion of stratified K\"ahler space has
700: been introduced and it has been shown that, under more general
701: circumstances, the K\"ahler structure on $\ctg
702: \group\cong\group^\CC$ explained in Subsection \rref{SSKaestr}
703: descends to a stratified K\"ahler structure on $\pha$ which is
704: compatible with the stratified symplectic structure. A detailed
705: discussion of this stratified K\"ahler structure can be found in
706: \cite{adjoint,bedlewo}. For completeness, we include a brief
707: description in Subsection \rref{SSstrfKae} below. Since this will
708: not be needed for quantization, the reader who is interested in
709: the quantum theory only may skip this subsection.
710:
711:
712:
713:
714:
715:
716: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
717: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
718:
719: \subsection{The stratified K\"ahler structure on the reduced phase space}
720: \label{SSstrfKae}
721:
722: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
723: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
724:
725:
726: The Weyl group $W$ acts on $\ctg T$ by pull back
727: and on $T^\CC$ by permutation of entries.
728: The trivialization \eqref{Gtrviz} and the polar decomposition
729: \eqref{polar} combine to a $W$-equivariant diffeomorphism $\ctg T
730: \to T\times \mf t \to T^\CC$. This diffeomorphism,
731: in turn, induces a homeomorphism between $\pha$ and the quotients $\ctg
732: T/W\cong T^{\mathbb C}\big/W$. Moreover, as explained in
733: Subsection \rref{SSKaestr}, the symplectic structure of $\ctg T$
734: and the complex structure of $T^\CC$ combine to give a K\"ahler
735: structure on $\ctg T \cong T^\CC$. In the sequel, we shall stick
736: to the notation $T^\CC$.
737:
738:
739:
740:
741: Viewed as the orbit space $\ctg T\big/W$, $\pha$
742: inherits
743: a stratified symplectic
744: structure by
745: singular
746: Marsden-Weinstein reduction.
747: That is to say: (i) The
748: algebra $C^\infty(T^\CC)^W$ of ordinary smooth $W$-invariant functions on
749: $T^\CC$ inherits a Poisson bracket and thus yields a Poisson
750: algebra of continuous functions on $\pha\cong T^\CC/W$, (ii) for
751: each stratum, the Poisson structure yields an ordinary symplectic
752: Poisson structure on that stratum, and (iii) the restriction
753: mapping from $C^{\infty}(T^\CC)^W$ to the algebra of ordinary
754: smooth functions on that stratum is a Poisson map.
755:
756:
757: Viewed as the orbit space $T^\CC\big/W$, $\pha$ acquires a
758: complex analytic structure in the standard fashion.
759: The complex structure and the Poisson structure
760: combine to give a stratified K\"ahler structure on $\pha$
761: \cite{kaehler}, \cite{adjoint}, \cite{bedlewo}. Here the precise
762: meaning of the term \lq\lq stratified K\"ahler structure\rq\rq\ is
763: that the Poisson structure satisfies (ii) and (iii) above and that
764: the Poisson and complex structures satisfy the additional
765: compatibility requirement that for each stratum, necessarily a
766: complex manifold, the symplectic and complex structures on that
767: stratum combine to give an ordinary K\"ahler structure.
768:
769: In the case $K=\SU(n)$, the complex analytic structure admits the
770: following elementary description: Let $\mathrm{Diag}(n,\mathbb C)$
771: be the group of diagonal matrices in the full linear group
772: $\GL(n,\CC)$. The Weyl group $W$ acts on $\mathrm{Diag}(n,\mathbb
773: C)$ by permutation of entries and the injection of $T^{\mathbb C}$
774: into $\mathrm{Diag}(n,\mathbb C)$ is compatible with this action.
775: The $n$ elementary symmetric functions $\sigma_1,\dots,\sigma_n$
776: furnish a map
777: \[
778: (\sigma_1,\dots,\sigma_n)\colon \mathrm{Diag}(n,\mathbb C)
779: \longrightarrow \mathbb C^n
780: \]
781: into complex $n$-space $\mathbb C^n$. The restriction
782: \begin{equation}\label{proj}
783: (\sigma_1,\dots,\sigma_{n-1})\colon T^{\mathbb C} \longrightarrow
784: \mathbb C^{n-1}
785: \end{equation}
786: of that map to $T^{\mathbb C}$ identifies the orbit space
787: $\pha\cong T^{\mathbb C}\big/W$ with the affine subspace of $\mathbb C^n$
788: given by the equation $\sigma_n =1$ which, in turn, may be identified with a
789: copy of $\mathbb C^{n-1}$. In this way, $\pha$ inherits an obvious complex
790: structure. Thus, affine complex $n$-space $\mathbb C^n$
791: appears here as the space of normalized complex degree $n$ polynomials,
792: and the orbit space $T^{\mathbb C}/W$ amounts to the subspace of
793: normalized complex degree $n$ polynomials with constant
794: coefficient equal to 1. Indeed, a normalized degree $n$ polynomial
795: $p(z)=z^n+a_{1}z^{n-1} + \ldots+ a_{n-1} z + a_n$ decomposes into
796: its linear factors $p(z) = \prod_j (z-z_j)$, and the coefficients
797: $a_j$ are given by
798: \[
799: a_{j}=(-1)^j\sigma_j(z_1,\ldots,z_n),\ 1\leq j \leq n;
800: \]
801: up to the signs $(-1)^j$, the map $\sigma$ may thus be viewed as
802: that which sends the $n$-tuple $\zz_1, \ldots, \zz_n$ to the
803: unique normalized degree $n$ polynomial having $\zz_1, \ldots,
804: \zz_n$ as its zeros, the coefficients of degree $\leq n-1$ being
805: taken as coordinates on the space of polynomials. A more profound
806: analysis shows that, indeed, in terms of
807: $\mathrm {SL}(n,\mathbb C)$ and $\mathrm {GL}(n,\mathbb C)$, the
808: passage to the quotient (which is here realized via the map
809: \eqref{proj}) amounts to the assignment to a matrix in $\mathrm
810: {SL}(n,\mathbb C)$ (or $\mathrm {GL}(n,\mathbb C)$) of its
811: characteristic polynomial.
812:
813:
814: We shall now describe the stratified K\"ahler structure on $\pha$
815: explicitly for $\group = \SU(2)$. Here, $T^\CC$ consists of the diagonal
816: matrices $\mathrm{diag}(z,z^{-1})$ where $z\in\CC^\ast$. The
817: non-trivial element of $W$ interchanges $z$ and $z^{-1}$. To determine
818: the complex structure we note that the map \eqref{proj} is given by
819: the restriction of the first elementary symmetric function
820: $\sigma_1$ on $\mr{Diag}(2,\CC)$ to the subgroup $T^{\mathbb C}$,
821: i.e.,
822: \begin{equation}\label{sigmaSU2}
823: \sigma_1 \colon T^\CC \longrightarrow \mathbb C
824: \,,~~~~~~
825: \sigma_1(\mathrm{diag}(z,z^{-1})) = z+ z^{-1}\,;
826: \end{equation}
827: this map identifies $T^\CC/W\cong \pha$ with a copy of $\CC$ and thus
828: provides a holomorphic coordinate on $\pha$. In particular,
829: topologically, the canoe shown in Figure \rref{FKanu} is just an
830: ordinary plane.
831:
832: To arrive at a description of the Poisson algebra
833: $C^\infty(T^\CC)^W$, we recall that, once a choice of
834: finitely many generators, say $p$, for the algebra
835: $\RR[T^\CC]^W$of real $W$-invariant polynomials on $T^\CC$
836: has been made,
837: the resulting Hilbert map induces
838: a homeomorphism from $T^\CC/W\cong\pha$ onto a semi-algebraic
839: subset of $\RR^p$.
840: According to a theorem in \cite{Schwarz}, any
841: element of $C^\infty(T^\CC)^W$ can be written as a smooth function
842: in these generators. Hence, to describe the Poisson algebra
843: $C^\infty(T^\CC)^W$ it suffices to list the Poisson brackets of
844: these generators. In the case at hand, a set of generators for
845: $\RR[T^\CC]^W$ can be obtained as follows. The complexification
846: $\mathbb R[T^{\mathbb C}]_{\mathbb C}$ of $\mathbb R[T^{\mathbb
847: C}]$ is generated by $z,z^{-1},\overline z, \overline z^{-1}$.
848: Since the non-trivial element of $W$ interchanges $z$ and $z^{-1}$
849: as well as $\overline z$ and $\overline z^{-1}$, the subalgebra
850: $\RR[T^\CC]^W_{\mathbb C}$ of $W$-invariants is
851: generated by the three elementary bisymmetric functions
852: \[
853: \sigma_1 = z+z^{-1},\ \overline \sigma_1 = \overline z+ \overline
854: z^{-1}, \
855: \sigma = z \ol z ^{-1} + \ol z z ^{-1}\,,
856: \]
857: and this algebra may be
858: identified with the complexification of $\RR[T^\CC]^W$ in an obvious manner.
859: These generators are subject to the single defining relation
860: \beq\label{Grelcpx}
861: (\sigma_1^2-4)(\overline \sigma_1^2-4) = (\sigma_1 \overline
862: \sigma_1 -2\sigma)^2\,,
863: \eeq
864: see \cite{adjoint}. Hence, $\RR[T^\CC]^W$ is generated by the
865: three real functions $X$, $Y$ and $\sigma$, where $\sigma_1 = X +
866: \mr i Y$. For convenience, instead of $\sigma$, we use $\tau =
867: \frac{2-\sigma}4$. In view of \eqref{Grelcpx}, the generators
868: $X$, $Y$, $\tau$ are subject to the relation
869: \begin{equation}\label{relation}
870: Y^2 = (X^2 + Y^2 + 4 (\tau -1)) \tau\,.
871: \end{equation}
872: In terms of the real coordinates $x$ and $y$ on
873: $T^\CC\cong\CC^\ast$ defined by $z = x + \mr i y$,
874: \beq\label{GXx}
875: X = x + \frac{x}{r^2}
876: \,,~~~~~~
877: Y = y - \frac{y}{r^2}
878: \,,~~~~~~
879: \tau = \frac{y^2}{r^2}\,,
880: \eeq
881: where $r^2 = x^2 + y^2$. The obvious inequality $\tau \geq 0$
882: brings the semialgebraic nature of the quotient $\mathbb
883: T^{\mathbb C}\big/W$ to the fore.
884: To determine the Poisson brackets among the generators $X$, $Y$ and
885: $\tau$, we recall that, in terms of the coordinates $x$
886: and $y$, the symplectic structure on $T^\CC\cong \CC^\ast$ is
887: given by $\frac{1}{r^2} \mr d x \wedge \mr d y$ whence
888: \[
889: \{x,y\} = r^2.
890: \]
891: A straightforward calculation involving
892: \eqref{GXx} yields the Poisson brackets
893: $$
894: \{X,Y\} = X^2 + Y ^2 + 4(2 \tau -1)
895: \,,~~~~~~
896: \{X,\tau\} = 2(1- \tau)Y
897: \,,~~~~~~
898: \{Y,\tau\} = 2 \tau X\,.
899: $$
900: The Poisson structure vanishes at the two points $(X,Y)=(2,0)$ and
901: $(X,Y)=(-2,0)$ representing the orbit type strata $\pha_+$ and
902: $\pha_-$, respectively. Hence, the resulting complex algebraic
903: stratified K\"ahler structure on $\pha$ is singular at
904: these two points. Furthermore, solving \eqref{relation} for
905: $\tau$, we obtain
906: \[
907: \tau = \frac 12 \sqrt{Y^2+ \frac {(X^2 + Y^2-4)^2}{16}} -
908: \frac{X^2 + Y^2 -4}{8} ,
909: \]
910: whence, at $(X,Y)=(\pm 2,0)$, $\tau$ is not smooth as a function
911: of the variables $X$ and $Y$. Away from these two points, i.e., on
912: the principal stratum $\pha_\prin$, the Poisson structure is
913: symplectic. We refer to the stratified K\"ahler space under
914: discussion as the exotic plane with two vertices. More
915: details and, in particular, an interpretation in terms of
916: discriminant varieties, may be found in \cite{bedlewo}.
917:
918:
919: \bre
920:
921: The algebra $\mathbb R[T^{\mathbb C}]^W$ is the real coordinate
922: ring of $T^{\mathbb C}/W$, viewed as a real semi-algebraic set.
923: Similarly, for the description of the Poisson structure on $\pha$
924: we could have used a set of generators of, e.g., the algebra
925: $\RR[T\times\mf t]^W$ of real $W$-invariant polynomials on
926: $T\times\mf t$. This is the real coordinate ring of $(T\times\mf
927: t)/W$, viewed, in turn, as a semi-algebraic set. Since the
928: diffeomorphism $T\times\mf t \cong T^\CC$ is
929: not algebraic,
930: $\mathbb R[T^{\mathbb C}]^W$ and $\RR[T\times\mf t]^W$ correspond
931: to different subalgebras of the Poisson algebra
932: $C^\infty(T^\CC)^W$ defining the Poisson structure on $\pha\cong
933: T^\CC/W$.
934:
935: \ere
936:
937:
938:
939:
940: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
941: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
942:
943: \section{The quantum picture}
944: \label{Squantum}
945:
946: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
947: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
948:
949: Our aim is to push further, in the context of stratified spaces,
950: the ideas which underlie the program of geometric quantization. As
951: our physical Hilbert space we take a certain space of
952: square-integrable holomorphic functions which arises by K\"ahler
953: quantization \cite{Sniatycki,woodhous}. Through an analogue of the
954: Peter-Weyl theorem, this space is related with the physical
955: Hilbert space arising by ordinary Schr\"odinger quantization on
956: $\group$. Within this Hilbert space we construct the additional
957: structure of a costratification. Thereafter, we discuss
958: observables.
959:
960: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
961: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
962:
963: \subsection{Holomorphic quantization}
964:
965: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
966: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
967:
968: Let $\varepsilon$ be the symplectic (or Liouville) volume form on
969: $\mathrm T^*\group\cong \group^{\mathbb C}$. In terms of the polar
970: decomposition \eqref{polar}, we then have the identity
971: $\varepsilon=dxdY$ where $dx$ is the volume form on $\group$
972: yielding Haar measure, normalized so that it
973: coincides with the Riemannian volume measure on $\group$,
974: and where $dY$ is the form inducing
975: Lebesgue measure on $\lieal$, normalized by the inner product on $\lieal$.
976: Next, let $\eta$ be the real $\group$-bi-invariant analytic function on $\group
977: \times\lieal\cong\group^{\mathbb C}$ defined by
978: \[
979: \eta(x\,\mathrm e^{iY})
980: =\sqrt{\mathrm{det}\left(\frac{\sin(\mathrm{ad}(Y))}{\mathrm{ad}(Y)}\right)},
981: \ x \in \group, \,Y \in \lieal,
982: \]
983: the square root being the positive one. We note that $\eta^2$ is
984: the density of Haar measure on $\group^{\mathbb C}$ relative to
985: Liouville measure $\varepsilon$, cf.\
986: \cite{bhallfou} (Lemma 5). To express $\eta$ in terms of a root system, we
987: choose a dominant Weyl chamber in the Cartan subalgebra $\mf t$ of $\lieal$
988: and denote by $R^+$ the corresponding set of positive roots. Then, on
989: $T\times\mf t\cong T^\CC$, $\eta$ is given by
990: $$
991: \eta(x\,\mathrm e^{iY})
992: =
993: \prod\nolimits_{\alpha\in R^+} \frac{\sinh(\alpha(Y))}{\alpha(Y)}
994: \,,~~~~~~
995: x \in T, \,Y \in \mathfrak t\,,
996: $$
997: cf.\ \cite{bhallone} (2.10). Here the $\alpha$'s are the real
998: roots, given by $-\mr i$ times the ordinary complex roots. Let
999: $\kappa$ be the $\group$-bi-invariant real analytic function on
1000: $\group \times \lieal \cong \group^{\mathbb C}$ defined by
1001: \eqref{kappa}. Half-form K\"ahler quantization on $\group^{\mathbb
1002: C}$ yields the Hilbert space $\mathcal HL^2(\group^{\mathbb
1003: C},\mathrm e^{-\kappa/\hbar}\eta \varepsilon)$ of holomorphic
1004: functions on $\group^{\mathbb C}$ which are square-integrable
1005: relative to the measure $\mathrm e^{-\kappa/\hbar}\eta
1006: \varepsilon$ \cite{bhallone}. The scalar product is given by
1007: \beq\label{GscaproKC}
1008: \scaproC{\psi_1}{\psi_2}
1009: =
1010: \frac{1}{\vol(\group)} \int_{\group^\CC}
1011: \ol{\psi_1}\psi_2 \mathrm e^{-\kappa/\hbar}\eta \varepsilon\,.
1012: \eeq
1013: For our purpose there is no need to write down the relevant
1014: half-forms explicitly. They are subsumed under the measure.
1015:
1016: Left and right translation turn the Hilbert space $\mathcal
1017: HL^2(\group^{\mathbb C},\mathrm e^{-\kappa/\hbar}\eta
1018: \varepsilon)$ into a unitary representation of $\group\times
1019: \group$. The Hilbert space associated with $\pha$ by reduction
1020: after quantization is the subspace $\mathcal HL^2(\group^{\mathbb
1021: C},\mathrm e^{-\kappa/\hbar}\eta \varepsilon)^\group$ of
1022: $\group$-invariants relative to conjugation.
1023:
1024: We will now describe the Hilbert space $\mathcal
1025: HL^2(\group^{\mathbb C},\mathrm e^{-\kappa/\hbar}\eta
1026: \varepsilon)^\group$ as a Hilbert space of $W$-invariant
1027: holomorphic functions on $T^{\mathbb C}$ that are
1028: square-integrable relative to a measure of the kind $\mathrm
1029: e^{-\kappa/\hbar}\gamma \varepsilon_T$ for a suitable density
1030: function $\gamma$ on $T^{\mathbb C}$ where $\varepsilon_T$ denotes
1031: the Liouville volume form on $T^{\mathbb C}\cong \mathrm \ctg T$.
1032: This Hilbert space may in fact be viewed as coming from
1033: quantization after reduction, i.~e., by quantization on $T^\CC/W$.
1034: Here and below we do not distinguish in notation between the
1035: function $\mathrm e^{-\kappa/\hbar}$ defined on $\group^{\mathbb
1036: C}$ and its restriction to $T^{\mathbb C}$.
1037:
1038: Let $m=\dim \group$ and $r=\dim T$. To construct the function
1039: $\gamma$, consider the conjugation mapping
1040: \begin{equation} q^{\mathbb C}\colon
1041: \left(\group^{\mathbb C}\big/ T^{\mathbb C}\right) \times
1042: T^{\mathbb C} \longrightarrow \group^{\mathbb C},\ (y T^{\mathbb
1043: C},t)\mapsto yty^{-1},\ y\in \group^{\mathbb C}, t\in T^{\mathbb
1044: C}, \label{conjugation}
1045: \end{equation}
1046: and integrate the induced $(2m)$-form $(q^{\mathbb C})^*(\mathrm
1047: e^{-\kappa/\hbar} \eta \varepsilon)$ over \lq\lq the fibers\rq\rq\
1048: $\group^{\mathbb C}\big/ T^{\mathbb C}$. Although the fibers are
1049: non-compact, in view of the Gaussian constituent $\mathrm
1050: e^{-\kappa/\hbar}$, this integration is a well defined operation.
1051: Let $\widetilde \gamma$ be the density of the resulting
1052: $(2r)$-form on $T^{\mathbb C}$ relative to the Liouville volume
1053: form $\varepsilon_{T}$ on $T^{\mathbb C}\cong \mathrm T^*T$, and
1054: let
1055: \begin{equation*}
1056: \gamma= \frac{\widetilde \gamma}{|W| \mathrm e^{-\kappa/\hbar}}
1057: \end{equation*}
1058: where $|W|$ is the order of the Weyl group.
1059: An explicit calculation of $\gamma$ can be found in Theorem 3 of Section 2 of
1060: \cite{Florentino}, see also Theorem 12 in \cite{HallMitchell}.
1061: The following is the
1062: analogue of Weyl's integration formula, spelled out for $\mathrm
1063: {Ad}(\group)$-invariant holomorphic functions.
1064:
1065: \begin{prop}
1066: Given two holomorphic $\mathrm {Ad}(\group)$-invariant functions
1067: $\psi_1,\psi_2$ on $\group^{\mathbb C}$ that are square-integrable
1068: relative to the measure $ \mathrm e^{-\kappa/\hbar} \eta \varepsilon$,
1069: \begin{equation}
1070: \int_{\group^{\mathbb C}} \overline {\psi_1} \psi_2 \mathrm
1071: e^{-\kappa/\hbar}\eta\varepsilon\,
1072: =
1073: \int_{T^{\mathbb C}} \overline {\psi_1} \psi_2 \mathrm
1074: e^{-\kappa/\hbar}\gamma\varepsilon_{T}\,.
1075: \end{equation}
1076:
1077: \end{prop}
1078:
1079:
1080: \begin{proof}
1081: Since $\psi_1$ and $\psi_2$ are $\Ad(\group)$-invariant and holomorphic, they
1082: are $\Ad(\group^\CC)$-invariant. Hence, their pullbacks under the conjugation
1083: mapping \eqref{conjugation} are constant along the constituent
1084: $\group^\CC/T^\CC$.
1085: Since the conjugation mapping has degree equal to
1086: the order $|W|$ of the Weyl group and since the complement of the image
1087: under the conjugation mapping has measure zero,
1088: \begin{equation*}
1089: \int_{\group^{\mathbb C}} \overline {\psi_1} \psi_2 \mathrm
1090: e^{-\kappa/\hbar} \eta \varepsilon =\frac 1{|W|}\int_{T^{\mathbb
1091: C}} \overline {\psi_1} \psi_2 \widetilde\gamma\varepsilon_{T}=
1092: \int_{T^{\mathbb C}} \overline {\psi_1} \psi_2 \mathrm
1093: e^{-\kappa/\hbar}\gamma\varepsilon_{T}\,.
1094: \end{equation*}
1095: \end{proof}
1096:
1097: \noindent
1098: The proposition implies that the restriction mapping induces an isomorphism
1099: \begin{equation}
1100: \mathcal HL^2(\group^{\mathbb C},\mathrm e^{-\kappa/\hbar}\eta
1101: \varepsilon)^\group
1102: \longrightarrow
1103: \mathcal H L^2(T^{\mathbb
1104: C},\mathrm e^{-\kappa/\hbar} \gamma \varepsilon_{T})^W
1105: \end{equation}
1106: of Hilbert spaces where, according to \eqref{GscaproKC}, the scalar
1107: product in $\mc H L^2(T^{\mathbb C},\mathrm e^{-\kappa/\hbar}
1108: \gamma \varepsilon_{T})^W$ is given by
1109: \beq\label{GscaproTC}
1110: \frac{1}{\vol(\group)}
1111: \int_{T^\CC} \ol{\psi_1}\psi_2 e^{-\kappa/\hbar}\gamma\varepsilon_{T}\,.
1112: \eeq
1113:
1114: A basis of $\mathcal HL^2(\group^{\mathbb C},\mathrm
1115: e^{-\kappa/\hbar}\eta \varepsilon)^\group$ and hence of $\mathcal
1116: H L^2(T^{\mathbb C},\mathrm e^{-\kappa/\hbar} \gamma
1117: \varepsilon_{T})^W$ is obtained as follows. For a highest weight
1118: $\lambda$ relative to the chosen dominant Weyl chamber, we will
1119: denote by $\chi^{\mathbb C}_{\lambda}$ the irreducible character
1120: of $\group^{\mathbb C}$ associated with $\lambda$. The holomorphic
1121: Peter-Weyl theorem established in \cite{holopewe},
1122: see Remark \rref{R-SBtrf} below for historical comments, implies that the
1123: total Hilbert space $\Hi$ contains the complex vector space which
1124: underlies the algebra $\mathbb C[\group^{\mathbb C}]^\group$ of
1125: $\mathrm{Ad}(\group)$-invariant polynomial functions on
1126: $\group^{\mathbb C}$ as a dense subspace. Hence the irreducible
1127: characters $\chi^{\mathbb C}_{\lambda}$ of $\group^{\mathbb C}$
1128: form a basis of $\mathcal HL^2(\group^{\mathbb C},\mathrm
1129: e^{-\kappa/\hbar}\eta \varepsilon)^\group$.
1130:
1131: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1132: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1133:
1134: \subsection{Schr\"odinger quantization}
1135:
1136: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1137: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1138:
1139: Half-form Schr\"odinger quantization on $\ctg\group$ yields the
1140: Hilbert space $L^2(\group,\mr d x)$ of ordinary square-integrable
1141: functions on $\group$ \cite{bhallone} with scalar product
1142: \beq\label{GscaproK}
1143: \scapro{\psi_1}{\psi_2}
1144: =
1145: \frac 1 {\vol(\group)} \int_\group \ol{\psi_1}\psi_2 \, \mr d x\,.
1146: \eeq
1147: We remind the reader that for reasons explained above we have
1148: normalized the Haar
1149: measure on $\group$ so that it coincides with the Riemannian volume measure.
1150: Left and right translation turn the Hilbert space $L^2(\group,\mr d x)$ into a
1151: unitary
1152: $(\group\times \group)$-representation. The Hilbert space
1153: associated with $\pha$ by reduction after quantization is the
1154: subspace $L^2(\group,\mr d x)^\group$ of $\group$-invariants. It
1155: also arises as the physical Hilbert space of the observable algebra
1156: \cite{qcd3} and by quantization via Rieffel induction
1157: \cite{LandsmanWren:Theta,LandsmanWren:Coherent,Wren:Rieffel,Wren:ThetaII},
1158: see also \cite[\S\S IV.3.7,8]{Landsman}.
1159:
1160: Similarly as $\mc H L^2(\group^\CC,\mr
1161: e^{-\kappa/\hbar}\eta\ve)^\group$, the space $L^2(\group,\mr d x)^\group$
1162: can alternatively be viewed as a Hilbert space of $W$-invariant
1163: functions which now live on $T$ rather than on $T^\CC$. Indeed,
1164: let $v\colon T\to\mathbb R$ be the real function given by
1165: $v(t)=\mathrm{vol}(\mathrm{Ad}(\group)t)/|W|$, $t\in T$, that is,
1166: $v(t)$ is the Riemannian volume of the conjugacy class
1167: $\mathrm{Ad}(\group)t$ in $\group$ generated by $t\in T$, divided
1168: by the order $|W|$ of the Weyl group. Restriction of
1169: $\mathrm{Ad}(\group)$-invariant functions from $\group$ to $T$ is
1170: well known to induce an isomorphism
1171: \begin{equation}
1172: L^2(\group,dx)^\group \longrightarrow L^2(T, v\mr d t)^W
1173: \end{equation}
1174: of Hilbert spaces where the scalar product on $L^2(T, vdt)^W$ is
1175: given by
1176: \beq\label{GscaproT}
1177: \frac 1 {\vol(\group)} \int_T \ol{\psi_1}\psi_2 \, v \, \mr d t\,.
1178: \eeq
1179: Given a highest weight $\lambda$, we will denote by
1180: $\chi_{\lambda}$ the corresponding irreducible character of
1181: $\group$, so that $\chi_{\lambda}$ is the restriction of
1182: $\chi^{\mathbb C}_{\lambda}$ to $\group$. The $\chi_\lambda$'s
1183: form an orthonormal basis of $L^2(\group,\mr d x)^\group$.
1184:
1185: Let $\rho=1/2\sum_{\alpha\in R^+} \alpha$ denote the half sum of the
1186: positive roots and let $C_{\lambda}$ be the constant
1187: \beq\label{G-Cl}
1188: C_{\lambda}=(\hbar\pi)^{\dim(\group)/2}\mathrm
1189: e^{\hbar|\lambda+\rho|^2},
1190: \eeq
1191: where $|\lambda+\rho|$ refers to the norm of $\lambda+\rho$
1192: relative to the inner product on $\lieal$.
1193:
1194: \begin{thm} \label{comparison}
1195:
1196: The assignment to $\chi_{\lambda}$ of
1197: $C_{\lambda}^{-1/2}\chi^{\mathbb C}_{\lambda}$, as $\lambda$
1198: ranges over the highest weights, yields a unitary isomorphism
1199: \begin{equation}
1200: L^2(\group,\mr d x)^\group
1201: \longrightarrow
1202: \mc HL^2(\group^\CC,\mr e^{-\kappa/\hbar}\eta\ve)^\group
1203: \label{H}
1204: \end{equation}
1205: of Hilbert spaces.
1206:
1207: \end{thm}
1208:
1209:
1210:
1211: \begin{proof}
1212:
1213: The holomorphic function
1214: $C_{\lambda}^{-1/2}\chi^{\mathbb C}_{\lambda}$ is the image of $\chi_\lambda$
1215: under the Segal-Bargmann transform
1216: \beq\label{Hallg}
1217: L^2(\group,\mr d x) \longrightarrow \mc HL^2(\group^\CC,\mr
1218: e^{-\kappa/\hbar}\eta\ve)
1219: \eeq
1220: which is a unitary isomorphism \cite{Hall94}.
1221: \todo{Gilt das exakt oder nur bis auf einen Faktor?}
1222: The assertion follows because $\chi_\lambda$ and $\chi_\lambda^\CC$ are bases in
1223: $L^2(\group,\mr d x)^\group$ and $\mc HL^2(\group^\CC,\mr
1224: e^{-\kappa/\hbar}\eta\ve)^\group$, respectively. Alternatively, the assertion
1225: is a direct consequence of Theorem 5.3 in \cite{holopewe}.
1226:
1227: \end{proof}
1228:
1229:
1230: \bre\label{R-SBtrf}
1231:
1232: The Segal-Bargmann transform \eqref{Hallg} and, therefore, the isomorphism
1233: \eqref{H}, rely on the description of the Hilbert spaces $L^2(\group,dx)$
1234: and $\mathcal H L^2(\group^{\mathbb C},\mathrm e^{-\kappa/\hbar}
1235: \eta \varepsilon)$ as half-form Hilbert spaces and involve the appropriate
1236: metaplectic correction \cite{woodhous}.
1237:
1238: Originally, in \cite{Hall94}, see also \cite{Hall97}, the Segal-Bargmann
1239: transform was developed via heat kernel analysis on $\group$ and $\group^\CC$.
1240: More recently, an alternative purely geometric description of this transform
1241: in terms of representative functions and independent of heat kernel
1242: analysis has been given in Theorem 5.3 of \cite{holopewe}. This description
1243: relies on the holomorphic Peter-Weyl theorem \cite{holopewe}.
1244: The holomorphic Peter-Weyl theorem yields a proof of Theorem \rref{comparison}
1245: above as well and the geometric methods in \cite{holopewe}
1246: also recover the heat kernel analysis. On the other hand,
1247: the holomorphic Peter-Weyl theorem can likewise be deduced from the
1248: Segal-Bargmann transform developed in \cite{bhallone}, combined with the
1249: ordinary Peter-Weyl theorem.
1250:
1251: Alternatively, we can describe the isomorphism \eqref{Hallg} as
1252: being induced by the corresponding BKS-pairing map
1253: from $L^2(\group,\mr d x)$ to $\mc HL^2(\group^\CC,\mr
1254: e^{-\kappa/\hbar}\eta\ve)$, multiplied by a factor
1255: $(4\pi)^{-\dim(\group)/4}$. For
1256: details, see \cite{bhallone} (description in terms of the
1257: heat kernel on $K$) or Theorem 6.5 in \cite{holopewe} (description in terms of
1258: representative functions).
1259:
1260: \ere
1261:
1262:
1263: Theorem \rref{comparison} entails that the complex
1264: characters $\chi^\CC_\lambda$ satisfy the orthogonality relations
1265: \beq\label{GogonrelchiC}
1266: \scaproC{\chi^\CC_\lambda}{\chi^\CC_{\lambda'}} = C_\lambda
1267: \delta_{\lambda\,\lambda'}\,.
1268: \eeq
1269: Hence, the vectors $C_\lambda^{-1/2} \chi^\CC_\lambda$, where
1270: $\lambda$ ranges over the highest weights, form an orthonormal
1271: basis of $\mc HL^2(\group^\CC,\mr
1272: e^{-\kappa/\hbar}\eta\ve)^\group$.
1273:
1274: From now on, we will take the Hilbert space of our model
1275: to be the Hilbert space $\Hi$ with orthonormal basis $\ket\lambda$ labelled by
1276: the highest weights. In the holomorphic representation, $\Hi$
1277: is then realized as $\mc HL^2(\group^\CC,\mr
1278: e^{-\kappa/\hbar}\eta\ve)^\group$ or, equivalently, as $\mc
1279: HL^2(T^\CC,\mr e^{-\kappa/\hbar}\gamma\ve_T)^W$ whereas, in the
1280: Schr\"odinger representation, $\Hi$ is realized as $L^2(\group,\mr d
1281: x)^\group$ or, equivalently, as $L^2(T,v\,\mr d t)^W$. The passage
1282: to the respective representation is achieved by substitution for
1283: $\ket \lambda$ of the function $C_\lambda^{-1/2} \chi^\CC_\lambda$
1284: or $\chi_\lambda$ as appropriate.
1285:
1286: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1287: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1288:
1289: \subsection{The costratified Hilbert space structure}
1290: \label{SScostrfHispa}
1291:
1292: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1293: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1294:
1295: We will now construct the additional structure of a costratification. To begin
1296: with, we recall from \cite{kaehler} the precise definition
1297: of a costratified Hilbert space. Let $N$ be a
1298: stratified space. Let $\mathcal C_N$ be the category whose objects
1299: are the strata of $N$ and whose morphisms are the inclusions $Y'
1300: \subseteq \overline Y$ where $Y$ and $Y'$ are strata.
1301:
1302: \begin{defi}\label{def2}
1303:
1304: A costratified Hilbert space relative to $N$ is a
1305: contravariant functor from $\mathcal C_N$ to the category of
1306: Hilbert spaces, with bounded linear maps as morphisms.
1307:
1308: \end{defi}
1309:
1310: In more down to earth terms, a costratified Hilbert space relative
1311: to $N$ assigns a Hilbert space $\mathcal C_Y$ to each stratum $Y$,
1312: together with a bounded linear map $\mathcal C_{Y_2} \to \mathcal
1313: C_{Y_1}$ for each inclusion $Y_1 \subseteq \overline {Y_2}$ such
1314: that, whenever $Y_1 \subseteq \overline {Y_2}$ and $Y_2 \subseteq
1315: \overline {Y_3}$, the composite of $\mathcal C_{Y_3} \to \mathcal
1316: C_{Y_2}$ with $\mathcal C_{Y_2} \to \mathcal C_{Y_1}$ coincides
1317: with the bounded linear map $\mathcal C_{Y_3}\to\mathcal C_{Y_1}$
1318: associated with the inclusion $Y_1 \subseteq \overline {Y_3}$.
1319:
1320: To construct a costratified Hilbert space relative to the reduced
1321: phase space $\pha$, we start with the Hilbert space $\mc HL^2(\group^\CC,\mr
1322: e^{-\kappa/\hbar}\eta\ve)^\group$ and single out subspaces
1323: $\Hi_{\tau,i}$ associated with the strata $\pha_{\tau,i}$ as
1324: follows. The elements of $\mc HL^2(\group^\CC,\mr
1325: e^{-\kappa/\hbar}\eta\ve)^\group$ are ordinary functions on
1326: $\group^\CC$, not classes of functions as in the $L^2$-case. Therefore,
1327: being $\group$-invariant, these functions define functions
1328: on $\pha$. Thus, we associate with each stratum $\pha_{\tau,i}$ of $\pha$ the
1329: subspace
1330: \[
1331: \vani_{\tau,i} =\{f\mc \in\mc HL^2(\group^\CC,\mr
1332: e^{-\kappa/\hbar}\eta\ve)^\group;\ f|_{\pha_{\tau,i}}=0 \}
1333: \]
1334: of $\mc HL^2(\group^\CC,\mr
1335: e^{-\kappa/\hbar}\eta\ve)^\group$ which consists of the functions that vanish
1336: on $\pha_{\tau,i}$. We then define the Hilbert space $\Hi_{\tau,i}$ associated
1337: with $\pha_{\tau,i}$ to be the orthogonal complement of $\vani_{\tau,i}$ in
1338: $\mc HL^2(\group^\CC,\mr e^{-\kappa/\hbar}\eta\ve)^\group$, so that
1339: $\mc HL^2(\group^\CC,\mr e^{-\kappa/\hbar}\eta\ve)^\group
1340: =\vani_{\tau,i}\oplus\Hi_{\tau,i}$.
1341:
1342: By construction, if $\pha_{\tau_1,i_1}\subseteq\ol{\pha_{\tau_2,i_2}}$ then
1343: $\vani_{\tau_2,i_2} \subseteq \vani_{\tau_1,i_1}$ and, therefore,
1344: $\Hi_{\tau_1,i_1}\subseteq\Hi_{\tau_2,i_2}$. Let
1345: $\Pi_{\tau_2,i_2;\tau_1,i_1}:\Hi_{\tau_2,i_2}\to\Hi_{\tau_1,i_1}$
1346: denote the orthogonal projection. The resulting system
1347: $\{\Hi_{\tau,i}\}$, together with the orthogonal projections
1348: $\Pi_{\tau_2,i_2;\tau_1,i_1}: \Hi_{\tau_2,i_2}\to
1349: \Hi_{\tau_1,i_1}$ whenever $\pha_{\tau_1,i_1} \subseteq \overline
1350: {\pha_{\tau_2,i_2}}$, is the costratified Hilbert space relative to
1351: $\pha$ we are looking for. When $\tau$ is the principal orbit type, $\Hi_\tau$
1352: plainly coincides with the total Hilbert space $\mc HL^2(\group^\CC,\mr
1353: e^{-\kappa/\hbar}\eta\ve)^\group$.
1354:
1355: While being defined in the holomorphic representation, the
1356: costratified Hilbert space structure may be transferred to the
1357: Schr\"odinger representation. In Section \rref{SSU2} we shall
1358: determine the costratified Hilbert space structure explicitly for
1359: the case $\group=\mathrm{SU}(2)$.
1360:
1361:
1362:
1363: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1364: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1365:
1366: \subsection{Observables}
1367:
1368: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1369: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1370:
1371: The prequantization procedure assigns to a classical observable
1372: $f\in C^\infty(\ctg\group)$ the operator $\hat f$ on the
1373: prequantum Hilbert space $L^2(\ctg\group,\ve)$ given by
1374: \begin{equation}
1375: \hat f = \mr i \hbar X_f + f-\frac 1 {\hbar} \theta(X_f)\,;
1376: \label{prequant} \end{equation} here $\theta$ is the symplectic
1377: potential \eqref{GsplPot}, so that $-d\theta$ coincides with the
1378: cotangent bundle symplectic structure $\omega$ on $\ctg\group$,
1379: and $X_f$ denotes the Hamiltonian vector field associated with
1380: $f$, determined by the identity
1381: $$
1382: \omega(X_f,\,\cdot\,) = df,
1383: $$
1384: in accordance with Hamilton's equations. The formula
1385: \eqref{prequant} is essentially the same as that given as (8.2.2)
1386: in \cite{woodhous}, save that the Hamiltonian vector field $X_f$
1387: and the symplectic potential $\theta$ are the negatives of the
1388: corresponding objects in \cite{woodhous}. Let $\{\,\cdot \, ,
1389: \,\cdot \,\}$ be the Poisson structure on $C^\infty(\ctg\group)$
1390: associated with the cotangent bundle symplectic structure
1391: $\omega$; this Poisson structure is given by \eqref{GPoibra}. Then $\hbar \{\,\cdot \,
1392: , \,\cdot \,\}$ is the Poisson structure on $C^\infty(\ctg\group)$
1393: associated with the symplectic structure $\frac{\omega}{\hbar}$.
1394: The formula \eqref{prequant} yields a representation of the Lie
1395: algebra underlying the Poisson algebra
1396: $\left(C^\infty(\ctg\group),\hbar \{\,\cdot \, , \,\cdot
1397: \,\}\right)$ which satisfies the Dirac conditions. This
1398: representation is not irreducible and, to arrive at an irreducible
1399: representation of at least a certain subalgebra, the standard
1400: procedure is to introduce a polarization. Observables in this
1401: subalgebra are then referred to as being quantizable in the
1402: polarization under discussion.
1403:
1404: In our situation, in the K\"ahler polarization, only the constants
1405: are quantizable. In the Schr\"odinger polarization, the
1406: topological obstruction to the existence of a half-form bundle
1407: vanishes for trivial reasons and, with the half-form correction
1408: incorporated, the relevant subalgebra of $C^\infty(\ctg\group)$
1409: contains the functions which restrict to polynomials of at most
1410: second order on the fibres of $\ctg\group$, i.~e., which are at
1411: most quadratic in the generalized momenta. Thus, it contains the
1412: (classical) Hamiltonian \eqref{GHaFn} of our model. The associated
1413: quantum observable, i.~e., the (quantum) Hamiltonian, is given by
1414: \beq\label{GHaOp}
1415: H = -\frac{\hbar^2}{2}\Delta_\group + \frac{\inco}{2}
1416: (3-\Re\chi_{\lambda_1})\,,
1417: \eeq
1418: where $\lambda_1$ denotes the highest weight of the defining
1419: representation of $\group$. The operator $\Delta_\group$ arises
1420: from the non-positive Laplace-Beltrami operator
1421: $\tilde\Delta_\group$ associated with the bi-invariant Riemannian
1422: metric on $\group$ as follows: The operator $\tilde\Delta_\group$
1423: is essentially self-adjoint on $C^\infty(\group)$ and has a unique
1424: extension $\Delta_\group$ to an (unbounded) self-adjoint operator
1425: on $L^2(\group,dx)$. The spectrum being discrete, the domain of
1426: this extensions is the space of functions of the form $f=\sum_n
1427: \alpha_n \vp_n$ such that $\sum_n |\alpha_n|^2 \lambda_n^2 <
1428: \infty$ where the $\vp_n$'s range over the eigenfunctions and the
1429: $\lambda_n$'s over the eigenvalues of $\tilde\Delta_\group$.
1430:
1431: Since the metric is bi-invariant, so is the operator
1432: $\Delta_\group$, whence this operator restricts to a self-adjoint
1433: operator on the subspace $L^2(\group,\mr d x)^\group$ which we
1434: continue to denote by $\Delta_\group$. A core for this operator,
1435: and hence for the Hamiltonian $H$, is given by
1436: $C^\infty(\group)^\group$.
1437:
1438: By means of the unitary transform \eqref{H} we now transfer the
1439: Hamiltonian and, in particular, the operator $\Delta_\group$ to
1440: the holomorphic representation, i.~e., to self-adjoint operators
1441: on $\mc HL^2(\group^\CC,\mr e^{-\kappa/\hbar}\eta\ve)^\group$.
1442: Concerning $\Delta_\group$, we may alternatively view
1443: $\tilde\Delta_\group$ as a differential operator on
1444: $\group^{\mathbb C}$ via the embedding of $\lieal$ into
1445: $\lieal^{\mathbb C}$, extend it to a self-adjoint operator on
1446: $\mathcal H L^2(\group^{\mathbb C},\mathrm e^{-\kappa/\hbar} \eta
1447: \varepsilon)$, and take the restriction to the subspace $\mathcal
1448: H L^2(\group^{\mathbb C},\mathrm e^{-\kappa/\hbar} \eta
1449: \varepsilon)^\group$.
1450:
1451: Next, we determine the eigenvalues and the eigenfunctions of
1452: $\Delta_\group$. The operator $\tilde\Delta_\group$ is known to
1453: coincide with the Casimir operator on $\group$ associated
1454: with the bi-invariant Riemannian metric, see \cite{taylothr} (A
1455: 1.2). That is to say, after a choice $X_1,\dots,X_m$ of
1456: orthonormal basis of $\lieal$ has been made,
1457: $$
1458: \tilde\Delta_\group = X^2_1 +\dots + X^2_m
1459: $$
1460: in the universal enveloping algebra $\mathrm U(\lieal)$ of $\lieal$, cf.
1461: e.~g. \cite{enelson} (p.~591). Since $\tilde\Delta_\group$ is
1462: bi-invariant, by Schur's lemma, each isotypical $(\group\times
1463: \group)$-summand $L^2(\group,dx)_\lambda$ of $L^2(\group,dx)$ in
1464: the Peter-Weyl decomposition is an eigenspace, and the
1465: representative functions are eigenfunctions for
1466: $\tilde\Delta_\group$. The eigenvalue of $\tilde\Delta_\group$
1467: corresponding to the highest weight $\lambda$ is known to be given
1468: explicitly by $-\varepsilon_{\lambda}$ where
1469: \begin{equation}
1470: \varepsilon_{\lambda}=(|\lambda+\rho|^2-|\rho|^2),
1471: \label{EWL}
1472: \end{equation}
1473: cf. e.~g.\ \cite{helgaboo} (Chap. V.1 (16)). The sign is chosen in such a way
1474: that the $\ve_\lambda$ can be interpreted as energy values. Hence,
1475: in particular, each character $\chi_\lambda$ is an eigenfunction
1476: of $\Delta_\group$ associated with the eigenvalue $-\ve_\lambda$.
1477: Consequently, $\Delta_\group$ being viewed as an operator on the
1478: abstract Hilbert space $\Hi$, the vectors $\ket\lambda \in \Hi$
1479: form an orthonormal eigenbasis of $\Hi$. In view of an
1480: observation spelled out above, the domain of $\Delta_\group$ is
1481: explicitly given by
1482: \beq\label{Gdomain}
1483: \left\{
1484: \sum\nolimits_\lambda \alpha_\lambda \ket \lambda \in \Hi
1485: :
1486: \sum\nolimits_\lambda |\alpha_\lambda|^2 \ve_\lambda^2 < \infty
1487: \right\}\,.
1488: \eeq
1489:
1490:
1491:
1492: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1493: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1494:
1495: \section{The costratified Hilbert space for $\SU(2)$}
1496: \label{SSU2}
1497:
1498: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1499: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1500:
1501:
1502:
1503: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1504: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1505:
1506: \subsection{Group theoretical data}
1507:
1508: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1509: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1510:
1511: The (real) root system of $\lieal = \su(2)$ consists of the two
1512: roots $\alpha$ and $-\alpha$, given by
1513: $$
1514: \alpha\big(Y) = 2 y,
1515: \,~~~~~~
1516: Y\in\mf t\,,
1517: $$
1518: where $Y=\diag(\mr i y ,- \mr i y)$, $y\in\RR$. Then $\halfsum =
1519: \frac12\alpha$. We label the irreducible representations
1520: by non-negative integers $n$ (twice the spin). The corresponding
1521: highest weights $\lambda_n$ are given by $\lambda_n = \frac n 2
1522: \alpha$. On $T\times\mf t \cong T^\CC$, the corresponding complex characters
1523: $\chi^\CC_n$ of $\group^\CC = \SL(2,\CC)$ are given by
1524: \beq\label{GchiC}
1525: \chi^\CC_n(t) = z^n + z^{n-2} + \cdots + z^{-n}
1526: \,,~~~~~~
1527: t\in T^\CC\,,
1528: \eeq
1529: where $t = \diag(z,z^{-1})$, $z\in \CC^\ast$. Restriction to
1530: $\group$ yields the real characters which, on $T$, can be written
1531: as
1532: \beq\label{Gchi}
1533: \chi_n(t) = \frac{\sin\big((n+1)x\big)}{\sin(x)}
1534: \,,~~~~~~
1535: t\in T\,,
1536: \eeq
1537: where $t=\diag(\mr e^{\mr i x}, \mr e^{- \mr i x})$, $x\in\RR$.
1538: Any invariant inner product $\langle \,\cdot \, , \,\cdot \,
1539: \rangle$ on $\lieal = \su(2)$ is proportional to the (negative
1540: definite) trace form. Hence, given $\langle \,\cdot \, ,\,\cdot
1541: \, \rangle$, we can define a positive number $\scale$ by
1542: $$
1543: \langle Y_1,Y_2 \rangle = -\frac{1}{2\scale^2} \tr(Y_1Y_2),\
1544: Y_1,Y_2 \in \lieal\,.
1545: $$
1546: For the Killing form, $\scale = \frac 1 {\sqrt 8}$. Relative to the given
1547: invariant inner product on $\lieal$, the two roots
1548: $\alpha$ and $-\alpha$ have norm $|\alpha|^2 = 4 \scale^2$. Hence
1549: $|\halfsum|^2 = \scale^2$ and $|\lambda_n+\halfsum|^2 =
1550: \scale^2(n+1)^2$ whence according to \eqref{G-Cl} and \eqref{EWL}
1551: \beq\label{GEWCn}
1552: \ve_n = \scale^2 n (n+2)
1553: \,,~~~~~~
1554: C_n = (\hbar\pi)^{3/2} \mr e^{\hbar\scale^2(n+1)^2}\,.
1555: \eeq
1556:
1557: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1558: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1559:
1560: \subsection{The costratified Hilbert space structure}
1561: \label{costratified}
1562: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1563: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1564:
1565: According to Section \rref{Squantum}, the appropriate Hilbert
1566: space for the holomorphic representation is the Hilbert space $\mc H
1567: L^2(\group^\CC,\mr e^{-\kappa/\hbar} \eta\ve)^\group$ or,
1568: equivalently, $\mc H L^2(T^\CC,\mr e^{-\kappa/\hbar} \gamma
1569: \ve_T)^W$. The Hilbert space for the Schr\"odinger representation is
1570: the space $L^2(\group,\mr d x)^\group$ or, equivalently,
1571: $L^2(T,v\,\mr d t)^W$. There is no need to spell out the functions
1572: $\kappa$, $\eta$, $\gamma$ or $v$ here, because we can work
1573: entirely in the basis given by the characters. For $n \geq 0$, let
1574: $|n\rangle := \ket{\lambda_n}$; then $\{\ket n : n=0,1,2,\dots\}$
1575: is an orthonormal basis of $\Hi$, and we can pass to the
1576: holomorphic and to the Schr\"odinger representation by replacing
1577: each $\ket n$ with the corresponding (normalized) character.
1578:
1579: We now determine the costratified Hilbert space structure
1580: constituents $\Hi_\pm$ and $\Hi_\prin$ associated with the strata
1581: $\pha_\pm$ and $\pha_1$ of $\pha$ and the subspace $\Hi_\sing$
1582: associated with the orbit type subset $\pha_0$. Recall the
1583: notation and the description of these strata from Subsection
1584: \rref{SSsymred}. As $\pha_\prin$ is the top stratum, $\Hi_\prin =
1585: \Hi$. To describe the subspaces $\Hi_\pm$ and $\Hi_\sing$, we pass
1586: to the holomorphic representation.
1587:
1588: \ble\label{Lvani}
1589:
1590: The systems \eqref{Gvanipbs1}, \eqref{Gvanimbs1}, and \eqref{Gvanisingbs1} below
1591: constitute bases of, respectively, the subspaces $\vani_+, \vani_-, \vani_\sing$ of $\Hi$
1592: corresponding to, respectively, the strata $\pha_+$, $\pha_-$ and $\pha_\sing:$
1593: \beqa\label{Gvanipbs1}
1594: \chi^\CC_n - (n+1) \chi^\CC_0
1595: \,,&~~~~~~&
1596: n = 1,2,3,\dots\,,
1597: \\ \label{Gvanimbs1}
1598: \chi^\CC_n + (-1)^n \frac{n+1}{2} \chi^\CC_1
1599: \,,&~~~~~~&
1600: n = 0,2,3,\dots\,,
1601: \\ \label{Gvanisingbs1}
1602: \chi^\CC_{2k} - (2k+1) \chi^\CC_0
1603: \,,~~~~~~
1604: \chi^\CC_{2k+1} - (k+1) \chi^\CC_1
1605: \,,&~~~~~~&
1606: k = 1,2,3,\dots\,.
1607: \eeqa
1608:
1609: \ele
1610:
1611: \begin{proof}
1612:
1613: We view the elements of $\Hi$ as functions on $T^\CC$ rather than
1614: on $\group^\CC$. Via the polar decomposition map $T\times\mf t \to
1615: T^\CC$, the points $(\pm\II,0)$ are mapped to $\{\pm\II\}$. Hence,
1616: $\vani_+$, $\vani_-$ and $\vani_\sing$ consist of the functions
1617: $\psi\in\Hi$ that satisfy, respectively,
1618: \beq\label{Gvanicond}
1619: \psi(\II) = 0
1620: \,,~~~~~~
1621: \psi(-\II) = 0
1622: \,,~~~~~~
1623: \psi(\pm\II) = 0
1624: \,.
1625: \eeq
1626: Due to $\chi^\CC_n(\pm\II) = (\pm 1)^n(n+1)$, we
1627: have
1628: $$
1629: \chi^\CC_{2k}(\pm\II) = 2k+1 = (2k+1)\chi^\CC_0(\pm\II)
1630: $$
1631: and
1632: $$
1633: \chi^\CC_{2k+1}(\pm\II) = \pm (2k+2) = (k+1)\chi^\CC_1(\pm\II)\,.
1634: $$
1635: Hence, all the functions given in
1636: \eqref{Gvanipbs1}--\eqref{Gvanisingbs1} satisfy the corresponding
1637: condition in \eqref{Gvanicond}. Conversely, given $\psi\in
1638: \vani_+$, expanding it in the basis of $\Hi$ given by the elements
1639: in \eqref{Gvanipbs1} together with $\chi^\CC_0$ we see that the vanishing of
1640: $\psi(\II)$
1641: implies that the coefficient of $\chi^\CC_0$ is zero. The
1642: reasoning for $\vani_-$ and $\vani_\sing$ is analogous. Finally,
1643: linear independence of the systems
1644: \eqref{Gvanipbs1}--\eqref{Gvanisingbs1} is obvious.
1645: \end{proof}
1646:
1647:
1648: \noindent
1649: We express the bases \eqref{Gvanipbs1}--\eqref{Gvanisingbs1}, up
1650: to a common factor $(\hbar\pi)^{3/4}$, in terms of $\ket n$:
1651: \beqa\label{Gvanipbs}
1652: \mr e^{\hbar\scale^2(n+1)^2/2} \ket n
1653: -
1654: (n+1) \mr e^{\hbar\scale^2/2} \ket 0
1655: \,,&&
1656: n = 1,2,3,\dots\,,
1657: \\ \label{Gvanimbs}
1658: \mr e^{\hbar\scale^2(n+1)^2/2} \ket n
1659: -
1660: \frac{n+1}{2} \mr e^{2 \hbar\scale^2} \ket 1
1661: \,,&&
1662: n = 0,2,3,\dots\,,
1663: \\ \nonumber
1664: \mr e^{\hbar\scale^2(2k+1)^2/2} \ket {2k}
1665: -
1666: (2k+1) \mr e^{\hbar\scale^2/2} \ket{0}
1667: \,, && k = 1,2,3,\dots\,,
1668: \\ \label{Gvanisingbs}
1669: \mr e^{2\hbar\scale^2(k+1)^2} \ket {2k+1}
1670: -
1671: (k+1) \mr e^{2\hbar\scale^2} \ket 1
1672: \,,&&
1673: k = 1,2,3,\dots\,.
1674: \eeqa
1675:
1676: \bpp\label{PHsing}
1677:
1678: The subspaces $\Hi_+$ and $\Hi_-$ have dimension $1$. They are
1679: spanned by the normalized vectors
1680: \beqa \label{Gpsip}
1681: \psi_+ & := & \frac{1}{N} \sum\nolimits_{n=0}^\infty (n+1)\, \mr
1682: e^{-\hbar\scale^2\,(n+1)^2/2}\, \ket n
1683: \,,
1684: \\ \label{Gpsim}
1685: \psi_- & := & \frac{1}{N} \sum\nolimits_{n=0}^\infty (-1)^n\, (n+1) \, \mr
1686: e^{-\hbar\scale^2 \, (n+1)^2/2} \, \ket n\,,
1687: \eeqa
1688: respectively. The subspace $\Hi_\sing$ has dimension $2$. It is
1689: spanned by the orthonormal basis
1690: \beq\label{Gpsising}
1691: \psi_\even
1692: :=
1693: \frac{1}{N_\even} \sum_{\text{\rm $n$ even}}
1694: (n+1) \, \mr e^{-\hbar\scale^2\,(n+1)^2/2} \ket{n}
1695: \,,~~~~
1696: \psi_\odd
1697: :=
1698: \frac{1}{N_\odd} \sum_{\text{\rm $n$ odd}}
1699: (n+1) \, \mr e^{-\hbar\scale^2\,(n+1)^2/2} \ket{n}\,,
1700: \eeq
1701: where the sum over the even $n$ includes $n=0$. The normalization factors are
1702: $$
1703: N^2 = \sum_{n=1}^\infty n^2 \, \mr e^{-\hbar\scale^2 \, n^2}
1704: \,,~~~~~~
1705: N_\even^2 = \sum_{\text{\rm $n$ odd}} n^2 \mr e^{-\hbar\scale^2 \,
1706: n^2}
1707: \,,~~~~~~
1708: N_\odd^2 = \sum_{\text{\rm $n$ even}} n^2 \mr e^{-\hbar\scale^2 \,
1709: n^2}\,.
1710: $$
1711:
1712: \epp
1713:
1714: \bpf
1715:
1716: The sums in \eqref{Gpsip}, \eqref{Gpsim} and \eqref{Gpsising}
1717: converge, their limits are normalized, and $\psi_\even$ and
1718: $\psi_\odd$ are mutually orthogonal. The vector $\psi_+$ together
1719: with the system \eqref{Gvanipbs}, $\psi_-$ together with the
1720: system \eqref{Gvanimbs}, and $\psi_\even$, $\psi_\odd$ together
1721: with the system \eqref{Gvanisingbs} provide bases of $\Hi$.
1722: Finally, it is straightforward to check that $\psi_+$, $\psi_-$
1723: and $\psi_\even$, $\psi_\odd$ are orthogonal to the corresponding
1724: system in \eqref{Gvanipbs}--\eqref{Gvanisingbs}. \epf
1725:
1726: \noindent
1727: Proposition \rref{PHsing} implies that, in Dirac notation, for $i=\sing,\pm$,
1728: the orthogonal projections $\Pi_i\equiv\Pi_{\prin i}\colon\Hi_\prin\to\Hi_i$
1729: are given by
1730: \beq\label{Goproj}
1731: \Pi_\pm
1732: =
1733: |\psi_\pm\rangle\,\langle\psi_\pm|
1734: \,,~~~~~~
1735: \Pi_\sing
1736: =
1737: |\psi_\even\rangle\,\langle\psi_\even|
1738: +
1739: |\psi_\odd\rangle\,\langle\psi_\odd|
1740: \,.
1741: \eeq
1742: The normalization factors $N$, $N_\even$ and $N_\odd$ can be
1743: expressed in terms of the $\theta$-constant $\theta_3$ with \lq nome\rq\ $Q$ as
1744: \beq\label{Gdeftheta}
1745: \theta_3(Q) = \sum_{k=-\infty}^\infty Q^{k^2}\,.
1746: \eeq
1747: For example,
1748: \beq\label{Gnormtheta}
1749: N^2
1750: =
1751: - \frac{\partial}{\partial (\hbar\scale^2)}
1752: \sum_{n=1}^\infty \mr e^{-\hbar\scale^2 \, n^2}
1753: =
1754: - \frac{1}{2} \frac{\partial}{\partial (\hbar\scale^2)}
1755: \theta_3(\mr e^{-\hbar\scale^2})
1756: =
1757: \frac 1 2 \mr e^{-\hbar\scale^2} \theta_3'(\mr e^{-\hbar\scale^2})\,.
1758: \eeq
1759: Then
1760: \beq\label{Gnorm01theta}
1761: N_\odd^2
1762: =
1763: 4 \mr e^{-4 \hbar\scale^2}\theta_3'(\mr e^{-4
1764: \hbar\scale^2})
1765: \,,~~~~~~
1766: N_\even^2
1767: =
1768: N^2 - N_\odd^2\,.
1769: \eeq
1770:
1771:
1772:
1773: \bre \label{R-cs}
1774:
1775: Let $\rho_\hbar(a) := \mr e^{\hbar\beta^2} N \psi_+$. Using \eqref{Gpsip} and
1776: plugging in for $\ket n$ the real characters $\chi_n$ we see that $\rho_\hbar$
1777: satisfies the heat equation $\frac{\mr d}{\mr
1778: d\hbar} \rho = \frac12\Delta_\group\rho_\hbar$ subject to the initial condition
1779: $\rho_0 = \delta_\II$, i.e., $\rho_t$ is the heat kernel of $\group$. The
1780: expansion of $\rho_\hbar$ obtained from \eqref{Gpsip} is the standard expansion
1781: of the heat kernel of a compact Lie group in terms of its characters \cite[p.\
1782: 38]{Stein}. According to \cite[\S 4, Prop.\ 1]{Hall94}, the function
1783: $\rho_\hbar$ has an analytic continuation to $K^\CC$. This analytic continuation
1784: does not consist in substitution of the character $\chi_n^\CC$ for the character
1785: $\chi_n$ in the standard expansion; in particular, the resulting formal series
1786: does not converge in $\mc H L^2(\group^\CC,\mr e^{-\kappa/\hbar}
1787: \eta\ve)^\group$. Thus $\rho_\hbar$ defines the complex-valued functions
1788: $$
1789: \psi^{(\hbar)}_g(a) = \ol{\rho_\hbar(ga^{-1})}
1790: \,,~~~~~~
1791: a\in \group\,,
1792: $$
1793: on $\group$, parametrized by the members $g\in\group^\CC$.
1794: According to \cite{Hall94}, these functions admit an interpretation as coherent
1795: states on $\group$. Indeed, the functions $\psi_\pm $ and
1796: $\psi^{(\hbar)}_{\pm\II}$ are related by the identity
1797: $$
1798: \psi_\pm = \frac{\mr e^{-\hbar\beta^2}}{N} \psi^{(\hbar)}_{\pm\II}\,,
1799: $$
1800: i.e., up to a normalization factor, the states spanning the subspaces
1801: $\Hi_\pm$
1802: are the coherent states labelled by the points of the corresponding
1803: strata. This observation is certainly not a coincidence; in fact, for physical
1804: reasons, the states which the functions $\psi_\pm $ represent
1805: should come down to coherent states because the (phase space) wave function
1806: orthogonal to all wave functions vanishing at a given point represents a state
1807: of optimal localization in phase space (i.e., minimal position-momentum
1808: uncertainty). This is exactly what is generally understood to be a coherent
1809: state.
1810:
1811: \ere
1812:
1813:
1814:
1815:
1816:
1817:
1818: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1819: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1820:
1821: \subsection{Tunneling between strata}
1822:
1823: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1824: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1825:
1826: Consider the constituents $\Hi_+$ and $\Hi_-$ of the
1827: costratified Hilbert space $\Hi$ relative to the orbit type stratification
1828: of $\pha$. A straightforward calculation yields
1829: $$
1830: \langle \psi_+,\psi_-\rangle
1831: =
1832: \frac{1}{N^2} \sum_{n=1}^\infty (-1)^{n+1} \,
1833: n^2 \, \mr e^{-\hbar\scale^2\,n^2}
1834: =
1835: \frac{N_\even^2 - N_\odd^2}{N^2}\,.
1836: $$
1837: As in subsection \ref{costratified} above, the scalar product can
1838: be expressed in terms of $\theta$-functions. Likewise, as in
1839: \eqref{Gnormtheta} for $N^2$, the alternating sum in the
1840: denominator can be rewritten as
1841: $
1842: - \mr e^{-\hbar\scale^2}\theta_3'(-\mr e^{-\hbar\scale^2})\,.
1843: $
1844: Together with \eqref{Gnormtheta} this yields
1845: \beq\label{G-scapro-theta}
1846: \langle \psi_+,\psi_-\rangle
1847: =
1848: -\, \frac{
1849: \theta_3'\big(-\mr e^{-\hbar\scale^2}\big)
1850: }{
1851: \theta_3'\big(\mr e^{-\hbar\scale^2}\big)
1852: }\,.
1853: \eeq
1854: The absolute square $|\langle \psi_+,\psi_-\rangle|^2$ is the
1855: tunneling probability between the strata $\pha_+$ and $\pha_-$,
1856: i.~e., the probability for a state prepared at $\pha_+$ to be
1857: measured at $\pha_-$ and vice versa.
1858:
1859: The numerical value of this quantity strongly depends on the
1860: combined constant $\hbar\scale^2$, see Figure \rref{AbbTunnelW}.
1861: For large values of $\hbar\scale^2$, $|\langle \psi_+,\psi_-\rangle|^2$
1862: is almost equal to $1$. This can
1863: also be read off from the expansions \eqref{Gpsip} and
1864: \eqref{Gpsim}: the first coefficient that distinguishes between
1865: $\psi_+$ and $\psi_-$ is $\mr 2e^{-4\hbar\scale^2}$; for
1866: large $\hbar\scale^2$, this coefficient
1867: is much smaller than the leading coefficient
1868: $\mr e^{-\hbar\scale^2}$, so that $\psi_+$ and $\psi_-$ have a
1869: large overlap. In fact, in the limit $\hbar\scale^2\to\infty$ they become both
1870: equal to $\ket 0$.
1871:
1872: On the other hand, for $\hbar\scale^2\to 0$ we have
1873: $|\langle \psi_+,\psi_-\rangle|^2 \to 0$. Thus, in the semiclassical limit,
1874: the tunneling probability
1875: vanishes.
1876:
1877: \begin{figure}
1878:
1879: \begin{center}
1880:
1881: \hspace*{-2.5cm}\epsfig{file=tunnelw.eps,width=10cm}
1882:
1883: \caption{\label{AbbTunnelW} Tunneling probability $|\langle
1884: \psi_+,\psi_-\rangle|^2$ as a function of $\hbar\scale^2$}
1885:
1886: \end{center}
1887:
1888: \end{figure}
1889:
1890: \bre
1891:
1892: Since the strata $\pha_+$ and $\pha_-$ together constitute
1893: the orbit type subset
1894: $\pha_\sing$, a tunneling between them should not be
1895: visible in the costratification given by $\Hi_\sing$, that is, in
1896: the costratification
1897: relative to the coarser decomposition $\pha=\pha_\sing\cup\pha_\prin$ by mere
1898: orbit types (and not by the connected components thereof). Indeed, we have
1899: \beq\label{GHsingdeco}
1900: \Hi_\sing = \Hi_+ \oplus \Hi_-\,,
1901: \eeq
1902: where the sum is direct but not orthogonal, and $\psi_\pm$ can be written as
1903: $$
1904: \psi_\pm = \frac{N_\even}{N} \psi_\even \pm \frac{N_\odd}{N} \psi_\odd\,.
1905: $$
1906: In other words, the subspaces
1907: $\Hi_+$ and $\Hi_-$ are swallowed by $\Hi_\sing$
1908: and there is no way to reconstruct them from $\Hi_\sing$
1909: alone.
1910:
1911: \ere
1912:
1913:
1914: \bre
1915:
1916: Expressing the scalar product in terms of the coherent states
1917: $\psi^{(\hbar)}_\II$ and $\psi^{(\hbar)}_{-\II}$, see Remark \rref{R-cs},
1918: we obtain the identity
1919: $$
1920: |\langle \psi_+,\psi_- \rangle|^2
1921: =
1922: \frac{
1923: \big|\big\langle \psi^{(\hbar)}_{\II},\psi^{(\hbar)}_{-\II} \big\rangle\big|^2
1924: }{
1925: \|\psi^{(\hbar)}_{\II}\|^2\,\|\psi^{(\hbar)}_{-\II}\|^2
1926: }\,.
1927: $$
1928: The quantity
1929: $
1930: \frac{
1931: \big|\big\langle \psi^{(\hbar)}_g,\psi^{(\hbar)}_h \big\rangle\big|^2
1932: }{
1933: \|\psi^{(\hbar)}_g\|^2\,\|\psi^{(\hbar)}_h\|^2
1934: }
1935: $
1936: is known as the overlap of the coherent states $\psi^{(\hbar)}_g$ and
1937: $\psi^{(\hbar)}_h$; it was studied in more general situations in great detail in
1938: a series of papers by Thiemann and collaborators
1939: \cite{Thiemann,ThiemannWinkler}. Among other
1940: things, they have shown that for $K=\SU(2)$ the overlap is related with the
1941: geodesic distance on $\group^\CC = \SL(2,\CC)$ and that, in general,
1942: for $g\neq h$ and $\hbar\to 0$, the overlap vanishes faster than any power of
1943: $\hbar$.
1944:
1945: The scalar product $\langle \psi^{(\hbar)}_g,\psi^{(\hbar)}_h \rangle$,
1946: viewed as a function of $g$ and $h$, is known as the reproducing kernel
1947: associated with the familiy of coherent states $\psi^{(\hbar)}_g$,
1948: $g\in\group^\CC$. It can be expressed in terms of the heat
1949: kernel $\rho_\hbar$. For $\SU(2)$, this leads to Formula
1950: \eqref{G-scapro-theta}.
1951:
1952: \ere
1953:
1954:
1955:
1956:
1957: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1958: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1959:
1960: \subsection{Adapted orthonormal bases}
1961:
1962: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1963: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1964:
1965: For $i=\pm,\sing$, we will now construct orthonormal bases of the subspaces
1966: $\vani_i$ of $\Hi$. To this end, let
1967: $$
1968: \hat\psi_\pm
1969: :=
1970: \frac{(1-\Pi_\mp)\psi_\pm}{\|(1-\Pi_\mp)\psi_\pm\|} .
1971: $$
1972: Then $\vani_\pm = \vani_\sing \oplus \CC\hat\psi_\mp$,
1973: the sum being orthogonal since $\hat\psi_\pm\in\Hi_\sing$.
1974: Hence, it suffices to construct an orthonormal basis of $\vani_\sing$.
1975: For that purpose, we orthonormalize the family \eqref{Gvanisingbs}. This can
1976: of course be done for the even and odd degree families separately.
1977:
1978:
1979: \begin{Lemma}\label{Logon}
1980:
1981: Let $\vp_n$, $n=0,1,2,\dots$ be an orthonormal basis of the Hilbert
1982: space $\mc E$ and let $\vanifac_n$, $n=0,1,2,\dots$, be real numbers with
1983: $\vanifac_0=1$. Then orthonormalization of the system $\vp_n -
1984: \vanifac_n \vp_0$, $n=1,2,3,\dots$, yields the system
1985: \beq\label{GvaniONB}
1986: \tilde\vp_n
1987: ~=~
1988: \frac{\vanisum_{n-1}}{\vanisum_n} \vp_n ~-~
1989: \frac{\vanifac_n}{\vanisum_n \vanisum_{n-1}} \sum_{k=0}^{n-1}
1990: \vanifac_k\,\vp_k
1991: \,,~~~~~~
1992: n=1,2,3,\dots\,,
1993: \eeq
1994: where $\vanisum_n^2 = \sum_{k=0}^n \vanifac_k^2$.
1995:
1996: \end{Lemma}
1997:
1998: \begin{proof}
1999:
2000: Straightforward calculation.
2001: \end{proof}
2002:
2003: Let $\psi_{2n}$, $n=1,2,3,\dots$, denote the basis elements
2004: obtained by application of Lemma \rref{Logon} to the even degree family of
2005: \eqref{Gvanisingbs}. Thus substituting
2006: $\ket{2k}$ for $\vp_k$ and $(2k+1) \mr e^{-\hbar\ve_{2k}/2}$
2007: for $\vanifac_k$ in \eqref{GvaniONB} yields $\psi_{2n}$.
2008: Likewise, let $\psi_{2n+1}$, $n=1,2,3,\dots$, denote the
2009: basis elements obtained by applying the lemma to the odd degree
2010: family of \eqref{Gvanisingbs}, so that substituting
2011: $\ket{2k+1}$ for $\vp_k$ and
2012: $(k+1) \mr e^{-2\hbar\ve_k}$ for $\vanifac_k$ in
2013: \eqref{GvaniONB} yields
2014: $\psi_{2n+1}$.
2015:
2016: The resulting vectors $\psi_n$, $n=2,3,4,\dots$ form an orthonormal basis of
2017: $\vani_\sing$. Adding $\hat\psi_-$, we obtain an orthonormal basis of
2018: $\vani_+$. Adding $\hat\psi_+$, we obtain an orthonormal basis of
2019: $\vani_-$.
2020:
2021: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2022: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2023:
2024: \subsection{Representation in terms of $L^2[0,\pi]$}
2025:
2026: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2027: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2028:
2029: From now on we will work in the Schr\"odinger representation,
2030: i.e., we realize $\Hi$ as
2031: $$
2032: L^2(\group)^\group \cong L^2(T,v\mr d
2033: t)^W.
2034: $$
2035: In order to produce plots of wave functions $\psi\in\Hi$ we choose
2036: a suitable parameterization of $\cfg$ and represent the elements
2037: of $\Hi$ by ordinary $L^2$-integrable functions on the parameter
2038: space. This representation will also be used in the discussion of
2039: the stationary Schr\"odinger equation of our model in Section
2040: \rref{Seigen}. A suitable parameterization of $\cfg$ can be
2041: obtained as follows. We parameterize $T$ by
2042: \beq\label{GparamT}
2043: \diag(\mr e^{\mr i x},\mr e^{-\mr i x})
2044: \,,~~~~~~
2045: x\in[-\pi,\pi]\,.
2046: \eeq
2047: Since the nontrivial element of $W$ acts by reflection $x\mapsto
2048: -x$, restriction of the parameter $x$ to the interval $[0,\pi]$
2049: yields a (bijective) parameterization of $\cfg$, where $\cfg_+$
2050: corresponds to $x=0$ and $\cfg_-$ to $x=\pi$.
2051:
2052: In the parameterization \eqref{GparamT}, the measure $v\,\mr d t$
2053: on $T$ is given by
2054: $$
2055: v\,\mr d t = \frac{\vol(\group)}{\pi} \sin^2(x) \,\mr d x \,.
2056: $$
2057: Hence, the assignment to $\psi\in C^\infty(T)^W$ of the function
2058: $x\mapsto \psi(\diag(\mr e^{\mr i x},\mr e^{-\mr i x}))$,
2059: $x\in[0,\pi]$ defines a Hilbert space isomorphism
2060: \beq\label{GGamma1}
2061: \Gamma_1\colon \Hi \to L^2([0,\pi],\sin^2(x) \mr d x)\,.
2062: \eeq
2063: Furthermore, multiplication by $\sqrt 2 \sin x$ defines a Hilbert
2064: space isomorphism
2065: \beq\label{GGamma2}
2066: \Gamma_2\colon L^2([0,\pi],\sin^2(x) \mr d x) \to L^2[0,\pi]\,.
2067: \eeq
2068: Here the scalar products in $L^2([0,\pi],\sin^2(x) \mr d x)$ and
2069: $L^2[0,\pi]$ are normalized so that the constant function with
2070: value $1$ has norm $1$.
2071:
2072: The composite isomorphism $\Gamma = \Gamma_2\circ\Gamma_1$
2073: identifies $\Hi$ with the space $L^2[0,\pi]$ of ordinary
2074: square-integrable functions on $[0,\pi]$. Plotting the
2075: function $\Gamma\psi$ rather than $\psi$ has the advantage that
2076: one can read off directly from the graph the corresponding
2077: probability density with respect to Lebesgue measure on the
2078: parameter space $[0,\pi]$.
2079:
2080: Plots of $\Gamma\psi_i$, $i=\pm,\even,\odd$, are shown in Figure
2081: \rref{Abbpsipm} for $\hbar\scale^2 = \frac 1 2, \frac 1 8, \frac 1
2082: {32}, \frac 1 {128}$. We remark that the value $\hbar\scale^2 =
2083: 1/8$ appears when we choose $\hbar =1$ and the negative of the
2084: Killing form as the invariant scalar product on $\mf g$. Moreover,
2085: according to \eqref{Gchi},
2086: \beq\label{GGammachi}
2087: \big(\Gamma\chi_n\big)(x) = \sqrt 2 \sin\big((n+1)x\big)\,,
2088: \eeq
2089: hence the expansions \eqref{Gpsip}--\eqref{Gpsising} boil down to
2090: ordinary Fourier expansions of the functions $\Gamma\psi_i$,
2091: $i=\pm,\even,\odd$.
2092:
2093: Plots of $\psi_2,\dots,\psi_5$ and $\hat\psi_\pm$ are shown in
2094: Figure \rref{AbbvaniONB} for $\hbar\scale^2 = 1, \frac 1 2, \frac
2095: 1 4, \frac{1}{16}$. For $\hbar\scale^2\to 0$, the outer nodes of
2096: the $\Gamma\psi_n$ run into the points $\cfg_\pm$ and thus
2097: decrease the number of nodes to $n-2$. Moreover, since for
2098: decreasing value of $\hbar\scale^2$ the overlap $\langle
2099: \psi_+,\psi_-\rangle$ decreases, the functions $\hat\psi_\pm$
2100: converge to $\psi_\pm$.
2101:
2102:
2103: \begin{figure}
2104: \begin{center}
2105:
2106: \unitlength1cm
2107:
2108: \begin{picture}(12,2.5)
2109: \put(0.1,0){
2110: \marke{0,0}{cc}{\psi_+}
2111: \marke{4,0}{cc}{\psi_-}
2112: \marke{8,0}{cc}{\psi_\even}
2113: \marke{12,0}{cc}{\psi_\odd}
2114: }
2115: \put(-2.8,-0.2){
2116: \put(0,0){\epsfig{file=psipw.eps,width=4.5cm,height=2.5cm}}
2117: \put(4,0){\epsfig{file=psimw.eps,width=4.5cm,height=2.5cm}}
2118: \put(8,-0.2){\epsfig{file=psi0w.eps,width=4.5cm,height=2.5cm}}
2119: \put(12,-0.2){\epsfig{file=psi1w.eps,width=4.5cm,height=2.5cm}}
2120: }
2121: \end{picture}
2122:
2123: \end{center}
2124:
2125: \caption{\label{Abbpsipm} Plots of images of the wave functions $\psi_i$,
2126: $i=\pm,\even,\odd$ under $\Gamma$, for $\hbar\scale^2 =
2127: 1/128$ (continuous line), $1/32$ (long dash), $1/8$ (short dash),
2128: $1/2$ (alternating short-long dash).}
2129:
2130: \end{figure}
2131:
2132:
2133: \begin{figure}
2134: \begin{center}
2135:
2136: \unitlength1cm
2137:
2138: \begin{picture}(12,5.5)
2139: \put(0,0.7){
2140: \put(0.1,2.5){
2141: \marke{0,0}{cc}{\psi_2}
2142: \marke{4,0}{cc}{\psi_3}
2143: \marke{8,0}{cc}{\psi_4}
2144: \marke{12,0}{cc}{\psi_5}
2145: }
2146: \put(-2.8,3){
2147: \put(0,0){\epsfig{file=psi2.eps,width=4.5cm}}
2148: \put(4,0){\epsfig{file=psi3.eps,width=4.5cm}}
2149: \put(8,-0.2){\epsfig{file=psi4.eps,width=4.5cm}}
2150: \put(12,-0.2){\epsfig{file=psi5.eps,width=4.5cm}}
2151: }
2152: \put(0.1,-0.7){
2153: \marke{2.5,0}{cc}{\hat\psi_+}
2154: \marke{9.5,0}{cc}{\hat\psi_-}
2155: }
2156: \put(-2.2,-0.2){
2157: \put(2.5,0){\epsfig{file=psiphat.eps,width=4.5cm}}
2158: \put(9.5,0){\epsfig{file=psimhat.eps,width=4.5cm}}
2159: }
2160: }
2161: \end{picture}
2162:
2163: \end{center}
2164:
2165: \caption{\label{AbbvaniONB} Plots of the images of the wave functions $\psi_2,
2166: \dots, \psi_4$ and $\hat\psi_\pm$, under $\Gamma$, for
2167: $\hbar\scale^2 = \frac{1}{16}$ (continuous line), $\frac 1 4$
2168: (long dash), $\frac 1 2$ (short dash), $1$ (alternating short-long
2169: dash).}
2170:
2171: \end{figure}
2172:
2173: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2174: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2175:
2176: \section{Energy eigenvalues and eigenstates for $\SU(2)$}
2177: \label{Seigen}
2178:
2179: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2180: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2181:
2182:
2183: We now determine the energy eigenvalues and the corresponding
2184: eigenfunctions of our model for $K = \SU(2)$. We start with a
2185: general discussion of the Hamiltonian.
2186:
2187: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2188: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2189:
2190: \subsection{The Hamiltonian}
2191: \label{SSHam}
2192:
2193: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2194: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2195:
2196: In the Schr\"odinger representation, the Hamiltonian is given by
2197: \eqref{GHaOp}. It is a self-adjoint operator on the Hilbert space
2198: $L^2(\group,\mr d x)^\group \equiv L^2(T,v\mr d t)^W$. For domain
2199: issues it suffices to consider the kinetic part, i.\ e., the
2200: Laplacian $\Delta_\group$. As a core we may take
2201: $C^\infty(\group)^\group \equiv C^\infty(T)^W$. According to
2202: \eqref{Gdomain}, the full domain is
2203: $$
2204: \left\{
2205: \sum\nolimits_{n=0}^\infty \alpha_n \chi_n \in L^2(\group,\mr d
2206: x)^\group
2207: :
2208: \sum\nolimits_{n=0}^\infty |\alpha_n|^2 n^2(n+2)^2 < \infty
2209: \right\}\,.
2210: $$
2211: The isomorphisms $\Gamma_1$ and $\Gamma_2$, see \eqref{GGamma1}
2212: and \eqref{GGamma2}, carry $\Delta_\group$ and $H$ to
2213: the selfadjoint operators
2214: \begin{align*}
2215: \Delta_1 & = \Gamma_1\circ\Delta_\group\circ\Gamma_1^{-1}
2216: \,, &
2217: \Delta_2 & = \Gamma_2\circ\Delta_1\circ\Gamma_2^{-1}
2218: \equiv
2219: \Gamma\circ\Delta_\group\circ\Gamma^{-1}
2220: \\
2221: H_1 & = \Gamma_1\circ H\circ\Gamma_1^{-1}
2222: \,, &
2223: H_2 & = \Gamma_2\circ H_1\circ\Gamma_2^{-1}
2224: \equiv
2225: \Gamma\circ H\circ\Gamma^{-1}
2226: \end{align*}
2227: on the Hilbert spaces $L^2([0,\pi],\sin^2x\mr d x)$ and
2228: $L^2[0,\pi]$, respectively. Then
2229: \beq\label{GH12}
2230: H_i = -\frac{\hbar^2}{2} \Delta_i + \frac \nu 2 (3-2\cos x)
2231: \,,~~~~~~
2232: i=1,2\,,
2233: \eeq
2234: where, formally,
2235: \beq\label{GDelta12}
2236: \Delta_1
2237: =
2238: \scale^2 \left(\frac{1}{\sin(x)} \, \frac{\mr d^2}{\mr d x^2}
2239: \sin(x) + 1\right)
2240: \,,~~~~~~
2241: \Delta_2
2242: =
2243: \scale^2 \left(\frac{\mr d^2}{\mr d x^2} + 1\right)\,.
2244: \eeq
2245: The formula for $\Delta_1$ follows from the general formula for
2246: the radial part of the Laplacian on a compact group, see \cite[\S
2247: II.3.4]{helgaboo}, or by explicitly applying this operator to the
2248: functions $\Gamma_1\chi_n$.
2249:
2250: Let $C^\infty[0,\pi]$ denote the space of Whitney smooth complex
2251: functions on the closed interval $[0,\pi]$. These are the smooth
2252: functions on the open interval $]0,\pi[$ that can be extended to
2253: smooth functions on $\RR$. In particular, the elements of
2254: $C^\infty[0,\pi]$ have well-defined derivatives of arbitrary order
2255: in $0$ and $\pi$.
2256:
2257: \begin{prop}\label{Pcore}
2258:
2259: A core for $\Delta_1$ is given by $D_1 = \{ \psi \in
2260: C^\infty[0,\pi] : \psi'(0) = \psi'(\pi) = 0 \}$. A core for
2261: $\Delta_2$ is given by $D_2 = \{ \psi \in C^\infty[0,\pi] :
2262: \psi(0) = \psi(\pi) = 0 \}$.
2263:
2264: \end{prop}
2265:
2266: \begin{proof}
2267:
2268: First, consider $\Delta_1$. We have to show that
2269:
2270: (a)~ $\Gamma_1\big(C^\infty(\group)^\group\big) \subseteq D_1$,
2271:
2272: (b)~ $\Delta_1(D_1) \subseteq L^2([0,\pi],\sin^2 x \mr d x)$,
2273:
2274: (c)~ $\Delta_1$ is symmetric on $D_1$.
2275:
2276: \noindent We may replace $\Delta_1$ with the operator
2277: $\tilde\Delta_1 = \frac{1}{\sin x}\,\frac{\mr d^2}{\mr d x^2} \sin
2278: x$. Concerning (a), we observe that the algebra of real invariant
2279: polynomials on $\group = \SU(2)$ is generated by the trace
2280: monomial $\rho(a) = \frac 1 2 \tr(a)$. A theorem in
2281: \cite{Schwarz} states that $C^\infty(\group)^\group = \rho^\ast
2282: C^\infty(\RR)$. Hence, for given $\psi\in C^\infty(\group)^\group$
2283: there exists $\vp\in C^\infty(\RR)$ such that $\psi = \vp\circ
2284: \rho$. Then $(\Gamma_1 \psi)(x) = \vp(\cos x)$ and thus
2285: $(\Gamma_1 \psi)'(x) = - h'(\cos x) \sin x$ vanishes for
2286: $x=0,\pi$.
2287:
2288: To check (b), let $\psi\in D_1$. It suffices to show that the
2289: values of the function $(\tilde \Delta_1 \psi)(x)$, $0<x<\pi$,
2290: converge for $x\to0$ and $x\to\pi$. Since $\psi(0)$ and
2291: $\psi''(0)$ exist,
2292: $$
2293: \lim_{x\to 0} \big(\tilde \Delta_1 \psi(x)\big)
2294: =
2295: \psi''(0) - \psi(0) + 2 \lim_{x\to 0} \frac{\cos x \psi'(x)}{\sin
2296: x}\,.
2297: $$
2298: Since $\lim_{x\to 0} \psi'(x) = 0$, we can apply the rule of
2299: Bernoulli and de l'Hospital. This yields
2300: $$
2301: \lim_{x\to 0} (\tilde\Delta_1\psi(x)) = 3 \psi''(0) - \psi(0)\,.
2302: $$
2303: The reasoning for $x\to\pi$ is analogous.
2304:
2305: To prove (c), let $\psi,\vp\in D_1$. Then, omitting the
2306: normalization factor $2/\pi$, we find
2307: \begin{align*}
2308: \int_0^\pi
2309: \ol{\psi(x)} \, (\tilde\Delta_1\vp)(x) \,
2310: \sin^2 x \mr d x
2311: = &
2312: \int_0^\pi
2313: \ol{(\tilde\Delta_1\psi)(x)} \, \vp(x) \, \sin^2 x \mr d x
2314: \\
2315: & +
2316: \sin x \, \psi(x) (\sin x \, \vp(x))'\big|^\pi_0
2317: -
2318: \sin x \, \vp(x) (\sin x \, \psi(x))'\big|^\pi_0 \,.
2319: \end{align*}
2320: The boundary terms vanish because $\psi(x)$, $\psi'(x)$, $\vp(x)$
2321: and $\vp'(x)$ exist for $x=0$ and $x=\pi$.
2322:
2323: Next, consider $\Delta_2$. We have to check conditions (a)--(c)
2324: with the subscript $1$ replaced with the subscript $2$,
2325: with $L^2([0,\pi],\sin^2 x\mr d
2326: x)$ instead of $L^2[0,\pi]$, and with $C^\infty(\group)^\group$
2327: instead of $D_1$. Conditions (a) and (b) are trivially satisfied
2328: and the verification of (c) is analogous to that for $\Delta_1$.
2329:
2330: \end{proof}
2331:
2332:
2333: \bre
2334:
2335: The operator $\Delta_1$ is discussed in \cite[\S 4]{Wren:Rieffel}
2336: as a specific example of a reduced Laplacian obtained by Rieffel
2337: induction. There, the same core is isolated. In our concrete
2338: situation the proof is much simpler than in the general setting of
2339: \cite{Wren:Rieffel}, though.
2340:
2341: \ere
2342:
2343: In view of the proposition, we will now discuss two items. First,
2344: we will relate our system with two standard elementary quantum
2345: mechanical systems. Thereafter, we will make a remark on the
2346: extension problem of the Hamiltonian in a \lq naive\rq\
2347: quantization-after-reduction procedure.
2348:
2349: The proposition implies that, for $\nu=0$, the Hilbert space
2350: isomorphism $\Gamma$ maps our original system to that of a
2351: particle of mass $m=\frac 1 {2\scale^2}$ moving in a
2352: one-dimensional square potential well of width $\pi$ with infinitely high
2353: walls. Inside the well the energy is shifted by $\scale^2 = \frac 1
2354: {2m}$. For $\nu\neq 0$, the potential inside the square well is further
2355: modified by a cosine. This corresponds to a planar pendulum that
2356: is bound to move in one half of the circle only and is reflected
2357: elastically at the two equilibria. It would be interesting to
2358: clarify the relevance of the subspaces $\Hi_\pm$ in both these systems.
2359:
2360: The relationship with the pendulum is in fact more intimate:
2361: Multiplication by the function $\sqrt 2 \sin x$, $x\in[-\pi,\pi]$,
2362: defines a Hilbert space isomorphism from $L^2(T,v\mr d t)\equiv
2363: L^2([-\pi,\pi],\sin^2 x\mr d x)$ onto $L^2(T,\mr d t) \equiv
2364: L^2[-\pi,\pi]$ which maps the subspace $ \Hi$ of $W$-invariants
2365: onto the subspace of odd functions. The Hamiltonian is
2366: given formally by the same expression as $H_2$. A core for this
2367: operator is given by the odd $2\pi$-periodic
2368: $C^\infty$-functions on $\RR$. Hence, this operator describes a
2369: planar pendulum of mass $m = \frac{1}{2\scale^2}$ and ratio of
2370: gravitational acceleration by length given by
2371: $\frac{\nu}{\hbar^2\scale^2}$ with the constraint that among the
2372: states of the pendulum only the odd ones emerge. Finally,
2373: restriction to $[0,\pi]$ defines a Hilbert space isomorphism from
2374: the subspace of $L^2[-\pi,\pi]$ of odd functions onto $L^2[0,\pi]$
2375: that carries the Hamiltonian of the planar pendulum to $H_2$.
2376: Hence, we arrive again at the square potential with cosine
2377: potential inside. By construction, the resulting isomorphism
2378: $\Hi\equiv L^2(T,v\mr d t)^W \to L^2[0,\pi]$ coincides with
2379: $\Gamma$.
2380:
2381: \bre
2382:
2383: The relation between our system and the quantum planar pendulum is
2384: the quantum counterpart of the observation made above that the
2385: reduced classical phase space of our system is isomorphic, as a
2386: stratified symplectic space, to that of a spherical pendulum,
2387: constrained to move with zero angular momentum, reduced relative
2388: to rotations about the vertical axis. This system is manifestly
2389: equivalent to that of a planar pendulum reduced relative to
2390: reflection about the vertical axis.
2391:
2392: \ere
2393:
2394: Now we discuss briefly the extension problem which arises in this context.
2395: Naive quantization after reduction on $\ctg\group$
2396: fails because of the presence of singularities on $\pha$ .
2397:
2398:
2399:
2400: The part of $\ctg \group$ to which
2401: regular cotangent bundle reduction applies
2402: is the cotangent bundle of the unreduced principal stratum
2403: $\group\setminus\{\pm\II\}$.
2404: On this part, symplectic reduction leads to the
2405: cotangent bundle of the quotient manifold, i.e., of the principal stratum
2406: $\cfg_\prin$ but, beware, $\ctg\cfg_\prin$
2407: is a proper subset of the principal stratum
2408: $\pha_\prin$ of the reduced phase space $\pha$
2409: rather than being the entire stratum.
2410: In the
2411: parameterization of $\cfg$ chosen above, $\cfg_\prin$ corresponds to the open
2412: interval $]0,\pi[$. Since the parameterization is an isometry when scaled via
2413: $\scale$, canonical quantization of the kinetic energy then yields the
2414: symmetric operator
2415: \beq\label{GnaiveDelta}
2416: \scale^2~\frac{\mr d^2}{\mr d x^2}\,
2417: \eeq
2418: on the Hilbert space $L^2[0,\pi]$ having as domain the compactly
2419: supported smooth functions on the open interval $]0,\pi[$.
2420: This leads to a naive quantization procedure away from the singularities
2421: of $\cfg$.
2422:
2423: To arrive at a
2424: well-defined quantum theory of the entire system including the singular subset
2425: $\cfg_\sing$, one faces the problem of determining the self-adjoint
2426: extensions of the operator \eqref{GnaiveDelta},
2427: each of which defines a different quantum theory,
2428: and to isolate one of these extensions as the
2429: \lq correct\rq\ one.
2430: Thus, among the different
2431: extensions, one has to pick
2432: one according to the
2433: boundary conditions imposed on the wave functions
2434: and the physical interpretation of the theory
2435: will depend on the choice of boundary conditions.
2436: This is the problem studied in \cite{emmroeme}
2437: in the situation where the classical configuration space is a cone over a
2438: Riemannian manifold; see also \cite{DeserJackiw} and \cite{KayStuder} where
2439: related questions are discussed under a more general perspective. When the
2440: classical configuration space arises by reduction, the extension problem
2441: does not really arise, though, since by
2442: reduction after quantization
2443: the kinetic energy operator is uniquely determined.
2444: This was already observed in \cite{Wren:Rieffel} in the context of
2445: quantization by Rieffel induction. Indeed, in our situation, up to the shift by
2446: $\scale^2$ which, in the case of \eqref{GnaiveDelta},
2447: can be obtained by the metaplectic
2448: correction, $\Delta_2$ is a self-adjoint extension of \eqref{GnaiveDelta}.
2449: According to Proposition \rref{Pcore}, this is the Friedrichs extension.
2450:
2451: To conclude we speculate that some deeper insight into quantization after
2452: reduction will, perhaps, make the kinetic energy
2453: operator unique in general as well.
2454:
2455: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2456: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2457:
2458: \subsection{Eigenvalues and eigenstates}
2459:
2460: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2461: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2462:
2463: For $\inco = 0$, i.\ e., in the strong coupling limit, in view
2464: of \eqref{EWL} and \eqref{GEWCn}, the energy eigenvalues are given
2465: by
2466: $$
2467: E_{n,\inco=0}
2468: =
2469: \frac{\hbar^2}{2} \ve_n
2470: =
2471: \frac{\hbar^2\scale^2}{2} n(n+2)
2472: $$
2473: and the corresponding normalized eigenfunctions are given by the
2474: characters $\chi_n$. To solve the eigenvalue problem for
2475: nonvanishing $\inco$ we carry $H$ via $\Gamma$ to $H_2$. Let
2476: $$
2477: \tinco = \frac{\inco}{\hbar^2\scale^2} \equiv \frac{1}{\hbar^2\scale^2 \coco^2}\,.
2478: $$
2479: According to \eqref{GH12} and \eqref{GDelta12}, on the core $D_2$ of $H_2$,
2480: $$
2481: H_2
2482: =
2483: -\, \frac{\hbar^2\scale^2}{2}
2484: \left(
2485: \frac{\mr d^2}{\mr d x^2}
2486: +
2487: 2 \tinco \cos(x) + \left(1 - 3\tinco\right)\right)\,,
2488: $$
2489: and so the stationary Schr\"odinger equation for $H_2$ reads
2490: \beq\label{GSreqf}
2491: \left(
2492: \frac{\mr d^2}{\mr d x^2}
2493: +
2494: 2 \tinco \cos(x) + \left(\frac{2E}{\hbar^2\scale^2} + 1 -
2495: 3\tinco\right)
2496: \right)
2497: \psi(x)
2498: =
2499: 0,
2500: \eeq
2501: $E$ being the desired eigenvalue and $\psi\in D_2$ the
2502: corresponding eigenfunction. The change of variable $y= (x - \pi)/2$
2503: leads to the Mathieu equation
2504: \beq\label{GMathieu}
2505: \frac{\mr d^2}{\mr d y^2} f(y) + (a-2q\cos(2y)) f(y) = 0\,,
2506: \eeq
2507: where
2508: \beq\label{Gaq}
2509: a = \frac{8 E}{\hbar^2\scale^2} + 4 - 12\tinco
2510: \,,~~~~~~
2511: q = 4\tinco
2512: \,,
2513: \eeq
2514: and $f$ is a Whitney smooth function on the interval $[-\pi/2,0]$
2515: satisfying the boundary conditions
2516: \beq\label{Gboundcond}
2517: f(-\pi/2) = f(0) = 0\,.
2518: \eeq
2519: For the theory of the Mathieu equation and its solutions, called
2520: Mathieu functions, see
2521: \cite{Arscott,McLachlan,MeixnerSchaefke}. All we need is this: for
2522: certain characteristic values of the parameter $a$, depending
2523: analytically on $q$ and usually being denoted by $b_{2n+2}(q)$,
2524: $n=0,1,2,\dots$, solutions satisying \eqref{Gboundcond} exist.
2525: Given $a = b_{2n+2}(q)$, the corresponding solution is unique up
2526: to a complex factor and can be chosen to be real-valued. It is
2527: usually denoted by $\se_{2n+2}(y;q)$, where \lq$\se$\rq\ stands for
2528: sine elliptic. For given $\tinco\geq 0$, let vectors
2529: $\xi_n\in\Hi$ be defined by
2530: \beq\label{Gdefxi}
2531: (\Gamma\xi_n)(x)
2532: =
2533: (-1)^{n+1}\sqrt 2
2534: \left(
2535: \se_{2n+2}\left(\frac{x - \pi}{2};4\tinco\right)
2536: \right)
2537: \,,~~~~~~~
2538: n=0,1,2,\dots\,.
2539: \eeq
2540: Since $\se_{2n+2}(y;0) = \sin((2n+2)y)$ the factor $(-1)^{n+1}$
2541: ensures that for $\tinco=0$ we get $\xi_n = \chi_n$ exactly, and
2542: not only up to a sign.
2543:
2544:
2545: \btm\label{Teval}
2546:
2547: For any $\tinco\geq 0$, the vectors $\xi_n\in\Hi$,
2548: $n=0,1,2,\dots$, form an orthonormal basis of eigenvectors of $H$.
2549: The corresponding eigenvalues are non-degenerate. They are given
2550: by
2551: $$
2552: E_n
2553: =
2554: \frac{\hbar^2\scale^2}{2}
2555: \left(
2556: \frac{b_{2n+2}(4\tinco)}{4} + 3 \tinco - 1
2557: \right)\,.
2558: $$
2559:
2560: \etm
2561:
2562: \bpf
2563:
2564: This follows at once from the fact that, for any value of the
2565: parameter $q$, the functions $\sqrt 2\,\se_{2n+2}(y;q)$, $n =
2566: 0,1,2,,\dots$, form an orthonormal basis in $L^2[-\pi/2,0]$, see
2567: \cite[\S 20.5]{AbramowitzStegun}. Moreover, the characteristic
2568: values satisfy $b_2(q)< b_4(q) < b_6(q) < \cdots$, see \cite[\S
2569: 20.2]{AbramowitzStegun}. Hence, for any value of $\tinco$ we have
2570: $E_0 < E_1 < E_2 < \cdots$. \epf
2571:
2572: \noindent Figure \rref{AbbEn} shows the energy eigenvalues $E_n$
2573: and the level separation $E_{n+1}-E_n$ for $n=0,\dots,8$ as
2574: functions of $\tinco$. The transition energy values manifestly
2575: reverse their order as $\inco$ increases. Figure \rref{Abbxin}
2576: displays the images of eigenfunctions $\xi_n$, $n=0,\dots,3$, under
2577: $\Gamma$ (i.e., the rescaled and shifted Mathieu functions
2578: themselves), for $\tinco = 0,3,6,12,24$. The plots have been
2579: generated by means of the built-in Mathematica functions {\tt
2580: MathieuS} and {\tt MathieuCharacteristicB}.
2581: \begin{figure}
2582:
2583: \hspace*{-1.5cm}\epsfig{file=energies.eps,width=8cm}
2584: \hspace*{-0.5cm}\epsfig{file=energydifferences.eps,width=8cm}
2585:
2586: \caption{\label{AbbEn} Energy eigenvalues (left) $E_n$ and
2587: transition energy values $E_{n+1}-E_n$ (right) for $n=0,\dots,7$ in units of
2588: $\hbar^2\scale^2$ as functions of $\tinco$.}
2589:
2590: \end{figure}
2591:
2592: \begin{figure}
2593: \begin{center}
2594:
2595: \unitlength1cm
2596:
2597: \begin{picture}(12,2.5)
2598: \put(0.1,-0.2){
2599: \marke{0,0}{cc}{\xi_0}
2600: \marke{4,0}{cc}{\xi_1}
2601: \marke{8,0}{cc}{\xi_2}
2602: \marke{12,0}{cc}{\xi_3}
2603: }
2604: \put(-2.8,0.2){
2605: \put(0,0){\epsfig{file=xi0.eps,width=4.5cm}}
2606: \put(4,0){\epsfig{file=xi1.eps,width=4.5cm}}
2607: \put(8,0){\epsfig{file=xi2.eps,width=4.5cm}}
2608: \put(12,0){\epsfig{file=xi3.eps,width=4.5cm}}
2609: }
2610: \end{picture}
2611:
2612: \end{center}
2613:
2614: \caption{\label{Abbxin} Images of the energy eigenfunctions $\xi_0,\dots,\xi_3$,
2615: under $\Gamma$, for $\tinco = 0$ (continuous line), $3$
2616: (long dash), $6$ (short dash), $12$ (alternating short-long dash),
2617: $24$ (dotted line).}
2618:
2619: \end{figure}
2620:
2621: \bre
2622:
2623: The Schr\"odinger equation for the planar pendulum is solved in an
2624: analogous way \cite{Condon}. From the discussion in Subsection
2625: \rref{SSHam} it follows that the only difference is that in the
2626: case of the pendulum, the function $f$ in \eqref{GMathieu} can be
2627: any $\pi$-periodic smooth function on $\RR$. Then, in addition to
2628: the family of $\pi$-periodic odd solutions given by the functions
2629: $\se_{2n+2}(y;q)$ and their characteristic values $b_{2n+2}(q)$
2630: there is a family of $\pi$-periodic even solutions which are
2631: usually denoted by $\ce_{2n+2}$ (for \lq cosine elliptic\rq). The
2632: corresponding characteristic values are usually denoted by
2633: $a_{2n+2}(q)$. For any value of $q$, $a_2(q) < b_2(q) < a_4(q) <
2634: b_4(q) <\cdots$. Thus, precisely every second eigenstate of the
2635: planar pendulum emerges in our system. In particular, the ground
2636: state of our system does not correspond to the ground state but to
2637: the first excited state of the planar pendulum.
2638:
2639: \ere
2640:
2641: \bre
2642:
2643: According to Remark \rref{Remmodel}, Theorem \rref{Teval} yields
2644: the solutions to the stationary Schr\"odinger equation for quantum
2645: Yang-Mills theory on $S^1$ when the self-interaction is described
2646: by the potential in \eqref{GHaFn}. In particular, in this simple
2647: model we have constructed the vacuum and all excited states, for
2648: arbitrary values of the coupling constant.
2649:
2650: \ere
2651:
2652: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2653: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2654:
2655: \section{Expectation values of the costratification orthoprojectors for
2656: $\SU(2)$}
2657:
2658: \label{Sexpect}
2659:
2660: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2661: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2662:
2663:
2664: The most elementary observables associated with the
2665: costratification are the orthoprojectors $\Pi_i$ onto,
2666: respectively, the subspaces $\Hi_i$, $i=\pm,\sing$. The
2667: expectation value of $\Pi_i$ in a state $\psi$ yields the
2668: probability that the system prepared in this state is measured in
2669: the subspace $\Hi_i$. We determine the expectation values of
2670: $\Pi_i$ in the energy eigenstates, i.~e.,
2671: $$
2672: P_{i,n} := \langle \xi_n | \Pi_i \xi_n \rangle = \| \Pi_i\xi_n
2673: \|^2
2674: \,,~~~~~~
2675: i=\sing,\pm\,.
2676: $$
2677: Then, we derive approximations for these expectation values for
2678: strong and weak coupling.
2679:
2680:
2681: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2682: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2683:
2684: \subsection{Expectation values} \label{SSexpect}
2685:
2686: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2687: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2688:
2689:
2690:
2691:
2692: According to \eqref{Goproj},
2693: \beq\label{GPin}
2694: P_{\pm,n}
2695: =
2696: |\langle \xi_n | \psi_\pm \rangle|^2
2697: \,,~~~~~~
2698: P_{\sing,n}
2699: =
2700: |\langle \xi_n | \psi_\even \rangle|^2
2701: +
2702: |\langle \xi_n | \psi_\odd \rangle|^2 .
2703: \eeq
2704: As $\se_{2n+2}$ is odd and $\pi$-periodic, it can be expanded as
2705: $$
2706: \se_{2n+2}(y;q) = \sum_{k=0}\nolimits^\infty B^{2n+2}_{2k+2}(q)
2707: \sin((2k+2)y)\,,
2708: $$
2709: where $B^{2n+2}_{2k+2}(q)$ refers to the Fourier coefficients. The
2710: Fourier coefficients satisfy certain recurrence relations, see
2711: \cite[\S20.2]{AbramowitzStegun}. Using \eqref{GGammachi}, we find
2712: \beq\label{Gscaproxichi}
2713: \langle \xi_n | k \rangle = (-1)^{n+k}
2714: B^{2n+2}_{2k+2}(4\tinco)\,.
2715: \eeq
2716: Then \eqref{Gpsip}--\eqref{Gpsising} yield
2717: \beqa\label{Gxipsip}
2718: \langle \xi_n | \psi_+ \rangle
2719: & = &
2720: \frac{(-1)^n}{N}
2721: \sum\nolimits_{k=0}^\infty
2722: (-1)^k\,(k+1)\,\mr e^{-\hbar\scale^2(k+1)^2/2} \,
2723: B^{2n+2}_{2k+2}(4\tinco) ,
2724: \\ \label{Gxipsim}
2725: \langle \xi_n | \psi_- \rangle
2726: & = &
2727: \frac{(-1)^n}{N}
2728: \sum\nolimits_{k=0}^\infty
2729: (k+1)\,\mr e^{-\hbar\scale^2(k+1)^2/2} \, B^{2n+2}_{2k+2}(4\tinco) ,
2730: \\ \label{Gxipsi0}
2731: \langle \xi_n | \psi_\even \rangle
2732: & = &
2733: \frac{(-1)^n}{N_\even}
2734: \sum\nolimits_{k=0}^\infty
2735: (2k+1) \mr e^{-\hbar\scale^2(2k+1)^2/2} B^{2n+2}_{4k+2}(4\tinco)
2736: \,,
2737: \\ \label{Gxipsi1}
2738: \langle \xi_n | \psi_\odd \rangle
2739: & = &
2740: -\,\frac{(-1)^n}{N_\odd}
2741: \sum\nolimits_{k=0}^\infty
2742: (2k+2) \mr e^{-\hbar\scale^2(2k+2)^2/2}
2743: B^{2n+2}_{4k+4}(4\tinco)\,.
2744: \eeqa
2745: Together with \eqref{GPin}, this yields formulas for the $P_{i,n}$'s,
2746: $i=\sing,\pm$.
2747: We do not spell them out, since they do not lead to
2748: significant simplification.
2749: The functions $P_{i,n}$ depend on the parameters $\hbar$,
2750: $\scale^2$ and $\inco$ only via the combinations $\hbar\scale^2$ and
2751: $\tinco = \inco/(\hbar^2\scale^2)$.
2752: Figure \rref{AbbPin} displays
2753: the $P_{i,n}$, $n=0,\dots,5$, as functions of $\tinco$ for three specific
2754: values of $\hbar\scale^2$, thus treating $\tinco$ and $\hbar\scale^2$
2755: as independent parameters. This is appropriate for the discussion of
2756: the dependence of $P_{i,n}$ on the coupling parameter $g$ for fixed
2757: values of $\hbar$ and $\scale^2$. The plots have been generated by
2758: Mathematica through numerical integration.
2759:
2760: \begin{figure}
2761:
2762: \unitlength1cm
2763:
2764: \begin{picture}(12,11.5)
2765: \put(-1,8){
2766: \put(0,0){\epsfig{file=pp0-5-2.eps,width=6cm,height=2cm}}
2767: \put(5,0){\epsfig{file=pm0-5-2.eps,width=6cm,height=2cm}}
2768: \put(10,0){\epsfig{file=p00-5-2.eps,width=6cm,height=2cm}}
2769: }
2770: \put(-1,4.5){
2771: \put(0,0){\epsfig{file=pp0-5-4.eps,width=6cm,height=2cm}}
2772: \put(5,0){\epsfig{file=pm0-5-4.eps,width=6cm,height=2cm}}
2773: \put(10,0){\epsfig{file=p00-5-4.eps,width=6cm,height=2cm}}
2774: }
2775: \put(-1,1){
2776: \put(0,0){\epsfig{file=pp0-5-6.eps,width=6cm,height=2cm}}
2777: \put(5,0){\epsfig{file=pm0-5-6.eps,width=6cm,height=2cm}}
2778: \put(10,0){\epsfig{file=p00-5-6.eps,width=6cm,height=2cm}}
2779: }
2780: \put(1,7.5){
2781: \marke{0,0}{cl}{P_{+,n}\,,~~\hbar\scale^2 = \frac 1 2}
2782: \marke{5,0}{cl}{P_{-,n}\,,~~\hbar\scale^2 = \frac 1 2}
2783: \marke{10,0}{cl}{P_{\sing,n}\,,~~\hbar\scale^2 = \frac 1 2}
2784: }
2785: \put(1,4){
2786: \marke{0,0}{cl}{P_{+,n}\,,~~\hbar\scale^2 = \frac 1 8}
2787: \marke{5,0}{cl}{P_{-,n}\,,~~\hbar\scale^2 = \frac 1 8}
2788: \marke{10,0}{cl}{P_{\sing,n}\,,~~\hbar\scale^2 = \frac 1 8}
2789: }
2790: \put(1,0.5){
2791: \marke{0,0}{cl}{P_{+,n}\,,~~\hbar\scale^2 = \frac 1 {32}}
2792: \marke{5,0}{cl}{P_{-,n}\,,~~\hbar\scale^2 = \frac 1 {32}}
2793: \marke{10,0}{cl}{P_{\sing,n}\,,~~\hbar\scale^2 = \frac 1 {32}}
2794: }
2795: \end{picture}
2796:
2797: \caption{\label{AbbPin} Expectation values $P_{+,n}$, $P_{-,n}$
2798: and $P_{\sing,n}$ for $n=0$ (continuous line), $1$ (long dash), $2$
2799: (short dash), $3$ (long-short dash), $4$ (dotted line) and $5$
2800: (long-short-short dash), plotted over $\log\tinco$
2801: for $\hbar\scale^2 = \frac 1 2, \frac 1 8, \frac 1 {32}$.}
2802:
2803: \end{figure}
2804:
2805: Perhaps the most impressive feature is the dominant peak of
2806: $P_{+,0}$ which is enclosed by less prominent maxima of the other
2807: $P_{+,n}$ and moves to higher $\tinco$ when $\hbar\scale^2$
2808: decreases. In other words, for a certain value of the coupling constant, the
2809: state $\psi_+$ which spans $\Hi_+$ seems to coincide almost
2810: perfectly with the ground state. If the two states coincided completely
2811: then \eqref{Gscaproxichi} would imply that, for a certain value of $q$,
2812: the coefficients $B^{2n+2}_{2k+2}(q)$ would be given by
2813: $(-1)^{n+k}\frac 1 N (k+1)\mr e^{-\hbar\scale^2(k+1)^2/2}$.
2814: However, this is not true; the latter expressions do not
2815: satisfy the recurrence relations valid for the coefficients
2816: $B^{2n+2}_{2k+2}(q)$ for any value of $q$. Another interesting phenomenon is
2817: that, for decreasing $\hbar\scale^2$, the maxima of $P_{-,n}$ move
2818: to lower $\tinco$ and the subsequent descent becomes steeper.
2819:
2820: Next, we will derive approximations for the $P_{i,n}$'s for small
2821: and large values of $\tinco$. When $\hbar$ and $\scale$ are fixed,
2822: this corresponds to strong and weak coupling, as appropriate. The
2823: strong coupling approximation will provide a resolution of the
2824: first crossings of the graphs of the $P_{i,n}$. The weak coupling
2825: approximation will allow us to analyze the position and the height
2826: of the dominant peak of $P_{+,0}$ as well as of the subsequent
2827: maxima of the other $P_{+,n}$'s. A detailed study of the maxima of
2828: the $P_{-,n}$'s and of the behaviour of the $P_{+,n}$'s in the intermediate
2829: region between strong and weak coupling remains as a future task.
2830:
2831: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2832: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2833:
2834: \subsection{Strong coupling approximation}
2835:
2836: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2837: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2838:
2839: In the region of strong coupling, i.~e., for large $g$, $\tinco$ is
2840: small, at least when the parameter $\hbar\scale^2$ is fixed. Power
2841: series expansions for the characteristic values $b_{2n+2}(q)$ in
2842: $q$ about $q = 4\tinco = 0$ can be found, e.~g., in \cite[\S
2843: 20.2.25]{AbramowitzStegun}. They immediately provide expansions
2844: for the energy eigenvalues. We do not spell out the latter here,
2845: because we are merely interested in approximations of the
2846: expectation values $P_{i,n}$, $i=\pm,\sing$. Quadratic
2847: approximations for the Fourier coefficients $B^{2n+2}_{2k+2}(q)$
2848: in $q$ can be read off from \cite[\S 2.25, (42)]{MeixnerSchaefke}:
2849: For the central coefficients, this yields
2850: $$
2851: \begin{array}{rclcrcl}
2852: B^2_2(4\tinco) & = & 1 - \frac{1}{18} \, \tinco^2 + O(\tinco^3)
2853: \,, & &
2854: B^{2n+2}_{2n+2}(4\tinco)
2855: & = &
2856: 1 - \frac{(2n+2)^2 + 1}{2((2n+2)^2-1)^2} \, \tinco^2 + O(\tinco^3)
2857: \,,~~ n \geq 1\,.
2858: \end{array}
2859: $$
2860: For the next-to-central coefficients,
2861: $$
2862: \begin{array}{rclcrcl}
2863: B^{2n+2}_{2n}(4\tinco) & = & \frac{1}{(2n+1)}\, \tinco +
2864: O(\tinco^3)
2865: \,, & \hspace*{-0.5cm} &
2866: B^{2n+2}_{2n+4}(4\tinco) & = & - \frac{1}{(2n+3)} \tinco +
2867: O(\tinco^3)
2868: \,,
2869: \\
2870: B^{2n+2}_{2n-2}(4\tinco) & = & \frac{1}{4n(2n+1)} \tinco^2 +
2871: O(\tinco^3)
2872: \,, & &
2873: B^{2n+2}_{2n+6}(4\tinco) & = & \frac{1}{2(2n+3)(2n+4)} \tinco^2 +
2874: O(\tinco^3)
2875: \,,~~ n\geq 0\,.
2876: \end{array}
2877: $$
2878: All the other coefficients are of order $O(\tinco^3)$. Using
2879: \eqref{GPin} and \eqref{Gxipsip}--\eqref{Gxipsi1} we obtain
2880: $$
2881: P_{\pm,n}
2882: =
2883: \frac{A_{n,0}^2}{N^2}
2884: \pm
2885: \frac{2 A_{n,0} A_{n,1}}{N^2}~ \tinco
2886: +
2887: \frac{A_{n,1}^2 + A_{n,0} A_{n,2}}{N^2}~ \tinco^2
2888: +
2889: O(\tinco^3)
2890: $$
2891: and, for $n$ even, we get
2892: $$
2893: P_{\sing,n} = \frac{A_{n,0}^2}{N_\even^2}
2894: + \left(
2895: \frac{A_{n,1}^2}{N_\odd^2} + \frac{A_{n,0}A_{n,2}}{N_\even^2}
2896: \right) \tinco^2 + O(\tinco^3)
2897: \,,
2898: $$
2899: whereas, for $n$ odd, in this expression,
2900: one has to interchange $N_\even$ and $N_\odd$. The coefficients are
2901: $$
2902: \begin{array}{rcl}
2903: A_{n,0} & = & (n+1) \mr e^{-\hbar\scale^2(n+1)^2/2}
2904: \,,~~~~~~
2905: n\geq 0\,,
2906: \\
2907: A_{n,1}
2908: & = &
2909: \frac{(n+2)\mr e^{-\hbar\scale^2(n+2)^2/2}}{2n+3}
2910: -
2911: \frac{n\mr e^{-\hbar\scale^2 n^2/2}}{2n+1}
2912: \,,~~~~~~
2913: n\geq 0\,,
2914: \\
2915: A_{0,2}
2916: & = &
2917: \frac{\mr e^{-9\hbar\scale^2/2}}{4} - \frac{\mr
2918: e^{-\hbar\scale^2/2}}{9}\,,
2919: \\
2920: A_{n,2}
2921: & = &
2922: \frac{(n-1)\mr e^{-\hbar\scale^2(n-1)^2/2}}{2n (2n+1)}
2923: +
2924: \frac{(n+3)\mr e^{-\hbar\scale^2(n+3)^2/2}}{(2n+3)(2n+4)}
2925: \\
2926: & &
2927: \phantom{\frac{(n-1)\mr e^{-\hbar\scale^2(n-1)^2/2}}{2n (2n+1)}}
2928: -
2929: (n+1) \mr e^{-\hbar\scale^2(n+1)^2/2}
2930: \left(
2931: \frac{1}{(2n+1)^2} + \frac{1}{(2n+3)^2}
2932: \right)
2933: \,,~~~~~~
2934: n\geq 1\,.
2935: \end{array}
2936: $$
2937: For $\hbar\scale^2 = \frac 1 8$,
2938: plots of the quadratic approximations of the $P_{i,n}$'s, $i=\pm,\sing$,
2939: are shown in Figure
2940: \rref{AbbPinstrong}, for $n=0,\dots,5$ and $\tinco$ ranging between
2941: $0$ and $0.2$. Here the approximation has a relative error of less
2942: than $0.01$. The plots yield, in particular, a resolution of the
2943: first crossings of the graphs of the $P_{i,n}$'s in the bottom line
2944: of Figure \rref{AbbPin}.
2945:
2946: \begin{figure}
2947:
2948: \epsfig{file=ppquad0-5.eps,width=5cm}
2949: \epsfig{file=pmquad0-5.eps,width=5cm}
2950: \epsfig{file=p0quad0-5.eps,width=5cm}
2951:
2952: \caption{\label{AbbPinstrong} Quadratic approximations for
2953: $P_{+,n}$, $P_{-,n}$, $P_{0,n}$ (from left to right),
2954: $n=0,\dots,5$, $\hbar\scale^2=\frac 1 8$, plotted over $\tinco$.}
2955:
2956: \end{figure}
2957: For very strong
2958: coupling, the state $\xi_2$ rather than the ground state has
2959: the highest probability to be measured in $\Hi_+$.
2960: In fact the ground state is excelled by all $\xi_n$ with $n\leq 4$.
2961: (This follows of course directly from consideration of the case $\inco
2962: = 0$, where $\xi_n = \chi_n$.) The precise order of the
2963: expectation values in this region is
2964: $$
2965: P_{i,2} \geq P_{i,1} \geq P_{i,3} \geq P_{i,4} \geq P_{i,0} \geq
2966: P_{i,5} \geq P_{i,6} \geq \cdots\,,~~~~~~ i=\sing,\pm\,.
2967: $$
2968: On the other hand, the probabilities $P_{i,0}$ of the ground state
2969: change most rapidly as $\inco$ increases. In particular, $P_{+,0}$
2970: has overtaken all other probabilities already for $\inco = 0.2$.
2971:
2972: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2973: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2974:
2975: \subsection{Weak coupling approximation}
2976:
2977: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2978: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2979:
2980: Similarly to the approximation of a classical planar pendulum by a
2981: classical harmonic oscillator, for excitations that are small
2982: compared with the length of the pendulum, the quantum planar
2983: pendulum can be approximated by a quantum harmonic oscillator for
2984: energy values that are small compared with the range of the potential
2985: \cite{AldrovandiLealFerreira, BakerBlackburnSmith, Condon, PradhanKhare}. We
2986: use this procedure to obtain approximations for the energy eigenfunctions
2987: $\xi_n$ and, from these, approximations for the expectation values $P_{i,n}$,
2988: $i=\pm,\sing$, for large $\tinco$ and small $n$.
2989:
2990: Consider the Schr\"odinger equation for $H_2$ in \eqref{GSreqf}.
2991: Making the change of variable $z = \sqrt[4]{\tinco}\,x$ we obtain the
2992: equation
2993: $$
2994: \left(\frac{\mr d^2}{\mr d z^2}
2995: +
2996: 2\tinco \cos(z/\sqrt[4]{\tinco})
2997: +
2998: \left(
2999: \frac{1}{\sqrt\tinco}
3000: \left(
3001: \frac{2E}{\hbar^2\scale^2}
3002: +
3003: 1
3004: \right)
3005: -
3006: 3 \sqrt\tinco
3007: \right)\right)
3008: f(z) = 0
3009: $$
3010: where $f$ may be a Whitney smooth function on the interval
3011: $[0,\sqrt[4]\tinco\pi]$ satisfying the boundary conditions $f(0) =
3012: f(\sqrt[4]\tinco\pi) = 0$. Replacing the cosine with its second
3013: order Taylor expansion we obtain the Schr\"odinger equation
3014: \beq\label{GSreqhO}
3015: \left(
3016: \frac{\mr d^2}{\mr d z^2}
3017: -
3018: z^2
3019: +
3020: 2\epsilon
3021: \right) f(z) = 0
3022: \eeq
3023: of the harmonic oscillator with unit frequency, where
3024: \beq\label{frequency}
3025: \epsilon
3026: =
3027: \frac{1}{\sqrt\tinco}
3028: \left(
3029: \frac{E}{\hbar^2\scale^2}
3030: +
3031: \frac 1 2
3032: \right)
3033: -
3034: \frac{\sqrt\tinco}{2}\,.
3035: \eeq
3036: For large $\tinco$ and small energies, the solutions of either
3037: equation are concentrated about $x = z/\sqrt[4]{\tinco} \sim 0$.
3038: Under these circumstances, restriction to the interval
3039: $[0,\sqrt[4]\tinco\pi]$ of solutions of \eqref{GSreqhO} satisfying
3040: $f(0) = 0$ yields satisfactory approximations for solutions of
3041: \eqref{GSreqf}. The appropriate solutions of \eqref{GSreqhO} are
3042: well known to be
3043: $$
3044: f(z) = H_{2n+1}(z) \mr e^{-z^2/2}
3045: \,,~~~~~~
3046: \epsilon
3047: =
3048: 2n + \frac 3 2
3049: \,,~~~~~~
3050: n=0,1,2,\dots\,,
3051: $$
3052: where
3053: $$
3054: H_{2n+1}(z)
3055: =
3056: \sum_{r=0}^n \frac{(-1)^{n+r} (2n+1)! (2 z)^{2r+1}}{(n-r)! (2r+1)!}
3057: $$
3058: are the odd degree Hermite polynomials. Define vectors
3059: $\xi^{(\infty)}_n\in\Hi$ by
3060: \begin{eqnarray}\label{Gdefxiinf}
3061: (\Gamma\xi^{(\infty)}_n)(x)
3062: & = &
3063: (-1)^n\,N^{(\infty)}_n
3064: ~
3065: H_{2n+1}\!\left(\sqrt[4]{\tinco}\, x\right)
3066: ~
3067: \mr e^{-\sqrt{\tinco}\, x^2/2}
3068: \,,
3069: \\ \label{GNnhO}
3070: N^{(\infty)}_n
3071: & = &
3072: \frac{\pi^{1/4}\,\tinco^{1/8}}{2^{n+1}\,\sqrt{(2n+1)!}}\,,
3073: \end{eqnarray}
3074: where the choice of sign is dictated by that for the $\xi_n$'s, see
3075: \eqref{Gdefxi}. The $\xi^{(\infty)}_n$'s form a
3076: basis of $\Hi$. Substituting for $\epsilon$ the
3077: right-hand side of \eqref{frequency}, we obtain the energy values
3078: $$
3079: E^{(\infty)}_n
3080: =
3081: \frac{\hbar^2\scale^2}{2}
3082: \left(
3083: \tinco + (4n+3) \sqrt\tinco -1\right)\,.
3084: $$
3085: The $E^{(\infty)}_n$'s and the $\xi^{(\infty)}_n$'s yield
3086: approximations for the true energy eigenvalues $E_n$ and for the
3087: eigenfunctions $\xi_n$ of our model for large $\tinco$ and small
3088: $n$. Note that the $\xi_n$'s are neither orthogonal nor normalized, because the
3089: functions $\Gamma\xi^{(\infty)}_n$ are orthogonal over the interval $[0,\infty]$
3090: rather than the interval $[0,\pi]$ and the normalization factor $N^{(\infty)}_n$
3091: is therefore also chosen so that the functions are normalized
3092: over the interval $[0,\infty]$. The deviation from being orthonormal is however
3093: negligible for small $n$ and large $\tinco$.
3094:
3095: To compute the scalar products $\langle\chi_k,\xi^{(\infty)}_n\rangle$, we use
3096: \eqref{GGammachi}
3097: and \eqref{Gdefxiinf} and move the upper bound of the resulting integral from
3098: $\pi$ to infinity, which is again justified for large $\tinco$ and small $n$.
3099: The result is
3100: \beq\label{Gscaproxichi2}
3101: \langle\chi_k,\xi^{(\infty)}_n\rangle
3102: =
3103: \frac{2^{-n} \, \pi^{-1/4}}{\sqrt{(2n+1)!}}
3104: \,
3105: \tinco^{-1/8}
3106: \,
3107: H_{2n+1}\big((k+1)\,\tinco^{-1/4}\big)
3108: \,
3109: \mr e^{-(k+1)^2 \, \tinco^{-1/2}/2}
3110: \,.
3111: \eeq
3112: This formula is also a consequence of \eqref{Gscaproxichi} and the
3113: asymptotic expansion of the Fourier coefficients
3114: $B^{2n+2}_{2k+2}(q)$ for large $q$ given in
3115: \cite[\S 2.333]{MeixnerSchaefke}. Using \eqref{Gscaproxichi2} and
3116: writing out the formula of the Hermite polynomials, we obtain
3117: \beq\label{GscaprohoHerm}
3118: \langle \psi_i,\xi_n^{(\infty)}\rangle
3119: =
3120: (-1)^n \frac{\sqrt{(2n+1)!}}{2^n \pi^{1/4} N_i}
3121: \sum_{r=0}^n
3122: \frac{(-1)^r \, 2^{2r+1}\, \tinco^{-(4r+3)/8}}{(n-r)! (2r+1)!}
3123: ~~\Sigma_i^r\!\!\left(\frac{\hbar\scale^2 +
3124: \tinco^{-1/2}}{2}\right)\,,
3125: \eeq
3126: where $i=\pm,\even,\odd$, $N_\pm \equiv N$, and
3127: \begin{align*}
3128: \Sigma_+^r(b) & = \sum_{k=1}^\infty k^{2r+2} \mr e^{-b k^2}
3129: \,, &
3130: \Sigma_-^r(b) & = \sum_{k=1}^\infty (-1)^{k+1} k^{2r+2} \mr e^{-b
3131: k^2}\,.
3132: \\
3133: \Sigma_\even^r(b) & = \sum_{k\text{ odd}} k^{2r+2} \mr e^{-bk^2}
3134: \,, &
3135: \Sigma_\odd^r(b) & = \sum_{k\text{ even}} k^{2r+2} \mr e^{-b
3136: k^2}\,.
3137: \end{align*}
3138: Expressing the sums in terms of the theta constant $\theta_3$, see
3139: \eqref{Gdeftheta}, we obtain
3140: \begin{align*}
3141: \Sigma_+^r(b)
3142: & =
3143: \frac 1 2 \frac{\mr d^{r+1}}{\mr d (-b)^{r+1}} \theta_3(\mr
3144: e^{-b})
3145: \,, &
3146: \Sigma_-^r(b) & = \Sigma_+^r(b) - 2^{2r+3} \Sigma_+^r(4b)
3147: \,,
3148: \\
3149: \Sigma_\even^r(b) & = \Sigma_+^r(b) - 4^{r+1} \Sigma_+^r(4b)
3150: \,, &
3151: \Sigma_\odd^r(b) & = 4^{r+1} \Sigma_+^r(4b)
3152: \,.
3153: \end{align*}
3154: Substituting in \eqref{GscaprohoHerm}, for $\Sigma_i^r$, the right-hand side of
3155: each of these identities as appropriate and taking the square, we
3156: arrive at the harmonic oscillator approximations of $P_{\pm,n}$ and
3157: $P_{0,n}$. These approximations are hard to handle, however, as they contain
3158: higher derivatives of the theta constant w.r.t.\ the nome. Instead, we
3159: use the approximation
3160: \beq\label{Gapproxtheta}
3161: \theta_3(e^{-y}) = \sqrt\pi y^{-1/2} + \dots\,,
3162: \eeq
3163: valid for small $y$ and hence for small $\hbar\scale^2$ and
3164: large $\tinco$. Even for $y=1$, this approximation has a relative error of only
3165: $10^{-4}$. In this approximation,
3166: \beq\label{GSigma+r}
3167: \Sigma_+^r(b)
3168: =
3169: \sqrt\pi\frac{(2r+1)!}{4^{r+1} r!} b^{-(2r+3)/2}
3170: \,,~~~~~~
3171: \Sigma_-^r(b) = 0
3172: \,,~~~~~~
3173: \Sigma_\even^r(b) = \Sigma_\odd^r(b) = \frac 1 2 \Sigma_+^r(b)\,,
3174: \eeq
3175: and
3176: $$
3177: N
3178: =
3179: \sqrt{\Sigma_+^2(\hbar\scale^2)}
3180: =
3181: \frac{\pi^{1/4}}{\sqrt 2} (\hbar\scale^2)^{-3/4}
3182: \,,~~~~~~
3183: N_\even = N_\odd = \frac{1}{\sqrt 2} N\,.
3184: $$
3185: In particular, $\Hi_+$ and $\Hi_-$ appear to be orthogonal.
3186: Moreover, \eqref{GscaprohoHerm} yields $P_{-,n} = 0$ and $P_{0,n}
3187: = P_{+,n}$, so that it suffices to determine $P_{+,n}$. Inserting
3188: $\Sigma_+^r$ from \eqref{GSigma+r} into \eqref{GscaprohoHerm} and writing
3189: $$
3190: \ratio
3191: =
3192: \sqrt\hbar\scale\tinco^{1/4} \equiv \sqrt{\frac{\scale}{g}}
3193: $$
3194: we obtain the identity
3195: $$
3196: \langle \psi_+,\xi_n^{(\infty)}\rangle
3197: =
3198: (-1)^n \frac{\sqrt{(2n+1)!}}{2^n}
3199: \sum_{r=0}^n
3200: \frac{(-1)^r 2^{(2r+3)/2}}{r! (n-r)!}
3201: \frac{\ratio^{3/2}}{(\ratio^2 + 1)^{(2n+3)/2}}\,.
3202: $$
3203: Taking the sum yields
3204: \beq\label{Gscaproprat}
3205: \langle \psi_+,\xi_n^{(\infty)}\rangle
3206: =
3207: (-1)^n \frac{\sqrt{(2n+1)!}}{2^n \, n!}
3208: \left(
3209: \frac{2\ratio}{\ratio^2 + 1}
3210: \right)^{3/2}
3211: \left(
3212: \frac{\ratio^2 - 1}{\ratio^2 + 1}
3213: \right)^n
3214: \,.
3215: \eeq
3216: Hence, in the harmonic oscillator approximation and the
3217: approximation of $\theta_3$ by \eqref{Gapproxtheta}, the
3218: expectation values $P_{+,n}$ are given by the rational functions
3219: \beq \label{GPprat}
3220: P_{+,n}^{(\infty)}(\tau)
3221: =
3222: \frac{(2n+1)!}{4^n \, (n!)^2}
3223: \left(
3224: \frac{2\ratio}{\ratio^2 + 1}
3225: \right)^3
3226: \left(
3227: \frac{\ratio^2 - 1}{\ratio^2 + 1}
3228: \right)^{2n}\,.
3229: \eeq
3230: It is interesting to note that, in this approximation, $P_{+,n}$ depends on the
3231: parameters $\hbar$, $\scale$ and $\nu$ only through the
3232: ratio $\scale/\coco$. Figure \rref{AbbPprat} shows plots of $P_{+,n}^{(\infty)}$
3233: and $P_{+,n}$ as functions of $\tinco$ on a logarithmic scale for
3234: $\hbar\scale^2 = 1, \frac 1 2, \frac 1 8$ and $n=0,\dots,3$. We
3235: see that for sufficiently small values of $\hbar\scale^2$ and
3236: sufficiently small $n$ the approximation of $P_{+,n}$ by
3237: $P_{+,n}^{(\infty)}$ is already satisfactory in the region of the dominant
3238: maximum of $P_{+,0}$ and even more so for larger $\tau$. Hence, this
3239: approximation can be used to study the behaviour of $P_{+,n}$ in this region.
3240: \begin{figure}
3241:
3242: \begin{center}
3243:
3244: \unitlength1cm
3245:
3246: \begin{picture}(12,2.7)
3247:
3248: \put(0.1,-0.2){
3249: \marke{0,0}{cc}{n=0}
3250: \marke{4,0}{cc}{n=1}
3251: \marke{8,0}{cc}{n=2}
3252: \marke{12,0}{cc}{n=3}
3253:
3254: \marke{-1.05,2.45}{cb}{\scriptscriptstyle a}
3255: \marke{-0.525,2.45}{cb}{\scriptscriptstyle b}
3256: \marke{-0.05,2.45}{cb}{\scriptscriptstyle c}
3257:
3258: \marke{3.65,2.1}{cb}{\scriptscriptstyle a}
3259: \marke{4,2.1}{cb}{\scriptscriptstyle b}
3260: \marke{4.5,2.1}{cb}{\scriptscriptstyle c}
3261:
3262: \marke{7.7,1.72}{cb}{\scriptscriptstyle a}
3263: \marke{8.15,1.72}{cb}{\scriptscriptstyle b}
3264: \marke{8.65,1.72}{cb}{\scriptscriptstyle c}
3265:
3266: \marke{11.8,1.8}{cb}{\scriptscriptstyle a}
3267: \marke{12.3,1.8}{cb}{\scriptscriptstyle b}
3268: \marke{12.8,1.8}{cb}{\scriptscriptstyle c}
3269: }
3270: \put(-3,0.2){
3271: \put(0,0){\epsfig{file=pp0rat.eps,width=4.5cm}}
3272: \put(4,0){\epsfig{file=pp1rat.eps,width=4.5cm}}
3273: \put(8,0){\epsfig{file=pp2rat.eps,width=4.5cm}}
3274: \put(12,0){\epsfig{file=pp3rat.eps,width=4.5cm}}
3275: }
3276: \end{picture}
3277:
3278: \end{center}
3279:
3280: \caption{\label{AbbPprat}Exact values of $P_{+,n}$ (continuous
3281: lines) and approximations $P_{+,n}^{(\infty)}$ (dashed lines) as
3282: functions of $\tinco$ on a logarithmic scale for $\hbar\scale^2 =
3283: \frac 1 2 (a), \frac 1 8 (b), \frac 1 {32} (c)$ and $n=0,\dots,3$.}
3284:
3285: \end{figure}
3286: Moreover, we claim that this approximation is consistent in the
3287: sense that, for any $\ratio>0$,
3288: $$
3289: P_{+,n}^{(\infty)}(\ratio) \geq 0
3290: \,,~~~~~~
3291: \sum_{n=0}^\infty P_{+,n}^{(\infty)}(\ratio) = 1
3292: \,.
3293: $$
3294: Indeed,
3295: \beq\label{GsumPp}
3296: \sum_{n=0}^\infty P_{+,n}^{(\infty)}(\ratio)
3297: =
3298: \left(
3299: \frac{2\ratio}{\ratio^2 + 1}
3300: \right)^3
3301: \sum_{n=0}^\infty \frac{(2n+1)!}{4^n \, (n!)^2}
3302: \left(
3303: \frac{\ratio^2 - 1}{\ratio^2 + 1}
3304: \right)^{2n}\,.
3305: \eeq
3306: Recall that the function $y\mapsto (1-y)^{-3/2}$ has the Taylor series
3307: $$
3308: (1-y)^{-3/2} = \sum_{n=0}^\infty \frac{(2n+1)!}{4^n \, (n!)^2}
3309: ~y^n\,,
3310: $$
3311: and this series is absolutely convergent for $|y|<1$. Replacing $y$ with
3312: $(\ratio^2 - 1)^2/(\ratio^2 + 1)^2$, where $\tau>0$, we deduce that the
3313: approximation is consistent in the asserted sense.
3314:
3315: We determine the extremal points of $P_{+,n}^{(\infty)}$ on the positive
3316: semiaxis. For $n=0$,
3317: $$
3318: \frac{\mr d}{\mr d \ratio}\,P_{+,0}^{(\infty)}(\ratio)
3319: =
3320: \frac{24 \ratio^2(1-\ratio^2)}{(\ratio^2+1)^4}\,.
3321: $$
3322: Hence, at $\ratio = 1$, $P_{+,0}^{(\infty)}$ has a maximum, the maximal value
3323: being
3324: $$
3325: P_{+,0}^{(\infty)} (\ratio = 1) = 1\,.
3326: $$
3327: This means that, for coupling constant $g=\scale$, up to the
3328: approximations we have made, the state $\psi_+$ spanning $\Hi_+$
3329: coincides with the ground state. In particular, the state $\psi_+$
3330: is then approximately stationary. As remarked in Subsection
3331: \rref{SSexpect}, the coincidence holds only in the approximation
3332: and is not exact though. A physical interpretation of this
3333: phenomenon has still to be found. For $n\geq 1$,
3334: $$
3335: \frac{\mr d}{\mr d \ratio}\,P_{+,n}^{(\infty)}(\ratio)
3336: =
3337: - \frac{(2n+1)!}{2^{2n-3} \, (n!)^2}~
3338: \frac{\ratio^2(\ratio^2 - 1)^{2n-1}}{(\ratio^2 + 1)^{2n+4}}
3339: \left(
3340: 3 \ratio^4 - (8n+6) \ratio^2 + 3
3341: \right)\,.
3342: $$
3343: Hence, $P_{+,n}^{(\infty)}$ has maxima at
3344: \beq\label{GPpratmax}
3345: \ratio_\pm
3346: =
3347: \sqrt{
3348: \frac{4n + 3 \pm 2 \sqrt{4 n^2 + 6 n}}{3}
3349: }
3350: \eeq
3351: and a minimum at $\ratio = 1$. The first maximum, $\ratio_-$, lies
3352: in a region where the approximation is reliable only for very
3353: small values of $\hbar\scale^2$, see Figure \rref{AbbPprat}. For
3354: increasing $n$,
3355: $\ratio_-$ approaches $\ratio=1$ from below and $\ratio_+$ moves towards
3356: larger
3357: values of $\ratio$. The maximal values of $P_{+,n}^{(\infty)}$ are
3358: $$
3359: P_{+,n}^{(\infty)}(\ratio_\pm)
3360: ~~=~~
3361: \frac{3^{3/2} (2n+1)!}{2^{2n-3} \, (n!)^2}
3362: ~
3363: \frac{
3364: \left(4n + 3 \pm 2 \sqrt{4 n^2 + 6 n}\right)^{3/2} \left(4n \pm 2
3365: \sqrt{4 n^2 + 6 n}\right)^{2n}
3366: }{
3367: \left(4n + 6 \pm 2 \sqrt{4 n^2 + 6 n}\right)^{2n+3}
3368: }\,.
3369: $$
3370: These values are independent of the parameters $\hbar$, $\scale$ and $\nu$ and
3371: decrease for increasing $n$.
3372:
3373: In the minimum $\ratio = 1$, $P_{+,n}^{(\infty)}$ vanishes. This
3374: is consistent with what we have found for
3375: $P_{+,0}^{(\infty)}$. The order of contact of $P_{+,n}^{(\infty)}$
3376: with the real axis is $2n$. This order of contact is reflected in a broadening
3377: of the valley between the two maxima, see Figure \rref{AbbPprat}.
3378:
3379: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3380: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3381:
3382: \section{Outlook}
3383:
3384: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3385: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3386:
3387: There is still more to say about the case of $\SU(2)$. The expectation values
3388: $P_{\pm,n}$ and $P_{0,n}$ in the region between the strong and weak coupling
3389: approximations and the dynamics relative to the costratified structure remain to
3390: be studied. The exploration of that dynamics will rely on a detailed
3391: investigation of the probability flow into and out of the subspaces $\Hi_\pm,
3392: \Hi_\sing$.
3393:
3394: The next step is the construction of the costratified Hilbert space and the
3395: subsequent analysis of various physical quantities for $\SU(3)$. Here, the orbit
3396: type stratification of the reduced phase space has a 4-dimensional stratum, a
3397: 2-dimensional stratum, and three isolated points. Thereafter the construction
3398: remains to be extended to an arbitrary lattice.
3399:
3400: Finally, the costratified Hilbert space structure exploited in this paper
3401: implements the stratification of the reduced classical phase space
3402: on the level of states but leaves open the question what the stratification
3403: might signify for the quantum observables, a question to be clarified in the
3404: future. Then the physical role of this stratification can, perhaps, be
3405: studied in more realistic models like lattice QCD \cite{qcd1,qcd2,qcd3}.
3406:
3407:
3408:
3409: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3410: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3411:
3412: \section*{Acknowledgement}
3413:
3414: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3415: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3416:
3417: The authors would like to express their gratitude to
3418: Szymon Charzy\'nski, Heinz-Dietrich
3419: Doebner, Alexander Hertsch, Jerzy Kijowski and Konrad Schm\"udgen for
3420: the stimulus of conversation,
3421: to Brian Hall for pointing out the relationship between the
3422: states spanning $\Hi_\pm$ and coherent states,
3423: to Christian Fleischhack for hints at Thiemann's work on
3424: the overlap of coherent states,
3425: and to Jim Stasheff for a number of comments
3426: which helped improve the exposition.
3427: J.\ H.\ and M.\ S.\ acknowledge funding by the German Research
3428: Council (DFG) under contract \ Le 758/22-1 and contract RU692/3,
3429: respectively.
3430:
3431:
3432: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3433: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3434:
3435: \begin{thebibliography}{19}
3436:
3437: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3438: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3439:
3440: {\footnotesize
3441:
3442: \bibitem{AbrahamMarsden}
3443: R.\ Abraham, J.E.\ Marsden:
3444: {\em Foundations of mechanics.}
3445: Benjamin/Cummings Publishing Co., Inc., Reading, Mass., 1978
3446:
3447: \bibitem{AbramowitzStegun}
3448: M.\ Abramowitz, I.A.\ Stegun:
3449: {\em Handbook of Mathematical Functions\/} (abridged edition).
3450: Verlag Harri Deutsch, 1984
3451:
3452: \bibitem{AldrovandiLealFerreira}
3453: R.\ Aldrovandi, P.\ Leal Ferreira:
3454: Quantum pendulum.
3455: Amer.\ J.\ Phys.\ {\bf 48}, 660--664 (1980)
3456:
3457: \bibitem{armcusgo}
3458: J.M.\ Arms, R.\ Cushman, M.\ J.\ Gotay:
3459: A universal reduction procedure for Hamiltonian group actions.
3460: In: {\em The geometry of Hamiltonian systems.} T. Ratiu, ed., MSRI Publ.\
3461: \textbf{20}, Springer 1991, pp.\ 33--51
3462:
3463: \bibitem{armamonc}
3464: J.M.\ Arms, J.E.\ Marsden, and V.\ Moncrief:
3465: {\pemphas Symmetry and bifurcation of moment mappings.}
3466: {\empha Comm. Math. Phys.} \textbf{78} (1981), 455--478
3467:
3468: \bibitem{armamotw}
3469: J.M.\ Arms, J.E.\ Marsden, and V. Moncrief:
3470: The structure of the space of solutions of Einstein's equations. II.
3471: Several Killing fields and the Einstein-Yang-Mills equations.
3472: Ann.\ Phys.\ \textbf{144} (1), 81--106 (1982)
3473:
3474: \bibitem{Arscott}
3475: F.M.\ Arscott:
3476: {\em Periodic Differential Equations. An Introduction to Mathieu, Lam\'e, and
3477: Allied Functions.}
3478: Pergamon Press 1964
3479:
3480: \bibitem{asfalolu}
3481: M.\ Asorey, F.\ Falceto, J.L.\ L\'opez, G.\ Luz\'on:
3482: Nodes, monopoles and confinement in $(2+1)$-dimensional gauge theories.
3483: Phys.\ Lett.\ B {\bf 349}, 125--130 (1995)
3484:
3485: \bibitem{BakerBlackburnSmith}
3486: G.L.\ Baker, J.A.\ Blackburn, H.J.T.\ Smith:
3487: The quantum pendulum: Small and large.
3488: Amer.\ J.\ Phys.\ {\bf 70}, 525--531 (2002)
3489:
3490: \bibitem{cfg}
3491: S.\ Charzy\'nski, J.\ Kijowski, G.\ Rudolph, M.\ Schmidt:
3492: On the stratified classical configuration space of lattice QCD.
3493: J.\ Geom.\ Phys.\ {\bf 55}, 137--178 (2005)
3494:
3495: \bibitem{cfgtop}
3496: S.\ Charzy\'nski, G.\ Rudolph, M.\ Schmidt:
3497: On the topological structure of the stratified classical configuration space of
3498: lattice QCD.
3499: {\tt hep-th/0512129}
3500:
3501: \bibitem{Condon}
3502: E.U.\ Condon:
3503: The physical pendulum in quantum mechanics.
3504: Phys.\ Rev.\ {\bf 31}, 891--894 (1928)
3505:
3506: \bibitem{CuBa}
3507: R.H.\ Cushman and L.M.\ Bates:
3508: {\em Global Aspects of Classical Integrable Systems.}
3509: Birkh\"auser 1997
3510:
3511: \bibitem{DeserJackiw}
3512: S.\ Deser, R.\ Jackiw:
3513: Classical and quantum scattering on a cone.
3514: Commun.\ Math.\ Phys.\ {\bf 118}, 495--509 (1988)
3515:
3516: \bibitem{Dimock}
3517: J.\ Dimock:
3518: Canonical quantization of Yang-Mills on a circle.
3519: Rev.\ Math.\ Phys.\ {\bf 8}, 85--102 (1996)
3520:
3521: \bibitem{DriverHall}
3522: B.K.\ Driver, B.C.\ Hall:
3523: Yang-Mills theory and the Segal-Bargmann transform.
3524: Comm.\ Math.\ Phys.\ {\bf 201} (1999) 249--290
3525:
3526: \bibitem{emmroeme}
3527: C.\ Emmrich and H.\ Roemer:
3528: Orbifolds as configuration spaces of systems with gauge symmetries.
3529: Commun.\ Math.\ Phys.\ {\bf 129}, 69--94 (1990)
3530:
3531: \bibitem{cdy}
3532: E.\ Fischer, G.\ Rudolph, M.\ Schmidt:
3533: A lattice gauge model of singular Marsden-Weinstein reduction. Part I.
3534: Kinematics.
3535: J.\ Geom.\ Phys.\ {\bf 57}, 1193--1213 (2007)
3536:
3537: \bibitem{Florentino}
3538: C.A.\ Florentino, J.M.\ Mour\~ao, J.\ Nunes:
3539: Coherent state transforms and vector bundles on elliptic curves.
3540: J.\ Funct.\ Anal.\ {\bf 204} (2003) 355--398
3541:
3542: \bibitem{Goresky}
3543: M.\ Goresky, R.\ MacPherson:
3544: {\em Stratified Morse theory.}
3545: Springer 1988
3546:
3547: \bibitem{Hall94}
3548: B.C.\ Hall:
3549: The Segal-Bargmann "coherent state" transform for compact Lie groups.
3550: J.\ Funct.\ Anal.\ {\bf 122} (1994) 103--151
3551:
3552: \bibitem{Hall97}
3553: B.C.\ Hall:
3554: The inverse Segal-Bargmann transform for compact Lie groups.
3555: J.\ Funct.\ Anal.\ {\bf 143} (1997) 98--116
3556:
3557: \bibitem{bhallfou}
3558: B.C.\ Hall:
3559: Phase space bounds for quantum mechanics on a compact Lie group.
3560: Commun.\ Math.\ Phys.\ \textbf{184}, 233--250 (1997)
3561:
3562: \bibitem{Hall:Coherent}
3563: B.C.\ Hall:
3564: Coherent states and the quantization of 1+1-dimensional Yang-Mills theory.
3565: Rev.\ Math.\ Phys.\ {\bf 13}, 1281--1306 (2001),
3566: {\tt quant-ph/0012050}
3567:
3568: \bibitem{bhallone}
3569: B.C.\ Hall:
3570: Geometric quantization and the generalized Segal-Bargmann transform
3571: for Lie groups of compact type.
3572: Commun.\ Math.\ Phys.\ \textbf{226}, 233--268 (2002), {\tt quant-ph/0012105}
3573:
3574: \bibitem{HallMitchell}
3575: B.C.\ Hall, J.J.\ Mitchell:
3576: The Segal-Bargmann transform for noncompact symmetric spaces of the complex
3577: type.
3578: J.\ Funct.\ Anal.\ {\bf 227} (2005) 338--371
3579:
3580: \bibitem{helgaboo}
3581: S.\ Helgason:
3582: {\em Groups and geometric analysis. Integral geometry, invariant differential
3583: operators, and spherical functions.}
3584: Academic Press 1984
3585:
3586: \bibitem{Hetrick}
3587: J.E.\ Hetrick:
3588: Canonical quantization of two-dimensional gauge fields.
3589: Int.\ J.\ Mod.\ Phys.\ A {\bf 9}, 3153--3178 (1994)
3590:
3591: \bibitem{locpois}
3592: J.\ Huebschmann:
3593: Pois\-son ge\-o\-me\-try of flat connections for {\rm SU(2)}-bundles on
3594: sur\-fa\-ces.
3595: Math.\ Z.\ \textbf{221}, 243--259 (1996), {\tt hep-th/9312113}
3596:
3597: \bibitem{modus}
3598: J.\ Huebschmann:
3599: Symplectic and Poisson structures of certain moduli spaces. Duke
3600: Math. J. \textbf{80}, 737--756 (1995), {\tt hep-th/9312112}
3601:
3602: \bibitem{kaehler}
3603: J.\ Huebschmann:
3604: K\"ahler spaces, nilpotent orbits, and singular reduction.
3605: Memoirs of the AMS \textbf{172 (814)}, Amer. Math. Soc., Providence R.I., ,
3606: 2004, {\tt math.DG/0104213}
3607:
3608: \bibitem{qr}
3609: J.\ Huebschmann:
3610: K\"ahler quantization and reduction.
3611: J.\ reine angew.\ Math.\ \textbf{591}, 75--109 (2006),
3612: {\tt math.SG/0207166}
3613:
3614: \bibitem{adjoint}
3615: J.\ Huebschmann:
3616: Stratified K\"ahler structures on adjoint quotients.
3617: Differential Geometry and its Applications (to appear),
3618: {\tt math.DG/0404141}
3619:
3620: \bibitem{bedlewo}
3621: J.\ Huebschmann:
3622: Singular Poisson-K\"ahler geometry of certain adjoint quotients,
3623: in: Proceedings, The mathematical legacy of C. Ehresmann, Bedlewo,
3624: 2005, Banach Center Publications 76 (2007) 325--347,
3625: {\tt math.SG/0610614}
3626:
3627: \bibitem{holopewe}
3628: J.\ Huebschmann:
3629: Kirillov's character formula, the holomorphic Peter-Weyl theorem, and the
3630: Blattner-Kostant-Sternberg pairing.
3631: J.\ Geom.\ Phys.\ 58 (2008), 833-848; {\tt math.DG/0610613}
3632:
3633: \bibitem{qcd1}
3634: P.D.\ Jarvis, J.\ Kijowski, G.\ Rudolph:
3635: On the structure of the observable algebra of QCD on the lattice.
3636: J.\ Phys.\ A {\bf 38}, 5359--5377 (2005)
3637:
3638: \bibitem{KayStuder}
3639: B.S.\ Kay, U.M.\ Studer:
3640: Boundary conditions for quantum mechanics on cones and fields around cosmic
3641: strings.
3642: Commun.\ Math.\ Phys.\ {\bf 139}, 103--139 (1991)
3643:
3644: \bibitem{qcd2}
3645: J.\ Kijowski, G.\ Rudolph:
3646: On the Gauss law and global charge for quantum chromodynamics.
3647: J.\ Math.\ Phys.\ {\bf 43}, 1796--1808 (2002)
3648:
3649: \bibitem{qcd3}
3650: J.\ Kijowski, G.\ Rudolph:
3651: Charge superselection sectors for qcd on the lattice.
3652: J.\ Math.\ Phys.\ {\bf 46}, 032303 (2005)
3653:
3654: \bibitem{qed1}
3655: J.\ Kijowski, G.\ Rudolph, C.\ \'Sliwa:
3656: On the structure of the observable algebra for QED on the lattice.
3657: Lett.\ Math.\ Phys.\ {\bf 43}, 299--308 (1998);~
3658:
3659: \bibitem{qed2}
3660: J.\ Kijowski, G.\ Rudolph, C.\ \'Sliwa:
3661: Charge superselection sectors for scalar QED on the lattice.
3662: Ann.\ Henri Poin\-car\'e {\bf 4}, 1137--1167 (2003)
3663:
3664: \bibitem{qed3}
3665: J.\ Kijowski, G.\ Rudolph, A.\ Thielmann:
3666: Algebra of observables and charge superselection sectors for QED on the
3667: lattice.
3668: Commun.\ Math.\ Phys.\ {\bf 188}, 535--564 (1997)
3669:
3670: \bibitem{Landsman}
3671: N.P.\ Landsman:
3672: {\em Mathematical topics between classical and quantum mechanics}.
3673: Springer 1998
3674:
3675: \bibitem{LandsmanWren:Theta}
3676: N.P.\ Landsman, K.K.\ Wren:
3677: Constrained quantization and $\theta$-angles.
3678: Nucl.\ Phys.\ B {\bf 502}, 537--560 (1997)
3679:
3680: \bibitem{LandsmanWren:Coherent}
3681: N.P.\ Landsman, K.K.\ Wren:
3682: Hall's coherent states, the Cameron-Martin theorem, and the quantization of
3683: Yang-Mills theory on a circle.
3684: {\tt math-ph/9812012}
3685:
3686: \bibitem{McLachlan}
3687: N.W.\ McLachlan:
3688: {\em Theory and Application of Mathieu Functions.}
3689: Dover Publications 1964
3690:
3691: \bibitem{MeixnerSchaefke}
3692: J.\ Meixner, W.\ Schaefke:
3693: Mathieusche Funktionen und Sph\"aroidfunktionen.
3694: Grundlehren Bd.\ 71, Springer 1954
3695:
3696: \bibitem{enelson}
3697: E.\ Nelson:
3698: Analytic vectors.
3699: Ann.\ of Math.\ \textbf{70}, 572--615 (1959)
3700:
3701: \bibitem{PradhanKhare}
3702: T.\ Pradhan, A.V.\ Khare:
3703: Plane pendulum in quantum mechanics.
3704: Amer.\ J.\ Phys.\ {\bf 41}, 59--66 (1973)
3705:
3706: \bibitem{Schwarz}
3707: G.W.\ Schwarz:
3708: Smooth functions invariant under the action of a compact Lie group.
3709: Topology {\bf 14}, 63--68 (1975)
3710:
3711: \bibitem{Sniatycki}
3712: J.\ \' Sniatycki:
3713: {\em Geometric quantization and quantum mechanics.}
3714: Applied Mathematical Sciences 30, Springer 1980
3715:
3716: \bibitem{Stein}
3717: E.M.\ Stein:
3718: {\it Topics in harmonic analysis related to the Littlewood-Paley theory.}
3719: Annals of Mathematics Studies, No.\ 63. Princeton University Press 1970
3720:
3721: \bibitem{taylothr}
3722: J.\ Taylor:
3723: The Iwasawa decomposition and limiting behaviour of Brownian motion
3724: on symmetric spaces of non-compact type.
3725: Cont.\ Math.\ \textbf{73}, 303--331 (1988)
3726:
3727: \bibitem{Thiemann}
3728: T.\ Thiemann:
3729: Gauge field theory coherent states (GCS). I. General properties.
3730: Classical Quantum Gravity {\bf 18} (2001) 2025--2064
3731:
3732: \bibitem{ThiemannWinkler}
3733: T.\ Thiemann, O.\ Winkler:
3734: Gauge field theory coherent states (GCS). II. Peakedness properties.
3735: Classical Quantum Gravity {\bf 18} (2001) 2561--2636
3736:
3737: \bibitem{woodhous}
3738: N.M.J.\ Woodhouse:
3739: {\em Geometric quantization.}
3740: Clarendon Press, Oxford 1991
3741:
3742: \bibitem{Wren:Rieffel} K.K.\ Wren:
3743: Quantization of constrained systems with singularities using Rieffel
3744: induction.
3745: J.\ Geom.\ Phys.\ {\bf 24}, 173--202 (1998)
3746:
3747: \bibitem{Wren:ThetaII}
3748: K.K.\ Wren:
3749: Constrained quantisation and $\theta$-angles. II.
3750: Nucl.\ Phys.\ {\bf B 521}, 471--502 (1998)
3751:
3752: }
3753:
3754: \end{thebibliography}
3755:
3756: \end{document}
3757:
3758:
3759: