Download as pdf or txt
Download as pdf or txt
You are on page 1of 106

Republic of Iraq

Ministry of Higher Education


And Scientific Research
University of Baghdad
College of Science

Microscopic Effective Charges and the


E2 Strength B (E 2;21 01 ) for eveneven Carbon Isotopes
A Thesis
Submitted to Council of the College of Science
University of Baghdad in Partial Fulfillment of
The Requirements for the Degree of Master of Science
in Physics
By

Dheyaa Alwan AbdulHussain AL-Ibadi


B.Sc., University of Babylon (1998)

Prepared Under the Supervision of


Prof. Dr. Raad A. Radhi

2013

&

Prof. Dr. Zaheda A. Dakhil

1434

()

Certificate
We certify that preparation of this thesis, entitled Microscopic Effective
Charges and the E2 Strength B (E 2;21 01 ) for even-even Carbon
Isotopes was made under our supervision by Dheyaa Alwan
AbdulHussaen at the College of Science, University of Baghdad, in partial
fulfillment of the requirements for the degree of Master of Science in
Physics (Nuclear Physics).

Signature:

Signature:

Name: Dr. Raad A. Radhi

Name: Dr. Zaheda A. Dakhil

Title: Professor
Date:

Title: Professor

/ 2013

Date:

/ / 2013

In view of the available recommendations, I forward this thesis for


debate by the examining committee.

Signature:
Name: Dr. Raad M. S. Al-Haddad
Title: Professor
Head of Physics Department, College of Science
Date:

/ 2013

We certify that we have read this thesis, entitled " Microscopic Effective
Charges and the E2 Strength B (E 2;21 01 ) for even-even Carbon
Isotopes'' and as examining committee, examined the student Dheyaa
Alwan AbdulHussaen on its contents, and that in our opinion it is
adequate for the partial fulfillment of the requirements for the degree of
Master of Science in Physics (Physics Nuclear).

Signature:
Name: Dr. Abdul hussain A. Mahmood
Title: Professor
(Chairman)
Date: / / 2013

Signature:
Name: Dr. Fadhil I. Sharrad
Title: Assistant Professor
(Member)
Date: / / 2013

Signature:
Name: Dr. Raad A. Radhi
Title: Professor
(Supervisor)
Date: / / 2013

Signature:
Name: Dr. Gaith N. Flaiyh
Title: Assistant Professor
(Member)
Date: / / 2013

Signature:
Name: Dr. Zaheda A. Dakhil
Title: Professor
(Supervisor)
Date: / / 2013

Approved by the University Committee of Postgraduate Studies.

Signature:
Name: Dr. Saleh M. Ali
Title: Professor
Dean of the College of Sciences, University of Baghdad
Date: / / 2013

Dedication
To
My Wife

ENAS

Dheyaa

Acknowledgement
Praise be to Allah, Lord of the worlds, blessings and peace be upon his
messenger Mohammed and his household.

Firstly I would like to express my sincere and great appreciation to my


supervisors Prof. Dr. Raad A. Radhi and Prof. Dr. Zaheda A. Dakhil for
suggesting this project, appreciable directions, help and support throughout the
work.
I would like to thank the Dean of the college of science, and the head of
the department of physics for their support to perform this work.

I wish I have a word better than thanks to express my feelings to the


Ministry of Industrial represented by Minister Mr. AHMAD ALKURBULI,
and Al-Rasheed Co. Represented by the general director
Eng. JALAL H. HASSAN for their permission, help and support to get
this graduate degree.

I am grateful to the staff of the library of the college of Science, staff of


the central library of the University of Baghdad, and IVSL program for
providing references.

My thanks, gratitude and appreciation to all my friends and colleagues in


the M.Sc. program for their support and encouragement along the whole period
of the research work, especially to Dr. Fadhil I. Sharrad , Dr. Laith Ahmed
Najam and Dr. Fouad Al-Ajeeli to effective help for us to get some research
and sources used in this thesis.

Also special thanks for Dr. Arkan Refah, Mr. Ahmed, Mr. Bhaa, Mr.
Natheer, Mr. Malek, Mr.Moustfa and my closed friend and my colleagues in
this work Mr. Noory Sabah.

In this moment I can see your tears, I feel your happiness, I hear your
congratulations words, I am sad too much for losing you, thanks for all. To my
dear father, may God have mercy on you.

My deep and sincere thanks are to my dear mother for her support and
patience during my study.

Finally, to the light of my eyes Brothers and sisters the all faithful hearts
which help me.

Dheyaa

Contents
Contents
List of figures
List of tables
Abstract

I
III
V
VI

Chapter One: Introduction


1.1
1.2
1.3
1.4
1.5

Introduction
Exotic Nuclei
Puzzle of Halo Nuclei
Literature Survey
Aim of the Present Work

1
5
14
19
34

Chapter Two: Theoretical Considerations


2.1
2.2
2.3

35
36

2.4
2.5
2.6
2.7
2.8

Electron Scattering
General Theory
The Reduced SingleParticle Matrix Elements of the Longitudinal
Operator
The Many-Particles Matrix Elements
SingleParticle Matrix Element in SpinIsospin Formalism
The One-Body Density Matrix Elements (OBDM)
Corrections to the Form Factor
Electron Scattering Form Factor and Nucleon Density

2.9
2.10
2.11
2.12
2.13

Normalization of ( )
Root mean square radius in terms of occupation number
The Reduced Electromagnetic Transition Probability
The Relation between B (EJ) and B (EJ)
Core-polarization effects and effective charges

48
49
51
52
55

39
42
42
43
44
46

Chapter Three: Results, Discussion and conclusions


3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
3.9

Introduction
The nucleus 10C
The nucleus 12C
The nucleus 14C
The nucleus 16C
The nucleus 18C
The nucleus 20C
Conclusions
Future work
Reference

59
63
65
66
67
69
70
73
75
76

II

List of Figures
CHAPTER ONE
Figure
1.1

The chart of nuclides. The valley of stability is indicated by the


black dots representing the stable nuclei in nature. The limits of
nuclear stability are indicated by the proton and neutron drip
lines, behind which no bound nuclei can exist. The double lines
indicate the magic numbers for the stable nuclei.

1.2

A chart of the nuclei in the light isotope region (adapted from


[4]). Unbound systems, being potentially populated through
one-proton removal reactions, form the outskirts of the neutronrich landscape.

1.3

Shows closed quantum systems (bound state) at left and Open


quantum systems (unbound state, nuclei far from stability) at the
right

1.4

The bound nuclear states come close to the continuum.

1.5

a) Stable nuclei, proton and neutron homogenously mixed, no


decoupling of proton and neutron distributions. (b) Unstable
nuclei, decoupling of proton neutron distributions (neutron-rich
system).

1.6

The size and granularity for the most studied halo nucleus 11Li.
The matter distribution extends far out from the nucleus such
that the rms matter radius of 11Li is large as 48Ca, and the radius
of the halo neutrons large as for the outermost neutrons in 208Pb

1.7

The behavior of density distributions halo nuclei.

10

1.8

Schematic view of the nuclear density as a function of distance


with the definition of the half density radius and the surface
thickness.

12

1.9

(a) Schematic view of 11Li as 9Li core and two loosely bound
neutrons. (b) Visualization of the three-body system in 11Li.

13

1.10 The Borromean structure in 11Li.

III

13

1.11 Shows the dependence of density distributions of a loosely


single proton and neutron (in a Woods-Saxon-type potential) on
binding energy (EB), orbital momentum quantum numbers (),
and charge. The effect of the Coulomb and the centrifugal
potential are clear. The figure is taken from.

IV

15

List of Tables
CHAPTER THREE
Table
3.1 The calculated root mean square matter radii of 1020C compared
with the experimental data.

63

3.2 The calculated effective charges and B(E2) values of 10C compared
with the experimental data.

64

3.3 The calculated effective charges and B(E2) values of 12C compared
with the experimental results.

65

3.4 The calculated effective charges and B(E2) values of 14C compared
with the experimental data.

67

3.5 The calculated effective charges and B(E2) values of 16C compared
with the experimental data.

68

3.6 The calculated effective charges and B(E2) values of 18C compared
with the experimental data.

70

3.7 The calculated effective charges and B(E2) values of 20C compared
with the experimental data.

71

3.8 The calculated B(E2) values using average effective charges


e peff 1.1e and e neff 0.27e , using psd MS for A=10,14 and spsdpf
(0+2) MS for A=16,18 and 20, compared with the experimental
data.

72

Abstract
Quadrupole transitions and effective charges are calculated for even-even
Carbon isotopes (10 A 20) based on shell model with p and psd shell model
spaces for 10 A 14 and with large basis no core shell model with (0 2)
truncations for 16 A 20. Calculations with configuration mixing shell
model with these limited model spaces usually under estimate the measured E2
transition strength. Although the consideration of a large basis no core shell
model with (0 2) truncations for 16 A 20 where all major shells s, p,
sd and pf are used, calculations fail to describe the measured reduced transition
strength B( E 2;21 01 ) without normalizing the matrix elements with effective
charges to compensate for the discarded space. Instead of using constant
effective charges, excitations out of major shell space are taken into account
through a microscopic theory which allows particlehole excitations from the
core and model space orbits to all higher orbits with 2 excitations, which are
called core-polarization effects. The two body Michigan sum of three ranges
Yukawa potential (M3Y) is used for the core-polarization matrix elements. The
simple harmonic oscillator potential is used to generate the single particle
matrix elements of all isotopes considered in this work. Two values of the size
parameter b are used, one for A=10-14 (group A) and the other one for A=1620 (group B), due to the difference in the root mean square (rms) matter radii
(Rm) between the two isotopes groups. The b value of each isotope is adjusted
to reproduce the experimental matter radius. The average value of b for group
A is 1.55 fm, while that of group B is 1.78 fm. These size parameters of the
harmonic oscillator almost reproduce all the rms matter radii for
isotopes within the experimental errors.

VI

10,12,14,16,18,20

Almost same effective charges are obtained for the neutron-rich C isotopes
which are smaller than the standard values. The calculated B(E2) values agree
very well with the experimentally observed trends of the recent experimental
data for the entire chain of even-even C isotopes.
The major contribution to the transition strength comes from the corepolarization effects. The present calculations of the neutron-rich
20

16

C,

18

C and

C isotopes show a deviation from the general trends of even-even nuclei in

accordance with experimental and other theoretical studies. The experimental

B( E 2;21 01 ) values are very well reproduced confirming the anomalous


suppression in

16

C and

18

C. The configurations arises from the shell model

calculations with core-polarization effects which reproduce the experimental


B(E2) values, and give small effective charges confirm the formation of protonshell closure for
14,16,18,20

14,16,18

C. The average microscopic effective charges for

C are 1.1e and 0.25e, for the proton and neutron, respectively which are

smaller than the standard empirical values for this mass region.

VII

CHAPTER ONE
Introduction

CHAPTER ONE

INTRODUCTION

CHAPTER ONE
1.1. Introduction
Many years ago beyond the exotic nuclei phenomena had been discovered,
this discovery is represented a split moment in the history of knowledge; it
was leading to a new era in the structure nucleus. In fact, this phenomenon
had been taken the attention of many physicists over the world with their
imaginations and efforts. This aid approached to understand the structures
and interpretations behavior of these nuclei. There have achieved in both
theoretical and experimental perspectives.
In the beginning, the few concepts should be understood due to exotic
nuclei. An atomic nucleus is a many-body Fermionic quantum system made
of neutrons and protons (called nucleons). Nuclear systems are extremely
small, with typical radii on the order of 10 -14 meters. Protons have a positive
electric charge and interact with one another through the Coulomb force,
which repels protons from one another and decreases in strength with 1/r2,
where r is the radial coordinate. Neutrons have no electric charge, and since
nuclear systems exist, there must be attractive forces stronger than the
Coulomb force at play on the length scale of nuclear existence. The
interaction between the protons and neutrons in the nucleus is called the
strong nuclear force: it is about 100 times stronger than the Coulomb force
on the length scale of a nucleus but is negligible for longer distances [1].
The properties of matter are determined by the number of protons and
neutrons in a nucleus. In particular, the number Z of protons characterizes
the different elements, from hydrogen to uranium; depending on the number
N of neutrons, each element can be present in nature in a variety of isotopes.
The stable elements that exist in nature are 92 and there are almost 300

CHAPTER ONE

INTRODUCTION

stable isotopes. When displayed in the chart of the nuclides, see figure 1.1,
where the number of protons Z is plotted against the number of neutrons N,
the stable nuclei lie approximately along the diagonal from the lower left to
the upper right, called the valley of stability; this figure also is called
"nuclear landscape".
In nuclear landscape, one can classify it into a major three areas, the first
one is indicated by black dots represented the stable nuclei: i.e., infinite
lifetime, stable nuclei can be found around the so-called stability line, where
NZ. The second one is neutrons drip line (at the bottom of stable line that
shows in figure. 1.1, red line). The nucleus, in this line, has being known the
neutrons-rich. The last area, in figure 1.1, is proton drip line (above the
stable line that shows as blue line in the figure). The nucleus, in this line, has
been known the protons-rich. Figure 1.2 is a similar chart for nuclear
systems (light nuclides which have been concentrated in the present work).
In figure 1.1, the unstable area indicated in green color is called the Terra
Incognita, meaning in Latin unexplored land. At the upper right of
landscape, superheavy nuclei had been appeared, this discrimination
between neutrons-rich and protons-rich depending on N/Z ratio [2, 3]. The
professional scientists in this field called the area that lies beyond neutronrich and proton-rich edge of stability, where nuclei after edge of stability
become unbound. Figure 1.3 shows the nuclei in stability (bound state) and
at far from stability i.e. unbound state [1]. So the separation energy for
nucleons in this edge goes to zero.
Besides the search for the exact position of the drip lines in the landscape,
other motivations to investigate nuclei far from stability are the quests for
which combination of protons and neutrons can make up a nucleus.

CHAPTER ONE

INTRODUCTION

Figure 1.1: The chart of nuclides. The valley of stability is indicated by the black dots
representing the stable nuclei in nature. The limits of nuclear stability are
indicated by the proton and neutron drip lines, behind which no bound nuclei
can exist. The double lines indicate the magic numbers for the stable nuclei.

These unstable isotopes lie away from the line of stability, in the region
within the neutron and proton drip lines, which defined previously, the limits
beyond which nuclei become unstable. Due to their distance from the valley
of stability, nuclei close to the drip lines are often referred to as "exotic
nuclei", indicating entities different from the most ordinary ones, available
in nature. The drip lines have experimentally been reached only for the
lightest 8 elements and their position in the nuclear chart is not exactly
known yet. Therefore, in order to test existing nuclear structure models, one
of the most important topics of research in nuclear physics is the exploration
of the nuclear chart to find out where the limits of nuclear binding energy.

CHAPTER ONE

INTRODUCTION

One of the most remarkable results of these studies was the discovery of
novel nuclear structures in nuclei far from stability in the last decades of the
20th century. In analogy with the luminous ring around the sun or the moon
seen under certain meteorological conditions, there are a type of nuclei was
referred to as "Halo Nuclei". Thus, the investigation of the properties of
exotic nuclei becomes one of the most important goals in nuclear physics
and it is now possible thanks by the development of modern technologies
available for radioactive beam production and heavy-ion accelerators [1].

Figure 1.2: A chart of the nuclei in the light isotope region (adapted from [4]). Unbound
systems, being potentially populated through one-proton removal reactions,
form the outskirts of the neutron-rich landscape.

CHAPTER ONE

INTRODUCTION

Figure 1.3: Shows closed quantum systems (bound state) at left and open quantum
systems (unbound state, nuclei far from stability) at the right.

1.2. Exotic Nuclei


In the beginning, what is an exotic nucleus? The word 'exotic' is referred to
anything which is out of ordinary and attract a wide interest to realize it.
Theoretically, nuclei are finite many-body quantum systems which exhibit a
rich variety of single-particle, few-body, and many-body phenomena. In
nuclear physics, the exotic nuclei are phenomena for some light nuclei and it
has special conditions far from stability line or near drip-lines (neutronrich or proton-rich). In other words, exotic nuclei are nuclei with an
extraordinary ratio of protons and neutrons Z/N. Typically; they are very
unstable and decay into more stable nuclei, i.e. exotic nuclei so-called rareisotopes, which have a loosely binding energy, short-lived isotopes and large
isospin. Actually, exotic nuclei are available as secondary beams at many
radioactive beam facilities around the world [5]. These unstable nuclei are
generally weakly bound with few excited states. These exotic nuclei have a
thin cloud of nucleons orbiting at large distances from the others, forming
the core [5].

CHAPTER ONE

INTRODUCTION

Neutron-rich nuclei have attracted much interest during the past decades
[6- 11], and this will continue to be so due to new generation radioactive
beam facilities in the world. These nuclei are characterized by a small
binding energy, and many new features originating from the weakly bound
property have been found, a halo and skin structures with a large spatial
extension of the density distribution [12], a narrow momentum distribution
[13].
Halo states, when approaching the drip lines the separation energy of the
last nucleon or pair of nucleons decreases gradually and the bound nuclear
states come close to the continuum, see figure 1.4. The combination of the
short range of the nuclear force and the low separation energy of the valence
nucleons results, in some cases, in considerable tunneling into the classical
forbidden region and a more or less pronounced halo may be formed. As a
result the spatial structure of the valence nucleons is very different from the
rest of the system and the valence and the core subsystems are to a large
extent separable [14]. Therefore, halo nuclei may be viewed as an inert core
surrounded by a low density halo of valence nucleon(s). They may therefore
be described in few-body or cluster models [10, 15]. The formation of halo
states is characteristic especially for light nuclei in the drip line regions,
although not all of these can form a halo. There is a large sensitivity of the
spatial structure and the separation energy close to the threshold. The
increase in size, which is due to quantum mechanical tunneling out from the
nuclear volume, will only take place if there is no significant Coulomb or
centrifugal barriers present. There are at present many well-established halo
states for light neutron-rich drip line nuclei consisting of a core plus one or
two neutrons.

CHAPTER ONE

INTRODUCTION

Figure 1.4: The bound nuclear states come close to the continuum, where (proton) at left and
(neutron) at right.

Nuclei at the drip lines are led with nucleons up to the limit set by the
combination of the nuclear kinetic energy and the depth of the nuclear
potential well, resulting in a bound state close to the continuum. The binding
energy for this state is very small and, due to the short range of the nuclear
force, threshold effects appear in figure 1.5, (see also figure 1.3). The last
nucleons undergo the quantum-mechanical effect of tunneling, so that their
probability of being at large distances from the core is appreciable. This
referred to as 'halo' state for the first time in 1987 [8] and since then the term
halo has been referred to all exotic nuclei manifesting those properties. The
manifestation of the halo phenomenon is less evident if the halo nucleons are
in a state of large angular momentum >1, in which case the centrifugal
barrier lowers the probability of tunneling far from the core. Analogously,
the Coulomb barrier hinders the formation of proton halos, because the
repulsion between the protons and the nuclear core makes it difficult for the
nucleons to tunnel. The Coulomb repulsion also affects the absolute binding,
so that the proton drip line actually lies closer to the valley of stability than
the neutron drip line.

CHAPTER ONE

INTRODUCTION

(a)

(b)

Figuer1.5: (a) Stable nuclei, proton and neutron homogenously mixed, no decoupling of
proton and neutron distributions. (b) Unstable nuclei, decoupling of proton
neutron distributions (neutron-rich system).

A halo state consists of a veil of dilute nuclear matter that surrounds the
core. This is in contrast to the nuclear skin [16], which is essentially
difference in the radial extent of the proton and neutron distributions. A
loose definition of a halo would be that the halo nucleon(s) spend about 50%
of the time outside the range of the core potential and thus in the classically
forbidden region. The necessary conditions for the formation of a halo have
been investigated [17- 19] and it was found that, besides the condition of a
small binding energy for the valence particle(s), only states with small
relative angular momentum may form halo states. Two-body halos can thus
only occur for nucleons in s- or p-states.
The advent of rare-isotope beams has made it possible to explore how the
properties of nuclei evolve along a chain of isotopes extending to both
proton-rich and neutron-rich nuclei. This capability had been confirmed the
discovery of halo systems. Figure 1.6 gives a schematic illustration of the
sizes involved in the case of the two-neutron halo nucleus 11Li. The binding
energy for the two halo neutrons is only about 300 keV [20] and they are
mainly in s- and p-states and can therefore tunnel far out from the core. It

CHAPTER ONE

INTRODUCTION

turns out that the root mean square (rms) of 11Li is similar to the radius of
48

Ca while the two halo neutrons extend to a volume similar in size to 208Pb.

In addition to weak binding, the orbital occupied by the halo neutrons in Li


also contribute to its large size. In particular, the low angular momentum 2s
orbital intrudes into the p-shell. Such a reordering of orbital also leads to a
breakdown of the N = 8 shell gaps for Li.The neutron density distribution
in such loosely bound nuclei shows an extremely long tail, called the neutron
halo, see figure 1.7 [21].

Figure 1.6: The size and granularity for the most studied halo nucleus

11

Li. The matter

distribution extends far out from the nucleus such that the rms matter radius
of

11

Li is large as

48

Ca, and the radius of the halo neutrons large as for the

outermost neutrons in 208Pb.

CHAPTER ONE

INTRODUCTION

Figure 1.7: The behavior of density distributions halo nuclei.

Due to the confining Coulomb barrier which holds them closer to the core,
the formation of proton halo is rather difficult. Decoupling of core and
valence particles and their small separation energy are the important
criterion for a halo and in addition to these there is one another criterion that
the valence particle must be in a relatively low orbit angular momentum
state, preferable an s-wave, relative to the core, since higher l-values give
rise to a confining centrifugal barrier. The confining Coulomb barrier is the
reason for proton halos not so extended as neutron halos [22].
As a consequence the radius of such a nucleus is much larger than
expected. Although the density of a halo is very low, it strongly affects the
reaction cross section and leads to new properties in such nuclei, e.g. a very
narrow momentum distribution. This reflects the Heisenberg uncertainty
principle, when the distribution in coordinate space is wide, that in
momentum space is narrow.
x p x

10

(1-1)

CHAPTER ONE

INTRODUCTION

In other word, the matter radius is much extended; there is a large


uncertainty in the position and thus a small uncertainty in the linear
momentum [22].
In the halo nucleus the three basic rules of nuclear density for nuclei close
to the valley of stability are broken [23]:
The first rule which is; the radius where the nuclear density is reduced to
half of the maximal density is expressed as R r0 A 1 3 , r0 1.2 fm is the
radius constant. The second one is protons and neutrons are homogenously
mixed in the nucleus, i.e. p (r ) n (r ) , in halo nuclei this is only true for
the core. Finally, the distance from where the nuclear density drops from
90% of the maximal value to 10% of its maximal value, called the surface
thickness t, is constant and is about t 2.3 fm, see figure 1.8. This feature is
true for nuclei in the valley of stability because of the nearly-constant
nucleon separation energy (6-8MeV) for stable nuclei. In general the surface
thickness or the surface diffuseness is expected to depend on the nucleon
separation energy. The neutron halo is the most pronounced case for small
separation energy, less than 1MeV [24].
Very heavier neutron-rich nuclei, rather than forming halos may instead
form neutron skin. The extension of the neutron skin can be expected to
have important influences on fusion cross sections because of the
localization of neutron matter at the surface of the nucleus. A complete
characterization of the neutron skin will require the knowledge of the singleparticle neutron states occupied in these nuclei in addition to their decay
properties and reaction rates. Further, the establishment of the thickness of
the skin determined by examining the difference between the nuclear matter

11

CHAPTER ONE

INTRODUCTION

radius and the charge radius is a crucial quantity for testing models of
nuclear matter [23].

Figure 1.8: Schematic view of the nuclear density as a function of distance with the
definition of the half density radius R and the surface thickness t.

The most prominent and most studied example is 11Li with a 9Li core and
two loosely bound neutrons see figure 1.9 early in 1975 Thibault et al. [24]
found very small two-neutron separation energy, but they did not attribute
this to a neutron halo. In 1985 Tanihata et al. [12] discovered the large
interaction cross section of 11Li and attributed it to a halo property.
In the fragmentation of

11

Li, narrow momentum distributions were found

for the 9Li fragments, but not for the other fragments such as 8Li or 8He,
indicating that only the two neutrons in the last orbital contribute to the
formation of a neutron halo [23]. 11Li was found to be a three-body system,
in which none of the internal two-body subsystems (dineutron and 10Li) were
bound, giving it the name "Borromean halo nucleus" (This name is given
by the analogy of the three subsystems to the three rings of the coat of arms

12

CHAPTER ONE

INTRODUCTION

of the Italian Borromean family) [8,10]. Figure 1.9 and figure 1.10 showed
three-body interactions are necessary for a full description of the nucleus.

Figure 1.9: (a) Schematic view of 11Li as 9Li core and two loosely bound neutrons.
(b) Visualization of the three-body system in 11Li.

Figure 1.10: The Borromean structure in 11Li.

13

CHAPTER ONE

INTRODUCTION

The obvious difference between a proton and a neutron is the Coulomb


interaction. A proton sees the Coulomb barrier at the surface of the nucleus.
Although the asymptotic form of the wave function may be the same, the
amplitude is damped because of the barrier. The height of the Coulomb
barrier exceeds 1 MeV, even if Z (atomic number) is as small as 5 (Boron),
if a normal Woods-Saxon-type density distribution is assumed. The other
type of the barrier that affects the wave function is the centrifugal barrier.
Since the centrifugal potential is proportional to (+1)/r2, its heights
depends heavily on the orbital angular momentum of the halo nucleon.
Figure 1.11 shows the density distributions of a loosely-bound single proton
and a single neutron in Woods-Saxon-type potential. Distributions are
shown separately for different orbitals: 2s, 1p and 1d. The effect of the
Coulomb interaction and the centrifugal potential is clearly visible. It is seen,
for example, for 2s, in which no centrifugal potential exists. The proton
density does not extend as much as that of neutrons. If one now compares
the neutron density tail for different , the tail is shorter for higher orbital
[23].

1.3. Puzzle of halo nuclei


In reference to halo stricter, the first hint came from the measurement of
the electric dipole transition between the bound states in

11

Be. Firstly, a

simple shell model picture of the structure of 11Be suggested that its ground
state should consist of a single valence neutron occupying the 1p1/2 orbital
(the other six having filled the 1s1/2 and 1p3/2 orbital). But it was found that
the 2s1/2 orbital drops down below the 1p1/2 and this 'intruder' state is the
one occupied by the neutron, making it a (1/2)

14

ground state. The first

CHAPTER ONE

excited state of

INTRODUCTION
11

Be, and the only other particle bound state, is the (1/2) -

state achieved when the valence neutron occupies the higher 1p1/2 orbital.
The very short lifetime for transition between these two bound states was
measured in 1983 [25] and corresponded to an E1 strength of 0.36 W.u
(Weisskopf Unit). It was found that this large strength could only be
understood if it is realistic single particle wave function is used to describe
the valence neutron in two states, which extended out to large distances due
to the weak binding.

Figure 1-11: The dependence of density distributions of a loosely single proton and
neutron (in a Woods-Saxon-type potential) on binding energy (EB), orbital momentum
quantum numbers (), and charge. The effect of the Coulomb and the centrifugal
potential are clear. The figure is taken from [23].

15

CHAPTER ONE

INTRODUCTION

In the mid-1980's, the Berkeley experiment was carried out by Tanihata


and his group [12]. The interaction cross section of Helium (He) and
Lithium (Li) isotopes were measured and they found that the value was
much larger than the expected for the case of 6He and

11

Li. These

corresponded to larger rms (matter density) than that is predicted by the


normal A1/3 dependence. After two years, Hansen and Jonson [8] proposed
that the large size of these nuclei is due to the halo effect. They explained the
large matter radius of 11Li by treating it as a binary system of 9Li core plus a
di-neutron. Dineutron is a hypothetical point particle, implying the two
neutrons are stuck together i.e. the n-n system is unbound. Hence they
showed that the weak binding between this pair of clusters could form an
extended halo density.
High-energy proton elastic scattering is considered to be the best method
to obtain matter density distributions of nuclei. Proton elastic scattering
cross sections on various halo nuclei were measured at GSI Darmstadt using
liquid hydrogen target system [26, 27]. The matter density distributions had
been determined for

6,8

He and for

8,9,11

Li. The rms matter radii Rm = 2.45

0.10 fm for 6He and Rm = 2.53 0.08 fm for 8He were consistent with the
values obtained by the interaction cross sections [28].
The first value that had been measured on exotic nuclei using secondary
radioactive beam was the interaction cross section (1 ) [29, 30]. It defined
as the total cross section for all processes in which projectile number of
nucleons is changed. From the interaction cross sections, one can define, the
interaction radius [29] using a simple geometric model:

1 (P ,T ) (R I (P ) R I (T ))2

16

(1-2)

CHAPTER ONE

INTRODUCTION

where RI (P) is radius of projectile and RI (T) is radius of target. It has been
shown that the interaction radius is more or less independent of the target.
Measuring the interaction cross section for different target one can obtain the
interaction radius of a projectile. In nuclear theory, it is known that stable
nucleus density is rather constant up to a certain radius from which it drops
to zero. The central density is quite similar from the lightest nuclei to the
heaviest. This led to the semiclassical liquid-drop model in which nuclei are
viewed as liquid droplet with a homogeneous density. In this model a
nucleus containing A nucleons is therefore seen as a sphere with a radius R
proportional to A1/3 [29].

R ro A

1
3

with ro 1.2 fm

(1-3)

If one assumes that the interaction radius is somehow a measure of the


nuclear size, it should follow the A1/3 law predicted from the liquid-drop
model. Some nuclei near the neutron drip line (like 6He, 8He, 11Li, 11Be and
14

Be) exhibit anomalously high interaction radii in comparison with their

neighbors. This indicates that those nuclei have large nuclei radii due to an
extended matter density and /or a major deformation [30].
In the last two decades, development of experiments with a secondary
radioactive ion beam has extensively helped the studies on light neutron-rich
nuclei. One of the most striking features in neutron-rich nuclei is the nuclear
deformation. The deformation can be investigated experimentally and
theoretically, through their electromagnetic transitions. The general trend of
the 2+ excitation energy E (21 ) and the reduced electric quadrupole transition
strength between the first excited 2+ state and the 0+ ground state,

B (E 2; 21 01 ) for even-even nuclei are expected to be inversely proportional

to one another [31]. However, recent experimental and theoretical studies of

17

CHAPTER ONE

INTRODUCTION

the neutron-rich 16C, 18C and 20C isotopes [32-35] show a deviation from the
general trends of even-even nuclei. A systematic study of transition rates for
these isotopes was recently conducted both theoretically and experimentally
[35-41].
The structure of light neutron-rich nuclei can be understood within the
shell model. Shell model within a restricted 1p model space was found to
provide good description for the

10

B natural parity level spectrum and

transverse form factors [42]. However, they were less successful for E2 form
factors and gave just 45% of the total observed E2 transition strength.
Expanding the model space to include 2 configurations in describing the
form factors, Cichocki et al. [42] had found that only 10% improvement was
realized. It had been long recognized [43] that these transitions have highly
collective properties. Those collective properties could be supplemented to
the usual shell model treatment by allowing excitations from the core and
model space orbits into higher orbits. The conventional approach to supply
that added ingredient to shell model wave functions is to redefine the
properties of the valence nucleons from those exhibited by actual nucleons
in free space to model-effective values [44]. Effective charges were
introduced for evaluating E2 transitions in shell-model studies to take into
account effects of model-space truncation. A systematic analysis had been
made for observed B(E2) values with shell-model wave functions using a
least-squares fit with two free parameters gave proton and neutron effective
charges, e peff 1.3 e

and e neff 0.5 e

[45] in sd-shell nuclei.

The role of the core and the truncated space can be taken into
consideration through a microscopic theory, which allows one particleone
hole (1p1h) excitations of the core and also of the model space to describe

18

CHAPTER ONE

INTRODUCTION

these E2 excitations. These effects provide a more practical alternative for


calculating nuclear collectivity. These effects are essential in describing
transitions involving collective modes such as E2 transition between states
in the ground-state rotational band, such as in

18

O [46]. Umeya [39, 47]

calculated effective charges for quadrupole moments and transitions by


using first order perturbation. These theoretical results show that the
effective charges are smaller than the standard value in light- neutron rich
nuclei and imply decoupled quadrupole motions between protons and
neutrons in neutron-rich nuclei.

1.4. Literature Survey


Different kinds of theoretical models are currently being used to
investigate the structure of exotic nuclei. One of the early works was an
intermediate between shell model calculations and fully microscopic ones, is
the so called cluster-orbital shell model [48- 50].
When shell model calculations are considered, single particle wave
functions are provided from a certain mean field potential assuming stable
single particle motion. The shell model calculation is then carried out to treat
the residual interaction between nucleons moving on such stable single
particle orbits. In neutron rich nuclei where the last neutrons are forced to
occupy loosely bound or even unbound orbits of the mean potential, the
validity of shell model calculations can be questioned.
Tanihata et al. (1985) [12] calculated the interaction cross sections (I) for
all known Li isotopes 6Li-11Li and 7Be, 9Be, and 10Be on targets Be, C, and
Al at 790 MeV/nucleon. Root mean square radii of these isotopes as well as
He isotopes had been deduced from the I by a Glauber-type calculation.
Appreciable differences of radii among isobars ( 6He-6Li, 8He-8Li, and 9Li19

CHAPTER ONE
9

INTRODUCTION

Be) had been observed for the first time. The nucleus

11

Li showed a

remarkably large radius suggesting a large deformation or a long tail in the


matter distribution.
Tanihata et al. (1988) [51] measured 8B's interaction cross section. The
experiment found an interaction cross section very similar to those of the
neighboring Boron isotopes, and the deduced interaction radius followed the
A1/3 dependence typically seen in 'normal' nuclei. Those results for the 8B
nucleus were in contrast to measure interaction radii of nuclei found to have
neutron halos (which showed much larger than expected interaction radii).
That early result caused many to dismiss the possibility that 8B might
contain a proton halo.
Raman et al. [52] in 1988, completed a compilation of experimental results
for the electric quadrupole transition probability B(E2) between the 0 +
ground state and the first excited 2+ state in even-even nuclei. For nuclei
away from closed shells, the SU(3) limit of the intermediate boson
approximation implies that the B(E2) values were proportional to
(epNp+enNn)2, where ep (en) is the proton (neutron) effective charge and Np
(Nn) refers to the number of valence protons (neutrons). The proportionality
was consistent with the observed behavior of B(E2) vs NpNn. For
deformed nuclei and the actinides, the B(E2) values calculated in a
schematic single-particle "SU(3)" simulation or large single-j simulation of
major shells successfully reproduced not only the empirical variation of the
B(E2) values but also the observed saturation of these values when plotted
against NpNn.
Minamisono et al. (1992) [53] measured the quadrupole moment of 8B and
used it as a probe of the structure. The quadrupole moment was found to be
|Q (8B) |= 68.3 2.1 mb, which was anomalously large, compared to shell
20

CHAPTER ONE

INTRODUCTION

model predictions. The large value of the quadrupole moment was


interpreted by the authors as evidence for a proton halo within the 8B
nucleus.
Riisager et al. (1992) [17] calculated the radius of the two-body system for
different separation energies. The radius of the system became larger and
larger as separation energy decreased.
Otsuka et al. (1993) [54] proposed a variation shell model in order to
describe the structure of such nuclei. The model was applied to 11Be, where
by using a Skyrme interaction the observed ground state of 11Be nucleus was
reproduced correctly. In general mean field approximations proved to be of
restricted validity because of the weak binding of the halo neutrons. It was
realized that a more realistic approach to the halo structure would rely on
microscopic many-body model.
In 1994, Nakada and Otsuka [55] studied the E2 properties of A = 6-10
nuclei, including those of nuclei far from stability, which were studied by a
(0 2) shell-model calculation which includes E2 core-polarization

effects explicitly. The quadrupole moments and the E2 transition strengths in


A = 6-10 nuclei were described quite well by the present calculation. In that
paper, the result indicated that the relatively large value of the quadrupole
moment of 8B could be understood without introducing the proton halo in
8

B. An interesting effect of the 2 core polarization was found for

effective charges used in the

0 shell model; although isoscalar effective

charges are almost constant as a function of nucleus, appreciable variations


are needed for isovector effective charges which play important roles in
nuclei with high isospin values.
Guimaraes et al. (1995) [56] had simultaneously measured the
longitudinal momentum distribution of
21

18

C,

17

C and

16

C following the

CHAPTER ONE

breakup of

19

INTRODUCTION

C, 18C and 17C. In the 19C data both the observation of narrow

momentum peak and a large cross section indicated the presence a one
neutron halo. The conclusion coincided with a shell-model calculation which
predicts an s state for the 19C ground state. The width observed for the

18

one neutron breakup had been interpreted in the framework of fragmentation


model and suggested an increase of the absorptive radius cutoff possible due
to a higher neutron density on the nuclear surface. The 17C data also showed
a narrow momentum peak, but it was wider and therefore incompatible with
the width deduced from the binding energy in the assumption of an s-state
neutron halo.
Experiments studying the momentum distributions of the breakup products
of 8B were also performed in order to study its structure. The breakup of 8B
into 7Be and a proton was studied, and the momentum distribution of the 7Be
products was measured. The first such experiment was done by Schwab et
al. (1995) [57] and the measured momentum distribution were found to be
very narrow when compared to those obtained from the breakup of
16

12

C and

O. The deduced rms radius for the valence proton in 8B of 6.83 fm was

much larger than the expected for a normally bound proton, and was thought
to be a clear sign of a proton halo.
Bertulani and Sagawa (1995) [58] investigated the use of elastic and
inelastic scatterings with secondary beams of radioactive nuclei as a mean to
obtain information on ground state properties and transition matrix elements
to continuum states. In particular they discussed possible signatures of halo
wavefunctions in elastic and inelastic scattering experiments. The eikonal
approximation together with the t approximation (the double folding
approximation) yielded the simple and transparent formula. The model gave
very reasonable results for elastic scattering cross sections at forward angles
22

CHAPTER ONE

INTRODUCTION

which could be used to test the ground state densities of radioactive nuclei.
The extended nuclear matter in exotic nuclei was manifest in the magnitude
of the elastic cross sections as well as in the position of the first minimum.
Pecina et al. (1995) [59] studied of quasi-elastic scattering of 8B from 12C
was performed with the goal of studying the structure of 8B. The measured
scattering cross sections were fitted using Optical Model potentials, and the
matter distribution determined. The deduced rms matter radius of 2.207 fm
for the 8B nucleus was small, and the authors interpreted this as evidence
against a substantial proton halo in 8B.
The studied of the momentum distributions of products from the breakup
of 8B was performed by Kelley et al. (1996) [60]. In that experiment, the
measured momentum distribution of the 7Be fragments was found to be
similar to that obtained by Schwab et al. (1995) [57] but after accounting for
absorption by the 7Be core, they deduced a rms radius for the valence proton
in 8B of 4.24 fm. Though the radius is larger than the systematic behavior of
nuclear radii (r0A1/3 =2.5 fm), it was much smaller than the rms halo radius
observed for the neutron halo nucleus

11

Li of 6.2 fm (1992) [61]. Those

results were interpreted as evidence against a large proton halo in 8B.


The first excited state of 17Ne had been discussed by Chromik et al. (1997)
[62] were it populated via relativistic Coulomb excitation with a radioactive
beam of

17

Ne on a

observed. The

3
2

197

Au target and the subsequent -ray decay had been

state was bound with respect to proton emission but

unbound to two-proton decay. The measured -ray yield accounts for


43+19
14 % of the predicted yield from an excitation cross section of 28 mb. It
was unlikely that the missing cross section could be attributed to two-proton

23

CHAPTER ONE

INTRODUCTION

emission because the lifetime of this branch would have to be a factor 1700
smaller than predicted by standard barrier penetration calculations.
Kolata and Bazin in (2000) [63] had measured the longitudinal momentum
of the 13B core fragment in one-neutron knockout from 14B, on both 9Be and
197

Au targets. The results of the experiment lend very strong support to the

idea that 14B is a neutron-halo system, the first odd-odd nucleus to display
this structure. It also appeared that, with the sole exception of 17C at N=11,
all of the lowest-mass, particle-stable isotones from N=7-13 were halo
nuclei.
Zheng et al. [64] (2002) studied the reaction cross sections for

12,16

C had

been measured at the energy of 83 MeV by a new experimental method. The


larger enhancement of the 16C reaction cross section at the low energy had
been used to study the density distribution of 16C. The finite-range Glaubermodel calculations for different density distributions had been compared
with the experimental data. A large extension of the neutron density
distribution to a distance far from the center of the nucleus suggested the
formation of neutron halo in the 16C nucleus.
Bazin et al. (2003) [65] used a new direct reaction: two-proton knockout
from neutron-rich nuclei, and they showed that a neutron-rich projectile
reacted with a light nuclear target, the knockout of two protons occurs as a
direct reaction. Consequently, the observed partial cross sections to
individual final levels conveyed selective information about nuclear
structure.
Zerguerras et al. (2004) [66] studied light proton-rich bound and unbound
nuclei by complete kinematics measurements which were produced by
means of stripping reactions of secondary beams of

20

Mg and

18

Ne. Their

work was the first measurement of its decay. As the decay scheme of the
24

CHAPTER ONE

INTRODUCTION

nucleus could not be determined, two possible scenarios were proposed and
discussed. In addition, the decay of excited states in

17

Ne via two-proton

emission was observed.


In 2004, Imai et al. [32] presented a study of the electric quadrupole
transition from the first excited 2+ state to the ground 0+ state in

16

C was

studied through measurement of the lifetime by a recoil shadow method


applied to inelastically scattered radioactive 16C nuclei. The measured mean
lifetime was 77 14(stat) 19 (sys.) ps. The central value of mean lifetime
corresponded to a B(E2; 21+ 0+) value of 0.63 e2fm4, or 0.26 Weisskopf
units. The transition strength was found to be anomalously small compared
to the empirically predicted value.
Sagawa et al. [36] in (2004) investigated static and dynamic quadrupole
moments of C and Ne isotopes by using the deformed Skyrme Hartree-Fock
model and also shell model wave functions with isospin-dependent core
polarization charges. It was shown at the same time that the quadrupole
moments Q and the magnetic moments of the odd C and odd Ne isotopes
depended clearly on assigned configurations, and their experimental data
will be useful to determine the deformations of the ground states of nuclei
near the neutron drip line. Electric quadrupole (E2) transitions in even C and
Ne isotopes were also studied using the polarization charges obtained by the
particle vibration coupling model with shell model wave functions.
Although the observed isotope dependence of the E2 transition strength was
reproduced properly in both C and Ne isotopes, the calculated strength
overestimates an extremely small observed value in 16C.
Stanoiu et al. (2004) [67] studied the drip line nuclei through two-step
fragmentation and concluded that the neutron-rich

1720

C and

2224

O nuclei

had been performed by the in-beam -ray spectroscopy using the


25

CHAPTER ONE

INTRODUCTION

fragmentation reactions of radioactive beams. The 2+ energy of

20

C was

determined. Its low-energy value hinted for a major structural change at


N = 14 between C and O nuclei. Evidence for the non-existence of bound
excited states in either of the

23,24

O nuclei had been provided, pointing to a

large subshell effect at N = 16 in the O chain.


The electron scattering was studied on halo nuclei by Bertulani (2005)
[68], he was using the inelastic scattering of electrons on weakly-bound
nuclei to study with a simple model based on the long range behavior of the
bound state wavefunctions and on the effective-range expansion for the
continuum wavefunctions. It was shown that the cross sections for electrodissociation of weakly-bound nuclei reach ten milibarns for 10 MeV
electrons and increase logarithmically at higher energies.
Campbell et al. (2006) [69] established the measurement of excited states
in 40Si and evidence for weakening of the N = 28 shell gap by detecting -rays
coincident with inelastic scattering and nucleon removal reactions on a
liquid hydrogen target. The large proton sub-shell gap at Z = 14 at means
that the evolution of excitation energies in the silicon isotopes was directly
related to the narrowing of the N = 28 shell gap.
Taqi and Radhi (2007) [70] had studied longitudinal form factors of the
low-lying, T=0, particle-hole states of

16

O,

12

C and

40

Ca using Random

Phase Approximation (RPA). The basis of single particle states was


considered to include 1s, 1p, 2s-1d and 1f-2p. The Hamiltonian was
diagonalized in the presence of Michigan three ranges Yukawa (M3Y)
interaction and compared with their previous results that depended on
Modified Surface Delta Interaction (MSDI). Admixture of higher
configuration up to 2p-1f was considered for the ground state. Effective
charges were used to account for the core polarization effect. Comparisons
26

CHAPTER ONE

INTRODUCTION

were made to experimental data where the theoretical significance of the


calculations and its results were discussed.
In 2007, Hagino and Sagawa [37] applied a three-body model consisting of
two valence neutrons and the core nucleus

14

C in order to investigate the

ground state properties and electric quadrupole transition of the 16C nucleus.
The calculated B(E2) value from the first 2 + state to the ground state showed
good agreement with the observed data with the core polarization charge
which reproduced the experimental B(E2) value for 15C. It was also showed
that their calculations account well for the longitudinal momentum
distribution of the 15C fragment from the breakup of the

16

C nucleus. They

pointed out that the dominant (d5/2)2 configurations in the ground state of 16C
played a crucial role in these agreements.
Radhi and Salman (2008) [46] had discussed Coulomb form factors for
collective E2 transitions in 18O taking into account core polarization effects.
These effects are taken into account through a microscopic perturbation
theory including excitations from the core orbits and the model space
valence orbits to all higher allowed orbits with 10 excitations. The twobody Michigan three range Yukawa (M3Y) interaction was used for the
core-polarization matrix elements. Two different model spaces with different
Hamiltonians were adopted. The calculations included the lowest four
excited 2+ states with excitation energies 1.98, 3.92, 5.25 and 8.21 MeV.
Ong et al. in 2008 [33] presented a studying of lifetime measurements of
first excited states in

16,18

C. In that article, the electric quadrupole transition

from the first excited 2+ state to the ground 0+ states in

18

C was studied

through a lifetime measurement by an upgraded recoil shadow method


applied to inelastically scattered radioactive 18C nuclei. The measured mean
lifetime was 18.9 0.9(stat) 4.4(syst) ps, corresponding to a B(E2; 21+
27

CHAPTER ONE

INTRODUCTION

2
4
0+
g.s. ) value of 4.3 0.2 1.0 e fm , or about 1.5 Weisskopf units. The mean

lifetime of the first excited 2+ state in 16C was re-measured to be 18.3 1.4
4.8 ps, about four times shorter than the value reported previously. The
discrepancy between the two results was explained by incorporating the ray angular distribution measured in that work into the previous
measurement. The transition strengths were hindered compared to the
empirical transition strengths, indicating that the anomalous hindrance
observed in 16C persists in 18C.
Wiedeking et al. in 2008 [38] studied the lifetime of the 21+ state in

16

which had been measured with the recoil distance method using the 9Be
(9Be, 2p) fusion-evaporation reaction at a beam energy of 40 MeV. The
mean lifetime was measured to be 11.7(20) ps corresponding to a B (E2;
21+0+) value of 4.15(73) e2fm4, consistent with other even-even closed
shell nuclei. Their result did not support an interpretation for decoupled
valence neutrons.
Radhi et al. (2008) [71] calculated large basis no core shell model to study
the elastic and inelastic electron scattering on

19

F. All major shells s, p, sd

and pf were considered with (0 + 2) truncations. Excitations out of major


shell space were taken into account through a microscopic theory that allows
particlehole excitations from the sd and pf shell orbits to all higher orbits
with 2 excitations. Excitations out the no core shell model space were
essential in obtaining a reasonable description of the longitudinal and
transverse electron scattering form factors.
Chatterjee et al. (2008) [72] investigated the role of higher multipole
excitations in the electromagnetic dissociation of one-neutron halo nuclei
within two different theoretical models were finite-range distorted-wave
Born approximation and another in a more analytical method with a finite28

CHAPTER ONE

INTRODUCTION

range potential. They also showed within a simple picture, how the presence
of a weakly bound state affected the breakup cross-section.
Radhi (2009) [73] investigated Coulomb excitations of open sd-shell
nuclei. Microscopic theory was employed to calculate the C2 form factors
for the first two excited 2+ states in

22

Ne,

26

Mg and

30

Si. Those collective

transitions were discussed, taking into account core-polarization effects.


Remarkable agreements were obtained between the measured and calculated
form factors for the first excited 2+ state. No strong conclusion could be
drawn for the second excited 2+ state.
Radhi et al. (2009) [74] analyzed elastic and inelastic electron scattering
from 9Be using large-basis shell model calculations. All major shells s, p, sd
and pf were considered. The calculations were performed with truncated
space, with configurations up to 2. Such calculations fail to describe the
electron scattering data without normalizing the matrix elements with
effective charges. Instead of using constant effective charges, one particle
one hole excitations were taken into account from all the major shell orbits
into all higher allowed orbits with excitations up to 10. Excitations up to
6 seemed to be large enough for sufficient convergence. Those
excitations were essential in obtaining a reasonable description of the data.
Calculations were presented for the transitions from JT = 3/2 1/2 to
JT = 3/21/2, 5/21/2 and 7/21/2.
Umeya et al. (2009) [39] presented the theoretical approach based on the
shell model to calculate effective charges of electric quadrupole transitions.
The shell-model calculation with single-particle 2 excitations in the first
order perturbation qualitatively reproduced existing experimental B(E2)
values for carbon isotopes with neutron number 5 N 16 and showed a
sudden change of the isovector effective charge beyond N = 8.
29

CHAPTER ONE

INTRODUCTION

The calculations of the neutron skin and its effect in atomic parity
violation adopted by Brown et al. (2009) [75] to atomic parity non
conservation (PNC) in many isotopes of Cs, Ba, Sm, Dy, Yb, Tl, Pb, Bi, Fr,
and Ra. Three problems were addressed: (I) neutron-skin-induced errors to
single-isotope PNC, (II) the possibility of measuring neutron skin using
atomic PNC, and (III) neutron-skin-induced errors for ratios of PNC effects
in different isotopes. In the latter case the correlations in the neutron skin
values for different isotopes lead to cancellations of the errors; this makes
the isotopic ratio method a competitive tool in a search for new physics
beyond the standard model.
Coraggio et al. (2010) [76] studied neutron-rich carbon isotopes in terms
of the shell model employing a realistic effective Hamiltonian derived from
the chiral N3LOW nucleon-nucleon potential. The single-particle energies
and effective two-body interaction had been both determined within the
framework of the time-dependent degenerate linked-diagram perturbation
theory. The calculated results were in very good agreement with the
available experimental data, providing a sound description of that isotopic
chain toward the neutron dripline. The correct location of the drip line was
reproduced.
Wuosmaa et al. [77] in 2010, studied the 15C (d, p)16C reaction in inverse
kinematics using the Helical Orbit Spectrometer at Argonne National
Laboratory. Neutron-adding spectroscopic factors gave a different probe of
the wave functions of the relevant states in

16

C. Shell-model calculations

reproduced both the present transfer data and the previously measured
transition rates.
Hagino et al. (2011) [78] proposed a simple schematic model for twoneutron halo nuclei. In that model, the two valence neutrons moved in a one30

CHAPTER ONE

INTRODUCTION

dimensional mean field, interacting with each other via a density-dependent


contact interaction. They investigated the ground state properties, and
demonstrate that the dineutron correlation could be realized with that simple
model due to the admixture of even- and odd-parity single-particle states.
They then solved the time-dependent two-particle Schrdinger equation
under the influence of a time-dependent one-body external field, in order to
discuss the effect of dineutron correlation on nuclear breakup processes. The
time evolution of two-particle density showed that the dineutron correlation
enhanced the total breakup probability, especially for the two-neutron
breakup process, in which both the valence neutrons were promoted to
continuum scattering states. They found that the interaction between the two
particles definitely favors a spatial correlation of the two outgoing particles,
which were mainly emitted in the same direction. Neutron halo deformed
nuclei from a relativistic Hartree-Bogoliubov model in a Woods-Saxon
basis.
Hammer and Phillips (2011) [79] computed E1 transitions and electric
radii in the Beryllium-11 nucleus using an effective field theory (EFT) that
exploited the separation of scales in that halo system. They fixed the
leading-order parameters of the EFT from measured data on the 1/2 + and
1/2 levels in

11

Be and the B(E1) strength for the transition between them.

Also they obtained predictions for the B(E1) strength for Coulomb
dissociation of the 11Be nucleus to the continuum. They computed the charge
radii of the 1/2+ and 1/2 states. Agreement with experiment within the
expected accuracy of a leading-order computation in the EFT was obtained.
That paper was discussed how next-to-leading-order (NLO) corrections
involving both s-wave and p-wave 10Beneutron interactions affected their
results, and displayed the NLO predictions for quantities which were free of
31

CHAPTER ONE

INTRODUCTION

additional short-distance operators at that order. Information on neutron 10Be


scattering in the relevant channels was inferred.
Zhou et al. (2011) [80] studied and discussed the halo phenomenon in
deformed nuclei by using a fully self-consistent deformed relativistic
Hartree-Bogoliubov model in a spherical Woods-Saxon based with the
proper asymptotic behavior at large distance from the nuclear center. Taking
a deformed neutron-rich and weakly bound nucleus 44Mg as an example and
by examining contributions of the halo, deformation effects, and large spatial
extensions, they showed a decoupling of the halo orbital's from the
deformation of the core.
Petri et al. (2011) [40] reported the first measurement of the lifetime of the
21+ state was in the near-dripline nucleus

C. The deduced value of 2+1 =

20

9.82.8()+0.5
1.1 (syst) ps gave a reduced transition probability of B(E2;
2
4
+3.0
+1.0
21+0+
g.s. ) = 7.51.7 ()0.4 (syst) e fm which was in a good agreement

with a shell model calculation using isospin-dependent effective charges.


In 2012, Voss et al. [35], presented the lifetime of the first excited 2+ state
which was measured with the Kln/NSCL plunger via the recoil distance
method to be (21+) = 22.4 0.9()+3.3
2.2 (syst) ps, which corresponds to a
reduced quadrupole transition strength of B(E2; 21+ 01+) = 3.64+0.15
0.14
2
4
()+0.40
0.47 (syst) e fm . In addition, an upper limit on the lifetime of a

higher-lying state feeding the 21+ state was measured to be < 4.6 ps. The
results were compared to the large-scale ab initio no-core shell model
calculations using two accurate nucleon-nucleon interactions and the
importance-truncation scheme. That comparison provided strong evidence
that the inclusion of three-body forces was needed to describe the low-lying
excited-state properties of that A = 18 system.

32

CHAPTER ONE

INTRODUCTION

Nakamura (2012) [81] discussed how the neutron halo nuclei were studied
by the breakup reactions at relativistic energies. In the Coulomb breakup of
halo nuclei, enhancement of the electric dipole strength at low excitation
energies (soft E1 excitation) was observed as a unique property for halo
nuclei. The mechanism of the soft E1 excitation and its spectroscopic
significance was shown as well as the applications of the Coulomb breakup
to the very neutron rich

22

C and

31

Ne, which was measured at 230-240

MeV/nucleon at the new-generation RI beam facility, RIBF (RI Beam


Factory), at RIKEN. Evidence of halo structures for those nuclei was
provided as enhancement of the inclusive Coulomb breakup, which was a
useful tool for the low-intense secondary beam. He also showed that the
breakup with a light target (C target), where nuclear breakup was a dominant
process, could be used to extract the spectroscopic information of the
removed neutron. The combinatorial analysis was found very useful to
extract more-detailed information such as the spectroscopic factor and the
separation energy. Prospects of the breakup reactions on neutron-drip line
nuclei at RIBF at RIKEN were also briefly presented.
In 2012, McCutchan et al. [41] studied lifetime of the 21+ state in

10

C. It

was measured using the Doppler shift attenuation method. That


measurement, combined with that determined for

10

Be 9.2(3) e2fm4 testing

the structure of those states, including the isospin symmetry of the wave
functions. By adopting Quantum Monte Carlo calculations, the reproduced
the 10Be B(E2) value by using realistic two- and three-nucleon
Hamiltonians can predict a larger 10C B(E2) probability. In these
calculations, the sensitivity to the admixture of different spatial symmetry
components in the wave functions and to the three-nucleon potential is
considered.
33

CHAPTER ONE

INTRODUCTION

1.5. Aim of the present work


The aim of the present work is using the fundamental relations to get the
reduced transition strength B(E2) from the first-excited 2+ state to the ground
state for some even-even carbon isotopes in the range A=10-20. Also, a
good imagination for the nuclear structure of these isotopes has adopted
using different model spaces and interactions. These B(E2) values represent
basic nuclear information complementary to the knowledgement of the
energies of low-lying levels in these nuclides. As well, the calculations are
depending on basic equations which explained the relation between different
parameters to reproduce the rms to fix the size parameter b of the singleparticle (HO) wave function. In this study, the size parameter plays the role
of a characteristic length of the harmonic-oscillator potential. Also, the rms
radius matter for these isotopes are reproduced and compared with the
available experimental data depending on the size parameters b for both halo
and core. For the quadrupole transition strength, excitation from the core and
model space will be taken into consideration through first-order perturbation
theory, where 1p-1h excitations are taken into considerations. These 1p-1h
excitations from the core and model space orbital are considered into all
higher allowed orbits with 2 excitations. Effective charges are calculated
in this work through the microscopic theory discussed above for the different
model spaces used, and compared with each of the stable nuclei and the
standard effective charges (e eff
1.3 e , e neff 0.5 e ) . The calculated B(E2)
p
values will be compared with the most recent experimental data. All
calculations presented in this work are performed by a Fortran 90 computer
program written by Prof. Dr. R. A. Radhi.

34

CHAPTER Two
Theoretical Considerations

CHAPTER TWO

THEORETICAL CONSIDERATIONS

CHAPTER TWO
Theoretical Considerations
2.1. Electron scattering
Electron scattering method is an excellent tool for studying nuclear
structure because there are two reasons; one of them is that the interaction is
known, as the electron interacts electromagnetically with the local charge
and current density in the target. Since this interaction is relatively weak,
one can make measurement without greatly disturbing the structure of the
target. The second one is the advantage of electrons for fixed energy loss to
the target, one can vary the three-momentum transfer q and map out the
Fourier transforms of the static and transition densities [82].With electron
scattering one can immediately relate the cross section to the transition
matrix elements of the local charge and current density operators and this is
directly to the structure of the target itself.
The scattering of electrons from a target nucleus can be distinguished two
ways. In one, the nucleus is left in its ground state after the scattering and the
energy of the electrons is unchanged which is known as "Elastic electron
scattering". In the other, the scattered electron leaves the nucleus in
different excited state and has a final energy reduced from the initial just by
the amount taken up by the nucleus in its excited state, it is called "Inelastic
electron scattering" [83-85].
Electron scattering process can be explained according to the first Born
approximation as an exchange of virtual photon, which carrying a
momentum between the electron and nucleus [86]. The first Born
Approximation is being valid only if 1 , where

is the fine

structure constant. According to this approximation two types of electron

35

CHAPTER TWO

THEORETICAL CONSIDERATIONS

scattering from the nucleus are recognized. The first is the longitudinal or
Coulomb scattering FJCo . in which the electron interacts with the charge
distribution of the nucleus where the interaction is considered as an
exchange of a virtual photon carries a zero angular momentum along the
direction of the momentum transfer q . This process gives all information
about the nuclear charge distribution. In the second type, the electron
interacts with the magnetization and current distributions. This process is
considered as an exchange of a virtual photon with angular momentum +1
along q direction. This type of scattering is called transverse scattering FJT
and it provides the information about the nuclear current and magnetization
distributions. The transverse form factor can be divided into two kinds;
transverse electric and magnetic form factors according to the parity
selection rules.

2.2. General Theory


In the plane-wave Born approximation (PWBA), the differential crosssection for the scattering of an electron from a nucleus of charge (Ze) and
mass (M) into a solid angle (d) is given by [86, 87]:
2
d d

f rec FJ (q , )

d d Mott
J

(2-1)

where d
is the Mott cross-section for high-energy electron from a
d Mott
point spineless nucleus, which is given by [88]:

Z cos( 2)
d

Mott 2 E sin 2 ( 2)
i

36

(2-2)

CHAPTER TWO

where e

THEORETICAL CONSIDERATIONS

c 1 / 137 is the fine structure constant, is the scattering

angle and Ei is the energy of incident electron.


The recoil factor of the nucleus is given by [88]:
2E

f rec 1 i sin 2 ( 2)
M

(2-3)

The total nuclear form factor for elastic scattering or for inelastic scattering
between an initial state i and final state f is FJ (q , ) of a given multipolarity
J, which is a function of momentum transfer q and the angle of scattering ,
L

contains two parts, longitudinal (Coulomb) part FJ and transverse (electric


T

and magnetic) part FJ which can be written as [86]:


q
FJ2 (q , )
q

2 q
2
L
FJ (q ) 2 tan 2 ( 2) FJT (q )

2q

(2-4)

The four-momentum transfer q is given by, (with =c=1):

q 2 q 2 (E i E f )2

(2-5)

where,
q 2 4E i E f sin 2 ( 2) (E i E f )2

(2-6)

q 2 4E i E f sin 2 ( 2) 2

(2-7)

where E i E f , E i and E f are the initial and final total energies of the
incident and scattered electron, respectively. The squared transverse form
factor can be expressed as the sum of the squared electric form factor and
squared magnetic form factor as follows:
2

FJT (q ) FJEl . (q ) FJmag . (q )

37

(2-8)

CHAPTER TWO

THEORETICAL CONSIDERATIONS

The form factor of multipolarity as a function of momentum transfer

is

written in terms of the reduced matrix elements of the transition operator


T J (q ) as [86]:

FJ (q )
2

4
J f T J (q ) J i
2
Z (2J i 1)

(2-9)

The symbol represents longitudinal L or transverse T (electric El or


magnetic mag.), Ji and Jf are the total angular momentum of the initial and

the final states, respectively. The TJ (q) is the electron scattering multipole

operator, and

is the reduced matrix element.

The nuclear states have a well-defined isospin. So using the Wigner-Eckart


theorem in isospin space, the form factor can be written in terms of the
matrix element reduced both in total angular momentum (spin) (J) and
isospin (T) (triple-bar matrix elements) [87].
FJ (q )
2

4
Z 2 (2J i 1)
T f T z

(1)
T 0,1

T z

Tf

Ti

MT

Tz

J f T f TJ T J i T i

. (2-10)
where,

TZ

Z N
2

(2-11)

. . .
The bracket
denotes the 3j-symbols. Since T z f T z i for electron
.
.
.

scattering, then MT=0.


The multipolarity J in equation (2-10) is restricted by the angular
momentum selection rule:
38

CHAPTER TWO

THEORETICAL CONSIDERATIONS

Ji Jf J Ji Jf

(2-12)

The parity selection rules [89]:


El (1)J

(2-13)

mag . (1)J 1

(2-14)

2.3. The Reduced SingleParticle Matrix Elements of the


Longitudinal Operator
The longitudinal scattering comes from the interaction of the electrons
with the charge distribution of the nucleus. The longitudinal operator is
defined as [86]:

T co
J M (q ) d r j J (qr )Y JM (r ) (r ,t z )

(2-15)

where,
j J (qr ) is the spherical Bessel function.
Y JM (r ) is the spherical harmonics function.

(r ,t z ) is the nucleon charge density operator which is given by:


z

(r ,t z ) ei (t z ) (r ri )

(2-16)

i 1

Here, the sum is over protons,

e (t z )

1 z (i )
, z 2t z and (r ri ) is Dirac delta function.
2

From equations (2.15) and (2.16), the longitudinal operator becomes:

TJcoM (q ) e (t z ) j J (qr )Y JM (r )

(2-17)

The reduced single-particle matrix element of the longitudinal operator


between the final state and the initial state can be written as:

39

CHAPTER TWO

THEORETICAL CONSIDERATIONS

TJtCoz . e (t z ) n l j J (qr ) n l l

1
1
j Y J (r ) l j
2
2
(2-18)

The reduced matrix element of spherical harmonic is given by [88]:


j 1
l l J
1
1
1
l j Y J (r ) l j (1) 2 1 (1)

2
2
2

1
(2 j 1)(2 j 1)(2J 1) 2

1
2

1
0
2

(2-19)

Equation (2.18) can be written as:

TJ ,t Z e (t Z ) PJ ( l , l )C J ( j , j )
n l

(2-20)

j (qr ) n l

where PJ and C J are the coefficients of electric parity-selection rules, which


given by [84]:

l l J
1
PJ ( l , l ) 1 (1)
2

(2-21)

C J ( j , j )

(2 j 1)(2 j 1)(2J 1) 2

j
j J

1
1
0

2
2

j 1
(1) 2

(2-22)

The radial parts n l R nl (r ) are normalized as:

2
2
R nl (r ) r dr 1

(2-23)

40

CHAPTER TWO

THEORETICAL CONSIDERATIONS

where,

n l j J (qr ) n l dr r 2 j J (qr ) R n l (r ) R n
0

(r )

(2-24)

By using the harmonic oscillator potential with the size parameter b, the
radial matrix elements of Bessel function can be solved analytically as [86]:
n l j J (qr ) n l

J
2J
y 2 exp( y ) (n 1) ! (n 1)! 2
(2J 1)!!

1
1 2

(n l ) (n l )
2
2

n 1

n 1

m 0 m 0

(1)

m m

m !m !(n m 1)!(n m 1)!

1
( (l l 2m 2m J 3))
2
3
3
(m l ) (m l )
2
2
3
1
F ( J l l 2m 2m ); J ; y (2-25)
2
2
2

bq
where y , is gamma function and F is the confluent
2
hypergrometric function which may be evaluated using [86, 90]:
F (A , B , y ) 1

A
B

A ( A 1) y

B ( B 1) 2!

where A and B are positive integers.

41

(2-26)

CHAPTER TWO

THEORETICAL CONSIDERATIONS

2.4. The Many-Particles Matrix Elements


The many-particles reduced matrix elements of T

J ,t z

, operator can be

expressed as the sum of the product of the elements of the one-body density
matrix (OBDM) times the single-particle matrix elements [84]:
J f |||TJ

||| J i OBDM

J ,T

( i , f , j , j ) |||TJ

where the reduced single-particle matrix element a

|||

(2-27)

|||TJ T ||| is given

in equation (2.20).
For inelastic scattering, the sum extends over all pairs of single-particle
states in the model space, but elastic longitudinal scattering, the sum
including the core orbits.

2.5. SingleParticle Matrix Element in SpinIsospin Formalism


The operator TJ ,t z can be written in terms of the isoscaler part and the
isovector part. Using Wickner-Eckart theorem, the single-particle matrix
element reduced in spin space can be written in terms of that reduced in
spin-isospin space as follows:1 t 1
z
1
1
j , t z TJ ,t z j , t z (1) 2 2

2
2
t z

1 t
z
(1) 2

1
2

t z

1
1
1
2 j , TJ ,T 0 j ,

2
2
0 tz
0

1
1
1
2 j , TJ ,T 1 j ,

2
2
0 tz

(2-28)

42

CHAPTER TWO

THEORETICAL CONSIDERATIONS

The reduced single-particle matrix element in spin-isospin space, given in


equation (2-27) becomes:

|||TJ T |||

2T 1
IT (t z ) ||TJ
2 tz

||

(2-29)

where,
=

= 0

(1)

1

2

(2-30)

= 1

The single-particle matrix element

||TJ

|| can be calculated
z

according to equation (2-18).

2.6. The One-Body Density Matrix Elements (OBDM)


In isospin representation, the value of OBDM JT is obtained from the
value of OBDM
OBDM

J tz

J , tz

as [91]:
T f T z

(1)

T
2 f
T z

T
(2 t z ) 6 f
T z

0 T i OBDM (T 0)
0 T z
2
1 T i OBDM (T 1)
0 T z
2

(2-31)

The OBDM contains all the information about transitions of given


multiplicities, which is imbedded in the model wave functions.

Jf
OBDM (i , f , j , j , J T )

a a

JT

(2J 1)(2T 1)

Ji
(2-32)

In words, this result says that the nuclear matrix element of this multipole
operator between any two states can be written as a sum of single particle

43

CHAPTER TWO

THEORETICAL CONSIDERATIONS

matrix elements multiplied by numerical coefficients, where the numerical

coefficient is simply the matrix element of the creation a and destruction a


operator [92].
The OBDM elements of the longitudinal C 0 (J zero ) form factor are
given by:

2T i 1 2J i 1

OBDM T 0 n j p n

(2-33)

2j 1

T i (2T i 1)(T i 1) 2J i 1

OBDM T 1 n j p n

6T z

2 j 1

(2-34)

where n j p n and n j p n are the occupation numbers.

2.7. Corrections to the Form Factor


Several corrections must be applied to the nucleus form factor given in
equations (2-9) and (2-10) to convert them into a representation appropriate
for a comparison with the experimental form factor. One of these corrections
is the center of mass. The conventional harmonic-oscillator approximation
for this correction is given by [93]:
Fc . m e q

2 b2

4A

(2-35)

where A is the nuclear mass number.


In the shell-model the nucleon is assumed to be a point, but nucleons are
actually of finite size, then the calculated form factors have to be corrected
by another correction which takes into account the finite size of the nucleon
and is given by [93- 95]:
Ff . s

(q ) 1 (
fm 1 )2
4.33

(2-36)

44

CHAPTER TWO

THEORETICAL CONSIDERATIONS

The free nucleon form factors are assumed to be the same for the proton
and the neutron.
For nuclei in which Z

1, the Plane Wave Born Approximation PWBA

is expected to describe the electron scattering data very well, except in the
region of diffraction minima, where the PWBA goes to zero. The Coulomb
distortion of the electrons increases the momentum transfer q , and an
effective momentum transfer q eff . can be used to include these effects. The
effective momentum transfer is given by [83, 82]:
q

eff

3Z e 2
q 1

2 E i Rc

where: R
c

rms

(2-37)

, and e 2 c 1.44 Mev fm.

The differences between the Distorted Wave Born Approximation DWBA at


q eff . and PWBA at q is in general smaller than the experimental errors,

therefore the form factors are calculated in PWBA and plotted vs. q , with
the data plotted vs. q eff . . Including these corrections, the form factor for a
given multipolarity J can be written in terms of matrix elements reduced
both

in

FJ (q )
2

angular

momentum

4
(1)
Z (2J i 1) T 0,1

and

T T
f

Fc .m (q ) Ff .s (q )

isospin

Tf
T
z

spaces

Ti

MT

Tz

as

[83]:
2

J f T f TJ T J i T i

2 38

45

CHAPTER TWO

THEORETICAL CONSIDERATIONS

2.8. Electron Scattering Form Factor and Nucleon Density


The reduced many-particle matrix element in equation (2-9) can be written
according to equation (2-27) as:
J f TJ , t z (qr ) J i OBDM

J ,t z

( i , f , j , j ) T J , t z

(2-39)

So, the form factor, in equation (2-9) can be written as:


FJ ,t (q )
Z

4
2
dr r j J (qr ) J ,t (r )
Z 0

(2-40)

From equation (2-9) and equation (2-39), the nucleon density can be found
to be:

J , t ( r )
Z

1
4

4
OBDM (J i , J f , J , a,b ,t Z )
2J i 1 ab

ja Y J j

0, t (r )
Z

1
4

R n l (r )R n l (r )
a a

(2-41)

b b

1
OBDM (J i , J f ,0, a,b ,t Z )
2J i 1 ab

ja Y 0 j

R n l (r )R n l ( r )
a a

(2-42)

b b

In general; the ground density distribution (Ji=Jf and J=0), equation (2-42)
can be simplified to:

0,t Z (r )

1
4

1
OBDM (J i , J i ,0, a,b ,t Z )
2J i 1 ab

j a Y 0 jb R n

a la

(l 1 ) j Y 0 r (l 1 ) j 1
2
2

1
2

(r )R n l (r )

(2-43)

j
(2 j 1) j 0

1 2 0 1 2
4

(2-44)

b b

46

CHAPTER TWO

THEORETICAL CONSIDERATIONS

j
j 0
j 1

1
1 2 0 1 2 2

where,

(l 1 ) j Y 0 r (l 1 ) j 1
2
2

2 j 1

1
2 j 1

2 j 1

2 j 1
4

(2-45)

The result in equation (2-43) satisfies the normalization condition given by:

(r )d 3r

0, t Z p / n

(2-46)

/N

where Z/N represents the total number of protons/neutrons in the nucleus,


respectively.
The root mean square radius for nuclei at J 0 is given by:

r 2

4 o , t z (r ) r 4 dr
0

(2-47)

4 o , t z (r ) r dr
2

The total matter density is given by:

0, m (r ) 0, p (r ) 0, n (r )

(2-48)

The correspondent neutron, proton, and matter rms radii can be written as
[1]:
1
2 2
r
n

4
4
0,n (r )r dr
N 0

(2-49)

1
2 2
r
p

4
4
0, p (r )r dr
Z 0

(2-50)

1
2 2
r
m

4
4
0,m (r )r dr
A 0

(2-51)

47

CHAPTER TWO

THEORETICAL CONSIDERATIONS

2.9. Normalization of ( )
From equation (2-42), one get:

0,t (r )
Z

1
4

1
OBDM (a,b ,t z , J 0)
2J i 1 ab

2 j 1 R n l (r ) R n l (r )

(2-52)

For a n l j ,b n l j , and OBDM is given by eq. (2-31), eq. (2-52) can be


written as:

4 d r r 2 0, t z (r )
0

4
4

1
2J i 1

OBDM (a,b ,t z , J

0)

ab

d r r 2 R n l (r ) R n l (r ) 2 j 1

(2-53)

where,

2
d r r R n l (r ) R n l (r ) n n

n n 0 for n n and n

4 d r r 2 0, t z (r )
0

4
4

1 for n n

1
2J i 1

OBDM (a,b ,t z , J 0) 2 j 1

(2-54)

ab

4 d r r 2 0, t z (r )
0

ab

2 j 1
OBDM (a,b ,t z , J 0)
2Ji 1

(2-55)

The factor inside the summation has been found to be the average
occupations number in each subshell j.
Normalization occ #( j ,t z )

(2-56)

Normalization Z for t z

1
2

(2-57)

48

CHAPTER TWO

THEORETICAL CONSIDERATIONS

Normalization N for t z

1
2

(2-58)

So the average occupations number in each subshell j is given by:

occ #( j ,t z ) OBDM (a,b ,t z , J 0)

2 j 1
2J i 1

(2-59)

The normalization factor is defined to be:


Normt z

1
OBDM (a,b ,t z , J 0) 2 j 1
2J i 1 a

(2-60)

2.10. Root mean square radius in terms of occupation number


From equation (2-43) and equation (2-55), one can get:
1
OBDM (a,b ,t z , J 0) 2 j a 1
ab
2
J

1
i
r 2 t z
1
OBDM (a,b ,t z , J 0) 2 j a 1
2J i 1 a

4
d r r R n l (r ) R n l (r )

d r r R n a l (r ) R n b l (r )

but,

2
d r r R n l (r ) R n l (r ) n n

1
OBDM (a,b ,t z , J 0) 2 j a 1
ab
2
J

1
i
r 2 t z
1
OBDM (a,b ,t z , J 0) 2 j a 1
2J i 1 a
49

(2-61)

CHAPTER TWO

THEORETICAL CONSIDERATIONS

d r r 4 R n l (r ) R n l (r )

(2-62)

Equations (2-60) and (2-62) become:


r 2 t z

1
1
OBDM (a,b ,t z , J 0) 2 j a 1
a
,b
Norm
(
t
)
2J i 1
z

d r r 4 R n l (r ) R n l (r )

(2-63)

but from [96]:

3
4
2
d r r R n l (r ) R n l (r ) b ( N )
0
2

(2-64)

Here N is represented the total number of oscillator quanta excited and b


is the size parameter.
r 2 t z

1
3
2
occ # (t z ) (b (N ))
N
2

(2-65)

r 2 t z

1
3
2
occ # (t z ) (b (N ))
Z
2

(2-66)

Equations (2-65) and (2-66) are the root mean square radius for (p/n),
respectively, and for matter is:
r 2 m

1
occ #
A

3
2
b
(
N

where (A) is matter number A=Z+N.

50

(2-67)

CHAPTER TWO

THEORETICAL CONSIDERATIONS

2.11. The Reduced Electromagnetic Transition Probability


The electromagnetic transition probability is defined at the photon point,
where the momentums transfer is q k

Ex
where E x the excitation energy.
c

The single particle electric multipole transition operator is given by [87, 97].
OJCO
M e (t z ) j J (qr )Y J M (r )

(2-68)

The longitudinal operator is defined as in equation (2-17).


So,
at q k

Ex
(kr )J 1 (kr )2
, j J (kr )

.......
1

c
(2J 1)!! 2 (2J 3)

(2-69)

Retaining only the leading term in the series expansion of j J (kr ) , is


obtained:
j J (kr )

(kr )J
(2J 1)!!

(2-70)

The single particle longitudinal operator becomes at q k :

(kr )J
Co .
TJM
(q ) e (t z ) Y JM (r )
(2J 1)!!

TJM OJM

kJ
(2J 1)!!

(2-71)
(2-72)

(2J 1)!!
OJM
T JM (k )
kJ

(2-73)

The many particle reduced transition electric multipole matrix element is


written as:
(2J 1)!!
J f OJ J i
J f TJ J i
J
k

The reduced transition probability B (CJ ) is defined as [96]:

51

(2-74)

CHAPTER TWO

B (CJ )

THEORETICAL CONSIDERATIONS
2

J f O J

i
2J i 1

(2-75)

So, in terms of the longitudinal multipole transition matrix element, this


becomes:
(2J 1)!!
B (CJ )
k 2J

Jf TJ Ji

(2-76)

2J i 1

The reduced transition probability B (CJ ) can be written, in term of the form
factor defined in equation (2-38) at q k , as:
(2J 1)!!
B (CJ )
k 2J

2
Z 2e 2 L
FJ (k q )
4

(2-77)

where FJL (k q ) can be calculated from equation (2-10).


The transition strength can be also calculated directly from the electric
transitions operator OJ

as:

T
B (EJ )
f
2J i 1 T 0,1 T z
1

Ti
J T
0 T z f f

OJ T J i T i

(2-78)

2.12. The Relation Between ( ) and ( )


Each matrix element permits only transitions between states of particular
angular momenta and magnetic sub state. For a transition from a state with
angular momentum J1 to a state with angular momentum J 2 with a transfer
of angular momentum J , the allowed transitions are given by [98]:
J1 J 2 J

(2-79)

52

CHAPTER TWO

THEORETICAL CONSIDERATIONS

This is a vector summation over these quantum numbers. In almost all


discussions of these transitions, the individual magnetic sub states are
removed from consideration by defining reduced transition matrix elements
using the Wigner-Eckart theorem [99]. The total transition probability from
a state J1 to a state J2 with a transfer of angular momentum J is given by the
reduced transition matrix elements. The magnetic sub states of these states
are M1 and M2, respectively, and the magnetic quantum number of the
transition is given by , which may have values J J .
The total transition probability as the result of the transition operator EJ,
which is of electric and transfers angular momentum J, is denoted by
B (EJ ; J1 J 2 ) and is given by [99]:

B (EJ ; J1 J 2 )

M 2

J 2 M 2 O (EJ ) J1M 1

(2-80)

The summation is over the magnetic quantum number of the photon and the
possible magnetic sub states M2 of the final nuclear state, with an unreduced
transition operator EJ that species the magnetic quantum number of the
photon. This reduces to [99]:

J 2 O (EJ ) J
1
B (EJ ; J1 J 2 )
(2J1 1)

(2-81)

This has no remaining dependence on the magnetic sub states M1, M2, and

of the initial state, final state, and the photon, respectively. The factor
1 2J 1 1 is present in the equation of the Wigner-Eckart theorem and

accounts for averaging over the initial magnetic sub states.


The electromagnetic operators describing the transitions between
quantized states of nuclei consist of matrices in an expanded multipole

53

CHAPTER TWO

THEORETICAL CONSIDERATIONS

sequence. These consist of electric and magnetic operators, commonly


identified as E and M, respectively. The multipole of the operator is
identified by a number (i) meaning dipole, (ii) meaning quadrupole, (iii)
meaning octupole, etc. The reduced transition matrix element for an electric
quadrupole transition from the state with spin J2 to the state with spin J1 is
therefore J 2 O (E 2) J 1 . The B (E 2 ) value for this transition, assuming
that J2 describes a state of higher energy than J1, is:

B (E 2; J 2 J1 )

J1 O (E 2) J 2

(2-82)

(2J 2 1)

In general, transition strength of any multipolarity is described by a B(E2)


value, where J is the multipolarity. The use of B(EJ) values removes the
need to specify a phase when discussing transition strengths. The B(EJ)
values are quoted as either up () or down (). For state J2 of higher energy
than state J1, the B (EJ ) value is given by equation (2-82) [99] and
equation (2-83). By using this relation:
J1 O (EJ ) J 2

J 2 O (EJ ) J1

J1 O (EJ ) J 2

(2-83)

So,

B (EJ ; J1 J 2 )

(2J1 1)

(2-84)

From equations (2.82), (2.83) and (2.84), one get:


(2J1 1) B (EJ ; J1 J 2 ) (2J 2 1) B (EJ ; J 2 J1 )

So,

54

(2-85)

CHAPTER TWO

THEORETICAL CONSIDERATIONS

B (EJ ) (2J1 1)

B (EJ ) (2J 2 1)

(2-86)

We give the relation between the B (EJ ) values for the emission and
absorption process. Suppose we have two states labeled i and f. The relation
between the reduced transition probabilities B (EJ ) B (EJ ; f i ) and
B (EJ ) B (EJ ; i f ) is given by [100]:
B (EJ )

2J i 1
B (EJ )
2J f 1

(2-87)

2.13. Core-polarization effects and effective charges


The role of the core and the truncated space can be taken into
consideration through a microscopic theory, which combines shell model
wave functions and configurations with higher energy as first order
perturbation to describe EJ excitations: these are called core polarization
effects. The reduced matrix elements of the electron scattering operator O
is expressed as a sum of the model space (MS) contribution and the core
polarization (CP) contribution, as follows:
f ||| O ||| i f ||| O ||| i

MS

f ||| O ||| i

(2-88)

CP

which can be written as:


f ||| O ||| i X ( , ) ||| O ||| ||| O |||

f i

where X is the OBDM elements. The states

and

(2-89)

are described by

the model space wave functions. Greek symbols are used to denote quantum
numbers in coordinate space and isospace, i.e. i J i T i , f J f T f
JT .

55

and

CHAPTER TWO

THEORETICAL CONSIDERATIONS

According to the first-order perturbation theory, the single particle corepolarization term is given by [96]:

||| O ||| ||| O

Q
Vres |||
Ei H 0
||| Vres

(2-90)

Q
O |||
Ef H 0

where the operator Q is the projection operator onto the space outside the
model space. The single particle core-polarization terms given in equation
(2-92) are written as [96]:

||| O |||

(1) 2
(2

1)

(1 1 )(1 2 )

1 2 1 2
2

1 | Vres | 2

2 ||| O ||| 1
(2-91)

+terms with 1 and 2 exchanged with an overall minus sign, where the
index 1 runs over particle states and 2 over hole states and is the singleparticle energy, and is calculated according to [96]:
1 / 2( 1) f (r )
1 / 2 f (r ) n

n j (2n 1 / 2)

for j 1 / 2
for j 1 / 2

(2-92)
with f (r )

20A 2/3 and 45A 1/3 25A 2/3

Higher energy configurations are taken into consideration through 1p-1h


2 excitations. For the residual two-body interaction Vres, the M3Y
interaction of Bertsch et al. [101] is adopted. The form of the potential is

56

CHAPTER TWO

THEORETICAL CONSIDERATIONS

defined in equations (1-3) in Ref. [101]. The parameters of 'Elliot' are used
which are given in Table -1 of the mentioned reference. A transformation
between LS and jj is used to get the relation between the two-body shell
model matrix elements and the relative and center of mass coordinates, using
the harmonic oscillator radial wave functions with Talmi-Moshinsky
transformation. The single particle matrix elements reduced in both spin and
isospin, is given by equation (2-29).
The reduced electric transition strength is given by:

T T Ti
B (EJ )
J f T f ||| OJT ||| J i T i
f

(2J i 1) T 0,1 T z 0 T z
1

(2-93)

which can be written as:

T T Ti
B (EJ )
eT f
M
(2J i 1) T 0,1 T z 0 T z JT
1

where

(2-94)

M JT J f T f ||| M JT ||| J i T i . The isoscalar (T=0) and isovector

1
1
(T=1) charges are given by e 0 e IS e , e1 e IV e .
2
2
The B(E2) value can be represented in terms of only the model space matrix
elements as:

B (EJ )

1
(2J i 1)

eff
eT
T 0,1

Tf T Ti
T 0 T
z
z

M JT

(2-95)

Then the isoscalar and isovector effective charges are given by:

eTeff

M JT M JT
2M JT

e p (1)T e n
2

The proton and neutron effective charges can be obtained as follows:

57

(2-96)

CHAPTER TWO

THEORETICAL CONSIDERATIONS

e p e 0eff e1eff and e n e 0eff e1eff .

The above effective charges work for mixed isoscalar and isovector
transitions. For pure isoscalar transition (for 12C), the polarization charge is
given by:
e

M J
2M J

(2-97)

and the effective charges for the proton and neutron becomes:
e peff e e , e neff e

(2-98)

58

CHAPTER Three
Results, Discussion and
Conclusions

CHAPTER THREE

Results, Discussion and Conclusions

CHAPTER THREE
Results, Discussion and Conclusions
3.1. Introduction
The shell model is the basic framework for nuclear structure calculations
in terms of nucleons. This model, which entered into nuclear physics more
than fifty years ago [102,103] is based on the assumption that, as a first
approximation, each nucleon inside the nucleus moves independently from
the others in a spherically symmetric potential including a strong spin-orbit
term. Within this approximation the nucleus is considered as an inert core,
made up by shells filled up with neutrons and protons paired to angular
momentum J=0, plus a certain number of external nucleons which are
"valence'' nucleons. As is well known, this extreme single-particle shell
model, supplemented by empirical coupling rules, proved very soon to be
able to account for various nuclear properties [104]. It was clear from the
beginning [105], however, that for a description of nuclei with two or more
valence nucleons, the "residual'' two-body interaction between valence
nucleons had to be taken explicitly into account, the term residual meaning
that part of the interaction which is not absorbed into the central potential.
This removes the degeneracy of the states belonging to the same
configuration, and produces a mixing of different configurations. A
fascinating account of the early stages of the nuclear shell model is given in
the comprehensive review by Talmi [106].
In any shell-model calculation, one has to start by defining a "model
space'', namely by specifying a set of active single particle (SP) orbits. It is
in this truncated Hilbert space that the Hamiltonian matrix has to be set up
and diagonalized. A basic input, as mentioned above, is the residual
59

CHAPTER THREE

Results, Discussion and Conclusions

interaction between valence nucleons. This is in reality a "model-space


effective interaction'', which differs from the interaction between free
nucleons in various respects. In fact, besides being residual in the sense
mentioned above, it must account for the configurations excluded from the
model space.
It goes without saying that a fundamental goal of nuclear physics is to
understand the properties of nuclei starting from the forces between
nucleons. Nowadays, the A-nucleons in a nucleus are understood as nonrelativistic particles interacting via a Hamiltonian consisting of two-body,
three-body, and higher-body potentials, with the nucleon-nucleon (NN) term
being the dominant one.
In this present work, equations (2-77) and (2-78) are used to get the
reduced transition properties B(E2) from the first-excited 2+ state to the
ground state for some even-even carbon isotopes in the range A=10-20, to
adopted a good imaginations for the nuclear structure for these isotopes
using different models space and interactions. Such calculations are
compared with available experimental data. These B(E2) values represent
basic nuclear information complementary to a knowledge of the energies of
low-lying levels in these nuclides. By applying equations (2-51) and (2-67),
the results of calculation for the (rms) are reproduced to fix the size
parameter b of the single-particle (HO) wave function.
The size parameter plays the role of a characteristic length of the
harmonic-oscillator potential.
All single-particle wave functions calculations are performed with a
harmonic oscillator potential (HO). Two values of the size parameter b are
used, one for A=10-14 (group A) and the other one for A=16-20 (group B),
due to the difference in the root mean square (rms) matter radii (Rm)
60

CHAPTER THREE

Results, Discussion and Conclusions

between the two isotopes groups. The b value of each isotope is adjusted to
reproduce the experimental matter radius. The average value of b for group
A is 1.55 fm, while that of group B is 1.78 fm. The experimental and
theoretical Rm radii are tabulated in table 3.1. The experimental values of
Rm are all taken from Ref. [107].
The 0 calculations for group A depend on two models space, the p
model space, with Cohen-Kurath interaction (CKPOT) [108] and the psd
model space, where CKPOT is used for p shell, Preedom-Wildenthal [109]
for sd shell and Millener-Kurath [110] for p-sd.
For group B Large-basis no core model space is used in this study. This
space covered the four shells 1s, 1p, 2s1d and 2p1f with (0 2)
truncations. Shell model interactions encompassing the four oscillator shells
have been constructed by Warburton and Brown [111]. These interactions
are based interactions for the 1p2s1d shells determined by a least square fit
to 216 energy levels in the A = 1022 region assuming no mixing of n
and (0 2) configurations. The 1p2s1d part of the interaction (cited in
Ref. [111] as WBP) results from a fit to two-body matrix elements and
single-particle energies for the p shell and a potential representation of the
1p2s1d cross shell interaction. The WBP model space was expanded to
include the 1s and 2p1f major shells by adding the appropriate 2p1f and
cross-shell 2s1d2p1f two-body matrix element of the WarburtonBecker
MillinerBrown (WBMB) interaction [112] and all the other necessary
matrix elements from the bare G-matrix potential of Hosaka, Kubo and Toki
[113]. The 2s1d shell interaction of Wildenthal [114] used in WBP
interaction is replaced in this study by a new interaction referred as USDB
(Universal sd-shell B) [115], where the derivation of the USD Hamiltonian

61

CHAPTER THREE

Results, Discussion and Conclusions

[114] has been refined with an up dated and complete set of energy data. The
new Hamiltonian USDB leads to a new level of precision for realistic shellmodel wave functions.
Shell model calculations were performed with the shell-model code
OXBASH [116], where the OBDM elements given in equation (2-89) were
obtained. The first term in this equation is the zero-order contribution, which
gives the single-particle matrix element for the model space (MS)
contribution. The second and third terms are the first-order contributions
which account for the higher energy configurations (core-polarization
effects). These configurations are taken through 1p1h excitations from the
core and MS orbits into all higher orbits with 2 excitations. For the
residual interaction Vres, the M3Y interaction of Bertsch et al. [101] is
adopted.

62

Results, Discussion and Conclusions

CHAPTER THREE

TABLE 3.1. The calculated root mean square matter radii of

1020

C compared with the

experimental data.

Root mean Square matter radii ( fm )


A

b (fm)

psd

Exp. [107]

10

1.55

2.25

2.35

2.270.03

12

1.55

2.28

14

1.55

2.30

b (fm)

16

1.78

2.74

2.70.03

18

1.78

2.81

2.820.04

20

1.78

2.87

2.980.05

2.350.02
2.37

2.300.07

spsdpf (0 2)

Exp. [107]

3.2. The nucleus 10C


The ground state of

10

C is with J T 0 1 , with half lifetime = 19.255s.

This nucleus has only one bound excited state, with J 2 , at 3354 keV
[41]. It becomes unbound at 4006 keV [41].
The calculations are performed with p-shell model space (0 ). The
calculated B( E 2;21 01 ) is 2.2 e2fm4 in comparison with the recent
precisely measured value 8.8 0.03 e2fm4 [41]. The 0 value under
predicts the experimental value by about a factor of 4. Including corepolarization effects with 2 excitations gives the value 6.3 e2fm4. The
core-polarization effects enhances the B( E 2;21 01 ) by about a factor of 3
over the 0 value. Previously measured value was 12.2 1.9 e2fm4 [117].
The calculated effective charges are 0.97 e and 0.67 e, for the proton and
neutron, respectively. The calculated value underestimates the measured
63

Results, Discussion and Conclusions

CHAPTER THREE

values even with core-polarization effects. The calculated B( E 2;21 01 )


depends strongly on the b value of the HO potential. Different values of b
are used previously, which give different values for the transition rate. The
present choice for b=1.55 fm gives Rm=2.25 fm which is within the
experimental value 2.270.03 fm [107]. Using the psd model space with the
configuration (1p3/2)4(1d5/2)2 enhances the B(E2) value to become 7.76 e2fm4.
Including CP effects gives the value 10.6 e2fm4 which is in agreement with
the previously measured value. The effective charges are calculated to be
equal to 1.09 e and 0.28 e, for the proton and neutron, respectively. This
result of B(E2) agrees with that of Forssen et al. (10 2 e 2 fm4 ) [118] using
a large-scale ab initio no-core shell model (NCSM). The results of B(E2)
and effective charges are tabulated in table 3.2. The analysis of the above
calculations shows a strong contribution of 1d5/2 for valence nucleons.
TABLE 3.2. The calculated effective charges and B(E2) values of 10C compared with the
experimental data.

B (E 2;21 01 ) e 2 fm 4

Effective
Model Space

Charges (e)
e eff
,e neff
p

Theo.

Theo.

Exp. [Ref]

(No CP) (With CP)

0.97, 0.67

2.2

6.3

psd [(1p3/2)4(1d5/2)2]

1.09, 0.28

7.76

10.6

64

8.80.03 [41]

Results, Discussion and Conclusions

CHAPTER THREE

3.3. The nucleus 12C


The ground state of 12C is with J T 0 0 , which is a stable nucleus. The
first excited 2+ state of this nucleus is at 4439 keV.
The calculations are performed with p-shell model space (0 ). The
calculated B( E 2;21 01 ) is 2.83 e2fm4 in comparison with the measured
value 7.94 0.66 e2fm4 [31]. The calculations with p-shell model space
including core-polarization effects with 2 excitations gives the value of

B( E 2;21 01 ) equal to 8.4 e2fm4 which reproduces very well the


experimental value. The effective charges are 1.36 e and 0.36 e for the
proton and neutron, respectively. This result of B(E2) agrees with that of
Forssen et al. 8.8(7) e2fm4 [118] using a large-scale ab initio no-core shell
model (NCSM). The results of B(E2) and effective charges are tabulated in
table 3.3.

TABLE 3.3. The calculated effective charges and B(E2) values of 12C compared with the
experimental data.

B (E 2;21 01 ) e 2 fm 4

Effective
Model Space

Charges (e)
e eff
,e neff
p

1.36, 0.36

Theo.

Theo.

Exp. [Ref]

(No CP) (With CP)


2.83

65

8.4

7.940.66 [31]

CHAPTER THREE

Results, Discussion and Conclusions

3.4. The nucleus 14C


The ground state of 14C is with J T 0 1 , with half life= 5730 y. The first
excited 2+ state of this exotic nucleus is at 7012(4) keV.
The calculations are performed with p-shell model space without core
polarization effects (0 ) which gives B( E 2;21 01 ) equal to 3.64 e2fm4
and with core-polarization effects with 2 excitations which gives the
value of B( E 2;21 01 ) equal to 5.9 e2fm4 in comparison with the measured
value 3.74 0.5 e2fm4 [31]. The calculated effective charges are 1.27 e and
0.52 e, for the proton and neutron, respectively. These values are very close
to the standard values. The calculated value with CP effects overestimates
the measured value by about a factor of 1.5. However, this low value is
expected due to the neutron N=8 closed shell effect, where only the protons
have contribution. Using the psd model space with the configuration
(1p1/2)4(1p3/2)4 (1d5/2)2 reduces the B(E2) value to become 0.66 e2fm4 without
CP and 3.0 e2fm4 with CP which is close to the measured value. The
effective charges are calculated to be equal to 1.1 e and 0.3 e. The result of
Forssen et al. is 5.3(7) e2fm4 [118] using a large-scale ab initio no-core shell
model (NCSM), which is close to the p-shell result including CP effects. The
results of B(E2) and effective charges are tabulated in table 3.4. The analysis
of the above calculations shows a strong contribution of 1d5/2 for valence
neutrons. The configuration given above that gives this small value of B(E2)
which agrees with the measured value, suggests a possible proton-shell
closure in 14C.

66

CHAPTER THREE

Results, Discussion and Conclusions

TABLE 3.4. The calculated effective charges and B(E2) values of 14C compared with the
experimental data.

B (E 2;21 01 ) e 2 fm 4

Effective
Model Space

Charges (e) Theo.


Theo.
Exp.
e eff
,e neff
(No CP) (With CP)
p

1.27, 0.52

3.64

5.9

psd[(1p1/2)4(1p3/2)4(1d5/2)2]

1.10, 0.30

0.66

3.0

3.740.5 [31]

3.5. The nucleus 16C


The ground state of 16C is with J T 0 2 , with half life= 0.747s. The first
excited 2+ state of this exotic nucleus is at 1766 (10) keV [33].
The recent measurements of the B( E 2;21 01 ) values are 4.150.73
e2fm4 [38] and 2.60.9 e2fm4 [33]. The calculations are performed with
spsdpf model space with (0 + 2) . For the ground state configurations, two
neutrons are distributed over the sd shell orbits with 62% over 1d5/2 orbit,
6% over 1d3/2 orbit and 32% over the 2s orbit. The calculated

B( E 2;21 01 ) is 0.51 e2fm4 without CP effects and that with CP effects is


3.3 e2fm4, which agrees very well with both experimental values. The
effective charges are calculated to be equal to 1.1 e and 0.24 e, for the proton
and neutron, respectively. The results of B(E2) and effective charges are
tabulated in table 3.5. The analysis of the above calculations shows a major
contribution of 1d5/2 for valence two neutrons. The configurations resulting
from the spsdpf model space with (0 + 2) gives a simple structure of
16

C=

14

C+n+n. A large extension of the neutron density distribution to a


67

Results, Discussion and Conclusions

CHAPTER THREE

distance far from the center of the nucleus suggests the formation of neutron
halo in the 16C nucleus [64]. Due to the relatively small value of B(E2), and
the distribution of the two neutrons over the sd shell out side

14

C may

support the decoupling of neutrons from protons to form neutron halo. The
result of Forssen et al. is 2.2(6) e2fm4 [118] using a large-scale ab initio nocore shell model (NCSM).
TABLE 3.5. The calculated effective charges and B(E2) values of 16C compared with the
experimental data.

B (E 2;21 01 ) e 2 fm 4

Effective
Model Space

Charges (e)
eff
p

e ,e

spsdpf (0+2)

eff
n

1.1, 0.24

Theo.

Theo.

Exp. [Ref]

(No CP) (With CP)


0.51

3.33

4.150.73 [38]
2.60.9

68

[33]

CHAPTER THREE

Results, Discussion and Conclusions

3.6. The nucleus 18C


The ground state of 18C is with J T 0 3 , with half life= 95 ms. The first
excited 2+ state of this exotic nucleus is at 1558 (19) keV [35].
2
4
The recent measurements of the B( E 2;21 01 ) values are 3.64 ..55
61 e fm

[35] and 4.31.2 e2fm4 [33]. The calculations are performed with spsdpf
model space with (0 + 2) . The ground state configurations show that four
neutrons are distributed over the sd shell orbits with 74.5% over 1d5/2 orbit,
8% over 1d3/2 orbit and 17.5% over the 2s orbit. The calculated

B( E 2;21 01 ) is 0.69 e2fm4 without CP effects and that with CP effects is


4.8 e2fm4, which agrees very well with experimental value of Ref. [33] and
very close to that of Ref. [35]. The effective charges are calculated to be
equal to 1.11 e and 0.25 e, for the proton and neutron, respectively. The
results of B(E2) and effective charges are tabulated in table 3.6. The analysis
of the above calculations shows a major contribution of 1d5/2 for valence
four neutrons. The configurations resulting from the spsdpf model space
with (0 + 2) gives a simple structure of

18

C=

14

C+4n. The result of

Forssen et al. is 4.2(4) e2fm4 [118] using a large-scale ab initio no-core shell
model (NCSM). The B(E2) values of

16

C and

18

C are less than those of

closed shell nuclei. Same conclusion can be drawn as in 14C and 16C for the
formation of proton-shell closure.

69

Results, Discussion and Conclusions

CHAPTER THREE

TABLE 3.6. The calculated effective charges and B(E2) values of 18C compared with the
experimental results.

B (E 2;21 01 ) e 2 fm 4

Effective
Model Space

Charges (e)
e eff
,e neff
p

spsdpf (0+2)

1.11, 0.25

Theo.

Theo.

Exp. [Ref]

(No CP) (With CP)


0.69

4.8

3.64.55
.61 [35]
4.31.2 [33]

3.7. The nucleus 20C


The ground state of 20C is with J T 0 4 , with half life= 14 ms, the first
excited 2+ state of this exotic nucleus is at 1618(6) keV [40].
The recent measurement of the B( E 2;21 01 ) value is 7.542.1 e2fm4 [40].
The calculations are performed with spsdpf model space with (0 + 2) . In
this case six neutrons are distributed over the sd shell orbits with about 72%
over 1d5/2 orbit, 8% over 1d3/2 orbit and 20% over the 2s orbit. The
calculated B( E 2;21 01 ) is 2.30 e2fm4 without CP effects and 9.4 e2fm4
with CP effects, which agrees very well with experimental value of Ref.
[40]. The effective charges are calculated to be equal to 1.13 e and 0.26 e,
for the proton and neutron, respectively. The results of B(E2) and effective
charges are tabulated in table 3.7. The analysis of the above calculations
shows a major contribution of 1d5/2 for valence six neutrons. The
configurations resulting from the spsdpf model space with (0 + 2) gives
a simple structure of

20

C=

14

C+6n. The result of Forssen et al. is 4.8(1.1)

e2fm4 [118] using a large-scale ab initio no-core shell model (NCSM).


Almost same effective charges are obtained as those of 14,16,18C. The average
70

Results, Discussion and Conclusions

CHAPTER THREE

effective charges for

14, 16, 18,20

C are 1.1 e and 0.25 e, for the proton and

neutron, respectively.
TABLE 3.7. The calculated effective charges and B(E2) values of 20C compared with the
experimental results.

B (E 2;21 01 ) e 2 fm 4

Effective
Model Space

Charges (e)
eff
p

e ,e

spsdpf (0+2)

eff
n

1.13, 0.26

Theo.

Theo.

Exp. [Ref]

(No CP) (With CP)


2.30

9.4

7.542.1 [40]

The calculated effective charges using psd model space with the
configurations discussed above and those calculated with spsdpf model
space with (0+2) are very close to each other and their average value is
1.1 e and 0.27 e for the proton and neutron, respectively. These values can
be used for the entire chain of Carbon isotopes for mixed isoscalar and
isovector transitions to get the required B(E2) values. These values are
tabulated in table 3.8.
The calculated B(E2) values agree very well with the recent experimental
data for the entire chain of Carbon isotopes, as seen in table 3.8.

71

Results, Discussion and Conclusions

CHAPTER THREE

TABLE 3.8. The calculated B(E2) values using average effective charges

e p 1.1 e and e n 0.27 e


eff

eff

, using psd MS for A=10,14 and spsdpf

(0+2) MS for A=16,18 and 20, compared with experimental values.

B (E 2;21 01 ) e 2 fm 4

A
Theory

Exp.

10

10.8

8.80.03a

14

2.7

3.740.5b

16

3.8

18

5.2

e
d
3.65.55
.61 0.73 , 4.31.2

20

9.4

7.542.1 e

4.150.73c ,

a) Reference [41].
b) Reference [31].
c) Reference [38].
d) Reference [33].
e) Reference [40].

72

2.60.9d

Results, Discussion and Conclusions

CHAPTER THREE

3.8. Conclusions
Shell model calculations are performed for even-even Carbon isotopes
(10 A 20 ) including core-polarization effects through first-order
perturbation theory, where 1p1h with 2 excitation are taken into
considerations. In general, there are some notes have been indicated from the
present work which can be explained as:
The 0 and (0 + 2) calculations which succeed in describing energy
levels and other static properties, are less successful for describing
dynamical properties such as transition strengths B(E2). The core
contributions cannot be ignored in such transitions and the core polarization
effects play a major role for describing such dynamical property.
The size parameters of the harmonic oscillator potential chosen for this
work almost reproduce all the rms matter radii for

10,12,14,16,18,20

C isotopes

within the experimental errors.


Now, the main conclusions are briefly summarized as:
1. The calculations for 10C show a strong contribution of 1d5/2 orbit for
the valence nucleons in order to describe the data reasonably.
2. The B(E2) value for the

12

C nucleus is described very well with p-

shell model space when core polarization effects are taken into
consideration. This nucleus is considered for many calculations as an
inert core, and only the valence nucleons outside this core are
considered. However, the calculations showed that the major
contribution to the transition strength comes from the core
polarization where excitations are considered from the 4He core and
the valence nucleons in p-shell orbits. Same effective charges are
obtained as the empirical values of other closed shell nuclei.

73

Results, Discussion and Conclusions

CHAPTER THREE

3. In

14

C, no essential contribution comes from the neutrons. The 8

neutrons (magic number) form a closed shell. The small value of


B(E2) shows a strong contribution of 1d5/2 for valence neutrons. This
small value of B(E2) which agrees with the measured value, suggests
a possible proton-shell closure in 14C. The value of transition strength
for the transition from the first 2 + state to the ground state shows a
clear exotic behavior for this neutron rich isotope. A simple 12C+n+n
structure are suggested for 14C.
4. Also the small value obtained for the transition strength in

16

C which

agrees very well with the measured values suggests a possible protonshell closure with a simple

14

C+n+n structure. These two neutrons

may form a halo around the 14C nucleus, but cannot be considered as a
Borromean since

14

C exists as a long lived isotope. A clear exotic

behavior is observed for this neutron rich isotope.


5. Also a clear exotic behavior is observed for 18C neutron rich isotope.
6. The calculated B(E2) value for the neutron rich

20

C isotope, which

agrees very well with the experimental value, is greater than those for
14,16,18

C.

The overall conclusions are: The major contribution to the transition


strength comes from the core-polarization effects. The present calculations
of the neutron-rich

16

C,

18

C and

20

C isotopes show a deviation from the

general trends of even-even nuclei in accordance with experimental and


other theoretical studies. The experimental values are very well reproduced
confirming the anomalous suppression in

16

C and

18

C. The configurations

arises from the shell model calculations with core-polarization effects which
reproduce the experimental B(E2) values, and give small effective charges
74

Results, Discussion and Conclusions

CHAPTER THREE

confirm the formation of proton-shell closure for


microscopic effective charges for

14,16,18,20

14,16,18

C. The average

C are 1.1 e and 0.27 e, for the

proton and neutron, respectively which are smaller than the standard
empirical values for in this mass region.

3.9. Future Work


1. Calculate the quadrupole moment for even-even Carbone isotopes.
2. Study even- odd Carbone isotopes and compare the results with
available experimental data.
3. Study other exotic nuclei for example: Li, Be, He and B isotopes.
4. Extending this work in an attempt to include heavy exotic nuclei, for
example: 56Ni and 126Sn nuclei.

75

References

REFERENCES

References
[1]

K. Heyde, Basic Ideas and Concepts in Nuclear Physics, IOP


Publishing Ltd, Second edition, (1999).

[2]

I. Talmi and I. Unna, Phys. Rev. Lett. 4 (1960) 469; P G Hansen; Nucl.
Phys. A 682 (2001) 310.

[3]

J. Dobaczewski, W. Nazarewicz, T. R. Werner, J. F. Berger, C. R. Chinn


and J. Decharge, Phys. Rev. C 53 (1996) 6.

[4]

National

Nuclear

Data

Center

2012

Chart

of

nuclides

www.nndc.bnl.gov/chart.
[5]

E. Kwan, C. Y. Wu, N. C. Summers, G. Hackman, T. E. Drake, C.


Andreoiu, R. Ashley, G. C. Ball, P. C. Bender, A. J. Boston, H. C.
Boston, A. Chester, A. Close, D. Cline, D. S. Cross, R. Dunlop, A.
Finley, A. Garnsworthy, A. B. Hayes, A. Laffoley, T. Nano, P. Navratil,
C. J. Pearson, J. Pore, K. Starosta4, I. J. Thompson1, P. Voss, S. J.
Williams, Z. M. Wang, Phys. Lett. B 650 (2007) 124.

[6]

G. Raimann, A. Ozawa, R. N. Boyd, F. R. Chloupek, M. Fujimaki, K.


Kimura, T. Kobayashi, J. J. Kolata, S. Kubono, I. Tanihata, Y. Watanabe,
K. Yoshida, Prog. Part. Nucl. Phys. A 35 (1995) 505.

[7]

B. Jonson, Phys. Rep. 389 (2004) 1.

[8]

P. G. Hansen and B. Jonson, Euro. Phys. Lett. 4 (1987) 409.

[9]

G. F. Bertsch and H. Esbensen, Ann. Phys. 209 (1991) 327.

[10]

M. V. Zhukov, B. V. Danilin, D. V. Fedorov, J. M. Bang, I. J. Thompson


and J. S. Vaagen, Phys. Rep. 231 (1993) 151.

[11]

A. S Jensen, K. Riisageer, D. V. Fedorov and E. Garrido, Rev. Mod.


Phys. 76 (2004) 215.

[12]

I. Tanihata, H. Hamagaki, O. Hashimoto, Y. Shida, N. Yoshikawa, K.


Sugimoto, O. Yamakawa, T. Kobayashi and N. Takahashi, Phys. Rev.
Lett. 55 (1985) 2676.
76

REFERENCES

[13]

T. Kobayashi, O. Yamakawa, K. Omata, K. Sugimoto, T. Shimoda, N.


Takahashi and I. Tanihata, Phys. Rev. Lett. 60 (1988) 25.

[14]

Y. Ogawa, K. Yabana and Y. Suzuki, Nucl. Phys. A 543 (1992) 722.

[15]

A.S. Jensen and M.V. Zhukov, Nucl. Phys. A 693 (2001) 411.

[16]

W.D. Myers and W.J. Swiatecki, Ann. Phys. 55 (1969) 395.

[17]

K. Riisager, A.S. Jensen and P. Mhller, Nucl. Phys. A 548 (1992) 393.

[18]

D.V. Fedorov, A.S. Jensen and K. Riisager, Phys. Lett. B 312 (1993) 1.

[19]

D.V. Fedorov, A.S. Jensen and K. Riisager, Phys. Rev. C 50 (1994)


2372.

[20]

K. Tanaka, T. Yamaguchi, T. Suzuki, T. Ohtsubo, M. Fukuda, D.


Nishimura, M. Takechi, K. Ogata, A. Ozawa, T. Izumikawa, T. Aiba, N.
Aoi, H. Baba1, Y. Hashizume, K. Inafuku, N. Iwasa, K. Kobayashi, M.
Komuro, Y. Kondo, T. Kubo, M. Kurokawa, T. Matsuyama, S.
Michimasa, T. Motobayashi, T. Nakabayashi, S. Nakajima, T.
Nakamura, H. Sakurai, R. Shinoda, M. Shinohara, H. Suzuki, E.
Takeshita1, S. Takeuchi1, Y. Togano11, K. Yamada1, T. Yasuno6, and
M. Yoshitake, Phys. Rev. Lett. 104 (2010) 062701.

[21]

Melva McLean, T. I. Meyer, Five- Year Plan 2010-2015. Copyright


2008 TRIUMF.

[22]

Dana Borremans Ph. D. Thesis submits to Institute voor Kernen


Stralingsfysica, (2004).

[23]

Isao Tanihata and J. G. Phys, J. Phys. G: Nucl. Part. Phys. 22 (1996)


157.

[24]

C. Thibault, R. Klapisch, C. Rigaud, A.M. Poskanzer, R. Prieels, L.


Lessard and W. Reisdorf, Phys. Rev. C 12 (1975) 644.

[25]

D.J Millener, J.W. Olness, E.K. Warburton and S. Hanna, Phys. Rev.
C 28 (1983) 497.

[26]

A.V. Dobrovolsky, G.D. Alkhazov, M.N. Andronenko, A. Bauchet, P.


Egelhof, S. Fritz, H. Geissel, C. Gross, A.V. Khanzadeev, G.A. Korolev,
77

REFERENCES

G. Kraus, A.A. Lobodenko, G. Mnzenberg, M. Mutterer, S.R.


Neumaier, T. Schfer, C. Scheidenberger, D.M. Seliverstov, N.A.
Timofeev, A.A. Vorobyov, V.I. Yatsoura, Nucl. Phys. A 766 (2006) 1.
[27]

P. Egelhof, G.D. Alkhazov, M.N. Andronenko, A. Bauchet, A.V.


Dobrovolsky, S. F ritz, G. E. Gavrilov, H. Geissel, C. Gross, A. V.
Khanzadeev, G. A. Korolev, G. Kraus1, A. A. Lobodenko, G.
Munzenberg, M. Mutterer, S. R. Neumaier, T. Schafer, C.
Scheidenberger, D. M. Seliverstov, N. A. Timofeev, A. A. Vorobyov,
and V. I. Yatsoura, Euro. Phys. J. A 15 (2002) 27.

[28]

G.D. Alkhazov, A. V. Dobrovolsky, P. Egelhof, H. Geissel, H. Irnich, A.


V. Khanzadeev, G. A. Korolev, A. A. Lobodenko, G. Mnzenberg, M.
Mutterer, S.R. Neumaier, W. Schwab, D. M. Seliverstov, T. Suzuki and
A. A. Vorobyov, Nucl. Phys. A 712 (2002) 269.

[29]
[30]

G. Audi, A. H. Wapstra and C. Thibout, Nucl. Phys. A 729 (2003) 337.


Y. Urata, K. Hagino and H. Sagawa, Phys. Rev. C 83 (2011)
041303(R).

[31]

S.Raman, C. W. Nestor, J.R. and P. Tikkanen, Atomic Data and Nuclear


Data table 78 (2001) 1.

[32]

N. Imai, H.J. Ong, N. Aoi, H. Sakurai, K. Demichi, H. Kawasaki, H.


Baba, Zs. Dombrdi, Z. Elekes, N. Fukuda, Zs. Flp, A. Gelberg, T.
Gomi, H. Hasegawa, K. Ishikawa, H. Iwasaki, E. Kaneko, S. Kanno, T.
Kishida, Y. Kondo, T. Kubo, K. Kurita, S. Michimasa, T. Minemura,
M. Miura, T. Motobayashi, T. Nakamura, M. Notani, T. K. Onishi, A.
Saito, S. Shimoura, T. Sugimoto, M. K.Suzuki, E.

Takeshita, S.

Takeuchi, M. Tamaki, K. Yamada, K. Yoneda, H. Watanabe and M.


Ishihara, Phys. Rev. Lett. 92 (2004) 6.
[33]

H.J. Ong, N. Imai, D. Suzuki, H. Iwasaki, H. Sakurai, T.K. Onishi,


M.K. Suzuki, S. Ota, S. Takeuchi, T. Nakao, Y. Togano, Y. Kondo, N.
Aoi, H. Baba, S. Bishop, Y. Ichikawa, M. Ishihara, T. Kubo, K. Kurita,
78

REFERENCES

T. Motobayashi, T. Nakamura, T. Okumura and Y. Yanagisawa, Phys.


Rev. C 78 (2008) 014308.
[34]

Z. Elekes, Zs. Dombrdi, T. Aiba, N. Aoi, H. Baba, D. Bemmerer, B.A.


Brown, T. Furumoto, Zs. Flp, N. Iwasa, . Kiss, T. Kobayashi, Y.
Kondo, T. Motobayashi, T. Nakabayashi, T. Nannichi, Y. Sakuragi, H.
Sakurai, D. Sohler, M. Takashina, S. Takeuchi, K. Tanaka, Y. Togano,
K. Yamada, M. Yamaguchi and K. Yoneda, Phys. Rev. C 79 (2009)
011302(R).

[35]

P. Voss, T. Baugher, D. Bazin, R. M. Clark, H. L. Crawford, A.


Dewald, P. Fallon, A. Gade, G. F. Grinyer, H. Iwasaki, A. O.
Macchiavelli, S. McDaniel, D. Miller, M. Petri, A. Ratkiewicz, W.
Rother, K. Starosta, K. A. Walsh, D. Weisshaar, C. Forssen, R. Roth
and P. Navratil, Phys. Rev. C 86 (2012) 011303(R).

[36]

H. Sagawa, X. R. Zhou, X. Z. Zhang and T. Suzuki, Phys. Rev. C 70


(2004) 054316.

[37]

K. Hagino and H. Sagawa, Phys. Rev. C 75 (2007) 021301(R).

[38]

M. Wiedeking, P. Fallon, A. O. Macchiavelli, J. Gibelin, M. S.


Basunia, R. M. Clark, M. Cromaz, M.-A. Deleplanque, S. Gros, H. B.
Jeppesen, P. T. Lake, I.-Y. Lee, L. G. Moretto, J. Pavan, L. Phair, E.
Rodriguez-Vietiez, L. A. Bernstein, D. L. Bleuel, J. T. Burke, S. R.
Lesher, B. F. Lyles and N. D. Scielzo, Phys. Rev. Lett. 100 (2008)
152501.

[39]

A. Umeya, G. Kaneko and K. Muto, Nucl. Phys. A 829 (2009) 13.

[40]

M. Petri, P. Fallon, A. O. Macchiavelli, S. Paschalis, K. Starosta, T.


Baugher, D. Bazin, L. Cartegni, R. M. Clark, H. L. Crawford, M.
Cromaz, A. Dewald, A. Gade, G. F. Grinyer, S. Gros, M. Hackstein, H.
B. Jeppesen, I. Y. Lee, S. McDanie D. Miller, M. M. Rajabali, A.
Ratkiewicz, W. Rother, P. Voss, K. A. Walsh, D. Weisshaar, M.
Wiedeking, and B. A. Brown, Phys. Rev. Lett. 107 (2011) 102501.
79

REFERENCES

[41]

E. A. McCutchan, C. J. Lister, Steven C. Pieper, R. B. Wiringa, D.


Seweryniak, J. P. Greene, P. F. Bertone, M. P. Carpenter, C. J. Chiara,
G. Gurdal, C. R. Hoffman, R. V. F. Janssens, T. L. Khoo, T. Lauritsen
and S. Zhu. Phys. Rev. C 86 (2012) 014312.

[42]

A. Cichocki, J. Dubach, R. S. Hicks, G. A. Peterson, C.W. de Jager, H.


de Vries, N. Kalantar-Nayestanaki and T. Sato, Phys. Rev. C 51 (1995)
2406.

[43]

D. E. Alburger, P. D. Parker, D. J. Bredin, D. H. Wilkinson, P. F.


Donovan, A. Gallmann, R. E. Pixley, L. F. Chase Jr. and R. E.
McDonald, Phys. Rev. 143 (1966) 692.

[44]

B. A. Brown, R. Radhi and B.H. Wildenthal, Phys. Rep. 101 (1983)


313.

[45]

H. Sagawa and B.A. Brown, Nucl. Phys. A 430 (1984) 84.

[46]

R.A. Radhi and E.A. Salman, Nucl. Phys. A 806 (2008) 179.

[47]

A. Umeya and K. Muto, Nucl. Phys. A 722 (2003) 558c.

[48]

Y. Suzuki and K. Ikeda, Phys. Rev. C 38 (1988) 410.

[49]

Y. Suzuki and Wang Jing JU, Phys. Rev. C 41 (1990) 736.

[50]

Y. Tosaka and Y. Suzuki, Nucl. Phys. A 512 (1990) 46.

[51]

I. Tanihata, T. Kobayashi, O. Yamakawa, S. Shimourai, K. Ekuni, K.


Sugimoto, N. Takahashi, T. Shimoda and H. Sato, Phys. Lett. B 206
(1988) 592.

[52]

S. Raman, C. W. Nestor and K. H. Bhatt, Phys. Rev. C 37 (1988) 2.

[53]

T. Minamisono, T. Ohtsubo, I. Minami, S. Fukuda, A. Kitagawa, M.


Fukuda, K. Matsuta, Y. Nojiri, S. Takeda, H. Sagawa and H. Kitagawa,
Phys. Rev. Lett. 69 (1992) 2058.

[54]

T. Otsuka, N. Fukunishi, and H. Sagawa, Phys. Rev. Lett. 70 (1993)


1385.

[55]

H. Nakada and T. Otsuka, Phys. Rev. C 49 (1994) 886.

80

REFERENCES

[56]

V. Guimaraes, J. J. Kolata, D. Bazin, B. Blank, B. A. Brown, T.


Glasmacher, P. G. Hansen, R. W. Ibbotson, D. Karnes, V. Maddalena,
A. Navin, B. Pritychenko, B. M. Sherrill, D. P. Balamuth and J. E.
Bush, Phys. Rev. Lett. 74 (1995) 18.

[57]

W. Schwab, H. Geissel, H. Lenske, K. -H. Behr, A. Brnle, K. Burkard,


H. Irnich, T. Kobayashi, G. Kraus, A. Magel, G. Mnzenberg, F.
Nickel, K. Riisager, C. Scheidenberger, B. M. Sherrill, T. Suzuki, B.
Voss. Z, Nucl. Phys. A 350 (1995) 283.

[58]

C. A. Bertulani, H. Sagawa, Nucl. Phys. A 588, Issue 3, (1995) 667.

[59]

I. Pecina, R. Anne, D. Bazin, C. Borcea, V. Borrel, F. Carstoiu, J. M.


Corre, Z. Dlouhy, A. Fomitchev, D. Guillemaud-Mueller, H. Keller, A.
Kordyasz, M. Lewitowicz, S. Lukyanov, A. C. Mueller, Yu.
Penionzhkevich, P. Roussel-Chomaz, M. G. Saint-Laurent, N.
Skobelev, O. Sorlin, and O. Tarasov, Phys. Rev. C 52 (1995) 191.

[60]

J. H. Kelley, Sam M. Austin, A. Azhari, D. Bazin, J. A. Brown, H.


Esbensen, M. Fauerbach, M. Hellstrm, S. E. Hirzebruch, R. A. Kryger,
D. J. Morrissey, R. Pfaff, C. F. Powell, E. Ramakrishnan, B. M.
Sherrill, M. Steiner, T. Suomijrvi and M. Thoennessen, Phys. Rev.
Lett. 77 (1996) 5020.

[61]

N. A. Orr, N. Anantaraman,

Sam M. Austin, C. A. Bertulani, K.

Hanold, H. Kelley, D. J. Morrissey, B. M. Sherrill, G. A. Souliotis, M.


Thoennessen, S. W nfield and A. Winger, Phys. Rev. Lett. 69 (1992)
2050.
[62]

M. J. Chromik, B. A. Brown, M. Fauerbach, T. Glasmacher, R.


Ibbotson, H. Scheit, P. G. Thirolf and M. Thoennessen, Phys. Rev. C 55
(1997) 1676.

[63]

V. Guimares, J. J. Kolata, D. Bazin, B. Blank, B. A. Brown, T.


Glasmacher, P. G. Hansen, R. W. Ibbotson, D. Karnes, V. Maddalena,

81

REFERENCES

A. Navin, B. Pritychenko, B. M. Sherrill, D. P. Balamuth, and J. E.


Bush, Phys. Rev. C 61 (2000) 064609.
[64]

T. Zheng , T. Yamaguchi , A. Ozawaa, M. Chibaa, R. Kanungo,T. Kato


, K. Katori , K. Morimoto , T. Ohnishi , T. Sudaa, I. Tanihata , Y.
Yamaguchi , A. Yoshida , K. Yoshida , H. Toki and N. Nakajima ,
Nucl. Phys. A 709 (2002) 103.

[65]

D. Bazin, B. A. Brown, C. M. Campbell, J. A. Church, D. C. Dinca, J.


Enders, A. Gade, T. Glasmacher, P. G. Hansen, W. F. Mueller, H.
Olliver, B. C. Perry, B. M. Sherrill, J. R. Terry and J. A. Tostevin, Phys.
Rev. Lett. 91 (2003) 1.

[66]

T. Zerguerras, B. Blank, Y. Blumenfeld, T. Suomijarvi, D. Beaumel,


B. A. Brown, M. Chartier, M. Fallot, J. Giovinazzo, C. Jouanne, V.
Lapoux, I. Lhenry-Yvon, W. Mittig, P. Roussel-Chomaz, H. Savajols,
J.A. Scarpaci, A. Shrivastava and M. Thoennessen, Euro. Phys. J. A 20
(2004) 389.

[67]

M. Stanoiu, F. Azaiez, F. Becker, M. Belleguic, C. Borcea, C.


Bourgeois, B.A. Brown, Z. Dlouhy, Z. Dombradi, Z. Fulop, H.
Grawe, S. Grevy, F. Ibrahim, A. Kerek, A. Krasznahorkay, M.
Lewitowicz, S. Lukyanov, H. van der Marel, P. Mayet, J. Mrazek, S.
Mandal, D. Guillemaud-Mueller, F. Negoita, Y.E. Penionzhkevich, Z.
Podolyak, P. Roussel-Chomaz, M.G. Saint Laurent, H. Savajols, O.
Sorlin, G. Sletten1, D. Sohler, J. Timar, C. Timis and A. Yamamoto,
Euro. Phys. J. A 20 (2004) 95.

[68]

C.A. Bertulani, Phys. Lett. B 624 (2005) 203.

[69]

C. M. Campbell, N. Aoi, D. Bazin, M. D. Bowen, B. A. Brown, J. M.


Cook, C. Dinca, A. Gade, T. Glasmacher, M. Horoi, S. Kanno, T.
Motobayashi, W. F. Mueller, H. Sakurai, K. Starosta, H. Suzuki, S.
Takeuchi, J. R. Terry, K. Yoneda and H. Zwahlen, Phys. Rev. Lett. 97
(2006) 112501.
82

REFERENCES

[70]

Ali H. Taqi and Raad. A. Radhi. Turk. J. Phys. 31 (2007) 253.

[71]

R. A. Radhi, A. A. Abdullah and A. H. Raheem, Nucl. Phys. A 798


(2008) 16.

[72]

R. Chatterjee, L. Fortunato and A. Vitturi. Euro. Phys. J. A 35 (2008)


213.

[73]

R. A. Radhi. J. Phys. A: Math. Theor. 42 (2009) 214047.

[74]

R. A. Radhi, N. M. Adeeb and A. K. Hashim, J. Phys. G: Nucl. Part.


Phys. 36 (2009) 105102.

[75]

B. A. Brown, A. Derevianko and V. V. Flambaum, Phys. Rev. C 79


(2009) 035501.

[76]

L. Coraggio, A. Covello, A. Gargano and N. Itaco, Phys. Rev. C 81


(2010) 064303.

[77]

A. H. Wuosmaa, B. B. Back, S. Baker, B. A. Brown, C. M. Deibel, P.


Fallon, C. R. Hoffman, B. P. Kay, H.Y. Lee, J. C. Lighthall, A. O.
Macchiavelli, S. T. Marley, R. C. Pardo, K. E. Rehm, J. P. Schiffer,
D.V. Shetty and M. Wiedeking, Phys. Rev. Lett. 105 (2010) 132501.

[78]

K. Hagino, A. Vitturi, F. Prez-Bernal and H. Sagawa, J. Phys. G: Nucl.


Part. Phys. 38 (2011) 015105.

[79]

H. W. Hammer and D.R. Phillips, Nucl. Phys. A 865 (2011) 17.

[80]

S. G. Zhou, J. Meng, P. Ring and E. G. Zhao, Journal of Physics:


Conference Series 312 (2011) 092067.

[81]

Takashi Nakamura, J. Phys.: Conf. Ser. 381 (2012) 012014.

[82]

T. W. Donnelly and J. D. Waleaka, Nucl. Phys. A 201 (1973) 81c.

[83]

H. Uberall, Electron Scattering From Complex Nuclei Part B,


Academic Press, New York, (1971).

[84]

R. A. Radhi, Ph. D. Thesis, Michigan's State University, USA, (1983).

[85]

B. Ghosh, S. K. Sharma, Phy. Rev. C 32 (1985) 643.

[86]

T. Deforest, Jr. and J. D.Walecka, Adv. Phys. 15, 57(1966) 1.

83

REFERENCES

[87]

T. W. Donnely and I. Sick, Reviews of Modern Physics 56, 3 (1984)


461.

[88]

B. A. Brown, B. H. Wildenthal, C. F. Williamson, F. N. Rad, S.


Kowisiki, H. crannell and J. T. O' Brien, Phys. Rev. C 32 (1985) 1127.

[89]

A. Deshalit and H. Feshbach, "Theoretical Nuclear Physics", Vol.1:


Nuclear Structure, John and Sons, New York (1974).

[90]

P. M. Morse and H. Feshbach; "Methods of Theoretical Physics"; New


York, Me Grow-Hill, (1953).

[91]

B. A. Brown, R. A. Radhi and B. H. Wildenthal, Phys. Lett. B 133


(1983) 3.

[92]

J. D. Walecka, " theoretical nuclear and sub nuclear physics",


Cambridge, (Second edition), World Scientific Printers (S) Pte Ltd,
Singapore (2004).

[93]

L. J. Tassie and F. C. Barker; Phys. Rev. 111 (1958) 940.

[94]

H. Chandra and G. Sauer; Phys. Rev. C 13 (1976) 245.

[95]

J. P. Glickman, W. Bertozzi, T. N. Buti, S. Dixit, F. W. Hersman, C. E.


Hyde-Wright, M. V. Hynes, R. W. Lourie, B. E. Norum, J. J. Kelly, B.
L. Herman and D. J. Millener. Phys. Rev. C 43 (1990) 1740.

[96]

P. J. Brussard and P. W. M. Glademans, "Shell-model Application in


Nuclear

Spectroscopy",

North-Holland

Publishing

Company,

Amsterdam (1977).
[97]

M. Carchidi, B. H. Wildenthal and B. A. Brown. Phys. Rev. C 34


(1986) 2280.

[98]

D. B. Mcdevitt, ph. D. thesis, University of North Carolina State (2003).

[99]

A. Bohr, B. R. Mottleson, Nuclear Structure (World Scientific,


Singapore 1998, first published 1969).

[100]

E .B. Dally, M. G. Croissiaux, B. Schweitz, Phys. Rev. C 2 (1970)


2057.

84

REFERENCES

[101]

G. Berstch, J. Borysowicz, H. McManus, W. G. Love, Nucl. Phys. A


284 (1977) 399.

[102]

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

[103]

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

[104]

M.G. Mayer, J.H.D. Jensen, "Elementary Theory of Nuclear Shell


Structure", John Wiley. New York, 1955.

[105]

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

[106]

I. Talmi, Adv. Nucl. Phys. 27 (2003) 1.

[107]

A. Ozawa,T.Suzuli and I. Tanihata, Nucl. Phys. A 693 (2001) 32.

[108]

S. Cohen and D. kurath, Nucl. Phys. A 73 (1965) 1.

[109]

B. M. Preedom, B. H. Wildenthal, Phys. Rev. C 6 (1972) 1633.

[110]

D. J. Millener and D. K. Kurath, Nucl. Phys. A 255 (1975) 315.

[111]

E. K. Warburton and B.A. Brown, Phys. Rev. C 46 (1992) 923.

[112]

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

[113]

A. Hosaka, K.-I. Kubo and H. Toki, Nucl. Phys. A 244 (1985) 76.

[114]

B.A. Brown, B.H. Wildental, Annu. Rev. Nucl. Part. Sci. 38 (1988) 29.

[115]

B.A. Brown, W.A. Richter, Phys. Rev. C 74 (2006) 034315.

[116]

B. A. Brown, A. Etchegoyen, N. S. Godwin, W. D. M. Rae, W. A.


Richter, W. E. Ormand, E. K. Warburton, J. S. Winfield, L. Zhao, C. H.
Zimmerman, MSU-NSCL report number 1289 (2005) version.

[117]

T. R. Fisher, S. S. Hanna, D. C. Healey, and P. Paul, Phys. Rev. 176


(1968) 1130.

[118]

C. Forssen, R. Roth,P. Navratil, J. Phys. G: Nucl. Part. Phys. 40


(2012) 055105.

85

E2
) B (E 2; 21 01

( )10 A 20

psd p

( ,)10 A 14 16 A 20
. (0 2)

E2.

(0 2) 16 A 20

pf sd p s

) B (E 2;21 01

.

,

2 .
.M3Y

. , b ( A=10-14 )
( A=16-20 ) , )rms(
( ) ( (Rm . b
. b ,1.55 fm
. 1.78 fm
C

10,12, 14,16,18,20


. ) B(E2
.
. 20C 18C ,16C
.B(E2)
) B (E 2;21 01 () ( )
. 18C 16C
)B(E2
, - . 14,16,18C
C

14,16,18,20

0.25 e 1.1 e ,

E2
) B (E 2; 21 01



1998


. .

. .

You might also like