hep-th0602169/46.tex
1: \documentclass[11pt]{article}
2: \usepackage{amssymb,latexsym,amsmath}
3: \usepackage[dvips]{graphicx}
4: %\usepackage{showlabels}
5: %\pagestyle{empty}
6: \headheight=0mm
7: \headsep=0mm
8: \oddsidemargin=-5mm
9: \evensidemargin=-5mm
10: \textheight=235mm
11: \textwidth=165mm
12: %
13: % definitions concerning automatic numbering of definitions, etc. %
14: \newtheorem{theo}{Theorem}
15: \newtheorem{defi}[theo]{Definition}
16: \newtheorem{lemm}[theo]{Lemma}
17: \newtheorem{coro}[theo]{Corollary}
18: \newtheorem{rema}[theo]{Remark}
19: \newtheorem{prop}[theo]{Proposition}
20: %
21: % a few of my own definitions
22: %
23: \def\ds{\displaystyle}
24: \def\nn{\nonumber}
25: \def\deg{\mathop{\rm deg}\nolimits}
26: \def\str{\mathop{\rm str}\nolimits}
27: \def\qdots{\mathinner{\mkern1mu\raise1pt\vbox{\kern7pt\hbox{.}}\mkern2mu
28:  \raise4pt\hbox{.}\mkern2mu\raise7pt\hbox{.}\mkern1mu}}
29: \def\Z{{\mathbb Z}}
30: \def\N{{\mathbb N}}
31: \def\C{{\mathbb C}}
32: \def\gl{\mathfrak{gl}}
33: \def\ssl{\mathfrak{sl}}
34: \def\h{\mathfrak{h}}
35: \def\so{\mathfrak{so}}
36: \def\osp{\mathfrak{osp}}
37: %
38: \def\a{\alpha}
39: \def\b{\beta}
40: \def\J{{\bf J}}
41: \def\hR{\hat{R}}
42: \def\hP{\hat{P}}
43: \def\hJ{\hat{J}}
44: \def\hbR{\hat{\bf R}}
45: \def\hbP{\hat{\bf P}}
46: \def\lb{[\![}
47: \def\rb{]\!]}
48: %
49: 
50: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
51: \setcounter{footnote}{1}
52: %
53: % the end-of-proof box
54: %
55: \def\mybox{\hfill$\Box$}
56: %
57: % numbering of equations per section:
58: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
59: %
60: 
61: \begin{document}
62: \begin{center}
63: {\Large \bf
64: Representations of the Lie Superalgebra $\gl(1|n)$ in a \\[2mm]
65: Gel'fand-Zetlin Basis and Wigner Quantum Oscillators}\\[5mm]
66: {\bf R.C.~King}\footnote{E-mail: R.C.King@soton.ac.uk}\\[1mm]
67: School of Mathematics, University of Southampton,\\
68: Southampton SO17 1BJ, U.K.;\\[2mm]
69: {\bf N.I.~Stoilova}~\footnote{E-mail: Neli.Stoilova@UGent.be; Permanent address:
70: Institute for Nuclear Research and Nuclear Energy, Boul.\ Tsarigradsko Chaussee 72,
71: 1784 Sofia, Bulgaria} {\bf and J.\ Van der Jeugt}\footnote{E-mail:
72: Joris.VanderJeugt@UGent.be}\\[1mm]
73: Department of Applied Mathematics and Computer Science,
74: Ghent University,\\
75: Krijgslaan 281-S9, B-9000 Gent, Belgium.
76: \end{center}
77: 
78: 
79: \vskip 10mm
80: 
81: %\addtolength{\baselineskip}{3mm}
82: %\addtolength{\abovedisplayskip}{1mm}
83: %\addtolength{\belowdisplayskip}{1mm}
84: %\addtolength{\parskip}{2mm}
85: 
86: \begin{abstract}
87: An explicit construction of all finite-dimensional irreducible representations of 
88: the Lie superalgebra $\gl(1|n)$ in a Gel'fand-Zetlin basis is given. 
89: Particular attention is paid to the so-called star type~I representations (``unitary representations''),
90: and to a simple class of representations $V(p)$, with $p$ any positive integer.
91: Then, the notion of Wigner Quantum Oscillators (WQOs) is recalled.
92: In these quantum oscillator models, the unitary representations of $\gl(1|DN)$ are 
93: physical state spaces of the $N$-particle $D$-dimensional oscillator. 
94: So far, physical properties of $\gl(1|DN)$ WQOs were described only in the so-called Fock
95: spaces $W(p)$, leading to interesting concepts such as non-commutative coordinates and
96: a discrete spatial structure. 
97: %However, a classical limit of these solutions was lacking.
98: Here, we describe physical properties of WQOs for other unitary representations,
99: including certain representations $V(p)$ of $gl(1|DN)$.
100: These new solutions again have  remarkable properties following from the spectrum
101: of the Hamiltonian and of the position, momentum, and angular momentum operators.
102: Formulae are obtained that give the angular
103: momentum content of all the representations V(p)
104: of $\gl(1|3N)$, associated with the $N$-particle 3-dimensional WQO.
105: For these representations $V(p)$ we also consider in more detail the
106: spectrum of the position operators and their squares, leading to interesting consequences.
107: In particular, a classical limit of these solutions is obtained, that is in agreement
108: with the correspondence principle.
109: \end{abstract}
110: 
111: 
112: 
113: \vfill\eject
114: 
115: %\newpage
116: %
117: %\renewcommand{\thesection}{\Roman{section}}
118: %\renewcommand{\theequation}{\arabic{section}.{\arabic{equation}}}
119: %\def\I{\uppercase\expandafter{\romannumeral 1}}
120: 
121: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
122: %%%%% section
123: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
124: \setcounter{equation}{0}
125: \section{Introduction} \label{sec:Introduction}%
126: 
127: In this paper we construct all finite-dimensional irreducible representations 
128: of the general linear Lie superalgebra $\gl(1|n)$~\cite{Kac1,Kac2}. 
129: To this end we introduce a Gel'fand-Zetlin basis (GZ-basis)
130: for $\gl(1|n)$ and write down explicit expressions for the transformation of 
131: the basis vectors under the action of the algebra generators. 
132: This is analagous to the GZ-basis for $\gl(n|1)$ and the action of its generators 
133: given by Palev~\cite{Palev2}.
134: The Lie superalgebra $\gl(1|n)$ is a central extension of the special linear 
135: Lie superalgebra $\ssl(1|n)$. Each finite-dimensional irreducible 
136: $\gl(1|n)$ representation remains irreducible when restricted to $\ssl(1|n)$, and
137: each of the resulting finite-dimensional irreducible representations of $\ssl(1|n)$ 
138: is either typical or atypical~\cite{Kac1,Kac2}. 
139: 
140: The motivation for the present work stems from the fact that a set of generating 
141: elements of $\ssl(1|n)$ satisfies the compatibility conditions of the so called 
142: Wigner Quantum Oscillator (WQO)~\cite{Wigner}-\cite{SJ}. 
143: Therefore to investigate the properties of these 
144: $\ssl(1|n)$ WQOs we must consider those representations that are of physical relevance, 
145: namely the star representations of type I~\cite{Gould}. These include certain 
146: representations $W(p)$~\cite{Palev3} having highest weight $\Lambda=(p;0,0,\ldots,0)$,
147: that have a Fock space realisation. 
148: A study of these in the case $n=3$ and $n=3N$ has revealed
149: WQO models that have a finite, equally spaced energy spectrum, with discrete
150: values of measurements of both spatial coordinates and linear momenta, as well
151: as angular momenta~\cite{Palev1, Palev5, K1,K2}. Moreover, the underlying geometry of the model is non-commutative~\cite{HS, C},
152: in the sense that coordinate operator components in general do not commute;
153: thus measurements of two different coordinates, say $x$ and $y$, 
154: of a single particle cannot be performed simultaneously. 
155: As such, the WQO models are to be
156: compared with other non-commutative models, see for example~\cite{NP}-\cite{MM}.
157: 
158: Here the intention is to set up the mathematical machinery enabling these 
159: properties to be explored for WQO models based on any $\ssl(1|n)$ star representation of type~I.
160: The results confirm that in all such cases there is a finite number of equally
161: spaced energy levels, that spatial coordinates and linear momenta are
162: discretely quantised, and that the model always exhibits non-commutative
163: geometry. A detailed analysis is presented for those WQO models based on
164: the $\ssl(1|n)$ representations $V(p)$ having highest weight $\Lambda=(1;p-1,0,\ldots,0)$,
165: and also for a representation of highest weight $\Lambda=(2;1,0,\ldots0)$
166: that belongs to neither class $W(p)$ nor $V(p)$.
167: 
168: 
169: The structure of the paper is as follows. In Section~2 we construct all the 
170: finite-dimensional irreducible representations of the Lie superalgebra $\gl(1|n)$, complete
171: with a specification of the GZ basis vectors $|m)$ and the explicit action of a set of 
172: $\gl(1|n)$ generators on these vectors. The section culminates with formulae
173: for the action on an arbitrary vector $|m)$ of all the odd elements, $e_{0j}$ 
174: and $e_{j0}$, of $\gl(1|n)$. These are of importance in Section~4, where 
175: for $n=DN$ these odd elements play the role of certain linear 
176: combinations of position  and linear 
177: momentum operators for the $N$-particle $D$-dimensional WQO. 
178: Before this, in Section~3, the formulae of Section~2 are applied to the case of 
179: the $\gl(1|n)$ representations $V(p)$ of highest weight $\Lambda=(1;p-1,0,\ldots,0)$.
180: The GZ-basis and the action of the $\gl(1|n)$ generators are given explicitly in 
181: a rather more succinct notation than was possible in the general case.
182: 
183: Returning to Section~4, it is here that the WQO is introduced. The Hamiltonian
184: $\hat{H}$ of an $N$-particle $D$-dimensional harmonic oscillator takes the form
185: \begin{equation}
186: \hat{H}=\sum_{\alpha=1}^{N} \Big( \frac{\hbP_\alpha^2}{2m}
187: + \frac{m\omega^2}{2} {\hbR}_\a^2 \Big), \label{Intro-H}
188: \end{equation}
189: with $\hbP_\alpha$ and $\hbR_\alpha$ $D$-dimensional vector operators
190: corresponding to the momentum and position of the particle~$\alpha$, each
191: of the same mass $m$ and natural frequency $\omega$. This 
192: Hamiltonian can be re-expressed in the form
193: \begin{equation}
194:      \hat{H}= \frac{\hbar\omega}{DN-1} \sum_{j=1}^{DN} \{A_j^+, A_j^-\},
195:      \label{Intro-HAA}
196: \end{equation}
197: where the operators $A_j^\pm$ are simple linear
198: combinations
199: of $\hat{R}_{\a k}$ and $\hat{P}_{\a k}$ for some $\a$ and $k$ determined by
200: $j$.
201: With this notation, 
202: the WQO requirement that Hamilton's equations and the Heisenberg equations 
203: coincide as operator equations leads to compatibility conditions
204: on the operators $A_j^+$ and $A_j^-$ that have a non-canonical solution
205: allowing them to be identifed, as stated above, with the odd generators,
206: $e_{j0}$ and $e_{0j}$, respectively, of $\gl(1|n)$. This identification
207: is then exploited to determine the physical properties of WQO models,
208: including their energy spectrum and the eigenvalues of their spatial 
209: coordinate operators, as well as their non-vanishing
210: commutator. The determination of the angular momentum content 
211: of the particular $\gl(1|3N)$ modules $V(p)$ is deferred to Section~5, 
212: where it is achieved for all $N$ and all $p$, by means of a generating 
213: function derived from Molien's Theorem~\cite{Molien, Sturmfels} applied to the case
214: of the embedding of the rotation group $SO(3)$, with Lie algebra $\so(3)$
215: in the group $GL(3N)$, with Lie algebra $\gl(3N)$. The outcome in 
216: the cases $N=1$ and $N=2$ is given in detail.
217: 
218: Section~6 is concerned with an understanding of the position
219: operator spectrum. To this end, its crucial features are
220: illustrated in the case $n=2$, both for a 
221: single particle  two-dimensional ($N=1, D=2$)
222: model based on the $\gl(1|2)$ representations $V(p)$, and for a two-particle 
223: one dimensional ($N=2,D=1$) model 
224: based on the same representations. In both cases a classical limit is recovered by 
225: taking $p\rightarrow\infty$ and $\hbar\rightarrow0$ in such a way that
226: $p\hbar\rightarrow C$, for some constant $C$. 
227: 
228: A final example of a WQO model, based on a representation belonging to 
229: neither the set $W(p)$ nor the set $V(p)$, is given for illustrative
230: purposes in Section~7. This time the results of Section~2,
231: that immediately give the allowed energy levels and spatial
232: coordinates, are augmented by the use of the subalgebra chain
233: $\gl(3N)\rightarrow \gl(3) \oplus \gl(N)\rightarrow \so(3) \oplus \gl(N)
234: \rightarrow \so(3)$ to obtain a complete description of
235: the angular momentum content.
236: 
237: We close with a few brief concluding remarks in Section~8.
238: 
239: Following customary usage in the mathematical physics literature we use in this paper the words
240: ``module'' and ``representation'' more or less interchangeably, and often refer to
241: ``simple module'' as ``irreducible representation.''
242: 
243: 
244: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
245: %%%%% section
246: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
247: 
248: 
249: \setcounter{equation}{0}
250: \section{The $\gl(1|n)$ representations} \label{sec:A}%
251: 
252: The Lie superalgebra $\gl(1|n)$ can be defined as the set of all squared 
253: $(n+1)$-dimensional matrices with rows and columns labelled by indices $i,j=0,1,\ldots,n$. 
254: As a basis in $\gl(1|n)$ we choose the Weyl matrices $e_{ij}, \; i,j=0,1,\ldots,n$, where 
255: the odd elements are $\{e_{i0}, e_{0i} | i=1,\ldots,n\}$,
256: and the remaining elements are even.
257: The Lie superalgebra bracket is determined by
258: \begin{equation}
259: \lb e_{ij}, e_{kl} \rb \equiv e_{ij} e_{kl}-(-1)^{{\deg(e_{ij})
260: \deg(e_{kl})}}e_{kl}e_{ij} = \delta_{jk} e_{il} - (-1)^{\deg(e_{ij}) \deg(e_{kl})}
261: \delta_{il} e_{kj}. \label{Weyl}
262: \end{equation}
263: 
264: Note that $\gl(1|n)_0 = \gl(1) \oplus \gl(n)$, where $\gl(1) =\hbox{span}\{ e_{00}\}$ and
265: $\gl(n)=\hbox{span}\{ e_{ij}|i,j=1,\ldots,n\}$. For elements $x$ of
266: $\gl(1|n)$, one defines the supertrace
267: as $\str(x)=x_{00}-\sum_{j=1}^nx_{jj}$. The Lie superalgebra $\gl(1|n)$ is
268: not simple, and one can define the simple superalgebra $\ssl(1|n)$
269: as the subalgebra consisting of elements with supertrace $0$. However,
270: the representation theory of $\gl(1|n)$ or $\ssl(1|n)$ is essentially
271: the same (the situation is similar as for the classical Lie
272: algebras $\gl(n)$ and $\ssl(n)$), and hence we prefer to work with
273: $\gl(1|n).$ As an ordered basis in the Cartan subalgebra $\h$ of $\gl(1|n)$, we choose
274: $e_{00}, e_{11}, \ldots, e_{nn}$, and denote by 
275: \begin{equation}
276: \epsilon,\delta_1,\ldots,\delta_n \label{ep-de}
277: \end{equation}
278: the dual basis in the space $\h^*$ of all linear functionals of $\h$.
279: 
280: The finite-dimensional simple modules of $\gl(1|n)$ are characterized by
281: their highest weight $\Lambda$. Often, one writes
282: \begin{equation}
283: \Lambda = m_{0,n+1}\, \epsilon + \sum_{i=1}^n m_{i,n+1}\, \delta_i .
284: \end{equation}
285: Then, the finite-dimensional simple modules $W([m]_{n+1})$ of the Lie 
286: superalgebra $\gl(1|n)$ are in one-to-one correspondence with the set 
287: of all complex $n+1$ tuples~\cite{Kac1,Kac2}
288: \begin{equation}
289: [m]_{n+1}=[m_{0,n+1}, m_{1,n+1}, \ldots , m_{n,n+1}],  \label{mn+1}
290: \end{equation}
291: for which
292: \begin{equation}
293: m_{i,n+1}-m_{j,n+1}\in \Z_+, \; \forall i\leq j =1,\ldots, n. \label{cond}
294: \end{equation}
295: Within a given $\gl(1|n)$ module $W([m]_{n+1})$ the numbers~(\ref{mn+1}) are fixed.
296: The possibility of introducing a Gel'fand-Zetlin basis in any finite-dimensional simple 
297: $\gl(1|n)$ module $W([m]_{n+1})$ stems from the following proposition.
298: \begin{prop}
299: Consider the $\gl(1|n)$ module $W([m]_{n+1})$ as a $\gl(n)$ module.
300: Then $W([m]_{n+1})$ can be represented as a direct sum of simple $\gl(n)$ modules,
301: \begin{equation}
302: W([m]_{n+1})=\sum_i \oplus V_i([m]_n), \label{gl(n)}
303: \end{equation}
304: where 
305: \begin{itemize}
306: \item[I.] All $V_i([m]_n)$ carry inequivalent representations of $\gl(n)$
307: \begin{equation}
308: [m]_{n}=[m_{1n}, m_{2n},\ldots , m_{nn}], \; m_{in}-m_{i+1,n}\in\Z_+  .\label{n}
309: \end{equation}
310: \item[II.] 
311: \begin{equation}
312:  \begin{array}{rl}
313: 1.& m_{in}-m_{i,n+1}=\theta_{i}\in\{0,1\},\quad 1\leq i\leq n,\\
314: 2.& \hbox{if for }\; k\in\{1,\ldots,n\}\; \;
315: m_{0,n+1}+m_{k,n+1}=k-1, \hbox{ then}\; \theta_k=0.
316:  \end{array}
317: \label{cond0}
318: \end{equation}
319: \end{itemize}
320: \end{prop}
321: The decomposition~(\ref{gl(n)}) and conditions~(\ref{cond0}) follow from the character formula for simple
322: $\gl(1|n)$ modules~\cite{LJ, JHKR}. If for some $k\in\{1,\ldots,n\}$ the condition $m_{0,n+1}+m_{k,n+1}=k-1$
323: is satisfied, then the representation is {\em atypical of type~$k$}. Otherwise, it is typical.
324: 
325: A GZ-basis for the $\gl(n)$ module is well known~\cite{GZ}. A set of basis vectors is given 
326: by a triangular array of
327: numbers satisfying an integral property and the ``betweenness conditions'':
328: \begin{equation}
329:  \left|
330: \begin{array}{lcllll}
331:  m_{1n} & \cdots & \cdots & m_{n-1,n} & m_{nn}  \\
332:  m_{1,n-1} & \cdots & \cdots &  m_{n-1,n-1}  &  \\
333: \vdots & \qdots & & & \\
334: m_{11} & & & &
335: \end{array}
336: \right).
337: \label{gln}
338: \end{equation}
339: These ``betweenness conditions'' take the form
340: \begin{equation}
341: m_{i,j+1}-m_{ij}\in\Z_+\hbox{ and }\; m_{i,j}-m_{i+1,j+1}\in\Z_+,\quad
342:     1\leq i\leq j\leq n-1 ,
343: \label{inbetween}
344: \end{equation}
345: and they automatically imply the integral property $m_{i,j}-m_{i+1,j}\in\Z_+$ for all meaningful
346: values of the indices.
347: 
348: Using this GZ-basis for simple $\gl(n)$ modules and Proposition~1 we have 
349:  
350: \begin{prop}
351: The set of vectors 
352: \begin{equation}
353: |m) = \left|
354: \begin{array}{lcllll}
355: m_{0,n+1} & m_{1,n+1}& \cdots & m_{n-2,n+1} & m_{n-1,n+1} & m_{n,n+1}  \\
356: & m_{1n} & \cdots & \cdots & m_{n-1,n} & m_{nn}  \\
357: & m_{1,n-1} & \cdots & \cdots &  m_{n-1,n-1}  &  \\
358: &\vdots & \qdots & & & \\
359: &m_{11} & & & &
360: \end{array}
361: \right)
362: \label{m}
363: \end{equation}
364: satisfying the conditions
365: \begin{equation}
366:  \begin{array}{rl}
367: 1. & m_{i,n+1} \hbox{are fixed and } m_{i,n+1}-m_{j,n+1}\in\Z_+ \quad
368:     1\leq i\leq j\leq n,\\
369: 2.& m_{in}-m_{i,n+1}=\theta_{i}\in\{0,1\},\quad 1\leq i\leq n,\\
370: 3.& \hbox{if for }\; k\in\{1,\ldots,n\}\; \;
371: m_{0,n+1}+m_{k,n+1}=k-1, \hbox{then}\; \theta_k=0,  \\
372: 4.& m_{i,j+1}-m_{ij}\in\Z_+\hbox{ and }\; m_{i,j}-m_{i+1,j+1}\in\Z_+,\quad
373:     1\leq i\leq j\leq n-1.
374:  \end{array}
375: \label{cond1}
376: \end{equation}
377: constitute a basis in $W([m]_{n+1})$.
378: \end{prop}
379: We shall refer to the basis~(\ref{m}) as the GZ-basis for $\gl(1|n)$.
380: The purpose is now to give the explicit action of a set of $\gl(1|n)$ generators 
381: on the basis vectors~(\ref{m}). For this purpose,
382: denote by $|m)_{\pm ij}$ the pattern obtained from $|m)$ by the replacement
383: $m_{ij} \rightarrow m_{ij}\pm 1$. Then this action is given by:
384: \begin{eqnarray}
385: e_{00}|m)&=&\left(m_{0,n+1}-\sum_{j=1}^n \theta_{j}\right)|m); \label{e_00}\\
386: e_{kk}|m)&=&\left(\sum_{j=1}^k m_{jk}-\sum_{j=1}^{k-1} m_{j,k-1}\right)|m), 
387: \quad (1\leq k\leq n); \label{e_kk}\\
388: e_{k-1,k}|m)&=&\sum_{j=1}^{k-1} \left(-
389: \frac{\prod_{i=1}^{k} (l_{ik}-l_{j,k-1})
390: \prod_{i=1}^{k-2} (l_{i,k-2}-l_{j,k-1}-1)}{\prod_{i\neq j=1}^{k-1} (l_{i,k-1}-l_{j,k-1})
391: (l_{i,k-1}-l_{j,k-1}-1) } \right)^{1/2}|m)_{+j,k-1},\quad (2\leq k\leq n); \nn\\[-3mm]
392: && \label{ek}\\
393: e_{k,k-1}|m)&=&\sum_{j=1}^{k-1} \left(-
394: \frac{\prod_{i=1}^{k} (l_{ik}-l_{j,k-1}+1)
395: \prod_{i=1}^{k-2} (l_{i,k-2}-l_{j,k-1})}{\prod_{i\neq j=1}^{k-1} (l_{i,k-1}-l_{j,k-1})
396: (l_{i,k-1}-l_{j,k-1}+1) } \right)^{1/2}|m)_{-j,k-1},\quad (2\leq k\leq n); \nn\\[-3mm]
397: && \label{fk}\\
398: e_{0n}|m)&=&\sum_{i=1}^n \theta_{i}
399: (-1)^{\theta_{1}+ \ldots +\theta_{i-1} }(l_{i,n+1}+l_{0,n+1}+1)^{1/2}
400: \left( \frac{\prod_{k=1}^{n-1}  (l_{k,n-1}-l_{i,n+1}-1 )}{\prod_{k\neq i=1}^n (  l_{k,n+1}-l_{i,n+1})}
401: \right)^{1/2} |m)_{-in}; \nn \\
402: && \label{en}\\
403: e_{n0}|m)&=&\sum_{i=1}^n (1-\theta_{i})
404: (-1)^{\theta_{1}+ \ldots +\theta_{i-1} }(l_{i,n+1}+l_{0,n+1}+1)^{1/2}\nn\\
405:  &\times&\left(
406: \frac{\prod_{k=1}^{n-1} ( l_{k,n-1}-l_{i,n+1} -1)}{\prod_{k\neq i=1}^n (  l_{k,n+1}-l_{i,n+1} )}
407: \right)^{1/2} |m)_{+in}.\label{fn}
408: \end{eqnarray}
409: In all these formulas $l_{ij}=m_{ij}-i$.
410: 
411: In order to deduce~(\ref{e_00})-(\ref{fn}), 
412: we have used the paper of Palev~\cite{Palev2} and the isomorphism $\varphi$ of $\gl(n|1)$ onto $\gl(1|n)$:
413: \begin{equation}
414: \varphi(E_{n,n+1})=e_{0n},\;\; \varphi(E_{n+1,n})=e_{n0},\;\;
415: \varphi(E_{ij})=e_{ij}, \; i,j=1,\ldots, n,
416: \end{equation}
417: where $E_{ij}$ ($i,j=1,\ldots,n+1$) and $e_{ij}$ ($i,j=0,1,\ldots,n$) are the Weyl matrices
418: of $\gl(n|1)$ and $\gl(1|n)$ respectively.
419: As compared with the $\gl(n|1)$ case, note that in the right hand 
420: side of~(\ref{en})-(\ref{fn}) instead of $|m)_{\pm in}$
421: we have $|m)_{\mp in}$. This is because of the conditions~2 and~3 in Proposition~2. 
422: Also the atypicality factor $(l_{i,n+1}+l_{0,n+1}+1)$ is different.
423: 
424: For our purposes, it is also necessary to know which of these representations are ``unitary
425: representations'' with respect to some star condition (or Hermiticity condition) on the 
426: Lie superalgebra elements. The star condition of relevance is the antilinear anti-involutive
427: mapping determined by
428: \begin{equation}
429: e_{ij}^\dagger = e_{ji}.
430: \label{star-condition}
431: \end{equation}
432: The unitary representations are then those for which there exist a
433: positive definite inner product $\langle \;| \; \rangle$ (in the representation space $W$)
434: such that
435: \begin{equation}
436: \langle e_{ij}v|w\rangle = \langle v|e_{ji}w\rangle,
437: \end{equation}
438: for all $v,w\in W$. These are, in the terminology of~\cite{Gould}, the {\em star type I} representations.
439: They have been classified in general for $\gl(m|n)$, and for $\gl(1|n)$ the conclusion is
440: as follows:
441: \begin{prop}
442: The representation $W([m]_{n+1})$ is a unitary representation if and only if
443: \begin{itemize}
444: \item[(a)] The highest weight is real and 
445: \begin{equation}
446:   m_{0,n+1}+m_{n,n+1}-n+1>0.
447: \end{equation}
448: In this case, the representation is typical.
449: \item[(b)] The highest weight is real and there exists a $k\in \{ 1,2,\ldots, n\}$ such that
450: \begin{equation}
451:   m_{0,n+1}+m_{k,n+1}=k-1,\quad m_{k,n+1}=m_{k+1,n+1}=\cdots =m_{n,n+1}.
452: \end{equation}
453: In this case, the representation is atypical of type~$k$.
454: \end{itemize}
455: \end{prop}
456: For those familiar with the terminology, note that the representations under (b)
457: essentially correspond to the covariant and contravariant representations of 
458: $\gl(1|n)$~\cite{JHKT}.
459: They could also by labelled by $m_{0,n+1}$ and a partition $\lambda$.
460: 
461: The representations $W([m]_{n+1})$ are unitary if the conditions of 
462: Proposition~3 are satisfied, for the inner product corresponding to
463: orthonormal GZ basis vectors, i.e.
464: \begin{equation}
465: \langle |m')\,|\; |m) \rangle \equiv (m'|m) = \delta_{m,m'},
466: \end{equation}
467: under the action~(\ref{e_00})-(\ref{fn}).
468: 
469: We end this section with a technical result. 
470: In principle, it is sufficient to have the action of the $\gl(1|n)$ generators
471: $e_{kk}$ ($k=0,\ldots,n$), $e_{k-1,k}$, $e_{k,k-1}$ ($k=1,\ldots,n$), and
472: $e_{0n}$, $e_{n0}$ as given in~(\ref{e_00})-(\ref{fn}), in order to compute the action of
473: any element $e_{ij}$ on the GZ basis vectors~(\ref{m}). 
474: For our later application, it is useful to know the explicit action of the odd elements
475: $e_{0j}$ and $e_{j0}$ of $\gl(1|n)$. It is found that this is given by the
476: following complicated formulae:
477: \begin{eqnarray}
478: e_{0j}|m)&=&\sum_{i_n=1}^n\sum_{i_{n-1}=1}^{n-1}\ldots \sum_{i_j=1}^j
479: \theta_{i_n}(-1)^{\theta_1+\ldots +\theta_{i_n-1}}(l_{i_n,n+1}+l_{0,n+1}+1)^{1/2}\nn\\
480: &\times&
481: \prod_{r=j+1}^nS(i_r,i_{r-1})
482: \left( \frac{\prod_{k\neq i_{r-1}=1}^{r-1}  
483: (l_{k,r-1}-l_{i_r,r} )\prod_{k\neq i_r=1}^r(l_{kr}-l_{i_{r-1},r-1}+1)
484:  }{ \prod_{k\neq i_r=1}^r (l_{kr}-l_{i_r,r})\prod_{k\neq i_{r-1}=1}^{r-1}(l_{k,r-1}
485:  -l_{i_{r-1},r-1}+1)}
486: \right)^{1/2} \label{e0j}\\
487: &\times &
488: \left({\prod_{k\neq i_n=1}^n } \frac{(l_{kn}-l_{i_n,n} )}{(  l_{k,n+1}-l_{i_n,n+1})}
489: \right)^{1/2}
490: \left( \frac{\prod_{k=1}^{j-1} (l_{k,j-1}-l_{i_j,j} )}{\prod_{k\neq i_j=1}^j(  l_{kj}-l_{i_j,j})}
491: \right)^{1/2}
492: |m)_{-i_n,n;-i_{n-1},n-1;\ldots ;-i_j,j}\nn
493: \end{eqnarray}
494: 
495: \begin{eqnarray}
496: e_{j0}|m)&=&\sum_{i_n=1}^n\sum_{i_{n-1}=1}^{n-1}\ldots \sum_{i_j=1}^{j}
497: (1-\theta_{i_n})(-1)^{\theta_1+\ldots +\theta_{i_n-1}}(l_{i_n,n+1}+l_{0,n+1}+1)^{1/2}\nn\\
498: &\times&
499: \prod_{r=j+1}^nS(i_r,i_{r-1})
500: \left( \frac{\prod_{k\neq i_{r-1}=1}^{r-1}  
501: (l_{k,r-1}-l_{i_r,r}-1 )\prod_{k\neq i_r=1}^r(l_{kr}-l_{i_{r-1},r-1})
502:  }{  \prod_{k\neq i_r=1}^r (l_{kr}-l_{i_r,r})\prod_{k\neq i_{r-1}=1}^{r-1}(l_{k,r-1}
503:  -l_{i_{r-1},r-1}-1)}
504: \right)^{1/2} \label{ej0}\\
505: &\times &
506: \left({\prod_{k\neq i_n=1}^n } \frac{(l_{kn}-l_{i_n,n} )}{(  l_{k,n+1}-l_{i_n,n+1})}
507: \right)^{1/2}
508: \left( \frac{\prod_{k=1}^{j-1} (l_{k,j-1}-l_{i_j,j}-1 )}{\prod_{k\neq i_j=1}^j(  l_{kj}-l_{i_j,j})}
509: \right)^{1/2}
510: |m)_{+i_n,n;+i_{n-1},n-1;\ldots ;+i_j,j},\nn
511: \end{eqnarray}
512: where $j=1,\ldots ,n$, each symbol $\pm i_k,k$ attached as a subscript to $|m)$ indicates a
513: replacement $m_{i_k,k}\rightarrow m_{i_k,k}\pm 1$, and 
514: \begin{eqnarray}
515: &&S(k,l) = \left\{ \begin{array}{lll}
516:  {\;\;1} & \hbox{for} & k\leq l  \\ 
517:  {-1} & \hbox{for} & k>l .
518:  \end{array}\right.
519: \end{eqnarray}
520: 
521: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
522: %%%%% section
523: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
524: 
525: \setcounter{equation}{0}
526: \section{A special class of representations} \label{sec:covariant}%
527: 
528: One interesting class of representations of $\gl(1|n)$ is that with 
529: $[m]_{n+1}=[p,0,\ldots,0]$, i.e.\ with highest weight $\Lambda = p \epsilon$,
530: where $p\in\Z_+$.
531: These are covariant representations labelled by the partition $(p)$. 
532: The representation space $W([p,0,\ldots,0])$ is simply denoted by $W(p)$.
533: The GZ basis vectors of $W(p)$ can be denoted by $|p;\varphi_1,\ldots,\varphi_n\rangle$, where
534: the relation to the GZ labels is determined by
535: \begin{equation}
536: \varphi_i= \sum_{j=1}^i m_{ji} - \sum_{j=1}^{i-1} m_{j,i-1}.
537: \end{equation}
538: The constraints for the GZ labels lead to: $\varphi_i\in\{0,1\}$ and 
539: $\sum_{i=1}^n \varphi_i \leq \min(p,n)$.
540: The representations $W(p)$ and the basis vectors $|p;\varphi_1,\ldots,\varphi_n\rangle$
541: have been constructed by means of Fock space techniques,
542: and the action of the $\gl(1|n)$ generators is very simple, see~\cite{Palev3}.
543: This class is referred to as the class of Fock representations.
544: 
545: In this paper, another special class of representations will be of importance,
546: namely those with $[m]_{n+1}=[1,p-1,0,\ldots,0]$, i.e.\ with highest weight 
547: $\Lambda = \epsilon+(p-1)\delta_1$, where $p\in\Z_+$. These are again 
548: covariant irreducible representations, labelled this time by the
549: partition $(1,1,\ldots,1)=(1^p)$.
550: The representation space $W([1,p-1,0,\ldots,0])$ will be denoted by $V(p)$.
551: We shall assume that $p>1$, since for $p=1$ it is actually a Fock representation.
552: Note that $V(p)$ is atypical of type~2, and by Proposition~3 it is also unitary.
553: 
554: One can now apply the general GZ-basis construction of the previous section
555: to this special class.
556: The GZ basis vectors of $V(p)$ are of the form 
557: \begin{equation}
558: \left|
559: \begin{array}{lcllll}
560: 1 & p-1 & 0 & \cdots & 0 &0   \\
561: & q_n & 0 & \cdots & 0 & 0  \\
562: & q_{n-1} & 0 & \cdots &  0  &  \\
563: &\vdots & \qdots & & & \\
564: &q_2 & 0& & &\\
565: &q_{1} & & & &
566: \end{array}
567: \right)
568: \label{q}
569: \end{equation}
570: with $q_n=p-\theta \geq q_{n-1} \geq \cdots \geq q_1 \geq 0$, where $\theta =0,1$
571: (note that $\theta=1-\theta_1$ in the description of~(\ref{cond1})).
572: Since most of the entries in the GZ-array are zero, it will be more convenient to denote 
573: this vector by
574: \begin{equation}
575: | \theta; q_n,q_{n-1},\ldots,q_1 \rangle \qquad 
576: (q_n=p-\theta \geq q_{n-1} \geq \cdots \geq q_1 \geq q_0 \equiv 0).
577: \end{equation}
578: The actions~(\ref{e_00})-(\ref{fn}) and~(\ref{e0j})-(\ref{ej0}) imply that
579: \begin{eqnarray}
580: && e_{00} |\theta; q_n,q_{n-1},\ldots,q_1 \rangle = \theta |\theta; q_n,q_{n-1},\ldots,q_1 \rangle; \\
581: && e_{kk} |\theta; q_n,q_{n-1},\ldots,q_1 \rangle = (q_k-q_{k-1}) |\theta; q_n,q_{n-1},\ldots,q_1 \rangle,
582: \qquad (1\leq k\leq n);\\
583: && e_{k,k-1} |\theta; q_n,q_{n-1},\ldots,q_1 \rangle =\sqrt{(q_k-q_{k-1}+1)(q_{k-1}-q_{k-2})} \nn\\
584: && \qquad\times\  |\theta; q_n,q_{n-1},\ldots,q_k,q_{k-1}-1,\ldots,q_1 \rangle, \qquad (2\leq k \leq n);\\
585: && e_{k-1,k} |\theta; q_n,q_{n-1},\ldots,q_1 \rangle =\sqrt{(q_k-q_{k-1})(q_{k-1}-q_{k-2}+1)} \nn\\
586: && \qquad \times\  |\theta; q_n,q_{n-1},\ldots,q_k,q_{k-1}+1,\ldots,q_1 \rangle, \qquad (2\leq k \leq n);\\
587: && e_{k0} |\theta; q_n,q_{n-1},\ldots,q_1 \rangle =\theta \sqrt{q_k-q_{k-1}+1} \nn\\
588: && \qquad\times\  |1-\theta; q_n+1,q_{n-1}+1,\ldots,q_k+1,q_{k-1},\ldots,q_1 \rangle,\qquad(1\leq k\leq n);\\
589: && e_{0k} |\theta; q_n,q_{n-1},\ldots,q_1 \rangle =(1-\theta)\sqrt{q_k-q_{k-1}} \nn\\
590: && \qquad \times\  |1-\theta; q_n-1,q_{n-1}-1,\ldots,q_k-1,q_{k-1},\ldots,q_1 \rangle, \qquad(1\leq k\leq n).
591: \end{eqnarray}
592: These formulas become even simpler in yet another notation for the basis vectors.
593: Let us put
594: \begin{equation}
595: r_n=q_n-q_{n-1}, r_{n-1}=q_{n-1}-q_{n-2}, \ldots, r_{2}=q_2-q_1, r_1=q_1,
596: \end{equation}
597: then all $r_i$ are nonnegative integers, with $|r|=r_1+\cdots+r_{n}=q_n$.
598: For the vectors with $\theta=1$, $q_n=p-1$, so $|r|=p-1$.
599: For the vectors with $\theta=0$, $q_n=p$, so $|r|=p$.
600: Thus all vectors of $V(p)$ are described by:
601: \begin{equation}
602: v(\theta;r)\equiv v(\theta;r_1,r_2, \ldots,r_n),\qquad
603: \theta\in\{0,1\},\ r_i\in\{0,1,2,\ldots\},\ \hbox{ and } \theta+r_1+\cdots+r_n= p.
604: \label{newbasis}
605: \end{equation}
606: In this notation the highest weight vector is $v(1;p-1,0,\ldots,0)$.
607: The action of the $\gl(1|n)$ generators on the new basis~(\ref{newbasis}) is now given by:
608: \begin{eqnarray}
609: && e_{00} v(\theta; r) = \theta v(\theta; r);\\
610: && e_{kk} v(\theta; r) = r_k v(\theta; r),\qquad (1\leq k\leq n);\\
611: && e_{k,k-1} v(\theta; r) =\sqrt{r_{k-1}(r_k+1)}\
612:   v(\theta; r_1,\ldots,r_{k-1}-1, r_k+1,\ldots,r_{n}),\quad(2\leq k\leq n);\\
613: && e_{k-1,k} v(\theta; r) =\sqrt{(r_{k-1}+1)r_k}\
614:   v(\theta; r_1,\ldots,r_{k-1}+1, r_k-1,\ldots,r_{n}), \quad(2\leq k\leq n);\\
615: && e_{k0} v(\theta; r) =\theta \sqrt{r_k+1}\ v(1-\theta; r_1,\ldots,r_k+1,\ldots,r_n), \qquad (1\leq k\leq n);\\
616: && e_{0k} v(\theta; r) =(1-\theta)\sqrt{r_k}\ v(1-\theta; r_1,\ldots,r_k-1,\ldots,r_n),
617: \qquad (1\leq k\leq n).
618: \end{eqnarray}
619: With respect to the inner product 
620: \begin{equation}
621: \langle v(\theta;r) | v(\theta';r') \rangle = \delta_{\theta,\theta'}\delta_{r,r'},
622: \end{equation}
623: the representation $V(p)$ is unitary for the star condition~(\ref{star-condition}).
624: 
625: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
626: %%%%% section
627: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
628: 
629: \setcounter{equation}{0}
630: \section{The $N$-particle $D$-dimensional WQO} \label{sec:WQO}%
631: 
632: \subsection{The WQO operators}
633: 
634: The rest of this paper is devoted to new solutions of the WQO and their physical interpretation.
635: Let us briefly recall the context of WQOs~\cite{Palev1}-\cite{SJ}. 
636: Let $\hat{H}$ be the Hamiltonian of an $N$-particle $D$-dimensional
637: harmonic oscillator, that is
638: \begin{equation}
639: \hat{H}=\sum_{\alpha=1}^{N} \Big( \frac{\hbP_\alpha^2}{2m}
640: + \frac{m\omega^2}{2} {\hbR}_\a^2 \Big), \label{H}
641: \end{equation}
642: with $\hbP_\alpha$ and $\hbR_\alpha$ $D$-dimensional vector operators
643: corresponding to the momentum and position of the particle~$\alpha$ ($\alpha=1,2,\ldots,N$), $m$
644: the mass and $\omega$ the frequency of each oscillator.
645: %The cases $D=1$, $D=2$ and $D=3$ will be most common.
646: We proceed to view this oscillator as a Wigner quantum system: this means
647: that the canonical commutation relations are not required, but are replaced 
648: by compatibility conditions between Hamilton's equations and the
649: Heisenberg equations. In other words, Hamilton's equations
650: \begin{equation}
651:     {\dot{\hbP}}_\a=-m\omega^2\hbR_\a, \ \ {\dot{\hbR}}_\a = \frac{1}{m}\hbP_\a
652:     \ ~ {\rm for} ~\ \a=1,2,\ldots,N,
653:      \label{Ham}
654: \end{equation}
655: and the Heisenberg equations
656: \begin{equation}
657:      {\dot{\hbP}}_\a = \frac{i}{\hbar}[\hat{H},\hbP_\a], \ \
658:      {\dot{\hbR}}_\a = \frac{i}{\hbar}[\hat{H},\hbR_\a]
659:      \ ~ {\rm for} ~ \ \a=1,2\ldots,N,
660:      \label{Heis}
661: \end{equation}
662: should be identical as operator equations. These compatibility conditions are
663: such that
664: \begin{equation}
665:    [\hat{H},\hbP_\a]=i\hbar m \omega^2\hbR_\a ,\ \
666:    [\hat{H},\hbR_\a]=-\frac{i\hbar}{m}\hbP_\a
667:     \ ~ {\rm for} ~ \ \a=1,2,\ldots,N.
668:      \label{comp}
669: \end{equation}
670: To make the connection with $\ssl(1|DN)$ we write the operators
671: $\hbP_\a$ and $\hbR_\a$
672: for $\a=1,2,\ldots,N$ in terms of new operators:
673: \begin{equation}
674: A_{D(\alpha -1) +k}^\pm= \sqrt{\frac{(DN-1)m \omega}{4\hbar}}
675:  \hR_{\a k} \pm i
676:   \sqrt { \frac{(DN-1)}{4m \omega \hbar}} \hP_{\a k}, \quad k=1,\ldots,D.
677:  \label{A}
678: \end{equation}
679: The Hamiltonian $\hat{H}$ of~(\ref{H}) and the
680: compatibility conditions~(\ref{comp}) take the form:
681: \begin{equation}
682:      \hat{H}= \frac{\omega \hbar}{DN-1} \sum_{j=1}^{DN} \{A_j^+, A_j^-\},
683:      \label{HAA}
684: \end{equation}
685: \begin{equation}
686: \sum_{j=1}^{DN}  [ \{A_{j}^+,A_{j}^- \},A_{i}^\pm]
687: =\mp (DN-1)A_{ i}^\pm , \quad i,j=1,2,\ldots , DN.
688: \label{comp1}
689: \end{equation}
690: As a solution to~(\ref{comp1}) one can choose operators $A_{ i}^\pm$ that satisfy the
691: following triple relations:
692: \begin{eqnarray}
693: && [\{A_{i}^+,A_{j}^-\},A_{k}^+]=
694: \delta_{jk}A_{i}^+
695: -\delta_{ij}A_{k}^+,  \nn\\
696: && [\{A_{ i}^+,A_{ j}^-\},A_{ k}^-]=
697: -\delta_{ik}A_{j}^-
698: +\delta_{ij}A_{k}^-, \label{sl}\\
699: && \{A_{i}^+,A_{ j}^+\}=
700: \{A_{ i}^-,A_{ j}^-\}=0. \nn
701: \end{eqnarray}
702: 
703: \begin{prop}
704: The operators $A_{j}^\pm$, for $j=1,2,\ldots,DN$,
705: are odd elements generating the Lie superalgebra
706: $\ssl(1|DN)$.
707: \end{prop}
708: This means that one can identify the operators $A^+_j$ and $A^-_j$ with the $\ssl(1|DN)$ 
709: generators $e_{j0}$ and $e_{0j}$ respectively:
710: \begin{equation}
711: A_j^+=e_{j0},\qquad A_j^-=e_{0j}.
712: \end{equation}
713: With this identification it should be noted that
714: \begin{equation}
715: \hat{H}=\frac{\hbar\omega}{(DN-1)} \sum_{k=1}^{DN} (e_{00}+e_{kk}) =
716: \frac{\hbar\omega}{(DN-1)} \left( DN\,e_{00}+ \sum_{k=1}^{DN} e_{kk} \right).
717: \label{Hee}
718: \end{equation}
719: 
720: To end this subsection, let us also recall that in the case $D=3$ it is possible to
721: introduce the angular momentum operator of each particle $\alpha$.
722: These single particle angular momentum operators
723: $\hJ_{\a j}$ are defined by~\cite{Palev5}
724: \begin{equation}
725:       \hJ_{\a j} = -\frac{3N-1}{2\hbar} \sum_{k,l=1}^3
726:     \ \epsilon_{jkl} \{\hR_{\a k},\hP_{\a l}\}
727:       \ \ \a=1,2,\ldots,N,\ \ j=1,2,3,
728:       \label{MRP}
729: \end{equation}
730: and take the following form:
731: \begin{equation}
732:       \hJ_{\a j} = -i \sum_{k,l=1}^3 \epsilon_{jkl}
733:    \{ A_{3(\a-1)+k}^+,A_{3(\a-1)+ l}^-\}.
734:       \label{MA}
735: \end{equation}
736: In terms of these operators the three components of the total angular
737: momentum operator $\hat{\J}$ are given by
738: \begin{equation}
739: \hJ_j=\sum_{\a=1}^N \hJ_{\a j},\quad j=1,2,3. \label{M}
740: \end{equation}
741: It is straightforward to verify that with respect to this choice
742: of angular momentum operator $\hat{\J}$ the operators $\hbR_\a$, $\hbP_\a$,
743: $\hat{\J}_\a$ and $\hat{\J}$ all transform as $3$-vectors.
744: 
745: \subsection{The WQO representations}
746: 
747: The Hilbert space (state space) of the WQO is a representation space $W$ of the
748: Lie superalgebra $\gl(1|DN)$ in such a way that the ``observables'' (the position and
749: momentum operators $\hat R_{\alpha,k}$ and $\hat P_{\alpha,k}$) are Hermitian operators.
750: This means that $(A_j^\pm )^\dagger = A_j^\mp$, or $e_{j0}^\dagger=e_{0j}$. 
751: In other words, the state spaces are unitary representations of $\gl(1|DN)$, in the
752: sense of Proposition~3.
753: Since all ``physical operators'' in this WQO model are expressed in terms of
754: the $A_j^\pm$, and their actions are known by~(\ref{e0j})-(\ref{ej0}),
755: all the relevant physical properties can in principle be deduced.
756: 
757: In previous papers, an investigation was made of the physical properties of the
758: $\gl(1|DN)$ solutions in the Fock representation spaces $W(p)$, see~\cite{Palev1, K1}
759:  for $D=3$, $N=1$ 
760: (the single particle 3-dimensional WQO) and~\cite{Palev5, K2}
761: for $D=3$ and $N$ arbitrary (the last case corresponding to a superposition
762: of $N$ single particle 3-dimensional WQOs).
763: The most striking properties are as follows:
764: the energy of each particle has at most four different eigenvalues (equidistant energy levels);
765: the geometry is non-commutative, in the sense that coordinate operators do not commute;
766: the position and momentum operators have discrete spectra.
767: In this paper, we shall consider more general solutions corresponding to
768: arbitrary unitary representations of $\gl(1|DN)$, and to the special class
769: of representations $V(p)$.
770: 
771: Some of these properties will be dealt with
772: in general (arbitrary $N$ and $D$, arbitrary unitary representations); 
773: for some others it will be useful to take particular values of $N$ and/or $D$,
774: or to restrict oneself to specific classes of unitary representations.
775: In the general case, all unitary representations of $\gl(1|DN)$ are of relevance.
776: In order to simplify notation, we shall use the abbreviation
777: \begin{equation}
778: n=DN
779: \label{n=DN}
780: \end{equation}
781: throughout this section.
782: 
783: \subsubsection{Energy spectrum}
784: 
785: The Hamiltonian $\hat{H}$ is diagonal in the GZ-basis,
786: i.e.\ the basis vectors $|m)$ are stationary states
787: of the system. This follows from the form of $\hat{H}$ given in~(\ref{Hee})
788: and the action of $e_{00}$ and $e_{kk}$ given in~(\ref{e_00}) and~(\ref{e_kk}).
789: In each space $W([m]_{n+1})$ there is a finite number of equally
790: spaced energy levels, with spacing $\hbar\omega$: 
791: \begin{equation}
792:      \hat{H} |m) = E_q |m) 
793: \end{equation}
794: with
795: \begin{equation}
796:   E_q={\hbar\omega}\left( \frac{n m_{0,n+1}+m_{1,n+1}+\ldots+m_{n,n+1}}{n-1}-q\right),
797: \end{equation}
798: where  $q=\sum_{j=1}^{n}\theta_j$.
799: For typical representations, $q$ takes the values $0,1,\ldots,n$.
800: For a unitary representation $W([m]_{n+1})$ atypical of type~$k$, $q$ takes the values $0,1,\ldots, k-1$.
801: Recall that in that case, the representation labels satisfy
802: \begin{equation}
803:      m_{0,n+1}+m_{k,n+1}=k-1, \qquad m_{k,n+1}=m_{k+1,n+1}=\cdots=m_{n,n+1}.
804: \label{Eq}
805: \end{equation}
806: So for a typical representation, there are $n+1$ equidistant energy levels; 
807: for an atypical representation of type~$k$
808: ($k=1,2,\ldots,n$) there are $k$ equidistant energy levels. 
809: The degeneracy of these levels is high, and can be determined explicitly from the GZ-basis labels,
810: or equivalently from the dimensions of the $\gl(n)$ representations in the decomposition
811: of $W([m]_{n+1})$.
812: 
813: For example, for the special class of representations $V(p)$, atypical of type $2$, 
814: there are only two distinct energy levels:
815: \begin{eqnarray*}
816: &&E_0={\hbar\omega}(\frac{p}{n-1})\hbox{ with degeneracy } 
817: \left(\begin{array}{c}{p+n-1}\\{n-1}\end{array}\right),\\
818: &&E_1={\hbar\omega}(\frac{p}{n-1}+1)\hbox{ with degeneracy } 
819: \left(\begin{array}{c}{p+n-2}\\{n-1}\end{array}\right).
820: \end{eqnarray*}
821: 
822: \subsubsection{Position and momentum operators}
823: 
824: The position operators $\hat R_{\a k}$ ($\a =1,\ldots, N$, $k=1,\ldots,D$)
825: of the oscillating particles do not commute with each other
826: \begin{equation}
827: [\hat R_{\a i},\hat R_{\b j}]\ne 0 \quad \hbox{ for } \quad \a i\ne \b j.
828: \label{3.3}
829: \end{equation}
830: Similarly
831: \begin{equation}
832: [\hat P_{\a i},\hat P_{\b j}]\ne 0 \quad \hbox{ for } \quad \a i\ne \b j.
833: \label{3.4}
834: \end{equation}
835: These relations imply that the WQO belongs to the class of models of non-commutative
836: quantum oscillators~\cite{NP}-\cite{MM}, or to theories with non-commutative 
837: geometry~\cite{HS, C}. 
838: In general, the action of $[\hat R_{\a i},\hat R_{\b j}]$ is difficult to describe and interpret,
839: even when acting on GZ basis vectors of particular unitary representations.
840: For example, for the representations $V(p)$, one finds:
841: \begin{eqnarray}
842: && [\hat R_{\a i},\hat R_{\b j}] v(\theta;r) = \frac{\hbar}{(n-1)m\omega} \Bigl( 
843: (-1)^\theta \sqrt{r_l(r_k+1)} v(\theta;r_1,\ldots,r_k+1,\ldots,r_l-1,\ldots,r_{n}) \nn\\
844: && \qquad - (-1)^\theta \sqrt{r_k(r_l+1)} v(\theta;r_1,\ldots,r_k-1,\ldots,r_l+1,\ldots,r_{n})\Bigr),
845: \end{eqnarray}
846: where $k=D(\a-1)+i$ and $l=D(\b-1)+j$ (and it is assumed that $k<l$).
847: 
848: However, the squares of the components of position and momentum operators commute:
849: \begin{equation}
850: [\hat R_{\a i}^2,\hat R_{\b j}^2]=[\hat P_{\a i}^2,\hat P_{\b j}^2]= 0 \quad \hbox{ for } \quad \a i\ne \b j .
851: \label{3.5}
852: \end{equation}
853: Furthermore, the GZ basis states $|m)$ are eigenstates of these operators,
854: \begin{eqnarray}
855: \hat R_{\a i}^2 |m)&=& \frac{\hbar}{(n-1)m\omega} (m_{0,n+1}+\ldots+m_{n,n+1}
856: -m_{1,n}-\cdots-m_{n,n}\nn\\
857: && +m_{1,k}+\cdots +m_{k,k}-m_{1,k-1}-\cdots -m_{k-1,k-1})|m),
858: \label{actionR2}
859: \end{eqnarray}
860: where $k=D(\a-1)+i$.
861: Thus the spectrum of the position operator component $\hat R_{\a i}$ is given by 
862: the set of values
863: \begin{equation}
864: \pm \sqrt{ \frac{\hbar}{(n-1)m\omega} (\sum_{j=0}^{n}
865:  m_{j,n+1}-\sum_{j=1}^{n}m_{j,n}+\sum_{j=1}^{k}m_{j,k}
866:  -\sum_{j=1}^{k-1}m_{j,k-1}}),
867:   \label{rev}
868: \end{equation}
869: where $k=D(\a-1)+i$ and the internal labels $m_{j,l}$ take all possible values 
870: allowed by~(\ref{cond1}).
871: For the special representations $V(p)$, (\ref{actionR2}) becomes
872: \begin{equation}
873: \hat R_{\a i}^2 v(\theta;r) = \frac{\hbar}{(n-1)m\omega}(r_k+\theta) v(\theta;r),\qquad k=D(\a-1)+i.
874: \label{R2onVp}
875: \end{equation}
876: This gives rise to a simple spectrum of the operators $\hat R_{\a i}$ in these representations.
877: 
878: \setcounter{equation}{0}
879: \section{Angular momentum content for $V(p)$}
880: 
881: As mentioned before, the case $D=3$ allows us to introduce angular momentum operators, 
882: see~(\ref{MRP})-(\ref{M}).
883: It is easy to verify that in general the stationary states $|m)$ are not 
884: eigenstates of $\hat{J}_{\a 3}$, $\hat{\J}_\a$, $\hat{J}_3$
885: and $\hat{\J}^2$. Thus in order to determine the
886: possible values of the total angular momentum $j$ for the $N$-particle 3-dimensional WQO,
887: it is best to use group theoretical methods rather than the explicit actions 
888: of these operators on basis states.
889: The components $\hat{J}_k$ in~(\ref{M}) are the generators of an $\so(3)$ subalgebra of
890: $\gl(1|n)=\gl(1|3N)$, that may be identified by means of the following chain of subalgebras:
891: \begin{equation}
892:     \gl(1|3N)\rightarrow \gl(1)\oplus \gl(3N)
893: \rightarrow \gl(1) \oplus \gl(3) \oplus \gl(N)
894: \rightarrow \gl(1) \oplus \so(3) \oplus \gl(N)
895: \rightarrow \gl(1) \oplus \so(3).
896:      \label{gl13n}
897: \end{equation} 
898: The first step in the branching rule, from $\gl(1|3N)$ to $\gl(1)\oplus \gl(3N)$, is
899: just determined by the GZ-pattern~(\ref{m}), where the $\gl(1)$ value is given by~(\ref{e_00}).
900: The branching rule for $\gl(3N)\rightarrow \gl(3)\oplus \gl(N)$
901: required in the second step and that for $\gl(3)\rightarrow \so(3)$ required in the third
902: step are both rather well known and have been implemented for example in
903: SCHUR~\cite{Schur}. Since they involve coefficients for which there is
904: no known general formula, we content ourselves with giving the results explicitly just
905: for the special representations $V(p)$.% (and $N=1$, $N=2$). 
906: 
907: Before doing this we should remark that the embedding of $\so(3)$ in $\gl(3N)$ is such that
908: the defining $3N$-dimensional representation $V^{\{1\}}_{\gl(3N)}$ of $\gl(3N)$ decomposes 
909: into a direct sum of $N$ copies of the defining $3$-dimensional representation $V^{1}_{\so(3)}$
910: of $\so(3)$, so that all states are of angular momentum $j=1$. The $p$th-fold tensor powers
911: of $V^{\{1\}}_{\gl(3N)}$ therefore contain only states whose angular momentum $j$ is bounded
912: by $p$. This includes all covariant irreducible representations $V^{\{\pi\}}_{\gl(3N)}$
913: where $\pi$ is any partition of weight $p$, including the special case $V^{\{p\}}_{\gl(3N)}$
914: of particular interest here.
915:  
916: With respect to the first step of (\ref{gl13n}), 
917: the representation $V(p)$ of $\gl(1|3N)$ decomposes into just two symmetric 
918: irreducible representations of $\gl(3N)$:
919: \begin{equation}
920: V(p) \rightarrow V^{0}_{\gl(1)} \otimes V^{\{p\}}_{\gl(3N)} + V^{1}_{\gl(1)} \otimes V^{\{p-1\}}_{\gl(3N)},
921: \label{step1}
922: \end{equation}
923: labelled by the partitions $\{p\}$ and $\{p-1\}$ respectively.
924: For symmetric representations labelled by a positive integer $p$, the branching from
925: $\gl(3N)$ to $\gl(3)\oplus \gl(N)$ is determined by the branching rule~\cite{K3}: % 
926: \begin{equation}
927: V^{\{p\}}_{\gl(3N)} \rightarrow \sum_{\lambda, |\lambda|=p} V^{\{\lambda\}}_{\gl(3)} \otimes V^{\{\lambda\}}_{\gl(N)},
928: \label{step2}
929: \end{equation}
930: where the sum is over all partitions $\{\lambda\}$ with $\lambda_1+\lambda_2+\cdots =p$. 
931: Furthermore, the labelling of representations of $\gl(3)$ and $\gl(N)$ requires the extra
932: condition that the length of $\{\lambda\}$ should be at most $\min(3,N)$.
933: 
934: In the case $N=1$, this step is absent, and for the branching $\gl(3)\rightarrow \so(3)$ we have immediately
935: \begin{equation}
936: V^{\{p\}}_{\gl(3)} \rightarrow V^{p}_{\so(3)} + V^{p-2}_{\so(3)}+\cdots V^{1\ \hbox{\tiny or }0}_{\so(3)}.
937: \end{equation}
938: Thus, using~(\ref{step1}), the total angular momentum values are given by $j=0,1,\ldots,p$, with no
939: multiplicities, so that
940: \begin{equation}
941: V(p) \rightarrow \sum_{j=0}^p V^{j}_{\so(3)}.
942: \label{VpN1}
943: \end{equation}
944: 
945: In the case $N\geq2$, it might appear that the second step, (\ref{step2}),
946: cannot be avoided. However, if our intention is simply to identify all
947: possible angular momentum states $j$ and their multiplicities
948: that arise for $V^{\{p\}}_{\gl(3N)}$, then this may be accomplished
949: by exploiting some classical results from invariant theory. 
950: For this purpose, we shall briefly use the Lie group embedding $GL(3N)\supset SO(3)$
951: rather than the Lie algebra embedding $\gl(3N)\supset \so(3)$.
952: For the current class of representations the branching rules associated
953: with these two embeddings are identical.
954: 
955: For any compact continuous subgroup $G$ of $GL(n)$,
956: with group elements $g$ and Haar measure $d\mu(g)$, Molien's function~\cite{Molien}
957: takes the form~\cite{Sturmfels}
958: \begin{equation}
959:  M(P) = \int \frac{d\mu(g)}{\det(I-P\,g)},
960: \label{Mol}
961: \end{equation}
962: where $I$ is the unit $n\times n$ matrix. The significance of this
963: function in invariant theory is that when expanded in the form:
964: \begin{equation}
965:   M(P) = \sum_{p=0}^\infty \, n_p\, P^p\,,  
966: \label{MP}
967: \end{equation}
968: the expansion coefficient $n_p$ is the number of linearly independent 
969: invariants of $G$ that are homogeneous polynomials in the matrix elements
970: of $g$ of degree $p$. Equivalently, $n_p$ is the multiplicity of 
971: the trivial 1-dimensional identity representation of $G$ 
972: in the restriction of $V^{\{p\}}_{GL(n)}$ to $G$. More generally~\cite{RRS},
973: if the irreducible representation $V^j_{G}$, specified by some
974: label $j$, has character $\chi^j(g)$, with complex
975: conjugate $\chi^{j\ast}(g)$, then 
976: \begin{equation}
977: M^j(P)=\int \frac{\chi^{j\ast}(g)\, d\mu(g)}{\det(I-P\,g)}
978: \label{MolChi}
979: \end{equation}
980: has an expansion of the form
981: \begin{equation}
982:   M^j(P) = \sum_{p=0}^\infty\, n_{p,j}\, P^p\,.
983: \label{MPj}
984: \end{equation}
985: where each coefficient $n_{p,j}$ is the multiplicity of 
986: the irreducible representation $V^j_{G}$ of $G$ 
987: in the restriction of $V^{\{p\}}_{GL(n)}$ to $G$. 
988: 
989: Specialising this to the case $n=3N$ and $G=SO(3)$ we immediately have a
990: means of calculating the angular momentum content of the
991: representation $V^{\{p\}}_{GL(3N)}$
992: as has been done in the case $N=2$ by Raychev {\it et al}~\cite{RRS}.
993: To this end it should be noted that
994: each element $g$ of $SO(3)\subset GL(3N)$ arises as the $N$th-fold
995: tensor power of a $3\times 3$ matrix which can itself be diagonalised,
996: with diagonal elements $(z,1,z^{-1})$ where $z=\exp(i\phi)$ for some real
997: $\phi$. With this parametrisation, taking into account the diagonalisation 
998: process, the relevant Haar measure is given by~\cite{RRS}
999: \begin{equation}
1000:       d\mu(g)= \frac{1}{2}\ (1-\cos\phi) \, d\phi\,,
1001: \label{mu-g}
1002: \end{equation} 
1003: with $0\leq\phi<2\pi$.
1004: Moreover, the diagonal form of $g$ is such that
1005: \begin{equation}
1006:       \det(I-P\,g)=\left( (1-P\,z)(1-P)(1-P\,z^{-1})\right)^N. 
1007: \label{det-Pz}
1008: \end{equation} 
1009: In addition it is well known that the character of the irreducible
1010: representation $V^j_{SO(3)}$ of $SO(3)$ is given by
1011: \begin{equation}
1012:      \chi^j(g) = \frac { z^{j+\frac12}-z^{-j-\frac12} }
1013:          { z^{\frac12}-z^{-\frac12} }\,,
1014: \label{chi-j}
1015: \end{equation}
1016: where it is to be noticed that $\chi^{j\ast}(g) =\chi^j(g)$.
1017: In our case, only integer (rather than half-integer) values of $j$ arise.
1018: This allows us to introduce a generating function for $SO(3)$ characters
1019: as follows:
1020: \begin{eqnarray}
1021: \hbox{char}(J,g)&=& \sum_{j=0}^\infty \, J^j\, \chi^j(g) 
1022: = \sum_{j=0}^\infty\,  \frac{ (J\,z)^j\, z^{\frac12}-(J\,z^{-1})^j\, z^{-\frac12} }
1023:         { z^{\frac12}-z^{-\frac12} }\cr
1024: &=& \sum_{j=0}^\infty \left( \frac {(J\,z)^j}{1-z^{-1}} - \frac{(J\,z^{-1})^j}{z-1}\right)\cr
1025: &=& \frac {1}{(1-J\,z)(1-z^{-1})}-\frac{1}{(1-J\,z^{-1})(z-1)}.
1026: \label{charJ}
1027: \end{eqnarray}
1028: Hence, we have
1029: \begin{eqnarray}
1030: &&M(P,J)_N=\sum_{j=0}^\infty M^j(P)\,J^j%=\sum_{p,j=0}^\infty n_{p,j}\, P^p\,J^j 
1031: = \int \frac{ \hbox{char}(J,g^{-1})\, d\mu(g)}{\det(I-P\,g)} \cr
1032: &&\ = \oint_{|z|=1} \left(\frac {(1-z^{-1})^{-1} }{(1-J\,z)}
1033:                -\frac{(z-1)^{-1}}{(1-J\,z^{-1}) } \right)
1034: \frac { \left( 1-\frac12(z+z^{-1})\right) } 
1035:       { \left( (1-P\,z)(1-P)(1-P\,z^{-1})\right)^N     }
1036: \frac{dz}{4\pi iz}\,,    
1037: \label{MPJ}
1038: \end{eqnarray}
1039: where the fact that $z=\exp(i\phi)$ has allowed the integral over 
1040: $\phi$ to be re-written as an integral around the unit circle in the 
1041: complex $z$-plane.
1042: 
1043: In fact, in view of the rather simple dependence on $N$, we can go even further 
1044: and introduce the master generating function:
1045: \begin{eqnarray}
1046: \hskip-1cm
1047: &&M(P,J,{\cal N})=\sum_{N=0}^\infty M(P,J)_N\,{\cal N}^N
1048: =\sum_{p,j,N=0}^\infty n_{p,j,N}\, P^p\,J^j\,{\cal N}^N \cr
1049: \hskip-1cm
1050: &&\ = \oint_{|z|=1} \left(\frac {(1-z^{-1})^{-1} }{(1-J\,z)}
1051:                -\frac{(z-1)^{-1}}{(1-J\,z^{-1}) } \right)
1052: \left( \frac{\left( 1-\frac12(z+z^{-1})\right)} 
1053:         {1-{\cal N}/((1-P\,z)(1-P)(1-P\,z^{-1}))} \right) \frac{dz}{4\pi  iz}\,,    
1054: \label{MPJN}
1055: \end{eqnarray}
1056: where $n_{p,j,N}$ is the multiplicity of the angular momentum state $j$ in the
1057: irreducible representation $V^{\{p\}}_{GL(3N)}$ of $GL(3N)$, or equivalently in
1058: the representation $V^{\{p\}}_{\gl(3N)}$ of $\gl(3N)$. This integral
1059: may be evaluated by determining the residues of the integrand at all poles within 
1060: the unit circle. The result is
1061: \begin{eqnarray}
1062: &&M(P,J,{\cal N}) = 
1063: 1 + \frac12\, \frac{{\cal N}(1-J)} {J{\cal N}-(1-P)(1-JP)(J-P)}\cr
1064: &&{\phantom{M(P,J,{\cal N}) =}}\  + \frac12\, \frac{{\cal N}(1+J)} {(J{\cal N}-(1-P)(1-JP)(J-P)}\,
1065:     \sqrt{ \frac {{\cal N}-(1-P)^3 } {{\cal N}-(1-P)(1+P)^2   }  }
1066:     \label{masterGF}
1067: \end{eqnarray}
1068: 
1069: Although not very illuminating, this formula may readily be used to recover the 
1070: generating functions $M(P,J)_N$ appropriate to any fixed $N$ by expanding~(\ref{masterGF})
1071: in powers of ${\cal N}$. In particular we find:
1072: 
1073: \begin{eqnarray}
1074: M(P,J)_{N=1}&=&\frac {1} {(1-JP)(1-P^2)}\\
1075: M(P,J)_{N=2}&=&\frac {1+JP^2} {(1-JP)^2(1- P^2)^3}\\
1076: M(P,J)_{N=3}&=&\frac { 1+3JP^2+P^3-J^2P^3-3JP^4-J^2P^6} {(1-JP)^3(1- P^2)^6}\\
1077: M(P,J)_{N=4}&=&\frac { 
1078: \begin{array}{c} 
1079: (J^{3}P^{10} + J^{3}P^{8} + 6J^{2}P^{8} + 4J^{3}P^{7} - 4JP^{7} 
1080: + J^{3}P^{6} - 10J^{2}P^{6} + P^{6}\cr
1081:  - 10JP^{4} + P^{4} + J^{3}P^{4} - 4J^{2}P^{3} + 4P^{3} + 6JP^{2} + P^{2} + 1)\cr
1082: \end{array} }
1083: {(1-JP)^4(1- P^2)^9 }
1084: \label{MPJn} 
1085: \end{eqnarray}
1086: 
1087: Finally, it follows from (\ref{step1}) that the required generating function for $\gl(1|3N)$ 
1088: is obtained by including an additional factor of $(1+P)$ in the numerator
1089: of each of these expressions. For example in the case $N=1$, the $\gl(1|3)$ generating 
1090: function is
1091: \begin{equation}
1092: G(P,J)_{N=1}=(1+P)M(P,J)_{N=1}=\frac{1}{(1-JP)(1-P)}\,,
1093: \label{G13}
1094: \end{equation}
1095: whose expansion takes the form
1096: \begin{equation}
1097: G(P,J)_{N=1}=1 + (1+J)P + (1+J+J^2)P^2 + (1+J+J^2+J^2+J^3)P^3 + \cdots \,.
1098: \label{G13exp}
1099: \end{equation}
1100: This corresponds of course to our earlier observation (\ref{VpN1}) that in the
1101: $N=1$ case $V(p)$ contains states of angular momentum $j=0,1,\ldots,p$,
1102: without multiplicity.
1103: 
1104: 
1105: By way of a more interesting example, in the case $N=2$ the complete
1106: generating function is (see also~\cite{RRS})
1107: \begin{equation}
1108: G(P,J)_{N=2}=\frac{(1+P)(1+P^2J)}{(1-P^2)^3(1-PJ)^2}.
1109: \end{equation}
1110: In the expansion of $G(P,J)_{N=2}$, the coefficient of $P^p$ yields the 
1111: angular momentum content of the
1112: $\gl(1|6)$ representation $V(p)$,
1113: \begin{eqnarray}
1114: G(P,J)_{N=2}&=&1 + (1+2J)P+(3+3J+3J^2)P^2 + (3+7J+5J^2+4J^3)P^3\nn\\
1115: &&+ (6+9J+11J^2+7J^3+5J^4)P^4 \nn\\
1116: &&+ (6+15J+15J^2+15J^3+9J^4+6J^5)P^5+\cdots
1117: \end{eqnarray}
1118: For example, for $V(3)$ the possible total angular momentum values are $j=0,1,2,3$ with multiplicities 
1119: $3,7,5$ and $4$ respectively.
1120: Note that in general for $\gl(1|6)$ the total angular momentum values in $V(p)$ are given,
1121: as expected, by $j=0,1,\ldots,p$,
1122: but each value appears with a certain multiplicity.
1123: 
1124: To conclude, the complete  angular momentum content (including multiplicities) for all representations
1125: $V(p)$ of $\gl(1|3N)$ is resolved by $G(P,J)_N=(1+P) M(P,J)_N$, where $M(P,J)_N$ is the
1126: coefficient of ${\cal N}^N$ in the expansion of the master generating function~(\ref{masterGF}).
1127: 
1128: \setcounter{equation}{0}
1129: \section{Position operator properties and a classical limit for $V(p)$}
1130: 
1131: The special class of representations $V(p)$ not only allows a complete analysis of the angular momentum content,
1132: it also allows a deeper analysis of the position and momentum operator properties.
1133: Let us consider in more detail the single particle ($N=1$) $D$-dimensional WQO 
1134: in the Hilbert space $V(p)$ as representation space.
1135: A set of basis vectors is given by
1136: \begin{equation}
1137: v(\theta;r)\equiv v(\theta;r_1,r_2, \ldots,r_D),\qquad
1138: \theta\in\{0,1\},\ r_i\in\{0,1,2,\ldots\},\ \hbox{ and } \theta+r_1+\cdots+r_D= p.
1139: \label{newbasis2}
1140: \end{equation}
1141: There are only two distinct energy levels:
1142: \begin{equation}
1143: \hat{H} v(\theta;r) = \hbar\omega (\frac{p}{D-1}+\theta) v(\theta;r).
1144: \end{equation}
1145: Furthermore, the squares of the position operator components satisfy:
1146: \begin{equation}
1147: \hat R^2_i v(\theta;r) = \frac{\hbar}{(D-1)m\omega} (r_i+\theta) v(\theta;r).
1148: \end{equation}
1149: Similarly, one has
1150: \begin{equation}
1151: \hat P^2_i v(\theta;r) = \frac{\hbar m\omega}{D-1} (r_i+\theta) v(\theta;r).
1152: \end{equation}
1153: 
1154: In order to illustrate some of the spatial properties, let us first assume that $D=2$, i.e.\
1155: we are dealing with a 2-dimensional one particle WQO.
1156: The squares of the position operator components commute, $[\hat R_1^2,\hat R_2^2]=0$, and the $2p+1$ basis vectors
1157: $v(\theta;r_1,r_2)$ are common eigenvectors:
1158: \begin{eqnarray}
1159: &&\hat R^2_1 v(\theta;r) = \frac{\hbar}{m\omega} (r_1+\theta) v(\theta;r), \quad
1160: \hat R^2_2 v(\theta;r) = \frac{\hbar}{m\omega} (r_2+\theta) v(\theta;r), \nn\\
1161: &&(\hat R^2_1+\hat R^2_2) v(\theta;r) = \frac{\hbar}{m\omega} (p+\theta) v(\theta;r).
1162: \end{eqnarray}
1163: Let us consider in Figure~1 a plot of the eigenvalues of 
1164: $\hat R_1^2$ and $\hat R_2^2$ with respect to these eigenvectors, for some values of $p$.
1165: \begin{figure}[htb]
1166: \caption{Eigenvalues of $\hat R_1^2$ (on the $x$-axis) and $\hat R_2^2$ (on the $y$-axis), 
1167: for $p=5$ and $p=50$ (in units of $\frac{\hbar}{m\omega}$). 
1168: For each commom eigenvector $v(\theta;r_1,r_2)$
1169: the corresponding eigenvalues are plotted.}
1170: \begin{center}
1171: \includegraphics{figsq-D2-p5.eps}\hskip 15mm
1172: \includegraphics{figsq-D2-p50.eps}
1173: \end{center}
1174: \end{figure}
1175: 
1176: Since $\hat R_1$ and $\hat R_2$ do not commute, the position of the oscillating particle cannot be determined. However,
1177: if the system is in one of the basis states $v(\theta;r_1,r_2)$, the above implies that the 
1178: particle can be found in four possible positions, with coordinates
1179: $\sqrt{\frac{\hbar}{m\omega}}(\pm \sqrt{r_1+\theta},\pm \sqrt{r_2+\theta})$ (independent $\pm$-signs).
1180: If we continue in this interpretation, where the positions are determined through measuring the eigenvalues of
1181: $\hat R_1^2$ and $\hat R_2^2$ in their common eigenstates, 
1182: then the possible positions of the particle in the Hilbert space $V(p)$ consists of a collection
1183: of such coordinates $\sqrt{\frac{\hbar}{m\omega}}(\pm \sqrt{r_1+\theta},\pm \sqrt{r_2+\theta})$. 
1184: This leads to Figure~2.
1185: \begin{figure}[htb]
1186: \caption{Possible positions of the WQO, for $p=5$ and $p=50$ (in units of $\sqrt{\frac{\hbar}{m\omega}}$).}
1187: \begin{center}
1188: \includegraphics{fig-D2-p5.eps}\hskip 15mm
1189: \includegraphics{fig-D2-p50.eps}
1190: \end{center}
1191: \end{figure}
1192: Thus the particle is situated on one of two circles, one with radius $\rho_0=\sqrt{\frac{p\hbar}{m\omega}}$ and 
1193: one with radius $\rho_1=\sqrt{\frac{(p+1)\hbar}{m\omega}}$. The spectrum of the coordinate operators is
1194: discrete, however. 
1195: 
1196: The current situation is interesting, since the classical limit of the WQO in this Hilbert space $V(p)$
1197: can be considered. For this purpose, one should examine the situation where 
1198: the quantum numbers of the system become very large (in the limit to infinity) and $\hbar$ becomes
1199: very small (in the limit to~0). So let us deal with the situation where 
1200: \begin{equation}
1201: p\rightarrow +\infty, \qquad \hbar \rightarrow 0, \qquad\hbox{but}\qquad p\hbar \rightarrow C,
1202: \label{lim}
1203: \end{equation}
1204: for some constant $C$.
1205: We shall examine what happens to the spectrum of the physical operators $\hat H$, $\hat R_1$, $\hat R_2$ and 
1206: $\hat R_1^2+\hat R_2^2$ under this limit. Using~(\ref{lim}),
1207: the energy of the system becomes constant, $C\omega$, as the two energy levels
1208: coincide under the limiting process. 
1209: The limit of the spectrum of $\hat R_1$ and $\hat R_2$ is the interval
1210: $[-\sqrt{\frac{C}{m\omega}},\sqrt{\frac{C}{m\omega}}]$.
1211: The eigenvalue of $\hat R_1^2+\hat R_2^2$ becomes $\frac{C}{m\omega}$, a fixed constant.
1212: So the particle is situated on a circle with radius $\sqrt{\frac{C}{m\omega}}$. 
1213: In the limit, all possible positions on the circle are possible. 
1214: This suggests that under this limit the solution of the WQO corresponds to one of the classical solutions
1215: \begin{equation}
1216: x(t)=A\cos(\omega t), \qquad y(t)=A\sin(\omega t)
1217: \end{equation}
1218: of a 2-dimensional simple homogeneous harmonic oscillator, determined by the differential equations
1219: \begin{equation}
1220: \ddot{x}(t) + \omega^2\, x(t) =0, \qquad \ddot{y}(t) + \omega^2\, y(t) =0,
1221: \end{equation}
1222: where $\omega^2=\frac{k}{m}$ ($k$ being the ``spring constant''). 
1223: The classical energy of such a system is $m\omega^2A^2$, and the classical
1224: oscillator moves on a circle with radius $A$. Under the identification of the constants $C=m\omega A^2$, 
1225: both the energy and position of the WQO (for this solution) are in agreement with those of 
1226: the particular solution of the classical oscillator. In other words, the ``correspondence principle''
1227: holds for this solution of the WQO.
1228: 
1229: It should be clear that the above considerations about possible particle positions
1230: remain valid for arbitrary $D$. For instance, when $D=3$, the possible positions
1231: of the particle in one of the basis states of $V(p)$ consists of a collection of coordinates 
1232: $\sqrt{\frac{\hbar}{m\omega}}(\pm \sqrt{r_1+\theta},\pm \sqrt{r_2+\theta},\pm \sqrt{r_3+\theta})$ 
1233: (independent $\pm$-signs). The collection of such coordinates leads to Figure~3,
1234: i.e.\ a set of points on two concentric spheres, one with radius $\sqrt{\frac{p\hbar}{2m\omega}}$
1235: and one with radius $\sqrt{\frac{(p+1)\hbar}{2m\omega}}$.
1236: \begin{figure}[htb]
1237: \caption{Possible positions of the 3-dimensional WQO, for $p=5$ (in units of $\sqrt{\frac{\hbar}{2m\omega}}$).
1238: In the left figure, the set of points is on a sphere with radius $\sqrt{p+1}$; 
1239: in the right figure, they are on a sphere with radius $\sqrt{p}$.}
1240: \begin{center}
1241: \includegraphics{fig3D1b.eps}\hskip 15mm
1242: \includegraphics{fig3D2b.eps}
1243: \end{center}
1244: \end{figure}
1245: 
1246: As far as the position or momentum operators for the WQO is concerned, we have not yet considered their
1247: time dependence. Following~\cite{Palev5}-\cite{K2}, this is given by
1248: \begin{equation}
1249: \hat R_k(t)=\sqrt{\frac{\hbar}{(D-1)m\omega}}(A^+_k e^{-i\omega t} + A^-_k e^{i\omega t}), \quad
1250: \hat P_k(t)=-i\sqrt{\frac{m\omega\hbar}{(D-1)}}(A^+_k e^{-i\omega t} - A^-_k e^{i\omega t}), 
1251: \end{equation}
1252: for the $D$-dimensional single particle case.
1253: For a general mixed state of $V(p)$,
1254: \begin{equation}
1255: |x\rangle = \sum_{\theta,r} c_{\theta;r} v(\theta;r),\qquad
1256: \sum_{\theta,r} c_{\theta;r}^2=1,
1257: \end{equation}
1258: one finds that the average value or mean trajectories of the coordinate and momentum operator components is
1259: \begin{equation}
1260: \langle x | \hat R_k | x\rangle =
1261: 2 \sqrt{\frac{\hbar}{(D-1)m\omega}} B_k \cos(\omega t), \qquad
1262: \langle x | \hat P_k | x\rangle =
1263: -2 \sqrt{\frac{m\omega\hbar}{D-1}} B_k \sin(\omega t),
1264: \end{equation}
1265: where $B_k$ is the constant
1266: \begin{equation}
1267: B_k= \sum_r \sqrt{r_k} c_{0;r_1,\ldots,r_D} c_{1;r_1,\ldots,r_k-1,\ldots,r_D}.
1268: \end{equation}
1269: If the system is in a stationary state (fixed energy eigenvalue), one has either all $c_{0,r}=0$ or else
1270: all $c_{1;r}=0$. In that case, all $B_k$ are zero. In other words, for stationary states $|x\rangle$ the
1271: average values $\langle x | \hat R_k | x\rangle$ and $\langle x | \hat P_k | x\rangle$ vanish.
1272: 
1273: The analysis given so far was for a single particle $D$-dimensional WQO in the representation space $V(p)$
1274: of $\gl(1|D)$. One can also consider the same representations 
1275: $V(p)$ of $gl(1|N)$ 
1276: as state spaces of a system of $N$ 1-dimensional
1277: WQO's ($N>1$, since a single 1-dimensional WQO does not have a solution in terms of $\gl(1|1)$).
1278: In this second interpretation, the basis states
1279: \begin{equation}
1280: v(\theta;r)\equiv v(\theta;r_1,r_2, \ldots,r_N),\qquad
1281: \theta\in\{0,1\},\ r_i\in\{0,1,2,\ldots\},\ \hbox{ and } \theta+r_1+\cdots+r_N= p.
1282: \label{newbasis3}
1283: \end{equation}
1284: are the same as in~(\ref{newbasis2}), and similarly
1285: \begin{equation}
1286: \hat{H} v(\theta;r) = \hbar\omega (\frac{p}{N-1}+\theta) v(\theta;r), \qquad
1287: \hat R^2_i v(\theta;r) = \frac{\hbar}{(N-1)m\omega} (r_i+\theta) v(\theta;r).
1288: \end{equation}
1289: In other words, all the previously obtained formulas remain valid, only the interpretation of the
1290: system is different: now it consists of $N$ one-dimensional oscillating particles, and $\hat R_i$ is
1291: the operator corresponding to the position of particle~$i$.
1292: In the representation space $V(p)$, the spectrum of the operator $\hat R_i^2$ is $\{0,1,\ldots,p\}$
1293: and that of $\hat R_i$ is $\{0,\pm{1},\pm\sqrt{2},\ldots, \pm\sqrt{p}\}$. 
1294: Let us again assume that $N=2$, so that we are dealing with
1295: %Let us again assume that $D=2$, i.e.\ we are dealing with 
1296: just two one-dimensional oscillating
1297: particles. If the system is in one of the basis states $v(\theta;r_1,r_2)$, our analysis
1298: leads to the interpretation that the position of the first one-dimensional oscillator is 
1299: $\pm \sqrt{\frac{\hbar(r_1+\theta)}{m\omega}}$, and that of the second one-dimensional oscillator 
1300: is $\pm \sqrt{\frac{\hbar(r_2+\theta)}{m\omega}}$. Since $\theta+r_1+r_2=p$, these positions
1301: are not independent but ``correlated''. 
1302: To see this from a different point of view, 
1303: recall that the positions of the two oscillators cannot be measured simultaneously, since 
1304: $[\hat R_1,\hat R_2]\ne 0$. But suppose the system is in an arbitrary state of $V(p)$, and the position
1305: of the first oscillator is measured at time~$t=0$. 
1306: This measurement has one of the eigenvalues $\pm \sqrt{\frac{\hbar r_1}{m\omega}}$ of $\hat R_1(t)$
1307: as an outcome, with $r_1\in\{0,1,\ldots,p\}$.
1308: As a consequence of the measurement, the system is then in an eigenstate of $\hat R_1$ for this eigenvalue.
1309: It is easy to verify that this eigenstate is unique, given by (we assume $r_1>0$)
1310: \begin{equation}
1311: \frac{1}{\sqrt{2}} v(0;r_1, r_2) \pm \frac{1}{\sqrt{2}} v(1; r_1-1, r_2),
1312: \end{equation}
1313: with $r_2=p-r_1$ (the plus sign for the positive eigenvalue,
1314: and the minus sign for the negative eigenvalue). 
1315: The time evolution of this system is then described by
1316: \begin{equation}
1317: |x\rangle = \frac{e^{i\omega t/2}}{\sqrt{2}} v(0;r_1, r_2) \pm \frac{e^{-i\omega t/2}}{\sqrt{2}} v(1; r_1-1, r_2).
1318: \label{state_x}
1319: \end{equation}
1320: Note that $|x\rangle$ is an eigenstate of $\hat R_1(t)$ for the eigenvalue 
1321: $\pm \sqrt{\frac{\hbar r_1}{m\omega}}$.
1322: The state~(\ref{state_x}) can be rewritten as follows:
1323: \begin{eqnarray}
1324: |x\rangle &=& \frac{1}{2\sqrt{2}} [e^{i\omega t/2}v(0;r_1,r_2)+e^{-i\omega t/2}v(1;r_1,r_2-1)]\nn\\ 
1325: &&+ \frac{1}{2\sqrt{2}} [e^{i\omega t/2}v(0;r_1,r_2)- e^{-i\omega t/2}v(1;r_1,r_2-1)] \nn\\
1326: && \pm \frac{1}{2\sqrt{2}} [e^{i\omega t/2}v(0;r_1-1,r_2+1)+e^{-i\omega t/2}v(1;r_1-1,r_2)]\nn\\
1327: &&\mp \frac{1}{2\sqrt{2}} [e^{i\omega t/2}v(0;r_1-1,r_2+1)- e^{-i\omega t/2}v(1;r_1-1,r_2)].
1328: \end{eqnarray}
1329: The four linear combinations in square brackets are four eigenvectors
1330: of $\hat R_2(t)$, with eigenvalues $\sqrt{\frac{\hbar}{m\omega}}\sqrt{p-r_1}$, 
1331: $-\sqrt{\frac{\hbar}{m\omega}}\sqrt{p-r_1}$, $\sqrt{\frac{\hbar}{m\omega}}\sqrt{p-r_1+1}$ and 
1332: $-\sqrt{\frac{\hbar}{m\omega}}\sqrt{p-r_1+1}$ respectively.
1333: This implies that a measurement of the position of the first oscillator, i.e.\ of 
1334: $\hat R_1$ at time~$t=0$ with outcome $\sqrt{\frac{\hbar r_1}{m\omega}}$, imposes strong
1335: restrictions on the position eigenvalues of the second oscillator, measured at time $t$. 
1336: The two one-dimensional oscillators are not free but correlated in some sense.
1337: 
1338: Let us also examine the classical limit~(\ref{lim}) for this second interpretation.
1339: In fact, the limit of the spectrum of the operators $\hat H$, $\hat R_1$, $\hat R_2$ and 
1340: $\hat R_1^2+\hat R_2^2$ is the same as before. So the total
1341: energy of the system becomes constant, $C\omega$.
1342: The limit of the spectrum of $\hat R_1$ and of $\hat R_2$ is the interval
1343: $[-\sqrt{\frac{C}{m\omega}},\sqrt{\frac{C}{m\omega}}]$, and
1344: the eigenvalue of $\hat R_1^2+\hat R_2^2$ is fixed, $\frac{C}{m\omega}$.
1345: Thus, under this limit the solution of the 2-particle 1-dimensional WQO corresponds to
1346: the classical solution
1347: \begin{equation}
1348: x_1(t)=A\cos(\omega t), \qquad x_2(t)=A\sin(\omega t).
1349: \end{equation}
1350: This simply describes a system of two uncoupled identical 1-dimensional harmonic oscillators,
1351: but with a fixed phase difference in their oscillations. In other words, the initial conditions
1352: of the two oscillators are related: also here there is a correlation.
1353: So the ``correspondence principle'' also holds for this interpretation of the WQO.
1354: 
1355: \setcounter{equation}{0}
1356: \section{Analysis for a general example}
1357: 
1358: The results of the previous sections were mainly devoted to the special representations $V(p)$.
1359: However, the computations of Section~2 allow us to make an analysis of the WQO properties for
1360: any unitary representation. 
1361: As an example, we consider a representation $W([m]_{3N+1})$ which
1362: belongs to neither the set $W(p)$ nor the set $V(p)$,
1363: namely the representation corresponding to the $\gl(1|3N)$ 
1364: irreducible representation $V^{\{2,1\}}_{\gl(1|3N)}$ with highest weight 
1365: $\Lambda=2\epsilon+\delta_1$. This can be identified 
1366: with the $\gl(1|3N)$ module $W([m]_{3N+1})$ with $[m]_{3N+1}=[2,1,0,\ldots,0]$.
1367: 
1368: The branching from $\gl(1|3N)$ to $\gl(1)\otimes \gl(3N)$ is given by
1369: \begin{equation}
1370:      V^{\{2,1\}}_{\gl(1|3N)}\rightarrow 
1371:      V^{2}_{\gl(1)}\times V^{\{1\}}_{\gl(3N)} 
1372:      +V^{1}_{\gl(1)}\times V^{\{2\}}_{\gl(3N)} 
1373:      +V^{1}_{\gl(1)}\times V^{\{1,1\}}_{\gl(3N)} 
1374:      +V^{0}_{\gl(1)}\times V^{\{2,1\}}_{\gl(3N)} \,.
1375: \label{V21}
1376: \end{equation} 
1377: The four terms arising here are associated with the 
1378: four sets of states:
1379: \begin{equation}
1380: {\scriptsize
1381: \left|
1382: \begin{array}{lcllll}
1383: 2 & 1 & 0 & 0 & \cdots & 0  \\
1384:   & 1 & 0 & 0 & \cdots & 0  \\
1385:   & * & * & 0 & \cdots &  \\
1386:   &\vdots & \vdots &\qdots & & \\
1387: & * & * & & & \\
1388: & * & & & &
1389: \end{array}
1390: \right)
1391: \quad
1392: \left|
1393: \begin{array}{lcllll}
1394: 2 & 1 & 0 & 0 & \cdots & 0  \\
1395:   & 1 & 1 & 0 & \cdots & 0  \\
1396:   & * & * & 0 & \cdots &    \\
1397:   &\vdots & \vdots & \qdots & & \\
1398: & * & * & & & \\
1399: & * & & & &
1400: \end{array}
1401: \right)
1402: \quad
1403: \left|
1404: \begin{array}{lcllll}
1405: 2 & 1 & 0 & 0 & \cdots & 0  \\
1406:   & 2 & 0 & 0 & \cdots & 0  \\
1407:   & * & * & 0 & \cdots &   \\
1408:   &\vdots & \vdots & \qdots  & & \\
1409: & * & * & & & \\
1410: & * & & & &
1411: \end{array}
1412: \right)
1413: \quad
1414: \left|
1415: \begin{array}{lcllll}
1416: 2 & 1 & 0 & 0 & \cdots & 0  \\
1417:   & 2 & 1 & 0 & \cdots & 0  \\
1418:   & * & * & 0 & \cdots &    \\
1419:   &\vdots & \vdots & \qdots & & \\
1420: & * & * & & &\\
1421: & * &   & & &
1422: \end{array}
1423: \right)
1424: }
1425: \label{21m}
1426: \end{equation}
1427: with the entries $*$ taking values determined by the various 
1428: betweennness conditions. 
1429: For all states within each of these four sets the corresponding 
1430: values of $(\theta_1,\theta_2,\ldots)$ are given by
1431: $(0,0,0,\dots,0)$, $(0,1,0,\ldots,0)$, $(1,0,0,\ldots,0)$ and $(1,1,0,\dots,0)$,
1432: respectively. In the notation of (\ref{Eq}) we have
1433: \begin{equation}
1434:    E_q=\hbar \omega \left( \frac{6N+1}{3N-1} -q \right)
1435: \label{Eq21}
1436: \end{equation}
1437: with $q=0$, $1$, $1$ and $2$, for our four sets of states. It follows that
1438: we have just three distinct energy levels with degeneracies as shown 
1439: below:
1440: \begin{equation}
1441: \begin{array}{ll}
1442:    E_0=\hbar\omega \ds\frac{6N+1}{3N-1}\,,&\quad d_0=3N\,;\cr\cr
1443:    E_1=\hbar\omega \ds\frac{3N+2}{3N-1}\,,&\quad d_1=9N^2\,;\cr\cr
1444:    E_2=\hbar\omega \ds\frac{3}{3N-1}\,,   &\quad d_2=N(9N^2-1)\,.\cr
1445: \end{array}   
1446: \label{E012}
1447: \end{equation}
1448: These correspond precisely to the three $\gl(1)$ representations, $V^{2}_{\gl(1)}$, 
1449: $V^{1}_{\gl(1)}$ and $V^{0}_{\gl(1)}$ appearing in (\ref{V21}). The degeneracies,
1450: $d_q$ have been obtained by using $N$-dependent formulae for the dimensions of 
1451: the irreducible representations of $\gl(3N)$ that appear in (\ref{V21}). 
1452: 
1453: The spectrum of $\hR^2_{\a i}$ may be obtained from (\ref{actionR2}) by 
1454: setting $n=3N$ and applying the formula to the above states (\ref{21m}).
1455: One finds that for states of these four types the factor 
1456: $(m_{0,n+1}+\ldots+m_{n,n+1})-(m_{1,n}+\cdots+m_{n,n})$ takes
1457: the values $2$, $1$, $1$ and $0$, respectively, while
1458: the second factor $(m_{1,k}+\cdots+m_{k,k})-(m_{1,k-1}+\cdots+m_{k-1,k-1})$
1459: lies in the sets $\{0,1\}$, $\{0,1\}$, $\{0,1,2\}$ and $\{0,1,2\}$,
1460: respectively. It follows that for each of the possible energy 
1461: levels $E_q$ the eigenvalues of $\hR^2_{\a i}$ are given by
1462: \begin{equation}
1463:    R^2_{\a i}=\frac{\hbar\ r^2_{\a i}}{(3\,N-1)\,m\,\omega}
1464:    \hbox{~~with~~} r^2_{\a i} = \bigg\{
1465:    \begin{array}{ll}
1466:        2,3&\hbox{~for~$q=0$}\,;\cr
1467:        1,2,3&\hbox{~for~$q=1$}\,;\cr
1468:        0,1,2&\hbox{~for~$q=2$}\,.\cr
1469:    \end{array}
1470: \label{R012}
1471: \end{equation}
1472: Note that, unlike all the $V(p)$ cases with $p>0$, there exists a configuration with
1473: $r_{\a i}^2=0$, namely one of the $q=2$ ground state configurations.
1474:  
1475: To determine the angular momentum content of these states, we can
1476: extend (\ref{V21}) by considering the chain of subalgebras:
1477: \begin{equation}
1478:     \gl(3N)\rightarrow \gl(3) \oplus \gl(N)
1479: \rightarrow \so(3) \oplus \gl(N)
1480: \rightarrow \so(3).
1481:      \label{gl3n}
1482: \end{equation} 
1483: The corresponding branchings take the form:
1484: 
1485: \begin{eqnarray}
1486: \hskip-1cm
1487: &V^{\{1\}}_{\gl(3N)}
1488: &\rightarrow V^{\{1\}}_{\gl(3)} \times V^{\{1\}}_{\gl(N)}\cr
1489: \hskip-1cm
1490: &&\rightarrow V^{1}_{\so(3)} \times V^{\{1\}}_{\gl(N)}\cr
1491: \hskip-1cm
1492: &&\rightarrow N\, V^{1}_{\so(3)}\\ \cr 
1493: %%
1494: \hskip-1cm
1495: &V^{\{2\}}_{\gl(3N)}
1496: &\rightarrow V^{\{2\}}_{\gl(3)} \times V^{\{2\}}_{\gl(N)}+
1497: V^{\{1,1\}}_{\gl(3)} \times V^{\{1,1\}}_{\gl(N)}\cr
1498: \hskip-1cm
1499: &&\rightarrow (V^{2}_{\so(3)}+V^{0}_{\so(3)})\times V^{\{2\}}_{\gl(N)}  
1500:   +V^{1}_{\so(3)}\times V^{\{1,1\}}_{\gl(N)}\cr
1501: \hskip-1cm
1502: &&\rightarrow  \frac12\, N(N+1)\,V^{2}_{\so(3)}
1503: + \frac12 N(N-1)\,V^{1}_{\so(3)}
1504: +\frac12 N(N+1)\,V^{0}_{\so(3)} \\ \cr
1505: %%
1506: \hskip-1cm
1507: &V^{\{1,1\}}_{\gl(3N)}
1508: &\rightarrow V^{\{2\}}_{\gl(3)} \times V^{\{1,1\}}_{\gl(N)}+
1509: V^{\{1,1\}}_{\gl(3)} \times V^{\{2\}}_{\gl(N)}\cr
1510: \hskip-1cm
1511: &&\rightarrow (V^{2}_{\so(3)}+V^{0}_{\so(3)})\times V^{\{1,1\}}_{\gl(N)}  
1512:   +V^{1}_{\so(3)}\times V^{\{2\}}_{\gl(N)}\cr 
1513: \hskip-1cm
1514: &&\rightarrow  \frac12 N(N-1)\,V^{2}_{\so(3)}
1515: + \frac12 N(N+1)\,V^{1}_{\so(3)}+\frac12 N(N-1)\,V^{0}_{\so(3)} \\ \cr
1516: %%
1517: \hskip-1cm
1518: &V^{\{2,1\}}_{\gl(3N)}
1519: &\rightarrow V^{\{3\}}_{\gl(3)} \times V^{\{2,1\}}_{\gl(N)}
1520: +V^{\{2,1\}}_{\gl(3)}\times (V^{\{3\}}_{\gl(N)}+V^{\{2,1\}}_{\gl(N)}+ V^{\{1,1,1\}}_{\gl(N)})
1521: +V^{\{1,1,1\}}_{\gl(3)}\times V^{\{2,1\}}_{\gl(N)}\cr
1522: \hskip-1cm
1523: &&\rightarrow (V^{3}_{\so(3)}+ V^{1}_{\so(3)})\times V^{\{2,1\}}_{\gl(N)}+
1524: (V^{2}_{\so(3)}+ V^{1}_{\so(3)})\times (V^{\{3\}}_{\gl(N)}+V^{\{2,1\}}_{\gl(N)}+ V^{\{1,1,1\}}_{\gl(N)})\cr 
1525: \hskip-1cm
1526: &&\qquad + V^{0}_{\so(3)}\times V^{\{2,1\}}_{\gl(N)}\cr
1527: \hskip-1cm
1528: &&\rightarrow \frac13 N(N^2-1)\,V^{3}_{\so(3)}+\frac13 N(2N^2+1)\,V^{2}_{\so(3)}
1529: + N^3\,V^{1}_{\so(3)}+ \frac13 N(N^2-1)\,V^{0}_{\so(3)} \,.
1530: %% 
1531: \label{br3N}  
1532: \end{eqnarray}
1533: % 
1534: Combining these results with (\ref{V21}) we can write down the generating function
1535: \begin{eqnarray}
1536: G(Q,J)_N &=& \sum_{q,j=0}^\infty n_{q,j}(N)\, Q^q\,J^j \cr
1537: & =& N\,J\,Q^0 + N^2\,(1+J+J^2)\,Q^1  \cr 
1538: && + \left(\frac13 N(N^2-1)+ N^3\,J+\frac13 N(2N^2+1)\,J^2
1539: + \frac13 N(N^2-1)\,J^3 \right)\,Q^2
1540: \label{GQJ}
1541: \end{eqnarray}
1542: where the coefficient $n_{q,j}(N)$ gives the multiplicity 
1543: of states of angular momentum $j$ and energy $E_q$ as a function of $N$.  
1544: To recover the degeneracy of each energy level one replaces each $J^j$ by 
1545: the corresponding dimension $2j+1$ of $V^j_{\so(3)}$ to give
1546: \begin{equation}
1547: G(Q)_N= 3N\,Q^0+9N^2\,Q^1+ N(9N^2-1)\,Q^2,
1548: \label{GQ}
1549: \end{equation} 
1550: in agreement with (\ref{E012}).
1551: Similarly, the full angular momentum content for specific values of $N$ may
1552: be obtained from~(\ref{GQJ}) by setting $Q=1$. For example, we obtain
1553: \begin{eqnarray}
1554: G(1,J)_{N=1} &=& 1+3J+2J^2; \\
1555: G(2,J)_{N=2} &=& 6+14J+10J^2+2J^3.
1556: \end{eqnarray}
1557: 
1558: Although this example is not very remarkable, it illustrates how the methods 
1559: constructed here can be used, and it shows that the material developed in this
1560: paper can in principle be applied to any unitary representation or class of unitary representations.
1561: 
1562: \setcounter{equation}{0}
1563: \section{Conclusions}
1564: 
1565: The theory of WQOs has a deep connection with representation theory of Lie superalgebras.
1566: Solutions of WQO systems can be considered not only for the Lie superalgebra $\ssl(1|n)$,
1567: but also for other Lie superalgebras. As typical examples, we mention the $\osp(3|2)$ 
1568: WQO~\cite{Palev6}
1569: and the recently studied $\ssl(3|N)$ WQO~\cite{Palev7}. 
1570: The physical properties of the WQO depend on the Lie superalgebra, and on the class of
1571: representations considered.
1572: 
1573: This paper has made a contribution to the study of physical properties of WQOs of type $\ssl(1|n)$,
1574: initiated in~\cite{Palev1} and performed in detail for the 3-dimensional WQO 
1575: in~\cite{Palev5}-\cite{K2}.
1576: First of all, our current approach is slightly more general by considering the $D$-dimensional
1577: $N$-particle WQO of type $\ssl(1|n)$.
1578: Secondly, in earlier papers only the solutions corresponding to Fock representations $W(p)$
1579: were investigated.
1580: Here, we have initiated the study of solutions related to all (unitary) representations of $\ssl(1|DN)$.
1581: This requires the construction of a proper basis (the GZ-basis) for these representations, with
1582: explicit actions of the $\ssl(1|DN)$ generators, as given here in Section~2.
1583: 
1584: In order to emphasize the difference with the previously  considered Fock spaces $W(p)$, we have 
1585: paid  particular attention to another special class of representations $V(p)$. 
1586: These representations are, once again, weight multiplicity free, 
1587: so they are easy to describe explicitly. The energy spectrum and the spectra of
1588: position and momentum operators in these representations have been  obtained without much effort. 
1589: Among the special features we mention: there are only two different energy levels (the difference 
1590: being $\hbar\omega$); the position operator components are non-commuting operators and have a discrete
1591: spectrum, 
1592: leading to
1593: spatial properties that are quite different from those of a canonical
1594: quantum oscillator.
1595: 
1596: In addition, for these representations $V(p)$ the complete angular momentum content has
1597: been obtained in the form of a generating function using powerful group theoretical methods.
1598: 
1599: Another particularly nice result is that, for those solutions related to
1600: $V(p)$, a classical limit has been obtained,
1601:  showing that the behavior of the WQO solution reduces to a classical solution
1602: of the harmonic oscillator. 
1603: For the previously considered Fock space
1604: solutions based on $W(p)$, such a limit does not exist.
1605: 
1606: 
1607: Wigner Quantum Oscillators, belonging to the class of non-canonical quantum systems, 
1608: possess many attractive and fascinating properties, both for mathematicians and physicists. 
1609: We believe that the correspondence
1610: principle observed here makes them even more interesting for physics.
1611: 
1612: \section*{Acknowledgments}
1613: The authors would like to thank Professor T.D. Palev for his interest.
1614: NIS was supported by a project from the Fund for Scientific Research -- Flanders (Belgium).
1615: In addition, RCK is grateful for the
1616: award of a Leverhulme Emeritus Fellowship in support of this work.
1617: 
1618: \begin{thebibliography}{99}
1619: 
1620: \bibitem{Kac1}
1621: Kac V G 1977 {\it Adv.\ Math.} {\bf 26} 8
1622: \bibitem{Kac2}
1623: Kac V G 1978 {\it Lect.\ Notes Math.} {\bf 676} 597 
1624: \bibitem{Palev2}
1625: Palev T D 1989 {\it J.\ Math.\ Phys.} {\bf 30} 1433
1626: \bibitem{Wigner}
1627: Wigner E P 1950 {\it Phys. Rev.} {\bf 77} 711 
1628: \bibitem{Palev1}
1629: Palev T D 1982 {\it J.\ Math.\ Phys.} {\bf 23} 1778; 1982 {\it Czech.\ J.\ Phys.} 
1630: {\bf B32} 680
1631: \bibitem{Palev4}
1632: Kamupingene A H, Palev T D  and Tsaneva S P 1986 {\it J.\ Math.\ Phys.} {\bf 27} 2067
1633: \bibitem{Palev5}
1634: Palev T D and Stoilova N I 1997 {\it J.\ Math.\ Phys.}
1635:         {\bf 38} 2506 
1636: \bibitem{K1}
1637: King R C, Palev T D, Stoilova N I and Van der Jeugt J 2003 
1638: {\it J.\ Phys.\ A: Math.\ Gen.} {\bf 36} 4337 
1639: \bibitem{K2}
1640: King R C, Palev T D, Stoilova N I and Van der Jeug J 2003 
1641: {\it J.\ Phys.\ A: Math.\ Gen.} {\bf 36} 11999
1642: \bibitem{SJ}
1643: Stoilova N I and Van der Jeugt J 2005
1644: {\it J.\ Phys.\ A: Math.\ Gen.} {\bf 38} 9681
1645: \bibitem{Gould} 
1646: Gould M D and Zhang R B  1990 {\it J.\ Math.\ Phys.} {\bf 31} 2552 
1647: \bibitem{Palev3}
1648: Palev T D 1980 {\it J.\ Math.\ Phys.} {\bf 21} 1293 
1649: \bibitem{HS}
1650: Hatzinikitas A and Smyrakis I {\it J.\ Math.\ Phys.} 2002
1651: {\bf 43}, 113 
1652: \bibitem{C}
1653: Connes A 1994 {\it Non-commutative geometry} (San Diego: Academic Press)
1654: \bibitem{NP}
1655: Nair V P and Polychronakos A P 2001 {\it Phys.\ Lett.} {\bf B 505} 267
1656: \bibitem{SS1}
1657: Smailagic  A and Spallucci E 2002 {\it Phys.\ Rev.}
1658: {\bf D 65}: 107701
1659: \bibitem{SS2}
1660: Smailagic A and Spallucci E  2002 {\it J.\ Phys.\ A}
1661: {\bf  35}, L363 
1662: \bibitem{MM}
1663: Mathukumar  B and Mitra P 2002 {\it Phys.\ Rev.}
1664: {\bf D 66}: 027701 
1665: \bibitem{Molien}
1666: Molien T 1897 {\it Uber die Invarianten der linearen Substitutionsgruppe}
1667: (Sitzungsber. Konigl. Preuss. Akad. Wiss.) 1152
1668: \bibitem{Sturmfels}
1669: Sturmfels B 1993 {\it Algorithms in Invariant Theory}
1670: (Wien: Springer-Verlag) p. 188
1671: \bibitem{LJ} 
1672: Bernstein I N and Leites D A 1980 {\it C.R. Acad. Bulg. Sci.} {\bf 33} 1049
1673: \bibitem{JHKR}
1674: Van der Jeugt J, Hughes J W B, King R C and Thierry-Mieg J 1990 {\it 
1675: Commun.\ Alg.} {\bf 18} 3453 
1676: \bibitem{GZ}
1677: Gel'fand I M and Zetlin M L 1950 {\it Dokl.\ Akad.\ Nauk SSSR} {\bf 71} 825
1678: \bibitem{JHKT}
1679: Van der Jeugt J, Hughes J W B, King R C and Thierry-Mieg J 1990
1680: {\it J\ Math\ Phys} {\bf 31} 2278
1681: \bibitem{Schur}
1682: SCHUR, an interactive program for calculating the properties of Lie groups and 
1683: symmetric functions, distributed by S Christensen. 
1684: \bibitem{K3}
1685: King R C 1975 {\it J.\ Phys. A: Math. Gen.} {\bf 8} 429 
1686: \bibitem{RRS}
1687: Raychev P P, Roussev, R P and Smirnov Yu F 1991
1688: {\it J. Phys. A: Math. Gen.} {\bf 24}  2943
1689: \bibitem{Palev6}
1690: Palev T D and Stoilova N I 1994 {\it J.\ Phys. A}
1691:         {\bf 27} 977; 7387
1692: \bibitem{Palev7}
1693: Palev T D 2006 $SL(3|N)$ Wigner quantum oscillator: examples of ferromagnetic-like
1694: oscillators with noncommutative, square-commutative geometry {\it Preprint} hep-th/0601201  
1695: 
1696: 
1697:         
1698: \end{thebibliography}
1699: \end{document}