Sunteți pe pagina 1din 134

A Photonuclear Study of the Halo Nucleus He

Mark James Boland B.Sc.(Hons)

Submitted in total fulfilment of the requirements of


the degree of Doctor of Philosophy
September 25, 2001

School of Physics
The University of Melbourne
Victoria, 3010, Australia

Abstract
The photonuclear reaction  Li 
He was studied using tagged photons in
the energy range of 50 to 70 MeV at three lab angles of 30 , 60 and 90 . By
measuring the proton missing-energy, the low-lying excited states in He were
identified. As well as the know ground state and first excited state, evidence was
found to support the existence of a new state which has been predicted by theory.
The He nucleus has a neutron halo surrounding a  He core. A soft dipole
resonance between the halo and the core has been predicted to occur at low excitation energies. This thesis compares the newly found state with the theoretical
parameters of the soft dipole.
In the data analysis of the present measurement, a well established and unambiguous background removal process was used. This technique is contrasted
with charge exchange experiments which have claimed to observe the soft dipole
resonance. Photonuclear techniques are shown to be a more reliable method of
observing the states in He.

This is to certify that:


1. the thesis comprises only my original work towards the PhD,
2. due acknowledgement has been made in the text to all other material used,
3. the thesis is less than 100,000 words in length, exclusive of tables, maps, bibliographies and appendices.

Acknowledgments
I would like to thank.....

Contents

1 Introduction

2 Motivation

2.1

2.2

2.3

Halo Nuclei . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.1.1

Basic Physics of Halo Nuclei . . . . . . . . . . . . . . .

2.1.2

General Properties of Halo Nuclei . . . . . . . . . . . .

10

2.1.3

Theoretical Descriptions of Halo Nuclei . . . . . . . . .

14

2.1.4

The Soft Dipole Resonance in Halo Nuclei . . . . . . .

16

The Halo Nucleus He . . . . . . . . . . . . . . . . . . . . . .

17

2.2.1

The Radius of He . . . . . . . . . . . . . . . . . . . .

18

2.2.2

New States Predicted in He . . . . . . . . . . . . . . .

19

2.2.3

Previous Measurements of He . . . . . . . . . . . . . .

22

Advantages of the  Li 


He Measurement . . . . . . . . . .

28

3 Experimental Method
3.1

3.2

31

Producing Photons . . . . . . . . . . . . . . . . . . . . . . . .

31

3.1.1

Photons from Bremsstrahlung . . . . . . . . . . . . . .

32

3.1.2

Photon Tagging . . . . . . . . . . . . . . . . . . . . . .

33

The MAX-lab . . . . . . . . . . . . . . . . . . . . . . . . . . .

34

3.2.1

The MAXINE Electron Accelerator . . . . . . . . . . .

35

3.2.2

MAX-I Beam Pulse Stretcher . . . . . . . . . . . . . .

36

vii

viii

Contents

3.2.3
3.3

3.4

3.5

The MAX-lab Photon Tagger . . . . . . . . . . . . . .

37

Detecting Protons . . . . . . . . . . . . . . . . . . . . . . . . .

38

3.3.1

The GLUE Chamber . . . . . . . . . . . . . . . . . . .

39

3.3.2

Solid State Detector Telescopes . . . . . . . . . . . . .

40

3.3.3

Charged Particle Identification Method . . . . . . . . .

44

3.3.4


Li Target . . . . . . . . . . . . . . . . . . . . . . . . .

46

3.3.5

Nuclear Experimental Hall (Cave) . . . . . . . . . . . .

46

Data Acquisition System . . . . . . . . . . . . . . . . . . . . .

48

3.4.1

Overview . . . . . . . . . . . . . . . . . . . . . . . . .

48

3.4.2

Hardware Circuit . . . . . . . . . . . . . . . . . . . . .

48

3.4.3

Event Trigger . . . . . . . . . . . . . . . . . . . . . . .

51

Summary of Experimental Parameters . . . . . . . . . . . . . .

52

4 Data Analysis

55

4.1

Analysis Overview . . . . . . . . . . . . . . . . . . . . . . . .

55

4.2

ROOT/CINT Software . . . . . . . . . . . . . . . . . . . . . .

56

4.3

Particle Identification . . . . . . . . . . . . . . . . . . . . . . .

58

4.4

Photon Energy Measurement . . . . . . . . . . . . . . . . . . .

60

4.5

Proton Energy Measurement . . . . . . . . . . . . . . . . . . .

61

4.5.1

Energy Loss Corrections . . . . . . . . . . . . . . . . .

61

4.5.2

Reaction Kinematics . . . . . . . . . . . . . . . . . . .

63

4.5.3

Missing-Energy . . . . . . . . . . . . . . . . . . . . . .

65

Correction for Accidental Tagging . . . . . . . . . . . . . . . .

67

4.6.1

TDC Timing Spectra . . . . . . . . . . . . . . . . . . .

67

4.6.2

Prompt Missing-Energy Spectrum . . . . . . . . . . . .

70

4.6.3

Accidental Missing-Energy Spectrum . . . . . . . . . .

71



Correction . . . . . . . . . . . . . . . . . . . . . . . . .

73

4.6

4.7

Contents

ix

5 Results and Discussion

77

5.1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . .

77

5.2

Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

79

5.2.1

Excitation Energy Spectra . . . . . . . . . . . . . . . .

79

5.2.2

States Identified . . . . . . . . . . . . . . . . . . . . . .

84

5.2.3

Angular Distribution . . . . . . . . . . . . . . . . . . .

85

Interpretations . . . . . . . . . . . . . . . . . . . . . . . . . . .

86

5.3.1

Comparison with Previous Measurements . . . . . . . .

86

5.3.2

The Low-Lying Region,  MeV . . . . . . . . .

89

5.3.3

The High Region, 

93

5.3



MeV . . . . . . . . . . . .

6 Conclusion

95

A Analysis of TDC Spectra

97

A.1 Structure of Uncorrelated Contribution . . . . . . . . . . . . . .

97

A.2 Timing Resolution . . . . . . . . . . . . . . . . . . . . . . . .

99

B Experiments Conducted at the MAX-lab

101

B.1

!

O #"$ %!& N . . . . . . . . . . . . . . . . . . . . . . . . . . 101

B.2

!

O '(#") *!& O . . . . . . . . . . . . . . . . . . . . . . . . . . 102

B.3

!

O ' %!& N . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

B.4

!+

C 
%!, B . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

B.5


Li -.
He . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

C Papers

105

C.1 Conference Papers . . . . . . . . . . . . . . . . . . . . . . . . 105


C.2 Journal Papers . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

List of Figures
2.1

Neutron and / components of the He matter distribution . . . .

2.2

Momentum distribution from fragmentation of He . . . . . . .

2.3

Comparison of !0! Li radius to ,0102 Pb and  2 Ca . . . . . . . . . . .

11

2.4

Neutron drip line on the table of isotopes . . . . . . . . . . . . .

13

2.5

Schematic of a neutron halo . . . . . . . . . . . . . . . . . . .

17

2.6

The Borromean rings . . . . . . . . . . . . . . . . . . . . . . .

18

2.7

Li 3
He calculation by Danilin et al. . . . . . . . . . . . .

22

2.8

Li  Li 4 Be He experiment by Sakuta et al. . . . . . . . . . .

23

2.9

Li  Li 4 Be He experiment by Janecke et al. . . . . . . . . . .

25

2.10 He fragmentation experiment by Aumann et al. . . . . . . . . .

26

2.11 Li  Li 4 Be He experiment by Nakayama et al. . . . . . . . .

28

2.12 Li -564+ He 0 He experiment by Nakamura et al. . . . . . . . . . .

29

3.1

bremsstrahlung energy spectrum . . . . . . . . . . . . . . . . .

32

3.2

Schematic of photon-tagging principle . . . . . . . . . . . . . .

34

3.3

Overview of the MAX-lab . . . . . . . . . . . . . . . . . . . .

35

3.4

The MAXINE accelerator . . . . . . . . . . . . . . . . . . . . .

36

3.5

The MAX-lab photon tagger . . . . . . . . . . . . . . . . . . .

38

3.6

The GLUE chamber top view . . . . . . . . . . . . . . . . . . .

39

3.7

Detector telescope . . . . . . . . . . . . . . . . . . . . . . . . .

41

3.8

The GLUE chamber side view . . . . . . . . . . . . . . . . . .

42

xi

xii

List of Figures

3.9

Energy Spectrum of ,0,02 Th . . . . . . . . . . . . . . . . . . . .

43

3.10 Particle identification principle . . . . . . . . . . . . . . . . . .

44

3.11

plot from a spectrometer . . . . . . . . . . . . . . . . .

45

3.12 Experimental hall layout . . . . . . . . . . . . . . . . . . . . .

47

78

-

coincidence circuit . . . . . . . . . . . . . . . . . . . .

49

3.14 Data acquisition circuit diagram . . . . . . . . . . . . . . . . .

50

3.15 X-trigger and tagger timing . . . . . . . . . . . . . . . . . . . .

52

4.1

Event-by-event analysis overview . . . . . . . . . . . . . . . .

57

4.2

Typical particle identification plot . . . . . . . . . . . . . . . .

59

4.3

The tagger calibration . . . . . . . . . . . . . . . . . . . . . . .

60

4.4

Proton energy loss correction function . . . . . . . . . . . . . .

62

4.5

Reaction kinematics . . . . . . . . . . . . . . . . . . . . . . . .

63

4.6

Raw proton missing energy spectrum . . . . . . . . . . . . . . .

66

4.7

Typical TDC spectrum . . . . . . . . . . . . . . . . . . . . . .

68

4.8

Accidental tagging removal . . . . . . . . . . . . . . . . . . . .

69

4.9

Prompt missing-energy spectrum . . . . . . . . . . . . . . . . .

70

4.10 Accidental missing-energy spectrum . . . . . . . . . . . . . . .

71

4.11 TDC timing regions . . . . . . . . . . . . . . . . . . . . . . . .

72

4.12 Background-subtracted missing-energy spectrum . . . . . . . .

73

background spectrum . . . . . . . . . . . . . . . . . . .

75

4.14 Corrected missing-energy spectrum . . . . . . . . . . . . . . .

76

5.1

Nuclear levels in He . . . . . . . . . . . . . . . . . . . . . . .

77

5.2

Excitation-energy spectra . . . . . . . . . . . . . . . . . . . . .

80

5.3

Integrated spectrum . . . . . . . . . . . . . . . . . . . . . . . .

81

5.4

Fitted excitation-energy spectra . . . . . . . . . . . . . . . . . .

83

5.5

Angular distribution of states in He . . . . . . . . . . . . . . .

85

3.13

4.13

78

-



List of Figures

xiii

5.6

He energy level diagram . . . . . . . . . . . . . . . . . . . . .

5.7

Comparison of

levels . . . . . . . . . . . . . . . . . . . . .

90

5.8

Comparison of <; levels . . . . . . . . . . . . . . . . . . . . .

92

5.9

Missing-energy spectrum from Li 


& He. . . . . . . . . . .

94

A.1 A comparison of tagging rates . . . . . . . . . . . . . . . . . .

98

:9

88

xiv

List of Tables
2.1

A table of various radii for He, Li and Be. . . . . . . . . . . . .

19

3.1

List of / -particle energies from a ,0,02 Th source . . . . . . . . . .

42

3.2

List of experimental parameters and their values. . . . . . . . .

53

5.1

Energy levels in He . . . . . . . . . . . . . . . . . . . . . . .

84

5.2

A summary of states in He . . . . . . . . . . . . . . . . . . . .

87

xv

xvi

Chapter 1
Introduction
One of the frontiers of todays nuclear science is the study of structure at the
limits of stability. The neutron drip-line is one such limit, beyond which nuclear
binding ends and the strong force no longer holds nucleons together, they literally drip out of the nucleus. Interesting phenomena take place in this region
of large neutron to proton ratio, for example the formation of neutron halos;
loosely-bound neutron distributions that extend far outside the bounds of the stable nuclear matter distribution.
A well established example of a halo nucleus is that of He. This system has
been successfully modeled as a  He core surrounded by a two-neutron halo [1],
and the neutron distribution has been measured to extend far beyond the normal
nuclear matter radius for a nucleus with

= 6 [2, 3]. For these reasons, He


=

has been used as a test case to study the behaviour of loosely-bound three-body
systems. As a consequence, He has been one of the most extensively studied
halo nuclei, both theoretically [413] and experimentally [1420]. The results
of these studies have been the prediction and measurement of new state in the
excitation spectrum of He. These exciting new results, which have emerged
over the past decade, have opened up a whole new area of research. However,
the calculations and experimental data are far from complete or conclusive, and
1

Chapter 1. Introduction

they require continued improvements and confirmation.


Experimental measurements of new states reported in the literature have predominantly been from reactions using hadronic probes, for example radioactive
beams or ion beams. In general these experiments have involved some nonrigorous analysis procedures. It is therefore of particular interest if the available
data can be improved using more transparent experimental techniques. This thesis reports on such an experiment; one that provides evidence of a new state in
He following a reaction with an electromagnetic probe. Measuring the missing

energy spectrum following the reaction  Li -.


0 He reveals the population of
states in the residual nucleus He. Significant improvements are obtained in the
background removal process using tagged photon techniques, compared with the
radioactive beam or ion beam experiments. In the measurement reported in this
thesis, well-known and unambiguous data analysis techniques are used to obtain
high-quality data that shows clear evidence of a new state in He.
The following chapters give a detailed description of how these results were
obtained. Chapter 2 presents the motivation for conducting this research. The
features of photonuclear reactions are discussed, including the specific reaction


Li -.
0 He and the information that can be extracted from it. An overview of

the physics of halo nuclei is given, and the literature on the recent theoretical
and experimental results in this field are reviewed. Previous measurements and
the techniques used to analyse them are critiqued, followed by a comment on the
improvements achieved in the current measurement.
Chapter 3 outlines the experimental method. Considerable technical detail is
given of the research facility and the detector system used to make the measurement. The data acquisition circuit and detector calibration are also described,
and the experimental parameters are tabulated.
Chapter 4 explains how the data was analysed. Importantly, the method used

to unambiguously remove the background is illustrated, highlighting the improvements that are made over other experimental techniques. The steps taken
to transform the raw data into excitation-energy spectra of He are covered.
The results of the data analysis are presented in Chapter 5 and thier significance is discussed in terms of the literature review from Chapter 2. This chapter
argues that the knowledge gained from the  Li -.
0 He reaction shows clear evidence of a new state in He at an energy level of 5 MeV and with a width of
3 MeV. In the final chapter, concluding remarks are made on the significance of
the research project in the understanding of halo nuclei.

Chapter 2
Motivation
It might be said that at this stage the understanding of the structure of nuclei,
that the lexicon of excited states was well established, and ostensibly consistent
with the understanding of the nucleon-nucleon force. One area where this is not
the case is where nuclei are close to the neutron drip-line. Attempts to predict
their level structure using models such as the Shell Model failed to explain the
known structure. Recently, as a result of particular interest in these neutron-rich
nuclei, more appropriate models have been developed. These are reviewed in
this chapter.
However, on the experimental front, attempts to verify these predictions have
been frustrated by the extreme instability of the nuclei themselves. The major
reaction mechanism used to probe their structure involved charge exchange. Ion
beam reactions have played a major role in this research [21, 22], since they can
result in production of nuclei with extreme neutron excess. They are, however,
somewhat non-specific, producing a range of residual nuclei. Charge-exchange
via the reaction Li -?>; He has been proposed [23] when the new 250-MeV
facility is available at MAX-lab. Even more exotic reactions, involving the use
of radioactive ions as projectiles, have made some contribution [17]. The shortcomings of these experiments are analysed in later sections of this chapter.
5

Chapter 2. Motivation

It is therefore reassuring that, at least for the He nucleus, examination of its


level structure can be done using the relatively simple, but totally appropriate,
photonuclear reaction  Li -.
0 He. A proposal was made to make this measurement at the MAX-lab, and a preliminary measurement by the Gent Group,
although failing to resolve the issue, did provide sufficient justification to allow
the experiment reported here to proceed.

2.1

Halo Nuclei

Two classes of nuclear halos exist; neutron halos and proton halos [21]. Of these,
the neutron halos have been studied in more detail, consequently they are better
understood. The focus of this research, is a neutron-halo nucleus He, and so
the following discussion will be limited to neutron-halo nuclei. In the next few
sections, an overview will be given of the discovery and some physical characteristics of nuclear halos, followed by a discussion of some recent theoretical
models and predictions.

2.1.1

Basic Physics of Halo Nuclei

Halo nuclei do not conform to some of the characteristics of mail-line nuclei.


For example, the nuclear radius @ of stable nuclei with mass = is found to obey
the relationship
@

where B

1HG

JIK

ACB
1

F
=ED 

fm. However, the two-neutron halo nucleus

(2.1)
!0!

Li has a radius of

3.2 fm, as large as the much heavier nucleus +0, S.


Similarly the surface diffuseness of halos differ markedly from the norm. At
the surface of stable nuclei, the nuclear density drops from about 70% to 30%
of its maximum over a range of 1 fm. The surface thickness of halo nuclei is

2.1. Halo Nuclei

far greater, and the neutron density decreases gradually over a range of several
fermi [5]. One of the most striking differences between halo nuclei and other
nuclei can be seen by comparing the radii of the proton and neutron distributions.
In the stable isotope
surface of

!0!

 2

Ca, for example, the difference is 0.1 fm, whereas the

Li is composed almost entirely of the neutron halo, which extends

approximately 1 fm beyond the proton distribution [24].


A qualitative analysis of halos reveals that the small separation energy of
the valence neutrons is predominantly responsible for their extended spatial distribution. The attractive force exerted by the nuclear core on a neutron, can be
approximated by a square-well potential and the radial wavefunction of a neutron
in this potential, can be written as [22]
L

where
@

N
BM 3A

e R6S

e9

P>Q

*UT

@ ! V ,XW

is the radius of the potential and


B

RZY
W

(2.2)

is the distance to the centre of the

nucleus. The spatial distribution of the wavefunction is determined by the value


O

of , which is given by the separation energy

H[

and the reduced mass of the

system \ using
^ O
*]
,

(2.3)

A:\_H[ZI

Combining Equations 2.2 and 2.3 it can be seen that the smaller the separation
energy, the greater the spatial distribution. In stable nuclei, the separation energy
H[

of the last neutron is typically 68 MeV, while in halo nuclei it is much

smaller, often less than 1 MeV. Consequently, the wavefunction for these weaklybound neutrons extends far beyond the wavefunction of a neutron in a stable
nucleus of similar mass.
Figure 2.1 shows the results from a calculation of the properties of He by
Zhukov et al. [5] in which the total density beyond

B`abIKc

fm is due mostly

Chapter 2. Motivation

to the valence neutrons. The matter distribution of the / -core,

dJe

, follows the

Figure 2.1: A calculation of the f and neutron components of the He matter


distribution. (Source: M. V. Zhukov et al., Phys. Rep. 231, 151 (1993).)

shape of the familiar nuclear matter distribution given by the Fermi profile

d#BX gA

where B is the radius, d


m

d
UT

eh Y

1
9 Sji%k)Vl

is the central density, @


1

(2.4)


is the radius at half density and

is a measure of the diffuseness of the nuclear surface. On the other hand, the

neutron distribution has a long tail which cannot be parameterised using Equation 2.4.
From the application of the Heisenberg uncertainty principle, the large spatial extent of the wavefunction implies that the momentum distribution of these
neutrons must be well defined. The momentum distribution

nop

is the Fourier

2.1. Halo Nuclei

transform of the wavefunction

BM

, and is given by

nop
gA

where


 , T

O
,

is the momentum of the neutron and

(2.5)


is a constant. This equation


q

together with Equation 2.5 and 2.3 indicates that as the separation energy decreases, the width of the momentum distribution becomes narrower.
This is found to be the case for two-neutron halo nuclei in radioactive beam
experiments. The momentum distribution of halo neutrons in He was measured
following the fragmentation reaction C  He Z:  He, and determined to have
a width

rsAut<v

MeV/c [14] (see Figure 2.2). In contrast, stable nuclei have

Figure 2.2: The fragment momentum distribution following the reaction


C w He x*y{z
|  He. (Source: T. Kobayashi, Nucl. Phys. A538, 343c (1992).)

a broader distribution close to 80 MeV/c. For more usual nuclei, this can be
parameterised in terms of the masses of the system as

r ,

where
J

A}r ,
1


Q~ P~

is the mass of the beam particle,

MeV/c.

(2.6)


the mass of the fragment and


r

1G

From this qualitative overview it is clear that halo nuclei have significantly

10

Chapter 2. Motivation

different spatial properties than normal nuclear matter. Investigation of these


properties has led to the discovery of new modes of nuclear excitation. The study
of nuclei at the extremes of stability promises many new research opportunities
[25]. Neutron-rich nuclei offer a possibility of studying essentially pure neutron
matter, which has never before been done in the laboratory. Results of this new
area of research range from understanding the bare nucleon-nucleon interaction,
to Big Bang and stellar nucleosynthesis [17, 26].

2.1.2

General Properties of Halo Nuclei

The original observation of what we now refer to as a halo nucleus was made
by Tanihata et al. [2] in experiments using radioactive nuclear beams [14, 26].
To make radioactive nuclear beams, high-energy beams of stable ions undergo
spallation reactions with heavy targets to produce a wide range of nuclei, including unstable neutron-rich nuclei such as

!0!

Li and He. These nuclei are then

separated, and accelerated to produce secondary beams, which become projectiles for reactions with secondary targets. Using this technique, Tanihata et al.
measured the interaction cross sections of neutron-rich nuclei. The interaction
cross section (r ) is the total probability that a projectile nucleus will undergo
transmutation after interacting with a target nucleus. From the interaction cross
section r the radius of the projectile can be found using the expression

rA}>@

where @

and @

T@ , 

(2.7)

are the interaction radii of the projectile and target, respectively.

After they carefully established that @

is a well defined size parameter, their data

revealed a considerably larger radius for !0! Li and He than for their neighbouring
nucleus. Initially the effect was thought to be cuased by a large deformation or a

2.1. Halo Nuclei

11

tail in the matter distribution. Figure 2.3 shows just how anomalous the size of
!0!

Li is by comparing it with heavier nuclei.

208

Pb

12 fm
9

Li

11

Li

7 fm

48

Ca

Figure 2.3: The large matter radius of !0! Li compared with ,0102 Pb and
(Source: J. S. Vaagen et al., Physica Scripta T88, 209 (2000).)

 2

Ca.

After Kobayashi et al. [27] experimentally confirmed the results by Tanihata


et al., theorists began to ponder the consequences of nuclear halos. Two papers,
one by Hansen and Jonson [28] and the other by Ikeda [29], independently suggested a collective resonance could be excited as an oscillation between the core
and the halo. They calculated that this so-called soft dipole resonance (soft DR)
would occur at lower energies than other collective resonances, such as the giant
dipole resonance (GDR), due to the weak restoring force of the loosely-bound
halo. Subsequent experimental and theoretical studies enthusiastically pursued

12

Chapter 2. Motivation

the soft dipole in their measurements and calculations. However, it proved to be


an elusive feature to observe [27, 30, 31]. Conflicting calculations and inconclusive data have plagued the unequivocal confirmation of the soft dipole, leaving
its existence to remain an open question (see Section 2.1.4).
After the ground-breaking idea of binary halos by Hansen and Jonson (in
their case they modeled !0! Li as a di-neutron and a Li core), developments moved
away from this highly clustered picture of the nucleus. Current interpretations
favour the image of a tightly-bound core with a thin veil of neutrons. These
neutrons tunnel into forbidden regions far removed from the normal nucleus,
forming dilute nuclear matter. This complicated co-existence of normal nuclear
matter and loosely-bound nucleons requires the inclusion of Pauli-blocking in
the wave-function describing the system [32]. Pauli-blocking takes into account
that there are fewer degrees of freedom for the halo neutrons, since the the core
neutrons already occupy the inner states.
Nucleon-nucleon correlation in the nuclear halo plays a crucial role in its
structure. For example,
!0!

!1

Li is unbound, but when a neutron is added to form

Li, the system becomes bound. It is thought that the three-body dynamics of

the core +  +  alters the spatial correlations so that, although the core +  and


+


interaction potentials are not individually sufficiently strong to bind the

two clusters together, the three-body system does form a loosely-bound states.
Interestingly, the theory of this type of three-body system was developed by
Zehn and Macek [33] in 1988 in the field of atomic physics, whereas it was not
until 1993 that Zhukov et al. [34] independently developed the theory for nuclear
halos.
Halo nuclei are of particular interest in the study of nuclei far from stability, in the neutron drip-line region. Figure 2.4 shows the region on the table of
isotopes where the light halo-nuclei are found. Along with the obvious interest

2.1. Halo Nuclei

13

B8

B10 B11

Be7

Be9 Be10 Be11 Be12

Be14

proton halo?
Li6

Li7

He3 He4 He5 He6


H1

H2

Li8

Li9

Li11 Li12

He8

He10

H3

naturally abundant
halo candidate
decay

n1
neutron dripline

Figure 2.4: The lower end of the chart of nuclides, showing the neutron dripline and candidates for halo nuclei

associated with nuclear binding, nucleon-nucleon potentials and diffuse nuclear


matter, there are applications of the study of neutron halos in astrophysics. In
particular, He plays a role in the theory of Big Bang and stellar nucleosynthesis [17, 26].
Nucleosynthesis calculations involves solving many coupled differential-equations
representing the rate, cross section and energy distribution of numerous reaction
chains. Most heavy elements are made from hydrogen and helium in a chain
of nuclear reactions that proceed through stable nuclei. However, nucleosynthesis can also occur via a series of rapid radiative neutron captures, called the
r-process, such as

He  P3? He :3? 2 He 

9 2

Li /U'

!0!

BI

(2.8)

In the case of Equation 2.8, short-lived nuclei are produced that are not stable to
particle emission. In order to calculate reaction rates and cross sections involving these neutron-rich nuclei, accurate information is needed on their nuclear
states. Recent calculations show that if the newly predicted states in He are included, there is a significant increase in the contribution to the creation of heavier

14

Chapter 2. Motivation

elements from the r-process chain [35]. Detailed studies are now underway to
determine what importance the new halo states in neutron-rich nuclei will play
in nucleosynthesis.

2.1.3

Theoretical Descriptions of Halo Nuclei

The successful shell model (SM) of nuclei has not been able to reproduce the
new features observed at the limits of bound nuclear matter. Clustering inside
halo nuclei appears to lend itself better to a few-body treatment, rather than a
full multi-nucleon calculation. For this purpose the cluster orbital shell model
(COSM) [36] was developed, which treats the nucleus as an inert core with valence neutrons. To calculate the states in He, realistic potentials were required
for the neutron-neutron and the core-neutron interaction potentials, along with
experimentally determined values of the core radius [1]. The valence-neutron
wave functions were also modified to account for the Pauli-forbidden states that
exist due to the reality of the substructure of the core.
Reasonable agreement was achieved between COSM calculations and scattering data, including the large electromagnetic dissociation (EMD) cross sections of He and

!0!

Li [5]. Nevertheless, a fundamental problem with the SM

wave functions is that they have incorrect asymptotic behaviour. Since halo nuclei have a large extended tail in their neutron distribution, the SM is an inadequate description of these weakly bound states. To deal with the extended wave
functions properly, the method of hyperspherical harmonics (HH) was developed by Danilin and Zhukov [37]. This method modifies the co-ordinate system
to calculate the three-body dynamics of halo nuclei, by defining the hyperradius
d

as
+
d ,

AC=

9 !

$:?
!

B , 

(2.9)

2.1. Halo Nuclei


,

where B

15

are the inter-particle distances and = is the particle mass. Calculations

of this type use He as a test case, since they are particularly good at reproducing
the well established

ground state and the <; first excited state [38].


;

Several other models have been developed to study specific reactions involving halo nuclei. One approach is the four-body distorted-wave method, which
models the halo-nuclear system as a core +  +  projectile nucleus plus a target
nucleus [11]. In this model, the ground-state wave-function of the halo nucleus,
L

dj

, is calculated using the HH method. The dissociation cross section, for

example, is then calculated using

r
bZM 3A

^
 >g~ ] , ,(

(2.10)

dj ? , 

is an expression containing the masses, momenta and spectroscopic

where
~

factors of the system and

is the transition amplitude for the reaction. It

is assumed that neither the target nor the other three bodies form excited state
states, a process referred to as diffractive breakup [10]. In order to compute

, an optical-model potential is used for the target nucleus, and core-neutron

( #j ) and neutron-neutron (  ) interaction potentials are required. Once again


the He system can be modeled with one of these bases, because
are well established for the  He +


and the


+


#j

and

systems from experimental

data (see [1, 38] and references therein). Current theoretical and experimental
studies seem to be in good agreement [13], although more work is required to
remove some of the assumptions in the model, and improve the fundamental
understanding of halo nuclei.

16

Chapter 2. Motivation

2.1.4

The Soft Dipole Resonance in Halo Nuclei

Collective excitations have played an important part in clarifying the structure of


the atomic nucleus. In early experimental work, the photonuclear effect revealed
the GDR, and a large body of data was generated on the subject. Real photons
with energies up to 30 MeV were used to excite a range of nuclei across the table
of isotopes. Using the liquid-drop model, the GDR is described as an oscillation
of the neutron fluid ralative to the proton fluid, with a characteristic resonant
energy of



S:

=
9

MeV [39]. Similarly, the soft DR is described as an


D

oscillation of a core relative to the valence neutrons, and may also be studied
using medium energy photons.
In principle, the GDR in halo nuclei can be split into two components due to
the loosely-bound valence neutrons. Figure 2.5 shows schematically the different
components, and the corresponding resonance energies. The soft DR occurs at
a lower energy than the GDR due to the weak binding energy of the valence
neutrons, which produces a weaker restoring force between the core and the
halo. An estimate of the excitation energy of the soft DR, made by Suzuki et
al. [1], is given by

[

where

=
g g_T'

the mass of the halo nucleus,

@o, ?

] ,
,6?

(2.11)


is a measure of the neutron distri-

bution radius,  and  are the number of core protons and neutron respectively
and


is the number of valence neutrons, so that

=Ag3T}3T

Equation 2.11 to He leads to an estimate of the soft DR energy of

. Applying
[

AvI

MeV.
In practice, the soft DR has been difficult to observe experimentally in halo
nuclei. Systematic studies are currently underway to characterise the dipole re-

2.2. The Halo Nucleus He

17

proton and neutron


saturation

low neutron density

Soft DR

GRD

transition
strength

E ex

Figure 2.5: A schematic diagram of a neutron halo formed in He and the three
types of collective oscillations possible. (Source: I. Tanihata, J. Phys. G, 22,
157 (1996).)

sponse in light neutron-rich nuclei (see for example Aumann et al. [40] and references therein). It is the intention of the present measurement to prodive an
improvement over the currently available data, and possibly identify the presence of a soft DR in He.

2.2

The Halo Nucleus He

The halo nucleus He has captured the imagination of the nuclear physics community. The curious feature that the nucleus is bound, yet none of its binary
subsystems are bound, has led to it being called a Borromean system [34].

18

Chapter 2. Motivation

Figure 2.6: The heraldic symbol of the Borromean Rings.

The Borromean rings shown in Figure 2.6, are the heraldic symbol of the Princes
of Borromeo, carved in stone on their castle located on an island in Lago Maggiore, in northern Italy. If any one of the interlocked rings are broken, all three
come apart. So too are the  He +


and the


+


systems unbound, while He

forms a stable halo-nucleus.

2.2.1

The Radius of He

The half life of He is 806.7 ms, decaying by

-emission to Li. This makes

it hard to perform any direct measurement of its properties. However, with the
development of radioactive beams it has been possible to perform scattering experiments with beams of He. The first observation of the He halo was made by
Tanihata et al. [2] who deduced the interaction radius @ from the cross section
(see Section 2.1.2 and Equation 2.7). Using the interaction radius, it is possi
ble to calculated the root mean squared (rms) value of the matter radius ( @rms
),

the charge radius ( @

rms )

and the neutron radius ( @

was found in the neutron radius


@

rms

rms ).

A significant difference

of He compared with Li. In itself this

is not surprising, since He has one more neutron than Li. The significance is
best illustrated by comparing the difference between the He- Li radii and the


Be-  Li. Table 2.1 shows that the =A

systems have essentially the same

2.2. The Halo Nucleus He

and
@

rms  

Be

@
G

rms  

19

Li , as might be expected. In contrast, Li has identi-

cal values for its matter ( @

),
rms

charge ( @

rms )

and neutron ( @

rms )

radii, whereas

He shows clustering of the charge distribution, and dispersion of the neutron


distribution. This provided the first evidence of the neutron halo in He.
Interaction Matter Charge

Neutron
@

rms

@rms

He
Li

2.18
2.09

2.73
2.54

2.46
2.54

2.87
2.54

Li
Be

2.23
2.22

2.50
2.48

2.43
2.52

2.54
2.41




rms

@H

Table 2.1: A table of root mean squared (rms) radii for various isotopes to
highlight the large extent of the He neutron radius. (All values are in units of
fermi.)

The extent of the He neutron halo was confirmed most recently by Shostak
et al. [3] using the reaction  Li p6( 0 He with 70 MeV protons. At this energy,
only the surface properties of He were being probed, as the wavelength of the
protons is approximately 5 fm. The data gave a value of
@

rms

= 2.85 fm, signif-

icantly larger than the experimentally deduced nuclear charge radius of


@

rms 

2.50 fm [3, and references therein]. These results imply the presence of an extended neutron-cloud surrounding a charged central-core, i.e. a neutron halo.

2.2.2

New States Predicted in He

The most recent listing of the known states in He [41] shows the first excited
state at 1.8 MeV, and then no states until above the + H + + H threshold at 12.3
MeV. Results from calculations of the nuclear levels in He, using the theoretical
methods described in Section 2.1.3, have challenged this picture, with predictions of new low-lying states. This section discusses these theoretical predictions, while Section 2.2.3 considers the experimental evidence for new states in

He.

20

Chapter 2. Motivation

The nuclear levels in a stable system of =A

nucleons, such as Li, can be

accurately calculated using traditional Shell-Model (SM) methods. For example


the calculations by Karataglidids et al. [42] are in good agreement with the experimentally determined levels [41]. However, SM calculations have not been
successful at explaining the level structure in He. One reason for this is that the
neutron separation energy ( ( ) of He is significantly lower that Li ( (j He =
1.86 MeV and (j Li = 5.66 MeV [43]). To model these loosely-bound neutron
systems, modifications were made to the bases of the SM calculations. By specifically taking into account the clustering in the nucleus, modified SM calculations
have been successful in modeling He. Generally, this is done by considering the

He nucleus as three clusters:  He +  +  .

The first calculation to produce a particle-stable bound state in the  He +


+


system, was performed by Suzuki et al. [1] using the COSM (see Section

2.1.3). To account for the weak binding energy, and the large spatial extension of the valence neutrons, single-particle orbits with large angular momentum
were included in the calculation. The model was able to calculate the ground
state of He, but the binding energy was underestimated by 0.5 MeV compared
with experimental values. No attempt was made to calculate excited states. The
electromagnetic dissociation (EMD) cross section was calculated, and a large
enhancement was found between an excitation energies of 47 MeV. This was
done by assuming that the  He core remained in its ground state, and calculating
the electric dipole transition-probabilities
with

A 9

:

for valence neutrons to states

. The large dipole strength at low excitation energy was in agree-

ment with the properties of the predicted soft DR. However, the calculated EMD
cross section underestimated the experimental result, so it was concluded that
more information was needed on the underlying reaction mechanism. Nonetheless, these results were a significant step in the understanding of the nuclear halo

2.2. The Halo Nucleus He

21

in He.
Aoyama et al. [44] extended the the COSM calculations of Suzuki et al.
using the complex scaling method (CSM) [45] to transform the wavefunction
of He. Corrections of this type avoid problems associated with the asymptotic
behaviour of the system, which is important when calculating the nuclear levels.
Significant improvements were achieved with this method over the results by
Suzuki et al. Not only was the ground-state energy successfully calculated (0.2
MeV discrepancy with experiment), but the

<;

1.8 MeV first excited state was

predicted at 1.81 MeV with a width of 0.26 MeV. These calculations were limited
to  -shell configurations of the valence neutrons in order to focus on the threebody resonances, of which the

<;

1.8 MeV states is one. No sign of the

 9

soft

DR could be found below 10 MeV using this model, possibly indicating that
more complicated orbits need to be included in the model.
Predictions of a low-lying

 9

soft DR in He were made by Danilin et al. [8]

using a  He +  +  cluster model. To calculate the ground and excited states, the
HH method was used (see Section 2.1.3) in the framework of charge-exchange
and inelastic scattering reactions. The distorted-wave impulse approximation
(DWIA) reaction theory, appropriately modified for dilute matter, was applied
to calculate the reaction cross section

r

of the reactions Li .


0 He and

He )j"$ 0 He, with an incident nucleon energy of

Ac

MeV in each case.

The uncertainty surrounding the existence of a true soft DR state is illustrated


by the curve labeled

 9

in the cross section to a


compared with the 
that the

 9

in Figure 2.7. Although a local maximum is predicted


 9

configuration in He, it is broad and poorly defined

strength. The lack of a distinct peak has led to suggestions

enhancement is due to dynamical effects from final-state interactions

[8, 45].
It has been suggested that any calculation which treats a  He cluster as in-

22

Chapter 2. Motivation

Figure 2.7: Reaction cross-section to the He bound-states and the three-body


continuum following the charge exchange reaction Li wz_x#| He. (Source: B.
V. Danilin et al., Phys. Rev. C 55, R577 (1997).)

ert is inadequate, considering that realistic ground state wave-functions of He


contain only a 20% admixture of the  He +


+


configuration [46]. Further-

more, a much more accurate calculation of the ground-state binding energy of


He was made by Csoto [47] (see also [48]) using admixtures of  He +




+


and 5.T5 configurations. Csoto calculated the binding energy of He relative to




He to be

A

MeV, compared with

A

MeV obtained from

experimental data [41]. Despite these flaws in three-body  He +




+


mod-

els, theorists continue to use this framework because of its success in calculating
charge-exchange reactions leading to He [9, 38].

2.2.3

Previous Measurements of He

In the last ten years, experimental studies of the neutron halo have been done
predominately using hadronic reactions. With the advent of radioactive beam
facilities, many new and exotic nuclei near the neutron drip-line have been created [21]. Halos have been studied in !0! Li, 2 Be,

! 

Be and He, but of these only

2.2. The Halo Nucleus He

23

He can be studied simply and effectively using photonuclear techniques. For


this reason, the discussion of other experiments with halo nuclei, will be limited
to measurements of He. What follows is a review of five key experiments on
He and specific critiques of some of the analysis techniques.

Figure 2.8: He excitation energy spectrum following the reaction


Li w  Li x  Be | He. (Source: S. B. Sakuta et al., Europhys. Lett. 22, 511
(1993).)

Li  Li K Be 0 He by Sakuta et al. 1993


Sakuta et al. [15] were the first to

claim to have found experimental evidence for the soft DR in He. Following
the charge exchange reaction Li  Li 4 Be 0 He, they measured the energy of the


Be reaction products using a magnetic separator. The excitation-energy spec-

trum of He that was measured is shown in Figure 2.8. Some notable features
are the hydrogen peaks contaminating the spectrum, and the well known

 ;

1.8

MeV first excited state. At a central energy of 6 MeV, there is a broad structure

24

Chapter 2. Motivation

that was identified by Sakuta et al. as a candidate for the soft DR. The angular
distribution data did not conclusively determine the spin and parity (

) of the

structure. Significantly, the background in the experiment was not measured,


rather it was deduced from a phase space calculation, and normalised to the experimental data. The smooth curve labeled 1 in Figure 2.8, shows the distribution
obtained from a calculation of the background, which was normalised to the data
at an excitation energy of about 23 MeV. A fit was made to the backgroundsubtracted data using four Gaussians (labeled curve 2), and the centroid energies
obtained were


G

JI

{b{

MeV. Despite the lack of a measured back-

ground, and the complications resulting from the hydrogen contamination, these
results do indicate a broad structure that might be evidence of the predicted soft
DR. They also have evidence for broad states at higher excitation eneries.

Li  Li K Be 0 He by Janecke et al. 1996


Janecke et al. [16] used the same

reaction as Sakuta et al., although a different technique was used to extract the
He excitation energy spectrum. In an attempt to clean up the spectrum,  Be

ejectiles were measured in coincidence with 430-keV de-excitation  -rays from




Be


Be6 [%T . Figure 2.9 shows the He excitation energy spectrum (a)

without, and (b) with, a de-excitation  -ray coincidence requirement. The states
identified in Figure 2.9 are at

A

and

JI

MeV, along with broad reso-

nances at 5.6, 14.6 and 23.3 MeV. The angular distribution data of the 5.6 MeV
resonance seems to indicate it is a

:;

state, but the fit is far from convincing.

Once again, like the Sakuta et al. measurement, the contributions from background reactions like p   Li   Be n were not measured, but were calculated. The
structure at

6 MeV found by Sakuta et al. was confirmed by Janecke et al.,

but it could still not be fully characterised.

2.2. The Halo Nucleus He

25

Figure 2.9: He excitation energy spectrum from Li w  Li x  Be | He (a) without


and (b) with a coincident de-excitation -ray requirement. (Source: J. Janecke
et al., Phys. Rev. C 54, 1070 (1996).)

He


He TT break-up off Pb and C by Aumann et al. 1999

The half-

life of He is only 806.7 ms, therefore it is not easy to perform experiments with
this nucleus. Despite this difficulty, Aumann et al. [17] successfully measured
the breakup reaction He

He

by scattering a secondary He beam

ToTC

off Pb and C targets. A primary beam of

!2

O was fragmented using a beryllium

target; the He fragments were then separated and transported to the secondary
target.
Figure 2.10 shows the excitation energy spectra of He deduced from the inelastic nuclear scattering off Pb and C. The familiar

<;

1.8 MeV resonance was

observed with a resolution of 0.2 MeV. There was evidence for a second 
at 4.4 MeV that conflicted with

 ;

state

state at 5.6 MeV reported by Janecke et al.:

the state found by Aumann et al. was 0.2 MeV wide, whereas that claimed by

26

Chapter 2. Motivation

Janecke et al. had a width of 10 MeV. Qualitative analysis of the data suggested
the presence of strength in the low-lying continuum from a mixture of monopole
and quadrapole resonances. No indication of the soft DR was reported, and
above the first excited state, the spectrum appears relatively smooth and featureless. This might be due to restrictions in the possible transitions available for
inelastic scattering between states of certain spin and isospin [18].

Figure 2.10: He excitation energy spectrum from the fragmentation of He off


Pb and C targets. (Source: T. Aumann et al., Phys. Rev. C 59, 1252 (1999).)

Li  Li K Be 0 He by Nakayama et al. 2000

Like Janecke et al. and Sakuta

et al. before them, Nakayama et al. [19] measured states in He via the charge
exchange reaction Li  Li K Be 0 He. However, their approach to the data analysis was completely different to the previous two measurements. Firstly, they
isolated the spin-flip ( 7A ) from the spin-nonflip ( 7
A

) excitations by

measuring de-excitation  -rays in coincidence with  Be ejectiles [49]. It was

2.2. The Halo Nucleus He

then assumed that the 7A

27

spectrum contained exclusively GDR excitations,

and the 7A spectrum contained a mixture of the soft DR and the spin dipole
resonance (spin DR). Citing previous studies on dipole resonances, the GDR and
the spin DR were assumed to have the same energy distributions and the same
strength. On this basis, the soft DR was observed by simply subtracting the
7A

spectrum from the 7A spectrum.

The resulting fit to the structure can be seen in Figure 2.11 as a shaded-in
Lorentzian curve, with an excitation energy of ACv MeV and a width of A
v

MeV. As with all the other measurements, the

 ;

1.8 MeV first excited state

is observed. Claims by Nakayama et al. that the results are a candidate for the
soft DR, may be compromised by not having considered any other background
channels in their analysis. Considering that the spin DR background was not
measured directly, and the assumptions made to account for it, the spectra are
not totally convincing evidence for the soft DR.

Li 56+ He He by Nakamura et al. 2000 Nakamura et al. [20] measured the

reaction Li 56+ He He with a secondary triton beam, using a method similar to


Aumann et al. A  He beam impinged on a beryllium target to produce a triton
beam, which in turn impinged on a Li target. They observed a broad asymmetric
structure at c MeV, along with the familiar ground and first excited states.
The He excitation-energy spectrum they measured can be seen in Figure 2.12. A
distribution momentum transfer was also found, and when compared with theory
was in agreement with a transition to negative parity states with

7A

. The

analysis procedure still contains some of the flaws of the previous experiments;
Specifically, the background was not measured, but calculationed and normalised
to the data. No other specific background channels were corrected for. Despite
these drawbacks, these data provide good evidence of the presence of low-lying
dipole strength in He.

28

Chapter 2. Motivation

Figure 2.11: He excitation energy spectrum following the reaction


Li w  Li x  Be | He. The two spectra shown are (a) the `s spin-flip spectrum and (b) the  spin-nonflip spectrum. (Source: S. Nakayama et al.,
Phys. Rev. Lett. 85, 262 (2000).)

2.3

Advantages of the Li 6 He Measurement

In the previous section, all the experiments showed some evidence of new structure in the excitation-energy spectrum of He. They all revealed ground state and
known first-excited state: essential features that must be seen for the results to
be credible. In the region of primary interest, from 3 to 10 MeV excitation, the
experimental data do not agree. Two of the  Li-beam experiments show broad
overlapping states, while the other shows a single narrow state. Similarly, the
tritium-beam experiment shows a mixture of broad states, and the He fragmentation reaction shows a single narrow state. One thread that runs through all
the previous experiments, is the lack of an unambiguous background removal
process.

2.3. Advantages of the  Li -


0 He Measurement

Figure 2.12: He excitation energy spectrum following the reaction


Li wx + He | He. (Source: T. Nakamura et al., Phys. Lett. B493, 209 (2000).)

None of the experiments using lithium-ions measured the background for the
data they present. Consequently, they do not have a consistent method for dealing
with the background reaction channels. On the other hand, the tagged-photon
technique used to measure the  Li 
He reaction, measures the uncorrelated
background contribution as part of the normal experimental procedure. In the
off-line data analysis, the uncorrelated spectrum is produced and subtracted from
the correlated spectrum to give a background corrected spectrum. This process is
well-known [5055], and has been used successfully in several important studies,
for example that by Kuzin et al. [55].
The importance of the new results presented in this thesis is not strictly related to the photonuclear reaction mechanism, as they are not explicitly compared with any photoabsorption models. However, it is serendipitous that the cur-

29

30

Chapter 2. Motivation

rent interest in halo nuclei, in particular He, and other areas of nuclear physics
should overlap in the reaction  Li 
He. Indeed, He is probably the only
neutron-halo nucleus that can be studied using photonuclear techniques: heavier
neutron-rich nuclei must be created by fragmentation reactions at a radioactive
beam facility. Thus, using a stable  Li target and a tagged photon beam, a very
clean picture of the nuclear levels in He has been obtained. Importantly, the
ground state and first excited state are clearly observed, along with new structure
above these well known states. The precision of the

-

measurement, includ-

ing a careful background-subtraction process, has given clear and unambiguous


evidence of the existence of at least one new state in He.
The chapters that follow, describe the experimental and data-analysis techniques used to obtain the nuclear levels in He.

Chapter 3
Experimental Method
A beam of tagged photons in the energy range A 5070 MeV was used to induce the reaction  Li -
0 He. Photon tagging is achieved using fast coincidencedetection electronics, which measures both the real events and the random events.
In this experiment, not only are protons produced, but a whole range of particles,
for example neutrons, deuterons, tritons and helium isotopes. All the charged
particles emitted by the photonuclear reactions, were detected using solid-state
detector telescopes. A charged-particle spectrometer with two components was
used to identify the protons from the other charged particles, on the basis that
particles with different mass, charge and energy, have a different range in each
detector component. This allows each particle type to be separated from the others by their energy-loss characteristics. All the data collected from the detector
systems were recorded by a computer and stored for off-line analysis.

3.1

Producing Photons

In order to study the



photonuclear reactions it is necessary to have a know

flux of photons with a known energy. Sources such as natural radioactive isotopes can be used, but they are limited in their use by the relatively low energy
31

32

Chapter 3. Experimental Method

photons they emit (of order 110 MeV), and the difficulty in measuring their
flux. To produce photons of higher energies ( 

10 MeV) electron accelerators

are required. The first few sections will discuss the general principles of the technique used for the experiment presented in this thesis. The subsequent sections
will discuss the specifics of the laboratory where the experiment was conducted.

3.1.1

Photons from Bremsstrahlung

When an electron is scattered by the Coulomb field of a nucleus, a photon is


created to conserve momentum. Photons created this way are called bremsstrahlung, the German word for braking radiation, i.e. deceleration radiation. The
energy spectrum of this type of radiation is continuous [56], as can be seen in
Figure 3.1.
N (MeV-1)

10

Tagging Range

10

10

10

10

10

20

30

40

50

60

70

80

90
100
E (MeV)

Figure 3.1: The Schiff energy spectrum of bremsstrahlung produced by an


 MeV.
electron beam of energy 

If a beam of electrons impinges on a thin target of high- material, for example gold, bremsstrahlung photons are emitted with energies from zero up to

3.1. Producing Photons

33

the incident energy of the electrons



. These high-energy photons can be used

to induce nuclear reactions in targets made from stable nuclear-matter. However,


such a measurement only provides an integrated yield, rather than an energy
dependent cross section. In order to perform high-resolution energy dependent
measurements that can resolve individual nuclear states, the interacting photon
energy must be known. One technique devised to achieve this is called photon
tagging and is described in the next section.

3.1.2

Photon Tagging

The technique of photon tagging is designed to indirectly measure the energy


of bremsstrahlung photons. A schematic diagram illustrating the principle of
photon tagging is shown in Figure 3.2. If an electron of energy
by a thin radiator, and its final energy is measured to be

 "



is scattered

, the energy of the

photon that is created is given by

AC

" I

(3.1)

The scattered electron is detected in a position sensitive spectrometer which is


discussed in detail in Section 3.2.3.
Assume that the photon then interacts with a target nucleus, resulting in the
emission of a reaction product-particle. If this particle is detected in coincidence with the associated scattered electron, the energy of the interacting photon
is determined by Equation 3.1. In the resulting experimental data, a complete
measurement of the kinematics of each reaction event is recorded.

34

Chapter 3. Experimental Method


Bending Magnet
Ee

Photon Beam

Electron Beam

Target

E = Ee E
e
Charged
Particle

Radiator
Electron

Ee

Detector

Detector
Timer
Coincidence

Figure 3.2: The principle of photon tagging, showing the electron spectrometer
which is used to detect recoil-electrons in coincidence with ejectiles from the
target.

3.2

The MAX-lab

The MAX-lab is the Swedish National Electron Accelerator Laboratory for Nuclear Physics and Synchrotron Radiation Research and is situated on campus
at Lund University in Sweden. Amongst the accelerator facilities is an injector
race-track microtron, MAXINE, and an electron storage/stretcher ring, MAXI. MAXINE is used as a source of energetic electrons that are injected into the
storage/stretcher ring. MAX-I is primarily used as a sources of high-luminosity
X-ray for synchrotron-light experiments. It can also be used as a pulse stretcher
to produce a continuous wave (CW) electron-beam for nuclear physics experiments. The tagging facility at the MAX-lab uses this CW electron beam to
produce bremsstrahlung with energies of up to

95 MeV. A schematic of the

equipment and the beam line relevant for nuclear physics is shown in Figure 3.3.
The following sections give a description of the MAX-lab tagging facility
and the configuration which was used for the experiment presented in this thesis.

3.2. The MAX-lab

35

Synchrotron Light
Beam Lines
Kicker Magnet

Dipole Magnets

Quadrupole Magnets

500 MHz
Cavity

MAXI
550 MeV Storage Ring

GROUND FLOOR

Undulator
Beam
Septum Magnet
Tagger
e Injector

Nuclear Physics Beam Line

e Beam
100 MeV Microtron

BASEMENT

Figure 3.3: A schematic overview of the MAX-lab showing the tagger in the
basement. (Source: J.-O. Adler et al., Nucl. Instr. and Meth. A294, 15 (1990).)

3.2.1

The MAXINE Electron Accelerator

The primary electron accelerator at the MAX-lab is a 100 MeV race-track microtron called MAXINE, a detailed description of which is presented in Reference [57]. MAXINE is a pulsed accelerator and is used to inject the MAX-I
storage/stretcher ring with a beam of energetic electrons. A pulsed beam is produced by an electron gun delivering a 100 keV beam into a buncher, which injects
into the linear accelerator (linac). The linac consists of a radio frequency (RF)
cavity that produces a standing wave to accelerate electron bunches. After the
electrons emerge from the RF cavity, a pair of magnets guides the bunch around
and back into the cavity (see Figure 3.4). Each pass through the linac, increases
the kinetic energy of an electron by

MeV, up to a maximum of 19 turns. A

small magnet is used to extract the beam in an evacuated transport beam-line.


The usual operating energy of MAXINE is

MeV at an average current of

30 nA.
The accelerator can be operated in two frequency modes; at 50 Hz in synchrotron-

light mode, and at 100 Hz in photon-tagging mode. For this experiment the
microtron was set to an energy of 92.45 MeV in 100 Hz mode.

36

Chapter 3. Experimental Method


Bending
Magnet

Extraction Magnet

Bending
Magnet

RF Cavity
Linac
Electron
Bunches

Displacing
Magnets

Buncher
Electron
Gun

Figure 3.4: A simplified view of MAXINE, the 100 MeV microtron at MAXLab.

A drawback of this acceleration method is that bunches of electrons are delivered in

1 \ s pulses every 10 ms, i.e. a duty factor of 0.01%. The arrival of

such a large bunch in a short time would flood an electron detector system, making tagging experiments very difficult to conduct. The solution to this problem
is to stretch the beam pulse to several milliseconds long, producing an almost
continuous beam of electrons. At the MAX-lab this is achieved with the MAX-I
stretcher ring as discussed in the next section.

3.2.2

MAX-I Beam Pulse Stretcher

The MAX-I ring [58, 59] functions as both a beam-pulse stretcher for nuclear
physics measurements, and a storage ring for synchrotron-light experiments. In
synchrotron-light mode it is capable of accelerating an electron beam up to 550
MeV and storing this beam for several hours. For the present experiment it was
used as a pulse stretcher to produce a CW beam. To achieve this, the ring is
injected every 1.3 ms with a 0.4 \ s long pulse from the microtron. After the
pulse-stretching process, the duty factor of the beam is increased from

0.01%

3.2. The MAX-lab

37

to 50%. The extraction process is controlled by the septum magnet (see Figure
3.3). As electrons lose energy in the ring from synchrotron radiation, they fall
into an orbit that the septum magnet extracts out of the ring. A beam-line then
transports the near continuous electron beam to the tagging hall, where bremsstrahlung photons are produced.

3.2.3

The MAX-lab Photon Tagger

The MAX-lab photon tagger  is capable of tagging photons in the energy range


= 2080 MeV. It consists of a magnetic spectrometer that focuses electrons

to a point along the focal plane (see Figure 3.5). The photon tagging energy
is calculated from the position of the electron detectors along the focal plane.
There are 64 electron detectors made from NE102 plastic scintillators material
and have an energy resolution of 78
G

300 keV. The tagger has two arrays of

32 detectors which can be moved independently, and cover an energy range of


about 10 MeV each. The tagging energy range is set by sliding the tagger along
rails aligned with the focal plane.
The spectrometer has a fixed magnetic field of approximately 0.3 T. An electron passing through this field will be deflected in a circular path, with a radius
proportional to its energy. Therefore, electrons of a given energy will all cross
the focal plane at the same position. The positions on the focal plane have been
calibrated to electron energy.
The efficiency with which photons are tagged is

25%. For some mea-

surements where the absolute cross-section of the reaction is required, a special


measurement of the tagging efficiency is made to determine  for each of the 64

detectors. The operating current of the beam that is extracted from the MAX-I

 The tagger described here was in operation at the MAX-lab from 1993 until 1999. A differ-

ent tagger was used before 1993 and a new tagger will be installed in 2001 which has different
characteristics from the previous two.

38

Chapter 3. Experimental Method

E = Ee E
e

Ee
Electron Beam

Target

Photon Beam
Collimator
Ejectile

Radiator
E = 0.8 Ee

Moveable
Focalplane
Detectors

Detector

Ee

Electron Paths

Exit
Flange

Magnetic Electron
Spectrometer

E = 0.1 Ee

To Electron
Beam Dump

Focal Plane
Under Vacuum
Not To Scale

Figure 3.5: The MAX-lab photon tagger.

ring is approximately 100 nA. This translates into a count rate in each focal plane
detector of about


 

electrons per second (see Appendix A). A complete

description of the tagger used in the experiment can be found in Reference [60].

3.3

Detecting Protons

In order to measure the reaction  Li  He, it is necessary to identify the


protons amongst a range of particles emitted from the  Li target. The photonuclear reactions  Li  ,  Li 

 Li )(

! ,  Li #" ,  Li #$% ,  Li '&

He  and

He  produce charged particles that are detected together with protons

from the  Li   He reaction. The detector system that was used, detected
charged particles in such a way that they can be sorted by type using computer

3.3. Detecting Protons

39

analysis. The following sections describe the principle of detecting protons and
the reaction chamber used in this experiment.

3.3.1

The GLUE Chamber

In previous collaborations between Gent University and Lund University [53,


61] the Gent group constructed the Gent Lund Universities Experiment (GLUE)
chamber. It consists of a metal vacuum chamber with target holders for an array
of *,+ - + detector telescopes, shown schematically in Figure 3.6. For the present
measurement, detectors were placed at angles of 30 - , 60 - and 90 - .

E
E

90

Cold Fingers
in Liquid Nitrogen
Cold Metal Plates

Photon Beam

Target
30

Beam Entrance Port

60

Beam Exit Port

Vacuum Chamber

Figure 3.6: Top view of the detector positioning in the GLUE chamber.

The photon beam enters and exits the chamber through thin mylar windows
attached at each end. In order to reduce any background caused by the beam
interacting with the mylar, the entrance and exit pipes are made as long as is
practicable. This configuration shields the detectors from a direct line-of-sight
view of charged particles emitted from the exit and entrance windows. Connected to the exit pipe is a turbo pump, that evacuates the chamber to a pressure

40

Chapter 3. Experimental Method

of approximately

 /. 

torr.

A target holder that can rotate through 360 degrees and move vertically, is
positioned in line with the beam. The spectrometers are arranged about the central axis of the target holder, on a circle with a radius of 100 mm. The product
particles are detected in the telescope array.
In order to distinguish between the different type reaction products, each
spectrometer consists of a thin

*
+

detector and a thick

detector. The

*,+

detector measures the partial energy loss of a particle as it passes through, while
the

detector is thick enough to stop all the charged particle of interest, and

measures the remaining energy. A comparison of the *,+ -signal to the + -signal
in off-line analysis allows the different types of charged particles to be identified,
and described in detail in the next two sections.

3.3.2

Solid State Detector Telescopes

Design The detector telescopes were designed to measure the

*,+

and the

of a charged particle for the purposes of particle identification. The principle of


particle identification is described in Section 3.3.3. Figure 3.7 shows the dimensions of a telescope and the materials used for construction. Each E detector was
mounted in a metal holder that was in thermal contact with a metal plate. The

*,+

detectors were mounted in aluminium holders and suspended from a rack

hanging from the top of the vacuum chamber.


The thin *,+

detector was made from silicon 500 0 m thick, and designed to

allow particles through into the thicker germanium detector, while losing only
part of their kinetic energy. In turn, the

detector was 15 mm thick HPGe,

and designed to completely stop the highest energy charged particles, in order
to measure their total kinetic energy. In particular, the highest energy protons
created in the  Li 1 He reaction with +32 = 5070 MeV have around 60 MeV

3.3. Detecting Protons

41

Signal Lead Connectors

15 mm Ge

35 mm

85 mm
500 m Si
35 mm
Brass Casing

Steel Holders

Figure 3.7: Charged particle detector telescope with a Si- 465

and a Ge-5 .

of kinetic energy, and a range of 9.5 mm in germanium.

Operation

In order to reduce the inherent random electrical noise the HPGe

detectors need to be operated at low temperatures, around -190 -%7 . This is


achieved by mounting them on a plate that is attached to a cold finger immersed
in liquid nitrogen (see Figure 3.8).
The bias voltages on each of the detectors were set to maximise the energy
resolution. The full width half maximum of the peaks produced by 8 -particles
from

99: Th was used as a measure of resolution. These peaks were also used to

calibrate the detectors.

42

Chapter 3. Experimental Method

Mylar Window
Beam

Mylar Window
27Al
7Li

Source

Cold Fingers

Liquid
Nitrogen
Dewars

Pump

Figure 3.8: Side view of the GLUE chamber.

Calibration An energy calibration of the

*,+

using 8 -particle from a 99: Th source. The *,+

and

detectors was performed

detectors were moved up from in

front of the + detectors, so each detector was irradiated by the source. A typical

-particle energy spectrum from the

99: Th decay-chain can be seen in Figure

3.9. The energy of the 8 -particles from this decay chain are tabulated in Table
3.1. The

*
+

and

detectors were calibrated on a regular basis to monitor the

stability of the gain.


Peak Number Parent Nucleus Energy (MeV)
1
5.42
99: Th
2
5.69
99 ( Ra
6.05
3
9<;9 Bi
4
9<;9 Bi
6.09
5
6.29
99= Rn
6
6.78
9<;  Po
8.78
7
9<;  Po
Table 3.1: A list of > -particle energies from a 99: Th source. The peak numbers
are labeled in Figure 3.9.

Counts (MeV-1)

3.3. Detecting Protons

43

A6

80

@5

B70
A60

@50

Th -Spectrum

228

F2

B7

E40
30
3,4
20
10

D0

@5

A6

B7

C
?Energy9 (MeV)

Figure 3.9: The energy spectrum of > -particles emitted from the calibration
source 99: Th, energy values are given in Table 3.1.

44

Chapter 3. Experimental Method

3.3.3

Charged Particle Identification Method

The two-component telescope detectors are designed to provide data that can
distinguish between different types of charged particles. This is achieved by
measuring the partial energy loss in the thin *
+ Si detector and the total energy
loss in the thick + HPGe detector, as shown in Figure 3.10.
t

E+ E

Charged Particle Path

500 m Si

15 mm Ge

Figure 3.10: Schematic of the 465 -5


cle identification

telescope detector system used for parti-

The energy loss per unit path length of a charged particle through matter, is
accurately described using the empirical Beta-Bloch relationship [62] given by

LMON
R S#U 9 V 9
^\ ]
Q( P 9 WYX[Z\^]Y_a` T
G "H+
b
c G <  G V 9  G V 9ed 
T
R
#
S
U
"/IKJ
%9 V9
where

RTS

is the electron rest mass, the particle has a rest mass

(3.2)

ghg RTS ,

U , energy loss "H+ along a path "/I in a medium with


V j
J W ilk
X
b
atomic number and atoms/cm & . The ionising potential of the absorbing

charge P , velocity

atoms is an experimentally determined parameter related to the electron charge


distribution. For particles with non-relativistic velocities, the energy loss in a
given medium can be reduced to

"H+
f P9
"/IKm + n

(3.3)

3.3. Detecting Protons

45

So for a given incident energy

and charge P , for example protons, deuterons

and tritons, particles can be separated out according to their mass.


The energy loss *
+ in the Si detector of thickness $ is given by

"H+sI
/" I
Jpor= q "/I
n

(3.4)

R P9
R P9
"/I
w$ 
mto = q +sI J +vu

(3.5)

*
+
Using Equation 3.3 we get

*,+
where

+s$%yxK+ u xz+s H .

In most cases, where particles are produced in pho-

tonuclear reactions with +32 = 5070 MeV,


relationship between

*,+

and

+ u~ +

. Thus, the

f P9$
m + n

*
+
f

and so

in the detector telescope can be approximated

by

The

*,+|{}+

(3.6)

dependence can clearly be seen in plots of

*,+

against

shown in

Figure 3.11. The ability to measure the *,+ and + of a reaction product, provides
a clean method of particle identification.
E (Channels)

He

1200

1000

He

800
600

tritons

400
200
0

200

400electrons600

deuterons
protons
800

1000
E (Channels)

Figure 3.11: A 465 -5 from a particle spectrometer, showing the different


bands of charged particles that are formed

46

Chapter 3. Experimental Method

3.3.4

Li Target

The lithium metal target was a 99% enriched  Li, measuring

L/

mm,

and covered by 8 0 m thick aluminium foil for protection. Lithium metal corrodes
readily in air and had to be transported in an air-tight container, filled with inert
argon gas. The container was open in the GLUE chamber which was also filled
with argon gas. While the target was being mounted, argon was continuously
flushed through the GLUE chamber to prevent it coming contact with air. Once
the target was mounted, the chamber was evacuated and could not be brought
back up to atmospheric pressure with air, until the experiment was complete.
The taregt was placed at an angle of 60 - to the incident photon beam.

3.3.5

Nuclear Experimental Hall (Cave)

An overview of the nuclear experimental hall, or the cave, is shown in Figure


3.12. The photon beam enters the cave through a port in the lead and concrete
shielding. A steel collimator was placed at the entrance port to define the beam.
The GLUE chamber was positioned as close to the collimator as possible to
minimise the beam spot size at the target. The beam spot was measured at the
target position using Polaroid film, and had a diameter of 20 mm.
Part of the electronics was assembled in the cave to reduce the noise pickup
on the cables. The noise arose from stray fields associated with the accelerator
and other electrical equipment. The pre-amplifiers and the amplifiers were as
close to the experimental chamber as possible, to ensure that a clean signal was
sent to the counting room for further processing. The counting room was near
the tagging hall, and contained the data acquisition system.

3.3. Detecting Protons

Tagger



Collimator

Electronics

Beam
Dump

47

GLUE

Photon Beam

Beam
Dump
Entrance
Concrete Pylons
Permanent Walls

Movable Pallets (Shielding)


Movable Detectors (Unused)

Figure 3.12: Schematic view of the layout of the experimental hall showing
the positioning of the experimental chamber and the tagging spectrometer.

48

Chapter 3. Experimental Method

3.4
3.4.1

Data Acquisition System


Overview

The data acquisition system (DAQ) processed the signals that were generated
by charged particles depositing their energy in the detectors. To process the
signals, an electronic circuit was assembled from NIM and CAMAC modular
electronics. If the signals triggered the correct response in the DAQ, the data for
that event was stored by the computer. The minimum trigger conditions were
a hit in both

*,+

and

of any one of the detector telescopes in the GLUE

chamber. The events stored by the computer could also be viewed on-line, to
tune the circuit and to monitor the progress of the experiment. Each step in the
acquisition process is describe in detail in the following sections.

3.4.2

Hardware Circuit

The hardware used in the DAQ system for this experiment consisted of modules
using different standards; both NIM and CAMAC. To optimise the signal quality,
the modules were split between the cave and the counting room.
The pre-amplifiers were located in the cave and connected to the detectors
with very short cables ( ~ 10 cm). Keeping them close to the detectors to reduce
their capacitive effect, and to minimised any noise they may add to the signal.
Each pre-amp had two outputs; an + -output and an integrated -output. The + output signals were connected to spectroscopy amplifiers and sent to the control
room. The

-output signals were connected to a timing filter amplifier (TFA)

to remove high frequency noise from the signal, and thereby reduce the timing
jitter. Constant fraction discriminators (CFD) were used to determine the arrival
time of the -signals. The delay times and thresholds of the CFDs were carefully
optimised for each detector, to minimise the walk-time of the logic output.

3.4. Data Acquisition System


The *,+

49

detectors provided a faster and more stable timing signal, and were

*,+

therefore used to establish the coincidence between the


The requirement of a hardware coincidence between the

*,+

and

and

detectors.
detectors

eliminated the need to process low energy particles that are stopped in the

*
+

detectors. Figure 3.13 shows how the narrow


falls within the wide output from the

*
+

signal is delayed so that it

CFD and establishes the coincidence in

the AND gate. In the counting room, the signals from the cave were attenuated
using decade boxes to match the dynamic range of the ADCs and CFDs.
Charged Particle
E
E

Amp

Delay

TFA

CFD

To ADC

PreAmp
T

AND
E

TFA

CFD

Amp

Delay

To XTrigger
Circuit

PreAmp
T

More stable Ge detector


determines the coin timing

Figure 3.13:

465 -5

To ADC

E
E

coincidence circuit for for a charged particle telescope

The timing between the tagger and the charged particle telescopes, i.e. the
photon tagging process, was not determined in hardware as was the

*,+ -+

tim-

ing, but rather in software during the off-line analysis. A time to digital converter
(TDC) recorded the time difference between the tagger and the telescopes. The
presence of the coincidence peak in the TDCs indicates that photons were successfully tagged (see Section 4.6.1).
Once an event triggered the DAQ, an interrupt signal was sent to the VME
computer to read the ADCs and TDCs that were in the CAMAC crate. After
the VME read the event, it sent a done signal to clear the CAMAC modules and
ready the DAQ for the next event. The VME operated on a Linux kernel and

50

Chapter 3. Experimental Method

was connected via an Ethernet network to a Sun workstation. The data rate was
low enough for the VME to read and temporarily store several events, before
transferring them to the workstation. The workstation saved the data to disk and
8 mm storage tape, as well as displaying it on-screen.
The DAQ circuit is shown in Figure 3.14. For clarity the figure shows only
schematically how the circuit was connected, omitting the delay modules, attenuators etc.
CFD
o

AND

30

60
o
90

CFD
E

NIM
OR

Machine
Trigger
Gate Generator
Inhibit
Gate

Input

ADC

XTrigger

CAMAC

Interrupt

VME
Computer

Busy

TDC

I/O

Start

Tagger

Read/Clear

Stop

Delay

FIFO

Ethernet

Data Stream


Data Tape
Workstation

Figure 3.14: The Logic circuit diagram for the data acquisition electronics.
Some modules explained in the text were left out for clarity.

A list of the modules and their function in the circuit follow:

3.4. Data Acquisition System

51

NIM Analog:
Ortec 140 A

Pre-amplifier

Ortec 590

Spectroscopy amplifier

Ortec 510

Timing filter amplifier (TFA)

Ortec 410
Constant fraction discriminator (CFD)
NIM Logic:
LeCroy 222 Gate and delay generator
LeCroy 622 Logic module
CAMAC:
LeCroy 2259 A Peak-sensing analogue-to-digital converter (ADC)
LeCroy 2229

3.4.3

Tagger time-to-digital converter (TDC)

Event Trigger

Determining if an event was of interest and should be kept, or be discarded, was


the role of the event trigger, or X-trigger. The essential criterion for accepting an
event was that a charged particle was observed in both the *
+
Signals from the

*,+ -+

and +

detectors.

coincidence module were fed into a gate generator to

produce the X-trigger. The gate could be inhibited by the presence two other
signals: the microtron inject signal and the VME busy signal.
RF-noise is produced during the injection stage in the microtron, and is
picked up by the electronics in the DAQ. In order to eliminate this noise, an
inhibit signal is generated from the inject signal.
Once the VME computer receives an interrupt, and while it is reading the
CAMAC, it send out a busy signal. To prevent other other events from trying to
trigger the DAQ, the busy signal is used to inhibit the X-trigger.
The X-trigger also provided a common start signal to the 64 tagger TDCs. If
there was a hit in any of the taggers, that detector signal would provide a stop for
one TDC. All 64 of the signals from the tagger detectors are delayed to arrive

52

Chapter 3. Experimental Method

after the X-trigger. Figure 3.15 shows a single tagging event and the relative
timing of the signals.
Tagger Signal

Xtrigger

Delayed

200 ns Delay

Tagger Signal
400 ns TDC Timeout

TDC Common
Start
TDC Single
Stop
Figure 3.15: The relative timing of the X-trigger and a single tagger signal.

This type of X-trigger resulted in essentially 64 separate experiments, one for


each tagger. Each detector tagged a narrow photon energy range ( ~

keV).

In the data analysis process, all the experiments were summed together to form
the final result. This approach is valid provided the proper care is taken with
the analysis process. The details of the data analysis are presented in the next
chapter.

3.5

Summary of Experimental Parameters

Table 3.2 lists some parameters of the beam and detector systems as they were
for the present experiment.

53

+ S

Parameter

Tagging Magnet Field

+2

Radiator

*
+32

Left tagger position


Right tagger position
Angles Measured,
Average Tagging Efficiency,
Detector Solid Angle, "H
 Li target thickness
 Li target purity

Value
92.45 MeV
0.31900 T
5070 MeV
50 0 m Al foil
310 keV
391 mm
71 mm
30 - , 60 - , 90 0.26
54 msr
915 0 m
99.9 %

Table 3.2: List of experimental parameters and their values.

54

Chapter 4
Data Analysis
The aim of the data analysis procedure is to extract the population of states in the
residual nucleus  He, following the reaction  Li ! He. In order to achieve
this, it is necessary to determine the energy of the emitted proton, and initiating
photon in each





reaction. A well established technique based on similar

experiments previously performed at the Max-lab [54, 63, 64] is used to

perform this analysis.

4.1

Analysis Overview

All of the off-line data analysis was performed on a Linux PC using the ROOT
analysis package. This system was fast, flexible, and provided a dedicated environment specifically tasked with the complex data reduction procedures. A
sequence of constraints was used to reduce the raw data, and each of these is
further discussed in detail in the subsequent sections. These steps consisted of:
1. Discriminating the protons from other emitted charged particles.
2. Determining the energy of the photon that induced the reaction.
3. Determining the proton energy and consequently the excitation energy in
 He.
55

56

Chapter 4. Data Analysis

4. Removing any background contribution in the excitation energy spectrum.

An event-by-event processing technique was used by the analysis software,


allowing the experiment to be rerun on the computer, with stringent triggering
requirements. As a result of each analysis step the data was reduced to smaller
subsets, thereby reducing the subsequent analysis time.

4.2

ROOT/CINT Software

The analysis software was compiled using the ROOT package, which was developed at CERN for the NA49 experiment [65]. ROOT is essentially a set of
C++ classes, specifically designed to process data generated in a nuclear (or particle) physics experiment. The data acquisition software, written by staff at the
MAX-lab [66], was also compiled using ROOT, and stored the data in a format
readable by ROOT. As a consequence, the analysis could proceed without the
need to convert the data into another format.
Integrated into the ROOT package is a program called CINT (C interpreter),
that can execute macros written in the C/C++ language. CINT reduces the time
associated with the debugging, linking and compiling of C-code into a program.
This is achieved by not requiring the full formalism of the C/C++ language
in the macros, and allowing real-time interaction with the code by means of a
command-line interface. To make the C-code run faster and more efficiently, it
can later be compiled into a program.
The programs that were used to analyse the data from the present experiment required code that was specific to photonuclear reactions. This specialised
code, written by the author, was first debugged and tested using macros in the
ROOT/CINT package. Later, to speed up the analysis, the macros were compiled
into programs that analysed the data to extract the population of states in  He.

4.2. ROOT/CINT Software

57

Other software designed to analyse physics experiments, such as the popular


package PAW (Physics Analysis Workstation) [67], also created at CERN, often
use row-wise ntuples (RWN) format to sore data. By contrast the data format
used by ROOT is a more convenient tree structure, and the data is stored in
compressed-binary files that can be read by any operating system with ROOT
installed. Data stored in a tree structure can be searched and retrieved more
quickly and more efficiently than data in a RWN structure. Each event recorded
during the experiment was stored in a data tree. The ADC and TDC modules
were stored in the tree as logical objects called branches, while the values read
from the modules were stored as leaves on the branches.
To extract the data from the tree structure stored in the data files, a branch
address is defined in memory and linked to the file containing the data tree. In the
example below, the tree event is linked to datafile.root and a histogram
he60 is defined. The macro then loops over all the events in the tree, and if the
ADC channel is greater than 163, the 60 proton data is stored in the histogram.
TFile *file = new TFile("datafile.root"); //open data file
Event *event = new Event();
file->SetBranchAddress("event", &event); //map event
TH1F *he60 = new TH1F("he60","E 60",256,0,1024);
Int_t nentries = file->GetEntries();
for(i=0; i<nentries;i++) { //start event loop
file->GetEvent(i);
fe60 = event->GetADC1(e60); //E detector 60 degrees
fde60 = event->GetADC1(de60); //dE detector 60 degrees
if(fe60 > 163){
he60->Fill(e60);}
event->Clear();}
// end event loop

The event loop in the complete analysis code, contained a far more compli-

Read event

cated set of trigger conditions than the


preceding macro. A flow diagram of

Choose detector angle


Particle identification
Proton?
Correct for energy loss
and determine T p

58

Chapter 4. Data Analysis

the triggers and calculations required to


generate the missing-energy spectrum
of Li  , is shown in Figure 4.1.
The steps shown were not performed by
a single program, but the process was
broken up into smaller sections using
dedicated programs. Each program reduced the data to a smaller subset, so
that subsequent analysis could be performed more rapidly. Derived values
that were calculated from the experimental data, such as missing-energy
(  ), were added as new variables
in the event structure, and saved in the
data file. Moreover, these values could
later be used in trigger conditions, without the need for them to be recalculated.

4.3

Particle Identification

Protons were separated from other


charged particles that were detected, using a

, -

particle identification (PI) method. The method relies on a varia-

tion in the energy loss of particles with different charge/mass ratios (see Section
3.3.3). Plotting the total energy , against the partial energy loss , , separates
the singly charged particles into four distinct bands of different mass: electrons,
protons, deuterons and tritons. A polygon cut was applied to the PI plot to iso-

4.3. Particle Identification

59

late a small data-subset that contained only proton events. Figure 4.2 shows a
2-D histogram of plotted against ,

for = 30 , with the polygon cut drawn

around the proton events. Further analysis was performed only on the events in
the proton data-subset.

E (Channels)

600
500
tritons

400

deuterons

300

protons

200

100

200

electrons

400

600

800

E1000
(Channels)

Figure 4.2: A typical particle identification plot for the = 30 detector. The
polygon represents the cut used to select the proton events.

The deuterons and tritons seen in the PI plot were produced by

#%

%

and

reactions in Li; however the number of these events was two orders of

magnitude smaller than the number of proton events. Therefore little significant
information could be obtained from these reaction channels. However, the superior performance of the , -

detection system is demonstrated by the clean

separation between the charged particle groups in the PI spectrum.

60

Chapter 4. Data Analysis

4.4

Photon Energy Measurement

The photon tagger consisted of two arrays of 32 electron detectors, located on the
focal plane of the spectrometer magnet (see Section 3.2.3). The magneto-optical
properties of the spectrometer magnet determines the energy of the electron that
reaches a particular position on the focal plane. Thus the response of a detector
on the focal plane identifies this energy, and hence the photon energy that is
tagged. To determine the tagging energy of each detector, the program POS [68]
was used. This program calculated the photon energy as a function of detector

E (MeV)

position using the known properties of the spectrometer magnet.


64

70
65
33
32

60

55
50

1
-200

-100

100
200
300
Focal Plane Detector Position (mm)

Figure 4.3: The calibration plot of photon energy corresponding to tagger detector

Figure 4.3 shows the results of the calculation used for the present measurement, with the incident electron beam energy set to apH/ MeV. Photons in
the energy range

6062 MeV were not tagged, due to a gap between the two

detector arrays. The electron detectors were all

mm wide, and the centre of

each electron detector defined its position on the focal plane. The precise tagging
range was from 50.81 MeV to 71.81 MeV. Over this tagging range, the photon
energy resolution varied from 250 keV at the lowest, to 270 keV at the highest
energy photons.

4.5. Proton Energy Measurement

4.5

61

Proton Energy Measurement

In order to measure the population of states in He following the Li 1 He


reaction, the emitted proton energy must be known. As a starting point, the , -

detector system was calibrated using an -particle source (see Section 3.3.2),

so that the energy deposited in the

detector by protons could be measured.

However, the energy deposited in the

detector is not equal to the energy with

which the protons were emitted from the nucleus. Protons emitted from Li
suffered energy losses as they passed through part of the target, and the
detector, before being stopped in the

detector. So, an energy-loss correction

was made, and applied to the detected energy, to determine the energy with which
the protons were emitted.

4.5.1

Energy Loss Corrections

To calculate the energy lost by a proton passing through the target and the
the thickness of each material must be known. The ,

detector was made from

silicon with a thickness of 500 m. However, the thickness of the target through
which a proton passes varies for each event, since the proton can be produced
anywhere in the irradiated region of the target. To calculate the energy loss in
the lithium, the approximation was made that protons were produced at the center
of the target. The thickness of the target material through which the proton had
to travel also depended on the emission angle, and was determined by the angle
at which the proton was detected.
Once the thicknesses were determined, the energy lost in each material was
calculated using a table of energy loss values [69]. First the proton energy loss
in the lithium target was calculated, then this reduced kinetic energy was used
to calculate the energy loss in the silicon

. To determine the original proton

62

Chapter 4. Data Analysis

energy

11

, the total calculated energy loss was added to the detected proton

energy <

11r,
where

Li

and

Li

Si

(4.1)

are the calculated energy losses of a proton in lithium

Si

and silicon respectively. At low proton energies the corrections were

4 MeV,

whereas at the highest proton energies they were 1 MeV.


The values of 11 were plotted against a range of  for each detector angle, an example of such a plot is shown in Figure 4.4. The curve shown is a
third-order polynomial fitted to the calculated points. This provides the correc-

Original Energy, Torig (MeV)

tion !1 and  and the equation is shown in Figure 4.4.

90

floss(x) = 3.979 + 0.8648 x + 0.001969 x2 - 9.817e-06 x3

80

70
60

50
40
30

20
10

10

20

30

40

50

60

70 80 90
Detected
Energy, T (MeV)
det

Figure 4.4: A plot of the function used to convert the detected proton energy
to the original proton energy for = 60 .

To check the consistency of the energy loss correction and the proton calibration, the expected value of the proton energy was calculated as described in
the next section, and compared with 11 .

4.5. Proton Energy Measurement

4.5.2

63

Reaction Kinematics

The kinematic equations of the reaction Li  He were used to determine the
expected proton emission energy (H ), to check the energy loss calculation and
the detector calibration. A schematic diagram of the Li ! He reaction is
shown in Figure 4.5. The masses used in the calculation were taken from Audi
et al. [43].

Tp

mp
E p pp
p

mLi

E p

m He
EHe pHe
THe
Figure 4.5: A schematic diagram of the reaction kinematics of photo-proton
emission from lithium.

To calculate the proton kinetic energy

, relativistic kinematics and con-

servation of energy and momentum were applied. The equations for relativistic
energy and momentum for a particle of mass

are given by

v

where

is the total energy,

momentum, and

(4.2)

(4.3)

Tt r

(4.4)

is the kinetic energy of the proton and

is its

. The energy and momentum conservation relations

64

Chapter 4. Data Analysis

are given by

p3

Li

p3r

(4.5)

He

(4.6)

He

 
Q% #/

(4.7)

Using Equations 4.2 to 4.7 the proton kinetic energy is given by [70]

l!#
#
 y
/
l #

(4.8)

where

y
y
H e ,
is the initial total energy,

Li

are the total energy and momentum of the incident photon,

is the mass of the Li target nucleus,

, , , #

are the total energy, momentum, mass and emission angle of

the proton,

He ,

He ,

He

are the total energy, momentum and mass of the residual He

nucleus.

 protons were calculated from Equa-

The expected kinetic energies of the

tion 4.8 above for each value of


Section 4.4 as to how

corresponding to the 64 tag channels (see

was determined). These values of



with the energy of the observed

were compared

energies, as determined using the -source

calibrations, and the agreement within

300 keV. However the poor statistics in

each proton spectrum limited the accuracy of this comparison.


In the experiment, 64 proton-energy spectra are produced by triggering on
single tag channels (i.e. one proton spectrum for each photon energy). Each of

4.5. Proton Energy Measurement

65

these spectra would under ideal conditions be analysed separately. In the present
measurement, none of the 64 proton spectra contained more than 10 counts in
the



peak. This made it very difficult to obtain meaningful results from the

individual spectra. To overcome this difficulty, proton energy spectra were converted into missing-energy spectra so they could be summed together. Since the
missing-energy is invariant with respect to photon energy, all the spectra were
summed together to produce a missing-energy spectrum integrated over

5070 MeV. The statistics in these missing-energy spectra were sufficient to accurately check the calibration of the , - detectors. The next section describes
how the missing-energy spectra were obtained.

4.5.3

Missing-Energy

Referring to Figure 4.5, the missing-energy,

  , of the Li 1 He reaction

is defined as

  3 H

He

(4.9)

Substituting in Equations 4.2 to 4.7 and rearranging gives



  <3

Li

H
l 
3

H  #

Li

(4.10)

The missing-energy is related to the excitation energy of the residual nucleus

He by

 


where

(4.11)

is the reaction threshold, and for Li ! He is given by



T y Li  He  e

 

MeV

66

Chapter 4. Data Analysis

where

is the mass of the target and



is the mass of the reaction products.

Therefore in a missing-energy spectrum of protons from the reaction Li  He,


the ground state proton peak should appear at   = 10 MeV.
In the initial



spectrum produced by summing together the data from

all the tag channels, the ground state peak appeared 0.5 MeV above the known
value. A discrepancy between the measured and known value was not unexpected, since the calibration of the detectors was only accurate to within a few



hundred keV (see Section 4.5.2). In order to align the

peak in the summed

missing-energy spectrum with   = 10 MeV, the calibration of the detector


was modified down by a few hundred keV.



After aligning the

peaks, missing-energy spectra were produced for data

at 30 , 60 and 90 . A typical missing-energy spectrum summed over all tagged

Counts (MeV-1)

photon energies 3 = 5070 MeV is shown in Figure 4.6.

400
350
300
250
200
150
100
50
0

10

20

30

40

50

60
Emiss (MeV)


 and  = 50 70 MeV.

Figure 4.6: A missing-energy spectrum at



The summed missing-energy spectrum shows the


at

 

peak at

peak correctly aligned

= 10 MeV. Protons emitted to the first excited state in He, form the

 a

12 MeV. The energy resolution of the present measurement is

clearly sufficient to resolve these two states.

4.6. Correction for Accidental Tagging

The fall off in the spectrum from 40 to 62 MeV is a result of summing together 64 spectra with different end-point energies. At   50 MeV a small
dip can be observed, that is caused by the gap between the two detector arrays in
the tagger (see Sections 4.4 and 3.2.3). The spectrum has been arbitrarily cut-off
at  = 0 MeV.
The majority of events in the spectrum in Figure 4.6 result from accidental
proton coincidences (accidentals). In order to correct for this accidental component, it was necessary to examine the TDC timing spectra that established the
coincidences. The next section explains the analysis of the TDC spectra and
how the accidental tagging component was subtracted from the missing-energy
spectra.

4.6
4.6.1

Correction for Accidental Tagging


TDC Timing Spectra

The TDCs measure the time between the detection of a tagging electron and the
detection of a proton. Accidental tagging events form an approximately constant
background across the whole spectrum. The exact shape of the accidental tagging background is discussed in more detail in Appendix A. Correlated tagged
events lie in the range 750 to 1000 leading to the formation of a peak in this
range. A typical sum of 64 TDC timing spectra is shown in Figure 4.7 (where
the full range of the TDC was 400 ns).
To produce the missing-energy spectrum described in Section 4.5.3 (see Figure 4.6), a cut was made on the prompt timing region shown by the dark shading
in Figure 4.7. However this region also contains accidental events, that could
not be distinguished from correlated events in event-by-event analysis. In order
to remove this contribution, an accidental missing-energy spectrum was pro-

67

68

Chapter 4. Data Analysis

Counts

9000

6000

4000

8000

&

50

100

150

200

250

'#

300

Time (ns)
350

64

Correlated
4 ns FWHM

7000

5000

3000
2000

"

1000
0

500

1000

1500

Uncorrelated

2000

2500
3000
3500
Time (0.098 ns per Channel)

Figure 4.7: A TDC spectrum showing the prompt timing region and the accidental timing region.

duced by cutting on the accidental tagging region shown by the light shading in
Figure 4.7. The accidental spectrum was then normalised and subtracted from
the prompt spectrum to produce a corrected missing-energy spectrum.
An overview of the correction procedure is shown in Figure 4.8, while a
detailed description of each step is presented in the following sections.

4.6. Correction for Accidental Tagging

500

1000

69

1500 2000 2500


Time (Channels)

Calculate missing energy


of protons

10

20

30
40
50
Energy (MeV)

60

3000

3500

Calculate missing energy


of accidentals and normalise

70

10

20

30
40
50
Energy (MeV)

60

Subtract accidentals
and rebin

10

20

30
40
Energy (MeV)

50

60

70

Figure 4.8: Overview of the technique used to remove the accidental tagging
contribution from the proton missing-energy spectra.

70

70

Chapter 4. Data Analysis

4.6.2

Prompt Missing-Energy Spectrum

Figure 4.9 shows the prompt missing-energy spectrum for


the

 and

90 . The

peaks can be seen at   = 1012 MeV. The counts between   = 0

10 MeV result from accidentally tagged events, since they are below the Li 
reaction threshold. Indeed only 22% of the counts in the proton spectrum are

Counts (MeV-1)

from correlated events.

350
300

250

(,p) Threshold

200
150
100

50
0

30
35
40
*Missing
Energy (MeV)
Figure 4.9: The missing-energy spectrum at = 90 showing the ,  and the ,
peaks.

10

15

20

25

One way to remove the contribution from accidentals to the missing-energu


spectrum is to accurately determine the shape of a missing-energy spectrum
constructed soley from accidentals. The subtraction of this spectrum after normilisation should remove all of the strength below the reaction threshold at   =
10 MeV, and reveal the true missing-energy spectrum.

4.6. Correction for Accidental Tagging

4.6.3

71

Accidental Missing-Energy Spectrum

The accidental missing-energy spectrum shown in Figure 4.10 was produced by


cutting on the accidental timing region shown lightly shaded in Figure 4.7, and

Counts (MeV-1)

characterised by a featureless exponential structure up to   = 40 MeV.

300
250

200
150
100

50
0

10

15

20

25

30
35
40
*Missing
Energy (MeV)

Figure 4.10: The characteristic shape of the random coincidence proton


missing-energy spectrum, normalised and fitted with a polynomial.

This spectrum was normalised, to ensure that the total number of counts in
the spectrum was equal to the number of accidentals in the prompt timing region.

-/. , was calculated from the counts in two different


timing regions of the TDC spectrum: accidental in the correlated region 021 ; and
accidentals in the uncorrelated region 043 . These regions are shown in Figure
The normalisation factor,

4.11. The functional form of the accidentals in the TDC spectrum has been
determined by Owens [50] to be weakly exponential. However in the present
experiment, there is an accelerator-related underlying substructure in the TDC
spectrum that can be approximated by a sinusoidal function [52]. This feature

72

Chapter 4. Data Analysis

021 , a function of the form -%6587:9O <; %>=@?BA DCF E was fitted to
0 .
the TDC data in the region v
of the TDC spectrum is discussed in more detail in Appendix A. In order to
estimate

The normalisation factor is given by

Counts

-G. 0420 31
(4.12)
where 041 is determined by the integral of 
- % in the prompt timing region.
M
4500

K
K

4000

3500

f(t)

3000
2500

NP

2000
1500

Nb

Na

1000

500
0

500

1000

1500

2000

2500

KH

3000
3500
Time (Channels)

Figure 4.11: TDC timing cuts containing the three different regions used to
calculate the normalisation factor for the accidental spectrum.

A third-order polynomial was fitted to the normalised accidental spectrum


spectrum between

 

040 MeV, as shown in Figure 4.10. This fit was then

used in the subtraction of the accidental tagging component from the prompt
missing-energy spectrum.
Figure 4.12 shows the accidental-corrected missing-energy spectrum at

60 . The error bars shown are the statistical errors from the uncorrected prompt
missing-energy spectrum (see Figure 4.9), and the data has been plotted in 400

Q

 O

4.7.

Correction

73

Counts (MeV-1)

keV energy bins.

SLi(+,p) THe

400

= 60
300

200

100

0
5

10

15

20

25

30

35
40
Emiss (MeV)

Figure 4.12: The background subtracted proton missing-energy spectrum for


= 60

Below the Li  reaction threshold at   = 10 MeV, the average counts


drop to zero as expected, providing justification of the method used to remove
the accidental tagging contribution. The systematic error in the accidental subtraction process is estimated to be about 2%.

UWVYX/Z\[^]

4.7

Correction

Protons produced in multi-nucleon emission reactions, such as



Q

  ,   ,

and % , also contribute a background in the missing-energy spectrum.

Protons emitted by such reactions are indistinguishable from those emitted from
a Li  reaction, and consequently must be removed.
The

Q

 O

Q reaction

channel is particularly important, since the Li 

threshold is 11.91, MeV and falls precisely in the missing-energy region of the

74

Chapter 4. Data Analysis


low-lying excited states in He. Since the cross sections of the

 

reaction is



and

100 time smaller than the

Q

 

reaction [71, 72], and the

cross sections are even smaller, only the


An estimate of the

Q

 

!

%

channel was considered.

contribution in the missing-energy spectrum was

made by MacGregor [73], using a Monte Carlo program to calculate the momentum of the emitted protons. This calculation used a quasi-deuteron model
of the nucleus in which photons were absorbed on proton-neutron pairs. The
initial pair momentum was selected from a distribution that was derived from an

_ He was left in its ground state, and that there were no final state interactions,

harmonic oscillator wavefunction. On the assumption that the residual nucleus

conservation of energy and momentum were applied to derive the momentum


with which protons were emitted. The angular distribution of the emitted pro-

Q

tons were determined by the angular distribution of the H 

reaction [74].

The calculation included all the experimental parameters of the GLUE chamber
detection system, and covered the full phase space of the experiment. A detailed
description of the calculation appears in Reference [75].
In order to compare the results of the calculation with the experimental missing-

Q reaction was


_
calculated assuming a recoil mass of He and not He.
Figure 4.13 shows the 
Q missing-energy distribution for = 60 (shaded
area). The Q contribution was normalised such that after subtracting it,

energy spectrum, the missing-energy of the protons from the 

the net missing-energy spectrum was positive at all energies. The normalisation
factor used for the 60 data was also used for the 30 and 90 data, since the
calculation was internally consistent for all angles.

QO contribution is essentially structureless, with its broad

Importantly, the 

resonance peaking at approximately  = 3035 MeV. This is well above the

region of interest where new states are predicted in He, and leaves no doubt

Counts (MeV-1)

4.7.

Q

 O

Correction

75

500
400

300

200
100

SLi(+,pn)

0
0

10

`badcD,:egf

15

20

25

30

E35

miss

40

(MeV)

Figure 4.13: The calculated proton missing-energy spectrum from the reaction
from a Monte Carlo 2 photon absorption model.
Li

ikjmlonn = 20 MeV is not due to structure in the prqtsuvtw


background. The final missing-energy spectrum at x = 60 y with the accidental
and the pzqtsuvtw contributions subtracted is shown in Figure 4.14. The spectra for

that any structure below

the other angles are presented in Chapter 5. From these missing-energy spectra,

the population of states in He can be determined.

Counts (MeV-1)

76

400
350
300

200
250

150

~50
}0

100

}0

~5

10

15

20

25

30

|E35R

miss

`bagcD f {

`bagcD:egf

40

(MeV)

Figure 4.14: The missing-energy spectrum of protons from the reaction


Li
He at = 60 with the accidental and
contributions removed.

Chapter 5
Results and Discussion
5.1

Introduction

The goal of the present experiment was to measure the nuclear levels of He
in order to observe new states that have been predicted by theory. A tagged

pzqtsudw{ He. The data from this experiment was analysed, and

photon experiment was conducted to measured the energy of protons emitted


by the reaction Li

missing-energy spectra were produced to determine the relative population of

states in the residual nucleus He.


Figure 5.1 shows the known low-lying excited
23.2

15.5
(1,2)

13.6

states of He up to an energy of

ik = 30 MeV [41],

while Table 5.2 presents a summary of the most


recent theoretical predictions and experimental re-

1.8

0
J

0
MeV

6He

sults. The basis of this chapter is to achieve a reciliation between the theoretical picture and the real-

Figure 5.1: He en-

ity of the experimental results. To achieve this, the

ergy levels [41].

results from the present measurement will be compared and evaluated against these other experiments and predictions. Through

these comparisons, a consistent understanding of the level structure of He will


77

78

Chapter 5. Results and Discussion

be presented.

The low-lying level structure in He has long been considered complete and
consistent. It was not until the experiments by Tanihata et al. [2], and the subsequent interpretations by Hansen and Jonson [28], that the picture shown in Figure

5.1 required revision. From these studies it emerged that the structure of He is

well described as a He core with a two-neutron halo. Using the halo model,
new states were predicted in the region between the known states at 1.8 and 13.6
MeV, including a new low-lying collective resonance, called the soft DR (see
Section 2.1.4). However, experiments designed to verify these predictions have
so far failed to provide conclusive measurements that clearly identify the new
states.

As mentioned in Chapter 2, the experimental studies of the structure of He


have primarily been conducted using beams of stable or radioactive ions. The
present experiment used a tagged-photon beam to provide a measurement of

the nuclear levels in He that is independent of the other experimental methods.


Furthermore, the analysis techniques used to extract the tagged-photon data are

pzqtsugw { He reaction, provide a clear picture of the

more rigorous than those used in some of the previous experiments. Therefore,
the results of measuring the Li

nuclear levels in He, and any new states that exist should be observed.
The interpretations section of this chapter present a discussion of the present
measurement in the context of halo nuclei as developed in Chapter 2. The discussion is separated into two sections: the low-lying states in the region
MeV, and the higher states in the region

ik4G

ik6G

MeV. This distinction is made

prqmsugw reaction leading to the


population of states in { He; while the high region is formed by the inclusive pzqtsudw
because the low region is formed by the exclusive

reaction involving more than two-particle emission. The high region contains a
complicated mix of contributions from several reactions that do not necessarily

5.2. Results

79

lead to states in He. The possible nature of both these regions is discussed in
Section 5.3.

5.2
5.2.1

Results
Excitation Energy Spectra

prqtsugw , measured at 30y , 60y and 90y . Prior to preseting these results,

This section will present the excitation-energy spectra of He following the reaction Li

the validity of making the conversion from missing-energy to excitation-energy


will be discussed.
Figure 5.2 shows the missing-energy spectra described in Section 4.7 for each
angle. There are three significant features in each of the missing-energy spectra:
1. The clear and strong population of the ground and first-excited state.
2. A dominant broad structure above

ikjmlonn = 22 MeV.

3. Evidence of strength between these features.

The thresholds of each of the inclusive

prqmsugw reactions are indicated on the

ikjmlonn = 1022 MeV is exclusively populated by protons leading to excited states in { He, since the pzqtsuvtw component has been removed from the spectrum (see Section 4.7). Above ikjtlnn = 22 MeV several
60 axis. The region from

other three- and four-body breakup reactions become possible, and the structure

in this region can no longer be uniquely identified as states in He. Therefore


converting form missing-energy to excitation-energy is only valid from
022 MeV. The region above

ikjmlonn =

ijtlnn = 22 MeV will be discussed in Section 5.3.3.

80

Chapter 5. Results and Discussion

Counts (MeV-1)

500
= 30

400

Li(,p)6He

300
200
100

400

10

15

20

25

30

35

40

15
(,pn)

25
30
35
40
(,pt)dn (,pd)4H (,pp)5H

15

= 60

300
200
100

250

10
(,p)6He

20
(,pt)t

200
150
100

50

bgD {

10

20

25

30

y y

35
40
Emiss (MeV)

Figure 5.2: The missing-energy spectra of He following the reaction


Li
He, integrated over
= 5070 MeV at = 30 , 60 and 90 .

5.2. Results

81

The excitation energy, ik , of the residual nucleus { He following the reaction
Li pzqtsudw { He, is related to the missing energy, ikjmlonn , by

ikikjmlnns
where

MeV is the reaction threshold. The missing-energy spectra

were converted to excitation-energy, and integrated over all photon energies (

y y

= 5070 MeV) and all angles (30 , 60 and 90 ). Figure 5.3 shows the resultant

Counts (MeV-1)

spectrum.

700
800

600
~500
400
300
200

}0

100

-10

~5

}0

-5

10

|E

ex

15
(MeV)

Figure 5.3: The excitation-energy spectrum integrated over all angles.

In order to determine the centroid energy of the peaks, Gaussian functions


were fitted to the spectrum. These fits specifically considered the large increase
in the cross section at

ik4

12. The values obtained for the centriod energies

were 0.0 MeV for the ground state, 2.0 MeV for the
broad state at 7.4 MeV.
The excitation-energy spectra at

y y

first excited state, and a

= 30 , 60 and 90 were also fitted with

82

Chapter 5. Results and Discussion

Gaussian functions, as shown in Figure 5.2. The energy and width of the peaks
in these fits were fixed to the values obtained from the angle-integrated spectrum,
and only the height of each Gaussian was left as a free parameter. The quality
of the data can be seen by the clear separation of the
excited state, at

ik

ground state and the

0 and 2 MeV respectively. The energy resolution at low

excitation energy in each spectrum is

ik2

1.0 MeV, which is mainly due to

the uncertainty in the energy loss straggling in the target and the

detectors.

The second largest contribution to the resolution is from the uncertainty in the
photon energy (

0.3 MeV). The remaining width is due to the kinematic

broadening caused by the angular acceptance of the detectors and the statistical
errors in the data.

Of the three angles, the 60 data has the highest resolution. This is to be

expected, since the target presents its thinest aspect to the 60 detector, and as a
consequence the energy straggling is at a minimum in that direction. The high
excitation-energy region of each spectrum, above

= 12 MeV, is formed by

low energy protons, for which the straggling effect is greater than for the higher
energy protons. Consequently the know states between

ik

= 1324 MeV (see

Figure 5.1) cannot be resolved, but are part of a broad structure in the high
excitation-energy region of the spectrum.
The features described above are all expected from this type of experiment,
and show that the results are of a high quality. Significantly, evidence of a new
broad state can be seen in each spectrum in the region between

ik

310

MeV. There are no well-defined states in this region, however there is significant strength above background suggesting the presence one or more broad resonances. The structure at this excitation energy confirms similar observations
made in other measurements, which are discussed in more detail in Section 5.3.

5.2. Results

-1

Counts (MeV )

400

350

83

= 30

Li(,p)6He

300
250
200
150

100

50
0

350

-5

300

= 60

10

15

10

15

10

250
200
150

100

50
0

120

-5

100

= 90

80
60
40

20
0

-20
-10

-5

15
Eex (MeV)

Figure 5.4: The excitation-energy spectra of He fitted with three Gaussian


functions.

84

Chapter 5. Results and Discussion

5.2.2

States Identified

The results of the fits to the data are presented in Table 5.1. The energy assignment to the new state at 7 MeV was on the assumption that it is a single broad
resonance. The quality of the fits can be seen by how closely the energy of
and the

states were determined compared with the accepted values. The ac-

curacy was determined quantitatively by calculating the reduced chi-squared of


the fits, which was approximately

d ik [76] i
0
1.8
0
1.8
0
1.8
{

1.5 for each angle.

(this measurement)
0.0 0.1
1.99 0.3
5.83 0.5
0.0 0.1
1.97 0.3
5.84 0.4
0.0 0.1
1.97 0.3
5.85 1.0

1.1
1.79
7.88
1.0
1.76
7.88
1.1
1.76
7.84

0.1
0.3
0.6
0.1
0.3
0.5
0.1
0.3
1.0

Table 5.1: A comparison of He energy levels from this measurement with


accepted values [76] (in units of MeV).

It was not possible to obtain a measurement of the energy and width of the
known 13.6, 15.5 or 23.2 MeV states, because of the onset of multi-particle
breakup reactions above

= 12 MeV (see Section 5.2.1). The

and

states are strongly populated in most reactions leading to He; consequently it


is essential for the validity of the present measurement that they are accurately
identified. On the basis of the good agreement of these known states, within
experimental errors, a new broad state was found at
width on

= 7.9

0.7 MeV.

i = 5.8

0.5 MeV with a

5.2. Results

5.2.3

85

Angular Distribution

pzqtsugw

The angular distribution of the protons emitted following the reaction Li

to residual states in He is shown in Figure 5.5. Each data point was obtained by
integrating the corresponding Gaussian function that was fitted to each peak in
the excitation-energy spectrum. These functions are indicated by the dashed lines

Li pzqtsudw .

Integrated Counts

in Figure 5.4, and represent the relative population of states in He following

Li(,p) He

10

}0

2+
+

10

20

40

60

80

100
120
Angle (LAB)

Figure 5.5: The angular distribution of the ground, first excited state and the
new resonance structure (lines are a guide only).

The lines are draw in to guide the eye, while the errors on the points include
the statistical errors in the integrated number of counts, and an estimate of the
error in the fits. Very similar angular distributions are observed for the
the

and

states, while the 7 MeV state deviate markedly from this trend. An anal-

ysis of this distribution on the basis of transitions from continuum states in Li

to states of various spin and parity in He, might reveal the nature of this new

86

Chapter 5. Results and Discussion

prqtsugw { He reaction [77]. Calculations

state. However, there are currently no calculations available that can predict the
angular distribution of protons from the Li

of photonuclear reaction, such as those by Ryckebusch [78], have been successful at determining many aspects of the reaction mechanism: for example the role
of meson exchange currents, direct knockout, pairing forces and final state interactions. However, these calculations are limited to even-even nuclei, and the

assumptions that are made about the symmetry in the nucleus are not valid for
photon absorption on Li [79]. So at this stage, no conclusion can be drawn from
the angular distributions presented here.

5.3
5.3.1

Interpretations
Comparison with Previous Measurements

Despite the present inability to interpret the angular distribution of the new state

in He, the energy levels that were measured can be compared with previous
measurements. Naturally a photonuclear reaction leading to the population of

states in He cannot be directly compared to the relative population of states


following other types of nuclear reactions. However the position of the states in

the excitation energy spectrum of He is independent of the reaction that is used


to measure it. Therefore the present results can be compared with those from
the five key experiments described in Section 2.2.3. Similarly, the calculation
discussed in Chapter 2 can also be used for comparison. Table 5.2 presents a

survey of the measured and calculated nuclear levels in He to which the current
measurement will be compared.

t t t
15

15

0.7
4

0
1.99
7.4

t t

Table 5.2: A summary of theoretically predicted and experimentally measured states in He (all values in MeV).

7.4

0.5

0
1.8
5( )

14.6

0.5

0.5

1.8
4.3

1.8

0.5
12.1

0
1.92
5.6

broad

1.04
11.1

0
1.8

Sakuta [15]

Boland [80]

E XPERIMENTAL M EASUREMENTS
Janecke [16] Aumann [17] Nakayama [19] Nakamura [20]

47 broad

Danilin [38]

0
0
0
0
1.81
0.26
1.75
0.04
1.72
0.04
1.78
0.1
3.43
4.75
4.3
1.2
4.0
1.2
3.7
1.2
3.75
6.39
4.5
wide 2
4.4
1.8
4.2
1.8
4.69
9.45
5
wide 4
6.0
6.0
6.4
6.0
broad non resonant broad non resonant broad non resonant broad non resonant

Suzuki [1]

T HEORETICAL C ALCULATIONS
Aoyama [44]
Danilin [8]
Ershov [9]

5.3. Interpretations
87

88

Chapter 5. Results and Discussion

Currently Known Energy Levels in He

The currently know structure of He is shown in Figure 5.6 next to the excitationenergy spectrum obtained from the present measurement. The ground state and

zp qtsudw{ He
g , while
measurement. The ground state has a half-life of 806.7 ms and
6 and a narrow width of
the first excited state is at k MeV with
:B MeV. These two narrow states, along with a series of broad states in the

continuum above \ MeV, dominate the cross section in the Li pzqtsudw{ He


the first excited state, which have been observed in all the experimental studies,
and are predicted by most calculations, are also clearly seen in the Li

25

30
Eex (MeV)

reaction.

20

23.2

15
5

10

(1,2)

15.5
13.6

2+
0
J

6He

MeV

1.8
0

Figure 5.6: An energy level diagram of He [41] compared with the excitationenergy spectrum from the present measurement, drawn with a smooth line.

The exact character of the broad states above 12 MeV have not been deter-

mined, as can be seen by the ambiguous assignment of the spin and parity of
the 13.6 MeV state, and the lack of any

As previously mentioned, the region above

for the 15.5 and 23.2 MeV states.

k

= 12 MeV cannot be uniquely

5.3. Interpretations

89

identified as states in He, but some contribution to the cross section is exspected
in this region (as indicated by the arrows in Figure 5.6).
States have been measured and predicted in the region

k = 212 MeV over

the past 10 years, but due to ambiguities in the results no definite assingments

have been made. In particular, the prediction of the soft DR in He due to the
nuclear halo [27,28] has led to several experimental attempts to find states in this
weakly populated region.
The broad nature of these resonances and the relatively low cross section of
the reactions employed, have made it difficult for any definitive conclusions to
be drawn about the soft DR [81]. However, the ion beam and radioactive beam
experiments which have thus far provided most of the data, have used some
ambiguous methods of background subtraction and make controversial assumptions in their analysis (see Section 2.2.3 and Reference [81]). In contrast, photon
tagging experiments, such as the present measurement, use a well established

ztg He data with the

method of background removal. The following sections will discuss the possible

nature of the new structure in He, by comparing the Li


previous theoretical and experimental results.

5.3.2

The Low-Lying Region,

Y 

MeV

Evidence for the Soft DR


The earliest calculation of the resonance energy of the soft DR, on the basis of

an energy weighted sum rule of electric dipole strength in He, was made by
Suzuki [1] and found to be 4.7 MeV. Subsequently Zhukov et al. estimated the
resonance energy of the soft DR, using a cluster Shell Model, to be

7.3 MeV.

Both these values agreed quite well with the data that was available at the time.
Figure 5.7 shows a comparison of a recent and more complicated calculation
by Ershov et al. [9] and three experiments. This calculation was of the reaction

90

Chapter 5. Results and Discussion

2+ 1+ 0+ 1

2+

4.3
1.8






1

2

14.6

0
J

6He
6

0
0
MeV

Li(n,p) 6 He

Ershov et al.

+
1.8 2

6He

Li(t , 3 He) 6 He

Nakamura et al.

6He

4
1.8





2
+

Li( 7Li, 7 Be)6 He

Nakayama et al.

6He

5.8
2.0
0

Li( ,p) 6 He
Present
Measurement
7

Figure 5.7: A comparison of the


levels predicted by Ershov et al. [9] with
experimental results from Nakamura et al. [20], Nakayama et al. [19] and the
present measurement, see text for details. (The height of the level diagrams
represents the region of the excitation energy spectrum presented by each of
the references.)

Li zg He, which was chosen because it was considered the most appropriate
reaction to observe the soft DR. Unlike previous results that were attempting to

calculate a resonant state in He, Ershov predicted a broad state resulting from
the overlap of continuum states of various spin and parity, including

,  , 

and   . The state shown at   = 4.3 MeV in Figure 5.7 is predominantly from
a

component, indicating the presence of a soft DR.

In this discussion, only the position of the levels found in He are considered,
because the relative population of states depends on the type of reaction used to
excite it. The 60
data from present measurement is in good agreement with
the results shown in Figure 5.7, particularly those by Nakamura et al. In both
experiments, as well as the known   1.8 MeV state, a new state is found at

5 MeV with a width of

8 MeV. On the basis of a distorted wave Born

approximation (DWBA), Nakamura et al. identified the state as predominantly


dipole in nature, with a strong negative parity component. They also argue that

5.3. Interpretations

91

the Li  He  He reaction is a good spectroscopic tool compared with other


charge exchange reactions, because the angular momentum transfer is limited to

the same type of transitions as the  reactions, but the energy and angular
resolution is better.
The state at

5 MeV in the Li   He reaction is also in close agree-

ment with that found by Nakayama et al. using the Li Li  Be  He reaction.


This reaction is very complicated, with many possible combinations of target and
projectile angular momentum transfers. However the analysis procedure demonstrated in Reference [49] was used to limit the degrees of freedom in the reaction,
and a

state was identified at  = 4 MeV with  = 4 MeV. The same reaction

was measured by Sakuta et al. [15], who obtained similar results and were the
first to claim to have observed the soft DR.
As mentioned in Section 5.2.3, the present measurement is unable to make
a ! assignment to the observed state. However it is consistent with the other
measurements in the energy level and width of the new state in He.
Mention the analogy between the soft DR in 6He and the pygmy resonance in
13C [82]. Refer to 5He and 12C which have no low energy collective resonance.

The Case for Other States


Despite the evidence for the soft DR discussed in the previous section, arguments are mounting against the existence of a

resonant state. The current

theory favours a complicated picture involving many states, and the simple interpretation by Nakayama et al. has come under some criticism (see for example
Vaagen et al. [81]). However the experimental data to support the alternative
formulation is still quite poor.
Resent calculations by Danilin et al. [38] suggest that the low-lying states
measured in He are not true resonances, but the result of the increase in the tran-

92

Chapter 5. Results and Discussion

sition strength to  and   continuum states that combine to form a pseudo


resonance. These calculations use the method of hyperspherical harmonics to
account for the large spatial extent of the He halo, and contain a strong presence of admixtures in the ground state. Figure 5.8 show the results of Danilin
et al. compared with measurements that support the " assignment to the state,
and the results from the present measurement.
/

0/

0/

0/

0/

/
('

0/

('

/ ('

21

2 ('
21 2+ 121 + 12
('
21 21 12 3
+
1.8
2 ('
0
0+
6He MeV

21
2+
0+
J
6

He( p,p)6 He

Danilin et al.

0/

('

&%
%

&%

%
'(
('

21

14.6

0/

'(

21

0/

5.6

- ('
.('

.-

6He

*)

*)

*)

2+
,+
1.8 2+
0 0+

,+

,+

*)

*)
,+

,+

6He

..
Janecke
et al.

Aumann et al.

He

&%

&%
$#

&%

$#

4 % &% $# %& #$
+
1.8 2

Li( 7Li, 7 Be)6 He

&%

He+n+n

0+

6He

5.8

$#

$#

2.0
0

Li( ,p) 6 He
Present
Measurement
7

Figure 5.8: A comparison of the 3  levels predicted by Danilin et al. [38]


with experimental results from Janecke et al. [16], Aumann et al. [17] and the
present measurement (at 4 = 60
, see text for details.

Analysis of the data by Janecke et al., on the basis of a DWBA calculation,


agree with Danlini [38] (and others, for example Myo et al. [83]) that the lowlying state between   = 35 MeV is a   resonance. However it has been
suggested by Timufeyuk [18] that the three resonances identified by Janecke
are positioned at the threshold energies and might be mistaken for non-resonant
background features. Despite not taking this non-resonant background into account, the results by Janecke are most often cited as the experimental benchmark
for charge exchange reactions.
The largest inconsistency in the experimental results has come from the breakup

5.3. Interpretations

93

reaction of He on Pb and C, measured by Aumann et al. [17]. The higher exci-

reaction appears smooth and fea



tureless, which seems incongruous with the Li   He and the Li Li  Be  He
tation energy region in this 57698;: He < =<

results. Aumann et al. is the only measurement that has seen a narrow low-lying
resonance, but some calculations have predicted such states [13, 83], and are in
reasonable agreement considering poor statistics of the data.
The probability of populating certain states in the residual nucleus depends
on how strongly its wave function overlaps with that of the target nucleus. Timofeyuk [18] suggests that for both the   and the single charge exchange reactions leading to He, the overlap integrals are very small and that no states should
be populated between   213 MeV. Indeed it is possible that the broad states
shown in Figures 5.7 and 5.8 are a combination of very weakly populated states
predicted by Aoyam et al. [7] and others.

If the latest theoretical interpretations are to be followed, the structure in He


following Li > He is more likely to be the   state predicted by Danilin et

al. [38], with small contributions from  ,  and @? unbound continuum states.

5.3.3

The High Region, ACBEDGF

HJI

MeV

large bump above 12 MeV made up of:


 JK reaction

14,16,23 MeV continuum states


GDR component

With the improvement of the experimental data a more consistent picture of


the halo nucleus He is emerging. The present measurement has provided clear
data on the nuclear levels in He. Future measurements should include a higher
resolution measurement of the angular distribution to provide a better picture of

94

Chapter 5. Results and Discussion

Figure 5.9: A missing-energy spectrum of protons following the reaction


Li LNMOQPSRT He. (Source: J. F. Dias et al., Nuclear Physics A 587, 434 (1995).)

the ! assignment to the new state.


The validity of the conversion from missing-energy to excitation-energy is

based on the assumption that the detected protons come from decays to He following exclusive Li   reactions: i.e. that the only missing energy is the excitation in the residual nucleus. This conversion is not valid if protons are detected

from reactions that involve multiple particle emission. The contaminations could

come from Li   U T He, Li   T H, Li  WVJ : H and Li  JK : H. In Sec-

tion 4.7 the   U contribution was considered, and a correction of 20% was

made to the missing-energy spectrum.

The Li  X total photoabsorption cross section was measured by Nefkens

et al. [72] to be 100 times smaller than the Li   Y cross section measured by
Stein et al. [71].
mention the relative cross sections for the (g,t)3H and (g,t)dn etc.

Chapter 6
Conclusion

A measurement of the reaction Li   He has revealed a new state in the excitation energy spectrum of He. By careful handling of the data, the characteristics of the new state was identified as follows:

MeV;

Z Excitation energy,  \[^]_


Z Energy width, 7[a`_

MeV.

The significance of this measurement, over the previous measurements of


He, is the unambiguous background removal process. Unlike the radioactive

beam and the ion beam measurement techniques, tagged photon experiments
measure the random background contribution. This enables the exclusive  
component of the data to be extracted very cleanly. Consequently, convincing
evidence has been found for a new state, predicted by the theory of neutron-rich
nuclei.
Clearly, radioactive beam experiments are the only method suitable for a
systematic and complete study of the properties of unstable neutron-rich matter.
It is serendipitous that photon tagging can be used to study He: very few halo
nuclei can be formed by impinging a photon beam on a stable and naturally
abundant target. Nevertheless, it is important where possible, to confirm the
95

96

Chapter 6. Conclusion

exciting new results from studies at the limits of stability.


Unfortunately the Li   He results are limited in what they can reveal
about the nature of the new state in He. Although an distribution was measured,

the lack of a theoretical calculation meas the spin and parity of the state cannot
be determined. Currently no calculation exists for the Li   He reaction with
correct quantum-mechanical treatment of the many-body problem [77].
Nonetheless the present experiment confirms the existance of a new lowlying state in the nulear energy levels of He. For the first time photonuclear
techniques have been used to support the findings made by ion- and radioactivebeam experiments. This thesis presents convincing evidence of the new state
using well known and unambiguous analysis techniques.

Appendix A
Analysis of TDC Spectra
A.1

Structure of Uncorrelated Contribution

In order to determine the uncorrelated contribution to the TDC timing spectra, it


is necessary to accurately determine the structure it produces. The structure is
determined by the count rate in the tagger and the microstructure of the electron
beam.
Figure A.1 shows three typical TDC spectra, each taken with different beam
currents. The uncorrelated background has an exponential structure that depends
on the count rate in the tagger [50]. At higher count rates the exponential decay
is more pronounced than for the lower rates. Furthermore the signal to noise
ratio is worse at high count rates, where the correlated events at channel 800 is
hard to distinguish above background. So for the present experiment the count

rate in the tagger was set to cbdfe  s ?cg , as shown in Figure A.1(c), in order to
optimise the signal to noise ratio.
Superimposed on the exponential shape of the TDC spectra is a sinusoidal
like variation of the uncorrelated background. This variation has been observed
in each detector and each tag channel, and in previous experiments at the MAXlab, and is due to microstructure in the electron beam.
97

98

Appendix A. Analysis of TDC Spectra

(a)

(b)

(c)
Figure A.1: Typical tagging spectra of 64 TDC channels summed together for
count rates of (a) hjik3mlnhpo s ?cg , (b) oikqrlChpo s ?cg and (c) oikstlChpo s ?cg .

A.2. Timing Resolution

The time scale on which the electron beam varies is approximately 50 ns,
half of the time it takes for the beam to make one full revolution in the stretcher
ring. An analysis of the microstructure in the beam by [52] revealed that this
translated into a periodic variation in the TDC spectra of 100 ns. The variation
can be interpreted as a probability distribution of the time difference between the
protons and electrons, and the prompt peak will always appear on a crest of the
uncorrelated background. Figure A.1 also shows a sinusoidal variation with a
period of 100 ns, and was modeled as an exponential plus a sine function. This
model was successful in estimating the uncorrelated contribution in the TDCs
(see fit in Figure 4.11), and consequently the correction factor was accurately
determined for the uncorrelated proton spectrum (see Section 4.6.3).

A.2

Timing Resolution

The timing resolution was determined from the width of the peak formed by correlated events in the TDC, called the prompt peak (see Figure 4.7). In the present
measurement a timing resolution of 4 ns full width half maximum was achieved.
Part of the width of the prompt peak, was due to a 1 ns time difference between
the time-of-flight of the highest and lowest energy protons. The remaining width
was caused by a small walk-time in the pick-off time of the detector signals by
the CFDs.
The focal plane electron detectors were made of NE102 scintillation plastic.
The pulses produced by electrons interacting with NE102 have a very fast characteristic rise time of u

2 ns. In comparison, the rise time of pulses produced

in the Si and Ge detectors is typically wv " ns, therefore it more difficult to


get accurate timing from these detectors. As a consequence, the overall timing
resolution is determined predominantly by the timing signal from the charged

99

100

Appendix A. Analysis of TDC Spectra

particle detectors.
To optimise the timing signal from the xy - 

detectors, the integrated z -

output from the pre-amplifiers was processed by a TFA. The TFA removed some
of the noise-jitter in the pulses. The walk-time was minimised by accurately setting the delay and threshold level of the CFDs. Previous experiments with the
GLUE chamber achieved a resolution of 69 ns [54]. Therefore the present measurement, with a resolution of 4 ns FWHM, achieved significant improvements
in the timing resolution.

Appendix B
Experiments Conducted at the
MAX-lab

B.1

{ O |}G~f} ] N

December 1996

This experiment was a full-scale implementation of a pilot

study performed by Kuzin et al. [63]. A novel technique was used to detect the
de-excitation  -rays following g O   to measure the population of states in
g T N. Since the resolution was determined by the NaI  -ray detector, the proton

energy resolution was sacrificed by using a thick target to maximise the count
rate.
The aim of the study was to determine the importance of sort range correlations in the nucleus, such as the meson exchange currents (MEC) and nucleonnucleon interactions in the photon absorption mechanism. Models such as the
direct knock-out model (DKM) and the quasi-deuteron model (QDM) are inadequate at describing the photonuclear cross section. If the correct admixture for
the nuclear wavefunctions is used, and the popultation of positive and negative
parity states is measured, it is possible to determine the relative importance of
101

102

Appendix B. Experiments Conducted at the MAX-lab

the short range correlations.

B.2

j{ O |p}~c} ] O

June 1997 The same de-excitation technique was employed on g O, but this
time the  -rays were detected in coincidence with the emitted neutrons. Once
again the resolution was determined by the NaI  -ray detector, so the neutron
detectors were placed a close to the target as possible to maximise the count rate.
Like the   reaction, the relative population of states in the residual nucleus can help determine the role of the short range correlations. The results can
also help to find the most appropriate admixture of states in the wavefunction for
g O, based on a calculation of the reaction cross section.

B.3

j{ O |p}~c ] N

December 1997

In order to measure the photoneutron cross section, the same

experimental configuration was used as in June 1997. However, this time only
the neutrons were detected in a continuation of the pilot study done by Sims [84].
The photonuclear cross section was measured at a range of forward and backward angles, so as to measure the angular asymmetry of the reaction. From
the asymmetry and the cross section, it is possible to measure the isovector
quadrupole resonance (IVQR), one of a number of collective nuclear resonances
that can be observed using photons. The aim of the experiment was to confirm
the predictions of the continuing downward trend of the backward-to-forward
angular asymmetry, with decreasing photon energy.
During this run a wedge-shaped electron detector, desinged for the focal
plane tagger, was also tested. It was made to trigger under test conditions, although the resolution achieved did not compare favourable with the conventional

B.4. g  C  ! g B

103

electron detectors. This test was a continuation of the work done for an honours
project [85] in 1996 at the Australian Radiation Protection and Nuclear Safety
Agency (ARPANSA), Yallambie, Victoria.

d C |p}G~  B

B.4

June 1998

The motivation for this experiment stemmed from the success-

full results of Kuzin et al. that measured the residual states in gg B following
g C  JSgg B. Since the population of states in gg B were so clearly resolved,

the aim of this measurement was to see if the g B structure resembled the states
in gg B coupled to a valence neutron. The experiment was conducted downstream of the main experiment on g : N [64], which meant that the beam quality
was poorer than usual. Consequently, the signal-to-noise ratio in the timing spectra was very poor and the tagging peak was difficult to define. The final analysis
did not result in an energy spectrum of the residual states in g B as planned, because the statistics were too poor. However, anomously low proton-to-deuteron
and proton-to-triton ratios were observed in the particle identification plots. It
was suggested that this could be caused by the presence of the valence neutron,
breaking-up the otherwise strongly clustered g C core. Another possibility is
that the neutron caused an interference with an outgoing reaction particle, causing unusual final state interactions.

B.5

Li |p}G~ { He

June 1999

This run was used to collect the data presented in this thesis. Un-

fortunately, the run was ended prematurely due to technical difficulties with the
electron accelerator. So, only four days of Li data were taken instead of the
scheduled two weeks. Nevertheless, enough data was taken to analyse the reac-

104

Appendix B. Experiments Conducted at the MAX-lab

tion successfully and obtain good results.

The experiment was motivated by one of the first experiments which was
performed at MAX-lab on a Li target. A subsequent measurements of Li using
the GLUE chamber showed an interesting asymmetry in the first excited state in
He. This structure was also being investigated in radioactive beam experiments

at the time, while theoretical calculations were predicting a new halo excitation
in the same region of the He excitation spectrum. Given the ability to easily
separate proton events from the random background using the tagged photon

technique, it was decided to carefully measure the excited states in He following


the Li   reaction.

Appendix C
Papers
C.1

Conference Papers

The author presented papers at the following conferences. An abstract of the talk
is given in each case.

4th Workshop on EM Induced Two-Hadron Emission


Granada, Spain, May 29, 1999.
Protons, deuterons and tritons from g C and g  C: What do they tell us?
M. J. Boland, R. P. Rassool, M. N. Thompson, P. D. Harty, M. A. Garbutt
School of Physics, The University of Melbourne, Melbourne, Australia
J. Jury
Trent University, Ontario, Canada
J-O. Adler, K. Hanson, B. Schroder, M. Lundin, M. Karlsson, D. Nilsson
Department of Nuclear Physics, Lund University, Lund Sweden
T. Davison, S. A. Morrow, K. Foehl
Department of Physics, University of Edinburgh, Scotland
J. R. M. Annand, J. C. McGeorge
Department of Physics and Astronomy, University of Glasgow, Scotland
Abstract:
A recent intermediate-energy photonuclear experiment on g  C at =[ 5070
MeV indicates significantly more deuterons and tritons are emitted than from
g C. The reason for this is not clear, but may well be related to the differences
induced in the g C ground-state wavefunction with the addition of the extra neutron. An understanding of the experimental observations may be relevant in
understanding the photonuclear reaction mechanism in the intermediate-energy
105

106

Appendix C. Papers

region. The data presented here results from a measurement made at the MAXLab at Lund University last year. It was made in collaboration with groups from
Lund University, University of Glasgow, University of Edinburgh and Trent University.

American Physical Society, Annual Meeting


Long Beach, USA, April 28, 2000.
Searching for States in the Halo Nucleus He Using Intermediate Energy
Tagged Photons
M.J. Boland, M.A. Garbutt, R.P. Rassool, M.N. Thompson, A.J. Bennett
The University of Melbourne
J.W. Jury
Trent University
J.O. Adler, B. Schoder, D. Nilsson, K. Hansen, M. Lundin, M. Karlsson
Lund University
Abstract:
A recent photonuclear experiment we conducted at the MAX-Lab in Sweden
appears to support calculations that predict new states in He between the known
first and second excited states. The reaction Li   He was measured using
tagged photons in the energy range  = 50 70 MeV and high resolution proton
telescopes at angles [d"
p{
p
 "
and ]"
. Preliminary results show
some evidence exists to support the presence of previously unseen states in He
that are predicted by Ershov et. al. on the basis of a distorted wave impulse
approximation (DWIA) reaction theory calculation.
S.N. Ershov et al., Phys. Rev. C 56, 1483 (1997).

C.2

Journal Papers

The following paper was published by the American Physical Society in Physical
Review C, Vol. 64, 031601(R), 2001. It is in the Rapid Communication section
of the journal, and contains a brief summary of the experimental method, the
analysis techniques and the new state observed in He.

C.2. Journal Papers

107

RAPID COMMUNICATIONS

PHYSICAL REVIEW C, VOLUME 64, 031601 R

Excitations in the halo nucleus 6 He following the 7 Li

,p 6 He reaction

M. J. Boland, M. A. Garbutt, R. P. Rassool, M. N. Thompson, and A. J. Bennett


School of Physics, The University of Melbourne, Victoria 3010, Australia
J. W. Jury
Trent University, Peterborough, Ontario, Canada K9J 7B8

J.-O. Adler, B. Schroder, D. Nilsson, K. Hansen, M. Karlsson, and M. Lundin


Department of Physics, University of Lund, P.O. Box 118, S-221 00 Lund, Sweden
I. J. D. MacGregor

Department of Physics and Astronomy, University of Glasgow, Glasgow G12 8QQ, Scotland
Received 22 April 2001; published 26 July 2001

A broad excited state was observed in 6 He with energy E x 5 1 MeV and width
3 1 MeV, following
the reaction 7 Li( , p ) 6 He. The state is consistent with a number of broad resonances predicted by recent
cluster model calculations. The well-established reaction mechanism, combined with a simple and transparent
analysis procedure confers considerable validity to this observation.

DOI: 10.1103/PhysRevC.64.031601

PACS number s : 25.20.Lj, 24.30.Cz, 27.20. n

poor resolution, and use the same background removal


The physics of nuclei approaching the neutron drip line is
from
of interest
process as the ( Li, Be) reactions 11 . In contrast, tagged
as a means of further refining our understanding of
the nucleon-nucleon
photon
have a relatively simple and unampotential. Amongst these so-called
biguous measurements
background removal procedure that is proven and
attention. The
halo nuclei, He has received considerable

established level structure of He 1 has been questioned for well established 1215 and references therein .
some years in a number of theoretical calculations. These
This paper reports the presence of a broad resonance at an
considered extended neutron distributions by modeling He excitation
energy of 5 MeV in He that has been observed
Li( , p ) He photonuclear reaction. The meaas a He n n three-body cluster. A common feature of following the

these calculations is low-lying structure, above the well surement was made
in the energy range of E 50 70 MeV,
known 2 first excited state. The nature of this structure was using the MAX-lab tagged
photon facility 16 at Lund University. The protons and other charged particles were dethought to be a soft dipole resonance 2,3 , with two
tected with solid-state spectrometers, each consisting of a
initially
halo neutrons oscillating against the core. However, more


recent calculations refute this and postulate that it is caused thick
HP-Ge E detector and a thin Si E detector. These
by three-body dynamics 4 6 .
Experimental measurements on the He system have so
far been concentrated on charge exchange reactions of the
type Li( Li, Be) He 710 and Li(t , He) He 11 . All
7

these results have reported low-lying strength in the reaction


6

cross section at roughly the energies predicted by calculations, but none are able to determine the nature of the observed structure.
In each case the analysis of these experiments has in-

volved several controversial assumptions in the background


removal
process.
In particular, the nonresonant background
in the ( Li, Be) reaction was calculated but not measured.
7

This process must include degrees of freedom due to the


excited states of both the projectile and the ejectile. In one
case 9 , nonresonant background contributions to the cross
section were not included at all.
Background subtraction is only one of the complications
involved with heavy-ion reactions. Another difficulty is that
many possible combinations of angular-momentum transfer
exist between projectile and target. One of the simplest
charge exchange reactions, namely (n , p ) , does not suffer the
same problem. However, the poor resolution of these (n , p )
experiments makes it difficult to see even the commonly
resolved 2 state. Reactions of the type (t , 3 He) also suffer

0556-2813/2001/64 3 /031601 3 /$20.00

FIG. 1. The time correlation spectrum between protons and

tagged
for 60 . The 6 ns wide prompt peak shaded is

clearly photons
visible on top of a random background labeled .

64 031601-1

2001 The American Physical Society

108

Appendix C. Papers

RAPID COMMUNICATIONS

M. J. BOLAND et al.
n

PHYSICAL REVIEW C 64 031601 R


r

2. Proton missing-energy spectrum at 60 showing i


60 following the
FIG. 3. Proton missing-energy spectrum at
the FIG.
) He with the background contributions

random background open dots with a polynomial fit dotted


reaction
,
p
Li(
line , ii the calculated ( , pn) background solid line , and iii the tracted. The He excitation energy scale is drawn for reference.subprompt protons filled dots .

pn) missing-energy
is located at
see Fig. 2 and asdistribution
were placed at angles of 30 ,60,90,120, and 150 to Eof the 29( ,MeV
such cannot account for
the photon beam, similar to the configuration described in all the strength observed
between E
13 20 MeV. The

pn background was normalized in a consistent manner for all


17 . A 1 mm thick target of 99.9% pure Li was placed at

angles, then subtracted such that the net missing-energy


60 to the photon beam.
spectrum
was positive at all energies. The resulting missingwere selected from other charged particle events
by Protons
energy spectrum
of protons emitted at
60 is shown in
use of a particle-identification plot of the energy lost in
the full-energy detector, versus that lost in the E detector. Fig. 3.
leading to the ground state and the first excited
Protons correlated with tagged photons were located in a

stateProtons
at E 1.8 MeV can be clearly seen. Evidence for the
narrow prompt timing peak, shown shaded in Fig. 1, sitting

on
known second excited state near E 14 MeV can be distina timing spectrum of random events. Missing-energy
spectra
guished at the onset of the high missing-energy region of the
were produced from a cut on the prompt peak at each

angle filled
spectrum. Significantly, the evidence for a broad state can be
2 . The missing energy is defined as
E E Tdots inT Fig.
, where T is the kinetic energy of the seen in the region between E 3 10 MeV. A fit of three
Gaussians to the data in Fig. 3 gives a width of
3 1
He nucleus, and T is the kinetic energy of the emitted
proton. The excitation energy, shown in Fig. 3, is related to MeV
and a centroid energy of E 5 1 MeV to the new
E by E E Q, where Q is the proton separation structure,
on the assumption that it is a single resonance.
energy, and for the reaction Li( , p) He, Q 10.0 MeV.
The present experiment, like those using charge exchange
reactions, is unable to define the exact nature of the observed
The contribution of random proton events in the prompt reresonance. The strongest candidates seem to be a 1 soft
gion, was measured by making a cut on the random back
mode and a second 2 state, predicted by Suzuki 3
ground region labeled in Fig. 1 . The resulting featureless
background
dipole
and others 1922 . A calculation of the E1 breakup of He
spectrum open circles in Fig. 2 was normalized
and fitted, before being subtracted from the spectrum of the 6 shows an enhancement to the 1 continuum at an energy
consistent with the measurement presented here. It is posprompt region.

sible that the strength we observe in the Li( , p) He cross


The contribution due to the ( , pn ) reaction threshold

at 5 MeV is evidence of the 1 dipole and the posiE


11.9 MeV also needed to be considered. The momentum distribution of this background channel was calcu- section
tive parity states, both of which were predicted by Danilin
et al. 5 . A complete analysis of our data, including the anusing a Monte Carlo model of direct two-nucleon emislated
gular distribution, may clarify the nature of the structure and
sion 18 , which included all the experimental parameters,

and
thereby validate some of the model assumptions.
covered the full phase space of the experiment. The peak
T

miss

miss

miss

<

>

miss

miss

"

"

&

'

miss

031601-2

C.2. Journal Papers

109

RAPID COMMUNICATIONS

EXCITATIONS IN THE HALO NUCLEUS 6 He . . .

15 A. Kuzin et al., Phys. Rev. C 58, 2167 1998 .


16 J.-O. Adler, B.-E. Andersson, K. I. Blomqvist, K. G. Fissum,
K. Hansen, L. Isaksson, B. Nilsson, D. Nilsson, H. Ruijter, A.
Sandell, B. Schroder, and D. A. Sims, Nucl. Instrum. Methods
Phys. Res. A 388, 17 1997 .
17 J. F. Dias, D. Ryckbosch, R. V. de Vyver, C. V. den Abeele, G.
D. Meyer, L. V. Hoorebeke, J.-O. Adler, K. I. Blomqvist, D.
Nilsson, H. Ruijter, and B. Schroder, Nucl. Phys. A587, 434
1995 .
18 J. C. McGeorge et al., Phys. Rev. C 51, 1967 1995 .
19 B. V. Danilin, T. Rogde, S. N. Ershov, H. Heiberg-Andersen, J.
S. Vaagen, I. J. Thompson, and M. V. Zhukov, Phys. Rev. C 55,
R577 1997 .
20 S. N. Ershov, T. Rogde, B. V. Danilin, J. S. Vaagen, I. J. Thompson, and F. A. Gareev, Phys. Rev. C 56, 1483 1997 .
21 M. V. Zhukov, D. V. Fedorov, and B. V. Danilin, Nucl. Phys.
A539, 177 1992 .
22 S. N. Ershov, B. V. Danilin, and J. S. Vaagen, Phys. Rev. C 62,
041001 R 2000 .
o

1 F. Ajzenberg-Selove, Nucl. Phys. A490, 1 1988 .


P. G. Hansen and B. Jonson, Europhys. Lett. 4, 409 1987 .
Y. Suzuki, Nucl. Phys. A528, 395 1991 .
4 A. Csoto , Phys. Rev. C 49, 2244 1994 .
5 B. V. Danilin, I. J. Thompson, J. S. Vaagen, and M. V. Zhukov,
Nucl. Phys. A632, 383 1998 .
6 I. J. Thompson, B. V. Danilin, V. D. Efros, J. S. Vaagen, J. M.
Bang, and M. V. Zhukov, Phys. Rev. C 61, 024318 2000 .
7 J. Janecke et al., Phys. Rev. C 54, 1070 1996 .
8 T. Annakkage et al., Nucl. Phys. A648, 3 1999 .
9 S. Nakayama et al., Phys. Rev. Lett. 85, 262 2000 .
10 S. B. Sakuta, A. A. Ogloblin, O. Y. Osadchy, Y. A. Gluukhov,
S. N. Ershov, F. A. Gareev, and J. S. Vaagen, Europhys. Lett.
22, 511 1993 .
11 T. Nakamura et al., Phys. Lett. B 493, 209 2000 .
12 R. O. Owens, Nucl. Instrum. Methods Phys. Res. A 288, 574
1990 .
13 I. J. D. MacGregor et al., Nucl. Phys. A533, 269 1991 .
14 L. V. Hoorebeke, Nucl. Instrum. Methods Phys. Res. A 321,
230 1992 .

23

PHYSICAL REVIEW C 64 031601 R

031601-3

110

Bibliography
[1] Y. Suzuki, Nuclear Physics A528, 395 (1991).
[2] I. Tanihata et al., Physical Review Letters 55, 2676 (1985).
[3] V. B. Shostak et al., Physical Review C: Nuclear Physics 63, 017602
(2000).
[4] M. V. Zhukov, L. V. Chulkov, B. V. Danilin, and A. A. Korsheninnikov,
Nuclear Physics A533, 428 (1991).
[5] M. V. Zhukov et al., Physics Reports 231, 151 (1993).
[6] B. V. Danilin, M. V. Zhukov, J. S. Vaagen, and J. M. Bang, Physics Letters
B 302, 129 (1993).
[7] S. Aoyama, S. Mukai, K. Kato, and K. Ikeda, Progress of Theoretical
Physics 93, 99 (1995).
[8] B. V. Danilin et al., Physical Review C: Nuclear Physics 55, R577 (1997).
[9] S. N. Ershov et al., Physical Review C: Nuclear Physics 56, 1483 (1997).
[10] S. N. Ershov, B. V. Danilin, T. Rogde, and J. S. Vaagen, Physical Review
Letters 82, 908 (1999).
[11] Y. T. Oganessian, V. I. Zagrebaev, and J. S. Vaagen, Physical Review C:
Nuclear Physics 60, 044605 (1999).
111

112

Bibliography

[12] Y. T. Oganessian, V. I. Zagrebaev, and J. S. Vaagen, Physical Review Letters 82, 4996 (1999).
[13] S. N. Ershov, B. V. Danilin, and J. S. Vaagen, Physical Review C: Nuclear
Physics 62, 041001 (2000).
[14] T. Kobayashi, Nuclear Physics A538, 343c (1992).
[15] S. B. Sakuta et al., Europhysics Letters 22, 511 (1993).
[16] J. Janecke et al., Physical Review C: Nuclear Physics 54, 1070 (1996).
[17] T. Aumann et al., Physical Review C: Nuclear Physics 59, 1252 (1999).
[18] N. K. Timofeyuk, Nuclear Physics A652, 132 (1999).
[19] S. Nakayama et al., Physical Review Letters 85, 262 (2000).
[20] T. Nakamura et al., Physics Letters B493, 209 (2000).
[21] K. Riisager, Reviews of Modern Physics 66, 1105 (1994).
[22] I. Tanihata, Journal of Physics General: Nuclear Particle Physics 22, 157
(1996).
[23] C. Bennhold, in MAX-lab Workshop (PUBLISHER, Lund University,
1997).
[24] J. S. Vaagen et al., Physica Scripta T88, 209 (2000).
[25] W. Nazarewicz, in Fronties in Nuclear Physics, edited by S. Kuyucak
(World Scientific, Singapore, 1999), p. 17.
[26] R. N. Boyd and I. Tanihata, Physics Today 45, 44 (1992).
[27] T. Kobayashi et al., Physics Letters B 232, 51 (1989).

Bibliography

[28] P. G. Hansen and B. Jonson, Europhysics Letters 4, 409 (1987).


[29] K. Ikeda, INS Report No. JHP-7 (1988), in Japanese.
[30] M. V. Zhukov et al., Journal of Physics General: Nuclear Particle Physics
20, 201 (1994).
[31] S. Karataglidis et al., Physical Review Letters 79, 1447 (1997).
[32] I. J. Thompson et al., Physical Review C: Nuclear Physics 61, 024318
(2000).
[33] Z. Zhen and J. Macek, Physical Review A 38, 1193 (1988).
[34] M. V. Zhukov, D. V. Fedorov, and B. V. Danilin, Nuclear Physics A539,
177 (1992).
[35] V. Efros et al., Z. Phys. A 355, 101 (1996).
[36] Y. Suzuki and K. Ikeda, Physical Review C: Nuclear Physics 38, 410
(1988).
[37] B. V. Danilin and M. V. Zhukov, Physics of Atomic Nuclei 56, 410 (1993).
[38] B. V. Danilin, I. J. Thompson, J. S. Vaagen, and M. V. Zhukov, Nuclear
Physics A632, 383 (1998).
[39] K. Heyde, Basic Ideas and Concepts in Nuclear Physics, Graduate Student
Series in Physics, 2nd ed. (Institute of Physics Publishing, Philadelphia,
1999).
[40] T. Aumann et al., Nuclear Physics A687, 103c (2001).
[41] F. Ajzenberg-Selove, Nuclear Physics A490, 1 (1988).

113

114

Bibliography

[42] S. Karataglidis, B. A. Brown, K. Amos, and P. J. Dortmans, Physical Review C: Nuclear Physics 55, 2826 (1997).
[43] G. Audi and A. H. Wapstra, Nuclear Physics A595, 409 (1995), see also
http://nucleardata.nuclear.lu.se/database/masses/.
[44] S. Aoyama, S. Mukai, K. Kato, and K. Ikeda, Progress of Theoretical
Physics 94, 343 (1995).
[45] A. Csoto, Physical Review C: Nuclear Physics 49, 2244 (1994).
[46] K. Amos, Private Communication, 2001.
[47] A. Csoto, Physical Review C: Nuclear Physics 48, 165 (1993).
[48] K. Rusek, K. W. Kemper, and R. Wolski, Physical Review C: Nuclear
Physics 64, (2001).
[49] S. Nakayama et al., Nuclear Instruments and Methods in Physics Research
A404, 34 (1998).
[50] R. O. Owens, Nuclear Instruments and Methods in Physics Research A288,
574 (1990).
[51] L. van Hoorebeke, Nuclear Instruments and Methods in Physics Research
A321, 230 (1992).
[52] L. V. Hoorebeke et al., Nuclear Instruments and Methods in Physics Research A326, 608 (1993).
[53] J. F. Dias et al., Nuclear Physics A587, 434 (1995).
[54] H. Ruijter, Ph.D. thesis, University of Lund, 1995.
[55] A. Kuzin et al., Physical Review C: Nuclear Physics 58, 2167 (1998).

Bibliography

[56] L. I. Schiff, Physical Review 83, 252 (1951).


[57] M. Eriksson, Nuclear Instruments and Methods in Physics Research 196,
331 (1982).
[58] L. J. Lindgren and M. Eriksson, Nuclear Instruments and Methods in
Physics Research 214, 179 (1983).
[59] L. J. Lindgren and M. Eriksson, Nuclear Instruments and Methods in
Physics Research A294, 10 (1990).
[60] J.-O. Adler et al., Nuclear Instruments and Methods in Physics Research
A388, 17 (1997).
[61] H. Ruijter et al., Physical Review C54, 3076 (1996).
[62] R. D. Evans, The Atomic Nucleus, International Series in Pure and Applied
Physics (McGraw-Hill, New York, 1955).
[63] A. Kuzin, Ph.D. thesis, School of Physics, University of Melbourne, Australia, 1997.
[64] S. A. Morrow, Ph.D. thesis, Department of Physics and Astronomy, University of Edinburgh, United Kingdom, 2000.
[65] R. Brun and F. Rademakers, Nuclear Instruments and Methods in Physics
Research A389, 81 (1997), see also http://www.root.ch/.
[66] D. Nilsson, Private Communication, 1999.
[67] PAW, Physics Analysis Workstation, CERN Program Library Q121, 1993.
[68] H. Ruijter, lund University (unpublished).

115

116

Bibliography

[69] ICRU,
for

ICRU

Report

Protons

and

49,
Alpha

Stopping

Powers

Particles,

1993,

and

Ranges

see

also

http://physics.nist.gov/PhysRefData/Star/Text/PSTAR.html.
[70] A. M. Baldin, V. I. Goldanskii, and I. L. Rosenthal, Kinematics of nuclear
reactions (Pergamon Press, New York, 1961).
[71] P. C. Stein, A. C. Odian, A. Wattenberg, and R. Weinstein, Physical Review
119, 348 (1960).
[72] B. M. K. Nefkens, Physics Letters 10, 55 (1963).
[73] I. J. D. MacGregor, Private Communication, 2000.
[74] A. E. Thorlacius and H. W. Fearing, Physical Review C: Nuclear Physics
33, 1830 (1986).
[75] J. C. McGeorge et al., Physical Review C: Nuclear Physics 51, 1967 (1995).
[76] F. Ajzenberg-Selove, Nuclear Physics A506, 1 (1990).
[77] S. Karataglidis, Private Communication, 2001.
[78] J. Ryckebusch et al., Physical Review C: Nuclear Physics 46, R829 (1992).
[79] J. Ryckebusch, Private Communication, 2001.
[80] M. J. Boland et al., Physical Review C: Nuclear Physics 64, 031601(R)
(2001).
[81] J. S. Vaagen, Nuclear Physics A690, 302c (2001).
[82] J. W. Jury et al., Physical Review C: Nuclear Physics 19, 1684 (1979).
[83] T. Myo, K. Kato, S. Aoyama, and K. Ikeda, Physical Review C: Nuclear

Physics 63, 054313 (2001).

Bibliography

[84] D. A. Sims, Ph.D. thesis, School of Physics, University of Melbourne, Australia, 1995.
[85] M. J. Boland, B.Sc. (Hons) Report, School of Physics, University of Melbourne, Australia, 1996.

117

118

S-ar putea să vă placă și