Download as pdf or txt
Download as pdf or txt
You are on page 1of 163

Such title, much dissertation

Jesper Halkjær Jensen

PhD Dissertation

Department of Physics and Astronomy


Aarhus University
Denmark
Jesper Halkjær Jensen
Department of Physics and Astronomy
Aarhus University
Ny Munkegade 120, Building 1520
8000 Aarhus C
Denmark

c Jesper Halkjær Jensen 2019



Revision (None) dated (None).
Such title, much dissertation

A Dissertation Presented to the


Faculty of Science and Technology of Aarhus University
in Partial Fulfillment of the Requirements
for the PhD Degree

by
Jesper Halkjær Jensen
October 2018
v

Some clever quote...

Clever guy,
“Clever guys book”, unpublished [Citation].
Contents

Contents i

Abstract v

Resumé vii

Acknowledgments ix

List of Publications xi

1 Preface 1

2 Reaction theory 3
2.1 Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2 Kinematic compression . . . . . . . . . . . . . . . . . . . . . 5
2.3 Scattering theory . . . . . . . . . . . . . . . . . . . . . . . . 10

3 Experimental technique 13
3.1 ISOLDE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2 Detectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3 ∆E − E telescope . . . . . . . . . . . . . . . . . . . . . . . 18

4 Beam diagnostics 21

i
ii CONTENTS

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.2 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.3 Setup and simulation . . . . . . . . . . . . . . . . . . . . . . 24
4.4 Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

5 AUSAlib 43
5.1 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.2 Energy loss . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.3 Analysis pipeline . . . . . . . . . . . . . . . . . . . . . . . . 47

6 Data acqusition 53
6.1 Triggers and acqusition control . . . . . . . . . . . . . . . . 54
6.2 Analogue chain . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.3 Readout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.4 Multicrate . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.5 Dead time ratio . . . . . . . . . . . . . . . . . . . . . . . . . 58
6.6 Pulsed beam . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.7 Back end . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
6.8 DaqC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.9 RunDB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6.10 Real time monitoring . . . . . . . . . . . . . . . . . . . . . . 66

7 Acceptance 69
7.1 MC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
7.2 Rutherford estimation . . . . . . . . . . . . . . . . . . . . . 73

8 Experimental setups 77

9 IS367 79
9.1 Data analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 80
9.2 Intensity estimation . . . . . . . . . . . . . . . . . . . . . . 85

10 IS561A 91
10.1 Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
CONTENTS iii

10.2 Calibration and beam properties . . . . . . . . . . . . . . . 93


10.3 Data analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 96

11 IS561C 101
11.1 Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
11.2 Main target . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
11.3 Secondary target . . . . . . . . . . . . . . . . . . . . . . . . 110

12 Results 115
12.1 Elastic scattering . . . . . . . . . . . . . . . . . . . . . . . . 116
12.2 10 Li transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
12.3 8 Li transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

List of Figures 137

Abbreviations 143
Abstract

v
Resumé

vii
Acknowledgments

Jesper Halkjær Jensen,


Aarhus, July 10, 2019.

ix
List of Publications
The following manuscripts form part of this thesis

1. M. Munch, J. H. Jensen, B. Löher, H. Törnqvist and H. T. Johansson


VME readout at and below the conversion time limit. Submitted to
IEEE Transactions on Nuclear Science. arXiv: 1810.03574

Other papers not included in this thesis

1. J. J. W. H. Sørensen, M. K. Pedersen, M. Munch, P. Haikka, J. H.


Jensen, T. Planke, M. G. Andreasen, M. Gajdacz, K. Mølmer, A.
Lieberoth and J. F. Sherson Exploring the quantum speed limit with
computer games. Nature, 532(7598) (2016), p. 210–213. doi: 10.1038/na-
ture17620 arXiv: 1506.09091
2. O. S. Kirsebom, O. Tengblad, R. Lica, M. Munch, K. Riisager, H. O.
U. Fynbo, M. J. G. Borge, M. Madurga, I. Marroquin, A. N. Andreyev,
T. A. Berry, E. R. Christensen, P. D. Fernández, D. T. Doherty, P.
Van Duppen, L. M. Fraile, M. C. Gallardo, P. T. Greenlees, L. J.
Harkness-Brennan, N. Hubbard, M. Huyse, J. H. Jensen, H. Jo-
hansson, B. Jonson, D. S. Judson, J. Konki, I. Lazarus, M. V. Lund,
N. Marginean, R. Marginean, A. Perea, C. Mihai, A. Negret, R. D.
Page, V. Pucknell, P. Rahkila, O. Sorlin, C. Sotty, J. A. Swartz, H. B.
Sørensen, H. Törnqvist, V. Vedia, N. Warr and H. De Witte First Ac-
curate Normalization of the β-delayed α Decay of 16 N and Implications

xi
xii CONTENTS

for the 12 C(α, γ)16 O Astrophysical Reaction Rate. Physical Review


Letters 121, 142701 (2018). doi: 10.1103/PhysRevLett.121.142701
1

Preface

A very introduction-kind of text

1
2

Reaction theory

2.1 Kinematics
The reactions we study are usually two body reactions. When we populate
unbounds states such a description will not suffice due to break up into
multiple constituents, but depending on the break up mechanics, we may still
be able to learn something from a simple two body picture. In this chapter
we will develop the equations we need to describe the reactions as well as
discuss the kinematic considerations required to design and undertake an
experiment in inverse kinematics, that is when the projectile is significantly
heavier than the target.
Since our experiments are carried out a relatively low energies, we will
treat the kinematics in a classical fashion.
Consider a beam particle, B, impinging on a target particle t, with some
kinetic energy TB in the laboratory frame. We assue the target is at rest
and thus no kinetic energy Tt = 0.
They react in some way and produce a heavy projectile-like recoil R and
a light target-like ejectile e (if we were in normal kinematics, it is customary
to denote the light particle as the projectile-like instead)

B + t → R + e. (2.1)
We assume that both the projectile and the target are initially in the
ground state. Conservation of energy gives

3
4 CHAPTER 2. REACTION THEORY


TB = TR + Te + ER + Ee∗ − Q, (2.2)
where Ee∗ is the excitation energy of the ejectile, ER
∗ is the excitation energy

of the recoil and


Q = (mB + mt − mR − me )c2 . (2.3)
The available energy to initiate reactions is
mt
Ech = TB (2.4)
mt + mB
Linear momentum conservation gives

p~B = p~R + p~e , (2.5)



where p~i = 2Ti mi n̂i and n̂i is the unit vector in the direction of motion.
Given the initial conditions of B and t we can rearrange 2.2 and 2.5 to
calculate the relationship between the energy of the ejectile angle θe and its
kinetic energy Te , called the kinematic curve

mB me TB cos θe ± mB me TB cos2 θe + (mR + me )[mR Q + (mR − mB )TB ]
p
Te =
1/2
.
mR + me
(2.6)
We can in in principle denote the particles as we like, and use equation
2.9 to calculate the kinematic curve for both the projectile and recoil.
If the beam energy is below a certain threshold there are no solutions,
and the reactions can not occur due to energy conservation. If we are above
the threshold we can have two situations
Single value solutions In this case each θ corresponds to single energy.
This occurs when the content of the square root is positive such that there
is a solution, and the second square root is larger than the first, such that
the negative solutions (the − sign between the two terms) is not a valid
solution.
Double value solutions This occurs when the content of the square
root is positive but the second square root is smaller than the first, such
that the negative solution (the − sign between the two terms) is a valid
2.2. KINEMATIC COMPRESSION 5

solution. In this case the a given angle corresponds to two energies. The
maximal angle this can occur is when the two terms inside the square root
just cancels and the two "branches" of solutions meets. Since the low-energy
solutions corresponds to large-angle ejectiles, we say that the kinematic
curve bends back.
An example of both single valued and double valued kinematic curves
can bee seen in Figure 2.1 (A) for elastic scattering of 9 Li (d,d)9 Li.
In the case of two body reactions, we can instead limit our measurement
to the ejectile and use energy and momentum conservation to infer the
kinematics of the recoil, which is usually the particle of interest. If we
assume that the ejectile stays in the ground state (which is resonable in the
case of deuterons and tritons since they would be unbound otherwise) we
can calculate the excitation energy of the recoiling nucleus with 2.2 and 2.5
if we know the masses of all four constituents


ER = TB − TR − Te + Q (2.7)
(~
pB − p~e )2
= TB − Te − + Q, (2.8)
2mR

where we know TB , p~B and Q and we measure Te and p~e .

2.2 Kinematic compression


If the mass of the beam is heavier that the target, the kinematic changes
accordingly. There is as such no new physics involved, but our intuition
may need a slight ajustment. A heavy beam leads to what we refer to as
kinematic compression. There is a nice discussion in [CATFORD] to gain
a better intuition about how to think about the transformations, but we
will limit us to discuss the consequences by calculating the kinematic curves
for different reactions. To guide this discussion we will use the concrete
reactions that we also study, namely 9 Li (d,t)8 Li, 9 Li (d,d)9 Li and 9 Li
(d,p)10 Li.
6 CHAPTER 2. REACTION THEORY

Let us consider elastic scattering 9 Li (d,d)9 Li. In this case the projectile
and the heavy recoil have the same mass mB = mR and Q = 0. The
kinematic curve reduces to
√ √
1/2 mB me TB cos θ ± mB me TB cos2 θ
Tb = , (2.9)
mR + me
which vanishes when θ ≥ 90◦ . We can see the resulting kinematic curves for
the ejectile and the recoil in Figure 2.1 (A).
All the deuterons are scattered in the first 90◦ in LAB and we have
effectively compressed the full angular range in CM to only 90◦ in LAB.
For the heavy recoil we see the back bend in the kinematic curve. The
heavy recoil is focused a small LAB angles, since it eventually will bend
back toward 0◦ . This has two important implications if we want to measure
both the ejectile and the recoil. Firstly, we need to put detectors close
to the beam axis cover a resonable area. Secondly, our angular resolution
suffers close to the turning points, since we have a lot of CM-angles that
are compressed into a small angular range of the LAB frame.
When we transfer a neutron we change the Q-value for the reaction.
Depending on whether we change transfer it to the target or the projectile,
we get two different situations.
In Figure 2.1 (B) we see the kinematic curves for 9 Li (d,p)10 Li. The
ejectile is single valued all the way to 180◦ . The energy at backward angles
are, however, much lower. We can increase the proton energy by increasing
the beam energy. An experimental setup will have some threshold that
determines how low energy protons we can measure. It is thus necessary to
increase the beam energy enough to be able to actually measure the protons
in the backward directions if desired. This asymmetric energy distribution
is important to consider when designing the setup.
The recoil is scattered even closer to the beam axis, and coincidence
measurements requires coverage even closer to the beam axis than in the
case of elastic scattering. Moreover will the angular resolution be even
worse, especially at the region around the maximally scattered recoils.
In Figure 2.1 (B) we see the kinematic curves for 9 Li (d,t)8 Li. In this
case we even see the back bend for the light ejectile. We notice that the
2.2. KINEMATIC COMPRESSION 7

25 30
Ejectile Ejectile Ejectile
25 Recoil Recoil Recoil
25
20
20
20
15
15
15
10
10 10

5 5
5

0 0 0
0 25 50 75 0 50 100 150 0 50 100

Figure 2.1: Kinematic curves for ejectile and recoil of the reactions 9 Li
(d,t)8 Li, 9 Li (d,d)9 Li and 9 Li (d,p)10 Li at 2.68 MeV/A.

recoil scatters further from the beam axis, making this situation better for
coincidens measurements.

Detector coverage
When designing an experiment it is important to consider what region of
4π that should be covered by the detectors. The physics of the reaction will
guide what angles are the most interesting. Usually, however, theoretical
calculations are performed in CM and thus we need to convert this to
LAB to design the detector setup, and this conversion is governed by the
kinematic of the reaction.
In the previous section we discussed how the angular resolution is
compressed in the forward LAB angles when we work in inverse kinematics.
This means that the angular resolution will be different across the angular
range if similar detectors are used. Take for instance the case of 9 Li (d,p)10 Li.
We can calculate the angular resolution as dθCM /dθLAB . The result at
2.68 MeV/A is shown in Figure 2.2. We can see that we need a much better
angular resolution to at forward angles compared to backward angles if we
want to have compareable CM resolution in both regions. The required
8 CHAPTER 2. REACTION THEORY

resolution, however, mostly depends on how fast the angular distributions


are expected to vary. Nontheless, is it important to keep the compression of
the resolution into account.
This dependence on the angles is even more significant in terms of the
solid angle transformation. If most of our CM angles are focused in the
forward angles, we need to cover a large region in backward LAB angles to
cover a decent region in CM. We can quantify this with the Jacobian that
transform angular cross sections from CM to LAB

dσ dσ
J(θLAB ) = J(θCM ) J(θCM ), (2.10)
dΩ LAB dΩ CM

where
(1 + γ 2 + 2γ cos θCM )3/2
J= , (2.11)
1 + γ cos θCM

and
s
mB me Ech
γ= . (2.12)
mt mR Ech + Q

The Jacobian transformation is shown in Figure 2.2. We see a strong


enhancement in the forward angles and a equally strong supression in the
backward LAB angles. That means we need much more coverage to get a
resonable integrated cross section.
In a real experiment we need a certain amount of counts to have enough
statistics to extract the physics. We should thus also consider the sin θ
effect. A differential solid angle is given as dΩ = sin θdθdφ. When we get
go away from θ = 90◦ we get a further suppression due to the sin θ factor.
If we consider J(θ) sin θ instead we get the result in panel (D).
When designing an experiment, it is important to keep in mind that the
number of measured particles is heavily suppressed in backward directions.
If we compare the maximum close to 50◦ where we have an factor of ≈ 2 with
150◦ we are suppresed with almost approximately two orders of magnitude.
2.2. KINEMATIC COMPRESSION 9

A B
2.0

150
1.5

dθLAB
100
dθCM
θCM

1.0

50 0.5

0 0.0
0 50 100 150 0 50 100 150
θLAB θLAB
C D
4 2.0
θLAB
θCM
3 1.5
J sin θ

2 1.0
J

1 0.5

0 0.0
0 50 100 150 0 50 100 150
θ θLAB

Figure 2.2: Kinematic effect on angular resolution and solid angle coverage.
All the plots corresponds to a proton from 9 Li (d,p)10 Li at 2.68 MeV/A.
A: The relation between the scattering angle in LAB and CM. B: The
differential, which gives a measure of the angular resolution in LAB needed
to recover a given angular resolution in CM. C: The Jacobian that transforms
from CM to LAB plottet. Since there is a 1-1 correspondance between
θCM and θLAB it is plotted as a function of both. D: The Jacobian times
sin θ plottet as a function of θLAB .
10 CHAPTER 2. REACTION THEORY

2.3 Scattering theory


To extract physics knowlegde from our experimental results, we need to
compare with theoretical models. These models should both be able to
describe the reaction mechanism as well as the internal structure of the
reaction constituents. It is typically the latter we are interested in and it is
thus important that we can include both in a consistent way.
The problem of scattering is general well understood. Consider our
previous situation with two particles scattering

p+t→R+e (2.13)
The scattering is described by the total wavefunction of the system and
the potential that describes the interaction between the projectile and the
target via the time-independent Schrödinger equation

[T̂R~ + Hp (xp ) + Ht (xt ) + Vα (R,


~ xp , xt ) − E]Ψ(R,
~ xp , xt ) = 0 (2.14)

where T̂R~ is the kinetic operator, Hp (xp ) and Ht (xt ) are the internal Hamil-
tonians of the projectile and target, Vα (R, ~ xp , xt ) is the interaction and
Ψ(R,~ xp , xt ) is the total wavefunction. The internal coordinates of the pro-
jectile and the target are denoted xp and xt . The total energy of the system
is the sum of kinetic and internal energies.
To calculate the cross sections of a given reaction, we need to solve ??
to obtain Ψ(R, ~ xp , xt ).
Since we measure the scattered particles far away from the reaction
site, we consider the asymptotically behavior of the solution. The incident
projectile can be represented as a plane wave approaching the target. During
the collision the wave will be distorted, due to the interaction described by
Vα (R,
~ xp , xt ). The outgoing wavefunction is assumed to be a radial wave
and we can write the total asymptotic wave function as
~ ~
~ R~ eiK·R
Ψ (+)
→e iK·
+ f (θ) , (2.15)
R
2.3. SCATTERING THEORY 11

where (+) refers to the fact that we have outgoing boundary conditions.
f (θ) is the socalled scattering amplitude and we can calculate the scattering
cross sections from this


= |f (θ)|2 . (2.16)
dΩ

Optical potential
The simplest approach to determine elastic cross sections is the Optical
Model (OM). In this model only the elasic channel is calculated, and all
other channels are treated as loss of flux from the the elastic channel.
In general the scattering problem is well understood.
3

Experimental technique

In this chapter we will discuss some more technical topics that were common
for all three experiments. There has been a varity of upgrades to the various
techniques and equipment between the experiments, but the main concepts
and objectives has remained the same.
Performing experiments with highly unstable isotopes are no trivial task,
due to their very short lifetime. In the case of 9 Li we have τ = 172ms to
perform the experiment before half the nuclei has decayed. This can not be
done in ordinary kinematics since we can not create a target with such short
lived isotopes. Instead we can inverse the kinematics and use a radioactive
ion beam (RIB) of 9 Li impinging on hydrogen isotopes. Deuterons can easily
be implanted in a plastic foil, taking the place of protons. A small percentage
of the protons will remain, giving rise to a contamination that must be
assessed in the analysis. Tritons can, on the other hand, be implemented
on a titanium backing [REF]. Several RIB-facilities across the world exists.
The experiments of this thesis were all performed at the ISOLDE facility
at CERN, in Switzerland. Between IS367 and IS561A there were major
upgrade of the facility, the HIE-ISOLDE upgrade, which meant we could
reach higher beam energies.
To detect the particles we used standard silicon detectors. The signals
were amplified with a series of analogue modules and finally digitized with
VME-compatible ADC’s and TDC’s. The readout of these methods were
optimized between the experiments. As I did not take part in the IS367

13
14 CHAPTER 3. EXPERIMENTAL TECHNIQUE

experiment I will not discuss the data acqusition system from this experiment,
however, it was by large the same modules and logic that was used in the
two latter experiments.

3.1 ISOLDE
The radioactive beams at ISOLDE are produced using the isotope seperation
on-line (ISOL) method[REF]1 . A pulse of ≈ 1.4 GeV protons from the PS
Booster impinges on a primary target of a heavy element, for instance UCx
or Tantulum. This produces a wide array of different isotopes through
fragmentation, spallation and fission. The target is heated which makes the
products diffuse into the ion source. Inside the ion source they are laser
ionized and subsequently accelerated to ≈ 60 keV through a high voltage.
They are sent to one of the two mass seperators GPS or HRS. The GPS
(General purpose seperator) have a single bending magnet and the HRS
(High resolution seperator) have two bending magnets. The mass seperators
seperate the particles according to their masses and the isotope of interest
can be sent to one of the experiments.
In order to do reactions we need a higher beam energy than 60 keV.
The first installation to perform post-acceleration of the beam was REX-
ISOLDE and it was intalled between 2001 and 2005. Firstly the beams reach
a trap where the beam particles are trapped, cooled down and bunched in a
Penning trap. A buffer gas is used in the cooling process. These bunches
are continued into an electron beam ion source (EBIS). Inside the EBIS the
beam is highy ionized before it is sent through another seperator, which
seperates according to A/q to get rid of the buffer gas residues. After the
seperator the beam is sent into a linear accelerator which can accelerate
them up to ≈ 3 MeV/A.
The REX-ISOLDE was used for post acceleration in IS367. In the decade
following this experiment a series of superconducting cavity moduels was
installed to increase the postacceleration capabilities. This was part of the
HIE-ISOLDE upgrade, which aims to provide High Intensity and Energy.
1
FiXme Note: Find reference!!!
3.1. ISOLDE 15

HRS

1.4 GeV
GPS protons

60keV

HIE-ISOLDE energy upgrade

Figure 3.1: Schematic overview of ISOLDE. 2

The superconducting cavities were parts of the energy upgrade and was
completed with the seventh module in (WHEN?)3 . In 2016 only three of
the modules were installed, and 6.72 MeV/A was the highest energy that
was reached until then.

TEBIS and TPROTON


The protons from the PS Booster are delivered across the CERN facility
with a period of 1.2 s. Since multiple facilities are running simultaneously
we only get a fraction of these protons, and in a very non-continuous fashion.
We refer to this as the super cycle. In our experimental setup, however,
we record when we get the proton pulses at the primary target. We can
thus calculate the time since proton impact and a measurement in our
3
FiXme Fatal: MMH!!
16 CHAPTER 3. EXPERIMENTAL TECHNIQUE

chamber. We call this time the TPROTON. Due to the large timescale of the
supercycle, we clearly see the decay characteristics of the beam. This is a
good way to access whether the beam is indeed 9 Li. In Figure 3.2 we see the
TPROTON-spectrum for 9 Li induced reactions. The clear steps in increments
of 1.2 s corresponds to the super cycle. The fast rise in the beginning is
the production, extraction and acceleration time, before they end up in the
chamber. This is followed by a exponential slope, which corresponds to
the decay of 9 Li. This is key indication the 9 Li composition of the beam.
This can also be used to limit background contributions. Backgrounds
from the buffer gases would be uncorrelated with TPROTON and give a flat
contribution. By restricting the TPROTON to the first 3-4 life times, we can
omit a considerable amount of background, however, we will also discard
some false negatives. This will, however, be a small fraction since to most
of the 9 Li arrives in the first second or so.

A better constraint can be achieved from the EBIS release. The trap
releases a bunch with a well defined repetition frequency (typically around
30 Hz) and each of these "spills" have a well defined length. We also record
a timestamp when the trap is opened, so we can calculate the time from
the begin of the spill and a given event. We refer to this time as TEBIS,
and a spectrum can be seen in Figure 3.2. The events are nicely constained
to a region of ≈ 700 µs, corresponding to the release time. We can thus
also use this to exclude background components which are recorded outside
this window. Background sources such as noise is in general uniformely
distributed and can by suppressed significantly in this way.

Since the EBIS release is relatively short and the release curve is concen-
trated in the first 100 µs, we effectively get very high instanous rates. With
a repitition rate of 30 Hz and a release of ≈ 700 µs we recieve all the beam
in only 700 µs · 30 Hz = 21 ms of each second. This mean that the average
intensity during a release will be two orders of magnitude larger than the
reported intensity. It will be even larger in the begining of the release. This
put extra requirements on the data acqusition system in order to keep up.
3.2. DETECTORS 17

105 103

104
102
Counts/20 ms

103

Counts/1 µs
1.2s
102 101
2.4s

101
100
100

10−1 10−1
0 1 2 3 4 5 6 0 100 200 300 400 500 600 700 800 900 1000
TPROTON [s] TEBIS [µs]

Figure 3.2: Left: TPROTON spectrum for 9 Li beam on deuterated plastic


in IS561A. Right: TEBIS spectrum for 9 Li beam on deuterated plastic in
IS561A.

3.2 Detectors
Our main workhorse for charged particle spectroscopy is silicon detectors.
The silicon detectors works as a pn-juntion in reverse bias. When a charged
particle enteres the detector an electron-hole pair is created. By applying a
bias across the detector, these carges generate a signal which we can pick
up an digitize. We use generally two types of silicon detectors.
To get position information we use double sided silicon strip detectors
(DSSD). The electrodes that read out the charges are segmented into a
number of strips. On one side of the detector these strips are oriented
horizontally and on the other side they are oriented vertically. When a
particle hit the detector, ideally the charge is split evenly between a strip
on each side (electrons goes to one side, holes to the other side) and we get
a signal in two orthogonal strips. This lets us determine which pixel, i.e.
which small area that overlaps with both strips that was hit.
The second kind of detectors are the simpler unsegmented detectors.
They consist of is a single silicon wafer, and we do not have any positional
information from these. We will refer to these as a pad.
DSSD’s comes in many different sizes and geometries, and I will shortly
describe the kinds we use. I will refer to their unique product names.
18 CHAPTER 3. EXPERIMENTAL TECHNIQUE

W1 Square DSSD with 16 strips on each side. This gives 256 pixels. Each
strip has a width of 3 mm and an interstrip region of 100 µm. They come in
different thicknesses, but during these experiments we only used 60 µm.

BB7 Another square DSSD but with 32 strips on each side. This gives us
a greater angular resolution compared to the W1. The strip width is 2 mm
and the interstrip region is 100 µm.

S3 A round DSSD with a hole in the middle. The front side consists of 32
radially oriented strips (spokes) and the back side consists of 24 rings. Each
ring is 886 µm wide and each radial spoke covers 11.25◦ . The radius of the
inner hole is 11 mm. This detector was used in two of the experiments, and
were 1 mm thick.

3.3 ∆E − E telescope
We can use a combination of two silcon detectors to distinguish between
different particle types. If a particle has enough energy it will penetrate
the detector and only deposit a part of its energy in the detector. If we
place a thick pad detector behind which stops the particle, we can use the
correlation between the energy deposited in the front detector and the back
detector to uniquely identify which particle it was while still measuring the
full energy. This relationship is described by the Bethe Bloch equation. We
call such a setup for a ∆E − E telescope.
If we plot the energy deposited in the DSSD vs the energy deposited in
the pad we get Figure 3.4 (A). The upper feature corresponds to α and the
lower feature corresponds to hydrogen isotopes.
The energy deposited in the front detector, however, dependes on the
distance traveled and thus on the incoming angle,
t
teff = , (3.1)
cos θ
where teff is the distance traveled, t is the thickness of the detector and θ is
the angle shown in Figure 3.3. Each particle, depending on the angle, will
3.3. ∆E − E TELESCOPE 19

teff
DSSD t

θ Particle

Figure 3.3: The angle, θ, between the normal of the detector and the particle
determines the actual length teff the particles travel in the detector.

(A) (B)
12.5 12.5

10 10
∆E · cos θ [MeV]
∆E [MeV]

7.5 7.5

5 5

2.5 2.5

0 0
0 5 10 15 20 25 0 5 10 15 20 25
E [MeV] E + ∆E · (1-cos θ) [MeV]

Figure 3.4: Energy correlation plots for a silicon telescope. In (A) we see
the energy deposited in the thin detector versus the energy in the back
detector. In (B) we have made the geometrical correction described in 3.4
to obtain a better resolution.

thus deposit a different amount of energy even if they are identical. This is
why we can not distinguish between the hydrogen isotopes in 3.4 (A).
Let us assume that the energy loss is constant through the front detector.
The energy deposited in the front detector will then be proportional to the
thickness,

dE
∆E = · t. (3.2)
dx
20 CHAPTER 3. EXPERIMENTAL TECHNIQUE

For a particle incident at θ, this translate to a larger energy loss

dE dE t ∆E 0
∆E = · teff = · = , (3.3)
dx dx cos θ cos θ
where ∆E 0 is the corresponding energy loss if the particle had entered
the front detector orthogonally. With this correction, we get an improved
∆E − E spectrum.
The remaining energy is deposited in the back detector, and the total
correction is

dE 0 = dE · cos θ
E 0 = E + dE · (1 − cos θ), (3.4)

where E is the measured energy in the back detector and dE is the energy
measured in the front detector and θ is the angle between the particle and
the normal of the detector. In Figure 3.4(B) we see an example of an with
this correction. Notice how the three particles in the lower regions are much
better resolved.
4

Beam diagnostics

4.1 Introduction
In experimental work with radioactive ion beams (RIB) the beam is often
assumed to be ideal point-like and incident in the target at a well known
position. This is, however, often not the situation and the beam’s emittance
and displacement are nonzero. Minimizing the emittance of the beam is a
major challenge in the design and operations of accelerator facilities.
In the resulting analysis of the experimental data, and in particular
single event measurement, the energy is deduced based on the assumption
that we know the geometrical profile and energy distribution of the beam.
When this is not the case we introduce a bias in the analysis leading to an
unnecessary decrease in resolution of our excitation spectra.
To correct this, we propose a method to deduce the beam’s prop-
erties from coincidence measurements. The method is partly based on
[johansen2013]. It has the advantage that it is applicable directly on the
primary experiment of scientific interest. The method applies to reactions
of the type
A+a→B+b
where both particles are detected. This of course includes elastic scattering.
We start with a theoretical description of the method. This is followed
by a demonstration of the method on simulations of the reaction 9 Li(d, t)8 Li
using a 5 MeV beam incident on a deteronated carbon foil.

21
22 CHAPTER 4. BEAM DIAGNOSTICS

4.2 Theory
Firstly, we present a short review of the method described in [johansen2013].
Thereafter we describe the new method. Both methods are derived for a
general reaction of the form

A + a → B + b,

where we take A as the beam and a as the target.

Old method
The azimuthal angle of the two outgoing particles are given as

y
tan(φ) = . (4.1)
x
In the center-of-mass frame, the two particles will move back to back,
due to momentum conservation, i.e.

|φb − φB | = π. (4.2)

This is, however, unaffected by the transformation along the z-axis, and
hence, combining (4.1) and (4.2) we get the relation

yb yB
= ,
xb xB
which holds as long as the beam was incident in (0,0). If this is not the
case, but it is incident in (xA , yA ) we get the following relation

yb − yA yB − yA
= .
xb − xA xB − xA
We now assume that the offset is as small as possible, that is we need
to minimize x2A + yA 2 . This results in the following two relations (note that

there is a sign error in the last parenthesis in [johansen2013])


4.2. THEORY 23

x b yB − x B yb
xA = (yB − yy )
(yb − yB )2 + (xb − xB )2
x b yB − x B yb
yA = (xb − xB )
(yb − yB )2 + (xb − xB )2

New method
In the new method we can calculate the momentum of the beam particles
from momentum conservation in the LAB frame (the target, b, is at rest)

p̄A = p̄B + p̄b


s s
2EB 2Eb
= mB v̂B + mb v̂b
mB mb
= 2EB mB v̂B + 2Eb mb v̂b ,
p p

where mi is the mass of the particles, Ei is the measured energy of the


particle in the detector and v̂i is the direction of the particles movement
given as

(xi − xA , yi − yA , zi − zA )
v̂i = p
(xi − xA )2 + (zi − zA )2 + (zi − zA )2
The energy of the incoming beam is given as

p2A
EA = (4.3)
2mA
1
(2EB mB +2Eb mb
= 2mA (4.4)
+2 4EB mB Eb mb v̂B · v̂b ).
p

Due to energy conservation, the measured energies of the two outgoing


particles should be equal to EA + Q, i.e.
24 CHAPTER 4. BEAM DIAGNOSTICS

EA = EB + Eb − Q. (4.5)
Subtracting (4.5) from (4.4) should thus equal zero. We can now do a
numerical minimization to find the position of the beam that best full fills
this criterion.

Angle determination
One we have estimated the displacement of the beam, we can estimate the
angle, since they are given as

q
p2A,x + p2A,y
tan(θA ) = (4.6)
pA,z
p2A,y
tan(φA ) = (4.7)
pA,x
In the analysis we will not use these angles, but the angles projected
onto the x an y-axes

θx = θA cos(φA ) (4.8)
θy = θA sin(φA ) (4.9)

4.3 Setup and simulation


The simulations is performed with the Monte Carlo method. An incoming
beam particle is propagated into the target where it undergoes the reaction.
The path length inside the target is uniformly distributed over the thickness
of the target. After the reaction, the two daughter nuclei are placed back
to back in the center-of-mass frame and rotated to a direction uniformly
sampled from the unit sphere. We model the energy loss in the various
materials using the GEANT range tables. We start off with an ideal point-
like beam with an energy of 45 MeV (5 MeV per nucleon). The beam is
4.4. ANALYSIS 25

incident on a deuteron target. We take the z-axis as the beam direction.


The deuteron target is embedded in a 1 mg/cm2 carbon foil which extends
10mm×10mm in the xy-plane.
We use the setup shown in Figure 4.1. The symmetry in the azimuthal
angle φ allow us to showcase some important features of the method which
will become clear in the later sections. We will now give a technical descrip-
tion of the detectors but restrict us to the DSSSD’s. Each DSSSD is in
a telescope configuration with a single sided 1.5 mm silicon pad detector
placed behind it. Since we are measuring coincidence event in two detectors,
we will call the axis that connects the center of the two detectors for the
"primary axis". The reason for this, will also become clear in the discussion
of the results. The axis orthogonal to this axis will be called the "secondary
axis"
The geometry of the setup is mainly motivated by the kinetic curves
of the reaction. We want to measure 11 Li in coincidence with tritons. The
11 Li is emitted up to θ ≈ 20◦ . We thus need a least one detector that is

close enough to the z-axis to measure the 11 Li.


The setup consists of five 16x16 double sideded silicon strip detectors
(W1’s). The surrounding W1’s are placed in a distance of 55mm from origin
and are angled such that their normal vector points to origin. The central
W1 is placed in a distance of 62.4 mm from origin.
In this setup we have four primary axes, each pairing the central W1
with one of the peripheral W1’s.
The W1’s detectors are 60 µm thick. They have a 100 nm deadlayer.
Their strip width is 3.0 mm and their strip pitch is 3.1 mm.

4.4 Analysis
To apply the method we iterate over all events, check for coincidence and
find the numerical solution to

p2A
= EB + Eb − Q. (4.10)
2mA
26 CHAPTER 4. BEAM DIAGNOSTICS

Figure 4.1: The setup used in the simulations.

This gives the position of the incident beam and thus the incoming
momentum. From this we calculate the corresponding angles, using (4.8)
and (4.9). We will firstly look at an ideal beam and restrict us to the
position estimate in order to learn a important feature of the method.
With this knowlegde we can continue to look at an angled beam. In real
application we will experience both displaced and angled beams, so we will
afterwards look at an combination of the both displacement and angling.
To assess the geometrical dependencies we will then look some of the factors
that determine the accuracy. Finally we will discuss energy estimation
distribution.
4.4. ANALYSIS 27

20 6

20 12 15
5

15 10
10
4
10 5

y[mm]
8
5 0 3
y[mm]

0 6 −5
2

−5 − 10
4
1
− 10 − 15

2
− 15 − 20 0
−20 − 15 −10 −5 0 5 10 15 20
x[mm]
− 20 0
−20 − 15 −10 −5 0 5 10 15 20
x[mm]
Figure 4.3: The total po-
Figure 4.2: The total po- sition distribution for all
sition distribution for all concidence measurements
concidence measurements. with a solution to (4.10).

500 X projection

400
Y projection

300
N

200

100

0
−20 − 15 −10 −5 0 5 10 15 20
[mm]

Figure 4.4: The projec-


tion of position distribu-
tion for all concidence
measurements with a solu-
tion to (4.10).
28 CHAPTER 4. BEAM DIAGNOSTICS

Beam position
We apply the method to a simulation of an ideal beam and the previous
described setup. This position estimates are shown in Figure 4.2. Firstly,
we notice that the setup structure is really pronounced. Secondly we notice
that the distribution is broad. It extends to approximately 15 mm in both
directions, even though our target only is 10 mm in both directions. This
can, however, be enhanced by noticing that (4.10) is quadratic in pA . There
may thus not be a solution or we can get two solutions. If there is no solution
we ignore the measurement, since it can not reproduce our conditions. If
we have two solutions we take the one closest to the z-axis, assuming we
are as close to the z-axis as possible.
If we disregard the coincidences that yields no solution to (4.10) we get
the plot in Figure 4.3. We removed many of the outer estimates, confining
the estimates within approximately 10 mm. In Figure 4.4 are the projections
on the x-axis and y-axis shown respectively. The distributions are centered
around (0 mm, 0 mm) but the distributions are so broad that it is hard to
tell whether the centers of the distributions are shifted slightly.
Let us consider what is actually happening in the reaction. Each reaction
spans a reaction plane, defined by three points: the reaction position (i.e.
the beam position) and the two detected particle positions. The method
estimates one of these points, namely the reaction position. Ideally we
should take the intersection of all these planes in the target plane. This is,
however, not feasible since we do not have an infinite resolution. Instead we
average over ensembles of measurements. We do this by defining specific
planes of interest, namely the primary axes of the detector pairs, and make
an averaged estimate from each of these. The intersection in the target
plane will then be our estimate of the beam’s position.
More specifically, we plot the position estimates from the method for
each detector pair separately. This is shown in Figure 4.5. We see that we
get the best estimate along the secondary axis of the two detectors. The
reason for this is twofold. It is partly because of the detectors geometry.
We have "more detector" along the primary axis than perpendicular to it.
The second reason is the method itself. We have much more freedom to
4.4. ANALYSIS 29

match momenta in the reaction plane than perpendicular to it. This is due
to the minimization criterion. A projection onto the primary axis and the
secondary axis clearly show this effect, see Figure 4.6.
To make an averaged estimate over the ensemble, we fit a straight line
with a fixed slope (the primary axis) for each of the pairs. This means we
only estimate the displacement along the secondary axes. This line is our
first estimate; the reaction took place somewhere on that line in the targets
xy-plane. If we plot each of these lines we see that their intersection is very
close to origin, see Figure 4.7. We can thus estimate the position of the
beam as the intersection point by minimizing the sum of the square of the
distances to each line, yielding (0.02 mm,0.01 mm).
If we displace the beam to be incident in (1 mm, 1 mm) we get the result-
ing lines in Figure 4.8. The displacement is estimated to be (0.99 mm,0.98 mm).
The method thus give a good estimate of the displacement.

Angled beam
We now turn to an angled beam. As described we can calculate the angle of
the beam from the estimated beam position and conservation of momentum.
For this analysis we also isolate each detector pair. Figure 4.10 shows
the angle spectrum from the detector pairs of a beam angled θ = 50 mrad
and φ = 45◦ , i.e. θx = 50 mrad · cos(45◦ ) = 35.3 mrad and θy = 50 mrad ·
sin(45◦ ) = 35.3 mrad. We notice that many counts are in (0,0). The method
are thus insensitive to coincidences angled in the inferred reaction plane, as
we also saw in the position estimate. This means we must look at the angle
relative to the reaction plane, i.e. the angle with the local x-axis, θx .
In Figure 4.11 are the projections onto the primary and secondary axes
shown. We see the y-projections are mainly peaked in zero, whereas the
x-projections show the same behaviour as with the displacement analysis.
That is, they are shifted and we can once again fit a line. This time it is,
however, the axis orthogonal to the primary axis. We also note that the
distributions are very broad. Gaussian fits to the x-axis give a σ in the order
of ≈ 45 mrad. This is the order of error we expect due to the resolution of
our detectors. This will be discussed in more detail later. We can, however,
30 CHAPTER 4. BEAM DIAGNOSTICS

Det1 Det2
15 7 15 6

6
10 10 5

5
5 5 4

4
y [mm]

y [mm]
0 0 3
3

−5 −5 2
2

−10 −10 1
1

−15 0 −15 0
−15 −10 −5 0 5 10 15 −15 −10 −5 0 5 10 15
x [mm] x [mm]

Det3 Det4
15 6 15 6

10 5 10 5

5 4 5 4
y [mm]

y [mm]

0 3 0 3

−5 2 −5 2

−10 1 −10 1

−15 0 −15 0
−15 −10 −5 0 5 10 15 −15 −10 −5 0 5 10 15
x [mm] x [mm]

Figure 4.5: The position distributions for coincidence measurements in the


central W1 and each of the surrounding W1’s.
4.4. ANALYSIS 31

Det1 Det2
200
200
X projection X projection
180
180
160
160
Y projection 140
Y projection
140
120
120
N

N
100 100

80 80

60 60

40 40

20 20

0 0
−15 −10 −5 0 5 10 15 −15 −10 −5 0 5 10 15
x [mm] x [mm]

Det3 Det4
200
X projection 180 X projection
180
160
160
Y projection 140 Y projection
140
120
120
100
N

100
80
80

60 60

40 40

20 20

0 0
−15 −10 −5 0 5 10 15 −15 −10 −5 0 5 10 15
x [mm] x [mm]

Figure 4.6: Projections of the position distributions for coincidence mea-


surements in the central W1 and each of the surrounding W1’s.
32 CHAPTER 4. BEAM DIAGNOSTICS

Position estimate Position estimate


4 4

3 3

2 2

1 1
y [mm]

y [mm]
0 0

−1 −1

−2 −2

−3 −3

−4 −4
−4 −3 −2 −1 0 1 2 3 4 −4 −3 −2 −1 0 1 2 3 4
x [mm] x [mm]

Figure 4.7: The position Figure 4.8: The position


estimate of a point beam estimate of a point beam
incident in (0mm,0mm). incident in (1mm,1mm).
100
Angle estimate
80

60

40
θ y [mrad]

20

− 20

− 40

− 60

− 80

− 100
−100 −80 −60 −40 −20 0 20 40 60 80 100
θ x
[mrad]

Figure 4.9: The angle es-


timate of a point beam in-
cident in (0mm,0mm) but
angled (θx = 35.3 mrad,
θy = 35.3 mrad).
4.4. ANALYSIS 33

Det1 Det2
100 100 35
35
80 80
30
60 30 60

40 40 25
25
20 20
θ y [mrad]

θ y [mrad]
20
20
0 0
15
−20 15 −20

−40 −40 10
10
−60 −60
5 5
−80 −80

−100 0 −100 0
−100 −80 −60 −40 −20 0 20 40 60 80 100 −100 −80 −60 −40 −20 0 20 40 60 80 100
θ x [mrad] θ x [mrad]

Det3 Det4
100 100 40
40
80 80
35
35
60 60
30 30
40 40

25 25
20 20
θ y [mrad]

θ y [mrad]

0 20 0 20

−20 −20
15 15
−40 −40
10 10
−60 −60
5 5
−80 −80

−100 0 −100 0
−100 −80 −60 −40 −20 0 20 40 60 80 100 −100 −80 −60 −40 −20 0 20 40 60 80 100
θ x [mrad] θ x [mrad]

Figure 4.10: The angle distributions from the new method. There are many
counts in (0,0), meaning the method is rather insensitive to coincidences
angled in the reaction plane. This means we should use the angle relative to
the inferred reaction plane, i.e. θx .
34 CHAPTER 4. BEAM DIAGNOSTICS

Det1 Det2
250 400

X projection X projection
350
200
300
Y projection Y projection
250
150
N

N
200

100
150

100
50
50

0 0
−100 −80 −60 −40 −20 0 20 40 60 80 100 −100 −80 −60 −40 −20 0 20 40 60 80 100
θ x [mrad] θ x [mrad]

Det3 Det4
300 350
X projection X projection

250 300

Y projection Y projection
250
200

200
N

150
150

100
100

50
50

0 0
−100 −80 −60 −40 −20 0 20 40 60 80 100 −100 −80 −60 −40 −20 0 20 40 60 80 100
θ x [mrad] θ x [mrad]

Figure 4.11: Projection of the angle distributions from the new method. The
x-projection shifts, and can thus be used to estimate the total angle of the
beam.
4.4. ANALYSIS 35

6000
µ=44.9MeV 1800 µ=44.9MeV
σ =0.0615MeV σ =0.255MeV
1600

5000
1400

1200
4000

1000
N

N
3000
800

2000 600

400
1000
200

0 0
36 38 40 42 44 46 48 50 52 54 36 38 40 42 44 46 48 50 52 54
E[MeV] E[MeV]

(a) Point beam (b) σ = 0.25 MeV

µ=44.9MeV
500 σ =0.979MeV

400

300
N

200

100

0
36 38 40 42 44 46 48 50 52 54
E[MeV]

(c) σ = 1 MeV

Figure 4.12: The energy estimates for a) a point beam b) a gaussian energy
distribution with σ = 0.25 MeV and c) a gaussian energy distribution with
σ = 1 MeV
36 CHAPTER 4. BEAM DIAGNOSTICS

improve these distribution by using the old method. If we use this instead
we get the projections show in Figure 4.13.
We see that our x-projections are more well defined. If we use these
spectra instead, fit straight lines and estimate the intersection we get an
estimate of (32.8 mrad, 33.0 mrad), see Figure 4.9. We underestimate slightly,
but we are close to the expected angling.

Combined displacement and angling


In a real experiment we must assume that the beam is both angled and
displaced so we will now apply the technique to such a situation. We let the
beam be angled with θ = 25 mrad and φ = 45◦ , resulting in θx = cos(45◦ ) ·
25 mrad ≈ 17.6 mrad and θy = sin(45◦ ) · 25 mrad ≈ 17.6 mrad. We start
the beam in (0,0,−100 mm) and we thus expect the incident position to be
x = tan(17.6 mrad) · 100 mm ≈ 1.77 mm and y = tan(17.6 mrad) · 100 mm ≈
1.77 mm.
We can now perform the full analysis. We use the new method to
estimate the position and the old method to estimate the angle and we get
the result shown in Figure 4.15. The estimates are for position and angle
are respectively (2.49 mm, 2.53 mm) and (40.6 mrad, 40.3 mrad).
The method can thus not distinguish between the two effects completely.
We clearly see that the beam is off, but we can not quantify it directly with
good accuracy. There is thus a bias in the method. In order to correct for
this bias, one have to do simulations of the particular setup. By simulating
different beam configurations it should be possible to converge towards the
results of the analysis of the actual experiment. The beam configuration
from the simulations is thus the best estimate.

Broad beam
Typically a beam is also spatially distributed. The exact shape is due to
various accelerator effects and therefore very complex. We will, however,
simulate a simple spatial distribution to see how our method responds.
In Figure 4.16 we see the distribution of one detector pair using a beam
4.4. ANALYSIS 37

Det1 Det2
800
120 X projection X projection
700
100
Y projection 600 Y projection

80 500
N

N
400
60

300
40
200
20
100

0 0
−100 −80 −60 −40 −20 0 20 40 60 80 100 −100 −80 −60 −40 −20 0 20 40 60 80 100
θ x [mrad] θ x [mrad]

Det3 Det4
140 800
X projection X projection
700
120

Y projection 600 Y projection


100
500
80
N

400
60
300

40
200

20 100

0 0
−100 −80 −60 −40 −20 0 20 40 60 80 100 −100 −80 −60 −40 −20 0 20 40 60 80 100
θ x [mrad] θ x [mrad]

Figure 4.13: The projection of the angle distributions using the old method.
We see that the x-projection shifts and we can thus use that for our angle
estimate.
38 CHAPTER 4. BEAM DIAGNOSTICS

Figure 4.14: We can look a narrower slices by only considering the center
half (the red and green area) or the center fourth (the green area).

with a radius of 2.5 mm. The position distributions get slightly broader on
the secondary axis, and the angle distribution is slightly elongated on the
primary axis. We can thus not quantify the beams spatial distribution. It
does, however, not affect the results of the position and angle estimates. We
can thus still estimate the primary component of the beam.

Resolution and spread in ensembles


As mentioned earlier, what we really do is analysing possible reaction planes
for each coincidence. This is what leads us to the conclusion that we are
4.4. ANALYSIS 39

Position estimate
100
Angle estimate
4

80
3
60
2
40
1

θ y [mrad]
20
y [mm]

0 0

− 20
−1
− 40
−2
− 60
−3
− 80

−4 − 100
−4 −3 −2 −1 0 1 2 3 4 −100 −80 −60 −40 −20 0 20 40 60 80 100
x [mm] θ x
[mrad]

(a) Position (b) Angle

Figure 4.15: Position and angle estimate of a beam which was both displaced
and angled. We do not recreate the correct values. The method are thus
biased when both effect are present.

most sensitive secondary axis when estimating position and angles. It thus
seems logical to make the analysed pairs "thinner". This can be done by
looking at slices of the segmented detectors. That is, we might only look
at the center strips of the surrounding W1’s in the setup. In this way we
should narrow the spread in the estimates from the coincidences.
Another limiting factor is the resolution of the detectors. We can only
detector particles within one pixel. This gives a positional resolution in the
order of the pitch and an angular resolution in the order of 2 · tan−1 ( p/2
d ),
where p is the pitch and d is the distance from the target. For the W1’s in
our setup the positional resolution is ≈ 3.1 mm and the angular resolution
is ≈ 100 mrad. By using detectors with a better resolution, we thus also
expect to lower the spread in the estimates.
To investigate the effect of looking at narrower slices, we perform the
analysis only using the center half and the center quarter of the back strips
of the surrounding detectors as sketched in Figure 4.14. To investigate the
effect or better resolution, this is done for both W1 and an identical setup
40 CHAPTER 4. BEAM DIAGNOSTICS

Det1 Det1
15 7 15 6

6
10 10 5

5
5 5 4

y [mm]

y [mm]
0 0 3
3

−5 −5 2
2

− 10 − 10 1
1

− 15 0 − 15 0
− 15 −10 −5 0 5 10 15 − 15 −10 −5 0 5 10 15
x [mm] x [mm]

(a) Point beam:(b) Broad beam:


Position Position

Det1 Det1
100 100 100

80 120 80

60 60 80
100
40 40

80
θ y [mrad]

θ y [mrad]

20 20 60

0 0
60
− 20 − 20 40

− 40 40 − 40

− 60 − 60 20
20
− 80 − 80

− 100 0 − 100 0
−100 −80 −60 −40 −20 0 20 40 60 80 100 −100 −80 −60 −40 −20 0 20 40 60 80 100
θ x
[mrad] θ x
[mrad]

(c) Point beam:(d) Broad beam:


Angle Angle

Figure 4.16: Position and angle distributions for a point beam and a spatial
uniformly distributed beam with a radius of 2.5 mm.

with BB7 1 detectors instead. The results are given in Table 4.1.
We see that an increase in resolution is to be preferred since it decreased
the spread in both position and angle. Looking a narrower slices decreases
the position, but not as much. We do, however, see that the spread in angle
increases. This means we need to do slices in the other direction if we want
it to narrow the angle interval.

Energy distribution
We can estimate the energy distribution of the beam by calculating back-
wards from our position and angle estimate. For each coincidence we
calculate the energy using (4.4). In Figure 4.12 we see the results for a
1
A square DSSSD detector with 2 mm pitch and 32x32 strips.
4.5. CONCLUSION 41

σθ,new σθ,old σxy


Full 42.9 mrad 15.8 mrad 1.38 mm
W1 Half 46.7 mrad 16.0 mrad 1.15 mm
Fourth (49.9 mrad) (15.5 mrad) 1.10 mm
Full 28.6 mrad 9.96 mrad 1.01 mm
BB7 Half 31.2 mrad 10.5 mrad 0.79 mm
Fourth (32.9 mrad) (10.6 mrad) 0.70 mm

Table 4.1: The sample deviation of a coincidence measurements, using


either the whole outer detector, only the center half or only the center fourth.
Furthermore this is tested with to types of detectors with different resolution.
We see that the main improvement is due to higher resolution detector.

µ σ
Point 44.9 MeV 0.06 MeV
σ = 250 keV 44.9 MeV 0.26 MeV
σ = 1 MeV 44.9 MeV 0.98 MeV

Table 4.2: The deduced energy distributions of three beams with different
energy distributions.

point beam, a gaussian shaped energy distribution with σ = 250 keV and a
gaussian shaped energy distribution σ = 1 MeV, respectively. Applying a
gaussian fit get the results in Table 4.2. There are clear very close to the
simulated values. We can thus accurately determine the beam energy from
coincidence measurements.

4.5 Conclusion
The proposed method can be used to estimate a beam’s properties and in
particular estimate a displacement and angling of a beam using coincidence
measurements in segmented detectors. The method can not, however,
distinguish between the effects. One can, however, use simulations of the
42 CHAPTER 4. BEAM DIAGNOSTICS

setup to find the beam that yields same results as the analysis of the actual
experiment. In order to get the best intersection estimates we need multiple
different primary axes. The setup should thus cover multiple directions in
the azimuthal angle. Four detectors are for instance sufficient.
5

AUSAlib

The Aarhus subatomic group are involved in many experiments, both in


house and at different international facilites such at ISOLDE. The experi-
mental setups are typically similar with multiple DSSDs, pads and telescope
configurations. The initial steps of an analysis are therefore also similar.
The raw data files are unpacked into a more suitable data format, with
ADC and TDC values mapped to the right detectors. A calibration is
applied and a matching step is performed for the double sided detectors
that ensure consistency between the front and back side of the detector.
The geometry of the detectors in a global coordinate system must be laid
out, in order to calculate positions and angles for the hits.
These are some of the tasks that every analysis starts with. To streamline
this process and make it less error prone, a common analysis library has been
developed to perform some of these tasks, the Aarhus University Subatomic
library (AUSAlib). It is written in C++ and is built on top of ROOT. The
development began before I started working in the group by M. Munch, O.
Kirsebom and A. Howard. During my PhD I contributed substantially to
the development of this project.

5.1 Geometry
A core feature of AUSAlib is the generalization of geometrical properties
of the detectors. When a particle hits a double sided detector, the front

43
44 CHAPTER 5. AUSALIB

3 2 1 2 3 ... Orientation

1
4 1
2
n
tio
ta
ri en 3
O
..
...

.
Normal Normal

0 Center Center
1
2
3
4
...

Figure 5.1: Geometry for a round DSSD and a square DSSD.

strip and the back strip is recorded in the data stream. This translate to a
position in space, but the transformation depends on the specific geometry
of the DSSD. In Figure 5.1 we see to very different geometries that can
still be described orthogonal front and back strips. The interface in the
analysis is thus the same. Given a front strip and a back strip, calculate a
position. We can thus have a single superclass that exposes the interface
for a DSSD. Concrete implementations, such as the round DSSD, can then
specify the exact transformation. For the experiments in this thesis we can
use the square DSSD to describe both the W1 and the BB7, whereas we
can describe the S3 with the round DSSD.
Even more general is the position and orintation in space. This is
required for all types of detectors, segmented or unsegmented, in order to
place them in a global coordinate system and rotate them correctly.
This leads to a class hierachy as shown in Figure 5.2.
This approach has to main advantages. Firstly the data is decoupled
from the geometry in the sense that a hit only specify the segments that were
hit, i.e. the strips. Then it is up to the concrete detector implementation to
5.1. GEOMETRY 45

Detector

SingleSidedDetector DoubleSidedDetector

SingleSidedSiliconDetector CloverDetector DoubleSidedSiliconDetector

SquareSSSD PadDetector Wedge RoundDSSD SquareDSSD

Figure 5.2: Detector hierachy in AUSAlib.

calculate what position this corresponds to. Secondly is it easy to add new
detectors. It just a matter of implementing the correct transformation, and
then it can be used in an existing analysis through the common interface.
This favors code reuse and modifications are done by addition, keeping the
codebase backward compatible.

Setup file
For a given experiment the position and type of the different detectors
must be specified to the analysis. To ease this process, AUSAlib uses a
socalled setup file, that encompass all the geometry specific details for each
detector, as well as which detectors that are meant to be a telescope. An
example of the description of a detector is shown in Figure 5.3. Each
detector has a name. The "file" contains details such as detector type
(from the detector hierachy), number of strip, deadlayer thickness, etc. The
"calibration" file contains calibrations for each strip. The "position",
"normal" and "orientation" specify the position and orientation in the
global coordinate system. The data mapping to the unpacked files are
described in "frontmapping" and "backmapping".
Once such a file is written, typically during the setup of an experiment,
it can be used throughout all analysis and even the online analysis. AUSAlib
simply loads this file making it easy to makes changes without having to
46 CHAPTER 5. AUSALIB

1 {
2 "name ": "Det1",
3 "file ": "DSSD1.json",
4 " calibration ": "Det1.cal",
5 " position ": {"x": "-4mm", "y": " -42.0 mm", "z": "39.0 mm"},
6 " normal ": {"x": "0.0 mm", "y": "1.0 mm", "z": "0.0 mm"},
7 " orientation ": {"x": "0.0 mm", "y": "0.0 mm", "z": "1.0 mm"},
8 " frontMapping ": {
9 " multiplicity ": "DET1F",
10 " segment ": " DET1FI ",
11 "adc ": " DET1F_E ",
12 "tdc ": " DET1F_T "
13 },
14 " backMapping ": {
15 " multiplicity ": "DET1B",
16 " segment ": " DET1BI ",
17 "adc ": " DET1B_E ",
18 "tdc ": " DET1B_T "
19 }
20 }

Figure 5.3: ""

recompile a large analysis codebase. Moreover is the full geometrical details


contained in few files that can easily be shared with collaborators.

5.2 Energy loss


From the reaction site to the particle is detected in a detector, it may loose
energy in inactive material layers such as deadlayers or target components.
These energies can not be measured, so they must be corrected for. Such
energy losses has been tabulated for a large number of charged particles in
different materials and AUSAlib includes tables from GEANT[], ICRU and
SRIM.
5.3. ANALYSIS PIPELINE 47

Custom analysis

.lmd .root .root .root


Readout Unpacker Sorter Identifier Eventbuilder

Calibrator .cal

Figure 5.4: ""

5.3 Analysis pipeline


A key operating principle in AUSAlib is the idea of an analysis pipeline.
The data starts out as raw data and the is piped through a series of steps
which increases the level of abstraction. The data is thus transformed from
ADC and TDC values to observables describing physical events.
The pipeline is shown in Figure 5.4, and the steps will be described
in more detail in the following. Each step outputs ROOT files that can be
investigated directly in ROOT or via custom written analysis program using
the AUSAlib infrastructure This is useful to make consistency checks along
the way. The setup-file is used in every step of the pipe line. This keeps
the geometry and detector specifications consistent throughout the pipeline.

Unpacking
The DAQ saves the files in a GSI list mode format (.lmd). The first step
is to convert these files to ROOT files that can be read into the rest of the
AUSAlib pipeline. This is done with ucesb. The user is required to make a
precise specification of which modules are used, how the data is layed out in
the data stream as well as the mapping from the individual ADC channels
to the detector segments. Typically this is done during the setup of the
DAQ, such that it can be used during the experiment and in the subsequent
analysis. One major strength of ucesb is that it checks that every single
48 CHAPTER 5. AUSALIB

bit in the data is consistent with the data layout of the modules. If the file
unpacks, the data is thus not corrupt.

Energy calibration
The first task after the unpacking is to energy calibrate the detectors. We use
an α-source with well defined peaks. In the ISOLDE experiments we used
a source with 244 Cm, 241 Am, 239 Pu and 148 Gd which all give well defined
nearly monoenergetic peaks. The source is placed in a well known position.
The energy response in a silicon detector is highly linear [REF??] and we
can now fit a linear relationship between the ADC channel corresponding
to the centroid of the peak and the well known tabulated energy. There are,
however, a few small corrections we need to make to get a better calibration.

Deadlayer
There is small deadlayer on both sides of a silicon detector. This is a small
layer of inactive material. The α-particle loses energy, but the created
electron-hole pairs are not collected. This means the measured energy is
slightly lower than the actual energy

dE ∆t
E 0 = Eα − , (5.1)
dx cos θ
where ∆t is the thickness of the deadlayer and θ is the angle shown in
Figure 3.3 and Eα is the tabulated α energy. The dead layer is assumed
to be so small that the energy loss is constant. It is thus important that
the geometry of the detector as well as the source position is known, so θ is
calculated properly.

Shift calculation
Each segment is calibrated individually, but extends over many angles. The
measured energies should therefore be averaged over the full surface of the
segment. Since our positional resolution are limited to the pixels the average
5.3. ANALYSIS PIPELINE 49

can be approximated with a weighted average with the solid angle of the
pixels a weight

X dE ∆t Ωpixel
E 0 = Eα − , (5.2)
pixel
dx cos θpixel Ωstrip

where θpixel is the angle at the center of the pixel.


Since there are many strips that has to be calibrated, AUSAlib provides
a calibration tool, the Calibrator. It implements a peak-finding algorithm
to find peaks, perform the correction in 5.2, perform a linear fit and display
the result to the user. The user can then quickly browse through and check
the calibrations. If the algorithm misidentified a peak, the user can easily
specify which peaks corresponds to the different energies for that specific
segment and redo the calibration. This reduces the time to calibrate several
hundred segments quite significantly.

Sorting
The sorting step has two main jobs, and is performed by a general purpose
tool, the Sorter. Firstly it applies the calibrations. These can come from
the Calibrator as described above, or other means of calibration. Only
linear calibrations are, however, implemented.
Secondly it performs matching to determine which pixels was hit by a
particle. When a particle hits a DSSD a signal is detected in both sides.
If there is only one hit in the back side and one hit in the front side they
must match. If, however, there are more than one hit in one or both sides
we need to determine which pairs that match. There might also be noise
that only shows up in one side.
The Sorter tool implement a greedy algorithm that computes the dif-
ference between all pairs. The pair with the smallest difference is selected
as the first pair and those hits are excluded from the pool of candidates.
The next pair is now the pair with the smallest energy difference from the
remaining hits. This continues until there are no more pairs with an energy
difference between a certain thresholds.
50 CHAPTER 5. AUSALIB

The majority of hits will hit a strip in both the front and the back, but
a minor part will hit in an interstrip region. This results in a sharing of the
energy between two strips. It is also possible that multiple particles hit the
same front strip but different backstrip and we see a summing of the energies.
Neither of these situations can be handled by the current algorithm, but they
can mostly be recovered by a more sophisticated algorithm that considers
all pairs and tries to reconstruct both normal, sharing and summing.

Identification
After the sorting, the energy and the position of the hits are determined.
This enables the next step, which is trying to identify what particle a
given hits corresponds to. There are different techniques to perform this
identification. When the detectors are organized as telescopes we can use
the ∆E − E method to seperate the different isotopes. If the energy is
too low to penetrate the ∆E detector, it may be possible to exclude some
particle types simply based on how much energy they lost in the detector.
This is, however, not unambigious, and a hit can have several candidate
identifications. Based on each of these identifications, the energy is corrected
for energyloss in deadlayers and the target.
The identification and energyloss is performed by another tool, the
Identifier. It requires details on the target and the particles too look for,
which are supplied either as a command line argument or throguh simple
text files.

Eventbuilding
From the files with identified particles, the eventbuilding stage attempts to
reconstruct acutal physical events from the detected hits. It requires one or
more final states to look for or reconstruct. If an event contains an identified
deuteron as well as hit that is consistent with a 9 Li, the four-momentum
for both are computed based on the position of the hit and the energy in
the detector. If only a deuteron is detected it can try to reconstruct the
5.3. ANALYSIS PIPELINE 51

four-momentum of the 9 Li based on the initial four-momentum of the beam


and the target.
A user supplied analysis can subsequently perform the final analysis
based on the four-momentum and particle identifications. This includes
extraction of excitation energies, CM angles, particle correlations, etc.
6

Data acqusition

The data acqusition system (DAQ) is an essential part of any experiment.


The DAQ is loosely speaking the set of electronics and software that converts
the incoming signals from the detectors to files on a computer that can be
analyzed. We will divide the description into two main parts, the backend
and the frontend. A schematic overview of the essential parts in the DAQ
is shown in Figure 6.1.
The backend handles amplification and digitizing of the analogue signals
from the detectors as well as deciding which signals are worth saving
(discrimination).
The frontend recieves the digitized data is first and foremost responsible
for storing the data. Moreover should the software give the operators of the
experiment some degree of control and monitoring such as an preliminary
online analysis.
The bridge between the two is the readout that extract the data from
the modules and transfer them to the frontend. The different components
will be described in more detail in the following sections.
During my phd I took part in the setup of the DAQ in IS561A and
did the majority of the DAQ for IS561B and IS561C. With guidance and
some help from M.Munch and H.T Johannson I designed and implemented
a two-crate system for IS561C. During the development of such a system I
contributed to the paper "VME Readout at the conversion limit".

53
54 CHAPTER 6. DATA ACQUSITION

Detectors Front-end Back-end

Discrimination Amplification File taking

Digitization Readout Relay ...

Acquisition control Online analysis

Figure 6.1: Figure by M.Munch

6.1 Triggers and acqusition control


In a nuclear physics experiment the reactions happens randomly and we
do not know precisely when to look to capture the result. Instead we must
make this decisicion as the reaction happens, that is we want to trigger
on signals we believe are of interest. We use either a simple threshold
or a constant fraction of the incoming signals to determine if we want to
digitize it. In the simplest case, when we accept to do so, we issue a MASTER
START which tells our ADCs to start digitizing and when they are done
we transport the data to the computer and store it. The ADC takes some
time to digitize the signal and it takes some time to transfer the data to
the computer. During this time we can not accept any new triggers, since
our system is busy. The DAQ is in dead time. Once we are done and the
ADC’s are ready again, we accept a new trigger and repeat. The DAQ runs
synchronously which means that after an accepted trigger, all modules start
the digitization process and the DAQ stays busy until all modules are ready
again. The slowest module limits to life time.
Each channel can produce a trigger, and with an experimental setup
that easily have ≥ 200 channels, it is important to to keep track of which
state the DAQ is in. If busy it rejects the trigger, otherwise it start the
digitization process. This acqusition control (AcqC) is implemented with a
6.2. ANALOGUE CHAIN 55

TDC trigger
TDCs
Busy

Trigger requests
Acquisition Gate
Discriminators ADCs
control Busy

Readout trigger Deadtime

Readout

Figure 6.2: Figure by M.Munch

VULOM 4b running the TRLO II firmware. The AcqC get trigger inputs
from all our input sources, i.e. the detectors. If the DAQ is ready, the
AcqC generates a MS that is sent to the digitizing modules. While they
are digitizing they assert a logic BUSY signal. As long as any module has
asserted BUSY, the AcqC can not accept a new trigger. The readout software
also requires time periods to check that all modules agree on the number of
events they have digitized. If they disagree, there is a fatal flaw somewhere
that needs to be addressed immidiately. During these times, the DAQ can
also not accept new triggers. Lastly, a module may need to be emptied
before it can store a new event. In this case the DAQ also needs to wait in
order to stay synchronized. We will refer to these times as dead time (DT).
The time the DAQ can accept a new trigger is the life time.

6.2 Analogue chain


The ionization from a charged particle creates a small signal in a silicon
detector. A 9 MeV particle will give rise to a in the order of 0.3 pC. This
charge production is converted into a voltage signal in a pre-amplifier which,
as the name suggest, also amplifies the signal. This amplification is done
close to the detector to avoid too much noise. The amplified signal has
56 CHAPTER 6. DATA ACQUSITION

a fast rise time and a long exponential tail and is sent into a secondary
combined amplifier and signal discriminator. The signal is further amplified
and bandwith limited, producing a signal of 1 µs − 2 µs1 . The discriminator
produces logic signal if the shaped signal passes a certain condition. We use
both constant fraction and simple threshold modules. For each channel a
shaped energy signal, a logical discriminator signal is produces. These are
sent to the digitization modules. Furthermore is a logical OR from all the
channels produced, and this is used by the trigger logic.

ADC
To measure the energy we use peak sensing Analog-to-Digital coverters.
They charge a capacator internally to capture the full peak, but then during
of this integration period is up to the experiment. This was done by the
AcqC that sent a logical GATE signal. After this periode the module converts
the value and stores in a buffer. During this time it asserts its BUSY to tell
the AcqC to wait. If the buffer gets full it can not and also asserts BUSY
and waits to be emptied. The GATE should be wide enough to capture the
peak of the signal. For charged particles in a silicon detector the width of
the shaped signals is a few µs, and the GATE was adjusted accordingly.
Both IS561A and IS561B was implemented with CAEN 7xx. They have
a 7.06 µs conversion time and 32 event buffer. In the IS561C they were
exchanged with MESYTEC ADC instead, since they had a faster conversion
time of 1.2 µs and significantly deeper buffers. Two different revisions were
used and the newer version could store 4096 events whereas the older version
could only store 1024 events in the buffer.

TDC
The leading edge from the discriminators is sent to Time-to-Digital converter
which timestamp the arrival time. For each channel it keeps the timestamp
from all leading edges it recieved within some time periode. When it

1
FiXme Fatal: Check det her
6.3. READOUT 57

recieves a trigger from the AcqC it saves all channels that fired within some
predefined time window.

6.3 Readout
The readout is responsible for reading the data from the modules, move
them over the network and save to it a physical disk. The DAQ was VME-
based in all cases meaning that all the modules that acquired data was
housed in a VME crate[ref] and the communication with the modules was
via the VME bus. To empty the modules in an orderly fashion, the readout
must communicate with the AcqC. The transport of the data as well as the
communication with AcqC is handled by available DAQ frameworks and
we used MBS[ref] for IS561A and drasi[ref] for IS561B and IS561C. They
require a piece of code from the user that have the following responsibilities

• Configure the modules at startup.

• Empty the module buffers when the AcqC requests it

• Verify the integrity of the data. If the data is corrupt, something is


most likely very wrong.

• Check module synchronizity by comparing the module event counters


with the AcqC.

These tasks were handled by the Nustar Readout Library [ref] and the
readout code ran on Single Board Computer (SBC), located in the VME-
crate. More specifically it was a CES RIO4-8072 [ref].

6.4 Multicrate
As the number of detectors increases, so does the required number of ADC
channels. At one point a single VME-crate can not house enough modules
and we have two use one more. This introduces the complexity of having
58 CHAPTER 6. DATA ACQUSITION

Slave
Gate
Readout trigger ADC
Slave
Readout Busy
AcqC
Deadtime TDC
Trigger
Event Data Readout trigger Deadtime
builder Master start Slave busy
Gate
Readout trigger ADC
Master
Readout Busy
AcqC
Deadtime TDC
Trigger
Master
Discriminators
Trigger requests

Figure 6.3: Figure by M.Munch

two crates, two SBC and two AcqC to commincate. The single crate design,
however, scales nicely and it is rather easy to extend to two crates running
in a master/slave configuration.
Both the master and the slave crate is more or less identical to the
single crate setup. Figure 6.3 shows an illustration of such a system. The
triggers logic is now only handled by the master. The master propagates
the generated MASTER START to the slave crate, much like if was just a
module. Moreover is the readout trigger propagated. The slave adheres to
this, however, the master must respect the busy and DT from the slave. If
it generated MASTER START in while the slave crate was busy, the modules
would get out of sync.
The readout in each crate does not know about the other, and an event
builder therefore collects the events and merge them. The events are tagged
with a crate number.
This multicrate design was used in IS561C.

6.5 Dead time ratio


Instead of the time it can be more useful to talk in terms of ratios. The life
time ratio L is the ratio of accepted triggers to the total incoming trigger
6.5. DEAD TIME RATIO 59

and the dead time ratio is the ratio of rejected triggers to the total incoming
triggers

Nacc
L= , (6.1)
Ntot
where Nacc is the accepted number of triggers and Ntot is the total incoming
number of triggers. These numbers were recorded by the AcqC for each
trigger input. In all three experiments this corresponded to one detector.
Throughout the analysis this is used to correct for dead time, by dividing
the actual count number in a given detector by L.
If a stream of Poisson distributed trigger the life time ratio can also be
evaluated analytically [ref]

1
L= , (6.2)
1 + fr ∆t
where fr is the request frequency and ∆t is the dead time per event. De-
pending on the operational mode of the readout, ∆t can be estimated.

Single cycle mode When the DAQ recieves a trigger it issues a MS. After
the GATE, the ADC’s converts the data and the readout empties the buffer
and check synchronicity. The main contributions to the total dead times is
the sum of the gate, the conversion time and the readout time. On top of
this, there is some readout overhead that adds up to a significant amount.

Multi event mode The readout overhead can be reduced by using the
internal buffers in the modules. If we wait until the buffers are full we can
effectively split the readout overhead across many events. The deeper the
buffers, the bigger the advantage. The DAQ ran in this mode in IS561A
and IS561B.

Shadow readout mode Depending on the experiment and the setup,


there may be times where the readout is completely idle. Instead of sitting
idle the readout could continously request data from the modules to keep
60 CHAPTER 6. DATA ACQUSITION

100 Shadow Single cycle


Livetime ratio, L [%]

Multi event

80

60

40

20

0
0 10 20 30 40 50 60 70
Trigger accept frequency, f a [kHz]
Figure 6.4: MMh

the buffers as empty as possible. If the events come in at a slower pace than
the readout can empty it from the buffer, a continous readout effectively
removes the dead time associated with readout, including the overhead.
This was implemented by M. Munch after the idea of H. T. Johansson and
was used in IS561C.
In particular the last scheme gives large improvements of L, and in
particular for large trigger frequencies. A more detailed investigation of this
was carried out and published[ref]. A simplified figure from the paper is
showed in Figure 6.4. The three different schemes are calculated with 6.2
for a crate with six CAEN V785s and a RIO4 SCB. The reader is referred
to the paper for the technical details. The data points are generated with a
poisson distributed pulser. The shadow reaout mode performs significantly
better, especially for large request frequencies. Close to 55 kHz the curve
drops - this is data bandwith limit. The DAQ can simply not move data
faster. For a typical ISOLDE reaction experiment the rate is typically much
lower. At 20 kHz the accepted trigger rate is increased by roughly 30 %,
which is not an unrealistic rate in a down stream detector where Rutherford
scattering dominates.
6.6. PULSED BEAM 61

105

104
Counts/1µs

103

102 12.1 µs

101

100
0
0 50 100 150 200 250 300
TEBIS [µs]

Figure 6.5: TEBIS spectrum from a down stream detector in IS561B. The
time period between the peaks corresponds to the effective DT of the DAQ.
The main contribution is the GATE and the conversion time.

6.6 Pulsed beam


The major source of DT was the combined GATE and conversion time. With
CAEN ADC’s this adds up to ≈ 10 µs. The challenge with a pulsed beam is,
however, that the majority of our events comes in a rather short timespan.
As discussed in Section 3.1 the EBIS releases the majority of it’s events in
the first 100 µs. The GATE and conversion thus take 10% of the time. This
leads to a significant ∆t. At high rates the DT increases due to conversion.
Figure 6.5 shows an EBIS spectrum from a measurement with a high rate
12 C beam, and it has very clear dips with a spacing corresponding to exactly

the effective conversion time.


The effect was only so extreme for detectors very close to the beam
axis. Since the DAQ is synchronous, however, the dead time is the same.
The number of events in the other detectors, however, may be much lower,
effectively lowering the dead time ratio. During the experiment we monitored
L regularly.
For IS561C the CAEN ADC’s were exchanged with MESYTEC ADC’s
62 CHAPTER 6. DATA ACQUSITION

to reduce the conversion time from 7 µs to 1.6 µs. The deeper buffers also
made sure the DAQ could accept a full EBIS release without having to be
emptied. To be sure that the modules were mostly empty at the begining
of an EBIS release, a readout trigger was produced at a fixed time interval
before the EBIS pulse. With a 30 Hz release cycle there were approximately
30 ms between each release which is plenty of time to empty the buffers.
With the lower intensity 9 Li beam there were no major issues with the
dead time. In the detectors of interest we had a dead time ratio in the order
of 5 %-10 % in IS561A and it was reduced to 1 %-2 % in IS561C.

6.7 Back end


When the data has been extracted from the modules, it is sent across the
network to be stored on a physical disk (file taking) as well as online analysis.
The file taking is naturally the most essential feature. The online analysis is
very important to continously monitor the different detectors, but it is not
essential to be up at all times. There might be other services that monitors
the data in real time which are not essential.
The majority of the backend services runs on a desktop computer that
is connected to the same network as the readout node. To decouple these
services, there is a relay service between the readout and the consumption
of the data.

Relay
The relay is an ucesb instance that connects to the readout data stream
from and fan the data out to two streams. One data stream is in "weak-hold"
mode, which means that if a client can not keep up with the data it blocks
the acqusition. This is used by the file taking.
The second stream is in "no-hold"-mode and does not require the clients
to consume all data. This is used for the online analysis.
The relay have the additional feature that it validates the data on the
fly. This means we do not have to save data and unpack it to detect corrupt
data.
6.8. DAQC 63

File taking
The file taking is another ucesb instance that connects to the relay and
writes the data to a file. It saves the data in the GSI list mode (.lmd)
format [ref]. It is furthermore compressed with gzip.
To ease the subsequent analysis the files are written in chunks of typically
100 MB. In case of a failure it also makes recovery easier, since all the data
from a measurement are not in the possibly corrupted file.

Online analysis
The online analysis presents the incoming data to the operators. This is
handled by the go4cesb project, which combines ucesb, GO4 and AUSAlib.
GO4 provides an advanced user interface that lets the operator interact with
different spectra produced by the analysis routines. The data is piped from
the relay to a ucesb process which performs the unpacking of the data.
This process is spawned by go4cesb and the unpacked data is subsequently
piped back into GO4 as an event stream.
This design has two main advantages. The unpacker only has to be
written once for both offline analysis and online analysis. Furthermore does
the unpacked data have the same AUSAlib-friendly format, so it is trivial
to apply AUSAlib analysis routines in the online analysis.
The simplest analysis includes mapping the ADC channels to detector
strips, applying calibrations and matching. Since this is built in to AUSAlib,
this can automatically be applied to the unpacked data. More sophisticated
analysis such as ∆E − E spectra and kinematic curves can now be imple-
mented by the user, but since everything is in the AUSAlib format, it is
easy to reuse such analysis components acrosse different experiments. Such
analysis components can be turned on and off in a simple configuration file.

6.8 DaqC
To have a fully functional DAQ all the described components must be running
and continously monitored. Since a typical experiment spans several days,
64 CHAPTER 6. DATA ACQUSITION

C
A

D
B
Figure 6.6: DAQC

non-experts should also be able to start, operate and even recover the DAQ.
To ease this, we have developed DaqC.
Internally it consists of interfaces start, stop and monitor all the dif-
ferent services. It implements the correct startup order and keeps track
of dependencies between the different services. When started the startup
sequence is executed. When all the services is up and running, the internal
watchdog monitors each service to check that is up and healthy. If a service
crashes it tries to revive it.
When running, an operator can interact via a simple web interface,
as shown in Figure 6.6. It provides controls to all services managed by
DAQC as well as simple status report. Panel (A) shows the status shown
6.9. RUNDB 65

for the different componenets. In panel (B) is a control panel for the trigger
logic. The operator can quickly include or exclude detectors in the global
trigger between measurements. In (C) is a simple overview of previous
measurements shown.
Panel (D) provide a button to start measurements, i.e. start the file
taking. When a new measurement is started, the operator is prompted with
a form to fill in a run sheet with information about target, beam, operators
on shift, etc. This is saved in log file, and effectively compiles a detailed
measurement log. Furthermore, when a measurement is started and stopped,
DaqC dumps the configuration of all the ADCs, TDCs, amplifiers as well as
other available meta data. This is saved together with the data files. Both
the data and the log files are continuously backed up to a remote storage
facility in case of hardware failure.
I started the development of DaqC together with M. Munch after IS561A.
It has been used for IS561B, IS561C and multiple other ISOLDE experiments
and it is the control program for the DAQ at the 5 MeV accelerator at Aarhus
University.

6.9 RunDB
The information from the run sheet is essential for the subsequent analysis of
the data, both during the experiment and after the experiment. Traditionally
this has been written by hand in a paper log by the operators, when a
measurement has been started and stopped. With the implementation of
DaqC, it was natural to compile this data in a database. This led to the
development of RunDB, which stores the meta data and file location of the
data in a SQL database.
RunDB thus provides an easy way of browsing through the different
measurement configuration and in particular the time spent. The automated
upload of the run sheet has moreover reduced the number of errors in the
log such as wrong time stamps.
A web interface makes it easy to search for specific targets, time periods
etc. It is shown in Figure 6.7.
66 CHAPTER 6. DATA ACQUSITION

Figure 6.7: RunDB

6.10 Real time monitoring


During an experiment the running conditions will change. The beam
intensity will fluctuate, a detector may pick up some electronic noise or
the accelerator may experience a short failure, resulting in no beam at all.
While the online analysis is essential to monitor the physics going on in the
detector, it does not provide a good overview of the conditions.
The typical way of dealing with this, has been to allocate one or two
shift people to continously perform yield estimates on the incoming data.
This process can, however, largely be automated.
We have implemented this using two open-source projects called In-
fluxDB[] and Grafana[]. InfluxDB is a time-series database and is thus
excellent at storing data with a timestamp. During an experiment we feed
all kinds of data to this database, such as trigger rates, dead time ratio,
leakage current from the detectors, pressure in the vacuum chamber and
6.10. REAL TIME MONITORING 67

Figure 6.8: Grafana

even results from simple automated analysis-scripts. This can be done on a


second-timescale.
In particular the trigger rates can be usefull to detect trends in the
overall yield. If the beam drops for some reason, the trigger rates will
also drop to zero, and this can easily be used for automated alerts for the
operator.
Grafana is a simple web frontend that can display the content of the
InfluxDB and provide the operators with all the data in an structured way.
An example is shown in Figure 6.8. On the left panel is the rate of two
detectors at small angles and in the right panel is the trigger rate of two
detectors at large angles. We notice how the trigger rate of only one detector
rose over the course of several hours. It turned out it was picking up noise
and was properly grounded before it became a problem.
7

Acceptance

In this chapter we will discuss how to do a proper treatment of the acceptance


of the setups. When we do an experiment we count reactions. The counts
does, however, not only reflect the reaction probablity, but also the target
properties, the geometric coverage of our detectors, the kinematics and
more. This makes it difficult to compare to other experiments with different
setups and in particular to theoretical calculations that does not care about
the geometrical particularities. Instead we need to transform the counting
number to cross sections.
In principle we can explore a cross section as a function of many variables,
but some makes more sense than other. We will thus limit this discussion
to differential angular cross sections and differential energy cross section, or
more generally the double differential cross section.
For a given reaction, we can calculate the expected numbers of detected
particles from the total cross section

NR = NI nt σ, (7.1)

where NR is the number of reaction products, NI is the number of


incident beam particles, nt is the number of target particles per cm2 and σ
is the integrated cross section for the reaction.
We can also express this as a differential cross section

69
70 CHAPTER 7. ACCEPTANCE

d2 Nr (θ, φ, E) d2 σ(θ, φ, E)
= NI nt , (7.2)
dΩdE dΩdE
and the total number of detected particles can be calculated by integrating
over the solid angle covered by our detectors, ΩS , and the energy range that
is allowed by the kinematics of our experiment Er

d2 σ(θ, φ, E) 0 0
Z Z
NR = NI nt dΩ dE . (7.3)
Er ΩS dΩdE
Since we want to calculate the cross section we rearrange 7.2 to get

d2 σ(θ, φ, E) 1 d2 Nr (θ, φ, E)
= . (7.4)
dΩdE NI nt dΩdE
In an actual experiment we can not measure infitesimal range in angle
and energy, and we therefore bin our data into a suitable bin size, usually
depending on our amount of statistics. Choosing too small bins, and we get
large uncertainties. Choosing too big, we average out possible interesting
features. In either way, we can approximate the cross section for each point
in a given bin by the average

d2 σ(θ, φ, E) 1 Nr (θ, φ, E)
≈ , (7.5)
dΩdE NI nt ∆Ω∆E
where Nre is the detected numbers in a given bin, ∆Ω is the solid angle
corresponding to the bin and ∆E is the energy range covered by the bin.
In reality we do not have full coverage of all the bins. Some bins will be
outside of detectors and some will suppressed due to kinematics. This will
alter ∆Ω and ∆E corresponding to a particular Nr , but proportionally. We
can thus write

d2 σ(θ, φ, E) 1 Nrd (θ, φ, E)


≈ , (7.6)
dΩdE ¯ ∆E
NI nt ∆Ω ¯
where Nrd (θ, φ, E) is the actual detected number of particles in a given bin
¯ ∆E
and ∆Ω ¯ is the effective size of the bin.
7.1. MC 71

7.1 MC
RAMBLE ABOUT WHY IT IS DIFFICULT
To get a reliable estimate of the cross section, each of the values on the
right side must be evaluated. NI and nt does not depend of the geometric
setup of our detectors or the reference frame we choose to investigate. Nre ,
∆Ω and ∆E are, however, more difficult. First of all these depends on the
geometry of the setup. The particular geometry of the detectors dictates
how large a solid angle it covers. This may not be the same in all angles,
and we thus have to take an angular dependence of the solid angle into
account. Since the detected particles scales with the solid angle, so does
Nre . If we try to do our analysis in the CM-frame the solid angles need to
be transformed correspondingly.
Moreover, the effects of a broad divergent beam, finite angular resolution
in the detectors and the like, will change the effective solid angle. These
effects are not easy to incorporate in an analytic way, however, we can solve
it with a Monte Carlo simulation.
Consider a simulation with a uniform distribution. We should see a
constant fraction of the the simulated particles in each bin, proportional to
the size of the bin
Ns (θ, E) ∆Ω∆E
= , (7.7)
Ns 4πEr
where Ns (θ, E) is the number of simulated particle in a specific bin, Ns
is the total number of simulated particles and Er is the energy range we
sample uniformly from.
If we also include our setup in the simulation and keep track of how
many particles we detect, we will instead get a measure of the effective bin
size
Nsd (θ, E) ¯ ∆E
∆Ω ¯
= , (7.8)
Ns 4πEr
where Nsd (θ, E) is now the detected amount of a the simulated reactions
and ∆Ω¯ ∆E
¯ is the effective bin size. It is, however, the same ∆Ω
¯ ∆E
¯ as in
¯ ¯
7.6. This means we can isolate ∆Ω∆E in 7.8 and substitute into 7.6 and
get
72 CHAPTER 7. ACCEPTANCE

d2 σ(θ, φ, E) 1 Ns Nrd (θ, φ, E) 1


≈ . (7.9)
dΩdE nt NI Nsd (θ, φ, E) 4πEr
Based on this we can get the projected single differential cross sections. In
this case we need to integrate over the other value. For angular distributions
we get

dσ(θ, φ) 1 Ns Nrd (θ, φ, E) 1


Z
≈ dE 0
dΩ E nt NI Nsd (θ, φ, E) 4πEr
1 Ns X Nrd (θ, φ, E) ∆E
!
= , (7.10)
nt NI i Nsd (θ, φ, E) 4πEr

where the sum is over all energy bins and we have assumed that each bin
have the same width ∆E.
Similarily for the single differential energy cross section we get

1 Ns X Nrd (θ, φ, E) 1
!
dσ(E)
≈ ∆Ω , (7.11)
dE nt NI i Ns (θ, φ, E)
d 4πEr

where ∆Ω is the solid angle of the bin.


Notice an important difference here. For the angular distribution we
could sum up all energies at a given angle, since each bin has the same weight
(∆E) and we can pull it outside the sum. For the energy distribution we
need to apply the weight (∆Ω), consequently also scaling the uncertainties.

Bound states
We can do a similar analysis to get the single differential angular cross section
of bound states. In this case we simulate a single energy, corresponding to
the resonance energy, and the energy dependence ∆E/Er in 7.8 vanishes.
This gives us

dσ(θ, φ) 1 Ns Nrd (θ, φ) 1


≈ . (7.12)
dΩ nt NI Nsd (θ, φ) 4π
7.2. RUTHERFORD ESTIMATION 73

7.2 Rutherford estimation


We can use the same ideas to estimate the intensity in an experiment based
on Rutherford scattering.
Consider now an experiment with a beam impinging on a heavy target
and we measure the scattered beam particles. Not all particles will react
with a target particle, and the distribution follows the well known Rutherford
cross section in CM

!2
dσR e2 zZ 1
(θ, φ) =
dΩ 4π0 E sin (θ/2)
2

α~czZ 1 2
 
=
4E sin2 (θ/2)
1.44MeV · fm
2
zZ

= (7.13)
4E sin2 (θ/2)

where z is the charge of the beam, Z is the charge of the target and E is
the available energy in the CM system
mtarget
E = E LAB (7.14)
mtarget + mbeam

The measured number of beam particles is thus

dσR
Z
NR = NI nt (θ, φ)dΩ0 (7.15)
Ωsetup dΩ

where NI is the total number of incoming beam, nt is the target density.


Notice we have integrated the Rutherford cross section over Ωsetup which
is the solid angle covered by our setup. Since we are interested in the
intensity, and therefore the total number of incoming beam particles we can
rewrite it to
NR
NI = R (7.16)
dΩ (θ, φ)dΩ
nt Ωsetup dσ R 0
74 CHAPTER 7. ACCEPTANCE

We can now use two different strategies to determine Ωsetup . The first
method is a MC methods like the one we used for the acceptance. This
method makes it easy to incorporate imperfect beams, etc. The second is a
simpler geometrical method, which is good to consistency check, but lacks
the sophistication we can achieve with a simulation.

Monte Carlo
Imagine we simulate Ns reactions. We distribute them accordingly to the
Rutherford formula in CM with the MC method and count how many we
detect, Nsd . Notice that it diverges when θ → 0, hence we only simulate in
the range [θmin , 180].
The ratio of detected scattered beam particles (analogous to 7.8) to
simulated particles is

dΩ (θ, φ)dΩ Ωsetup dΩ (θ, φ)dΩ


R dσR 0
R dσR 0
Nsd Ω
= R setup = . (7.17)
dσR
(θ, φ)dΩ0
R θmin dσR
Ns θ>θ min dΩ
2π 180 dΩ (θ, φ)sinθdθ 0

We can rearrange to

Nd
Z θmin
dσR dσR
Z
(θ, φ)dΩ0 = s 2π (θ, φ)sinθdθ0 (7.18)
Ωsetup dΩ Ns 180 dΩ

and plug it back into 7.16 to get

Nr Ns 1
NI = (7.19)
2πnt Nsd 180
R θmin
dΩ (θ, φ)sinθdθ
dσR 0

Geometric
We can also calculate the solid angle of the individual pixels with a simple
geometrical approximation where we project the pixel onto the tangent
plane of the unit sphere and calculate the area on this plane. If the pixels
are small enough, this is a sufficiently good approximation.
7.2. RUTHERFORD ESTIMATION 75

We can calculate the differential Rutherford cross section in the LAB


frame and simply multiply with the solid angle of the pixel. The Rutherford
cross section in LAB is given as

2 2
r !

LAB 2 cos θ + 1− m1
m2 sin θ
dσR zZα~c

(θ, φ) = ·
2E
r
dΩ  2
sin4 θ 1− m1
m2 sin θ

We thus get
Z
dσR X dσ LAB
(θ, φ)dΩ0 ≈ R
(θpixel )Ωpixel (7.20)
Ωsetup dΩ pixels
dΩ

Plugging back into 7.16 we get


Nr
NI = LAB (7.21)
dσR
pixels dΩ (θpixel )Ωpixel
P
nt
8

Experimental setups

77
9

IS367

[!!!Probably something about the significance at this energy...?!!!]


The first experiment in this series was carried out in 2004 and the results
from that experiment was published in a series of papers by Jeppesen. Based
on the experiences, it was followed up by another experiment in 2005 with
an improved setup, mainly manifested as twice the detector coverage. In
this chapter I will discuss the analysis of these data. I did not take part in
the actual experiment, but have analyzed the data based on the logbook
from the experiment.
Parts of the analysis was also part of my progress report, but will also
be included here for completeness.
The experiment was carried out at ISOLDE, using the REX post accel-
erator, the predecessor to HIE-ISOLDE with a beam energy of 2.68 MeV/A.
The primary target was a deuterated polyethelene target. The primary
background contributions were measured with a pure carbon target and a
regular polyethylene target.
To estimate the intensity we can unfortunately not rely on direct current
measurements. This is mainly due to the low intensities of our beam and
the short release time of the EBIS. Instead we can use a target that is heavy
enough to keep the impinging beam particles below the Coulomb barrier
and scatter according to the Rutherford formula. From the counted number
of scattered beam particles we can infer the number of total incident beam
particles using the methods described in 7.2.

79
80 CHAPTER 9. IS367

An overview of the specifications of the targets and their purpose can


be seen in Table 9.1.

Table 9.1: The different target used in IS367.

Target Thickness Usage


CD2 15 µm The main target.
CD2 7.5 µm A thinner main target.
CH2 15 µm Background.
12 C 2 mg/cm2 Background.
9 Be 1.95 mg/cm2 Not considered.
107 Ag 1.1 mg/cm2 Yield estimates.

The experimental setup is shown in Figure 9.1. It consists of two 32x32


DSSD’s with a strip pitch of 2.0 mm and a thickness of 60 µm. Each of these
was backed an unsegmented silicon detector with a thickness of 1.5 mm.
This setup enables us to do particle identification based on the correlation
between the energy loss in the DSSD and the Pad detector. The pad is
thick enough to stop the light particles so we measure the full energy.
Both detectors were facing the target, but we only covered the LAB
angles below 90◦ . Due to the kinematic compression discussed in Section
2.2, the energy of the ejectiles that will reach the detectors in the backward
angles would be too low to be detected properly.

9.1 Data analysis


Before we can start to understand the data, we need to perform an energy
calibration of the detectors. With a good energy calibration, we can start
identifying the different particles, based on the correlation between the
energy deposited in the front DSSD and the back Pad. Once we have
identified the different particle types, we need to do a background subtraction
before we can start to understand the physics of the different reaction
channels we observe.
9.1. DATA ANALYSIS 81

12000

10000

8000

∆E · cos θ [MeV]
6000

4000

2000

0
0 5000 10000 15000 20000 25000 30000 35000
E + ∆E · (1-cos θ) [MeV]

Figure 9.1: Left:Schematic representation of the setup from IS367. The


setup consists of two telescope configurations covering the forward direction
in the laboratory. Right: ∆E-E plot for both detectors.

Energy calibration
The calibration of the detecters was done with a standard α source, as
described in Section ??. Each segment was calibrated individually and the
pads were calibrated with a source from the back side.
Since the energy of the particles were that stops in the pad, are up to
4 times higher than the highest calibration point, the calibrations must
be consistency checked at higher energies. Luckily, as we will see in the
following analysis, the kinematics and the ∆E − E spectra are consistent,
so we accept the energy calibrations.

Particle identification
Using the methods described in ??, we can identify the different particles
detected. The scattered 9 Li form both the heavy target and the primary
target does not have enough energy to penetrate the ∆E-detector. We can
82 CHAPTER 9. IS367

thus only identify these based on kinematics, which we will discuss as part
of the kinematic analysis.
In the right panel in Figure 9.1 we see the ∆E − E plot for the physics
runs. We clearly see protons, deuterons, tritons and α. There is also a small
contribution of 6 He. With proper gates, we can pick out the individual
particles.

Kinematics
The kinematic curves for protons, deuterons and tritons are show in the
upper panel in Figure 9.2. For the protons we see a clear component
from elastic scattering of protons. This is expected, since the target have
≈ 1 % protons. Since we do not know the exact amount, we instead did
a measurement of CH2 which we can subtract. Also notice that is well
seperated from the continuum expected from 9 Li (d,p)10 Li reactions.
The deuteron spectra have two well defined components, correspond to
the ground state and 1st excited state respectively.
The tritons have components from the three first excited states. At
larger excitation energies we exceed the neturon seperation energy and we
see a continuum. There a various channels that may contribute to this
continuum, but this has not been investigated in detail. The corresponding
excitation spectra, calculated with 2.8 are shown in the lower panel in Figure
9.2.
The width of the 9 Li ground state peak can be used to estimate the
resolution of our setup. A simple gaussian fit gives a resolution of FWHM=
551(4) keV.

Heavy particles For the intensity measurements we scatter 9 Li off of


silver and thus measure 9 Li which stops completely in the DSSD, and we
can not identify them using ∆E − E. Instead we can gate on them simply
using kinematics. In Figure ?? we see the ∆E energy as a function of angle
with the expected kinematic curve for 9 Li scattering. The data follows the
kinematics nicely, and we can use the kinematic region as the gate. Notice
9.1. DATA ANALYSIS 83

Protons Deuterons Tritons


25
Li9(p,p)Li9 Li9(d,d)Li9 Li9(d,t)Li8
Li9(d,p)Li10 (0.38MeV) Li9(d,d)Li9 (2.7MeV) Li9(d,t)Li8 (0.98MeV)
20 Li9(d,t)Li8 (2.26MeV)
E [MeV]

15

10

0
0 10 20 30 40 50 60 70 80
0 10 20 30 40 50 60 70 80
0 10 20 30 40 50 60 70 80
θ LAB [deg]

800
900
700
800

200 700 600


Counts/20 keV

600 500
500
400
400
100 300
300
200
200
100 100

0 0 0
-4 -2 0 2 -4 -2 0 2 4 -4 -2 0 2 4 6
Ex [MeV]

Figure 9.2: Upper panel: Kinematic curves for identified protons,


deuterons and tritons. Kinematic curves for reactions with different exci-
tation energies of the recoil are also shown. Lower panel: Excitations
spectra for the different particles, assuming 9 Li (d,p)10 Li, 9 Li (d,d)9 Li and
9 Li (d,t)10 Li respectively.
84 CHAPTER 9. IS367

25

25000
d
9 Li
20

20000

15
E [MeV]

15000

E [MeV]
10
10000

5
5000

0 0
0 10 20 30 40 50 60 70 80 90 0 20 40 60 80
θCM [deg] θCM

Figure 9.3: Left: Kinematic curve of 9 Li scattering off of silver. Right:


Coincident deuterons and 9 Li with the kinematic curves corresponding to
the ground state of 9 Li.

the feature below the main curve. This is an artifact of a faulty strip in one
of the detectors.
We also measure some amount of coincident recoils from the light par-
ticles. An example is shown in the left panel in Figure 9.3. The spectra
here fits the curves nicely, and they are very clean when the coincidence
requirements have been made. Once again, we can use the kinematic region
as a gate.

Background subtraction
In the 10 Li spectrum we see a peak just below 0 followed by large tail at
negative excitation energies. These events are unphysical in the sense, that
they can not origin from the 9 Li (d,p)10 Li reaction. They must instead come
from various background contributions. The peak below 0 corresponds to 9 Li
(p,p)9 Li as we saw in the kinematic curve. Since we did a measurement with
both CH2 and C we can pipe these measurements through the same analysis
9.2. INTENSITY ESTIMATION 85

(A) (B) (C)


CD2 CD2
CH2 250
Background
500
C
200
Counts/20 keV

400 200 150

300 100

50
200 100
0
100
−50
0 0
-4 -2 0 2 -4 -2 0 2 -4 -2 0 2
Ex [MeV]

Figure 9.4: A: Excitations spectra for protons for the main target (CD2 )
and the two background targets (CH2 and C). B: The combined background
on top of the signal. C: The resulting background subtracted spectrum.

to get an understanding of the background. The different components are


shown Figure 9.4 (A). We can now do a simple superposition fit of the two
background components to the negative excitation energies. We also want
to describe the elastic peak, so we fit the left edge and down to −4 MeV to
a sum of the two background components and we get

BG = 0.25(4) · CHx + 0.35(1) · C. (9.1)


Since the conditions have been the same independent of which particles
we detect, we can use the same fit to background subtract deuterons and
tritons.

9.2 Intensity estimation


In the following discussion we will refer to a measurement with such a heavy
target as a intensity measurement and we will refer to a measurement with
the deuterated target as a physics measurement.
During the experiment, several intensity measurements were made,
sandwiched in between two physics measurements. Such three measurements
86 CHAPTER 9. IS367

can be used to get an estimate of the intensity in terms of scatter particles in


the physics measurements and thus in principle enable us to extrapolate the
total incoming beam particles during the full set of physics measurements.
In practice, however, we need to do multiple of these due to systematic
effects which will be discussed in the following.
We denote the intensity of beam particles in the intensity measurement,
as calculated by the methods in Section 7.2, by I. The intensity across all
sandwiches in the experiment can be seen in Figure 9.5. We notice that the
intensity fluctuates. However, given constant beam properties, the number
of scattered particles in any given reaction channel should scale accordingly.
For instance will the number scattered deuterons scale with the number of
incoming beam particles. This means that if we know how many deuterons
we see per incoming beam particle, we can in principle count deuterons in a
given measurement and calculate the corresponding intensity.
To calculate this ratio, we use the sandwiched measurements. From the
intensity measurement we get I, and from the adjecent physics measure-
ments we can get the rate of scattered deuterons, Ri , where i indicates the
measurement. The ratio is thus

Nd Rd
Nd = = . (9.2)
N9 Li I

In the above ratio, we have assumed that the intensity did not change
during two adjecent measurements, which is generally not correct. In
particular will the super cycle at CERN determine how many proton bunches
that reaches ISOLDE. This changes on a timescale that is comparable to
the length of the runs. However, by looking at the distribution across all
intensity measurements, we may be able to treat this systematic error as a
statistical error.
In Figure 9.6 N is plotted as a function of measurement number. The
upper panel is the deuterons and the lower panel is for α. The median as
well as the 1σ interquartile range is shown. We use the median, since it is
more robust to outliers. In practice, however, the median turned out to be
very close to the weighted average.
9.2. INTENSITY ESTIMATION 87

The distribution does not seem to have a clear systematic trend. We can
make a simple χ2 test, to test is they are consistent with being from the same
mean with a gaussian distribution. For deuterons we get a pval = 0.048.
This is on edge of being consistent with coming from the same mean, but
due to the expected systematic errors we accept it, with a correspondingly
larger uncertainty on the integrated intensity.
The medians for the two particle types are respectively

Nd = 8.50 · 10−6 ± 6.67%


Nα = 3.08 · 10−5 ± 7.90%

Using the ratio we can calculate the integrated beam across all the
measurements, by simply counting scattered particles.

N9 Li = Np Nip , (9.3)
X

where p can be either protons, deuterons, tritons or α. For deuterons and α


we get

N9dLi = 3.97 · 109 ± 6.7%


N9αLi = 4.07 · 109 ± 7.9%,

which is consistent within the statistical uncertainties. We can do the


same for both tritons and protons, and they are all consistent within the
uncertainties, so we do a weighted average of all four

N9dLi = 4.04 · 109 ± 1.2%


88 CHAPTER 9. IS367

40
Beam intensity [kHz]

35

30

25

20

120 140 160 180 200 220 240


Measurement number

Figure 9.5: Intensity measurements throughout the experiment. Notice the


increase in intensity after measurement 180, due to a reoptimization of the
beam by the operators.
9.2. INTENSITY ESTIMATION 89

0.000010

0.000008
N

0.000006

120 140 160 180 200 220


Measurement nr

0.000035
N

0.000030

120 140 160 180 200 220


Measurement nr

Figure 9.6: Particle per incident 9 Li for d and α respectively.


10

IS561A

The second experiment in this series was carried out in 2016 and this was
the first experiment I took part in. This was the first experiment with the
HIE-ISOLDE upgrade. The full upgrade was not completed at this time, but
we were still able to get a significantly higher beam energy of 6.72 MeV/A.
This was requested to be able to measure the protons in the backward
directions, which has quite low energy due to the kinematic compression.
With 6.72 MeV/A the protons should reach 0.5 MeV-1.0 MeV, which should
be possible to measure. Therefore the setup was redesigned and in particular
included a telescope to cover the backward directions.
During the run we encountered problems both with beam alignment
and low yields from the ISOLDE part. This resulted in significantly lower
statistics than expected. The beam alignment problems also led us to steer
the beam around during the experiment, which in turn requires a more
careful analysis.
The experimental program for this experiment was very similar to the
older IS367. We had a deuterated polyethylene target as the main target.
For background measurements we used a pure carbon target and a normal
polyethylene target. For intensity measurements we used a gold target. A
table of the targets is shown in Table 10.1.
The operators used 12 C as a pilot beam to steer the beam into the
chamber. We were allowed to take some data with this beam, which turned
out to be very useful since the intensity is rather large, which means we can

91
92 CHAPTER 10. IS561A

quickly get a significant amount of statistics. By doing reactions with this


beam we could get good constaints on the geometry and beam properties,
and we will discuss later.

Table 10.1: The different target used in IS561A.

Target Thickness Usage


CD2 8 µm The main target.
CH2 6 µm Background.
12 C 2 mg/cm2 Background.
197 Au 0.1 µm Yield estimates. Backed by 3.5 µm mylar

10.1 Setup
The experimental setup for IS561A was updated compared to IS367. A
picture is shown in ??. The setup had detector coverage in both forward
and backward directions. In the forward direction (1) there was 5 telescopes
placed symmetrically around the beam axis in a pentagon structure. Each
of these were 60 µm thick and had 16x16 strips. Perpendicular on the beam
axis were a round DSSD with a thickness of 1 mm. The idea with this design
was inspired by the method discussed in Section ??. The thick DSSD should
capture the recoils and the 5 pentagon detectors would capture the light
ejectiles. With the high degree of symmetry we should be able to capture
the beam spot and beam direction. These detectors were mounted in a
3D-printed plastic structure, which provided good geometrical constraints
on the detector positions.
In the backward direction was a single telescope place (3). The DSSD
was 60 µm. The purpose of this detector was to measure the ejectiles in
the backward lab direction, which corresponds to the forward CM angles
due to the inverse kinematic. We expect there to be larger cross sections
in this area, and we should be able to get a better resolution due to the
higher angular resolution. Due to the kinematic compression the energy of
10.2. CALIBRATION AND BEAM PROPERTIES 93

10000

8000

∆E · cos θ [MeV]
6000

4000

2000

0
0 5000 10000 15000 20000 25000 30000 35000 40000 45000
E + ∆E · (1-cos θ) [MeV]

Figure 10.1: Left: Setup Right: Banana plot

the ejectiles from the primary channels will be 500 keV-1000 keV, and will
thus not penetrate the DSSD in the backward telescope.
The targets was places on a moterized wheel (2) which could be controlled
from outside of the vacuum chamber. This means we could change the
target easily without having to open up the vacuum chamber.

10.2 Calibration and beam properties


The calibration of the five DSSD’s in the pentagon and the backward DSSD
was done with an α source as described in Section ??. An initial calibration
for the pads were done in the same way. It, however, became clear that
these were not precise enough at higher energies. One pad, in particular,
had undergone a change between the calibration and the physics runs,
and the calibration could thus not be trusted. Besides the calibrations, it
was also clear that the beam axis was angled with respect with the setup.
Unfortunately we had too few coincidences to use the method described in
Section ??. Instead we can fit the kinematic curves of a well know reaction
94 CHAPTER 10. IS561A

X
theta_h pd
d
theta

Z pbs

to determine the beam spot and the beam direction. There is five parameters
we need to determine

p~bs = (x, y, z)
dˆ = (θ, φ),

where p~ is the position where the beam hits the target in a global coordinate
system and dˆ is unit vector in the direction of motion of the beam. See Figure
4.14. When we measure a particle in a detector and want to determine
the kinematics of the reaction it originated from, we need to use the angle
relative to the beam axis

pd − p~bs ) · dˆ
(~
cos θ = , (10.1)
||~
pd − p~bs ||
where p~ is the vector to the detection position.
An initial guess of p~bs = (0, 0, 0) and dˆ = ẑ is good enough for us
to identify the elastic scattered deuterons from the 12 C(d,d)12 C. We use
the 12 C beam since the intensity was sevel orders of magnitude larger. In
Figure ?? (A) we can see the initial guess and a kinematic curve. The fit
10.2. CALIBRATION AND BEAM PROPERTIES 95

of the curve is not particulary good, but we can clearly identify the elastic
component. We can now calculate the a χ2 value for each point, based on a
given set of beam parameters

X (Ei − Td (θi ))2


χ2 = , (10.2)
i
σ2

where Ei is the total measured energy in the telescope (corrected for dead
layers), θi is the angle calculated with 10.1 and Td (θi ) is the kinematic curve
from 2.9. We can now minimize this function to get an estimate of the
beam parameters. Notice that this methods does not account for divergence,
spatial distribution or energy distribution. It turned out, however, that
we are limited by the detector resolution, unless the beam is unphysically
large. This resolution also means that the uncertainties on the estimated
parameters have a similar magnitude.
Another crucial part is the symmetrical coverage from the pentagon.
With just one telescope, the orthogonal direction were too unconstraint,
but having five telescopes covering most of φ gave nice constraints on the
parameters.
The result is shown in Figure ?? (B), with one important extension. The
pad calibrations were also adjusted, since they did not extrapolate nicely to
large energies. This imposes an issue, since Ei depends on the pad energy.
We can, however, also adjust the pad calibrations based on reactions.
Consider again the elastic scattering of deuterons and 12 C. The scattered
deuterons penetrate the DSSD, and the energy deposited depends on the
distance traveled in the DSSD, which in turn dependes on the geometry of
the detector and the angle of the deuteron, θ̂.
We can, however, determine what the initial energy of a deuteron would
have been, given the energy deposited in the DSSD based on energy loss
tabulations.
We can calculate the effective distance it travels in the detector in the
same way as in Section ??, teff (θ̂) which depends on the angle with the
detector θ̂. We can then numerically find Einit , the energy that corresponds
to the energy we measure in the DSSD
96 CHAPTER 10. IS561A

Z teff (θ̂)
dE
∆E(θ̂) = (x) (10.3)

dx.
0 dx Einit

The expected energy in the pad is now what is left

Eexp = Einit (θ̂) − ∆E(θ̂) − ∆Edeadlayer , (10.4)

where ∆Edeadlayer is the various loses in dead layers.


For each detected deuteron we can now compare with the ADC channel,
and fit a linear relationship. We can of course do this for all the different
particles types. In Figure ?? we can see the linear linear relationship.
There is one issue with these two procedures, that unfortunately make
them dependent on each other. In order to do the pad calibration, we need
to know θ̂ which depends on p~bs . To get p~bs we do the kinematic fit, which
relies on the pad calibrations. We are caught in a vicious cycle. By iterating
the two methods, it was possible to converge, even though it required some
hand tuning.
The resulting beam spot and direction was determined to be

p~bs = (6.0(2) mm, −3.1(3) mm, 4.3(2) mm)


dˆ = (3.2(6)◦ , −135(1)◦ )

10.3 Data analysis


With calibrations, beam spot and beam direction determined, we can
evaluate the data. Due to the suboptimal run conditions, is the amount of
statistics rather limited. If we include all the good runs, we get the ∆E − E
spectrum shown in Figure 10.1 (B), in the pentagon detectors. As in IS367
we see protons, deuterons, tritons and α. With appropriate gates on the
different particle types we get the excitations spectra shown in Figure 10.2.
10.3. DATA ANALYSIS 97

CD2 CD2 CD2


C 80 C C
40
80 CH2 (sim)
70
Counts/200 keV

60
60 30
50

40 20
40
30

20 20 10
10

0 0 0
-6 -4 -2 0 2 4 6 8 -6 -4 -2 0 2 4 6 8 -6 -4 -2 0 2 4 6 8
Ex [MeV]

Figure 10.2: Erik

Forward telescopes
The background subtraction in this experiment turned out to be more
difficult, due to the low statistics. In all three plots are the background
from the carbon target is shown. The statistics from CH2 was unfortunately
essentially zero.
For the deuterons and the tritons, we expect the background from
protons to be vanishing, and the C background should be sufficient. For
the protons, however, they have a significant contribution. Moreover, due
to the larger beam energy, the background from elastic protons is now in
the middle of our excitations spectrum for 10 Li. To perform a subtraction
of this, we have instead simulated the elastic contribution and estimated
the contribution. The result of the simulation is shown in Figure 10.2 (A).
In the protons spectrum we see a rather large feature without a lot of
structure. The elastic proton peak is, however, visible around 3 MeV as a
peak on top of a broader structure beneath.
In the deuteron spectrum we see a the ground state around 0 MeV and
a broader structure at higher energies. The 1st excited state at 2.7 MeV is
easily embedded in this region. The neutron seperation energy for 9 Li is
4.06 MeV[TILLEY], so start to get into the continuum, and thus expect little
structure, even if there are 9 Li resonances up to 6.43 MeV. The resolution
98 CHAPTER 10. IS561A

(A) (B) (C)


20 10 10
9 Li(d,p)10 Li 12 C(d,p)13 C

8 8
15
∆E · cos θ [MeV]

E [MeV]

E [MeV]
6 6
10
4 4

5
2 2

0 0 0
5 10 15 20 120 125 130 135 140 145 150 120 125 130 135 140 145 150
E + ∆E · (1-cos θ) [MeV] θLAB [deg] θLAB [deg]

Figure 10.3: Results from the backward telescope. (A): The ∆E − E


spectrum. (B): The kinematic spectrum for events that stops in the DSSD
with 9 Li beam. (C): The kinematic spectrum for events that stops in the
DSSD with 12 C beam.

is clearly worse than in IS367, but this is expected from the higher beam
energy.
In the triton spectrum we see much less structure, than we did in IS367.
!!!MMMH; WHY!?

Backward telescope
In the backward telescope we only expect to see the protons from the 9 Li
(d,p)10 Li reaction. In Figure 10.3 we see the ∆E − E plot for the telsecope
to the left and the kinematic plot in the middle with the condition of no
particle in the pad. The protons we are interested in are so low in energy,
that they should stop in the DSSD. The kinematic curve for 9 Li (d,p)10 Li
is drawn on top, to show in which region we should see the protons. On the
right we see corresponding kinematic plot for 12 C(d,x).
There is not much structure in either of the two. The features down in
the right corner is unfortunately a noise peak. This effectively renders half
of the detector incapable of detecting the protons from 10 Li. In the other
half we see very few events and in particular not enough to extract any
structure in terms of angular distributions. This could be a DAQ related
10.3. DATA ANALYSIS 99

problem. Since we do see noise in half a detector, which corresponds to one


shaper, we may have put the trigger-thresholds too high to suppress this
noise. This may, however, unfortunately have been such that we also ignore
physics events. We can, however, try to put an upper limit on reactions in
this region. [UPPER LIMIT!!].

Intensity
In this experiment we only performed a single measurement on a heavy
target to estimate the intensity. The scattered 9 Li are very forward focused,
so we only see them in the downstream DSSD. The spectrum is shown in
Figure 10.4. Notice that the beam energy have been reduced to account
for energy loss in the target of the kinematic curve. We perform the same
analysis as we did in Section 7.2, but only for a single run. During the
measurement we get an intensity of

I = 4.6 kHz ± 6.4 %,


with a ratio
Nd = 1.53 · 10−6 ± 11.2 %,
and a total number of 9 Li

N9dLi = 2.87 · 108 ± 8.7 %.


100 CHAPTER 10. IS561A

70
65
60
55
50
45
E [MeV]

40
35
30
25
20
15
10
5
0
0 5 10 15 20 25 30
θCM [deg]

Figure 10.4: Kinematic spectrum for elastic scattered 9 Li off of a gold target.
The kinematic curve has been corrected for energy loss in the target.
11

IS561C

The last experiment in this series was carried out in the fall of 2018. The
main goal of this experiment was the two neutron transfer with a tritium
target to study 11 Li. The tritium target was backed by titanium so, we also
used a pure titanium target to estimate the background. For this experiment
the energy upgrade of HIE-ISOLDE was complete and we managed to get
a beam energy of 8.0 MeV/A. This energy was required to be able to
measure the protons from the reaction in the backward direction. Due to
the kinematic compression, would their energy be relatively low.
In case of problems with the primary target, we also had a deuterated
plastic target in the chamber. In case this was required, a normal plastic
target was also present for background.
This, unfortunately, turned out to be necessary. The main target turned
out to be damaged from an earlier experiment, and the repair made it mor
or less unuseable for our purpose. I will discuss this in more detail in Section
11.2.
For the stable pilot beam a heavy target was also used, in this case a
lead target. It was, however, not in the chamber during the radioactive
beam.
A table of all the targets can be seen in 11.1.
With the high intensity of the stable beam, we could use the same method
for calibrating pads and determine the beam spot and beam direction as we
did in 10.2. The beam spot and beam direction was estimated to be

101
102 CHAPTER 11. IS561C

p~bs = (−0.1(2) mm, 0.4(3) mm, −3.7(2) mm)


dˆ = (0.4(3)◦ , −23(1)◦ )

Table 11.1: The different targets used in IS561C.

Target Thickness Usage


3 H(48 Ti) 450 µg/cm2 Primary target
48 Ti 900 µg/cm2 Background for primary target
CD2 15 µm Secondary target.
CH2 15 µm Background for secondary target.
208 Pb 1 mg/cm2 Heavy target. Only used with the stable 12 C beam.

11.1 Setup
The setup for this experiment is very similar to the one used in IS561A.
The only major difference is another backward telescope, that was placed
symmetrically to the previous one, effectively doubling our coverage in the
backward directions. A sketch of the setup is shown in Figure ??.
The target wheel was replaced by a target stick, that could still be
operated from outside the chamber.
At the beam dump, we installed a small silicon telescope. The motivation
was to differentiate between beam components. The large instantanous
intensities meant that it could not be used for intensity measurements, the
electronics could simply not keep up. Moreover were there very severe pileup
effects during the experiment and the detectors deteriorated significantly
during the experiment. It was thus useful as a dianostics tool during the
experiment, but not for quantitative analysis.
An alternative beam dump was a Faraday Cup, which we used as a
diagnostics tool during the experiment. It was, however, not sensitive
enough to give a reliable estimate of the beam intensities.
11.2. MAIN TARGET 103

Figure 11.1: The target ladder slot with the 3 H foil. The three regions we
investigated is indicated.

11.2 Main target


During the run we realized there was a significant contribution from an
unknown component in our target. Due to a remount of the 3 H target in a
previous experiment, we suspected that it was due to glue that was used for
the remount. To sort this out and possibly get a background subtraction we
steered the beam to, what we believe, was the outer regions of the target.
An image of the target is shown in Figure 11.1.

Pentagon
If the target only had the expected 3 H and 48 Ti components, all the protons
we would see would come from a varity of reactions with 9 Li. If there is
a contaminant with a lot of proton (glue has a major organic part) we
would see a lot of elastically scattered protons. We can thus get an idea
of the background by looking at the kinematics of the protons we see in
104 CHAPTER 11. IS561C

Steel H3 Glue
20

18

16

14
E [MeV]

12

10

2
40 50 60 70 40 50 60 70 40 50 60 70
θLAB [deg]

Figure 11.2: Kinematic curves for 9 Li beam on the main target. The red
curves corresponds to elastic protons.

the forward direction. Kinematic curves for the three regions are shown in
Figure 11.2 together with the expected kinematic curve for elastic scattered
protons.
There is a big contaminant in both the region we expect to be the "good"
region and where we expect to see the glue. Notice the energy offset to the
kinematic curves. This indicated that the glue component is rather thick,
but we will return to this.
If we assume the two neutron transfer reaction we can plot the corre-
sponding excitation energy for 11 Li based on the measured protons, see
Figure 11.3. Unfortunately they are too similar in the region of interest and
we can not do a simple background subtaction.

S3
To get a more clear picture of the situation, we can also look in the round
DSSD placed downstream. We can not identify protons in this detector,
but we can see the elastic scattered 9 Li. To simplify the analysis we restrict
us to a single ring. The uncalibrated energy spectra from the three regions
as well as an energy spectrum from a measurement on the pure 48 Ti is
shown in Figure 11.4. Notice that the pure 48 Ti target was not modified in
11.2. MAIN TARGET 105

H3
103
Glue
Steel
Counts/81 keV

102

101

100

0
-8 -6 -4 -2 0 2 4 6
Ex [MeV]

Figure 11.3

preperations to this experiment, it was simply mounted and it is free from


glue.
In the spectrum for 3 H we see a clear peak, corresponding to scattered
9 Li. Below there is a feature, that resembles the feature from channel 2000

in the Glue spectrum. This is consistent with smeared out elastic peak due
to a large energy loss, for instance a thick layer of flue.
At the low channel numbers, up until channel 1000, we see a feature
that resembles the same region in the Steel spectrum. This feature falls off
to quickly to also be present in the Glue spectrum.
Neither of these two features are present in the pure 48 Ti spectrum. The
three spectra from the 3 H target thus all seem to be largely dominated by
reactions with the glue and steel.

Backward detectors
The elastic protons will not scatter beyond 90◦ , and they will therefore not
contaminate the backwards detectors. Reactions from 9 Li with steel, the
106 CHAPTER 11. IS561C

103
H3 Glue
103

102
Counts/4ch

102

101
101

100 100
0 0
0 1000 2000 3000 4000 0 1000 2000 3000 4000

Steel 102 Ti
103
Counts/4ch

102
101

101

100
100
0 0
0 1000 2000 3000 4000 0 1000 2000 3000 4000
Channel Channel

Figure 11.4

unknown components of glue and 48 Ti may however give a sizeable but


unknown contribution in the backward direction.
Since we did not have particle identification for such low-energetic
protons, we must rely on energy cuts. The kinematics of the two neturon
transfer reaction allow the protons to have in the order of 800 keV-1000 keV
in the angular range of that detector for the groundstate of 11 Li. We
therefore require the energy in the DSSD to be within 450 keV-1500 keV
with no hit in the pad.
The resulting excitation energy for the two backward detectors are shown
in Figure 11.5. The situation here are more or less the same, and we can
not distinguish the background components from any signal.
11.2. MAIN TARGET 107

60 BB7L
H3
50 Glue
Steel
40
30
20
Counts / 100 keV

10
0
BB7R
30 H3
Glue
Steel
20

10

0
-2 -1 0 1
E [MeV]

Figure 11.5

Trigger thresholds
We experienced another issue with the backward detectors, limiting our
effective coverage region. The trigger thresholds was, due to noise issues
during the setup phase, simply set to high for the backward detectors. Our
trigger effiency, that is the fraction of hits that we trigger on at a certain
energy, can be calculated from the TDC and ADC values. The TDC have a
value whenever a channel had a signal to trigger on. The ADC can, however,
digitize non-zero values if the DAQ was trigger from another source. In a
physics run this could be a coincident particle or random noise.
For a more precise measurement of this we can use a pulser, and make
sure we trigger the DAQ every time we send in a pulse. By varing the pulse
height we can simulated different energies and we can see if a particular
channel produces a trigger as well, i.e. a TDC value. We can calculate the
108 CHAPTER 11. IS561C

0.8
Counts/2keV

0.6

0.4

Pulser Pulser
0.2
Physics Physics

0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7 8
E [MeV]

Figure 11.6

trigger effeciency for a given channel as

N (T DC > 0 and ADC > 0)


Rch = . (11.1)
N (ADC > 0)
For both detectors we did a pulser measurement. The result is shown in
Figure 11.6 together with the effiency measured from physics. We can see
that the slopes are very similar, which we expect. Unfortunately it vanish
below 1 MeV, which is the area we need to be looking in. We can thus not
use these detectors to get a measurement of the protons. When the effiency
is so close to zero, attempts to put an upper limit had huge uncertainties,
and was therefore effectively useless.

Thickness of contaminant
We can try to estimate the thickness of the glue component based on the
energy loss. Since we do not know exactly what the glue is made of we model
it as plastic. We thus assume that the target consists of a think titanium
foil and a thick layer of plastic. As usually we identify all the protons, and
do the energy correction, assuming that that the reaction happend in the
middle of the thick plastic part.
11.2. MAIN TARGET 109

20

15 (A) 9Li on Ti (B) 12C on Ti

10

5
E [MeV]

200

15 (C) 9Li on Ti+PE (D) 12C on Ti+PE

10

0
40 50 60 70 80 40 50 60 70 80

θLAB [deg]

Figure 11.7

If we ignore any glue component we get the kinematic curves as shown


in Figure 11.7 (A) and (B) for 9 Li and 12 C respectively. The graphs are the
elastic scattering of protons in each case. We clearly see that the energy is
too low, i.e. we are missing some energy in our analysis.

We can now try to get the kinematic curve for the elastic proton for
both 12 C(Glue) and 9 Li(Glue) to fit, by adding a layer of plastic. We should
be able to agree on a thickness which fits both beams. With a thickness of
90µm we get Figure ?? (C) and (D). This looks consistent, considering the
material is not the correct one.
110 CHAPTER 11. IS561C

p d t
CD2 CD2
120 CD2
200
CH2 CH2 CH2
250
175 100

150 200
80
125
150
100 60

75 100 40
50
50 20
25

0 0 0
-4 -2 0 2 4 6 8 -4 -2 0 2 4 6 8 -4 -2 0 2 4 6 8
Ex [MeV]

Figure 11.8: Excitation spectra from the identified protons, deuterons and
tritons.

11.3 Secondary target


The secondary target was deuterated plastic as in case of the two previous
experiments. Due to the limited space on the target ladder, did we only
have a CH2 target for background measurements and no pure carbon target.
The ∆E − E plot for the forward detectors is shown in Figure ??, and we
have detected the same particle types as in the previous experiments. The
resulting excitation spectra is shown in Figure 11.8.
The protons spectrum shows largely the same features as we saw at
6.72 MeV/A. The background is again dominated by the elastic peak. Since
we do not have a measurement with pure carbon we can not explicitly
subtract the carbon component. However, looking at the spectra for both
deuterons and tritons, where the backgrond should mainly be dominated by
carbon reactions, we get approximately a factor of two, which is more or
less negleglible for the protons.
In the deuteron spectrum we see a large ground state peak and at the
first excited state at 2.7 MeV. We also see two fetures above the neutron
seperation threshold, which corresponds to the second, third and even fourth
excited state.
11.3. SECONDARY TARGET 111

In the triton spectrum, we have a prominent ground state and second


excited state. The first excited state is not resolved with the resolution of
this setup.

Intensity
In this experiment we did not have a heavy target due to target ladder
limitations. This also means that we do not have a good measurement
of the intensity. The titanium target is too light for the the 9 Li beam to
scatter according to the Rutherford distribution in our coverage region. The
grazing angle for 9 Li is only 9.1◦ in LAB and we do not have coverage here.
We can make a crude estimate by extrapolating the scattering on ti-
tanium into the Rutherford region. Beyond the grazing angle, we see
Fresnel and Fraunhofer types of diffraction patterns. We do not have
the angular resolution to resolve these, but they do follow an exponential
trend[HODGESON, pp533]. At angles up to the grazing angle the the
ratio is just one. At angles above the the grazing angle the ratio follow an
exponential decay

dσ/dΩ
∝ e−λ(θ−θgz ) ,
dσR /dΩ

for angles θ ≥ θgz .


If we calculate the intensity based on the assumption that we are below
the grazing angle, according to the method in 7.2, we will in fact carry on
the exponential factor and instead calculate

I 0 = I · e−λ(θ−θgz ) . (11.2)
We can now measure in a number of different angles and fit to an
exponential to determine λ and in particular the real intensity I, which
corresponds to the value at θ = θgz .
We had a varity of problems with the down stream detector, and in par-
ticular the calibrations. This meant that the matching algorithm discarded
112 CHAPTER 11. IS561C

most of the elastic events. I therefore performed this analysis using just
the rings which are mostly symmetric in θ. If we assume that each ring
corresponds to a single angle we can plot the deduced intensity as a function
of the central angle in each ring. See Figure 11.9 for the plotted data and
the subsequent fit to 11.2. The fitted value is

Ifit = 41(3) kHz, (11.3)

where the error comes from the fit. There are, however, several systematic
errors, which must also be accounted for. First of all are the rings not fully
symmetrical, and the central angles does not reflect this. Secondly are we
averaging over the ring, which becomes more problematics as we approach
0◦ . Most importantly, however, is our model crude, and there are many
details about the diffraction pattern that we have left out. We do also not
fit over many angles. Due to these systematics we take slightly conservative
approach and report an uncertainty on 25 %.

I = 41(10) kHz, (11.4)


We did have a Faraday cup in, but it was not very accurate due to the
low current. The reading, according to the operators of this, is however
consistent with the order of magnitude.
11.3. SECONDARY TARGET 113

102
Fit
Intensity [kHz]


θgz = 9.1◦

101

7 8 9 10 11 12 13 14
θLAB

Figure 11.9: Calculated intensity based on 9 Li scattered off of 48 Ti. The


data has been fitted to an exponential fit and extrapolated to the grazing
angle at 9.1◦ .
12

Results

The first simplest reaction is the elastic scattering of deuterons. Due to the
inverse kinematics the deuterons only scatter at θLAB ≤ 90◦ in LAB, and
we detect a large range of the CM angles. We only observed the ground
state at 2.68 MeV/A, but at 6.72 MeV/A and 8.0 MeV/A we also observed
inelastic scattering to the excited states in 9 Li and even break-up channels.
The width of the ground state can give us an estimate of the resolution of
the experiment. The FWHM yields a resolution of INSERT RESOLUTION.
The elastic channel can be described by the optical model, and we
can calculate the expected angular distribution as described in Section ??.
Furthermore, we can try to describe the scattering using a CDCC approach,
where we treat the scattering of a neutron and a proton respectively and
introduce a secondary deuteron interaction. Such a description would be
similar to our model of the neutron transfer to 10 Li, and may provide some
good consistency checks.
With the elastic scattering results, we can continue to the more difficult
task of describing 10 Li, which will be done in the framework described in
Section ??. We can, however, use the optical potentials for 9 Li-p scattering,
which we measured with the plastic target.
Another open channel in our experiments is one neutron transfer from
9 Li to the deuteron, and thus the measurement of 8 Li. For this reaction we

have several bound states, and will use a DWBA approach, as described in
Section ??. This has, however, not been a focus point.

115
116 CHAPTER 12. RESULTS

12.1 Elastic scattering


Optical model
In the optical model, we need to determine the the optical model parameters
to best describe the data. We can try to different approaches. There has
been several attempts to make systematic parameterizations of optical model
parameters, however, many of these are mostly valid at higher beam energies
and heavier systems. Nontheless is it worth to try some of these to see how
they perform in this energy range. For 2.68 MeV/A we can compare to the
fit from the 2004 experiment. The beam energy is very similar, so we can
expect the cross sections to be similar. Lastly we can try to perform a fit.
In Figure 12.1(a) we see the fit and the two fits from [Jeppesen] and
in Table 12.1, we see the corresponding parameters. The agreement seems
resonable with P1. The fitted parameters are also close to this fit, with a
slightly lower Wi , meaning slightly less strength in the inelastic channels.
We notice that between 140 and 160 there seems to be a small feature
in the data that is not reproduced by the fit.
When we increase the beam energy, the structure of 9 Li should not
change. In the optical model this is mainly described by the real part
of the potential, which dictates the mean field of the nuclear potential.
The loss of flux to other channels, may however change due to more open
channels. This is mainly described by the imaginary surface terms. A
simple extrapolation from 2.68 MeV/A to 6.72 MeV/A is thus to free the W
parameter and perform a new fit. The result is shown in Figure 12.1(b), with
Wi = 59.05. Freeing more parameters did, however, give a better result. We
can do the same excercise with 8.0 Mev/A, and the result is shown in Figure
12.1(c), with Wi = 53.4. Since the optical model is a phenomenological
model, all of the parameters may have an energy dependence.
We can do the same analysis for elastic protons, since we need the
potentials for 9 Li-p for our following CDCC calculations. We have measured
elastic scattering of protons as part of our background measurement.
The result is shown in Figure 12.2(a), (b) and (c) for 2.68 MeV/A,
6.72 MeV/A and 8.0 MeV/A respectively. The three parametrizations, Pow-
12.1. ELASTIC SCATTERING 117

ell, Watson and KD can not reproduce the correct form, but the magnitude
is not too far of. Making a fit to the data, we get a set of parameters that
reproduces the data better. This would in principle be the correct input
for the CDCC calculation. Since we do not have a corresponding fit for the
neutrons, and we need to use some parametrization to describe the 9 Li-n
interaction.

CDCC
For the CDCC calculation we need to include both the 9 Li-n and the 9 Li-p
interaction. Further more we need to specify the bins we use for the p-n
continuum. In the present calculation we make sure to include energies
high enough to let the solution converge. In practice we used 6 MeV for
2.68 MeV/A and 10 MeV for 6.72 MeV/A and 8.0 MeV/A. Furthermore we
need to specify the spin and parity states for the individual bins. Once again,
we simply need to include enough partial waves to make the calculation
converge. The deuteron ground state is J π = 1+ , and with the coupling
scheme discussed in Section ?? we include states up to l = 3.
For the 9 Li groundstate we use the Reid-Soft-Core as implemented in
FRESCO by AM. We include the coupling to each of the bins in the p-n
continuum.
The result of the CDCC calculations with different optical potentials is
shown in Figure 12.3 for the three different energies. In the 2.68 MeV/A case,
the Watson parametrization is far off. From the elastic protons calculation,
Watson was the parametrization which was furthest from reproducing the
shape. If the 9 Li-n interaction have a similair descrepency, it is not surprising
that the CDCC calculation using these optical models are worse. It is,
however, a noteworthy that it overestimates the magnitude by a large factor.
The KD parametrization, on the other hand, does a good job reproducing
the shape and magnitude of the data. The first minimum is particularly
underestimated.
If we use the fitted potential for the 9 Li-p interaction we get closer in
this region but at the expense of agreement in the larger angles. Since
118 CHAPTER 12. RESULTS

20
(a) 2.68 MeV/A

15 P1
DA1P
P2
10
Fit

250
(b) 6.72 MeV/A
20 P2
dΩ [mb/sr]

P1
15
P3
10

300
(c) 8.0 MeV/A
25
P2
20 P1

15

10

0
20 40 60 80 100 120 140 160 180
θCM

Figure 12.1: Angular cross section for elastic deuterons at three different
energies. Data for 2.68 MeV/A is shown in (a) together one parametrizations
of the optical potentials as well as two fits from [Jeppesen]. The last curve
corresponds to the a fit to the data. In (b) we see the data for 6.72 MeV/A
with three fits to the optical potentials. In (c) we see the data for 8.0 MeV/A
with two fits. P1 us fitted to the full dataset whereas P2 is fittede to the
first feature below 60◦ . The parameters used in the calculations are listed
in 12.1.
12.1. ELASTIC SCATTERING 119

100
(a) 2.68 MeV/A
80 Powell
Watson
60
KD
40 Fit

20

2000
(b) 6.72 MeV/A

150 Watson
dΩ [mb/sr]

KD

100

50

2000
(c) 8.0 MeV/A

150 Watson
KD

100

50

0
20 40 60 80 100 120 140 160 180
θCM

Figure 12.2: Angular cross section for elastic protons at three different
energies. Data for 2.68 MeV/A is shown in (a) together with different
parametrizations of the optical potentials. The last curve corresponds to
the a fit to the data. In (b) and (c) we see the data for 6.72 MeV/A and
8.0 MeV/A respectively together with two parametrizations of the optical
potentials. The parameters used in the calculations are listed in 12.1.
120 CHAPTER 12. RESULTS

we do not have a measurement of the 9 Li-n interaction, it is taken from


parametrizations.
For 6.72 MeV/A the Watson parametrizations performs much better, and
the KD underestimates by a factor of ≈ 2. It is not clear why this performs
worse at higher energies, since we in principle are getting closer to the proper
energy range. We are, however, still close to the limit. Furthermore, does
the elastic protons seem to be well reproduced by both, which could indicate
that the 9 Li-n interaction is not well described.
The same trend continues as we go to 8.0 MeV/A for KD. Watson is
now also down by the same factor. In both cases are the 9 Li-p cross
sections underestimated, which may explain why the CDCC calculation
underestimates both.
It is worth noting that the shape seems to be reproduced correctly,
which could indicate that we are missing some reaction channels such as
CN-reactions, that would contribute with a relatively flat background level,
and thus an offset to the calculations. However, since the background level
was low, this would mean that the majority of the CN-contribution should
be kinematically similar to elastic scattering, which seems unlikely
When we want to model 10 Li we need both the 9 Li-p and 9 Li-n interaction.
Therefore it would in principle be good to "measure" the 9 Li-n interaction
by means of measuring the 9 Li-p and fit the 9 Li-n part of the potential
via a CDCC calculation. The agreement of both KD and KD+Fit are
resonable it is not clear whether we will learn much by performing a more
detailed investigation of the 9 Li-n interaction. The computation time for
each iteration of the CDCC is calculation relatively long, making a fit
unfeasible for this thesis.

10
12.2 Li transfer
The analysis of the one-neutron transfer is more difficult, since the resulting
10 Li is unbound. This means that the calculation is more involved due to

fact that we need to describe both the reaction mechanism as well as the
12.2. 10 LI TRANSFER 121

20
(a) 2.68 MeV/A

15 Watson
KD
KD+Fit
10

250
(b) 6.72 MeV/A
20 Watson
dΩ [mb/sr]

KD
15

10

300
(c) 8.0 MeV/A
25
Watson
20 KD

15

10

0
20 40 60 80 100 120 140 160 180
θCM

Figure 12.3: Angular cross section for elastic deutrons calculated with CDCC.
Data for 2.68 MeV/A is shown in (a) together with two parametrizations of
the optical potentials for 9 Li-n and 9 Li-p. The KD+Fit uses the 9 Li-p that
was fitted to elastic protons. In (b) and (c) we see two parametrizations of
the 9 Li-n and 9 Li-p optical potentials.
122 CHAPTER 12. RESULTS

Table 12.1: Parameters for optical potentials for elastic scattering. Parame-
ters in bold were fitted otherwise they were fixed.

Vr rr ar Wi ri ai Vls rls als


9 Li (d,d)9 Li
2.68 MeV/A
P1 104.6 1.2 0.65 12.4 1.2 0.65 6 1.4 0.7
P2 134.6 1.0 0.64 12.2 1.0 0.72 6 1.4 0.7
FIT 104.3 1.2 0.62 10.5 1.2 0.68 6 1.4 0.7
6.72 MeV/A
Freev 104.3 1.2 0.65 59.05 1.2 0.65 6 1.4 0.7
Freevw 93.04 1.2 0.65 61.85 1.2 0.65 6 1.4 0.7
Freevwaa 103.23 1.2 0.61 57.88 1.2 0.63 6 1.4 0.7
8.0 MeV/A
FIT 104.3 1.2 0.62 53.4 1.2 0.68 6 1.4 0.7
FIT2 104.3 1.2 0.62 58.2 1.2 0.68 6 1.4 0.7
9 Li (p,p)9 Li
2.68 MeV/A
Watson 68.86 1.15 0.57 4.86 1.15 0.5 5.5 1.5 0.57
Powell 49 1.25 0.65 7 1.25 0.47 6 1.4 0.7
KD 61.74 1.11 0.68 7.98 1.31 0.52 5.69 0.87 0.59
Fit 62.9 1.2 0.65 13.6 1.2 0.65 6 1.4 0.7
6.72 MeV/A
Watson 67.77 1.14 0.57 7.19 1.14 0.5 5.5 1.14 0.6
KD 59.95 1.11 0.68 9.68 1.31 0.5 5.6 0.87 0.6
12.2. 10 LI TRANSFER 123

10 Li structure via couplings to the p-n continuum as described in Section


??.

Acceptance
The correction of the acceptance is also slightly more complicated since we are
not looking at bound states and thus not on single excitation energies. This
requires a more explicit treatment of the the double differential acceptance
as described in 7.9.
One challenge with the double differential acceptance is the amount of
statistics. Nrd (θ, φ, E) and in particular Nsd (θ, φ, E) can get quite small,
which can lead to large uncertainties in those bins.
We have large kinematic and geometric areas that our setup does not
cover. If we follow the procedure outlined in Section ?? we can investigate
our acceptance region by simply plotting the number of detected particles
given a uniform distribution. In the case of 10 Li we thus perform the
simulation uniformly in 4π and uniformely on Er = 2.01 MeV which is the
available energy in CM. We can now analyze the simulation and plot the
number of detected particles as a function of CM angle and excitation energy
and the result for 2.68 MeV/A and 6.72 MeV/A is show in Figure 12.4.
We see that it is a non-triviel region that is covered. Traditionally
theorist are aware and will report the integrated cross section in some
region, for instance angular distributions between 0 MeV-2 MeV. There are,
however, a complicated correlation between and angle and energy in terms
of the acceptance.
We can approach this in two ways. The simples way is to gate on a
"simple" region. In this case we would have coverage and we can safely
integrate the theoretical calculations to get single differential cross sections.
An alternative approach is to use the acceptance to calculate the expected
number of counts and then compare this the experimental result. This has
the advantage that we can use the full dataset, however, it makes it difficult
to compare to other experimental datasets and models that have not been
folded with our acceptance.
124 CHAPTER 12. RESULTS

2.68 MeV/A 6.72 MeV/A 8.0 MeV/A

7 8

1.5 6
6
5
E [MeV]

1.0 4
4
3
0.5 2
2
1
0.0 0 0

75 100 125 150 40 60 80 100 120 40 60 80 100 120


θCM

Figure 12.4: Acceptance for 9 Li (d,p)10 Li at 2.68 MeV/A 6.72 MeV/A and
8.0 MeV/A respectively. Notice lines in the two first panels. They corre-
sponds to dead strips of the detectors.

Let us consider the case of 2.68 MeV/A, the left panel in Figure 12.4.
Suppose we want to get the differential cross section the angular axis. The
upper left area in the acceptance plot makes it impossible to define a "simple"
region that includes all the data. We can choose θCM = 90◦ − 150◦ and
E = 0 MeV − 1.5 MeV. In this case we have full coverage, and we can
compare to theoretical predictions integrated over either θCM = 90◦ − 150◦
or E = 0 MeV − 1.5 MeV. This would, however, be on the expense of some
data. If we instead include angles down to θCM = 70◦ and project onto the
angle axis, we are missing a lot of cross section at the low angles but high
energy due to the vanishing acceptance.
Given a range we now have to choose our bin size. To the lowest
approximation, we may simply choose one energy bin to increase the statistics
in each of the angular bins. This effectively changes (7.10) to

dσ(θ, φ) 1 Ns Nrd (θ) ∆E


≈ , (12.1)
dΩ nt NI Nsd (θ) 4πEr

where ∆E is now the energy range we chose from the acceptance plot. Nrd (θ)
12.2. 10 LI TRANSFER 125

is the number of detected particles detected in a given angle, summed over


all energies in the range. Nsd (θ) is the corresponding number of detected
simulated particles summed over the energy range.
In the same way (7.11) become

dσ(θ) 1 Ns Nrd (E) ∆Ω


≈ , (12.2)
dE nt NI Nsd (E) 4πEr

where Ω is the solid angle covered in region.


In this approximation we have assumed that we can treat the accep-
tance across an energy bin as an average. If there is a large variation in
acceptance as well as in cross section accross the energy bin, this is not a
valid approximation. In this case we need to use the correct version with
the statistics problems it may give.
For 2.68 MeV/A we can try both methods to see how they perform,
depending on the range we choose. Consider first we make a conservative
range with θCM = 90◦ − 150◦ and E = 0 MeV − 1 MeV. We can now choose
to use 1 bin (S) or 15 bins (C). The result for differential energy cross
sections as well as differential angular cross sections are shown in Figure
12.5.
When we choose a region that is well covered(S90 and C90), we can
see that the two methods are very similar. We can thus use either and get
equally good results. When we increase the range to include areas that
are not covered by our acceptance (S70 and S90) we start to overestimate
the cross section. The simpler method, in particular. If we compare the
angular distributions of C70 and S70 we see that we get qualitatively very
different behaviours at low angles. This underlines the importance of have
a good understanding of the acceptance behaviour when we compare with
experiment.
The exact behaviour depends on the specific acceptance, but we can
conclude that if we stay within a save region we can safely use either
methods.
126 CHAPTER 12. RESULTS

S 90
0.010 0.05 S 70
C 90
C 70
0.008 0.04
Arb. units

Arb. units
0.006 0.03

0.004 0.02

S 90
0.002 S 70 0.01
C 90
C 70
0.000 0.00
80 100 120 140 0.0 0.5 1.0 1.5 2.0
θ CM E [MeV]

Figure 12.5: Acceptance corrections for 2.68 MeV/A for both angular and
energy differential cross sections.

Singles
Based on the acceptance we limit our analysis to θCM = 90◦ − 150◦ and
E = 0 MeV − 1 MeV. The final result is shown in Figure 12.6 (A). There
are two main observations in the excitation spectrum.
The first is the significant strength at the threshold in the excitation
plot (right panel). This is on the debated points, but it is clear that these
data show a significant contribution from a s1/2 state in order to reproduce
the data. The overall shape of the spectrum seems to fit the calculation
nicely, however, the magnitude if off by a factor of ≈ 2. One possible
explanation is a sizeable contribution from CN reactions. There is multiple
ways of estimating the expected contribution of compound nucleus channels,
however, we will not pursue such calculations in this works.
The second main observation is the shape of the peak. Even though the
data give a clear peak just below 0.5 MeV, we can not directly conclude that
we need two resonant p1/2 contributions to reproduce it. The resolution of
the setup is not good enough to distinguish two close resonant states. This
will instead require a more detailed invesitagtion of the model.
For the d-wave contributions, we still have to little energy available
12.2. 10 LI TRANSFER 127

Angular cross section (0-1 MeV) Energy cross section (90-150 ◦ )


30
s1/2 1-
s1/2 2-
25 p1/2 1+

dσ/dEx [mb/MeV]
p1/2 2+
dσ/dΩ [mb/sr]

101 20 Sum

15
A)
10
100

0
0 25 50 75 100 125 150 175 0.0 0.5 1.0 1.5 2.0
θCM [deg] Ex [MeV]

Angular cross section (0-6 MeV) Energy cross section (60-100 ◦ )


s1/2 1-
s1/2 2-
102 20
p1/2 1+
dσ/dEx [mb/MeV]

p1/2 2+
dσ/dΩ [mb/sr]

d5/2 1-
15 d5/2 2-
101
d5/2 3-

B) 10
d5/2 4-
Sum
100
5

10−1
0
0 25 50 75 100 125 150 175 0 1 2 3 4 5 6 7
θCM [deg] Ex [MeV]

Angular cross section (0-6 MeV) Energy cross section (60-100 ◦ )


14 s1/2 1-
s1/2 2-
102 12 p1/2 1+
dσ/dEx [mb/MeV]

p1/2 2+
dσ/dΩ [mb/sr]

10 d5/2 1-
d5/2 2-
101 8 d5/2 3-

C) 6
d5/2 4-
Sum

100 4

2
10−1 0
0 25 50 75 100 125 150 175 0 2 4 6 8
θCM [deg] Ex [MeV]

Figure 12.6: Differential cross for 2.68 MeV/A (A), 6.72 MeV/A (B) and
8.0 MeV/A (C). Right panel shows the angular cross section and the left
panel shows the energy cross section. The title of each panel indicates the
integration range in the other variable.
128 CHAPTER 12. RESULTS

in CM at 2.68 MeV/A. We can instead look at the results from the two
experiments at higher energies. In these two experiments we have a few more
limitations. For 6.72 MeV/A the statistics are limited and the the intensity
are more uncertain, giving a larger systematic error. For 8.0 MeV/A we
have decent statistics, but the intensity estimate is very uncertain. For
both cases, the energy resolution is not as good as the low energy, which
is mainly an effect of the higher beam energy, giving a larger kinematic
compression. This means that we can not say much about the low energy
region (0 MeV-2 MeV), but we can use it to investigate the indications of a
d-wave which may lie in the 2 MeV-10 MeV range.
We need to choose a safe range like we did before. For 6.72 MeV/A
AND 8.0 MeV/A we chose θCM = 65◦ − 100◦ and E = 0 MeV − 6 MeV. The
results are shown in Figure 12.6 (B) and (C), respectively. We see a broad
feature above 2 MeV in both cases. For 6.72 MeV/A the shape of the angular
distribution seems to match well, but the magnutide is too low. The missing
cross section in the model comes mainly from the high energy region, and
in particular in the area where we could expect to see the d-wave resonance.
This could indicate the exsistence of such a resonance, and this is backed
by a similar feature at 8.0 MeV/A. The normalization of this data set is,
however, rather uncertain, so it is difficult to conclude much from magnitude.
Without the measurement of coincidences, it is difficult to determine how
much of this belongs to other reactions mechanisms.

Slices
We can learn more about the data if we only look at the angle integrated or
energy integrated differential cross sections in certain bins. This is basically
an attempt to look at the double differential cross section. We can divide
our acceptance region into bins of 10◦ and calculate the angle integrated
cross section. Likewise can we divide the energy range into bins of 200 keV
for 2.68 MeV/A and 1 MeV for 6.72 MeV/A and 8.0 MeV/A and calculate
the energy integrated cross sections. The results are shown in Figure 12.7,
12.8 and 12.9 respectively.
12.2. 10 LI TRANSFER 129

2.68 MeV/A We notice that the main contribution of the strength at


0 MeV comes from the small angles. According to the calculation, this is
not suprising, since the s1/2 (2−) are dominating is this region. When we go
further out to larger angles, we see very similar curves, indicating that the
different components does not evolve significantly. The angular distribution
does not have a lot of structure. They are all rather flat, contrary to what
the model predicts. This could be an indication of a compound nucleus
component.

6.72 MeV/A The angle integrated cross sections (left column) shows an
interesting dip in the 60◦ -70◦ bin, clearly indicating some structure. This
could be due to the suppresion of the s-waves even though this does not fit
with the position from the current calculation. The energy integrated cross
sections shows a similar dip in the 3 MeV-4 MeV bin. In general, however,
the angular distribution looks more consistent with the model, since all the
bins increases as we approach lower angles.

8.0 MeV/A The energy integrated bins (right column) looks more struc-
tureless. At the higher energy bins, the angular distributions again increases
at lower angles, but it does look more flat at low energies. The angle
integrated (left column) cross sections looks structurally very similar.

Coincidences
In IS367 we did measure a small amount of 9 Li in coincidens with protons.
We now consider all protons that was detected together with one other hit.
If we assume that the other hit is a 9 Li, we can calculate the missing energy
and momentum and deduce the neutrons energy as well as direction. More
importantly, however, can we look at the correlation between the neutron
and the 9 Li or the neutron and the proton.
To explain the descrepency we see in the cross section we may explore
some other possible channels that may contribute to our reaction, and to
which extend we can understand the contributions from an experimental
point of view.
130 CHAPTER 12. RESULTS

5 1

0.0-0.2MeV
θCM =80-90 0
0 1
5

0.2-0.4MeV
0
0
θCM =90-100 1
5
0.4-0.6MeV
0
1
θCM =100-110
0
dσ/dΩ [mb/sr]

5
0.6-0.8MeV
0
dσ/dE [mb/MeV]

1
θCM =110-120
0
5 0.8-1.0MeV
0
1
θCM =120-130
0
1.0-1.2MeV
5 0
1

θCM =130-140
0 1.2-1.4MeV
0
5 1

θCM =140-150 1.4-1.6MeV


0 0
0.0 0.5 1.0 1.5 80 100 120 140
E [MeV] θCM [deg]

Figure 12.7: Differential cross section for angular (left panel) and energy
(right panel) bins for 2.68 MeV/A. The distributions seems very similar
across both axes, indicating that we are, in fact, not measuring a direct
reaction.
12.2. 10 LI TRANSFER 131

5 5

θCM =50-60
0 0 0.0-1.0MeV

5 5

θCM =60-70
0 0 1.0-2.0MeV

5 5
dσ/dE [mb/MeV]

dσ/dΩ [mb/sr]

θCM =70-80
0 0 2.0-3.0MeV

5 5

θCM =80-90
0 0 3.0-4.0MeV

5 5

θCM =90-100
0 0 4.0-5.0MeV

5 5

=100-110
0 5.0-6.0MeV
θ
0 CM
0 2 4 6 50 60 70 80 90 100 110
E [MeV] θCM [deg]

Figure 12.8: Differential cross section for angular (left panel) and energy
(right panel) bins for 6.72 MeV/A.
132 CHAPTER 12. RESULTS

5.0 5

2.5
0 0.0-1.0MeV
θCM =50-60
0.0 5
5.0
0 1.0-2.0MeV
2.5
5
θCM =60-70
0.0
5.0 0 2.0-3.0MeV
5
2.5
dσ/dΩ [mb/sr]

θCM =70-80
0.0 0 3.0-4.0MeV
5.0 5

2.5
0 4.0-5.0MeV
dσ/dE [mb/MeV]

θCM =80-90
0.0 5
5.0
0 5.0-6.0MeV
2.5
5
θCM =90-100
0.0
5.0 0 6.0-7.0MeV
5
2.5
=100-110
0 7.0-8.0MeV
θ
0.0 CM
0 2 4 6 8 50 60 70 80 90 100 110
E [MeV] θCM [deg]

Figure 12.9: Differential cross section for angular (left panel) and energy
(right panel) bins for 8.0 MeV/A.
12.2. 10 LI TRANSFER 133

50 Singles
Coincidences (Phase space)
Coincidences (Breakup)

40
dσ/dEx [mb/MeV]

30

20

10

0
0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75
Ex [MeV]

Figure 12.10: Differential cross section for 10 Li for singles and coincidens
measurements. The singles spectrum is only based on protons. The second
dataset (squares) are coincidences with the acceptance of C. The last dataset
(triangles) are coincidences but with the acceptance of 10 Li breakup A.

In the current scheme we look at

A) 9
Li(d, p)10 Li 
n + 9 Li
B) 9
Li(d)11 Be 9 
Li + n + p
C) 9
Li(d, Li)d 9

n+p
D) 9
Li(d, n)10 Be 9 
Li + p

• A This is the scheme in which we have developed our model. The


reaction is a sequence of a neutron transfer followed by a breakup into
a neutron and 9 Li.
134 CHAPTER 12. RESULTS

• B The deuteron may create a compound nucleus that eventually


breaks up. This is not the only channel that can eject a proton, but if
we measure a consistent 9 Li as well, this is the only reaction that has
the correct final state.

• C The deuteron may scatter of and break up into a proton and a


neutron.

• D We may in fact not transfer a neutron, but rather a proton. In this


case, the resulting 10 Be may be unstable to proton emission and we
end up in the same final state as previous.

In Figure 12.10, we see the singles spectrum together with the coincidens
spectrum. We have corrected for acceptance with two assumed reaction
mechanisms, A and C. Our first observation is that both magnitude and
shape is consistent with each other and the singles spectrum. The chan-
nels that contribute must be within our acceptance for both singles and
coincidences. A simulation of D showed that we have no acceptance for
coincidences in this reaction, and we can thus rule out D (At least at
the precision of our uncertainties). Otherwise we would see a descrepency
between singles and coincidences. It should be noted that our acceptance
region is the same as for the singles, and we can use the same region and
compare the integrated cross sections. 1
To investigate the reaction mechanism, we can exploit the kinematic
differencens and therefore the difference in acceptance. We can simulate the
three first different situations, uniformly in CM, and analyse the results as

9
Li + d → 9 Li + p + n, (12.3)
where we measure 9 Li and p and deduce n.
The result of the simulation is shown in Figure 12.11, where we have
plotted the correlation between the deduced neutron as a function of either
the proton or 9 Li. The first row, corresponding to situation A, have a

1
FiXme Fatal: En figur?
12.3. 8 LI TRANSFER 135

different acceptance region compared to B and C, which are more similar.


In the framework of 10 Li, it is easier to think about the proton, since it is
ejected opposite in CM (left column). If we now compare the data, we see
that it have the same cut off at ≈ 100◦ as A. Both B and C does, however,
extend to ≈ 140◦ . To a first approximation, the data is mainly compatible
with A.
The cross section of the deuteron break-up could, however, drop very
fast close to 100◦ for the proton, and thus explain why we do not see it
in this correlation plot. It would then still contribute to measurement at
lower angles. We do, however, actually have an estimate of this, due to
our CDCC calculation of the elastic cross section in Section 12.1. Based
on this calculation, see Figure ??, the cross section does not vanish in this
region, and we can thus conclude that the deuteron break up have a small
contribution.
We can use the same argument for situation B. In this case, the angular
distribution should be largely flat (WHY?), and we would expect to see a
component beyond θp > 100◦ . Since this is not the case, we also conclude
that this mechanism is mostly a small component.
The dominating reaction mechanism for the coincidence measurements
is thus A, and since both the magnitude and the shape is consistent with
the singles, we also expect the dominating reaction mechanism in this case
to be A.

8
12.3 Li transfer
136 CHAPTER 12. RESULTS

150 (A) (A)

100

50

0
150 (B) (B)

100

50

0
150 (C) (C)
θn

100

50

0
150 Data Data

100

50

0
0 50 100 150 0 50 100 150
θp θ9Li

Figure 12.11: Acceptance for angle correlations between the neutron and
either the ejected proton or the ejected 9 Li based on different simulated
kinematic situations. The letters correspond to the situations outlined in
the text. The left column is the θn versus θp and the right columns is θn
versus θ9 Li . All angles are in center of mass.
List of Figures

2.1 Kinematic curves for ejectile and recoil of the reactions 9 Li


(d,t)8 Li, 9 Li (d,d)9 Li and 9 Li (d,p)10 Li at 2.68 MeV/A. . . . . . 7
2.2 Kinematic effect on angular resolution and solid angle coverage.
All the plots corresponds to a proton from 9 Li (d,p)10 Li at
2.68 MeV/A. A: The relation between the scattering angle in
LAB and CM. B: The differential, which gives a measure of the
angular resolution in LAB needed to recover a given angular
resolution in CM. C: The Jacobian that transforms from CM
to LAB plottet. Since there is a 1-1 correspondance between
θCM and θLAB it is plotted as a function of both. D: The Jacobian
times sin θ plottet as a function of θLAB . . . . . . . . . . . . . 9

3.1 Schematic overview of ISOLDE. 2 . . . . . . . . . . . . . . . . 15


3.2 Left: TPROTON spectrum for 9 Li beam on deuterated plastic
in IS561A. Right: TEBIS spectrum for 9 Li beam on deuterated
plastic in IS561A. . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3 The angle, θ, between the normal of the detector and the particle
determines the actual length teff the particles travel in the detector. 19

2
FiXme Fatal: Tegn ind hvor eksperimental setup er!

137
138 List of Figures

3.4 Energy correlation plots for a silicon telescope. In (A) we see the
energy deposited in the thin detector versus the energy in the
back detector. In (B) we have made the geometrical correction
described in 3.4 to obtain a better resolution. . . . . . . . . . . 19

4.1 The setup used in the simulations. . . . . . . . . . . . . . . . . 26


4.2 The total position distribution for all concidence measurements. 27
4.3 The total position distribution for all concidence measurements
with a solution to (4.10). . . . . . . . . . . . . . . . . . . . . . 27
4.4 The projection of position distribution for all concidence mea-
surements with a solution to (4.10). . . . . . . . . . . . . . . . 27
4.5 The position distributions for coincidence measurements in the
central W1 and each of the surrounding W1’s. . . . . . . . . . . 30
4.6 Projections of the position distributions for coincidence measure-
ments in the central W1 and each of the surrounding W1’s. . . 31
4.7 The position estimate of a point beam incident in (0mm,0mm). 32
4.8 The position estimate of a point beam incident in (1mm,1mm). 32
4.9 The angle estimate of a point beam incident in (0mm,0mm) but
angled (θx = 35.3 mrad, θy = 35.3 mrad). . . . . . . . . . . . . . 32
4.10 The angle distributions from the new method. There are many
counts in (0,0), meaning the method is rather insensitive to
coincidences angled in the reaction plane. This means we should
use the angle relative to the inferred reaction plane, i.e. θx . . . 33
4.11 Projection of the angle distributions from the new method. The
x-projection shifts, and can thus be used to estimate the total
angle of the beam. . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.12 The energy estimates for a) a point beam b) a gaussian en-
ergy distribution with σ = 0.25 MeV and c) a gaussian energy
distribution with σ = 1 MeV . . . . . . . . . . . . . . . . . . . . 35
(a) Point beam . . . . . . . . . . . . . . . . . . . . . . . . 35
(b) σ = 0.25 MeV . . . . . . . . . . . . . . . . . . . . . . . 35
(c) σ = 1 MeV . . . . . . . . . . . . . . . . . . . . . . . . 35
List of Figures 139

4.13 The projection of the angle distributions using the old method.
We see that the x-projection shifts and we can thus use that for
our angle estimate. . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.14 We can look a narrower slices by only considering the center half
(the red and green area) or the center fourth (the green area). . 38
4.15 Position and angle estimate of a beam which was both displaced
and angled. We do not recreate the correct values. The method
are thus biased when both effect are present. . . . . . . . . . . . 39
(a) Position . . . . . . . . . . . . . . . . . . . . . . . . . . 39
(b) Angle . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.16 Position and angle distributions for a point beam and a spatial
uniformly distributed beam with a radius of 2.5 mm. . . . . . . . 40
(a) Point beam: Position . . . . . . . . . . . . . . . . . . 40
(b) Broad beam: Position . . . . . . . . . . . . . . . . . . 40
(c) Point beam: Angle . . . . . . . . . . . . . . . . . . . . 40
(d) Broad beam: Angle . . . . . . . . . . . . . . . . . . . 40

5.1 Geometry for a round DSSD and a square DSSD. . . . . . . . . 44


5.2 Detector hierachy in AUSAlib. . . . . . . . . . . . . . . . . . . . 45
5.3 "" . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.4 "" . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

6.1 Figure by M.Munch . . . . . . . . . . . . . . . . . . . . . . . . 54


6.2 Figure by M.Munch . . . . . . . . . . . . . . . . . . . . . . . . 55
6.3 Figure by M.Munch . . . . . . . . . . . . . . . . . . . . . . . . 58
6.4 MMh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
6.5 TEBIS spectrum from a down stream detector in IS561B. The
time period between the peaks corresponds to the effective DT of
the DAQ. The main contribution is the GATE and the conversion
time. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.6 DAQC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.7 RunDB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
6.8 Grafana . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
140 List of Figures

9.1 Left:Schematic representation of the setup from IS367. The


setup consists of two telescope configurations covering the forward
direction in the laboratory. Right: ∆E-E plot for both detectors. 81
9.2 Upper panel: Kinematic curves for identified protons, deuterons
and tritons. Kinematic curves for reactions with different excita-
tion energies of the recoil are also shown. Lower panel: Excita-
tions spectra for the different particles, assuming 9 Li (d,p)10 Li,
9 Li (d,d)9 Li and 9 Li (d,t)10 Li respectively. . . . . . . . . . . . 83
9.3 Left: Kinematic curve of Li scattering off of silver. Right:
9

Coincident deuterons and 9 Li with the kinematic curves corre-


sponding to the ground state of 9 Li. . . . . . . . . . . . . . . . 84
9.4 A: Excitations spectra for protons for the main target (CD2 )
and the two background targets (CH2 and C). B: The combined
background on top of the signal. C: The resulting background
subtracted spectrum. . . . . . . . . . . . . . . . . . . . . . . . 85
9.5 Intensity measurements throughout the experiment. Notice the
increase in intensity after measurement 180, due to a reoptimiza-
tion of the beam by the operators. . . . . . . . . . . . . . . . . 88
9.6 Particle per incident 9 Li for d and α respectively. . . . . . . . 89

10.1 Left: Setup Right: Banana plot . . . . . . . . . . . . . . . . . 93


10.2 Erik . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
10.3 Results from the backward telescope. (A): The ∆E−E spectrum.
(B): The kinematic spectrum for events that stops in the DSSD
with 9 Li beam. (C): The kinematic spectrum for events that
stops in the DSSD with 12 C beam. . . . . . . . . . . . . . . . . 98
10.4 Kinematic spectrum for elastic scattered 9 Li off of a gold target.
The kinematic curve has been corrected for energy loss in the
target. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

11.1 The target ladder slot with the 3 H foil. The three regions we
investigated is indicated. . . . . . . . . . . . . . . . . . . . . . . 103
11.2 Kinematic curves for 9 Li beam on the main target. The red
curves corresponds to elastic protons. . . . . . . . . . . . . . . 104
List of Figures 141

11.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
11.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
11.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
11.6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
11.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
11.8 Excitation spectra from the identified protons, deuterons and
tritons. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
11.9 Calculated intensity based on 9 Li scattered off of 48 Ti. The data
has been fitted to an exponential fit and extrapolated to the
grazing angle at 9.1◦ . . . . . . . . . . . . . . . . . . . . . . . . . 113

12.1 Angular cross section for elastic deuterons at three different


energies. Data for 2.68 MeV/A is shown in (a) together one
parametrizations of the optical potentials as well as two fits from
[Jeppesen]. The last curve corresponds to the a fit to the data.
In (b) we see the data for 6.72 MeV/A with three fits to the
optical potentials. In (c) we see the data for 8.0 MeV/A with two
fits. P1 us fitted to the full dataset whereas P2 is fittede to the
first feature below 60◦ . The parameters used in the calculations
are listed in 12.1. . . . . . . . . . . . . . . . . . . . . . . . . . . 118
12.2 Angular cross section for elastic protons at three different ener-
gies. Data for 2.68 MeV/A is shown in (a) together with different
parametrizations of the optical potentials. The last curve corre-
sponds to the a fit to the data. In (b) and (c) we see the data
for 6.72 MeV/A and 8.0 MeV/A respectively together with two
parametrizations of the optical potentials. The parameters used
in the calculations are listed in 12.1. . . . . . . . . . . . . . . . 119
12.3 Angular cross section for elastic deutrons calculated with CDCC.
Data for 2.68 MeV/A is shown in (a) together with two parametriza-
tions of the optical potentials for 9 Li-n and 9 Li-p. The KD+Fit
uses the 9 Li-p that was fitted to elastic protons. In (b) and
(c) we see two parametrizations of the 9 Li-n and 9 Li-p optical
potentials. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
142 List of Figures

12.4 Acceptance for 9 Li (d,p)10 Li at 2.68 MeV/A 6.72 MeV/A and


8.0 MeV/A respectively. Notice lines in the two first panels.
They corresponds to dead strips of the detectors. . . . . . . . . 124
12.5 Acceptance corrections for 2.68 MeV/A for both angular and
energy differential cross sections. . . . . . . . . . . . . . . . . . 126
12.6 Differential cross for 2.68 MeV/A (A), 6.72 MeV/A (B) and 8.0 MeV/A
(C). Right panel shows the angular cross section and the left
panel shows the energy cross section. The title of each panel
indicates the integration range in the other variable. . . . . . . 127
12.7 Differential cross section for angular (left panel) and energy (right
panel) bins for 2.68 MeV/A. The distributions seems very similar
across both axes, indicating that we are, in fact, not measuring
a direct reaction. . . . . . . . . . . . . . . . . . . . . . . . . . . 130
12.8 Differential cross section for angular (left panel) and energy (right
panel) bins for 6.72 MeV/A. . . . . . . . . . . . . . . . . . . . . 131
12.9 Differential cross section for angular (left panel) and energy (right
panel) bins for 8.0 MeV/A. . . . . . . . . . . . . . . . . . . . . 132
12.10Differential cross section for 10 Li for singles and coincidens mea-
surements. The singles spectrum is only based on protons. The
second dataset (squares) are coincidences with the acceptance
of C. The last dataset (triangles) are coincidences but with the
acceptance of 10 Li breakup A. . . . . . . . . . . . . . . . . . . 133
12.11Acceptance for angle correlations between the neutron and either
the ejected proton or the ejected 9 Li based on different simulated
kinematic situations. The letters correspond to the situations
outlined in the text. The left column is the θn versus θp and the
right columns is θn versus θ9 Li . All angles are in center of mass. 136
Abbreviations

143
145

Rettelser
Note: Find reference!!! . . . . . . . . . . . . . . . . . . . . . . . . . . 14
Fatal: Tegn ind hvor eksperimental setup er! . . . . . . . . . . . . . 15
Fatal: MMH!! . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
Fatal: Check det her . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
Fatal: En figur? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
Fatal: Tegn ind hvor eksperimental setup er! . . . . . . . . . . . . . 137

You might also like