1: \documentclass[aps,twocolumn,prl]{revtex4}
2: \usepackage{graphicx}
3:
4:
5: \begin{document}
6: \title{A description of the yrast states in $^{24}$Mg
7: by the self-consistent 3D-cranking model}
8: \author{Makito Oi}
9: \affiliation{Department of Physics, University of Surrey, Guildford, GU2 7X,
10: Surrey, United Kingdom}
11:
12: \begin{abstract}
13: With the self-consistent 3D-Cranking model,
14: the ground-state rotational band in $^{24}$Mg is analysed.
15: A role of triaxial deformation is discussed, in particular, in
16: a description of the observed two $I^{\pi}=8^+$ states.
17: \end{abstract}
18:
19: \maketitle
20:
21: \section{Introduction}
22: $^{24}_{12}$Mg has been studied very well
23: as a typical case of well-deformed light-mass nuclear systems.
24: After self-consistent microscopic calculations
25: of the non-relativistic \cite{BFH87}
26: and relativistic methods \cite{KR88} in the late 1980s,
27: the nucleus is believed to have an axially-symmetric
28: and prolate shape in the ground state.
29: Assuming the core of $^{16}$O,
30: the ground state configuration is supposed to have
31: eight valence particles occupying the d$_{5/2}$ orbits
32: (four neutrons and four protons).
33: %
34: Since the Fermi level of this nucleus is in the beginning of the
35: sd-shell (corresponding to the $N=2$ harmonic oscillator shell),
36: there are many open valence orbitals above the Fermi level, which
37: may induce deformation, and consequently a collective rotation.
38: %
39: Such a rotational band has been already identified in experiments
40: (for instance, see Ref. \cite{RB68,WWJ01}).
41: A particular interest is an existence of the
42: two $I^{\pi}=8^+$ states observed in experiment.
43: These two states are energetically close to each other
44: (the difference is about 2 MeV).
45: After the study of Sheline et al. \cite{SRA88},
46: it is believed that the second $8^+$ belongs to the ground-state
47: rotational band (g-band).
48: %
49: Valor, et al. analysed the g-band with the cranked Skyrme HF + BCS approach
50: as well as the configuration mixing approach based on the generator coordinate
51: method (GCM) \cite{VHB00}.
52: However, their calculations assume axial symmetry for descriptions of
53: intrinsic states.
54: Up to $I^{\pi}=4^{+}$, they were
55: able to reproduce the experimental data very well.
56: The cranked mean-field calculation gave a fairly good agreement for
57: $I^{\pi}=6^+$, while the configuration mixing approaches returned
58: larger energies for the state. This is simply because
59: these states at high spin
60: were projected out from the non-cranked mean-field state.
61: Interestingly, the cranked mean-field calculation for $I^{\pi}=8^+$
62: matches the observed energy of the first $8^+$ state, but the authors
63: dismissed the agreement on the basis of the Sheline's analysis \cite{SRA88},
64: and they speculated that the disagreement might mainly come from
65: the triaxial effects at high spin induced by the disappearance
66: of the pairing correlation in their calculation.
67:
68: \section{3D-cranked HFB method}
69: Inspired by the study by Valor et al. \cite{VHB00},
70: we performed the self-consistent cranking calculation
71: allowing triaxial deformation in a self-consistent manner.
72: As a new aspect in our study, not only 1D-cranking but also
73: 3D-cranking calculations were carried out.
74: %
75: An advantage of the 3D-cranking model
76: is that low- and high-K intrinsic structures can be systematically
77: studied \cite{OWA02}.
78:
79: The Hamiltonian used in our study reads
80: \begin{equation}
81: \hat{H}=\hat{H}_0 + \hat{V},
82: \end{equation}
83: where the first term describes the one-body part, which is
84: the spherical Nilsson Hamiltonian in this study,
85: while the second part is the two-body interactions, which
86: is the pairing-plus-quadrupole force (so-called P$+$QQ force).
87: The model space (valence space) to diagonalise the two-body part
88: contains two major shells ($N=2,3$) in the spherical Nilsson model,
89: in accordance with the Kummar-Baranger criteria for the P$+$QQ force
90: \cite{KB65}.
91: %
92: The variational state is the Hartree-Fock-Bogoliubov ansatz,
93: which is a generalised product state. With quasi-particle
94: annihilation operators $\beta_q$, the ansatz is expressed as
95: \begin{equation}
96: |\text{HFB}\rangle=\prod_q\beta_q|0\rangle,
97: \end{equation}
98: where $|0\rangle$
99: is the vacuum for the canonical basis $a_m$ and $a^{\dag}_m$.
100: (In our case, the canonical basis correspond to the spherical Nilsson
101: basis.) The canonical basis and the quasi-particle basis are connected by
102: a unitary transformation called the general Bogoliubov transformation
103: \cite{On86}.
104: %
105: The variational equation is derived for
106: \begin{equation}
107: \delta\langle\text{HFB}|\hat{H}-\sum_{i=1}^{3}\left(\omega_i\hat{J}_i
108: +\mu_i\hat{B}_i\right)-\sum_{\tau=\text{p,n}}\lambda_{\tau}\hat{N}_{\tau}|
109: \text{HFB}\rangle = 0.
110: \end{equation}
111: In the above equation,
112: $\hat{J}_i$ is the $i$-th component of the angular momentum operator
113: (the index $i$ takes $i=1,2,3$),
114: and $\hat{N}_{\tau}$ describes the number operator for proton ($\tau=\text{p}$)
115: and neutron ($\tau=\text{n}$).
116: $\hat{B}_i$ is an off-diagonal component of the quadrupole operator,
117: defined as
118: \begin{equation}
119: \hat{B}_i = \frac{1}{2}\left(\hat{Q}_{jk}+\hat{Q}_{kj}\right),
120: \end{equation}
121: where the indices $(i,j,k)$ should be placed in a cyclic manner.
122: %
123: Each term with the Lagrange multipliers ($\omega_i,\mu_i$ and $\lambda_{\tau}$)
124: is necessary to put constraints in intrinsic states:
125: $(\langle\hat{J}_1\rangle,\langle\hat{J}_2\rangle,\langle\hat{J}_3\rangle)
126: = (J\cos\theta, J\sin\theta\sin\phi,J\sin\theta\cos\phi);
127: \langle\hat{N}_{\tau}\rangle = N_{\tau};
128: \langle\hat{B}_{i}\rangle = 0.$
129: The last constraint is necessary so as to keep the orientation of
130: the angular momentum vector against the intrinsic coordinate axes \cite{On86}.
131: %
132: The variational equation is solved by means of the method of steepest descent.
133: Details of the method are presented in Ref.\cite{On86}.
134: %
135: A deformed Nilsson + BCS state is used for an initial trial state at $J=0$.
136: Deformation parameters and gap energies for the trial state are
137: determined by referring to the calculations of the liquid drop model
138: by the M\"oller and Nix \cite{MN95}.
139: In the present study, the deformation parameters for the ground state
140: are chosen to be $\left(\beta,\gamma)=(0.347,0.0^{\circ}\right)$ and
141: the pairing gap energies are $\left(\Delta_{\text{p}},\Delta_{\text{n}}\right)
142: =(1.840 \text{ MeV},1.962 \text{ MeV})$.
143: %
144: All the physical quantities, such as energy, quadrupole moments (deformation),
145: single-particle spin components, and gap energies,
146: are self-consistently calculated
147: under the above constraints in this framework.
148:
149: \section{Results and Discussions}
150:
151: First of all, let us report the results from the 1D-cranking calculation.
152: Despite a use of a simple separable interaction,
153: the ground-state rotational spectra is reproduced reasonably well
154: (Fig.\ref{fig-ene}).
155: %
156: As Valor, et al. commented in Ref.\cite{VHB00},
157: the gap energies disappear both for protons and neutrons
158: before $J=6\hbar$ (Fig.\ref{fig-ene}).
159: %
160: Triaxial deformation gradually decreases from $\gamma=0^{\circ}$.
161: However, in $J\le 4\hbar$, triaxial deformation can be still regarded
162: negligible. In other words, axial symmetry is kept fairly well.
163: On the other hand, at high spin ($J=8, 10\hbar$),
164: substantial triaxial deformation is formed ($\gamma \agt -10^{\circ}$),
165: and axial symmetry is clearly broken.
166: %
167: It should be noted here that
168: the convention for the quadrupole deformation parameters ($\beta,\gamma$)
169: in this study follows the Hill-Wheeler coordinates, which gives
170: the opposite sign in $\gamma$ to the so-called Lund convention.
171: A fact that the gamma deformation becomes negatively larger implies
172: that the nucleus is reaching the non-collective rotational state
173: ($\gamma=-60^{\circ})$, where the rotational axis corresponds
174: to the shortest principal axis of the deformation.
175:
176: %
177: \begin{figure}
178: \includegraphics[width=0.45\textwidth]{fig1a}
179: \includegraphics[width=0.45\textwidth]{fig1b}
180: \includegraphics[width=0.45\textwidth]{fig1c}
181: \caption{(top) Rotational energy obtained in the 1D cranking calculation
182: (solid line), and the experimental data (cross).
183: (middle) Gap energies for protons and neutrons, obtained in the 1D cranking
184: calculation. (bottom) Gamma deformation obtained in the 1D cranking
185: calculation.}
186: \label{fig-ene}
187: \end{figure}
188: %
189: \begin{table}[tb]
190: \begin{tabular}{ccccccc}
191: \hline
192: $J$ \quad\quad & $0$& $2$& $4$& $6$& $8$& $9$\\
193: \hline
194: $\beta$ \quad\quad & 0.37 & 0.38 & 0.37 & 0.35 & 0.31 & 0.29\\
195: \hline
196: \end{tabular}
197: \caption{Evolution of the deformation in $\beta$ as a function of the total spin $J$, which are obtained from the 1D-cranking calculation.}
198: \end{table}
199:
200: From this result,
201: it is understandable why the cranking calculation by Valor, et al. \cite{VHB00}
202: was successful up to $J=6\hbar$ and why the deviations from the
203: experimental values become larger at higher spins.
204: As mentioned earlier,
205: they solved the HFB equation within the axial symmetry constraint.
206:
207: \begin{table}[htb]
208: \begin{tabular}{cccccc}
209: \hline
210: & $J=2$& $J=4$& $J=6$& $J=8$& $J=9$\\
211: \hline
212: $\pi$d$_{5/2}$ &
213: 1.0 ($50\%$) & 2.0 ($50\%$) &
214: 2.9 ($48\%$) & 3.6 ($45\%$) &
215: 3.9 ($43\%$)\\
216: $\pi$d$_{3/2}$ &
217: 0.0 ($0\%$) & 0.0 ($0\%$) &
218: 0.1 ($2\%$) & 0.4 ($5\%$) &
219: 0.6 ($7\%$)\\
220: \hline
221: $\nu$d$_{5/2}$ &
222: 1.0 ($50\%$) & 2.0 ($50\%$) &
223: 2.9 ($48\%$) & 3.6 ($45\%$) &
224: 3.9 ($43\%$)\\
225: $\nu$d$_{3/2}$ &
226: 0.0 ($0\%$) & 0.0 ($0\%$) &
227: 0.1 ($2\%$) & 0.4 ($5\%$) &
228: 0.6 ($7\%$)\\
229:
230: \hline
231: \end{tabular}
232: \caption{Single-particle components of the total spin in the 1D cranking
233: calculation.The first two rows correspond to proton orbitals, while
234: the last two to neutron orbitals. The numbers in brackets are ratios of
235: single-particle spins against the total spin. The unit of spin is
236: $\hbar$.}
237: \label{tb2}
238: \end{table}
239:
240: In the HFB theory, the total spin is expressed as the sum
241: of single-particle spins, that is,
242: \begin{equation}
243: \langle\hat{J}_i\rangle
244: = \sum_{m}\langle j_i^{(m)}\rangle = \sum_{mn}(j_i)_{mn}\rho_{nm},
245: \end{equation}
246: where $\rho$ is the density matrix and $j_i$ is the single-particle
247: angular momentum operator. The indices $m$ and $n$
248: are for the canonical basis, while the index $i$ for the coordinate axes,
249: that is, $i=1,2,3$. Using this information of angular momentum,
250: we can discuss nuclear structure with single-particle spins.
251: %
252: In the Table \ref{tb2} are the calculated main components of single-particle spins
253: for the different total spin $J$. The result reflects a fact that
254: $^{24}$Mg is a $N=Z$ nucleus, that is, the way of single-particle excitations
255: is the same both for protons and neutrons.
256: %
257: For low-spin members in the rotational band, the total spin
258: consists mainly of the d$_{5/2}$. The higher the total spin,
259: the more the d$_{3/2}$ orbit is occupied.
260: Therefore, the ``collectivity'' in this nucleus is attributed
261: to gradual excitations into the d$_{3/2}$ orbit.
262:
263: Next, let us present the results of the 3D-cranking calculations.
264: In this paper, we focus on the analysis of the $J^{\pi}=8^{+}$ state,
265: where its triaxial deformation becomes substantial ($\gamma\simeq -10^{\circ}$)
266: in the 1D-cranking calculation.
267: %
268: \begin{figure}[t]
269: % \includegraphics[width=0.5\textwidth]{e3dj08}
270: \includegraphics[width=0.5\textwidth,angle=0]{fig2}
271: \caption{Energy surface of the $J=8\hbar$ state, calculated by means
272: of the 3D-cranked HFB method. There are two minima as intrinsic states,
273: which are seen at $(\theta,\phi)=(0^{\circ},0^{\circ})$ and
274: $=(90^{\circ},0^{\circ})$.}
275: \label{3d-ene}
276: \end{figure}
277: %
278: Obviously from the Fig.\ref{3d-ene},
279: two configurations compete with each other.
280: This competition can be considered as a kind of ``level crossing''
281: between two different states (or configurations), but
282: in the 3D-cranking calculation each ``level'' is represented by a curved
283: surface.
284: There are mainly two minima in the energy surface for the $J=8\hbar$ state
285: (Fig.\ref{3d-ene}): $(\theta,\phi)=(0^{\circ},0^{\circ})$ and
286: $=(90^{\circ},0^{\circ})$, and they characterize the two configurations.
287: %
288: The former minimum corresponds to the 1D-cranking
289: solution where triaxiality is calculated to be $\gamma\simeq -10^{\circ}$.
290: %
291: The rotation axis is found to be along the shortest axis,
292: and the corresponding state is expected be of low-$K$ character.
293: %
294: The latter minimum is the energetically lowest state (yrast state)
295: at this spin, which is about 2.5 MeV lower than the 1D-cranking solution.
296: This yrast state at $J^{\pi}=8^+$ is found to be axially symmetric because
297: the triaxiality is calculated to be $\gamma\simeq 0^{\circ}$
298: ($\beta\simeq 0.24$).
299: %
300: In this case, the rotational axis
301: points along the longest axis of the axially symmetric shape,
302: so that the rotation is of single-particle character.
303: As a result, the major components of the state should be of high-$K$
304: characters.
305: %
306: Studying in detail the microscopic structure,
307: the total spin is found to be constructed almost purely by
308: the d$_{5/2}$ orbits (in both protons and neutrons):
309: $3.97\hbar$ each by the d$_{5/2}$ orbits of protons and neutrons.
310: In addition to the difference in the deformation,
311: a lack of the d$_{3/2}$ component in the first $8^+$ state
312: implies the yrast state is surely different from rotational members
313: of the g-band, from a microscopic point of view.
314: %
315: The shell model calculation by Wiedenh ver, et al. \cite{WWJ01}
316: says that such a configuration, that is, (d$_{5/2}$)$^8$,
317: corresponds to the first $8^+$ state (which does not belong to
318: the g-band) observed in experiment. Therefore, our result is
319: consistent with the shell model calculation, too.
320:
321: From this result, we can conclude that the yrast state found in our
322: calculation at $J=8\hbar$ is a high-$K$ state with $K^{\pi}=8^{+}$.
323: It was experimentally observed to be
324: energetically lower than the second $8^+$ state by about 2 MeV.
325:
326: \begin{figure}[th]
327: \includegraphics[width=0.45\textwidth]{fig3a}
328: \includegraphics[width=0.45\textwidth]{fig3b}
329: \caption{(Top panel) A cross section of the energy surface at $\phi=0^{\circ}$.
330: (Bottom panel) Triaxial deformation at $\phi=0^{\circ}$.}
331: \label{fig-ph0}
332: \end{figure}
333:
334: \begin{table}[th]
335: \begin{tabular}{ccccccccccc}
336: \hline
337: $\theta$ \quad \quad &0 & 10 & 20 & 30 & 40 & 50 & 60 & 70 & 80 & 90 \\
338: \hline
339: $\beta$ \quad \quad &0.31& 0.25 & 0.21 & 0.18& 0.14 & 0.11& 0.17& 0.21& 0.23& 0.24\\
340: \hline
341: \end{tabular}
342: \caption{Change of $\beta$ as a function of the tilt angle $\theta$, obtained
343: in the 3D-cranking calculation for $J=8\hbar$ amd $\phi=0^{\circ}$.}
344: \end{table}
345:
346: In the lower panel of Fig.\ref{fig-ph0}, the triaxial deformation
347: is seen to have $\gamma\simeq 120^{\circ}$ at $20^{\circ}\alt \theta \alt
348: 45^{\circ}$. Because the ``level crossing'' happens at $\theta\simeq 20^{\circ}$,
349: we cannot exactly see how the graph continues toward $\theta\rightarrow 0$.
350: However, from the trend of the graph,
351: it is possible to guess that the graph forms a symmetric
352: shape with respect to $\theta=45^{\circ}$ in the upper panel of Fig.\ref{fig-ph0}
353: , and that $\gamma\rightarrow 120^{\circ}$ for $\theta\rightarrow 0^{\circ}$
354: in the lower panel of Fig.\ref{fig-ph0}. This reflection symmetry around
355: $\theta=45^{\circ}$ indicates that the
356: state in $0^{\circ}\le \theta \le 45^{\circ}$ and the state
357: in $45^{\circ}\le \theta \le 90^{\circ}$ have the same intrinsic structure.
358:
359: \section{Conclusion}
360: The ground-state rotational band was studied with the self-consistent
361: 1D-cranking calculation. It was confirmed in this study that
362: an effect of triaxiality in the nature of the rotational band
363: is important at high spin, as previously anticipated by Valor, et al.
364: In addition, two high-spin states at $J=8^+$ observed in experiment
365: were analysed by means of the self-consistent 3D-cranking calculation.
366: It concludes (in a qualitative manner)
367: that the yrast $8^+$ state is an axially-symmetric
368: high-$K$ state created by the deformation-aligned protons and neutrons
369: in the d$_{5/2}$ orbitals while the second $8^+$ is a rotational member
370: of the ground-state rotational band with substantial triaxial deformation.
371:
372: For the first time, in the framework of the self-consistent and
373: microscopic method, the two $I^{\pi}=8^+$ states in $^{24}$Mg,
374: which correspond to low- and high-$K$ states respectively,
375: are explained on the same footing, that is,
376: through the self-consistent 3D-cranking model.
377:
378: \section{acknowledgments}
379: The author thanks Professors P. Ring and P.-H. Heenen for their suggestions
380: to calculate this $^{24}$Mg nucleus using the self-consistent 3D-cranking
381: method. Useful discussions with Dr W. Catford,
382: Professor P. W. Walker and Dr P. Regan
383: are also acknowledged. This work is supported by EPSRC
384: with an advanced research fellowship GR/R75557/01
385: as well as a first grant EP/C520521/1.
386:
387:
388: \begin{thebibliography}{99}
389: \bibitem{BFH87} P. Bonche, H. Flocard, P.-H. Heenen,
390: Nucl. Phys. A 467 (1987) 115.
391: \bibitem{KR88} W. Koepf, P. Ring,
392: Phys. Lett. B 212 (1988) 397.
393: \bibitem{RB68} S. W. Robinson, R.D. Bent,
394: Phys. Rev. 168 (1968) 1266.
395: \bibitem{WWJ01} I. Wiedenhover, et al.,
396: Phys. Rev. Lett. 87 (2001) 142502.
397: \bibitem{SRA88} R. K. Sheline, I. Ragnarsson, S. Aberg, and A. Watt,
398: J. Phys. G 14 (1988) 1201.
399: \bibitem{VHB00} A. Valor, P.-H. Heenen, and P. Bonche,
400: Nucl. Phys. A 671 (2000) 145.
401: \bibitem{OWA02} M. Oi, P. M. Walker, and A. Ansari,
402: Phys. Lett. B 525 (2002) 255.
403: \bibitem{KB65} K. Kumar and M. Baranger,
404: Nucl. Phys. A 110 (1968) 529; M. Baranger, and K. Kumar,
405: Nucl. Phys. 62 (1965) 113.
406: \bibitem{On86} N. Onishi,
407: Nucl. Phys. A 456 (1986) 279; T. Horibata, and N. Onishi,
408: Nucl. Phys. A 596 (1996) 251.
409: \bibitem{MN95} P. M\"oller, J. R. Nix, W. D. Meyer, and W. J. Swiatecki,
410: At. Data Nucl. Data Tables 59 (1995) 185.
411: \end{thebibliography}
412:
413: \end{document}