You are on page 1of 177

Nu

cle
ar
Ph
ys
ics

Manual for M.Sc. (P) Nuclear Physics


Lab.

La

M
an
ua
l

Shobhit Mahajan

February 24, 2016


Version 1.01

Nu
cle
ar
Ph
ys
ics

Contents
3

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

4
4
5
6
6
8
9
10
10
13
13
19
19
24
30
38
39
51
51
53
55
58
60

2 RADIOACTIVITY
2.1 Radioactivity . . . . . . . . . . .
2.1.1 Measure of radioactivity .
2.1.2 Activity Law & Half Life
2.2 Nuclear Decay . . . . . . . . . .
2.2.1 Alpha Decay . . . . . . .
2.2.2 Beta Decay . . . . . . . .
2.2.3 Gamma Decay . . . . . .
2.3 References . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

61
61
61
62
64
65
67
69
71

. . . . .
. . . . .
. . . . .
matter
. . . . .
. . . . .
. . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

72
72
72
74
75
83
87
97

La

M
an
ua
l

1 STATISTICS & ERROR ANALYSIS


1.1 Probability . . . . . . . . . . . . . . . . .
1.1.1 Random Variables . . . . . . . . .
1.2 Uncertainty in Measurement . . . . . . . .
1.2.1 Uncertainity, Accuracy & Precision
1.2.2 Systematic & Random Errors . . .
1.2.3 Significant Digits . . . . . . . . . .
1.3 Statistical Analysis of Data . . . . . . . .
1.3.1 Histograms & Distribution . . . .
1.3.2 Parent & Sample Distribution . . .
1.3.3 Mean & Deviation . . . . . . . . .
1.4 Distributions . . . . . . . . . . . . . . . .
1.4.1 Binomial Distribution . . . . . . .
1.4.2 Poisson Distribution . . . . . . . .
1.4.3 Normal Distribution . . . . . . . .
1.5 Error Estimation . . . . . . . . . . . . . .
1.5.1 Propagation of Errors . . . . . . .
1.6 Estimation and Error of the Mean . . . .
1.6.1 Method of Maximum Likelihood .
1.6.2 Estimated Error in the Mean . . .
1.7 Method of Least Squares . . . . . . . . . .
1.8 Goodness of Fit . . . . . . . . . . . . . . .
1.9 References . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

3 INTERACTION WITH MATTER


3.1 Introduction . . . . . . . . . . . . . . . . . . . . .
3.1.1 Cross Section . . . . . . . . . . . . . . . .
3.2 Interaction of Charged Particles with Matter . .
3.2.1 Interaction of Heavy charged particle with
3.2.2 Interaction with matter of electrons . . .
3.2.3 Interaction of gamma rays with matter . .
3.3 References . . . . . . . . . . . . . . . . . . . . . .

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

98
98
98
102
102
105
106
106
108
110
113
115
117

5 Experiment: GM Characteristics
5.1 Introduction . . . . . . . . . . . . .
5.2 Precautions . . . . . . . . . . . . .
5.2.1 Health Effects of Radiation
5.3 Experiment . . . . . . . . . . . . .
5.3.1 Purpose . . . . . . . . . . .
5.3.2 Method . . . . . . . . . . .
5.3.3 Sample Data . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

118
118
118
118
120
120
121
122

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

7 Experiment: Absorption
7.1 Introduction . . . . . .
7.2 Experiment . . . . . .
7.2.1 Purpose . . . .
7.2.2 Method . . . .
7.2.3 Sample Data .

La

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

for
. .
. .
. .
. .
. .
. .

&
. . .
. . .
. . .
. . .
. . .
. . .

rays
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

133
133
133
133
135
137
141

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

144
144
145
145
145
145

the Inverse Square Law for


. . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . .

.
.
.
.
.

rays
. . . .
. . . .
. . . .
. . . .
. . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

150
150
151
151
151
153

of
. . .
. . .
. . .
. . .
. . .

8 Experiment: Verification of
8.1 Introduction . . . . . . . .
8.2 Experiment . . . . . . . .
8.2.1 Purpose . . . . . .
8.2.2 Method . . . . . .
8.2.3 Sample Data . . .

.
.
.
.
.
.
.

Efficiency
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .

M
an
ua
l

6 Experiment: GM Counter: Counting


6.1 Introduction . . . . . . . . . . . . . .
6.2 Experiment . . . . . . . . . . . . . .
6.2.1 Purpose . . . . . . . . . . . .
6.2.2 Method . . . . . . . . . . . .
6.2.3 Sample Data . . . . . . . . .
6.2.4 Error Estimation . . . . . . .

.
.
.
.
.
.
.

Nu
cle
ar
Ph
ys
ics

4 G-M COUNTER
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . .
4.2 Detector Models . . . . . . . . . . . . . . . . . . . . .
4.3 Ionisation of Gases . . . . . . . . . . . . . . . . . . . .
4.3.1 Townsend Avalanche . . . . . . . . . . . . . . .
4.3.2 Kinds of Detectors & Detector Regions . . . .
4.4 GM Counter . . . . . . . . . . . . . . . . . . . . . . .
4.4.1 Geiger Discharge . . . . . . . . . . . . . . . . .
4.4.2 Quenching . . . . . . . . . . . . . . . . . . . . .
4.4.3 Dead Time & Recovery Time . . . . . . . . . .
4.4.4 Geiger Counting Plateau & Operating Voltage
4.4.5 Counting Efficiency . . . . . . . . . . . . . . .
4.5 References . . . . . . . . . . . . . . . . . . . . . . . . .

rays
. . .
. . .
. . .
. . .
. . .

9 Experiment: To Determine the


Energy
9.1 Introduction . . . . . . . . . . .
9.2 Experiment . . . . . . . . . . .
9.2.1 Purpose . . . . . . . . .
9.2.2 Method . . . . . . . . .
9.2.3 Sample Data . . . . . .

in
. .
. .
. .
. .
. .

Iron
. . .
. . .
. . .
. . .
. . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

Range of rays in Aluminum and to determine the End Point


157
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

10 Experiment: Scintillation Counter


10.1 Introduction . . . . . . . . . . . . .
10.2 Theory . . . . . . . . . . . . . . . .
10.2.1 Inorganic Scintillators . . .
10.2.2 Organic Scintillators . . . .
10.2.3 Photomultiplier Tube . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

165
165
165
166
169
171
174

La

M
an
ua
l

Nu
cle
ar
Ph
ys
ics

Index

.
.
.
.
.

Chapter 1

1.1

Nu
cle
ar
Ph
ys
ics

STATISTICS & ERROR ANALYSIS


Probability

Since we are going to be dealing with probability in the subsequent sections, let us give a brief background. The probability of an event refers to the likelihood that the event will occur. Mathematically,
the probability that an event will occur is expressed as a number between 0 and 1. The sum of probabilities in any statistical experiment is always 1, a statement of the fact that something will certainly
happen. Let us illustrate how one can calculate probabilities.

M
an
ua
l

Consider first the case of an experiment with n possible outcomes which are each equally likely. Now
if we take a subset r of these and call them successes then clearly the probability of success in the
experiment is nr . Thus, if there are 10 balls, 7 white and 3 black in a bag, then the probability of getting
3
a black ball if one is picked out at random from the bag is 10
.

La

There is another approach to probability where one talks about relative frequencies. Suppose I count
the number of cars passing a particular point on a road at a particular interval of time and notice how
many of them are white. Suppose on the first day, I see 5 white cars out of a total of 20 cars, while on
the second day I count 9 white cars out of 30 while on the third day I find 3 white cars out of 5 and so
on. Clearly, the relative frequency of white cars is different on different days. However, one could find
for instance that if I repeat this experiment many many times, then the relative frequency is 0.26. Then
the Law of Large Numbers says that the relative frequency of an event will converge on the probability
of that event as the number of trials increases.
Some definitions in probability theory are useful:

1. Two events are mutually exclusive or disjoint if they cannot occur at the same time.
2. The probability that Event A occurs, given that Event B has occurred, is called a conditional
probability. The conditional probability of Event A, given Event B, is denoted by the symbol

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

P (A|B).
3. The complement of an event is the event not occurring. The probability that Event A will not
occur is denoted by P (A0 ).
4. The probability that Events A and B both occur is the probability of the intersection of A and B.
T
The probability of the intersection of Events A and B is denoted by P (A B). If Events A and B
T
are mutually exclusive, P (A B) = 0.

Nu
cle
ar
Ph
ys
ics

5. The probability that Events A or B occur is the probability of the union of A and B. The probability
S
of the union of Events A and B is denoted by P (A B) .
6. If the occurrence of Event A changes the probability of Event B, then Events A and B are dependent. On the other hand, if the occurrence of Event A does not change the probability of Event
B, then Events A and B are independent.
These definitions allow us to write down the rules for probability.

Subtraction: The probability that event A will occur is equal to 1 minus the probability that event
A will not occur.
P (A) = 1 P (A0 )

M
an
ua
l

Multiplication: The probability that Events A and B both occur is equal to the probability that
Event A occurs times the probability that Event B occurs, given that A has occurred.

P (A

B) = P (A)P (B|A)

La

Addition: The probability that Event A or Event B occurs is equal to the probability that Event
A occurs plus the probability that Event B occurs minus the probability that both Events A and B occur.

P (A

B) = P (A) + P (B) P (A

B)) = P (A) + P (B) P (A)P (B|A)

These rules are fairly obvious intuitively and the easiest way to prove them is to use Venn diagrams
where the results quoted above are immediately clear.

1.1.1

Random Variables

When the value of a variable is determined by a chance event, that variable is called a random variable. Random variables can be discrete or continuous.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Discrete Random Variable

Continuous Random Variable

Nu
cle
ar
Ph
ys
ics

Within a range of numbers, discrete variables can take on only certain values. Suppose, for example,
that we flip a coin and count the number of heads. The number of heads will be a value between 0
and +. Within that range, though, the number of heads can be only certain values. For example,
the number of heads can only be a whole number, not a fraction. Therefore, the number of heads is a
discrete variable. And because the number of heads results from a random process - flipping a coin - it
is a discrete random variable.

1.2
1.2.1

M
an
ua
l

Continuous variables, in contrast, can take on any value within a range of values. For example,
suppose we randomly select an individual from a population. Then, we measure the age of that person.
In theory, his/her age can take on any value between 0 and +, so age is a continuous variable. In
this example, the age of the person selected is determined by a chance event; so, in this example, age
is a continuous random variable.
Note that discrete variables can be finite or infinite. Thus, for instance the number of heads in coin
flips can be infinite while the number of aces that I can choose from a deck of cards is finite (0, 1, 2, 3, 4).
Continuous variables can always take an infinite number of values while some discrete variables can take
infinite number of values.

Uncertainty in Measurement

Uncertainity, Accuracy & Precision

La

All measurements that we do have some degree of uncertainty. This uncertainty might come from a
variety of sources about which we will talk later. But the fact that needs to be emphasised is that all
measurements have some uncertainty and an analysis of this uncertainty is what we call error analysis.
Any measured value that we quote must be accompanied by our estimate of the level of certainty or confidence associated with that measurement. This fact is absolutely essential since without this the basic
question of science, namely does the result of our experiment agree with the theory can not be answered. To decide whether the proposed theory is valid or not, this question would need to be answered.
When we carry out an experiment to measure a quantity, we of course assume that some true or
exact value exists. We of course may or may not know this value but we always attempt to find the best
value possible given the limitations of our own experimental setup. Typically, every time we carry out
the experiment, we will find a different value and so the question is how do we report our best estimate
of this true value? Usually, this is done as

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

measurement = best estimate uncertainty

Nu
cle
ar
Ph
ys
ics

For example, let us assume you want to find the weight of your mobile phone. By simply putting
in your hand, you can estimate it to be between 100 and 200 grams. But that is not good enough. So
you go to a balance in the laboratory and it gives you a reading of 145.55 grams. This value is much
more precise than the original estimate you obtained, but how does one know that it is accurate?
One way is to repeat your measurement several times and suppose you get the values 145.59, 145.53
and 145.51 grams. Then one could say that the weight of the phone is 145.55 .04 grams. But
now suppose you go to another balance and find a value of 144.15 grams? Now one is faced with a
problem since your original best estimate is very different from this measurement. So what does one do?
To understand this, we need to understand first the difference between precision and accuracy .

Precision & Accuracy

Accuracy is how close the measured value is to the true or the accepted value of a
quantity.

M
an
ua
l

Precision is a measure of how well the result can be measured, irrespective of the
true value. It reflects the degree of consistency and agreement between repeated, independent measurements of the same quantity as well as the reproducibility of the results.

La

Any statement of uncertainty associated with a measurement must include factors


which affect both accuracy and precision. After all, it is a waste of time if we determine
a result which is very precise but highly inaccurate or its converse, that is, a result
which is very accurate but highly imprecise.

In our example above, we have no way of knowing whether our result is accurate or not unless we
compare it with a known standard. For instance, we could use a standard weight to determine if the
balances used in our measurement are accurate or not.
Precision is often reported experimentally as relative uncertainty defined as


uncertainty

Relative Uncertainty =
measured value
Thus in our example, the relative uncertainty is
Relative Uncertainty =

0.04
= 0.027%
145.55

(1.1)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Accuracy on the other hand is reported usually as relative error which is defined as
Relative Error =

measured value - expected value


expected value

In our example above, if we think that the expected value is 145.50 grams, then the relative error is
0.05
145.55 145.50
=
= 0.034%
145.50
145.50
Thus we see that any measurement needs to be both precise and accurate for it to be good. The
idea of making good measurements is directly related to errors in measurement. Errors in measurement
can be broadly classified into two categories- random and systematic.

1.2.2

Systematic & Random Errors

Nu
cle
ar
Ph
ys
ics

Relative Error =

Systematic Errors are errors which will make the results obtained by us differ from the true value
of the quantity under consideration. They are reproducible inaccuracies which are difficult to detect and
also cannot be analysed statistically. An important thing to realise is that systematic errors cannot
be reduced by repeated measurements.

M
an
ua
l

Random Errors are errors which are fluctuations in the observations which are statistical in nature.
They can be analysed statistically and furthermore, they can be reduced by repeated measurements
and taking averages as we shall see later.

La

To illustrate this distinction think of the following experiment. Suppose I wish to find the time
period of a pendulum by timing some number of oscillations with the help of a stop watch. There could
be several sources of error- One source of error will be my reaction time, that is the time between my
seeing the pendulum bob reaching the extreme position and my starting the watch and again at a later
point my observing the bob and my stopping the watch. Obviously, if my reaction time was always
exactly the same, then the time delay wouldnt matter since they would cancel. However, we know
that in practice, my reaction time will be different. I may delay more in starting, and so underestimate
the time of a revolution; or I may delay more in stopping, and so overestimate the time. Since either
possibility is equally likely, the sign of the effect is random. If I repeat the measurement several times,
I will sometimes overestimate and sometimes underestimate. Thus, my variable reaction time will show
up as a variation of the answers found. By analyzing the spread in results statistically, I can get a very
reliable estimate of this kind of error.
Now suppose that the watch that I use is slow or fast. Then, no matter how many times I repeat the experiment (of course with the same watch), I can never know the amount of such an error.
Further, the errors sign will always be the same- either the watch will be fast or slow leading to ei-

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

ther an overestimate or underestimate of the rate of revolution. This is an example of a systematic error.

Nu
cle
ar
Ph
ys
ics

In general, there is no set prescription for eliminating systematic errors and mostly one has to use
common sense to know if there are any systematic errors and to eliminate them. Random errors on the
other hand are usually easier to study and eliminate or reduce. But one should remember that in many
situations, the accuracy of a measurement is dominated by possible systematic errors in the instrument
rather than the precision with which we can make the measurement.
To summarise

Systematic & Random Errors

Systematic Errors are reproducible inaccuracies that are difficult to detect and cannot
be analyzed statistically. If a systematic error is identified when calibrating against a
standard, applying a correction or correction factor to compensate for the effect can
reduce the bias.

M
an
ua
l

Random Errors are statistical fluctuations in the measured data due to the precision
limitations of the measurement device. Random errors can be evaluated through
statistical analysis and can be reduced by averaging over a large number of observations

The fundamental aim of an experimentalist is to reduce as many sources of error as s/he can, and
then to keep track of those errors that cant be eliminated.

Significant Digits

1.2.3

La

Whenever one writes the results of an experiment, the precision in the experiment is normally indicated
by the number of digits which one reports the result with. The number of significant figures depends
on how precise the given data is. The following rules are helpful:

Significant Figures
1. The leftmost NONZERO digit is ALWAYS the MOST significant digit.
2. When there is NO decimal point, the rightmost NONZERO digit is the least
significant digit.
3. In case of a decimal point, the rightmost digit is the least significant digit EVEN
IF IT IS A ZERO.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

10

4. The number of digits between the most and least significant digits are the number
of significant digits.

Nu
cle
ar
Ph
ys
ics

Thus for instance, 22.00, 2234, 22340000, 2200. all have four significant digits. When one is adding,
subtracting, multiplying or dividing numbers, then the result should be quoted with the least number
of significant figures in any one of the quantities being used in the operation of adding, multiplying etc.
In your intermediate calculations, always keep ONE MORE significant digit than is needed in the
final answer. Also when quoting an experimental result, the number of significant figures should be one
more than is suggested by the experimental precision.
Two things that need to be always avoided are
Writing more digits in an answer (intermediate or final) than justified by the number of
digits in the data.
Rounding-off, say, to two digits in an intermediate answer, and then writing three digits
in the final answer.

1.3.1

Statistical Analysis of Data


Histograms & Distribution

La

1.3

M
an
ua
l

While dropping off some figures from a number, the last digit that one keeps should be rounded
off for better accuracy. This is usually done by truncating the number as desired and then treating
the extra digits (which are to be dropped) as decimal fractions. Then, if the fraction is greater than
1
, increment the truncated least significant figure by one. If the fraction is smaller than 21 , then do
2
nothing. If the fraction is exactly 12 , then increment the least significant digit only if it is odd.

The fundamental problem in reporting the results of an experiment is to estimate the uncertainty in a
measurement. It is reasonable to think that the reliability of an estimate of uncertainty in a measurement can be improved if the measurement is repeated many times. The first problem in reporting the
results of many repeated measurements is to find a concise way to record and display the values obtained.
Suppose we measured the weights of all the new one rupee coins minted since they were introduced.
It is clear that all the coins will not have the same weight. The actual weight would depend on several
things including when it was minted and how much it has been in use etc. One way to display the
results of our measurement would be a histogram as shown in the Figure 1.1. This is for a sample of
25 coins and we have divided the weights into bins of width = 0.01 gm.

11

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Figure 1.1: Binned histogram

We have plotted the data in such a way that the fraction of measurements that fall in each bin is
given by the area of the rectangle above the bin. That is to say that the height P (k) of the k th bin is
such that
Area = P (k) = fraction of measurements in the k th bin

Thus, for instance, the area of the rectangle between 2.50 2.51 is 20 0.01 = .2. This means that
20% of the coins fall in this weight range.

La

M
an
ua
l

We can see that such a plot gives us a good way to represent the data namely how the weights of the
coins in our sample are distributed. In most experiments, as the number of measurements increases, the
histogram begins to take on a definite simple shape, and as the number of measurements approaches
infinity, their distribution approaches some definite, continuous curve, the so-called limiting distribution as in Fig(1.2).

Figure 1.2: Limiting Distribution

An important concept that we will need to understand is that of a probability distribution. A


probability distribution is a table or an equation that links each outcome of a statistical
experiment with its probability of occurrence. Recall that when the value of the variable is an

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

12

outcome of a statistical experiment, then it is a random variable. Thus, for instance we can think of
a statistical experiment of tossing a coin twice. We can get 4 possible outcomes- HH,HT,TH and TT.
Now let the variable X represent the number of heads in this experiment. Then X is a random variable
and it can take 3 values, 0, 1, 2. We can construct a table for the values x of the random variable X
and the probability associated with that value.
p

0.25

0.50

0.25

Nu
cle
ar
Ph
ys
ics

Table 1.1: Discrete Probability Distribution

This is a probability distribution. Clearly we can see that there will discrete and continuous probability distributions depending on whether the variable is discrete or continuous. In the example above
of the coin toss, the variable X is a discrete variable and hence this is a discrete probability distribution.
Examples of discrete distributions are the Binomial distribution and the Poisson distribution which we
shall examine shortly.

M
an
ua
l

On the other hand, if the random variable is continuous then the the probability distribution associated with it is called continuous probability distribution. Note that a continuous probability distribution
differs from a discrete probability distribution in that for a continuous distribution, the probability that
the variable assumes a particular value is zero and hence it cant be represented by a table. Instead,
we represent it with a function. Such a function is called a probability distribution function.

A probability distribution function has the following properties.

La

1. Since the continuous random variable is defined over a continuous range of values
(called the domain of the variable), the graph of the density function will also be
continuous over that range.
2. Since the continuous random variable can take an infinite number of values, the probability that it takes a specific value, say a is zero.
3. Furthermore, the area bounded by the curve of the density function and the x-axis is
equal to 1, when computed over the domain of the variable.
4. Finally, the probability that a random variable assumes a value between a and b is
equal to the area under the density function bounded by a and b. Note that the area
below the line x = a say, is equal to the probability that the variable X can take any
value value less than or equal to a.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

13

Examples of continuous probability distribution that we will study will be the Normal or Gaussian
distribution.

1.3.2

Parent & Sample Distribution

Nu
cle
ar
Ph
ys
ics

Any measurement of a quantity is usually expected to approximate the quantity though not be exactly
equal to it. We have already seen that every time we make a measurement, we expect some discrepancy
between them because of random errors and so every measurement we expect to be different. However,
as we increase the number of measurements, we see that the data is more and more distributed around
the correct value of the quantity being measured. (Of course all this is true if we can neglect or correct
for systematic errors).

M
an
ua
l

Suppose we make an infinite number of measurements. Then in principle, we would know the exact
nature of the distribution of the data points. If we had such a hypothetical distribution, then we could
determine the probability of getting any particular value of the measurement by doing a single measurement. This hypothetical distribution is called the parent distribution . Thus, in the above example,
the parent distribution for the one rupee coins minted in a particular period would be the weights of
ALL such coins. However, in practice, we always have a finite set of measured values, as in the example
above where we have a sample of 25 coins on which we carried out the measurements. The distributions
that are obtained from the samples of the parent distribution are called sample distribution. Of
course, in the limit of infinite observations, the sample distribution becomes the parent distribution.

La

We can define the probability distribution function P (x) which is normalised to a unit area. This
function is defined in such a way that for the limiting distribution (that is in the limit that the number
of observations N is very large, the number of observations of the variable x between x and x + x is
given by

1.3.3

N = N P (x)x

Mean & Deviation

The parent and the sample distributions discussed above can be characterised by several quantities.
We can define a mean of the sample distribution in exactly the same way as we understand it- as the
average value of the quantity. Thus the mean of the sample distribution, x is
x

N
1 X
1 X
xi
xi
N i=1
N

(1.2)

where xi are the different observed values of the variable x. Clearly, the mean of the parent distri-

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

14

bution, is

lim


1 X
xi
N

(1.3)

Nu
cle
ar
Ph
ys
ics

If the measurement of interest can be made with high precision, the majority of the values obtained
will be very close to the true value of x, and the limiting distribution will be narrowly peaked about
the value . In contrast, if the measurement of interest is of low precision, the values found will be
widely spread and the distribution will be broad, but still centered on the value . Thus, we see that
the breadth of the distribution not only provides us with a very visual representation of the uncertainty
in our measurement, but also, defines another important measure of the distribution. How can we
characterise this measure?
The most often used parameter for characterising the dispersion is called the standard deviation,
. We can define the variance 2 of the parent distribution as
2

= lim

which is easily seen to be


2


1 X
2
(xi )
N

(1.4)


1 X 2
xi 2
N

(1.5)

= lim

M
an
ua
l

from the definition of in Eq(1.3).

This is the measure of dispersion for the parent distribution. What about the sample distribution?
The variance here is defined in an analogous way, except that the factor in the denominator is N 1
instead of N .

La

"

1 X
s2 =
(xi x)2
N 1 i=1

#
(1.6)

Note that as N approaches , N 1 and N are the same. But for any finite N the difference comes
in because though the initial set of measurements were N independent measurements since all the N xi
were independent, however, in calculating x, we have used one independent piece of information. The
rigorous proof of this statement is a bit tricky though not required for our purposes.
The importance of these two parameters, the mean and the standard deviation (or variance 2 ) is
that this is precisely the information we are trying to extract from the experiment that we perform. For
the sample distribution, s2 characterises the uncertainty associated with our experiment to determine
the true and actual values. As we shall see, the uncertainty in determining the mean of the parent
distribution is proportional to the standard deviation. Thus we conclude that for distributions which

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

15

are a result of statistical or random errors , these two parameters describe the distribution well.

Nu
cle
ar
Ph
ys
ics

How do we determine the mean and standard deviation of distributions? For this, we define a
quantity called the expectation value . The expectation value of any function f (x) of x is defined as
the weighted average of f (x) over all the values of x weighted by the probability density function p(x).
Thus, the mean is the expectation value of x and the variance is the expectation value of square of the
deviations from . Thus for a discrete distribution from Eq(1.3), we need to replace the observed values
xi by a sum over the values of possible observations multiplied by the number of times we expect the
observation to occur. Thus
N
N
X
1 X
1 X
E(X) = lim
[xj P (xj )]
xi = lim
[xj N P (xj )] = lim
N N
N N
N
i=1
j=1

(1.7)

In a similar way, the variance can be written as

N
N
X
X


 2

2
= lim
(xj ) P (xj ) = lim
xj P (xj ) 2 = E(X 2 ) E(X)2
2

j=1

(1.8)

j=1

For a continuous distribution, the analogous quantities are


Z

M
an
ua
l

E(X) =
and

2 =

(1.9)

x2 p(x)dx 2 = E(X 2 ) E(X)2

(1.10)

La

(x )2 p(x)dx =

xp(x)dx

Example 1.3.3.1
I throw a dice and get 1 rupees if it is showing 1, get 2 rupees if it is showing 2, get 3 if it
is showing 3, etc. What is the amount of money I can expect if I throw the dice 150 times?
For one throw, the expected value is
E(X) =

xi P (xi )

The probability of getting any of the digits in one roll is 61 . Thus

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

E(X) = 1

16

1
1
1
1
1
1
7
+2 +3 +4 +5 +6 =
6
6
6
6
6
6
2

Thus if I roll the dice 150 times, my expected payoff is 525 rupees.

Nu
cle
ar
Ph
ys
ics

Example 1.3.3.2
I am given a probability distribution as below

X P(X)
1
8
4
1
12
6
3
16
8
1
20
5
1
24
8

Find the variance of this distribution?

M
an
ua
l

We know that the variance is given by Eq(1.8). But to use this, we first need to find the
expectation value of x or the mean. This is
E(X) =
Now

xi P (xi ) = 8

X
[xi E(X)]2 P (xi ) = 32.71

La

2 =

1
3
1
1
1
+ 12 + 16 + 20 + 24
= 17
4
6
8
5
18

Example 1.3.3.3
At a pediatricians clinic, the age of the children, x years, coming to clinic, is given by
the following probability distribution function:

3
x(2 x) 0 < x < 2
4
= 0 otherwise

f (x) =

(1.11)

If on a particular day 100 children are brought to the clinic, how many are expected to
be under 16 months old?

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

17

16 months is 43 years. So the probability of finding a child under 16 months is given by


the area under the curve of the probability distribution function between 0 and 43 . This
is

4
P (x < ) =
3

Z3
f (x)dx
0
4

3
x(2 x)dx
4

Nu
cle
ar
Ph
ys
ics

Z3

3
4

Z3

x(2 x)dx

0


3 80
=
4 81
20
=
27

(1.12)

M
an
ua
l

Thus for 100 children, the number we expect to be under 8 months is


100

20
74.07
27

La

What about the mean age of the children brought to the clinic? For this,we need to use
Eq(1.9). Thus

Z2
E(X) =

xf (x)dx
0


Z2 
3
x(2 x) dx
=
x
4
0

= 1
(1.13)
This result is not surprising if we try to see how the distribution looks graphically as in
Figure 1.3.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

18

1
f(x)

0.8

0.6

0.4

0.2

0
0.5

1.5

Nu
cle
ar
Ph
ys
ics

Figure 1.3: Graph of function in Eq(1.11)

We can see that the distribution is symmetrical about x = 1 and therefore the mean must
be in the middle of the range that is x = 1. This is always true and can be used when we
know that the distribution is symmetrical.
Finally, what about the variance of the distribution of the age of the children?
We know that the variance for a continuous distribution is given by Eq(1.10). Thus

M
an
ua
l

2 = E(X 2 ) E(X)2

But

Z2

E(X ) =
0

b
La

x2 f (x)dx

Z2
=


3
x(2 x) dx
4

6
=
5
(1.14)

Thus
2 =

6
1
12 =
5
5

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

1.4

19

Distributions

We have already seen that the results of a statistical experiment result in a distribution (either discrete
or continuous). We would be interested in three kinds of distributions, Binomial, Poisson and Gaussian
or normal distributions. It is important to note where these distributions are used.

Nu
cle
ar
Ph
ys
ics

The Gaussian distribution is the one which we encounter most frequently since this describes the
distribution of random observations in many experiments.
The Poisson distribution is generally used for counting experiments of the kind used in nuclear
physics where the data is the number of events per unit interval. In the study of random processes like
radioactivity, Poisson distribution is important as it is whenever we sort data in bins to get a histogram.
Finally, the binomial distribution is a discrete distribution which is used whenever the possible number of final states is small. This is true for instance in coin tossing experiments or even in scattering
experiments in particle or nuclear physics.

1.4.1

Binomial Distribution

M
an
ua
l

A Binomial or Bernoulli trial is basically a statistical experiment which has the following properties:

1. The experiment consists of n repeated trials.


2. Each trial can result in just two possible outcomes. We call one of these outcomes a success and
the other a failure.

3. The probability of success, denoted by P , is the same on every trial.

La

4. The trials are independent, that is, the outcome on one trial does not affect the outcome on other
trials.
A typical example would be repeated tosses of a coin and counting the number of heads that are
turn up. Clearly the properties mentioned above are satisfied.
We can define a Binomial random number as the number of successes, say x in a binomial
experiment with n trials. The probability distribution of such a binomial random number
is called the binomial distribution. An example of such a distribution we have already seen in the
discussion of the discrete distribution where the probability distribution is given as a table in Table 1.1.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

20

We can find an expression for the probability P (x; n) for x successes in a binomial experiment with n
trials by analysing an experiment of coin tosss. Suppose we want to know the probability of x coins with
heads and n x coins with tails. For this purpose, we know that there are n Cx different combinations
in which we can get the set of observations. In each of these combinations the probability of x heads
x
coming is px which in this case is 21 and the probability for n x tails is (1 p)nx = q nx which here
nx
is 12
. With this, we can write down the probability P (x; n) of getting x successes in an experiment
with n trials, each with probability p as

Nu
cle
ar
Ph
ys
ics

 
n!
n x nx n
P (x; n, p) =
p q
= Cx px q nx =
px (1 p)nx
x
x!(n x)!

Mean of Binomial Distribution

(1.15)

To find the mean of the binomial distribution, recall that the mean is just the expectation value of
the random variable. In this case, the random variable is x and the probability distribution function
Eq(1.15) is the probability of x successes in n independent trials when the probability of success in each
trial is p. Thus by the definition of expectation value, we have

xP (x; n, p)

M
an
ua
l

E(x) =

n
X

x=0
n
X
x=0

n
X
x=1

n!
px (1 p)nx
x!(n x)!

n!
px (1 p)nx
(x 1)!(n x)!

(1.16)

La

since the x = 0 term does not contribute. Now substitute m = n 1 and y = x 1. Then
m
X
(m + 1)!
E(x) =
(p)y+1 (1 p)my
y!(m y)!
y=0

= (m + 1)p

m
X
y=0

= np

m
X
y=0

m!
(p)y (1 p)my
y!(m y)!

m!
(p)y (1 p)my
y!(m y)!

But we know that the binomial theorem states that

(1.17)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

(p + q)m =

m
X
y=0

21

m!
py q my
y!(m y)!

Thus
m
X
y=0

m!
(p)y (1 p)my = (p + 1 p)m = 1
y!(m y)!

and so

Variance of Binomial Distribution

Nu
cle
ar
Ph
ys
ics

E(x) = np

(1.18)

Recall that the variance of a distribution is defined as the difference of the Expectation value of the
square of the random variable and the square of the expectation, that is
2 = E(x2 ) E(x)2
Consider

M
an
ua
l

n
X

E(x(x 1)) =

x=0
n
X

x(x 1)n Cx px (1 p)nx


x(x 1)

x=0

n
X

La

x=2

n!
px (1 p)nx
x!(n x)!

n!
px (1 p)nx
(x 2)!(n x)!

= n(n 1)

n
X
x=2

= n(n 1)p2

(n 2)!
px (1 p)nx
(x 2)!(n x)!

n
X
x=2

= n(n 1)p2

m
X
y=0

(n 2)!
p(x2) (1 p)nx
(x 2)!(n x)!
(m)!
p(y) (1 p)my
(y)!(m y)!

= n(n 1)p (p + 1 p)m


= n(n 1)p2
where we have used the substitution, y = x 2 and m = n 2. Now variance is

(1.19)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

22

2 = E(x2 ) E(x)2
= E(x(x 1)) + E(x) E(x)2
= n(n 1)p2 + np n2 p2
= np(1 p)

Nu
cle
ar
Ph
ys
ics

(1.20)

Example 1.4.1.1
An unbiased coin is tossed ten times. What is the probability of getting less than 3 heads?
The probability of finding less than 3 heads in 10 tosses, is the probability of finding less
than or equal to 2 heads, P (H 2). This will be the sum of probabilities of finding no
heads, 1 head and two heads. Thus
P (H 2) = P (H = 0) + P (H = 1) + P (H = 2)

M
an
ua
l

Now the probability of finding x heads in n tosses is given by Eq(1.15). In our case,
n = 10, p = q = 21 . Thus
 
   10
n x nx
10
1
1
P (H = 0) =
p q
=
=
x
0
2
1024

La

Similarly

 
     9
n x nx
10
1
1
10
P (H = 1) =
p q
=
=
x
1
2
2
1024

 
   2  8
n x nx
10
1
1
45
P (H = 2) =
p q
=
=
x
2
2
2
1024

Thus the total probability of getting less than 3 heads in 10 tosses is


P (H = 0) + P (H = 1) + P (H = 2) =

1
7
[1 + 10 + 45] =
1024
128

Example 1.4.1.2
Here is a game to test sixth sense. Take 4 cards numbered 1 to 4. One person picks a card
at random and another person tries to identify the card. What is the probability distribution that the second person would identify the card correctly if the test is repeated 4 times?

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

23

Let P (X = x) be the probability of correctly identifying x cards after 4 attempts. Then,


by the binomial probability distribution function, this is given by
 
 
n x nx
4
P (X = x) =
p q
==
(0.25)x (0.75)4x
x
x

Nu
cle
ar
Ph
ys
ics

since the probability of identifying the correct card in 4 attempts, out of 4 cards is
1
= 0.25. Here x = 0, 1, 2, , 4. Thus the probability of getting one card right in 4
4
attempts is
 
4
P (1) =
(0.25)(0.75)3 = 0.4219
1
The probability distribution is given by

P(x)
0.3164
0.4219
0.2109
0.0468
0.0039

M
an
ua
l

x
0
1
2
3
4

If this is done with say 100 people, we can see that the number of people getting 1 card
correct is 100 0.4219 42.

La

Example 1.4.1.3
A biased that is an unfair dice is thrown fifty times and the number of sixes seen is ten.
If the dice is thrown a further fifteen times find:

(a)the probability that a six will occur exactly thrice;

(b)the expected number of sixes;


(c)the variance of the expected number of sixes.
The experiment is clearly a Bernoulli or Binomial trial. If the success is taken to be
getting a six, then the probability p is given by
p=

10
= 0.2
50

Now if X is defined as the number of sixes in 15 trials, then

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

24

X = B(15, p)
We want the probability for getting exactly 3 sixes in 15 trials. Thus, x = 3 and
10
B(15, ) =
50

   3  12
15
1
4
0.2475
3
5
5

Nu
cle
ar
Ph
ys
ics

The expected number of sixes will be the expectation value E(X). This is, from Eq(1.18),
E(X) = np = 15

1
15
=
=3
5
5

Finally, the variance is given by Eq(1.20)

2 = np(1 p) = 15
1.4.2

Poisson Distribution

1 4
= 2.4
5 5

A Poisson distribution is the probability distribution that results from a Poisson experiment. A Poisson
experiment has the following properties:
1. The experiment results in outcomes that we can call successes or failures.

M
an
ua
l

2. The average number of successes that occur in a specified region is known.


3. The probability that a success will occur is proportional to the size of the region.
4. The probability that a success will occur in an extremely small region is virtually zero.

La

A Poisson random variable is the number of successes that result from a Poisson experiment. The
probability distribution of a Poisson random variable is called a Poisson distribution. A Poisson distribution is an approximation to the binomial distribution when the average number of successes, that
is is much smaller than the possible number, that is when  n. In such cases, evaluation of the
binomial probability is extremely complicated and tedious.
As an example of a Poisson distribution, consider the flux of cosmic rays reaching the earth. This
is known to be around 1 per cm2 per minute. Now consider a detector with a surface area of 40 cm2 .
We expect to detect 40 cosmic rays per minute in this detector. Now suppose we record the number
of cosmic rays detected in 20 second interval. On an average we then expect about 13.3 cosmic rays.
However, when we do the experiment over many 20 second intervals, we will detect something like
13, 15, 14, 12 etc and occasionally even 9 or 18 cosmic rays. We can plot a histogram of this, that is
plot the number of times nx that we observe x rays in this fixed interval of time. Or, if we divide
the number of times nx by the total number of intervals N , then we can get the probability Px of

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

25

observing x cosmic rays in this experiment. If our number of intervals N is large, then this probability distribution will be a Poisson distribution. Whenever we observe independent random events that
occur at a constant rate such that the expected number of events is , we get a Poisson distribution.
In the case of cosmic rays, the events are obviously random and clearly independent since the arrival
of one cosmic ray does not depend on the arrival of others. Further, the rate of arrival is almost constant.

P (x; n, p) =

1
n!
px (1 p)x (1 p)n
x! (n x)!

M
an
ua
l

But

Nu
cle
ar
Ph
ys
ics

Another example can be a scattering experiment with a beam of B particles incident on a thin foil
and the probability p of any one interaction taking place is very small. Then we know that the number
of observed interactions r will be binomially distributed where we will take the number of trials as B
and the probability of success (interaction) as p. It turns out that when p is very small, then the values
of Px will be like a Poisson distribution with a mean given by Bp (which we have seen is the mean of
the binomial distribution Eq(1.18)).
Thus we see that a binomial distribution goes to a Poisson distribution when the number of trials
N increases while the probability of success p decreases in such a way that N p is a constant.
Consider the case of a binomial distribution where p  1 and we consider the situation where n
but np remains finite. Recall that np is the mean of the binomial distribution (Eq1.18). The probability
function for the binomial distribution is

n!
= n(n 1)(n 2) (n x + 2)(n x + 1)
(n x)!

Now since x  n, each of these x factors is very nearly n and hence this becomes

La

Now

n!
nx
(n x)!

(1 p)x (1 + px) 1

since p 0. Thus we now have the probability function as


P (x; n, p) =

1 x x
n p (1 p)n
x!

Consider now
 
1
(1 p) = lim[(1 p) ] =
= e
p0
e
n

1
p

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

26

Since the mean, = np for a Binomial distribution.


Thus we get the Probability function for the Poisson distribution
PP (x; ) = lim PB (x; n, p) =
p0

x
e
x!

(1.21)

Mean of Poisson Distribution

Nu
cle
ar
Ph
ys
ics

This is the probability of obtaining x events in the given interval. Remember that x is positive
integer or zero.

The mean of the distribution is the expectation value of the random variable. Thus


X
x
E(x) =
x e
x!
x=0


X
x1

= e
(x 1)!
x=1
= e

X
y

y!

M
an
ua
l

y=0

(1.22)

Thus we see that the mean of the Poisson distribution is


(1.23)

E(x) =

La

Variance of the Poisson Distribution


We know that the variance is defined in terms of difference of the expectation value of the square of
the variable and square of the expectation value. That is
2 = E(x2 ) E(x)2
Consider

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

E(x ) =


X

27

e
x!


x
X
2
0+
x
e
x!
x=1


y+1
X
2

(y + 1)
e
(y + 1)!
y=0


X
y
2

(y + 1)
e
(y + 1)(y)!
y=0


X
y

(y + 1)
e
(y)!
y=0
" 
 X
#
 y
X
y

(y)
e
e
+
(y)!
(y)!
y=0
y=0

=
=
=
=

Nu
cle
ar
Ph
ys
ics

x=0

= E(x) +
= ( + 1)

(1.24)

M
an
ua
l

The sum of the probability distribution function over all x is unity and therefore the second term in
the square bracket is unity. Thus we see that the variance is simply
2 = E(x2 ) E(x)2 = ( + 1) 2 =

(1.25)

La

This is a remarkable result. The mean and the variance of the Poisson distribution is the
same. This gives rise to the famous square root rule. In an experiment where the distribution
satisfies Poisson distribution conditions, for instance in the counting of N independent events in a fixed
interval, we can estimate the mean of the distribution to be N . Then, as we saw above, the variance

is also N and therefore = N . The statistical errors in then would be N and we would quote our

results as N N . We will return to this later in the Chapter when we discuss error estimation.

Example 1.4.2.1
A factory produces resistors and packs them in boxes of 500. If the probability that a
resistor is defective is 0.005, find the probability that a box selected at random contains
at most two resistors which are defective.
If we take X as the number of defective resistors in a box of 500, then

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

28

X = B(500, 0.005)
since the trials are obviously Binomial. Now in this case, we can see that the number of
trials, n = 500 is large and the probability of success p = 0.005 in each trial is low, so the
Binomial distribution can be approximated by a Poisson distribution with a mean

Nu
cle
ar
Ph
ys
ics

= np = 500 0.005 = 2.5


Then the probability of finding two or less defective resistors in a box of 500 is
P (X 2) = P (0) + P (1) + P (2)
where

2.50 e2.5
P (0) =
0!
2.51 e2.5
1!

P (2) =

2.52 e2.5
2!

M
an
ua
l

or

P (1) =

P (X 2) = 6.625 e2.5 0.543

La

Example 1.4.2.2
A bank manager opens on an average 3 new accounts per week. Use Poissons distribution
to calculate the probability that in a given week she will open 2 or more accounts but less
than 6 accounts.
To use the Poisson distribution function, we need to know the mean. In this case, the
mean is given as 3 accounts per week. Thus we can use Eq(1.21) with = 3. Then the
probability of opening 2 or more accounts but less than 6 in a week is simply

P (2 < x < 6, 3) = P (2, 3)+P (3, 3)+P (4, 3)+P (5, 3) =

e3 32 e3 33 e3 34 e3 35
+
+
+
= 0.7167
2!
3!
4!
5!

Example 1.4.2.3
Thirty sheets of plain glass are examined for defective pieces. The frequency of the number

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

29

of sheets with a given number of defects per sheet was as follows:


Frequency
8
5
4
7
4
2
P
= 60

Nu
cle
ar
Ph
ys
ics

No. of defects
0
1
2
3
4
5
P
= 30

What is the probability of finding a sheet chosen at random which contains 4 or more
defects?
To use the Poisson distribution function, we need to find the mean number of defects. We
know that there are 60 defects in 30 sheets of glass. Thus the mean number of defects
per sheet is
=

60
=2
30

M
an
ua
l

The probability then of finding a sheet with 4 or more defects is

La

P (x 4) = 1 P (x < 4)
= 1 P (0) P (1) P (2) P (3)
 2 0

e2 (2)1 e2 (2)2 e2 (2)3
e (2)
= 1
+
+
+
0!
1!
2!
3!
= 0.1431

Example 1.4.2.4
At the ITO intersection, vehicles pass through at an average rate of 600 per hour.

(a)Find the probability that none passes in a given minute.


(b)What is the expected number passing in five minutes?

The average number of vehicles per minute is simply

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

30

600
= 10
60

Thus the probability that no vehicle passes in a given minute is


P (0, 5) =

e10 100
= 4.53 105
0!

Nu
cle
ar
Ph
ys
ics

The expected number of vehicles passing in five minutes is


E(X = 5) = 10 5 = 50

1.4.3

Normal Distribution

M
an
ua
l

The Normal Distribution is extremely important in probability theory and statistics and forms the
cornerstone of most of statistical analysis. For our purposes, its importance lies in the fact that many
real-world phenomena involve random quantities that are distributed in an approximately normal fashion. For instance, the errors in a scientific measurement are approximately normal. It is often called
Gaussian distribution and also referred to as bell-shaped distribution, because the graph of its probability density function resembles the shape of a bell.
The Normal or Gaussian Distribution is an approximation to the Binomial distribution for the limiting case when the number of possible observations, that is n goes to infinity AND the probability of
success in each measurement is finite and remains constant, that is when np  1. It is also the limiting
case of the Poisson distribution when the mean becomes large.

La

The Gaussian probability density is defined as


"

2 #
1
1 x
PG = exp
2

(1.26)

As one can see this is a continuous distribution function and thus describes the probability of getting
a value x from a random observation from a parent distribution of mean and standard deviation .
Properly defined, we should talk about a Gaussian Probability Distribution Function, such that
the probability dPG of the random observation having a value between x and x + dx i.e.
dPG (x; , ) = pG (x; , )dx
With this probability distribution function, we can now see how a Poisson distribution goes to a
Gaussian distribution for large mean. Consider first a Poisson distribution and a normal distribution

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

31

both with mean 1. Thus = 1 and in the case of the Gaussian distribution, = 1. Then the probability
that n 2 that is n + 1 for the Poisson distribution is
P (1 1) = P (0) + P (1) + P (2) = .736
where P (n) is the Poisson distribution function of Eq(1.21). With the Gaussian distribution function
(Eq(1.26)), we get

Nu
cle
ar
Ph
ys
ics

P (0 < x < 2) = 0.68


Thus we see that for small means, the two distributions are fairly different. What about for large

mean? Let us take = 15


= 7.5. Then = 7.5 = 2.7. Then the corresponding probabilities for
2
+ 1 are
P (7.5 2.7) = P (0) + P (1) + + P (10) = 0.64
and for the Gaussian distribution

P (0 < x < 10.2) = 0.68

M
an
ua
l

which is very similar. In general, for > 5, the Gaussian distribution is a good approximation to the
Poisson distribution. This fact will be important to us since we will see that in counting experiments,
it will become easier to use the Gaussian distribution in cases where the mean number of counts is large.
A normal distribution has the following properties:

1. The total area under the normal curve is equal to 1.

2. The probability that a normal random variable x equals any particular value is 0.

La

3. The probability that x is greater than some value a equals the area under the normal curve bounded
by a and .
4. The probability that x is less than a equals the area under the normal curve bounded by a and
.
5. About 68% of the area under the curve falls within 1 standard deviation of the mean.
6. About 95% of the area under the curve falls within 2 standard deviations of the mean.
7. About 99.7% of the area under the curve falls within 3 standard deviations of the mean.
A convenient form of the normal distribution is the Standard Normal Distribution. To obtain
this, we simply use the substitution

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

32

in Eq(1.26). Then the probability distribution function becomes


z=

 2
1
z
dz
pG (z)dz =
exp
2
2

(1.27)

(1.28)

Nu
cle
ar
Ph
ys
ics

It is important to see that since all the values of X, the normal variable falling between x1 and x2
have corresponding Z (the standard normal variable) values between z1 and z2 , it means that the area
under the X curve between X = x1 and X = x2 equals the area under the Z curve between Z = z1 and
Z = z2 . Therefore, we have, for the probabilities,
P (x1 < X < x2 ) = P (z1 < Z < z2 )

Mean of the Standard Normal Distribution

Z
E(x) =

xpG (x)dx

M
an
ua
l

1
=
2


dx

Z0

 2
 2
Z
x
x
x exp
dx + x exp
dx
2
2

1
=
2

= 0

(1.29)

La

x2
x exp
2


Thus we see that for a standard Normal distribution, the mean is 0.

Variance of the Standard Normal Distribution


We know that
2 = E(x2 ) E(x)2
Now

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

E(x ) =

33

x2 pG (x)dx

1
=
2

 2
x
x exp
dx
2
2

Z0



 2 
Z 
x
dx + x x exp
dx
2

Nu
cle
ar
Ph
ys
ics

x
1
= x x exp
2
2


 2 0
 2

 2  Z
 2
Z0
x
1
x
x
x
= x exp
+
exp
dx + x exp
+
exp
dx
2
2
2
2
2

1
=
2


exp

= 1


2

x
2

dx

(1.30)

R
R
using Integration by parts ( udv = uv vdu) and also the fact that the probability distribution
function is normalised to 1.

M
an
ua
l

Therefore

2 = E(x2 ) E(x) = 1 0 = 1

(1.31)

Thus we see that the standard normal distribution has a mean equal to 0 and a variance
equal to 1.

La

The difference between the standard normal distribution and the normal distribution can be seen
from the curves for probability distribution. Consider a normal distribution with = 2, = 13 . The
probability distribution function will look like Figure 1.4

34

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Figure 1.4: Normal Distribution with mean = 2 and =

1
3

La

M
an
ua
l

The corresponding standard normal distribution with = 0, = 1 will resemble Figure 1.5

Figure 1.5: Standard Normal Distribution with mean =0 and = 1

The two graphs obviously have very different and but have identical shapes and a shifting of
the axes will give one from the other. It is also easy to see that the area under the two curves between two equivalent points is the same. Thus, for instance, the area of the Normal distribution (with
= 2, = 13 between 0.5 to 2 to the right of the mean will be the area from x1 = + 2 = 2 + 16
to x2 = + 2 = 2.66. The area under the Standard normal distribution would be the area from
z1 = + 2 = 0 + 0.5 = 0.5 to z2 = + 2 = 0 + 2 = 2.0.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

35

Nu
cle
ar
Ph
ys
ics

The Standard Normal distribution with the probability distribution function given by Eq(1.28), gives
us the probability of finding the value. This is also the area under the curve from 0 to the value. This is
usually tabulated in a z-table which can be looked up as a standard reference as given below in Figure
1.6.

Figure 1.6: z-tables for Standard Normal Distribution

M
an
ua
l

(Source: http:// www.katyanovablog.com/ picsgevs/ normal-distribution-table)

La

Example 1.4.3.1
The mean weight of 1000 parts produced by a machine was 30.05 gm with a standard
deviation of 0.05 gm. Find the probability that a part selected at random would weigh
between 30.00 mm and 30.15 mm?
30.00 is 1 that is 0.05 below the mean. Similarly, 30.15 is 2 = 0.1 above the mean.
Thus
P (30.00 < X < 30.15) = P (1 < Z < 2)
since recall that the area under the normal gaussian curve between two points x1 and x2
is the same as under a standard normal curve between two points z1 and z2 which are
related to x1 and x2 by the transformation Eq(1.27). And 1 for the standard normal
distribution is 1 from the mean that is 0 and 2 is 2. These values can be looked up from
the standard tables.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

36

P (1 < Z < 2) = 0.3413 + 0.4772 = 0.8185


So the probability is 0.8185.

E(x) =
2
1

=
2

Nu
cle
ar
Ph
ys
ics

What about the Gaussian distribution? The Mean of the Gaussian distribution is obtained easily
now either by a substitution as given above into the standard normal distribution or by direct calculation.

(x )2
x exp
2 2


dx

(x )2
(x ) exp
2 2


1
= 0+
2

1
dx +
2



(x )2
exp
dx
2 2


(x )2
exp
dx
2 2


(1.32)

M
an
ua
l

since the distribution function is normalised to 1.


Thus we see that for a Gaussian distribution,
E(x) =

(1.33)

La

We can also find the Variance of Gaussian Distribution easily.

2 = E((x )2 )


Z
1
(x )2
2

=
dx
(x ) exp
2 2
2

2
=
2
=

(y)2
y exp
2
2


dy

(1.34)

since the integral is simply the variance of the Standard Normal Distribution given in Eq(1.30) which
is 1. Thus we see that the Variance of the Gaussian distribution is 2 .

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

37

The probability distribution function for the Normal or Gaussian distribution (Eq(1.26) is normalised
to 1. That is
Z
PG (x)dx = 1

Interval
0.674 < x < + 0.674
<x<+
1.282 < x < + 1.282
1.645 < x < + 1.645
1.960 < x < + 1.960
2.576 < x < + 2.576
3.291 < x < + 3.291

M
an
ua
l

Probability
0.50
0.68
0.80
0.90
0.95
0.99
0.999

Nu
cle
ar
Ph
ys
ics

From the definition of the probability distribution function, we know that the probability that any
one x value lies between the limits x = and x = + is simply the the area under the Gaussian
curve between these limits. If one computes (by integration) such areas for various choices of , one
can show that the probability of finding any one measurement of x between various limits, measured
as multiples of the standard deviation, , is given by the data given in Table 1.2.

Table 1.2: Normal Distribution: Probabilities with intervals

La

How can one interpret this table? The table indicates that we can be 95% confident that any one measurement that we make in the experiment (assuming all measurements are distributed normally) will
lie within approximately 2 of the mean. Thus, the probability column can be taken as the confidence
level and the interval column as the confidence interval.
This interpretation and analysis looks very straightforward. However, there is a problem- the problem is that the and the that we are using in the Gaussian distribution is the parent mean and
standard deviation. This means, as we have discussed above, that this will only be valid if we make an
infinite number of measurements!
We will address this issue in Section 1.6.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

1.5

38

Error Estimation

Nu
cle
ar
Ph
ys
ics

The basic aim in our experiment is to measure a quantity and also estimate the uncertainties in the
measurements. We also need to understand the sources of the uncertainties. Lastly we need to know
how to combine uncertainties in measurements of more than one quantity into the error in the quantity
which is calculated from these measurements. This is what we now discuss. Throughout this section,
we will only be dealing with statistical or random errors. Systematic errors will be assumed to have
been taken care of.

M
an
ua
l

First of all, it is important to understate a crucial fact which allows one to use the statistical methods
discussed above to analyse the errors in any experiment. The crucial result is that any measurement
subject to many small random errors will be distributed normally. This follows from the
Central Limit Theorem which states that the distribution of the sum of a large number of random
variables will tend towards a normal distribution. We may think of a measurement as being the result
of a process namely our carrying out many small steps in the experiment. Each step in the process
may lead to a small error with a probability distribution. When we sum the error over all steps to get
final error, the Central Limit Theorem guarantees that this will lead to a normal distribution no matter
what the error on the individual steps works out to be. So as a result of this, we generally expect
normal distributions to describe errors. Note that this simple, yet powerful fact allows us to use the
whole machinery of normal distributions and statistics to analyse errors.

La

In the case of observations that we are taking are collections of finite number of counts over finite
intervals, then the underlying distribution we know is Poisson. In this case, we know that the observed
values would be distributed around the mean in a Poisson distribution. (Recall that the random variable
in a Poisson distribution can only take positive values, including zero since it is defined as the number
of successes in a Poisson experiment.) In fact, in any experiment where data is grouped in bins to form
a histogram, the number of events in each bin will be distributed according to a Poisson distribution.
This allows us a tremendous simplification. We know that for a Poisson distribution, the standard
deviation is, Eq(1.25), simply
=

Thus the relative uncertainty, Eq(1.1), that is ratio of standard deviation to the mean is simply
relative uncertainty =

1
=

In our counting experiments, for instance, this means, that as we increase the number of counts per
interval (that is increase the mean ), the relative uncertainty goes down as the square root of the mean.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

39

Nu
cle
ar
Ph
ys
ics

This is actually referred to in general as the Square root rule for Counting Experiments which
states that the uncertainty in the any counted number of random events, which is used as an estimate
of the true average number, is equal to the square root of the counted number. For our purposes, this is
extremely important in our counting experiments since the process which we are measuring, namely the
counts are random and distributed in a Poisson distribution in any time interval. However, care should
be taken to note that this uncertainty is only in the counted variable and not in any derived
quantity. Thus, for instance if we were to measure N counts in an interval of T seconds, to get the
rate R per second,
N
T
Now to find the uncertainty in R, we know that the uncertainty is in the measured random variable

N and it is N . Thus the number of counts in time T is really


R=

From this the rate can be seen to be simply

N N
R=
T

and NOT

M
an
ua
l

r
N
N

R=
T
T
since only the quantity N is counted and hence the uncertainty in N is the square root of N .

1.5.1

La

This example above then leads us to the issue of how to estimate the error in any derived quantity
from the errors in the measured quantities?

Propagation of Errors

Consider an experiment where we measure some quantities, for instance the number of counts and the
distance and use these measured quantities to determine a quantity which is a function of the two
measured quantities. Let the desired quantity by x and the measured quantities, u and v be such that
x = f (u, v)

(1.35)

Suppose further that the most probable value of x, x is such that


x = f (
u, v)

(1.36)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

40

To determine the most probable value of x, we take the measurements of u and v, that is ui and vi and
determine the different xi . That is
xi = f (ui , vi )

(1.37)

x2

Nu
cle
ar
Ph
ys
ics

We have already seen that in the limit of an infinite number of measurements, the sample distribution
will go to the limiting distribution or parent distribution and the average of x, that is x will be the
mean of the distribution. In that limit, we can use the calculated sample mean, x to find the variance
x2 as


= lim


1 X
2
(xi x)
N

(1.38)

Also, expanding the function in Eq(1.35) in a Taylor series around the averages of u and v,


xi x ' (ui u)

x
u

+ (vi v)

x
v

(1.39)

Of course the partial derivatives have to be evaluated when the other variable is kept fixed at the
mean value.
Now combining Eq(1.38) and Eq(1.39), we have


 
 2
x
x
1 X
(ui u)
+ (vi v)
' lim
N N
u
v
"


   #
 2
2
1 X
x
x
x
x
(ui u)2
(1.40)
' lim
+ (vi v)2
+ 2(ui u)(vi v)
N N
u
v
u
v

M
an
ua
l

x2

Clearly, the first two terms are related to the variance of u and v, that is

La

and

u2

v2


= lim


= lim


1 X
2
(ui u)
N

(1.41)


1 X
2
(vi v)
N

(1.42)

We also define a new quantity covariance between the two variables u and v as
2
uv


= lim

Thus the variance for x is given by


1 X
(ui u)(vi v)
N

(1.43)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

x2

'

u2

x
u

2
+

v2

x
v

41

2
+

2
2uv

x
u



x
v


(1.44)

This is known as the error propagation equation.

x2

'

u2

Nu
cle
ar
Ph
ys
ics

The covariance term which is a measure of the correlation of the variations in u and v. In most
experiments the fluctuations in the measured variables are uncorrelated and so on the average the
covariance term vanishes for a large number of observations. We shall neglect this in our further
discussion. Thus we have in general, the error propagation equation as


x
u

2

v2

x
v

2

(1.45)

where the quantity x could be a function of any number of uncorrelated, independent variables.
Let us see what the error propagation equation looks like in some common cases.

1. Sums & Differences

M
an
ua
l

Let

where a is a constant. Then since

x
u

x=u+a

(1.46)

= 1, we get
x = u

(1.47)

La

We can also find the relative uncertainty (Eq(1.1))


x
u
u
=
=
x
x
u+a

2. Weighted Sums & Differences


Suppose x is the weighted sum of two variable u and v.
x = au + bv
Then, since


x
u


=a

(1.48)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

x
v

42


=b

we get using Eq(1.44)


x2 = a2 u2 + b2 v2

(1.49)

3. Multiplication & Division

Nu
cle
ar
Ph
ys
ics

assuming no correlation.

Suppose the quantity of interest x is defined by

x = uv

where u and v are measured quantities. In this case, we can see that


and

M
an
ua
l

x
u

x
v

=v

=u

and therefore the error propagation equation tells us that


(1.50)

x2
u2 v2
=
+ 2
x2
u2
v

(1.51)

La

or

x2 = v 2 u2 + u2 v2

For the case of division, we have


x=

u
v

then
x
1
=
u
v

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

43

x
u
= 2
v
v
and therefore
x2 =

u2 u2 v2
+ 4
v2
v

Nu
cle
ar
Ph
ys
ics

or
x2
u2 v2
=
+ 2
x2
u2
v

(1.52)

Note the very important difference between Eq(1.52, 1.51) and Eq(1.49). In the case of weighted
sums and differences the absolute errors are relevant while in this case, it is only the fractional
errors in u and v which are related to the fractional error in x.

4. Powers
Suppose

M
an
ua
l

x = aub

Then

La

and

x
u

= abub1 =

x2 = u2

bx
u

bx
u

The relative uncertainty in x is


x2
b
= u2
x
u

(1.53)

Example 1.5.1.1
Consider an experiment where we count N1 = 945 counts in a 10 second interval in an
experiment and then N2 = 19 counts in a 10 second interval. We have already found
a background reading of NB = 14.2 counts for the same 10 second interval by carrying

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

44

out a separate experiment carefully. Thus we assume that there is no uncertainty in the
background reading. Now in the first time interval, the corrected counts are
x1 = N1 NB = 930.8 counts
with an uncertainty of

and a relative uncertainty of

945 ' 30.7 counts

Nu
cle
ar
Ph
ys
ics

x1 = N1 =

x1
30.7
=
= 0.032 ' 3.1%
x
930.8

On the other hand, in the second interval, the figures are

x2 = 19 14.2 = 4.8 counts

with an uncertainty of

x2 = N2 =

19 ' 4.4

M
an
ua
l

and a relative uncertainty of

4.4
x2
=
= 0.91
x
4.8

La

Example 1.5.1.2
Now suppose in the previous example, the background reading also was subject to some
uncertainty. Then we have the formula
x=u+v

and thus the error propagation becomes


x2 = u2 + v2
since the partial derivatives are unity and the uncertainty is
x =

u2 + v2

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

45

In the above experiment for the first reading, if the background was not error free, then
we will have the error in the net count to be
x1 =

2
2
N
+

=
30.7 + 3.7 = 5.6
N
1
B

and so we should report our net counts as

Nu
cle
ar
Ph
ys
ics

930.8 5.6
Example 1.5.1.3
Suppose we want to find the ratio of the activity of two sources by measuring the counts
from them. Then suppose the counts from the first source is N1 = 586 and that from the
second is N2 = 265. Then the ratio of activity is given by
R=

586
N1
=
= 2.211
N2
265

The error in R by Eq(1.52) is given by

x2
u2 v2
N1
N2
1
1
=
+ 2 = 2+ 2 =
+
= 2.3 102
2
2
x
u
v
N1
N2
586 265

M
an
ua
l

and therefore

=R

2.3 102 = 2.211 0.151 = 0.333

R = 2.211 0.333

La

and therefore the ratio of activities should be quoted as

Example 1.5.1.4
Consider the simple experiment of finding g, the acceleration due to gravity. For this, I
throw a mass from top of the telescope tower and measure h the height of the tower and
also t, the time it takes for the mass to reach the ground. I get
t = 1.9 0.1 s
and
h = 18.5 0.2 m

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

46

Now I can use the formula h = 12 gt2 to determine g as


2h
= 10.24 m s2
t2

g=

I need to know the uncertainty in h and t which are given and then use the error propagation rules to find the uncertainty in g. We have

and

Nu
cle
ar
Ph
ys
ics

h
0.2
=
= 0.011 = 1.1%
h
18.5

t
0.1
=
= 0.052 = 5.2%
t
1.9

Now using the error propagation formula, we get




and therefore

g
g

2
=

 2
h

+2

 2
t

= (1.1)2 % + 2 (5.2)2 % = 55.29%

M
an
ua
l

g = .074 10.24 m s2 = 0.76 m s2


Therefore we should quote our result as
g = 10.24 0.76 m s2

La

We can see that the accepted value of g, that is 9.8 ms2 is within the margin of error.
Also, since the contribution to the error in g from t is much more, one should improve
the measurement of t rather than try to make the measurement of h better since that
contributes very little to the error.

Example 1.5.1.5
A radioactive sample is counted for one minute and we observe 900 counts. We then
remove the sample and count the background rate for 10 minutes to observe 100 counts.
What is the net count rate and the uncertainty in it?
There are two measured quantities in this experiment- the gross count rate and the background count rate. (We assume that the time has no uncertainty). The uncertainty in each

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

47

of these measured quantities, we have seen is simply the square root of the measurement.
Thus
Gross count rate = RG = 900

900 = 900 30 counts per minute

Nu
cle
ar
Ph
ys
ics

Background count rate = RB = 100 100 = 10010 in 10 minutes = 101 counts per minute

Net count rate = RN = RG RB = 890 counts per minute


What about the uncertainty in RN ? For this we need to use the error propagation equation
since remember, RN is a derived quantity and NOT a measured quantity. The defining
relation for RN is given above and so we get
RN =

R2 G + R2 B =

p
(30)2 + (1)2 30

Thus we should quote our result as 890 30.

M
an
ua
l

Example 1.5.1.6
A Geiger counter is placed near a suspected source of radioactivity and it records 58
counts in 30 seconds. The source is removed and the background count is found to be 48
counts in 30 seconds. Can we be sure that the source is truly radioactive?

The total count rate is

La

RT =

NT
58
=
= 116 counts per minute
T
0.5

The uncertainty in this measurement is

RT

N
=
=
T

NT
58
=
15 counts per minute
T
0.5

The background count rate is


RB =
and the uncertainty is

NB
48
=
= 96 counts per minute
T
0.5

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

RB

N
=
=
T

48

NB
48
=
14 counts per minute
T
0.5

Our net computed source count rate is thus


RN = RT RB = 20 counts per minute

RN

Nu
cle
ar
Ph
ys
ics

Once again, this net count rate is NOT a measured but a derived quantity. So we need to
use the error propagation equation to find the uncertainty in this given the uncertainties
in the total count rate and the background count rate. Thus
q

= R2 T + R2 B = 152 + 142 21 counts per minute

Thus we have to report our result as 20 21 counts per minute and conclude that we can
not say for sure whether the source is radioactive or not.

Example 1.5.1.7
Finally, let us consider a very simple example where there are both systematic and random
errors. Suppose I want to measure an unknown resistance R2 . I can for instance use the
following circuit.

La

M
an
ua
l

(Adapted from A Practical Guide to Data Analysis for Physical Science Students by L. Lyons)

Figure 1.7: Resistance Measurement

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

49

We measure the resistance R1 , voltages V1 and V2 using meters and then deduce the value
of R2 using Ohms law
R2 =

V2 V1
R1
V1

(1.54)

Nu
cle
ar
Ph
ys
ics

Clearly, the experiment involves measurement of 3 quantities and each of these will have
random and systematic errors. Random errors we already know happen because of statistical fluctuations. What about the systematics? In this particular case, there could be
many sources of systematic errors. Thus, for instance, the voltmeters V1 and V2 may not
be calibrated properly and/or the resistance meter might also not be calibrated. We could
also have the resistance being temperature dependent and hence not be very accurate.
Or there could be some stray capacitances in case we are using an AC source. We would
need to determine or estimate the amount of these systematic errors.
Now suppose, we performed the experiment and obtained the readings for R1 , V1 and V2
as follows:
R1 = (2.0 0.1) 1%

V1 = (1.00 0.02V) 10%

M
an
ua
l

V2 = (1.30 0.02V) 10%

La

The results of the measurements are reported with the average value plus/minus the random error and the second error in percentage is the estimate of the systematic errors. The
random errors come as previously discussed from statistical fluctuations and are known
accurately. The systematic errors quoted above are assumed to have been estimated using
a variety of methods.
Now as we have been discussing, the nature of random and systematic errors is different.
However sometimes we may want the error in our measurement as a single figure. In this
case, since the two sources of error, random and systematic are uncorrelated, we should
add them in quadrature just as we saw in the error propagation equation. Thus we get
R1 = 2.0 0.1
V1 = 1.00 0.10V
V2 = 1.30 0.13V
But note that the quantity of interest is not V2 or V1 but V2 V1 . What about the error in

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

50

this? This depends crucially on whether we measure the voltages using the same equipment or separate equipment. If the same equipment is used, then the errors are obviously
correlated while if separate instruments are used, then the errors are uncorrelated. If
different meters are used, then
V2 V1 = (1.30 0.13) (1.00 0.10) = 0.30 0.16 V

Nu
cle
ar
Ph
ys
ics

where we have added the errors in quadrature. Now if the same meter is used then the
errors are correlated and we have to include this effect. Thus, if there is a 10% systematic
uncertainty in each measurement, then the overall result will also have a 10% uncertainty.
Thus,

V2 V1 = (1.30 0.02) (1.00 0.02) 10% = 0.30 0.03 10% = 0.30 0.04 V
where the errors have been added in quadrature.

We are finally ready to see what the error in the quantity of interest, namely R2 is.

M
an
ua
l



V2 V1
V2
R2 =
R1 =
1 R1
V1
V1

i
Think of this expression as the product of two measured quantities,
1 and R1 . If
we know the errors in both these, then we can
to find the error in R2 . The
i
h use Eq(1.51)
V2
error in R1 is known as above. The error in V1 1 is simply the error in VV21 . Again the
error in this quantity can be found from Eq(1.52) once the errors in V2 and V1 are known.
V2
V1

La

The error in VV21 can be calculated assuming that the same meter is used to measure both
these quantities. Then
V2
(1.30 0.02)
=
= 1.30 3% = 1.30 0.03
V1
(1.00 0.02)

since the systematic errors cancel out using the same meter. Thus the value of



V2
1 = 0.30 0.03
V1

Therefore
R2 = [0.30 0.03] [2.0 0.1] = 0.60 0.07

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

51

1.6

Nu
cle
ar
Ph
ys
ics

What if we had done the following: We have the error in V2 V1 above. We have the
errors in R1 and V11 . Why couldnt we have used the error propagation equation for these
three quantities and found the error in R2 ? This would be wrong because we cannot
assume, under any circumstances that the errors in V2 V1 and V11 are uncorrelated since
V1 comes in both of these and no matter how we measure V1 , the errors will always be
correlated. Thus one has to be careful when using related quantities in determining the
errors in a calculated quantity.

Estimation and Error of the Mean

Recall that when we do any measurement, we typically end up with a sample distribution of the data
points of the quantity being measured. We saw that in the limit of an infinite number of measurements, this sample distribution goes to the parent distribution. Of course, the aim of any experiment
is ultimately to determine the parameters of the parent distribution. We also saw that a random set of
observations or measurements are distributed according to a Gaussian (or Poisson distribution). The
question we ask is what is the best estimate of the mean of the parent distribution?

Method of Maximum Likelihood

M
an
ua
l

1.6.1

La

Suppose we conduct an experiment where we obtain N data points which are randomly selected from
the parent distribution. This set of N points of course is our sample distribution. The question
we need to answer is how are the parameters, x and s, that is the average and standard
deviation of the sample distribution related to and , the mean and standard deviation of
the parent population? If the parent distribution is Gaussian with mean and standard deviation
, then the probability of finding x1 (one of the data points in the sample distribution) in the range x1
and x1 + dx1 , which we will define as the probability of finding x1 , is simply
"

2 #
1 x1
1
P (x1 ) =
exp
2

(1.55)

Similarly, the probability of finding x2 in the range x2 and x2 + dx2 is


"

2 #
1 x2
1
exp
P (x2 ) =
2

(1.56)

and so on. Note that each of these observations are independent and hence these probabilities are
uncorrelated. Thus, the probability of finding x1 between x1 and x1 + dx1 , x2 between x2 and x2 + dx2
etc is simply

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

52


P (x1 , x2 , x3 , , xN ) = P (x1 )P (x2 )P (x3 ) P (xN ) =

N

"

1X
exp
2

xi

2 #
(1.57)

Nu
cle
ar
Ph
ys
ics

It is important to stress that this probability is for a specific value of and . But we dont know
what that value is! So we start off by guessing a value of and , say 0 and . We compute the
probability P0 , given Eq(1.57). Then we guess another value of , say 00 with the same and compute
P00 , . We do this for several values of with the same . Then we choose that value of for which
the probability is a maximum. This is known as the Method of Maximum Likelihood.
Now it is clear that for the probability (as a function of ) in Eq(1.57) to be a maximum, the
argument of the exponential should be a minimum since the probability is a constant times a negative
exponential. Let
"

2 #
1 X xi
Y =
2

Then, we want to find the value of for which


"


2
1 X xi

d
dY
=
d
d 2

M
an
ua
l

or

X  xi 

=0

xi = N = 0

xi
= x
N

(1.58)

La

which gives us

Thus, we see that the most probable value of the population mean is precisely the average or sample mean x of the sample distribution.
This is an extremely powerful result. It basically allows us to get an estimate of the parent population mean from a sample distribution which is all what we can possibly obtain from any experiment.
What about the error in this mean?

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

1.6.2

53

Estimated Error in the Mean

We know that the sample standard deviation, s, is the average uncertainty associated with each of the
measurements x1 , x2 , , xN . A legitimate question from our point of view is also, what is the estimated
uncertainty or standard deviation of in our determination of the mean for the parent distribution or
x for the sample distribution? Recall that our assumption was that each of the data points xi is from
the same parent distribution and hence it is characterised by the same standard deviation .

Nu
cle
ar
Ph
ys
ics

To find the uncertainty in the mean, we can use the error propagation equation Eq(1.45) to find
using Eq(1.58). Thus we have
"

2 =

i2

xi

2 #

(1.59)

Now, as we have supposed, all the data points have the same , that is i = for all i. We also have

=
xi
xi
Thus, we get

"

i2

1
N

2 #

M
an
ua
l

2 =


1 X
1
xi
N
N

2
N

(1.60)

Thus we see that the uncertainty in the mean is

=
N

(1.61)

La

The quantity is called the standard deviation of the mean or the standard error . Thus we
see that if we take N measurements of some quantity x and obtain x1 , x2 , xN , then we can state
our result as the best estimate of x which we have seen is the mean of the sample population and the
uncertainty as Eq(1.61), that is
x = x
where
x =

1 X
xi
N

and
x
=
N

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

54

Recall that x is simply the sample standard deviation s.

Nu
cle
ar
Ph
ys
ics

This result has an enormous significance for us. The standard deviation x or s represents the average uncertainty in the individual measurements x1 , x2 , , xN . Thus, if we were to make some more
measurements (using the same technique), the standard deviation x = s would not change appreciably. On the other hand, the standard deviation of the mean, would slowly decrease as we increase
N . This decrease is just what we would expect. If we make more measurements before computing an
average, we would naturally expect the final result to be more reliable, and this improved reliability is
just what the denominator guarantees. Stated another way, by taking more and more measurements,
we smoothen the distribution (the histogram of the data points) and also can determine the peak (the
mean) in an improved fashion. But note that the increase in precision is only growing as the square
root of the number of measurements and so there is a limitation in by how much we can increase the
precision by taking more and more measurements.
We can now return to the question that we had raised in the Section on Gaussian distribution? How
does one establish confidence limits in the absence of an infinite number of observations. Now that we
have the best estimate of the standard deviation of the mean, we can see that the best way to report our
experimental observations of a parameter x whose mean x has been determined experimentally from
taking N measurements is

M
an
ua
l

best value of x = x

Using the table for Gaussian probability intervals, we can say that this way of reporting the result
is at 68% confidence level. We can increase the confidence level to 95% by reporting it as
best value of x = x 2

La

Note once again that the values for confidence limits used are for the Gaussian distribution and not
for the sample distribution which is obtained by a finite set of measurements whose standard deviation is
s. The relationship between confidence limits and the sample deviation s was obtained by W.S. Gosset
and is known as the Students t-distribution. Basically, there exists a function which can only be
evaluated numerically that allows us to compute a single number which relates s to the confidence level.
What we can then say is that the true value of the parameter x fell in the interval
t
t
x < true value of x < x +
N
N
where N is the number of measurements and t can be found from the tables at various confidence
levels and values of N . We will not go into the details of the t- distribution but only note that in
the limit that N , the values of t at various confidence levels approaches that of the Gaussian
distribution as it should.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

55

1.7

Method of Least Squares

Nu
cle
ar
Ph
ys
ics

Basically, for our purposes it is enough to note that for a small number of observations, the predictions from Students t-distribution are more accurate since the Gaussian probability distribution
overestimates the confidence level associated with a given range. To put it another way, the Students
t-distribution probability requires a larger uncertainty estimate than the Gaussian probability and the
two only coincide for a very large number of measurements.

M
an
ua
l

In an experiment suppose we find a collection of data points (xi , yi ) and we want to find a relationship
between them. The theoretical relationship between these two quantities, x and y is supposed to be
linear. On plotting the points (xi , yi ), we see that the relationship looks somewhat linear. In this case,
we can ask two questions- firstly, assuming the relationship to be linear, can we find the actual relationship, that is y = M x + C which best fits the data. This of course means finding the best estimates
of the constants M , the slope of the straight line and the C, the y intercept. The finding of the best fit
straight line to the data is called linear regression or the least square fit for a line. Once we have
found the best fit straight line, then we can also ask the question about the goodness of fit, that is to
say, how well does our data fit the assumed straight line. Let us first see how to find the best fit linear
relation between the two quantities x and y.

La

We assume that the uncertainty in measuring the variable x is negligible while there is an appreciable
uncertainty in the measuring the variable y. In most experiments, this assumption is a reasonable one
since usually, the uncertainty in one of the variables is much larger than that in the other one. For instance, when we measure speed, the distance measurement has much larger uncertainties than the time
variable. Or, in our case for instance, when we plot the characteristic of the number of counts versus
voltage, the error in the voltage is negligible. We also assume that the measurements of the variable
y is governed by a Gaussian distribution of the same standard deviation, y for all measurements yi of y.
Now the linear relationship we have assumed between x and y is
y = Mx + C
If we knew M and C, then for any value of x, say xi , we would know the true value of the quantity
y, that is
yiT True Value of yi = M xi + C
But we have assumed that the measurements of y are distributed according to a Gaussian distribution,

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

56

and therefore all the measurements of yi will be distributed in a Gaussian distribution centered on this
true value which would be the mean, . It also follows that the probability of finding a particular
value of yi is
PM,C (yi ) =

"

1
exp
2
2

yi M xi C
y

2 #
(1.62)

Nu
cle
ar
Ph
ys
ics

Similarly, we can obtain the probabilities of obtaining all the values of y, namely yi , y2 , , yN . Since
these values are all independent, the probability of obtaining our complete set of measurement, that is
yi , y2 , , yN is simply the product of each of these probabilities.
"

PM,C (y1 , y2 , , yN )
Let us define a quantity 2 as
2

N
1 X (yi M xi C)2

1
exp
yN
2

y2

i=1

N
X
(yi M xi C)2

y2

i=1

(1.63)

(1.64)

Then the probability can be easily written as

 2
1

PM,C (y1 , y2 , , yN ) N exp


y
2

(1.65)

M
an
ua
l

We now use the same technique as we did in the previous section to find the best values of M and
C, that is the method of Maximum Likelihood. The best values of these quantities would be those for
which the probability in Eq(1.65) is a maximum. Or, since the probability is proportional to a negative
exponential, those values of M and C for which 2 in Eq(1.64) is a minimum. To get those values of
the parameters M and C, we differentiate 2 w.r.t them and equate to zero. Thus

and

La

2
M

2
C

N
2 X
= 2
xi (yi M xi C) = 0
y i=1

(1.66)

N
2 X
= 2
(yi M xi C) = 0
y i=1

(1.67)


M

These equations can be rewritten in a more suggestive form as

xi + M

CN + M

Solving these simultaneous equations, we get

x2i =

xi y i

xi =

yi

(1.68)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

M=

57

P
P P
xi yi xi yi
P
P
N x2i ( xi )2

(1.69)

P
P P
y i xi xi y i
x2i
P
P
N x2i ( xi )2

(1.70)

and
P
C=
Defining a quantity as

we get
M=
and

P
C=

x2i

2

X

xi

Nu
cle
ar
Ph
ys
ics

=N

x2i

xi yi

xi

yi

P P
y i xi xi y i

(1.71)

(1.72)

With these estimates of the slope M and the intercept C, we get the Least Square fit of the straight
line or the line of regression of y on x.

La

M
an
ua
l

The question now remains as to what is the uncertainty in the measurements of the quantity y?
Before we do this, it is important to remember that the measurements y1 , y2 , , yN are NOT the
measurements of the same quantity. So, a spread in their values is NOT a measure of the uncertainty in
their values. The measurement of each yi is, we have already assumed normally distributed around the
best fit value which we have seen to be M xi + C with a standard deviation which is assumed to be y .
The deviations of the measured value from the true value or best fit value would thus be yi M xi C
and these too would be normally distributed around a mean of 0 and with a standard deviation of y .
To obtain the best estimate for y , we again need to use the Method of Maximum Likelihood with
the probability of finding the measured values y1 , y2 , , yN which is given in Eq(1.65). But this time,
the parameter we are interested in is y and so we need to find the most likely value of y which will
maximise the probability in Eq(1.65). Thus we differentiate Eq(1.65) with respect to y to get
N y2 =

X
(yi M xi C)2

or
r
y =

1 X
(yi M xi C)2
N

(1.73)

It turns out that just as in Eq(1.6), we needed to divide by N 1 instead of N because the number
of independent values had been reduced in computing the mean, in this case too, since we need to find

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

58

two parameters M and C (from the measurements) we need to divide by N 2 instead of N . In any
statistical calculation, we can find the number of degrees of freedom by taking the number of independent
measurements and subtracting the number of parameters determined using these measurements. Thus
we finally get the expression for y , the uncertainty in the measurements yi , y2 , , yN is
r
y =

1 X
(yi M xi C)2
N 2

(1.74)

Nu
cle
ar
Ph
ys
ics

This finally leaves us with the issue of the uncertainty in the quantities that we have determined,
that is M and C. For this, we know the expressions for M and C in terms of yi , namely Eq(1.71) and
Eq(1.72). Thus, knowing the uncertainty in yi , that is y which we have just determined, we can easily
use the error propagation equation Eq(1.44) to determine the uncertainty in M and C. These are given
by
r

M = y

and

(1.75)

rP

C = y

Goodness of Fit

(1.76)

M
an
ua
l

1.8

x2i

La

Let us consider a typical counting experiment where we obtain counting statistics. The experimental
data consists of a series of measured or observed quantities. Let us assume that we have N independent
measurements of the same quantity. These could be anything- in our case, these could be the number
of counts in a specified time interval that we have taken repeatedly. This set of observations allow us to
prepare a sample distribution which as we have already seen can be characterised by two quantitiesthe sample mean, x and the sample variance s2 .
The underlying distribution which describes the process is, as we have seen, called the parent distribution which is characterised by the mean, and the variance 2 . This distribution is either Poisson
or Gaussian, depending on the value of the mean since we recall that for large values of typically
larger than 20, these two distributions become almost identical. But the question is that we dont know
these parameters of the parent distribution. So how do we proceed?
We take the sample mean x and assume it to be the mean of the underlying parent distribution.
This is something which is a good approximation as we saw in the section on Method of Maximum
Likelihood. With this mean for the parent distribution, we compute the actual variance 2 . We also
have the sample variance s2 . If our estimate of the sample distribution being a good representative of

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

59

the parent distribution is good, then we expect that the two, that is the sample variance and the true
variance should be close to each other. This comparison is done quantitatively by a procedure known
as the Chi-squared test.

Nu
cle
ar
Ph
ys
ics

Basically, what we are trying to see is whether an obtained or measured set of frequencies in a random
sample and what we expect from an assumed statistical hypothesis match and how well they match. In
our present case, the Chi Squared test allows us to determine how well the observed sample distribution and the assumed parent distribution (in this case a Poisson distribution with a mean = x ) match.
We define 2 as

1X
=
(xi x)2
x i=1
2

(1.77)

If we recall the definition of the sample variance s2 from Eq(1.6), we can rewrite this in terms of the
sample variance as
2 =

(N 1)s2
x

(1.78)

M
an
ua
l

In our case of the underlying distribution being a Poisson distribution,we know that the variance
of the parent distribution 2 is simply the mean of the distribution . But we have already chosen
2
the mean to be the same as the sample mean x. Thus, the deviation of the ratio sx from unity is
a measure of how much the sample variance differs from the variance. In terms of the 2 , we can say
that if the sample distribution is truly Poisson, then 2 = N 1. Any departure from this would be a
measure of how much the sample distribution differs from a Poisson distribution.

La

Thus we see that 2 is a statistic that tells us about the dispersion of the observed frequencies from
the expected frequencies. It is usually convenient to define a quantity called the degrees of freedom
as
= N Nc

(1.79)

where N is the number of sample frequencies and Nc is the number of constraints. One way to think
of the number of constraints is that it is the number of parameters which have been calculated from
the data to determine the probability distribution. Thus, in the case above of the Poisson distribution,
we have calculated one parameter, the sample mean x from the data and therefore in that case, the
number of degrees of freedom is simply N 1. With this, we can define a quantity called reduced
Chi sqaured or 2 as

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

2 =

60

(1.80)

This reduced chi squared clearly has an expectation value equal to 1. If the calculated values of 2
are much larger than 1 then we can say that either our measurements are not good, or the underlying
probability distribution that we have assumed is incorrect. If the value of 2 is very small then again
there is some problem with the experiment.

Nu
cle
ar
Ph
ys
ics

Another way to think of the the 2 squared test is to think of trying to fit a model to some given
data. Recall the quantity 2 that we encountered while discussing the Method of Least Squares in
Eq(1.64). There, we had defined
N
X
(yi M xi C)2
i=1

y2

for the case of the least square fitting of a straight line. It is obvious that in the general case, suppose
we have some observations yobs and we assume an underlying model for the data which yields the values
yth for the data points, then we can define 2 as
2 =

N
X
(yobs yth )2
i=1

i2

References

La

1.9

M
an
ua
l

Typically, we can have various competing models for our data or, what is the same thing, several different values of some model dependent parameter. To decide which model (or the value of a parameter)
fits the data best, we compute the 2 from the data and from that determine the reduced chi squared,
2 . From our discussion above, it is clear that we should choose that model, or the value of the model
parameter, for which the reduced chi squared 2 is closest to 1.

1. Data Reduction & Error Analysis for the Physical Sciences, D. Keith Robinson & Phillip R.
Bevongton, Mcgraw Hill (2003).

2. An Introduction to Error Analysis: The Study of Uncertainties in Physical Measurements, John


R. Taylor, University Science Books (1997).

3. Radiation Detection & Measurement, Glenn F. Knoll, Wiley India (2009). Chapter 3.

Chapter 2

2.1

Nu
cle
ar
Ph
ys
ics

RADIOACTIVITY
Radioactivity

M
an
ua
l

Radioactivity, discovered in 1896 by Becquerel has played an important role in our understanding
of the nature of matter at the subatomic level. Radioactivity has certain characteristic features which
were inexplicable when it was discovered but can be easily explained within the context of quantum
mechanics and relativity. Thus, for instance, the fact that a radioactive nucleus can spontaneously
decay and liberate energy without any excitation from outside can only be understood by thinking of
the equivalence of mass and energy. The completely random or probabilistic nature of the decay process
cannot be understood classically but comes out naturally within the framework of quantum mechanics.
There are basically five kinds of radioactive decay as listed in the Table (2.1).
Transformation

Example

Reason for Instability

Alpha Decay

A
A4
4
ZX Z-2Y + 2He
A
A

ZX Z+1Y + e +
A
+
A
X

Y
+
e
+

Z
Z-1
A
A
X
+
e Z-1
Y
Z
A *
A
ZX ZX +

238
234
4
92U 90Th + 2He
14
14

6C 7N + e
64
+
64
29Cu 28Ni + e
64

64
29Cu + e 28Ni
87 *
87
38Sr 38Sr +

Nucleus is too large

Beta Decay
Positron Emission
Electron Capture

La

Gamma Decay

Decay

Nucleus has too many neutrons relative to protons


Nucleus has too many protons relative to neutrons
Nucleus has too many protons relative to neutrons
Nucleus has excess energy

Table 2.1: Radioactive Decays

(Adapted from Concepts of Modern Physics by Beiser,Mahajan & Rai Choudhury)

We shall discuss these different kinds of phenomenon in some detail later.

2.1.1

Measure of radioactivity

A quantitative measure of the radioactivity of a sample is activity which is defined as the rate at
which the atoms of the sample decay. If N is the number of atoms (or nuclei) present at time t, then
61

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

62

its activity R is defined as


dN
(2.1)
dt
Clearly, since the derivative is negative, the negative sign in the definition makes the activity a
positive quantity. The SI unit of activity is becquerel (Bq) which is defined as
R=

1 Bq = 1 decay s1

Nu
cle
ar
Ph
ys
ics

The traditional unit of activity is the curie (Ci) which was defined originally as the activity of 1
gram of radium 236
88Ra, but is now defined as
1 Ci = 3.70 1010 decays s1 = 37 GBq

As we shall see, the radiations which come out when a radioactive nucleus decays are ionizing in
nature and when they pass through living tissue, they can damage the tissue. Sometimes the damage
maybe slight and the body heals itself. But sometimes the damage can be severe and have long term
disastrous consequences which include cancer and other illnesses.

M
an
ua
l

Radiation dose is measured in a unit called sieverts (Sv) which is defined as the amount of radiation
(of any kind) which has the same biological effect as the absorption of 1 joule of X-rays or gamma rays.
Typically, a safe exposure to radiation is taken to be about 1 milliSv per year. This does not include the
background radiation to which we are anyway exposed. The background radiation that we experience
is from the radionuclides in the rocks and earth and buildings as well as due to cosmic rays hitting the
atmosphere. To put this in perspective, a typical X-ray exposes us to about 0.02 mSv while a CT scan,
which is basically several X-ray exposures, typically can expose us to 5 8 mSv.

Activity Law & Half Life

La

2.1.2

When we measure the activities or the rate of decay of radionuclides, we find that the rate falls off
exponentially with time. This can be encapsulated mathematically as
R(t) = R0 et

(2.2)

where the constant is characteristic of the radionuclide and is called the decay constant. It is
convenient to define another quantity called the half-life T1/2 as the time in which the activity drops
to one half of its initial value. Thus, we see that by definition,
R(T1/2 ) =
or

R0
= R0 eT1/2
2

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

63

T1/2 = ln 2
and so we get
T1/2 =

ln 2

(2.3)

It is clear that for a radionuclide which has a large decay constant, the half life is small and vice versa.

Nu
cle
ar
Ph
ys
ics

The empirical activity law (Eq(2.2)) can be obtained if we assume a constant probability per unit
time for the decay of every nucleus in the sample. Then, in time dt, the probability for decay of any
one nucleus is simply dt. This is the probability for any one nucleus to decay. If we have N nuclei,
the number dN that will decay in time dt will be clearly
dN = N dt

Now since the number of nuclei (of a particular kind which we had initially) is decreasing, this
expression must have a negative sign for it to make sense. Thus we get
dN = N dt

(2.4)

M
an
ua
l

Integrating this expression gives us the Radioactive decay law as


N (t) = N0 et

(2.5)

La

It is important to note that the whole phenomenon of radioactivity is statistical in nature. As noted
above, every single nucleus has a definite probability of decay but which particular one decays in a
particular interval of time is essentially random in nature. All we can say is that if we had many nuclei
present (which is always the case in any experiment that we do), the fraction that will decay in any time
period will be approximately the same as the probability of an one nucleus to decay. As we have seen in
Chapter 1, we can model this statistical phenomenon as a Poisson distribution and the probabilities will
then be given by the distribution function for the Poisson distribution. Thus if we say that a particular
sample has a half life of 1 hour, all it means is that every single nucleus in that sample has a 50%
probability or chance of decaying in 1 hour. And recall that the decay probability is constant and thus
if a particular nucleus does not decay in 1 hour, it has a 75% probability of decaying in 2 hours and
NOT a 100% probability.
We can also define a quantity called Mean Lifetime which is simply the reciprocal of the decay
constant. Thus

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

64

T1/2
1
= 1.44T1/2
(2.6)
T = =

ln 2
Finally, a quantity which is used sometimes to describe a radioactive source is specific activity
which is defined as the activity per unit mass. The mass of the radioactive nuclide with N nuclei and
having a molecular weight (or atomic weight) M , will be given by
mass =

NM
AAv

(2.7)

Nu
cle
ar
Ph
ys
ics

Thus, using Eq(2.7, Eq(2.1), Eq(2.2) and Eq(2.4), we have

N AAv
NM
AAv
=
M

specific activity =

(2.8)

where AAv is the Avogadros number or 6.02 1023 nuclei mole1 .

2.2

M
an
ua
l

We now have the definitions of quantities used to describe radioactivity. We next turn to a discussion
of the nature of particles and radiation emitted by the radionucleus during the process of its decay by
radioactivity. These we have seen can be of several types as given in Table 2.1.

Nuclear Decay

There are several facts that we know about the nucleus. Let us recall them.

La

1. Nuclei consist of positively charged protons and neutral neutrons. These are collectively known as
nucleons.
2. The size of the nucleus is of the order of 1 femtometer or 1015 m. This unit is also called a Fermi.
3. The density of nucleons is roughly the same in the inside of the nucleus and hence the nuclear
radius is proportional to A1/3 where A is the mass number. This relationship is usually written as
R = R0 A1/3 with R0 1.2 1015 m
4. Nuclear densities are enormous- a typical nucleus will have a mass density of 1017 kg m3 .
5. Protons and neutrons are fermions and carry spin 21 . They also possess a magnetic moment. The
unit of nuclear magnetic moment, in analogy to the Bohr magneton is the nuclear magneton

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

which is defined as
N =

65

e~
= 3.15 108 eV T1
2mP

and is smaller than the Bohr magneton because of the presence of the proton mass mP in the
denominator. The proton magnetic moment is P = 2.793N and the neutron magnetic moment
is n = 1.913N .

Nu
cle
ar
Ph
ys
ics

6. The nucleons experience two kinds of forces- the positively charged protons experience repulsive
electric force which is a long range force. This is balanced by the short range nuclear force which
attractive and is experienced by all the nucleons. The short range nuclear forces arise from the
fundamental force of strong interaction.
7. The nucleons also experience weak interactions which is also a short range force.

8. The sum total of the masses of the nucleons in a nucleus is more than the mass of the nucleus.
The balance is called binding energy which can be thought of as the energy required to keep the
nucleus together. The range of binding energies is from a few MeV (as in the case of deuterium)
to more than 1.5 GeV (in the case of an isotope of Bismuth).

M
an
ua
l

9. Some combinations of neutrons and protons form stable nuclei. For light nuclei, generally the
number of protons and neutrons are equal. As we go to heavier nuclei, the number of neutrons
becomes greater since neutrons only experience the short range strong nuclear force and this is
required to balance the electric repulsion of protons.
10. Nuclear force is short range and nucleons interact via this force only with their nearest neighbours.
On the other hand, the electric force is present throughout the nucleus and so there comes a point
when the neutrons cannot prevent the break up of the nucleus. This is the limit of stable nuclide
which is 209
83Bi.

Alpha Decay

La

2.2.1

Alpha particles are basically helium nuclei which are emitted by some radionuclides during radioactivity
decay. The process is
A
ZX

A4
Z2Y

+ 42He

The fundamental reason for alpha particle emission is the fact that some nuclei are too large to be
stable since the short range nuclear force cannot sufficiently counteract the electric repulsion. Nuclei
which contain more than 210 nucleons are such nuclei and they decay by emitting alpha particles to
reduce their size and thereby increase their stability. The disintegration energy or the Q factor is
basically the mass difference between the parent nuclei and the sum of masses of the daughter nucleus
and all the other decay products. Since the nucleus, both the parent and the daughter nucleus in the

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

66

case of alpha decay are so much heavier than the alpha particle, there is very little recoil of the daughter
nucleus. Energy momentum conservation then gives us the kinetic energy of the alpha particle as
A4
Q
A
Clearly, since almost all alpha particle emitters have A > 210, the kinetic energy of the alpha particle
is roughly the disintegration energy. Typical alpha particle energies are in the range of 4 6 MeV.
Since the energy of the alpha particle is directly related to the Q value, the alpha particles are
monoenergetic.

Nu
cle
ar
Ph
ys
ics

K.E

La

M
an
ua
l

Although a heavy, unstable nucleus can become more stable by emitting an alpha particle, the problem still remains as to how the alpha particle can escape from the nucleus. The strong nuclear forces
dominate at very short distances inside the nucleus and this leads to a potential barrier. If we model
this barrier and the alpha particle as a particle in a box, the height of the box (or the potential barrier )
turns out to be around 25 MeV. This is much more than the kinetic energy of the alpha particle. Thus,
classically, there is no possible mechanism for the alpha particle to escape from the nucleus as shown
in Figure 2.1.

Figure 2.1: Nuclear Potential Barrier and particle tunnelling

However, we know that alpha particles do come out and are observed. It was Gamow who first
proposed the mechanism whereby this could be possible. Essentially, we can think of the alpha particle
tunnelling through the potential barrier, something which is permitted by quantum mechanics. Under
some very reasonable assumptions, the Gamow Theory of Alpha decay agrees remarkably well with
experimental observations regarding the behaviour of the decay constant with the energy of the alpha

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

67

particle. The higher the energy of the alpha particle emitted, the higher is the decay constant or the
shorter is the half life of the parent nuclide. Typically, when the energy is more than 6.5 MeV, the half
life is of the order of a few days. On the other hand, if the energy is less than 4 MeV, the probability
of the alpha particle tunnelling through the nuclear potential barrier is very small and the half life is
very long. This is shown in Table 2.2.

Half-Life

Alpha particle Energy (MeV)

Branching Ratio

148

93 years

3.1827

100%

4.196
4.149

77 %
23 %

5.168

76 %

5.124

24 %

5.80

76.4 %

5.76

23.6 %

Gd

238

238

U
240
Pu

4.5 10 years
4.5 109 years
6.5 103 years

240

Pu

6.5 10 years

244

Cm

18 years

244

18 years

Cm

Nu
cle
ar
Ph
ys
ics

Source

Table 2.2: Alpha Particle Sources


(Adapted from Radiation Detection & Measurement by Knoll )

2.2.2

Beta Decay

M
an
ua
l

Another process by which an unstable nucleus can become more stable is by the process of beta decay.
This is
A
ZX

A
Z+1Y

+ e +

La

Beta particles are electrons which are produced when a neutron goes to a proton. Initially, it was
thought that the only particles produced in this reaction were a proton and an electron. However, it was
observed that the beta particles come out with a range of energies from 0 upto a maximum energy called
the end point energy instead of having a monoenergetic spectrum. This was surprising since if the
only particles produced were the proton and the electron, then energy momentum conservation tell us
that the electron will be monoenergetic and that the electron and the recoiling nucleus would be moving
back to back. This was not the case. Further, the neutron, proton and electron are all fermions and
have spin 21 and hence the reaction with a neutron going to a proton and electron would not conserve
spin. Finally, as we know, the electron carries a charge called lepton number. The nucleons have 0 lepton number and thus producing only an electron in the decay would violate lepton number conservation.
All these anomalies were solved by introduction of a new particle called the neutrino by Pauli.
The neutrino is assumed to be massless, neutral and carries spin 21 and lepton number of 1. Thus the
process that we have is a neutron going to a proton, an electron and an antineutrino (the antiparticle
of the neutrino). With this, all the anomalous observations could be accounted for. Thus, the Q value

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

68

Nu
cle
ar
Ph
ys
ics

of the reaction, that is the mass difference between the parent nuclei and the daughter nuclei is the end
point energy of the electron. The anti neutrino carries some kinetic energy and the actual energy of
the electron is thus the Q value minus the kinetic energy of the anti neutrino. This being a three body
decay now (that is there are three particles in the final state), we get a continuous energy spectrum
for the electron upto the maximum of Q value. Again, since there are now three particles in the final
state, there is no reason for the electron and the recoiling nucleus to be moving back to back since the
anti neutrino carries some momentum too. Finally, lepton number conservation and spin conservation
is also taken care of because of the anti neutrino.

M
an
ua
l

n p + e +

Figure 2.2: Energy Distribution in Beta decay

(Adapted from Concepts of Modern Physics by Beiser, Mahajan & Rai Choudhury)

La

Each beta decay is characterised by a fixed Q value as we have seen above. This is normally the
case when the transition of the nucleus takes place between the ground states of both the parent and
the daughter nuclei. However, if the transition is between excited states and/or ground states, then the
spectrum will change because the Q value will be different. In practice, in some beta emitters, because
of the presence of excited states which could be populated, one gets several components with different
end-point energies.

Source
14

32

Half-Life

Endpoint Energy (MeV)

5730 years

0.156

14.28 days

1.710

36

Cl

3.08 10 years

0.714

45

Ca

165 days

0.252

Table 2.3: Beta particle Sources


(Adapted from Radiation Detection & Measurement by Knoll )

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

69

Another process which can take place inside the nucleus is positron emission. This is the conversion
of a proton into a neutron and a positron and a neutrino.
p n + e+ +

Nu
cle
ar
Ph
ys
ics

Obviously, since a proton is lighter than a neutron, this process can only take place inside a nucleus
and not for free protons. On the other hand, a free neutron does decay by emitting a proton and an
electron and antineutrino with a half life of around 10 minutes.
Finally, another related process is electron capture. This is when a nucleus absorbs an inner shell
electron and a proton and the electron go to a neutron and a neutrino.
p + e n +

Usually, since the K shell electrons are the ones which are captured, an electron from the outer
shell falls to fill the vacancy thereby releasing X-rays which are characteristic of the daughter nuclide.
Electron capture is more likely than positron emission in heavy nuclides since the inner shell electrons
are closer to the nucleus in heavy elements.

2.2.3

Gamma Decay

La

M
an
ua
l

We are familiar with atoms existing in excited states as well as ground state. When an electron in an
excited state returns to the ground state, a photon is emitted. In essentially the same way, a nucleus
can also exist in ground as well as excited states. When an excited nuclei returns to its ground state, it
emits photons of energies equal to the difference in the energies of the excited and ground state. We can
easily estimate this energy by using the Uncertainty Principle. A typical nucleus size is x 1015 m,
while a typical mass for a nucleon would be 1 GeV c2 . This gives us an estimate of the energy as
~2
10
2(x)
m and mass
2 m few MeV. Incidentally, a similar calculation for an atom, with x 10
2
1 MeV c for an electron would give us the energy estimate to be a few eV. Clearly, atomic transitions occur between energies separated by a few eV and those in the nucleus by a few MeV. Thus, as
we have seen, alpha and beta particles also carry energies in the MeV range.
Electromagnetic radiation which carries a few MeV of energy is said to be in the Gamma ray region
of the electromagnetic spectrum.
In most cases, some form of beta decay results in the creation of an excited state of the daughter
nuclei. This process happens in a time scale which is characteristic of the half life of beta emitters as
in Table 2.3. However, the excited states of the daughter nuclei thus created are short lived states and
decay to the ground state by emitting gamma rays. Thus what we see typically is that a parent nuclei

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

70

emits a beta particle and gamma ray photons with a time of the order of the half life. However, and
this is crucial, the energy of the beta particle is characteristic of the parent nuclei while that of the
gamma rays is determined by the energy levels of the daughter nuclei.

Nu
cle
ar
Ph
ys
ics

Two typical examples of gamma decay schemes are shown in Fig(2.3) and Fig(2.4).

Figure 2.3: Decay Scheme for

Co- Gamma decay

Figure 2.4: Decay Scheme for

27

Mg - Gamma decay

La

M
an
ua
l

(Source: Wikicommons)

60

(Adapted from Concepts of Modern Physics by Beiser, Mahajan & Rai Choudhury)

For instance, in the decay of 27


12Mg shown in Fig (2.4), the half-life of the decay is 9.5 minutes and
27
*
it can take place to either of the two excited states of 27
13Al. The excited aluminum nucleus, 13Al can
then decay by emitting one or two gamma rays and come to the ground state.
An excited nucleus can sometimes also give its energy to one of the atomic electrons around it. In that
case, the electron then is emitted with a kinetic energy equal to the nuclear excitation energy minus
the binding energy of the electron. This process, a sort of photoelectric effect for nuclear photons is
called internal conversion.
Thus we see that certain nuclei are radioactive and emit one or more of the above mentioned par-

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

71

ticles/radiation. These radiations and particles are what we use to study the properties of the parent
and daughter nuclei. However, to detect and measure the properties of these radioactive emissions, we
need to have them interact with matter. This is what we shall now turn to.

2.3

References

Nu
cle
ar
Ph
ys
ics

1. Radiation Detection & Measurement, Glenn F. Knoll, Wiley India (2009).

La

M
an
ua
l

2. Concepts of Modern Physics, Arthur Beiser, S. Mahajan & S. Rai Choudhury, Mcgraw Hill
(2015).

Chapter 3

3.1

Introduction

Nu
cle
ar
Ph
ys
ics

INTERACTION WITH MATTER

3.1.1

M
an
ua
l

The experiments that we perform in this laboratory are all concerned with the detection and measurement of radiation and particles, namely gamma rays and beta particles (We do not carry out experiments
with alpha particle emitters in this laboratory). Clearly, we need detectors for this purpose- in our laboratory, we use two kinds of detectors. Geiger-Muller counters (GM counters) and Scintillation counters.
We shall be studying in detail about the workings of these detectors in a later Chapter. In this Chapter,
we would like to understand in general how particles and radiation interact with matter. After all, if
radiation or a particle is to be detected, it must interact with the material of the detector. This might
be a gas, a liquid or even a solid. Before we do this, let us define some terms which we shall be using.

Cross Section

La

Cross section is a way to express the probability of interaction. Consider a beam of incident particles
impinging on a target material. We assume that each particle in the target material has a certain area,
which we call cross section. Any incident particle which passes within this area will interact with the
target particle. Clearly, the larger the area, the larger is the possibility of interaction. Of course, this
cross section which we think of as an area of influence in a way, depends on the nature of the process,
the nature and energy of the incident particle etc. In principle, it could be very different from the
geometric cross section.
We know that flux of a beam is defined as the number of particles crossing a unit area in unit time.
Suppose now we have a slab of some target material with area A and thickness x. Let the number of
atoms per unit volume of the target be n, that is the total number of nuclei or atoms in the slab are
nAx. If each nuclei has a cross section of , then the aggregate cross section for the slab is nAx.
Now consider a beam of incident particles, N of them hitting the slab. The number N which will
interact with the nuclei in the slab will be
72

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

73

N
Aggregate cross section
nAx
=
=
= nx
N
Target area
A
Since one incident particle is assumed to interact only once with any nuclei, and get deflected, it is
removed from the beam. Then for a finite slab thickness, we have
dN
N

N0

dN
N

Zx

Nu
cle
ar
Ph
ys
ics

ZN

= ndx
= n

dx

0
nx

N = N0 e

(3.1)

We see that the number of surviving particles decreases exponentially with the slab thickness. Recall
that this is exactly what we see when we pass a beam of light through some absorber. The intensity
falls exponentially with thickness.

La

M
an
ua
l

We can think of the cross section as in Figure (3.1) where an incident beam of flux F hitting a target.
d
(, ).
The differential cross section is d

Figure 3.1: Differential Cross Section

(Source: ScatteringDiagram by Original uploader was JabberWok at en.wikipedia - Transfered from en.wikipedia.
Licensed under CC BY-SA 3.0 via Wikimedia Commons
- https://commons.wikimedia.org/wiki/File:ScatteringDiagram.svg#/media/File:ScatteringDiagram.svg)

Then,

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

d
1 dNs
(, ) =
d
F d

d
(, )
dNs = F d
d

74

(3.2)

Nu
cle
ar
Ph
ys
ics

We should think of this equation as the number of particles that scatter, Ns , into a portion of solid
angle per unit time is equal to the flux of incident particles per unit area per unit time multiplied by
the probability (represented by a cross section area) that would scatter into that portion of solid angle.
We can integrate Eq(3.2) over all solid angles to get the total cross section . The usual unit for cross
section is a barn which is defined as

1 barn = 1028 m2

Please remember that this discussion of cross section, though in the context of scattering, is equally
applicable for absorption of incident particles or radiation by matter.

M
an
ua
l

If Nt is the number density of the target particles, then the probability that a single interaction
through a volume with thickness x is simply Nt x. A convenient parameter in this discussion is the
mass thickness. This quantity is often more convenient to use instead of thickness. Thus
mass thickness = x gm cm2

(3.3)

La

This is convenient because it tells us directly the effect of the absorber on the incident beam. Thus,
for instance, a beam travelling through 2 gm cm2 of air (of density = 0.012 gm cm3 ) has the same
effect as the beam passing through 2 gm cm2 of water, even though it passes through 1.67 m of air
and just 2 cm of water. Essentially, what we have done is to factor out the density of the absorber and
encapsulate that in the mass thickness.
We are interested in the passage of alpha, beta and gamma radiation through matter. Thus we can
divide our discussion into two parts- interaction of charged particles with matter and interaction of
radiation with matter. For charged particles, once again, we need to consider two cases of passage of
heavy particles (like alpha particles) and the passage of electrons or beta particles.

3.2

Interaction of Charged Particles with Matter

When a charged particle interacts with matter, two things happena) the particle loses energy traversing matter and

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

75

b)particle is deflected from its initial direction.


In general, there are two main processes which cause this. These are inelastic collisions with atomic
electrons in the material and also elastic scattering off the nuclei. There are some other processes which
contribute to the energy loss namely, Bremsstrahlung and nuclear reactions which are extremely rare.

3.2.1

Nu
cle
ar
Ph
ys
ics

As mentioned above, the processes which are dominant causes of energy loss are different for light and
heavy charged particles. For heavy charged particles it is basically the inelastic collision with atomic
electrons which are responsible for energy loss.

Interaction of Heavy charged particle with matter

M
an
ua
l

Consider a charged particle like an alpha particle entering a medium. The atoms of the absorbing
medium will have atomic electrons which will interact with the incoming charged particle. It is important to remember that the alpha particle interacts simultaneously with many atomic electrons. In
this process, some energy is transferred to the atomic electron and depending on the amount of energy
and the nature of the absorber, the atomic electron either gets into an excited state or in some cases,
may even become free that is, the atom gets ionised. This energy transferred is of course exactly the
energy that is lost by the incoming particle. However, as we shall see below, the maximum amount of
energy that is transferred in such a collision is very small compared to the kinetic energy of the incoming particle. This means that the incoming particle continuously loses small amounts of energy as it
passes through the absorber. However, the number of such encounters in any macroscopic length is large.

La

Let us consider one such encounter in some detail. Consider a particle of mass M and velocity V
incident on another particle at rest of mass m, M  m. This could be the case, for instance of an alpha
particle colliding with an electron. Classically, using non-relativistic energy momentum considerations,
we know that
1
1
1
M V 2 = M V12 + mv22
2
2
2
M V = M V1 + mv2

where V is the initial velocity of the incoming particle, V1 is the velocity of the alpha particle after
the collision and v2 is the velocity of the electron after the collision. Solving these two equations for the
velocity after collision of the incoming particle, V1 , we get
M m
V
M +m
Thus, the loss of energy of the incoming particle is
V1 =

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

76


1 
4M mE
E = M V 2 V12 =
2
(M + m)2

Nu
cle
ar
Ph
ys
ics

where E is the initial kinetic energy of the incoming particle. Thus, we see that in the case when
M  m, we get that the maximum energy transferred to the electron is 4m
E. This justifies our stateM
ment above that in the case of alpha particles (or protons) and electrons, each collision only diminishes
1
the incoming particles energy by a small amount, in this case about 500
of the initial energy. It also
implies that the electrons velocity after the collision can be at most 2V since the electron was at rest
initially and all this energy E is the electrons kinetic energy after collision.

M
an
ua
l

Given that in any one collision, the energy transferred is small, we can also see that the momentum
P
transferred in any one collision, p m(2V ) 2m M
which is again small since M  m. Thus we can
assume that the incoming particle does not get deflected in any one collision and continues along the
undeflected.

Figure 3.2: Heavy particle in matter

La

Consider a heavy particle of mass M and charge ze entering a cylinder of absorber material with a
velocity v along the x direction. Consider a cylinder of radius b and length dx placed along the x axis as
shown in Fig 3.2. Let us look at the interaction of the heavy particle with a single atomic electron at a
distance b from the path of the incoming particle. This interaction is the Coulomb interaction between
the heavy particle and the electron and so we need to find out the field at the surface of the cylinder,
which is the location of the electron. The field is easily computed using Gausss Law as the symmetry
considerations tell us that the x component vanishes.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

77

ze
0
Z
ze
Ey 2bdx =
0
Z
2ze
Ey dx =
40 b
k2ze
=
b
where k =

Nu
cle
ar
Ph
ys
ics

Ey dA =

1
.
40

(3.4)

Now consider the impulse produced on the electron as a result of this Coulomb force.
Z

Fy dt
Z
= e Ey dt
Z
dx
= e Ey
v
2ze2 k
=
vb

M
an
ua
l

I =

using Eq(3.4).

(3.5)

Now the energy gained by the electron (which is at a distance b from the incoming particle) is simply
I2
2me
4z 2 e4 k 2
=
2v 2 b2 me
2z 2 e4 k 2
= 2 2
b v me

La

E(b) =

(3.6)

This is the energy loss to a single electron. We need to find out the energy loss of the incident
particle when it travels a distance dx in the absorber material. For this we need to add the energy lost
to all the electrons in the annular thickness of the cylinder between b and b + db. If Ne is the electron
density in the material, then this energy loss is given by

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

78

dE(b) = E(b)Ne dV
= E(b)Ne 2bdbdx
2z 2 e4 k 2
= 2 2 Ne 2bdbdx
b v me
4z 2 e4 k 2 db
Ne dx
=
me v 2
b

(3.7)

Nu
cle
ar
Ph
ys
ics

To find the total loss, we need to integrate over values of b and thus we get
dE
4z 2 e4 k 2

=
Ne ln
dx
me v 2

bmax
bmin

(3.8)

Clearly, the values of bmax and bmin will be determined by physical considerations. These are easily obtained. Consider the maximum energy lost by the incoming particle (or the maximum energy
2
transferred to the electron). This, we have already seen is given by me (2v)
. This will happen when the
2
impact parameter is the minimum. Thus we have, using Eq(3.6),
Tmax = 2me v 2 =

2z 2 e4 k 2
b2min v 2 me

(3.9)

M
an
ua
l

where Tmax is the maximum energy that can be transferred.


bmin =

kze2
me v 2

La

Now, for the maximum impact parameter, bmax , clearly, if the distance is very large and the energy
transfer is smaller than the ionization or excitation energy Ie , then no energy is transferred. Thus,
again using Eq(3.6), we get

Ie =

2z 2 e4 k 2
b2max v 2 me
(3.10)

Thus we have for the energy loss


dE
4k 2 z 2 e4
=
Ne ln
dx
me v 2

Tmax
Ie

4k 2 z 2 e4
=
Ne ln
me v 2

2me v 2
Ie


(3.11)

since as we have seen Tmax = 2me v 2 . This is the classical formula for the energy loss of a charged
particle in an absorbing material as given by Bohr. Of course this formula assumes that classical,
non-relativistic, particle must be heavy compared to me , that the interaction time is short thus the
electrons are stationary and also does not account for binding of atomic electrons. We can find the

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

79

number density Ne of the electrons in the material as


Ne =

ZNA
AmN

where is the density of the absorber material with atomic weight A and atomic number Z and NA
is the Avogadros number.

Nu
cle
ar
Ph
ys
ics

The classical formula of Bohr was extended by Bethe and Bloch to include relativity and other effects
like density effects and the effect of atomic electrons etc. When one uses relativity to find the maximum
energy transferred (and hence bmin ) by a particle with mass M interacting with a particle of mass m,
we get instead of Eq(3.9),
Tmax =
where = 1

1 2

and = vc .

2m 2 2

m
1 + 2 M
+


m 2
M

(3.12)

The basic result is the Bethe-Bloch formula

M
an
ua
l

dE
4k 2 e4 z 2 ZNA
=
B(v)
dx
me c2 2 A


2me v 2
B(v) = ln
ln(1 2 ) 2
Ie

(3.13)
(3.14)

Clearly, as we have seen, for non-relativistic particles, only the first term in the factor B(v) is significant (that is  1).

La

This is the basic formula to estimate the energy loss by a charged particle in matter. We can rewrite
the Bethe-Block formula in another way. Remember that the classical electron radius re is defined as
re =

ke2
m e c2

In terms of re , we can rewrite Eq(3.13) as

1 dE
z2 Z
= 4re2 me c2 NA 2 B(v)
dx
A

(3.15)

Let us see what this equation for the energy loss by a charged particle in matter tells us in general.
Z
12 except for hydrogen for which it is 1. Thus we can
Note that most materials have the same A
2
immediately say that apart from the corrections due to B(v), the energy loss depends on z 2 since all
the other factors are constants. That is, for a given non-relativistic particle, the energy loss
varies inversely with particle energy since it varies inversely with v 2 . This can be easily

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

80

Nu
cle
ar
Ph
ys
ics

understood since if the particle has higher energy, it spends less time near any one electron
and hence the impulse on the electron is less and the energy transfer is lower. Slow moving particles lose more energy and as their momentum increases and thus their velocity
approaches c, we expect a flattening of the dE
curve. We can see this in the energy loss
dx
curves for various charged particles as shown in Fig 3.4. The second point is the variation with
z 2 . Energy loss depends directly on z 2 . Thus alpha particles with the greatest charge will have
the highest energy transfer (among particles with the same energy or velocity). Finally,
when we compare different absorbers, the quantity which is entering the energy loss is the
product N Z. This obviously is a measure of the electron density in the absorber. Hence,
materials with high N Z or higher electron density will have a higher stopping power or
energy loss for a given beam.

La

M
an
ua
l

Another interesting
 thing
 about the Bethe-Bloch formula is the logarithmic term. The first term in
2me v 2
the factor B(v) is ln
shows that there is a slow rise in the energy loss with increased momentum
Ie
or velocity. The reason for this is actually the behaviour of a relativistic particles electric field as the
velocity increases. We know that the electric field of the incoming charged particle becomes more and
more squashed as the velocity increases. The reason for this is that the transverse field, Ey , is larger
by Ey relative to the isotropic case of the charge at rest while the longitudinal field Ex decreases by a
factor of 12 relative to the isotropic case as shown in Figure 3.3. The typical whiskbroom pattern of the
electric field of a moving charge is shown in the Figure 3.3. This leads to a slow increase in ionization
at farther and farther distances from the particle track.

Figure 3.3: Electric field of moving charge

(Source: http:// sernam.ru/ lect f phis6.php? id=65 )

From Figure 3.4, we see that the typical specific energy loss at the minimum ( for most particles
except the electron) is around 2 MeV cm2 gm1 . Since most relativistic particles have very similar
behaviour (in terms of energy loss as a function of velocity or energy), we refer to these relativistic
particles ( > 3) as minimum ionizing particles.

81

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Figure 3.4: Specific energy loss in air versus energy of charged particles
(Source: Review of Modern Physics, Vol 24, page 273 )

M
an
ua
l

The Bethe-Bloch formula is a good guide to understand the energy loss of charged particles in matter.
However, there is another factor which comes into play as the charged particle moves through matter.
This is reduction in the effective charge. To understand this, consider Figure 3.5.

La

Figure 3.5: Energy loss of Alpha particles of energy 5.49 MeV

(Source: Wikicommons)

There are several things about this figure that one needs to understand. Firstly, as the alpha particles
move through matter, we expect from the Bethe-Bloch formula that the energy loss increases as E1 . We
see this in the initial part of the plot. However, as the particle keeps moving in the material, its velocity
and hence energy decreases and the energy loss increases to a maximum near the end of the track. At
that point, we see that the energy loss sharply falls off. This is because the alpha particles are now
travelling slow enough that they can pick up electrons from the material and thus the effective charge is
reduced. We expect that charged particles with the maximum nuclear charge will pick up the electrons
earlier than those with less nuclear charge. This is indeed seen as we can see in Figure 3.6.

82

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Figure 3.6: Energy loss for helium & hydrogen ions.


(Source: B.Wilken, T.A. Fritz, Nuclear Instrumentation Methods, 138, pp 331 (1976))

La

M
an
ua
l

Finally, we consider the phenomena of straggling. When we pass a beam of charged particles of a
fixed energy (instead of a single particle) through different materials and study the number transmitted,
we find an interesting feature. We see that energy loss is not continuous but statistical in nature and
some particles undergo less/more energy loss and their range will be larger/smaller than the typical,
expected range. This is known as straggling.

Figure 3.7: Energy Straggling

(Source: DEVELOPMENT And IMPLEMENTAZIONE IN C++ OF ALGORITHMS FOR The CALCULATION OF PLANS
OF TREATMENT IN ADROTERAPIA Giovanni Nicco, Bachelors Thesis, University of Turin, 2000. )

As we can see from Figure 3.7, the variation in the amount of collisions/energy loss is approximately
Gaussian (because of the central limit theorem). We define the mean range as the distance at which
50% of the particles are transmitted, as shown in Figure 3.7. We can find the range of transmission by
extrapolation to the distance at which we expect zero transmission as in the Figure.
The discussion above is strictly valid for heavy charged particles and hence will be a good description

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

83

for the energy loss for a beam of alpha particles. However, things are different for beta particles.

3.2.2

Interaction with matter of electrons

Nu
cle
ar
Ph
ys
ics

In our discussion of the Bethe-Bloch formula, we had assumed that the incoming charged particle is
much heavier than the atomic electron in the absorber and therefore does not suffer any significant deviation from its straight line path. However, when one is interested in the interaction of beta particles
or electrons with matter, this assumption clearly is invalid. Here, the beam and the target particles are
identical and hence large deviations can be expected from the beam path. In addition, the interaction
between the beam and the nucleus of the absorber can lead to sudden changes in the path. The target
and the beam particles are identical and indistinguishable and this needs to be taken into account.
Finally, at the energies that we are interested in, namely nuclear energies (remember the beta particles
are in the MeV range), the electrons are always relativistic.

M
an
ua
l

The above differences relate to the considerations of energy loss by collisions with atomic electrons
as in the case of alpha particles. However, since the electrons are light and therefore suffer substantial
acceleration (or deceleration) because of their interaction with the absorber, they emit radiation as we
know from the well known Larmor formula in classical electrodynamics. The emission of radiation because of deflections (and hence acceleration) leads to energy loss. These are of two kinds: Cherenkov
radiation and Bremsstrahlung. To find out the energy loss of an electron beam in matter, we need
to take into account all these processes.

with

La

To find out the energy loss, we need to modify the Bethe-Bloch formula to take into account all the
above mentioned differences between a heavy particle interaction and electron interaction. Recall the
Bethe-Bloch formula Eq (3.13) which we had for heavy particle interaction with matter,
dE
4k 2 e4 z 2 ZNA
=
B(v)
dx
m e c2 2 A


B(v) = ln

2me v 2
Ie

ln(1 2 ) 2

This gets modified in the case of electrons to


dE
4k 2 e4 z 2 ZNA
=
B(v)
dx
m c2 2 A
 e

2 
p
p
me v 2 E
1
2
2
(ln 2)(2 1 2 1 + ) + (1 ) +
B(v) = ln
1 1 2
(3.16)
2I 2 (1 2 )
8

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

84

Nu
cle
ar
Ph
ys
ics

This is a much more complicated formula for energy loss due to collisions of electrons in a beam
with atomic electrons in an absorbing material. Another interesting difference is in the straggling
for electrons as compared with heavy charged particles. Since the electron mass is small, there is a
significant fractional energy loss in each collision than for heavier particles. Hence we see a lot more
straggling for electrons, Figure 3.8.

Figure 3.8: Energy Straggling for particles


(Source: www.physics.queensu.ca/ phys352/ )

La

M
an
ua
l

Apart from the collisions suffered by the beta particles, they also radiate and hence lose energy
because of the two processes mentioned above. Cherenkov radiation is the electromagnetic shock wave
generated by a particle in a medium when it moves faster than the velocity of light in that medium.
That is when the velocity of the particle, v > nc where n is the refractive index of the material. This radiation, which is the familiar blue radiation that one sees in pictures of nuclear reactors water shield, goes
as 12 and thus peaks at small wavelengths or in the ultraviolet. In the case of electrons in a medium, the
loss due to Cherenkov radiation is small, around 1% and so we will ignore this. As it turns out, in the full
Bethe-Bloch formula for energy loss, this radiative loss due to Cherenkov radiation is already accounted.
The main radiative loss for an electron is due to Bremsstrahlung. This braking radiation is what
the electron emits when it is accelerated due to the strong nuclear electric field, though electron-electron
Bremsstrahlung is also possible. One can use the classical theory of radiation by an accelerated charge
(Larmors formula) to calculate the radiative loss due to this process. It turns out that the energy loss
is given by


dE
dx


r

N EZ 2 e4
=
m2 c4

2E
4
4 ln

2
mc
3


(3.17)

Clearly then, the total energy loss is the sum of the energy loss due to collisions (Eq(3.16)) and that
due to Bremsstrahlung (Eq (3.17)). That is

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

dE
=
dx

85

dE
dx


+

dE
dx


(3.18)
r

Nu
cle
ar
Ph
ys
ics

If we compare the two expressions for energy loss of beta particles, Eq(3.16) and Eq(3.17), we notice
some important features. Firstly, the collisional energy loss increases logrithmically with the energy E
and linearly with the atomic number of the absorber, Z while the energy loss due to radiation increases
linearly with the energy E and as the square of the atomic number Z. The radiative energy loss also
has a mass factor in the denominator which implies that the radiative energy losses are most for lighter
particles (like beta particles) than for heavier particle (like alpha particles). Also radiative energy losses
are most for high energy particles and in materials of high Z.
This also allows us to define a critical energy, Ecrit for which


dE
dx

dE
dx

An approximate formula for Ecrit is given by

1600mc2
Z
The value of the critical energy for various materials is given in Table 3.1.
Ecrit

Ecrit (MeV)

M
an
ua
l

Material
Cu

24.8

Pb

9.51

air (STP)

102

plastic
water

100
92

La

Table 3.1: Critical energy for various materials

We can also take the ratio of the energy losses by the two processes and we find that
dE
dx r

dE
dx c

EZ
700

where the energy E is in MeV. This clearly shows that for typical beta particles with energies typical
of nuclei (a few MeV), the radiative losses are much smaller than those by collisions except in the case
of very high Z materials.
As we have discussed above, the electron being of the same mass as the scattering particles (atomic
electrons), can suffer large deviations from its original path. Thus, if one has a beam of mono-energetic

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

86

electrons from a source and we pass it through an absorber, even a thin absorber can lead to a loss of
electrons from the detector. This scattering therefore means that the intensity of the transmitted beam
drops immediately and goes to zero as the absorber thickness is increased.

Nu
cle
ar
Ph
ys
ics

A corollary to this random motion of the electron is therefore that the range of the electron is hard
to define since it might have travelled a much larger distance within the absorber than simply the
thickness of the absorber. What is normally done is to extrapolate the range from the linear portion of
the graph.
To summarise then, we can say that the energy loss of electrons is much smaller than that of heavy
charged particles of the same energy. This means that they have a much larger range. What is observed
experimentally is that for a wide variety of absorber materials, the product of the range and the density
of the absorber is a constant for any particular electron energy.

M
an
ua
l

The situation is very different for the beta particles emitted by a radioactive source. This is because,
as we have seen in Section 2.2.2 in Fig 2.2, the energy spectrum of the beta particles is continuous.
What is seen therefore is that the beta particles at the lower end of the spectrum are absorbed even
with a very thin absorber. However, for the most part of the spectrum, the transmission of the beta
particles shows an exponential decrease with thickness. This is an experimental fact which cannot be
derived easily from fundamental physics. What we see is that the counting rate (or intensity) falls
off exponentially with an attenuation coefficient which depends on the end point energy of the beta
particle.
C = C0 end

(3.19)

La

where C is the counting rate with the absorber material, C0 is the counting rate without the absorber
and d is the mass thickness in units of mass per unit area. The coefficient n is the attenuation coefficient.
This behaviour is shown in Fig 3.9.

87

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Figure 3.9: Range-Energy for 1.17 MeV beta particles from

210

Bi in Al absorbers

(Source: Radiation Protection, course at Oregon State University Extended Campus )

3.2.3

Interaction of gamma rays with matter

M
an
ua
l

The interaction of electromagnetic radiation like gamma rays with matter is fundamentally different
than what we have studied so far. The interaction in this case is between radiation and the charged
particles like atomic electrons unlike between two charged particles that we have been investigating
so far. The interaction of electromagnetic radiation with matter depends on its frequency (and hence
energy). Gamma rays, as we have seen in Section 2.2.3 have typical energies in the MeV range since their
origin is in the de-excitation of the nucleus. In general, the interaction of electromagnetic radiation with
matter can be of four kinds depending on the energy of the radiation. These are Rayleigh Scattering,
Photoelectric Absorption, Compton Scattering and Pair Production. The relevant photon
energies for which these processes are operative are given in Table 3.2 ( EI is the ionization energy of
the atom).
Rayleigh Scattering

Photoelectric Absorption

Compton Scattering

Pair Production

h > EI

h me c

h > 2me c2

eV

keV

MeV

MeV

Visible

X-rays

rays

Hard rays

h < EI

La

Table 3.2: Interaction of Photons with matter

The fact that the three processes listed above are dominant in different energy regimes can be clearly
seen if we plot the variation of the cross section for the three processes with the gamma ray energy.
This is shown in Fig 3.10.

88

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Figure 3.10: Variation of the cross sections with energy for different interaction processes of photons with
matter

To understand these processes of course one needs to develop a fully relativistic theory of the interaction of photons and electrons, which is known as quantum electrodynamics. However, for our purposes,
a simple classical description is sufficient to provide some insights into the mechanisms. We think of
the electron in an atom as a dipole which is attached to the nucleus by a spring and is oscillating with
a frequency 0 around its mean position. This natural frequency of oscillation can be modelled by a
linear restoring force kx on the electron, with the spring constant k and 0 being related as

M
an
ua
l

02 =

k
me

La

Of course, k is related to the attractive electric force between the electron and the nucleus and is
related to the binding energy of the electron. Now when the photon or electromagnetic wave is incident
upon such an oscillating dipole, the electric field of the wave exerts an added force on the electron,
eE(t) where E(t) is the oscillating electric field of the electromagnetic wave given by E(t) = E0 sin t
where is the frequency of electromagnetic wave. The equation of motion for the electron is then

me x = kx eE(t)
e
x + 02 x = E(t)
me

(3.20)

The harmonic solutions to this equation of a forced oscillator are well known as x(t) = A sin t which
we can substitute in Eq(3.20) and get

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

89

e
E0
me
e
1
A =
E0
2
2
0 me
1
e
x(t) =
E0 sin t
2
2
0 me

(02 2 )A =

(3.21)

Nu
cle
ar
Ph
ys
ics

We know from classical electromagnetic theory that an accelerated charge radiates energy. The
power radiated is given by Larmors formula as
2 e2 2
P =
a
3 c3
where a is the acceleration. In our case, the time averaged acceleration squared is simply

2
a =
which gives us the radiated power as
1
P =
3

e
2
E0
2
2
0 me

e2
me c2

2

M
an
ua
l

Now we can think of the power radiated

2

1
2

4
cE 2
( 2 02 )2 0

(3.22)

P = I

where is the cross section of the interaction and I is the intensity of the incoming radiation. We
know that

La

I=

since I = uc with u =

E02
2

cE02
8

being the energy density of the radiation. Thus we get


cE02
P =
2

or
8
=
3

e2
me c2

2

4
( 2 02 )2

(3.23)

This classical cross section can be expressed in terms of the classical radius of the electron re =
as

e2
me c2

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

2
4re2

90

2
( 2 02 )

2
(3.24)

This is the expression which describes the interaction of an electromagnetic wave with an electron.
Consider Rayleigh Scattering first. This, as we have seen happens when the incoming energy is
much less than the binding energy of the electron,  0 . In this limit,
2
( 2 02 )

2

4
04

Nu
cle
ar
Ph
ys
ics


and the cross section is simply

8re2 4
3 04

Ray =

(3.25)

This is the famous Rayleigh scattering cross section where we can see the dependence on the wavelength of the electromagnetic wave. This is what gives us that blue colour of the sky where blue light
is scattered the most.

M
an
ua
l

Rayleigh Scattering occurs when the electromagnetic wave has a much smaller energy than the
binding energy. What about if the electromagnetic radiation is energetic enough to ionize the atom but
not energetic enough for it to accelerate it to relativistic speeds. This is the domain ~0  ~  me c2 .
This kind of scattering is called Thomoson Scattering In this limit,


2
( 2 02 )

1
02
2

Thus the cross section for Thomson Scattering becomes

La

8re2
(3.26)
3
This is a remarkable expression since it is totally independent of the frequency of the incoming radiation provided the condition above is met. In fact, it has a fixed value of about T 23 barn. In fact,
Thompson scattering is basically the low energy limit of Compton scattering which we shall explore a
little later. In Thompson scattering, the photon is scattered elastically while Compton scattering is the
inelastic scattering of the photon from an electron. (Recall that in elastic scattering of particles, the
kinetic energy of the particle is conserved in the center of mass frame though not in the lab frame.)
T =

The energies what we have encountered above are clearly too low for the gamma rays that we are
interested in since we have seen that the gamma ray energies are typically in the MeV range (Figure 2.3
for instance). Thus, neither of these two processes of elastic collision of photon and electron are impor-

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

91

tant for gamma rays. For gamma rays, the three processes which play a significant role in the energy
loss in matter. These are Photoelectric Absorption, Compton Effect and Pair Production. It
is important to note that unlike the case of alpha and beta particles, where the energy of the incoming
particle is lost gradually due to continuous interaction with the electrons in the material, in the case of
gamma rays, the photon can simply disappear or get scattered significantly abruptly.
Photoelectric Absorption

Nu
cle
ar
Ph
ys
ics

The classical theory that we have described above of the interaction of the photon and the electron
breaks down when 0 since the denominator blows up. This is the resonance condition where
the theory outlined above is not valid. In this case, the photon interacting with the atomic electron
transfers all its energy to the electron and disappears. If the photon energy was more than Eb , the
binding energy of the electron it interacts with, then of course the balance energy manifests itself as
the kinetic energy of the free electron, Ee . That is
E = h = Ee + Eb

M
an
ua
l

It turns out that to properly understand the process, one needs quantum mechanics to calculate
the cross section for photoelectric absorption. If a proper quantum mechanical calculation is done, we
see that the cross section PE Z 5 where Z is the atomic number of the absorber. Thus we can see
that if we want an absorber to absorb gamma rays of the relevant energies, then we need to use high Z
materials like lead etc. In fact, though no analytical expression exists for the probability of photoelectric
absorption per atom for all photon energies, an approximate empirical formula can be obtained which
shows the relationship of the probability and the energy E and atomic number of the absorber Z as

Zn
E3.5

(3.27)

La

where n is between 4 and 5.


Clearly, the probability of interaction will increase when the gamma ray energy is close to the energy
of one of the bound electrons. Thus we should see peaks in the cross section of interaction near the
energies corresponding to the K, L, M shells of the absorber. This is indeed what is seen. When a
photoelectron is ejected, typically from an inner shell, it creates a vacancy in that shell. This vacancy is
filled either by a free electron which is captured or by an outer electron falling into the vacant position.
This results in the production of characteristic X-rays which it turns out are mostly absorbed in the
material itself.
To conclude, photoelectric absorption is the main process by which gamma rays of relatively low

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

92

energy interact with matter (the cross section in Eq(3.27) varies inversely with energy and so at high
energies is small). The probability depends strongly on the atomic number of the absorber and also at
the binding energies of the various shells in the absorber atoms.
Compton Scattering

Nu
cle
ar
Ph
ys
ics

For gamma rays of energies that we typically encounter in radioactive decay, the dominant process
for interaction is Compton Scattering. This, as we know, is the interaction of an atomic electron with
the incoming photon, in which the electron acquires enough energy to become free and also relativistic. The scattering is therefore inelastic ( as opposed to Rayleigh and Thomson scattering which were
elastic). The photon transfers a part of its energy to the atomic electron and thereby loses energy to
the recoil electron. The process can be analysed using relativistic energy momentum conservation as
follows. We perform the calculation for the case of a free, unbound electron for simplicity.
i.
Let the initial 4-momentum of the electron be Q

i = (me , 0)
Q

f . Now
Let the final 4-momentum of the electron after being hit by the photon be Q

M
an
ua
l

2i = Q
2f = m2e
Q

Let Pi and Pf be the initial and final four momenta of the photon. Then

La

and

Pi = (pi , pi n
i)

Pf = (pf , pf n
f )

where n
i and n
f are the unit vectors in the direction of the initial and final photon 3-momentum
respectively.
For the photon, we know that
Pi2 = Pf2 = E 2 p2 c2 = 0
Now

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

93

f = Q
i + Pi Pf
Q
2f = Pi2 + Pf2 + Q
2i 2Pi Pf + 2Q
i (Pi Pf )
Q
i (Pi Pf )
Pi Pf = Q

Nu
cle
ar
Ph
ys
ics

pi pf pi pf cos = me (pi pf )
1
1

2 sin2 (/2) = me
pf
pi
 1
1 
2
2
2 sin (/2) = me c

hf
hi

(3.28)

where is the scattering angle of the photon. We can use the above expressions to get the energy of
the scattered photon as
hf =

hi

1 + mhe ci2 (1 cos )


(3.29)

or, the recoiling electron energy which is simply the kinetic energy


(1 cos )



Ee = hi hf = hi
hi
1 + me c2 (1 cos )
hi
me c2

(3.30)

M
an
ua
l

The calculation of the cross section for Compton Scattering once again needs to take into account
relativistic and quantum mechanical effects. The result is a very complicated formula called the KleinNishina formula which is
2
d
1
2 1 + cos
= re
d
2
[1 + hi (1 cos )]2


1+

hi2 (1 cos )2
(1 + cos2 )[1 + hi (1 cos )]


(3.31)

La

by

This is of course a very complicated expression. A useful way to remember this is to approximate it

me c2
h
Thus we expect the cross section to decrease at high energies.
Comp T

(3.32)

It is instructive to consider two extreme cases for Compton scattering- one in which the photon is
hardly scattered, that is 0 and the other of back scattering, that is . For 0, we can see
from Eq(3.26) and Eq(3.30), that
hf
= hi

(3.33)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

94

and
Ee 0

(3.34)

There is hardly any energy transfer as is expected in such a case.


For , the photon is back scattered and
hf =

(3.35)

Nu
cle
ar
Ph
ys
ics

and

hi
1 + 2 mhe ci2

Ee = hi

2hi
m e c2
2hi
+m
2
ec

(3.36)

M
an
ua
l

This is the case for the maximum energy transfer from the photon to the electron. Figure 3.11 shows
how the energies of the scattered photon and recoil electron vary.

Figure 3.11: Compton Effect energies as a function of scattering angle

La

(Source: Wikipedia)

Of course, when gamma rays pass through any absorber, scattering will occur at all angles and thus if
one was to measure the energy spectrum of the recoiling electron, we would find a continuous spectrum
till the maximum energy given by Eq(3.36). If one takes a monoenergetic gamma ray, then the energy
spectrum will look like that in Figure 3.12.

95

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Figure 3.12: Compton Effect- Energy spectrum of recoiling electron

(Source:By flutefreek (Flutefreek) (Own work) [GFDL (http://www.gnu.org/copyleft/fdl.html), CC-BY-SA-3.0


(http://creativecommons.org/licenses/by-sa/3.0/)

or CC BY-SA 2.5-2.0-1.0 (http://creativecommons.org/licenses/by-sa/2.5-2.0-1.0)],


via Wikimedia Commons)

In this figure, the difference between the Compton edge and the full energy peak is Ec and is given
by

M
an
ua
l

Ec = hi hi

2hi
me c2
2hi
+m
2
ec

hi
2hi
1+ m
2
ec

(3.37)

or for large gamma ray energies (hi  me c2 /2), we get


Ec

me c2
= 0.256 MeV
2

La

Pair Production
For high energy gamma rays, another process can result in the loss of gamma ray photons. This
is the production of an electron-positron pair by the photon in the presence of a nucleus. An isolated
photon cannot produce a particle-antiparticle pair because of conservation of energy and momentum.
To see this, just go to the center of mass frame in which the electron and positron fly back to back
with equal momentum. The final momentum in this frame is zero. However, the photon, in any frame
always moves with the velocity c and has a momentum equal to the energy of the photon (in units
where c = 1). Thus it is not possible. If there is another nucleus or another particle then the situation
is completely different and that nucleus or particle can recoil and balance the momentum.
By energy considerations, the minimum energy of the photon to be able to produce a pair of electron
and positron would be equal to twice the mass of the electron, that is 1.02 MeV. Of course, this is
the threshold energy and it is only when the photon energies are of several MeV that pair production

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

96

becomes significant. The excess energy of the photon, above 1.02 MeV is shared by the electron-positron
pair as kinetic energy.

Nu
cle
ar
Ph
ys
ics

An expression for the pair production cross section is not easy to find. However, it is known that the
cross section in the field of an atomic nucleus varies as Z 2 and so absorbers with high atomic number
will give an increased energy loss due to this process. It is also known that cross section at low energies goes as the logarithm of the photon energy. Finally, since we have seen above that the Compton
Scattering cross section varies as 1 , pair production is significant at higher energies when the the
Compton effect cross section falls off.
In practice, when a high energy gamma ray passes through an absorber and pair production takes
place, what we see are two photons emerging as secondary products of the interaction. This is because
the positron which is produced in the pair production process, typically interacts with an electron soon
after being produced and produces two photons (again, it cannot produce one photon due to energy
momentum conservation).

La

M
an
ua
l

Thus far, we have studied five processes which are operative when electromagnetic radiation is incident on a material. We have already seen that for gamma rays produced in radioactive decays, Rayleigh
and Thomson scattering are unimportant. The three process of importance are the photoelectric absorption, Compton Effect and Pair Production. All of these are operative but depending on the energy
of the incoming gamma rays, their cross sections vary. This is shown in Figure 3.13.

Figure 3.13: Relative importance of interactions of gamma rays

We now have some understanding of how


alpha, beta and gamma rays interact with matter. We next turn to a study of the Gieger-Muller counter
which is the mainstay of our laboratory. We will study its working and see how it can be used to detect
these radioactive emissions.
(Source: Encyclopedia of Occupational Health & Safety, www.icocis.org)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

3.3

References

La

M
an
ua
l

Nu
cle
ar
Ph
ys
ics

1. Radiation Detection & Measurement, Glenn F. Knoll, Wiley India (2009).

97

G-M COUNTER
4.1

Introduction

Nu
cle
ar
Ph
ys
ics

Chapter 4

In our laboratory, we use the Geiger-Muller counter (GM Counter) to study the radiation from
radioactive sources. The GM counter is a kind of counter or detector which uses the ionization produced
in a gas by the radiation to detect and study the radiation from the sources. It is a very convenient
detector or counter but can not be used for studying the energy characteristics of the radiation and is
only used to detect and count.

Detector Models

La

4.2

M
an
ua
l

Before we study the GM counter, it is instructive to understand the general principles of ionisation
of a gas by radiation since these are used in the GM counter as well. In addition to the GM counter,
other kinds of detectors also use this phenomenon. Thus, ion chambers and proportional counters are
also based on this process. We first look at a generalised model of a detector and then investigate the
process of ionisation by radiation in gas filled detectors before studying the GM counter in detail.

Fundamentally, all detectors of nuclear radiation work on the principle of the radiation (alpha, beta
or gamma rays) interacting with matter as discussed in Chapter 3 and depositing its energy in some
form in the material. This could lead to an ionisation of the material, or the production of photons etc.
Typically we detect and measure this change in the detector material and from that infer the properties
of the incoming radiation.
In most detectors of interest to us, the net result is the production and collection of charge in the
detector. The charge Q produced by the interaction at some time t = 0 is collected by the presence of the
field which separates the two kinds of charges produced in the detector and collects them. Depending
on the detector, the time taken to collect the charge will obviously vary. The flow of this charge will
98

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

99

lead to a current which last till the time the charge is collected and whose integral over the time from
t = 0 to the charge collection time tc will give us the amount of charge deposited and collected.
Ztc
Q=

I(t)dt

Nu
cle
ar
Ph
ys
ics

Figure 4.1: Current versus time

La

M
an
ua
l

A pulse produced in the detector is due to one single interaction and we normally assume that the
rate of the ionising radiation entering the detector is low enough that we can distinguish between various
pulses as shown in Figure 4.2. Clearly, the size and the duration of each pulse depends on the actual
interaction taking place between the detector and the ionising radiation.

Figure 4.2: Current versus time

We can distinguish between three different modes of the operation of detectors:

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

100

1. Current Mode
2. Pulse Mode
3. Mean Square Voltage Mode

Nu
cle
ar
Ph
ys
ics

In the Current Mode, one basically measures a detector signal as a current measurement. In this
mode, a current meter is connected to the detector output. Since the current level is in picoAmperes
or nanoAmperes, a precise meter is required. Given that the current meter has a response time T , the
observed current from a sequence of events at time t will be
= 1
I(t)
T

Zt

I(t0 )dt0

tT

The response time is usually longer than the time between individual detection events, so that an
average current is recorded at a time t. The current mode is used when event rates are very
high, which makes a stable current.

M
an
ua
l

In the Pulse Mode, the information on energy and timing of individual events, i.e. the
information on the signal amplitude and time of occurrence is usually recorded. In this
mode, the detector records the charge from each individual ionising event. This mode is used when
we need to get the information on the timing and the amplitude of individual pulses. Thus, given this
property, this mode is not suitable for high count rates since then the time between neighbouring events
may not allow for getting that information. Because the information on the charge collected in each
individual event is recorded, this mode is used for energy measurements and spectrum.

La

The signal shape from a radiation detector depends on the electronics to which the detector is
connected as well as the detector response. In most cases, the the input stage of the electronics is
an RC circuit where the resistance is the total input resistance and the capacitance C is the total
capacitance of the detector, cables, electronics etc as shown in Figure 4.3. Here V (t) is the time
dependent voltage across the load R and is the signal produced.

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

101

Figure 4.3: Detector Circuit

It is easy to see that the time dependent voltage V (t) is given by


V (t) = V0 (1 et/RC )

(4.1)

where = RC is the time constant of the RC circuit.

M
an
ua
l

We can distinguish two kinds of operations : When the time constant = RC  tc , the charge
collection time. In this case, the current through R is the instantaneous value in the detector and
V (t) = I(t)R

In this case, the detector can collect charge from a single event with time tc .

La

In the second case, we have = RC  tc . Now we can see that very little current flows through
R during the time tc and the current from the detector gets integrated in the capacitor which charges
to Vmax = Q
. If the time between the two events is long, this charge will discharge through R. In this
C
case, note that for a fixed C (which is determined by the electronics and the geometry and construction
of the detector etc.), Vmax is directly proportional to the charge Q deposited by the event. Thus,
measuring the height of the pulse that is Vmax , we can determine the charge deposited in
the detector and hence the energy produced by the incoming radiation in its interaction
with the detector material. Furthermore, a measurement of the pulse rate gives us a rate for the
interactions of the incoming radiation with the detector material. The two cases are shown in Fig 4.4

102

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Figure 4.4:

Ionisation of Gases

4.3

M
an
ua
l

Finally, the Mean Square Voltage Mode (MSV) Mode is used for certain kinds of measurements where we need to measure radiation from sources which produce charges which very different
from each other. Basically, what we do is that in a current mode operation, the varying current which
can be regarded as the superposition of a steady current and a varying component, we block the steady
component. The varying component is then squared and the signal is thus proportional to the square
of the charge that is created by the incident radiation. This enhances the differences between the two
kinds of radiation entering the detector. We shall not discuss this much since it is not used in our lab.

La

The basic idea in any ionisation based detector is that of gas multiplication. In its most rudimentary
form, consider a cylindrical enclosure filled with some gas. The enclosure has an electric field which
increases from the center to the walls of the cylinder. A radiation in the form of alpha, beta or gamma
rays enters the gas and interacts with the atoms or molecules of the gas, ionizing the atom and producing a positive ion and a free electron. The main characteristic of any gaseous ionisation detector is to
amplify this initial ionization event. Let us see how this happens.

4.3.1

Townsend Avalanche

The ionizing radiation upon interacting with the gas produces an ion and an electron. If there was no
field present, then of course the positive ion and the electron would just drift till they lose all their

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

103

energy. However, in the presence of the field, they move towards the electrodes, the ion towards the
cathode and the electron towards the anode.

Nu
cle
ar
Ph
ys
ics

The motion of the electrons and the ions is of course very different because of their different masses.
Both the ions and the electrons at any given temperature are always in some random thermal motion.
In the presence of the field, there is a force on the charges and they also get a drift velocity. (This is
very similar to the motion of electrons in a conductor when a voltage is applied between the two ends
of the conductor). It is seen that in a gas, the drift velocity v in the presence of a field is given by
E
P
E
v =
P

(4.2)

La

M
an
ua
l

where E is the electric field and P is the pressure. The proportionality constant is called the
mobility and it depends on the nature of the gas and it turns out to be fairly independent of the value
of E and P . For ions, typical value of the mobility is around 104 m2 s1 atm V1 , For electrons,
because of their much smaller mass, the mobility is typically higher by a factor of 103 than the ions.
The variation of the drift velocity with E
for some gases is shown in Figure 4.5.
P

Figure 4.5: Variation of drift velocity of electrons in various gases with

E
P .

(Source: http:// encyclopedia2.thefreedictionary.com/ Mobility+of+Ions+and+Electrons)

Now suppose we increase the field. The ions being heavier move somewhat faster than before but
still have low mobility. The electrons on the other hand could gain significant kinetic energy. In their
path towards the anode, the electron would collide with other neutral atoms and if it transfers more
than the ionisation energy to the electron in a neutral atom, it will ionise it. Of course, exactly when
this secondary ionisation caused by the electrons in motion actually happens, depends on the density
and pressure of the gas as well as the electric field strength. It is seen that in most gases at atmospheric
pressure, the threshold value of the field required for this is around 106 V m1 .

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

104

Nu
cle
ar
Ph
ys
ics

So after the first ionisation by the radiation, we now have a second ionisation produced by the ionizing
electron from the first ionisation. The two electrons now again move towards the anode because of the
electric field and gain kinetic energy. When this energy is large enough, they could collide with two other
neutral atoms and ionise them thereby producing 4 electrons now. These four electrons move towards
the anode and after gaining enough energy could ionise four neutral atoms and thereby producing 4 more
electrons and so on. This process continues as a chain reaction or a cascade as depicted schematically
in Figure 4.6 and is known as the Townsend Avalanche.

Figure 4.6: Schematic Diagram of Townsend Avalanche

M
an
ua
l

(Source: Electron avalanche by Dougsim - Own work. Licensed under CC BY-SA 3.0 via Wikimedia Commons http:
//commons.wikimedia.org/wiki/File:Electron avalanche.gif#/media/File:Electron avalanche.gif)

We can quantify this process of cascade ionisation by the Townsend Equation which gives us the
increase in the number of electrons per unit length as

La

dN
= dx
(4.3)
N
where is the first Townsend coefficient which depends on the field strength and the nature as well
as the density and pressure of the gas. It is clearly zero for fields below the threshold value of the field
referred to above. The solution to this equation is obviously an exponential increase in the number of
electrons
N (x) = N (0)ex

(4.4)

The charges are finally deposited on the electrodes and this leads to a pulse in the associated electronics and whose characteristics give us some information on the nature and properties of the original
ionizing radiation.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

4.3.2

105

Kinds of Detectors & Detector Regions

We mentioned that there are several kinds of detectors which use the principle of ionisation. Proportional Counters and GM counters are two examples which though using the same basic principle of
ionisation, work in different regions as we shall see below. Both these work in the pulse mode.

M
an
ua
l

Nu
cle
ar
Ph
ys
ics

Consider a gas filled detector where, as we have seen above, we have a chamber filled with a gas
and an electric field. Suppose in such a detector, we have one ionisation event due to the interaction
of radiation. If we increase the applied voltage to the electrodes, hence increasing the electric field
between them and note the amplitude of the pulse generated (which is a measure as we have seen of
the charge which has been generated by the ionisation and is collected by the electrodes), we can see
the differences in the response of the detector as shown in Figure 4.7.

Figure 4.7: Detector Regions

(Source: https://www.science.mcmaster.ca/medphys/images/files/courses/4R06/ )

La

Starting from very low values of the voltage, we see that the electric field experienced by the charges
is so small that some of the original ions can recombine and the charge that is collected by the detector
is less than that produced by the original radiation. This region is of course of no use for detection of
radiation. As we increase the applied voltage, we reach a saturation. This is when the charge collected
by the electrodes is equal to the charge created by the original radiation. This is the region where
another class of detectors called ion chambers operate.
When the voltage is increased even further, at some point the field becomes strong enough that the
threshold is reached for gas multiplication which as we have seen above, then produces secondary ions
and electrons by collisions. The charge collected by the detector increases and is proportional to the
energy of the ionising radiation. As the voltage increases, the collected charge increases and we see an
increase in the pulse amplitude. Initially, for a range of applied voltage, the gas multiplication is linear
and we see that the charge collected is proportional to the original number of ion pairs (and hence to

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

106

the energy of the incoming radiation). This region is called the region of true proportionality and is
the region in which proportional counters operate. We can achieve a multiplication of about 106 in
this region.

Nu
cle
ar
Ph
ys
ics

If we increase the field further, then this proportionality breaks down. The reason for this is the
accumulation of a sheath of positive charges formed from the ions which are produced in the primary
and secondary ionisation. The positive ions being much heavier have smaller mobility and hence take a
much longer time to move towards the cathode. The space charge that is formed, distorts the electric
field experienced by the electrons in the gas and this leads to some nonlinearity. Thus in this region of
the applied voltage, we will not see a strictly linear increase of pulse amplitude with voltage though there
is still an increase. This region is known as the region of limited proportionality. We can say that
in this region, the pulse amplitude becomes more dependent on the voltage than on the initial ionisation.

M
an
ua
l

As the field is increased even more, we reach what is called the Geiger Region. Now the conditions
are there for an avalanche to occur in the gas and a maximum number of electrons are produced in the
gas due to the cascade effect as described above in Section 4.3.1. This is the region where the ion pair
count is now independent of the original ionisation and the detector cannot distinguish between different
ionising radiations or their energies. However, the detector efficiency is very good in this region. What
we see because of this independence is that the pulse amplitude is independent of the applied voltage
over a range of voltage and we get a plateau in the graph.
Finally, when the voltage increases beyond a certain value ( which depends on the nature and the
pressure of the gas as well as the geometry of the detector), the gas breaks down into a plasma and we
get a continuous discharge. This is bad for the detector and can damage the detector.

GM Counter

La

4.4

We have already seen how in general the detectors which are based on ionisation of gases work. The
GM counter, which is what we use in the laboratory is a kind of ionisation based detector which has
certain special characteristics which we shall examine now.

4.4.1

Geiger Discharge

The Townsend avalanche which we discussed above (Section 4.3.1) consists of secondary ions and electrons forming after the initial ionisation event. However, the electrons produced in the avalanche also
sometimes produce excited gas molecules, that is when the energy transferred to the molecule is less
than that required for ionisation. These excited molecules deexcite with a characteristic time of around

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

107

M
an
ua
l

Nu
cle
ar
Ph
ys
ics

109 seconds and in doing so, emit a photon. This photon could be in the visible or the ultraviolet,
depending on the excitation state of the molecule. Now when this photon interacts (by photoelectric
absorption as we discussed in Section 3.2.3) with a less tightly bound electron somewhere else in the gas,
the electron might be released. This electron will move towards the anode and once again be a source
of a new avalanche. The photon could also hit the walls of the tube and release an electron when it is
absorbed. Typically, in a GM counter, the gas multiplication factor is very large, 109 1010 , and thus
we have an increasing number of avalanches created after a single ionisation event. The avalanches begin
when the electron is close to the anode wire and thus the time needed for producing all the ions and
electrons is less than a microsecond. The ultraviolet photons produce more electrons near the original
avalanche and these electrons also move towards the anode and create further avalanches. Thus, what
we see is that the discharge grows along the central anode wire in both directions from the position of
the original ionising event. The speed of this growth along the anode is very high, 2 4 cm ( s)1 .
This process is shown in Figure 4.8. The electrons collected by the anode constitute a pulse that is an
increased voltage across the external load resistor and this is counted as an event or a count.

Figure 4.8: Geiger Discharge

(Source: Spread of avalanches in G-M tube by Dougsim - Own work. Licensed under CC BY-SA 3.0 via Wikimedia Commons

jpg)

La

- http://commons.wikimedia.org/wiki/File:Spread of avalanches in G-M tube.jpg#/media/File:Spread of avalanches in G-M tube.

So we have a scenario where a single ionizing event has in a very short time, created many avalanches
throughout the tube. How does this process end? To understand this, we need to realise that whenever
a free electron is created in the avalanche or otherwise, a positive ion is also created. The positive ions
have a much lower mobility because of their larger mass. Thus they hardly move while all the electrons
created in the various avalanches are collected by the anode. Clearly, this concentration of positive
charges near the anode causes the electric field to reduce since we know that in a cylindrical tube the
electric field at a distance r from the anode will be

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

E(r) =

V
r ln ab

108

(4.5)

4.4.2

Nu
cle
ar
Ph
ys
ics

where V is the constant voltage between the cathode and the anode and b and a are the inner radius
of the cathode and the radius of the anode wire respectively. (To see this, think of the positive space
charge acting as a sheath around the anode, thereby increasing the effective radius of the anode wire , a.
Now obviously, with an increase in a, the field at a fixed r will increase. However, and this is important,
the r where the field is relevant now for ionisation, increases because of the space charge. This effect is
much larger and so overall the field will decrease. To demonstrate this, let V = 100 V, b = 10 cm and
a = 1 cm. Then the field at r = 1 cm, that is just above the anode wire, will be 43.4 V cm1 . Now let
a = 2 cm, then the field at r = 2 cm, that is just above the effective anode radius now is 31.1 V cm1 .
) With the decrease in the electric field below some threshold value, no further avalanches are possible
and the Geiger discharge ends. For a particular tube (that is a particular geometry and composition of
the gases), with a fixed external voltage the termination of the Geiger discharge is always after a fixed
number of avalanches or, what is the same thing, a fixed amount of total charge. This does not depend
on the original ionizing event at all. To put it another way, the size of the pulse is always the same no
matter what the nature of the incoming radiation which caused the initial ionizing event is.

Quenching

La

M
an
ua
l

What we have seen till now is that incoming radiation generates an ionisation induced Geiger discharge
in the GM counter. We saw that the mechanism for this is a Townsend avalanche which, in a GM tube
generates multiple avalanches till the discharge ends because of the space charge due to the positive
ions. In all of this, it is clear that the role of the gas in the GM tube is very important. Firstly, we must
make sure that the gas used in the counter, called the fill gas is such that there is no possibility of it
forming negative ions. This is obvious since the whole operation depends on the generation of electrons
and positive ions. Thus, we need the fill gas to have a low electron attachment coefficient like hydrogen
or the rare gases and not a high electron attachment coefficient like oxygen. Secondly, the properties
of the fill gas should be such that the energy gained by the free electrons is enough for the multiple
avalanches to be generated. Recall that the drift velocity and hence the average energy depends on the
ratio E
(Eq(4.2)). Thus we try to have a gas mixture which gives the maximum of this ratio so that
P
we can generate the minimum electron energy to generate multiple avalanches. Typically, we use argon
in the GM counters used in our labs.
The discussion above on Geiger discharge began with the single ionising event creating a cascade of
avalanches and then the avalanches travel down the anode wire very quickly and a pulse is created by
the charge collected at the anode which then essentially gets stored in the capacitance of the circuit and
discharges through the load resistor. We also discussed how the discharge ends because of the positive

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

109

Nu
cle
ar
Ph
ys
ics

space charge formed by the low mobility positive ions which reduces the electric field near the anode as
we discussed above (Eq(4.5)). Now let us consider what happens to the positive ions. These ions, say of
argon or some other inert gas with a high ionisation potential, are much heavier and therefore have low
mobility. Their motion towards the cathode is slow but they at some point will arrive at the cathode.
In this motion, they will gain kinetic energy. Now when they hit the cathode they will neutralise by
picking up an electron from the cathode. But suppose their energy was larger than twice the work
function of the cathode, that is the energy required to liberate an electron from the surface of the
cathode. In that case, the positive ion could, with some probability, liberate more than one electron
from the cathode. If the total number of positive ions is large (which it will be as we saw above since in
a GM tube, a large number of ion-electron pairs are produced) there could be at least one extra electron
liberated from the cathode. This electron, like the other electrons in the tube, would accelerate towards
the anode and can cause another Geiger discharge producing more positive ions and the process can go
on endlessly producing a continuous pulse.

M
an
ua
l

There are several ways of dealing with this problem of multiple pulses. In the initial days of the
GM counter, a method known as External Quenching was used. In this, the electronics of the device
was arranged in such a way so as to reduce the voltage for some time after a pulse. The lower voltage
is such that no gas multiplication can take place and therefore no Geiger discharge happens after the
initial pulse. The time that the voltage needs to be reduced obviously depends on the time that the
positive ions take to travel to the cathode, typically 10 1 milliseconds. One way to do this is to
choose a large enough R in the RC circuit external to the tube. Then, as we have seen, (Eq(4.1)), the
time constant of the RC circuit becomes much larger than the charge collection time. Clearly, this also
means that the GM counter will only be able to record events which are separated by more than a few
milliseconds. This method therefore can only be used with low count rates.

La

In the GM counters used in our lab, we use Internal Quenching. In this method, a small percentage, typically 5 10% of another gas, called the quench gas is added to the tube. The quench gas
has the property that it has a lower ionisation potential than the fill gas and is also a more complex
molecular structure. Popular quench gases are ethyl alcohol and halogens, both of which satisfy these
conditions compared to argon, a typical fill gas. Now as the positive argon ions are drifting towards the
cathode, they will collide with the halogen molecules and since these have a lower ionisation potential,
the argon ions will transfer their energy to the quench molecules and neutralise. The quench gas ions
will now start drifting towards the cathode. When they strike the cathode, they will neutralise by
picking up an electron but, and this is the crucial point, the extra energy instead of liberating another
electron, will go towards dissociating the quench gas molecules.
As mentioned above, both halogen gases and complex organic gases like ethyl alcohol are used as
quench gases. However, the organic gases are generally consumed and so the tubes which use them

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

110

have a finite lifetime. Halogens on the other hand can recombine spontaneously and so can be used
since they are self replenishing.

4.4.3

Dead Time & Recovery Time

La

M
an
ua
l

Nu
cle
ar
Ph
ys
ics

After a Geiger discharge takes place, as mentioned above, we still have the positive ions of the fill gas
drifting away from the anode towards the cathode. Initially of course, when they are closer to the
anode, they reduce the electric field experienced by the electrons to below the level required for creating
another avalanche. At this time, if another particle of radiation enters the GM tube, and causes a single
ionising event, that will not lead to a Geiger discharge because the field is lower than that required by
the ionising electron to create an avalanche. This time is called the dead time and if another ionising
event occurs in this period, it will not be recorded. However, slowly the positive ions drift towards
the cathode and so the field near the anode rises. In this time, there could be some avalanches formed
and therefore some pulses may be observed depending on the sensitivity of the counting circuit. These
pulses would not be of the same amplitude as the original Geiger discharge pulse though. Finally, all
the positive ions do reach the cathode and at this point the electric field near the anode becomes strong
enough again for the Geiger discharge of the same amplitude to take place. The behaviour is shown in
Figure 4.9

Figure 4.9: Time behaviour of a GM tube & Dead Time

(Source: Dead time of geiger muller tube by Dougsim - Own work. Licensed under CC BY-SA 3.0 via Wikimedia Commons http://commons.wikimedia.org/wiki/File:Dead time of geiger muller tube.png#/media/File:Dead time of geiger muller tube.png)

In most tubes, the dead time is around 50 100s. As mentioned above, the tube takes a longer
time than the dead time to return to its original configuration, that is when it can produce a second
pulse of the same magnitude as the first one. This time is called recovery time.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

111

Clearly, in cases where the count rates are high, determination of dead time is of critical importance.
There are several methods of determining the dead time. Here we discuss the commonly used two
source method.

Nu
cle
ar
Ph
ys
ics

Let us define the following variables:


, the dead time of the counter
tr , the real or actual time the counter is in operation. This is the total time that we take to measure
the counts. It is clearly something which does not depend on .
tl , the live time of the detector, or the time that the detector is able to record the counts. This obviously
depends on .
N , the total number of counts recorded by us during tr .
m, the measured counting rate, that is N
.
tr
N
n, the true counting rate, that is tl .
So we have

N
tr

n=

N
tl

M
an
ua
l

Thus

m=

tl
m
=
n
tr

(4.6)

La

The interpretation of this is clear- the fraction of the counts we record is equal to the ratio of the live
time to the real time or to put it another way, it is the fraction of the time that the detector can actually
record the counts. The fraction of the time that the detector cannot record the counts is given by m .
It turns out that to a very good approximation, the live time is simply the real time minus N . This is
easy to see since the detector is unable to record any event during the total counting time tr is simply N .
So we have

tl = tr N

(4.7)

tl
N
= 1 = 1 m
tr
tr

(4.8)

Using Eq(4.6) and Eq(4.7), we see that

Or

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

m
tl
=
tr
n
= 1 m
m
n =
1 m

112

(4.9)

Nu
cle
ar
Ph
ys
ics

Eq(4.9) tells us the relationship between the measured count rate m , the dead time and the true
count rate n. Note that n is always greater than m as it should be. Further, for small values of m , the
dead time is not important since the difference between n and m is small as a percentage. Finally, note
that the parameter of importance is not just but m . So, we can keep the product small by either a
small value of or that of m.
To find the dead time of the counter, we use the fact that the count rate from two sources individually
do not add up to the count rate from both of them together. This is because when the count rate is
high, the counter is dead for a longer period and therefore the counts dont add up. If m1 and m2 are
the measured counts for the two sources and m12 is the measured count rate for both of them together,
m12 6= m1 + m2

M
an
ua
l

Of course, the true count rates do add up, that is

n12 = n1 + n2

La

In carrying out any counting experiment with a source, we also need to take into account the measured
background count rate, that is the count rate caused by ambient sources of activity. Let the true count
rate for the background be nb and the measured one be mb . Then we have

n12 nb = n1 nb + n2 nb
n12 + nb = n1 + n2

(4.10)

Using Eq(4.9) for each of these true count rates n12 , n1 and n2 , we get
m12
mb
m1
m2
+
=
+
1 m12
1 mb
1 m1
1 m2

(4.11)

In this equation all the quantities except is a measured quantity and therefore we can use this
equation and the measured values of the count rates for the sources individually, together and the
background to get the value of . Solving for , we get

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

A(1

1 B)

C
A = m1 m2 mb m12

113

(4.12)

C = m1 m2 (m12 + mb ) mb m12 (m1 + m2 )


C(m1 + m2 m12 mb )
B =
A2
is

Nu
cle
ar
Ph
ys
ics

This is obviously a very complicated expression. An approximate solution to the equation (Eq(4.11))

m1 + m2 m12
2m1 m2

(4.13)

The above expression, in the presence of significant background counts is replaced by

m1 + m2 m12 mb
2(m1 mb )(m2 mb )

(4.14)

4.4.4

La

M
an
ua
l

Similarly, many other approximations are sometimes used instead of the more complicated expression Eq(4.12).The experiment is usually carried out by measuring the counts from source 1, placing the
second source 2 close by to count the combined counts and finally removing source 1 to get the counts
for source 2. To take care of absorption and scattering, when measuring counts from individual sources
1 or 2, a dummy source without activity is placed in place of the other source. One needs to be sure
that the positions of the sources does not change when measuring them together. We also know that
since there is not much difference between m12 and m1 + m2 , the count rates need to be measured very
accurately. We already know from Chapter 1 that the error in counting experiments goes as 1N
where N is the count rate. Thus to get good results we need to have the fractional dead time m12 in
Eq(4.9) to be around 15 20%.

Geiger Counting Plateau & Operating Voltage

Now that we have discussed the general working of a GM counter, we need to understand in detail as
to how exactly one needs to use the GM counter in our experiments. The first thing to determine is the
Operating Voltage of the tube. From the discussion above, we know that the GM tube will only work
if the electric field inside it is above a certain value. If the electric field (or the applied voltage) is too
low, there will no counts recorded because the field cannot produce any pulse. As we raise the applied
voltage above the starting voltage , which is defined as the voltage at which the GM counter just
begins to count, we start observing counts. As the voltage is increased further, we observe a very rapid
increase in the counting rate which is almost like a step function. This rising region is called the knee
because of its resemblance to the knee. The counting rate then rises till we reach a threshold voltage

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

114

Nu
cle
ar
Ph
ys
ics

Vt . Above this voltage, all ionising events produce the same output pulses. The threshold voltage depends on the GM tube and the circuit components of the counter. Above the threshold value, the graph
levels off for a broad range of applied voltage. This region, after the knee or threshold value, where
the counting rate is level is called the plateau region. As we increase the voltage, the counting rate
remains essentially constant, that is the shape of the counting rate versus applied voltage is a straight
line almost parallel to the x axis. This continues till the voltage is high enough that a continuous
discharge takes place in the tube and there is a breakdown. The reason for the continuous discharge are
clearly related to the inability of the quenching mechanism to stop runaway avalanches because of the
high energy gained by the positive ions. The operating voltage is a value of the voltage in the middle
region of the plateau, roughly equidistant from the knee and the point where continuous discharge starts.

La

M
an
ua
l

To illustrate this, let us consider a GM counter where the starting voltage is Vs . If the applied voltage
is less than Vs , then the pulse is not recorded as we have seen. As the voltage is increased to beyond
Vs , to reach the plateau region, the pulse can be measured. However, if the voltage is somewhat close
to the knee, then the low amplitude tail of any pulse will cause a slight rise in the slope of the plateau
region. The plateau region could also have a finite slope because of the quenching mechanism failing
sometimes.

Figure 4.10: Counting Plateau of GM tube

The operating voltage of the GM tube depends therefore on the fill gas and the quenching gas used.
For instance, the typical operating voltage for an argon filled tube with alcohol as the quenching gas is
around 1000 1200 V while those with argon as a fill gas and bromine as a quenching agent is around
200 400 V.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

4.4.5

115

Counting Efficiency

Nu
cle
ar
Ph
ys
ics

Now that we have looked at how the GM counter works in theory, let us see what are the design considerations that go into the making of a GM counter. Firstly, we need a gas filled enclosure with a
window. The window and the walls of the enclosure need to be strong enough to withstand the pressure
difference between the inside and outside the tube since most GM tubes operate below the atmospheric
pressure. The window while being strong enough must be thin enough so as not to have significant
effect on the flux of radiation which one is trying to measure. The gas, as we have seen is typically a
rare gas along with a quenching agent.
The operation of the tube depends crucially on the existence of a high enough field near the anode.
This to a large extent determines the geometry of the counter. Thus, for instance, a parallel plate kind
of arrangement of electrodes will be less efficient than a cylindrical geometry. This is evident from
the expressions for the fields in the two geometries. For a parallel plate geometry, the field would be
uniform and inversely proportional to the distance between the electrodes. On the other hand, the field
for a cylindrical geometry, for the same voltage would be inversely proportional to the distance from
the anode (Eq(4.5)) and so near the anode, where we need a large field to ensure the avalanche, we can
use a much smaller external voltage.

M
an
ua
l

With the above mentioned design of the GM tube, when a charged particle like an alpha or a beta
particle enters the tube, we can detect it. The reason for this, as we have seen is that in a GM tube,
a single ionising event can cause a pulse. So if the charged particle enters the tube, the efficiency in
its detection will be close to 100%. However, we know that alpha particles have a low penetrating
power and so if we need to detect alpha particles, we need to ensure that the window is extremely thin.
For beta particles, we use a thicker window though here too some particles could be reflected or back
scattered from the window material also.

La

The situation with gamma rays is very different from that for charged particles. The gamma ray
photons do not cause direct ionisation of the fill gas atoms. Instead, the photons interact with the
walls of the detector to produce a secondary electron. If the electron is able to penetrate the material of the wall (this depends on the thickness and the material used for the walls), and enter the gas
in the tube, it will essentially behave like an ionising electron from an ionising event produced by a
charged particle in the GM counter and cause a Geiger discharge as discussed above. Thus we need
to make the walls of the tube thinner than the range of the electrons which are produced by the interaction of the gamma ray photons with the wall material. Further, to increase the cross section of
this interaction, material with a higher atomic number needs to be used. Even with all this, for high
energy gamma rays, the typical GM counter efficiency is a few percent. However, for low energy gamma
rays, the probability of a direct interaction with the fill gas atoms increases and if we use appropriate gasses of high atomic number (Xenon for instance) at high pressures, the efficiency can be very high.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

116

Nu
cle
ar
Ph
ys
ics

Gamma sources are also are typically isotropic, that is the radiation is emitted isotropically in all
directions. Since the GM counter has a window of a finite area, only a fraction of these emitted gamma
rays will hit the counter window. Geometrically, it is clear that if we take the counter window to be
a circle of radius r and place the counter at a distance d from the gamma source, then the fraction of
gamma rays hitting the counter window will be the ratio of the area of the spherical cap of radius r
and the total area of the sphere of radius R. The area of the spherical cap with height x can be easily
seen to be 2Rx using double integrals or by any other method. Thus we have

fraction of gamma rays incident on counter window =

2Rx
Rd
1
d
x
=
=
=
(4.15)
2
2
4R
2R
2R
2 2 r + d2

La

M
an
ua
l

This expression for the fraction of gamma rays incident on the counter window is valid for all
distances of the source from the counter. We can intuitively see it to be true since when the distance
of the source is 0, that is the source is next to the window, we expect one half of the emitted gamma
rays to enter the counter, which is what we get. On the other hand, when the source is very far away,
then d  r and we get no gamma rays entering the counter.The geometry of the arrangement is shown
in the Figure 4.11.

Figure 4.11: Isotropic source with GM counter

We can define two kinds of efficiency of the counter/detector in principle. Absolute Efficiency and
Intrinsic Efficiency . Absolute efficiency, Abs can be thought of as the number of particles detected

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

117

as a fraction of the total number of particles emitted. That is


Abs =

Number of particles detected


Number of particles emitted

(4.16)

Clearly, the absolute efficiency will depend on the counter geometry and the distance from the source.
Intrinsic efficiency  int factors in these parameters and gives us a measure of efficiency for that particular
detector
Number of particles detected
Number of particles incident on the detector

Nu
cle
ar
Ph
ys
ics

int =

(4.17)

We are now familiar with the mechanism by which a GM counter works as well as its operational
details. We also have learnt about the various parameters associated with the GM counter like operating voltage, dead time, recovery time and its efficiency. With this knowledge, we can now proceed to
use the GM counter for various experiments.

4.5

References

La

M
an
ua
l

1. Radiation Detection & Measurement, Glenn F. Knoll, Wiley India (2009).

Chapter 5

5.1

Introduction

Nu
cle
ar
Ph
ys
ics

Experiment: GM Characteristics

5.2

M
an
ua
l

Now that we have discussed the theoretical basis for the experiments in the laboratory, let us look at
the actual experiments that we perform in this laboratory. The basic equipment that we use in the
laboratory is the GM counter. We use it to find out the count rates from various nuclear sources and
thus try and understand the characteristics of the radiation coming out from them. For this purpose,
we first need to study the operational characteristics of the GM counter itself. This experiment
investigates the working of a GM counter and finding its parameters like the optimum
operating voltage, dead time etc. In addition, it also demonstrates the statistical nature
of radioactive processes.

Precautions

La

Since all the experiments in this laboratory deal with radioactive sources, it is important to follow certain guidelines and precautions for working in this laboratory. Radioactive sources pose a substantial
health hazard and so it is essential that you follow these guidelines very carefully.

5.2.1

Health Effects of Radiation

As we have already seen in earlier chapters, radioactive materials decay spontaneously to produce ionizing radiation like alpha, beta and gamma rays. These radiations have sufficient energy to strip away
electrons from atoms (as we saw in Sections 3.2.1, 3.2.2 & 3.2.3) or to break some chemical bonds. Any
living tissue in the human body can be damaged by ionizing radiation in a unique manner. The body
attempts to repair the damage, but sometimes the damage is of a nature that cannot be repaired or it
is too severe or widespread to be repaired. Also mistakes made in the natural repair process can lead to
cancerous cells. All types of radiation to which the person is exposed and the pathway by which they
118

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

119

Nu
cle
ar
Ph
ys
ics

are exposed influence health effects. Different types of radiation vary in their ability to damage different
kinds of tissue. Radiation and radiation emitters (radionuclides) can expose the whole body (direct
exposure) or expose tissues inside the body when inhaled or ingested. All kinds of ionizing radiation
can cause cancer and other health effects. The main difference in the ability of alpha and beta particles
and gamma and x-rays to cause health effects is the amount of energy they can deposit in a given space.
Their energy determines how far they can penetrate into tissue. It also determines how much energy
they are able to transmit directly or indirectly to tissues and the resulting damage. Although an alpha
particle and a gamma ray may have the same amount of energy, inside the body the alpha particle will
deposit all of its energy in a very small volume of tissue. The gamma radiation will spread energy over
a much larger volume.
There are two kinds of health effects that radioactive sources can cause - Stochastic and Nonstochastic.

La

M
an
ua
l

Stochastic health effects are those that are associated with long term and low level exposure to radiation. Increased levels of exposure make these health effects more likely to occur, but do not influence
the type or severity of the effect. Cancer is considered by most people the primary health effect from
radiation exposure. Simply put, cancer is the uncontrolled growth of cells. Ordinarily, natural processes
control the rate at which cells grow and replace themselves. They also control the bodys processes for
repairing or replacing damaged tissue. Damage occurring at the cellular or molecular level, can disrupt
the control processes, permitting the uncontrolled growth of cells cancer This is why ionizing radiations ability to break chemical bonds in atoms and molecules makes it such a potent carcinogen. Other
stochastic effects also occur. Radiation can cause changes in DNA, the blueprints that ensure cell
repair and replacement produces a perfect copy of the original cell. Changes in DNA are called mutations. Sometimes the body fails to repair these mutations or even creates mutations during repair. The
mutations can be teratogenic or genetic. Teratogenic mutations are caused by exposure of the fetus in
the uterus and affect only the individual who was exposed. Genetic mutations are passed on to offspring.
Non-stochastic health effects appear in cases of exposure to high levels of radiation, and become
more severe as the exposure increases. Short-term, high-level exposure is referred to as acute exposure. Many non-cancerous health effects of radiation are non-stochastic. Unlike cancer, health effects
from acute exposure to radiation usually appear quickly. Acute health effects include burns and radiation sickness. Radiation sickness is also called radiation poisoning. It can cause premature aging
or even death. If the dose is fatal, death usually occurs within two months. The symptoms of radiation sickness include: nausea, weakness, hair loss, skin burns or diminished organ function. Thus, for
instance, some medical patients receiving radiation treatments often experience acute effects, because
they are receiving relatively high bursts of radiation during treatment.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

120

Nu
cle
ar
Ph
ys
ics

The unit of effective or equivalent dose or radiation is the rem or Roentgen Equivalent in Man.
This is the older though much used unit of the radiation dose exposure. A more modern unit is the
Sievert which is defined as 100 rem. A rem is a very large unit and so typically millirem is used for
ordinary exposures. 1 rem or 0.01 sievert translates to typically a 0.055% chance of the development of
cancer eventually. Typically,the exposure from background sources is around 0.01 millisieverts (mSv)
per day while the exposure to a chest x-ray is 0.06 mSv during the exposure to the x-ray. For a CT
scan, the exposure can be as high as 2 mSv. To compare, during the Fukushima nuclear accident, near
the accident site, a level of 1000 mSv hour1 has been reported.

General Precautions

1. Handle the radioactive sources with utmost care and respect. Dont bend or try to break them.
2. Although the sources in the laboratory are always in their holders, it is in general important
to never touch the source using bare hands. Always use forceps to handle sources.
3. Do not eat or drink during the lab. Please do not keep any edible material or even drinking
water on your work table. Keep your bags with the food and water on the table on the side.

M
an
ua
l

4. When bringing or returning the source to the source room, please be extremely careful to not
let it fall.
5. Do not leave the source lying around the work table. Always keep it carefully while using it
and return it promptly after you are finished.
6. Do not keep the source in your pocket or in close contact with your body.

La

7. Wash your hands after the experiment and before eating or drinking anything.

5.3
5.3.1

Experiment
Purpose

The purpose of this experiment is to study the GM counter and determine its characteristics. Specifically, we will be plotting the counts versus the external voltage curve to determine the plateau region
(Section 4.4.4) and hence the operating voltage. We will also be determining the dead time (Section
4.4.3) of the counter. Finally, we will investigate the statistical nature of the radioactive process. For
this purpose we would need a GM counter (which comprises of the GM tube, associated electronics for
counting purposes and an external high voltage source), two radioactive sources (which in our case are

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.


60
27Co)

60
27Co

60
27Co

60

28Ni

60

28Ni

60

28Ni

60

28Ni

60
28Ni

is given in Figure 2.3. The reactions are

+ e + e

Method

Nu
cle
ar
Ph
ys
ics

5.3.2

and a dummy source. The decay scheme for

121

M
an
ua
l

1. GM Characteristics:
To determine the characteristics of the GM counter, we first set the voltage to zero. Place the
source provided near the counter window and set the preset time to 50 seconds. This is the
time during which the counter will measure the counts. We increase the voltage gradually, initially
increasing it by a larger amount (say 50 V) till the counts start. This voltage at which the counts
start is the threshold voltage. Once the counts start, the voltage should be increased in smaller
steps (say 10 V). You will note that after a sharp sudden rise in the counts (the knee of the characteristic), the counts remain constant with an increase in the voltage. This is the plateau region
discussed in Section 4.4.4. Finally, at some point, the count rate starts to increase. This is the
region where the continuous discharge starts and one should not increase the voltage beyond this
to avoid damage to the counter.

La

2. Measurement of Dead Time:


We use the two source method to determine the dead time of the GM counter (Section 4.4.3). For
this we need two sources. First, we set the voltage in the GM counter at the operating voltage.
This, as we have seen is the value at which the GM counter should be operated and is the value
in the middle of the plateau region of the GM characteristic. After setting the voltage, we set the
preset time. For this experiment now, we set the preset time to a large value, say around 5001000
seconds. Recall that for determining the dead time, we need to find the counts from two sources
separately and then together and also need to know the background counts (Section 4.4.3, Eq
(4.14)). First we remove all sources and measure the background count rate only with the dummy
source. Then we place one source and the dummy source (to ensure that all the absorption and
other effects from the source holder are equalised) and measure the counts. We repeat the same
with the second source and a dummy. Finally, we take both the sources together and determine
the combined count rate for the two sources.

3. Counting Statistics:
To see the statistical nature of the counts, we remove all sources. We set the operating voltage
and then for various preset times (say 5, 10, 20 seconds), we note the number of counts to get the

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

122

background reading. We can repeat the measurement for 150 200 times to get a viable statistical
sample.
5.3.3

Sample Data

1. GM Characteristics:

La

M
an
ua
l

Voltage (V)
320
324
326
328
330
332
334
336
338
340
345
350
360
370
390
400
420
430
440
450
460
470
480
490
500
510
520
530
540
550
560
570
580
590
600

Nu
cle
ar
Ph
ys
ics

Given below is the sample data for determining the GM characteristics.

Counts (N)
0
16
111
411
539
795
930
909
950
912
998
933
1005
964
947
983
1009
980
993
1035
1037
971
991
1054
938
1029
987
1043
1085
952
998
1026
1049
1058
1069

Error
0.00
4.00
10.54
20.27
23.22
28.20
30.50
30.15
30.82
30.20
31.59
30.55
31.70
31.05
30.77
31.35
31.76
31.30
31.51
32.17
32.20
31.16
31.48
32.47
30.63
32.08
31.42
32.30
32.94
30.85
31.59
32.03
32.39
32.53
32.70

Figure 5.1: Sample data for GM Characteristics

The first two columns give us the voltage (V) in Volts and the number of counts N . Since this is
a counting experiment with a fixed time interval (the preset time) and the events are random (ra-

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

123

dioactive decay is a random event), the conditions of the distribution being a Poisson distribution
are satisfied as we saw in Section 1.4.2. Therefore, square root rule (Section 1.5) tells us that the

error in each measurement of the counts is N . This is tabulated in column 3. As expected, the
estimated error is lower for lower count rates.
We can plot the date as shown in Fig 5.2.

Nu
cle
ar
Ph
ys
ics

GM Characteristic

1200

1000

Counts

800

600

400

200

350

400

M
an
ua
l

0
300

450

500

550

600

Voltage

Figure 5.2: Sample GM Characteristics

La

We see that the graph looks very similar to what we expect theoretically as in Figure 4.10. The
sudden rise after the threshold voltage, the knee, the plateau are all clearly visible in this sample
data. The plateau region is not a line of zero slope as expected in an ideal GM tube but has some
finite slope. Nevertheless, the constancy of the count rate beyond the knee of the characteristic is
clearly visible in the graph. The error bars represent our estimate of the error in each measurement
of N . We can draw a smooth line through the points and we obtain the characteristic. Alternatively,
we can assume that the plateau region is a straight line and fit the points to a straight line using
Method of Least Squares (Section 1.7, Eq(1.69) & Eq(1.70)). We can use the linear part of the
curve to find the operating voltage of our GM tube. For this, recall that the operating voltage is
typically taken to be in the middle of the plateau region. We can take two values near the extreme
ends of the straight line in the plateau region and compute the middle point. In this case, the
operating voltage turns our to be 450 V. The slope of the plateau region should ideally be 0.
However, we can determine the slope from the data given as

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

m=

124

y
952 933
19
=
=
0.1
x
550 350
200

2. Dead Time:

N1
28401
28589
N1 = 28495
p
N1 = N1 = 168.80

Nu
cle
ar
Ph
ys
ics

Preset time is set at 600 seconds. The observed counts for source 1 + dummy (N1 ), source 2 +
dummy (N2 ), source 1 + source 2 (N12 ) and the background (NB ) are are follows.

N2
N12
23926
51957
23970
51827
N2 = 23948
N12 = 51892
p
p
N2 = N2 = 154.75 N12 = N12 = 227.80

NB
268
236
NB = 252
p
NB = NB = 15.88

Table 5.1: Dead Time Determination

We can calculate the dead time using Eq(4.14)

m1 + m2 m12 mb
2(m1 mb )(m2 mb )

M
an
ua
l

To get the proper dead time,we need to multiply with the preset time since the counts m1 , m2 ,
are the counts per unit time. In our case N1 , N2 , are the total number of counts in the preset time
interval.

La

Putting in the numbers we get,

1.34 104 seconds

The error in can also be calculated. We know that is a derived quantity from the measured quantities
N1 , N 2, N12 and Nb . We can use the error propagation equation (Eq(1.44) to calculate the error in
if the errors in the measured quantities are known. But the measured quantities are simply the counts
in a fixed time interval and the events are random, the distribution will be Poissononian and hence the
error in the measurements will go as the square root of the counts.
3. Counting Statistics:
Preset Time: 5 seconds
Operating Voltage: 460 Volts

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

125

In Table 5.2, xi is the number of counts in the preset interval and fi is the number of times out of a
P
total of
fi = 200 that we get xi counts.

0
1
2
3
4
5
6

fi

xi f i

27
0
55
55
60
120
31
93
16
64
06
30
05
30
P
P
fi = 200
xi fi = 387

(xi x) (xi x)2


-1.94
-0.94
0.06
1.06
2.06
3.06
4.06

3.76
0.880
0.086
1.12
4.24
9.86
16.48

pe =

Pfi
fi

(xi x)2 p

0.135
0.507
0.275
0.242
0.300
0.010
0.155
0.173
0.080
0.339
0.030
0.280
0.025
0.412
P
P
2
pe = 1 E = (xi x)2 p = 1.96

Nu
cle
ar
Ph
ys
ics

xi

Table 5.2: Counting Statistics for Preset time = 5 seconds

We note that the sum of the probabilities

pe = 1 as it should be.

M
an
ua
l

We can easily calculate the sample mean as in Eq(1.2) as


P
x i fi
387
x = P
=
= 1.94
fi
200

The corresponding sample variance can be calculated from the expectation value of the square of the
deviations from the sample mean

La

or

(xi x)2 p = 1.96

E2 =

E = 1.40

Our theoretical expectation is that the distribution of the counts would follow a Poisson distribution.
However, we do not know the mean of the distribution. We can estimate the mean of the parent
distribution by taking it to be the sample mean given above. This means that
= x
Then we expect that the probability distribution function to be (Eq(1.21))

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

126

PPoisson P (xi , ) =

xi
e
xi !

We can tabulate these values also


PPoisson
0.144
0.279
0.271
0.175
0.084
0.030
0.011

pe
0.135
0.275
0.030
0.155
0.080
0.030
0.025

Nu
cle
ar
Ph
ys
ics

xi
0
1
2
3
4
5
6

Table 5.3: Experimental and Poisson probabilities: Preset Time 5 seconds

The standard deviation of the Poisson distribution is simply


Poisson =

= 1.39

La

M
an
ua
l

The graph of the experimental data and the theoretical Poisson distribution with the same mean is
shown in Figure 5.3.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

127

GM Counting Statistics: Preset time 5 seconds


0.35
Poisson
Data
0.3

0.25

0.15

0.1

0.05

0
0

Nu
cle
ar
Ph
ys
ics

0.2

Figure 5.3: Counting Statistics : Preset 5 seconds

M
an
ua
l

We can see that the theoretically expected Poisson distribution and the experimental distribution are
fairly similar though there are deviations. However, the error bars on the experimental values of the
derived probabilities are such that the theoretical distribution falls within the range. The error bars
are the error on pe . However, we know that pe is actually a derived quantity, since
fi
pe = P
fi

La

P
Now
fi = 200 and obviously there is no error in this. We take the error in fi as fi and get the

fi
error bars by taking the error in pe to be 200
.
Preset Time: 10 seconds
Operating Voltage: 460 Volts

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

0
1
2
3
4
5
6
7
8
9
10

fi

x i fi

06
0
18
18
33
66
38
114
36
134
27
135
18
108
13
91
05
40
04
36
02
20
P
P
fi = 200
xi fi = 762

(xi x)

(xi x)2

-3.81
-2.81
-1.81
-0.81
0.19
1.19
2.19
3.19
4.19
5.19
6.19

14.52
7.890
3.280
0.660
0.036
1.420
4.790
10.180
17.560
26.940
38.320

pe =

Pfi
fi

(xi x)2 p

0.030
0.435
0.090
0.710
0.165
0.541
0.190
0.125
0.180
0.006
0.135
0.191
0.090
0.431
0.0650
0.661
0.0250
0.439
0.020
0.538
0.010
0.383
P
P
pe = 1 E2 = (xi x)2 p = 4.46

Nu
cle
ar
Ph
ys
ics

xi

128

Table 5.4: Counting Statistics for Preset time = 10 seconds

We can easily calculate the sample mean as in Eq(1.2) as


P
x i fi
762
x = P
= 3.81
=
fi
200

M
an
ua
l

The corresponding sample variance can be calculated from the expectation value of the square of the
deviations from the sample mean
E2 =

La

or

(xi x)2 p = 4.46

E = 2.11

Our theoretical expectation is that the distribution of the counts would follow a Poisson distribution.
However, we do not know the mean of the distribution. We can estimate the mean of the parent
distribution by taking it to be the sample mean given above. This mean that
= x
Then we expect that the probability distribution function to be (Eq(1.21))
PPoisson P (xi , ) =
We can tabulate these values also

xi
e
xi !

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

129

pe
0.030
0.090
0.165
0.190
0.180
0.135
0.090
0.065
0.025
0.020
0.010

Nu
cle
ar
Ph
ys
ics

xi PPoisson
0
0.022
1
0.083
2
0.159
3
0.202
4
0.193
5
0.147
6
0.093
7
0.050
8
0.024
9
0.010
10 0.003

Table 5.5: Experimental and Poisson probabilities: Preset Time 10 seconds

The standard deviation of the Poisson distribution is simply


Poisson =

= 1.95

M
an
ua
l

The graph of the experimental data and the theoretical Poisson distribution with the same mean is
shown in Figure 5.4.

GM Counting Statistics: Preset time 10 seconds

0.25

Poisson
Data

0.15

La

0.2

0.1

0.05

0
0

Figure 5.4: Counting Statistics : Preset 10 seconds

10

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

130

We can see that the theoretically expected Poisson distribution and the experimental distribution are
fairly similar though there are deviations. However, the error bars on the experimental values of the
derived probabilities are such that the theoretical distribution falls within the range.
Preset Time: 20 seconds
Operating Voltage: 460 Volts

5
10
5
15
15
60
12
60
23
138
25
175
21
168
30
270
26
260
15
165
09
108
05
65
02
28
04
60
02
32
01
17
P
P
fi = 200
xi fi = 1631

(xi x)

(xi x)2

pe =

Pfi
fi

(xi x)2 p

Nu
cle
ar
Ph
ys
ics

xi f i

-6.155
-5.155
-4.155
-3.155
-2.155
-1.155
-0.155
0.845
1.845
2.845
3.845
4.845
5.845
6.845
7.845
8.845

37.88
26.57
17.26
9.95
4.64
1.33
0.024
0.714
3.4
8.09
14.78
23.47
34.16
46.55
61.54
78.23

0.025
0.947
0.025
0.664
0.075
1.294
0.060
0.597
0.115
0.533
0.125
0.166
0.105
0.002
0.150
0.107
0.130
0.442
0.075
606
0.045
0.665
0.025
0.586
0.010
0.341
0.020
0.937
0.010
0.615
0.005
0.391
P
P
pe = 1 E2 = (xi x)2 p = 8.893

2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17

fi

M
an
ua
l

xi

La

Table 5.6: Counting Statistics for Preset time = 20 seconds

We can easily calculate the sample mean as in Eq(1.2) as


P
xi f i
1631
x = P
=
= 8.155
fi
200
The corresponding sample variance can be calculated from the expectation value of the square of the
deviations from the sample mean
E2 =

X
(xi x)2 p = 8.893

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

131

or
E = 2.98
Our theoretical expectation is that the distribution of the counts would follow a Poisson distribution.
However, we do not know the mean of the distribution. We can estimate the mean of the parent
distribution by taking it to be the sample mean given above. This mean that

Nu
cle
ar
Ph
ys
ics

= x
Then we expect that the probability distribution function to be (Eq(1.21))
PPoisson P (xi , ) =

xi PPoisson
2
0.009
3
0.025
4
0.052
5
0.080
6
0.117
7
0.136
8
0.139
9
0.130
10 0.100
11 0.076
12 0.051
13 0.032
14 0.010
15 0.010
16 0.010
17 0.003

La

M
an
ua
l

We can tabulate these values also

xi
e
xi !

pe
0.025
0.025
0.075
0.060
0.115
0.125
0.105
0.150
0.130
0.075
0.045
0.025
0.010
0.020
0.010
0.005

Table 5.7: Experimental and Poisson probabilities: Preset Time 20 seconds

The standard deviation of the Poisson distribution is simply


Poisson =

= 2.85

The graph of the experimental data and the theoretical Poisson distribution with the same mean is
shown in Figure 5.5.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

132

GM Counting Statistics: Preset Time 20 seconds


0.18
Poisson
data
Gaussian

0.16
0.14
0.12

0.08
0.06
0.04
0.02
0
-0.02
2

Nu
cle
ar
Ph
ys
ics

0.1

10
x

12

14

16

18

Figure 5.5: Counting Statistics : Preset 20 seconds

La

M
an
ua
l

We can see that the theoretically expected Poisson distribution and the experimental distribution are
fairly similar though there are deviations. From our theoretical considerations (Section 1.4.3), we know
that when the mean of the Poisson distribution is large, the distribution tends towards a Gaussian or
normal distribution. In Figure 5.5, we have also plotted the Gaussian distribution with the sample
mean as the mean and the sample standard deviation as the standard deviation. We can see the
similarities or differences between the experimental distribution and the normal distribution. Indeed
one sees that the normal distribution approximates the experimental distribution much better than the
Poisson distribution.

Chapter 6

Nu
cle
ar
Ph
ys
ics

Experiment: GM Counter: Counting


Efficiency for & rays
6.1

Introduction

6.2
6.2.1

La

M
an
ua
l

We have already studied the GM counter and recorded its characteristics. These include finding the
operating voltage for the GM counter by plotting its characteristic and taking the middle point of the
flat plateau region. We have also determined its dead time. In this experiment we will study the
efficiency of the GM counter for detecting gamma and beta radiation. As already discussed
in Section 4.4.5, alpha particles are difficult to detect with the GM counters available to us because
of the thickness of the window (which needs to be thick for structural reasons to be able to withstand
the difference in pressure inside and outside the tube). Beta particles on the other hand have a larger
penetrating power and so can penetrate the window and being charged can easily cause an ionising
event in the tube which leads to a Gieger discharge. Gamma rays on the other hand, rarely interact
directly with the fill gas in the tube but instead cause the emission of a photoelectron from the inner
walls of the tube which leads to an avalanche as explained in Section 4.4.5.

Experiment
Purpose

To determine the efficiency of the detector, we clearly need to measure the number of particles emitted
by the source and the number detected. The efficiency can be defined in two ways, Absolute Efficiency,
Abs which does not take into account geometric factors like the distance from the source etc. (Eq(4.16))
and Intrinsic efficiency , int which does take into the account the number of particles which actually
enter the counter, Eq(4.17). For gamma ray sources, since the emission is roughly isotropic, we need to
take into account the number of particles actually striking the counter window while for beta rays, we

133

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

134

assume that all the particles emitted by a source are actually striking the window.
For this experiment, we would need a GM counter setup, a range of beta and gamma sources, which
in our case are 60Co, 57Co, 133Ba, 204Tl and 147Pm. We would also need some aluminum sheets.

Nu
cle
ar
Ph
ys
ics

The decay schemes for these sources are given below.

60

Co

57

Co

La

M
an
ua
l

Figure 6.1: Decay scheme for

Figure 6.2: Decay scheme for

135

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Figure 6.3: Decay scheme for

133

Figure 6.4: Decay scheme for

204

Ba

La

M
an
ua
l

Tl

Figure 6.5: Decay scheme for

6.2.2

147

Pm

Method

As mentioned above, we need to basically know three numbers- one, the total number of particles
emitted by the source; the number of particles detected by the source and finally, the number of
particles entering the detector (which, as discussed in Section 4.4.5 may be different from the number
emitted.). The total number of particles emitted by the source depends on the activity of the source.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

136

The activity of a source is defined as the rate of decay of the radionuclides in a source and is of course
a time dependent quantity as we know from Section 2.1.2. The activity follows an exponential law
(Eq(2.2))
R(t) = R(0)et
where is the decay constant. The decay constant is related to the half life T 1 by
2

ln 2
T1

Nu
cle
ar
Ph
ys
ics

For the sources mentioned above, we are given the activity of the source at the time of its fabrication,
R(0). The time of fabrication ti is also given. From this, we can determine the time elapsed since the
source with the given activity R(0) was fabricated. This time is t. Knowing the source, we know the
half life T 1 of the particular radioisotope. This allows us to calculate , the decay constant. Finally,
2
knowing R(0), t and , allows us to calculate the present activity R(t).

La

M
an
ua
l

For the purely gamma sources, 133Ba, 57Co, we need to take into account the geometric factor also
as discussed in Section 4.4.5 since the sources are isotropic and the number of particles entering the
counter window will depend on the distance from the source and the radius of the counter window.
In addition to the pure gamma sources, we can also use 60Co which is a gamma and beta source. We
place an aluminum sheet between the source and the detector window which effectively blocks all the
beta particles, its thickness being more than the range of the 0.514 MeV beta particle but is not thick
enough to significantly attenuate the gamma rays.
For the pure beta sources, 204Tl, 147Pm, the geometry is not a significant issue since we assume all the
beta particles emitted enter the window. This is because the source is placed in a holder which has only
a small aluminum window through which the beta particles can escape. The source is otherwise enclosed
by thick plastic casing which absorbs the beta particles. Nevertheless, we will use the geometric
factor for determining the number of particles striking the window of the counter, instead
of taking it to be equal to the number emitted. Of course as we increase the distance between
the source and the counter, there is some absorption of the beta particles in the air.
We start by determining the operating voltage of the GM tube as discussed in Chapter 5 and choosing
an appropriate preset time, say 180 seconds. The radius of the GM counter window r is known to be
7.5 mm.We also determine the background count rate NB in the absence of all sources. With the GM
counter set up at the appropriate voltage, we then use the pure gamma sources at various distances d,
and measure the count rate N . We repeat this for various distances for the gamma sources and beta
sources. The detected count rate per second has to be decreased by the background count rate per
second to get the net counts detected per second (cps). The counts emitted per second or the
number of decays per second (dps) which are hitting the counter window are obtained from the
calculation of the activity today R(t).

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.


dps = R(t)

137

1
d

2
2 2 r + d2

where the geometric factor is important for the gamma sources but not so important for the beta
sources. The efficiency is then simply
=
Sample Data

Operating Voltage: 425 V


Preset Time: 180 seconds

Nu
cle
ar
Ph
ys
ics

6.2.3

cps
100%
dps

nB = 85 counts

85
= 0.47 counts per second
180
Error estimation in the tables is discussed in the next section.
NB ( average) =

Ba

133

La

1.

M
an
ua
l

sources:

R(0)
ti
t
T1

111 KBq
August 2013
1.833 years
10.5 years

R(t)
r

0.06601
98.346 103 Bq
0.75 cm

Table 6.1: Data for

133

Ba

(6.1)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

R(t) Bq

d (cm)

98.346 103
98.346 103
98.346 103

2.75
4.75
7.0

dps = R(t)

1
2

d
2 r2 +d2

138

N (counts)

1732.66
601.73
279.83

3045
1201
629

Table 6.2: Count rate and efficiency data for

57

Co

R(0)
ti
t
T1

76.2 KBq
August 2013
1.833 years
0.74 years

R(t)
r

0.936
13.64 103 Bq
0.75 cm

d (cm)

13.64 103
13.64 103
13.64 103

2.75
4.75
7.0

dps = R(t)

La
60

1
2

240.1
83
38

R(t) Bq

M
an
ua
l

Table 6.3: Data for

3.

N
180

NB

=

cps
dps

100% 

0.95 0.084
1.03 0.097
1.08 0.112

16.44
6.2
3.02
133

Ba

Nu
cle
ar
Ph
ys
ics

2.

cps =

d
2 r2 +d2

57

Co

N (counts)

cps =

239
117
101

Table 6.4: Count rate and efficiency data for

Co with Aluminum sheet

N
180

NB

0.86
0.18
0.091
57

Co

=

cps
dps

100% 

0.36 0.052
0.21 0.09
0.24 0.21

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

R(0)
ti
t
T1

17 KBq
August 2013
1.833 years
5.27 years

R(t)
r

0.131
13.37 103 Bq
0.75 cm

R(t) Bq

d (cm)

13.37 103
13.37 103
13.37 103

2.75
4.75
7.0

dps = R(t)

1
2

235.3
81.5
37.4

60

d
2 r2 +d2

N (counts)
1010
426
277

Table 6.6: Count rate and efficiency data for

Tl

204

La

1.

60

cps =

N
180

NB

5.14
1.89
1.07

R(0) 0.11 Ci = 0.11 3.7 104 = 4070 Bq


ti
August 2011
t
3.89 years
T1
3.9 years
2

R(t)
r

0.178
2036.48 Bq
0.75 cm
Table 6.7: Data for

204

=

Co with aluminum sheets

M
an
ua
l

sources:

Co

Nu
cle
ar
Ph
ys
ics

Table 6.5: Data for

139

Tl

cps
dps

100% 

2.18 0.20
2.31 0.25
2.86 0.38

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

R(t) Bq

d (cm)

2036.48
2036.48
2036.48

2.75
4.75
7.0

dps = R(t)

1
2

d
2 r2 +d2

140

N (counts)

35.84
12.42
5.70

cps =

1972
579
406

Pm

=

204

cps
dps

100% 

29.24 2.6
22.06 2.3
31.22 3.6

Tl

Nu
cle
ar
Ph
ys
ics

147

NB

10.48
2.74
1.78

Table 6.8: Count rate and efficiency data for

R(0)
13 KBq
ti
Decemeber 2014
t
0.493 years
T1
2.6 years
2

R(t)
r

0.266
11.40 103 Bq
0.75 cm

d (cm)

11.40 103
11.40 103
11.40 103

2.75
4.75
7.0

dps = R(t)

1
2

200.6
69.5
31.9

R(t) Bq

M
an
ua
l

Table 6.9: Data for

La

2.

N
180

d
2 r2 +d2

147

Pm

N (counts)

cps =

3238
907
341

Table 6.10: Count rate and efficiency data for

N
180

NB

17.52
4.566
1.42
147

=

cps
dps

100% 

8.74 0.77
6.57 0.63
4.45 0.54

Pm

From the data and the results obtained above, we see that the efficiency of the counter for rays is
indeed very small as expected while that for the particles is much higher. Theoretically, we expect an
almost 100% efficiency for beta particles. There could be several reasons why this was not seen in this
experiment. One, the efficiency depends on the dead time and that might have reduced the efficiency for
some source. This would be particularly noticeable for high activity sources where the time between the
particles entering the detector is small. Secondly, as the distance increases, the probability of the beta
particles being absorbed in the air or being scattered off increases. This also might be a contributing
factor which our experiment does not take care of. It only takes care of the distance via the geometric

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

141

factor.

6.2.4

Error Estimation

Nu
cle
ar
Ph
ys
ics

The experiment aims to determine the efficiency  of the GM counter in measuring the counts for beta
and gamma rays. We calculate the efficiency by finding the counts per second and the detected counts
per second as in Eq(6.1). Clearly, both these quantities are derived quantities and so to estimate the
errors in these, we need to estimate the errors in the measured quantities from which they are derived
and then use the error propagation equation Eq(1.45).
We have

=

or

e=

Thus

cps
dps

c
dp

M
an
ua
l

2
c2 dp
e2
=
+
e2
c2
dp2

Let us consider the numerator first.

N nB
180
Each of these quantities is a measured quantity, that is the total number of counts in 180 seconds
with the source and without the source (with the background only). We have seen already in Chap 1
that the errors will be simply the square root of the counts. Thus

La

c=

p
2
2
N
+ (nB
)
c =
180
We assume that there is no uncertainty in the time measurement. Then

= nB

N =
nB
Therefore we get
c2 =

N + nB
1802

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

142

and
c2
=
c2

N +nB
1802
(N nB )2
1802

N + nB
(N nB )2

What about the denominator? Here again, the quantity is a derived quantity and we need to estimate
the error in each of the measured quantities. We have


Nu
cle
ar
Ph
ys
ics

1
d
dps = dp = R(t)

2
2
d + r2

Now R(t) which we have calculated from the original activity and the time elapsed since the fabrication of the source does not have any error since we assume we know these quantities precisely. The error
really is in the geometric factor since that involves r and d. r, the radius of the GM counter window
also is specified by the manufacturer and so we assume no error in it. The error in dp is thus derived
from the above expression from the error in the measurement of the length d, the distance between the
source and the window. Once again, using the error propagation equation, we can see that
2
dp
d2
d2 d2
=
+
dp2
d2 (r2 + d2 )2

Putting it all together, we get

M
an
ua
l

2
c2 dp
e2
=
+
e2
c2
dp2
N + nB
d2
d2 d2
=
+
+
(N nB )2 d2 (r2 + d2 )2

(6.2)
(6.3)

La

The only estimation we need therefore to determine the error in the efficiency  is the error in d.
We assume that the error is equal to the least count of the scale used to measure the distance, namely
0.1 cm. As an example, consider the measurement for the beta source 204Tl. We have
N = 1972
nB = 85
r = 0.75 cm
d = 2.75 cm
d = 0.1 cm
 = 0.2924

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

143

Then a calculation of e2 gives us


e2

= 0.2924

1972 + 85
+ 0.12
(1972 85)2

1
2.752
+
2.752 (2.75 + 0.75)2


= 0.0006

or
e = 0.024

Nu
cle
ar
Ph
ys
ics

Thus we have for this case


 = 29.24 2.4%

La

M
an
ua
l

A similar analysis can be carried out for all the other readings to get an estimate of the error involved
in our experiment.

Chapter 7

7.1

Introduction

Nu
cle
ar
Ph
ys
ics

Experiment: Absorption of rays in Iron

We have already seen how gamma rays interact with matter in Chapter 3, Section 3.2.3. We saw there
that at energies of MeV, the three processes of Photoelectric absorption, Compton Effect and Pair
Production are operative. The relative importance of these depends on the energy of the gamma rays
(Figure 3.13). We also saw that when gamma rays pass through an absorber, there is an exponential
attenuation of the incoming beam (Section 3.1.1). The number of transmitted particles after crossing
a distance x is (Eq(3.1))

M
an
ua
l

N = N0 enx = N0 ex

(7.1)

N = N0 e

(7.2)

La

where is called the linear attenuation coefficient. Clearly, this attenuation coefficient, being
a product of the cross section and n, the number of scatterers per unit volume, will depend on the
density of the scattering material. A more convenient measure to quantify absorption is the mass
attenuation coefficient which is simply . Consequently, we replace Eq(7.1)by

where x is the mass thickness of Eq(3.3). This, as we have seen is much easier to measure and
also much more informative in comparing different absorbers.
We can define a quantity called the half thickness, or the mass thickness where the intensity of the
incoming gamma rays is reduced to one half the initial value. Thus
N0
= N0 em d1/2
(7.3)
2
where m is the mass attenuation coefficient and d1/2 is the half thickness of that particular
absorber for the particular energy gamma rays used. Clearly
N=

144

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

d1/2 =

7.2
7.2.1

0.693
m

145

(7.4)

Experiment
Purpose

7.2.2

Nu
cle
ar
Ph
ys
ics

The aim of the experiment is to find the half thickness of monoenergetic gamma ray source 137Cs, which
emits gamma rays with energy 0.6617 MeV, using iron plates of varying thickness.
The experiment uses a GM counter, a gamma ray source, 137Cs and iron plates of varying thickness.

Method

As with all experiments using a GM counter, the first step is always to find the operating voltage of
the counter by plotting its characteristic as in Chapter 5. We also obtain the background counts for a
fixed preset time, say 180 seconds.

Sample Data

La

7.2.3

M
an
ua
l

We then put the source in the lead stand and note down the number of counts during the preset time.
The lead absorber stand is then placed between the source and the counter window and an increasing
number of iron plates of known thickness are placed in the absorber stand and the counts noted. A
plot of the logarithm of the net number of counts versus the mass thickness of the absorbers gives us a
straight line whose slope allows us to calculate m and d1/2 .

Operating Voltage = 425 V


Preset Time = 180 seconds
Background counts NB = 97 in 180 seconds
Density of Iron = 7.86 gm cm3
Atomic Number of Iron = 56

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Counts (Ni ) Net Counts (Ni NB )


337
240
273
176
280
183
250
153
250
153
216
119
192
95
172
75
173
76
167
70
179
82
141
44
132
35
117
20
131
34

Error ( Ni + NB )
20.83
19.23
19.41
18.62
18.62
17.69
17.00
16.40
16.43
16.24
16.61
15.42
15.13
14.62
15.09

Nu
cle
ar
Ph
ys
ics

Mass Thickness (ti = xi ) gm cm2


0
2.28
3.76
4.60
6.08
8.36
9.86
11.41
13.92
15.26
16.20
19.12
23.72
25.29
27.61

146

Table 7.1: Count rate and thickness data for

137

Cs with iron plates

M
an
ua
l

We need to plot the graph between ln(N NB ) and t = x. This, as we expect will be straight line.
However, given the statistical nature of the data, we would need to find the best fit straight line using
the Method of Least Squares (Section 1.7, Chapter 1).

La

To determine the slope and the intercept of the best fit straight line, we need to use Eq(1.71) and
P
P 2
P
P P
P
P
Eq(1.72) and therefore need
xi =
ti , yi =
ln(Ni NB ), x2i =
ti and
xi y i =
P
ti ln(Ni NB ). We can calculate them as below.

xi = ti
0
2.28
3.76
4.60
6.08
8.36
9.86
11.41

x2i = t2i
0
5.10
14.13
21.16
36.97
69.89
97.22
130.19

yi = ln(Ni NB )
5.48
5.17
5.21
5.03
5.03
4.78
4.55
4.32

xi yi = ti ln(Ni NB ) ln( N + NB )
0
3.03
11.79
2.58
19.59
2.60
23.14
2.51
30.58
2.51
39.96
2.39
44.86
2.28
49.29
2.16

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

4.33
4.25
4.41
3.78
3.46
2.99
3.53
P
yi = 66.34

60.83
64.09
71.44
72.27
82.07
75.62
97.46
P
xi yi = 742.99

2.17
2.12
2.20
1.89
1.78
1.50
1.76

Nu
cle
ar
Ph
ys
ics

13.92
193.77
15.26
232.83
16.20
262.44
19.12
365.57
23.72
562.64
25.29
639.58
27.61
762.31
P
P 2
xi = 187.47
xi = 3393.80

147

Table 7.2: Least Square Fitting for graph of ln(N NB ) vs t

The number of observations, N = 15.

With these values, we can compute the best fit straight line using Method of Least Squares. The results
are
m = 0.0819, c = 240

La

M
an
ua
l

The graph of N NB vs t is plotted on a semi-log scale with the error bars as indicated. We have also
drawn the best fit line with the above mentioned slope and intercept.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

148

Gamma Ray absorption

1000

data
best fit

10

1
0

Nu
cle
ar
Ph
ys
ics

N-NB

100

10

15

20

25

30

t (gm cm-2)

Figure 7.1: ray absorption

M
an
ua
l

With the given slope, we can calculate the mass attenuation coefficient m , which is precisely
the slope calculated as 0.0819 gm cm2 . The half thickness can be calculated from Eq(7.4) as
= 8.55 gm cm2 .
d1/2 = 0.693
m

La

What about the error in our calculation? The error in N NB is clear. We know that a weighted sum

or difference has an error given by Eq(1.49). The error in N is simply N and that in NB is NB .

These two quantities are uncorrelated and hence the overall error in N NB will be N + NB . The

error bars are then simply ln( N + NB ). But what about the errors in the slope and the intercept of
the graph which is what we are using to determine the desired derived quantity, d1/2 ? As you would
recall from Section 1.7, the uncertainty in the slope determined by this method can be calculated using
Eq(1.75). Since we have the values of m and c, for each xi = ti , we calculate y = mxi + c . We also
P
P
have the observed yi as well as the values of = N x2i ( xi )2 . Putting it all together, we then
know the uncertainty in y which is given by
rP
(yi mxi c)2
(7.5)
y =
N 2
With this uncertainty, we can calculate the uncertainty in the slope determined by the Least Square
method as

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

r
m = y

N
= y

149

s
N

x2i

N
P
( xi )2

The values for the uncertainty in the slope turns out to be


m = 0.00018

Nu
cle
ar
Ph
ys
ics

Thus, our value for the mass attenuation coefficient is


m = 0.0819 0.0002

The corresponding error in the half thickness can also be calculated using Section 1.5.1 is
d = d1/2
Thus we can quote the half thickness as

m
= 0.017
m

La

M
an
ua
l

d1/2 = 8.55 0.02 gm cm2

(7.6)

Chapter 8

Nu
cle
ar
Ph
ys
ics

Experiment: Verification of the Inverse


Square Law for rays
8.1

Introduction

La

M
an
ua
l

The intensity of radiation from a point source emitting isotropically is known to fall of as the square of
the distance from the source. We know that this is a purely geometric effect since the area of the surface
increases as d2 and the same energy therefore spreads over four times the area leading to a decrease in
intensity by a factor of four. This can be clearly seen in Figure 8.1.

Figure 8.1: Inverse Square Law for radiation

(Source: http:// hyperphysics.phy-astr.gsu.edu/ hbase/ forces/ isq.html )

We expect that gamma rays being electromagnetic radiation will exhibit similar behaviour.

150

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

8.2
8.2.1

151

Experiment
Purpose

The objective of this experiment is to determine the relationship between the intensity
of gamma radiation and the distance from the source to the detector. We expect that the
inverse square law would be valid.

Method

Nu
cle
ar
Ph
ys
ics

8.2.2

If we have a point source of gamma radiation, which we assume emits isotropically at a rate of N0
particles per second, and we observe the intensity, I at a distance r from the source, we expect a
relationship like
N0
(8.1)
4r2
If we use a long lived source such that N0 is constant through the duration of the experiment, we then
will see that as we increase r, the intensity will go down as the r12 . That is a graph of I vs r12 will be
a straight line with a slope N40 . Alternatively, a graph of log I vs log r will give us a straight line with
slope 2.
I=

137

Cs,

60

Co,

22

Na and

133

Ba. The decay schemes for these are given below.

La

The sources we use are

M
an
ua
l

We use a variety of long lived gamma sources and determine the relationship of the intensity and
distance.

Figure 8.2: Decay scheme for

137

Cs

152

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Figure 8.3: Decay scheme for

Co

133

Ba

La

M
an
ua
l

Figure 8.4: Decay scheme for

60

Figure 8.5: Decay scheme for

22

Na

We also need a GM counter to measure the intensity at various distances as well as a scale to measure
the distance from the source to the counter window.
As with all experiments which use a GM counter, we first need to determine the GM characteristics
and find the operating voltage. We then choose a preset time and determine the background radiation
rate so that it can be subtracted from the count rate with the sources to get the net or true count

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

153

rate. Finally, with each of the sources placed at different distances, the count rate is measured and
tabulated to see if the inverse square law is valid or not.

8.2.3

Sample Data

137

Cs source with Activity 3.1 Ci

Nu
cle
ar
Ph
ys
ics

Operating Voltage = 425 V


Preset Time = 60 seconds
Average Background count NB = 18.8 in 60 seconds

Counts (N)
8728
4858
2976
1969
1418
1061
867
711
572
485
407

M
an
ua
l

d (cm)
2.0
2.5
3.0
3.5
4.0
4.5
5.0
5.5
6.0
6.5
7.0

N NB
8709.2
4839.2
2957.2
1950.2
1399.2
1042.2
839.2
692.2
553.2
466.2
388.2

+NB
Error= N60
1.55
1.16
0.91
0.74
0.63
0.55
0.50
0.45
0.41
0.37
0.34

La

Table 8.1: Measured Count rates at various distances from

137

Cs with error estimates

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Counts (N)
8728
4858
2976
1969
1418
1061
867
711
572
485
407

N NB
8709.2
4839.2
2957.2
1950.2
1399.2
1042.2
839.2
692.2
553.2
466.2
388.2

B
yi = R = N N
60
145.15
80.65
49.28
32.50
23.32
17.37
13.98
11.53
9.22
7.77
6.47
P
yi = 397.29

xi = d12 m2
x2i
2500
625 104
1600
256 104
1111
123.4321 104
816
66.5856 104
625
39.0626 104
493
24.3049 104
400
16 104
330
10.89 104
278
7.7284 104
236
5.5696 104
204
4.1616 104
P
P 2
xi = 8593
xi = 1178.74 104

Nu
cle
ar
Ph
ys
ics

d (cm)
2.0
2.5
3.0
3.5
4.0
4.5
5.0
5.5
6.0
6.5
7.0

154

Table 8.2: Count rates at various distances from

137

Cs & Least Square fitting

M
an
ua
l

We can plot the curves for R vs d as shown in Figure 8.6.

Rate vs distance

160

data
best fit

140

120

La

Rate

100

80

60

40

20

0
2

d (cm)

Figure 8.6: Graph of R vs d

xi y i
362875
129040
54751
26520
14575
8563
5592
3937
2563
1834
1320
P
xi yi = 611571

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

We can also plot a graph of R vs


we get

1
d2

155

using the Method of Least Squares. Using Eq(1.71) and Eq(1.72),

m = 0.0593, c = 10.250

Rate vs 1/d^2

Nu
cle
ar
Ph
ys
ics

160

data
best fit

140

Rate counts/sec

120

100

80

60

40

20

0
500

1000

M
an
ua
l

1500

2000

2500

1/d^2 (m)^-2

Figure 8.7: Graph of R vs

1
d2

La

Finally, we need to compute the error in our estimation of the slope of this graph. In Section 1.7, we
saw that the uncertainty in the slope determined by this method can be calculated using Eq(1.75).
Since we have the values of m and c, for each xi , we calculate y = mxi + c . We also have the observed
P
P
yi as well as the values of = N x2i ( xi )2 . Putting it all together, we then know the uncertainty
in y which is given by
rP
(yi mxi c)2
y =
(8.2)
N 2
With this uncertainty, we can calculate the uncertainty in the slope determined by the Least Square
method as
r
m = y

N
= y

s
N

x2i

N
P
( xi )2

(8.3)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

156

Putting in the values, we get a value of


m = 0.002
We can thus quote the result for the slope of the best fit line as
m = 0.0593 0.002

La

M
an
ua
l

Nu
cle
ar
Ph
ys
ics

A similar exercise would need to be done with all the other 3 sources and similar graphs would be
obtained.

Chapter 9

Nu
cle
ar
Ph
ys
ics

Experiment: To Determine the Range of


rays in Aluminum and to determine the
End Point Energy
9.1

Introduction

M
an
ua
l

In Chapter 3, we saw how electrons, both mono-energetic and those which are produced in nuclear
beta decay interact with matter. Section 3.2.2, we saw that for a beam of mono-energetic electrons the
energy loss of electrons is much smaller than that of heavy charged particles of the same energy. This
means that they have a much larger range. What is observed experimentally is that for a wide variety
of absorber materials, the product of the range and the density of the absorber is a constant for any
particular electron energy.

La

The situation is very different for the beta particles emitted by a radioactive source. This is because, as
we have seen in Section 2.2.2 in Fig 2.2, the energy spectrum of the beta particles is continuous. What
is seen therefore is that the beta particles at the lower end of the spectrum are absorbed even with
a very thin absorber. However, for most part of the spectrum, the transmission of the beta particles
follows an exponential curve with thickness. This is an experimental fact which cannot be easily derived
from fundamental physics. What we see is that the counting rate (or intensity) falls off exponentially
with an attenuation coefficient which depends on the end point energy of the beta particle.
C = C0 ed

(9.1)

where C is the counting rate with the absorber material, C0 is the counting rate without the absorber
and d is the mass thickness in units of mass per unit area. The coefficient is the attenuation
coefficient. Thus, a graph between ln CC0 and d would give us a straight line whose slope will be
the attenuation coefficient. This behaviour is shown in Figure 3.9 where the flat part of the curve

157

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

158

corresponds to the count rate going to the constant background value.

Nu
cle
ar
Ph
ys
ics

We also know from Chapter 2, that the beta particle spectrum is a continuous one with a maximum
energy determined by the Q value of the nuclear reaction producing the beta particle. Thus, we can infer
that beta particles from a radioactive source will have different penetrating power and the maximum
penetration depth, the range will be for particles with the maximum or end-point energy. It turns out
that for aluminum absorbers, a single range energy relationship for both monoenergetic electrons and
beta particles with energy range 0.01 E 2.5 MeV exists as was shown empirically by Katz and
Penfold (Reviews of Modern Physics, 24, page 28, 1952). They found that the relationship is given by

R = 412E0n

n = 1.265 0.0954 ln E0

(9.2)

Here E0 is in MeV and the range R is in units of mass thickness, that is mg cm2 . Eq(9.2) allows
us to determine the end point energy of the beta rays from a radioactive source. If we can find the
maximum range R0 experimentally, then that will give us the value of Emax which will be the end point
energy since the maximum energy will correspond to the maximum range.

9.2.1

Experiment

M
an
ua
l

9.2

Purpose

9.2.2

La

In this experiment, we will study the absorption of rays in aluminum and investigate the exponential
attenuation of Eq(9.1). We will also determine the end point energy of beta rays from 90Sr using the
range-energy relationship Eq(9.2).

Method

For this experiment, we would need a GM counter, a radioactive source ( 90Sr), and aluminum absorber
foils of varying thickness.
We first need to determine the operating voltage of the GM counter as in Chapter 5. We then determine
the background counts for a preset time of say 120 seconds. Next we take a 90Sr source with known
activity. The decay scheme for the source is given in Figure 9.1.

159

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Figure 9.1: Decay scheme for

90

Sr

We see that the dominant emission is with a maximum energy of around 2.28 MeV and so we can
assume that the range energy relation given above will be valid.

Sample Data

La

9.2.3

M
an
ua
l

We first find the count rate without any absorber and then use different aluminum foils to block the
beta particles and measure the count rate with the varying thickness of the absorber which is aluminum.
We continue this process till the number of counts reaches a constant which is the background value
obtained earlier.

Background Count

Operating Voltage = 420 V


Preset Time = 120 seconds
90
Sr source with Activity 3.7 KBq

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

S.No.
1
2
3
4
5

NB
(NB NB )
45
-1.8
47
-0.2
47
-0.2
49
1.8
48
0.8
NB = 47.2

160

(NB NB )2
3.24
0.04
0.04
3.24
0.64
P
(NB NB )2 = 7.16

Nu
cle
ar
Ph
ys
ics

Table 9.1: Background Count rate for 120 seconds

With these 5 values of the background counts, we determine the mean to be 47.2 counts in 120 seconds.
To determine the error in this, we can think of using use Eq(1.61) to determine the error in the mean.
For the of the parent distribution, we can use the sample standard deviation s given by
rP

s=

(x x)2
N 1

In our case, we get s = 1.33. From this, we can estimate the error in the mean (from Eq(1.61)) to be
s
=
N

M
an
ua
l

or

1.33
NB = = 0.59
5

Thus one would think that one should report the result as
NB = 47.20 0.59

La

However, there is a subtle point here that one needs to understand. Let us see what we are doing- we
are taking the total number of background counts in 120 seconds and reporting it as a number. Then
we are repeating the same procedure 5 times. Note that in each of the values of NB , there is an error.
This is the inherent statistical error that is associated with the event which as we know is a result of a
Poisson distribution. We can therefore think of each value of NB as a mean of a Poisson distribution.

The associated error with each value of NB is thus NB . Therefore, the correct procedure to exhibit
this inherent statistical error would be to take the expression for the mean value of NB , that is NB and
apply the appropriate error prorogation equation to it. In our case this means

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

5
P

2
=
N
B

=
=
NB =
=

(NB )i

i=1

5

1  2
2
2
(NB )1 + (N
+ + (N
B )2
B )5
25
1
[45 + 47 + 47 + 49 + 48]
25
236
25

236
5
3.072

Nu
cle
ar
Ph
ys
ics

NB =

161

Thus our background count should be reported as

NB = 47.20 3.07

Next we take several values of the counts with the source and without any absorbers.

N0
(N0 N0 )
(N0 N0 )2
8082
47.8
2284.84
7975
-59.2
3504.64
8010
-24.2
585.64
8188
153.8
23654.54
7916
-118.2
13971.24
P
N0 = 8034.2
(N0 N0 )2 = 44000.9

Table 9.2: Count rates without absorbers for

La

M
an
ua
l

S.No.
1
2
3
4
5

The sample standard deviation is therefore


rP
r
(x x)2
44000.9
=
= 104.9
s=
N 1
4
Thus the error in the mean is
s
104.9
= = = 46.9
5
N
Thus our count rate is
N0 = 8034.2 46.9

90

Sr

(9.3)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

162

The data for various thicknesses of aluminum plates is as below.


2.7 gm cm3 .

Thickness(cm) t ( mg cm )

8034.2
4396
2768
2346
1938
1946
1796
1188
1041
910
556
586
513
402
276
212
199
136
110
103
62
72
65
59
51
50
51
53
51
52

7987.2
4349
2721
2299
1891
1899
1749
1141
994
863
509
539
466
355
229
165
152
89
63
56
15
25
18
12
4
3
4
6
4
5

M
an
ua
l

0
100
207
256
307
332
356
432
443
461
539
543
551
588
650
699
707
758
775
807
883
888
914
963
988
994
1039
1151
1258
1331

La

0
0.037
0.077
0.095
0.114
0.123
0.132
0.160
0.164
0.172
0.200
0.201
0.204
0.218
0.241
0.259
0.262
0.281
0.287
0.299
0.327
0.329
0.339
0.357
0.366
0.368
0.385
0.426
0.466
0.493

No. of Counts N N NB

Error Transmission Coefficient


46.9
66.65
53.05
48.91
44.55
44.64
42.93
35.14
32.98
30.93
24.55
25.16
23.66
21.19
17.97
16.09
15.68
13.52
12.53
12.25
10.44
10.91
10.58
10.29
9.89
9.85
9.89
10.
9.9
9.9

N NB
N0 NB

1
0.5445
0.3407
0.2878
0.2368
0.2378
0.2190
0.1428
0.1244
0.1081
0.0640
0.0670
0.0580
0.0440
0.0290
0.0210
0.0190
0.0110
0.0080
0.0070
0.0019
0.0031
0.0022
0.0015
0.0005
0.0004
0.0005
0.0006
0.0005
0.0005

Nu
cle
ar
Ph
ys
ics

The density of aluminum is

Table 9.3: Count rates with absorbers from

90

Sr, Error Estimation & Transmission Coefficient

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

163

We can plot the graph of N NB vs t on a semilog scale. The error bars for N NB are obtained
as discussed in Section 1.5.1. We see that as the net count rate becomes small, the errors increase
drastically and at very small net count rates, the errors are much more than the count rate itself. We
also plot the transmission coefficient (multiplied by a 100, to get a percentage) against t. We also plot

the graph for N vs t with the error bars given by N . The graphs are shown in Figure 9.2.

Beta Rays Range


N-N_B: data
N-N_B data
N-N_B: best fit
Tramsission Data
Transmission Best Fit
N: data
N: data with errorbars

1000

100

10

0.1

0.01
0

200

Nu
cle
ar
Ph
ys
ics

N, N-N_B, transmission coeff

10000

400

600

800

1000

1200

1400

M
an
ua
l

t( mg cm^-2)

Figure 9.2: Graph of N NB vs t

La

We can see that the net count rate approaches a constant value which is close to zero while the count
rate approaches the background value after a point. This is the point where essentially all the beta
particles have been stopped by the absorber and thus gives us the range for the beta particles from
this particular 90Sr source. To determine the range, we take the N vs t graph and determine the point
where it turns to become the constant background value. The error in the determination of the range

is then the difference in the values between the range obtained from N and that from N N . For
this sample data, we obtain
R = 900 15 mg cm2
Using this range, we can now calculate the end point energy for the beta rays from this source, using
Eq(9.2). Thus

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

164

R = 412E0n
n = 1.265 0.0954 ln E0
ln R = ln 412 + n ln E0
= ln 412 + (1.265 0.954 ln E0 ) ln E0
R
412


= 0
1.265 1.141
0.1908
= 1.92, 299539

Nu
cle
ar
Ph
ys
ics

0.0954(ln E0 )2 1.265 ln E0 + ln

ln E0 =
E0

E0 = 1.92 MeV

(9.4)

The equation to determine ln E0 is a quadratic and the two solutions are given above. Clearly, the only
reasonable value is E0 = 1.92 MeV.
We can also determine the lower and upper limits of the end point energy corresponding to the error
in the determination of the Range. Thus
R = 900 15 mg cm2 , E0 = 1.88 MeV

M
an
ua
l

R = 900 + 15 mg cm2 , E0 = 1.94 MeV

Thus

E0 = (1.92 + 0.02 0.04) MeV

La

The theoretical value of the end point energy for this source is known to be 2.28 MeV (Figure 9.1).
Thus the percentage error in our determination is simply
% error =

2.28 1.92
= 15.8%
2.28

Chapter 10

10.1

Introduction

Nu
cle
ar
Ph
ys
ics

Experiment: Scintillation Counter

10.2

Theory

M
an
ua
l

In all the previous experiments, we have been using a GM counter to detect and measure radiation
from radioactive nuclides. In this experiment we will use a different kind of detector, the Scintillation
Counter. This works on a very different principle than a GM counter which is a gas filled detector
working on the principle of an ionising particle or radiation causing an avalanche. We shall study the
working of this counter as well as its method of operation.

The Scintillation counter basically can be thought of a scintillating material which emits light when
radiation or an ionising particle interacts with it, and a mechanism for collecting and measuring the
light which is produced. We shall study these separately. The incoming particle loses its energy and it
is this energy which is ultimately converted into light.

La

Actually the use of light produced in certain kinds of materials when radiation or a particle interacts
with it is one of the oldest ways of detection of such radiation. In the historic Rutherford experiment for
instance, the scattered particles were detected by the light flashes they produced on a Zinc Sulphide
screen. In fact, the first detection of X-rays by Roentgen also used scintillation- the platino-barium
cyanide crystals began to glow when the rays from his apparatus interacted with them. So how is this
light produced and what are the properties that we want the material of the detector to have in order
for it to be useful?
The emission of light by a material which is excited can be of several kinds. The most common one,
fluorescence is what we get when a material which has been excited, emits light immediately after
excitation and the light is in the visible region of the electromagnetic spectrum. Phosphorescence
165

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

166

is basically the same as fluorescence but here the light emitted is of a much longer wavelength and is
emitted usually on a time scale much longer than fluorescence.
We can define some terms which would be useful later on:

1. Luminescence is the process of exciting a material (not thermally) and the subsequent emission
of light.

Nu
cle
ar
Ph
ys
ics

2. How the material is excited determines the type of luminescence (e.g. photoluminescence, chemiluminescence, triboluminescence)
3. Fluorescence is photoluminescence or scintillation (i.e. excitation produced by ionizing radiation)
that has a fast decay time (nanoseconds or s)
4. Phosphorescence is the same as fluorescence, but with a much slower decay time (milliseconds
to seconds)
There are basically two types of scintillating materials- inorganic and organic . These two kinds of
materials have very different mechanisms of production of light. In our laboratory, we use inorganic
scintillators. We shall study these two kinds separately.

Inorganic Scintillators

M
an
ua
l

10.2.1

La

To understand the phenomenon of scintillation in inorganic materials, we first need to understand the
energy structure of crystalline materials since the mechanism for scintillation depends crucially on the
structure of the crystal. Recall that an atom has discrete energy levels or orbitals. When several
atoms form a molecule we get molecular orbitals which as molecules aggregate into a solid, combine
and become more and more dense. Finally, when a large number of molecules combine to form a
solid, the energy levels become so close to each other that they can be considered to form a continuum
which is called an energy band. The width of energy bands depends on the atomic orbitals which are
superposed to form the band. It also happens that there are some energies where there is no overlap at
all and we then get band gaps.The width of the bands of course can vary and depends on the overlap
between the underlying atomic orbitals.
Typically, a solid has an infinite number of allowed bands but most of the them have very high
energies to be of any relevance. It turns out that the electronic properties of solids depends on
the bands which are near the Fermi level. Fermi level (NOT Fermi energy) in the band theory
is a hypothetical level such that at equilibrium, it has a 50% probability of being filled. It is not
necessarily an actual energy level. In the language of Fermi-Dirac statistics, it corresponds to the total
chemical potential of the electrons. The closest band above the Fermi level is called the conduction
band and the one closest below is called the valence band. The band gap is large in insulating

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

167

Nu
cle
ar
Ph
ys
ics

materials, somewhat smaller in semiconductor materials and very small in conductors, as seen in Figure
10.1. The electrons will never have energy in the forbidden region or the band gap. Electrons in a
lower energy state are in the valence band and these are tightly bound to lattice sites. On the other
hand, electrons in the conduction band have a higher energy and are more mobile throughout the crystal.

M
an
ua
l

Figure 10.1: Band gap in different class of materials

La

Now let us consider what happens when radiation (we shall use the term radiation to denote both
ionising radiation and charged particles which can ionise the material), hits a pure inorganic crystal
like Sodium Iodide. The radiation can deposit its energy in the crystal and one of the electrons in the
valence band can gain enough energy to move to the conduction band, leaving a hole in the valence
band. When this excited electron returns to the valence band, it will emit a photon though in a pure
crystal, this is a very inefficient process. Furthermore, the energy gap between the conduction and
valence band is typically so large that the photon emitted is of short wavelength and not in the visible
region.
To get the crystal to emit light in the visible range, we obviously need to decrease the energy gap that
the electron experiences when it falls or de-excites from the conduction band. This is usually done by
adding a trace amount of impurity to the pure crystal. These impurities are called activators. The
activator sites modify the energy band structure of the pure crystal locally while the overall energy
band structure remains unmodified. The net effect of the activator sites is to create energy levels in
the forbidden region (the region between the conduction and the valence band, where ordinarily in a
pure crystal, the electrons cannot be). This is shown in Figure 10.2.

168

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Figure 10.2: Energy Band Structure of Pure crystal & Activated Crystal

Now let us consider what happens when radiation or a charged particle enters the crystal with activator
sites. The incoming particle deposits its energy and creates an electron hole pair in the valence band.
This primary electron-hole pair, through a cascade effect creates many secondary electron-hole pairs.
When the energy of the electronic excitations becomes below the ionization energy, thermalization
takes place. At the end of this stage, all the electrons are at the bottom of the conduction band and
the holes at the top of the valence band. This whole process takes place on a time scale of a picosecond.

La

M
an
ua
l

After the thermalization stage, the free electron hole pairs migrate through the material. The hole will
migrate to the activator site and ionise it. The electron in the conduction band continues migrating
till it meets one of the ionised activator site and neutralises it. Now we have a neutral activator, but
one which depending on the energy of the electron will be an excited state (of the neutral activator).
This excited state will de-excite to its ground state and in the process give out a photon which, now
since the energy difference between the activator excited and ground state is smaller than the original
band gap of the crystal, will be in the visible region. This process takes place on a time scale of around
1010 1011 seconds. This is essentially then the time scale of the scintillation which one observes.
Sometimes, the electron in the conduction band when it encounters the ionised activator or impurity
site, neutralises it but goes into an excited state from where the transition to the ground state is
forbidden. In that case, the electron typically gains more energy from thermal motion and moves to
a higher excited state from which it can de-excite to the ground state emitting light. Obviously, this
process is on a much longer time scale and sometimes therefore we get an after glow in scintillating
materials. This component of light emitted, as we saw above is phosphorescence.
It can also happen that the de-excitation of the electron from the activator excited state to its ground
state is such that no visible light is emitted. This process is called quenching.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

169

Nu
cle
ar
Ph
ys
ics

The processes we have outlined above depend on the dynamics of an electron-hole pair created in the
valence band, which are essentially independent of each other. In semiconductors, there is another kind
of process which can take place. Instead of the electron and hole moving independently, sometimes it
happens that the electron and the hole form a loosely bound state called an exciton. The exciton
band is typically just below the conduction band and therefore the energy of the electron in an exciton
is somewhat lower than that in the conduction band. Here again, the electron hole exciton moves
together to the activator site and the hole gets neutralised while the electron in the exciton band
neutralises the ionised activator and when it de-excites to the ground state, can emit visible light. It
should be noted that the electron hole pair in an exciton can also recombine at the site of impurities or
traps in a crystal without any external doping of the kind used to create activator sites. The time scale
for the exciton recombination is typically much faster than that of the electron-hole pairs which move
independently since here the electron and hole move together. So we have two kinds of components
in most materials- the fast component which is caused by the recombination of excitons and the slow
component which is when the electrons in the conduction band and holes in the valence band are
captured successively by the activator sites. The fast and slow components can be resolved for most
scintillating materials.

10.2.2

La

M
an
ua
l

What we have thus seen is how an incoming radiation interacting with a crystal which has been doped
with an activator produces light through luminescence. One can do an elementary analysis of the
energy efficiency of such materials and we find that for every electron-hole produced by the incoming
radiation, there is roughly one photon produced. Further, note that the emitted light can essentially
pass right through the bulk of the crystal. This is because remember that the energy difference in the
pure crystal bulk between the conduction and valence bands was such that a de-excited electron from
the conduction gap produced short wavelength or high energy photons. The energy difference between
the activator excited states and its ground state is much less and so the light produced is of a longer
wavelength or smaller energy. There is thus usually no absorption of this light by the bulk of the crystal material since its emission(and hence absorption spectrum ) is peaked at a much shorter wavelength.

Organic Scintillators

The scintillation mechanism in organic materials is quite different from the mechanism in inorganic
crystals that we studied in the previous section. In inorganic scintillators, we saw that the scintillation
arises because of the structure of the crystal lattice and the impurities which we introduce. The
fluorescence mechanism in organic materials arises from transitions in the energy levels of a single
molecule and therefore the fluorescence can be observed independently of the physical state. Practical
organic scintillators are organic molecules which have symmetry properties associated with the electron
structure.

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

170

The molecular energy levels of organic molecules which exhibit scintillation are separated by a few
electron volts and they get closer to each other as we go up. The singlet energy levels are subdivided
into a series of levels with a much finer structure because of the vibrational modes of the molecule.
The typical spacing of these is around a tenth of an electron volt.

Nu
cle
ar
Ph
ys
ics

Now let us consider the case when radiation interacts with an organic scintillator. The average energy
at room temperature is around 0.025 eV (recall that an energy of 1 eV is roughly equal to the thermal
energy at 104 K, thus a room temperature of 300 K is about 0.03 eV) and so all the molecules are in
the ground state (called the S00 state where the first subscript indicates the singlet spin state and the
second the fine structure state). When radiation deposits its energy into the material, the electrons
are excited to the upper levels. The higher states like S2 , S3 etc. de-excite in a matter of picoseconds
to the S1 state via transitions which do not produce any radiation. The S1 states like S11 , S12 etc
with higher vibrational energy also lose energy and soon we have all the excited molecules in the S10
state. When these electrons in the S10 state de-excite to the S0 state, we get luminescence. Again,
essentially all the emission light is of a lower energy than that required for absorption and therefore
the organic material is transparent to the luminescence produced like in the case of inorganic scintillator.

La

M
an
ua
l

In addition to the transitions in the singlet states, there are also transitions from the triplet states to
the ground state. The triplet states are typically longer lived and therefore the typical time scale of
this transition is longer leading to a phosphorescence. A schematic illustration of the states is shown
in Figure 10.3.

Figure 10.3: Energy States in an organic scintillator


(Source:

Pistates by Napy1kenobi - Own work.

Licensed under CC BY-SA 3.0 via Wikimedia Commons -

https://commons.wikimedia.org/wiki/File:Pistates.svg#/media/File:Pistates.svg)

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

10.2.3

171

Photomultiplier Tube

We have seen now how ionising radiation or a charged particle from a radionuclide when it interacts with
a scintillator would produce light because of fluorescence. However, to convert this light into something
which can be detected or its properties measured requires some kind of sensor which is sensitive to the
light. The most often used sensor is a Photomultiplier Tube (PMT). Let us now see how this works.

La

M
an
ua
l

Nu
cle
ar
Ph
ys
ics

The basic purpose of a photomultiplier tube is to convert the light signal into an electrical signal
that is, ultimately convert a photon into one or more electrons which can be detected and whose
properties measured. The PMT basically consists of three components- a photocathode which
will produce the initial or primary electron on interaction with the photon; an arrangement to
accelerate the electron(s) produced and an arrangement to measure the current produced by these
electrons. The whole arrangement has to be in a vacuum tube. As we will see, the efficiency of the
photocathode to produce the primary electrons is not very high. Thus, what is usually done in a
PMT is to have an arrangement which multiplies the number of electrons that is produce a number of
secondary electrons from the primary electrons. Let us see how each of these three components function.

Figure 10.4: Schematic of a Photomultiplier Tube

(Source: Wikipedia)

The light from the scintillator is made to fall on a photocathode. This is, as the name suggests, made
of a material which by using the photoelectric effect, produces electrons when photons impinge on it.
Obviously, we need to choose the material and the design of the photocathode in such a way that the
energy of the photon is transferred efficiently to the primary electron which is produced. Firstly, recall
that when a photon transfers its energy to an electron, for the electron to emerge from the material
and be detected, it needs to overcome not only the collisions with other electrons and the lattice within
the material but also the surface potential barrier. Clearly then, there has to be a minimum energy of
the photon when this is possible. This means that any photocathode will have a long wavelength cutoff

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

172

depending on the material and the geometry of the photocathode. For our purposes, typical light
given off in the scintillation is in the blue region and thus has an energy of around 3 eV. ( 400 410
nanometers). It turns out that the surface potential barrier or work function for most metals is more
than 3 eV while that for some semiconductors is only around 2 eV. Thus we see that to detect the light
from the scintillator, we need a photocathode made from a suitably prepared semiconductor.

M
an
ua
l

Nu
cle
ar
Ph
ys
ics

But this is not enough- as we saw above, the electron on absorbing the energy from the photon,
needs to travel to the surface of the material to be ejected out. In its motion to the surface, it loses
energy also by collisions with other electrons. Clearly, the probability of collisions with other electrons
increases with increased electron density in the material as well as the distance travelled. Thus, metals
have a high energy loss and so the electrons travel only a small distance before losing enough energy to
be unable to escape from the surface. The maximum depth from which an electron can travel to the
surface of a material and still escape is called the escape depth. In metals it is only a few nanometers.
This fact then determines the geometry of the photocathode since we thus need a very thin layer of the
photosensitive material or else most of the electrons produced will not be able to escape. The situation
in semiconductors is better where the escape depth is a few tens of nanometers. Here again, we need to
have a very thin layer in order to maximise the probability of the electrons produced to escape from the
surface. This however, leads to another issue which effects the efficiency of the photocathode- a very
thin layer of photosensitive material means that it will allow a large fraction of the incident radiation
through, thereby reducing the efficiency of the photocathode. Thus, there is a trade off which needs to
be made between these two factors while determining the thickness of the photosensitive material.
The efficiency of the photocathode is usually described by a quantity called quantum efficiency which
is defined as
number of photoelectrons emitted
number of incident photons

Quantum Efficiency =

La

Clearly, this is a function of the wavelength of the incident light. Most PMTs have an efficiency of
around 15 25%.
The photomultiplier tube, as the name suggests, does more than just produce photoelectrons- it
also has a multiplying effect which we now turn to. First the primary electrons produced in the
photocathode layer are focussed onto a narrow region. In this process, the electrons are also accelerated
in an electric field of a few hundred volts. The focussing is done by using a focussing electrode.
The accelerated primary electrons are then made to produce secondary electrons by the process of
secondary emission using a series of electrodes called dynodes.
A dynode is basically an electrode in a vacuum tube that serves as an electron multiplier through
secondary emission. When the accelerated primary electrons are focussed on a dynode, they transfer

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

173

Nu
cle
ar
Ph
ys
ics

their energy to the electrons in the dynode material which are ejected out. Of course, just like the
photoelectric effect, the electrons in the dynode material need to have enough energy to overcome the
surface potential barrier which, as we have seen is around 3 4 eV. However, the primary electrons
produced by the photocathode were emerging with very low energies, ( 1 eV) but these are accelerated
through around 100 V and so when they strike the dynode, have an energy of 100 eV. Thus, if all this
energy was transferred to the electrons in the dynode, we could in principle get around 30 secondary
electrons for each primary electron striking the dynode. Clearly, this is the maximum number of
secondary electrons that can be produced per primary electron. The actual number is significantly less
because once again, for the electrons to emerge from the material, they need to travel to the surface
and in this process lose energy. Only those electrons which reach the surface with energy more than the
work function can escape. Typically, around 6 8 secondary electrons are produced for each primary
electron impinging on the dynode.

La

M
an
ua
l

The next stage in the PMT is to multiply this number of secondary electrons. This is done by using
different geometries of an array of dynodes. Each dynode when struck by the secondary electrons from
the previous dynode produces more secondary electrons which then impinge on the next dynode to
produce an even larger number, leading to a cascade effect. Various kinds of geometries are used to
achieve this. A fairly simply, box and grid type of arrangement is shown in Figure 10.4. Finally, the
multiplication achieved is around 107 by a PMT. The electrons which finally emerge from the dynode
arrangement are collected and analysed using the associated electronics to which the PMT is connected.

Index

Absorption of Gamma Rays in Iron, 144


accuracy, 7

Propagation of errors, 39
expectation value, 15

Nu
cle
ar
Ph
ys
ics

Interaction with matter


Rayleigh Scattering, 87
Students t-distribution, 54

flux, 72

La

M
an
ua
l

GM Counter, 98
Absolute Efficiency, 116
barn, 74
Counting Efficiency, 115
Beta Ray end point energy, 157
Dead Time, 110
Binding Energy, 65
Detector Regions, 105
Binomial Distribution, 19
External Quenching, 109
mean, 20
fill gas, 108
probability, 20
gas multiplication, 102
variance, 21
Geiger discharge, 106
Geiger Region, 106
Central Limit Theorem, 38
Internal Quenching, 109
Chi Squared test, 59
Intrinsic Efficiency, 116
classical electron radius, 79
Ion Chamber, 105
Continuous Random Variable, 6
Operating Voltage, 113
Covariance, 40
Plateau, 113
proportional counters, 106
degrees of freedom, 59
Quench gas, 109
Detector Models, 98
Quenching, 108
Discrete Random Variable, 6
region of limited proportionality, 106
error analysis, 6
Region of true proportionality, 106
Error Estimation, 38
starting voltage, 113
Error propagation equation, 41
time constant, 101
error propagation for multiplication & division,
Townsend Avalanche, 102
42
Townsend Equation, 104
error propagation for Powers, 43
two source method, 111
error propagation for sums and differences, 41
work function, 109
error propagation for weighted sums and dif- GM Counter Characteristics, 118
ferences, 41
preset time, 121
174

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

Mass attenuation Coefficient, 144


mean, 13
Method of Least Squares, 55
Method of Maximum Likelihood, 51

half thickness, 144


Health effects
Non-Stochastic health effects, 119
rem, 120
sievert, 120
Stochastic health effects, 119
Health effects of radiation, 118
histogram, 10

La

M
an
ua
l

Interaction with matter


Bethe-Bloch for electrons, 83
Bethe-Bloch formula, 79
Bohr formula, 78
Bremsstrahlung, 83
Cherenkov radiation, 83
Compton Scattering, 87
critical energy, 85
cross section, 72
electrons, 83
gamma rays, 87
Heavy particles, 75
Klein Nishina Formula, 93
mass thickness, 74
mean range, 82
minimum ionizing particles, 80
Pair Production, 87, 95
Photoelectric Absorption, 87
radiative loss, 84
Straggling, 82
Thomson Scattering, 90
internal conversion, 70
Inverse Square law for Gamma Rays, 150
limiting distribution, 11
Linear attenuation coefficient, 144
linear regression, 55

neutrino, 67
Normal Distribution, 30
Mean, 36
Mean of standard normal, 32
probability density, 30
probability density function, 30
standard form, 31
Variance, 36
Variance of standard normal, 32
nuclear magneton, 64

Nu
cle
ar
Ph
ys
ics

GM Counting Efficiency, 133


Goodness of Fit, 58

parent distribution, 13
Poisson Distribution, 24
Mean, 26
Probability function, 26
Variance, 26
Precautions, 118
precision, 7
probability, 4
probability distribution function, 12
Radioacitivity
Nuclear Decay, 64
Radioactivity, 61
Alpha decay, 65
Becquerel unit, 62
Beta decay, 67
curie unit, 62
decay constant, 62
Decay law, 63
definition of activity, 61
electron capture, 69
Gamma Decay, 69
Gamow theory of alpha decay, 66
Half Life, 62

175

Mean Lifetime, 63
positron emission, 69
sieverts unit, 62
specific activity, 64
Activity Law, 62
Radioctivity
half-life, 62
random errors, 8
random variable, 5
reduced Chi squared, 59
relative uncertainty, 7
rules of probability, 5

La

M
an
ua
l

sample distribution, 13
Scintillation Counter, 165
activators, 167
dynode, 172
fluorescence, 165
Inorganic Scintillators, 166
Organic Scintillators, 166
Phosphorescence, 166
Photomultiplier Tube, 171
secondary emission, 172
significant figures, 9
Square root rule, 39
standard deviation, 14
standard error, 53
systematic errors, 8

Nu
cle
ar
Ph
ys
ics

MANUAL FOR M.Sc.(P) NUCLEAR PHYSICS LAB.

variance, 14

176

You might also like