You are on page 1of 44

Chapter 2

Theoretical aspects

The hadrons inside an atomic nucleus interact through the fundamental interac­
tions, viz., the strong, the electromagnetic and the weak interactions. The nuclear
structure can be studied using these interactions as probes and the nuclear phenom­
ena are essentially the observable response to the probes that act on the nucleus
through these interactions [1]. In gamma-ray spectroscopy, the probes are gen­
erally hadrons that interact with the nucleus through strong interaction and the
subsequent gamma decay is a result of electromagnetic interaction that responds
to the electric charges of the hadrons and their electric currents. A large part of
the knowledge of nuclei comes from the study of electromagnetic transitions since
the electromagnetic interaction is well understood compared to the nuclear forces.
Hence using the features of electromagnetic interaction, gamma-ray spectroscopy
has been extensively applied to probe various aspects of nuclear structure, such
as, energy (E), angular momentum (J), parity (zr), lifetime of the nuclear states;
branching ratios and mixing ratios of the transitions; transition probabilities be­
tween two nuclear states etc.
The first section of the present chapter deals with the theory of gamma decay.
The second section is a discussion on the Shell Model (one of the successful models
of nuclear structure) that was used to interpret the experimental results obtained
through gamma-ray spectroscopy in the present thesis.

2.1 Gamma decay

The nucleus may be regarded as made of point nucleons, each carrying a magnetic
dipole moment and in case of protons, a net charge as well [2]. The charge distri-
Chapter 2. Theoretical aspects 15

bution interacts with the radiation field near the nucleus causing “electric” tran­
sitions and the intrinsic magnetism of each nucleon together with that generated
by current loops set up by the orbital motion of the protons induce “magnetic”
transitions in nuclei. The interaction between the nuclear currents and charges
with the radiation field is given by the coupling term:

He.m. = “ / j(r>f)- A(r> + / P(r> > (2-1)

where, j(r,i) and p(r, t) represent the current and charge densities in the nucleus
respectively. A(r, t) and ip(r, t) are the vector and scalar potentials of the radiation
field [3]. The interaction i?e.m. is considered as a perturbation causing transition
between stationary states of the nuclear Hamiltonian.

2.1.1 Transition probability


The transition amplitudes are determined in perturbation theory by the matrix
elements of He.m. between the initial and final states considered [3]. Using Fermi’s
golden rule, the transition probability is given by
2tt
w = yK^(r)|ft.„.|^(r))|V(£/) , (2.2)

where <po(r) and </>k(r) are the wave functions of the initial and final states respec­
tively and are products of nuclear and electromagnetic parts. p(Ef) is the density
of final states and is a product of the number of nuclear and electromagnetic states
per energy interval at Ef. In order to calculate the transition probability using the
above equation, it is necessary to make a multipole expansion of the perturbation
He rn_. This requires the decomposition of the potentials into electric and magnetic
multipole parts for the fields they describe. This results in the following expression
for the transition probability for multipole A from an initial nuclear state \JiMiQ
to a final nuclear state | [2]

mM JiC -> JfO = Aplri)!!P *T m JiC ^ Jf() ■ (2'3)

Here k is the wave number and B(X; J*£ —» J/£) is the reduced transition probabil­
ity which can be expressed in terms of the reduced matrix element of the multipole
Chapter 2. Theoretical aspects 16

operator for either electric (0\p (EX)) or magnetic (0\p {MX)) transition.

B{A; Ji( Jf0 = £ \iJfMft\Oxn\JiMiO\2 = ™IWfl|OA||«)|2 ■ (2.4)


IxM;

The electric transition operator is given by

Oxf{EX) = £ e(i)r}Yx,(9h *) , (2.5)


S=1

where the summation is taken over all A nucleons. are the spherical harmonics.
e(i) represents the electric charges for proton and neutrons in units of e. The free-
nucleon values are: e(i) = le for proton and 0 for neutron. In a shell model
calculation (Sec. 2.2.9) where some nucleons form a part of an inert core and a few
active nucleons in a truncated active space are used to determine the properties of
the nucleus, the above free-nucleon values cannot be used. Rather effective charges
have to be considered to account for such truncations. For example, the effective
charge of an active proton is taken to be 1.5e and that of an active neutron is taken
to be 0.5e in a typical shell model calculation. The magnetic transition operator
is given by

0Xfl(MX) = ± j ' V^Ya^, 4>i)) , (2.6)

where gs(i) and gt(i) are the spin and orbital g-factors respectively. The free-
nucleon values are: gs(i) = 5.586/rN for a proton and -3.826/7,w for a neutron; gt(i)
= lpM for a proton and 0 for a neutron.
The transition probability IY(A), is expressed as number of decays per unit
time. When B(EX) is expressed in units of e2/m2A and B(MA) is expressed in
units of p%fm2X~2 the numerical values of W and B{A) are related as

ahcli0mtUTc)2X+lE?+lB(X in e2fm2X)
W(A) =
^c(2^)2lfSriiFl(i)2A+1^A+^(A in nlfm2X~2) ,

where, the relation, e2 = ahc — 1.44 MeVfm has been used, a = is the
fine structure constant, Mp is the mass of proton, pN = eh/2Mpc2 is the nuclear
magneton. Incorporating the numerical values of all the constants in the above
Chapter 2. Theoretical aspects 17

equation, the electromagnetic transition probabilities for the lowest three multi­
poles are obtained as

W{E 1) = 1.59 x 1015 E3 B{E 1) ,

W(E2) = 1.23 x 109 E\ B{E2) ,

W(E3) = 5.71 x 102 £7 B(E3) ,

W(M1) = 1.76 x 1013 E* B(M1) ,

W(M2) = 1.35 x 107 E^ B(M2) ,

W(M3) = 6.31 x 10° £7 B{M3). (2.7)

The reduced transition probabilities can be obtained from the experimentally de­
termined mean lifetime (r), branching ratio (BR), mixing ratio (5), conversion
coefficient (a) (defined in the following sub-sections) and the gamma-ray energy
(Ery) using the following relations. Here, the reduced transition probabilities for
electric transitions are given in units of e2fm2X and those for magnetic transitions
are in units of p2Nfm2X~2. The gamma-ray energy E7 is in units of MeV.

0.629
BR 52
B(E1) =
r(l +1a) 1 + S2 ’
816 BR 62
"^ t(1 + a) 1 + 52’

1760 BR 52
B{E3) = (r is in units of ps),
' E1r r(l +1a) 1 + S2 ’
56.8 BR 1
B(M1) (r is in units of fs),
"^ r(l +1a) 1 + S2 ’
74.1 BR 1

- ^ r(l +1a) 1 + S2 ’
0.1585 BR 1
B(M3) , (r is in units of s). (2.8)
E1 t(1 + a) 1 + 52

Weisskopf estimates

Electromagnetic interaction operators have well known structure and only weakly
perturb the strong nuclear interaction. Hence it is possible to compare experi­
mentally determined 7-ray transition probabilities between nuclear states with the
theoretical prediction of nuclear models. Considering the extreme independent
Chapter 2. Theoretical aspects 18

particle model of nuclear structure (discussed in the next section) and considering
the nuclear transitions to be taking place by a nucleon moving from one single­
particle orbit to another without affecting the rest of the nucleons in the nucleus,
the “Weisskopf single-particle estimates” for the EX and MX reduced transition
probabilities are given by

B(E\) = T9X (JL)V/V/m«l , (2.9)

BIMX) = — (1.2)“-2 (tM) . (2.10)


7T \A-f-3/
These equations are only approximate estimates and are used as units (Weisskopf
or single-particle units, W.u.) for the actual B(EX) and B(MX) values. They
provide a basis with which one can compare observed and calculated values. If the
observed decay rate for a certain transition is several orders of magnitude smaller
than the Weisskopf estimate, it is likely that the overlap between the initial and
final wave functions is poor. Alternatively, if the observed transition rate is much
greater than the Weisskopf estimate, it is likely that the transition is a result of
collective motion of several nucleons.

2.1.2 Selection rules

The angular momentum selection rule for the A-th multipole electromagnetic tran­
sition from an initial state i to a final state / is given by

| Jj — Jj| < A < Jf + Jj. (2.11)

The lowest possible 7-ray multipole order when Jj = Jf is dipole (A = 1) since


there are no monopole (A = 0) transitions. For Jj = Jf = 0, the selection rules
will give only A = 0 and hence radiative transitions are not permitted in such cases
and the states decay through internal conversion in which the excitation energy
is emitted by ejecting an orbital electron. This is accompanied with the emission
of characteristic X-rays. In some cases like above, internal conversion is favoured
over 7 emission while in other cases it may be completely negligible. The internal
conversion coefficient (a) is defined as the probability of electron emission relative
Chapter 2. Theoretical aspects 19

to 7 emission. This coefficient decreases rapidly with increasing transition energy


in contrast to the probability for 7 emission. It increases with increase in multipole
order and is more important for heavier nuclei than for lighter nuclei [4],
The parity selection rule for a 7 transition is given by

TTiKf = (-1)A for EX ,

TTiiTf = (—1)A+1 for MX . (2.12)

A transition is said to be stretched if the change in angular momentum is the


largest that is allowed by the selection rule between two nuclear states. Because
of the selection rule, EX and MX transitions of the same multipolarity cannot
occur between same pair of nuclear states. For a given pair of nuclear states, if
both EX and M(X + 1) are allowed by angular momentum and parity selection
rules, the EX mode usually dominates the transition by a large factor. On the
other hand, if both MX and E(X +1) transitions are allowed, the higher multipole
order electric transition may be competitive in terms of transition rates with the
magnetic transition in spite of the hindrance factor due to energy dependence
(Eq. (2.7)).

2.1.3 Lifetime
The lifetime or mean life of an excited nuclear state is the average amount of
time it takes for a radioactive nucleus to decay. It is connected to the transition
probability and the half-life by the relation:

1 _ h/2
(2.13)
W ~ 0.693

2.1.4 Branching ratio


A given excited state may decay to more than one final states via different decay
paths. If the transition probability to each final state is denoted by W(i), the total
transition probability for the initial state is given by

W = Y,W{i) . (2.14)
i
Chapter 2. Theoretical aspects 20

The relation between the half-life ti/2 and the partial half-lives ti/2(i) = In2/W(i)
is given by the relation:
—=Y 1 (2.15)
*1/2 i *1/2(1)

The branching ratio is defined as the partial transition probability to a particular


final state as a fraction of the total transition probability from a specific initial
state, that is for the ith decay path, the branching ratio is given by:

W(i)
Branching ratio — (2.16)
W

2.1.5 Mixing ratio


Mixed transitions such as E(A)+M(A+1) or M(A)+E(A+1) allowed by selection
rules between a pair of nuclear states are characterised by the mixing ratio 5,
defined as:
W (M (A + 1))
82{M( A + 1)/E(X)) = (2.17)
W(E( A))
W(E(X +1))
82(E(X + 1)/M(X)) (2.18)
W (MX)
Conventionally the rate for higher multipole is in the numerator in the expression
for 8. For a mixed transition the branching fraction, b, for a given type of transition
is related to 8. For example, for a Ml +E2 transition, the E2 branching fraction is:

W{E2) S2
b(E 2) (2.19)
W(M1) + W(E2) “ 1 + 82

and the Ml branching fraction is:

W(M 1) 1
b(Ml) (2.20)
W(M1) + W(E2) ~ 1 + 82

2.1.6 Directional correlation and linear polarization


The probability of emission of a quantum of radiation depends on the angle between
the nuclear spin axis and the direction of emission. An ordinary source of radiation
consists of an ensemble of nuclei which are randomly oriented in space. Hence the
emitted radiation is isotropic. An anisotropic distribution of radiation can be
observed only if the angular momentum vectors of the radiating nuclei favour a
Chapter 2. Theoretical aspects 21

definite direction. An, oriented nuclei can be produced either as a result of nuclear
reaction or by a Coulomb excitation process or by a low-temperature orientation
technique. In the first two cases, the orientation-axis “z” is the beam direction.
In the latter case, the orientation-axis “z” is determined by the orienting external
fields. An oriented state can also be realized if the nuclei decay through successive
emission of two radiations 71 and j2 (Fig. 2.1). The observation of ji in a fixed
direction kx selects an ensemble of nuclei that has a non-isotropic distribution
of spin orientations. The succeeding radiation 72 then shows a definite angular
correlation with respect to ki [5]. In 7-7 directional correlation measurements only
the directions of the two radiations 71 and 72 are observed. This gives information
on the angular momenta carried away by the radiation, that is they yield spins of
nuclear levels. However they do not yield parities which can be determined from
polarization measurements.
The directional correlation function between 71 and 72 with multipolarities
Li + L[ and L2 + L2 emitted by a nucleus decaying through the cascade J—> J—> Jf
(Fig. 2.1) is given by

W( k1,ka)=W(fl)= 22 AktPt(cos«), (2.21)


k even

where 9 is the angle between ki and k2 and Pk(cos9) are Legendre polynomials [5].

------J,
, -,+^)
7 0
'___ J
72(L2+l!2)
■—~ Jf

Figure 2.1: A nucleus decaying through the cascade of two gamma transitions, 71 and
72 with multipolarities L% + l!l and L2 + l!2 from state to J/ via the intermediate
state J.

The selection rule for the summation index k:

0 < k < Min(2J, Lx + L'vL2 + L'2) . (2.22)


Chapter 2. Theoretical aspects 22

For a pure multipole radiation:

0 < k < Min(2J, 2LU2L2) . (2.23)

Dividing each coefficient A'kk by and expanding Eq. (2.21), one obtains

w(e) = \ + A^p2(cose) +... + (cose). (2.24)

Here Akk are normalised coefficients which can be broken into two similar factors,
each factor depending on only one transition of the cascade. Thus for the general
case of mixed transitions,

Akk — Ak(LiL1JiJ)Ak(L2L2JfJ) , (2.25)

where,

4 /T tr 7 t\ Fk(L1L1Ja)+2S1Fk(L1L[ja)+5fFk(L'1L,lJiJ)
= --------------------------- v", --------------------------- (2.26)
1 + Of

Here is the mixing ratio of 71 and Fk are the generalised F-coefficients when
the initial state is randomly oriented. A similar expression can be written for
AkiL2L2JfJ). The tables of the generalised F-coefficients can be found in Ref. [5,
6] and references therein. The generalised F-coefficients for mixed transitions
satisfy:
Fq(LL' JiJ) = 8ll> . (2.27)

It is assumed L' = L + 1. Theoretical values of only A22 and A44 are important
since usually only these can be observed experimentally.
In fusion-evaporation reactions (employed in the present thesis work), the initial
state (state J, in Fig. 2.1) is oriented, with the beam-direction as the orientation
axis. The projectile brings in orbital angular momentum in the m = 0 substate
with respect to the beam direction as the axis of quantization. The compound
nucleus formed in such reactions is strongly aligned with respect to the beam
axis. The orientation of an axially symmetric state is described by orientation
parameters B\ [7, 6] which are defined with respect to the orientation (symmetry)
axis. The index A refers to the tensor rank of the radiation field and its observation
Chapter 2. Theoretical aspects 23

which is described by an efficiency tensor; for example, A = 0 corresponds to a


non-observation of the radiation field in a particular direction. The directional
correlation of two radiations 71 and 72 (Fig. 2.2) that are emitted from an initial
state Ji that is axially symmetrically oriented is referred to as the DCO (Directional
Correlation from oriented states). The directions ki and k2 of the emission of the

/
/
/

Figure 2.2: The angles in a directional correlation of two successive radiations 71 and
72 emitted from an axial symmetric oriented source “S”.

7-rays are defined with respect to the orientation axis (chosen as the z-axis of the
reference system) (Fig. 2.2). The DCO function is then defined as (from Ref. [6]):

W{9U92,$) = £ BXl(Ji)^xXl(lMx2(72)HXlXX2(91^) , (2.28)


A1AA2

where,

A/ 4tt
HXlxxA91*2®) £ ^TT(Ai°Ag|A2g)yA(?(0i,O)r;2?(02,$) (2.29)

9i, 02 and $ are defined in the Fig. 2.2. BXl(Ji) are the orientation parameters
of the state Ji defined with respect to the orientation axis (z-axis). The indices
Ax and A2 are the tensor ranks of the statistical tensors that describe states Ji
Chapter 2. Theoretical aspects 24

and J. Since only directional correlations are considered and polarization is not
detected, Ai and A2 are even integers. The numerical values of the Clebsch-Gordan
coefficients (AiOAg|A2g) are given in [6]. A^2Al (71) are the generalized distribution
coefficient for 71 that connects two oriented states with orientation tensors of rank
Ai and A2.
FtAl (LihJJj) + + SjFf'
4!,a,(7i) = (2.30)
1 + i?

The tables of the above ^-coefficients is given in Ref. [6]. When the initial state
is random (Ai = 0), the Aa2Ai(7i) coefficients are given by Eq. (2.26) and the
i?AiAA2($i$2<h) reduces to ordinary Legendre Polynomial. Eq. (2.28) then represents
the ordinary directional correlation function given by Eq. (2.24).
The Ax2{rf2) are the ordinary directional distribution coefficients for 72. The
ordinary directional distribution coefficients for a transition from a state Ji to a
state Jj and with multipolarity L+L' is given by

, , A Fx(LLJfJi) + 251Fx(LL,JfJi) + 621Fx(L'L,JfJi} ,001,


^M7)- 1+52 ’ (2‘31)

where, the E-coefficients are ordinary E-coefficients tabulated in Ref. [8].


The directional correlation measurement is not sensitive to the electric or mag­
netic nature of the radiation. The information on this is obtained from the linear
polarization measurements. A nuclear state is said to be “oriented” if the rel­
ative populations P(m) of the angular momentum substates are unequal; i.e.,
E(m)/P(m'). If P(m)=P(—m), the state is “aligned”. If P(m)^P(—m), it is
“polarized”.
The angular distribution of linearly polarized gamma rays from an axially ori­
ented ensemble of nuclei is given by [8, 10, 7]

W(e, i/>) = g- E BXUX [AxPx{cos9) + 2Ax>2P^{cos9) cos 2$] , (2.32)


X—even

where B* are orientation tensors describing the degree of orientation of the parent
nucleus and UA are deorientation coefficients. PA are the ordinary Legendre poly­
nomials and P*2 are the unnormalized associated Legendre polynomials. Ax are
angular distribution coefficients which depend on the spin of the initial and final
Chapter 2. Theoretical aspects 25

state and the multipolarity of the 7-transition. The coefficients Aa,2 depend on
the electric/magnetic character of the radiation [8]. 0 is the angle that the emitted
quanta makes with the orientation axis whereas ip is the angle between the electric
vector E of the emitted quanta and the reaction plane (the reaction plane being
defined as the plane containing the orientation axis of the parent nuclear ensemble
and the direction of the primary radiation) [8].
The degree of linear polarization P(i9) of a 7 ray is defined as the difference
between the intensities of the radiations presenting an electric vector parallel to
the reaction plane (ip — 0°) and those with an electric vector perpendicular to that
plane (ip = 90°) [8, 11, 12}:

W(6, ip = 0) - W(9, ip = 7r/2)


[ ] W(9, ip = 0) + W(6, ip = tt/2) ’ 1 '

where the normalization is such that -1 < P(0) < +1.


P(0)— 0 for an unpolarized 7-ray and has a maximum value at 9 = 90°. Defining
2B\U\A\2 = (±)L'a.^ (where, (±)L> = 1 if (2)L -pole radiation is electric and it is
-1 if (2)l -pole radiation is magnetic), B^U\A\ = a\ and H\(LlJ) = 2a^/a a [13],
one obtains the polarization at 90° and for Aeuen<4, from Eq. (2.32) and Eq. (2.33),
as:
0 = 3a2ff2 - 7J3«4g4
K ' 2-a2 + 0.75a4 v ’
a\ are called the angular distribution coefficients. Considering the initial state to
be completely aligned by the fusion-evaporation process, the angular-distribution
coefficients (following Yamazaki’s notation [8]) are labeled as A™ax = Bk(jP)Ak
where Bk are the orientation tensors and Ak are the ordinary directional distri­
bution coefficients defined in Eq. (2.31). From the tables of the B(k)F(k) values
listed in Refs. [8, 14];, A™ax are determined. However, in actual cases where the
alignment is partial, attenuation coefficients (afc) [8, 14] have to be taken into
consideration. Thus the coefficients a\ in Eq. (2.34) are given by,

„ __ _ Amax
0,2 — CX.2A2 , (2.35)

a4 = a4A™ax . (2.36)
Chapter 2. Theoretical aspects 26

The attenuation coefficients depend on J, and the distribution of the nuclear state
over its m substates. According to Yamazaki [8], it is experimentally justifiable to
assume that partial alignment may be represented by a Gaussian distribution of
m-states characterized by a parameter a which is the half-width of the assumed
Gaussian distribution; in particular, the m-state population parameter is given by,

Pm{J) = exp(—m2/2a2)/ exp(—m/2/2o2) . (2.37)


m/ =—j

With this assumption otk are expressed as a function of J and a/J and are listed
by der Mateosian and Sunyar [14].
H2 and H4 are called linear-polarization mixing coefficients. They depend on
the initial and final spin and the mixing ratio (5) [15]. They can be derived for
mixed dipole-quadrupole (L = 1, L' = 2) and mixed quadrupole-octupole (L = 2,
L' = 3) transitions using the formalism given in Ref. [15, 10].
F2(ll) - (2/3)AF2(12) + 52F2(22}
H2(L — 1, L — 2) — (2.38)
F2(ll) + 28F2{12) + 52F2(22) ’
H4(L = 1, l/ = 2) = —1/6 . (2.39)
F2(22) - 6F2(23) + (2/3)52F2(33)
H2(L = 2,L — 3) = - (2.40)
F2{22) + 25F2(23) + <52F2(33)
5F4(22) - 28F4{23) + 208F4(33)
H4(L = 2, L' = 3) = (2.41)
30(F4(22) + 28F4(23) + 52F4{33)) '
One attempts to reproduce all the observable physical quantities of the nuclei
discussed in this section through mathematical models. A nuclear model is simply
a way of looking at the nucleus that gives a physical insight into as wide a range of
its properties as possible. The three important models are the liquid-drop model
that treats the nucleus as a macroscopic droplet of water, the shell model that
emphasizes the orbits of individual nucleons in the nucleus and the collective model
which complements the shell model by including the motions of entire nucleus. The
next section discusses the shell model that was used in the present thesis work.

2.2 The Shell Model


A good part of our understanding of the nature of the nuclear force comes from
binding energy measurements. Fig. 2.3 shows the binding energy per nucleon as
Chapter 2. Theoretical aspects 27

a function of atomic number A. The short range and saturation property of the

9
0*
Mean binding energy per nucleon (MeV)
-4 T
w
wo>
y*

Figure 2.3: Binding energy per nucleon. (The figure is taken from the Internet).

nuclear force can be deduced from the behaviour of the binding energy curve. If a
nucleon were to interact with every other nucleon then the binding energy should
have roughly increased as A(A — l)/2 and consequently the density of the nucleus
should have increased. But the binding energy is almost independent of A (~8.5
MeV/nucleon for A>12) and the electron scattering measurements indicate the
nuclear density to be fairly constant. This implies that a nucleon interacts with
only its nearest neighbours, and it follows that the nuclear force must be short-
ranged and therefore becomes saturated. The saturation of the binding energy
prompted the assumption of localised interactions. Also in spite of the Coulomb
repulsive force the existence of the nucleus consisting of positively charged protons
and neutral neutrons indicate that the nuclear force is much stronger than the
Coulomb force and is also attractive for all nucleons at least in the range of the
nuclear diameter. Coulomb repulsion becomes important for heavier nuclei. The
nuclear force saturates but not the Coulomb force, so there is a slight decrease in
binding energy at higher A. The binding energy curve at low mass region where it
Chapter 2. Theoretical aspects 28

increases with increase in A shows several kinks which correspond to sudden rise in
binding energy per nucleon at certain numbers of neutron and/or protons, called
“magic numbers”. It was observed that nuclei with magic number of nucleons
(N = 8,20,28,50,82,126) of protons and/or neutrons (N = 126 is experimentally
established magic number for neutrons only) exhibit the following properties:

• A larger total binding energy of the nucleus.

• A larger energy is required to separate a nucleon, i.e., larger Sp (proton


separation energy) and Sn (neutron separation energy) which leads to higher
stability and higher natural abundance for such nuclei.

• The first excited state occurs at higher excitation energy.

• Larger number of isotopes or isotones with same magic number of protons


or neutrons respectively.

• Relatively low neutron absorption cross-sections.

The shell model was successful in explaining these magic numbers in terms of shell
structure in analogy with the atomic shell model and derives its name from this.
This model describes the the structure of the nucleus as consisting of energy levels
grouped into shells and filled up by nucleons in order of increasing energy. The
magic numbers are nucleon numbers that correspond to shell closures. The energy
spacing between two shells is larger than that between levels within the same shell.
Although this is in analogy to the structure of the atom there are some some major
differences. In case of the atom the potential is supplied by the external Coulomb
field of the nucleus, in the nuclear case there is no external source, but the mean
field is exclusively due to the nucleon-nucleon interaction. Also there are two kinds
of nucleons giving rise to a new quantum number, isospin. Further, the nucleon-
nucleon interaction is attractive unlike Coulomb interaction and contains a strong
spin-orbit term. However, in spite of these differences the nucleons appear to move
in definite spatial orbits like electrons without collisions due to the weakness of
the nuclear long-range interaction and the Pauli exclusion principle. The Pauli
Chapter 2. Theoretical aspects 29

exclusion principle limits the number of nucleons in a particular level. Nuclear


collisions cannot occur as there are no levels available below the valence level to
which the nucleons can get scattered and Pauli exclusion principle prevents further
filling of these levels. The nuclear force is weak compared to the kinetic energy
of the relative motion of two nucleons. Hence in a nuclear collision it is unlikely
that enough energy would be transferred to excite nucleons to valence band (above
Fermi surface). Therefore to a first approximation, it is assumed in the shell model
theory that each nucleon moves independently in a potential that represents the
average interaction with the other nucleons in the nucleus. The model that is based
exclusively on this concept of central potential and independent motion of nuclei
therewith is called the “Independent Particle Model”. The Schrodinger equation
for single-particle levels:

£ + V(r,)) A, = wA, , (2.42)

where the first term represents the kinetic energy of the individual nucleons and
the second term represents the mean field potential. ipTi represents single-particle
wave function and a represents the single-particle energy of the ith particle. The
choice of the mean field or single particle potential is equivalent to requiring that
the potential should be spherically symmetric. It should be relatively constant
inside heavier nuclei as indicated by the constant density observed in such nuclei
and should fall off rapidly outside the nuclear surface. There are several potentials
available to model the nuclear potential, these include the Woods-Saxon poten­
tial, the square well potential, and the harmonic oscillator potential. These are
schematically represented in Fig. 2.4. The harmonic oscillator potential is widely
used due to its simplicity and ease of computation. It has the advantage that many
mathematical operations can be performed analytically. This potential has been
used in all shell model calculations in the present work and hence we discuss this
in detail.
Chapter 2. Theoretical aspects 30

----- Harmonic Oscillator


Vt
0-
/

Woods-Saxon

•* ---------Square Well

Figure 2.4: A comparison of two-dimensional representations of the harmonic oscillator


potential, the Woods-Saxon potential and the square-well potential.

2.2.1 The Harmonic Oscillator Potential


The harmonic oscillator potential is given by:

V(r) = -Va + \mu2r2


(2.43)
<u

Here, Vo is the well depth, m is the mass of the nucleon, and oo is the oscillator
frequency of the simple harmonic motion of the particle, hu) is approximately
41A-1/3 MeV. The oscillator frequency oo can be written as,

u J™!L (2.44)
\jrnR2 ’

where R is the nuclear radius. Substituting Eqs. (2.43) and (2.44) in Eq. (2.42),
the energy eigenvalues for a three-dimensional harmonic oscillator obtained are
given by
Enl = (N + ^)huo , (2.45)

where,
iV = 2(n-l) + i (2.46)

N labels the oscillator shell, with n = 1,2,3,... and l = 0,1,2, ...n — 1. The l
(orbital angular momentum) values are labeled using the spectroscopic notation
of s,p, d, f for l = 0,1,2,3.... From the above equation one can deduce the levels
belonging to a given oscillator shell. All levels with same value of N (oscillator
Chapter 2. Theoretical aspects 31

shell number) has the same energy given by Eq. (2.45), i.e. they are degenerate
and the degeneracy is given by 2(21 + 1). By Pauli principle, each nl level can
contain 2(21 + 1) particles. The parity is given by

7T = (-1)' . (2.47)

Therefore, one oscillator shell contains only states with the same parity. Total
number of particles for all levels up to AT is l/3(N+l)(N+2)(N+3). However only
first three magic numbers are reproduced correctly this way, as shown to the left
of Fig. 2.5.
The inclusion of a l2 term splits each oscillator level into levels with different
l as seen in the middle portion of Fig. 2.5. The effect of the l2 term increases
with increase in orbital angular momentum. So orbitals with higher values of l are
attracted more and their energies get lowered compared to low l orbits with same
oscillator shell number N. The addition of the l2 term gives a potential intermedi­
ate between the harmonic oscillator and the square well since it is equivalent to
a more attractive potential at larger radii and comes closer to the desired effect of
a more constant interior potential [17]. But even with this correction, the magic
numbers are not reproduced correctly.

2.2.2 The spin-orbit coupling

The addition of a strong spin-orbit potential by Goeppert-Mayer and indepen­


dently by Haxel, Jensen and Suess [18,19, 20, 21] was able to accurately reproduce
the magic numbers and therein lay the great success of the spherical shell model.
It couples the spin and orbital angular momentum of each individual nucleon and
so corresponds to the jj-coupling limit of atomic theory, though unlike the atomic
case this is not of electromagnetic origin [22], The total angular momentum

J=l + s. (2.48)

A single nucleon has s = ±|, so the possible values of the total angular momentum
are j = l±\ (except for Z = 0, in which case only j = \ is allowed). The spin-orbit
Chapter 2. Theoretical aspects 32

•3d 3/2
\ _--------------------------,4c
-4s - -------------------------------- in
\___________________________ .'**^870
-3d-
........... "-3Jm
-2g- -< -.v _______________ "^lSfl
‘tx-T.______________ lill/2

N=6 — ,x-" 4 ^ 2g9/2


•' 112 : -li - -<' 126
______________________________ 3Pl/2
*--• / 3p ---------------
1m / '------------ ■“x?3°
// X — 2f ■ ------------------------ ^ 1113/2
y \ lh 9/2

N=5 82 'Ztm
—lh
{70 } ___________"lhna
1m ______ ---------------------------- 2d 3/2
****** y— —3s • -•" ''3sl/2
----- 2d
✓ *8 7/2
N=4 ____ ° 5/2

—Ig- < 50
hm
; 40; ------------------------ ig9/2
-----------------------2p1/2

-2p- r' lf 5/2


2p3/2
N=3
-lf- 28
If.7/2
20; 20
Id 3/2
-2s- 2s 1/2
N=2 -ld-
Id 5/2

I&G) : 8 : 8
‘P 1/2
N=1 -lp-
lp 3/2

■: 2 ;

N=0 -ls- Is 1/2

Harmonic spin-orbit
l2 term +
oscillator term (l.s)

Figure 2.5: Single-particle energies for a simple harmonic oscillator, a modified harmonic
oscillator with l2 term, and a realistic shell model potential with l2 and spin orbit (l ■ s)
terms. Here N labels the oscillator shell
Chapter 2. Theoretical aspects 33

interaction is written as

Vr^r) = -V„^-T-g. (2.49)

Here V(r) is the central potential and Vis is a strength constant. The (l • s) term
causes the degenerate levels with same l to split into j = l± | levels with degeneracy

(2j + 1) which is the number of the magnetic substates for a given j. Hence by
Pauli principle the maximum number of nucleons that can be accommodated in

each orbit is 2j + 1. For example the 2p orbit splits into 2p3/2 and 2pi/2 (since
/ = 1 in this case). The maximum number of nucleons allowed in the 2p3/2 level is
2 x | + 1 = 4 and in 2pi/2 level is 2 x | + 1 = 2.

The expectation value of the operator l- s for states with definite l, s, j is given

by

(f- s) = ^[j(j +1) - 1(1 + 1) - s{s + 1 )}h2

= forj = l + 7,
= ^(J + l)*2, forj = l~\- (2-50)

For any pair of states with l > 0, the energy difference between the states is given

by:

<r-^,+i-(f-fUrs(2'+1)s2' (2-51)
The sign of the radial part of the spin-orbit potential is chosen to be negative so
that the member of the pair with larger j is pushed downward as observed exper­

imentally. The right part of Fig. 2.5 shows the effect of the spin-orbit coupling on
the level system. Now the shell closures appear at the observed magic numbers.
The real shells are thus not necessarily comprised of the levels comprising a har­
monic oscillator shell, particularly at higher levels. The parity of a wave function
is given by (—l)i as mentioned in Eq. (2.47). With only the harmonic oscillator

potential, the 2n +1 degeneracy leads to shells containing sets of l values differing


by even numbers. Hence as mentioned earlier each harmonic oscillator shell con­
tains levels of same parity. The spin-orbit term causes the lowering of the j = l+ \
orbit sufficiently so that, it is brought down among the levels of the next lower
Chapter 2. Theoretical aspects 34

shell. Thus a real shell bounded by magic numbers contains a majority of levels
of one parity called normal parity orbits and one level of opposite parity called
non-normal or unique parity or intruder orbit (originating from the next major
harmonic oscillator shell).

2.2.3 Application of the Independent Particle Model


According to the independent particle model:

• The nucleons in the nucleus are assumed to move in a common (mean) po­
tential.

• Most of the nucleons are paired so that a pair of nucleons contribute zero
spin and zero magnetic moment.

• The paired nucleons thus form an inert core, and therefore the predicted spin
and magnetic moment of even-even nuclei is zero.

• In the case of odd-A nuclei the properties of the nucleus are characterised by
the unpaired nucleon.

• For odd-odd nuclei, the properties of the nucleus are characterised by the
unpaired proton and neutron.

• The total parity of a multiparticle system of N independent particles in orbits


with orbital angular momentum h,h, •••, In is:

7r = nC-1)'1' = (-1)^' • (2-52)


i-l

As an example of application of the single-particle shell model, the nucleus 170


is considered. This has 8 protons and 9 neutrons. Applying the methodology
outlined in the previous sections and taking help of Fig. 2.5 the ground-state con­
figuration is:
for protons: (ls1/2)21 (lp3/2)4(lPi/2)21,
for neutrons: (ls1/2)2 |i(lp3/2)4(ljPi/2)2[ (l^s/2)1 -
Chapter 2. Theoretical aspects 35

From this one can see that all the protons are paired to give total angular momen­
tum zero and hence according to the model, should not affect the ground-state spin
of the nucleus. The neutrons are paired up to lpi/2 orbital with one odd neutron
in ld^/2 orbital. Thus the theory predicts the ground state spin-parity as 5/2+ and
experimentally it was found to be the same. If there are three particles in a j shell
then two of them can couple to zero angular momentum and the resultant spin is
j due to the unpaired nucleon. This J — 0 coupling of pairs of identical nucleons
in same orbit can be used to predict spin-parity of excited states formed by elevat­
ing one particle to higher energy orbits. However there are discrepancies observed
between prediction and observation when the prediction is based only on the rules
mentioned above. As an example, 33AS42 is considered where the neutrons are all
paired and the proton configuration is:
(lsi/2)21 (IP3/2)4 (lpi/2)21 (lcfe/2)6 (2si/2,)2ld3/21 (I/7/2)81 (2P3/2)4 (I/5/2)1 •
From the Single-particle shell model or the Independent particle model, the pre­
dicted ground state spin of 33A.S42 should be § . But experimentally the ground
state is |~. This required a modification in the rule which states that if a high
spin orbital (in this case I/5/2) comes after a low spin orbital (2p3/2 in this case),
the high-spin orbital fills faster, pairing its particles before the low-spin orbital can
be filled completely. Following this rule the ground-state configuration of 33A.S42
becomes:
(l^X/2)21 (I-P3/2)4 (lPl/2)21 (1^5/2)6 (2Si(/2)2ld|^21 (I/7/2)81 (2p3/2)3 (1/5/2)2-

This modification in the model is put in the form of pairing potential, which gives
paired nucleons a lower energy than unpaired ones, and which increases with in­
crease in l [16]. The pairing interaction is a “residual interaction” not envisaged
within the central potential.

2.2.4 Isospin
The proton and the neutron are regarded as two different charge states of the same
particle, the nucleon, which are indistinguishable under the strong interactions. In
analogy with intrinsic spin, each nucleon is assigned an isospin T — 1/2. Protons
Chapter 2. Theoretical aspects 36

and neutrons are distinguished by the projection of the isospin on an imaginary


isospin z-axis. For protons, this projection is -1/2; for neutrons this is +1/2.
In multinucleon system, the Tz projections add algebraically, and T values add
vectorially. For a nucleus with A nucleons (N neutrons and Z protons):

= X+< = - 2) (2.53)

and
(2.54)

n~z<t<a.
(2.55)
2 ~ “ 2
A given T has projection quantum numbers, Tz = T,T — 1,.. . — T, for example,
if T = 1, then Tz = 1,0, —1. T can never be less than its projection, i.e., T > Tz.
The p-p, n-n and p-n systems have:

• n-n'. Tz — +1/2 + 1/2 = +1, T —1,

• p-p:Tz — —1/2 — 1/2 = —1, T = 1,

• n-p: Tz = +1/2 - 1/2 = 0, T = 0or 1.

Figure 2.6 shows the level occupation and the spin-directions for the above three
systems. The n-n and p-p systems are allowed to exist only in S = 0 (anti-parallel
spin orientation) state by the Pauli principle. Tz calculated from the definition
given above comes out to be 1 and -1 respectively and hence T = 1 for these
two systems. The n-p system has Tz = 0. Experimentally the n-n, p-p and n-p
interactions are found to be identical (charge independence of the nuclear force).
Therefore the S = 0 state for the n-p system must be identical to the n-n and p-p
systems and hence T = 1 in this case. But the Pauli principle allows n—p system to
exist also with parallel spins (5 = 1). This state then must have T — 0 since if the
only substate of a spin-like quantum number has a value zero, then the spin itself
has that value. In nature the p-p and n-n systems are unbound but the deuteron
which is a n-p system has one bound state with spin-parity 1+. Experimental
observations reveal the bound state wave function to be in predominantly 3 Si
Chapter 2. Theoretical aspects 37

Figure 2.6: Illustration of level occupation and and spin directions for the 2 nucleon
system. Protons are indicated by filled circles, neutrons by open circles. The total spin
and isospin quantum number are given on the left [17].

state with a small admixture of 3Di. Thus the only bound state of deuteron has
S = 1 and is therefore an isospin singlet state (Tz = 0, T = 0). This can also
be deduced from the fact that the p-p and n-n systems with T = 1 are unbound
and hence the identical n-p configuration with 5 = 0 and T = 1 in deuteron is
also unbound. The concept of isospin can be extended to heavier systems which
have several n-n, p-p, n-p pairs, by studying the properties of the isobaric or the
isospin multiplets, i.e., nuclei with same number of nucleons (same A), and the
same isospin, but different z-components of isospin (Tz). From the case of the
deuteron and isospin triplets the following observations can be made:

1. The charge independence of the nuclear force, i.e., the equivalence of n-p,
p-p and n-n forces applies specifically and only to the T = 1 coupling for the
p-n system.

2. Different spin couplings (isospins) have different energies, and the strength
of the interaction of two unlike nucleons in a T = 0 state is stronger (more
attractive) than for a T = 1 state.

3. Different isospins represent different spin couplings and can also represent
states in which nucleons occupy different orbits. Within the same nucleus
Tz remains the same but T can be different and such states with different T
Chapter 2. Theoretical aspects 38

differ in the orbits occupied by the protons and the neutrons or in the spin
couplings.

4. States in different nuclei of an isospin multiplet with same T are called iso-
baric analogue states. Such states differ in interchange of protons into neu­
trons and vice-versa. The absolute binding energies of these states are identi­
cal once they are corrected for Coulomb energies and the n-p mass difference.
The excitation energy and other properties such as spectroscopic factors and
electromagnetic transition strengths of these states are also similar.

5. In isobaric multiplets (in light nuclei where Coulomb effects are small) the
highest orbit symmetry and the lowest total energy occurs for the smallest
Tz nucleus of the multiplet. N = Z nuclei have the lowest Tz and hence
are most stable. That is why the valley of stability has N ~ Z nuclei until
a mass A ~ 30-40 where Coulomb effect shifts it to the neutron-rich side.
Generally, states of lowest isospin lie lowest and are most proton-neutron
symmetric states in the space and spin coordinates: that is, Pauli principle
favours low T.

In shell model calculations both isospin and proton-neutron formalism can be


used. The proton-neutron (p-n) formalism, unlike the isospin formalism treats the
protons and neutrons as distinguishable particles. The isospin formalism often
enormously simplifies complex shell model calculations in nuclei with both valence
protons and neutrons which otherwise will be of considerable complexity.

2.2.5 Residual Interactions


The experimental data cannot be explained by considering a central potential
alone. The residual interactions become important when dealing with nuclei con­
taining several particles outside closed or magic configurations. Residual interac­
tions affect the energies of multiparticle configurations by lifting the degeneracy of
all the J states formed by coupling of particles outside the core. Residual interac­
tion are responsible for all configuration mixing, collectivity and other correlations
Chapter 2. Theoretical aspects 39

in the nucleus.
The basic assumption in shell model of the existence of a central potential in
the nucleus under which nucleons orbit as independent particles is a good approxi­
mation of the actual motion. But in reality, there is no pure central potential. The
actual Hamiltonian consists of kinetic-energy terms Ylk T(fc) and the two particle
interaction ^2k<lW(k,l). Schrodinger equation (Eq. (2.42)) now reads [3]:
'a a
zm+ £ iv(k,i) *(1,2, ...A) = £4(1,2, ...A) . (2.56)

The Hamiltonian can be written as,

H = £[T(*) + 17(*)] + £ V/(k,l)]--£U(k) if(0) + (1) , (2.57)


k=1 1=k<l k-1

where U(k) is the single-particle potential or central potential. is the residual


interaction which is regarded as a perturbation. H^ is the unperturbed Hamilto­
nian. The basis of single-particle states 4>a{r) is determined by the single-particle
Schrodinger equation
T<f)a (r) + U<f>a (r) = ea(j)a (r) . (2.58)

Here the eigenvalue ea represents the single-particle energy with a labeling the
single-particle state \nljm). Any product of A single-particle wave functions <f>W
= ))—.<paA(f'(A)) satisfies the Schrodinger equation

ff(°)$(0) = E(o)$(o) _ (2.59)

The unperturbed energy is given by

Ef! = £ e.„ . (2.60)


jfc=l
The product of A-particle wave functions has to be antisymmetrized to satisfy
Pauli exclusion principle. This is done by taking appropriate linear combinations
of function $t0)(r(1),....,r(A)). Let an antisymmetrized A-particle wave function
possessing a well-defined total angular momentum (J) and isospin (T) be denoted
by <&r(r(l), ...(r(A)). In the first order perturbation theory,

l*r> = |40)> + I®?1) (2.61)


Chapter 2. Theoretical aspects 40

and
Et = 40) + £r} , (2-62)

where |$[X) and E^ represent the supposedly small changes in the wave function
and the energy of the unperturbed state. Substitution of Eq. (2.57), (2.61) and
(2.62) in Eq. (2.56) yields

(if(0) + tf(1))(|40)> + I#)) = (40) + ^Xl#) + I#)) • (2-63)

Disregarding the second-order quantities and separating the zeroth- and fist- order
quantities one gets:
if(0)|40)> = 40)|40}) , (2.64)

0)|41)) + 77(1)|40)) = ^0)|41)) + EPi®?) • (2.65)

Multiplying Eq. (2.65) by (4°^l on the left and using Eq. (2.59) one obtains:

41} = <40)[i?{1)|40>) . (2.66)

This is the energy shift in presence of residual interaction HW from the degenerate
case. The energy of the state |44 is given by:

Ev = EP + EP = (40)|^(0) + H(1)|40)) = Ee*k + (40)|ff(1)l40)) • (2-67)


k=1
The first term is the contribution from the single-particle energies and the second
term is that from the residual interaction. An example of a two-particle config­
uration consisting of a closed-shell inert core (no excitations allowed from core)
with spin zero and two nucleons in equivalent orbit outside the core is considered
here [3]. The Hamiltonian in such a case is

H = Hcore + H\2 ■ (2.68)

Hcore refers to interaction between core particles (numbered k = 3, ...A) and is


taken to be constant for an inert core. Hcore is given by Eq. (2.57) with the
summation going from k = 3, ...A particles. H\2 describes the contribution of the
two extra-core particles.
H12 = 't[T(k)+U(k))+[i:'tw(k,l)+W(l,2)-'tu(k)} = , (2.69)
k=l k=11=3 k=1
Chapter 2. Theoretical aspects 41

where,

#£’ = PX1) + £7(1)1 + pr(2) + V(2)] = H„.( 1) + H^X2) , (2.70)

denotes the single-particle Hamiltonian and

= E W(l,.I) - [7(1)1 + E W(2, () - £/(2)] + ^(1,2)1 , (2.71)


{=3 1=3

specifies the residual interaction. Generally the single-particle potential is taken


to be a Harmonic-oscillator potential and the residual interaction is represented
by a two-body interaction V(l, 2). So the total Hamiltonian is given by

H = Hcore + Hs.p.{ 1) + 2) + V(l, 2) . (2.72)

The information on the properties of the two-body interaction, F(l,2), can in


principle be obtained by analyzing the contribution of the two particles to the total
energy of a nuclei which is a inert core plus two nucleon system. The wave function
(1, ...A) can be written as an antisymmetrized product of the core wave function
$oo(core) and the wave function describing the extra two nucleons $p^(l, 2). The
simple product function also suffices. Therefore from the orthonormality of the
wave functions one obtains:

($oo(core)$^0)(l, 2) |jH|$0o (core) $£0)(1, 2)) = ($oo(core)|i7core|$00(core))

+($<?>(1,2)|ffs.p.(l) + ff,,.(2)|#(l, 2))

+(40>(1,2)|V(1,2)|4°>(1,2)) . (2.73)

Here the first term is the binding energy of the core. The second term expresses the
single particle energies of the two extra-core particles and the third term expresses
the two-body matrix elements of the residual interaction. For more than two
particles outside the inert core the multiparticle interaction can be expressed in
terms of the two-body matrix elements.

Multipole decomposition of residual interaction

A two-body interaction V{f\ — rjj) can be expanded as

V(fi - r2) = Yl vkPk{cos6) , (2.74)


k
Chapter 2. Theoretical aspects 42

where the expansion coefficients Vk are given by:

2k + 1 r
vk = —-— J V(ri- f2)Pk{cosO)dO (2.75)

Each multipole k (order of the multipole), has an angular dependence Pk(cos9).


The multipole orders relevant to the shell structure are the monopole and the
quadrupole component. The monopole part (k = 0, Po(cos9) = constant) of the
interaction is constant over all (angular) space. It is given by the average of the
interaction over all directions [17, 23, 24]

yT _ Ej(2J + l){jlj2\V\jlh)jT
(2.76)
juh~ Ej(2/+1)
where, {jij2\V\jij2)JT stands for the matrix element of the two-body interaction
V for a state coupled to angular momentum J and isospin T. The monopole term
is the same for all relative orientations of j\ and j2, that is, all J values. So
it does not contribute to splittings of a multiplet but only gives an overall shift
to the multiplet. It is viewed as the mean energy brought to the nucleus by
the addition of two interacting nucleons irrespective of the relative orientation
of their orbits. The evolution of single-particle levels and subsequently of shell
gaps are to a first order governed by monopole interactions. It is the principal
origin of the mass dependence of relative single-particle energies. In particular, the
monopole p-n interaction leads to the changes in neutron single-particle energies
as a function of proton number and vice-versa. Thus it influences the evolution of
shell closures. The J-dependent part of the two-body matrix elements is contained
in the multipole components. The quadrupole component leads to quadrupole
deformation. The quadrupole moment is found to be sizable, it is strongest at
9 = 0° and 90°. The quadrupole component contributes to the splitting of the
multiplet. The long-range component of the residual interaction which is crucial
in producing the collective properties and non-spherical nuclei is simulated by the
quadrupole interaction. The multipole expansion of any interaction can be used to
estimate the effects of an arbitrary two-body residual interaction between nucleons
in any pair of single particle orbits [17].
Chapter 2. Theoretical aspects 43

2.2.6 Single-particle energies


The single-particle energies, in the independent particle model are the eigen values
obtained by solving Eq. (2.42) or Eq. (2.58). These are called the bare (spherical)
single-particle energies and can be experimentally determined from a number of
nuclei that are one particle or hole removed from various major closed shells, with
one-nucleon transfer reactions, such as, (d,p) and (d, t) [17]. If the kinematical
aspects of the reaction collision are ignored, the cross-sections in these reactions
give a nuclear matrix element which relates to the purity of the single-particle state.
Spectroscopic factor in a (d,p) reaction, is a measure of this purity. S(d,p) = 1
corresponds to a pure single-particle neutron wave function coupled to the target
nucleus.
The single-particle energies and the sequence of the single-particle levels as pre­
dicted by the independent particle model and depicted in Fig. 2.5 are not constants.
They depend on the number of nucleons in the nucleus since the single-particle po­
tential arises from these same nucleons. One of the reasons behind the dependence
of the single-particle energies on N and Z is the potential and its boundary condi­
tions. For example, in case of a square-well potential, the energy goes as E ~
Therefore, in the shell model, approximately,
Ea.P.e. ~ ^ ~ (2.77)

This dependence on A is depicted in Fig. 2.7. The spacings get compressed for
heavier nuclei. There is a simultaneous decrease in all energy levels with increase
in A hence this effect does not alter the shell model predicted sequences of the
levels significantly. The other effect on single-particle energies comes from the
residual interactions. This can cause drastic changes in the energy and sequence
of the shell model levels. The residual interaction between closed shells of filled
orbits and a single valence nucleon outside that shell in an orbit j and magnetic
substate m alters the SPE of that orbit. The wave function of a closed shell is
spherically symmetric (J = 0), that is, the closed shell has no preferred direction
in space. Hence its interaction with the valence nucleon is independent of m, i.e.,
independent of direction. This implies that the shift in SPE is exclusively due to
Chapter 2. Theoretical aspects 44

Figure 2.7: Changes of single-particle energies with nucleon number, reflecting the de­
pendence of energies of confined particles on the size of the containment volume [17].
Chapter 2. Theoretical aspects 45

the monopole part of the interaction between the closed and open shell nucleons
since that is the only multipole that is 9 independent (Pq[Cos9) = constant).
A major shell consists of orbits of different j values, and each of the j values of a
closed major shell can have a different (spherically symmetric) effect on each of the
valence j orbits, thus shifting them to different extents. Thus the relative single­
particle energies in a given major shell depend critically on the specific lower-lying,
filled closed shells. The interaction of a given shell j\ with another shell j2 (closed
or open) depends on the extent of overlap between the two; orbits with same n, l, j
are likely to have higher overlaps. If one of the two interacting shells is closed then
the overlap is radial since as stated above, the interaction in this case in angle
independent.
The filling of neutron (proton) orbits can strongly change the proton (neu­
tron) level order due to strong attractive p-n interaction. The proton and neutron
single-particle sequences are slightly different because of Coulomb potential. The
monopole interaction is given by Eq. (2.76). The monopole shift of a certain orbital
jx increases linearly with increasing number of nucleons in the orbital j2. As the
orbit j2 is occupied, the single-particle energy of the orbit ji, is changed by

Af* = V^j2nh , (2.78)

where nj2 is the occupation number (operator) of orbit j2 [25]. At shell closures, all
other components except the monopole component vanish. The monopole compo­
nent governs (spherical) SPEs on top of closed (sub)shells according to the above
equation. In open shell systems, its effects can be viewed through Eq. (2.78) as
effective single-particle energies.
The separation energy of an orbit calculated with the bare SPE of the config­
uration space and including the effects of the monopole interaction is called the
effective single-particle energy (ESPE) and depends on the configurations. Thus
the spacing between the ESPEs defines the energies for the excitation of the indi­
vidual nucleons, and thus the effective shell gaps [24, 26].
Chapter 2: Theoretical aspects 46

2.2.7 Multiparticle configuration


Multiparticle configuration implies valence configurations of two or more particles
outside the inert core. The Pauli exclusion principle requires that a many-nucleon
wave function be antisymmetric in all particle coordinates for identical nucleons. In
the isospin formalism where neutrons and protons are considered different charge
states of the same particle, one has the generalised Pauli exclusion principle that
requires a multi-nucleon wave function to reverse its sign upon an odd permutation
of all coordinates (space, spin and isospin) of any two nucleons. The following two
subsections deal with the prediction of spin for two-particle and more than two-
particle configurations outside a closed shell applying the above principle.

J values in two-particle configurations

Two particles outside inert core in orbits ja and jb can couple to give total J =
integer values from \ja — jb\ to (ja + jb)- A two-particle wave function (space-spin
part) is given by [3]:

$jM(ja(l)jb{2)) = (jamajbmb\JM)(t).Jama{l)4>jimh{2) , (2.79)


TTlaTTlb

where, (jamajbrnb\JM) is a Clebsch-Gordon coefficient and j>jama{ 1) and <pjbmb(2)

denote single-particle wave functions for particle 1 in orbit ja and particle 2 in


orbit jb, respectively. When P12 denotes the operator that interchanges particles
1 and 2, one obtains

Pl2*h JM {ja{l) jbity) = P\2 (jamajbmb\J M)(f)jaina (l)0j(,m6 (2)


7Yl(i 77^h

= J2 \3o,rna]bTnb| JM)(jjama (2)j>jbrjlb (1) . (2.80)


mamb

From the symmetry property of the Clebsch-Gordon coefficients,

{jarnajbmb\JM) = (-l)J~3a~M(jbmbjama\JM) , (2.81)

one finds

TYIq 771 b

= (~l)J^jb$JM(jb{l)ja(2)) ■ (2.82)
Chapter 2. Theoretical aspects 47

• Two-particles in same orbit:


For identical orbits, i.e. na = rib, la = and ja = jb = j,

= (-i)J-2i$jM(i(i)i(2)) = -(-i)j^M(j(i)j(2)),
(2.83)
since 2j is always odd. Thus <hjM is antisymmetric for J even and symmetric
for J odd. Similarly the isospin part of the two-particle wave function is
antisymmetric for T even and symmetric for T odd. Thus the total two-
particle wave function for two nucleons will be antisymmetric if

J + T — odd . (2.84)

Next we consider two cases:

— Two identical nucleons in the same orbit:


Here T — 1. Therefore on imposing the condition, J+T = odd, we find
that the only allowed J states for two identical nucleons in equivalent
orbits are those of even total angular momentum. J = 0,2,4, ...(2j — 1).

— Two nonidentical nucleons in the same orbit:


In this case T = 0 or 1. For T = 1 all even values of J are allowed and
for T = 0, all odd values of J are allowed.

• Two-particles in different orbits:


A two-particle wave function, with two particles in different orbits can be
antisymmetrized for any combination of the total J and T values.

Many predictions for \jnJ) configurations are identical to those for \j2J) config­
urations if seniority is conserved (where seniority (j/) is defined as the number of
unpaired particles in a state of angular momentum J in the configuration jn, so
that the number of unpaired particles is (n-v)). There are several ways of predict­
ing the total angular momentum for a multiparticle configuration. The m-scheme
approach is one such approach that has been employed in the code used in the
shell model calculations in the present work.
Chapter 2. Theoretical aspects 48

J values in multiparticle configuration: the m scheme

The m scheme approach is based on the consideration of the possible m-substates


allowed to be occupied by the particles under the the Pauli exclusion principle.
The case of three identical nucleons in d5/2 orbit is considered as an example [17].
The maximum allowed J value cannot be 15/2 (J ^ ji + j2 + j3 = 5/2+S/2+5/2
= 15/2). This is because J = 15/2 implies M =15/2 substate which can only
be formed if all three particles are in m = 5/2 substates. This however is not
allowed by the Pauli exclusion principle. By the same logic J = 13/2 or 11/2 are
also not possible The maximum J value possible is 9/2 which can be formed by
placing three particles in the states jirrii = 5/2, 5/2, j2m2 = 5/2, 3/2 and /3m3
= 5/2, 1/2. Here, Mmax = 5/2 + 3/2 + 1/2 = 9/2. Therefore Jmax = 9/2 (since
M = J, J — 1,..., — J). The allowed J and M values are listed in Table 2.1 for the
|(5/2)3«7) configuration. The M < 0 cases are symmetrical to M > 0 cases and
hence have not been listed in the Table 2.1.

Table 2.1: m scheme for the configuration |(5/2)3J) [17].

ji = 5/2 h = 5/2 k = 5/2 M J


mi m2 m3
5/2 3/2 1/2 9/2 J = 9/2, 5/2, 3/2
5/2 3/2 -1/2 7/2
5/2 3/2 -3/2 5/2
5/2 3/2 -5/2 3/2
5/2 1/2 -1/2 5/2
5/2 1/2 -3/2 3/2
5/2 1/2 -5/2 1/2
5/2 -1/2 -3/2 1/2
3/2 1/2 -1/2 3/2
3/2 1/2 -3/2 1/2

Configuration mixing

Excited states generally do not have pure configurations but are complex linear
combinations of multiparticle wave functions. The residual interaction mixes con­
figurations and for a given state the nucleus has only a certain probability to be
Chapter 2. Theoretical aspects 49

in a particular configuration. The mixing depends on the initial separation and

on the matrix element. There may be several states with close-lying energies that

can mix to give an observed nuclear state. A series of two-state mixing calcula­

tions can however approximate such a situation. Two initial states with energy Ex

and E% and wave functions </>i and 02 with the same spin-parity and isospin are

considered [17]. Under the influence of the residual interaction V they mix to give

the observed nuclear states with energies Ej and Eu and wave functions ipi and

'ipu (Fig. 2.8). The task is to find the appropriate linear combination of (pi and </->2

%
Eii
02 AEs
E,

AE UNP

E,
0i AES
Et
Vi

Figure 2.8: Two-state mixing: definition and notations [17].

that gives ipj and 'tpi;.

i>i = a0i + Pfa , (2.85)

: fpll = «02 - P<t> 1 , (2.86)

where a2 + f32 = 1 . The wave functions and final energies are obtained by diago­

nalizing the matrix


( El V
(2.87)
\v e2 j
where V denotes the mixing matrix element {^>i\V\<j>2j. The perturbed energies in

terms of the unperturbed energies and the mixing matrix element are given by

E,j, = i(Ei + E2) ± P(E2-E1y + 4P .


(2.88)
Chapter 2. Theoretical aspects 50

The amplitude f) is given by

D= 1 . (2.89)

Eqs. (2.88) and (2.89) are universal expressions completely independent of the
nature of interaction or the initial splitting [17].

2.2.8 Limitations of the spherical shell model


The spherical shell model is capable of explaining the magic numbers and also
many other nuclear properties such as spin, magnetic moment and energy levels
but only for nuclei near shell closures. Experimental electromagnetic transition
probabilities are reproduced near major shell closures but otherwise the experi­
mental values are much larger than predicted values. Again, there is reasonable
agreement between the predicted and the experimental quadrupole moment for
nuclei near the major shell closures. Theoretically the quadrupole moment Q0
should vanish for j = \ and should be negative otherwise. But experimentally in
most nuclei the quadrupole moment are positive and about hundred times larger.
For these nuclei the model has to be modified to account for nuclear deformations
and this gives rise to Deformed Shell Model. These models have not been used
in the present work and hence the discussion is restricted to the Spherical Shell
Model only.

2.2.9 Shell Model Calculation


In a shell model calculation, the energy spectrum of the nucleus can be predicted
by solving the Schrodinger equation Hty = AT. As mentioned earlier, this requires
the energy matrix to be constructed where the matrix elements of
the given interaction between many particle states are expressed in terms of two-
body matrix elements. This matrix is then diagonalized to obtain the eigenvalues
that yield the required energies. The eigenfunctions give the wave functions of the
state.
Chapter 2. Theoretical aspects 51

In general, an inert core is chosen along with a configuration/model space out­


side this core consisting of limited number of single-particle orbits for the valence
nucleons. Higher-lying orbits which are unlikely to be occupied and are irrelevant
to the properties of the nucleus under consideration, are neglected. The core is
treated as inert, i.e., no excitations from the core are considered in the calcula­
tion. The two-body matrix elements of the residual interaction together with the
single-particle energies of the active orbits constitute the inputs to the shell model
calculation. The single-particle energies are bare (spherical) single-particle ener­
gies which are obtained from the experimental spectra observed with the same core
and one nucleon outside the core. The two-body matrix elements can be evaluated
by either empirical or realistic or schematic method. In the empirical approach the
TBME are determined empirically by comparing the experimental data of several
nuclei in a given mass region, once the SPEs are determined. The optimum values
are obtained from a least-squares fit to the experimental data. In this approach
there is no need to explicitly specify the two-body interaction. One example of
empirical interaction is the universal (2s ld)-interaction (USD) [27, 28]. In the
realistic approach the TBMEs are not obtained from a fit to spectroscopic data
but are derived from a bare NN potential for free nucleons in vacuum. That is why
they are referred to as realistic interactions. One such interaction used frequently
in shell model calculations is the one obtained by Hamada-Johnston that has been
used by Kuo and Brown for the (1/ 2p) shell nuclei [29]. However to be applied
to shell model calculations they are derived taking into account medium effects,
the Pauli exclusion principle and truncated model space. Then such an interaction
is called an effective interaction. The effective interaction is expected to account
for all the processes occurring outside the chosen configuration space and also the
short-range repulsion and the core polarization. In the schematic approach, certain
rather simple two-body interactions such as the surface delta interaction are used
to successfully correlate many observed nuclear properties when a truncated shell
model basis is used [3]. A few parameters, characterizing a schematic interaction
are adjusted to reproduce low-energy spectra of a few neighbouring nuclei of inter­
est. These parameters are then supposed to change from one region of the nuclear
Chapter 2. Theoretical aspects 52

chart to another.

2.2.10 Present calculations


In the present work, large-scale shell model calculations were carried out to inter­
pret the experimental observables 0f 32j33’34P and 33S (Chapters 6 and 7). The cal­
culations were carried out with the code NUSHELL@MSU [30] using the “sdpfmui'
interaction, taken from the Warburton, Becker, Millener, and Brown (WBMB) sd-
pf shell Hamiltonian [33] assuming a 160 core.

The sdpfmw Interaction

The sdpfmw interaction is derived from an effective one body plus two-body Hamil­
tonian and is composed of three parts: the USD (2s, Id) interaction, the McGrory
(1/, 2p) interaction and the Millener-Kurath cross-shell interaction [33]. It was
designed to operate in the (2s, ld)A“16"(1/, 2p)n model space with a single-value
of n and was developed to explain the 32 region. The core is assumed to be
160.
In the universal (2s, Id) interaction denoted by USD [34, 27, 28] the parameters
consist of 63 two-body matrix elements and three single-particle energies relative to
the 16 0 core that were determined from a least-squares fit to ~440 binding energies
in A = 18 - 39 nuclei. The McGrory’s interaction [35] consists of 195 TBME and
four SPE relative to 40Ca core. The starting point of the interaction is the (If, 2p)
effective interaction of Kuo and Brown [36]. McGrory performed a least-squares fit
to 29 binding energies in the A = 42 - 44 region with only eight (/7/2/7/2IUI/7/2/7/2)
TBME variable. The Millener-Kurath particle-hole interaction [37] was developed
to describe the non-normal parity states in the A = 16 region. A fairly good fit
was obtained to the T = 0 and 1, lp-lh states of 160 with the parameters of this
interaction.
Warburton et al. [33]: used the Millener-Kurath potential to generate 510 TBME
needed to describe the interaction connecting the (2s, Id) and (If, 2p) shells. Out
of these, selected TBMEs were varied to give a better agreement with the the T =
Chapter 2. Theoretical aspects 53

0 and 1, lp-lh states of 40Ca. For A <40 nuclei the A-0-3 dependence was adopted
for all the three interactions while for .4>40 nuclei the TBME of the Millener-
Kurath and McGrory’s interactions are fixed at their >1=40 values. In calculations
with (2s, ld)A~16~n(lf, 2p)n model space, the TBME of the USD are given an A
dependence appropriate to A — n rather than A. The single-particle energies are
assumed to be independent of A. The SPEs of the ld5/2, ld3/2 and lsi/2 orbits
are those of the USD interaction and SPEs of the four fp orbits are the relative
values of McGrory. The energy gap between (2s, Id) and (1/, 2p) was initially
determined from the difference between the experimental binding energies of41 Ca
and 40Ca. Later, this energy gap was fixed so as to best describe binding energies
of selected levels in the A = 35 - 43 region.

The Shell model code: Nushell@MSU

The Nushell@MSU code [30, 31, 32] is an m-scheme code at heart but uses angular
momentum and isospin projection to produce a basis with good J and good T if
desired. A basis state with given value of M does not in general have a definite
value of the total angular momentum J. However the angular momentum projec­
tion operator can be used to construct explicit linear combinations of the m-scheme
Slater determinants which have good total angular momentum. The Hamiltonian
is evaluated in the reduced dimension (J,T) basis. This is much smaller than the
m-scheme matrix. This matrix is then diagonalized and the eigenvectors are then
generated in the full m-scheme basis before observables are calculated. The central
potential is a Harmonic oscillator potential. There is a library of model spaces and
interactions. The model space indicates the orbitals and the truncation within that
set of orbitals which is assumed for a given calculation. The basic inputs are the
single-particle matrix elements and the two-body matrix elements. The program
calculates reduced transition probabilities from which the transition probabilities,
lifetimes, branching ratios and mixing ratios can be calculated using Eqs. (2.7),
(2.13), (2.16), (2.17) and (2.18). The sequence of programmes for prediction of
energies of the nuclear states is given below:
Chapter 2. Theoretical aspects 54

1. SHELL makes a batch file that coordinates the program sequence and their
inputs.

2. NUBASIS makes a list of all possible m-scheme basis states for a given model
space together with a given set of restrictions. The single-particle states for
the calculation are contained in a data file, and restrictions can be placed in
both individual J orbits and major shells.

3. NUPROJ makes linear combinations of the m-scheme basis states (within a


given partition) that have good J values in p/n formalism or good J and T
in isospin formalism by using the angular momentum and isospin projection.
The number of linear combinations is the J (or JT) basis dimension.

4. NUMATRIX constructs the Hamiltonian matrix corresponding to the J (or


JT) basis dimension of the problem.

5. NULANCZOS finds the lowest N eigenvalues and eigenvectors for the matrix
by the Lanczos diagonalization method. The output of this program is the
set of eigenvalues and the eigenvectors in the projected basis.
Bibliography

[1] H. Ejiri and M. J. A. de Voigt, Gamma-ray and Electron Spectroscopy in


Nuclear Physics, Clarendon Press, Oxford (1989).

[2] Samuel S. M. Wong, Introductoy Nuclear Physics, Prentice-Hall of India Pri­


vate Limited

[3] P. J. Brussaard and P. W. M. Glaudemans, Shell-Model Applications


in Nuclear Spectroscopy, North-Holland Publishing Company, Amsterdam-
New York-Oxford, 1977.

[4] Kenneth S. Krane, Introductory Nuclear Physics, John Wiley & Sons, Inc.,
New York, 1988.

[5] Kai Siegbahn, Alpha- Beta- and Gamma-Ray Spectroscopy, Volume 2, North-
Holland Publishing Company, Amsterdam-New York-Oxford.

[6] K. S. Krane, R. M. Steffen and R. M. Wheeler, Nucl. Data Tables 11, 351
(1973).

[7] R. M. Steffan and K. Alder, The Electromagnetic Interaction in Nuclear Spec­


troscopy, North-Holland, Amsterdam, (1975).

[8] T. Yamazaki, Nucl. Data. A 3, 1 (1967).

[9] J. K. Deng, W. C. Ma, J. H. Hamilton, A. V. Ramayya, J. Rikovska, N. J.


Stone, W. L. Croft, R. B. Piercey, J. C. Morgan, and P. F. Mantica, Jr., Nucl.
Instrum. Methods Phys. Res. A 317, 242 (1992).

[10] L. W. Fagg and S. S. Hanna, Rev. Mod. Phys. 31, 711 (1959).
56

[11] P. M. Jones, L. Wei, F. A. Beck, P. A. Butler, T. Byrski, G. Duchene, G. de


France, F. Hannachi, G. D. Jones and B. Kharraja, Nucl. Instrum. Methods
Phys. Res. A 362, 556 (1995).

[12] K. Starosta et aL, Nucl. Instrum. Methods Phys. Res. A 423, 16 (1999).

[13] R. Palit, H. C. Jain, P. K. Joshi, S. Nagaraj, B. V. T. Rao, S. N. Chintalapudi


• and S. S. Ghugre, Pramana 54, 347 (2000).

[14] E. Der Mateosian and A. W. Sunyar, At. Data. Nucl. Data. Tables 13, 391
(1974).

[15] T. Aoki, K. Furuno, Y. Tagishi, S. Ohya and J. Ruan, At. Data. Nucl. Data.
Tables 23, 349 (1979).

[16] R. P. Roy and B. P. Nigam, Nuclear Physics Theory and Experiment, New
Age International Private Limited, (2001).

[17] R. F. Casten, Nuclear Structure from a Simple Perspective, Oxford University


Press, Second Edition (2000).

[18] M. G. Mayer, Phys. Rev. 74, 235 (1948).

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

[20] M. G. Mayer, Phys. Rev. 78, 22 (1950).

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


(Wiley, New York 1985).

[22] W. Greiner and J. A. Maruhn, Nuclear Models, Springer-Verlag Berlin Hei­


delberg New York.

[23] T. Otsuka, R. Fujimoto, Y. Utsuno, B. A. Brown, M. Honma, T. Mizusaki,


Phys. Rev. Lett. 87, 082502 (2001).

[24] T. Otsuka, T. Suzuki, R. Fujimoto, H. Grawe, Y. Akaishi, Phys. Rev. Lett.


95, 232502 (2005).
57

[25] T. Otsuka, T. Suzuki, M. Honraa, Y. Utsuno, N. Tsunoda, K. Tsukiyama and


M. Hjorth-Jensen, Phys. Rev. Lett. 104, 012501 (2010).

[26] R. Kriicken, Cont. Phys. 52, 101 (2011).

[27] B. A. Brown and B. H. Wildenthal, Ann. Rev. Nucl. Part. Sci. 38, 29 (1988).

[28] B. A. Brown and W. A. Richter, Phys. Rev. C 74, 034315 (2006).

[29] T. T. S. Kuo and G. E. Brown, Nucl. Phys. A 114, 241 (1968).

[30] B. A. Brown and W. D. M. Rae, MSU-NSCL Report, (2007).

[31] B. A. Brown, Lecture Notes in Nuclear Structure Physics, November 2005.

[32] B. A. Brown, Prog. Part. Nucl. Phys. 47, 517 (2001).

[33] E. K. Warburton, J. A. Becker and B .A. Brown, Phys. Rev. C 41, 1147
(1990).

[34] B. H. Wildenthal, Prog. Part. Nucl. Phys. 11, 5 (1984).

[35] J. P. McGrory, Phys. Rev. C 8, 693 (1973).

[36] T. T. S. Kuo and G. E. Brown, Nucl. Phys. A 85, 40 (1967).

[37] D. J. Millener and D. Kurath, Nucl. Phys. A 255, 315 (1975).

You might also like