Вы находитесь на странице: 1из 29

Chapter 2

Theoretical Models

2.1 Introduction
Atomic nuclei are complex many-body quantum system composed of mainly two kinds
of particles namely protons and neutrons which are bound together with strong nuclear
force between them. The protons and neutrons are fermions bounded in a small nuclear
volume, which obey Pauli exclusion principle. The nuclear force exists between these
fermions is complicated and not an easy task to understand like other fundamental forces.
Study of nucleus presents many aspects which challenge our understanding to many-body
manifestations of nuclear properties and their relationship to the underlying fundamental
interactions. Study of properties of nuclear excited states gives information about the
forces between these nucleons and the behavior of this many particle system. The nucleus
can be excited to higher rotational states by pumping energy to the nucleons and excite
them to higher orbits. These excitations are due to collective motion of nucleons or single
particle excitations. To understand these nuclear excitations and associated aspects of
nuclear structural effects many nuclear models were developed over the past few decades
but theoretical models of the nucleus encounter two principal problems:
(1) There is no exact mathematical expression that accounts for the nuclear force,
unlike the atomic case, for which the electromagnetic force is well-defined by coulombs
law, and
(2) There is no mathematical solution to the many-body problem, a limitation shared
by both nuclear and atomic systems

19
Fortunately, the power of computer permits calculations that use sophisticated ap-
proximations to minimize these problems and provide increasingly accurate models for
describing nuclear and atomic properties. At the macroscopic extreme is the Liquid Drop
Model, which examines global properties of nuclei, such as energetics, binding energies,
sizes, shapes and nucleon distributions.
The Liquid drop model is one of the primary model of nucleus, formulated by Bohr
and Wheeler, in which nucleus is considered as a liquid drop. Protons and neutrons in a
nucleus are analogous to molecules in a liquid drop and are held together by short-range
interactions and surface-tension effects. This model describes many nuclear properties like
nuclear masses and binding energies in terms of volume energy, surface energy, compress-
ibility, etc. These parameters are usually associated with liquid. This model successfully
describes how a nucleus can deform and undergo fission and large class of nuclear reac-
tions. The most basic property of a nucleus is its mass or total energy in its ground state.
The binding energy of the nucleus is defined as

B(A, Z) = (ZMp + N Mn ) − M (A, Z) (2.1)

where A, is mass number of the nucleus, Z is number of protons and N = A - Z is


number of neutrons. Mp , Mn and M(A, Z) are the proton mass, neutron mass and
mass of nucleus respectively. Experimentally the observed values of Binding energies per
nucleon increase sharply as A increases, peaking at iron (Fe) and then decreasing slowly
for the more massive nuclei with an average value in between 7.5 to 8.8 MeV except for
few lighter and for heavy elements. Figure 2.1 is the representative plot showing the
variation of experimental binding energy per nucleon with respective to mass number of
nuclei. Several attempts were made to reproduce the B/A curve, but the liquid drop
model proposed by Bethe and Weizsacker is one of the successful model in reproducing
this result. This model provides the mathematical equation which attempts to predict
the binding energy of nucleus in terms of the number of protons and neutrons it contains.
This equation have five terms corresponding to volume energy, surface energy, coulomb
energy, asymmetry energy and pairing energy.
The complete formula of Wiezsacker-Bethe, known as semi-empirical mass formula,

20
Figure 2.1: Variation of BE/A with respect to the mass number of nuclei is shown.

which describes the binding energy of a nucleus of mass number A with Z protons and N
neutron as:
Z2 (Z − N )2 ((−1)Z + (−1)N ) aP
B(A, Z) = aV A − aS A2/3 − aC − a A + (2.2)
A1/3 A 2 A1/2
From fitting to measured nuclear binding energies, the values of the parameters are
aV = 15.56 MeV, aS = 17.23 MeV, aC = 0.697 MeV, aA = 23.285 MeV and aP = 12.0
MeV.
Although this model was quite successful in calculating the nuclear masses and
binding energies, it failed in certain regions. The binding energies predicted by liquid
drop model underestimate the actual binding energies of magic nuclei for which either
the number of neutrons N = (A - Z) or the number of protons, Z is equal to one of the
following magic numbers 2, 8, 20, 28, 50, 82, 126. This is particularly the case for doubly
magic nuclei in which both number of neutrons and number of protons are equal to magic
numbers. Since it was based on classical approach and therefore lacked the important
microscopic quantal effects, so it had limited success and was able to explain only bulk
properties of the nuclei.

21
15
Volume Energy

Binding Energy per nucleon (MeV)


10

Total Energy
5

0
Surface Energy

Coulomb Energy
−5

− 10
50 100 150 200 250
Mass Number (A)

Figure 2.2: Showing the variation of different terms in BE equation of liquid drop model
with the plot between BE/A and mass number(A).

2.2 Nuclear shell model


The first most successful nuclear model, which included the microscopic quantal effects,
was nuclear shell model proposed by Mayer and Jensen [1, 2, 3] in 1949. This model
describes nucleus as a system of A-independent fermions, interacting with each other
through many-body nucleon-nucleon forces. The basic assumption of nuclear shell model
is that to a first approximation each nucleon moves independently in the average potential
V(r). This independent motion of nucleons can be understood qualitatively from a combi-
nation of weakness of the nuclear long - range attraction and the Pauli exclusion principle.
As such, any chosen mean-field should possess rotational invariance associated with the
conservation of total angular momentum (i.e. [Htot ,Jtot ] = 0) and should allow occupation
of proton and neutron single-particle levels separately. The shell model employs as its
Hamiltonian H,

X 1X
H= [Ti + U (i)] + v(ij) (2.3)
i=1
2 i6=j

22
and is comprised of the unperturbed Hamiltonian,

X X
H0 = [Ti + U (i)] = h(i) (2.4)
i=1

as well as the two-body residual interaction, v(ij), which acts on H0 as a first order
perturbation to break symmetries embodied within. In many cases, the study of such
configuration-mixing interactions yields valuable information on the underlying nuclear
structure. However, extraction of details on the residual interaction from total Hamilto-
nian requires a model-dependent form of U(i), which is often taken as the phenomenolog-
ical Woods-Saxon potential, Vws

−V◦
Vws = (2.5)
(1 + exp(r − R◦ /a))

where V◦ ≈ 50 MeV, R◦ ≈ 1.2A1/3 fm and a ≈ 0.6 fm. It was realized, however, that
use of such a phenomenological potential alone cannot even reproduce the empirical shell
closures without the inclusion of a spin-orbit coupling term li .si , as shown by Mayer et al
[4]. Application of Schrodinger equation to zeroth-order (i.e. no v(ij)) Hamiltonian plus
the coupling term,

−~2 2
( ∇ + U (i) + C(li .si )) = E (2.6)
2M

then yields independent single-particle levels and the observed magic numbers (at Z
= N = 2, 8, 20, 50, 82, and 126 as shown in figure 2.3). The degeneracy of oscillator level
was partially removed with the introduction of this li .si term and the levels split up as j
1
= l ± , where j is the total angular momentum. The levels are now characterized by
2
|n, l, ji. An attractive spin-orbit potential assures the experimentally observed fact that
1 1
the l + levels are energetically always below the l - . The shell model was successful
2 2
in reproducing the spherical magic numbers and other properties like prediction of the
spins and parities of the ground states of almost all spherical nuclei with inclusion of spin-
orbit term to the Woods-Saxon potential. But this model has limited success in case of
deformed nuclei. Another success of the nuclear shell model was the prediction of islands
of nuclear isomerism in regions around the closed shells. The shell model also correctly

23
1i 13/2 14 {126}
3p 1/2 2
p (l = 1) 3p 3/2 4
N=5 2f 5/2 6
f (l = 3) 2f 7/2 8
1h 9/2 10

h (l = 5)

1h 11/2 12 {82}
s (l = 0) 3s 1/2 2
N=4 2d 3/2 4
d (l = 2) 2d 5/2 6
1g 7/2 8

g (l = 4)

1g 9/2 10 {50}

2p 1/2 2
p (l = 1) 6
1f 5/2
N=3
2p 3/2 4
f (l = 3)

1f 7/2 8 {28}

1d 3/2 4 {20}

2s 1/2 2
N=2 s (l = 0)

d (l = 2) 1d 5/2 6

1p1/2 2 {8}
N=1
p (l = 1) 1p3/2 4

N=0
s (l = 0) 1s1/2 2 {2}
Major Oscillator l = N, N−2, ...... Spin−Orbit No. of Nucleons Magic
Shell Coupling(l.s) in shell Numbers

Figure 2.3: Shell model orbitals with respective to different potential assumed in nuclear
shell model.

24
predicts the sign and roughly the magnitude of ground state magnetic moments of nuclei.
The shell model predictions of the quadrupole moments are generally in agreement with
experiments for nuclei near the magic numbers. As number of protons and neutrons
departs from the magic numbers it becomes indispensable to include in some way the
residual two-body interaction, to break the degeneracies inherent in the filling of orbits
with two or more nucleons. At this point difficulties accumulate, these difficulties were
explained and overcome by Nilsson with Deformed shell model.

2.3 Nuclear shape parametrization


For classically rotating systems, a fundamental physical quantity is the moment of in-
ertia. As nuclear matter is essentially incompressible, this is directly related to shape
and deformation of the nucleus. This shape is determined by the characteristics of the
orbitals occupied. The contribution from macroscopic binding energy and electrostatic
repulsion of the constituent protons dictates nuclear shape. For example, macroscopic
binding energy favours spherical shapes since this reduces surface energy whereas elec-
trostatic repulsion of protons favours elongated shapes. For deformed nuclei, the nuclear
shapes are often parametrized in terms of a spherical harmonic multipole expansion and
are related to observed nuclear moments or transition rate in a model dependent way.
The length of a radius vector R(θ, φ) pointing from origin to nuclear surface is given by
[5]
" λX λ
#
max X
R(θ, φ) = R◦ 1 + αλ,ν Y (θ, φ) (2.7)
λ=1 ν−λ

where R◦ represents the radius at spherical equilibrium with the same volume and αλ,ν
are the shape parameters. Where λ represents different modes of deformations. The
λ = 0 term corresponds to breathing mode of the nucleus. The λ = 1 term does not
really correspond to a deformation of the nucleus but rather to a shift of center of mass
and it corresponds only to a translation of nucleus, and hence disregarded for nuclear
excitations. In many cases, nuclear shapes are assumed to be invariant with respect to
the three symmetry planes, so that

25
Figure 2.4: Schematic representation of kind of nuclear shapes for different values of
deformation parameters like spherical, prolate and oblate shown in order.

αλ,ν = αλ,−ν αλ,ν = 0 (λ=odd and/or ν=odd )

For axially symmetric shapes with respect to the major axis all deformation parame-
ters except ν = 0 disappear and the remaining deformation parameters αλ,0 are described
as βλ , i.e. βλ = αλ,0
If one fixes the center of mass, the coefficients α1,µ = 0 as these terms are proportional
to displacements. Therefore considering only the lowest order terms, the nuclear surface
can be parametrized as
" 2
#
X
R(θ, φ) = R◦ α2,µ Y2,µ (θ, φ) (2.8)
µ=−2

where R◦ is mean radius. The five parameters α2,µ are not all independent as three
parameters θk are required to specify the orientation of the nucleus. If one fixes the
coordinate system to lie along the principle axis of the nucleus. Then α2,1 = α2,−1 = 0
and α2,2 = α2,−2 .
The two remaining parameters (α2,0 , α2,2 ) are commonly expressed in Hill-Wheeler
coordinates β2 and γ

α2,0 = β2 cos(γ) (2.9)

1
α2,2 = √ β2 sin(γ) (2.10)
2

26
where the deformation is parametrized by β2 > 0 and the triaxiality by the angle γ. With
these definitions we can write R(θ, φ) as

R(θ, φ) = R◦ [1 + a20 Y20 (θ, φ) + a22 Y22 (θ, φ) + a2−2 Y2−2 (θ, φ)] (2.11)

simplifying above equation length of three semi-axis of ellipsoid are then given by
" r #
5
Rk = R◦ 1 + β2 cos(γ + 2πk/3) (2.12)

where k = 1, 2, and 3. A pure prolate shape corresponds to γ = 0 while pure oblate


shapes occur when γ = π/3. Intermediate values of correspond to triaxial shapes. As can
be seen from the above each value does not lead to a unique shape and incrementing γ by
an integral multiple of 2π/3 results in a simple permutation of the axis labels. Thus, for a
given direction of rotation. this permutation of axis labels differentiates between collective
and non-collective rotation. Advantage is taken of this fact in the Lund parametrization
of nuclear deformation, where γ has a range of -2π/3 to π/3 allowing one to present both
collective and non-collective rotation.

2.4 Deformed shell model


The deformed shell model was described by a phenomenological model potential, first
proposed by Nilsson [6, 7], called modified harmonic potential, to investigate the effect
of deformation on the single particle orbits. Limited only to the harmonic-oscillator
potential, the zeroth order Hamiltonian, Hintr , can be expressed as

X ~2 1 X 2 2
Hintr = − ∇2i + M [ωx (x1 + yi2 ) + ωz2 zi2 ] (2.13)
i
2M 2 i

where wx,y,z are the oscillator frequencies in x-, y-, z-direction. The energy corresponding
to the above Hamiltonian is

1
E(N, nz , δ) = ~ωx (N − nz + 1) + ~ωz (nz + ) (2.14)
2

where N = nx +ny +nz and is the total number of harmonic quanta in x-, y-, z- direction.
The number of oscillator quanta can be set to be a function of deformation δ, so that
27
Non−Collective Oblate
o
60

γ
o
0

Collective Prolate

o o
−120 −60

Non−Collective Prolate Collective Oblate

Figure 2.5: The lund conventions: Schematic of nuclear shapes with respect to the defor-
mation parameters (β, γ).

28
Figure 2.6: Schematic of single-particle coupling to core and the Nilsson quantum num-
bers.

1
ωx + ωy = ωδ ∼ ω0 [1 + δ] (2.15)
3

2
ωz ∼ ω0 [1 − δ] (2.16)
3

∆R
δ= (2.17)
R0

The above relations illustrate that for a spherical single-particle configuration, i.e.
δ, N = nx +ny +nz is degenerate and the shell gaps appear in the single particle spectrum
at usual magic numbers. As the potential becomes deformed, i.e. δ > 0, this degener-
acy is lost, but a high degree of degeneracy reappears whenever the ratio of rotational
frequencies, ωx,y : ωz , are simple integers. This is a general feature of the anisotropic
harmonic oscillator potential Hintr , which leads to the deformed magic numbers
Assumptions of rotational and reflection invariance of Hintr allow the eigen-states of
Hintr to be characterized by the Nilsson quantum numbers (shown in figure 2.6) as

Ωπ [N, nz , Λ] (2.18)

29
Figure 2.7: The partial schematic of Nilsson single particle level diagram is shown.

30
where N is the principal quantum number, Ω is projection of single-particle angular mo-
mentum on the symmetry axis (z), Λ is projection of the orbital angular momentum on
the symmetry axis and nz is number of oscillator quanta along the symmetry axis where
z-projection of total angular momentum of the particle satisfies

1
Ω=Λ± (2.19)
2

The parity satisfies

π = (−1)N (2.20)

and the orbital angular momentum projection quantum number, Λ satisfies,

|Λ| = N − nz , N − nz − 2, N − nz − 4, .......... ± 1 or 0 (2.21)

These quantum numbers are known as asymptotic, because these are approximately good
quantum numbers at low deformation. However, at very large deformation, the defor-
mation term is much larger compared to spin-orbit term. One of the features in Nilsson
diagrams is that by breaking spheroid symmetry, the 2j+1 degenerate shell model state
(2j + 1)
(ε2 =0) is split into levels for ε2 6= 0. Each of these is two-fold degenerate with
2
eigenvalues ±Ω. The levels with lower values are shifted downwards for positive defor-
mations (prolate shapes), and upwards for negative deformations (oblate shapes). This
effect is enhanced for high-j, low-Ω orbits and plays a significant role in modifying the
shapes of nuclei. The plot of energy of different levels as a function of the deformation
parameter ε2 is known as the Nilsson diagram. In the Nilsson diagrams, the spherical
magic numbers appear at 2, 8, 20, 28, 50, 82, 126 appear at ε2 =0. In addition, there are
other magic numbers which appear at different deformations. For example, in the A ≈
70-80 region, shell gaps occur for particle numbers 34, 38 for prolate deformation and at
36 for oblate deformation. These numbers are called the deformed magic numbers. These
play an important role in stabilizing the nuclear shape.
Just as in spherical shell model where the zeroth-order single-particle Hamilton,
H0 , is required to have a spin-orbit term Cli .si and Dl2i , to reproduce splitting between
levels with different values of l. Well deformed nuclei may be expressed in the Nilsson
Hamiltonian of the form, 31
X ~2 1 X 2 2 X X
Hintr = − ∇2i + M [ωx (x1 + yi2 ) + ωz2 zi2 ] + C li .si + D li2 (2.22)
i
2M 2 i i i

The constants C and D are empirical values, and are different for each major shell.
The fact that the Nilsson Model provides two or three alternatives for the lowest state
of a given nucleus indicates the limit of validity. Constants, such as ω0 , C, and D,
represent only the average properties of the underlying real nucleon-nucleon interactions,
so that more detailed examination of experimental data requires the inclusion of residual
interactions, such as pairing.

2.5 Rotation of nucleus


Breaking of spherical symmetry of a quantal system leads to an anisotropy that allows for
the specification of an orientation. An attribute of existence of such a deformation is the
nuclear rotation. The spectra of deformed nuclei manifest rotational bands characteristic
of quantum mechanical rotors. The total angular momentum of a rapidly rotating nuclei
is given by,

I¯ = R̄ + J¯ (2.23)

where R̄ and J¯ are collective and intrinsic rotational angular momentum, respectively. It is
related to symmetry axis by its projection k̄ . Since no collective rotation is possible about
the symmetry axis, R̄ must always be perpendicular to the symmetry axis. This is because
it is not possible to quantum mechanically distinguish the different orientations when the
nucleus rotates about symmetry axis. The system is invariant under rotations about the
symmetry axis and the component K of I is a constant of motion. The Hamiltonian, H for
such a system can be written in terms of hamiltonians describing its collective rotation
and intrinsic motions,

H = Hcol + Hintr (2.24)

where
~2 ¯ ¯ 2
Hcol = (I − J) (2.25)
2ℑ

32
¯ with in the body-fixed system of
and Hintr describes the nucleonic motion (with spin J)
reference, which is valid only for adiabatic case, i.e., when the rotation is slow compared
to single-particle motion. The state of motion is completely specified by I2 , M and K
with,

p
R=~ I(I + 1) − K 2 (2.26)

1
This expression is valid for j = K and for K6= . In case of no intrinsic excitation (K =
2
0), the energies of a collectively rotating rigid nucleus are given by,

~2
Erot = I(I + 1) (2.27)
2ℑ0

where ℑ0 is the moment of inertia, which is constant for an ideal rigid body rotor, but
changes slowly within a rotational sequence. A rotational sequence may be built upon
ground state or an excited single particle state of a nucleus. For an even-even nucleus, the
nucleons fill up each of the orbitals having single particle angular momenta ±ωi in pairs.
As a result, the ground state has the parity π = +. In this case the intrinsic angular
momentum K = 0 and according to equation Erot , the nuclear rotation is manifested as a
series of levels with excitation energies proportional to I(I + 1) where I = 0, 2, 4, . . . . .
For an odd-A nuclei, the intrinsic spin j is not equal to zero, the rotational Hamil-
tonian can be written as,

~2 ~2 ¯ ¯ 2
Hrot = ( )R̄2 = ( )[I − J] (2.28)
2ℑ0 2ℑ0

Therefore, the coupling of intrinsic and rotational angular momentum should be


conserved. Depending upon relative effects of Coriolis effect and core deformation, two
types of particle-core couplings can be defined. In the deformation aligned scheme (strong
coupling) the odd particle adiabatically follows the rotations of the core while in the
rotation aligned scheme (decoupled), the odd particle essentially moves on spherical shell
model levels without facing any significant disturbance from the quadrupole vibrations
of the core. The validity of these schemes depend upon magnitude and direction of J¯
and magnitude of angular momentum of the system. For states with large Ω and low
value of total angular momentum I, the strong coupling scheme can be applied and the
Hamiltonian is given by, 33
~2
H = Hsp + Hrot = Hsp ( )[(I 2 − Ix2 ) − δk,1/2 (I+ J− + I− J+ )] (2.29)
2ℑ0

Here the adiabatic approximation has been applied by assuming that the effect of
rotational motion is small. The term I+ J− + I− J+ is the Coriolis term and gives the
coupling of the single particle and collective motion. In the situation, when the total
angular momentum I and the intrinsic spin j are large, the effect of rotational motion on
the single particle motion is significant and the total energy is minimized when the intrinsic
spin, j is aligned along the direction of the total angular momentum (I) and momentum,
I are parallel to the rotation axis. This alignment of particle angular momenta along the
rotation axis leads to several interesting phenomena at high spins.

2.5.1 Cranking model

The alignment of the nucleons and the associated nuclear shape effects with nuclear
rotation can be understood more clearly in the light of cranking model calculations. The
first mathematical formulation of Cranking Model was made by Inglis [8, 9] and has been
further developed by Bengtsson and Frauendorf. The effects of rotation on the single-
particle states are described within the framework of the cranking model. The behavior
of nucleons is studied by rotating the nucleus with rotational frequency, ω i.e. to crank the
nucleus, hence the name cranking model. The model provides a microscopic description
of the influence of rotation on single-particle motion. The rotation is treated classically
and the nucleons are considered as independent particles moving in an average rotating
potential (Woods-Saxon and modified oscillator). To simplify the theoretical treatment,
the intrinsic (body-fixed) coordinate system (x1 x2 x3 ) is used which has a fixed orientation
at all times with respect to the nuclear potential. It rotates with rotational frequency, ω,
relative to the laboratory coordinate system (xyz), shown in figure 2.8
If we assume that the intrinsic coordinates rotate around an axis perpendicular to
the symmetry axis and the symmetry axis coincides with the intrinsic axis x3 and the
intrinsic axis x1 coincides with the laboratory axis x, the total cranking Hamiltonian is
given by

34
x

x1

x3

wt
z

y x2

Figure 2.8: Schematic of the body-fixed (intrinsic) coordinates (x1 x2 x3 ) and the laboratory
coordinates (xyz) are shown.

H ω = H 0 − ω.Jx = Σµ [h0µ − ωJxµ ] (2.30)

where Hω and hωµ =h0µ -ωJxµ are the total and single particle Hamiltonians in the rotating
frame. H0 and h0µ are the total and single-particle Hamiltonians in the laboratory system,
and Jx and jµ are the total and single particle angular momentum projections onto the
rotation axis. The -ωJxµ term contains the Coriolis and centrifugal forces [10, 11, 12].
The Coriolis force acts to align angular momentum of nucleons with the rotation axis [13].
The eigenvalues of the Hamiltonian, Hω , are known as Routhians. The total energy in
the laboratory system is obtained as
X X
ε= eωµ + ω hµω |Jx |µω i (2.31)
µ µ

where eωµ denote eigenvalues in the rotating frame. Projection of total angular momentum
onto the rotation axis can be determined by
X
Ix = hµω |Jx |µω i (2.32)
µ

35
The slope of routhian is related to the alignment ix , that is the angular momentum gained
by aligning a nucleon along the rotation axis. It is given by

deωµ
ix = − (2.33)

This relation makes it possible to compare the experimental aligned angular mo-
menta with theoretical routhians. Inclusion of -ωJx term, breaks time-reversal symmetry.
Hamiltonian is only invariant under space reflection, parity π and under a rotation of 180◦
about x axis. The rotation operator is given by

ℜx = e−iπJx (2.34)

with the eigenvalues [14, 15] r=e−iπα where α is an additive quantum number known as
signature. It is related to total angular momentum by I = α mod 2. For a system with
an even nucleon number

½
0, 2, 4, 6, · · · if r = +1 (α = 0)
I=
1, 3, 5, 7, · · · if r = −1 (α = 1)

while for a system with odd nucleon number

½
1/2, 5/2, 9/2, · · · if r = −i (α = +1/2)
I=
3/2, 7/2, 11/2, · · · if r = +i (α = −1/2)

The rotational states are characterized by the parity π and the signature α

2.5.2 Pairing interaction

The paring interaction is short-range interaction, described by a δ function. The starting


point is to assume that the paring force acts between nucleons with same quantum num-
bers, but nucleons move in time reversed orbits. If these nucleon pairs are close enough to
Fermi level, they may scatter from one pair of time reversed orbits to another unoccupied
level as a pair. In a classical approach, one could imagine that nucleon pairs are orbiting
around nucleus in opposite directions and each time they collide, they scatter into another
pair of orbits. Orbits close to Fermi surface can be partially occupied which smoothens

36
Figure 2.9: A representative energy spectra of cranking Hamiltonian with Nilsson poten-
tial (left), with pairing interactions (middle) and with rotation (right) as a function of
rotational frequency (~ω). Spectra was taken from Ref.[16].

out the Fermi surface. Including pairing interaction leads to the concept of quasi-particles
where particles and holes are substituted by quasi-particles and particle-hole excitation
by the creation of one quasi-particle and destruction of another quasi-particle.
The total wave function from BCS formalism takes the form [17, 18]

ψ0BCS = Πν Uν + Vν a+ +
jm aj m̄ aj m̄′ ajm′ |0 > (2.35)

where pairing hamiltonian employed is


X
Hpair = a+ +
jm aj m̄ aj m̄′ ajm′ (2.36)
ν

where m refers to the time-reversed orbits of m, and a+ and a refer to creation and
destruction operators for single particles. The resulting two-particle wave function in
equation 2.35 is a product over all states ν, and Uν and Vν are subject to the normalization
conditions

Uν2 + Vν2 = 1 (2.37)

X
2 Vν2 = N (2.38)
ν

37
where U2ν refers to probability of an orbital ν not being occupied, V2ν refers to the proba-
bility that it is occupied, and N is total number of nucleons. The required wave-function
of equation 2.35 is the one which minimizes the total energy under conditions of equation
2.37 and 2.38. The solution of Hamiltonian in equation 2.36 then yields probability of an
orbit being occupied

1 (ǫν − λ)
Vν2 = [1 − ] (2.39)
2 Eν

where Eν is the quasi-particle energy, and is

p
Eν = (ǫν − λ)2 + ∆2 (where ∆ = GΣν Uν Vν ) (2.40)

Here, ǫν represents single-particle energy, G-pairing strength, λ-chemical potential, and


∆-pairing gap, which is often taken as odd even mass difference. According to the equation
2.39, each quasi-particle orbital represents a mixture of Nilsson single-particle orbitals,
and the quasi-particle energies Eν , are dependent on both ǫν and λ. Figure 2.9 demon-
strates effects of paring interaction and rotation on independent single-particle motion
in a deformed nuclear potential. However, significant changes to the gap equation (2.40)
have to be made in moderately deformed odd-even nuclei (40 < N ≈ Z < 150) due in part
to their low level densities. Such nuclei have only a small number of summations over
states ν in equations 2.37 and 2.38, so that Pauli principle, which forbids the occupancy
of a quasi-particle level already occupied by an odd particle (by a particle pair), and the
replacement of the number N in equation 2.38 with N-1 modify the equations 2.38 and ∆
as

X
2 Vν2 = N − 1 (2.41)
ν

∆ = GΣν6=ν1 Uν Vν (2.42)

where the ν1 represents quasi-particle level already occupied by the odd particle in the
ground state wave function ψBCS . These changes in ∆ and in the Uν /Vν are called the
blocking effect.

38
2.5.3 Total routhian surface calculations

Total Routhian Surface calculations are the shape calculations which are used to describe
the nuclear shape changes with respective to rotational frequency. The total nuclear
energy is calculated in the rotating frame and is minimized with respective to deformation.
The total routhian of a nucleus (Z, N), rotating with angular frequency ω, can be obtained
as a function of deformation (β) following the Strutinsky shell correction method as
explained below.

2.5.4 Nilsson-Strutinsky method

The Nilsson-Strutinsky method considers the nuclear binding energy of a nucleus to have
a shell structural contribution Esh , superimposed on to a smoothly varying liquid drop
part ELD , both with respect to nuclear mass;

E = △Esh + ELD (2.43)

The total shell model energy Esh , is further decomposed into a smoothly varying Êsh and
a fluctuating Esh part,

X
Esh = εvi = Êsh + △Esh (2.44)
i

so that total energy of the nucleus can be expressed as where ELD is the macroscopic
contribution describing the bulk properties, and

E = ELD + △Esh = ELD + [Esh − Êsh ] (2.45)

is the microscopic quantal shell correction. With inclusion of term △Esh , ground state
nuclear energies can be determined as a function of deformation parameters and nuclear
rotation. Usually, a pairing correction is included in the total routhian, so that routhian
may be constructed from an expression

E ω (Z.N.β̂) = Emacro
ω ω
(Z.N.β̂) + △Eshell ω
(Z.N.β̂) + △Epair (Z.N.β̂) (2.46)

39
where β̂ represents all deformation parameters. The shell correction term is calculated
using the deformed Nilsson potential at zero rotational frequency (ω = 0), and pairing
term is determined using self-consistent solution of BCS equations. The microscopic-
macroscopic form of total routhians can therefore be constructed as

 ω=0
 E (Z.N.β̂)
ω
E (Z.N.β̂) = +[hν ω |H ω (Z.N.β̂)|ν ω i − hν ω |H ω=0 (Z.N.β̂)|ν ω i]
 1 2
− 2 ω [ℑmacro (A, β̂) − ℑstruct (Z.N.β̂)]
where Eω=0 corresponds to sum of liquid drop energy, single-particle shell correction en-
ergy, and pairing energy at zero frequency. The total routhian is calculated in a lattice of
deformation space (β̂ = β2 , β4 , and γ) at a fixed rotational frequency, and then minimized
with respect to shape parameters to obtain the equilibrium deformation. In relation to
deformed harmonic oscillator potential, the shell structure is not unique to spherical nu-
clei, but also for certain finite deformations of nuclear potential. Although stable liquid
drops are always spherical, Strutinsky averaged energy of above equation could therefore
possess a minimum at finite values of deformation. The results are usually displayed as
Total Routhian Surface (TRS) plots, which are a contour map of energy in β2 , γ plane.
The total potential energy is minimized with respect to higher order shape parameters
like β4 . To make energy contour plot, the following transformation has been made,

X = β2 cos(γ + 30) (2.47)

Y = β2 sin(γ + 30) (2.48)

A typical example Total Routhian Surface (TRS) map of 70 Ge calculated for positive par-
ity sequence at rotational frequency ~ω = 0.5 MeV and 0.7 MeV is shown in figure 2.10.
The triaxiality parameter γ describes deviation of a nucleus shape from the axial sym-
metry. This parameter is defined such that γ = 0◦ and γ= 60◦ correspond to axially
symmetric prolate and oblate deformations, respectively. The quadrupole deformation
defines the parameter β2 and at each point in the β2 -γ plane. This model was quite
successful in describing nuclear alignments.

40
Figure 2.10: Total Routhian Surface calculations for the positive parity sequence of 70 Ge at
rotational frequencies of 0.5 MeV (left) and 0.7 MeV (right). The contours of equipotential
surfaces are separated by 0.20 MeV.

2.6 Comparison of experimental data with theoreti-


cal calculations
It is useful to compare experimental results with the theoretical calculations. But it
is not possible to compare experimental observables like excitation energies, spins and
transition probabilities with theoretical calculation in rotating frame. For this purpose one
need transformation of experimental excitation energies, Eex , and experimental angular
momentum I, to intrinsic rotating frame of the nucleus.
The experimental energies in the form experimental routhians defined by

E ′ (ω) = E(ω) − ωI (2.49)

where the rotational frequency ω, for a given state with angular momentum I, defined as
dE(I)
ω(I) = ~−1 (2.50)
dIx (I)
and Ix (the component of angular momentum along the cranking axis) will relate to
measured angular momentum (I), which is the average value of the angular momenta of
the two values I ± 1 relevant to experimental energy levels E (I ± 1) and the angular
momentum perpendicular to the cranking axis (K) is given by
r
p 1
Ix (I) = I(I + 1) − K 2 ≈ (I + )2 − K 2 (2.51)
2

41
The derivative in equation ω(I) is approximated by a quotient of finite differences,
therefore it can be expressed as

∆E(I) E(I + 1) − E(I − 1) −1


ω(I) ≈ = ~ (2.52)
∆Ix (I) Ix (I + 1) − Ix (I − 1)

The total experimental Routhian, E′ , as a function of measured angular momentum


is defined by

E ′ (I) = E(I) − ω(I)Ix (I) (2.53)

where E(I) is the average value of E(I + 1) and E(I - 1). The total routhian, E′ , includes
the energy associated both collective and quasiparticle excitations. Also projection of
angular momentum along the cranking axis contains with rotational and quasiparticle
components. When investigating small effects like the effect of single particle excitations
in energy or spin, it is convenient to subtract out collective part (larger effect) of the
rotor from the total routhian and total angular momentum. One way to subtract the
rotor effect is by using the Harris formula ℑref = ℑ0 + ℑ1 ω 2 with parameters ℑ0 and
ℑ1 are called Harris parameters [19]. Values for ℑ0 and ℑ1 can be chosen based on the
ground state band of the nucleus under study, or can be adopted from another nucleus
for comparison. A good reference nucleus would be the nearest even-even nucleus with
fewer or equal N and Z to the nucleus under study because it is less likely to show strong
single-particle effects.
The reference energy and reference angular momentum are defined by

1 1 1
Z
Eref (ω) = − Ixref (ω)dω = − ω 2 ℑ0 − ω 4 ℑ1 + (2.54)
2 4 8ℑ0

Ixref = ω(ℑ0 + ℑ1 ω 2 ) (2.55)

Therefore the experimental routhian e′ , and the aligned angular momentum ix , of


the excited single particles are obtained by subtracting the above reference values, given
by

e′ (ω) = E ′ (ω) − E ref (ω) (2.56)


42
ix (ω) = Ix (ω) − Ixref (ω) (2.57)

These transformations provides a way for finding e′ , Ix , ω, ℑ1 , ℑ2 from experiment


and comparing to quantities in the cranking model. From these results, conclusions can be
drawn about the configuration and parameters of the nucleus. In particular for rotational
cascades connected with γ-ray transitions of energy Eγ of quadrupole nature (∆J = 2) the
rotational frequency ω, the kinematic moment of inertia ℑ1 , and the dynamical moment
of inertia ℑ2 , can be calculated from experimental data using the following expressions
[20]


ω≈ [J + 2 → J](M eV /~) (2.58)
2

2
ℑ1 ≈ J[~/M eV ] (2.59)
Eγ [J → J − 2]

4
ℑ2 ≈ [~2 /M eV ] (2.60)
∆Eγ

where ∆Eγ = Eγ [J+2 → J] - Eγ [J → J-2]. In above equation it is clear that dynamical


moment of inertia ℑ2 is independent of spin(J). This parameter will be useful in getting
information about bands for which the spin and parity are not known. As an illustrative
example, the plots for the experimental routhians e′ and aligned angular momentum Ix as
a function of rotational frequency (ω) for positive parity band of 73 As are shown in figure
2.11 and figure 2.12. The plots represents that at particular rotational frequency known
as crossing frequency (ωc ), the rotational frequency decreases with increasing angular
momentum, the phenomenon known as back-bending. At this crossing frequency the
ground state band crossed by a new configuration which means the nucleus changes its
intrinsic structure. In the case of positive parity band of 73 As, the crossing frequency (ωc )
occurs around 0.5 MeV and the corresponding alignment gain (∆ix ) is shown in figure
2.12.

43
0

73As−postive parity band

Experimental routhian (e’)


−1
ωc

−2

−3

−4

−5
0.2 0.3 0.4 0.5 0.6 0.7 0.8
Rotational frequency (MeV)

Figure 2.11: Experimental routhians (e′ ) of positive parity ground state band in 73 As as
a function of rotational frequency (ω). This band undergoes a crossing around ωc ≈ 0.5
MeV.

15

73As−postive parity band

10
Alignment( ix )

∆ix

0
0.2 0.3 0.4 0.5 0.6 0.7 0.8
Rotational frequency (MeV)

Figure 2.12: Experimental aligned angular momentum (ix ) of positive parity ground state
band in 73 As as a function of rotational frequency (ω). The alignment gain is represented
as ∆ix .

44
2.7 Summary
In summary, the features of various nuclear models are discussed in this chapters. The
properties of Liquid drop model and shell model are outlined. Basic postulates of deformed
shell model and cranking model are described, prediction of which have been used for
direct or indirect comparison of observed data. The Total Routhian Surface calculations
are discussed at length, which are used to describe the shape evolution as a function
of rotational frequency in the present work. Finally various methods in comparing the
experimental data with theoretical calculations are discussed.

45
Bibliography

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

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

[3] M. G. Mayer and J. H. D. Jensen, Elementary Theory of Nuclear Shell Structure,


Wiely, New York (1955).

[4] M. G. Mayer, Phys. Rev. 78 16 (1950).

[5] Ring and P. Schunk: The nuclear many-body problem (Springer- Verlag, New York,
Inc.US, 1980).

[6] S. G. Nilsson, Mat. Fys. Medd. Dan. Vid. Selsk. 29 (1955).

[7] S. G. Nilsson, C. F. Tsang, A. Sobiczewski, Z. Szyma′nski, S. Wycech, C. Gustafson,


I. Lamm, P. Möler and H. Nilsson, Nucl. Phys. A131, 1 (1969).

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

[9] D. R. Inglis, Phys. Rev. 97, 701 (1955).

[10] M. J. A. de Voigt, J. Dudek and Z. Szyma′nsky, Clarendon Press. Oxford, 1983.

[11] M. J. A. de Voigt et al., Rev. Mod. Phys. 55, 949 (1983).

[12] D. Garrett: Collective phenomena in Atomic Nuclei, page 193, World Scientific,
Singapore, 1984.

[13] F. S. Stephens et al., Nucl. Phys. A183, 257 (1972).

[14] A. Bohr and B. R. Mottelson, Nuclear Structure, vol. II.

46
[15] W. A. Benjamin, Inc., New York, 1975.

[16] D. D. Garrett Nucl. Phys. A 409, 259c (1983).

[17] A. Bohr, B. R. Mottelson and D. Pines, Phys. Rev. 110, 936 (1958).

[18] J. Bardeen, L. N. Cooper and J. R. Schrieffer, Phys. Rev. 108, 1175 (1957).

[19] S. M. Harris et al., Phys. Rev. B 138, 509 (1965).

[20] R. Bengtsson, S. Frauendorf and R. R. May, At. Data Nucl. Data Tables 35, 15
(1986).

47

Вам также может понравиться