Вы находитесь на странице: 1из 28

Deformations and theoretical nuclear

models

Girija. K.K. “Studies of shape changes of deformed nuclei and its effects on
cluster emission” Thesis. Department of Physics , University of Calicut, 2012
Chapter 2

Deformations and theoretical

nuclear models

2.1 Nuclear deformations

The nuclear shapes and deformations have been interesting topics to the nuclear physi-

cists, since the identification of nucleus. The highly complicated nature of structure

and properties of nuclei may be one of the reasons for this. Moreover, shape is one of

the fundemental properties of nuclei. Till date, no theory is developed to describe the

nuclear structure and properties completely, since the knowledge about the forces which

shape the nucleus is very limited. The configuration dependent forces present inside

the nucleus are mainly the nuclear force between nucleons and Coulomb force between

protons. The shell effects and pairing correlation also contribute to the determination

of nucleonic configuration. The atomic nuclei exhibit spherical, quadrupole and higher

14
2: Deformations and theoretical nuclear models

order multipole deformed shapes, even though the quadrupole deformed shapes are

mostly discussed. Due to the interplay between single particle and collective degrees

of freedom, the coexistence of different shapes at the same spin and similar energies is

also not rare [1],[2],[3],[4].

Nuclei having spherical shape in their ground state (g.s) are few in number [2]. The

deformed nuclei are classified into prolate, oblate and triaxial. Prolate and oblate nuclei

are axially symmetric. If the third axis of the nucleus is longer than the others, the

nucleus is prolate and if it is shorter, the nucleus is oblate. For triaxial nuclei, the three

axes are different. In nature, prolate nuclei dominate over oblate ones [3]. It is found

that 86 % of the even-even nuclei are prolate in the ground state [5] and triaxial shapes

are very rare for them. The effect of Coulomb repulsion between protons is to deform

the nucleus more into an elongated shape than to a flattened shape. The difference

in the volume element of the collective coordinates between prolate and oblate shapes

is pointed out to be another reason for the prolate dominance over oblate shape. The

spin-orbit potential (coupling) between nucleons plays a role favouring stable prolate

shape for nuclei [5], [6]. The shell structure of nuclei is also responsible for the variety of

shapes, depending on the position of Fermi level between two closed shells [4]. Prolate

shape occurs just after closed shells and towards the end of closed shells, oblate shape

is observed.

The nuclear deformation is characterized by two collective parameters, the defor-

mation parameter β or ε and the triaxiality γ. Positive and negative quadrupole

deformations (β2 or ε2 ) correspond to prolate and oblate shapes respectively. For

15
2: Deformations and theoretical nuclear models

γ = 60o , 0, −60o and −120o , the nucleus is axially symmetric and it is triaxial for

all other γ values. γ = 0 and 60o represent prolate and oblate shapes respectively. The

variation of nuclear shapes with respect to deformation and triaxiality parameters are

illustrated in Figure 2.1.

Depending on the extent of deviation from spherical symmetry, the deformed nuclei

fall into different groups. Nuclei with major to minor axis ratio around 1.3 : 1 are

normally deformed and those with 1.5 : 1 are highly deformed. If the ratios are 2

: 1 and 3 : 1, the nuclei are superdeformed (SD) and hyperdeformed respectively

[2]. Nuclear deformation which is the departure from spherical shape without density

change is expressed in terms of the shape parameters αλµ and spherical harmonics

Yλµ (θ, φ) [7],[8],[9], as:

X
R(θ, φ) = R0 [1 + αλµ (t)Yλµ (θ, φ)] (2.1)
λµ

where R(θ,φ) is the distance of the nuclear surface at angles θ and φ from the centre

and R0 is the radius at spherical equilibrium. For each mode of order λ, µ has (2 λ + 1)

values i.e., from −λ to +λ. λ = 1 corresponds to dipole oscillation, λ = 2 to quadrupole

oscillation and λ = 3 to octupole oscillation. For quadrupole shapes,

X
R = R0 [1 + α2µ Y2µ (θ, φ)] (2.2)
µ

16
2: Deformations and theoretical nuclear models

Figure 2.1: Variation of nuclear shapes with deformation and triaxiality parameters.

The quadrupole deformation parameter β2 and the triaxiality γ are defined as [8],[10]:

α20 = β2 cosγ (2.3)

and

1
α22 = α2−2 = √ β2 sinγ (2.4)
2

so that

X
|α2µ |2 = (β2 )2 (2.5)

17
2: Deformations and theoretical nuclear models

since, α21 = α2−1 = 0. ε2 is defined as [11]:

ε2 = (Rmajor − Rminor )/R0 (2.6)

and the quadrupole deformations β2 and ε2 are related by [11], [12] :

ε2 = 0.95β2 or β2 = 1.06ε2 (2.7)

The deformation leads to change in potential energy surface (PES) of the nucleus.

Thus the calculation of PES can give information about the nature of shape evolution

taking place in nuclei at high angular momentum. At the minimum potential energy

(PE), the nucleus will be in equilibrium. Hence the deformation corresponding to the

minimum PE decides the shape of the nucleus in its ground state.

In the case of a spherical nucleus, according to the shell model, the energy states are

grouped into different shells and there is a large separation between these shells. Nuclei

with closed shells are having magic number of protons and neutrons and are stable in

their ground state. Then solving the Shrodinger equation for a spherical potential well,

all the energy states are known. When there are partially filled shells, the nucleus is

deformed and energy states are different from those for a spherical well. But as the

deformation is increased to superdeformation, it is found that new energy levels are

grouped together showing shell closure property with a different set of magic numbers.

Stable deformed nuclei are common in the rare earth (lanthenides: 50 < Z < 82, 82

18
2: Deformations and theoretical nuclear models

< N < 126) and transuranic elements (actinides: Z > 82, N >126). Light nuclei having

partially filled shells are also deformed. Superdeformed nuclei were first found in the

region of nuclear mass A=150, 190 and A > 220. In the periodic table, nuclei with

mass number A > 220 are superdeformed in their ground state. Many superdeformed

nuclei are discovered in distinct regions with mass number around 60, 90, 130, 150, 190

and 240 [2]. In the mass range 150<A<190 and A>220, the nuclei are found to be

deformed in their ground state [13].

2.2 Nuclear models

No single theory can describe the structure and properties of a complex nucleus com-

pletely. Still, many attempts were made and being made to reveal its structure. Various

phenomenological models have been used to describe the observed behaviour of atomic

nuclei.

2.2.1 Liquid drop model

Liquid drop model was first proposed by Neils Bohr and Kalckar in 1937 [14]. According

to this model, the nucleus is very similar to a liquid drop and in the absence of external

forces, it is spherical in shape. Major part of the forces present is due to the nuclear

and Coulomb interactions. This model suggests a semiempirical mass formula for the

19
2: Deformations and theoretical nuclear models

binding energy of the nucleus, which is known as Weizsacker formula as:

Z(Z − 1) (A − 2Z)2 ap
EB = av A − as A2/3 − ac 1/3
− a a + (±, 0) 3/4 (2.8)
A A A

The first term is the volume energy which is due to the short ranged nuclear forces,

assuming that all nucleons contribute equally to the binding energy. The second term

gives a reduction in binding energy due to the difference in the force experienced by the

nucleons in the interior and on the surface of the nucleus. The Coulomb repulsive force

between protons is to reduce the binding energy which is given by the third term. The

fourth term arises due to the asymmetry in the proton and neutron numbers. The last

term gives the pairing energy which gives an additional strength for binding in even-

even nuclei and a reduction in odd-odd nuclei. This shows that nuclear force favours

pairing of similar nucleons.

A set of coefficients that gives a good fit with the experimental data is: av = 14.1

MeV, as =13 MeV, ac =0.595 MeV, aa = 19 MeV and ap = 33.5M eV [15].

Liquid drop model was successful in giving an account of the systematic behaviour

of binding energy per nucleon as a function of mass number and it could justify the

observed fission barrier. But it could not reproduce the observed extra stability of

nuclei having magic number of similar nucleons, spin, parity etc.

20
2: Deformations and theoretical nuclear models

2.2.2 Nuclear shell model

Shell model was an attempt to reproduce the observed magic numbers and to explain

the unusual stability of nuclei having these number of similar nucleons. The basic

assumption of this model is that each nucleon is moving inside the nucleus in an average

potential due to the other nucleons i.e., the interaction between the nucleons is very

weak.

The energy levels are filled in accordance with the Pauli’s exclusion principle. Each

energy level has an upper limit 2(2l+1) for the number of nucleons that can be ac-

commodated. This model gives large energy gaps between particular groups of levels

forming closed shells which exhibits extra stability. The proton number and neutron

number corresponding to the shell closures are known as magic numbers. All the paired

nucleons form an inert core and the nuclear properties are attributed to the unpaired

valence nucleons.

Schrodinger equation for the motion of a particle with mass M in a mean spherically

symmetric potential is of the form [16]:

2M
[▽2 + (E − V (r))]Ψ(r) = 0 (2.9)
~2

where E is the energy eigen value. The general solution of this equation is of the form:

Ψnlm (r) = Unl (r)Ylm (θ, φ) (2.10)

21
2: Deformations and theoretical nuclear models

with n, l and m being the quantum numbers to determine eigen states corresponding

to Enl .

Since the exact nature of nuclear potential is unknown, various types of potentials

are taken as approximations.

2.2.2.1 Square well potential

Square well potential is of the form [9]:

V (r) = −U0 , r ≤ R (2.11)

V (r) = ∞, r ≥ R (2.12)

Using this infinite square well potential, the radial part of the solution of Schrodinger

equation are spherical Bessel functions [16]:

Unl (r) = jl (kr) = r−1/2 Jl+1/2 (kr) (2.13)

with

p
k= 2M E/~2 (2.14)

The eigen values are obtained as [9] :

~2 knl
2
Enl = (2.15)
2M

22
2: Deformations and theoretical nuclear models

The particle occupancy for a level l is 2(2l+1) and the closed shells occur with
P
proton and neutron numbers 2(2l + 1). According to the shell model with square

well potential, the predicted proton and neutron numbers corresponding to shell closure

are 2, 8, 18, 20, 34, 40, 58, 68, 70, 92, 106, 112, 138 and 156 [16]. But the experimentally

observed magic numbers are 2, 8, 20, 50, 82 and 126 i.e., this model could reproduce

only the three observed magic numbers 2, 8 and 20.

2.2.2.2 Harmonic oscillator potential well

Harmonic oscillator potential is given as [16]:

1
V (r) = M ω 2 r2 (2.16)
2

where ω is the classical frequency of the oscillator. With this potential, Schrodinger

equation leads to the differential equation for the radial part [16]:

d2 l(l + 1) 2M
[ 2− + 2 (Enl − V (r))]Rnl (r) = 0 (2.17)
dr r2 ~

with

Rnl (r) = r unl (r) (2.18)

23
2: Deformations and theoretical nuclear models

The solutions are

1
Rnl (r) = Nnl exp(− νr2 )rl+1 vnl (r) (2.19)
2

where, ν = (M ω)/~ and vnl (r) is the Laguerre polynomial. Then the normalized eigen

functions are:

Rnl (r)
Ψnlm (r) = Ylm (θ, φ) (2.20)
r

and the corresponding eigen values are:

Enl = ~ω(2n + l − 1/2) or EΛ = ~ω(Λ + 3/2) (2.21)

with n=1, 2, 3,..., l = 0, 1, 2,... and

Λ = (2n + l − 2) (2.22)

The degeneracy corresponding to each l value is 2(2l+1). The eigen states corresponding

to the same value of 2n+l are also degenerate. Since 2n=Λ-l+2= even, for a given value

of Λ, the degenerate eigenstates are [16]:

Λ+2 Λ
(n, l) = ( , 0), ( , 2), ...., (2, Λ − 2), (1, Λ) f or Λ = even
2 2
Λ+1 Λ−1
(n, l) = ( , 1), ( , 3), ..., (2, Λ − 2), (1, Λ) f or Λ = odd (2.23)
2 2

24
2: Deformations and theoretical nuclear models

so that the number of neutrons or protons with the eigen value EΛ for Λ=even (or odd)

is (setting l=2k or 2k+1 according to whether Λ is even or odd)

Λ/2
X
NΛ = 2[2(2k) + 1] f or even Λ
k=0
(Λ−1)/2
X
NΛ = 2[2(2k + 1) + 1] f or odd Λ
k=0

NΛ = (Λ + 1)(Λ + 2) in either case. (2.24)

Then the accumulating total number of particles for all levels upto Λ is:

X 1
NΛ = (Λ + 1)(Λ + 2)(Λ + 3) (2.25)
Λ
3

In the shell model using Harmonic oscillator potential well, the nucleon numbers

corresponding to the shell closures are calculated as: 2, 8, 20, 40, 70, 112 and 168 [16].

Here also the experimentally observed magic numbers above 20 are not reproduced.

2.2.2.3 Spin-orbit potential

In order to reproduce the observed magic numbers 28, 50, 82 and 126, Mayer [17] and

Haxel, Jensen and Suess [18] introduced an additional spin-orbit term to the centrally

symmetric potential. The spin-orbit potential is proprtional to l.s and its introduction

causes the splitting of the j ± 1/2 levels [16]. The sign of the radial part of spin-orbit

potential is so selected that the j + 1/2 level is lowered to the lower band and the

j − 1/2 level is raised. This rearrangement could reproduce all the observed magic

25
2: Deformations and theoretical nuclear models

numbers upto 126.

The spherical shell model with spin-orbit coupling could not reproduce the actual

energy eigen values. To rectify this, a term proportional to l2 is taken into account in

the Hamiltonian and it takes the form:

−~2 2 1
H= ▽ + M ω 2 r2 + Cl.s + Dl2 (2.26)
2M 2

and the energy eigen values are obtained as:

E = (Λ + 3/2)~ω + C(l.s) + Dl2 (2.27)

The spherical shell model with spin-orbit coupling and an l2 term succeeded in justi-

fying the observed shell closures and the energy eigen values. But in nature most of the

nuclei show deviation from spherical symmetry and appreciable quadrupole moments

are noticed in different regions of the periodic table. The reason is the polarization of

the closed core by the valence nucleons. In order to describe such nuclei, the present

shell model is to be modified. Nilsson succeeded in tackling the problem of deformed nu-

clei by some modifications to the spherical shell model, the details of which is discussed

under the section Collective models.

Due to the enormous dimension of the configuration space involved, it is not prac-

tically possible to use spherical shell model for heavy nuclei. A variational approach to

the shell model represented by Monster and Vampir calculations [19] may be adopted to

26
2: Deformations and theoretical nuclear models

get rid of from this situation. This variational shell model calculations can be performed

in the case of medium mass and heavy nuclei [10].

2.2.3 Collective models

A collective excitation is characterised by the movement of a large number of nucleons.

Rotation is a typical example of collective degree of freedom in nuclei [12]. The shape

variation of nuclei can be considered as another collective degree of freedom. Since

nuclear deformation is a collective property, only collective models can successfully

describe the behaviour of deformed nuclei. One can not define collective rotation around

a symmetry axis, since such a rotation would change only a trivial phase factor in the

wave function. Such an unchanged wave function is in contrast to collective rotation.

Instead, collective rotation is characterised by small angular momentum contributions

from a large number of particles, i.e., the wave functions of these particles change slowly

with increasing angular momentum. This implies that only deformed nuclei can rotate

collectively and if the nucleus is axially symmetric, the only possibe rotation axis is

perpendicular to the symmetry axis. For collective rotation of a nucleus [12],

L2
Hrot = (2.28)
2

where  is the moment of inertia and L is the collective angular momentum, which

equals the total angular momentum I (total spin) in pure collective rotation. Then the

27
2: Deformations and theoretical nuclear models

rotational spectrum is represented by [12]:

~2
EI = I(I + 1) (2.29)
2

Since only deformed nuclei exhibit rotational spectra, it should be possible to determine

deformation from the occurance of rotational bands. The moment of inertia can also

be extracted from measured rotational bands. A rotational band is a group of closely

spaced states differing in angular momentum slightly. In deformed nuclei, the valence

particles add their spin to the angular momentum due to rotation.

2.2.3.1 Nilsson potential

To incorporate the effects of deformation in nuclear properties, Nilsson modified the

shell model by introducing a self consistent deformed potential instead of a spherically

symmetric harmonic oscillator potential. The modified single particle oscillator poten-

tial Nilsson used is the axially symmetric oscillator potential with spin-orbit coupling

and a term proprtional to l2 . Thus the triaxial single particle Nilsson Hamiltonian is

given as [16]:

H = H0 + Cl.s + Dl2 (2.30)

where,

~2 ′ 2 1 ′ ′ ′
H0 = − ▽ + m(ωx2 x 2 + ωy2 y 2 + ωz2 z 2 ) (2.31)
2m 2

28
2: Deformations and theoretical nuclear models

′ ′ ′
with x , y and z being the coordinates in a frame fixed with the nucleus. Taking

cylindrical symmetry, the deformation parameter ε2 is introduced in such a way that

[12]:

1
ωx = ωy = ω0 (1 + ε2 ) (2.32)
3

and

2
ωz = ω0 (1 − ε2 ) (2.33)
3

with

ωx ωy ωz = a constant (2.34)

which is the condition for constant volume of the nucleus. Equations 2.32 to 2.34 relate

ω0 and ε2 as [16]:

4 16
ω0 (ε2 ) = ω00 (1 − (ε2 )2 − (ε2 )3 )−1/6 (2.35)
3 27

with ω00 being the value of ω0 for spherical nucleus. The deformation parameters ε2 and

β2 are related as:

r
3 5
ε2 = β2 (2.36)
2 4π

29
2: Deformations and theoretical nuclear models

The Hamiltonian H0 may be written as the sum of a spherical part and a deformation

term as:

H0 = H00 + Hε2 (2.37)

where,

1
H00 = ~ω0 [− ▽2 +r2 ] (2.38)
2

and

r
4 π 2
Hε2 = −ε2 ~ω0 r Y20 (2.39)
3 5

with

r
mω0 ′
r= r (2.40)
~

The base vectors used by Nilsson are |N lΛΣ > with N being the total number of oscil-

lator quanta, l, Λ, and Σ being the quantum numbers corresponding to the operators

l2 , lz and sz respectively. Now, the total Hamiltonian takes the form:

H = H00 + Hε2 + Cl.s + Dl2 = H00 + κ~ω00 R (2.41)

30
2: Deformations and theoretical nuclear models

with

1 C
κ=− (2.42)
2 ~ω00

and

R = ηU − 2l.s − µl2 (2.43)

where,

D
µ=2 (2.44)
C

ε2 ω0 (ε2 )
η= (2.45)
κ ω00

and

r
4 π 2
U =− r Y20 (2.46)
3 5

µ determines the depression of energy levels and κ decides the spin-orbit coupling. The

diagonalization of the matrix R leads to the energy eigen values:

E = (N + 3/2)~ω0 (ε2 ) + κ~ω00 rN Ω (2.47)

31
2: Deformations and theoretical nuclear models

with Ω = Λ + Σ, which represents the component of the nucleon total angular

momentum along the nuclear axis. Here, the spacing of oscillator level is ~ω0 . Nilsson

19 23
model succeeded in reproducing the energy levels, especially of O and N a.

2.2.3.2 Cranked Nilsson-Strutinsky shell correction method

Cranked Nilsson-Strutinsky shell correction method enables one to do systematic com-

putation of nuclear potential energy as a function of angular momentum and deforma-

tion. This model accounts for the collective rotation of the deformed nucleus around an

axis perpendicular to the symmetry axis, as suggested by Inglis [20]. Here the rotation

of an average field, unsymmetric with respect to the rotation axis introduces a time

dependence to the Schrodinger equation. Considering a rotating frame, with Z as the

symmetry axis and X as the cranking axis, the problem may be reduced to a stationary

equation [21]:


i~ ψintr = H ω ψintr (2.48)
∂t

where,

H ω = Hintr − ~ωIx (2.49)

is called the cranking Hamiltonian or Routhian. Here, intr stands for intrinsic i.e.,

Hintr is the Hamiltonian and ψintr - the corresponding eigen vector in the rotating frame

(body fixed frame rotating with an angular frequency ω). Ix denotes the X component

32
2: Deformations and theoretical nuclear models

of total angular momentum. Since H ω does not depend on time, the solution of the

equation can be reduced to the eigen value problem of H ω .

Nilsson made use of the modified single particle oscillator potential [12]:

p2 1
h0 = + m(ωx2 x2 + ωy2 y 2 + ωz2 z 2 ) + Cl.s + D(l2 − < l2 >) (2.50)
2m 2

where C = - 2κ~ω0 and D = -κµ~ω0 , κ and µ being the Nilsson parameters. The

introduction of the term − < l2 > is to cancel the effect of l2 i.e., the widening of radial

shape due to l2 is compensated by the compression due to the average value − < l2 >.

The deformation produces difference in the oscillator frequencies in the X, Y, and Z

axes.

The constraint incompressibility of nuclear matter is taken into account by restrict-

ing a constant volume to the deformed nucleus. This is made possible by varying the

cranking frequency ω0 (ε2 , γ) from its value at spherical shape ω00 [22] and the three

oscillator frequencies are given as [12], [22]:

2 2π
ωx = ω0 [1 − ε2 cos(γ + )] (2.51)
3 3

2 2π
ωy = ω0 [1 − ε2 cos(γ − )] (2.52)
3 3

2
ωz = ω0 [1 − ε2 cosγ] (2.53)
3

33
2: Deformations and theoretical nuclear models

so that

ωx ωy ωz = (ω00 )3 (2.54)

Here the calculations are carried out in a stretched coordinate system [22], [23].

The diagonalisation of the Hamiltonian gives the eigen values eωi and the eigen vector

ψiω . The single particle energies in the laboratory system and the single particle spin

contributions mi are obtained as [22]:

ei =< ψiω |h0 |ψiω > (2.55)

and

mi =< ψiω |jx |ψiω > (2.56)

with h0 being the single particle hamiltonian. The total single particle energy and spin

are defined as:

X X X
Esp = ei = eωi + ~ω mi (2.57)
occ occ occ

X
I= mi (2.58)

Here, summation is over all occupied orbitals. Even though this model predicts the

ground state spins and ground state deformations very well, the absolute binding energy

34
2: Deformations and theoretical nuclear models

can not be reproduced. The reason is that, binding energy is calculated by summing

over the single particle energies, but it is a bulk property. A small change in single

particle energy brings considerable error in the binding energy. Then to get the correct

value we have to use a combination of rotating liquid drop model and the deformed

shell model. The shell effects are deformation dependent, and a shell correction as

suggested by Strutinsky [12],[24],[25],[26] is required. In the liquid drop model, the

smooth behaviour of binding energy is due to an average level density and the observed

oscillating part is due to shell effects. i.e., the total energy:

E = ELDM + Eosc (2.59)

The shell model energy:

Esh = Ẽ + Eosc (2.60)

Taking

X
g= δ(e − ei ) (2.61)
i

as the level density, the total spin:

Z λ X
I= g2 deω = < mi > (2.62)
−∞ i

35
2: Deformations and theoretical nuclear models

and

Z λ
Esp = g1 eω deω + ~ωI (2.63)
−∞

where λ is the Fermi energy. Considering Strutinsky smoothed level density g̃, we have:

Z λ̃
I˜ =
X
g̃2 deω = < m̃i > (2.64)
−∞ i

and

Z λ̃
Ẽsp = g̃1 eω deω + ~ω I˜ (2.65)
−∞

with λ̃ being a well-defined Fermi energy. The protons and neutrons are coupled, with

identical potentials and frequency and the total spin is

N Z
I = I˜x = < I˜x >ων + < I˜x >ωπ
X X
(2.66)
ν=1 π=1

Here, ω may be selected for a particular spin. Now, the shell energy is given as:

Eshell (I) = Esp (I) − Ẽsp (I) (2.67)

with the total energy which is dependent on I and ε̄:

Etot (ε̄, I) = ERLD (ε̄, I) + EShell (ε̄, I) (2.68)

36
2: Deformations and theoretical nuclear models

where,

ε̄ = (ε2 , γ, ε4 ) (2.69)

The pairing effect is ignored, since for I > 30~, it is not significant. ERLD is the energy

in the rotating liquid drop model [22],[27],[28], which is given as:

1
ERLD = ELD − Irig ω 2 + ~ω I˜ (2.70)
2

The energy in the liquid drop model is given as [27]:

N −Z 2 3 e2 Z 2 5π 2 d 2
ELD = −av (1 − κv ( ) )A + [Bc (ε̄) − ( )]
A 5 Rc 6 Rc
N − Z 2 2/3 √
+as (1 − κs ( ) )A Bs (ε̄) (±, 0)12/ A (2.71)
A

with +, - and 0 are for odd-odd, even-even and odd-even nuclei respectively.

ELD = [2χ(Bc − 1)as + (Bs − 1)] (2.72)

with Bc = ECoul (ε̄)/ECoul (ε̄ = 0) and Bs = Esurf (ε̄)/Esurf (ε̄ = 0) and the fissility

parameter χ = (Z 2 /A)/45. The moment of inertia about the X axis is:

Irig + 2M b2 1 AM (Ry2 + Rz2 ) 2M b2


= + (2.73)
~2 5 ~ ~2

where 2M b2 is the diffuseness correction to the moment of inertia [29]. Energy value

37
Bibliography

corresponding to each combination of I, ω, ε2 and γ were then estimated.

Bibliography

[1] P.M. Walker, F.R. Xu and D.M. Cullen, Phys. Rev. C 71, (2005) 067303.

[2] Renee Lucas, Europhysics news 32 (1), (2001) 5.

[3] I. Maqbool, P.A. Ganai and J.A. Sheikh, Proceedings of the International

Symposium on Nuclear Physics 54, (2009) 164.

[4] H.Sagawa, X.R. Zhou and X.Z. Zhang, Phys. Rev. C 72, (2005) 054311.

[5] Naoki Tajima and Norifumi Suzuki, Phys. Rev. C 64, (2001), 037301.

[6] Satoshi Takahara, Naoki Onishi, Yoshifumi R. Shimizu and Naoki Tajima,

Physics Letters B702 (5), (2011) 429.

[7] Samuel S. M. Wong, Introductory Nuclear Physics, (Prentice - Hall of India, Pvt

Ltd, New Delhi 2007).

[8] Ashok K. Jain and P. Arumugam, Mean field description of nuclei, ed. Y.K. Gamb-

hir (Narosa Publishing House, New Delhi 2006) p.115.

[9] K. Heyde, Basic ideas and concepts in Nuclear physics - An introductory approach,

ed. Douglas F Brewer (Overseas Press, India, Pvt. Ltd. 1998).

[10] W. Nazarewicz and I. Ragnarsson in Handbook of Nuclear Properties, ed. Dorin

N.Poenaru, Walter Greiner, (Oxford University Press, New York 2008) p. 80.

38
Bibliography

[11] Raymond A. Sorensen, Rev. of Modern Physics 45 (3), (1973) 353.

[12] Sven Gosta Nilsson and Ingemar Ragnarsson, Shapes and shells in nuclear struc-

ture, (Cambridge University Press, Cambridge 1995).

[13] John Lilley, Nuclear Physics - Principles and applications, ed. D.J. Sandiford, F.

Mandl, A.C. Phillips (John Wiley & Sons, Ltd, New York 2002).

[14] Aage Bohr, Ben R. Mottelson, Nuclear structure, Vol. II, (World Scientific Pub-

lishing Co. Pte.Ltd. Singapore 1997).

[15] Arthur Beiser, Concepts of Modern Physics, (TATA Mc Graw-Hill Publishing Com-

pany Ltd. New Delhi 1997).

[16] R.R. Roy and B.P. Nigam, Nuclear Physics - Theory and Experiment, (New Age

International (P) Ltd., Publishers 1996).

[17] M.G. Mayer, Phys. Rev. 75, (1949) 1969.

[18] O. Haxel, J.H.D. Jensen and H.E. Suess, Phys. Rev. 75, (1949) 1766.

[19] K.W. Schmid, F. Grummer and A. Faessler, Phys. Rev. C29, (1984) 291.

[20] D.R. Inglis, Phys. Rev. 96, (1954) 1059.

[21] M.J.A. de Voigt, J. Dudek, Z. Szymansky, Rev. of Modern Physics 55, (1983) 949.

[22] T.Bengtsson, I.Ragnarsson and S.Aberg in Computational Nuclear Physics I, ed.

K. Langanke, J.A. Maruhn, S.E. Koonin (Springer-Verlag, Berlin 1991) p. 51.

39
Bibliography

[23] H. Miri-Hakimabad and A. Kardan, AIP Conference Proceedings-Carpathian

summer school of Physics-Exotic nuclei and Nuclear/Particle/Astrophysics 1304,

(2010) 369.

[24] V.M. Strutinsky, Nucl. Phys. A95, (1967) 420.

[25] V.M. Strutinsky, Nucl. Phys. A122, (1968) 1.

[26] J.N. De in Physics of rotating nuclei, ed. S.N. Mukherjee and Y.R. Waghmare,

(New age International publishers, New Delhi 1995) p.70.

[27] W.D. Myers and W.J. Swiatecki, Ark. Fys. 36, (1967) 343.

[28] G.Vijayakumari and V. Ramasubramanian, Int. Journal of pure and applied

sciences 2 (2), (2008) 36.

[29] Chandrasekaran Anu RADHA, Velayudham RAMASUBRAMANIAN and Em-

manuel James Jebaseelan SAMUEL, Turk J Phys 34, (2010) 159. .

40

Вам также может понравиться