1: %\documentstyle[12pt,graphicx,aps]{revtex}
2: %\documentstyle[12pt]{article}
3: \documentstyle[graphicx,aps,twocolumn]{revtex}
4: \sloppy
5: \parindent0em
6: \newcommand{\ga}{\gamma}
7: \newcommand{\th}{\vartheta}
8: \newcommand{\al}{\alpha}
9: \newcommand{\De}{\Delta}
10: \newcommand{\la}{\lambda}
11: \newcommand{\de}{\delta}
12: \newcommand{\f}{\varphi}
13: \newcommand{\eps}{\varepsilon}
14: \newcommand{\om}{\omega}
15: \newcommand{\k}{\kappa}
16: \newcommand{\bea}{\begin{eqnarray}}
17: \newcommand{\beq}{\begin{equation}}
18: \newcommand{\eea}{\end{eqnarray}}
19: \newcommand{\eeq}{\end{equation}}
20: \begin{document}
21: \title{Dynamical moment of inertia and quadrupole vibrations
22: in rotating nuclei}
23: \author{\sc R.G. Nazmitdinov$^{1,2}$
24: D. Almehed $^{3}$, F.~D\"onau $^4$}
25: \address{\sl
26: $^1$ Max-Planck-Institut f\"ur Physik Komplexer Systeme, D-01187 Dresden,
27: Germany\\
28: $^2$ Bogoliubov Laboratory of Theoretical Physics
29: Joint Institute for Nuclear Research, 141980 Dubna, Russia\\
30: $^3$ Department of Physics, UMIST, P.O. Box 88,
31: Manchester M60 1QD, United Kingdom\\
32: $^4$Institut f\"ur Kern- und Hadronenphysik,
33: FZ Rossendorf, 01314 Dresden, Germany}
34: \maketitle
35:
36: \vspace{0.3cm}
37:
38: \begin{abstract}
39: {\sf The contribution of quantum shape fluctuations
40: to inertial properties of rotating nuclei has been analysed
41: within the self-consistent one-dimensional cranking oscillator
42: model. It is shown that
43: in even-even nuclei the dynamical moment of inertia calculated in
44: the mean field approximation is equivalent to the Thouless-Valatin moment
45: of inertia calculated in the random phase approximation
46: {\it if and only if} the self-consistent conditions for the mean field
47: are fulfilled.
48: }
49: \end{abstract}
50:
51: \vspace{0.2in}
52:
53: PACS numbers: 21.60.Ev, 21.60.Jz, 21.10.Re
54:
55: \vspace{0.2in}
56:
57: A description of rotational states
58: is one of the oldest, yet not fully solved, problem in
59: nuclear structure physics.
60: While various microscopic models based on the
61: cranking approach \cite{RS,Fr01} describe reasonably well
62: the kinematical moment of inertia
63: ${\cal J}^{(1)} = -(dE/d\Omega)/\Omega $
64: for a finite angular frequency $\Omega$,
65: there is still a systematic deviation of the dynamical moment of inertia
66: ${\cal J}^{(2)}=-d^2E/d\Omega^2$ (E is the energy in the rotating frame)
67: from the experimental data at high spins \cite{Af}.
68: Since the moments of inertia are the benchmarks
69: for microscopic models of collective motion in nuclei,
70: the understanding of the source of the discrepancy becomes
71: a challenge for a many body theory of finite Fermi systems.
72:
73: The pairing correlations introduced in nuclear
74: physics by Bogoliubov \cite{Bog} and Bohr, Mottelson and
75: Pines \cite{BMP} improved the description
76: of the inertial nuclear properties, especially
77: in the low spin region \cite{Bel}.
78: It was conjectured that at high spins
79: the Coriolis and centrifugal forces break
80: Coopers pairs from the pairing condensate and cause
81: the transition from the superfluid
82: to the normal (unpaired) fluid \cite{MV}.
83: As a consequence, the rigid body (kinematical)
84: moment of inertia should be reached
85: at the normal phase. However, even for superdeformed nuclei where
86: the pairing condensate is expected to be strongly quenched,
87: the moment of inertia is usually lower than the rigid body
88: value.
89:
90: In fact, the conjecture was based on
91: the analogy between moment of inertia of a rotating nucleus
92: and magnetic susceptibility of a macroscopic superconductor
93: under a magnetic field. Therefore, the conjecture may lose its validity
94: for finite Fermi systems like nuclei, while remaining
95: correct for an infinite number of particles.
96: For example, the deviation from the rigid body value
97: could be partially explained due to shell effects \cite{Fr}.
98: In present paper we demonstrate that correlations caused
99: by shape oscillations of a system are another important
100: ingredient which is missing in all state of the art
101: calculations of the moments of inertia for rotating nuclei.
102: We focus our analysis upon the dynamical moment of inertia, since
103: the ${\cal J}^{(2)}$
104: contains more information about different properties of the system
105: due to the obvious relation
106: ${\cal J}^{(2)}={\cal J}^{(1)}+
107: \Omega d{\cal J}^{(1)}/d\Omega$.
108:
109: The first attempt to take into account the contribution of the quadrupole
110: oscillations to the correlation energy and to the moment of inertia
111: within the cranking+random phase approximation (RPA) approach \cite{Eg}
112: suffered from an inconsistency between
113: the mean field and the residual interaction. In addition,
114: the calculations were done in a restricted configuration
115: space (only 3 shells have been considered).
116: For a time the solution of the problem was postponed,
117: since there was no practical
118: recipe to calculate the contribution of the correlation energy,
119: even in a restricted configuration space.
120: Using the integral representation method developed recently
121: in \cite{dan}, a full energy $E_{RPA}$ can be calculated in the
122: RPA order \cite{RS,BR} with a high accuracy and with
123: minimal numerical efforts. Consequently, this energy can be used
124: to calculate the dynamical moments of inertia ${\cal J}^{(2)}$.
125: On the other hand, in literature it is stated
126: that the dynamical moment of inertia should be equivalent to
127: the Thouless-Valatin moment of inertia \cite{Th}.
128: Since the realistic application of the Thouless-Valatin theory requires the
129: self-consistent solution of the mean field and the RPA equations,
130: until now this point is not clarified.
131:
132: To understand the role of shape oscillations upon the value of the moment of
133: inertia and to calculate the Thouless-Valatin moment of inertia in a simple
134: but still realistic model, we use a self-consistent cranked
135: triaxial harmonic oscillator model as a rotating mean field.
136: The self-consistent residual interaction constructed
137: according to the recipe \cite{Ab} is added to describe the
138: collective excitations in a rotating system.
139: Since all shells are mixed, we go beyond the
140: approximation used in \cite{Eg}
141: (for a cranking harmonic oscillator see also \cite{Nil}).
142: Notice that the contribution of
143: the pairing vibrations to the correlation energy aside
144: of the one from the shape vibrations is also important
145: (see \cite{A0,Ber} and references there). However, there are some open
146: problems with the choice of the self-consistent pairing interaction.
147: Therefore, the combined effect
148: of the both types of vibrations is beyond the scope of
149: the present investigation
150: and we leave this problem for the future.
151:
152:
153: The mean field part of the many-body Hamiltonian (Routhian)
154: in the rotating frame is given by
155: \beq
156: \label{mf}
157: H=\sum_{i=1}^N(h_0-\Omega l_x)_i
158: =H_0-\Omega L_x
159: \eeq
160: where the single-particle triaxial harmonic oscillator Hamiltonian $h_0$ is
161: aligned along its principal axes and reads
162: \beq
163: h_0={1\over 2m}{\vec p}^2+{m\over 2}
164: (\omega _x^2 x^2+\omega _y^2 y^2+\omega _z^2 z^2).
165: \eeq
166: The eigenmodes and the total energy of
167: the mean field Hamiltonian Eq.(\ref{mf})
168: are well known \cite{Val,RBK,Zel},
169: \beq
170: \label{fr}
171: \omega_{\pm}^2=\frac{1}{2}\left(\omega_y^2+\omega_z^2+2\Omega^2\pm
172: [(\omega_y^2-\omega_z^2)^2
173: +8\Omega^2(\omega_y^2+\omega_z^2)]^{1/2}\right)
174: \eeq
175: \beq
176: \label{en}
177: E_{MF}=\hbar\left(\omega_x{\sum}_x+\omega_+{\sum}_+
178: + \omega_-{\sum}_-\right).
179: \eeq
180: Here,
181: $\sum_k=\langle \sum_j^N(n_k+1/2)_j \rangle$ and $n_k=a_k^+a_k$ ($k=x,+,-$)
182: where $a_k^+$, $a_k$ are the oscillator quanta operators.
183: The lowest levels are filled from the bottom,
184: which give the ground state energy in the rotating frame.
185: The Pauli principle is taken into account such that
186: two particles occupy one level.
187: The minimisation of the total energy Eq.(\ref{en})
188: with respect to all three frequencies,
189: subject to the volume conservation condition
190: $\omega_x\omega_y\omega_z=\omega_0^3$, yields the
191: self-consistent condition \cite{TA,HN}
192: for a finite rotational frequency
193: \beq
194: \omega_x^2\langle x^2 \rangle =\omega_y^2 \langle y^2 \rangle =\omega_z^2
195: \langle z^2 \rangle .
196: \label{cond}
197: \eeq
198: It should be pointed out that the condition Eq.(\ref{cond}) provides
199: generally the absolute minima in comparison with the local minima obtained
200: from the condition of the {\it isotropic velocity distribution} \cite{RBK,Zel}
201: \beq
202: \label{con1}
203: {\sum}_x\omega_x={\sum}_+\omega_+={\sum}_-\omega_-
204: \eeq
205: at large rotational frequency.
206:
207: To analyse the contribution of the quadrupole
208: shape oscillations we add to the mean
209: field Hamiltonian Eq.(\ref{mf}) the self-consistent
210: interaction resulting from small angular rotations
211: around the $x-$, $y-$, $z-$ axes and small variations
212: of two the intrinsic shape parameters $\varepsilon$ and
213: $\gamma$ \cite{Ab}.
214: Consequently, the total Hamiltonian can be expressed as
215: \beq
216: H_{{\rm RPA}}=H_0-\Omega L_x
217: - {\kappa \over 2}\sum _{\mu =-2}^2 Q_{\mu }^{\dagger} Q_{\mu }
218: ={\tilde H} - \Omega L_x.
219: \label{hrpa}
220: \eeq
221: The effective interaction restores the rotational
222: invariance of the Hamiltonian $H_0$ such that now
223: $[{\tilde H},L_i]=0 \quad (i=x,y,z)$ in the RPA order.
224: The self-consistency condition
225: Eq.(\ref{cond}) fixes the quadrupole strength
226: $\kappa=\frac{4\pi}{5}\frac{m\omega_0^2}{\langle r^2 \rangle}$.
227: Here a mean value
228: $\langle r^2 \rangle = \langle \bar x^2+\bar y^2+\bar z^2 \rangle$
229: and quadrupole operators $Q_{\mu}={\bar{r^2Y_{2\mu}}}$ are expressed
230: in terms of the double-stretched coordinates
231: $\bar q_i=\frac{\omega_i}{\omega_0} q_i$, $(q_i=x,y,z)$.
232: We remind that the self-consistent residual interaction does
233: not affect the equilibrium deformation obtained from the
234: minimisation procedure.
235:
236: Using the transformation from $p_i,q_i$ variables to the
237: quanta $a_k^+,a_k$ \cite{Zel} ,
238: all matrix elements are calculated analytically.
239: We solve the RPA equation of motion for the generalised coordinates
240: ${\cal X}_{\lambda}$ and momenta ${\cal P}_{\lambda}$
241: \bea
242: [H_{{\rm RPA}},{\cal X}_{\lambda}]&=&
243: -i\omega_{\lambda}{\cal P}_{\lambda},\quad
244: [H_{{\rm RPA}},{\cal P}_{\lambda}]=
245: i\omega_{\lambda}{\cal X}_{\lambda}, \\ \nonumber
246: [{\cal X}_{\lambda},{\cal P}_{\lambda}]&=&
247: i\delta_{\lambda, \lambda^\prime }.
248: \eea
249: where $\omega_{\lambda}$ are the RPA eigen-frequencies in
250: the rotating frame and the associated phonon operators are
251: $O_{\lambda}=({\cal X}_{\lambda}-i{\cal P}_{\lambda})/\sqrt{2}$.
252: Here
253: ${\cal X}_{\lambda}=\sum_s X_s^{\lambda} {\hat{f}}_s$,
254: ${\cal P}_{\lambda}=i\sum_s P_s^{\lambda}{\hat{g}}_s$ are
255: bilinear combinations of
256: the quanta $a_k^+,a_k$ such that
257: $\langle [{\hat{f}}_s, {\hat{g}}_{s^{\prime}}]\rangle=V_s\delta_{s,s^{\prime}}$ where
258: quantities $V_s$ are proportional to different combinations of $\sum_i$
259: $(i=x$,$+$,$-)$. Further, $\langle ... \rangle$ means the averaging over mean field states.
260: Since the mean field violates the rotational invariance, among the
261: RPA eigen-frequencies there exist two spurious solutions.
262: One solution with zero frequency is associated with the rotation
263: around the $x$-axes, since $[H,L_x]=0$. The other "spurious" solution
264: at $\omega\equiv \Omega$
265: corresponds to a collective rotation, since
266: $[H,L_{\pm}]=[H,L_y{\pm}iL_z]=\mp\Omega L_{\pm}$ \cite{m77}.
267: The Hamiltonian Eq.(\ref{hrpa}) possesses the signature symmetry, i.e.
268: $[R_x,H_{\rm RPA}]=0$ ($R_x=e^{-i\pi {\hat{L}}_x}$), such that it
269: decomposes into positive and negative signature terms
270: \beq
271: H_{\rm RPA}=H(+)+H(-)
272: \eeq
273: which can be separately diagonalized \cite{m77,JM,KN}. The negative signature
274: Hamiltonian contains the rotational mode and
275: the vibrational mode describing
276: the wobbling motion \cite{JM,m79}.
277: We focus on the positive signature Hamiltonian.
278: It contains the zero-frequency mode
279: defined by
280: \beq
281: \label{an}
282: [H(+),\phi_x]=\frac{-iL_x}{{\cal J}_{TV}}, \quad
283: [\phi_x, L_x]=i
284: \eeq
285: and allows one to determine the
286: Thouless-Valatin moment of inertia ${\cal J}_{TV}$ \cite{MW}.
287: Here, the angular momentum operator $L_x=\sum_s l_s^x{\hat{f}}_s$
288: and the canonically conjugated angle $\phi_x=i\sum_s\phi_s^x{\hat{g}}_s$
289: are expressed via ${\hat{f}}_s$ and ${\hat{g}}_s$ which obey the condition
290: $R_x{\hat{d}}_sR_x^{-1}=\hat{d}_s$ $(\hat{d}_s={\hat{f}}_s$ or ${\hat{g}}_s)$.
291: Solving Eqs.(\ref{an}) for the Hamiltonian $H(+)$,
292: \beq
293: H(+)=\sum_{k=x,+,-}\hbar\omega_k(a_k^{\dagger}a_k+1/2)-
294: \frac{\kappa}2(Q_0^2+Q_1^{(+)2}+Q_2^{(+)2})
295: \eeq
296: where
297: \bea
298: &&Q_0 = \sqrt{\frac{5}{16\pi}}(2{\bar z}^2-{\bar x}^2-
299: {\bar y}^2) = \sqrt{\frac{5}{16\pi}}
300: \sum_s q_s^0{\hat{f}}_s\\
301: &&Q_1^{(+)} = \sqrt{\frac{15}{4\pi}}{\bar y}{\bar z} = \sqrt{\frac{15}{4\pi}}
302: i\sum_s q_s^1{\hat{g}}_s\\
303: &&Q_2^{(+)} = \sqrt{\frac{15}{16\pi}}({\bar x}^2-{\bar y}^2) =
304: \sqrt{\frac{15}{16\pi }}\sum_s q_s^2{\hat{f}}_s
305: \eea
306: \vskip 1cm
307: we obtain the expression for the Thouless-Valatin moment of inertia
308: \beq
309: \label{tv}
310: {\cal J}_{TV}={\cal J_{I}}+\frac{[2S_{x0}S_{x2}S_{02}-S_{x0}^2(S_{22}-\frac{1}{\kappa_2})
311: -S_{x2}^2(S_{00}-\frac{1}{\kappa_0})]}{[(S_{00}-\frac{1}{\kappa_0})
312: (S_{22}-\frac{1}{\kappa_2})-S_{02}^2]}
313: \eeq
314: Here, the term ${\cal J_{I}}$ corresponds to the Inglis moment of
315: inertia
316: \beq
317: \label{ing}
318: {\cal J_{I}}=\sum_s \frac{(l_s^x)^2V_s}{E_s}.
319: \eeq
320: The second term in Eq.(\ref{tv}) is a
321: contribution of the quadrupole residual interaction in the cranking
322: model. In the cranking harmonic oscillator it consists of the terms which
323: have the following structure
324: \beq
325: S_{xm}=\sum_s \frac{l_s^xq_s^mV_s}{E_s},\quad S_{nm}=\sum_s
326: \frac{q_s^nq_s^mV_s}{E_s}, \quad n,m=0,2
327: \eeq
328: where $E_s$ are the energies of particle-hole excitations:
329: $E_1=2\hbar \omega_+$, $E_2=2\hbar \omega_-$, $E_3=2\hbar \omega_x$,
330: $E_4=\hbar\omega_+ + \hbar \omega_-$
331: and $E_5=\hbar \omega_+ - \hbar \omega_-$.
332: We also introduced the following notations:
333: $\kappa_0=\frac{5}{16\pi}\kappa$ and $\kappa_2=\frac{15}{16\pi}\kappa$.
334:
335: The above results are the starting point for our numerical analysis.
336: To take into account shell effects we consider two systems with number
337: of particles $A=20, 64 \quad (N=Z)$.
338: For $\hbar \Omega=0$ MeV the global minimum occurs for a prolate
339: shape and for a near oblate triaxial shape for $A=20$ and $64$,
340: respectively \cite{HN}. If we trace the configurations
341: which characterise the ground states,
342: with increasing rotational frequency both systems
343: become oblate. At this point the moment of inertia vanishes, since there is
344: no a kinetic energy associated with such a rotation.
345:
346: In order to compare various moments of inertia, i.e.
347: the Thouless-Valatin, Eq.(\ref{tv}),
348: the Inglis, Eq.(\ref{ing}), and
349: ${\cal J}^{(2)}_{MF}=-d^2E_{MF}/d\Omega^2$ with
350: ${\cal J}^{(2)}_{RPA}=-d^2E_{RPA}/d\Omega^2$ ,
351: we calculate the RPA correlation energy
352: $E^{RPA}_{\it corr}=\frac{1}{2}(\sum_{\lambda}\omega_{\lambda}-\sum_s E_s)$
353: which includes the positive and negative signature contributions.
354: In Figs.\ref{fig2},\ref{fig3}
355: the results of calculations for different moments of inertia are presented
356: for $A=20$ and $64$, respectively. For our knowledge this is
357: a first numerical demonstration of
358: the equivalence between the dynamical moment of inertia
359: ${\cal J}_{MF}^{(2)}$
360: calculated in the mean field
361: approximation and the Thouless-Valatin moment of inertia ${\cal J}_{TV}$
362: calculated in the RPA. For the both systems the Inglis moment of inertia
363: ${\cal J}_{I}$ is smaller than the ${\cal J}_{TV}$ and
364: ${\cal J}_{MF}^{(2)}$
365: and has a different rotational dependence.
366: While the Inglis moment of inertia
367: characterises the collective properties of non-interacting fermions,
368: the dynamical moment of inertia reflects the changes in the rotating
369: self-consistent mean field due to an inter-nucleon interaction.
370: As it was pointed out in \cite{m90}, the volume conservation
371: condition, used as a constraint in the mean field calculations,
372: can be interpreted as a Hartree approximation applied to an
373: interaction that involves the sum of one-, two- etc forces.
374:
375: \begin{figure}
376: \noindent
377: \centering
378: \includegraphics[width=7.0cm,angle=270]{HOvol10.eps}\\
379: \caption{Moments of inertia for $N=Z=10$ system as a function
380: of the rotational frequency $\omega \equiv \Omega$.
381: The definitions of different moments of inertia are given in the text.
382: }
383: \label{fig2}
384: \end{figure}
385:
386: \begin{figure}
387: \noindent
388: \centering
389: \includegraphics[width=7.0cm,angle=270]{HOvol32.eps}\\
390: \caption{As in Fig.\ref{fig2} for $N=Z=32$
391: system
392: }
393: \label{fig3}
394: \end{figure}
395:
396: The sharp drop of all moments of inertia in Fig.2
397: is caused by the onset of the oblate shape where the collective rotation does
398: not exist. For $A=64$ the onset of the oblate deformation occurs at
399: a smaller rotational frequency in contrast to the one for the system $A=20$.
400:
401: The dynamical moment of inertia ${\cal J}_{RPA}^{(2)}$ is larger
402: than the Thouless-Valatin moment of inertia.
403: However, from our calculations it follows that the contribution of the
404: RPA ground state correlations decreases with an increase of the
405: number of particles.
406: The difference between the ${\cal J}_{RPA}^{(2)}$ and the
407: ${\cal J}_{TV}$ is
408: due to the following reason.
409: The Inglis moment of inertia
410: is smaller than the Thouless-Valatin (or ${\cal J}_{MF}^{(2)}$)
411: value, since the ${\cal J}_{TV}$ contains the effect of the
412: residual particle-hole interaction.
413: On the other hand, the Thouless-Valatin moment of inertia manifests
414: the rotational dependence of the residual interaction.
415: Thus, we may speculate that inclusion of the phonon
416: interaction could help to reproduce
417: the behaviour of the ${\cal J}_{RPA}^{(2)}$ which characterises
418: the rotational dependence of the phonon-phonon interaction.
419:
420:
421: In summary, using the self-consistent cranking harmonic oscillator model, we
422: have numerically proved the equivalence of the dynamical moment of inertia
423: calculated in the mean field approximation to the Thouless-Valatin moment of
424: inertia calculated in the RPA. Our result is a consequence of the
425: self-consistent condition Eq.(\ref{cond}) which minimises
426: the expectation value of
427: the mean field Hamiltonian, Eq.(\ref{mf}).
428: This condition is equivalent to the stability condition
429: of collective modes in the RPA \cite{T21}, i.e.
430: $\omega_{\lambda}$ to be real and
431: non-negative, and has been used to calculate
432: different moments of inertia.
433: The rotational dependence of the both
434: dynamical moments of inertia,
435: ${\cal J}_{MF}^{(2)}$ and ${\cal J}_{RPA}^{(2)}$, is similar, however,
436: the ${\cal J}_{RPA}^{(2)}$ is larger than the
437: ${\cal J}_{MF}^{(2)}$ due to
438: the contribution of the ground state correlations.
439: This difference between the moments of inertia
440: is less important for heavy systems.
441:
442:
443:
444:
445: \vspace{.5cm}
446:
447: {\it Acknowledgement:}
448: This work was supported in part by the Heisenberg-Landau
449: program of the JINR.
450:
451:
452: \begin{thebibliography}{99}
453: \bibitem{RS}
454: P.~Ring and P.~Schuck,
455: {\it The Nuclear Many-Body Problem}
456: (Springer-Verlag, Heidelberg, 1980)
457: \bibitem{Fr01}
458: S.~Frauendorf, Rev.Mod.Phys. {\bf 73}, 463 (2001)
459: \bibitem{Af}
460: A.V.~Afanasjev, J.~K\"onig, P.~Ring, L.M.~Robledo and J.L.~Egido,
461: Phys.Rev.C {\bf 62}, 054306 (2000)
462:
463: \bibitem{Bog}
464: N.N.~Bogoliubov,
465: Doklady Akademii Nauk SSSR {\bf 119}, 52 (1958)
466: \bibitem{BMP}
467: A.~Bohr, B.R.~Mottelson and D.~Pines,
468: Phys.Rev. {\bf 110}, 936 (1958)
469: \bibitem{Bel}
470: S.T.~Belyaev,
471: Mat.Fys.Medd.K. Dan. Vidensk. Selsk. {\bf 31} (11),
472: 1 (1959)
473: \bibitem{MV}
474: B.R.~Mottelson and J.G.~Valatin,
475: Phys.Rev.Lett. {\bf 5}, 511 (1960)
476: \bibitem{Fr}
477: S.~Frauendorf, K.~Neergard, J.A.~Sheikh and\\
478: P.M.~Walker,
479: Phys.Rev.C {\bf 61}, 064324 (2000)
480: \bibitem{Eg}
481: J.L.~Egido, H.J.~Mang and P.~Ring,
482: Nucl.Phys.A {\bf 341}, 229 (1980)
483:
484: \bibitem{dan}
485: F.~D\"onau, D.~Almehed and R.G.~Nazmitdinov,
486: Phys.Rev.Lett. {\bf 83}, 280 (1999)
487:
488: \bibitem{BR} J.-P.~Blaizot and G.~Ripka,
489: {\it Quantum Theory of Finite Systems}
490: (MIT, Cambridge, MA, 1986)
491: \bibitem{Th}
492: D.J.~Thouless and J.G.~Valatin,
493: Nucl.Phys. {\bf 31}, 211 (1962)
494: \bibitem{Ab}
495: S.~\AA berg, Phys.Lett.B {\bf 157}, 9 (1985);
496: H.~Sakamoto and T.~Kishimoto,
497: Nucl.Phys.A {\bf 501}, 205 (1989)
498:
499: \bibitem{Nil}
500: S.G.~Nilsson and I.~Ragnarsson,
501: {\it Shapes and Shells in Nuclear Structure}
502: (Cambridge University Press, Cambridge, England 1995)
503: \bibitem{A0}
504: D.~Almehed, F.~D\"onau, S.~Frauendorf and R.G.~Nazmitdinov,
505: Phys.Scripta {\bf T88}, 62 (2000);
506: D.~Almehed, S.~Frauendorf and F.~D\"onau,
507: Phys.Rev.C {\bf 63}, 044311 (2001)
508:
509: \bibitem{Ber} J.~Dukelsky and P.~Schuck,
510: Phys.Lett.B {\bf 464}, 164 (1999);
511: K.~Hagino and G.F.~Bertsch,
512: Nucl.Phys.A {\bf 679}, 163 (2000)
513:
514: \bibitem{Val}
515: J.G.~Valatin, Proc.Roy.Soc.(London) {\bf 238}, 132 (1956)
516: \bibitem{RBK}
517: G.~Ripka, J.P.~Blaizot, and N.~Kassis,
518: in {\it International Extended Seminar, Trieste, 1973}
519: (IAEA, Vienna, 1975) Vol.1, p.445
520: \bibitem{Zel}
521: V.G.~Zelevinsky, Sov.J.Nucl.Phys. {\bf 22}, 565 (1975)
522:
523: \bibitem{TA}
524: T.~Troudet and R.~Arvieu,
525: Ann.Phys. (NY) {\bf 1}, 134 (1981);
526: E.R.~Marshalek and R.G.~Nazmitdinov,
527: Phys.Lett.B {\bf 300}, 199 (1993)
528:
529: \bibitem{HN}
530: W.D.~Heiss and R.G.~Nazmitdinov,
531: Phys.Lett.B {\bf 397}, 1 (1997)
532:
533: \bibitem{m77}
534: E.R.~Marshalek, Nucl.Phys.A {\bf 275}, 416 (1977)
535:
536: \bibitem{JM}
537: D.~Janssen and I.N.~Mikhailov,
538: Nucl.Phys.A {\bf 318}, 390 (1979)
539: \bibitem{KN}
540: J.~Kvasil and R.G.~Nazmitdinov,
541: Sov.J.Part.Nucl. {\bf 17}, 265 (1986)
542: \bibitem{m79}
543: E.R.~Marshalek, Nucl.Phys.A {\bf 331}, 429 (1979)
544: \bibitem{MW}
545: E.R.~Marshalek and J.~Weneser, Ann.Phys.(NY) {\bf 53}, 569 (1969)
546:
547: \bibitem{m90}
548: E.R.~Marshalek, Phys.Lett. B {\bf 244}, 1 (1990)
549: \bibitem{T21}
550: D.J.~Thouless, Nucl.Phys. {\bf 21}, 225 (1960)
551: \end{thebibliography}
552: \end{document}
553:
554: