1: \documentclass[12pt]{article}
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: \textheight 22.0cm \textwidth 16cm \tolerance=10000
4: \baselineskip=24pt \topmargin=-0.15cm \oddsidemargin=0cm
5: \renewcommand{\baselinestretch}{1.3}
6: \def\bt{\bibitem}
7: %\def\ii{\'{\i}}
8: \def\beq{\begin{equation}}
9: \def\eeq{\end{equation}}
10: \def\beqa{\begin{eqnarray}}
11: \def\eeqa{\end{eqnarray}}
12: \def\ban{\begin{eqnarray*}}
13: \def\ean{\end{eqnarray*}}
14: \def\bi{\begin{itemize}}
15: \def\ei{\end{itemize}}
16: %\def\bm{\fam9}
17: %\def\d{\mbox{d}}
18:
19:
20: \begin{document}
21: %
22: %
23: %
24: \begin{center}
25: \Large{\bf Applications of the Quon Algebra: 3-D Harmonic
26: Oscillator and the Rotor Model} \vspace{1.0cm}
27:
28: {S.S. Avancini, J.R. Marinelli and C. E. de O. Rodrigues} \\
29: {Depto de F\'{\i}sica - CFM - Universidade Federal de Santa Catarina \\
30: Florian\'{o}polis - SC - CP. 476 - CEP 88.040 - 900 - Brazil}
31: \end{center}
32: \footnote{E-mail address: ricardo@fsc.ufsc.br}
33: %
34: \begin{abstract}
35: {In this work we present a method to build in a systematic way a
36: many-body quon basis state. In particular, we show a closed
37: expression for a given number N of quons, restricted to the
38: permutational symmetric subspace, which belongs to the whole
39: quonic space. The method is applied to two simple problems: the
40: three-dimensional harmonic oscillator and the rotor model and
41: compared to previous quantum algebra results. The differences
42: obtained and possible future applications are also discussed.}
43: \end{abstract}
44: %
45: \vspace{0.50cm} \noindent PACS number(s): 03.65.Fd: 21.60.Fw \\
46: keywords: quon algebra, quonic oscillator, rotor model.
47: %
48: %
49: \newpage
50: \vspace{1.0cm}
51:
52: \section{Introduction}
53:
54: \bigskip\
55:
56: Quons are particles that violate statistics by a small amount, which is
57: controlled by a single parameter q\cite{GreenPRD},\cite{Green6/7/2000}. The
58: range of variation of the parameter is between -1 and +1, and the limits of
59: the interval correspond respectively to fermionic an bosonic statistics. The
60: particular commutation relations obeyed by quons define an algebra (the so
61: called quon algebra), which, for a single degree of freedom, gives results
62: very similar to the ones obtained using deformed (or quantum) algebras\cite
63: {Chaichian}, once we keep the q interval as above defined. For more than a
64: single degree of freedom however, there are some important differences
65: between both algebras. One consequence of those differences is that it is
66: possible to define quonic operators that behave as irreducible su(2) tensors%
67: \cite{PhysLetA}. In other words, it is possible to assume that the quon
68: algebra follow the usual angular momentum coupling rules. This last result
69: opens up the possibility of applying the quon algebra to the study of
70: many-body systems, with some important technical advantages over deformed
71: algebras. However, those very same differences also introduce some
72: complications when we try to build many-body quonic states. The ''q-mutation
73: '' relation that defines the quon algebra is given by $[a_i,a_j^{\dagger
74: }]_q=a_ia_j^{\dagger }-qa_j^{\dagger }a_i=\delta _{ij}$ , with the
75: additional condition $a_i|0>=0$ and $|0>$ being the quon vacuum state.
76: Unlike quantum algebras, this relation prevent us to establish any
77: commutation relation between two creation or two annihilation operators,
78: unless q is either $+1$ or $-1$. Although no such a rule is needed to
79: calculate vacuum matrix elements of polynomials in the $a^{\prime }s$ or $%
80: a^{\dagger \prime }s$ \cite{Greenphysica}, the lack of those commutation
81: relations introduces the necessity to enlarge the basis whenever we try to
82: define a many-quon state, as we discuss latter in this paper.
83:
84: It is then our intention here to provide some basis to the use of the quon
85: algebra as a tool to improve our approximated descriptions of many-body
86: problems, and at the same time to pose some of the possible differences that
87: arise with the introduction of that algebra, as compared to the more usually
88: applied quantum algebras. In fact, the quonic Fock-like Hilbert space
89: contains all possible symmetry permutations of the polynomials in the $%
90: a^{\dagger \prime }s$ . As we shall see, even when we restrict ourselves to
91: the symmetric representation, some important differences between both
92: algebras can be noticed. Actually, a series of applications using a restrict
93: quonic subspace was already done in the context of boson mappings\cite{qbos1}%
94: ,\cite{qbos2}. In section 3, we make those differences more explicit here,
95: through the solution of the three-dimensional harmonic (quon) oscillator as
96: well as the analysis of the spectrum of a quonic version of the quantum
97: rotor. Of course, these two examples can be viewed as a simple laboratory to
98: test our main results in the construction of a many-body quon basis, as
99: discussed in section 2. We believe that other interesting applications can
100: be tackled in the future using our present results.
101:
102: \bigskip\
103:
104: \section{Many-Body Quon States}
105:
106: \bigskip\
107:
108: We start our discussion following the reasonings presented in reference\cite
109: {Green6/7/2000} for the case of a two quons state. In that case we may write
110: the following normalized states:
111:
112: \begin{equation}
113: \frac 1{\sqrt{1+q}}(a_1^{\dagger })^2|0>,~\frac
114: 1{\sqrt{1+q}}(a_2^{\dagger })^2|0>,~a_1^{\dagger }a_2^{\dagger
115: }|0> ~\mathrm{{and}~a_2^{\dagger }a_1^{\dagger }|0>.}
116: \label{2quons}
117: \end{equation}
118:
119: The last two states defined in (\ref{2quons}) can be expanded in terms of a
120: symmetric and a antisymmetric state in the form:
121:
122: \begin{equation}
123: a_1^{\dagger }a_2^{\dagger }|0>=\sqrt{\frac{1+q}2}|\phi _s>+\sqrt{\frac{1-q}2%
124: }|\phi _a>, \label{2sym}
125: \end{equation}
126:
127: \begin{equation}
128: a_2^{\dagger }a_1^{\dagger }|0>=\sqrt{\frac{1+q}2}|\phi _s>-\sqrt{\frac{1-q}2%
129: }|\phi _a>, \label{2anti}
130: \end{equation}
131:
132: where,
133: \[
134: |\phi _s>=\frac 1{\sqrt{2(1+q)}}(a_1^{\dagger }a_2^{\dagger }+a_2^{\dagger
135: }a_1^{\dagger })|0~>~~,~~~|\phi _a>=\frac 1{\sqrt{2(1-q)}}(a_1^{\dagger
136: }a_2^{\dagger }-a_2^{\dagger }a_1^{\dagger })|0~>
137: \]
138:
139: We may then conclude that the two-quon basis can be formed by one
140: antisymmetric and three symmetric states. Also, once any observable must be
141: represented by a symmetrical operator, the two states in equations (\ref
142: {2sym}) and (\ref{2anti}) should be considered the same, in the sense that
143: they give us the same observables. Another way to put this is to recognize
144: that we can obtain the two-quon (orthonormal) basis by forming the overlap
145: matrix from the states defined in (\ref{2quons}) and diagonalize it. That
146: procedure automatically lead us to the three symmetric and one antisymmetric
147: states above. Then we can diagonalize any observable within the symmetric an
148: the antisymmetric subspaces separately. Of course, this important property
149: can be readily generalized to any number of quons. For three quons for
150: example, besides the well known symmetric and antisymmetric states, there
151: are more exotic mixed symmetric states. To shorten the corresponding
152: expressions we adopt the convention $a_i^{\dagger }a_j^{\dagger
153: }a_k^{\dagger }|0>\equiv |ijk>$ . Then we have for the (normalized) basis
154: states in that case:
155:
156: \begin{equation}
157: |S>=\frac 1{\sqrt{1+2q^2+2q+q^3}}\frac 1{\sqrt{6}%
158: }[|ijk>+|jik>+|ikj>+|jki>+|kij>+|kji>] \label{S}
159: \end{equation}
160: \begin{equation}
161: |A>=\frac 1{\sqrt{1+2q^2-2q-q^3}}\frac 1{\sqrt{6}%
162: }[|ijk>-|jik>-|ikj>+|jki>+|kij>-|kji>] \label{A}
163: \end{equation}
164: \begin{equation}
165: |MS1>=\frac 1{\sqrt{1-q^2+q-q^3}}\frac 1{\sqrt{12}%
166: }[|ijk>-|jik>+2|ikj>+|jki>-2|kij>-|kji>] \label{MS1}
167: \end{equation}
168:
169: \begin{equation}
170: |MS2>=\frac 1{\sqrt{1-q^2+q-q^3}}\frac 12[-|ijk>-|jik>+|jki>+|kji>]
171: \label{MS2}
172: \end{equation}
173:
174: \begin{equation}
175: |MS3>=\frac 1{\sqrt{1-q^2-q+q^3}}\frac 12[|ijk>-|jik>-|jki>+|kji>]
176: \label{MS3}
177: \end{equation}
178: \begin{equation}
179: {}|MS4>=\frac 1{\sqrt{1-q^2-q+q^3}}\frac 1{\sqrt{12}%
180: }[|ijk>+|jik>-2|ikj>+|jki>-2|kij>+|kji>] \label{MS4}
181: \end{equation}
182: \bigskip\
183:
184: where $i,j,k=1,2,3$ . Also, the cases $i=j$, $i=k$, $j=k$ and $i=j=k$ are
185: automatically included in expressions (\ref{S}) to (\ref{MS4}) , unless to a
186: normalization factor which is q-independent. Evidently, the above basis
187: states can be built from the well known procedure based in the Young tableux
188: method \cite{Greiner},or, as said before, through the diagonalization of the
189: overlap matrix obtained from all possible order permutations from the state $%
190: a_i^{\dagger }$ $a_j^{\dagger }$ $a_k^{\dagger }|0>$ . In fact, the
191: q-polynomials which appear in the square roots in equations (\ref{S}) to (%
192: \ref{MS4}), correspond to the eigenvalues of the overlap matrix and measure
193: the degree of violation of statistics in the system. If we then choose q
194: sufficiently close to $1(-1)$, we may restrict ourselves to the symmetric
195: (antisymmetric) subspace, once the observables of the theory do not mix
196: subspaces corresponding to different symmetries. At this respect it would be
197: interesting to generalize our above expressions for the symmetric part of
198: the quonic space. This has a two-fold motivation: first of all many
199: applications of the deformed algebras (for which only symmetric states are
200: considered) to many-body problems are restricted to small deformations of
201: the usual Lie algebra, i.e., q close to 1\cite{BonnaRev} . We would like to
202: compare some of those results with the equivalent solutions using the quon
203: algebra. Secondly, the value of q very close ( but not equal) to 1 has the
204: quite appealing idea to try to take in to account possible violations of
205: bosonic statistics for systems in which the degrees of freedom are related
206: to particles with a integer spin value but that are in fact composed by
207: ''fundamental '' fermions.
208:
209: It is then possible to show (see Appendix) that the most general symmetric
210: state for a system of N quons can be written as:
211: \begin{equation}
212: |n_in_jn_k...;S>=\sqrt{\frac{n_i!n_j!n_k!...}{N![N]!}}{\widehat{S}}%
213: _N(a_i^{\dagger })^{n_i}(a_j^{\dagger })^{n_j}(a_k^{\dagger })^{n_k}...|0>
214: \label{NSym}
215: \end{equation}
216: where ${\widehat{S}}_N$ is a operator that generates all possible
217: combinations that are symmetric under the permutation of any of the creation
218: operators( as defined in the Appendix), $n_i+n_j+n_k+...=N$ and \cite{Kibler}%
219: :
220: \begin{equation}
221: \lbrack N]=\frac{1-q^N}{1-q}, \label{caixote}
222: \end{equation}
223: with $[N]!=[N][N-1]....[2][1]$ and [0]!=1. Another important result that we
224: are going to use next and which is also obtained in the Appendix, is the
225: following:
226:
227: \begin{equation}
228: a_i|n_in_jn_k...;S>=\sqrt{\frac{[N]}N}\sqrt{n_i}|n_i-1,n_jn_k......;S>
229: \label{aNSym}
230: \end{equation}
231:
232: This last expression allows us to calculate matrix elements between
233: symmetric quonic states with any number of quons. In the following section
234: we discuss two simple examples and compare them to the deformed algebra
235: results.
236:
237: \bigskip\
238:
239: \section{Applications and Results}
240:
241: \bigskip\
242:
243: In order to discuss some examples, it is important to recall that, according
244: to reference\cite{PhysLetA}, given a set of operators $a_m,a_m^{\dagger }$
245: for which $m=-j,...,+j$ , and such that they obey the quon commutation
246: relations, it can be proved that each $a_m^{\dagger }$ behave as a su(2)
247: irreducible tensor. In other words they obey the expected commutation
248: relations with angular momentum operators, built from the corresponding
249: number operators. However, those number operators present a complicated
250: structure, and are written as an infinite series of the quonic creation and
251: annihilation operators\cite{GreenPRD}. Also, it is not difficult to obtain a
252: su(2) scalar from the quons. For example, we may define a quonic three
253: dimensional harmonic oscillator by the Hamiltonian:
254:
255: \begin{equation}
256: H_{osc}^q=\frac{\hbar \omega }2\{(a_{+}^{\dagger }{}a_{+}+a_{-}^{\dagger
257: }{}a_{-}+a_0^{\dagger }{}a_0)(1+q)+3\} \label{qosc}
258: \end{equation}
259:
260: where
261: \begin{equation}
262: a_{+}=\frac 1{\sqrt{2}}(a_1+ia_2)~,~a_{-}=\frac 1{\sqrt{2}
263: }(a_1-ia_2)~,~a_0=a_3~.
264: \end{equation}
265: \noindent First of all we note that, if $a_1,a_2$ and $a_3$ obey quon
266: commutation relations, so does the set $a_{+},a_{-}$ and $a_0$ . Secondly,
267: the factor $(1+q)$ comes from the fact that the Hamiltonian must be
268: symmetrized and finally we can easily recover the usual harmonic oscillator
269: Hamilton operator by simply choosing $q=1$, from the above expression. In
270: order to get the corresponding spectrum we should now diagonalize (\ref{qosc}%
271: ) inside each subspace formed by the states of a given symmetry of the whole
272: permutation symmetry group and for a given number of quanta $%
273: N=n_{+}+n_{-}+n_0.$ Alternatively we could proceed with the diagonalization
274: from the basis formed by all order permutations obtained from the state $%
275: (a_{+}^{\dagger })^{n_{+}}$ $(a_{-}^{\dagger })^{n_{-}}$ $(a_0^{\dagger
276: })^{n_0}|0>$. This last procedure amounts however to a diagonalization in a
277: non-orthonormal basis. On the other hand, the prior diagonalization of the
278: overlap matrix, as done in the previous section for the $N=3$ case,
279: corresponds to a partial diagonalization of the Hamiltonian, which then
280: becomes block diagonal, each block corresponding to a given permutation
281: symmetry.
282:
283: As we said before, it is our intention here to make some comparisons to the
284: deformed algebra results. Then we content ourselves with only the symmetric
285: part of the solution, which is justifiable once we keep q close enough to 1.
286: In that case, using equation (\ref{aNSym}) and its Hermitian conjugate, we
287: readily find for the eigenvalues of our quonic harmonic oscillator, the
288: remarkable simple result:
289:
290: \begin{equation}
291: E_{osc}^q=\frac{\hbar \omega }2\{[N](1+q)+3\}
292: \end{equation}
293:
294: Our quonic harmonic oscillator give us then a spectrum which is not equally
295: spaced but still depends on just one quantum number, the total number $N$.
296: This is not the case in quantum algebras (see equation (28.38) in reference
297: \cite{BonnaRev}) , for which the spectrum depends on $N$ and on a additional
298: quantum number $l$, related to the su$_q$(2) angular momentum. To get an
299: idea of the effect of the deformation in the spectrum, we present in figure
300: I a comparison to the regular oscillator for some selected values of the
301: parameter q.
302:
303: To better spot the angular momentum structure within the quon algebra we
304: consider now the quantum rotor. A natural choice in this case is to follow a
305: Schwinger type of prescription\cite{Sakurai} for the definition of the
306: angular momentum components:
307:
308: \begin{equation}
309: L_{+}=a_{+}^{\dagger }a_{-},~L_{-}=a_{-}^{\dagger
310: }a_{+},~L_0=\frac 12\{a_{+}^{\dagger }a_{+}-a_{-}^{\dagger
311: }a_{-}\} \label{AngDef}
312: \end{equation}
313:
314: We again restrict our results to the symmetric quon subspace. Using once
315: more equation (\ref{aNSym}), we get the results:
316:
317: \begin{equation}
318: <n_{+}^{\prime },n_{-}^{\prime };S|[L_{+},L_{-}]|n_{+},n_{-};S>=\frac{[N]}%
319: N(n_{+}-n_{-})\delta _{n_{+}^{\prime }n_{+}}\delta _{n_{-}^{\prime }n_{-}}
320: \end{equation}
321:
322: and,
323:
324: \begin{equation}
325: <n_{+}^{\prime },n_{-}^{\prime };S|2L_0|n_{+},n_{-};S>=\frac{[N]}%
326: N(n_{+}-n_{-})\delta _{n_{+}^{\prime }n_{+}}\delta _{n_{-}^{\prime }n_{-}}
327: \end{equation}
328:
329: where now $N=n_{+}+n_{-}$. Using the same type of calculation we may also
330: prove that:
331:
332: \begin{equation}
333: <n_{+}^{\prime },n_{-}^{\prime };S|[L_0,L_{+}]|n_{+},n_{-};S>=<n_{+}^{\prime
334: },n_{-}^{\prime };S|L_{+}|n_{+},n_{-};S>
335: \end{equation}
336:
337: \begin{equation}
338: <n_{+}^{\prime },n_{-}^{\prime
339: };S|[L_0,L_{-}]|n_{+},n_{-};S>=-<n_{+}^{\prime },n_{-}^{\prime
340: };S|L_{-}|n_{+},n_{-};S>
341: \end{equation}
342:
343: \noindent The above results show that the operators defined in (\ref{AngDef}%
344: ), behave as genuine angular momentum components within the symmetric
345: subspace. All we need now is to obtain the matrix element of the operator $%
346: L^2$, which gives:
347:
348: \begin{equation}
349: <n_{+}^{\prime },n_{-}^{\prime };S|L^2|n_{+},n_{-};S>=\frac{[N]}2~(\frac{[N]}%
350: 2+1)~\delta _{n_{+}^{\prime }n_{+}}\delta _{n_{-}^{\prime }n_{-}}
351: \end{equation}
352:
353: \noindent Assuming the correspondence $n_{+}=l+m$ and $n_{-}=l-m$ we finally
354: obtain for our q-rotor spectrum:
355:
356: \begin{equation}
357: E_l^q=A\frac{[2l]}2\left\{ \frac{[2l]}2+1\right\} \label{qrotor}
358: \end{equation}
359:
360: with $A$ being the inertia constant. Again, we see that, although we get the
361: right limit for q=1, the spectrum given by equation (\ref{qrotor}) is
362: different from the one obtained through quantum algebra techniques (see
363: equation (19.3) in \cite{BonnaRev}). At this respect, one interesting result
364: that emerges from the deformed algebra rotor, is its ability to describe
365: stretching effects as experimentally observed in the spectra of heavy nuclei
366: and molecules, with the introduction of a single parameter\cite{BonaRotor}.
367: In our case, we could test the applicability of our previous results doing
368: the same sort of analysis, using expression (\ref{qrotor}). In figure II we
369: show the experimental spectrum of the fundamental rotational band in the $%
370: ^{240}$Pu nucleus, chosen here as typical sample, together with the one
371: obtained from our quonic rotor. We choose a q-value that minimizes the
372: differences between the theoretical spectrum and the experimental one,
373: within the interval allowed by the quon algebra. It is interesting to
374: observe that a q-value slightly smaller than $1$ is enough to produce the
375: desired modifications in order to take in to account the stretching effects
376: just mentioned, as we can observe particularly for higher values of the
377: angular momentum, compared to the rigid rotor result. Also shown in the same
378: figure is the best fit spectrum obtained using the quantum algebra approach
379: from reference\cite{BonaRotor}.
380:
381: \bigskip\
382:
383: \section{Conclusions}
384:
385: \bigskip\
386:
387: In this letter we have discussed and presented a method to build
388: in a systematic way, a basis of states that represent a system of
389: identical quons. The novel feature of that type of states lie in
390: the fact that, once quons obey commutation relations which
391: interpolate between bosons and fermions, all kind of permutation
392: symmetries can be accommodated in a many-quon state. The
393: probability that each type of symmetry occur is then controlled
394: by a single parameter q. Once any observable should be symmetric
395: by any particle exchange, a classification of the states by their
396: permutational symmetry amounts to a partial diagonalization of the
397: corresponding operator in the whole quonic space. In order to
398: make some contact with the deformed algebra results, we have kept
399: our attention to the totally symmetric subspace, for which we
400: could find a closed general expression for a state with any
401: number of quons, as well as for the action of an operator on it.
402: Two simple examples were then considered here within that point
403: of view: a quonic version of the three-dimensional harmonic
404: oscillator and a rotor model based on the quon algebra. As a by
405: product we have found out the interesting result that the angular
406: momentum operator written in terms of quons and within the
407: symmetric subspace behaves as usual su(2) angular momentum
408: operators, having the same functional form as in the case of
409: regular bosons. Also, a comparison of both examples with the
410: results previously obtained within quantum algebras, show a very
411: distinct energy spectrum distribution in the case of the harmonic
412: oscillator where the same degree of degeneracy observed in the
413: usual bosonic oscillator is recovered, contrary to what happens
414: when we use the quantum su$_q$(2) algebra. As for the rigid
415: rotor, though we have found an energy spectrum very close in both
416: cases, the angular momentum operator properties are quite
417: different from the su$_q$(2) properties.
418:
419: Quantum or q-deformed algebras are by now considered a very powerful tool to
420: deal with physical systems for which usual algebras can not take in to
421: account some of their properties. The quon algebra, which is considered in
422: the literature as a particular case\cite{Kibler}of those algebras, present
423: subtle differences that may reflect some important and even fundamental
424: differences in what concerns the interpretation of the final results. As
425: mentioned in the Introduction, an interesting point that deserves some
426: investigation in the near future is the analysis of bosonic systems that are
427: in fact composed by fermions, once some observed deviations from a true
428: boson behavior could in principle be realized by the quon algebra in a
429: natural way.
430:
431: \vskip 0.35in
432: \begin{center}
433: \textbf{Acknowledgments}
434: \end{center}
435: This work was partially supported by CNPq - Brazil. \vspace{0.5cm}
436: %------------------------------------------------------------------
437: \begin{center}
438: \textbf{Appendix}
439: \end{center}
440:
441: In this appendix we prove two important results, eqs.(\ref{NSym},\ref{aNSym}%
442: ), given in the main text. We begin with the most general symmetric (not
443: normalized) N quons state in an arbitrary basis:
444: \begin{equation}
445: \widehat{S}_{\mathrm{N}}(a_i^{\dagger })^{n_i}(a_j^{\dagger
446: })^{n_j}(a_k^{\dagger })^{n_k}...|0>~\equiv ~{\frac 1{{n_i!n_j!n_k!...}}}%
447: \sum_{P_{\mathrm{N}}}a_{{\alpha }_1}^{\dagger }a_{{\alpha }_2}^{\dagger
448: }...a_{{\alpha }_{n_i}}^{\dagger }a_{{\alpha }_{n_i+1}}^{\dagger }...a_{{%
449: \alpha }_{n_i+n_j+1}}^{\dagger }...a_{{\alpha }_N}^{\dagger }|0>~~,
450: \label{symm}
451: \end{equation}
452: where N=$n_i+n_j+n_k+...$ and the summation runs over all the N!
453: permutations, $P_{\mathrm{N}}$, in the indices $\alpha _1,\alpha
454: _2,...,\alpha _N$. We order these indices such that $\alpha
455: _1,\alpha _2,...,\alpha _{n_i}$ corresponds to the i-state,
456: $\alpha _{n_i+1},\alpha _{n_i+2},...,\alpha _{n_i+n_j}$ to the
457: j-state and so on. The factorials
458: under the denominator accounts for repeated terms in the summation and eq.(%
459: \ref{symm}) is the definition of the $\widehat{S}_{\mathrm{N}}$ operator. We
460: now prove by induction the following result:
461: \begin{equation}
462: a_i~\widehat{S}_{\mathrm{N}}(a_i^{\dagger })^{n_i}(a_j^{\dagger
463: })^{n_j}(a_k^{\dagger })^{n_k}...|0>~=~[N]~\widehat{S}_{\mathrm{{N}-1}%
464: }(a_i^{\dagger })^{n_i-1}(a_j^{\dagger })^{n_j}(a_k^{\dagger })^{n_k}...|0>~,
465: \label{hypo}
466: \end{equation}
467: where [N] is given in eq.(\ref{caixote}). It is easy to show that the
468: relation above is valid for N=1($\widehat{S}_{\mathrm{0}}=I$, the identity
469: operator). We assume that it is also true for a N-1 quons state, i. e.,
470: \begin{equation}
471: a_i~\widehat{S}_{\mathrm{{N}-1}}(a_i^{\dagger })^{n_i^{\prime
472: }}(a_j^{\dagger })^{n_j^{\prime }}(a_k^{\dagger })^{n_k^{\prime
473: }}...|0>~=~[N-1]~\widehat{S}_{\mathrm{{N}-2}}(a_i^{\dagger })^{n_i^{\prime
474: }-1}(a_j^{\dagger })^{n_j^{\prime }}(a_k^{\dagger })^{n_k^{\prime }}...|0>~,
475: \label{induc}
476: \end{equation}
477: where N-1=$n_i^{\prime }+n_j^{\prime }+n_k^{\prime }+...$ . One can shows
478: the following property of the symmetrization operator:
479: \[
480: \widehat{S}_{\mathrm{N}}(a_i^{\dagger })^{n_i}(a_j^{\dagger
481: })^{n_j}(a_k^{\dagger })^{n_k}...|0>~=~a_i^{\dagger }\widehat{S}_{\mathrm{{N}%
482: -1}}(a_i^{\dagger })^{n_i-1}(a_j^{\dagger })^{n_j}(a_k^{\dagger
483: })^{n_k}...|0>
484: \]
485: \begin{equation}
486: ~+~a_j^{\dagger }\widehat{S}_{\mathrm{{N}-1}}(a_i^{\dagger
487: })^{n_i}(a_j^{\dagger })^{n_j-1}(a_k^{\dagger })^{n_k}...|0>~+a_k^{\dagger }%
488: \widehat{S}_{\mathrm{{N}-1}}(a_i^{\dagger })^{n_i}(a_j^{\dagger
489: })^{n_j}(a_k^{\dagger })^{n_k-1}...|0>~+~...~~. \label{SNpro}
490: \end{equation}
491: This property follows from the definition of the symmetrization operator, eq.(%
492: \ref{symm}), and its rearrangement as given below:
493: \[
494: \widehat{S}_{\mathrm{N}}(a_i^{\dagger })^{n_i}(a_j^{\dagger
495: })^{n_j}(a_k^{\dagger })^{n_k}...|0>~=~{\frac 1{{n_i!n_j!n_k!...}}}
496: \]
497: \[
498: \cdot \big( a_{{\alpha }_1}^{\dagger
499: }\sum_{P_{\mathrm{{N}-1}}}{\hat{a}}_{{\alpha }_1}^{\dagger
500: }a_{{\alpha }_2}^{\dagger }...a_{{\alpha }_N}^{\dagger
501: }|0>~+~a_{{\alpha }_2}^{\dagger }\sum_{P_{\mathrm{{N}-1}}}a_{{\alpha }%
502: _1}^{\dagger }{\hat{a}}_{{\alpha }_2}^{\dagger }a_{{\alpha }_3}^{\dagger
503: }...a_{{\alpha }_N}^{\dagger }|0>~
504: \]
505: \begin{equation}
506: ~+~...~+~a_{{\alpha }_{n_i}}^{\dagger }\sum_{P_{\mathrm{{N}-1}}}a_{{\alpha }%
507: _1}^{\dagger }...{\hat{a}}_{{\alpha }_{n_i}}^{\dagger }a_{{\alpha }%
508: _{n_i+1}}^{\dagger }...a_{{\alpha }_N}^{\dagger }|0>~+~...~+~a_{{\alpha }%
509: _N}^{\dagger }\sum_{P_{\mathrm{{N}-1}}}a_{{\alpha }_1}^{\dagger }...a_{{%
510: \alpha }_{N-1}}^{\dagger }{\hat{a}}_{{\alpha }_N}^{\dagger }|0>~\big)~~,
511: \end{equation}
512: where the hat symbol on the creation operator means that it is
513: omitted in that position. The first $n_i$ terms above are equal,
514: since $\alpha _1,\alpha _2,...,\alpha _{n_i}$ are associated to
515: the i-state, the same argument may be used for the next $n_j$
516: terms and so on. So the property given in eq.(\ref{SNpro}) is
517: proved.
518:
519: From eq.(\ref{SNpro}) and the q-mutation relation, we obtain for
520: the action of the annihilation operator, $a_i$, in the N quons
521: symmetric state the result as follows:
522: \[
523: a_i\widehat{S}_{\mathrm{N}} (a_i^{\dagger })^{n_i}(a_j^{\dagger
524: })^{n_j}(a_k^{\dagger })^{n_k}...|0>~=~\widehat{S}_{\mathrm{{N}-1}}
525: (a_i^{\dagger })^{n_i-1}(a_j^{\dagger })^{n_j}(a_k^{\dagger })^{n_k}...|0>
526: \]
527: \[
528: ~+~q\big(~ a_i^{\dagger} a_i~\widehat{S}_{\mathrm{{N}-1}} (a_i^{\dagger
529: })^{n_i-1}(a_j^{\dagger })^{n_j}(a_k^{\dagger })^{n_k}...|0>~+~
530: a_j^{\dagger} a_i~\widehat{S}_{\mathrm{{N}-1}} (a_i^{\dagger
531: })^{n_i}(a_j^{\dagger })^{n_j-1}(a_k^{\dagger })^{n_k}...|0>
532: \]
533: \begin{equation} \label{aisn}
534: ~+~ a_k^{\dagger} a_i~\widehat{S}_{\mathrm{{N}-1}} (a_i^{\dagger
535: })^{n_i}(a_j^{\dagger })^{n_j}(a_k^{\dagger })^{n_k-1}...|0> \big) ~
536: \end{equation}
537: We now use eq.(\ref{induc}) to rewrite the term between parenthesis and get
538: \[
539: a_i \widehat{S}_{\mathrm{N}} (a_i^{\dagger })^{n_i}(a_j^{\dagger
540: })^{n_j}(a_k^{\dagger })^{n_k}...|0>~=~\widehat{S}_{\mathrm{{N}-1}}
541: (a_i^{\dagger })^{n_i-1}(a_j^{\dagger })^{n_j}(a_k^{\dagger })^{n_k}...|0>
542: \]
543: \[
544: ~+~q[\mathrm{{N}-1]\big(~ a_i^{\dagger} \widehat{S}_{{N}-2} (a_i^{\dagger
545: })^{n_i-2}(a_j^{\dagger })^{n_j}(a_k^{\dagger })^{n_k}...|0>~+~
546: a_j^{\dagger} \widehat{S}_{{N}-2} (a_i^{\dagger })^{n_i-1}(a_j^{\dagger
547: })^{n_j-1}(a_k^{\dagger })^{n_k}...|0> }
548: \]
549: \begin{equation} \label{aif}
550: ~+~ a_k^{\dagger} \widehat{S}_{\mathrm{{N}-2}} (a_i^{\dagger
551: })^{n_i-1}(a_j^{\dagger })^{n_j}(a_k^{\dagger })^{n_k-1}...|0> \big)~
552: \end{equation}
553: Finally using eq.(\ref{SNpro}) in the term between parenthesis and
554: the q-number property, [N]=1+q[N-1], we may rearrange the above
555: expression as:
556: \begin{equation} \label{eqla}
557: a_i~\widehat{S}_{\mathrm{N}} (a_i^{\dagger })^{n_i}(a_j^{\dagger
558: })^{n_j}(a_k^{\dagger })^{n_k}...|0>~=~[N]~\widehat{S}_{\mathrm{{N}-1}}
559: (a_i^{\dagger })^{n_i-1}(a_j^{\dagger })^{n_j}(a_k^{\dagger
560: })^{n_k}...|0>~.~~
561: \end{equation}
562: and the proof by induction is finished.
563:
564: Now we obtain the normalized symmetric state. Let us write:
565: \begin{equation}
566: |n_i,n_j,n_k,...;S>~\equiv ~A_{n_in_jn_k...}\widehat{S}_{\mathrm{N}%
567: }(a_i^{\dagger })^{n_i}(a_j^{\dagger })^{n_j}(a_k^{\dagger })^{n_k}...|0>~.
568: \end{equation}
569: So the normalization constant,$A_{n_in_jn_k...}$ is determined by:
570: \[
571: 1=<n_i,n_j,n_k,...;S|n_i,n_j,n_k,...;S>=A_{n_in_jn_k...}^2~{\frac{N!}{{%
572: n_i!n_j!n_k!}}}
573: \]
574: \[
575: \cdot ~<0|...(a_k)^{n_k}(a_j)^{n_j}(a_i)^{n_i}\widehat{S}_{\mathrm{N}%
576: }(a_i^{\dagger })^{n_i}(a_j^{\dagger })^{n_j}(a_k^{\dagger
577: })^{n_k}...|0>~=~A_{n_in_jn_k...}^2~{\frac{{N![N]!}}{{n_i!n_j!n_k!}}}~~,
578: \]
579: where in order to get the result above, eq.(\ref{eqla}), was iterated. So
580: the normalization factor in eq.(\ref{NSym}) is obtained. From eq.(\ref{eqla}%
581: ) and the normalization factor obtained above, eq.(\ref{aNSym}) follows
582: trivially.
583:
584:
585: \bigskip\
586:
587: \begin{thebibliography}{99}
588: \bibitem{GreenPRD} O.W. Greenberg, Phys.Rev. \textbf{D43} (1991) 4111.
589:
590: \bibitem{Green6/7/2000} O.W. Greenberg, hep-th/0007054 6Jul 2000.
591:
592: \bibitem{Chaichian} M.Chaichian, R.Gonzalez Felipe, C. Montonen, J. Phys
593: \textbf{A26} (1993) 4017.
594:
595: \bibitem{PhysLetA} S. S. Avancini, F. F. de Souza Cruz, J. R. Marinelli and
596: D. P. Menezes, Phys. Letters \textbf{A267} (2000) 109.
597:
598: \bibitem{Greenphysica} O.W. Greenberg, Physica \textbf{A180} (1992) 419.
599:
600: \bibitem{qbos1} S. S. Avancini, F. F. de Souza Cruz, J. R. Marinelli, D. P.
601: Menezes and M. M. Watanabe de Moraes, J. Phys. \textbf{G25} (1999) 525.
602:
603: \bibitem{qbos2} S. S. Avancini, J. R. Marinelli and D. P. Menezes, J. Phys.
604: \textbf{G25} (1999) 1829.
605:
606: \bibitem{Greiner} M. Hamermesh, Group Theory and its Application to
607: Physical Problems, Dover Pub.(1989), and W. Greiner, B.
608: M\"{u}ller, Quantum Mechanics Symmetries, Second Edition,
609: Springer(1994).
610:
611: \bibitem{BonnaRev} D. Bonatsos and C. Daskaloyannis, Prog. in Part. and
612: Nucl. Phys. \textbf{43} (1999) 537.
613:
614: \bibitem{Kibler} M.R. Kibler, Introduction to Quantum Algebras, Second
615: International School on Theoretical Physics, Poland (1992), World Scientific.
616:
617: \bibitem{Sakurai} J. J. Sakurai, Modern Quantum Mechanics, Addison-Wesley,
618: Reading MA (1985).
619:
620: \bibitem{BonaRotor} P. P. Raychev, R. P. Roussev and Y. F. Smirnov, J.
621: Phys. \textbf{G16} (1990) L137-L141.
622: \end{thebibliography}
623:
624: \newpage
625:
626: FIGURE\ CAPTIONS
627:
628: \bigskip\
629:
630: Figure I. Harmonic oscillator spectrum (\textbf{A}) compared to the quonic
631: harmonic oscillator spectra obtained for two different values of the
632: deformation parameter : $q=0.99$ (\textbf{B}) and $q=0.98$ (\textbf{C}).
633:
634: \bigskip\
635:
636: Figure II. The experimental (\textbf{A})\ spectrum for the
637: $^{244}$Pu
638: fundamental rotational band compared to the quonic rotor result (\textbf{B}%
639: ), deformed algebra result from reference \cite{BonaRotor}
640: (\textbf{C}) and the usual quantum rigid rotor result
641: (\textbf{D}). Spectrum \textbf{B} was obtained using $q=0.99478$.
642:
643: \end{document}
644: