Вы находитесь на странице: 1из 21

Alpha-Gamma Correlations

Reiss Moore (in collaboration with Luke Arnell)

Abstract
By looking at the correlations between the alpha and gamma decays in heavy nuclei
it was possible to deduce a number of their nuclear properties. Such nuclei used were
americium-241 and thorium-228. The lifetime of the intermediate state of the most
common correlated alpha-gamma decay in Am-241 was determined by using a fast
timing circuit. This resulted in data representing a series of exponentially decaying
gaussian curves from which we could deduce a half-life by optimising a convolution to
the raw data. A half-life of 70.7 6.5 ns was measured. By also observing the angular correlation of the same decays, the spin of the intermediate state was determined,
although results were inconclusive. Finally the coincidences of alpha and gamma transitions were observed in Th-228. By comparing coincidence energy spectra with the
general energy spectrum of Th-228, combined with measurements at different angles,
it is possible to observe the relaxation of spin alignment with time, although these
results were also deemed inconclusive.

Introduction

For a long time there have been a variety of ways to determine the quantum mechanical
properties of many heavy nuclei. Many of the older methods prove to be just as effective
today. By looking at the correlated alpha and gamma decays of heavy nuclei, it is possible
to determine properties such as the lifetime of particular states and their spin.1 This
experiment does exactly that by observing the time between correlated alpha and gamma
decays to measure the lifetime of the intermediate state that the alpha decay populates.
Triggered by the initial alpha decay, a timing circuit is used to detect the number of
correlated decays for different time intervals. A similar timing circuit is used to also look
at the angular correlation between these transitions. By fixing the axis of an initial alpha
decay, a subsequent gamma decay has some angular distribution, which can be observed by
recording the same information required to determine the lifetime of the intermediate state,
but at different angles between the respective alpha and gamma detectors. This allows
determination of the gamma transition character. Finally, coincidence measurements were
used to look at spin relaxation. Details will be given on all of the equipment used, how it
was used, and the results and discussion of the experiment.

2
2.1

Theory
Alpha-Gamma Correlation

When a nucleus decays by alpha emission, it can populate many different states as well as
the ground state, in a newly created excited nucleus, with each emission having a different
decay probability. The daughter states are at a higher energy than the ground state, so
in order to reach the ground state gamma emission will also occur via a number of lower
energy states, or often straight to the ground state. By combining a charged particle
detector with a gamma ray detector, it is possible to detect the gamma ray that is the
result of a particular alpha decay.
+

Americium-241 ( 25 ) decays by alpha emission, with the majority of decays populating the

excited states of neptunium-237. The most prominent of these is the decay to the 25 59.54
keV level, at a probability of 84.5%. Subsequently, gamma radiation is emitted in order to
+
reach the 52 ground state from the 59.54 keV level. These two decays can be described as
correlated. The intermediate state has a half-life of 67 ns. Similarly, thorium-228 decays
totally by alpha emission, populating the excited states of radium-224, which itself alpha
decays to the ground and 240.99 keV levels of radon-220. The alpha emission of Th-228 is
primarily to the ground state (73.4%), with 26% of the decays to the 84.37 keV state of Ra224 that in turn gamma decays to the ground state. Although 94.7% of the alpha decays

of Ra-224 are to the ground state, 5.25% are to the 240.99 keV level. Th-228 therefore
shows two significant correlated alpha-gamma decays that can readily be observed.

2.2

Lifetime

The lifetime of a particular populated state is determined by Heisenbergs uncertainty


principle:
~
Et
(1)
2
where E is the uncertainty in the states energy, and t is the uncertainty in the states
associated time. The uncertainty in energy is representative of the energy width of the
state, , which itself is related to the mean lifetime of the state, :
E =

~
=
2
2

(2)

Rearranging this relationship, we can see the equation used to measure the lifetime of
excited states:
~
=
(3)

In alpha decay, the associated state lifetimes are relatively long when compared to gamma
ray transitions that tend to occur for short-lived states.2 For these shorter lifetimes it is
not possible to calculate the lifetime from the exponential behaviour of the activity as a
function of time. By measuring the time interval between the formation of an intermediate
state and its subsequent decay instead, we get the same exponentially decaying behaviour
for the survival probability of a state, and we see the second decay transition shortly after
the intermediate state is formed.3 By comparing the time interval and the number of
counts for each time interval, a graph similar to that in Figure 1 can be produced.
The graph allows for the calculation of the mean lifetime of an intermediate nuclear state.
It represents a sum of exponentially decaying Gaussian peaks, and obeys the following
equation:
X
(t t0 )2
N (t) = A + B
exp t0 exp
(4)
2
0
t

where N is the number of counts, A represents the background radiation, B is some


normalisation factor, is the state decay constant, t0 represents a point along the curve,
t is the time of the interval mentioned previously and is the time resolution. From the
decay constant it is possible to calculate the mean lifetime of the nuclear state, as seen in
Equation 5:
1
(5)
=

Figure 1: An example of a plot for the time interval (x-axis) vs. the number of counts per each time
interval (y-axis). The time resolution shown should be less than or similar to the state half-life.

The state lifetimes are often quoted as half-lives, t1/2 , so the following equation can be
used for conversion between the two:
t1/2 = ln2 =

2.3

ln2

(6)

Angular Correlation

For gamma radiation, it is possible to work out the multipole order of a particular transition
by looking at the angular correlation between two subsequent decays.4 For a situation
where there exists an equal substate population (e.g in a large sample of nuclei), when
we average over the angular distribution we find that it is isotropic for a single transition.
For a nucleus that decays by two successive radiations to the ground state, however, when
the first decay is observed it implies that the angular momentum of the nuclear state
between the two has some orientation. As such, the following decay will have some angular
distribution about the initial decays direction.5
The angular correlation between the alpha and gamma decays, W(), can be represented
by:
X
W () =
Ak Pk (cos)
(7)
k=0

where k = 0, 2, ..., kmax (i.e even due to conservation of parity), is the angle between the
two decays, and Pk cos() are the Legendre polynomials. Ak accounts for the multiplicity of
the observed radiation and the angular momenta that the involved nuclear levels possess. If
we introduce conditions that result in unequal populations of the nuclear states, W() is no
longer constant. One such condition is when a strong magnetic field and low temperatures
4

are introduced. A reduction in temperature causes the populations to become unequal as a


direct result of the Boltzmann distribution.6 Another way is to look at the two successive
radiations and observe the angular distribution of the second around some axis defined by
the first transition. This definition of a fixed axis means that for the first transition =0,
and as a result a decay between states of the same angular momentum cannot occur in that
particular direction.7 The resulting secondary transitions have a non-isotropic distribution
in order to conserve angular momentum, a result of the spin orientations in the nucleus.
The shape of the correlation function is determined by these spins.8 As such it is possible
to deduce the spins of the states involved in the correlated transitions from the angular
correlation function.

2.4

Spin Relaxation

Spin relaxation occurs when spins interact with the surrounding thermal environment to
restore equilibrium. It can be split into two types: spin-lattice (longitudinal) relaxation,
and spin-spin (transverse) relaxation, each of which has a respective time constant. The
former relates to the Boltzmann distribution of the spin populations, and the latter is
related to the decay of equal spin state populations, which form spin coherences.9 To
observe both of these, similar methods to those in Section 2.3 can be used. The creation
of unequal substate populations produces a non-zero net magnetisation of the spins in
an excited nucleus. This magnetisation will decay over the same time constant described
above, which defines the length of time it takes for the spin orientations to return to the
equilibrium state.10

Equipment

This section will introduce and explain the equipment that has been used to carry out this
experiment.

3.1

Surface Barrier Detector

Silicon diode detectors are used to detect charged particles due to their availability and
operation at normal temperature (some other semiconductor detectors require cooling e.g.
HPGe). A surface barrier detector is an example that, in this case, contains a silicon PIN
diode. A PIN diode contains a p-type region, an n-type region, and an intrinsic layer
between the two. The p- and n-type regions are heavily doped, creating an acceptor level
near the valence band, and a donor level near the conduction band respectively. This
creates a large potential barrier between them, giving a depletion region that is depleted of
5

electrons and holes, used for the detection of charged particles. The intrinsic layer is lightly
doped, usually with silicon that is either slightly n- or p-type. PIN diodes are favoured over
PN junctions due to the intrinsic region providing a larger volume over which electron-hole
pairs can be produced, as well as offering a low capacitance and high breakdown voltage.11
By altering the size of the intrinsic region, once can change the capacitance and quantum
efficiency of the diode. The signal produced is a quantity of charge, Q, determined by the
following equation:
E e 106
Q=
(8)

where E is the energy if the alpha radiation (in MeV), e is the electronic charge and
is the energy needed to create an electron-hole pair. It is important that a high-voltage
reverse bias is applied to the detector. The reverse bias acts to increase the width of the
depletion region through to the underlying layer, making it the same width as the intrinsic
region.12 Any bias applied must not be too high as this will deplete the PIN junction of
all charge carriers.13
Radiation damage can limit the operation of any semiconductor detector. Incoming radiation can give rise to defects in the detector material, which allows different levels to
form that can trap decays and limit the semiconductors charge carriers. As such energy
resolution is decreased with the possibility of leakage current increasing.14

3.2

Plastic Scintillation Detector

Figure 2: Schematic of a scintillation detector.

Plastic scintillation detectors are most commonly used to detect gamma rays. A schematic
of a typical detector can be seen in Figure 2. When gamma radiation enters the scintillator,
their energy is absorbed by exciting the scintillator molecules. As these states decay back
to the ground state fluorescence occurs, which is then transmitted to a photomultiplier
tube (PMT) that multiplies the current produced by the fluorescence for suitable detection
and analysis. It is important to note that all of the fluorescence transitions occur at
a lower energy than that required to initially excite the electronic configuration. This
6

means that there is little, to no self-absorption of the produced fluorescence. Generally


scintillation detectors have a fast response time, making them excellent for the study of
different gamma decays that occur within short time periods of one another. Plastic is
often used for due to its flexibility, fast signal processing and high light output.15 A
plastic that has cross sections of the photoelectric and pair production effects larger than
that of Compton scattering should be used to maximise efficiency, since gamma rays are
not totally absorbed when they undergo Compton scattering, giving rise to detections at
different energies to those true to the nucleus begin studied.16
In order to discriminate between different radiation types, a scintillator relies on its dependence on the energy loss of particles, dE/dx, as they enter the material. The light emission
that occurs in a scintillator can be split into relative fast and slow decay components, each
dependent on dE/dx, which determine the decay time of the light emission. This provides
the discrimination between different radiation types, since different radiations have different
dE/dx, and therefore different ratios between the fast and slow components. For plastic,
a particle that goes through a large energy loss produces excited molecules at a high density. This results in more intermolecular interactions, and radiative states produced (fast
component) can lose their energy in a number of ways. The slow component is represented
by metastable excited states. For gamma rays, the fast component is significantly larger
than the slow component.17
The primary components of a PMT are a photocathode, a dynode and an anode.18 The
photocathode converts incident light from the scintillator into electrons by transferring the
energy of a photon, h, to an electron, which drifts and escapes from the photocathode
surface if h is greater than the work function of the photocathode material. Acceleration of
the electrons is then applied by an electrode, directing them to strike the dynode. As they
do, secondary electrons are released that are in turn accelerated to another dynode so more
electrons are released, and this continues until they reach the anode. An amplified current is
produced for analysis. Mu-metal shielding protects the PMT from low frequency magnetic
fields that may influence the low energy electrons that travel through the PMT system.19
Typically, NaI is used as the scintillator material, since it easily grown to large sizes and
shapes. It also has a relatively large atomic number, giving a reasonable interaction cross
section. A dopant of thallium is often used to provide wavelength shift, which reduces the
likelihood of excitation emissions being reabsorbed.20
The transmission of light between a scintillator and PMT must be maximised to maintain
efficiency. As such, silicon is used as the contact for organic scintillators and a PMT since
the refractive indices of all three are similar.21 For timing measurements, the PMT is
held at a high negative voltage. The size of the voltage determines the gain of the signal
produced.

3.3

Charge Sensitive Preamplifier

Signals from semiconductor detectors are generally weak, so they are amplified by preamplifiers. Charge sensitive preamplifiers have a lack of sensitivity to varied capacitances
associated with temperature changes making them ideal for use with semiconductor detectors.22 This can be demonstrated by looking at a schematic of the charge sensitive
preamplifier circuitry in Figure 3.

Figure 3: Charge-sensitive preamplifier schematic.

The governing equation is simply related to the capacitance of the preamplifier, Cf :


V0 =

Q
Cf

(9)

where V0 is the output voltage, and Q is the charge from the surface barrier detector. As
such there is no dependence on detector capacitance. The preamplifier collects charge on
the capacitor Cf and converts it to a voltage. Combining Equations 8 and 9, a relationship
between the incoming charged particle energy and voltage produced as a result can be seen
in the following equation:
E e 106
V0 =
(10)
Cf
The amplitude of this voltage is therefore directly proportional to the incoming alpha
energy, and as such one can then either view the output on an oscilloscope or amplify
the signal for further uses. Equation 11 shows the equivalent noise charge, EN C, for a
preamplifier. Noise must be limited to maximise detector resolution.
Vrms
EN C = e
C
(11)
w
where C is the total detector and preamplifier capacitance, w is the energy required to
produce an electron-hole pair, and Vrms is the output voltage noise level.
Bias for the detector is supplied via the preamplifier. A connection between the preamplifier
and detector facilitates the voltage supply to the detector plus the signal from the detector
to the input of the preamplifier.23
8

3.4

Fast Timing Amplifier

For timing experiments, a fast timing amplifier must be used. It is an example of a


Nuclear Instrument Module (NIM), which also applies to the remainder of equipment to be
explained. Timing amplifiers act to reproduce the incoming signal with a higher amplitude
whilst retaining the primary shape. The rise times of the signal produced are in the low
or sub-nanosecond region, making them perfect for timing measurements. The resulting
pulse also has its polarity reversed, although a switch can provide the opposite. Other
controls on a timing amplifier include gain adjustment, and integration and differentiation
time constant selection. The former allows the size of the incoming signal to be adjusted
on a coarse or fine scale, and the latter provide pulse shaping to optimise the signal for
timing measurements

3.5

Slow Spectroscopy Amplifier

Slow spectroscopy amplifiers, or linear amplifiers, provide optimisation of the size and shape
of incoming signals for energy spectroscopy, whilst achieving a high energy resolution.
Typically the input is a preamplifier signal, which can often be irregular. Pile-up can
occur whereby signals occur before the tail of an initial signal has reached the baseline,
thus altering the amplitudes of the signal. A shaping time can be defined to ensure that
the tail is shortened and this overlapping effect does not occur.24

3.6

Discriminators: Leading Edge and Constant Fraction

Fast timing discriminators generate an output dependent upon whether or not we are interested in a certain signal that the discriminator receives. For example, it will only output a
signal when the alpha decay signal from a fast timing amplifier is above a chosen threshold
value25 . A logic pulse is then created that is related in time to an event occurrence. The
timing resolution is affected by three main elements. The first is walk, where variations in
the incoming pulses shape and amplitude cause the output pulse to move from the pick-off
element, relative to the input pulse. The second is drift, a result of temperature variations
and component degradation, giving a long-term timing error. The third is jitter, caused by
noise and statistical fluctuations that increase the timing uncertainty of the pick-up signal.
Both the jitter and walk can be seen in Figure 4.26
Leading edge discriminators look at the leading edge of a signal and check when it crosses
a particular threshold level. Once the threshold is reached the output logic pulse is given.
The threshold is adjusted via a potentiometer. The main problem, however, is that walk
is very prominent since only the leading edge is used. Frequent changes in the pulse
amplitudes coming into the discriminator mean that the corresponding outputs occur at
9

Figure 4: Graphic of the jitter and walk effects in discriminators.

different times. For detectors that produce a large number of signals of varying amplitudes,
the result is poor time resolution and a lack of accuracy in timing measurements.
To overcome this, a constant fraction discriminator (CFD) can be used. Instead of just
using the leading edge of the incoming signal, a CFD takes the whole signal and separates
it, with one part being reduced in amplitude by a factor f, and the other being inverted and
delayed. The sum of the two is then taken to produce the constant fraction signal, with
a zero-crossing time corresponding to the time at which the pulse reaches the fraction, f
(same as the amplitude reduction factor), of its final amplitude. The zero-crossing time and
pulse amplitude are independent, and so the crossing gives a time marker at the optimum
pulse height fraction.27 As such, there is little walk with a CFD and the timing resolution
is greatly improved when jitter is reduced, making them much more suitable for timing
experiments. A CFD unit will have an input for a signal from a timing amplifier, two
inputs to introduce the delay that is required for a CFD to work, a monitor output that
shows the combination of the two initial, and then altered split signals, and a number of
outputs for use with further equipment. Also included are three potentiometers: the T
screw provides adjustment of the threshold level, the Z screw provides adjustment of the
walk, and the W screw allows the output width to be adjusted.
In order to correctly set a leading edge discriminator, the threshold level should be set as
low as possible so as to minimise walk, but also at a steep slope to ensure that jitter is also
minimised. To optimise a CFD, first the threshold should be adjusted to minimise the noise
and low-energy pulses. Next, the walk should be optimised by viewing the signal from the
CFD monitor output and adjusting the walk potentiometer so that all of the signals that
are seen have the same zero-crossing time28 . Figure 5 shows the ideal threshold and walk
adjustments.

10

(a)

(b)

Figure 5: CFD optimisation, where (a) shows the ideal threshold adjustment, and (b) shows the
ideal walk adjustment.

3.7

Time-to-Amplitude Convertor (TAC)

A TAC is a timing module that outputs a pulse with an amplitude that is proportional to
the time interval that it records.29 A start pulse triggers the TAC to start, for example the
CFD output from an alpha detector, which is then stopped by another signal, for example
the CFD output from a related gamma detector. For recording decay correlations, it is
important that a delay is applied to the second decay to ensure that the TAC is stopped as
a direct result of the same decay that triggered the TAC to start. The primary control on
a TAC is the range selection, which is used to select the time interval between the signals
that start and stop the unit. Given options of 50, 100 and 200 ns, a multiplier allows
further range selection. A range should be selected that is related to the time intervals
that will be measured.

3.8

Coincidence Unit

A coincidence unit requires at least two signal inputs from a discriminator. Provided the
incoming signals are of a good width (roughly 100 ns), and they overlap when the rise time
of the second input is more than 5 ns after the first signal, the coincidence condition is
met (see Figure 6).30 The result is a logic pulse, which indicates that a coincidence has
occurred.

11

(b)
(a)
Figure 6: (a) shows the steps and signals required to correctly calibrate the coincidence unit. (b) is
an example of the coincidence output signal.

3.9

Gate and Delay Generator

Gate and delay generators produce a logic pulse with adjustable width and delay. They are
triggered by an incoming pulse, giving a logic output that is wider than typical logic pulses
making them useful for event windowing. The output can be used as a gating signal for
an MCA (see following section) when combined with the signal from a spectroscopy amplifier.31 The gating signal should overlap the signal from the spectroscopy amplifier.

3.10

Multichannel Analyser (MCA)

An MCA is an Analog-to-Digital Convertor (ADC) that stores the information in its internal memory, based on a count-by-count system. An input pulse has some amplitude and
the ADC converts this amplitude into a digital number. The number of pulses that the
analyser records over a given time gives the total number of counts over all memory. By
connecting it to a computer it is possible to record quantitative information from detector
set-ups and analyse the recordings using MCA emulation software. It is also possible to
gate the MCA, which prevents certain pulses from being recorded by the ADC when it is
in the process of recording a previous pulse. A logic signal holds the input gate open when
the ADC is not busy.32
For timing measurements the MCA should be calibrated so that the relationship between
the channel number and the time interval recorded is known.

3.11

Arrangement

The NIM modules are slotted into a NIM bin that supplies them with power. Different
NIMs require different voltages, defined by the pin arrangements of the connectors on
12

the back of the modules. Depending on the function of each module and the desired
arrangement, they can be connected together by using coaxial cables, which are suitable
to carry the signals involved in typical nuclear experimental systems.

Method

Initially it was essential to determine the correct bias that should be applied to the charge
sensitive pre-amplifier that was connected to the surface barrier detector. This was simply
done by measuring the background fluctuations and the amplitude of the signal produced
as the bias voltage is incrementally increased. By striking a balance between the two, the
optimal bias voltage can be found. The MCA was also calibrated by comparing an initial
recording from the TAC, with a delayed version of the original. The shift in the peak
position corresponds to the delay time used and as such the number of nanoseconds per
channel can be deduced.

PMT

Timing Amp

Discriminator

Source

Detector

Delay

TAC

Preamplifier

Timing Amp

MCA

Discriminator

Figure 7: Schematic of the first and second experiments.

To measure the lifetime of Am-241, it is possible to look at the coincidence between two
decays linked as described in the Section 2.1. The surface barrier detector, which has to
be connected to a charge sensitive preamplifier first, and PMT are each connected to a
fast timing amplifier. These signals are then connected to a discriminator each. Initially
they were connected to leading edge discriminators, but for reasons explained in Section
3.6 CFDs were used for the remainder of the experiment. The gamma signal must then
be delayed to ensure that the gamma decay that is definitely detected after the initial
13

alpha decay. By using a TAC, it is possible to measure the difference in time at which the
two decays occur. The TAC is started when an alpha decay to the intermediate state is
detected, and subsequently stopped when the correlated gamma decay to the ground state
is detected. By running the TAC signal through an MCA for a chosen amount of time
(the longer the time, the better the results), a distribution of the counts against time can
be analysed to determine the lifetime of the state (as discussed in Section 2.2). The raw
data given by the MCA can be interpreted using the relationship in Equation 4. Initially
a rough fit of the equation to the data can be applied, with a code designed to apply a
chi-squared analysis used to minimise the equation to the raw data by changing the A, B,
and components of Equation 4.
To determine the best threshold for the leading edge discriminators, recordings were taken
for different settings and the results compared by looking at the overall shape and resolution
of each. The same method was used to optimise the settings of the constant fraction
discriminators, although more data was needed due to there being more settings to control
with the CFDs. A combination of leading edge and CFDs was also used to provide a
middle-ground comparison between the two types of discriminator.
For the second part of the experiment it was necessary to alter the angle between the two
detectors in order to deduce the spins of the nuclear states involved in the correlated decay
of Am-241. The set-up was exactly as it was in the first part of the experiment, with angles
of 180 , 135 and 90 between the two detectors being measured. It was important that
the experiment for each angle was run for a lengthy amount of time to ensure that suitable
comparison could be achieved between the different angles, since the data created for each
can be very similar.
The third part of the experiment involved looking at the coincidences of the alpha and
gamma decays at different angles, from which the spin relaxation time can be deduced. In
this case Th-228 was used since it is easier to differentiate between the singles and coincidence conditions than with Am-241. All of the equipment previously used is kept the same
except for the TAC and delay units. By first tuning the CFDs for the alpha and gamma
decays from the Th-228 source, the outputs were connected to a coincidence unit, which
in turn was connected to a Gate Delay Generator. This signal was then used to gate the
ADC within the MCA unit, with a signal from a slow spectroscopy amplifier attached to
the alpha detector used as the ADC input. A schematic is shown in Figure 8.

14

PMT

Timing Amp

Discriminator

Gate
Source

Detector

Coincidence

Preamplifier

Timing Amp

GDG

MCA

Discriminator

Spec. Amp

ADC input

Figure 8: Schematic of the third part of the experiment.

Results and Discussion

Biases of 80.5 V and 1.33 kV were used for the alpha and gamma detectors respectively.
An initial calibration of the MCA showed that each channel number had to be multiplied
by 0.0265 to convert the scale to time.

5.1

Lifetime Measurement

Setting the gain of the timing amplifiers at 20.6 and 20.6, the integration time at 20 ns
and 20 ns, and the differentiation time at 100 ns and 500 ns for the alpha and gamma
detectors respectively, with the TAC set at a range of 200 ns and a delay of 48 ns applied
to the gamma signal, the leading edge discriminator was set-up using the steps described
in Section 3.6, with it optimised as described in Section 4. An optimum threshold of 0.7
V was deduced for both the alpha and gamma detectors. The experiment was run for 24
hours and the results show in Figure 9.
Minimising the fit of Equation 4 to the data, values of 1.52 0.02, 0.054 0.001, 0.012
0.001 and 5.54 0.08 for A, B, and respectively were produced, with the red line

15

Figure 9: Leading edge lifetime measurement. The blue lines show the measured data, and the red
line represents the minimised fit to the data.

showing the optimised fit. The errors were produced by the applied minimising function in
the form of an error matrix. By using Equation 6, a half-life of 57.8 7.2 ns was calculated,
with the uncertainty propagated form the error produced on . The actual lifetime of the
measured state is given in Section 2.2. Comparing the two we see a percentage difference
of 13.7%. The actual value does not lie within the error range and this can be accounted
for by the low time resolution of the leading edge discriminators.
CFDs were calibrated for the alpha and gamma detectors, with threshold values of 1.5 V
and 1.5 V, and walk values of 23 mV and 67 mV respectively determined. All other settings
remained the same, apart from the range, which was 500 ns. The results are shown in Figure
10. Values of 7.15 0.05, 1.39 0.02, 0.098 0.001, and 1.31 0.02 were deduced for
A, B, and respectively. The resolution is markedly improved when comparing the
graphs and values, with a steeper rise shown in Figure 10 and CF D < LE . CFDs are
clearly more suitable for timing measurements. Equation 6 gives a half-life of 70.7 6.5
ns, which is an improvement in comparison to the leading edge results. This clearly shows
that CFDs are the best discriminators to use when measuring lifetimes of transition states
due to their superior time resolution.
To confirm the resolution improvement, a leading edge was used for the gamma detector,
and a CFD for the alpha detector, with all other settings as they were. The fit produced
a value of 3.24 for , which clearly lies between the values found for the leading edge and
CFD measurements.

16

Figure 10: CFD lifetime measurement. The blue lines show the measured data, and the red line
represents the minimised fit to the data.

5.2

Angular Correlation Measurements

To compare the results of the angular correlation measurements, they had to be normalised
for the number of counts. The data for the 180 measurement gave the most counts due
to the lack of attenuation, and so the 135 and 90 data was normalised to match the
counts of the 180 data by adjusting the normalisation factor B. The results are shown in
Figure 11. After normalising the 135 and 90 data, the results from the 90 arrangement
clearly do not line up with the rest of the data. This is a result of attenuation. At angles
of 180 and 135 , the emitted gamma rays have to propagate through less of the material
containing the source. The time taken for the gamma rays in the 90 arrangement to reach
the detector is therefore increased, which we see as a reduced number of counts as the
time interval between the correlated decays is increased. To correct this, the data was
normalised relative to the other graphs, with the corrected data shown in Figure 12. The
measured
number of counts, N , is subject to some statistical uncertainty, determined by

N . Accounting for this, it is clear to see that the results produced


could be a result of

statistical fluctuations, with values of roughly 6 counts given for N at the peaks for each
angle.
It is important to note that the number of counts reduced significantly compared to previous
parts of the experiment, as seen by comparing Figures 10 and 11/12, which were run for a
similar amount of time. The distance between the two detectors was adjusted but the count
rate was not improved by brining them closer together. There is a possibility of equipment
degradation and it would have been useful to analyse the resolution and efficiency of both

17

Figure 11: Angular correlation measurement, comparing the data for different angles between the
two detectors. The data has been normalised for the number of counts at each angle.

Figure 12: The normalisation to remove attenuation effects seen initially seen in Figure 11.

18

detectors to determine if the problem was down to any changes in these at the later stages
of the experiment. The results may have been improved by restoring the apparatus back
to its original count rate. This could be achieved by testing each experimental component
to ensure that it is working correctly and receiving the correct amount of power. At times
there may have been too many NIM modules in one power bin meaning that some modules
may have suffered slight power losses. It may have also been more advantageous to use the
Th-228 source for this part of the experiment since the states involved stay aligned for a
longer period of time that in Am-241. This means it is easier to see discrepancies in the
peaks of the different angles measured.

5.3

Coincidence Measurement

Results for the coincidence measurements produced energy spectra with unresolved peaks
and a large amount of noise. As such, results were deemed inconclusive due to a reduction
in the energy resolution of the alpha detector. The most likely source of this is radiation
damage (Section 3.1). By the time that this part of the experiment was undertaken, the
apparatus had been used for an extensive amount of time. Ideally, the detector would be
replaced between experiments to ensure that the resolution remains as high as possible,
particularly given that it involves the comparison of energy spectra.

Conclusions

Timing measurements were taken to measure the lifetime of the intermediate state of a
correlated alpha-gamma decay, as well as the angular correlation between the two and
the observation of energy spectra in coincidence conditions. The lifetime measured was
comparable to the actual value, but the remaining two parts of the experiment produced
inconclusive results. To improve on this, the equipment should be optimised to allow for
a higher count rate by testing the equipment used. Also, the detector should be replaced
when it becomes heavily radiation damaged, and a source whose intermediate states remain
aligned longer, specifically for the angular correlation measurements.

19

References
1 T.

Hoagland, A Brief Description of Discriminators, Michigan State University (2004).

2 K.

Siegbahn, Alpha, Beta and Gamma Ray Spectroscopy, Elsevier Science (1965).

3 K.

Krane, Introductory Nuclear Physics, Wiley (1987), p.232.

4 Krane,

p.335.

5 Gamma-Gamma
6 Krane,

p.338.

7 Krane,

p.338.

Angular Correlations, Rutgers University (2008).

8 Ortec:

for

AN34 Experiment 19 - Gamma-Ray Decay Scheme and Angular Correlation


AMETEK Corporation.

60 Co,
9 M.H.

10 G.

Levitt, Spin Dynamics: Basics of NMR, Wiley (2013), Chapter 20.

Ehrenstein, H. Lecar, Biophysics, Academic Press (1982), p.14.

11 L.M.

Simms, Hybrid CMOS SiPIN Detectors as Astronomical Imagers, Stanford University (2009), pp.24-25
12 D.R.

Kirkby, A Picosecond Optoelectronic Cross Correlator using a Gain Modulated


Avalanche Photodiode for Measuring the Impulse Response of Tissue, University of London
(1999), p.90.
13 W.R.

Leo, Techniques for Nuclear and Particle Physics Experiments: A How-To Approach, Springer (1994), p.226.
14 Leo,

p.237.

15 Leo,

p.156.

16 Leo,

p.157

17 Leo,

pp.163-164.

18 Leo,

p. 169.

19 G.F.

Knoll, Radiation Detection and Measurement, Wiley (2010), p.274.

20

20 Knoll,

p.246.

21 Leo,

p.194.

22 Leo,

p.236.

23 Knoll,

p.594.

24 Knoll,

p.571

25 Hoagland,
26 Ortec:

p.2.

Fast-Timing Discriminator, AMETEK Corporation, p.3.

27 Fast-Timing

Discriminator, p.4

28 Gamma-Gamma
29 Knoll,
30 622

Coincidence, GSI Helmholtzzentrum f


ur Schwerionenforschung, p.11.

p.600.

Quad 2 Input Logic Input Manual, LeCroy.

31 J.P.

Omtvedt, User Manual for NIM Gate and Delay Generator, University of Oslo
(2010), p.1.
32 Knoll,

p.713.

21

Вам также может понравиться