Quinn 2010 354
Quinn 2010 354
GERMANIUM-BROMINE REGION
A Dissertation
Doctor of Philosophy
by
Matthew A. Quinn,
April 2010
BETA DECAY HALF-LIVES OF NEUTRON RICH ISOTOPES IN THE
GERMANIUM-BROMINE REGION
Abstract
by
Matthew A. Quinn
Understanding the origin of the heaviest elements in the universe is one of the
great outstanding problems in physics. More than half of these elements are thought
to have been created by a process called the rapid neutron capture, or R process.
This process involves many nuclei that are very far from the line of stability. Because
these nuclei have such short beta decay half-lives, they are very difficult to produce
and study. The vast majority of r-process nuclei have not been observed at all.
This work details the measurement of several r-process nuclei that have never
been produced before. These nuclei lie in the r-process path near the Z = 40,
N = 60 region. This region of the chart of the nuclides is also interesting because a
rapid change in the shape of nuclei is seen. With the addition or removal of just a
the beta decay half life and beta-delayed neutron emission ratio Pn values are impor-
tant parameters. This experiment determined these values for the r-process nuclei
in the Germanium-Bromine region. The experimental work was carried out at the
88 89
As, and As, were measured for the first time. In addition, we have confirmed
previously measured half-lives for Y, Sr, Rb, Kr, Br, Se, and As isotopes. The new
92−90 96,94
measurements were used to calculate the Pn values for Br and Rb.
An r-process simulation was performed with the results from this measurement.
Because of the possible change in shape of the Se nuclei in particular, the uncertainty
factor of two. This work determined the shape of the Se isotopes in this region to
FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv
TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii
CHAPTER 1: NUCLEOSYNTHESIS . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Big Bang . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.1 Big Bang Timeline . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.2 Big Bang Nucleosynthesis . . . . . . . . . . . . . . . . . . . . 6
1.3 Stellar Nucleosynthesis . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3.1 Charged Particle Nucleosynthesis . . . . . . . . . . . . . . . . 7
1.3.2 S Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3.3 R Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3.4 R Process Abundance Observations . . . . . . . . . . . . . . . 15
1.3.5 R Process Models . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.3.6 R Process Sites . . . . . . . . . . . . . . . . . . . . . . . . . . 18
CHAPTER 3: EXPERIMENT . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.1 Goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.2 Experimental Efforts . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.3 Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.4 Particle ID . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.5 Detector setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
ii
3.6 Electronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
CHAPTER 5: DISCUSSION . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
5.1 Half-life Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
5.1.1 Comparison to Data . . . . . . . . . . . . . . . . . . . . . . . 91
5.2 Deformation Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 93
5.3 R-Process Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . 97
5.3.1 R-Process results . . . . . . . . . . . . . . . . . . . . . . . . . 98
BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
iii
FIGURES
iv
3.2 K1200 Dees and trim coils during installation. From NSCL image
library. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.3 Layout of the NSCL showing the K500 and K1200 cyclotrons, A1900
fragment separator and location of the experimental vault used for
this work. From NSCL image library. . . . . . . . . . . . . . . . . . . 47
3.4 SEGA germanium detectors surrounding the implantation plate. . . . 51
3.5 Schematic drawing of the setup of the SEGA detectors and aluminum
implant plate used for particle id. Pin 0 was used as the dE detector
in this setup, while the same scintillators (not shown) used in the rest
of the experiment provided the particle time-of-flight. . . . . . . . . . 52
3.6 BCS detectors. The four PIN detectors can be seen in addition to
the ribbon cable connected to the DSSD. . . . . . . . . . . . . . . . . 53
3.7 NERO detector. The inner ring of 3 He detectors and the outer rings
of BF3 detectors can be seen. Though not pictured here, the BCS
detectors sit inside central cavity during the experiment. . . . . . . . 55
3.8 Schematic drawing of the time-of-flight, BCS, and NERO detectors
used in this experiment. . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.9 Schematic of the beta-detection electronics setup. . . . . . . . . . . . 57
v
4.11 Decay curve for 93 Kr with mother, daughter, granddaughter and
background components. . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.12 Decay curve for 94 Kr with mother, daughter, granddaughter and
background components. . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.13 Decay curve for 95 Rb with mother, daughter, granddaughter and
background components. . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.14 Decay curve for 96 Rb with mother, daughter, granddaughter and
background components. . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.15 Decay curve for 96 Sr with mother, daughter, granddaughter and back-
ground components. . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.16 Decay curve for 97 Sr with mother, daughter, granddaughter and back-
ground components. . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.17 Decay curve for 98 Y with mother, daughter, granddaughter and back-
ground components. . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.18 Results for Y isotopes, along with previous measurements [?] and
FRDM-QRPA calculations [50]. Note that FRDM deformed and
spherical resulst overlap in some cases. . . . . . . . . . . . . . . . . . 77
4.19 Results for Sr isotopes, along with previous measurements [?] and
FRDM-QRPA calculations[50]. . . . . . . . . . . . . . . . . . . . . . 78
4.20 Results for Rb isotopes, along with previous measurements [?] and
FRDM-QRPA calculations[50]. . . . . . . . . . . . . . . . . . . . . . 78
4.21 Results for Kr isotopes, along with previous measurements [?] and
FRDM-QRPA calculations[50]. . . . . . . . . . . . . . . . . . . . . . 79
4.22 Results for Br isotopes, along with previous measurements [?] and
FRDM-QRPA calculations[50]. . . . . . . . . . . . . . . . . . . . . . 79
4.23 Results for Se isotopes, along with previous measurements [?] and
FRDM-QRPA calculations[50]. . . . . . . . . . . . . . . . . . . . . . 80
4.24 Results for As isotopes, along with previous measurements [?] and
FRDM-QRPA calculations[50]. . . . . . . . . . . . . . . . . . . . . . 80
4.25 ADC spectra for NERO quadrant A. Y axis is the number of counts in
each detector. X axis is the uncalibrated neutron energy. Detectors
1-4 are 3 He detectors, and show their characteristic spectrum, with a
sharp peak at the Q value of 0.764MeV. Behind that peak are events
where the entire Q value of the reaction was not deposited into the
detector. When the outgoing deuterons strike the tube walls a sharp
drop in the spectra can be seen. Below that, the events for which
the 3 H nuclei strike the tube walls can also be seen. Detectors 5-15
are BF3 detectors, which show a broader peak at the full Q value of
2.792MeV. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
vi
4.26 ADC spectra for NERO quadrant B. . . . . . . . . . . . . . . . . . . 81
4.27 ADC spectra for NERO quadrant C. . . . . . . . . . . . . . . . . . . 81
4.28 ADC spectra for NERO quadrant D. . . . . . . . . . . . . . . . . . . 82
4.29 TDC spectrum from NERO ring 1 detectors. Channels are propor-
tional to time neutron is detected after a beta decay event. Ring 1
is the innermost ring of detectors, and therefore sees more neutrons
shortly after beta decay with relatively few neutrons coming at later
times. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4.30 TDC spectrum from NERO ring 2 detectors. Channels are propor-
tional to time neutron is detected after a beta decay event. Ring 2
is the middle ring of detectors, and sees more neutrons at later times
than does ring 1. This is due to the fact that neutrons reaching this
ring have to travel through more moderating material. . . . . . . . . 83
4.31 TDC spectrum from NERO ring 3 detectors. Channels are propor-
tional to time neutron is detected after a beta decay event. Ring 3 is
the outermost ring of detectors, and sees the most neutrons at later
times of all the NERO rings. . . . . . . . . . . . . . . . . . . . . . . . 83
4.32 β background rate versus run number. . . . . . . . . . . . . . . . . . 84
4.33 β background rate versus DSSD front strip number. . . . . . . . . . . 84
4.34 Neutron background rate versus run number. . . . . . . . . . . . . . . 84
4.35 Neutron background rate versus DSSD front strip number. . . . . . . 85
4.36 Background neutrons per β as a function of run number. . . . . . . . 85
vii
TABLES
viii
CHAPTER 1
NUCLEOSYNTHESIS
1.1 Introduction
Understanding the elements that make up our universe has been a goal of hu-
mankind since ancient times. The Greeks gave us the concept of atoms (from the
Greek “Atomos” meaning un-cuttable); an idea which states that the universe is
today, and while we know that there are particles more fundamental than atoms,
we still can not say why atoms exist in the abundances that we see in nature. The
lightest nuclei, including Hydrogen, Helium, and Lithium are thought to have been
created shortly after the Big Bang. Therefore, elements heavier than hydrogen and
helium must be created inside of stars through thermonuclear reactions [1]. Light
nuclei fuse together to create heavier nuclei until a nucleus with maximal binding
Beyond the Iron region, nuclear fusion is unlikely to occur because of the high
electrostatic repulsion between nuclei. This fact seems to indicate that nuclei cap-
turing neutrons create the heaviest elements. Stellar abundance patterns of heavy
nuclei suggest that several neutron capture mechanisms may be responsible. Direct
evidence of any such mechanisms has been elusive. Models created to reproduce the
tron captures on seed nuclei. From these models two processes are responsible for
1
most of the production of heavy nuclei. They are called the rapid (r) and slow (s)
has been around for nearly fifty years [1], the sites where they occur still remain
unknown. The problem of determining how elements heavier than iron are created
remains one of the great unsolved problems in physics [2], and is at the heart of this
project.
the ngam code by Bradley Meyer. The ingredients of the code can be divided into
measure the evolution of nuclear structure in the region near A=90. This region is
of a nucleus affects the mass of the nucleus and its beta decay rates. Also, it is a
nearly impossible task to measure all or a significant portion of nuclei that may
occur along the r-process. We do however aim to measure as far as possible from
nuclear stability. The NSCL facility is one such place where we can produce some
This work seeks to measure the beta-decay halflives and beta-delayed neutron
emission ratios of new nuclei in the r-process path around the region of yttrium to
selenium. These nuclei are interesting because they lie in the r-process path just
past a bottle neck in the process at the nucleus 78 Ni, which can have a great influence
on the abundance peak at A = 130 and higher. The nuclei in this region may even
be the seed nuclei for the r-process. Their properties may reveal how the r-process
starts and how other astrophysical processes end. This work will show the effects
of these new halflives on the solar system abundances via an r-process abundance
are also a few nucleons away from nuclei that exhibit fantastic changes in shape.
2
The beta-decay halflives of these isotopes will give the first indication of whether
additional nuclei also change shape rapidly, and will further the understanding of the
mechanisms that govern this and other nuclear properties. This work will explore
the effects of these new measurements on the predictions made for nuclear shapes
in this region.
to review how the light elements that make up stars are thought to have been cre-
ated. The Big Bang is a model used to describe the evolution of the universe from a
singularity to its present (and future) state. Monsignor Georges Lemaitre, a physi-
cist and Catholic priest from Belgium, and Russian physicist Alexander Friedmann
the idea a “hypothesis of the primeval atom” [3]. Both Lemaitre and Friedmann’s
solutions stemmed from the fact that Einstein’s general theory of relativity did not
permit static solutions for the universe. This meant that instead of being perpet-
ually stable, the universe must have begun at some time in the past. From this
initial point, the universe exploded outward and grew into its present enormous
size. Two years before Lemaitre’s paper, Edwin Hubble had discovered that every
other galaxy is receding from our own at a speed proportional to its distance from
constant known as the Hubble constant. That all galaxies in the universe appeared
to be moving away from one another seemed to indicate that Big Bang concept was
correct.
In 1965 the discovery of the Cosmic Microwave Background radiation gave exper-
3
imental support for the Big Bang [5] theory. The CMB is a field of radiation, peaked
at 160.2GHz, that is seen isotropically throughout the universe. Its existence was
when the universe was only 379,000 years old. After the Big Bang the universe was
extremely hot, and took that long to cool enough for Hydrogen atoms to be formed.
The presence of the CMB and its extremely isotropic distribution could only be ac-
counted for by the Big Bang hypothesis. The Cosmic Background Explorer (COBE)
experiment, begun in 1989, measured the temperature and anisotropy of the CMB
with great precision. It revealed that the spectrum of the CMB was described nearly
perfectly by a blackbody emitter, as was predicted by the Big Bang theory. In ad-
dition to the spectral shape, COBE also revealed slight anisotropies in the CMB
distribution throught the universe. This anisotropy in the early universe is believed
to be what made it possible for matter condense into galaxies and solar systems
instead of being smeared out evenly throughout the universe. These results were
seen to be strong enough evidence for the Big Bang hypothesis that the Nobel Prize
was awarded to COBE scientists John Mather and George Smoot in 2006.
The Big Bang theory includes a detailed hypothesis of how the universe evolved
from a very short time after it came into existance [6]. In the first instant after the
Strong and, Weak, were indistinguishable. The universe was extremely hot and
dense, and was expanding outward. This time is known as the Planck Epoch, and
lasted only 10−43 seconds. At this time gravity separated from the other fundamental
forces. The universe had cooled enough so that two particles, the X and Y bosons
could be created. These theoretical particles could have decayed into two quarks
4
or one anti-quark and one anti-lepton. Decaying more favorably into two quarks
could explain why we see far more matter than anti-matter in the universe today.
This period, called the Grand Unification Epoch, lasted until 10−36 seconds after
After the Grand Unification Epoch, the universe underwent a rapid expansion.
Its size went from approximately 10−50 meters in diameter to about 1 meter in
about 10−33 seconds. A scalar field settling into a lower energy state caused inflation.
The result of the rapid expansion was that inhomogeneities and anisotropies present
in the universe were smoothed out, which is how we observe the universe today. At
the end of the Inflation Epoch the universe stopped its rapid expansion, and instead
converted the potential energy of the scalar field into radiation and particles such
as quarks, electrons, and neutrinos. The universe went from being largely empty to
The universe continued to expand and cool and eventually, at a time 10−12 sec-
onds after the beginning of the universe, its temperature was cool enough for the
electromagnetic and weak forces to decouple. At this time then all of the four funda-
mental forces existed in the manner in which we observe them today. The universe
was still too hot to allow the combination of quarks into heavier particles. It was
not until 10−6 seconds after the Big Bang that quarks could form heavier parti-
cles known as hadrons, which include protons and neutrons. From this time until
one second after the Big Bang hadrons dominated the universe, being produced in
hadron/anti-hadron pairs until the universe had cooled to the point where these
pairs could no longer be created. After that point, the hadrons and anti-hadrons
collided and annihilated each other, leaving only a small amount of hadrons re-
maining. From one second to three minutes after the Big Bang, particles known
5
as leptons were the dominant matter in the universe. At the three-minute point,
Although at three minutes, the universe was largely made up of radiation, the
building blocks of nuclei and atoms are thought to have present. The temperature of
the universe had lowered to the point where protons and neutrons could combine to
form heavier nuclei, and for only about seventeen minutes light nuclei were created.
This process, known as Big Bang Nucleosynthesis or BBN, while creating only a
few isotopes, is thought to have been responsible for the production of 98 percent of
the nuclei we see today. Protons and neutrons combined to form deuterium, which
often captured an additional proton and neutron to form 4 He. As can be seen in
Figure 1.1, small amounts of deuterium, tritium, 3 He, 6 Li, and 7 Li were created via
heavier elements from 4 He must take place through the combination of three 4 He
nuclei to form 12 C. This is called the triple α process, and is known to be extremely
slow. Because of the length of time the triple α process takes and because of the
small abundances of other isotopes, heavier nuclei were not created in the Big Bang.
After only seventeen minutes of activity, the universe had cooled to the point that
nuclear fusion was no longer possible. It would not be until 100 million years later
that any new nuclei would be created inside the first stars. But it was from the
nuclei created in the Big Bang that stars themselves would be created, and would
6
1.3 Stellar Nucleosynthesis
1.3.1 Charged Particle Nucleosynthesis
A look at the present elemental abundances of our solar system reveals clues
about the how those elements were produced in the past, and how they continue
to be produced today. Hydrogen and helium are the most abundant, as they were
produced in the Big Bang. Elements heavier than H, He, and Li are produced in
extremely low abundances. These elements were created only minutes after the Big
Bang, over 13 billion years ago. Changes to these abundances are the result of
The first stage of nuclear burning consumes hydrogen to create helium. This is
done through the p − p chain burning with proton-proton reactions, and through
CNO cycle proton reactions on carbon, nitrogen, and oxygen nuclei. The p −p chain
4p + 4e → 4 He
The process must proceed through intermediary reactions though, because the prob-
ability of four protons reacting at the same time in a stellar environment is essentially
p + p → 2 He
will not occur because the nucleus 2 He is unstable. This is also true for reactions
p + p + p → 3 Li
2
He + p → 3 Li
reaction:
7
p+p→d+e
is governed by the weak interaction and is therefore very slow. The mean lifetime
years. This is quite fortunate for us though, as it has allowed our sun to burn long
Deuterium burning takes place via several reactions, all of which are governed
by the strong force and proceed much more quickly than hydrogen burning. The
lifetime for a deuterium nucleus in a stellar environment is about 1.6 seconds. The
d + p → 3 He + γ
d + 3 He → 4 He + p
Therefore, burning usually proceeds through the 3 He nucleus, rather than the
d + d → 4 He
reaction. This cycle is called the p −pI chain. The process can also proceed through
the chains:
3
He + 4 He → 7 Be
7
Be + e → 7 Li
7
Li + p → 4 He + 4 He
and
3
He + 4 He → 7 Be
8
7
Be + p → 8 B + γ
8
B + e → 8 Be + ν
8
Be → 4 He + 4 He
These are known as the p−pII and p−pIII chains, respectively. The p−pI chain is
responsible for 86 percent of 4 He production, while the p − pII and p − pIII chains
The energy released in these nuclear reactions opposes the gravitational pressure
exerted on the stellar core by the outer layers. When hydrogen burning ceases, the
star will contract due to lack of thermonuclear energy release. The pressure on
the He in the core will increase until burning is possible. As 8 Be is unstable, α-α
12
reactions will not contribute to creating heavier nuclei. Stable C will be created
3α → 12 C
From the carbon nuclei produced in the triple alpha reaction, oxygen and neon
12
C + 4 He → 16 O
16
O + 4 He → 20 Ne
12
C + 12 C → 20 Ne + 4 He
12
C + 12 C → 23 Na + p
12
C + 12 C → 23 Mg + n
neon burning, and oxygen burning occur. The final stage of charged particle reac-
tions is silicon burning. Because of silicon’s high Coulomb barrier, direct burning
9
(28 Si + 28 Si) does not take place. Instead, as the temperature inside the star in-
creases, nuclei are photodisintegrated by the high energy gamma rays present. The
28
result is that Si and 4 He nuclei are available to interact and create 32
S. 32
S will
has a higher binding energy per nucleon than heavier nuclei. This means that further
charged particle reactions are unlikely to occur. In order to create heavier elements,
neutron captures will be the dominant mechanism. The two neutron capture pro-
cesses responsible for most of the heavy element production are called the s- (slow)
In addition to the s- and r- processes, there are other processes that contribute
to the total heavy element production [1]. This was realized because there are
190 168
proton-rich nuclei such as Pt or Yb that cannot be created by the neutron-
rich r- and s-processes. The p- (proton capture) process creates heavy proton-
rich nuclei by a series of γ-n or γ-α reactions on heavy seed nuclei. Thus, the
p-process is not responsible for creating heavy elements from lighter constituents
as in charged particle reactions or the s- and r-processes, but it does change the
for the photodisassociations, possible sites for the p-process include neutron star
mergers and supernova shock fronts [8]. The rp- (rapid proton capture) process
does create heavy nuclei from lighter seeds. It entails making heavy proton rich
decays. Sites for the rp-process are thought to include X-ray bursts on accreting
The results of Big Bang and stellar nucleosynthesis can be seen in the elemental
abundances present inside of our own sun. Nuclei created in the Big Bang and
10
in older generations of stars are still present today. Figure 1.2 shows these abun-
dances. It is important to note that these abundances show the effects of nuclear
structure. Especially stable nuclei like 4 He, 12 C, 16 O, and 56 Fe are produced in large
abundances. In addition the r- and s-process peaks occur at the points where these
abundances of the elements, we must understand the nuclear structure that guides
their creation.
1.3.2 S Process
The first process that dominates production of elements heavier than iron is the
neutron fluxes (compared to the r-process) and a path close to stability. For a stable
or slightly unstable nucleus of charge Z and mass A, the capture of a neutron can
will simply wait to capture an additional neutron (or neutrons) before undergoing
beta decay. If the product nucleus is unstable, it may beta decay to its Z+1 isobar,
or it may capture another neutron before beta decay has a chance to occur. The
neutron flux, beta decay half-lives and neutron capture cross sections determine how
long the process takes, where nuclei will be involved, and what the final abundances
will be. The observed solar abundances show large peaks at masses 90, 135, and
210. These are related to where the s-process path intersects the N = 50, 82, and
dNA (t)
= Nn (t)NA−1 (t)hσνiA−1 − Nn (t)NA (t)hσνiA − λβ(A) NA (t)
dt
11
The first term on the right side of the equation describes its production by neu-
tron capture on the lighter A-1 nucleus. The second term describes its destruction
by capturing a neutron to form the A+1 nucleus. The third term describes its
destruction by beta decay. With low capture cross sections and a large enough neu-
tron flux a steady flow will be achieved, with NA × hσνiA = const. The system of
equations can then be solved analytically by making the following substitutions [6]:
λn = 1/tnγ = Nn hσνiA .
λ1 >> λ2 . For the best fit to the data multiple exponential neutron exposures are
needed.
This model breaks down for cases where λ1 = λ2 . These are the so called s-
134 148 151 154 170
process branching points which include Cs, Pm, Sm, Eu, Yb , and
176
Lu. A network using the temperature dependent beta decay and neutron capture
correctly reproduce the solar system s-process abundances. The first is the Main
component. It is responsible for producing most of the nuclei with A 90-208. It takes
place in the He burning shells in low mass (< 4M⊙) asymptotic giant branch (AGB)
stars. The neutron source comes from 13 C “pockets” created in the H-He intershell.
Recurrent thermal instabilities above the He burning shell cause the H-He intershell
13
to become convective during the thermal pulse. Pockets of C are created in the
intershell by protons injected from the hydrogen envelope during each Third Dredge
12
Up (TDU) phase capturing on C. After the thermal pulse, convection mixes the
13 16
He, C and s-process material. The C+α → O + n reaction then produces the
12
nn = 108 cm3 for a period of about 20000yr.
The third component is the strong component introduced by Clayton and Rass-
208
back in 1967. It is responsible for producing of the Pb abundance, and occurs in
dence from grains of stardust found in meteors. These grains contain s-process
elemental abundances that agree closely with predicted s-process abundances [?] [?]
our solar system is taken as evidence that these grains really do come from outside
the solar system. This material may be reliably called s-process material, and indi-
cates that the s-process abundance predictions are largely reliable. Having reliable
the r-process abundances that this project and other works depend upon will be
1.3.3 R Process
The abundance peaks at masses 80,130, and 190 must be produced by a process
separate from the s-process. The intersection of these masses with the closed neutron
shells indicates that the nuclei involved in this process are much farther from stability
than the s-process nuclei. Because of this, the r- or rapid neutron capture process is
13
defined by beta-decay lifetimes that are longer than the time necessary to capture a
lifetimes that are much shorter than the neutron capture times. A path far from
stability with lifetimes ∼ 100ms necessitates a large flux of neutrons, which suggests
Burbidge, Burbidge, Fowler and Hoyle (B2FH) in 1957 theorized that the solar
however the astrophysical conditions and sites where an r-process may take plance
are still unknown [1]. Their paper describes how the double peaks seen in abundance
plots are due to the superposition of slow and rapid neutron capture processes. It
also shows how these peaks are due to the neutron shell closures that each process
path encounters. They were able to make calculations for the abundances from
these processes, and were able to estimate timescales for the process as well. These
classical models used 56 Fe as a seed for the r-process. Newer models attempt to take
into account the physics of the proposed site in a more realistic way. For models
using core-collapse supernova as the site of the r-process, the tremendous amount
of energy released by the collapse ( 1056 ergs) is enough to disassociate the nuclei
in the core into protons and neutrons. As the temperature drops, the protons and
neutrons combine to form alpha particles. Because of the high density, electrons
in the area are degenerate, and it becomes energetically favorable for protons to
capture electrons and form neutrons. Alpha particles fuse together to form heavier
nuclei, and in the presence of excess neutrons, can form neutron rich nuclei heavier
than iron. This scenario is called an ”alpha rich” freezeout, and provides seeds for
14
1.3.4 R Process Abundance Observations
not a trivial matter. The elemental abundances present in a given star include
many p-, s-, and r-processes. For the elements heavier than iron, the solar system
As the s-process path occurs close to the line of stability, beta-decay rates and
neutron capture cross sections can be measured for these nuclei. With this data
and information about the stellar environment, it is possible to generate the solar
abundance due to s-process nuclei. The r-process abundance is then taken as the
difference between the solar observed abundances and the s-process abundances.[10]
and s-process calculations have been treated by [11]. The resulting error bars are
small for r-process only elements such as Ba and Ir, while being larger for the
Recent observations of very old galactic halo stars have shed new light onto
the r-process, while creating new questions at the same time[12]. A class of stars
called Metal-poor halo stars show remarkable agreement with solar system r-process
abundances. These stars are called metal poor because they contain less iron relative
to hydrogen than our sun. Although in astronomy any element past helium can be
The log of the ratio of iron to hydrogen is generally referred to when determining
15
metallicity. Stars with values of less than one are considered metal poor (log F e/H <
−1).
Halo stars are thought to contain the remnants from only one or a few r-processes.
This is because they are very old and contain very little heavy element abundance.
These halo stars were produced from the remnants of the first stars in the galaxy and
would contain r-process material only from the nearby stars of that first generation.
Thus, unlike the solar abundances, they contain very pure r-process contributions
stars.
process abundances of elements with Z > 55 agree nicely with the solar system
abundances [13]. This suggests that the mechanism for creating the r-process el-
ements above Z = 55 is quite robust, and does not differ very much from event
to event. However for elements with Z < 55, there is a large under-production of
r-process elements in the halo stars. It is unclear then if the r-process outputs can
differ for different stellar conditions, or if there are two (or three) different types of
r-processes that occur with different temperature and neutron density conditions.
Suggestions of a weak r-process [15] or of a Light Element Primary Process [14] have
been made in order to describe these abundance patterns. It is clear, though, that
because Ge-Br isotopes are at the center of this disputed region, measuring their
Because the nuclei involved in the r-process are so far from stability, measure-
ments of even their most basic properties has been rare. Early models of the r-
16
process [17] [1] [16] set up a system of differential equations that describe the change
delayed neutron emission, photodisassociation, and fission for all nuclei involved in
the process.
This system of equations is greatly simplified if one assumes that the abundance
of each isotopic chain is contained largely in one nucleus. This is because the
neutron captures in an isotopic chain will occur before any of the nuclei will have
a chance to beta-decay. As more and more neutrons are added the last neutron
will be less strongly bound to the nucleus. At some point the gamma rays present
in this (presumably) violent scenario will photodisassociate the last neutron added,
preventing further neutron captures. The nucleus will then wait at a particular
mass until it beta-decays to the next isotope. This “waiting point” nucleus contains
the majority of the matter abundance for its isotopic chain. The waiting points are
determined by the temperature of the scenario and the binding energy of the last
neutron, which can be obtained from the nuclear mass. The series of differential
equations describing the r-process can then be simplified greatly, using only the
abundance for each isotopic chain. The dependence on the neutron capture rates is
also removed, meaning that the simulation will rely only on the beta-decay rates,
nuclear masses, neutron density, and temperature. The equation for the abundance
3/2 3/2
2πh2
Y (Z, A + 1) G(Z, A + 1) A+1 Sn (Z,A+1)
= nn e( kT ) .
Y (Z, A) 2G(Z, A) A mu kT
Where Y(Z,A) is the abundance of the nucleus A Z, Y (Z, A + 1) is the abundance of
A+1
the nucleus Z, G(Z, A) and G(Z, A + 1) are the partition functions for A Z and
A+1
Z, nn is the neutron number density, Sn is the neutron separation energy, and
T is the temperature. [19] and [18] showed what neutron density and temperature
17
range the waiting point approximation is valid for, and that this region successfully
1GK and neutron densities larger than 1020 /cm3 are required to successfully repro-
The location of the r-process is currently unknown. From an early time [1] it was
recognized that explosive scenarios such as core collapse supernova or neutron star
mergers were good candidates for the site of the r-process. This is because of the
need for high neutron number densities (> 1020 ) and high temperatures (> 1GK).
Core collapse supernova start out as massive stars (M = 8 − 100 M⊙). Through
nuclear burning they reach a point where their cores are composed of iron group
nuclei. Because the iron group nuclei have the highest binding energy per nucleon
further charged particle reactions will be endothermic. The core stops burning
and can no longer support the gravitational pressure from upper layers of the star,
which causes the core to collapse. As the temperature in the core increases, nuclei
are photo-disintegrated to alpha particles and bare nucleons. The density of the
core increases until it reaches nuclear density and a proto-neutron star is formed. A
shock wave rebounds from the now incompressible core and is sent out through the
Unfortunately, in models of the explosion, this shock stalls in the outer layers of
the star and does not produce an explosion that would transfer material throughout
the interstellar medium. Energy released by the proto-neutron star in the form of
neutrinos might be a solution to this shock stall, but the amount of energy deposited
the form of two- and three- dimensional simulations, with the inclusion of magnetic,
18
acoustic, convective, rotational, and asymmetric effects have been made with no
Perhaps the most promising explanation has been the neutrino-driven wind sce-
nario. In it, neutrinos from the cooling center of the proto-neutron star deposit
energy into its surface and ablate material.[20] [21] This high entropy (S > 400kb),
neutron rich material, called the neutrino driven wind, is driven from the proto-
neutron star into the low density region behind the shock. There it recombines
mainly through alpha captures to form neutron rich seed material heavier than
iron. The r-process then occurs with these neutron rich seeds instead of iron group
19
CHAPTER 2
NUCLEAR STRUCTURE
2.1 Structure
function of the number of protons and neutrons that make it up. Since the 1940’s
the shell model has been accepted as way to successfully describe the properties of
a nucleus[22]. Its largest success has been the description of the so called magic
numbers. These are the numbers of protons and neutrons for which exciting a
particle to the next excited state takes a large amount of energy. Starting with a
gives a bound solution with a non-zero ground state energy, and a series of excited
states with large energy gaps between certain numbers of nucleons. These energy
states are generated by solving for the eigenvalues of the Schrodinger equation, using
the square well potential in the Hamiltonian. The levels are filled in order due to
the Pauli principle, which says that two nucleons can not have the same quantum
numbers[23]. Unfortunately, the magic numbers produced are not the same as seen
in experiments, and the energy gaps are identical in energy also unlike what is
seen in experiments. The magic numbers 2,8,16,20,28,50,82, and 126 were able to
of the orbital angular momentum and a term proportional to the orbital angular
20
momentum dotted with the spin.
Even though the correct magic numbers are predicted by the shell model, it is
important to note that this is not the full story. The spacing and ordering of the
orbital levels changes as nucleons are added to a system. This is obvious from the
fact that the potential will change with the addition of a proton or neutron. In
addition to the change in the central potential, nucleons also have interactions with
other nucleons. These interactions depend on the overlap of the individual nucleon
orbits and their energy spacing. These interactions include not only proton-proton
the p-n interactions are largely responsible for the changes in the shape of a nucleus,
Further from stability the shell structure can change a great deal. Evidence for
21
the breakdown of the N = 8 shell gap has been seen near the neutron drip line in 12 Be
[24]. The emergence of a gap at N = 14 in 22 O has also been seen. Some predictions
have suggested that for heavier nuclei, the familiar magic numbers will be quenched
far from stability and new magic numbers will appear, while others indicate that the
old magic numbers will remain. Using r-process simulations some have shown that
some evidence for this effect[27], but there is still no conclusive picture. Determining
the role of shell closures and magic numbers in the neutron rich region is extremely
important for nuclear structure and nuclear astrophysics. New magic numbers or
the quenching of shell effects test our understanding of the nucleus and will change
22
Figure 2.2. Nilsson-model calculation showing evolution of single neutron energy
levels as l2 term is decreased. Decreasing this term in the modified harmonic oscil-
lator Hamiltonian represents more neutron rich nuclei, therefore the plot shows the
evolution of shell structure as nuclei become more neutron rich. The N = 50, 82,
and 126 shell gaps are all reduced in this calculation because the low j states above
the shell gaps are reduced in energy, while the high j state below the gaps are raised
in energy.
23
2.1.1 Vibrational Nuclei
As the number of valence nucleons increases, the number of residual p-n inter-
overwhelm the single particle shell model structure. The mixing of many nearly
degenerate levels produces one lowered level with a highly coherent wave function.
Such a state will have collective properties for the entire nucleus. Expanding the
residual interaction into multipoles will give monopole, dipole, quadrapole, and
higher terms. The electric quadrapole term plays a strong role in the low lying
nucleus to oscillate in shape. The situation can be understood by treating the nu-
cleus as a liquid drop and describing the shape of the nucleus with a deformation
2
1 1 dαλµ
H = Cλ Σµ |αλµ |2 + Dλ Σµ
2 2 dt
where Cλ is a parameter related to the surface and Coulomb energies of the
5
1/2
nuclear liquid drop, and Dλ = ρRλ0 . Cλ and Dλ are related by ωλ = D
Cλ
λ
, and
will produce a 2+ excited state. A second phonon can produce a triplet of states
twice the energy of the one phonon state. The ratio E4+ /E2+ or the ratio of the
energy of the first 4+ level to the energy of the first 2+ level is used as a guide
24
to identify vibrational collective nuclei. In addition to this vibrational structure,
enhanced collectivity will also lower the first 2+ state in relation to its position in
a closed-shell nucleus. Thus the energy of the first 2+ state can be used as a check
The addition of more nucleons to a nucleus already between shell closures will
produce more residual p-n interactions. These interactions will result in more config-
uration mixing, and a further departure from a spherical shape. Eventually even the
multipole vibrations will not be able to accurately describe the system, and a new
formalism is needed. These nuclei are seen as having a permanent deformed shape
in the ground state and excited states. Deformed nuclei can have quadrapole and
higher deformations, with the quadrapole being the most common at low energies.
Nuclei can either be prolate shaped (football) or oblate shaped (frisbee). Because
the deformed shape breaks angular symmetry the nucleus will be free to rotate. The
~2 2
H= J .
2I
where I is the moment of inertia, and J is the rotational angular momentum oper-
2
~2
∂H
I=
2 ∂J(J + 1)
25
in powers of J(J +1). For the simple case of considering the nucleus a rigid ellipsoid,
this becomes:
2
I = MR02 (1 + 0.31β)
5
4
p π ∆R
where β = 3 5 Rav
For the ground and low-lying states in even-even nuclei all of the angular mo-
mentum will be due to rotation. The expression for the energies of these states is
then:
h2
E= J(J + 1).
2I
The first few excited states will have energies of:
6h2
E2+ =
2I
20h2
E4+ =
2I
42h2
E6+ =
2I
Therefore, the ratio E4 /E2 will be 20/6 = 3.33. We can get an approximate idea
Spherical: < 2
Vibrational: 2
Rotational: 3.33
Because the deformation is a result of collective behavior, the first 2+ will also
26
2.1.3 Np Nn Scheme
it looks at the product of the number valence protons and neutrons[30]. It is assumed
that the onset of collectivity and deformation is due solely to the proton-neutron
interactions. The product of the number of protons and neutrons (or holes) counted
from the nearest closed shell is plotted versus the E2+ or E4+ /E2+ values for a series
of nuclei. These plots can show trends across regions that give important clues to
cleon can give good rules of thumb concerning the general nuclear structure. This
P = Np Nn /Np + Nn .
Plotting the E4+ /E2+ ratio versus P shows us that nuclei are deformed (E4+ /E2+ =
3.33) for P > 5, and may be deformed for P = 4. This means that for a nucleus to
The reordering of levels can have consequences for the ground state properties
as well. Nuclei in the A = 100, N = 60 region show a sudden and severe change in
the ground state E2+ levels, E4+ /E2+ ratios, quadrapole deformations, and half-lives.
This behavior is unlike the gradual change from spherical to vibrational to deformed
properties. The presence of two configurations, one spherical, one deformed account
27
proton-neutron interaction[31]. The monopole component of the proton-neutron in-
teraction does not depend on the angle between the proton and neutron orbits (as
radial distance between the proton and neutron orbits, so that the proton-neutron
interaction will be strong for orbits with the same orbital angular momentum re-
gardless of what their total angular momentum is. So for instance, the ν1g7/2 (and
ν1h11/2 ) orbit and the ν1g9/2 orbit will have a strong proton-neutron interaction,
which will have the effect of lowering the orbits. As the ν1g7/2 (or ν1h11/2 ) neutron
orbit is filled and the 1g9/2 proton orbit is lowered, the the shell gap at Z = 40 will
be reduced or destroyed.
If the nucleons fill the reordered shells from the bottom up as usual, the nucleus
will be spherical. However, if a pair of protons from the 2p1/2 orbit is excited
across the reduced shell gap a deformed configuration will result. This is because
the excitation of a proton pair creates more valence protons, which increase the
shell gap.
Deformed “intruder states” have been seen in other regions of the nuclear land-
scape [32]. In such regions near proton shell closures, the lowest state of the deformed
configuration is a 0+ level that is seen to lower in energy as neutrons are added to-
ward mid-shell. The 0+ state is lowest in energy at the middle of the neutron shell,
and increases in energy as neutrons are added past the middle of the shell. The
difference for nuclei in the A = 100 region is that the deformed 0+ state is lowered
beneath the 0+ spherical ground state. Thus the ground state becomes deformed,
The rapid transition to deformed ground states has been seen in Mo isotopes
28
above the Z = 40 shell closure at Zr, and Sr isotopes below. Models predict strong
Kr isotopes do not indicate strong deformation [33]. Because the beta-decay rates
nuclei in this region will help to indicate if the nuclei below Zr and Kr around
Figure 2.3. First 2+ energy levels in neutron rich nuclei around A = 100. A rapid
drop in energy is seen between N = 56 and N = 60. This drop corresponds to the
rapid onset in deformation due to the presence of intruder states.
29
Figure 2.4. Calculation of ground state energy as a function of deformation for
neutron rich Kr, Sr, and Zr isotopes. The onset of a deformed minimum for N = 60
in the Zr isotopes can be seen, while deformed ground states are predicted below N
= 60 for the Sr and Kr isotopes. This suggests that deformation may be seen in the
Se and Ge isotopes with N < 60. From [45]
30
Figure 2.5. Wood Saxon single particle neutron (top) and proton (bottom) energy
levels as a function of deformation parameter β. From [45].
31
2.2 Nuclear Physics Inputs for R-Process Simulations
2.2.1 R-Process Codes
tions for the abundances of each isotope. These equations can include terms for
neutrino captures, and fission. The network solves for the abundance of each iso-
tope given initial choices of parameters such as: seed nuclei, neutron number density,
56
temperature, and duration of the simulation. In classical r-process models Fe is
rich seed nuclei. The neutron rich nuclei in the Ge-Br region of interest in this
work are among these seed nuclei. For a canonical r-process simulation considering
neutron capture, beta decay, and beta-delayed neutron emissions each differential
Where Y (Z, A) is the abundance of the nucleus A Z, σ is the neutron capture rate for
the nucleus A Z-1. The first term on the right hand side of the equation includes beta-
decays and photodisassociations, while the second term includes neutron induced
reactions.
For large enough neutron number densities and high temperatures (i.e. photon
densities), neutron capture and photodisassociation will occur much more quickly
than beta decay. This situation is known as an (n, γ) ⇔ (γ, n) equilibrium, where
the nucleus with the maximum abundance for each isotopic chain must wait for
the slower beta decay before proceeding to the next isotopic chain. The change in
32
An equilibrium condition ensures that Ẏ (Z, A) = 0 The reaction rates λ between
neutron capture and photodisassociation are related by detailed balance, and are
given by :
3/2 3/2
2G(Z, A) A mu kT Sn (Z,A+1)
λA+1 = hσνiA e( kT ) .
G(Z, A + 1) (A + 1) 2πh2
where G is the partition function, T is the temperature, and Sn is the neutron
separation energy for the nucleus A+1 Z. Thus, the location of the waiting point nuclei
depends largely on the nuclear mass (via the Sn ), and the temperature. Combining
the abundance equation and rate equation above, the abundance ratios for each
addition to new values of nn and T. The simulation is ended when the neutron
number density or temperature drop below a certain level. Static network codes
calculate abundances the abundances in each isotopic chain for the full simulation
time, then factor in the contribution to each abundance due to beta decays. The
most important nuclear physics inputs then are the nuclear masses (Sn values),
beta-decay halflives, and the beta-delayed neutron emission ratios. For nuclei far
from stability these values are unknown and must be calculated. There are often
sured. Because the astrophysics surrounding the location and mechanism for the
r-process are unknown, accurate nuclear physics information can help to constrain
2.2.2 Masses
Because so few r-process nuclei have been measured, performing r-process sim-
ulations requires the use of predictions for nuclear properties. Nuclear masses are
33
perhaps the most important nuclear data needed for r-process calculations. Masses
determine the neutron separation energies, Q values for beta decays, and define the
neutron drip line. With the temperature, the nuclear masses dictate what the r-
process path will be. For a given temperature of 1 − 2 × 109 K, the r-process waiting
points will occur at nuclei with a Sn of 1.4MeV. Because very few masses have been
measured for r-process nuclei, abundance calculations rely heavily on nuclear mass
extrapolations or calculations.
The first mass model was the semiemperical model from von Weizsacker [34]
and Bethe and Bacher [35], which treated the nucleus as liquid drop. It contains
terms for nuclear volume, volume symmetry, surface energy, and coulomb energy
in addition to a surface symmetry term added by Meyers and Swiatecki [36]. Four
coefficients (all terms except Coulomb) are then adjusted to fit the measured mass
values and create a formula to predict unknown masses. This formula produces a
surprisingly good fit for only five terms with an overall rms deviation of 2.97MeV
[37], but fails badly for closed shell nuclei, with an error of around 10MeV.
Figure 2.6. Difference between nuclear mass from the Semi-empirical mass model
and experimentally measured nuclear masses as a function of neutron number. The
overall RMS fit is good, but large deviations at N = 28, 50, 82, and 126 can be seen.
From [37].
34
the liquid drop model. These macroscopic-microscopic models include the Finite
Range Droplet Model (FRDM) and the Extended Thomas Fermi plus Stutinski
Intergral (ETFSI) formulas. The FRDM [38] differs from earlier liquid drop mod-
els in that it takes into account the effect on the surface energy due to the finite
range of the nucleon-nucleon interaction. It also treats the effect of nuclear shape
on the Coulomb field and the diffuseness of the charge distribution, and introduces
overstate the central nuclear density. The microscopic corrections include a poten-
tial for use in the Strutinsky integral in order to introduce shell effects into the
The FRDM also uses a version of the BCS treatment for paring corrections, a
term to account for the energy difference between mirror nuclei. The results for
the FRDM are much improved over earlier liquid-drop models, as it produces a rms
The ETFSI mass formula [39] is based on the extended Thomas Fermi approx-
imation. Both the macroscopic and microscopic parts of the calculation center
around a Skyrme-type force. This means that unlike the FRDM calculations, the
microscopic and macroscopic parts of the calculation are done self-consistently. The
results from ETFSI calculation are similar to the purely microscopic Hartree-Fock
The latest mass formulas treat the nucleus microscopically, in that they attempt
to solve the Schrodinger equation for some choice of nuclear mean-field force. The
Hartree Fock formula solves the Schrodinger equation for an initial choice of the
mean field (usually either zero-range Skyrme or finite-range Gogny forces) and wave
35
functions. The equation is then solved iteratively in order to achieve self consistency
~2 2
− ∇ + U φi = ǫi φi
2M
Pairing effects are state dependent, and are added by either the BCS or Bo-
golyubov methods. The HFBCS and HFB mass models of the Brussels group [40]
produce an rms error of just over 700keV, but produce worse results for doubly
An additional mass model known as the P-Scheme model is based on the re-
lationship between the microscopic portion of the nuclear mass with the average
mass is obtained by taking the difference between measured nuclear masses and
as:
P = Np Nn /(Np + Nn ),
where Np is the number of valence protons or proton holes, and Nn is the number
of valence neutrons or neutron holes. Particles are counted when the shell is less
than half full, and holes are counted when the shell is more than half full. The
SEM is plotted against the P values for a series of nuclei in an isotopic chain, and
the resulting curve is fit with a polynomial. With the polynomial calculated from
known nuclei, the SEM value can be obtained for nuclei of unknown mass.
36
Figure 2.7. SEM vs P value for nuclei between Z = 28-50 and N = 50-82. The be-
havior of each element can be seen clearly. By fitting each isotopic chain, predictions
can be made for the masses of unmeasured nuclei [41].
This representation breaks down at or near the closed shells, where the number
of valence particles and holes goes to zero. In these cases, one can use a related
1
F0 = (Nπ Nν ),
2
where Nπ and Nν are the number of proton and neutron valence pairs. This
overcomes the problems of using the P parameterization at closed shells, but is re-
stricted to even-even nuclei, as only pairs of nucleons are considered. Again the SEM
value is plotted, this time against the F0 value for each nucleus, and a polynomial
37
2.3 Beta Decay Halflives
(or vice versa), while keeping the nuclear mass number A constant. A beta decay
also emits a beta particle (electron or positron) and a neutrino. There are several
types of beta decay: β + , β − , and electron capture. β + and electron capture occur
β+ : A A +
Z XN → Z−1 YN +1 + β + ν + Qβ
β− : A A −
Z XN → Z+1 YN −1 − β + ν + Qβ
Electron Capture : A − A
Z XN + e → Z−1 YN +1 + ν + Qβ
Q is the amount of energy released in the decay and corresponds to the difference
in mass between the parent and daughter nuclei. If Q is greater than zero, the decay
The beta-decay rate is a function of the transition strengths from the ground
state of the parent nucleus to the ground state and excited states in the daughter
m5e c4
λ= f (Z, E0)|ΣµMf hJf Mf ξ|Oλµ (β)|Ji Mi ζi|2
2π 3 ~7
where
f (Z, E0 )
38
is the Fermi function, which describes the effect on the beta particle’s wave function
d[N]
= −Nλ
dt
N
= eλ·t
N0
ln 2
t1/2 =
λ
where N0 is the number of nuclei originally present in a sample, and N is the
number of nuclei still present at time t. The emitted electron and neutrino both
have intrinsic spin = 1/2. There are two possible alignments of the spins then, 1, and
0. These different types of beta-decays are known as Fermi (spins anti-aligned) and
Gamow-Teller (spins aligned). If the beta particle and neutrino are emitted with
decays. The parity change for both types of decays is ∆π = (−1)ℓ so that for allowed
decays, the parity will be the same in the initial and final states. The change in
Therefore, the Fermi-type decays proceed mainly through isobaric analogue states
(IAS) in the daughter nucleus. Because of the large neutron to proton ratio in r-
process nuclei, the parent and daughter generally do not have analogue states near
one another, which means that Fermi type decay is unlikely to occur.
decays. Decays with ℓ = 1 are known as first forbidden decays and have ∆J = 0, 1
39
for Fermi-type decays, and ∆J = 0, 1, 2 for Gamow-Teller-type decays. The parity
change for both types is ∆π = (−1)ℓ = 1. Because of the higher angular momentum,
first forbidden decays are less probable to occur than allowed decays by a factor of
1000.
As with nuclear masses, the beta-decay rates for most nuclei involved in the r-
process have not been measured. This requires the use of beta-decay rate predictions
for use in r-process simulations. Also like with nuclear masses, beta-decay rates
have been calculated with either global or microscopic formulas. The global Gross
integrates the strength function over the low-lying excited states in the daughter
nucleus to get the half-life. Microscopic approaches based on the Random Phase
Approximation (RPA) [?] treat low lying excited states in the daughter nucleus
spherical or deformed, with the values of the half-life being strongly dependent on
a mass model. In our case, we have compared our measured values with a RPA
calculation using the FRDM mass model. Table 4.1 shows the wide range of half-
lives for the region of interest depending on the choice of ground state shape.
40
CHAPTER 3
EXPERIMENT
3.1 Goals
nuclear orbitals and magic numbers change far from the line of stability, and how
these changes in nuclear structure affect the r-process abundances. The neutron
rich isotopes in the region of Ge-Br approaching N = 60 sit in the middle of both
of these questions. It has been shown for lighter nuclei that the magic numbers
disappear and new magic number appear if enough neutrons are added[24].
As the structure of nuclei change, other nuclear properties also change. This
means that beta-decay half-lives, beta-delayed neutron emission ratios, neutron cap-
ture rates, and masses can all be very different from their values in more stable nu-
clei. Changes in these values will affect the abundances produced by the r-process.
There have been suggestions that the energy gaps of the neutron closed shells at
R-Process calculations done with mass models similar except for their shell
strength have shown that a significant abundance difference develops. The ETFSI
and ETFSI-Q mass models are similar calculations except that the ETFSI-Q model
used quenched shell strength, which has the effect of filling in the abundance troughs
around the A = 80, 130, and 190 peaks. Because the shells are not as strong the
beta-decay half-lives are shorter and material will flow through the region more
41
smoothly, eliminating the abundance troughs. For alpha-rich freezeout scenarios,
neutron rich nuclei in the Ge-Br region may even be the seeds for the r-process.
the presence of different “weak” and “main” r-process abundances for Z < 50 re-
quires additional nuclear data. Information about the neutron rich Ge-Br isotopes
will help to disentangle the contributions from these two different processes, in addi-
tion to any other processes that may contribute to heavy element production across
the universe.
42
3.2 Experimental Efforts
The production and study of nuclei far from stability requires specialized fa-
cilities. Generally, two methods of production are used: the In-Flight and Ion
beam of stable nuclei that impinges on and reacts with a target. Exotic nuclei are
and uninteresting products from the nuclei of interest. This method can produce a
wide range of proton- and neutron-rich nuclei. It is advantageous because the nu-
clei are produced with velocities near the primary beam velocity, meaning that the
production, separation and detection can be done very quickly. In addition there is
no dependence on the chemistry of the products as in the ISOL method. There are
number of very different facilities that produce nuclei via the In-Flight technique
Ion Separation On-Line uses a high energy proton beam and a thick target.
Exotic nuclei are produced via spallation in the target and must be extracted from
it. While this technique can not be used for the shortest-lived isotopes, it can
provide high intensities of certain elements, and has the added feature that the
isotopes produced may be easily reaccelerated for use in further reactions. ISOL
43
3.3 Setup
cility (NSCL) at Michigan State University. The NSCL uses two coupled cyclotrons
through an electric field. To produce a high energy beam either the field must be
very strong, or the particle must be passed through it many times. Cyclotrons work
by using a magnetic field to bend particles into a circular orbit so that they can be
The particle is injected into the center of the cyclotron and is accelerated in a
spiral shaped path. The electrodes in the center are known as “dees” because they
have the shape of the letter D. The dees are connected to an alternating current
which has a period equal to the orbital period of the particle. This is done so that
the particle will be accelerated through both gaps between the dees on each orbit.
The time for one orbit is the circumference divided by the velocity of the particle.
Where B is the strength of the magnetic field, m is the mass of the particle, and
ω = qB/m.
The magnets above and below the dees are held at a constant field, so that as
the particle is accelerated its orbital radius increases. The force due to the magnetic
field is:
44
F = qvxB = mv 2 /r
so that the force is perpendicular to the direction of the velocity and produces
r = mv/Bq.
When the (nonrelativistic) particle reaches the edge of the dees it is extracted
2
Tmax = 1/2mvmax = (Bqρ)2 /2m = (Bρ)2 /2(q 2 /m)
where ρ is the maximum physical radius of the dees. The term (Bρ)2 /2 is a
parameter that categorizes the capabilities of each particular cyclotron and is known
as the K value for the cyclotron. This is where the names of the two cyclotrons at
NSCL, the K500 and the K1200, get their names. The K value is often given in units
At NSCL the cyclotrons have magnets broken up into high- and low-field sectors.
This is done in order to provide vertical focusing for the beam. Additionally, trim
coils are added to increase the magnetic field, in order to deal with the increase in
mass of the beam particles due to relativistic velocities. The ratio B/m appears in
the equation governing the cyclotron resonant frequency, so for an increase in mass
of:
m0
m= p
(1 − (v/c)2)
45
the magnetic field must be increased by the trim coils in order to maintain the
cyclotron frequency.
Figure 3.2. K1200 Dees and trim coils during installation. From NSCL image
library.
The first cyclotron, the K500, accelerates the primary beam of 136 Xe to an energy
of 10.85MeV/U. The beam is then passed through a stripper foil and injected into
the second cyclotron, the K1200. Next, the second K1200 cyclotron accelerates the
beam to an energy of 120MeV/U. The average intensity of the primary beam was
2pnA.
The primary beam strikes a 240mg/cm2 Be target and fragments into many
lighter neutron-rich nuclei. The fragments are produced with a kinetic energy that
is a large fraction of the primary beam, and proceed in forward direction to the
A1900 fragment separator [48]. Along with the fragments there is also degraded
primary beam that must be removed by the separator. This method of producing a
46
and transport of radioactive nuclei, so that isotopes with very short half-lives may
be studied. Additionally, unlike the slower ion separation on line (ISOL) technique,
Projectile Fragmentation does not depend on the chemistry of the primary beam
or secondary products. While not utilized in this experiment, the radioactive beam
could be used on a secondary target to produce even more exotic tertiary products.
Figure 3.3. Layout of the NSCL showing the K500 and K1200 cyclotrons, A1900
fragment separator and location of the experimental vault used for this work. From
NSCL image library.
In order to separate nuclei of interest from unwanted fragment nuclei and unre-
acted primary beam, a magnetic separator system called the A1900 is used at the
and forty quadrapole and higher multipole magnets used to focus the beam. A
dipole magnet with radius (r) and field strength (B) , bends an ion with a given
γmv = qBr
Two dipoles will produce a focused beam with a momentum to charge ratio
47
determined by the magnetic rigidity value obtained for a particular choice of the
magnetic field B. The location where this focus occurs is called the intermediate
focal plane. Nuclei with a different momentum to charge ratio are blocked by slits
at the focal plane and will not pass through. Nuclei produced in the fragmentation
of nuclei with particular mass to charge ratios. To separate nuclei with similar m/q
ratios but different Z than the nucleus of interest, a degrader is also placed at the
focal plane. The energy loss for each nucleus in the degrader is given by the Bethe
formula:
dE 4πe4 z 2
− = NA ρB(v)
dx m0 v 2 A
with
2m0 v 2 v2 v2
B(v) = Z ln − ln 1 − 2 − 2
I c x
where z is the charge of the nucleus, v is its velocity, ρ is the density of the degrader
Therefore each isotope will have a different momentum after passing through the
degrader. A second set of dipoles will again select a particular momentum, filter out
unwanted nuclei and transmit those that are of interest. For this experiment the
magnetic rigidity of the dipole magnets was tuned to pass 88 Se. A position sensitive
plastic scintillator was placed at the dispersive focal plane to act as a degrader,
and to provide a measure of the horizontal position of each isotope. The scintillator
was also used as the start for time-of-flight measurements. In addition an aluminum
“finger” was placed at the dispersive focal plane in order to block 136 Xe charge states
that were present there. Nuclei that were passed through the A1900 were then sent
48
3.4 Particle ID
To identify individual nuclear species, two methods were used. The first was
Isomers are states in a nucleus that exist for a long time before decaying via gamma
decay. The decay emits a gamma ray of a unique energy that can be used as a
unique signature from one particular nucleus. Once one nucleus was unambiguously
identified, the rest could be identified through a ∆E − T oF plot. This sort of plot
separates nuclei according to A and Z. The energy loss of a nucleus passing through
a detector is also given by the Bethe formula, so that for a nucleus fully stripped of
electrons:
dE ∝ Z 3 /A
T oF = d/v = dA/p
where d is the distance from the the A1900 focal plane to the N3 experimental vault
and p is the momentum of the nucleus. Thus a ∆E − T oF plot will separate nuclei
by Z and A, which enables gating on a single nucleus in order to study its beta
decay properties.
Isomeric decays are used for the ID because the distance from the target to the
N3 vault is approximately 40 meters. The time of flight for the secondary beam will
then be around 500ns. This time delay ensures that prompt gamma decays have
already occurred, but that isomeric states will still be populated. An isomeric level
of over 1µs was desired to ensure a sufficient counting rate. For this experiment a
99
8.6µs isomeric state in Y was used. With this isotope identified, the remaining
49
nuclei could be identified as well.
In order to measure the 126keV gamma ray from the isomeric state of 99 Y, three
HPGe detectors were used. These detectors come from the Segmented Germanium
work by transmitting the energy of an incoming photon to the valence band electrons
mainly by Compton scattering. The increased energy of the electrons promotes them
to the conduction band, where they can be collected by an anode at a high potential.
Germanium detectors are preferable because they have a high detection efficiency
A different implant station was used for this part of the experiment. It consisted
of a thick aluminum plate placed perpendicular to the beam in the N3 vault, with
a Silicon PIN ∆E detector placed in front of the plate. A PIN detector is a silicon
diode made up of three layers: a p-type (hole doped) layer, an intrinsic (no doping)
layer, and an n-type (electron doped) layer. The intrinsic layer acts as a large
depletion region in which electrons and holes are created by the incoming radiation.
A reverse bias is applied to the detector, which causes the electrons to move toward
the p-type material where they are collected by a terminal. These electrons compose
The three Ge detectors were placed around the beam pipe at the location of the
plate. Nuclei passed through the two plastic scintillators giving a T oF signal, and
through the PIN detector giving a ∆E signal. They were then implanted into the
aluminum plate, without being implanted into the Beta Counting System (BCS).
Once implanted, the isomeric states decayed and were detected by the Ge detectors.
The ∆E−T oF spectrum could be gated on events that produced the 126keV gamma
99
rays. This gated spectrum then showed which events were from Y.
50
Figure 3.4. SEGA germanium detectors surrounding the implantation plate.
In the N3 vault, nuclei were implanted into the BCS. The BCS consists of four
silicon PIN detectors, one double sided silicon strip detector (DSSD) and one single
sided silicon strip detector (SSSD)[49]. The DSSD and SSSD detectors are similar
to the PIN detectors except that they include multiple terminals that are separated
from one another. Therefore these detectors are able to be read out in individual
strips. Our setup also included a 1000µm thick Ge detector behind the silicon
detectors. The Ge detector was tested for use in identifying gamma rays from
implanted nuclei, but this proved to be ineffective. Instead, the Ge detector was
51
Figure 3.5. Schematic drawing of the setup of the SEGA detectors and aluminum
implant plate used for particle id. Pin 0 was used as the dE detector in this setup,
while the same scintillators (not shown) used in the rest of the experiment provided
the particle time-of-flight.
The four PIN detectors had thicknesses of 991µm, 303µm, 309µm, and 966µm.
The DSSD had a thickness of 979µm, and the SSSD had a thickness of 988µm.
These thicknesses were chosen so that nuclei were implanted into the DSSD, which
was segmented into forty 1mm strips on the front and forty perpendicular 1mm
strips on the back, giving 1600 pixels. The output from the DSSD was sent to a
nuclei and beta decays. Each implant or decay event was recorded with the pixel
location and time stamp in addition to the energy of the event. The SSSD was made
up of sixteen strips and was used to veto light particle implantations not stopped
in the DSSD, as well as detect beta decay events that did not deposit all of their
52
energy in the DSSD. Typically implantation rates were 1 × 10−3 implants/s/pixel.
Figure 3.6. BCS detectors. The four PIN detectors can be seen in addition to the
ribbon cable connected to the DSSD.
In order to detect the beta delayed neutron, the Neutron Emission Ratio Ob-
server (NERO) detector system was placed around the BCS setup. NERO is made
and 24(outer) BF3 detectors. The design was chosen to maximize the detection ef-
53
ficiency of the detector. Neutrons are detected via the 3 He(n, d)3 H and 10 B(n, a)7 Li
reactions. The cross section for these reactions is highest in the thermal neutron re-
the polyethylene block they are seated in. The anode of the detectors measures the
energy from the deuteron and alpha particles respectively. The energy of the neu-
tron is much smaller than the Q value for these reactions, which are 0.764MeV(3 He)
and 2.792MeV(BF3 ), so the energy detected is essentially these Q values. Thus the
energy information the neutrons is lost, but we are concerned with the number of
neutrons detected, not their energy. When a beta decay was detected in the DSSD,
NERO was gated to look for neutrons for 200µs. This amount of time was chosen
to allow sufficient time for the neutrons to be moderated by the polyethylene block.
3.6 Electronics
The experimental logic is shown in Figure 3.11. The master trigger for the
experiment was either a signal in PIN1 from an implant nucleus or a beta decay
event from the DSSD. An additional master trigger was created from the three
germanium detectors timing signals when the SeGA setup was used. The master
trigger live was created by using the master trigger and a NOT busy signal from
the acquisition computer, and was used to trigger the acquisition computer and to
Signals from the PIN detectors were sent to preamplifiers and then to amplifiers
with energy(slow) and timing(fast) outputs. The energy outputs were sent to VME
ADC modules, while the timing outputs were sent to a VME TDC, a scaler module
Energy signals from the SeGA detectors were sent to a spectroscopy amplifier
and then to a CAMAC ADC. The timing signals were sent to a scaler module and
54
Figure 3.7. NERO detector. The inner ring of 3 He detectors and the outer rings of
BF3 detectors can be seen. Though not pictured here, the BCS detectors sit inside
central cavity during the experiment.
The signals from the forty front DSSD strips and the forty back DSSD strips
were sent in three groups (1-16, 17-32, 33-40) to Multichannel Systems preamplifiers.
These preamplifiers were able to output signals with two different gains in order to
distinguish between implant events and beta decay events. The low gain signals
were sent directly to VME ADCs, while the high gain signals were sent first to
shaper/discriminator modules and then to VME ADCs and scaler modules. The
high gain signals were also combined in a logical OR and sent to a VME bit register.
This bit was used to readout DSSD strips that had a signal from one strip in their
55
Figure 3.8. Schematic drawing of the time-of-flight, BCS, and NERO detectors used
in this experiment.
NERO signals were read out in four quadrants of fifteen detectors each. The
shaper signals were sent to a VME ADC. The fast discriminator signals were sent to
VME TDC and scaler modules. The NERO master gate was created by the master
gate generated by a beta decay event in the high gain DSSD signals. A master gate
live was made by the AND of the master gate, a 200µs gate generated after a master
gate live signal, and a not bust signal from the acquisition computer. This master
gate live was used to set a window to readout the NERO ADCs so that neutrons
56
Figure 3.9. Schematic of the beta-detection electronics setup.
57
CHAPTER 4
DATA ANALYSIS
Particle ID was done with the dE-ToF technique. Both the energy and time-of-
had a momentum acceptance of about 4%. The energy loss, when velocity corrected
for the time of flight, gives elemental (Z) separation. The time of flight, when
momentum corrected via the focal plane position, gives isotopic (A) separation.
dE 4πe4 z 2
− = NA ρB(v)
dx me v 2 A
With:
2me v 2 v2 v2
B(v) = Z ln − ln 1 − 2 − 2
I c x
where e is the electron charge, Z is atomic number of the detector material, me is
the electron mass, v is the velocity of the projectile nucleus, A is the atomic weight
of the detector material, NA is Avogadro’s number, rho is the mass density of the
58
detector material, and I is the average excitation and ionization of the detector
material.
Because the energy of the projectile nuclei has a velocity dependence, a correction
r
dE me v 2 A 4 2
Z= πe z NA ρB(v)
dx 4
In the experiment the time-of-flight was measured, not the projectile velocity, so
the simple relation v = D/t is used to obtain the velocity. The distance D is the
total flight path from the focal plane of the spectrometer to the experimental vault,
which is essentially the same for all nuclei. A linear correction to the energy was
dEcorr = dE − M × T oF + B
where M is the slope and B is the intercept. The raw dE-ToF spectrum is shown
59
Figure 4.1. PIN1 vs ToF, both uncorrected.
60
4.1.2 Time of Flight
of each isotope. From the following equation we can see the mass and momentum
T oF = D/v = Dm/p
The momentum of nuclei is not directly measured in our experiment, but the
the equation:
Bρ = p/q
so that
mD
T oF =
Bρq
The 4% acceptance of the spectrometer means that there is a distribution in
time-of-flight for each mass. Using a correction for the momentum gives a unique
ToF for each mass m. We used a linear correction for the momentum dependence
of the time-of-flight. This is based on the horizontal position of each nucleus at the
61
Figure 4.3. ToF vs focal plane position for Yttrium isotopes. Tof uncorrected. Lack
of counts between the two sets is due to the presence of aluminum “finger” to block
out unreacted primary beam.
Figure 4.4. ToF vs focal plane position for Yttrium isotopes. Tof corrected for
momentum dependence.
62
4.2 Gain Matching
The high gain outputs of the DSSD and SSSD were gain matched using a 241 Am
source before the experiment. The source was placed just in front of the DSSD and
SSSD.
241
Figure 4.5. Spectrum from Am source.
4.3 Thresholds
Low energy thresholds for the high gain outputs of the DSSD and SSSD were
90
set using a Sr source. Figure 4.6 shows the spectrum from a typical DSSD strip.
90
Figure 4.6. Spectrum from Sr source.
Implant events were defined as events that had a signal in PIN1, and low gain
signal in the front and back of the DSSD. Events that produced a low gain signal in
63
the SSSD were rejected. For events with a signal in more than one strip on the front
or back sides, the strip with the maximum energy was recorded as the strip in which
the event took place. The pixel location (front and back strip numbers) and the time
of implantation were recorded for each event. Decay events were defined as events
that did not have a signal in PIN1 or PIN2, and produced a low gain signal in the
front and back of the DSSD. Events that produced an overflow signal in the small
Ge detector behind the SSSD were rejected. This was done to reject light particles
coming with the beam that could be confused with beta decay events from implant
events. For decay events, like with implants, when a signal in more than one strip on
the front or back sides occurred, the strip with the maximum energy was recorded
as the strip in which the event took place. To be considered as decays correlated
with implantations, decay events had to come within a ten seconds of an implant.
Decay events to one pixel away from the implant were considered valid. Up to three
decay events (ie. mother, daughter, granddaughter) within the ten second window
were recorded. The decay time for each decay event was the difference between the
implantation event time and the decay event time. A window from forty seconds
after an implant event in a particular pixel until the next implantation in the same
pixel was used to determine background. Any decay event that occurred in this
hydrogen-like nuclei with mass A+2,3 the total kinetic energy of each nucleus was
looked at. In our experiment the pre-amplifiers for the DSSD were saturated for
many implant nuclei, thus the DSSD could not be used to determine the total kinetic
energy. However,the sum of the energy in the four PIN detectors was sufficient to
64
distinguish between fully stripped nuclei and charge state contaminants. Gates were
drawn in the image 2 focal plane position versus the PIN energy sum plot in order
Figure 4.7. PIN1,PIN2,PIN3, and PIN4 energy sum vs focal plane position for 99 Y.
Charge state contaminants have a lower total kinetic energy, and were therefore
excluded by the gate shown.
made on the PIN1 vs PIN2 spectrum. Figure 4.8 shows this plot, and the gate used.
Figure 4.8. PIN1 energy vs PIN2 energy. Counts outside the gate were excluded.
65
4.4.3 Time-of-Flight Gate
Two photo tubes, one on each side of the image 2 scintillator were used to
measure the time-of-flight. Because of a high primary beam rate at the image 2
focal plane, the image 2 scintillator began to fail during the experiment. This was
manifest through failing to produce a signal for some particles. This resulted in
some events with spurious ToF signals. These events were rejected by creating a
gate in the N-ToF vs S-ToF spectrum. This spectrum is the time-of-flight from one
photo tube (N-ToF) versus the time-of-flight from the other photo tube (S-ToF).
Figure 4.9. N-ToF vs S-Tof. Counts outside of the gate shown were excluded.
For the isotopes with high statistics, the beta-decay half life could be obtained
through a fit of the decay curve. The decay times for each isotope were put into
66
bins of 100-400 ms to form decay curves. The curves were then fit using exponential
equations give the activity at a time t, and are known as Bateman equations.
n
X
An = N0 ci e−λi t
i=1
where
Qn
λi
i=1
cm = Qn
i=1 ′(λi − λm )
The expression for the total number of decays seen at time t becomes:
with
C1 = N0
λd N0
D1 =
λd − λp
λd N0
D2 =
λp − λd
λd λg N0
G1 =
(λd − λp )(λg − λp )
λd λg N0
G2 =
(λp − λd )(λg − λd )
λd λg N0
G3 =
(λp − λg )(λd − λg )
where λp is the decay constant of the parent nucleus, λd is the decay constant of the
daughter nucleus, and λg is the decay constant of the granddaughter. B is the value
minimizing the chi-squared of the fit, with the parent halflife, initial parent scaling,
and background being free parameters. The daughter and granddaughter halflives
67
Halflife of Br-91 h2
Entries 49
Mean 3133
Counts
RMS 2893
120
100
80
60
40
20
0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Time (ms)
Figure 4.10. Decay curve for 91 Br with mother, daughter, granddaughter and back-
ground components.
Halflife of Kr93 h2
Entries 49
Mean 3655
Counts
RMS 2797
200
180
160
140
120
100
80
60
40
20
0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Time (ms)
Figure 4.11. Decay curve for 93 Kr with mother, daughter, granddaughter and back-
ground components.
68
Halflife of Kr94 h2
Entries 99
Mean 1364
Counts
RMS 1381
70
60
50
40
30
20
10
0
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Time (ms)
Figure 4.12. Decay curve for 94 Kr with mother, daughter, granddaughter and back-
ground components.
Halflife of Rb95 h2
Entries 99
Mean 1310
Counts
RMS 1353
600
500
400
300
200
100
0
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Time (ms)
Figure 4.13. Decay curve for 95 Rb with mother, daughter, granddaughter and back-
ground components.
69
Halflife of Rb96 h2
Entries 99
Mean 1416
Counts
220 RMS 1418
200
180
160
140
120
100
80
60
40
20
0
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Time (ms)
Figure 4.14. Decay curve for 96 Rb with mother, daughter, granddaughter and back-
ground components.
Halflife of Sr96 h2
Entries 28
1800 Mean 3450
Counts
RMS 2815
1600
1400
1200
1000
800
600
400
200
0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Time (ms)
96
Figure 4.15. Decay curve for Sr with mother, daughter, granddaughter and back-
ground components.
70
Halflife of Sr97 h2
Entries 99
Mean 1546
Counts
RMS 1401
700
600
500
400
300
200
100
0
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Time (ms)
97
Figure 4.16. Decay curve for Sr with mother, daughter, granddaughter and back-
ground components.
Halflife of Y98 h2
Entries 49
1200 Mean 1599
Counts
RMS 1403
1000
800
600
400
200
0
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Time (ms)
98
Figure 4.17. Decay curve for Y with mother, daughter, granddaughter and back-
ground components.
71
4.5.2 Beta Detection Efficiency from Decay Curve Fits
The DSSD beta detection efficiency could be determined from the parameters of
the decay curve fits. This is an input parameter for the Maximum Likelihood fit.
The efficiency is taken as the number of parent beta decays detected divided by the
number of parent implants detected. The number of implants can be simply counted
from the particle-id spectrum, but because the number of beta decays detected for
total number of decays for each isotope gate can not be used for calculating the
DSSD efficiency. The number of parent decays can be obtained from the parent
(Ao )f it
NDparent =
λf it
therefore
NDparent
ǫβ = 100 ×
NI
For the isotopes with few events, fitting a decay curve is not possible, so another
method of determining the half-lives was needed. We used the Maximum Likelihood
Method (MLH), which has been used before for cases of poor statistics[?][?]. This
method uses the Bateman equations to determine the probability that beta decays
with a given half-life will occur at the times seen in the experiment. The half-life
ter, granddaughter, and background events for a particular value of the beta decay
72
constant. The beta detection efficiency, beta decay background, daughter and grand-
daughter half-lives, beta delayed neutron emission ratios (Pn ), and delayed neutron
function for the observance of zero, one, two, or three decays at a given time t. The
probability density for a parent decay at a time t, given the decay constant λ1 is:
f1 (λ1 , t) = λi e−λ1 t
Z t
F1 = f1 (λ1 , t′ ) dt′ = 1 − e−λ1 t
0
The probability density for observing a daughter decay with a decay constant
λ1 λ2
f2 (λ1 , λ2 , t) = (e−λ1 t − e−λ2 t )
λ2 − λ1
The probability for that daughter decay to occur within a time t is:
λ1 λ2 1 −λ1 t 1
F2 (λ1 , λ2 , t) = 1 − ( e − e−λ2 t )
λ2 − λ1 λ1 λ2
The probability density for observing a granddaughter decay with a decay con-
stant λ3 from a granddaughter with decay constant λ2 and a parent with decay
constant λ1 is :
λ1 λ2 λ3
f3 (λ1 , λ2 , λ3 , t) = ×
(λ2 − λ1 )(λ3 − λ1 )(λ3 − λ2 )
(λ3 − λ2 )e−λ1 t − (λ3 − λ1 )e−λ2 t + (λ2 − λ1 )e−λ3 t
The probability for that granddaughter decay to occur within a time t is:
73
λ1 λ2 λ3
F3 (λ1 , λ2 , λ3 , t) = 1 − ×
(λ2 − λ1 )(λ3 − λ1 )(λ3 − λ2 )
(λ3 − λ2 ) −λ1 t (λ3 − λ1 ) −λ2 t (λ2 − λ1 ) −λ3 t
e − e + e
λ1 λ2 λ3
Additionally, the probability for observing exactly r background events in a time
(btc )r e−btc
Br =
r!
To perform the MLH analysis, the probability of actually seeing a decay must be
calculated. This includes the probability for decay events to occur as shown above,
as well as incorporating the efficiency of detecting the beta decays. This can be
shown simply for the detection of no beta decays within a correlation time tc . The
equations for one, two, or three observed beta decays follow in the same manner but
are too lengthy to be listed here. To determine the probability for detecting a beta
decay, we must look at the probability that a decay from a particular generation i
happens Di , and the probability that it is observed Oi . The efficiency for detection
event comes from four terms: 1) that a parent does not decay, 2) that a parent decays
but is not observed and the daughter does not decay, 3) that the parent decays and
is not observed, the daughter decays and is not observed, and the granddaughter
does not decay and 4) the parent, daughter, and granddaughter decay but are not
observed.
74
P0 = (D 1 + D1 O1 D2 + D1 O 1 D2 O2 D 3 + D1 O1 D2 D 2 D3 O3 ) × B0
Plugging in the probability functions F from above and the detection efficiencies,
By including the possibility that one, two, or three neutrons may be emitted
For each isotope it is possible to calculate the background from the number of
events where no decay was observed N0 , relative to the number of events where at
N0 P0
=
N123 1 − P0
N0 and N123 are both measured in the experiment, N0 is the number of implants
for which no decay event is seen, while N123 is the number of implant events for
which at least one decay is seen. The background rate for r = 0 is then:
75
− ln B0
b=
tc
Using the equation for P0 , one can calculate the background rate from the num-
N0
B0 = × (1 − F1 ǫ1 − F2 ǫ1 ǫ2 − F3 ǫ1 ǫ2 ǫ3 )−1
N0 + N123
The total likelihood function for decay events is the sum of the probability func-
N
Y 123
4.6 Errors
4.6.1 Statistical Error Contributions
The reported errors include both statistical and systematic error contributions.
The statistical error is obtained directly from the Maximum Likelihood analysis,
which has been used before in cases of poor statistics. Because the likelihood prob-
ability density is not symmetric about the maximum, an interval was chosen such
that the interval distance was the minimum to contain 68% of the area under the
curve.
The uncertainties in the half-lives and Pn values of the daughter and granddaugh-
ter as well as the N-1 daughter and granddaughter contribute to the systematic error
76
for the parent half-life calculation. In addition, the background and beta-detection
The contribution of these factors to the overall error was obtained by varying
each of these parameters within their own uncertainty limits. The plus and minus
one sigma values for each were used for each parameter, while the nominal values
Y isotopes
Present values
1500 Previous values
FRDM deformed
FRDM spherical
Halflife (ms)
1000
500
0
98Y 99Y 100Y 101Y
Nucleus
Figure 4.18. Results for Y isotopes, along with previous measurements [?] and
FRDM-QRPA calculations [50]. Note that FRDM deformed and spherical resulst
overlap in some cases.
77
Sr isotopes
1500
Present values
Previous values
FRDM deformed
FRDM spherical
Halflife (ms)
1000
500
0
96Sr 97Sr 98Sr 99Sr
Nucleus
Figure 4.19. Results for Sr isotopes, along with previous measurements [?] and
FRDM-QRPA calculations[50].
Rb isotopes
3000
2700 Present values
Previous values
FRDM deformed
2400 FRDM spherical
Halflife (ms)
2100
1800
1500
1200
900
600
300
0
94Rb 95Rb 96Rb 97Rb
Nucleus
Figure 4.20. Results for Rb isotopes, along with previous measurements [?] and
FRDM-QRPA calculations[50].
78
Kr isotopes
2000
1800 Present values
Previous values
FRDM deformed
1600 FRDM spherical
Halflife (ms)
1400
1200
1000
800
600
400
200
0
92Kr 93Kr 94Kr 95Kr
Nucleus
Figure 4.21. Results for Kr isotopes, along with previous measurements [?] and
FRDM-QRPA calculations[50].
Br isotopes
2000
1800 Present values
Previous values
FRDM deformed
1600 FRDM spherical
1400
Halflife (ms)
1200
1000
800
600
400
200
0
90Br 91Br 92Br
Nucleus
Figure 4.22. Results for Br isotopes, along with previous measurements [?] and
FRDM-QRPA calculations[50].
79
Se isotopes
1800
1600 Present values
Previous values
FRDM deformed
1400 FRDM spherical
Halflife (ms)
1200
1000
800
600
400
200
0
88Se 89Se 90Se 91Se
Nucleus
Figure 4.23. Results for Se isotopes, along with previous measurements [?] and
FRDM-QRPA calculations[50].
As isotopes
1400 Present values
Previous values
FRDM deformed
1200 FRDM spherical
1000
Halflife (ms)
800
600
400
200
Figure 4.24. Results for As isotopes, along with previous measurements [?] and
FRDM-QRPA calculations[50].
80
Figure 4.25. ADC spectra for NERO quadrant A. Y axis is the number of counts
in each detector. X axis is the uncalibrated neutron energy. Detectors 1-4 are 3 He
detectors, and show their characteristic spectrum, with a sharp peak at the Q value
of 0.764MeV. Behind that peak are events where the entire Q value of the reaction
was not deposited into the detector. When the outgoing deuterons strike the tube
walls a sharp drop in the spectra can be seen. Below that, the events for which the
3
H nuclei strike the tube walls can also be seen. Detectors 5-15 are BF3 detectors,
which show a broader peak at the full Q value of 2.792MeV.
81
Figure 4.28. ADC spectra for NERO quadrant D.
Figure 4.29. TDC spectrum from NERO ring 1 detectors. Channels are proportional
to time neutron is detected after a beta decay event. Ring 1 is the innermost ring of
detectors, and therefore sees more neutrons shortly after beta decay with relatively
few neutrons coming at later times.
82
Figure 4.30. TDC spectrum from NERO ring 2 detectors. Channels are proportional
to time neutron is detected after a beta decay event. Ring 2 is the middle ring of
detectors, and sees more neutrons at later times than does ring 1. This is due to
the fact that neutrons reaching this ring have to travel through more moderating
material.
Figure 4.31. TDC spectrum from NERO ring 3 detectors. Channels are proportional
to time neutron is detected after a beta decay event. Ring 3 is the outermost ring
of detectors, and sees the most neutrons at later times of all the NERO rings.
83
Beta background (counts per second)
0.1
0.09
0.08
0.07
0.06
0.05
0.04
0.03
0.02
0.01
0
0 50 100
Run
0.04
0.03
0.02
0.01
0
0 10 20 30 40
Strip
0.003
0.0025
0.002
0.0015
0.001
0.0005
0
0 50 100
Run
84
Neutron Background Rate per DSSD Strip
0.0045
0.004
0.0035
0.003
0.0025
0.002
0.0015
0.001
0.0005
0
0 10 20 30 40
Strip
Figure 4.35. Neutron background rate versus DSSD front strip number.
0.08
Neutrons per beta decay
0.06
0.04
0.02
0
0 50 100
Run
85
TABLE 4.1
NERO was allowed to count neutrons after a high-gain (beta) DSSD event. This
was done by opening a gate for 200µs after any such event. The neutron energy
signal was recorded via an ADC, and the neutron time was recorded by a TDC.
The ADC only recorded the first neutron event during the 200µs window, while the
86
TDC recorded all events. The individual ADC spectra and TDC spectra for each
Obtaining the Pn value for a particular nucleus begins with the equation:
Nβ−n
Pn = 100 ×
Nβparent
This equation states that the Pn value (in %) is the number of beta-delayed neutrons
detected divided by the number of parent nucleus beta decays. Because the beta de-
must calculate the number of parent beta decays that should be detected. Nβparent
is the number of parent nucleus beta decays, and is calculated from the total number
where λ is the decay constant of the parent nucleus, and tcorr is the correlation time
for beta decay events (either 5 or 10 seconds). Nimp is the total number of implants
for the efficiency of NERO and the BCS, as well as for the neutron background
Nβn−detected − Nβn−background
Nβ−n =
ǫβn
dow, and ǫβn is the total efficiency for detecting beta-delayed neutrons. ǫβn is given
by: ǫβn = ǫN ERO ǫβ where ǫN ERO is the detection efficiency of NERO, and ǫβ is the
87
detection efficiency of the BCS. The number of beta-delayed neutrons detected in
implants of a particular nucleus, and Rateβn is the per pixel neutron background
4.9 Pn Results
TABLE 4.2
PREVIOUS Pn MEASUREMENTS
88
CHAPTER 5
DISCUSSION
The evolution of shell structure away from the line of stability is a major open
question in nuclear physics. The results of shell gaps disappearing or new gaps
appearing, affect our understanding of the nucleus and of stellar processes like the
r-process. While not a direct measure of the shape of a nucleus, the beta decay
halflife can give insight into the nuclear shape, and the evolution of shape in a
The present results have been compared to QRPA halflife calculations using
the FRDM mass model. QRPA calculations are able to provide information on a
wide range of nuclei far from the line of stability, and have been used previously
nuclear masses and deformations with the FRDM. The FRDM, or Finite Range
potential energy of the nucleus. It treats the nucleus as a liquid drop, but also
takes into account the effect on the surface energy due to the finite range of the
the Coulomb field and the diffusness of the charge distribution, and introduces
89
potential for use in the Strutinski integral in order to introduce shell corrections,
and the Lipkin-Nogami version of the BCS method for pairing corrections.
surface. With the deformations determined, values that depend on the shape of the
With the ground state shapes and masses determined, a quasi random phase
approximation (QRPA) calculation is then done to obtain the beta decay halflives
VGT = 2χGT β 1− β 1+
β 1± = Σi σi t±
i
The beta decay halflife is obtained from the beta strength function, which is the
Qβ
1 X
= Sβ (Ei ) × f (Z, Qβ − Ei )
T1/2 Ei
the daughter nucleus, and f (Z, Qβ −Ei ) is the Fermi function, which is proportional
to (Qβ − Ei )5 .
90
The beta delayed neutron emission ratio can be calculated by looking at the
ratio of the beta strength above the neutron separation energy Sn to the total beta
strength.
PQβ
Sβ (Ei ) × f (Z, Qβ − Ei )
Pn = PSQnβ
0 Sβ (Ei ) × f (Z, Qβ − Ei )
The results of this experiment are compared with QRPA calculations and with
previous results in Figs 5.1 and 5.2. There is generally good agreement with previ-
98 100
ously measured values, with the possible exception of Y and Y. These nuclei
have low-lying states from which beta decay is possible. Because we can not separate
these states, the observed halflives are superpositions of these two states.
For the Se and As nuclei, there is good agreement with the QRPA deformed
calculations. The half-lives are slightly larger than the predicted values, but the
overall trend is consistent, and the predicted values for spherical shapes are much
higher.
91
Se isotopes
1800
1600 Present values
Previous values
FRDM deformed
1400 FRDM spherical
Halflife (ms)
1200
1000
800
600
400
200
0
88Se 89Se 90Se 91Se
Nucleus
As isotopes
1400 Present values
Previous values
FRDM deformed
1200 FRDM spherical
1000
Halflife (ms)
800
600
400
200
92
5.2 Deformation Discussion
The agreement of the measured beta decay halflife with the FRDM-QRPA pre-
dictions for a deformed shape, as opposed to a spherical shape agrees with predic-
tions for the ground state deformation[44],[45]. The deformed ground state shape is
predicted because of a shell gap at Z = 34 due to the splitting of the 1f 5/2 orbitals.
The E(2+ ) systematics for the Se isotopes are unclear as N approaches 56, as can
be seen in Fig 5.3. Generally, the first 2+ state in a spherical nucleus will be at a
higher energy than for deformed nuclei. Looking at Fig 5.7, one can pick out the
high E(2+ ) for the N = 50 closed shell. The trend past N = 50 looks as though
the E(2+ ) at N = 56 could either increase to over 1MeV, which would indicate a
spherical shape; or may decrease to below 800 keV, which would indicate a more
deformed shape.
93
Figure 5.4. Measured E(2+ ) values for Kr isotopes.
94
The suggestion of N = 56 as a spherical sub shell closure comes largely from
96
the case of Zr, which has a large E(2+ ) value. The trend of the N = 56 isotones
suggests shell closures at Z = 40 and 50 but not away from those nuclei(Fig 5.7).
The suggested means for creating the N = 56 subshell closure is by the lowering of
94
the ν2d5/2 level, which can hold six neutrons. Sr nucleus (Fig 5.5).
does not support such a closure. The proposed mechanism for the emergence of
this subshell closure is the lowering of the 2p1/2 below the 1f 5/2 due to a weaker
seems unlikely that the π2p1/2 and π2p3/2 states would strongly interact with the
ν2d5/2 state necessary to create the N = 56 closure because of the large difference
in spin-orbit coupling ( 12 vs 52 ).
95
Figure 5.6. Measured E(2+ ) values for Zr isotopes.
96
5.3 R-Process Calculations
measured halflives and Pn values were put into a classical r-process simulation code.
56
This code starts with a Fe seed and uses the assumption of a
(n, γ) ⇋ (γ, n)
equilibrium to calculate the isotopic abundances for each element. The isotopic
abundances are determined by the Saha equation, so that the neutron number den-
3/2
Y (Z, A + 1) G(Z, A + 1) A + 1
= Nn
Y (Z, A) G(Z, A) A
3/2
2π~2
Sn (Z,A+1)
× e kT
mu kT
Where G are the partition functions. The abundance for each element is then:
X
Y (Z) = Y (Z, A)
A
. Once the abundances for each isotopic chain are calculated, the flow from one
dY (Z, A) (Z−1,A)
= Y (Z − 1, A)(1 − Pn(Z−1,A) )λβ
dt
(Z−1,A+1) (Z,A)
+Y (Z − 1, A + 1)Pn(Z−1,A+1)λβ − Y (Z, A)λβ
After a process time t freeze-out occurs, and only β decays are considered, until the
nuclei have decayed back to stability. After freeze-out occurs, beta-delayed neutron
97
5.3.1 R-Process results
A simulation using the present experimental results is shown in Fig 5.8 along
with the observed solar system r-process abundances. Also shown are the r-process
abundances for the FRDM-QRPA predictions for spherical Se nuclei. The simulation
TABLE 5.1
The results show that the halflife of 90 Se has a strong effect on the abundances of
the nuclei up to the A = 130 peak. The difference between a spherical and deformed
shape for this nucleus changes the abundances by a factor of two. The measured
90
halflife of 284ms for Se produces more material for A = 90 -130 than does the
that region. Longer halflives at the waiting points is the reason that the nuclei at or
near closed shells (such as N = 82, 126) have larger r-process abundances. Results
From the simulations, it is clear that the Se isotopes have a larger effect on
the r-process at low neutron number densities. The exposure of Nn = 1020 was
the only one to be greatly effected by the variation in Se halflives. Because of this
98
R-Process Abundances
Spherical vs. Deformed Se Halflives
10
Solar Abundances
Experimental Values
Spherical Values
Abundance
0.1
0.01
100
A
Figure 5.8. Results of R-Process simulation using measured half-life values (black
line) and values predicted for spherical ground state shapes (red line).
capture process may be of concern. It was seen that neutron number densities of
between 1018 and 1021 can be affected by the halflives in this region. At Nn = 1016
though, the change in halflives did not strongly affect the final abundances.
99
CHAPTER 6
CONCLUSION
The goal of this work was to provide a better understanding of the mechanisms
that are responsible for creating the matter that makes up our solar system, and
the world around us. This knowledge has been sought by civilizations across the
world for thousands of years. However, it has only been within the last century
The connection between nuclear fusion and stellar fuel was first made by Arthur
Eddington in 1920 and extended by George Gamow, and Hans Bethe (among many
others). They realized that the energy provided by the fusion of hydrogen into
stellar material. This link was the first step to explaining how heavier elements are
made.
We know that after the Big Bang mostly hydrogen and helium nuclei were
present. It was not until the first generation of stars began to convert hydrogen
into helium that the nuclear abundances changed. The p-p chains and CNO cycle
account for conversion of hydrogen into helium. Helium is converted into carbon
through the triple-alpha reaction, while oxygen and neon are created by the burn-
ing of helium and carbon. Further nuclear burning can produce nuclei as heavy as
iron. Carbon, oxygen, neon, and silicon burn until iron group nuclei are produced.
Because these nuclei have the highest binding energies per nucleon, heavier nuclei
100
are generally not produced by charged particle reactions. In some cases though, the
rapid proton capture process can create elements heavier than iron.
To make nuclei heavier than iron, reactions involving uncharged neutrons are
required. Nuclei of any Z can capture neutrons, so we know that elements beyond
iron must be created through neutron reactions. The shape of the elemental abun-
dance curve provides some clues about how different processes contribute to the
total abundances. The abundance curve is dominated by three sets of double peaks
past the iron region. These double peaks seem to indicate two separate processes
driven by similar physics. These two processes are known as the slow (s) and rapid
Currently, the reason for two sets of peaks is thought to be that they are the
s and r processes reaching neutron shell closures at different points. The s-process
moves along closer to the line of nuclear stability, and therefore has a greater number
of protons at each neutron shell closure. The r-process path is further from stability
which means that it will have fewer protons at each neutron shell closure. A nucleus
in the r-process will capture several neutrons, until the point where any additional
neutrons will be emitted as soon as they are captured. This nucleus, known as
a waiting point nucleus will beta decay to its daughter (with Z+1). Thus the β
decay rate is extremely important in determining the final abundances from the
r-process. Longer β decay halflives at particular waiting points will result in greater
nuclear abundance around those waiting points, with reduced abundance beyond.
Shorter halflives will result in a quicker r-process and greater abundance of the
heaviest nuclei. In addition, the beta-delayed neutron emission ratio Pn is also very
important to determining the final nuclear abundances from the r-process. After the
process ends and nuclei begin to beta decay back to stability, beta- delayed neutron
101
Because of the importance of β decay halflives and Pn values to the r-process,
it was the aim of this work to measure the β decay halflives Pn values of several
r-process nuclei. The isotopes chosen were just below Z = 40 near N = 60 and
include: Y , Sr, Rb, Kr, Br, Se, and As isotopes. These nuclei are interesting
100
because they lie in the r-process path and are near Zr, which is characterized by
its unexpected ground state deformation. With the addition of only a few neutrons,
Zr goes from a spherical shape to a deformed shape. This change in shape also
produces a sharp increase in the beta-decay halflife. The nuclei we sought to measure
might also have such a rapid change in shape, but a corresponding change in the
90
halflife and P n value would have consequences for the r-process. The nucleus Se
96
in particular is interesting because it has N = 56 like the spherical nucleus Zr.
There are predictions that it too might have a spherical shape and a longer half-life.
A computer simulation shows that show that if the neutron rich Se isotopes were
In this work, the beta decay halflives were measured and Pn values were cal-
culated for neutron rich isotopes below Z = 40 produced at the NSCL coupled
cyclotron facility at Michigan State University. Because these nuclei are so far from
stability, they are extremely difficult to produce. The NSCL is one of only a few
facilities around the world that can produce nuclei so far from stability. To measure
the properties of these rare nuclei a system of silicon, germanium, and scintillator
for the nuclei with enough counts to allow binning. For the nuclei with too few
counts for this procedure to work, the Maximum Likelihood Method was used. This
method has been used in other low count experiments, including similar experiments
at the NSCL.
102
90
Twenty-two halflives were determined in our measurments. Three nuclei, Se,
88 89
As, and As, were measured for the first time. In addition, we have confirmed
previously measured half-lives for Y, Sr, Rb, Kr, Br, Se, and As isotopes. The new
measurements were used to calculate the Pn values for 92−90 Br and 96,94 Rb shown in
Table 4-2.
Beta decay halflives are a good first indication of nuclear structure. As stated
90
above, Se is a good testing ground for the strength of the N = 56 subshell. Based
on FRDM-QRPA calculations for the halflives depending on the nuclear shape, our
90
results suggest that Se is deformed in shape. This agrees with predictions for
the ground state shape based on a Z = 34 deformed shell gap. However, this
above, the results show that if the neutron rich Se isotopes were spherical in shape,
two. Because of their importance in neutron exposures with Nn ≤ 1021 , the isotopes
in this region may be important for the weak r-process and other proposed neutron
to set new limits on the end points of processes such as the alpha-rich freezeout or
90
the LEPP. The two possible shapes of Se added a factor of two uncertainty in the
Many of the nuclei that were measured in this work have been measured previ-
ously, often with greater precision. The reason for this has to do with the methods
the method known as ion separation on line (ISOL) was used. This method pro-
103
duces nuclei far from stability as fission products from a thick target. The products
are then extracted from the target and reaccelerates them to be stopped in a detec-
tor system. ISOL works very well for some nuclei, as can be seen in the previous
measurements. This method has two large drawbacks though. The first is that
the nuclei of interest must be fission products of the target. The fragment mass
distributions for most fissioning nuclei are peaked at A = 90 and A = 140. Away
from these masses the fission yeilds drop off sharply. This means that attempting to
produce nuclei with masses much different from these values is extremely difficult.
In addition to the limits of production, ISOL must also extract the product nuclei
from the target and reaccelerate them. Because of the chemistry involved, some
isotopes are difficult to extract at all. More importantly, this process takes time,
which means that very short lived nuclei can not be extracted and reaccelerated
These limitations mean that, especially for r-process studies, fragment separation
is the method of the future. As can be seen from this work’s region of interest, many
of the nuclei were first produced twenty or thirty years ago, with little progress since
then. This work was the first study in this area using the method that will eventually
take us beyond even the r-process nuclei. Despite having production rates well below
what were estimated, this experiment was able to produce new nuclei beyond what
the more mature ISOL techinque has been able to. There is no inherant reason
that fragmentation cannot deliver nuclei in this region and beyond in much greater
quantities. Using the knowledge gained from this experiment, future experiments
can optimize magnet settings and make even greater discoveries. The history of
r-process experiments has been a slow and steady one. This work has delivered new
While we have a general idea of how the heavy elements are made, we do not
104
know the specific details in either the nuclear structure side or the astrophysics side.
There have been very few r-process nuclei produced and measured. We currently rely
from stability. New shell closures can open, while known shell closures can disappear
as extra neutrons are added. The fact that the r-process exists far from stability
makes it exciting to study. However the extreme neutron excesses of the r-process
means that it is difficult to produce and measure the properties of r-process nuclei.
The future will provide many more answers to the questions of heavy element
creation. Today, only about ten percent of the nuclei thought possible to exist
have been studied at all. Most of these unknown nuclei are extremely neutron rich.
This work has extended the understanding of a section of the neutron rich nuclei
involved in the r-process. Further study at current facilities will provide additional
to obtain experimental information about all r-process nuclei. The current plan for
the Facility for Rare Isotope Beams (FRIB) will be able to produce neutron rich
nuclei via projectile fragmentation with sufficient intensities to study nearly all the
r-process nuclei. While the future is bright for the study of the r-process, it is critical
that new facilities are built, in order to continue to answer the questions about how
105
BIBLIOGRAPHY
[1] W.A. Fowler, E.M. Burbidge, G. R. Burbidge, and F. Hoyle Synthesis of the
Elements in Stars. Rev. Mod. Phys., 29(4):547–650, 1957.
[2] National Research Council of the National Academies Connecting Quarks with
the Cosmos: Eleven Science Questions for the New Century The National
Academies Press, Washington, D.C., 2003.
[3] G. Lemaitre Expansion of the Universe. Monthly Notices of the Royal Astro-
nomical Society, 91:490-501, 1931.
[4] E. Hubble A Relation Between Distance and Radial Velocity Among Extra-
galactic Nebulae. PNAS, 15(3):168-173, 1929.
[5] A. A. Penzias and R. W. Wilson A Measurement of Excess Antenna Temperature
at 4080 Mc/s. ApJ., 142(1):419, 1965.
[6] Claus E. Rolfs and William S. Rodney Cauldrons in the Cosmos. The University
of Chicago Press, Chicago, Illinois, 1988.
[7] Aage Bohr and Ben R. Mottelson Nuclear Structure W.A. Benjamin, Reading,
Massachusetts, 1975.
[8] M. Arnould and S. Goriely The P-Process of Stellar Nucleosynthesis. Phys.
Rep., 384(1):1-84, 2003.
[9] M. Busso, R. Gallino, and G.J. Wasserburg Nucleosynthesis in Asymptotic Giant
Branch Stars. Ann. Rev. Astron. Astrophys., 37(1):239-309, 1999.
[10] F Kappeler, H Beer, and K Wisshak S-process Nucleosynthesis-Nuclear Physics
and the Classical Model. Rep. Prog. Phys., 52: 945-1013, 1989.
[11] S. Goriely Uncertainties in the Solar System R-abundance Distribution. Astron.
Astophys., 342:881-891, 1999.
[12] S.G. Ryan, J.E. Norris, and T.C. Beers Extremely Metal-poor Stars. II. El-
emental Abundances and the Early Chemical Enrichment of the Galaxy. ApJ.,
471(1):254-278, 1996.
[13] C. Sneden et. al The Extremely Metal-poor, Neutron Capture α rich Star CS
22892-052: A Comprehensive Abundance Analysis. ApJ., 591(2):936-953, 2003.
[14] F. Montes, et. al Nucleosynthesis in the Early Galaxy. ApJ., 671(2):1685-1695,
2007.
106
[15] G.J. Wasserburg, M. Busso, and R. Gallino Abundances of Actinides and
Short-lived Nonactinides in the Interstellar Medium: Diverse Supernova Sources
for the r-Processes. ApJ., 466(2):109-113, 1996.
[16] D.D. Clayton, W.A Fowler, T.E Hull, and B.A Zimmerman Neutron Capture
Chains in Heavy Element Synthesis. Annals of Physics, 12(3):331-408, 1961.
[17] P.A. Seeger, W.A. Fowler, and D.D. Clayton Nucleosynthesis of Heavy Ele-
ments by Neutron Capture. ApJS., 11(1):121, 1965.
[18] K.L. Kratz Nucear Physics Constraints to Bring the Astrophysical R-Process
to the Waiting Point. Reviews in Modern Astronomy, 1(1):184-209, 1988.
[19] A.G.W. Cameron, J.J. Cowan, H.V. Klapdor, J. Metzinger, T. Oda, and J.W.
Truran Steady Flow Approximations to the Helium R-Process. ApSS., 91(1):221-
234, 1982.
[20] S.E. Woosley, and E. Baron The Collapse of White Dwarfs to Neutron Stars.
ApJ., 391(1):228-235, 1992.
[21] B.S. Meyer et al. R-Process Nucleosynthesis in the High-Entropy Supernova
Bubble. ApJ., 399(2):656-664, 1992.
[22] Maria Goeppert Mayer and J. Hans Jensen Elementary Theory of Nuclear Shell
Structure John Wiley and Sons, New York, 1955.
[23] R.F. Casten Nuclear Structure From a Simple Perspective Oxford University
Press, New York, 2000.
[24] A. Navin et al. Phys. Rev. Lett., 85:266, 2000.
[25] B. Pfeiffer, K.-L. Kratz, F.-K. Thielemann, Z. Phys., A357:235, 1997.
[26] K. Farouqi et al. Astrophysical conditions for an r-process in the high-entropy
wind scenario of type II supernovae. Nucl. Phys. A, 758:631-634, 2005.
[27] I. Dillmann et al. Phys.Rev.Lett., 91:162503, 2003.
[28] K. Heyde Nucl. Phys. A, 507:149-154, 1990.
[29] K.S. Krane Introductory Nuclear Physics Wiley, New York, 1988.
[30] R.F. Casten Nucl. Phys. A, 443:1-28, 1985.
[31] P. Federman and S. Pittel Phys. Rev. C, 20:820, 1979.
[32] A. Aprahamian et al. Phys. Lett., 140B:22, 1984.
[33] U.C. Bergmann et al. Nucl. Phys. A, 714:21, 2003.
[34] CF Von Weizscker Theory of Nuclear Masses Z. Phys., 1935.
[35] H. Bethe and R. Bacher Stationary States of Nuclei Rev. Mod. Phys., 8:165,
1936.
107
[36] W.D.Myers and W.J.Swiatecki Nuclear Masses and Deformations Nucl. Phys.,
81:1, 1966.
[37] D. Lunney, J.M. Pearson, and C. Thibault Recent Trends in the Determination
of Nuclear Masses Rev. Mod. Phys., 75, 2003.
[38] P. Moller et al. Nuclear Ground-State Masses and Deformations ADNDT,
59:185, 1995.
[39] A.K. Dutta et al. Nucl. Phys. A, 458:77,1986.
[40] J.M. Pearson and S. Goriely Nuclear Mass Formulas for Astrophysics Nucl.
Phys. A, 777:623,2006.
[41] A. Aprahamian et al. Hyperfine Interactions, 132:417424, 2001.
[42] T. Takahashi, M. Yamada and T. Kondoh ADNDT, 12:101,1973.
[43] J. Halbleib and R. Sorensen Nucl. Phys. A, 98:542,1967.
[44] P. Moller, J.R. Nix, and K.L.Kratz ADNDT, 66:131, 1997.
[45] J. Skalski, S. Mizutori and W. Nazarewicz Nuclear Physics A, 617:282, 1997.
[46] R.V.F. Janssens et al. Phys. Lett. B, 546:55, 2002.
[47] P.T. Hosmer et al. Phys. Rev. Lett., 94:112501, 2005.
[48] A1900 Nucl. Instrum. Meth. B, 204:90, 2003.
[49] J.I. Prisciandaro, A.C. Morton and P.F. Mantica Nucl. Instrum. Meth. A,
505:140, 2002.
[50] B.Pfeiffer, K.L. Kratz, and P.Moller Status of delayed neutron precursor data:
half-lives and neutron emission probabilities Prog. Nucl. Energy, 41:39-69, 2002.
Cowan J.J. Cowan, F.K. Thielemann, and J.W. Truran Physics Reports, 208:267,
1991.
108