Sie sind auf Seite 1von 12

A microscopic nuclear collective rotation-vibration model: 2D submodel

P. Gulshani

NUTECH Services, 3313 Fenwick Crescent, Mississauga, Ontario, Canada L5L 5N1
Tel. #: 647-975-8233; matlap@bell.net

The previous microscopic collective rotation-vibration model is improved to include


interaction between collective oscillations in a pair of spatial directions, and to remove many of
the previous-model approximations. As in the previous model, the nuclear Schrodinger equation
(instead of the Hamiltonian) is canonically transformed to obtain a Schrodinger equation for
collective rotation and vibration of a nucleus coupled to an intrinsic motion, with the related
constraints imposed on the wavefunction (rather than on the particle co-ordinates). The resulting
equation is then effectively linearized into three self-consistent, time-reversal invariant,
cranking-type equations using a variational method. The relation of the equations to the
phenomenological hydrodynamic collective Bohr-Davydov-Faessler-Greiner model is discussed.
To facilitate the solution of the equations and enhance physical insight, we consider in this article
the collective oscillations in only two space directions. For harmonic oscillator mean-field
potentials, the equations are then solved and applied to some light and rare-earth nuclei. The
computed ground-state rotational band excitation energy, quadrupole moment and reduced
electric quadrupole transition probabilities are found to agree favourably with measured data and
the results from mean-field and Sp(3,R) models.

PACS number: 21.60.Ev, 21.60.Fw, 21.60.Jz

I. INTRODUCTION
In a previous publication [1,2], we used a canonical transformation to derive from first
principles a model for collective rotation and vibration of a nucleus coupled to an intrinsic
motion. One of the objectives of the derivation was to circumvent the implementation
difficulties associated with other collective models derived from an application of a canonical
transformation to the description of the collective motion [3-16, refer to others in 17]. These
difficulties arose primarily from the unknown nature of the intrinsic or redundant coordinates,
and the complicated nature of the coupling between collective-intrinsic motions in the models.
The other objective of the derivation in [1,2] was to show a microscopic justification for, and
hence reveal the assumptions, approximations, and limitations that underlie the
phenomenological hydrodynamic collective rotation-vibration model of Bohr-Davydov-Faessler-
Greiner (BDFG) [18-22]. The BDFG model has had remarkable success in explaining the
observed collective features in deformed nuclei using adjustable parameters for the vibration and
rotation masses and the first two excitation energies.

Collective Hamiltonians have also been derived using group theory [23,24] and mean-field
approaches, such as time-dependent (Hartree-Fock-Bogoliubov) HFB, constrained cranked HFB,
and generator co-ordinate methods [25,26,27, and references therein] where the collective co-
ordinates and Hamiltonian are identified with the expectation values of respectively multipole
moment operators and microscopic nuclear Hamiltonian, and subsequently quantizing the
1
collective Hamiltonian using Hamilton's principle and usual quantization rules. Angular
momentum projection methods and microscopic interactions have also been used to investigate
automatic emergence of nuclear rotational spectrum [25,28]. Another successful approach,
discussed in [17,29,30,31], diagonalizes the A-nucleon kinetic energy plus a phenomenological
potential energy, with two to six fitting parameters including a pairing interaction in one case, in
a suitably selected subspace of the harmonic oscillator irreducible representation of the non-
compact group Sp(3,R), which is a generalization of the compact group SU(3).

The first objective of the derivation in [1,2] was achieved by transforming the Schrodinger
equation rather than the Hamiltonian, and by imposing the transformation-related constraints on
the wavefunction rather than on the particle coordinates. The model in [1,2] simplified the
governing equations in a number of ways, including restricting them to an axially symmetric
motion, and ignoring the interaction between oscillations in a pair of spatial directions, to mimic
the BDFG model. It predicted reasonably well the excitation energies in the ground-state
rotational band. However, the intrinsic quadrupole moment and reduced electric quadrupole
transition probabilities (B(E2)'s) were predicted to increase monotonically with the angular
momentum (J), whereas the measured and the Sp(3,R) model-predicted [29,30] quadrupole
moment decreased with J, and the corresponding B(E2) increased at low values of J and
decreased at higher values of J.

In this article, we remove many of the approximations used in the previous model [1,2], and
for simplicity and to obtain physical insight, we restrict the equations governing the collective
oscillations (but not the rotation and intrinsic motions) to two space dimensions.

II. DERIVATION OF MICROSCOPIC COLLECTIVE-INTRINSIC MODEL


The model is derived by transforming the nuclear Schrodinger equation to collective
Euler angles and vibration co-ordinates with constraints imposed on the wavefunction. The
transformation is performed in two steps.
In the first step, we use the microscopic rotational model derived in [32,33] and described
briefly here. We use the rotational-model product wavefunction:
J s 3
JM DMJ K ( )
K J
s JK ( xni ) , where J A JK ( xn i ) = 0 ,
xn j

k 1
s
jk x nk (1)

where J
MK is the Wigner rotation matrix, J A is the Ath component of the total angular momentum
operator, s ( s =1,2,3) are three non-zero, arbitrary anti-symmetric 3x3 matrices, and xn j
( n 1,..., A; j 1, 2,3, where A nuclear mass number) are the space-fixed nucleon co-ordinates. The
zero-angular momentum constraint on the non-rotational wavefunction JK together with the
choice of the Euler angles s in Eq. (1) ensures that, in the transformed Schrodinger equation, the
Coriolis interaction, coupling the rotation and intrinsic motions, vanishes. Transforming the
nuclear Schrodinger equation to the co-ordinates s and its conjugate angular momentum J s ,
integrating over s , ignoring the relatively small non-axial rotation terms and the terms
associated with the oscillations in the spatial direction 2, we obtain [1,2,32,33] the effective
rotation-intrinsic axially-symmetric rotor Schrodinger equation:

2
2 2 2 2 K 2
x V J ( J 1) K 2 JK E JK (2)
2M 2 MR1 2 MR3
2
n, j nj

A A
where M is the nucleon mass and R1 ( yn2 zn2 ) , R3 ( xn2 yn2 ) . In Eq. (2), the rigid-flow
n 1 n 1

kinematic moment of inertia components R1 and R3 appear as a consequence of the zero-angular-


momentum constraint and the choice of the Euler angles in Eq. (1).

In the second step, we transform Eq. (2) to the collective vibration co-ordinates R1 and R3
using the product wavefunction JK F1 ( R1 ) F3 ( R3 ) JK ( xnk ) , where the spherically symmetric
intrinsic (such as shell-model or HFB) wavefunction JK is subject to the constraints
JK R1 JK R3 0 . In place of the nuclear potential V in Eq. (2), we use the harmonic
A
oscillator mean-field potential bint2 rn2 bv21 R1 bv23 R3 , with the reduced strengths bint for the intrinsic
n 1

system, and bv1 and bv 3 for the collective oscillations. We then apply, as in [1,2], a constrained
variational method to linearize the transformed Schrodinger equation to obtain the following
three coupled self-consistent, time-reversal invariant, cranking-type Schrodinger equations and a
self-consistency equation:

2 d2 1 d J ( J 1) K 2 1 bv21 2
R1 R1 a1 3 R1 1 R1 R1 F1 0 (3)
dR1 dR1
2
4 8 4 4
2 d2 1 d K2 1 bv23 2
R
3 R a
3 3 R
1 3 R
3 3 R3 F3 0 (4)
dR3 dR3 4 4
2
4 4
A 2 A

n 1 B3 3 B1 bint rn int JK 0
2 2
(5)
n 1 n 1
a3 1 a1 3 s (6)

where:

1 1 d
a1 a3 3 , a3 a1 1 , ak JK Bk JK , k 4 Fk Rk Fk , k 1,3 (7)
4 4 dRk

d 1 A
k 4 Fk Fk , Bk x n k x nk (8)
dRk
2 n 1 xnk xnk

and 1 , 3 , int , and s are functions of the reduced energy 2ME 2 and the system parameters
in Eqs. (7) and (8).

3
III. SOLUTION OF EQS (3)-(6)
We readily obtain, from the literature as discussed in [1,2, 34,35], the solutions of Eqs. (3)-
(5) in closed forms1. In particular, Eq. (5) is solved exactly in Cartesian co-ordinate system
avoiding the approximate solution obtained in [1,2] using spherical co-ordinate system. The
solutions of Eqs. (3)-(5) are used to evaluate the parameters in Eqs. (7) and (8) and 1 , 3 , int ,
and s 2. In particular, we obtain the reduced excitation energy:

2b1 1 2b3 3 bv1 2 n1 a1 2k1 bv 3 2n3 a3 2k3


1 1 1 1 (9)
3 a1 1a3 1 a3 3 a1 1 3 3 1
2 2 4 4

where:

2k1 (a1 1) (a1 1)2 J ( J 1) - K 2 2 , 2k3 a3 1 ( a3 1) 2 K 2 (10)

mkf
3 1
b1 bint2 32 4 , b3 bint2 12 4 , a1 , a3 1 3 , k mk 1 2 (11)
2b1 2b3 mk 0

1 a 2k 3 1 a 2 k1
a1 a3 a3 1 3 , a3 a1 a1 3 1 (12)
2 8 bv 3 2 8 bv1

8bv1 a1 1 4bv 3 a3 1 8bv 3 a3 1 4bv1 a1 1


1 , 3 (13)
3 a1 2k1 1 3 a3 2k3 1 3 a3 2k3 1 3 a1 2k1 1

1 2 1 2
bv21 3 bv21 , bv23 1 bv23 (14)
16 16

n1 ,n3 0,1,2,3,... are the quantum numbers for the collective oscillations in respectively 1 and 3
spatial directions (and may be identified with the so-called beta and gamma band heads), k is
the total oscillator particle-occupation number in kth direction, mk is a harmonic oscillator
quantum number, and mkf is the value of mk for the last particle-occupied (Fermi) level.

First we solve Eqs. (13) and (14) for bv21 , bv23 , 1 , and 3 in terms of bv21 , bv23 , a1 , a3 , 2k1 , and
2k3 . We then solve iteratively Eqs. (12) for a1 and a3 and hence a1 , a3 , b1 , b3 , and in terms of
b , bv23 , bint2 , k , n1 ,n3 , J, and K. The solution also yields the value of the rotational-band cutoff
2
v1

1
In this article, we, for simplicity and to obtain physical insight, ignore the two constraints on the intrinsic
wavefunction JK mentioned above, because they may have relatively small effects as the model's reasonable
predictions and the calculations in [34,35] seem to indicate.
2
The solutions of Eqs. (3) and (4) given in this article differ from those of Faessler-Greiner rotation-vibration model
[20,21] in three respects: (i) our solutions are not limited to small amplitude oscillations about a mean deformation,
(ii) the kinematic moment of inertia is not an adjustable parameter but rather is a dynamical variable (specifically is
the rigid-flow moment), (iii) Eqs. (3) and (4) includes interaction between the two collective oscillations, and (iv)
the interaction between rotation-vibration and intrinsic motions is included.
4
angular momentum Jc, which occurs when the self-consistency among the Eqs. (3)-(6) is
violated, as discussed in [2].

In Eq. (9), the first two terms are the intrinsic energies, and they decrease gradually in value
with J. The third and fourth terms are the energy eigenvalues for the collective oscillations
(including the effects of centrifugal stretching J(J+1) and K 2 in 2k1 and 2k3 parameters in Eq. (10)
in respectively 1 and 3 spatial directions), and they increase in value with J. The fifth to eighth
terms are the energies arising from the interaction between the collective-vibration displacement-
dilation and intrinsic dilation-compression (i.e., the second terms in Eqs. (3) and (4)). The ninth
and tenth terms are purely the energies arising from the interaction between the collective
vibrations in the two spatial directions. The fifth to tenth terms in Eq. (9) are responsible for
reducing the value of the rigid-flow kinematic moment of inertia in the first 2 excited state to the
measured value, and increasing the moment in the higher excited states. A comparison between
the collective model predictions and those of the others collective and mean-field models is
given in [2].

The excitation energy of a member of the ground-state rotational band is defined by:
2
E J J J 0 (15)
2M

Generally, the moment of inertia J for a given member of a rotational band with angular
momentum J is defined by [36]:

2 J 4J 2
( MeV ) 1
(16)
2 EJ EJ 2

where EJ is either the predicted or measured excitation energy.

For the ground-state rotational band, K 0 , 2k3 0 , n1 0 , and n3 0 .

IV. APPLICATION OF MODEL TO SOME NUCLEI


We now present the results of an application of the model to the ground-state rotational
bands in 84 Be , 12
6 C , 10 Ne , 12 Mg , 14 Si , 66 Dy , and 68 Er . For each of the nuclei, (i) we determine
20 24 28 162 168

k in Eq. (11) from its Nilsson's self-consistent deformed-oscillator particle configuration [37],
and (ii) we choose the parameters bint , bv1 , and bv3 in Eqs. (3)-(5) to match as closely as possible
the predicted and measured excitation energies and B(E2)'s (or the quadrupole moment) of the
first excited 2 state while ensuring that the rotational-band cutoff angular momentum J c is as
high as possible.

The results in Tables 1 and 2 show that the model predicts the excitation energy EJ within
12
+3% and -7% for 84 Be , +2% and +9% for 6 C , -25% and +33% for 20
10 Ne , -14% and +18% for
24
12 Mg , -28% and +26% for 28
14 Si , -11% and +25% for 162
66 Dy , and within 0% and 3% for 168
68 Er .

5
These results show that EJ is reasonably-well predicted at the current stage of the model
development, but EJ is progressively over-predicted with J. This systematic over-prediction
appears to be also a feature of other models [17,27,30,38], and it may be eliminated in our model
by, among other things, imposing the zero angular momentum constraint in Eq. (1) on the
intrinsic system to suppress any spurious angular momentum excitations. As a consequence of
the aforementioned over-prediction, the predicted moment of inertia J decreases with J rather
than increases (as measured). The predicted and measured moment of inertia are about 1 to 3
times smaller than the rigid-flow moment.

Tables 1 and 2 show that the predicted rotational-band cut-off angular momentum J c is
realistically much higher than that predicted by the previous model [1,2], which neglected the
interaction between the oscillations in a pairs of spatial directions (discussed in Section I).

Tables 1 and 2 show that the quadrupole moment eQo and B(E2)'s are reasonably-well
predicted when the measurement uncertainties are considered. The quadrupole moment is
predicted to decrease with J as in the Sp(3,R) model [29,30], whereas it was predicted to increase
with J in the previous model [1,2] because it neglected the interaction between oscillations in a
pairs of spatial directions. Therefore, the interaction among the collective oscillations in
different spatial directions, and its consequential sharing of the momentum and energy among
the oscillations, seems to reduce the nuclear deformation. For this reason, the B(E2)'s are
predicted to increase at low values of J and decrease at higher values of J, as in the Sp(3,R) model
[29,30].

6
Table 1. Predicted/measured excitation energy ( EJ ), Cut-off J ( J c ), moments of inertia ( J ), eQo , B(E2)

EJ 2 J 2 2 rigflow 2 Predicted Measured


(MeV) eQo /B(E2) eQo /B(E2)
J ( MeV ) 1 ( MeV ) 1
model/exp e fm 2 / e 2 fm4 e fm 2 / e 2 fm4
model/exp
2 3.0/2.9 2.0/2.0 2.1 29/17 41 / 34 10.5
8
Be
Qo is for J=0
4 4 10.6/11.4* 1.9/1.7 2.1 24/15
J c 6
(HF-BCS pre-diction for 2 [39])
(28/- at J=0)
2 4.5/4.4 1.3/1.4 4.1 -18/6 -21 10.5 /11-99 Qo is for J=0 [40]
/ 8.5 for 2 [41,42]
12
C
6
4 15.3/14.1* 1.3/1.5 4.1 -27/7
J c 6 (-19/- at J=0)

2 1.2/1.6 4.9/3.7 9.4 59/68 70 17.5 J 0 /274-762 [40], 57 8 [42],


480 8 [43],
4 4.2/4.3 4.7/5.4 9.4 55/85 / 71 7 [42]
20
Ne
6
10

9.0/8.8 4.6/4.9 9.4 50/24 / 66 8 [42]


8
J c 10 16.0/12.0* 4.3/9.5 9.4 42/18 / 24 8 [42]
(58/- at J=0)
2 1.3/1.4 4.6/4.4 7.3 47/44 84 21 J0 /395-1097 [40] 119.3 25 [42],
425 29 [43]
24
Mg 4 4.3/4.1 4.8/5.1 7.4 47/51 / 95, + 21,-16 [42]
12
6 8.5/8.1 5.2/5.5 7.7 37/42 / 140, +193, - 49 [42]

8 13.6/13.2* 6.0/5.9 8.2 27/25 / 74, +148, - 29 [42]


J c 10 (46/- at J=0)
2 1.5/1.8 4.0/3.4 7.3 -60/72 38.5 21 J0 /30.5-352.2 [40], 72 9 [42],
317 17 [43]
28
14 Si 4 4.1/4.6 4.4/4.9 7.6 -72/147 / 96 8 [42]

6 8.8/8.5* 5.4/5.6 8.3 -90103/254 / 106 55 [42]


J c 8 (-61/- at J=0)
* No measured ground-state rotational-band energy level above this energy is reported in the Table of Isotopes and
Nuclear Data Sheets.

7
Table 2. Predicted/measured excitation energy ( EJ ), Cut-off J ( J c ), moments of inertia ( J ), eQo , B(E2)

EJ 2 J 2 2 rigflow 2 Predicted Measured


(MeV) eQo /B(E2) eQo /B(E2)
J ( MeV ) 1 ( MeV ) 1
model/exp e fm 2 / w.u.3 e fm 2 / w.u.
model/exp
2 0.08/0.09 78/69 289 788/225 -/ 204 3 [43]

4 0.26/0.27 78/78 289 787/321 -/ 289 12 [43]

6
0.54/0.55 78/78 289 786/353 -/ 301 17 [43]

8 0.92/0.92 78/81 289 786/368 -/ 346 17 [43]

10
1.40/1.38 78/84 289 784/377 -/ 350 23 [44]

12 1.99/1.90 78/88 289 783/383 -/ 330 40 [44]

162
66 Dy 14
2.68/2.49 78/91 289 782/386 -/ 330 40 [44]

16 3.47/3.14 78/96 289 780/388 N/A

18 4.37/3.83 78/101 289 778/389 N/A

20 776/389 N/A
5.37/4.58 78/105 289
N/A
22 6.47/5.35 78/111 289 773/388
N/A
24 7.67/6.15 78/117 289 771/387
J c 98
(788/- at J=0)

2 0.08/0.08 80/75 308 718/189 w.u.4 -/ 213 4 [45]

4
0.25/0.26 80/76 308 717/266 w.u. -/ 319 9 [45]

6
0.53/0.55 80/77 308 717/293 w.u. -/ 424 18 [45]

8
0.91/0.93 80/79 308 716/306 w.u. -/ 354 13 [45]
168
68 Er
10 1.38/1.40 80/81 308 714/313 w.u. -/ 308 13 [45]

12 1.96/1.94 79/84 308 713/317 w.u. -/ 345 18 [45]

14 2.65/2.57* 79/86 308 711/320 w.u. -/ 336 + 20, - 69 [45]


J c 94 (718/- at J=0)

* No measured ground-state rotational-band energy level above this energy is reported in the Table of Isotopes and
Nuclear Data Sheets.

V. CONCLUDING REMARKS
In this article, we improve the previous microscopic collective model version [1,2] of the
remarkably successful phenomenological hydrodynamic Bohr-Davydov-Faessler-Greiner
nuclear collective rotation-vibration model to include the interaction between collective
oscillations in a pair of spatial directions. We also remove many of the approximations used in
the previous model.

As done previously, the current model is derived from a canonical transformation of the
nuclear Schrodinger equation to collective rotation angles chosen, together with an spinless

3
In weisskopt unit, 1 w.u. = 52.3 e 2 fm 4 .
4
In weisskopt unit, 1 w.u. = 54.9 e 2 fm 4 .
8
intrinsic state such that the Coriolis interaction term in the transformed Schrodinger equation
vanishes, yielding a tri-axial rigid-flow rotor Schrodinger equation. For simplicity and to obtain
physical insight, this equation is then restricted to collective oscillations in two spatial directions,
and transformed to two collective vibration co-ordinates chosen to be two of the three principal-
axis components of the rigid-flow moment of inertia tensor. The resulting equation is then
linearized using a constrained variational method to obtain three coupled self-consistent, time-
reversal invariant cranking-type Schrodinger equations for the intrinsic and two rotation-
vibration motions, and a self-consistency equation. The isotropic spinless intrinsic wavefunction
is also subjected to zero collective-vibration displacement constraints.

The above transformations differ from the those in previous collective-model derivations in
the following ways: (i) the Schrodinger equation rather the Hamiltonian is transformed, (ii) the
associated constraints are imposed on the wavefunction rather on the nucleon co-ordinates, (iii)
one deals with the space-fixed particle coordinates and avoids intractable intrinsic co-ordinates
and constraints, and (iv) the Coriolis interaction term in the transformed Schrodinger equation is
eliminated by a judicious choice of the rotation angles and the intrinsic state.

For harmonic oscillator-type mean-field potentials for the vibration and intrinsic systems, we
solve the model cranking-type Schrodinger and the model parameters are evaluated in closed
forms. The resulting algebraic equations for the model parameters are solved iteratively. We
apply the model to the ground-state rotational band in 84 Be , 12
6 C , 10 Ne , 12 Mg , 14 Si , 66 Dy , and
20 24 28 162

168
68 Er . The results are encouraging.

The excitation energies are reasonable-well predicted, but they are progressively over-
predicted with the angular momentum J, as in other models, possibly because our model
currently neglects zero angular momentum constraint on the intrinsic system. This will be
investigated in a future study. The rotational-band cutoff angular momentum is predicted more
realistically at much higher values of J than that in the previous model [1,2]. The quadrupole
moment and hence the nuclear deformation is predicted to decrease with J (unlike that in the
previous model) because of the interaction among the collective oscillations in the different
spatial directions and the resultant sharing of momentum and energy among the oscillations. The
B(E2) is predicted to increase at low values of J and decrease at higher values of J , unlike that in
the previous model. These results emphasize the importance of the interaction among the
collective oscillations in the different spatial directions.

As discussed in [1,2], the compression-dilation interaction between the collective


oscillation-rotation and the intrinsic motions in the model reduces the kinematic rigid-flow
moment of inertia to the measured value in the first excited 2 state. This reduction is achieved
without using a pairing interaction, which is commonly used [46,47] to reduce the rigid-flow
value of the moment of inertia. The dilation interaction in our model is also responsible for
increasing the moment of inertia toward the rigid-flow value in the higher excited states.
However, the mechanism for the gradual and large changes in the moment of inertia in the model
9
closely parallels that in mean-field models as discussed in [1,2]. The dilation-compression
mechanism for reducing the kinematic rigid-flow moment of inertia in our model differs from the
shearing mechanism in other canonical-transformation-related models, where the shearing
mechanism is expected to increase the kinematic irrotational-flow moment of inertia.

In a future study, we will examine the effects of the collective oscillations in the third spatial
direction, the zero angular momentum constraint on the intrinsic system, the neglected
triaxiallity, and the model application to excited rotational bands.

VI. REFERENCES
[1] P. Gulshani, arXiv 1502.07590v4 [nucl-th] 12 May 2015.

[2] P. Gulshani, Can. J. Phys. 93, 1 (2015), dx.doi.org/10.1139/cjp-2015-0371.

[3] H.J. Lipkin, A. de Shalit, and I. Talmi, Nuovo Cimento 2, 773 (1955).

[4] R. Skinner, Can. J. Phys. 34, 901 (1956).

[5] W. Scheid and W. Greiner, Ann. Phys. (N.Y.) 48, 493 (1968).

[6] F.M. Villars and G. Cooper, Ann. Phys. 56, 224 (1970).

[7] D.J. Rowe, Nucl. Phys. A152, 273 (1970).

[8] F.M. Villars, Ann. Phys. (N.Y.) 56, 224 (1970).

[9] B. Fink, D. Kolb, W. Scheid, and W. Greiner, Ann. Phys. 69, 375 (1972).

[10] P. Gulshani and D.J. Rowe, Can. J. Phys. 54, 970 (1976).

[11] O.L. Weaver, R.Y. Cusson, and L.C. Biedenharn, Ann. Phys. 102, 493 (1976).

[12] B. Buck, L.C. Biedenharn, and R.Y. Cusson, Nucl. Phys. A317, 205 (1979).

[13] P. Gulshani, Phys. Letts. 77B, 131 (1978).

[14] H. Herold, M. Reinecke, and H. Ruder, J. Phys. G: Nucl. Phys. 5, 907 (1979).

[15] P. Gulshani, J. Phys. A: Math. Gen. 14, 97 (1981).

[16] P. Gulshani and A.B. Volkov, J. Phys. A: Math. Gen. 15, 47 (1981).

[17] D. J. Rowe, Rep. Prog. Phys. 48, 1419 (1985).

[18] A. Bohr, Kgl. Dan. Mat. Fys. Medd. 26, 14 (1952).

[19] A.S. Davydov and G. F. Fillippov, Nucl. Phys. 8, 237 (1958).

[20] A. Faessler and W. Greiner, Z. Physik 180, 425 (1962).


10
[21] J.M. Eisenberg and W. Greiner, Nuclear Theory (North-Holland, Amsterdam, 1970).

[22] P. Ring and P. Schuck, The Nuclear Many-Body Problem (Springer-Verlag, N.Y., 1980).

[23] H. Ogura, Nucl. Phys. A207. 161 (1973).

[24] G. Rosensteel and N. Sparks, J. Phys. A: Math. Theor. 48, 445203 (2015).

[25] R.M. Lieder and H. Ryde, in Adv, Nucl. Phys.10, chapter 1, edited by M. Baranger and E.
Vogt (Plenum Press, N.Y., 1978).

[26] L. Prochniak and S.D. Rohozinski, J. Phys. G: Nucl. Part. Phys. 36, 123101 (2009).

[27] J.P. Delaroche, M. Girod, J. Libert, H. Goutte, S. Hilaire, S. Peru, N. Pillet, and G.F.
Bertsch, Phys. Rev. C81, 014303 (2010).

[28] M.A. Caprio, P. Maris, and J.P. Vary, Phys. Letts. B719, 179 (2013).

[29] J.P. Draayer, K.J. Weeks, and G. Rosensteel, Nucl. Phys. A413, 215 (1984).

[30] G. Rosensteel, J.P. Draayer, and K.J. Weeks, Nucl. Phys. A419, 1 (1984).

[31] A. Dreyfuss, K. D. Launey, Tomas Dytrych, J. P. Draayer, and C. Bahri, J. Phys.:


Conference Series 403, 012024 (2012).

[32] P. Gulshani, Nuclear Physics A 832, 18 (2010).

[33] P. Gulshani, Nuclear Physics A852, 109 (2011).

[34] P. Gulshani, arXiv: 1407.3635v2 [nucl-th] (October 24, 2014).


[35] P. Gulshani, arXiv:1401.5328 [nucl-th] (January 21, 2014).

[36] R.A. Sorensen, Rev. Mod. Phys. 45, 353 (1973).

[37] S.G. Nilsson and I. Ragnarsson, Shapes and Shells in Nuclear Structure (Cambridge, U.K.,
University Press, 1995).

[38] A. Faessler, in Proc. Conf. on High Angular Momentum Properties on Nuclei, Oak Ridge,
Tennessee, November 2-4, 1982, edited by N.R. Johnston (Harwood Academic Publishers,
N.Y.).

[39] S. Raman, C.W. Nestor, Jr., and P. Tikkanen, At. Nucl. Data Tables 78, 1 (2001).

[40] Table of Isotopes, Seventh edition, edited by C.M. Lederer and V.S. Shirley, (John Wiley
and Sons, Inc., New York, U.S.A., 1978).

[41] Y. Abgrall, G. Baron, E. Caurier, and G. Monsonego, Nucl. Phys. A131, 609 (1969).

11
[42] Y. Abgrall, B. Morand, and E. Caurier, Nucl. Phys. A192, 372 (1972).

[43] Y. Horikawa, Progr. Theor. Phys. 47, 867 (1972).

[44] C.W. Reich, Nucl. Data Sheets 108, 1807 (2007).

[45] C.M. Baglin, Nucl. Data Sheets 111, 1807 (2010).

[46] S.G. Nilsson and O. Prior, Mat. Fys. Medd. Dan. Vid. Selsk. 32, No. 16 (1961).

[47] X. Guan, K. D. Launey, Y. Wang, F. Pan, and J. P. Draayer, Phys. Rev. C92, 044303
(2015).

12

Das könnte Ihnen auch gefallen