You are on page 1of 202

Advanced Experiments in Physics.

Physics 306
University of Wisconsin-Parkside
Jeffrey R. Schmidt
October 1997, 1999, 2001, 2003, Revision: December 2005

NaI

Photocathode

Dynodes

Contents
1 Introduction

2 Nuclear Counting Statistics


2.1 Apparatus . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Background . . . . . . . . . . . . . . . . . . . . . . .
2.3 Geiger tube calibration . . . . . . . . . . . . . . . . .
2.4 Experiment I; Normal distributions . . . . . . . . . .
2.5 Experiment II. Hypothesis testing and goodness of fit
3 The
3.1
3.2
3.3
3.4
3.5

Half-life of Ba137
Background . . . . .
Apparatus . . . . . .
Procedure . . . . . .
Data and analysis . .
Background material;

. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .
the Nuclear Shell

. . . .
. . . .
. . . .
. . . .
Model

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

5
5
5
6
8
14

.
.
.
.
.

20
20
22
23
23
27

4 Determination of the Speed of Light

36

5 Determination of CCvp for a gas


5.1 Background . . . . . . . . . . . . . . . . . . . .
5.2 Apparatus . . . . . . . . . . . . . . . . . . . . .
5.3 Procedure . . . . . . . . . . . . . . . . . . . . .
5.4 Data and analysis . . . . . . . . . . . . . . . . .
5.5 Molecular spectra and the equipartition theorem
5.6 Safe Handling of High Pressure Gas Cylinders .
5.7 Lissojous program . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

40
40
42
42
44
45
49
49

.
.
.
.
.

53
53
54
57
61
64

Photoelectric Effect
Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Procedure and data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

69
69
70
71

6 Electron Diffraction and Bragg Scattering


6.1 Apparatus . . . . . . . . . . . . . . . . . . . .
6.2 Background . . . . . . . . . . . . . . . . . . .
6.3 Data and analysis . . . . . . . . . . . . . . . .
6.4 Crystal structure determination by scattering
6.5 Simulation of scattering . . . . . . . . . . . .
7 The
7.1
7.2
7.3

8 Measuring Plancks constant (solid state


8.1 Theory . . . . . . . . . . . . . . . . . . .
8.2 Procedure . . . . . . . . . . . . . . . . .
8.2.1 Finding roots by bisection . . . .
8.2.2 Newtons method . . . . . . . . .
2

.
.
.
.
.

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

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

version)
. . . . . .
. . . . . .
. . . . . .
. . . . . .

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

.
.
.
.

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

.
.
.
.

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

.
.
.
.

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

.
.
.
.

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

.
.
.
.

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

.
.
.
.

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

.
.
.
.

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

.
.
.
.

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

.
.
.
.

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

.
.
.
.

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

.
.
.
.

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

.
.
.
.

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

.
.
.
.

.
.
.
.

73
73
78
79
80

9 Helium Ionization Potential; The Franck-Hertz


9.1 The old version of the experiment; apparatus . .
9.2 Data and analysis . . . . . . . . . . . . . . . . .
9.3 Second version . . . . . . . . . . . . . . . . . .
10 The
10.1
10.2
10.3
10.4
10.5

Rydberg Constant
Background . . . . .
Atomic Spectra . . .
Apparatus . . . . . .
Data and analysis . .
Diffraction gratings .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

Experiment
. . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . .
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

83
84
86
87
89
89
94
98
100
102

11 Measurement of unknown spectral


11.1 The Mercury data . . . . . . . . .
11.2 First order calibration curve . . .
11.3 Measuring Sodium lines . . . . .

lines
106
. . . . . . . . . . . . . . . . . . . . . . . . 107
. . . . . . . . . . . . . . . . . . . . . . . . 108
. . . . . . . . . . . . . . . . . . . . . . . . 108

12 Stefan-Boltzmann Radiation Law


12.1 Background . . . . . . . . . . . .
12.2 Luminosity or radiance . . . . . .
12.3 Apparatus . . . . . . . . . . . . .
12.4 Procedure . . . . . . . . . . . . .
12.5 Data and analysis . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

109
109
115
116
122
122

13 Measurement of the Compton Edge


13.1 Background . . . . . . . . . . . . . . . . . .
13.2 Apparatus . . . . . . . . . . . . . . . . . . .
13.3 Scintillation counters and photomultipliers .
13.4 Stage 1; calibration . . . . . . . . . . . . . .
13.5 Stage II; photopeak and edge measurements

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

126
126
129
130
134
135

Electrons
. . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . .

135
136
136
137

14 Energy versus Momentum


14.1 Apparatus . . . . . . . .
14.2 Procedure . . . . . . . .
14.3 Data and analysis . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

for Relativistic
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .

15 Michelson-Interferometer and the Index of Refraction of Air

138

16 Angular distribution of emitted radiation


142
16.1 Polarization and angular distributions . . . . . . . . . . . . . . . . . . . . . . 145
16.2 Angular correlations of emissions . . . . . . . . . . . . . . . . . . . . . . . 148
16.3 A simulated experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

17 A simulated experiment; Geiger-Marsden experiment


17.1 The numerical experiment . . . . . . . . . . . . . . . . .
17.2 Raw data . . . . . . . . . . . . . . . . . . . . . . . . . .
17.3 Problem 1. Analysis of simulated lab data . . . . . . . .
17.4 The Thompson model . . . . . . . . . . . . . . . . . . .
17.5 Monte Carlo scattering simulation . . . . . . . . . . . . .
17.6 Energy dependence in the Geiger-Marsden experiment .
17.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . .
17.8 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
18 Appendix
18.1 Appendix
18.2 Appendix
18.3 Appendix
18.4 Appendix
18.5 Appendix
18.6 Appendix

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

I. Reformatting Lab Data with Sed and Awk . . .


II; Nonlinear Least Squares error matrix method .
III. The Plotting of Lab Data with GNU plotutils
IV. Preparing a lab report with LaTeX . . . . . .
V. The Oscilloscope . . . . . . . . . . . . . . . . .
VI. Support software . . . . . . . . . . . . . . . .

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

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

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

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

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

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

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

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

.
.
.
.
.
.
.
.

154
163
164
165
167
168
170
174
175

.
.
.
.
.
.

178
178
184
187
191
195
196

Introduction

This lab book is a synopsis of the setup, required apparatus, and data collection and analysis for each of our Modern Physics Laboratory experiments. The data for each experiment
are real, taken in 1996, 1997, 1999, and 2001. The book can serve as an outline and basic
reference for the setup of the experiments and their proper analysis.
I do not generally teach this course at the present time, so this has been prepared as a
courtesy to the students and any person who must teach the course for the first time, but
mostly in preparation for the day when I will be teaching it more regularly.
In addition to twelve basic experiments, and a variation on one or two of these, there are a
number of simulated experiments; things that would be interesting to do if we had the equipment. These include the Geiger-Mardsen experiment, Laue scattering, and angular
correlations. There are two numerical labs in the Geiger-Marsden section, one verifying the
energy dependence of the scattering cross section, another verifying the Rutherford formula
for the number of particles scattered from a gold foil. Hopefully I will add more simulations
as time allows.

c
1997,
2003, 2005 Jeffrey R. Schmidt

The experiments

Nuclear Counting Statistics

Nuclear decays are random processes that provide an excellent context in which to study
basic statistics and distribution functions. We will determine whether or not nuclear decays
are random events by counting radioactive decays with a Geiger tube and a scalar-counter.

SARGENTWELCH | SCALAR \ TIMER

Count

Geiger Tube Voltage

Mode
Ext

50/60 Hz

Freq

Volts

Preset

Freq

Time

2.1

Start

Count Time

Preset

Time

10M
100k

1s
10s

10k
1k
100
10

1min
10min
1hr
10hr

Reset Power

Apparatus

1. Geiger tube. Welch Scientific #1216.


2. Scaler-timer. Sargent-Welch model S-72095-10.
3. Radioactive source. 0.1c Cs137 , Amerschorn-Searle 184471.

2.2

Background

In at least four experiments in 306 you will study radiations emitted from unstable isotopes.
Some of these are given in the table below
Isotope
137
,
55 Cs
27 Co

58 Ce

60

144

11 N a

22

137
56 Ba

58 N i

60

, P r 144

Emission
; 0.514 M eV
; 0.662 M eV
; 0.312 M eV
; 1.172 M eV
; 1.333 M eV
; 0.320 M eV
; 0.184 M eV
; 0.134 M eV
; 0.081 M eV
+ ; 0.542 M eV
; 1.277 M eV
; 0.511 M eV
(From + )
5

Half-life
27 year
5.2 year

285 days

2.6 year

The Geiger tube is used to detect the decay particles emitted by the Cesium sample. The
scalar-timer counts the number of decays from the source in ten second intervals. This is the
data that will be statistically analyzed. A selection of our Geiger counters is shown below.
The counter is a cylindrical glass tubewith a thin tungsten wire running down the center,
and a coaxial metal cylinder enclosed in the glass. The tube is filled with air at a very low
pressure, a few torr. A voltage typically around 1000 V is applied between the wire and
cylinder, with wire positive and cylinder negative. This voltage is slightly below dielectric
breakdown voltage for the gas in the tube.

Tungsten wire

cylinder

1000 V

Ionizing radiation entering the tube will free ions from air molecules, which will in turn be
accelerated by the potential difference in the tube. This causes more collisions, liberating
more ions and creating a current flow from wire to cylinder and through the external resistor
R. The back-voltage across R combines with the voltage on the tube and diminishes it,
causing the net potential difference between wire and cylinder to drop below the critical
value needed to sustain the cascade of ions. This shuts off the current, but not before it can
be registered on the supporting electronics connected to the tube.
Proper operation of the tube depends critically on the voltage applied to it. Too high a voltage and the tube discharges spuriously from dielectric breakdown, registering false counts
even with no ionizing radiation present. To find the optimal operating voltage, the tube
must be calibrated by counting a constant source of ionizing radiation and generating the
characteristic curve of counts versus voltage for the particular tube.

2.3

Geiger tube calibration

A constant source is placed about 20 cm from the tube, and is counted at a variety of applied voltages. For voltages below the threshold value there will be no counts detected by
the tube, as the voltage is increased above this point. the number of registered counts will
increase and reach a plateau. The center of this plateau is the desired operating voltage.
As the voltage continues to climb, the number of counts blows up as the tube registers spurious counts from dielectric breakdowns.
6

Sample calibration data This data was taken by two students in 1995 for one of the
Sargent-Welch Geiger-Muller tubes used in the Modern Physics lab course.
Voltage (volts)
507
550
603
653
705
750
774
805
853
898

Counts
0
0
0
0
0
0
158
886
1120
1185

Voltage (volts)
954
1003
1058
1099
1162
1202
1263
1309
1354
1399

Counts
1192
1238
1258
1206
1221
1256
1295
1371
1507
1712

the plateau is fairly evident from the graph

Geiger Plateau
2000

Counts

1500

1000

500

0
400

600

800
1000
Voltage V
7

1200

1400

which indicates an operating voltage around 1038 volts.

In the experiment we will attempt to verify that the frequency with which the Geiger counter
gets X counts per time T is distributed as a random variable, demonstrating that nuclear
decays are random.
If this hypothesis is true, in time dt there is probability p that a nucleus in the sample will
decay, (1 p) that it will not, so in time T = N dt we should have a probability of getting
M counts of
!
M
N M
N M
p (1 p)

(M ) =
e
M
M!
where = N p, in other words a Poissonian distribution.

2.4

Experiment I; Normal distributions

We can use nuclear decays to learn a little bit about the Normal distribution, assuming that
the decay process is random and Normally distributed.
For data that is distributed Normally, meaning by a frequency determined by a PDF equal
to that of a Normal distribution, how many data should occur within one standard deviation
of the mean?
Z
x2
1

e 22 dx = 0.34135
0
2 2
Z 2
x2
1

e 22 dx = 0.47725
0
2 2
Z 3
x2
1

e 22 dx = 0.4985
0
2 2
In other words, within one standard deviation of the mean you will find 68.26% of all of the
data (both sides of the mean), and within 2 of the mean you will find 95.45% of the data.
Consider the following set of 104 Geiger counter counts for a Barium sample, each taken
over the same time period.

459
449
430
387
410
396
458
427

422
398
449
405
415
436
459
407

434
418
437
409
363
461
444
435

410
407
421
406
438
402
381
405

434
399
460
449
422
377
389
442

438
445
414
437
377
443
438
423

454
424
447
433
410
411
435
398

421
440
396
444
408
383
374
423

390
430
394
396
406
433
440
417

422
391
430
394
431
420
409
423

438
421
418
389
438
393
431
428

436
433
392
426
434
424
433
423

444
401
414
453
413
433
455
406

We can write a little C program to read this data (as a long list of numbers, one per line)
from a file into an array, from which an unbiased estimator for the mean and standard deviation can be gotten, as well as a largest and smallest value.
/* simple data sorting, mean and standard deviation program */
/* gcc sort.c -lm
*/
#include<stdio.h>
#include<stdlib.h>
#include<math.h>
int numbers[200];
double mean,sigma;
int largest,smallest,n,i,sum;
FILE *fptr;
main(int argc, char *argv[]){
if(argc!=2){
printf("./sort datafile\n");
exit(0);
}
fptr=fopen(argv[1],"r"); /* read data file named on command line */
n=0;
do{
fscanf(fptr,"%d", &numbers[n]);
n=n+1;
}while(!feof(fptr));
n=n-1;
/* we overshot */
/* get largest, smallest, do stats */
smallest=numbers[0];
largest=numbers[0];
for(i=1;i<n;i++){
if(numbers[i] < smallest) smallest=numbers[i];
if(numbers[i] > largest) largest=numbers[i];}
mean=0.0;
9

for(i=0;i<n;i++)
mean=mean+(double)numbers[i];
mean=mean/(double)n;
sigma=0.0;
for(i=0;i<n;i++)
sigma=sigma+((double)numbers[i]-mean)*((double)numbers[i]-mean);
sigma=sqrt(sigma/(double)n);
/* print the number of data, smallest, largest, mean and standard deviation */
printf("%d\t%d\t%d\t%f\t%f\n",n,smallest,largest,mean,sigma);
/* how many are within sigma of mean, to left ? */
sum=0;
for(i=0;i<n;i++)
if(numbers[i]<= mean && numbers[i]>= mean-sigma) sum=sum+1;
printf("There are %d between mean=%f and mean-sigma=%f\n",sum,
mean,mean-sigma);
/* how many are within sigma of mean, to right ? */
sum=0;
for(i=0;i<n;i++)
if(numbers[i]>= mean && numbers[i]<= mean+sigma) sum=sum+1;
printf("There are %d between mean=%f and mean+sigma=%f\n",sum,
mean,mean+sigma);

fclose(fptr);
}
This program can be downloaded from the 499 archive. If we run it on our data we obtain
gcc -o sort sort.c -lm
./sort data2
104
363
461
420.846154
21.848591
There are 25 between mean=420.846154 and mean-sigma=398.997562
There are 42 between mean=420.846154 and mean+sigma=442.694745
Notice that there are 25 + 42 = 67 of the 104 data within one standard deviation of the
mean, confirming the notion that the number of nuclear decays per fixed time interval is most
likely distributed Normally; this is 64.4%, agreeing quite well with the values computed for
a Normal distribution.
If we want to sort these data to create a frequency plot, graphing the number of data that fall
within a certain lower and upper limit, suitable for graphing, we use another short program.

10

/* simple data sorting, to create histogram plot


/* gcc pidgeon.c -lm
#include<stdio.h>
#include<stdlib.h>
#include<math.h>

*/
*/

int numbers[200],bins[100];
FILE *fptr;
main(int argc, char *argv[]){
float div;
int largest,smallest,n,N,i,which,sum;
if(argc!=5){
printf("./pidgeon datafile min max number\n");
exit(0);
}
fptr=fopen(argv[1],"r"); /* read data file named on command line */
smallest=atof(argv[2]);
largest=atof(argv[3]);
N=atoi(argv[4]);
div=((float)largest-(float)smallest)/(float)N;
n=0;
do{
fscanf(fptr,"%d", &numbers[n]);
n=n+1;
}while(!feof(fptr));
n=n-1;
/* we overshot */
/* zero all bins */
for(i=0;i<N;i++)
bins[i]=0;
/* pidgeon hole each bit of data */
for(i=0;i<n;i++){
which=(int)floor(((float)numbers[i]-(float)smallest)/div);
bins[which]=bins[which]+1;
}
for(i=0;i<N;i++)
printf("%f\t%d\n",smallest+(float)i*div+div/2.0,bins[i]);
}
This we can pass parameters to. For example, our data above could be made into a histogram
with 22 divisions between 360 and 470 with
gcc -o pidgeon pidgeon.c -lm
11

./pidgeon data2
365.000000
375.000000
385.000000
395.000000
405.000000
415.000000
425.000000
435.000000
445.000000
455.000000
465.000000

360 470 11
1
3
5
12
12
11
16
24
12
6
2

or piped directly into a graph with


./pidgeon data2 360 470 11 | graph -T X
This program is giving the midpoint of each division, versus the population of that division.
You may wish to determine exactly what fraction of the total data falls between xmin and
xmax for a Normal distribution of variance 2 . For zero mean this is the total number of
data points times the probability
=

xmax
xmin

x2
1
e 22 dx
2 2

which is computed with the program below.


#include<stdio.h>
#include<stdlib.h>
#include<math.h>
#define PI 3.1415926
double dx,x,sum;
double xmin,xmax,sigma;
int n;
main(int argc, char *argv[]){
if(argc != 4){
printf("./normal xmin xmax sigma\n");
exit(0);
}
xmin=(double)atof(argv[1]);
xmax=(double)atof(argv[2]);
sigma=(double)atof(argv[3]);
12

dx=(xmax-xmin)/10000.0;
sum=0.0;
for(n=0;n<10000;n++){
x=xmin+(double)n*dx;
sum=sum+exp(-x*x/(2.0*sigma*sigma));
}
sum=sum*dx/(sigma*sqrt(2.0*PI));
printf("%f\n",sum);
}
You can easily shift the mean of the distribution to the point of your choice. You might want
to use this program for the next nuclear counting experiment, to fit your count frequency to
a Gaussian Normal distribution rather than the Poissonian example.
You may wish to simulate the experiment before you actually perform it. The C program
below will generate N Normally distributed random numbers with mean mean and standard
deviation sigma by the accept-reject algorithm. You can use it to simulate the outcomes
of he nuclear counting experiments, and use the other programs to analyze the resulting data.
/* gcc -o accept_reject accept_reject.c -lm */
/* creates Normal deviates
*/
#include <stdlib.h>
#include <math.h>
#include <stdlib.h>
#define PI 3.1415926
int n;
float f( float x, float mean, float sigma);
main(int argc, char *argv[]){
float m1,m2,test;
float min,max,sigma,mean;
int N, crap;
if(argc != 4){
printf("./random mean sigma number\n");
exit(0);}
mean=(float)atof(argv[1]);
sigma=(float)atof(argv[2]);
N=atoi(argv[3]); /* how many to generate */
srand(17); /* initialize random number generator */
min=mean-5.0*sigma;
max=mean+5.0*sigma;
n=0;
13

do{
/* get a random number between mean-4*sigma, mean+4*sigma */
m1=min+(max-min)*(float)rand()/(float)RAND_MAX;
m2=f(m1,mean,sigma);
test=((1.2533141/sigma))*(float)rand()/(float)RAND_MAX;
if(m2>test){
crap=floor(m1);
printf("%d\n", crap);
n=n+1;}
}while(n<N);
}
float f( float x, float mean, float sigma){
/* Normal distribution */
float y;
y=(x-mean)/sigma;
return( ((1.2533141/sigma))*exp(-y*y/2));
}

2.5

Experiment II. Hypothesis testing and goodness of fit

Below is a table of the number of counts for 600 five second intervals. This data was taken
in 1997.
The Geiger tube voltage has been set at 1040 volts, the scaler-timer counts for 5 seconds at
a time with 5 second intervals between counting periods, so that we have time to record our
numbers. The more counts you make, the better will be your results. The number
400 is a minimum, I suggest that you take twice that many. This requires counting for over
two hours. Do it in 30 minute shifts with your lab partners.

14

459 422 434 410 434 438 454 421 390 422 438 436 444 422 427
449 398 418 407 399 445 424 440 430 391 421 433 401 425 424
430 449 437 421 460 414 447 396 394 430 418 392 414 423 432
387 405 409 406 449 437 433 444 396 394 389 426 453 423 422
410 415 363 438 422 377 410 408 406 431 438 434 413 432 429
396 436 461 402 377 443 411 383 433 420 393 424 433 390 450
458 459 444 381 389 438 435 374 440 409 431 433 455 455 392
427 407 435 405 442 423 398 423 417 423 428 423 406 421 424
418 453 418 398 408 438 443 412 433 440 417 425 427 403 414
398 401 408 401 408 454 390 437 410 416 408 446 409 425 396
388 423 441 394 414 420 424 396 417 405 429 395 381 438 403
445 393 424 431 419 404 398 423 410 445 430 402 428 407 453
429 405 389 482 369 406 417 426 402 393 456 428 402 407 406
435 433 435 417 450 426 445 443 446 424 393 433 438 424 405
409 428 410 400 408 412 428 435 431 395 419 431 430 428 461
434 433 406 404 403 404 420 425 417 433 392 391 439 431 419
417 405 398 435 407 446 437 458 402 415 432 408 383 374 418
400 387 410 360 385 445 427 439 423 465 434 413 435 416 430
442 437 418 447 456 426 413 396 435 430 433 360 434 428 454
396 408 413 406 406 391 431 437 450 382 413 422 406 437 393
438 387 430 425 415 429 430 393 381 437 479 409 396 412 428
426 423 362 424 411 432 399 432 397 422 424 410 449 415 478
420 416 421 417 417 408 466 436 451 412 433 407 419 460 394
427 415 432 429 410 449 421 381 423 435 402 419 432 453 416
447 404 386 434 445 399 393 461 399 417 434 432 425 429 445
450 402 412 407 414 423 420 404 405 425 385 448 452 435 418
419 436 425 475 418 426 378 469 429 438 453 432 418 368 400
419 439 423 356 428 382 415 422 402 437 444 417 364 448 443
383 435 411 415 415 441 393 394 424 371 399 446 407 433 412
373 437 416 410 432 432 409 468 399 407 425 404 434 399 401
402 401 446 393 388 445 459 369 409 429 441 390 439 436 438
408 449 430 417 447 391 404 412 414 409 424 439 412 394 419
428 408 426 401 442 428 412 425 408 425 434 423 400 444 388
406 376 447 429 427 409 372 416 404 439 419 455 401 411 389
399 431 435 410 393 435 407 439 398 433 454 443 431 426 397
405 409 436 432 414 434 472 458 399 408 396 397 465 423 442
446 421 447 423 452 440 420 412 404 459 438 432 415 422 402
414 398 442 390 393 442 436 392 407 437 429 412 419 428 448
426 386 397 381 449 418 437 384 429 389 418 434 455 452 380
418 433 405 400 420 421 418 420 417 396 423 420 435 414 429
We use this data to test a hypothesis; nuclear decays are random processes, with a
frequency that is a Poisson random variable. We could use a Normal PDF as well.
To analyze this data, I ran it through binsort in the support software, which sorts the data
into 20 equivalence classes and predicts the correct Poissonian parameter (using the unbiased
15

estimator). Below we have the output.


Best-fit Poisson parameter is = 10.148333 Total count = 600, number of bins (classes) = 20
Bin j
0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19

lower bound
355.5
361.5
367.5
373.5
379.5
385.5
391.5
397.5
403.5
409.5
415.5
421.5
427.5
433.5
439.5
445.5
451.5
457.5
463.5
469.5

upper bound
361.5
367.5
373.5
379.5
385.5
391.5
397.5
403.5
409.5
415.5
421.5
427.5
433.5
439.5
445.5
451.5
457.5
463.5
469.5
475.5

frequency mj
3
3
6
6
14
22
37
42
62
49
59
64
71
64
31
27
18
12
5
2

theoretical freq. mj,th


0.023485
0.238331
1.209332
4.090901
10.378958
21.065824
35.630501
51.655743
65.527462
73.888281
74.984291
69.178689
58.504033
45.670648
33.105783
22.397901
14.206336
8.480625
4.781345
2.553825

(mj mj,th )2
mj,th

377.250773
32.000915
18.977833
0.890918
1.263320
0.041427
0.052638
1.804898
0.189890
8.383285
3.407348
0.387675
2.669033
7.356260
0.133944
0.945594
1.013061
1.460505
0.009999
0.120103

To fit to a Normal curve, use the estimators for mean and standard deviation of the last
section.
Notice that when mtheor is small, such as for bins far from the most populated, the contribution to the 2 statistic is huge. This is a well known phenomenon; the relative error in an
essentially null measurement can be huge. Null measurements are to be avoided.
P
(mj mj,th )2
= 458.359406
2 = 19
j=0
mj,th
2

( 458.359406) =

458.359406

16

18
2

( 18
)
2

18

x 2 1 e 2 dx = 0.000000

71

63.9

56.8

49.7

42.6

35.5

28.4

21.3

14.2

7.1

0
0

10 11 12 13 14 15 16 17 18 19 20 21

Notice that the first three data bins really throw off the whole goodness of fit criterion, and
just for a few counts. If we discard this data, the result is the reduction of 2 to 2 = 30.1299,
and we find then that
(2 30.1299) =

30.1299

18
2

2 ( 18
)
2

18

x 2 1 e 2 dx = 0.036486

If instead we sort into 15 equivalence classes, we see a better result


Best-fit Poisson parameter is = 7.488333 Total count = 600, number of bins (classes) =
15

17

Bin j
0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
2 =

lower bound
355.5
363.5
371.5
379.5
387.5
395.5
403.5
411.5
419.5
427.5
435.5
443.5
451.5
459.5
467.5
P14

j=0

(mj mj,th )2
mj,th

upper bound
363.5
371.5
379.5
387.5
395.5
403.5
411.5
419.5
427.5
435.5
443.5
451.5
459.5
467.5
475.5

frequency mj
5
5
8
19
40
56
77
76
81
97
56
40
25
8
4

theoretical freq. mj,th


0.335745
2.514170
9.413470
23.497067
43.988468
65.880062
82.221978
87.957939
82.332296
68.503520
51.297719
34.921311
21.791868
12.552675
6.714187

= 88.394485

(2 88.394485) =

88.394485

13

13
2

2 ( 13
)
2

x 2 1 e 2 dx = 0.000000

97

87.3

77.6

67.9

58.2

48.5

38.8

29.1

19.4

9.7

0
0

18

10

11

12

13

14

15

16

(mj mj,th )2
mj,th

64.797043
2.457811
0.212238
0.860687
0.361637
1.481717
0.331652
1.625690
0.021559
11.854127
0.431042
0.738606
0.472291
1.651190
1.097201

In this case discarding the first two bins reduces 2 to 2 = 21.13964, and using the chisquared table generator in the support software, this gives
(2 21.1396) =

21.1396

13
2

( 13
)
2

13

x 2 1 e 2 dx = 0.070014

We read this as saying that if the deviations from the best fit Poissonian of less than 7% are
accepted to be insignificance, we have a good fit, and our hypothesis is verified to be true;
nuclear decays are random processes.
It should be clear that this experiment is sensitive to the size of the data set, the larger, the
better. The analysis should be performed by regarding the contributions of data
classes far from the most populated classes in the proper light. These classes will
contribute most to 2 but will account for the smallest percentage of the actual data.

19

The Half-life of Ba137

The purpose here is to gain more experience with statistics and to learn how to use the
Geiger counter.

3.1

Background

The Cesium sample will be used to determine the operating range (voltage) for the GeigerMuller tube. The scaler-timer simply records decays produced by ionizing radiation emitted
by the Barium and detected by the Geiger-Muller tube.
Radioactive Barium such as that used in this experiment decays by gamma ray emission,
each decay releasing a gamma ray of 0.662 M eV
137
m
56 Ba

137
56 Ba +

in which m denotes a meta-stable isotope of the element, and the right hand side contains
the stable isotope.
Other types of nuclear decay modes are decay
p
mX

4
p4
m2 Y +2 He

in which an particle or Helium nucleus is emitted, electron capture


p+ + e n0 + e
in which a neutrino is produced with a neutron and decay
p+ n0 + e + + e
a decay mode of the proton in an unstable nucleus producing a positron and neutrino, and
its inverse process
n0 p+ + e + e
the common decay mode of the neutron.
A typical decay chain for radioactive isotopes, in particular the one leading to the formation
of Radon gas, is illustrated below.

20

We hypothesize that the probability of decay for a particle within time dt is


d = dt
the rate at which the number of particles in a sample of size N is changing due to decay is
then
dN = N dt
which integrates to an exponential decay law

N (t) = N0 et
for the number of particles left after time t , beginning with N0 . The time it takes for the
sample to be cut in half by radioactive decays is
N0
T 1
= N0 e 2
2
21

or

ln 2
2

We can compute the mean lifetime by finding the probability that a particle has not
decayed by time t. Since N (t) = N0 et is the number remaining at time t, the fraction of
particles that have not decayed is
T1 =

f (t) =
and so the mean lifetime is

N (t)
= et
N0

0
which is what we will compute in the experiment. During the experiment we will not be
able to directly measure the number of remaining nuclei, but rather the number of decays,
N0 N (t) which also drops with the same exponential rate. Let N be the number of
decays detected between t and t + t, then
=

tet dt =

N = N (t) N (t + t) = N0 (1 et )et = N0 et
the quantity N0 is fixed by the size of the interval t and so we see that the actual number
of decays detected will also exponentially drop as the number of particles capable of decay
drops.
The standard measure of radioactivity of a sample is called its activity which is the number
of decays per second , one decay per second being called a Bq or Becqueral. We use the
Ci or curie instead
1 Ci = 3.7 1010 Bq
The activity R(t) at a given time is
R(t) =

dN
= N0 et = R0 et
dt

Since the activity is actually the decay rate, we see that our experiment will actually measure
this quantity, since we count for ten second periods, spaced by twenty second intervals. The
count rate itself exponentially decays along with N .
Our Barium sample has an activity of 9 Ci. To put this into perspective, the GeigerMarsden experiment was conducted with a 0.1 Ci radium sample!

3.2
1.
2.
3.
4.

Apparatus

Geiger tube. Sargent-Welch, catalog number S-72095-85.


Scaler-timer. Sargent-Welch catalog number S-72095-10.
Cesium Sample. Amersham-Searle, 0.1 c.
Barium sample . Redco mini-generator, miniature radio-isotope generator, 9c.

22

3.3

Procedure

The first step is to collect a Barium sample from the mini radio-isotope generator. An acid
solution is dripped into the generator which washes the surface of a Cesium/Barium sample,
and the residue collects on a small disc of blotter paper placed below the generator. Barium
is produced from the Cesium by a beta decay
137
55 Cs

136
e
56 Ba + e +

The Geiger counter operating voltage is now determined by counting the other radioactive
source, Cesium. The Cesium isotope used has a 27 year half-life and so provides a constant
rate of decays for the counter.The Barium half-life is 2.6 min.
The Barium sample is placed under the counter and the experiment begins. We will count
for ten second intervals separated by twenty second periods. In other words, count for ten,
rest for twenty, count for ten, rest for twenty and so on.

3.4

Data and analysis

Time
0
30
60
90
120
150
180
210
240
270
300
330
360
390
420
450
480
510
540
570
600

Cnts per 10 sec


139
103
87
72
60
66
41
37
41
26
22
18
19
14
11
13
11
10
6
8
5

We now need to correct for background radiation. Wait for one hour, and count the
sample for 600 seconds. We find a total of 198 counts, or 3.3 counts per 10 second interval
23

can be attributed to background radiation. We now correct the data for this and work it up
using the awk script
#analyze.awk
{
print $1,"&",$2,"&",$2-3.3,"&",sqrt($2-3.3),"&",log($2-3.3),"\\\","\\hline"}
This results in the table below
Time
0
30
60
90
120
150
180
210
240
270
300
330
360
390
420
450
480
510
540
570
600

Cnts per 10 sec


139
103
87
72
60
66
41
37
41
26
22
18
19
14
11
13
11
10
6
8
5

Cnts
135.7
99.7
83.7
68.7
56.7
62.7
37.7
33.7
37.7
22.7
18.7
14.7
15.7
10.7
7.7
9.7
7.7
6.7
2.7
4.7
1.7

Cnts0
ln |Cnts0 |
11.649
4.91045
9.98499 4.60217
9.14877 4.42724
8.28855 4.22975
7.52994 4.03777
7.91833 4.13836
6.14003 3.62966
5.80517
3.5175
6.14003 3.62966
4.76445 3.12236
4.32435 2.92852
3.83406 2.68785
3.96232 2.75366
3.27109
2.3702
2.77489 2.04122
3.11448 2.27213
2.77489 2.04122
2.58844 1.90211
1.64317 0.993252
2.16795 1.54756
1.30384 0.530628

with
Cnts0 = Cnts Cntsback

is the corrected number of counts per 10 second interval,

Cnts0 = Cnts0

is the standard deviation in the number of counts. We can compute the error in the log of
the number of counts
ln Cnts0 =

Cnts0
1
Cnts0
=
=
0
0
Cnts
Cnts
Cnts0

and produce a table of file of data suitable for error-bar graphing with the awk script
24

#analyse2.awk
{
print $1," ",log($2-3.3), " ", 1.0/sqrt($2-3.3)}
Time
0
30
60
90
120
150
180
210
240
270
300
330
360
390
420
450
480
510
540
570
600

ln |Cnts0 |
4.91045
4.60217
4.42724
4.22975
4.03777
4.13836
3.62966
3.5175
3.62966
3.12236
2.92852
2.68785
2.75366
2.3702
2.04122
2.27213
2.04122
1.90211
0.993252
1.54756
0.530628

ln |Cnts0 |
0.085844
0.10015
0.109304
0.120648
0.132803
0.126289
0.162866
0.17226
0.162866
0.209888
0.231249
0.26082
0.252377
0.30571
0.360375
0.321081
0.360375
0.386334
0.608581
0.461266
0.766965

The graph of which is below, created by lg in the support software. You could use gnuplot which has error bar graphing.

25

4.6

4.2

3.8

3.4

2.6

2.2

1.8

1.4

1
0

51

102

153

204

255

306

357

408

459

510

With log of corrected counts of the vertical axis and time in seconds on the horizontal
axis. We can run a linear regression on the data , but first we will discard the three data
points indicated in the table with an asterisk. These arepoints with count totals so close to
the background level that they are most probably overwhelmed by the fluctuations in the
background count rate. It is questionable scientific procedure to discard data, unless
there is very good reason. The linear regression program in the appendix gives us a slope
for these 17 points of
y = 4.811657 + 0.005963x,

2 = 0.114469,

2
m
= 0.000002

which results in
m = 0.005963 0.001414,

1 =
2

ln 2
= 116 s
0.005963 1s

which places the half-life between the limits


ln 2
ln 2
m

(1 +
+ ) = 152.28 sec
|m + m |
|m|
|m|
and

ln 2
ln 2
m

(1
+ ) = 93.95 sec
|m m |
|m|
|m|
26

which is in fair agreement with the accepted 2.6 min = 156 sec value. When we perform this
experiment we should be sure that enough isotope is collected that we have at least 600
or more counts initially. In this way we will have enough counts at subsequent times that
background fluctuations do not overwhelm our counts.

3.5

Background material; the Nuclear Shell Model

The actual potential felt by nucleons in a nucleus is complex and the Schroedinger equation
is impossible to solve analytically. However for deeply bound nucleons the potential looks a
lot like a three dimensional oscillator and the nuclear shell model is based on this potential.
It is very useful for predicting stable nuclei configurations.
Begin with the nonrelativistic, three dimensional oscillator in polar coordinates. Since the
potential is spherically symmetric we find
~ 2 ] = [H, Lz ] = 0
[H, L
and so we decompose the wavefunction into a radial part and a spherical harmonic
= R(r) Y`,m (, )
and the radial equation becomes
2m
h
2 `(` + 1) m 2 r 2
d2
2 d
+ 2 (E
))R(r) = 0
( 2 =

dr
r dr
2mr 2
2
h

Perform the variable change


=

mr 2
h
2

d
m q d

=2
dr
d
h
2
and we find
d2

`(` + 1)
3 d
+(

))R() = 0
+
2
d
2 dr
2
h
4
4
We now impose the boundary conditions on the wavefunction
(

R(0) = 0
R() = 0
by redefining the wavefunction
l

R() = 2 e 2 u()
and we find that
27

d2
3
d
E
3 `
+
(

)
+
(

))u = 0
d 2
2
d
2
h 4 2

and again we recognize the same confluent hypergeometric equation


x f 00 (a; c, x) + (c x)f 0 (a; c, x) af (a; c, x) = 0
that has appeared in every potential problem that we have studied in Physics 441. We know
that a series solution exists
f (a; c, x) =

(a)j xj
,
j=0 (c)j j!

(a)j = a(a + 1) (a + j 1)

which will truncate at a polynomial (satisfying the condition at ) if


E
3 `
=N
2
h 4 2
where N = 0, 1, 2, 3, . The spectrum that we obtain is
a =

3
E =h
(2N + ` + )
2
Each ` value has 2`+1 degenerate m` values and a little arithmetic shows that the degeneracy
of a level of energy
3
E =h
(M + )
2
with M = 0, 1, 2, is
(M + 1)(M + 2)
2
For a fixed value of M , ` can be M , M 2, M 4, , however ` must remain positive. The
plot of energy versus angular momentum forms straight lines called Regge trajectories.
D(M ) =

You will often see this written in the literature in terms of the oscillator radial quantum
number n as
E
` 1
n=
+
2
h 2 4
with solutions that can be written in terms of the Laguerre polynomial, which is a hypergeometric function
1 2

`+ 1

n,`,m (r, , ) = An,` e 2 Ln12 (2 ) Y`,m (, )


with
r
= ,
r0

r0 =

28

,
m

= 2

and the Laguerre polynomial satisfies the recursion


Lm
k (x) =

1 x m dk k+m x
e x
(x
e )
k!
dxk

These functions are orthonormal on [0, ] with respect to kernel xm ex ;


Z

m
xm ex Lm
k (x) Lk 0 (x) dx =

(m + k + 1)
k,k0
k!

which will allow one to show that the normalization factor for the wavefunctions is
An,` =

v
u
u
t

2n+`+1 (n 1)!

(2n + 2` 1)!! r03

We now try to find the oscillator frequency that best fits the experimental data as far as
the size of the nucleus and the binding energy of a nucleon in the nucleus is concerned. To
do this we compute
Z
n,`,m r 2 n,`,m d3 x = hr 2 i
and fit this to the experimental radius-squared of a nucleus containing a total of A nucleons.
This computation is a good excerise in quantum mechanics and the various recursion relations
satisfied by the Laguerre polynomials, and results in
1
hr 2 i = (2n + ` ) r02
2
and from the rather extensive experimental data, we have the result that the nucleus containing A nucleons subjects its nucleons to a potential pretty well described by a harmonic
oscillator with
q
34.7 M eV
1
h
= 1
,
rr = 1.19 A 3 0.44 f ermi
A 3 0.44
This is pretty useful for computing ray energies for nuclei that decay through emission.
Analogous to noble gasses with filled atomic shells, in the nuclear shell model we assume
that nuclei are stablest when they have nucleons filling each shell . Each state
|N, `, m` >
can contain two protons and two neutrons, one spin up,one spin down of each. This gives
rise to stable nuclei of atomic number (Z)
2=Z
2+6= 8=Z
2 + 6 + 2 + 10 = 20 = Z
29

2 + 6 + 2 + 10 + 6 + 14 = 40 = Z
These are called Magic Number nuclei. Indeed the Z = 2, 8, 20 nuclei are known to be
very stable. After that the nucleon spin-orbit coupling disrupts the energy levels and breaks
the degeneracy, and alters the magic number pattern. We define the term symbol of a single
nucleon in a way very similar to that of an electron in an atom, with nomenclature
(N + 1) Lj = n Lj
and the nuclei of various elements are constructed by a nuclear Aufbau process; we begin
adding nucleons to a shell model manifold of single particle states, until we reach Z protons
and A Z neutrons.
Even better predictions of the magic numbers A of stable nuclei come from including the
effects of the spin-orbit interaction beteen nucleons, altering the potential to
2 ~
1
~s
V (r) = V0 + m 2 r 2 2 L
2
h

The eigenstates of the full Hamiltonian become


1 2

`+ 1

n, `, j, mj (r, , ) = An,` e 2 Ln12 (2 ) Vj=` 1 ,`,mj (, )


2

in which the angular part of the wavefunction is an eigenstate maximal set of commuting
operators derivable from the total angular momentum operator
~ + ~s = L
~ + 1h
J~ = L
~
2
containing contributions from nucleon orbital and spin angular momenta.
You hopefully remember that that means
Jz Vj,`,mj = h
mj Vj,`,mj
and
from quantum theory. But
and so

~ 2 Vj,`,m = h
(J)
2 j(j + 1) Vj,`,mj
j
~ 2 = (L)
~ 2 + 2L
~ ~s + ~s2
(J)

~ ~s = j(j + 1) `(` + 1) s(s + 1)


L
2
and by the rules of angular momentum addition, since s(s + 1) = 21 ( 12 + 1),
1
j =`+ ,
2

or

j =`

1
2

We find then that the energy En,j,` of these states will be


1
h V0 `
En,`+ 1 ,` = (2n + ` )
2
2
30

and

1
h V0 + (` + 1)
En,` 1 ,` = (2n + ` )
2
2
This results in the single-nucleon states shown below, which is in very good agreement with
experimentally established stable nuclei-nuclei with filled shell-model shells.
The state of lowest energy in the shell model is
n Lj = 1s 1 ,
2

1
V0
E1, 1 ,0 = h
2
2

and the table below together with the formula for can be used to determine nuclear energy
level spacings, and the frequencies of rays emitted when a nucleon changes levels.
` State En,j,` E1, 1 ,0 Nj
2
0 1s 1
0
2
2
first magic filled shell; 2 particles
1 1 1p 3
h

4
2
1 1 1p 1
h
+ 2
2
2
second magic filled shell; 8 particles
1 2 1d 5
2
h 2
6
2
2 0 2s 1
2
h
2
2
1 2 1d 3
2
h + 3
4
2
third magic filled shell; 20 particles
1 3 1f 7
3
h 3
8
2
fourth magic filled shell; 28 particles
2 1 2p 3
3
h
4
2
2 1 2p 1
3
h + 2
2
2
1 3 1f 5
3
h + 4
6
2
1 4 1g 9
4
h 4
10
2
fifth magic filled shell; 50 particles
n
1

The superstable magic filled shells are well separated in energy from the next highest set of
(empty) one particle states.
Furthermore, experimental data shows that depends on `, but is small, for example (in
MeV);
n,` ; 1,2 = 1.15, 1,3 = 0.93, 1,4 = 0.76
2,1 = 0.68, 2,2 = 0.58, 3,1 = 0.19
and it continues to decrease for both increasing n and `.
An energy level diagram similar to that used to perform the Aufbau process for atomic states
is illustrated below.

31

Remember that we fill in neutrons and protons independently.


Example Apply the nuclear Aufbau process to build up the most stable Helium isotope.
Such an isotope will have a filled shell for both types of nucleons.

32

1d 5/2

1p 1/2
1p 3/2
1s 1/2
-1/2

1/2

-1/2

protons

1/2

neutrons

In order to use the shell model, which describes only single nucleon states, to describe real,
multinucleon nuclei, we use the Aufbau process to add nucleons to the shell model states
computed with the appropriate A, and further take into acount the fact that paired, like
nucleons have an additional pairing energy of
Epairing = 2

11.2 M eV

By pairing we mean; the two combine angular momenta to 0. Remember that two particles of angular momentum j can have two-particle wavefunctions of angular momentum 2j,
2j 1, 2j 2, , |j j| = 0. This pairing increases the binding energy of the resultant
state.
For a filled shell, this pairing amounts to the combination of all angular momenta of the
shells nucleons to j = 0. A filled shell has no contribution to the nuclear spin.
This means that we need to distinguish between three nuclear types;
even-p, even-n; or Z even, N = A Z even, so all protons are paired and all neutrons are
paired.
A-even, Z-odd; or Z odd, N = A Z odd, so all protons but one are paired and all
neutrons but one are paired. This is also called odd-odd.
odd-even nuclei; Z even, N odd or Z odd, N even.
The pairing is with respect to mj , so we expect pairing between |n, j, `, mj i and |n, j, `, mj i.
The shell model can be used to determine properties of the entire multinucleon nucleus,
which are going to be largely determined for odd-even nuclei by the last unpaired nucleon.
Predictions;
The spin of an even-even nucleus in the ground state should be zero. This conforms to
experimental evidence.
Due to cancellations by the pairing, the nuclear spin of odd-even nuclei is determined by
33

the wavefunction of the last nucleon, the ground state nuclear spin being the angular
momentum j of the last unpaired nucleon.
Several nucleons in an incomplete shell often do combine their angular momenta to produce
a different value J 6= j. Suppose that there are k 2j + 1 nucleons in a shell, they may
produce a state whose label is
(n(`)j )kJ
and a perfect example is the Sodium isotope used in several of our experiments
22
N e21
10 , N a11

Z
1
3
5
9
11
19
19
25
27
27

AZ
0
4
6
10
12
20
22
30
30
32

(1d 5 )33

are in state

Some odd-even nuclei


A Name J Last unpaired nucleon
1
1 H
1s 1
2
2
3
3
7 Li
1p
2
2
3
11 B
1p 3
2
2
1
19 F
2s 1
2
2
3
23 Na
1d 5
2
2
3
39 K
1d 3
2
2
3
41 K
1p 3
2
2
5
55 Mn
1f 7
2
2
7
57 Co
1f 7
2
2
7
59 Co
1f 7
2
2

Example O816 is doubly magic, and superstable.

1d 5/2

1p 1/2
-1/2

1/2

-1/2

1/2

1p 3/2
-3/2 -1/2

1/2

3/2

-3/2 -1/2

1/2

3/2

1s 1/2
-1/2

1/2

protons

-1/2

1/2

neutrons

O715 is doubly magic except for a proton hole in 1P 1 . Since in a filled shell all nuclear spins
2
are paired, this hole has spin one half, and therefore so does this nucleus. We define the j
value of a filled shell to be zero, a consequence of the nucleon urge to pair. This means that a
nucleus will have the quantum numbers of holes in filled shells. The N715 does experimentally
34

have j = 12 . The nucleus O817 is doubly magic except for an extra nucleon in the 1D 5 and so
2
the shell model predicts a nuclear spin j = 25 again in agreement with experiment. Nuclear
parity under inversion of the coordinate system is given by
P = (1)(
for example the nucleus

(nucleon ` values)

O816
has a filled s shell , each nucleon having ` = 0, and a filled p shell with each nucleon having
` = 1 and so has even parity, a fact verified experimentally. Nuclear spin and parity actually
plays a very imporant role in the spectra of diatomic molecules, and the parity of the wavefunction will determine which rotational modes exist for the molecule. This will be evident
even in the thermodynamic functions for gases of such molecules.
We can use the concept of pairing energy to understand decay processes of some of the
22
nuclei used in this course, such as N a22
11 . Consider a N a11 with an unpaired proton and
neutron. Remember that the neutrons are bound a little tighter than the protons due the
the raising of the proton energy caused by the Coulombic repulsion of like charged particles.
The total electrostatic potential energy due to a nuclear charge Z is
Ucoul

1
= 4
2

r0
0

Zr 2
1
r (
) dr + 4
3
40 r0
2
2

r0

r2 (

Z
3 Z2
2
)
dr
=
40 r 2
5 40 r0

in which r0 is the mean nuclear radius. Using the shell model value we see that this is
3
0.714 Z 2 M eV
Z2
1 =
1
5 40 (1.21 F ) A 3
A3

Ucoul =

which evaluates to 30.83 M eV for N a22


11 .
The most likely decay therefore, in order to achieve pairing, is for the unpaired proton to
inverse- decay

1d 5/2

1d 5/2
-5/2 -3/2 -1/2

-5/2 -3/2

1p 1/2

-5/2 -3/2 -1/2 1/2

1p 1/2
-1/2

1/2

-1/2

-1/2

1/2

1p 3/2

1/2

-1/2

1/2

1p 3/2
-3/2 -1/2

1/2 3/2

-3/2 -1/2 1/2

-3/2 -1/2

3/2

1s 1/2

1/2 3/2

-3/2 -1/2 1/2

1s 1/2
-1/2

1/2

protons

-1/2

1/2

-1/2

neutrons

1/2

protons

35

-1/2

1/2

neutrons

3/2

A proton first decays to a neutron, emitting a positron that quickly captures an electron from
the innermost shell (K shell) of the electron cloud, the resulting annihilation produces two
s at 0.511 M eV that are not shown in the figure. Recall that the total nuclear spin of the
Sodium N a22
11 was 3 before the decay. Remember that filled shells have nucleon spins paired
to zero, and in the valence shell there is one unpaired proton and one unpaired neutron with
jp = jn = 25 , and so the nuclear spin is one of the spins in the decomposition
5 5
=543210
2 2
Conservation of angular momentum requires that instantly after the decay, it be 31, in fact
it is experimentally found to be 2. The decay leaves a Neon nucleus with an even number
of protons and neutrons, which now quickly pair, leaving a spin-0 nucleus and shedding the
excess energy as another . This line of reasoning is probably qualitatively correct, but I
have not been able to compute the energy of the last -emission correctly from the shell
model.
A useful source of information is Nuclear Forces by Gernot Eder, MIT Press, 1968.

Determination of the Speed of Light

At one time the speed of light was thought to be infinite. A correct determination of the
speed of light is a crucial experiment in modern physics and can be performed with relatively
simple apparatus.
Transformer

120 V AC

Rotating
mirror
Photoelectric
tube

Fixed mirror

Beam splitter
Meter stick
scale

120 V AC

Frequency
meter

120 V AC

Power
supply

Laser

36

Apparatus
1. Power supply. Spectra Physics laser exciter model #249.
2. Laser. Spectra Physics Stabilite Helium-Neon laser model #120.
This is a powerful laser. Do Not Look Into The Beam!
3. Beam splitter. A glass slide.
4. Rotating mirror. Leybold 476-41.
5. Photoelectric tube.
6. Transformer. Variac type W5M73.
7. Flat mirror and stand.
8. Magnifier lens.
9. Frequency meter. Fluke 1900A multi-counter.
The rotating mirror is shown below.

The apparatus is set up with about 5 meters between focusing lens and laser. The distance
from the meter stick to the beam splitter is the same as from beam splitter to rotating
mirror. In this case the emitted beam from the laser and reflected beam from the splitter to
the meter stick scale have the same path length, and the returning beam reflected from the
splitter will focus to a point more or less on the scale. The Photoelectric tube must be set up
in such a position that each time the rotating mirror undergoes a single rotation, it reflects
the laser beam into the photo-tube at some point, and the number of pulses per second is
the inverse of the rotation frequency of the rotating mirror. This may require that the beam
sprayed into the tube be out of line with the beam reflected back into the beam splitter to
avoid interference of the two beams.
Consider the figure below, showing two beams arriving at the scale, one bounces off of
the rotating mirror and heads directly to the scale. The current position of the mirror is
position 2. At an earlier time, the mirror was at position 1, and the beam (very finely dotted
beam) from the laser was directed by the rotating mirror towards the distantsecond mirror.
Now it is returning and hits the mirror in its current position but does so at an angle such
that it is not directed back along its original path. These two beams arrive at the scale
at the same time, but were emitted by the laser at different times separated by the time it
takes the light to travel back and forth between rotating and fixed mirrors. by an angle.

37

position 1
position 2

S=distance between beam traces on scale


L=distance between fixed and rotating mirrors
D=distance from rotating mirror to laser.
d=distance from laser to beam splitter=distance from beam splitter to scale
= change in position of mirror
=rotation frequency of mirror in Hz
= natural rotation frequency of mirror.
We have the basic relations
= 2,

= dt,

and
tan 2 =
so that
c=

dt =

S
D

4 L D

Data
In this run we measured d five times and obtained
d = 53.1 0.03cm
and measured D five times finding
D = 309.9 0.5cm
38

2L
c

and finally L five times obtaining


L = 3111.0 3.0cm
When the separation S = 0.2 cm it was found that the rotating mirror had frequency 502 Hz
and so
m
4(3111 cm)(309.9 cm)(502 Hz)
= 3.041 108
c=
0.2 cm
s
8m
which is not far off of the accepted value of c = 2.9979 10 s .
Things to watch out for. Calculate correctly; the mirror is two-sided!

39

Determination of

Cp
Cv

for a gas

The specific heat ratio for a gas is known as the adiabatic constant and can be measured by
studying a typical adiabatic process such as sound wave propagation in the laboratory.

5.1

Background

Sound is a longitudinal compression wave in a material medium such as a gas or solid. We


will first derive the wave equation in order to find the velocity of sound and the relation
between the displacement and pressure waves.
Consider a layer of molecules of gas of cross sectional area A and thickness dx.
P(x+dx)

P(x)
A

x+dx

x+

x+dx+d +

Let passage of the wave through the gas cause the left wall to displace by an amount to
the right and the right wall of the layer displace by + d. This will cause the cell to expand
and so the air density and pressure inside of the cell will change. The old cell volume was
V0 = A dx
and the mass of the gas in the cell was and still is
dm = 0 A dx
in which 0 is the density of the gas at equilibrium. The new cell volume is
V0 + V = A(dx + d)
and so the volume change is
V = Ad
and we find that
40

d
V
=
V0
dx
The force causing these displacements is supplied by pressure gradients, the pressure on
the left wall of the cell, P (x), pushes to the right and that on the right wall
P
dx +
x
pushes to the left. We find that the net force to the right is
P (x + dx) = P (x) +

Fx = A(P (x) P (x + dx)) = Adx

P
x

Since the cell has acceleration


ax =

d2
dx2

the equation Fx = m ax reads


0 dxA

d2
P
= Adx
2
dx
x

or
d2
1 P
=

dx2
0 x
This is the acoustic equation. We need to establish the thermodynamics of the process,
after all we are dealing with the expansion of an ideal gas presumably. Sound propagation
is a fast process, and no heat is exchanged between adjacent gas cells and so propagation of
sound is adiabatic. In that case
P0 V0 = (P0 + P ) (V0 + V )
in which is the ratio of specific heats
=

Cp
Cv

We can expand the right hand side using the Binomial theorem
(1 + x)n = 1 + nx +
for small x, which in our case becomes
(V0 + V ) = V0 (1 +

V
)
V0

V
+ )
V0
and if we again neglect the product V P we find that
= V0 (1 +

41

P =

V
P0 = P0
V0
x

and the total gas pressure is

x
Insertion of this into the acoustic equation gives all the same results as before except now
we find that sound waves travel (adiabatically) at speed
P (x) = P0 + P = P0 P0

v=

P0
=
0

R T
M

where M is the molecular weight in grams per mole.


A full treatment of the specific heats CP and CV for polyatomic gasses is given at the end
of this experiment.

5.2
1.
2.
3.
4.
5.

Apparatus

Kundts tube apparatus.


High pressure gas cylinders of Ar, N2 , CO2 .
Oscilloscope.
Audio oscillator and speaker.
Thermometer.

This apparatus is essentially canned and lives in the physical chemistry laboratory.
Thermometer

Gas Out
Gas In
Speaker

Kundts tube
Moveable Piston
Rod
Microphone

Meter stick or scale

Audio
Oscillator

5.3

Oscilloscope

Procedure

An oscilloscope will be used to monitor the phase relationship between the input sound signal
into the gas in the tube and the signal picked up by a microphone. Such a phase relation
42

can be displayed as a Lissojous figure by displaying one signal on the horizontal input, and
the other on the vertical input to the scope. Typical Lissojous figures for example with
y 0 (t) = Ay sin(y t )

y(t) = Ax sin(t),
look like the following

This will require that the apparatus be calibrated by holding a tuning fork near the microphone with the speaker disconnected from the scope input. The apparatus needs to be
calibrated at several frequencies around 1 to 2 kilohertz. The program used to create these
figures is included at the end of this experiment.
The piston should be positioned near the end of the tube as seen in the figure, and the
tube flushed for about 10 minutes with the gas to be studied. Before data is ready to be
collected, reduce the flow rate to a low value, but not zero, to prevent air from diffusing into
the tube, which is not sealed. Adjust the gain on the scope until both input and microphone
signals are at the same magnitude.
Set the audio oscillator on the desired frequency, 1 kHz for Nitrogen and Carbon dioxide,
2 kHz for He or Ar, and slowly move the piston towards the speaker end of the tube. When
a half wave is standing in the tube the waveform being input will be
y(x, t) = A sin

2x

while the microphone receives


y 0 (x, t) = A0 sin

2(x + 2 )
2x
= A0 sin

eliminating x we should expect the curve


A 0
y
A0
or a straight line to be displayed on the scope. These runs need to be repeated and statistical
analysis performed.
y=

43

5.4

Data and analysis

This data was taken by a student in spring 1997.

Trial
1
2
3
4
5
6
Ave.

in cm
Air
CO2
Ar
17.40
13.65
16.10
17.25
13.45
16.10
17.20
13.55
16.20
17.50
13.60
16.05
17.10
13.55
16.10
17.35
13.45
16.05
17.30 0.15 13.54 0.08 16.10 0.05

All measurements were taken at temperature


T = 25.0C = 298K
and at frequency
= 1000.0Hz
Using the ideal gas constant
R = 8.31

J
moleo K

we obtain
Quantity

, cm
2
v, ms
gm
M, mole
p
=C
Cv

Air
CO2
Ar
17.30 0.15 13.54 0.08 16.10 0.05
346.0 3.0
270.8 1.6
322.0 1.0
28.97
44.01
39.948
1.400 0.024 1.303 0.015 1.673 0.01
44

The exact values are as follows. Diatomic molecules have one vibrational, two rotational
and three translational degrees of freedom., monatomics have only three translational, and
Carbon dioxide has vibrational modes, you should read the next section thoroughly
in order to see where the theoretical values come from.
Quantity
exp
theory
% difference

5.5

Air
CO2
Ar
1.400 0.024 1.303 0.015 1.673 0.01
1.400
1.285
1.667
0.0%
1.4%
0.38%

Molecular spectra and the equipartition theorem

) and CV = ( U
) in a quantum mechanically correct
The precise form of both CP = ( U
T P
T V
determination are both temperature dependent, both look like steps or plateaus, indicating that certain molecular motion modes are excited at various temperature regimes. What
modes are we talking about? There are three types of molecular motions; translations and
rotations (purely kinetic), and vibrations (possessing both kinetic and potential energy).
We can compute the specific heat of a molecule with classical formalism for a simple illustrative case such as the linear molecule CO2 . Consider then the lagrangean for the molecule
fixing all atoms to lie on a line;
1
1
1
1
1
L = mC z 2 + mO (z x 1 )2 + mC (z + x 2 )2 k(x1 `)2 k(x2 `)2
2
2
2
2
2
in which z fixes the position of the Carbon in space, and xi is the distance between each
Oxygen and the Carbon. The equations of motion are
d L
L
( )=
dt z
z
or
mC z + 2mO z mO x1 + mO x2 = 0
d L
L
(
)=
dt x 1
x1
or
mO x1 mO z = kx1

and finally

L
d L
(
)=
dt x 2
x2

or
mO x1 + mO z = kx2

Following the usual T and V matrix methods for normal modes, we can write this as

mC + 2mO mO mO
z
0 0
0
z

mO
mO
0 x1 = 0 k 0 x1

mO
0
mO
x2
0 0 k
x2
45

We assume that each coordinate is an amplitude times eit , which results in a secular equation
for the amplitudes and frequencies
(mC + 2mO ) 2
mO 2
mO 2
Az

mO 2
mO 2 + k
0
A x1 = 0

mO 2
0
mO 2 + k
A x2

The frequencies are determined by the vanishing of the determinant


0 = (mC + 2mO ) 2 (mO 2 + k)2 (mO 2 + k)(mO 2 )2
This gives us three frequencies;
02 = 0,

k
,
mO

12 =

22 =

(mC + 2mO ) k
mC + m O mO

In terms of these normal modes, we could perform a coordinate transformation

z
O1,1 O1,2 O1,3
0

1 = O2,1 O2,2 O2,3 x1


x2
O3,1 O3,2 O3,3
2
in terms of which the classical lagrangian would be
1
1
1
1
1
L0 = 02 + 12 12 12 + 22 22 22
2
2
2
2
2
The classical Hamiltonian will be
1
1
1
1
1
H0 = p20 + p21 + 12 12 + p22 + 22 22
2
2
2
2
2
which results in the classical partition function
Z=

dp0 dp1 dp2 d0 d1 d2 H0


e
,
h3

1
kT

It is a good exercise to compute this function by doing the integrals, each is from to
, to compute the average energy of such a molecule in contact with a heat reservoir at
temperature T
ln Z
5
U =
= kT

2
and the specific heat at constant volume
CV = (

5
U
)V = k
T
2

In a one dimensional world, we find that our gas of such molecules will have
=

CV + k
CP
=
= 1.4
CV
CV
46

The Equipartition theorem is essentially just a rule of thumb gotten by inspecting the
Hamiltonian written in normal mode coordinates; CV gets a contribution of 12 k from
each quadratic term in H0 . In this case there are five terms.
In three dimensions, we will need three coordinates (x, y, z) to specify the position of the
Carbon in space, the two bond distances, and two Euler angles and to specify the
orientation of the molecule in space. In that case the classical Hamiltonian will be (in
normal mode coordinates)
P2
1 ~2
1
1 2 1 2 2 1 2 1 2 2
2
H = P + (P +
2 ) + p 1 + 1 1 + p 2 + 2 2
2
2I
2
2
2
2
sin
0

Counting quadratic degrees of freedom, we find nine, and so CV = 92 k, and CP = CV + k


making the adiabatic constant 1.2222.
In the real world we find that stiff molecular vibrational modes are generally not excited
unless the temperature is quite high. This can only be seen from a quantum mechanical
treatment of the problem. If we toss out the four vibrational degrees of freedom, we are left
with five excitable molecular modes at low (room) temperature. This gives us = CCVP = 75 =
1.4 for CO2 . The full quantum treatment of only the vibrational modes of this molecule
will give us an average energy
h
1
h
2
= 5 kT +
U
h
1 +
h
2
2
2 sinh 2kT
2 sinh 2kT
since the quantum Hamiltonian will be
P2
1
1
1
1
H0 = P~ 2 + (P2 + 2 ) + h
1 (n1 + ) + h
2 (n2 + )
2
2I
2
2
sin
and the partition function for the quantized oscillation terms will be

ehi (ni + 2 )

ni =0

(sums rather than integrals). The rotational modes also will have a temperature dependence,
an issue that we will not go into here. The corresponding specific heat will be
h
1
h
2
5
h
2 22 cosh 2kT
h
2 12 cosh 2kT
CV = k +
h
1 +
h
2
2
4kT 2 sinh2 2kT
4kT 2 sinh2 2kT

This has the adversised plateau structure, here plotted for 2 = 6.0 1 ;

47

CV/k

10

2kT/h
You can see that the specific heat is roughly constant at some multiple of k, as classically
predicted, but only in certain temperature regimes. This is a clear indication that at sufficiently low temperatures, the oscillatory modes are not excited and cannot absorb heat.
It turns out that CO2 has an additional oscillatory mode, a flexing out of the linear geometry, that has a much lower frequency (because of weal restoring forces) than either
bond-stretching mode. This comes out of a group theory treatment of molecular vibrations.
If we work at a low enough temperature that all three translational modes, both rotational,
and the flexing mode (Potential and kinetic degrees) are excited, the classical prediction for
CO2 will be CV = 27 k, and the adiabatic constant is then = 97 = 1.2857, which is reported
as the theoretical value in the writeup. You should verify that for Ar, a monatomic, CV = 32 k
and = 53 = 1.666, and for N2 or O2 , if we neglect the vibrational modes of the extremely
rigid double bonds, CV = 25 k and = 57 = 1.4.

48

5.6

Safe Handling of High Pressure Gas Cylinders


High pressure
gauge
Low pressure
gauge
Shutoff
valve

Regulator
Outlet valve

Regulator
handle
Gas hose

Gas
Cylinder

Locate the shutoff valve on top of the cylinder, the regulator handle and outlet valve. Be sure
the outlet valve is closed firmly but never forced. Set the regulator to zero flow by turning
the regulator handle counter-clockwise until you feel resistance diminish. If it is already set
to zero flow it will turn freely. Open the shutoff valve. The high pressure gauge will now
read tank pressure.
Slowly turn the regular handle clockwise, screwing it in, until the low pressure gauge rises
slightly above zero. This amount will be adequate for the experiment.
Now turn the outlet valve. Even if the low pressure gauge reading drops there is still gas flow.
When you are done, first turn off the outlet valve (close it), set the regulator to zero, then
close the shutoff valve. Open the outlet valve,then the regulator slightly, allowing gas trapped
in the regulator to escape. The high pressure gauge should drop to zero. Once again close
the regulator and outlet valve.
Gas cylinders should always be secured to the workbench, and never use tools to force open
or close a valve.

5.7

Lissojous program

/* Draw lissojou figures for illustration


*/
/* gcc lissojou.c -L/usr/X11R6/lib -lplot -lXaw -lXmu -lXt -lSM
49

-lICE -lXext -lX11 -lm

#include<stdio.h>
#include<strings.h>
#include<stdlib.h>
#include<math.h>
#include<plot.h>
#define N 1000
#define PI 3.1415927
float X[N],Y[N]; /* f[n] is a sine function for cuts */
float dt,t,Ax,Ay,Wx,Wy,phi;
int n,m,k,l;
int pl_handle;
char buffer[6],*labels,cmd[3],*cmdptr;
main (int argc, char *argv[])
{
char Axl[14]="A\\sbx\\eb=";
char Ayl[14]="A\\sby\\eb=";
char Wxl[14]="\\*w\\sbx\\eb=";
char Wyl[14]="\\*w\\sby\\eb=";
char phil[6]="\\*f=";
dt = 2.0*PI/(1000.0);
if(argc !=7){
printf("./lissojou Amp_x omega_x Amp_y omega_y phase display\n");
exit(0);}
Ax=atof(argv[1]);
Wx=atof(argv[2]);
Ay=atof(argv[3]);
Wy=atof(argv[4]);
phi=atof(argv[5]);
cmdptr=argv[6];

pl_handle = pl_newpl (cmdptr, stdin, stdout, stderr);


pl_selectpl (pl_handle);
if (pl_openpl () < 0)
/* open Plotter */
{
fprintf (stderr, "Couldnt open Plotter\n");
return 1;
}
pl_fspace (-4.0, -4.0, 4.0, 4.0);
pl_flinewidth (0.005);
pl_erase();
pl_filltype(0);
pl_pencolorname ("black");
50

pl_fline(0.0, -3.8,0.0,3.8);
pl_fline(-3.8,0.0,3.8,0.0);
for(n=0;n<1000;n++){
t=(float)n*dt;
X[n]=Ax*sin(Wx*t);
Y[n]=Ay*sin(Wy*t-phi);
}
for(n=0;n<999;n++)
pl_fline(X[n+1], Y[n+1],X[n],Y[n]);
pl_fline(X[999], Y[999],X[0],Y[0]);
/* labels */
pl_ffontname("HersheySerif");
pl_ffontsize(0.25);
pl_fmove(3.4,-0.25);
pl_label("y(t)");
pl_fmove(0.15,3.7);
pl_label("y(t)");
pl_fmove(-3.5, -3.5);
labels=gcvt(Ax,4,buffer);
strcat(Axl,labels);
pl_label(Axl);
pl_fmove(-3.5, -3.8);
labels=gcvt(Ay,4,buffer);
strcat(Ayl,labels);
pl_label(Ayl);
pl_fmove(-2.0, -3.5);
labels=gcvt(Wx,4,buffer);
strcat(Wxl,labels);
pl_label(Wxl);
pl_fmove(-2.0, -3.8);
labels=gcvt(Wy,4,buffer);
strcat(Wyl,labels);
pl_label(Wyl);
pl_fmove(0.15, -3.5);
labels=gcvt(phi,4,buffer);
strcat(phil,labels);
pl_label(phil);

pl_closepl ();
pl_selectpl (0);
51

pl_deletepl (pl_handle);
}

52

Electron Diffraction and Bragg Scattering

The purpose is to not only measure typical inter-atomic spacings between atoms in a typical
crystalline solid (graphite) but to study as well the wave properties of electrons. The crystal
lattice will form a diffraction grating from which electronic waves will scatter.

6.1

Apparatus

1. Power supply A; a Teltron limited kV power supply, model 813. This consists of both a
0 5000 volt DC and a 6.3 volt AC supply.
2. Power supply B; a Hewlett-Packard 0 60 volt DC power supply, model 6218A.
3. A Keithly 175 Auto-ranging multi-meter, used as an ammeter.
4. The diffraction tube (Tel-Atomic)
The 6.3 volt AC supply will provide current to a filament that heats the cathode of the
diffraction tube. Once the cathode temperature is high enough, it will eject electrons. These
electrons will be controlled the 60 volt DC power supply, connected to the cathode. This is
done in such a way that electrons with insufficient energy to escape this voltage difference
are returned to the cathode. Such a measure controls the number of electrons that will form
the beam; too many in the beam will burn a hole in the graphite target. The total current
must be kept below 100A to prevent damage to the target.
The electrons will be accelerated by the 0 5000 volt DC power supply, connected between
the cathod can and anode, the actual voltage value determines the electron momentum ac1 2
cording to eV = mc2 ( 1) 2m
p

The graphite target is powdered graphite bonded to a nickel mesh. The layer is extremely
thin. Powder is used so that many different scattering planes are presented to the beam,
resulting in a circular ring of scattered electrons striking a fluorescent screen in the tube.

53

6.3 VAC
Cathode can
Cathode

Graphite target
Electron
beam

0-60
V DC

Anode
I

0-5000 V
DC

6.2

Ammeter

Background

Graphite is a carbon allotrope consisting of very large, parallel planes of benzene rings, with
an interplanar separation of about 3.5
A, with inter-atomic spacings d1 . This distance is
about 1.40
A, which is intermediate between a single carbon-carbon bond of 1.54
A and a

double bond of 1.33 A.

3.5 A

that perfectly superimpose, separated by planes


There are planes separated by about 7A
halfway in between. These half-way planes have Carbon atoms in locations that superimpose
54

on the centers of the hexagons in the planes above and below. The stereogram below may
help you see this.

There is an upper (green) plane, a lower (blue), and an interleaving (red) plane in the figure,
each separated by 3.55
A with intercarbon spacings in each plane of 1.4
A to scale. See Solid
State Physics, by Ashcroft and Mermin, p. 304 for details.
The spacings between atoms do not serve aperatures through which the electrons diffract;
the spacing is too big for that. What we see instead is reflection of the electron beam from
consecutive planes of atoms. The angle of incidence of the beam with respect to the plane
normals must be just right in order that beams reflected from consecutive planes interfere
constructively by time they reach the detector (glass shell of the diffraction tube). This
condition is called the Bragg condition From the figure we can see that when the incoming
beam arrives at the first plane of atoms, some of the beam is reflected, some transmitted to
reflect from the next plane. It is these two reflected beams that interfere. The optical path
difference between the two beams must be an integer number of wavelengths in order that
the beams constructively interfere when they reach a detector, this path condition is
2h s = m
according to the figure. From the geometry alone we see that
s = 2x sin ,

x = h sin ,

and the inter-planar spacing is


d = d1 sin

55

d = h cos

incoming
e

to screen


s
d

d1

=2

from which we obtain


m = 2h 2x sin = 2h(1 sin2 ) = 2d cos = 2d1 sin cos
or
m = d1 sin 2 = d1 sin
in which d1 is the spacing between atoms, and is the scattering angle, the angle between
incoming and outgoing beams. We will look for the m = 1 bright interference band, the
primary.
The voltage used to accelerate the electrons gives the kinetic energies that are nowhere near
relativistic, and so we can use
p = mv,

1 2
1
p
K = mv 2 =
2
2m

and the deBroglie condition


=

h
p

together with conservation of energy


1
eV = mv 2
2
to obtain the electron wavelength in terms of the accelerating voltage

56

h
1.2268 109 mV olt 2

=
=
2meV
V
we find that after using the Bragg condition
s

1
2me
=
d1 sin
h2
V
is the relation between accelerating voltage and the scattering angle for the electron beam.
If we plot 1V versus sin , the slope will give us d1 , the inter-atomic spacing.
Examine the picture of the apparatus. We can measure the scattering angle by measuring
the radius of the ring that the scattered electrons leave on the fluorescent tube. Notice
tan =
and
x=

D
2

Lr+x

r2 (

D 2
)
2

or
D

= tan1 (

Lr+

q2

Douter
cm
4.59
4.77
5.08
5.31
5.59
6.00
6.51

V
kV
3.9
3.4
3.1
2.8
2.5
2.2
1.9

r 2 ( D2 )2

We see that we must produce a plot of sin versus 1V and run a linear regression on the
data to obtain the slope of the line of best fit in order to get the inter-atomic spacing. For
this apparatus we find the physical parameters are r = 6.6cm, the radius of curvature of the
bulb, and L = 14.0cm, the length of the bulb.

6.3

Data and analysis


Dinner
cm
2.68
2.79
2.93
3.05
3.22
3.44
3.69

I
A
40.7
39.7
34.0
50.0
42.0
48.6
52.3

We have recorded the diameters of two rings, inner and outer, and angular displacement of
the diffraction rings.
We process the data by first put this data into a file as follows

57

2.68
2.79
2.93
3.05
3.22
3.44
3.69

4.59
4.77
5.08
5.31
5.59
6.00
6.51

3.9
3.4
3.1
2.8
2.5
2.2
1.9

called data and run it through several Awk scripts to compute the average ,
awk { print sin(atan2(($1)/2.0, 7.4+sqrt(43.56+($1)*($1)/4.0)))," ",
1/sqrt(1000.0*$3)} data
which results in the data for the inner ring
0.0943793
0.0981397
0.102907
0.106975
0.112709
0.120078
0.128375

0.0160128
0.0171499
0.0179605
0.0188982
0.02
0.0213201
0.0229416

which we can LaTeX format as a table


Inner

sin()
0.0943793
0.0981397
0.102907
0.106975
0.112709
0.120078
0.128375

ring
1
V

0.0160128
0.0171499
0.0179605
0.0188982
0.02
0.0213201
0.0229416

The graph of this is shown below, and linear regression

58

Inner ring
0.023
0.022

1/V

0.021
0.020
0.019
0.018
0.017
0.016
0.09

0.10

0.11
0.12
sin()

0.13

0.14

results in a line of best fit with 1V = 0.002513 + 0.0.198905 sin , 2 = 0.000007 which is
a very good fit. From this we find that
0.198905
d1 =
= 0.0244 108 m = 2.44 Angstroms
8
8.14367 10
with an accepted inter-carbon in-plane separation of 1.40 Angstroms, it is pssible that this
ring represents scattering off of the large parallel planes of benzene rings, for which our value
is in error by 31 %.
It is more likely that we are measuring twice the 1.21
A separation illustrated below

1.21

1.4
for which our results are in excellent agreement of 0.8 %.
For the second ring we run
59

awk { print sin(atan2(($2)/2.0, 7.4+sqrt(43.56+($2)*($2)/4.0)))," ",


1/sqrt(1000.0*$3)} data
resulting in the data
Outer ring

1
sin()
V
0.157521 0.0160128
0.163204 0.0171499
0.17287 0.0179605
0.179941 0.0188982
0.188429
0.02
0.200617 0.0213201
0.215368 0.0229416
and run the linear regression for the line of best fit;
1 = 0.002244 + 0.117366 sin with 2 = 0.000008, resulting in
V
d2 =

0.117366
= 0.0144 108 m = 1.44 Angstroms
8.14367 108

which is evidently scattering off of a different set of planes, and is a good value for the
in-plane inter-carbon spacing of 1.4
A.

Outer ring
0.023
0.022

1/V

0.021
0.020
0.019
0.018
0.017
0.016
0.14

0.16

0.18
sin()

60

0.20

0.22

6.4

Crystal structure determination by scattering

In a crystal, the spatial structure is invariant under translations by vectors


T = na + mb + pc,

m, n, p

integers

forming a periodic structure falling into 14 types called Bravais lattices. The notation for
this is usually an ordered triple with no parenthesis m n p or commas. The atoms in the
crystal form a periodic structure that superimoses upon itself if it is moved by any of these
translation vectors.

System
Triclinic

14 Bravais lattices
Number of types Symbol
Characteristics
1
P
a 6= b 6= c, 6= 6=

Monoclinic

P, C

a 6= b 6= c, = 90 = 6=

Orthorhombic

P, C, I, F

a 6= b 6= c, = 90 = =

Tetragonal

P, I

a = b 6= c, = 90 = =

Cubic

P=sc, I=bcc, F=fcc

a = b = c, = 90 = =

Trigonal

R, I

a = b = c, = = < 120, 6= 90

Hexagonal

a = b 6= c, = 90 = , = 120

For the important example of a body-centered cubic structure, the translation vectors T are
illustrated below. There are two cells in these figures, the conventional cell, which is the
rectangular cubic structure,

61

and the primitive cell which is the parallelopiped defined by the three lattice vectors a, c
and c, which has volume |(a b) c|. A stereogram of the primitive cell for a body-centered
cubic lattice (in red) is illustrated below.

The Wigner-Seitz primitive cell is gotten by picking a lattice site, drawing a connecting
line to each nearby lattice site, and bisecting each of these lines with a plane. The smallest
volume contained by this structure is the primitive cell.
A stereogram of the primitive cell for a face-centered cubic lattice (in red) is illustrated
below.

If you look at these pairs correctly, you will be able to see the three dimensional structure
of the crystal. Be careful not to break your brain.
Solid state physicists propagate some confusion by reporting atomic positions for cubic lattices using the conventional cell coordinates.
Aluminum is a good example of a face-centered-cubic structure with the length of a side of
its conventional cell being 4.04
A and an atomic nearest neighbor distance of 2.86
A.

62

Once primitive cells have been established, Atoms can be placed at various points within
them, with the convention that the center of a primitive cell is at 21 12 12 . A common error that
people make is to think that the atoms of the lattice are located on the vectors T, when the
atoms are merely separated by these vectors. For example, in a NaCl crystal, which is fcccubic, each primitive cell contains four Sodium and four Chlorine atoms, Sodiums being at
000, 12 12 0, 12 0 12 , 0 12 12 and Chlorines at 21 21 12 , 00 21 , 0 21 0, 12 00 in conventional cell (Cartesian)
coordinates. These atoms and there positions in the conventional and primitive cells are
illustrated in the stereogram below.

The length of the side of the conventional cell in the NaCl structure is 5.63
A.
Various types of waves incident on crystals interact with the atoms at lattice points, and
with planes upon which families of atoms lie. The Miller indices of planes of atoms appear
in the scattering formulas for these waves. Consider for example families of planes of atoms
parallel to a plane defined by three atoms that lie on it, for example 3a, 2b, and 3c. The
Miller indices of the plane are gotten by taking the reciprocals of these integers, namely
1 1 1
, , and constructing the smallest set of integers in the same ratio, for our example this
3 2 3
is (232). This triplet is called the set of Miller indices. In applications of these rules, a plane
that intersects a crystal axis at infinity has the corresponding index equal to zero.
Consider waves of k-vector k = (0, 0, 2
) incident on an array of atoms at positions T =

ma + nb + pc. The waves will arrive with a phase of


(n, m, p) = eikT
relative to the wave arriving at the atom at the origin. Consider spherical waves arriving
at a distant observation point in direction n from the atom at the origin, and a distance
R away. The phase of the wave scattering from the atom at T, arriving at the observation
point is
(n, m, p) = eikT+ikrn,m,p
in which rn,m,p is the vector between the atom at T and the distant point, so that
T + rn,m,p = Rn
Then
2
rn,m,p
= (T Rn) (T Rn) R2 2Rn T,

63

rn,m,p R n T

and our single arriving wave has phase


(n, m, p) = ei(kkn)T+ikR
We now add up all of the scattered waves coming from all of the atoms in the lattice from
which the waves scatter;
X i(kkn)(ma+nb+pc)+ikR
=
e
m,n,p

=(

eim(kkn)a+ikR )(

ein(kkn)b+ikR )(

= eiN kR (

eip(kkn)c+ikR )

eim(kkn)a )(

ein(kkn)b )(

eip(kkn)c )

Each sum appearing in this formula is in fact a geometrical series, very reminiscent of the
amplitudes for multiple slit interference, or diffraction gratings. If we use the sum
M
X

p=0

ipA

1 ei(M +1)A
,
=
1 eiA

M
X

ipA

p=0

sin( M2+1 A)
|=
sin( A2 )

we discover a wave amplitude of


|| =

sin( M2+1 (k kn) a) sin( M2+1 (k kn) b) sin( M2+1 (k kn) c)


sin( (kkn)a
)
2

sin( (kkn)b
)
2

sin( (kkn)c
)
2

in which M is some large integer. As M becomes very large, in other words as the size of the
crystal that we are scattering off of increases, this amplitude becomes very sharply peaked
around its central maximum, being just a product of three diifraction grating amplitudes.
The maximum occurs for n directions for which all three factors in the product are one;
(k kn) a = 2ma ,

(k kn) b = 2mb ,

(k kn) c = 2mc

in which ma , mb , mc are integers. These are the Laue conditions, and in fact are equivalent
to the Bragg condition.

6.5

Simulation of scattering

A typical crystal diffraction pattern for waves scattering off of a single crystal has bright
spots in directions in which these three conditions are met. We can easily simulate this. For
example, the (photographic negative) of the diffraction pattern for scattering from a 999
hexagonal, close packed crystal looks like the following

64

In polycrystalline scattering, like off of a powder consisting of many single crystals with
arbitrary orientations, the diffraction pattern is similar to that above, but is gotten from the
spot pattern by rotation to form concentric rings.
Below you will find a simple C program for simulating this experiment and testing the Laue
conditions. You may make changes in it, to output the angles and of the bright spots,
so that you can verify the conditions by hand.
#include<stdio.h>
#include<stdlib.h>
#include<math.h>
#include<plot.h>
#define PI 3.1415926
__complex__ double phase,eye;
__complex__ double cexp(__complex__ double num);
double modulus(__complex__ double num);
int m,n,p,M,N,pl_handle,num;
double a[3],b[3],c[3];
double r[3],Q,R,S;
double k[3],kp[3],spot[1000][3];
double phi,theta,x,y,dx,dy,amp,R,lambda;
main(){
/* lattice vector components */
a[0]=2.0,a[1]=0.0,a[2]=0.0;
b[1]=-sqrt(3.0),b[0]=1.0,b[2]=0.0;
c[2]=2.0,c[0]=0.0,c[1]=0.0;
eye=0.0+1.0i;
dx=0.05;
dy=0.05;
/* x-ray wavelength */
lambda=0.05;
k[2]=2.0*PI/lambda;k[0]=0.0;k[1]=0.0;

65

pl_handle = pl_newpl ("ps", stdin, stdout, stderr);


pl_selectpl (pl_handle);
if (pl_openpl () < 0)
/* open Plotter */
{
fprintf (stderr, "Couldnt open Plotter\n");
return 1;
}
pl_fspace (-4.0, -4.0,4.0, 4.0);
pl_filltype(1);
pl_fillcolorname("white");
/* pl_fbox(-4.0, -4.0,4.0, 4.0); */
pl_flinewidth (0.005);
for(M=-40;M<40;M++){
for(N=-40;N<40;N++){
/* using over 10,000 points on viewing screen */
R=sqrt(((double)N*dx)*((double)N*dx)+((double)M*dy)*((double)M*dy));
theta=atan2(R,30.0);
phi=atan2((double)M*dy,(double)N*dx);
/* this is k- the elastically scattered wavevectors */
kp[0]=(2.0*PI/lambda)*cos(phi)*sin(theta);
kp[1]=(2.0*PI/lambda)*sin(phi)*sin(theta);
kp[2]=(2.0*PI/lambda)*cos(theta);
phase=0.0+0.0i;
/* lattice will be 9x9x9, pretty big */
for(m=-4;m<=4;m++){
for(n=-4;n<=4;n++){
for(p=-4;p<=4;p++){
/* generate a lattice scattering site */
r[0]=(double)m*a[0]+(double)n*b[0]+(double)p*c[0];
r[1]=(double)m*a[1]+(double)n*b[1]+(double)p*c[1];
r[2]=(double)m*a[2]+(double)n*b[2]+(double)p*c[2];
phase=phase+cexp(eye*(r[0]*(k[0]-kp[0])+r[1]*(k[1]-kp[1])+\
r[2]*(k[2]-kp[2])));
}}}
amp=modulus(phase);
/*printf("%f\t%f\t%f\n",(double)N*dx, (double)M*dy, amp);*/
num=(int)(65535.0*(amp/729.0));
if(num>4000){
pl_color(num,num,num);
/* output a spot (hopefully) corresp. to reciprocal wavevector */
pl_fbox((double)N*dx-dx/2.0,(double)M*dy-dy/2.0,(double)N*dx+dx/2.0,(double)M*dy+dy/2.0);
/* check for laue conditions */
Q=(a[0]*(k[0]-kp[0])+a[1]*(k[1]-kp[1])+a[2]*(k[2]-kp[2]))/(2.0*PI);
R=(b[0]*(k[0]-kp[0])+b[1]*(k[1]-kp[1])+b[2]*(k[2]-kp[2]))/(2.0*PI);
S=(c[0]*(k[0]-kp[0])+c[1]*(k[1]-kp[1])+c[2]*(k[2]-kp[2]))/(2.0*PI);
66

/* output Q=m_a, R=m_b, S=m_c for brightest spots */


printf("%f & %f & %f & %f\\\\\n",100.0*(float)num/(float)65535.0,Q,R,S);
}
}}
pl_closepl ();
pl_selectpl (0);
pl_deletepl (pl_handle);
}
__complex__ double cexp(__complex__ double num){
__complex__ double tmp;
__real__ tmp=exp(__real__ num)*cos(__imag__ num);
__imag__ tmp=exp(__real__ num)*sin(__imag__ num);
return(tmp);
}
double modulus(__complex__ double num){
double tmpx,tmpy;
tmpx=__real__ num;
tmpy=__imag__ num;
return(sqrt(tmpx*tmpx+tmpy*tmpy));
}
The centers of the very brightest spots are organized in the table below, and the integers
for the Laue conditions are displayed. You can see that even for a mere 729 atoms, the
scattering maxima obey the Laue conditions quite well.

67

Laue bright spots


% brightness ma
mb
43.862058
0.998751 -0.999860
43.837644
-0.998751 -1.998611
57.804227
-0.000000 -1.038763
84.116884
-0.000000 -0.981102
82.714580
0.999575 -0.019607
83.109789
-0.999575 -1.019182
59.916075
0.999599 0.038104
59.301137
-0.999599 -0.961494
43.884947
1.997505 0.998752
100.000000
0.000000 0.000000
43.884947
-1.997505 -0.998752
59.301137
0.999599 0.961494
59.916075
-0.999599 -0.038104
83.109789
0.999575 1.019182
82.714580
-0.999575 0.019607
84.116884
-0.000000 0.981102
57.804227
-0.000000 1.038763
43.837644
0.998751 1.998611
43.862058
-0.998751 0.999860

mc
0.049962
0.049962
0.017988
0.016046
0.016989
0.016989
0.016046
0.016046
0.049906
0.000000
0.049906
0.016046
0.016046
0.016989
0.016989
0.016046
0.017988
0.049962
0.049962

Read through the program and determine how far the crystal sample is from the screen on
which these spots are cast, and find the wavelength. Use the data to actually determine
the lattice spacings. This might give you some insights into your experiment in electron
diffraction.

68

The Photoelectric Effect

In this experiment, the interpretation of which Einstein received the Nobel prize, we measure
Plancks constant and verify the particulate nature of light. It provides an excellent contrast
to the electron diffraction experiment that illustrates the wave nature of matter.

7.1

Apparatus

1. A light source. Welch Scientific model 36 Universal light source.


2. Planck constant apparatus; Sargent-Welch Scientific, catalog number 2120.

3.
4.
5.
6.

10 cm and 2.5 cm focusing lenses.


Voltmeter. Keithley 610c solid state electro-meter.
Ammeter. Keithley 485 auto-ranging picoammeter.
3 volt power supply. 2 1.5 volt size D batteries in series.

The 3 volt power source provides a fixed anode/cathode voltage difference and will be used
to measure the maximum energy of the ejected photo-electrons. There are a pair of variable
resistors R1 and R2 built into the Planck apparatus. They are used to precisely control the
anode/cathode voltage. A voltmeter is used to exactly measure this anode/cathode voltage.
The role of the ammeter is to measure the current of the photo-electrons. When the current
is zero, no electrons have sufficient kinetic energy to reach the cathode.
Filters built into the apparatus restrict the frequencies of light entering the device to a very
narrow bandwidth. We really only want light of one frequency entering the device, to be
sure that we know precisely which wavelengths cause photo-emission. The lense is used to
eliminate other light sources andto direct the beam from the universal light source. The
source itself is as close to white light as money (a little, not a lot) can buy. It provides the
photons that cause photo-emission of electrons from the target metal (anode).

69

White light
Lense
Changeable
filter

Electric field
Cathode

Anode

electron beam

I
Ammeter
V
Voltmeter

R1

R2
3 V DC

7.2

Background

In the Photoelectric effect, light incident on the anode can be absorbed by electrons , and
if they absorb enough energy, the electrons can be liberated from their potential wells and
escape the metal. The kinetic energy of the electron plus the energy required to liberate
them from the anode must equal the energy of the photon, here assumed to be a discrete
particle with energy related to its frequency
K + = h
is called the work function, and naturally varies from metal to metal. In terms of light
wavelength this is
hc

The kinetic energy K can be measured by applying sufficient voltage to stop the electron
K +=

K = eVstop
and we find that
Vstop =

e
e

70

stopping voltage versus light frequency is a straight line whose slope gives the value of
Plancks constant and intercept determines the work function of the metal.

7.3

Procedure and data

We will apply various voltages to the anode/cathode and monitor the current for several
light wavelengths.
577 nm filter 546 nm filter 435.8 nm filter
V
I
I
I
volt
A
A
A
0.050 .3475
.6206
1.6568
0.150 .2009
.3882
1.2375
.250
.0971
.2070
.8615
.350
.0373
.0928
.5461
.450
.0098
.0329
.3227
.550
-.0001
.0066
.1768
.650
-.0027
-.0018
.0898
.750
-.0034
-.0045
.0404
.850
-.0036
-.0052
.0127
.950
-.0036
-.0055
-.0016
1.05
-.0037
-.0056
-.0090
1.15
-.0038
-.0057
-.0119
1.25
-.0039
-.0058
-.0130
1.35
-.0040
-.0058
-.0134
1.45
-.0041
-.0059
-.0136
1.55
-.0041
-.0138
1.65
-.0042
-.0060
-.0140
1.75
-.0042
-.0141
2.00
-.0061
-.0143
2.10
-.0043
-.0143
2.50
-.0043
-.0062
-.0145
2.75
-.0043
We now graph this data ( voltage versus current, 435.8 nm, 546 nm, 577 nm respectively)
and determine the stopping voltage by looking for the place where the data curves start to
deviate from the straight line portion, since no current passes due to flowing electrons once
this point is reached.

71

2.0

0.7
0.6

1.5
0.5
0.4

1.0

0.3
0.5

0.2
0.1

0.0
0.0
0.5
0.0

0.5

1.0

1.5

2.0

2.5

0.1
0.0

0.5

1.0

1.5

2.0

2.5

0.35
0.30
0.25
0.20
0.15
0.10
0.05
0.00
0.05
0.0

0.5

1.0

1.5

2.0

2.5

3.0

By doing a spline fit we find that for 577 nm light ( = 5.1957 1014 Hz)
Vstop = 0.7167 0.0866 volts
for the 546 nm ( = 5.4907 1014 Hz) data we obtain
Vstop = .8667 0.0866 volts
and for the 435.8 nm ( = 6.8791 1014 Hz) data
Vstop = 1.33 0.0866 volts
This we now graph, stopping voltage versus frequency or else do a linear regression on the
data, to get the slope of the best fit line

72

1.4

1.2

1.0

0.8

0.6

0.4

4.5

5.0

5.5

6.0

6.5

7.0

The linear regression gives us a slope of


0.356822 0.032069

volt
h
=
14
10 Hz
e

from which we get


h = 5.17163 1.0913 1034 J sec

which is a rather good result , close to the accepted value of 6.6262 1034 J sec

Measuring Plancks constant (solid state version)

In this version of the experiment you will use the Klinger LED apparatus, containing a selection of six LED (light emitting diodes) in series with a 100 resistor. Hopefully you will
be able to obtain five or even six data points for the Planck line.
In addition you will need a 2 12 V AC variac and two voltmeters (DMMs).

8.1

Theory

When free atoms are brought together to form a periodic-lattice solid, the multiply degenerate energy levels of the free atoms blend into energy bands, near continuous collections of
electron levels. The valence electrons of the seperated atoms tend to form into one band,
the valence band, in the solid, and bonding electrons often compose a seperate higher energy band, the conduction band. The presence of electrons or other charge carriers in the
conduction band determine electrical conductivity of the solid, and there is usually a gap
between the valence and conduction bands.

73

Conduction band
Eg
Valence band

Seperate atoms

Solid state

We know that an n-type semiconductor has electrons in its conduction band, and a p-type
has some holes in the valence band. Before an n-type and p-type are joined to form a pnjunction, the electron energy diagrams for the two look like the following;
ptype

ntype

Conduction

Valence

Distance

The gray areas represent filled electronic band states. Once the two materials are physically
joined, there are electron and hole concentration gradients at the junction; on the n-side,
there is a higher concentration of electrons than on the p-side, and on the p-side the hole
concentration exceeds that of the n-side. Concentration gradients drive diffusion, and so
electrons will diffuse from the n-type into the p-type, and holes diffuse from p-type to ntype. Holes and electrons then annihilate one another, this forms a thin layer around the
junction in which there will be excesses of the appropriate donor or acceptor ions.

74

E
ptype

ntype

Conduction

Valence

Distance

This thin layer is called the space-charge region, and on the n-side of it there will be
an imbalance of acceptor ions, on the p-side an imbalance of electron donor ions, and this
creates an electric field and potential in the space-charge region. This is called the built-in
field. The field exerts forces on electrons and holes that opposes the diffusion, allowing an
equilibrium the result.
Examine the electrostatic potential energy of electrons on both sides of the junction, since
and outside of the space-charge region the two materials are electrically
Ebuiltin = dV
dx
neutral, we see that a solution to the Poisson equation is a constant potential Vp in the
p-type material, a constant Vn in the n-type, and the ramping potential due to the built-in
field connects the two in the space-charge region.
E

V(x)

ptype

ntype

Conduction band

Vn
EF

Ebuiltin

Valence band

Vp

Multiplication of this potential by the electron charge (negative) results in the electronic
energy structure illustrated above for a pn-junction at equilibrium, here illustrated with a
Fermi level indicated by a dotted line.
75

Consider the energy level scheme of a pn-junction forming the boundary between an n-type
solid with a partially filled conduction band (filled electron states) up to some Fermi level
(indicated on the right), and a p-type solid with holes acting as charge carriers. On the
p-side, empty electron states (holes) in the valence band are not shaded. The gap is drawn
on the p-side for convenience; there are no empty hole states in the conduction band on
the p-side, and the filled hole states fall short of the valence-to-conduction band gap on the
p-side.
If the junction is forward-biased, the energies of the n-type bands will be raised to a point
such that the barrier presented to the flow of electrons from the n-side is dramatically reduced, and electrons can migrate into the lower levels of the conduction band on the p-side
from the n-side, where they combine with holes.
E

ptype

ntype

Filled e states
(nconduction band)
EF

Empty states (holes)


not shaded
(pvalence band)

This results in the creation of a photon whose energy is the band-gap energy(between valence
and conduction bands).

76

E
h=Egap
ntype

ptype

|e| Vmin

If the band-gap is at least Eg 1.5 eV , the emitted photon will be visible. In addition
holes can migrate from the p-side to the n-side, combine with electrons and emit the same
frequency light. This process may be a bit easier to understand; the hole migrating to the
n-type side is a vacancy in the valence band on the n-side, which annihilates an electron
by providing a vacancy that an electron in the conduction band can fill (to oversimplify the
matter somewhat).

E
ptype

ntype
h=Eg
|e| Vmin

In any event, when the applied forward bias is at the threshold for a charge carrier transfer
from one side to another, one can see that we have
|e| Vmin = h
= h,
77

|e| Vmin Eg

which allows us to measure Plancks constant and verify the Planck Hypothesis by measuring
the threshold voltage for forward-biased conduction and plotting it versus LED dominant
radiated frequency. In the real world the diode emits a variety of frequencies, but the distribution is fairly sharp at a precise value. The wavelengths of the dominant photons are
printed right on the apparatus.

8.2

Procedure

It should be fairly clear what needs to be done; for each diode a plot of current versus voltage
across the diode needs to be produced. The current will be very small, but not zero (due to
tunneling) until the voltage drop reaches the required Vmin for forward conduction, at which
point it will climb rapidly.

100

mA

V
DMM

Do this for each diode, and use the computer program for fitting data to polynomial curves
to fit that portion of your data for which I > 0 to a high order polynomial such as
I = a4 V 4 + a3 V 3 + a1 V + a0 , and use the method of bisection program to find the
roots or zeros of this function. This will give you Vmin in each case. Both programs can be
obtained from the course website.
Example Suppose that I took the (ficticious) current versus voltage data below.

78

7
6

I(mA)

5
4
3
2
1
0
0.0

0.5

1.0
V

1.5

2.0

I have divided the data into two sets, that with current equal to zero (squares) and that
with current greater than zero (circles). I now run a least-squares polynomial fit on the
circle-data, obtaining
I(V ) = 0.7 Amp + 0.3

Amp 2
Amp 4
V + 0.5
V
2
V olt
V olt4

as the best fitting quartic, the fifth and sixth order fits being worse by a Chi-squared criterion. I obtain Vmin by the method of bisection or Newtons method, beginning with a guess
near 1.0 V .

8.2.1

Finding roots by bisection

To apply the bisection method, pick two points, x1 and x2 . If there is a root between x1 and
x2 , then
f (x1 )f (x2 ) < 0
otherwise
f (x1 )f (x2 ) > 0
The first step in the algorithm is to test which case prevails.
2
. Now test if there
If there is a root between x1 and x2 , bisect the segment and let z = x1 +x
2
is a root between x1 and z. If there is , reset x2 = z and iterate. If there was not, then there
is a root between z and x2 and so we reset x1 = z and iterate. We continue to do this until
we have narrowed down the interval in which there is a root to whatever accuracy we desire.
The simple C + + program below accomplishes this.

79

#include<iostream.h>
#include<math.h>
float f(float x); /* define your function in this routine */
main()
{
float x0,x1;
cout<<" Input x0 and x1, two points around a root"<<"\n";
cin>>x0>>x1;
for(int n=0;n<20;n++){
if (f(x0)*f((x0+x1)/2.0)<0.0)
x1=(x0+x1)/2.0;
else x0=(x0+x1)/2.0;
}
float root=(x0+x1)/2.0;
cout<<"the root is = "<<root<<"\n";
}
float f(float x){
return(x*x-3.0*x+1.0);}
Once you have the least squares polynomial fit, you can create a subroutine that computes the polynomial, and point the function pointer to your routine.

8.2.2

Newtons method

The Method of Newton is another iterative scheme for finding or refining roots that can
be potentially unstable, and so should be used with caution. If
f (xn+1 ) f (xn ) + (xn+1 xn )f 0 (xn ) 0
then
xn+1 xn

f (xn )
f 0 (xn )

so if xn is close to a root, xn+1 should be closer. The disadvantage, other than stability, is
that you need the derivative function as well as the function itself in order to evaluate this
expression. If you have both functions, this can be used to polish or refine an estimate of a
root gotten by bisection.

80

One distinct advantage of the Newton method over bisction is that since it is non-geometrical,
it applies equally well to real or complex roots. One needs to make sure that if a complex
root is sought, a complex initial guess is used, otherwise polynomials with real coefficients
are always real-valued. For example

/* Newton-Rafeson computation of roots of cubic */


/* for complex root, make sure x0 (first guess) */
/* is a complex number
*/
#include<stdio.h>
int n;
__complex__ double x0,x1,y;
__complex__ double f(__complex__ double x);
__complex__ double df(__complex__ double (*pf)(__complex__ double x), \
__complex__ double x); /* derivative routine accepts pointer to function */
double modulus(__complex__ double z);
main(){
x0=1.0+0.9i;
do{
y=x0;
/* test for precision */
x1=x0-f(x0)/df(f,x0); /* pass derivative a pointer to f */
printf("%f+%fi\n",__real__ x0,__imag__ x0);
x0=x1;}while(modulus(y-x0)>=1.0e-15);
return;
}
__complex__ double f(__complex__ double x){
return(x*x*x+1.0);
}
__complex__ double df(__complex__ double (*pf)(__complex__ double x),
__complex__ double x){
__complex__ double del,y;
81

del=0.1;
y=(8.0*(*pf)(x+del)-8.0*(*pf)(x-del)-(*pf)(x+2.0*del)- \
(*pf)(x-2.0*del))/(12.0*del);
return(y);
}
double modulus(__complex__ double z){
double re,im;
re=__real__ z;
im=__imag__ z;
return((re*re+im*im));
}
At any rate these programs give you at least two options for computing Vmin from
I(Vmin ) = 0, and you can devise other methods as well.
Your lab report will contain the current versus diode voltage drop for all six diodes in tabular
form, graphs of all six sets of data, the set of points used for the polynomial fit must be
clearly marked since I will perform the fit to check your work, and the actual polynomial, fourth or higher order. You will also report a table of Vmin versus diode wavelength
and frequency or each diode, and finally produce a graph of Vmin versus f = , along with
a least squares fit to determine Plancks constant from the slope, and an error estimate for
it. You must compare your value of Plancks constant with the accepted value.
All of these programs compile on any Linux or FreeBSD computer, and on CYGWIN or
MINGW32 as well. The lab computers are equipped with CYGWIN, and LaTeX for your
convenience.

82

Helium Ionization Potential; The Franck-Hertz Experiment

How do electrons interact with ordinary atomic matter? In 1914 J.Franck and G. Hertz
performed an experiment in which a beam of electrons is passed through a gas of atoms.
Some of the electrons will lose energy to the atoms (originally a gas of mercury vapor) . The
electrons are first accelerated by a voltage Vc between the apparatus cathode and a grid, and
also by a decelerating voltage Va between the anode and grid. Electrons should arrive at
the anode with energy e(Vc Va ), unless of course they interact with the gas atoms. In this
case they arrive at the anode with a final energy Ekin = e(Vc Va ) E. As anode current,
which measures the energy of electrons arriving at the anode, is monitored as a function of
Vc , we see fluctuations in the current at precise intervals.

Hg sample
25

I milliamps

20

15

10

10
15
V, volts

20

25

As the accelerating voltage is increased, electrons arrive at the collector ring in the apparatus
at an increasing rate, and so we expect the collector ring current to increase with accelerating
voltage. This does occur, until the incoming electrons have enough kinetic energy to excite
the Mercury atom in the tube. Then an inelastic collision takes place, the incoming electron
virtually stops dead, the Mercury atom is excited, and the ring current drops since electrons
are not reaching it.
This indicates that there is a threshold energy for electron/atom interactions, and that the
thresholds occur at energies regular intervals of 4.9eV for mercury. What we are seeing is
83

the projectile electron giving up energy to promote bound electrons in the mercury atom to
higher atomic energy levels. This is reinforced by the fact that each 4.9eV energy loss is
accompanied by photon emissions from the mercury, of light with a wavelength corresponding to this energy difference. This is very strong evidence for the discrete nature of bound
electron energy states.
We perform a variation on the classic experiment here at Parkside. Instead of mercury,
we excite a gas of helium atoms with the electron beam. This particular apparatus may be
phased out, since it is a bitch to set up. A short description with data for a newer apparatus
follows this section.

9.1
1.
2.
3.
4.
5.
6.
7.
8.
9.

The old version of the experiment; apparatus

1 MegaOhm resistor. Cornell-Dublier electronics decade resistor model RDC.


1.5 V power supply. Single 1.5 volt size D battery.
20 V power supply. Heathkit IP-2718.
Signal generator. Wavetek model 182A function generator.
Ammeter. Keithley 175 auto-ranging multi-meter.
Voltmeter. Keithley 178 digital multi-meter.
12 volt power supply. 12 volt automotive battery, Farmn Fleet model UL-300CA.
Rheostat. Cenco 22 ohm, 4.4 amp , catalog number 8291013.
Franck-Hertz tube. Tektronics Franck-Hertz apparatus.

84

These are assembled as shown below.


Ammeter

collecting ring

Filament

A
anode

Rheostat
V
Voltmeter
+ 12 V Power
supply

6 V peak-to-peak
signal generator
1M

1.5 V DC

0-20 V power supply

Oscilloscope

We briefly describe the role each item plays in the experiment.


Oscilloscope monitors the voltage across the 1M resistor, through which runs the current
of electrons that arrive at the collector ring. Increases in the electrons making it to the ring
increase the current, and the voltage across the resistor. The scope also monitors times
at which the voltage reaches a value that indicates absorptions by the helium gas. This is
signaled by sudden drops of voltage across the resistor, decreases in electron flow translate
into drops in beam kinetic energy due to energy absorption by the gas.
The resistor is in the circuit simply to aid in measuring the voltage or current from the
collection ring.
1.5 volt battery is used to raise the collection ring voltage slightly above the anode voltage.
Electrons that lose energy in collisions with helium atoms will be moving slowly enough to
be collected.
20 volt power supply is used raise the saw-tooth voltage from the signal generator to such
a point that the peak to peak value of the total voltage input spans the full energy range for
excitation of the helium, from 20 to 26 volts.
The signal generator provides a predictable voltage that increases at a determined rate.
Rather than performing the entire experiment with several cathode voltages, we perform it
with an AC sawtooth voltage that sweeps through the entire excitation range periodically.
The ammeter monitors the filament current, which cannot exceed 1.5 amps without being
damaged.
The voltmeter monitors voltage across the filament, which needs to be kept below 5 volts
85

in order not to damage the filament. Another safety precaution.


12 volt power supply provides a constant DC voltage to the filament. A battery is more
constant than an electronic power supply, no ripples or fluctuations. This produces a constant production of thermionic electrons.
The rheostat limits the flow of current to the filament, again to prevent burn out.
The Franck-Hertz tube contains a filament from which electrons are expelled by thermionic
emission. These will be accelerated and will collide with helium in the tube. Varying the
voltage on the anode creates the accelerating potential difference for the emitted electrons,
which are collected by the collector ring.

9.2

Data and analysis

Once the apparatus is assembled and all voltages are in the required range, we can take data.
We adjust the 20 V power voltage until three excitation peaks are visible on the oscilloscope,
and the time of the peaks are recorded in milliseconds

E1
E2
E3

trial 1
trial 2
trial 3
ms
ms
ms
1.9 0.4 1.9 .4 1.8 .4
3.9 0.4 3.9 0.4 3.8 0.4
6.1 0.4 6.2 0.4 6.0 0.4

In order to transform these times into voltages, we need to know the slope of the signal
generator waveform
V = 6.0 0.1 volts
t = 3.8 0.1ms times 2

since the scale on the scope has 2 ms per division. This gives us a slope of
0.789 0.025

V
ms

and we can compute the energies of the excitations


E1 = e(0.789

V
)(1.87ms) = 1.48eV
ms

V
)(3.87ms) = 3.05eV
ms
V
E3 = e(0.789
)(6.1ms) = 4.81eV
ms
and we can perform standard Gaussian error analysis on our data to arrive at
E2 = e(0.789

E1 = 1.5 0.6eV, E2 = 3.05 0.55eV, E3 = 4.81 0.57eV


which agree well with 1.1, 3.1, 4.8 eV respectively, the commonly accepted values.

86

9.3

Second version

This version uses the TEL 2533.01 Helium Hertz critical potential tube, or the This version
uses the TEL 2533.02 Neon tube, and a chart recorder. Recommended as well are the TEL
2800 LV and TEL 2801 HV DigiRamp power supplies, and a picommater.
The tube containing the gas sample has a weakly conducting inner coating connected to the
anode, otherwise it functions pretty much the same as in the previous description; accelerated
electrons undergo inelastic collisions with gas atoms once they reach sufficient energy to cause
the atoms to undergo an atomic transition, otherwise the collisions are elastic. As the beam
undergoes further acceleration by the field in the tube, it may reach another critical value
and reveal another atomic energy level. For the Neon tube, a typical chart recorder output
looks like this;

Neon sample
50

40

30

20

10

10

20
30
V, volts

40

50

recording collector ring current (from electrons that give up all or most energy to excite
atoms) versus accelerating voltage.
In this figure the current peaks at 18.30, 21.22, 39.08, and 41.93 V , corresponding to Neon
atomic absorptions of 18.30, 21.22, 39.08, and 41.93 eV . These compare very favorably with
the published peak voltages of 17.0, 21.5, 34.0, 38.5 V . Remember that it is the sudden drop
in current at the peak that signals the prpjectile electron energy loss. This energy loss equals
the difference in two atomic energy levels of the Neon, and the Neon will then spontaneously
decay by emitting light at these energies. You can see this with a diffraction grating.

87

Examine the first pair of peaks for this run. The first peak drops off rapidly at E = 18.30 eV ,
if the atom were to shed this energy as light, it would be at wavelength
1240 eV nm
hc
=
= E = 18.30 eV

or = 67.8 nm, which compares well with the known Neon line at N e = 73.5 nm. The
second peak at 21.5 eV corresponds to Neon ionization by the incoming electron.
Correcting for the contact potential will improve the accuracy. I would suggest determining
which spectral line is being excited for each of your peak dropoffs.
Neon has spectral lines at = 73.5, 73.4, 585.2, 640.2, 671.7 nm.

88

10

The Rydberg Constant

This constant determines the spacing between Hydrogenic Bohr energy levels, and was originally determined spectroscopically. In our modern lab experiments course, you will perform
the same experiment. This simply requires a diffraction grating, the theory of which we have
studied in Physics 202. The grating will be incorporated into a spectrometer, and we will
study the measure the Rydberg constant for hydrogen.

10.1

Background

The Old Quantum Theory that supplanted classical mechanics as a description of the
electron in an atom was based
We now state the previously derived quantization rule as a
H
postulate Postulate W = p dq = nh where h is Plancks constant.

Apply this now to the problem of finding the quantum energy levels of the hydrogen atom,
with lagrangean
m
L = (r 2 + r 2 2 + r 2 sin2 2 ) V (r)
2
The momenta are
pr = mr
p = mr 2 sin2
p = mr 2
and the Hamiltonian is then, by eliminating the derivatives
p2
p2r
p2
+ V (r)
+ 2+
2m 2mr
2mr 2 sin2
Impose the quantization conditions
E=H=

or p = `
h and similarly p = m
h and
I

pr dr =

p d = `h

2m(E V (r)

(`
h)2
) dr = nh
2mr 2

Now insert the form of the potential


k
r
and note that the contour integral has two poles, one for large r and one for small, which
we evaluate using the Cauchy theorem
V (r) =

89

dz
= 2i
z
z k dz = 0

,for any k value other than 1. For small r we neglect all but the last term under the radical
to get
s

2m(n
h)2
= 2i `2 h
2 = `h
2mr 2
For large values of r we instead throw out the last term in the radical and expand using the
binomial theorem, but only one term will have a residue
I

m
k
k
+ )dr = 2ik
2m(E )dr = 2mE (1
r
2Er
2E
r

The final result is


m
= nh
2E
This can be inverted to get E, recall that E is in fact negative
I

pr dr = `h 2ik

mk 2
E= 2
2
h (` + n)2
The interpretation of this result is very simple; there is an energy level degeneracy due
to the fact that several different states with different angular momenta can have the same
energy. These states differ in their orbital eccentricity.
From our classical mechanics experience we know that planetary orbits are elliptical with a
semimajor axis a given by
k
a=
2|E|
and semiminor axis b determined by both energy and angular momentum
b=q

L
2m|E|

such that the orbital eccentricity is


=

b2
1 2 =
a

2|E|L2
mk 2

For the Sommerfeld calculation these become


q

h
` = b 2m|E|,

90

m
= (n + `)
h
2|E|

Multiply these;
h
`
q

b 2m|E|
or

m
h
`a
=
= (n + `)
h
2|E|
b

n+`
N
a
=
=
b
`
`
in which the energy quantum number is
N = n + `,

E=

mk 2
2
h2 N 2

This means that our orbits have quantized semimajor and semiminor axes.
k
h
2 2
aN =
N = a0 N 2
=
2|E|N
mk
and

`
aN = `N a0
N
in which a0 = 0.53 Angstroms is the Bohr radius. Some of these orbits are illustrated below
b=

N=1, l=1

N=2, l=2

N=2, l=1

N=3, l=3
N=3, l=2
N=3, l=1

Sommerfeld was able to go one step further, by beginning with a relativistically correct
expression for the energy, all of the integrations are pretty much the same, and the end
result illustrates the fine structure splittings in the energies of certain orbits with the same
N but different `, orbits that would be degenerate (have the same energy) without relativity
91

taken into account.


The total energy according to relativity is
E=

c2 p~2 + m2 c4

k
r

which we write as

k
(E + )2 m2 c4 = c2 p~2
r
and we now insert or old expression for the momentum in polar coordinates from our first
Sommerfeld computation;
k 2
p2
2 4
2 2
(E + ) m c = c (pr + 2 )
r
r

but again

p d = `h

and since p is conserved (constant) this gives us


p = `
h
again. We conclude that relativistically
s

`2 c2 h
2
k k2
1
(E 2 + 2E + 2 ) m2 c4
pr =
c
r r
r2
and therefore its corresponding action quantization is
I

1
`2 c2 h
2
k k2
(E 2 + 2E + 2 ) m2 c4
dr = nh
c
r r
r2

This is exactly what we had before, with a few changes. Integrating exactly the same way
as before we obtain
s
k2
Ek
i 2 `2 h
2 + i 2
= h
h
c
c E m 2 c4

If we call

k
Zc
h

e2
40 h
c

= , the fine structure constant, we can rewrite this as

EZ
=n
i Z 2 2 `2 + i 2
E m 2 c4

and we expand in powers of the fine structure constant, which is very small;

and so
i

Z 2 2

`2

Z 2 2
+ )
i(`
2`

Z 2 2
Z 2 2
EZ
(n + `)
=N
2`
2`
E 2 m 2 c4
92

or
m2 c4 = E 2 (1 +

Z 2 2
)
2 2
)2
N 2 (1 Z2`N

This gives us the final relativistically correct energy quantization rule


E=r
mc2 q

mc2
1+

Z 2 2
2 2 2
N 2 (1 Z2`N
)

1
1+

Z 2 2
N2

Z 4 4
N 3`

1 Z 2 2 Z 4 4
3 Z 2 2
mc2 (1 ( 2 + 3 ) + ( 2 )2 + )
2 N
N `
8 N
2 2
2 2
Z
Z 1
3
mc2 (1
(1
+
(

)) + )
2N 2
N ` 4N
The first term we recognize as the rest energy of the electron, and all of the rest is the
quantized electronic energy
EN,` =

mc2 Z 2 2 mc2 Z 4 4 1
3

(
)+
2
3
2N
2N
` 4N

This shows that the different ` elliptical orbits allowed for a given N value are not degenerate
but are split by some very small energies. The fine structure constant is
=

e2
1
= 0.0072974
40 h
c
137

and this provides us with a very valuable tool for getting an idea as to how large a correction
to the energy each term represents. The basic electron energy is mc2 = 511 keV . The Bohr
keV
orbitals represent states whose energies differ from this by 2 mc2 = 511
0.02723 keV =
(137)2
27.23eV , and these fine structure corrections are four orders of magnitude smaller still, or
keV
mc2 4 = 511
1.451 106 keV = 1.451 103 eV . A most remarkable fact is that we
(137)4
know that Bohr-Sommerfeld theory is very flawed, yet this expression for the fine structure
spectrum is exactly correct, even when compared to computations performed with the
relativistic version of the Schrodinger equation (the Dirac equation), with only the proviso
that ` is replaced by `+ 21 to correct for electron spin, a factor unknown at Sommerfelds time.

93

E
N=4
N=3

-0.85 eV
-1.50 eV

N=2

-3.38 eV
Balmer Transitions (vis)

Lyman Transitions (UV)

N=1

-13.6 eV

UV

IR
Lyman

10.2

Balmer

Atomic Spectra

By hypothesis, Hydrogenic atoms only radiate light when they jump from one Bohr-Sommerfeld
stable orbit to one of lower energy. When this occurs, light of frequency f and wavelength
is emitted with
hc
= EN,` EM,`0
hf =

The spectrum of emitted radiation is determined by the differences in the energy levels.
Example Consider a simple quantized atomic system in which an electron can have four
energy levels E1 < E2 < E3 < E4 as shown below. There can be six wavelengths of light
emitted when a gas of such atoms is thermally excited.

94

2
1
Levels

Transitions

R
Emission Spectrum

These wavelengths are (from most energetic or most blue, to least energetic or reddest)
hf4,1 = E4 E1 , hf4,2 = E4 E2 , hf3,1 = E3 E1 , hf3,2 = E3 E2
hf4,3 = E4 E3 ,

hf2,1 = E2 E1

a better representation of what one would see of this light were passed through either a prism
or diffraction grating would be

R
Emission Spectrum

Keep in mind that we will see spectral lines because the light from the atoms will be passed
through a slit, and so the line on the film or grating is an image of the source or slit.
This type of emission spectrum is called discrete line, as opposed to a continuous spectrum
in which all colors or wavelengths of light are present, as in the spectrum of light emitted by

95

a very hot opaque body or a heated cavity in a conductor (cavity or BlackBody radiation
ala Planck). For light emitted by Hydrogen this expression becomes
hfn,m =
in which R =

mc2
2h

13.6 eV
13.6 eV

,
2
m
n2

1
n,m == R (

1
1
2)
2
m
n

is called the Rydberg constant


R = 109737.309 cm1

What we actually see when we analyze light from thermally excited atoms differs from the
predictions of Bohr-Sommerfeld theory only in two details, some of which are beyond the
ability of that theory to explain;
1. Not all predicted spectral lines (emitted wavelengths) are actually seen.
2. Not all emitted spectral lines have the same intensity or brightness.
Basic statistical mechanics can partially explain why some spectral lines are not seen if the
gas emitting the light is excited to temperature T . The probability that an given atom in
the gas will have its electron in energy level En is
En

(En ) = D(En ) e kT

in which D(E) is the density of states; the number of configurations of energy E. This is
also called the degeneracy of the energy level E. Notice that this tells us that if En >> kT
for a given temperature, we are very unlikely to find any atoms in energy state E n . If no
atoms are to be found in that state, we will see no light emissions from atoms falling out of
this state, and so many spectral lines emitted when atoms drop out of high energy states
will not be seen unless the gas is hot enough to have an appreciable number of atoms in the
requisite high energy starting-states to begin with.
Example At fairly low temperatures T , gas with the energy level spectrum of the previous
example may not be excited enough to have many atoms in E4 , so we will see almost no
light of the frequencies f4,3 , f4,2 , f4,1 emitted. At low T this gas will have emission spectrum
like that shown below.

R
Low T Emission Spectrum

However we will see in our excursions into wave mechanics, a far more correct picture of the
atom, that certain transitions between energy levels are forbidden on the basis of angular
96

momentum conservation, and other selection rules. Light carries angular momentum of
1h
, called spin angular momentum, and so in the emission of light by excited atoms, certain transitions might not be possible because the initial and final states may have angular
momenta that differ by the wrong amount for angular momentum conservation. This fact
was not known at the time of Bohr-Sommerfeld theory, and cannot be properly taken into
account except within the framework of wave mechanics, the successor-theory to the BohrSommerfeld model.
The inability of Bohr-Sommerfeld theory to explain the relative intensity of spectral lines was
one of its major failings. Different spectral lines are of different intensities because of several
factors, one of which is the degeneracies of the two quantum energy levels involved in the
transition. Suppose that We have four quantum energy levels E1 < E2 < E3 < E4 and they
are degenerate. An electron in E1 can have angular momentum 1 h
, and electron in E2 can
have angular momentum 1 h
or 2 h
and so one. We say that En is (n 1)fold degenerate.
Let us couple this together with the notion that since lignt carries angular momentum 1 h
,
the only allowd atomic transitions must obey the selection rule
` = 1,

We now list some transitions and the frequency of light emitted.


Initial N
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
..

Initial ` Final N
4
1
3
1
2
1
1
1
4
3
4
3
4
3
3
3
3
3
3
3
2
3
3
3
3
3
2
3
2
3
2
3
1
3
1
3
1
3
..
..

Final `
1
1
1
1
3
2
1
3
2
1
3
2
1
3
2
1
3
2
1
..

Allow/Forbid
forbid
forbid
allow
allow
allow
forbid
forbid
allow
allow
forbid
allow
allow
forbid
allow
allow
allow
forbid
allow
allow
..

f
f4,1
f4,1
f4,1
f4,1
f4,3
f4,3
f4,3
f4,3
f4,3
f4,3
f4,3
f4,3
f4,3
f4,3
f4,3
f4,3
f4,3
f4,3
f4,3
..

One can see the basic picture; there are two allowed transitions that will cause light of
frequency f4,1 to be emitted, if the temperature is high enough, and ten transitions of emission
97

frequency f4,3 and so the ratio of intensities should be


If4,3 : If4,1 =

10
=5
2

and we should see the spectral line of frequency f4,3 five times as bright at the f4,1 line.
Bohr-Somerfeld theory does not give us the correct energy level degeneracies, and so its
predictions of spectral line intensities are not correct. However it does successfully predict
the actual frequencies of the spectral lines to excellent precision, and so many of its core
ideas and assumptions must have some truth to them. The theoretical successor, wave mechanics, is incomparably more accurate, more general, and more applicable to a wider range
of problems.

10.3
1.
2.
3.
4.
5.

Apparatus

Sodium light source. Gaertner Scientific.


Hydrogen light source. Gas discharge tube powered up by a
Variac. Type W5MT3.
Diffraction grating.
Spectrometer. Gaertner Scientific.

The figure shows how the setup will look, with the variac connected to the hydrogen discharge
tube. The diffraction grating used in this lab (by me) has
m = 1, N = 25, 000, R = 25, 000
in which
m = d sin

R=
= mN

where R is the resolution in first order (m = 1). The quantity d in the first expression is
the distance between slits, the reciprocal of the number of lines per centimeter in the grating.
98

The first step is to determine the number of lines per centimeter by calibration using the
two Sodium lines
1 = 5889.953
A, 2 = 5895.923
A
We will first measure the angles at which the lines will be scattered by the grating.

Protractor
(angle scale)
Grating

H source

Slit colimator

Telescope
with cross
hair
Next we determine the angles at which maxima occur when light from hydrogen discharge
tubes is passed through the grating. We then solve for the wavelengths by
m = d sin
Do not confuse m with an atomic quantum number, it is the diffraction order of
the resolved spectral line.
Finally we use the fact that for the Bohr model of hydrogen, emissions occur with wavelengths
hc
c2 2 1
1
E=
=
( 2 2)

2
n
m
mp
me and = hkc is the fine structure
in which is the electron reduced mass = mmee+m
p
constant
1
= 0.0072974
137
e2
with k = 40 . Remember me c2 = 0.511 M eV and hc = 1239.8eV nm.
There will only be a few visible wavelengths from which we will deduce

2 2 k 2 e4 1
1
1
=
(

)
2

h3 c
ni
n2f
for ni nf principle quantum number transitions, with
= R (

1
1
2)
2
ni
nf

99

the Rydberg constant R .


What you should expect to see as you move the collimator from angle to angle is something
like the following

m=-2

m=-1

m=0

m=1

m=2

10.4

Data and analysis

1
2

Trial 1
36.1833o
36.233o

Trial 2
36.2o
36.233o

Trial 3
36.1833o
36.232o

from which we obtain the averages


1 = 36.1833o , 2 = 36.233o
Of course when you actually perform the experiment, you will need to do standard error
analysis. These give line spacings for the grating of
a1 =

5889.953
A
= 9975.59
A
sin(36.183o )

a2 =

5895.923
A
= 9974.99
A
sin(36.233o )

and

for an average linespacing of


a
= 9975.299 558.76
A

The Rydberg constant from Hydrogen spectra


With ni = 2 and nf = n;
color quantum number
red
n=3
green
n=4
blue 1
n=5
blue 2
n=6

trial 1 (deg)
40.1833
28.7167
25.5500
24.5500

trial 2 (deg)
40.1833
28.750
25.5500
24.6333

100

trial 3 (deg)
40.2000
28.7167
25.5667
24.6167

average (deg)
40.1889
28.7278
25.5556
24.6000

From this data and a


we find that
3 = 6437.16
A = 643.7 nm, 4 = 4793.82
A = 479.38 nm, 5 = 4303.24
A, 6 = 4152.52
A
You can compare these with the exact, known values reported in the next experiment, agreement with known wavelengths is very good, to within 2%.
These lines are transitions from the m = 2 level into one with quantum number in the table.
n12
.13889
.18750
.2100
.22222

1
22

n
3
4
5
6

.000155348
.000208601
.000232382
.000240817

The average value for the Rydberg constant from these four data can be gotten by plotting
the reciprocal wavelength versus the difference in the squared reciprocal quantum number
difference, and doing a linear regression.

0.24
0.22

(n24)/4n2

0.20
0.18
0.16
0.14
0.12
0.14

0.16

0.18

0.20 0.22
1000.0/

R = 0.001057 0.00010792A1

0.24

which is in good agreement ( error is 4%) with the accepted value of


.00109737A1
101

0.26

10.5

Diffraction gratings

Diffraction gratings are essentially multiple-slit aperatures that can be either reflection or
transmission devices. In a transmission grating, very fine grooves are cut into a transparent
glass or polymer slide. These grooves act like aperatures and light transmitted through them
undergoes multiple-source interference. From the figure below representing the grooves or
cuts seen edge-on

1
r

P
0

r-1

we can use superposition to construct the intensity I at the point P on a screen upon which
the transmitted light will fall. We represent a light wave of polarization P and wavelength
as
2
1
E0
2
~ = p
E
sin ( r t) = p Im ei( rt)
r

r
and notice that if we label the central slit j = 0, then the j th slit is approximately
rj = r0 + jd sin
away from point P , where r0 is the distance from point P to the central slit, and d is the
slit seperation. We can add up all of the waves arriving from all slits at point P ,
~T =
E

N
X

2
E0
pj Im ei( rj t)
rj
j=N

N
X

2
E0
p0 Im ei( rj t)
r0
j=N

102

N
X
E0
2
r0 t)
i( 2

p0
Im e
ei jd sin
r0
j=N

This sum is not difficult, call

x = ei d sin
then it is

N
X

j=N

xj = xN (1 + x + x2 + + xN ) = xN
1

=
=

1 xN +1
1x

xN + 2 x(N + 2 )
1

x 2 x 2

sin ((2N + 1) d sin


)

d sin
sin ( )

and we can assemble the intensity from the time averaged Poynting vector
I() =

) 2
) 2
sin ((2N + 1) d sin
1 E02 sin ((2N + 1) d sin

(
)
=
I
(
)
0
d sin
d sin
20 c r0
sin ( )
sin ( )

In this case our grating has 2N + 1 slits, the corresponding expression for an N -slit grating
is
sin ( N dsin ) 2
I = I0 (
)
sin ( d sin
)

For N = 11 the intensity of transmitted light versus angle looks like the following figure.
Notice that the bright fringes or maxima have an amplitude of about N 2 times the brightness
of the N 2 secondary maxima that lie between consecutive primary fringes. As N increases
the disparity in relative brightness grows, until we see only very sharp, bright primary maxima with complete darkness between them as N grows very large.

103

Diffraction grating, 11 slit


140
120
100

I()

80
60
40
20
0
2

0
(d sin)/

Typical gratings will be quoted as having a certain number of rulings or grooves cut into
them per centimeter. For example a grating with 10, 000 rulings per centimeter has a d-value
or slit-to-slit seperation of
d=

1.0 102 m
= 1.0 106 m = 1000.0 nm
4
10

If we pass a light sample containing several distinct wavelengths of light through the grating,
we will see one diffration pattern per wavelength present in the sample. These will not superimpose perfectly, and by measuring the spread we can determine the actual wavelengths.
This is what is done in the Rydberg constant experiment.
One must first establish the value of d for the grating being used by measuring the angular
position of a well known or standard wavelength of light, such as that from a laser or from
a mercury lamp. Once d is known, the light to be measured is passed through the grating
104

and the angular positions of the various wavelengths present is measured.


Consider a spectrum containing two colors of visible light passed through a diffraction grating, producing the intensity pattern seen below

Resolution in first order


300
250

I()

200
150
100
50
0
15

10

0
5
, degrees

10

15

We see two fringes resolved in first order since the m = 1 interference fringes for each
color are well separated. Carefully measuring with a ruler we see that the m = 1 fringes
have angular positions

1
2

(degrees)
7.697
9.236

105

Supose that we have previously measured our diffaction gratings d value, and found that it
had 10, 000 rulings per centimeter, and so
d = 1000.0 nm
Then since the positions of the primary maxima of the grating are
d sin = m
we discover that
1 = (1000.0 nm) sin(7.697) = 133.9 nm
and
2 = (1000.0 nm) sin(9.236) = 160.5 nm

11

Measurement of unknown spectral lines

We can use the same apparatus to identify an unknown sample of gas from its visible emission
spectrum.
This is done by creating a calibration curve for the instrument using a known sample, the
best being Mercury due to the number of visible spectral lines. The idea is that we use the
instrument to measure the angular location of particular spectral lines passed through the
grating, or through a prism, and match these lines to known wavelengths. We can make a
plot of versus angle , but this will be nonlinear, so instead we will plot the sine of the
angular displacement of a line versus .
Mercury spectral lines
6
2
Color
, nm 1.0102 nm
yellow-1
579.0 2.983
yellow-2
577.0 3.004
green
546.1 3.353
blue-1
496.0 4.065
blue-2
491.6 4.138
blue-violet 435.8 5.265
blue-violet 434.7 5.265
blue-violet 433.9 5.265
deep-violet 407.8 6.013
deep-violet 404.7 6.106
Hydrogen spectral lines
red
656.3 2.322
blue-green 486.1 4.232
violet
434.0 5.309
deep-violet 410.2 5.943

106

faint
faint
faint
faint
faint
faint
faint

11.1

The Mercury data

Using the same apparatus as in the previous experiment, we calibrate the grating spectrometer with a Mercury source, using the four brightest lines.
n=2,L

n=1, L

n=1, R

n=2, R

displacement
182 50

I found the instrument to be centered at 182o 500 = 182.8333o , and the angular positions and
displacements for the first three diffraction orders to be
Line
violet
green
yellow-2
yellow-1
violet
green
yellow-2
yellow-1
violet
green
yellow-2
yellow-1

Left
Right
n = 1 first order
190.4
175.3
192.333 173.333
192.883 172.8
192.917 172.75
n = 2 first order
198
167.55
201.867 163.6
202.967 162.45
203.017 162.4
n = 3 first order
205.667 159.6
211.667 153.217
213.383 151.367
213.483 151.25

Average Displacement
of diffraction
7.55
9.5
10.0417
10.0833
of diffraction
15.225
19.1333
20.2583
20.3083
of diffraction
23.0333
29.225
31.0083
31.1167

In this table all angles have been converted to decimal degrees from degrees and minutes,
for example 190o 240 = 190.4o , 175o 180 = 175.3o , and so forth.
We now could compute the diffraction grating spacing a from
n = a sin
but we have already done such a thing in the previous experiment. If we were to do it now
for this data and this grating, we obtain a
= 33800
A.
Instead we will correlate these lines with true wavelengths to produce calibration curves
relating sin to for each diffraction order n, which we can use to determine the wavelength
of an unknown spectral line from its angular displacement.

107

11.2

First order calibration curve


Line
violet
green
yellow-1
yellow-2

true nm
435.8
546.1
577.0
579.0

Average Displacement rad


0.13177
0.16581
0.17526
0.17599

sin
0.13139
0.16505
0.17436
0.17508

A linear regression using the program lg in the support software or from the 499 website
archive on rustam.uwp.edu on this versus sin results in a straight line
= 4.655966 + 3281.222900 sin
with a 2 = 0.000151 for the fit, which is a very good fit.

11.3

Measuring Sodium lines

The Sodium source produces the following lines in the first order set of diffraction maxima;
red; right angular position; 172o 460 , left; 194o 140 with the instrument centered on 183o 300 .
This results in an average angular displacement of = 10.7333o = 0.18733 rad, and we
measure this lines wavelength from the calibration curve to be
= 4.655966 + 3281.222900 sin(0.18733) = 615.74 nm
which is in superb agreement with the known red-Sodium line at true = 615.43 nm!
The yellow Sodium line is seen in first order n = 1 diffraction at angular positions right;
173o 140 and right; 193o 470 , with an average displacement of = 10.275o = 0.179333 rad.
We determine its wavelength to be
= 4.655966 + 3281.222900 sin(0.179333) = 589.94 nm
again in very good agreement with the true value true = 589.00 nm.

108

12

Stefan-Boltzmann Radiation Law

The purpose is to verify the radiated power versus temperature for a blackbody radiator,
=A

2hc2

hc

5 (e kT

1)

d = AT 4

in which A is the surface area of the radiator, is the Stefan-Boltzmann constant


= 5.67 108

12.1

W
m2 K 4

Background

If a metallic box enclosing a cavity is slowly heated ,and the wavelengths of light emitted
through a small hole in one wall is monitored, the proportion emitted at various wavelengths
could not be successfully predicted by classical physics. This is our first basic problem;
computation of the fraction of energy radiated within a given wavelength range. The box
itself, unheated, is a perfect absorber since once light enters the hole its odds of bouncing
back out are slim. The box is a perfect absorber since if its temperature is stable, any light
that enters will be absorbed and re-emitted by the walls. A perfect absorber is also a perfect
emitter (Kirchoffs law), since light that enters the hole will be absorbed and re-emitted,
and ultimately this energy must escape or the box temperature would rise.
Electromagnetic portion. The whole problem of heat radiation is complicated by the
fact that it lies at the intersection of several very distinct fields of physics, in particular
electromagnetism and thermodynamics. We begin this investigation with its electromagnetic
component.
You are familiar with Maxwells equations in integral form from elementary physics
I

E n dS =

qinS
,
0

B n dS = 0

and
d
dt

B n dS =

C=S

E d`,

0 IthruC

d
+ 0
dt

0 E n dS =

C=S

B d`

In empty space both the static charge q and conduction current I, the currents of moving
charges, are zero, leaving Faradays and Amperes laws in a highly symmetrical form
d
dt

B n dS =

C=S

d
0
dt

E d`,

0 E n dS =

We now apply the well known vector calculus identity


I

C=S

V d` =

109

V n dS

C=S

B d`

in which S is a two-dimensional surface in space, with normal n and right-handed bounding


curve C. This allows us to write both sides of the two dynamical laws of electromagnetism
as
d
d
B = E,
0 0 E = B
dt
dt
To proceed further we apply another vector calculus identity
I

S=V

V n dS =

V d3 x

in which S is a closed surface enclosing volume V . We apply this to the two static equations
(Gausss laws) to obtain, in empty space
E = 0,

B =0

Maxwell used these four equations to prove that light is an electromagnetic wave, a fact
confirmed experimentally by Hertz. To do this, first differentiate the two dynamical equations
to get
d
d2
d
B = ( E),
0 0 2 E = B
dt
dt
dt
and substitute one into another to arrive at
( E) = 0 0

d2
E
dt2

and now use a third vector identity to simplifiy the cross product
( E) = (( E) 2 E) = 0 0

d2
E
dt2

and use the differential form of Gausss law to arrive at the electromagnetic wave equation
1 2
2
( 2 2 )E = 0
c t
in which the speed of propagation of the waves so constructed is
c=

m
1
= 3 108
 0 0
s

the previously established speed of light.


We will now take for granted that light is an electromagnetic wave, and contemplate a hollow
metallic box containing electromagnetic standing waves, in thermal equilibrium with the hot
walls. By the time of Plancks heat radiation experiments, it was widely accepted that
radiant heat was a form of electromagnetic radiation with long wavelenghts compared to
visible light.
We now have two tasks; first to determine the spectrum of frequencies of standing ling waves
that can exist inside the box, without being annihilated by interference, and second, to
determine how many distinct waves have the same frequency, or how many distinct waves
have frequencies between f and f + df . We will need to do this in order to compute the
110

spectrum of light emissions from a small hole cut in the wall.


If the walls of the box are for example perfect conductors, so that no electric field can
penetrate into them, and if the walls are at x = 0, x = a, y = 0, y = a, and z = 0, z = a,
then we must have E = 0 on the walls we find the solution
Ex = E0 sin(

my
pz it
nx
) sin(
) sin(
)e
a
a
a

for a cubic box of side a such that


(n2 + m2 + p2 )

2
2
=
a2
c2

and each mode in the box is specified by these three integers n, m, p.


For example in two
dimensions let N () equal the number of modes in a radius of r = n2 + m2 around zero
frequency. Then as r becomes very large the number of combinations(n, m) becomes the
area of a quarter circle.
10

10
10

10

In two dimensions a light wave propagating an a given direction can have only one possible
polarization direction perpendicular to its flow of energy, so

1
N () = 1 r 2 = 2 a2 f 2
4
c
The number of modes per unit frequency interval is then
dN (f )
a2
= 4f 2
df
c
In three dimensions we find that N is twice ( because any light wave propagating in direction n can have two independent polarizations p such that p n = 0 in three dimensions)
111

one-eigth the volume of a sphere of radius r and so


1 4 3 8a3 f 3
N (f ) = 2 r =
83
3c3
so that
dN
8a3 f 2
=
df
c3
This is referred to as the density of states for electromagnetic waves in the box.
The next stage is to compute the total energy possessed by the radiation within the box
according to
Z 2
X dN
dU
U
d U (f, T )
= 3 =
u
(f ) =
df
dV
a
dV df
f df
in other words we multiply the number of modes of frequency f by the average energy of
such a mode when in equilibrium with the hot walls at T , and sum over all of the modes.
This requires input from thermodynamics to compute the average energy that a light wave,
which we will regard as an oscillator since it satisfies an oscillator equation, can
extract from the walls at equilibrium.
Thermodynamic portion. According to the classical equipartition of energy each quadratic
when in equilibrium with
degree of freedom in the Hamiltonian has average energy of kT
2
a heat bath. Using the Gibbs Canonical Ensemble,
=
E

(E) E dE =

E
2
E e kT dE

R 2 E
kT dE
0 e

R
0

= kT

This amount of energy then determines the amplitude of the wave, since in classical electrodynamics the energy density of electromagnetic radiation is proportional to the electric
field amplitude squared. The average energy stored as electromagnetic waves in the box per
volume per frequency interval is
dN (f )
8kT f 2
d2 U (f )
=
u(f ) =
df dV
dV
c3
This is the Rayleigh-Jeans law. It is of course not in agreement with experiment, as
Planck was to show. This law predicts an ultraviolet catastrophy; the box would contain
an infinite total energy, since it contains ever increasing amounts of energy at higher and
higher frequencies.

112

RayleighJeans curves for T=T0, 1.5 T0, 2.0 T0


800

d2U(f,T)/df dV

600

400

200

10
hf/kT

15

20

The problem lies in the assumption that the equipartition law is valid. This cannot be
true, there must be dependence upon the frequency, or the theory is doomed to ultraviolet
catastrophe.
Planck noticed that the experimental data can be fit with tremendous accuracy by instead
assuming that each mode, which is in fact an oscillator by virtue of being waves,can only have
a discrete spectrum of excitation energies rather than all possible energies, with an
average determined by the temperature. He proposed that the energy of each mode is n (hf )
in which n is 0, 1, 2... (the Planck hypothesis) and the constant h needs to be determined by
fitting the experimental data.
This assumption of discrete oscillator energies as opposed to continuous simply translates
into the need to sum rather than integrate in order to compute the average energy per mode.
We find now that
E =

(E) E dE =

E
2
E e kT dE

R 2 E
kT dE
0 e

R
0

n
E
kT
n=0 En e
P
n
E
kT
n=0 e

Fortunately this is no harder to compute than the integrals of the Equipartition version,
since the sums are geometric
P

hf
n kT

hf
ne
E = Pn=0 hf
n kT
n=0 e

hf
=

hf
kT
hf

(1e kT )2

1
1e

113

hf
kT

hf
e

hf
kT

This is the celebrated Bose-Einstein distribution function, and inserting this into the
expression for the energy density per frequency interval in the box yields
dN (f )
8f 2 hf
d2 U (f )
=
u
(f ) = 3 hf
df dV
dV
c e kT 1
which is in exact agreement with the experimental data, and correctly predicts the total
energy stored in the box as a function of temperature,the Stefan-Boltzman law
dU
=
dV

8hk 4 T 4
d2 U
df =
df dV
c 3 h4

8k 4 4 4 4
x3 dx
=
T = T 4
ex 1
c3 h3 15
c
2

U
versus
in which = 5.67 108 m2WK 4 is the Stefan-Boltzmann constant. The graph of dfd dV
f has a maximum value at a wavelength predicted by the Wein law of displacement

Planck curves for T=T0, 1.5 T0, 2.0 T0


12

d2U(f,T)/df dV

10
8
6
4
2
0

10
hf/kT

15

20

hc
k
Which can be found by maximizing the energy per frequency interval with respect to frequency.
max T = 0.2012

The Stephan-Boltzmann law has been used to estimate the size of the universe as it cools,
since cooling must be at constant total energy, assuming that the dominant constituent of
the universe is the electromagnetic field filling it. From U = 4c V T 4 we obtain the expansion
rule
V1 T14 = V2 T24
114

12.2

Luminosity or radiance

Consider now the problem of the luminosity of a small hole cut into the wall of the cavity.
As a function of the temperature of the box, what is the total flow of power in the form
of heat radiation through the hole? In the figure below we see a hole of area A. Consider
a volume element dV = r 2 dr sin d d within the cavity containing electromagnetic field
energy. The energy per unit volume in the cavity is
8hf 3
d2 U
=
(Planck),
hf
df, dV
c3 (e kT 1)

8f 2 kT
(Rayleigh-Jeans)
c3

and so the little volume element contains energy within freqency interval [f, f + df ] of
dU
8hf 3
=
r 2 dr sin d d (Planck),
hf
3
df
c (e kT 1)

8f 2 kT 2
r dr sin d d (Rayleigh-Jeans)
c3

all in the form of photons, all moving isotropically away from the box at speed c. The box
is a distance r < c dt away from the hole of area A.

c dt=R
dr

dV

-z

The fraction of the radiant energy in the volume element heading in a direction n that will
take it through the hole is

Azn
angle subtended by hole
A cos
2
= r =
4
4
4r 2
and so the energy from this particular volume element that will leave through the hole
within time dt is
A cos
8hf 3
8f 2 kT 2
A cos
dUthru
r 2 dr sin d d
=
(Planck),
r dr sin d d
(Rayleigh-Jeans
hf
2
3
df
4r
c
4r 2
c3 (e kT 1)

115

We now add up the contributions from all of the volume elements within a hemisphere of
radius R = c dt against the wall of the box containing the hole; for the Planck case
Z R Z Z 2
8hf 3
8hf 3 c dt A
A cos
dUthru,total
2
2
=
=
r dr sin d d
hf
hf
df
4r 2
c3 (e kT 1) 0 0 0
c3 (e kT 1) 4

and for Rayleigh-Jeans

8f 2 kT Z R Z 2 Z 2 2
8f 2 kT c dt A
A cos
dUthru,total
=
=
r
dr
sin

d
d
df
c3
4r 2
c3
4
0
0
0

Now divide by dt and use P =


radiated through the hole

dU
dt

to get the power at frequencies between f and f + df

8hf 3 c A
dP
=
Planck
hf
df
c3 (e kT 1) 4

dP
8f 2 kT c A
=
Rayleigh-Jeans
df
c3
4

The ultraviolet catastrophe is now pretty obvious for Rayleigh-Jeans; integration over all
frequencies results in an infinite power flow for the Rayleigh-Jeans law, but results in the
well-known and correct Stephan-Boltzmann law for the Planck version
Ptotal =

Z
8(kT )4 4 c A
8hf 3 c A
dP
df =
df
=
= AT4
hf
3 h3
3
df
4
c
15
4
0
kT
c (e 1)

for the power flowing from the hole. This is true for all Blackbody radiators, such as the
metal filament used in this experiment.

12.3

Apparatus

1. Stefan-Boltzmann Lamp. Pasco Scientific model TD-8555.

2. Thermal radiation sensor (thermopile). Pasco model TD-8553.


3. 3 Subtronics model 2000 multi-meters, to be used as voltmeters and an ammeter.
4. Power supply, 0-20 V. Micronta dual tracking DC power supply.

116

The apparatus is set up as in the figure.

Ammeter

0-20 V
Power

Model TD-8555
Stefan-Boltzmann
Lamp
CAUTION
Voltmeter

Voltmeter

It is very important that the voltage applied to the lamp never exceed 13 V, and that the
current through the lamp never exceed 3 A.
The power received by the heat sensor which detects thermal radiation in the range
0.5m 25.0m
will be monitored by placing the lamp a fixed distance from the detector. The lamp is
powered up and the voltage and current V , and I through the filament are recorded along
with the detector voltage. The resistance
R=

V
I

117

of the filament is a well known function of its temperature,


T =

R Rref
+ Tref
Rref

at low temperatures, and at high temperatures scales in a very precise way with respect to
a reference resistance.
The first step in determining the temperature function is to measure the resistance of the
filament at room temperature. To do this, carefully measure the room temperature, hopefully
around 300K, and measure the resistance R300 of the filament with a multi-meter. When
the filament is hot, its resistance will be related to the temperature by finding the ratio
R(T )
R300
and looking up the corresponding temperature in Kelvin in the table below, provided by the
manufacturer.

118

R
R300

1.0
1.43
1.87
2.34
2.85
3.36
3.88
4.41
4.95
5.48
6.03
6.58
7.14
7.71
8.28
8.86
9.44
10.03
10.63
11.24
11.84
12.46
13.08
13.72
14.34
14.99
15.63
16.29
16.95
17.62
18.28
18.97
19.66

Temperature (K)
300
400
500
600
700
800
900
1000
1100
1200
1300
1400
1500
1600
1700
1800
1900
2000
2100
2200
2300
2400
2500
2600
2700
2800
2900
3000
3100
3200
3300
3400
3500

We could also interpolate the graph below, made from this data.

119

3500
3000
2500
2000
Temp.
1500
1000
500
0

10
R/R_300

15

20

Instead we will use a Lagrange interpolation program that has this table coded into it. It
uses
T (R[n] + r) =

(r 1)(r 2)(r 3)
r(r 2)(r 3)
T [n] +
T [n + 1]
6
2

r(r 1)(r 3)
r(r 1)(r 2)
T [n + 2] +
T [n + 3]
2
2
and is given below

/* temperature interpolator for SB experiment */


#include<stdio.h>
#include<stdlib.h>
#include<math.h>
double R[34],T[34],alpha,y;
float r,c1,c2,c3,c4;
int n,m;
main(int argc, char *argv[]){
R[0]=1.0;T[0]=300.0;
R[1]=1.43;T[1]=400.0;
R[2]=1.87;T[2]=500.0;
R[3]=2.34;T[3]=600.0;
120

R[4]=2.85;T[4]=700.0;
R[5]=3.36;T[5]=800.0;
R[6]=3.88;T[6]=900.0;
R[7]=4.41;T[7]=1000.0;
R[8]=4.95;T[8]=1100.0;
R[9]=5.48;T[9]=1200.0;
R[10]=6.03;T[10]=1300.0;
R[11]=6.58;T[11]=1400.0;
R[12]=7.14;T[12]=1500.0;
R[13]=7.71;T[13]=1600.0;
R[14]=8.28;T[14]=1700.0;
R[15]=8.86;T[15]=1800.0;
R[16]=9.44;T[16]=1900.0;
R[17]=10.03;T[17]=2000.0;
R[18]=10.63;T[18]=2100.0;
R[19]=11.24;T[19]=2200.0;
R[20]=11.84;T[20]=2300.0;
R[21]=12.46;T[21]=2400.0;
R[22]=13.08;T[22]=2500.0;
R[23]=13.72;T[23]=2600.0;
R[24]=14.34;T[24]=2700.0;
R[25]=14.99;T[25]=2800.0;
R[26]=15.63;T[26]=2900.0;
R[27]=16.29;T[27]=3000.0;
R[28]=16.95;T[28]=3100.0;
R[29]=17.62;T[29]=3200.0;
R[30]=18.28;T[30]=3300.0;
R[31]=18.97;T[31]=3400.0;
R[32]=19.66;T[32]=3500.0;
/* find R in table closest to input r */
n=0;
r=atof(argv[1]);
for(m=0;m<31;m++){
if(fabs(r-R[m]) <= fabs(r-R[n]))
n=m;
}
/* now we have bracketting values */
alpha=r-R[n];
c1=-(alpha-1.0)*(alpha-2.0)*(alpha-3.0)/6.0;
c2=alpha*(alpha-2.0)*(alpha-3.0)/2.0;
c3=-alpha*(alpha-1.0)*(alpha-3.0)/2.0;
c4=alpha*(alpha-1.0)*(alpha-2.0)/6.0;
y=c1*T[n]+c2*T[n+1]+c3*T[n+2]+c4*T[n+3];
printf("R/R(300) = %f at T = %f\n",r,y);
}
121

We simply run it by passing it the RR300 that we want T for. You will find this program as a
download on the 499 website archive.

12.4

Procedure

First measure the background reading of the thermal sensor with no power applied to the
lamp. This background must be subtracted from all of the subsequent detector readings.
With the lamp and detector a fixed distance apart, apply a voltage to the lamp, wait until
all meter readings stabilize, and record filament current, voltage and detector voltage, which
is proportional to the power it receives within the given wavelength band.

12.5

Data and analysis

In 1996 I took the following data for this experiment.


Fil. V (volts)
0.289
0.867
2.57
4.26
5.08
6.29
6.99
7.56

Dark
Background

Fil. I (amps)
0.636
0.933
1.36
1.72
1.88
2.09
2.21
2.24

Fil. V (volts)
0.043
0.0005

Sns V (mvolts)
0.220
0.830
4.95
11.36
15.28
21.78
25.99
31.10

Fil. I (amps)
0.144
0.001

Sns V (mvolts)
-0.005
0.03

This data is corrected with the awk script


awk {print $1-0.0005, " ", $2-0.001, " ", $3-0.03} data>newdata
and the resistances found with

awk {print $1, "&", $2, "&", $3, "&", $1/$2, "\\\\", "\hline"} newdata
which results in the table

122

Fil. V (volt)
0.2885
0.8665
2.5695
4.2595
5.0795
6.2895
6.9895
7.5595

Fil. I (amp)
0.635
0.932
1.359
1.719
1.879
2.089
2.209
2.239

Sns. V (mvolt)
0.19
0.8
4.92
11.33
15.25
21.75
25.96
31.07

R (ohm)
0.454331
0.929721
1.89073
2.47789
2.7033
3.01077
3.1641
3.37628

The resistance at room temperature is gotten by measuring the current and resistance
through the filament at very low applied voltage, essentially the dark filament line of
data above,
0.0005V
= 0.5
R300 =
0.001A
and we then process our data file newdata, the corrected voltages and currents, with
awk { print $3, "&", $1/(0.5*$2), "\\\\", "\\hline"}newdata
which gives us the table
Sns. V (mvolt)
0.8
4.92
11.33
15.25
21.75
25.96
31.07

R(T )
R300

1.85944
3.78146
4.95579
5.4066
6.02154
6.3282
6.75257

T (K)
500
875
1100
1175
1300
1360
1430

We now compare our results to the theoretical predictions. The total power radiated within
the wavelength band sensed by the IR detector is
=A

25.0m

0.5m

2hc2
hc

5 (e kT

1)

d = AT 4

which we compute for each temperature using the following C program

#include <stdio.h>
#include <stdlib.h>
#include <math.h>
int n; /* for looping */
double lambda, dlambda, constant1, constant2, integral,factor;
123

float T;
/* constant1 = 2 pi h c^2 */
/* constant2 = hc/k
*/
main(int argc, char *argv[]){
if(argc < 2 || argc >2){
printf( ./sb temp\n);
exit(0);}
T=(double)(atof(argv[1]));
dlambda=0.0000001;
/* dlambda=1.0E-7 */
constant2=0.014413;
constant1=3.749176e-16;
integral=0.0;
for(n=50; n<=2500;n++){
lambda=(float)n*dlambda;
factor=lambda*lambda;
factor=factor*factor;
factor=factor*lambda; /* lambda^5 */
factor=factor*(exp(constant2/(lambda*T))-1.0);
integral=integral+1.0/factor;
}
integral=integral*dlambda*constant1;
printf(%f\t%f\n, T, integral);
}
Our sensor does not recieve all of the radiated power, but a fixed fraction depending on its
surface area and the distance from the filament, however the sensor voltage is proportional
to the received power, and so the plot of ln Vsens versus ln should be a straight line of slope
one, since by hypothesis they are proportional to one another. The C program gives the data
Vsense (mV) T (K)
0.8
500
4.92
875
11.33
1100
15.25
1175
21.75
1300
25.96
1360
31.07
1430

theory
3350.55
32695.8
82104.5
107002
160530
192369
235243

The log-log plot of the theoretical power versus sensor voltage is below

124

106

105
logP
104

103
0.1

10

100

logV_sens
which has a slope of 0.933, confirming within an accuracy of 7% the Stefan-Boltzmann
Radiation law.

125

13

Measurement of the Compton Edge

The purpose of this experiment is to test the validity of the Compton Scattering expression
for rays scattered through large angles (180o ).

13.1

Background

The concept that light can be treated like a particle in some applications is not off the wall,
there is evidence that in collisions with matter light behaves like any other particle. In the
Compton effect a stationary electron is bombarded with photons and the scattering angles
and energies satisfy standard relativistic versions of energy and momentum conservation
laws.

The relativistic dispersion relation


E=

p 2 c2 + m 2 c4

suggests that for massless particles such as light the magnitude of the momentum vector is
E/c. The De Broglie relation
E
hc
p=
=
c

allows us to write the conservation of energy-momentum equation


pbef ore = paf ter
q
hf 0
hf 0
hf
2
0
cos ,
sin , 0, hf )+(p cos , p sin , 0, p2 c2 + m2 c4 )
( , 0, 0, hf )+(0, 0, 0, mc ) = (
c
c
c
This gives us two momentum conservation

hf
hf 0
=
cos + p cos ,
c
c

0=

hf 0
sin p sin
c

and one energy conservation law


hf + mc2 = hf 0 +

126

p 2 c2 + m 2 c4

Take the last relation and square it


(hf + mc2 hf 0 )2 = p2 c2 + m2 c4 = (hf hf 0 )2 + 2mc2 (hf hf 0 ) + m2 c4
and square the first two relations and add them
(
to get
p2 = (

hf
hf 0

cos )2 = p2 cos2 ,
c
c

hf 0
sin )2 = p2 sin2
c

hf 0
hf 0
hf
hf 0 2 2h2 f f 0
hf

cos )2 + (
sin )2 = ( )2 + (
)
cos
c
c
c
c
c
c2

or
p2 c2 = (hf )2 + (hf 0 )2 2h2 f f 0 cos

We have two equations for p2 c2 , equate them

(hf )2 + (hf 0 )2 2h2 f f 0 = (hf hf 0 )2 + 2mc2 (hf hf 0 )


reduce them algebraically to
2mc2 (hf hf 0 ) = 2h2 f f 0 (1 cos )
or

and use

mc2 f f 0
(
) = (1 cos ),
h
ff0
c
f

mc2 1
1
( 0 ) = (1 cos )
h f
f

= to arrive at the Compton scattering formula


= 0 =

h
(1 cos )
mc

in which is the angle that the light is scattered through.


This equation makes a prediction about what direction light will be scattered into when it
is involved in a collision with static electrons, and therefore can be verified (or invalidated)
by experimentation. From our conservation law for energy we see that
h(f f 0 ) =

p2 c2 + m2 c4 mc2 = Ek

which is exactly the kinetic energy of the electron. But we also know that
mc2 1
1
( 0 ) = (1 cos )
h f
f
which we can solve for f 0 ;
1
h
mc2 + hf (1 cos )
1
=
+
(1

cos
)
=
f0
f
mc
f mc2
127

or
f0 =

f mc2
=
mc2 + hf (1 cos )
1+

f
hf
(1
mc2

The electron kinetic energy is then

Ek = hf
= hf

1+

cos )

hf
cos )

hf
(1
mc2

hf
(1 cos )
mc2
hf
+ mc2 (1 cos )

which is clearly maximized by a direct back-scattering of the light, = for which 1cos =
2, giving us
hf
2 mc
1
2
Ek,max = hf
2
hf = hf
1 + 2 mc2
1 + mc
2hf
This is the maximal kinetic energy that a stationary electron can obtain in a collision with
a photon, and is called the Compton edge energy Ee . It is easy to set up the experiment,
and measure the stopping voltage needed to keep post-collision electrons from arriving at a
collection plate, where they will complete a circuit and create a current in the apparatus. All
experiments indicate that this result, gotten by assuming that light behaves like a particle
in its interactions with matter, is correct.
A common version of this experiment is the measurement of the Compton Edge at which
the photon is backscattered at angle = . From our work above we see that this results in
photons of energy
hf
E
E0 = hf 0 =
2hf =

1 + mc2
1 + 2E
mc2
The actual experiment is carried out with a -ray source, producing primary radiation
of energy E , a scintillation detector, and a multichannel analyzer. Gamma rays
entering the detector undergo Compton scattering from electrons in a crystal (NaI doped
with Thallium), liberating electrons. The detector produces electronic pulses of intensity
proportional to the electrical energy dissipated in the crystal. Using a series of
photoelectric dynodes, the scintillator produces a signal (photopeak) due to the primary
radiation, by using it to photoelectrically eject electrons. Another signal, the Compton
Plateau, is due to primary radiation scattering electrons out of the conduction band of the
crystal. The detector produces pulses proportional to the energy carried by these electrons,
which of course can acquire any energy from the primary gamma ray, right up to the maximal
value at the Compton edge
0 Ek Ek,max = hf

1
2
1 + mc
2hf

These signals are sorted by the MCA, which records the number of signals recieved versus
intensity or channel. The channel ch of the MCA output is linearly related to the energy
of the event that caused the signal
ch = a + b E
128

and so the device can be calibrated by recording the channels of two gamma ray sources of
known energy,
ch1 = a + bE,1 ,
ch2 = a + bE,2
from which
b=

ch1 ch2
,
E,1 E,2

a=

A typical MCA output looks like the following;

ch2 E,1 ch1 E,2


E,1 E,2

MCA output
2500

Counts

2000

1500

1000

500

200

400
Channel ch

600

800

The ideal output would be a sharp spike at a channel corresponding to the primary gamma
ray E , and a sharp-edged plateau extending from 0 to Ek,max , but these features are broadened by secondary scattering and other phenomena. The actual Compton edge occurs about
half-way up the plateau (at ch 330 on the figure).

13.2
1.
2.

Apparatus

137
60

Cs sample.
Co sample.
129

3.
4.
5.
6.

Multichannel Analyser. Canberra Series 30.


NaI Scintillation detector.
Power supply. Hewlett-Packard 0-200 V, 0-6 mA, 61516 A DC power supply.
Lead bricks and sheeting.

PHA M
N x10

4
N

Dead Time
%

count

1024

TL
cursor

512

MCS M+1
Nx10
3 4
2

1/2 2/2
5
6

1/1

com

0
ADC offset

vert.

vertical

2048

ADC gain
1024
512
256
2048

X1
ACQ
mcs
pha
mcsr
add
sub
col
time

memory

Amp
gain

X2
I/O
out
in

sel out

tty

ext
I/O

LLD

adc zero

ULD
high
low
input lev.

threshhold

channel
Power
on

Intensity

Roi
Expand
Scan
on
off
roll

Enter
off

Clear
on off
on

Gate

HV

Signal

off

120
VAC
Scintillation
counter

Lead bricks

Power
supply
120
VAC

13.3

Scintillation counters and photomultipliers

When a large collection of atoms in a gas or liquid are condensed to form a solid (crystalline),
the energy levels spread out to form energy bands as shown below. Valence electrons will
fill the lower energy valence bands, two electrons per band level (spin up, spin down). Conduction electrons will occupy the upper energy conduction bands, and be very delocalized
in the crystal, in a metal this gives rise to electrical conductivity. In an insulator there
are no electrons in the upper conduction bands, and there is an appreciable gap in energy
between valence and conduction bands, so it is very hard to promote a valence electron into
a conduction band level. In semi-conductors this band gap is very much smaller, so electrons
can be promoted into the conduction bands much more easily.
130

E
E
E

2
1

r
crystalline (small r)

gaseous (large r)

The electrons with which photons produced by the 137 Cs and 60 Co gamma-ray sources used
in this experiment will collide belong to the N aI crystal doped with Thallium in the scintillation counter. A gamma ray photon can collide with an electron in the valence band of the
NaI crystal, giving it enough energy to be ejected into the band of conduction energy levels
or even higher.

Incident
E
Conduction
band
Band gap
(6 eV for NaI)

Thallium
levels

Valence
band
ejected electron leaves a hole
The ejected electron leaves a hole in the valence band . Neighboring atoms try to share
electrons in attempts to fill in the hole,and this causes the hole to move through the crystal.
If the hole encounters a Thallium atom impurity, it can ionize the Thallium atom and the
previously ejected electron can fall into one of the excited states of the thallium atom. In
the drop from the conduction band intothe excited thallium electronic level, a photon of a
131

few eV energy is emitted, that starts the entire scintillation process.


The electron ejected from the valence band by the -ray is called the primary electron,
and doping with Thallium provides intermediate energy levels in the forbidden energy region between valence and conduction bands. The primary electron rejoins the valence band
by falling into these intermediate levels, performing an otherwise forbidden transition and
emitting light. The light produced is reflected by alumina or M gO reflective layers onto a
photocathode that emits electrons by the photoelectric effect.

NaI

Photocathode

Dynodes

The phototube then increases the number of liberated electrons 106 fold by producing secondary electron emissions with a series of dynodes. This electronic pulseis delivered to an
amplifier. To a very good approximation the pulse created is directly proportional to the
energy dissipated in the scintillator, and so the number of pulses displayed versus amplitude
(channel or energy) corresponds closely to the energy spectrum of the primary radiation.
The photoelectric effect is used in an important bit of laboratory apparatus called the photomultiplier. Schematically, they look like the following figure. This device has a window
coated with a photosensitive material that ejects an electron when a photon of sufficient
energy hits the window. This electron is accelerated by around 150 V by a high voltage
132

electrode, which it strikes. This is the first stage of a multi-stage collection of dynodes.
The accelerated electron proceeds to the next, higher voltage dynode, where it liberates
more electrons. The accelerated electrodes cascade deeper into the nest of dynodes, liberating more electrons. Eventually the single electron has become 106 electrons, a typical gain
factor for PM tubes operating at about 2000 V . This many electrons produces a signal big
enough for electronics to process and record.

Gamma ray
Photo-cathode

Dynode

Distributor
(high voltage to dynodes)

High voltage
input

Signal output

The probability that the origin photon entering the tube actually ejects an electron from
the photocathode is called the quantum efficiency. For light in the visible range, a good
PM tube has a quantum efficiency of around 0.25.
The figure above shows the spectrum recorded by a NaI(Tl) crystal when exposed to 0.6616
MeV -rays. Ideally we should get a continuous signal extending from zero energy to
133

Kmax = Ee , and a sharp peak at E for a single gamma ray energy incident on the scintillator. In practice the Compton continuum is smeared out, as is the photopeak, due to
secondary collisions and e+ e pair creations. The Compton Edge or the channel at which
electrons are ejected with Kmax = Ee by the -ray, is halfway up to the Compton plateau.
This is the energy that we are looking for in this lab.
The Multichannel Analyser displays 1023 channels and plots the number of gamma rays
detected per channel as a graph. The relation between the channel at which the photon is
detected and the corresponding energy deposited by the event in the scintillator is linear
ch = a + bE
in which ch is the channel, E is the energy and a and b are constants to be determined by
calibrating the device as illustrated in a previous section.
For this we use the

60

Co source which emits rays of energies


E1 = 1173.210 0.002KeV

and
E2 = 1332.470 0.002KeV
To calibrate the MCA, we will count the cobalt source for about 4 hours, and the count
number versus channel will look like the figure below.

13.4

Stage 1; calibration

When the experiment was performed in 1995, the following was found; the energy E1 photopeak occured at
ch1 = 837, 11010 counts
and
ch2 = 947, 8408 counts
We compute the standard deviations of the values of ch1 , ch2 by looking for the channels on
either side of the peak that hold half of these numbers of counts,
ch11 = 797, ch12 = 873
ch21 = 907, ch22 = 985
Assuming that each peak is a normal curve, the standard deviations given by
(chchp )2
1
A = Ae 22
2

or, the half-height channels are deviated from the photopeak channels by

ch = chp 2 ln 2 = chp 2.35


134

and so
ch1 =

ch12 ch11
= 32.34, ch1 = 837 32.34
2.35

and
We then set up the linear equations

ch2 = 947 33.19

837 = a + b(1173.210), 947 = a + b(1332.470)


and determine that the calibration constants are
a = 26.67, b = 0.6907

13.5

Stage II; photopeak and edge measurements

The 137 Cs sample was then set up and the following photopeak spectrum was observed after
counting for 30 minutes.
The main photopeak corresponding to energy
E3 = 661.64keV
was found with
ch3 = 476, 60, 000counts
ch31 = 453, ch32 = 496
and so
and a Compton edge at
corresponding to an energy

ch3 = 476 18.3


che = 345 10
Eedge = 461keV

which is quite close to the accepted value of 477.318 keV . The photopeak corresponds to a
-ray of energy
E = 650.54 keV
agreeing well with the 661.64 keV accepted value.

14

Energy versus Momentum for Relativistic Electrons

When an electron in a scintillators valence band is struck by a -ray from a radioactive


source, the electron can acquire energies of several MeV, making them highly relativistic.
By examining the energy of the Compton edge and the photopeak from the scintillator, one
can easily verify the form of the relativistic dispersion law
E 2 = p 2 c2 + m 2 c4
135

14.1

Apparatus

The same as the Compton edge experiment, with additional 22 N a, 137 Cs, 54 M n, and 60 Co
-ray sources. The energies E of the rays emitted by each source can be found below.

14.2

Procedure

The Compton edge and photopeak energies of all of the samples are measured. The actual
photopeaks should occur at
Isotope
22
Na
137
Cs
54
Mn
60
Co

E
0.511 MeV
0.622 MeV
0.835 MeV
1.178 MeV

From this data the channel (ch) versus energy (E) calibration constants for the multichannel
analyser are determined
ch = a + bE
We found in the last expriment that the electron kinetic energy at the Compton edge is
2E2
Ee =
mc2 + 2E
and at the Compton edge, the backscattered photon has energy
E 0 =

mc2 E
mc2 + 2E

Adding we find that


Ee + E 0 = E
equaling the energy of the pre-collision photon. Since the photon is backscattered we find
that
E (E 0 )
Pe =
c
Eliminating E 0 we find that after the collision the electron momentum is
Pe =

2E Ee
c

where Ee is the Compton edge energy and E is the photopeak energy.


The total energy of the electron is kinetic plus rest, or Ee + mc2 , which presumably satisfies
(Ee + mc2 )2 = Pe2 c2 + m2 c4 = Ee2 + 2Ee mc2 + m2 c4

136

or
Ee2 + 2mc2 Ee = Pe2 c2 ,

Pe2
Ee2
= Ee +
2m
2mc2
2

Pe
In order to determine the true dispersion law, we will measure Ee and E and plot 2m
versus
e
Ee , assuming that
Pe2
= a0 + a1 Ee + a2 Ee2
2me
and run a nonlinear least squares fit to determine a0 , a1 , a2 . Classically we should find that
a0 = 0, a1 = 1, and a2 = 0, however what we will see is relativistic departures from this
dispersion law.

14.3

Data and analysis

The channels and corresponding energies for photopeaks and Compton edges for all four
sources are in the table below.
Isotope
22
Na
137
Cs
54
Mn
60
Co

c
E M eV
107.5
0.511
143.0
0.622
183.5
0.835
260.0
1.178

ce
69.5
93.7
137.7
212.9

Ee M eV
0.341
0.447
0.640
0.969

We are using the MCA calibration constants


E = (0.00438 0.000035 M eV )c + (0.0366 M eV )
found by matching the photopeaks to their known energies. We now compute the classical
energy
Isotope
22
Na
137
Cs
54
Mn
60
Co

Ee M eV
0.341
0.447
0.640
0.969

Pe2
2me

M eV
0.453778
0.621535
1.03806
1.88236

We now run a nonlinear least squares fit of this data to try to fit the best curve to it. I used
the program polyfit to perform the regression, obtaining
Pe2
= 0.002837 + 0.957349 Ee + 1.015000Ee2
2me
in extremely good agreement with a supposition of a relativistic dispersion relation. We
obtain the estimate
1
me c 2 =
M eV = 0.493 M eV
2(1.015)
which is in error by only 3.6%.
137

15

Michelson-Interferometer and the Index of Refraction of Air

Apparatus
1.
2.
3.
4.
5.

Beck Interferometer M300/6407.


Vacuum pump, air cell.
He-Ne Laser.
Phototransistor.
Chart recorder.

The department has its own Michelson interferometer, however this apparatus belongs to
the chemistry department and is used in the physical chemistry lab course. The Beck interferometer used in the Michelson mode looks like the following.

138

Air
Cell

Source

To Vacuum pump

M3
Compensator

M
1

Operation of the interferometer is simple. The eye is aligned with the viewing axis about 10
inches from the instrument. There will be three reflected images of a pointer in the beam
divider head, as in A below.
Adjust the tilt controls on M1 to see which image can be moved, and the control should be
adjusted so that the moveable image superimposes on the stationary one B.

139

When exact vertical on horizontal coincidence of the images is achieved the interference
fringes will become visible. For unequal path lengths and nonparallel mirrors, the fringes
will be sections of circles surrounding a center displaced from the field of view. If the mirrors
are parallel but the path lengths are unequal, the fringes will be circular with a center within
the field of view, order n in the middle, order n 1, n 2 and so forth of increasing radii. If
the path lengths are equal to within 5 fringes, interference fringes will even be visible with
white light from a tungsten filament.
Zero path difference can be calibrated with a Mercury vapor lamp. Turn the path length
control dial in the upper right corner of the top figure in whatever direction causes the
fringes to move inward, in the direction of decreasing radius. As the path difference becomes
zero, the fringes contract to a central point. Now adjust the tilt controls to produce about
10 nearly vertical fringes as in D, and adjust the path length control until the fringes are
straight. Now switch to a tungsten lamp and continue to turn the path length control very
slowly until a group of bright fringes with a distinctive central dark band in the middle
appears in the field. Center this collection of fringes. The instrument is now ready and the
path length is zero.

Procedure
The He-Ne laser is split and recombined after passing through a pressurized cell. A photocell
and chart recorder are used to count the number of wavelength shifts the insrutment cycles
through as the cell is evacuated, and then slowly repressurized.

Sample data
The length of the air cell used is 0.10 m, the laser has a wavelength of
= 632.8173 109 m
140

We find that there will fit


0.2m
2L
=
= 316, 047 = N

632.8173 109 m
waves in the cell in vacuum, and so
2L = N v =

(N + N 0 )c
Nc
= ((N + N 0 )air ) =

nair

where N 0 is the number of wavelengths the interferometer cycles through as the cell is refilled.
Four separate trials revealed 83 fringes were shifted through as the cell was refilled with air,
and so
316, 047 + 83
nair =
= 1.0002626
316, 047
which agrees very well with the accepted value of 1.0002684 at 70o F , the air temperature of
the experiment.

141

16

Angular distribution of emitted radiation

The energies, and therefore frequencies, of light emitted by the Hydrogen atom when it
undergoes a transition from one energy level to another, can be gotten from the Bohr formula
h = E =

Z 2 mc2 2 1
1
( 2 2)
2
Ni
Nf

however not all energy level transitions that one might naively assume occur do not in fact
ever happen, because the transition violates certain conservation laws. We not only have an
energy balance to preserve
Ei = Ef + h
but angular momentum must be preserved as well as linear momentum. The problem of
energy emission by the atom is most appropriately addressed through quantum field theory,
a theory developed to study processes in which particle number or number of quanta is not
preserved. Emission of light is a creation process,
|n, `, mi |n0 , `0 , m0 i |k, i
in which the rightmost ket is that describing a photon of wavevector k and polarization state
.
We can treat this problem using perturbation theory and the Golden Rule developed earlier,
if we model the interaction between the photon and the atom with
HI =

A J d3 x

( ( )) is a field operator for the electronic current and


in which J = q 2im
A is the vector potential of the photon field. Both of these field operatorscan be Fourier
expanded in terms of solutions to the appropriate Schroedinger or Maxwell equation, and we
will do so in a very schematic manner, because we are in fact getting far ahead of ourselves
right now as far as subject matter is concerned.
Let an,`,m create an electron with wavefunction n,`,m (r, , ) from the vacuum, and an,`,m
be the corresponding annihilation operator, so that

[an,`,m , an0 ,`0 ,m0 ] = n,n0 `,`0 m,m0


and let bk, create a photon with wavevector k and (transverse) polarization vector v k, .
Recall that in empty space the independent polarization vectors must be perpendicular to
the wavevector
vk,1 k = 0,
vk,2 k = 0,
vk, vk,0 = ,0
The part of a properly quantized photon vector potential responsible for creating a planewave photon will then look like
A=

d3 k 0

0 =1,2

Nk0 ,0 vk0 ,0 eik r bk0 ,0


0

142

which when acting on the vacuum creates a photon with wavefunction


Z

hk, |A|0i =

d3 k 0

0 =1,2

Nk0 ,0 vk0 ,0 eik r hk, |k 0 , 0 i = vk, eikr

There will be a similar Fourier expansion for the electron field


=

(n 0 ,`0 ,m0 (r, , )an0 ,`0 ,m0 + n0 ,`0 ,m0 (r, , )an0 ,`0 ,m0 )

n0 ,`0 ,m0

When acting on the vacuum, this will create a Hydrogenic bound state particle with wavefunction
hn, `, m||0i =

n0 ,`0 ,m0

n0 ,`0 ,m0 (r, , )hn, `, m|n0 , `0 , m0 i = n,`,m (r, , )

We have performed a rather subtle conversion to the Bargmann-Fock space in which states
are labeled by the number of quanta of each type occupying the universe, and wavefunctions
are matrix elements of field operators in the basis of such particle number states.
We can now apply Fermis Golden Rule which states that the probability of transition from
initial state |ii to final state |f i is proportional to the modulus of the matrix element hf |HI |ii.
We find then that allowed transitions resulting in the emission of light will have nonzero
values of
Z
d3 x hnf , `f , mf ; k, |A J|ni , `i , mi i
When we exand J in terms of annihilation/creation operators, we only need to reatin the
one term in four that has a nonvanishing matrix element between the one electron in and
out states, we find then that the expression above reduces to
Z

dx

n0 ,`0 ,m0 n00 ,`00 ,m00

d3 k 0

0 =1,2

Nk0 ,0 eik r (

q
h
)vk0 ,0 (n 0 ,`0 ,m0 n00 ,ell00 ,m00 n 0 ,`0 ,m0 n00 ,`00 ,m00 )
2mi

hnf , `f , mf ; k, | bk0 ,0 an0 ,`0 ,m0 an00 ,`00 ,m00 |ni , `i , mi i


which we now evaluate using
hnf , `f , mf ; k, | bk0 ,0 an0 ,`0 ,m0 = h0|nf ,n0 `f ,`0 mf ,m0 k,k0 ,0
and
an00 ,`00 ,m00 |ni , `i , mi i = |0ini ,n00 `i ,`00 mi ,m00
and we finally obtain the matrix element
|

d3 x eikr (

q
h
)vk, (n f ,`f ,mf ni ,`i ,mi n f ,`f ,mf ni ,`i ,mi )|2
2mi

which is basically just the Fourier transform of the electronic current matrix element between
the initial and final electron states.

143

The Schroedinger equation itself allows us to evaluate this matrix element in the dipole
approximation,
eikr 1
within the volume over which the bound state wavefunction probability density is appreciable. For two different solutions to the Schroedinger equation
(

h
2 2
+ V (r))f = Ef f ,
2m

h
2 2
+ V (r))i = Ei i
2m

we multiply the first by i and the second by f and subtract to obtain


h
2
(i f f i ) = (Ef Ei )f i

2m
Multiply both sides and integrate

h
2
2m

d3 x r (i f f i ) = (Ef Ei )

d3 x rf i

However for any vector f that vanishes sufficiently strongly at infinity;


Z

rf d x =

(x, y, z) (

fx fy fz 3
+
+
)d x
x
y
z

which can be integrated by parts to give


=

(xfx |
,

yfy |
,

zfz |
)

(fx , fy , fz ) d x =

f d3 x

Applying this now to our current operator shows that


Z

(i f f i ) d3 x =

2m(Ef Ei )
h
2

f r i d3 x

the transition probability is then, in the dipole approximation


2m(Ef Ei )
q
h
|
)vk,
(
2
2mi
h

d3 x (n f ,`f ,mf r ni ,`i ,mi ) |2

which is proportional to the matrix element of the dipole operator qr between


the initial and final states.
Transitions resulting in the emission of light between two states i and f are said to be
forbidden if the matrix element of the electric dipole operator between these two states is
zero. Whether or not this matrix element is zero has little or nothing to do with the radial
part of the wavefunction, it is the angular wavefunctions that really determine the selection
rules since
s
8
<e Y1,1 (, )
x = r cos sin = r
3
144

y = r sin sin = r
and

8
=m Y1,1 (, )
3

4
Y1,0 (, )
3
and we can use the Clebsch-Gordon recursion relations deried much earlier in our study of
angular momentum to deduce that
z = r cos = r

x|ni , `i , mi i = A|n , `i + 1, mi 1i + B|n , `i 1, mi 1i


y|ni , `i , mi i = A0 |n , `i + 1, mi 1i + B 0 |n , `i 1, mi 1i
and
z|ni , `i , mi i = A00 |n , `i + 1, mi i + B 00 |n , `i 1, mi i
These tell us that the final state of the atom must differ from the initial by
`f = ` i 1
and
mf = mi 1,

m f = mi

These are known as the electric dipole selection rules and they govern what we see in
the spectrum of atomic transitions. You may have noticed that Fermis Golden Rule even
gives us some indication as to the polarization state of the light emitted in a given transition.

16.1

Polarization and angular distributions

Using the formula


|

2m(Ef Ei )
q
h
(
)vk,
2
2mi
h

d3 x (n f ,`f ,mf r ni ,`i ,mi ) |2

we can determine the most probable direction of emission for an oriented atom or nucleus
undergoing a transition, and we can determine the probability that a photon emitted in a
certain direction n = (sin cos , sin sin , cos ) will have a given polarization state. For
light with k-vector k = kn, there will be two independent polarizations vk, , = 1, 2, which
we will take to be
vk,1 = (cos cos , cos sin , sin )
and
vk,2 = ( sin , cos , 0)
which satisfy vk, k = 0.
If our atom or nucleus begins in a state with wavefunction
i (r, , ) = Ri (r) Y0,0
145

we have seen from the dipole selection rules that the final state after photon emission must
be
f (r, , ) = Rf (r) Y1,m
with m = 0, 1. If we use
s

2
x=r
(Y1,1 + Y1,1 ),
3

2
y = ir
(Y1,1 Y1,1 ),
3

z=r

4
Y1,0
3

then our matrix elements will be


Z
Z

and

1
d3 x (n f ,`f ,mf x ni ,`i ,mi ) =
6

i
y ni ,`i ,mi ) =
6

x (n f ,`f ,mf

1
d
z ni ,`i ,mi ) =
3
and we will use the abbreviation
Z

x (n f ,`f ,mf

Rif =

Ri (r) rRf (r)r 2 dr(m,1 + m,1 )


Ri (r) rRf (r)r 2 dr(m,1 m,1 )
Z

Ri (r) rRf (r)r 2 dr m,0

Ri (r) rRf (r)r 2 dr

We can now construct the probability that an oriented atom will emit light in the k = kn
direction with polarization states 1 or 2. First we compute
1
i
vk,1 d3 x (n f ,`f ,mf r ni ,`i ,mi ) = Rif ((cos cos ) (m,1 +m,1 )+(cos sin )( )(m,1 m,1 )
6
6
Z

1
sin m,0 )
3

Rif
= ( cos ei m,1 + cos ei m,1 2 sin m,0 )
6
and for the other polarization
vk,2

1
i
d3 x (n f ,`f ,mf r ni ,`i ,mi ) = Rif ( sin (m,1 + m,1 ) + cos ( )(m,1 m,1 )
6
6

Rif
= ( iei m,1 + iei m,1 )
6
If we place a detector to receive light emitted in the n direction, regardless of polarization,
the probability that the transition into the m = 0 state will result in light of any polarization
arriving at the detector is
X

=1,2

00 = |

q
h 2 Rif 2
2m(Ef Ei )

2 sin |2
|
|

(
)|
|
2mi
h
2
6
146

and the probability that the transition into the m = 1 state will result in light of any
polarization arriving at the detector is
X

=1,2

01 = |

q
h 2 Rif 2
2m(Ef Ei )
(
)| | | (| cos ei |2 + |iei |2 )
2
2mi
h

2m(Ef Ei )
q
h 2 Rif 2 1 + cos2
)| | | (
)
(
2mi
2
h
2
6
and for the m = 1 we get the same result
=|

=1,2

0+1 = |

2m(Ef Ei )
q
h 2 Rif 2
)| | | (| cos ei |2 + | iei |2 )
(
2
2mi
h

2m(Ef Ei )
q
h 2 Rif 2 1 + cos2
)| | | (
)
(
2mi
2
h
2
6
If the transitions were from levels or into levels with ` 6= 0, 1, we would of course have different matrix elements, and could end up with more complicated angular dependencies on
our emission probabilities.
=|

The rather good text Experiments in Modern Physics by Melissinos gives the following concrete example, for which we can use our computations above. Consider a nucleus or atom
that decays from one quantum state to another in two stages as shown below.
a

(l,m)=(0,0)
ab transition

(l,m)=(1, m), m=0,-1,+1


bc transition

(l,m)=(0,0)

The intermediate state has three possible orientations with respect to an arbitrary quantization axis, about which we have no knowledge. The first for a b will be emitted
isotropically.
If we set up a detector as illustrated below, the direction between the sample of emitters and
the detector, receiving the first photon, establishes a quantization axis which we take to be
the z-axis. When this detector gets a of energy Ea Eb , it came from a nucleus left with
its angular momentum vector having z-component m = 1 with respect to this axis (since
sin2 0 = 0). Now this nucleus performs its second decay, which must be a m = 1 0 or
m = 1 0, and so has an anisotropic distribution
1 + cos2
() (
)
2

or

()
1 + cos2
)
=
(
( 2 )
2
147

16.2

Angular correlations of emissions

What we have said above applies to an oriented emitter. How can we orient an atom or
nucleus? We could freeze it down to quench rotations and vibrations, and apply a strong
magnetic field. In the nuclear angular correlation experiment, we will use nuclei that
emit two gamma rays. The direction of the first emitted will be used as the z-axis
for a study of the angular distribution of the second .
Cobalt-60 decays to Nickel-60, which in turn emits first a 1.172 M eV and then a 1.333 M eV
, with a delay of only 1 1012 s between these emissions.
Photomultiplier 1

Gamma source
z

Photomultiplier 2

The decays that produce these rays are between high-angular momentum levels, and a detailed computation of the angular distribution of the second photon would yield a theoretical
angular distribution
1
1
() = 1 + cos2 +
cos4
8
24
This could be measured by moving the angular placement of a second photomultiplier tube
from one angular position to another, 10 15 degrees apart, and counting at each location
for equal times.
How do we know that when one of the first emission energy arrives at the fixed counter
(photomultiplier 1), and when the second arrives at the moveable counter, that they cme
from the same nucleus? The signals from both PM tubes are fed into a coincidence circuit,
which only registers simultaneous arrivals. The odds are pretty good that simultaneously arriving s of the right energy are correlated, meaning related; coming from the same
source. The delay for 60 Co is about 1 1012 s, and so the circuit must be tuned to count
coincident s with this time delay between arrivals at the respective PM tubes.

148

16.3

A simulated experiment

I have not performed this experiment myself, but we can easily simulate it to see precisely
what the data should look like.
Apparatus would consist of
1. Coincidence counter.
2. Two photomultiplier tubes, equidistant from a 60 Co source. Tube 1 is fixed, tube 2 is on
a traveller that can be stopped at 10 degree position intervals.
We will need to know the distance of the PM tubes from the source, in order to compute
the angular width of the tubes as subtended from the source location. The PM tubes each
subtend an angle d = 8.0 degrees as measured from the source.
The coincidence circuit will count all events within a certain channel number of the photopeak. This means that it is integrating the signal between one full peak standard deviation on
either side of the peak. The peak occurs at a channel corresponding to energy E = 178 keV ,
and we integrate then from 150 to 201 keV for each PM. The counting time is set at 100 s
for each angular position, and it gives us the following data, the angle is the position of
the moveable PM.
Angle
9.000000
18.000000
27.000000
36.000000
45.000000
54.000000
63.000000
72.000000
81.000000
90.000000
99.000000
108.000000
117.000000
126.000000
135.000000
144.000000
153.000000
162.000000
171.000000
180.000000

PM-1 integration
13471
13315
13173
13105
13254
13420
13286
13146
13465
13249
13454
13451
13417
13217
13422
13269
13260
13086
13221
13555

PM-2 integration
13555
13433
13208
13378
13423
13204
13291
13413
13204
13590
13395
13314
13394
13450
13127
13358
13247
13340
13350
13570

C() coincident counts


14
10
9
11
14
16
20
26
26
36
45
73
90
88
121
199
339
604
1386
2967

The simulation confirms the hypothesis that simultaneously emitted rays are emitted from
the same nucleus, at at 180 degrees apart, consistent with conservation of momentum for
pair annihilation.
The program used for the simulation is listed below. It uses the accept-reject algorithm to
produce a random number sharply peaked (Lorentzian peak) about zero, which is used to
149

create a direction for the second photon once the first photon random direction has been
computed.
/* gamma-gamma correlation simulation */
#include<stdio.h>
#include<stdlib.h>
#include<math.h>
#define PI 3.1415926
#define dtheta 8.0 /* angular width in degrees */
long seed;
double d,dp,theta,del;
int n,m;
/* d, dp are two correlated directions */
/* dtheta is angular width of detectors */
/* theta is moveable detector position */
long c,c1,c2;
/* ci is counter i tally, c is coincidence tally */
double lorentz();
int trigger1(double angle); /* detect if PM is triggered */
int trigger2(double angle, double pos);
/* get a lorentzian random between 0, 360 peaked at 180 with respect to th*/
main(){
seed=1342;
srand(seed);
del=9.0;
/* angular increments for moveable PM */
/* counters set to 0 */
for(m=1;m<=20;m++){
theta=(double)m*del;
c=0;
c1=0;
c2=0;
for(n=0;n<300000;n++){
d=360.0*(double)rand()/(double)RAND_MAX;
dp=180.0+d+lorentz();
while(dp>360.0){dp=dp-360.0;}
/*printf("%f\t%f\t%f\n", d,dp,fabs(d-dp));*/
if(trigger1(d)==1 && trigger2(dp,theta)==1)
c=c+1;
150

if(trigger1(dp)==1 && trigger2(d,theta)==1)


c=c+1;
if(trigger1(d)==1 || trigger1(dp)==1)
c1=c1+1;
if(trigger2(d,theta)==1 || trigger2(dp,theta)==1)
c2=c2+1;
}
printf("%f & %d & %d & %d\\\\\n", theta,c1,c2,c);
}/*end loop over angles */
}
int trigger1(double angle){
if(angle >= 0.0 && angle <= dtheta)
return(1);
else
return(0);
}
int trigger2(double angle, double pos){
if(angle >= pos-dtheta/2.0 && angle <= pos+dtheta/2.0)
return(1);
else
return(0);
}
double lorentz(){
double num,num2,test;
/* get a random angular change between -180, 180 */
/* peaked sharply around 0
*/
start:
num=180.0*(2.0*(double)rand()/(double)RAND_MAX-1.0);
test=100.0/(100.0+num*num);
num2=(double)rand()/(double)RAND_MAX;
if(num2<=test)
return(num);
else
goto start;
}

double y11cor(){
double num,num2,test;
/* get a random angular change between -180, 180 */
/* peaked sharply around 0
*/
start:
151

num=180.0*(2.0*(double)rand()/(double)RAND_MAX-1.0);
test=0.5*(1.0+cos(num)*cos(num));
num2=(double)rand()/(double)RAND_MAX;
if(num2<=test)
return(num);
else
goto start;
}
A second unused routine is provided to generate random angles in degrees with PDF f () =
1
(1 + cos2 ).
2
We now work up our data by producing a polar plot of coincidence versus angle, which
translates into coincidence being the radial coordinate. The therefore plot the points
(x, y) = (() cos(), () sin ) (C() cos(), C() sin )
Which should be a spike directed to the left from the origin.
22

Na correlations

1.0

C() sin()

0.8

0.6

0.4

0.2

0.0
1.0

0.8

0.6
0.4
C() cos()

0.2

0.0

As a good exercise you are urged to modify the simulation to produce data consistent with
the angular correlations of emissions from the Cobalt source. The polar plot should look like
the following.

152

60

Co correlations

C() sin()

2.0

1.5

1.0

0.5

0.0

1.0

0.5

0.0
0.5
C() cos()

1.0

and those for the example used in the computations above, with
1
C() = (1 + cos2 )
2
would look like the following;

dipolar correlations
2.0

C() sin()

1.5

1.0

0.5

0.0

1.0

0.5

0.0
C() cos()

153

0.5

1.0

17

A simulated experiment; Geiger-Marsden experiment

In nuclear and high energy physics, which deal with probing the structure of very small objects, the tool of exploration is scattering. We see the shapes of objects by observing either
light directly emitted by luminous objects, or by seeing light scattered off of the object from
some other source. In order to see into very small objects, light or other particles of very
small de Broglie wavelength must be used. Fairly long ago the technology to produce beams
of particles such as alpha-particles (Helium nuclei) was developed and used to probe into
the interior of atoms, in order to determine how the atom is constructed. The definitive
experiments were conducted by Rutherford and Marsden.
Consider a beam of hard sphere particles colliding with a collection of targets, with a fixed
target density per unit volume. If a projectile can hit a certain area A() around any target
in the entire length L of the target sample, it will be scattered by a scattering angle between
p
and + d. If we have an incoming flow of projectiles at rate Rp = dN
distributed unidt
formly and flowing through an area A, we have a particle beam

N
target

of flux F = RAp . The rate then with which projectiles scatter by angle to +d is determined
by how many will hit the necessary area Ntarget A(). This will be
dN ()
= Rscatt () = F Ntarget A()
dt

154

target

The area A() is called the differential cross section, and is an annular region within the
confines of a certain impact parameter b and b + db, so that A() = 2b db, which we have
projected onto the targets in the two figures above.
Consider classical scattering of particles of mass m, initial speed v0 and impact parameter b
by a static central force field. If all particles with impact parameter in the range b b + db
are scattered into solid angle

b
m

+ d. What is the meaning of the so called scattering cross section? It


represents the size of the target that an incoming projectile must hit in order to be scattered
through an angle between and + d. For a given target then if a projectile hits the
155

ring of area 2b db in the figure, it will scatter through such an angle. Suppose that an
incoming particle beam has a flux of F particles per second per unit area. The rate with
which projectiles scatter into + d is
d2 N
= F 2b db
dt
We translate this from a function of b to a function of with
db
d2 N
= F 2b
d
dt
d
and now find the number of particles per unit time scattered into the cone of outer angle
+ d and inner angle , which subtends a solid angle d = 2 sin d;
db 1
b db
d2 N
= F b
(2 sin d) = F
d
dt
d sin
sin d
and finally we define that quantity loved by particle physicists; the differential cross section
1 d2 N
b db
d()
=
=
d
F dt d
sin d
The trajectory of the particle can be gotten from the energy expression
m 2
(r + r 2 2 ) + V (r)
2
but we can eliminate t in favor of using the angular momentum equation
E=

` = mr 2

d
dt

or by inverting
d
` d
=
dt
mr 2 d
We find that
E=

`2
`2 dr 2
(
)
+
+ V (R)
2mr 4 d
2mr 2

or
d =
Now set u =

1
r

`dr
mr 2

and
1 0 =

2
(E
m

V (r)

u1
u0

2mE
`2

156

du

2mV
`2

`2
)
mr 2

u2

If we let the initial and final values of r be and the point of closest approach be r0 , then
we find that 1 = , 0 = + and = 2 + and
Z

=2

bdu

umax

with the angular momentum given by

V (u)
E

b 2 u2

` = mbv0 = b 2mE
We now specialize to the case of the coulombic potential V (u) = ku and compute the point
of closest approach
1

r0 =

umax
The radial speed is zero at this point and so r = 0 and
E = kumax +

l2 u2max
2m

k 2 + 4b2 E 2
2b2 E
We can now compute the integral for the scattering angle and obtain
umax =

k +

= 2 sin1 (1) 2 sin1 (

k2

k
)
+ 4b2 E 2

which can be solved for b in terms of

k
cot( )
2E
2
which in turn can be inserted into the cross section formula
b=

k2

d()
=
csc4 ( )
2
d
16E
a
An alternative approach that does not involve integral transformation is to use the orbital
equation to find the angle between the asymptotes of the hyperbolic path taken by the projectile.
For Coulombic scattering from a nucleus, V (r) =

k
r

with k > 0,

d
`
k
1
E = m(( r)2 + r 2 ( 2 )2 ) +
2
dt
mr
r
the point of closest approach is,
s

mk
mk
2mE
=
= 2 + ( 2 )2 + 2
rmin
umax
`
`
`
We know that paths will be hyperbolas, such as in the figure
1

157

0
x

max

1
= A B cos
r
with asymptotes at = 0 so use conservation of angular momentum
u=

d
` d
=
dt
mr 2 d
to write

1
d
d
k
1
` dr
`2
k
E = m(( r)2 + (r )2 ) + = m(( 2 )2 + 2 2 ) +
2
dt
dt
r
2
mr d
mr
r
1
again use u = r to get
` du 2
`2
1
) + 2 u2 ) + ku
E = m((
2
m d
m
Insert our hyperbola solution to get
E=

`2
`2 2
B (1 cos2 ) +
(A2 2AB cos + B 2 cos2 ) + k(A B cos )
2m
2m

The coefficient of cos must be zero;


0 = kB
or
A=
and so

B= (

`2
AB
m

mk
`2

mk 2 2mE
) + 2
`2
`
158

and we have our complete path. To get the asymptotes we note that at a point where r ,
u 0 and so at =
0 = A B cos 0
and so

cos 0 =
or

A
1
=q
2
B
1 + 2E`
mk 2

2E`2
mk 2
but the scattering angle is related to the asymtote angle by
tan2 0 =

20 + =
so
`=

mk 2
tan 0 =
2E

mk 2
cot
2E
2

Remember that

1
E = m v02 ,
2
where b is the impact parameter.

` = m v0 b

The cross section derived above applies to one target. Suppose a beam of projectiles is shot
at a target that consists of many nuclei, such as a gold sheet? The number of gold nuclei
per unit volume is , and so the number of targets per unit volume is
ntarget =

NA
M

where M is the atomic mass of gold, and NA is Avagadros number. If the target is gold foil
of cross-sectional area A and thickness L, we have
Ntarget =

NA
AL
M

gold nuclei, which we multiply our cross section computed earlier by, to arrive at
d2 N
db 1
= F Ntarget b
(2 sin d)
dt
d sin
This quantity can be measured experimentally; if we bombard the target for time T we will
have accumulated
db 1
dN = F Ntarget T b
(2 sin d)
d sin
scattered projectiles in the detector positioned to count projectiles scattered into the cone
of solid angle 2 sin d centered on angle .
Rutherford, Marsden and Geiger bombarded thin gold foil with alpha particles. They were
watching for substantial back scattering at angles close to 180 degrees, since the plum pudding model lacked a hard, localized positive scattering center there should be practically
159

no backscattered alphas. They found instead that 1 in 10,000 alphas scatter through an2
gles exceeding 90 degrees, rather than the 1 in e(90) predicted by the plum pudding model.
Rutherford himself was surprised and compared it to firing a 15 inch naval artillary shell
at a sheet of tissue and having the shell bounce back at you. To explain this required all
of the positive charge in the atom to be localized at a single very massive point with the
electrons orbiting in a cloud in sufficient numbers to make the atom neutral. The classical
scattering formula for a pure Coulombic nuclear potential gives very good agreement with
the experimental data.
The details of this very thorough and brilliantly executed experiment are quite remarkable
considering the era in which it was performed. The source used, a sample of purified
radium, was very potent by modern standards; a butt-kicking 0.1 Curie. This is enough to
emit several billion particles per second in the decay process
222

Rn 218 P o +

The detector used was a micrometer-mounted zinc-sulfide screen that would briefly flash
when an struck it. The screen could be moved from one angular position to another, and
was visually monitored with a microscope.
The experimenters studied the scattering rate as a function of incoming particle energy,
scattering angle, and several other factors. Rutherford analyzed the data and confirmed that
it supported the hypothesis that the atom possesses a concentrated and very small nucleus.
Bombardment of a thin film of gold, such as that used by Rutherford, Geiger and Marsden, with particles (42 He nuclei) requires an apparatus very similar to that used in the
Bragg scattering experiment. Chemists often determine the lattice spacings of crystalline
samples using a powder diffraction apparatus, in which a fine powder of the crystal is
bombarded with X-rays, essentially an X-ray version of Bragg scattering. The sample is held
at the center of a flat cylindrical disk container with a removable top plate. The entire inner
wall is lined with photographic film, and X-rays enter through a port. The scattered X-rays
expose the film, and by careful measurements of the positions of exposed areas, the angular
displacements of diffraction maxima can be found.

160

Photographic film liner

s=R

entry
port

sample

Scattering of electrons requires more sophistication, such as in the spherical detector illustrated below, which uses perhaps hundreds of scintillation detectors to count the scattered
electrons. Two colliding beams scatter from the center point of the machine, and the electronics of the scintillation detectors count events and report them versus the scattering angle
cosine.

161

4
5

6
5

3
4

10

Sum to get
counter 1
signal

counts

104

103

102

101

100

50

100
counter

150

Six million particles bombarding total particles incident on a stationary Au nucleus at the
center of the detector produces the following plot of the number of counts dN gotten by the
detectors located at angle , versus cos . Out of 6.0 106 incident particles (total counts),
there were 1976 particles scattered through angles with > 2 = 90o , 3 particles scattered
at angles exceeding 169o ! If the gold atom was well described by the Plum-Pudding Model
162

of Thompsen, this would be completely impossible.

105

dN ()

104

103

102

101

100
1.0

0.5

0.0
cos

0.5

1.0

Plotted against this is the theoretical curve obtained from the assumption that there is a
nucleus, namely
b db
dN () = F
d T Ntarget
sin d
db
d T Ntarget
= 2 F b
d
in which T is the time of exposure (the time for which the sample was bombarded). This
makes F T equal to the total number of projectiles (6.0 106 ) divided by the cross sectional
area of the beam.
You can see that the data conforms extremely well with the hypothesis of a concentrated
nucleus.
In the simulated experiment below you will be asked to analyze a similar data set, and compare it to the theoretical predicted dN () for several models of the atom.

17.1

The numerical experiment

The figure illustrated above is the data analysis of a simulated scattering experiment. The
Rutherford-Geiger-Marsden experiment is a bit too costly to permit its performance on our
163

campus, but we can certainly analyze a generated data set.


Apparatus. We imagine that an apparatus utilizing scintillators such as that on the preceeding page was used, with a detector density of one scintillator per degree in the direction,

.
making d = 180
The incident beam. In this simulation, my incident beam was made by placing a purified
radium sample of 1.0 104 Curies behind a series of collimating aperatures to cut down
the the isotropically emitted particles into a tight beam. The particles are found to have
a kinetic energy of 113, 760 eV , making them deeply nonrelativistic.
Radium sample

Lead shield
Collimators

The runtime of the experiment is T = 60.0 seconds, during which time the scintillators record
6.0 106 events (incident particles), and so Np = 6.0 106 . The machine reports counts
in a manner similar to the operation of a scaler-counter; each event triggers the scintillator
that detects it to report its number (1 180) to a computer, which writes the number to a

file. Counter n has therefore has angular location = n 180


in radians. The data file must
then be analyzed to create a histogram or frequency plot of counts versus cosine of the angle
at which the event was detected.
The target. The target is a very thin gold foil of thickness L = 4.25 1010 m. This is
impractically thin, about ten close-packed atoms deep, but manufacturing techniques have
evolved considerably since the time of Geiger and Marsden. Gold layers a few atoms thick
can be laid down on a substrate by sputtering or ballistic deposition. The density of gold is
kg
g
1.93 104 m
3 , and its atomic mass is 197 mole . Since the gold atoms are close-packed, and
have a radius of about 1 1010 m, no individual is likely to interact with a gold atom
with an impact parameter exceeding 1.0 1010 m.

17.2

Raw data

The actual data is given in the table below. Each scintillator array for angle = n 180
is
counted by an MCA, which tabulates the counts and reports them by channel n. There is
as usual a linear relation between channel n and in this case the angle .

164

17.3

Problem 1. Analysis of simulated lab data

You should now work up this data, and determine which hypothesis it conforms to; a nuclear
atom, or a Thompson model atom. To do so you must first plot the data as dNscatt (), the
number of projectiles scattered into a cone of angle to + d, versus the cosine of the
scattering angle. You should use a log plot on the y-axis because of the tremendous variation
in dNscatt with angle.
Next you should use the information given regarding beam energy, target composition and
thickness, and scattering time, to find each of the following items;
F Ntarget =

Rp NA AL

A
M

Notice that if we integrate


over all scattering angles, since every incident is scattered by
R
some angle, Nscatt = 0 dNscatt () d;
Nscatt = Np = Rp T = b2max F Ntarget T = b2max

Np NA AL

T
AT
M

This should be used to verify the supposed value of bmax . You will also need
b
d,

db
,
d

for the Rutherford model

the angular separation of scintillators

Next you will need to compute the theoretical number of projectiles scattered into each
scintillator using
db
dNscatt () = 2 F b
d T Ntarget
d
for the Rutherford nuclear model, plot the experimental data against the theoretical curve
of dN () versus cos .
You should answer the questions below as well.
1. Explain how you can measure the energy of the particles used in the simulated experiment.
2. Compute the distance of closest approach to a gold nucleus of one of the particles used
in the simulated experiment in a head on collision b = 0.
3. Compute the particle beam rate Rp .

165

Channel
0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41

E = 113760 eV
Counts Channel Counts Channel
474473
42
160
85
3978818
43
153
86
858939
44
137
87
301879
45
137
88
138943
46
139
89
75729
47
105
90
45606
48
126
91
29514
49
96
92
20396
50
116
93
14383
51
81
94
10634
52
91
95
8093
53
72
96
6476
54
77
97
4933
55
73
98
3947
56
76
99
3385
57
70
100
2747
58
65
101
2313
59
50
102
1902
60
52
103
1634
61
47
104
1466
62
53
105
1242
63
51
106
1077
64
34
107
989
65
46
108
893
66
42
109
754
67
29
110
661
68
44
111
584
69
38
112
558
70
36
113
464
71
35
114
432
72
34
115
400
73
42
116
357
74
30
117
314
75
35
118
292
76
29
119
297
77
34
120
281
78
24
121
253
79
26
122
222
80
23
123
184
81
31
124
179
82
21
125
165
83
18
126

166

Counts
16
12
14
15
13
17
10
14
13
16
10
11
10
11
16
8
4
7
12
4
10
7
11
9
10
5
6
8
6
7
5
5
7
8
2
6
7
5
5
3
9
5

Channel
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
152
153
154
155
156
157
158
159
160
161
163
164
165
166
167
169
170
171

Counts
6
6
5
1
3
6
4
5
2
5
5
3
3
8
3
3
3
2
3
3
2
5
3
1
2
1
2
1
2
1
1
1
2
1
1
2
2
1
1
1
1
1

17.4

The Thompson model

Demonstrate using Gausss law that within the Thompson atom, the force exerted on an
particle due to the positive charge alone is
F=

2Ze2
(xi + yj)
40 R3

R
b

1
The simply blows the electrons, 8000
times its mass, away, and so interacts mostly with
the positively charged part of the atom. If the incident
with impact parameter b < R and
velocity v = v0 i enters the atom at point r(0) = R2 b2 i + b j, it will be pushed away
by the force above. Show that the solutions to the Newtonian equations of motion are

v0

r(t) = ( sinh t R2 b2 cosh t)i + b cosh t j

in which

2Ze2
40 m R3
giving the position vector of the after entering the atom. This is only valid as long as
|r(t)| R, since after the leaves the atom, it experiences no force, and so travels in a
straight line.
Prove that the time at which the leaves the atom is

2v0 R2 b2
tanh( t` ) =
v2
(R2 + 0 )
=

and demonstrate that the scattering angle is given by

2b R2 b2
vy (t` )
= 2
tan =
vx (t` )
v0 (R2 2b2 )
167

Demonstrate that the electrons embedded in the atom will execute elliptic orbits, with the
centers of the ellipses being at the center of the atom, rather than the focal points being at
the center of the atom, as is the case in planetary motion.
The final phase is to establish a reasonable value for R. It was thought in 1912 that the
atom was about R = 1.0 1010 m in radius.
If we suppose that R = 1.0 1010 m, then for gold (Z = 79)
Ze2
= 1137.6 eV
40 R
Geiger and Marsden used particles of about 5.5 M eV , which is very high. In our simulated
experiments we will use much lower energies.
Show that for the atomic size of R = 1.0 1010 m,
tan =

2 Rb
E
1137.6 eV

b2
R2

(1

2b2
)
R2

For the energy of 113760 eV , compute the deBroglie wavelength. You can see that this is
certainly small enough to reveal details within the atom, given its estimated size.

17.5

Monte Carlo scattering simulation

One of the most commonly performed numerical experiments today in high energy physics is
the Monte Carlo simulation of scattering. Runtime on big experimental hardware is expensive, so experiments are first designed by simulation before hardware is built or modified, and
runtime is paid for. Typically one generates scattering events that obey the dynamics of some
theoretical model, and precise projectile parameters such as impact parameter are simulated
using random number generation. You will now run a Monte Carlo for the Thompson model.
The actual Monte Carlo portion of the simulation is the beam generation. To simulate a
beam of spatially uniform flux, consider a phosphorescent screen placed perpendicular to the
beam. Projectile impacts should be randomly and uniformly distributed, as illustrated below.

168

1.0

0.5

0.0

0.5

1.0
1.0

0.5

0.0

0.5

1.0

This uniform beam was simulated by choosing two randomly and uniformly distributed
numbers between 1 and 1, one to act as x, one
to act as y. The PDFs of these numbers
are both 21 . The resulting impact parameter b = x2 + y 2 is randomly distributed, but not
uniformly;
Z a
Z
a2
dx dy
b db d
=
=
[b a] =
4
4
8
ba
0
and so b has an unnormalized PDF of
fb (b) =

d
b
[b0 b] =
0
db
4

Suppose that Rb is a randomly distributed variable on the interval from 0 to 1 with such a
PDF. What is the probability that an will be scattered by an angle greater than 2 radians?
This would require that the denominator be negative.
Write a simple computer program to generate 600, 000 scattering events with such a random
beam of particles.
1. Generate a uniform random deviate r, interpreted as Rb , with a normalized PDF based
on that given above. The easiest way
be to generate two random deviates x and y,
2would
2
both between 1 and 1, and if r = x + y 1, proceed, otherwise pick two more and try
again. You could use the inverse method or the accept-reject algorithm instead.
2. Compute the appropriate scattering angle from the Thompson scattering formula. Do
this for several particle energies.
3. Compute which scintillator would detect the scattered from n = f loor(180 ) where
f loor is the function that rounds a floating point number down to the nearest integer. Print
out this integer.
169

4. Now that you have a data file of detector events, create a frequency plot of the number
of counts gotten by each detector. Does this look anything like that seen by Geiger and
Marsden?

17.6

Energy dependence in the Geiger-Marsden experiment

Geiger and Marsden were very thorough. They also investigated the dependence of the
scattering rate on incident particle energy.
Beginning with
dN () = 2 F Ntarget T b db
and

k
cot
2E
2
we can determine the energy dependence of the scattering rate on incident energy by counting
the number of particles that should be scattered into angles exceeding 2 radians. These
must be incident on gold nuclei with impact parameters less than
b=

bmax =

k
k

cot =
2E
4
2E

and so the total number of scattering events logged by all detectors located at >
be
Z bmax
k2

2 F Ntarget T b db = F Ntarget T b2max = F Ntarget T


N> 2 =
4E2
0

should

This inverse-energy-squared behaviour is almost universal in high energy physics scattering


experiments. In this problem you will verify the hypothesis that the scattering rate has this
behaviour as a function of incident energy.
On the next few pages you will find the detector counts for several incident particle energies. The data for E = 113760 eV was displayed several pages back. The radium source
emits particles with a variety of energies. One energy can be singled out by passing the
beam through a velocity selector; crossed electric and magnetic fields. This sub-beam can
then be accelerated further with a high voltage.
For each energy for which you have the scattering data, determine the number of particles scattered through more than 2 radians. Plot this versus the energy. Now compute the
number of such events predicted by the Rutherford formula, given above, four our beam and
target parameters. In each experiment, we have logged 6.0 106 scattering events.
Plot the experimental counts and the theoretical curve on the same figure, compute 2 for
the four data points. Is the hypothesis verified to be true?

170

Channel
0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34

Counts
4453047
1160804
214677
75138
34747
18774
11402
7330
5061
3524
2701
2098
1626
1249
1017
828
692
564
529
438
341
299
274
239
224
207
170
150
122
118
128
99
81
81
70

E = 227520 eV
Channel Counts Channel
35
67
71
36
61
72
37
55
73
38
60
74
39
54
75
40
49
76
41
50
77
42
52
78
43
42
79
44
35
80
45
39
81
46
35
82
47
25
83
48
28
84
49
21
85
50
18
86
51
16
87
52
18
88
53
19
89
54
16
90
55
15
91
56
18
92
57
11
93
58
15
95
59
11
96
60
14
97
61
16
98
62
5
99
63
8
100
64
7
101
65
8
102
66
9
103
67
10
104
68
11
105
69
9
106

171

Counts
6
6
6
8
5
6
4
4
9
4
3
5
7
2
8
3
1
9
4
5
1
10
1
2
4
4
3
4
2
2
4
3
4
1
4

Channel
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
127
130
131
132
133
134
135
137
138
139
142
143
147
149
151
154
155
159
162
163

Counts
3
2
3
1
2
2
1
2
1
2
2
2
3
1
3
2
1
1
1
1
1
1
1
1
1
2
1
1
2
2
1
1
1
1
1

Channel
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44

Counts
474793
2776956
1202451
556203
302768
183063
118825
81131
57792
42979
32755
25560
20046
16380
13175
11103
9286
7731
6606
5644
4953
4323
3802
3499
2979
2590
2342
2033
1993
1748
1514
1500
1301
1221
1069
988
875
906
789
759
698
664
586
596

Channel
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88

E = 56880 eV
Counts Channel
508
89
520
90
493
91
433
92
406
93
368
94
368
95
334
96
324
97
317
98
267
99
279
100
263
101
238
102
247
103
224
104
204
105
195
106
183
107
195
108
177
109
194
110
163
111
165
112
145
113
147
114
122
115
112
116
125
117
112
118
113
119
119
120
90
121
102
122
100
123
95
124
112
125
81
126
82
127
98
128
66
129
99
130
61
131
68
132
172

Counts
69
58
61
63
51
79
49
56
48
44
45
44
50
35
41
43
39
26
41
40
35
40
33
37
34
37
30
36
29
31
31
23
22
24
13
20
17
15
19
15
18
19
13
15

Channel
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176

Counts
16
18
5
10
13
12
15
12
13
11
9
11
6
13
6
11
10
6
11
8
9
6
7
8
7
8
10
5
5
6
5
6
2
1
4
2
2
3
1
4
2
1
2
1

Channel
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46

Counts
476108
1569314
1210099
729550
472849
324276
231905
172040
130800
101969
81012
65076
53825
44150
37007
31211
26514
23014
20010
17451
15256
13624
11972
10545
9500
8568
7665
6969
6444
5712
5277
4840
4427
3981
3712
3411
3199
2947
2759
2522
2343
2326
2025
2066

Channel
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90

E = 28440 eV
Counts Channel
1840
91
1768
92
1597
93
1474
94
1461
95
1320
96
1239
97
1176
98
1166
99
1125
100
1044
101
966
102
885
103
924
104
848
105
822
106
742
107
752
108
681
109
617
110
620
111
565
112
617
113
563
114
538
115
540
116
466
117
493
118
460
119
441
120
400
121
444
122
393
123
350
124
389
125
374
126
339
127
322
128
303
129
272
130
320
131
260
132
260
133
266
134
173

Counts
268
222
236
219
216
214
217
202
175
182
167
175
171
179
185
151
158
153
137
123
125
120
121
118
122
103
120
103
108
111
88
120
81
98
82
83
75
82
68
97
76
60
61
57

Channel
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178

Counts
69
52
62
50
55
57
53
56
54
54
53
51
44
37
34
27
29
31
29
28
26
22
22
20
24
22
17
19
13
20
19
13
16
13
16
10
12
7
5
4
4
6
3
1

17.7

Problems

1. Consider a simplified atomic model in which the positive charge +Ze is spread out into
a uniform ball of radius R, and the Z electrons are embedded in this ball in such a way
that the electrostatic potential outside of the ball is zero, and inside is V0 > 0. Incoming
particles of impact parameter b and speed v0 will slow down once they get into the ball,
to speed v1 . The angular momentum will be conserved, requiring that the beam bend
sharply at r = R as illustrated below.

s
b

The impact parameter of the bent beam is s.


a. If the incoming beam has angle of incidence with respect to the normal of the surface
of the ball, and has kinetic energy exceeding V0 , the transmitted beam will be bent to angle
with respect to the normal. Use conservation of energy and angular momentum together
with the figure to show that
sin
v1
b
=
=
sin
v0
s
b. Use this to find the scattering angle as a function of b. Compute the maximum scattering angle as a function of m ( particle mass), b, and V0 . Perhaps you can see now why
Rutherford was so surprised when particles literally bounced back into his face.
c. Find the scattering angle in the case of 21 mv02 < V0 .
174

2*. For the case 21 mv02 > V0 in the first problem, find the differential cross section
is a slightly more challenging problem.

d
.
d

This

3. For a stream of 6.0 M eV alpha particles incident on a gold foil of thickness 1 106 m,
determine the rate at which particles are scattered through angles greater than 0.1 radian.

17.8

Notes

The data was generated using energies


E =

k
Rc

with Rc = 0.2 1010 meters for the 113760 electron volt . Maximum impact parameter
was in practicality around 1 1010 m. This makes
F Ntarget T
(

k2
=
4E2

6 106
(1.93 104 )(6.023 1023 )(4.25 1010 ) Rc2
)(60 s)
= 1.182 1026 Rc2
60 s
(0.197)
4

which will give 118, 473, 1890, and 7563 particles scattered by more than 2 radians for
Rc = 0.01, 0.02, 0.04 0.08 1010 m which gives the four energies used to generated the
energy dependence data. Physically counting in each case we find 116, 471, 1885 and 7540,
so that the simulations are quit good.
/* Monte Carlo that creates data for Thompson atom scattering */
#include <stdlib.h>
#include <math.h>
#include <stdlib.h>
#define PI 3.1415926
int n;
main(){
float m1,m2,test,max,b,theta,R_cl, E,E0,num,den;
int crap,det;
/* seed the random number generator, establish MAX=2^{31} */
srand(17);
max=pow(2.0,31.0);
/* constants and the energy */
E=11000.0;
E0=1137.6;
for(n=0;n<600000;n++){
175

/* here is the Monte Carlo step */


generate:{
m1=2.0*(float)rand()/max-1.0;
m2=2.0*(float)rand()/max-1.0;
b=sqrt(m1*m1+m2*m2);
if(b>1.0) goto generate;
/* now create the event */
num=2.0*b*sqrt(1.0-b*b);
den=(E/E0)-(1.0-2.0*b*b);
theta=atan2(num,den);
/* print out detector logging event */
det=floor(180.0*theta/PI);
printf("%d\n", det);
}
}
Below is the program to generate data for the theoretical Rutherford scattering, using
E=

k
0.02 1010

cos 2
sin
=
2
sin4 2
sin3 2
db
d F T Ntarget
d
(1.93 104 )(6.023 1023 )(4.25 1010 )
6 106
(0.197)
180
2b

= 2 (

k
k
4 0.0210
10

= 2

1
1
)2 cot
2
2 sin2

4
23
10
(0.02 1010 )2 sin
)
6 (1.93 10 )(6.023 10 )(4.25 10

10

4
44
(0.197)
180
sin 2

= 2 (37.62)

sin
sin4 2 180

#include<stdio.h>
#include<math.h>
#define PI 3.1415926
int n;
float dtheta,theta,sig;
main(){
dtheta=1.0;
for(n=1;n<180;n++){
theta=(float)n*dtheta*PI/180.0;
sig=sin(theta/2.0);
sig=sig*sig;
176

sig=sig*sig;
sig=2.0*PI*(37.62)*sin(theta)*(PI)/(180.0*sig);
printf("%f\t%f\n", cos(theta),sig);
}
}
Summing over all output values, we should come up close to six million. The Monte Carlo
program to create the experimental data is below;

/* creates data for rutherford scattering */


#include <stdlib.h>
#include <math.h>
#include <stdlib.h>
#define PI 3.1415926
int n;

main(){
float m1,m2,test,max,b,theta,R_cl;
int crap,det;
srand(17);
max=pow(2.0,31.0);
R_cl=0.02;
for(n=0;n<6000000;n++){
generate:{
m1=2.0*(float)rand()/max-1.0;
m2=2.0*(float)rand()/max-1.0;
b=sqrt(m1*m1+m2*m2);}
if(b>1.0) goto generate;
theta=2.0*atan2(R_cl,2.0*b);
/*printf("%f\n", theta); */
/* modify to print out detector number */
det=floor(180.0*theta/PI);
printf("%d\n", det);
}
}

177

18
18.1

Appendix
Appendix I. Reformatting Lab Data with Sed and Awk

Sed and awk are two Unix editting tools that are particularly useful in processing data taken
in lab. Awk is programmable and uses the syntax of C and shell programming.
If you have little or no programming experience, and it is difficult for you to use the support
programs included in the manual, try to use Awk to process your data. It is easy to learn,
and there are lots of Awk scripts strewn throughout the manual. Awk is a free download for
Win32 machines from the physics department website, and is part of any Linux installation.
As an example we will process data for the Nuclear Counting Statistics experiment.
Suppose that we count nuclear decays for 10 second periods, obtaining the following 20
measurements
1016
1207
1186
1244
1110
1099
1099
1185
1220
1286
1117
1280
1190
1083
1177
1189
1200
1201
1188
1291
which we enter into a file called list , one number per line as shown above, so the command
cat list
produces the output above. We need to find the largest and smallest values, easy for 20 data
points, but what if we had hundreds? We can sort the data using the Unix program sort,
with the n switch telling sort to sort by arithmetic value.
sort -n < list > sorted_list
which looks like this
178

1016
1083
1099
1099
1110
1117
1177
1185
1186
1188
1189
1190
1200
1201
1207
1220
1244
1280
1286
1291
This file will have a first line that is blank, a feature of some versions of sort, this means that
the second line of the file sorted list and the last lineare the smallest and largest entries
in our data file. Check for this behaviour when running sort for the first time .We can pipe
the output of this program into an awk script min max.awk that will read the file and
report the second and last records to the console. We need to tell awk that it will read a file
whose records are single lines in column format, one record per line. The record separator
RS will be a blank line, beginning with a null character, the field separator FS is a newline
character.
The normal operation mode of awk is that awk acts on each line of a file in turn, treating
each line as a record with several data fields, each separated by a field separator symbol that
is user defined. The fields are accessed by refereing to them by the names $1, $2, and so
forth. The last data field is called $NF. Here is the awk script min max.awk
#min_max.awk
BEGIN { FS = "\n"; RS = ""}
{ print "smallest is " , $2, " " , " largest is", $NF}
We can obtain the largest and smallest numbers in our data set by running sort on the file
list and piping the output into this awk script
sort < list | awk -f min_max.awk
which produces the output line
.
Smallest is

1016

largest is 1291
179

The next stage in processing the data for this experiment is to sort the data into bins of a
given width in count-space. The range of counts is
Nmax Nmin = 1291 1016 = 275
We can sort our data into 11 bins of width 25 counts by running the data in list through the
following awk script called bin sort.awk
#bin_sort.awk
{
i=0
while ( i<11 ) {
if ( $1 >= 1016+25*i && $1 <= 1016+25+25*i )
print i, " ", $1
++i}
}
This will run through the loop for each line of the file list in turn, assign each data point to
a bin, and print the bin first, then the data point separated by a blank space. The output of
awk -f bin_sort.awk list
is
0
7
6
9
3
3
3
6
8
10
4
10
6
2
6
6
7
7
6
10

1016
1207
1186
1244
1110
1099
1099
1185
1220
1286
1117
1280
1190
1083
1177
1189
1200
1201
1188
1291

We can sort this data by running


180

awk -f bin_sort.awk list | sort -n


which will produce the sorted output below. Note that sort will perform numeric sorting
based on the first field of each data line.
0
2
3
3
3
4
6
6
6
6
6
6
7
7
7
8
9
10
10
10

1016
1083
1099
1099
1110
1117
1177
1185
1186
1188
1189
1190
1200
1201
1207
1220
1244
1280
1286
1291

We can do even better by telling bin sort.awk to print only the bin number of each data
point and piping the output through sort into the uniq command that will list the number
of occurances of each bin label in the list, and output two columns, bin population followed
by bin number.
#bin_sort.awk
{
i=0
while ( i<11 ) {
if ( $1 >= 1016+25*i && $1 <= 1016+25+25*i )
print i
++i}
}
We run the command
awk -f bin_sort.awk list | sort -n | uniq -c
which produces output

181

1
1
3
1
6
3
1
1
3

0
2
3
4
6
7
8
9
10

We can save this data to a file called bin pop with


awk -f bin_sort.awk list | sort -n | uniq -c >bin_pop
and print out the contents of the file in a nice format that can be imported into a LaTeX
lab report as a table
awk { print $2, "&", $1, "\\\\", "\\hline"} bin_pop
which produces
0 & 1 \\ \hline
2 & 1 \\ \hline
3 & 3 \\ \hline
4 & 1 \\ \hline
6 & 6 \\ \hline
7 & 3 \\ \hline
8 & 1 \\ \hline
9 & 1 \\ \hline
10 & 3 \\ \hline
We now add a few lines to this and we have our processed data in the form of a nice LaTeX
table.
\begin{tabular}{|c|c|}\hline
Bin label & Bin population \\
\hline
0 & 1 \\ \hline
2 & 1 \\ \hline
3 & 3 \\ \hline
4 & 1 \\ \hline
6 & 6 \\ \hline
7 & 3 \\ \hline
8 & 1 \\ \hline
9 & 1 \\ \hline
10 & 3 \\ \hline
\end{tabular}
182

and this prints as seen below


Bin label Bin population
0
1
2
1
3
3
4
1
6
6
7
3
8
1
9
1
10
3
We can include another column for the center of the bin with
awk -f table.bin bin_pop
using the script
#table.awk
{ print $2, "&", $1, "&", 1016+12+25*$2, "\\\\", "\\hline"}
which produces the output
0 & 1 & 1028 \\ \hline
2 & 1 & 1078 \\ \hline
3 & 3 & 1103 \\ \hline
4 & 1 & 1128 \\ \hline
6 & 6 & 1178 \\ \hline
7 & 3 & 1203 \\ \hline
8 & 1 & 1228 \\ \hline
9 & 1 & 1253 \\ \hline
10 & 3 & 1278 \\ \hline
After adding afew LaTeX lines this becomes the table
Bin Bin pop. Bin center
0
1
1028
2
1
1078
3
3
1103
4
1
1128
6
6
1178
7
3
1203
8
1
1228
9
1
1253
10
3
1278
Awk and sed can not only be used to easily process huge data files for analysis, but also to
prepare data processed into convenient reports.

183

18.2

Appendix II; Nonlinear Least Squares error matrix method

Suppose that we have a set of data points with average values and standard deviations
{
xi i |I = 1, 2, N }
taken in the lab. These are graphed with error bars versus a parameter upon which they
depend, for example consider scattering data, in which the number of counts gotten by a
scintillation counter placed at position 1 is x1 (1 ) 1 . Suppose that the theoretical count
number versus is a function
N () = a0 + a1 cos + a2 cos2
or that we wish to fit our data to such a theoretical curve, and must find the coefficients ai .
This can be done by finding the best fit to such a function using a minimization process,
called nonlinear least squares regression.
Our experimental data is of the form
xi (i ) i
The theoretical values corresponding to these experimental values will be
i = a0 + a1 cos i + a2 cos2 i
and in general have the form
i =

n
X

Cim am

m=1

if there are n constants ai involved in the theoretical curve expression.


Suppose that the likelihood of xi being the measured value is given by a Gaussian distribution
L=

1
q

2i2

(
xi i )2
2 2
i

then the probability that the entire set of experimental data actually being found to have
the values gotten in the experiment will be
L=

1
N

(2) 2 1 N

PN

i=1

(
xi i )2
2i

We can maximize this probability with respect to the parameters ai , which will compute the
parameters describing the theoretical form most closely agreeing with our data out of all
possible theoretical curves in the same family.
Maximizing with respect to the parameters {ai } leads to the equations
X
i

(
xi i ) i
=0
i2
am
184

or
X
i

X Cim Cil
Cim
xi =
al
2
i
i2
i,l

We can define the vector of experimental results


Xm =

X
i

and the error matrix


Mml =

X
i

Cim
xi
i2
Cim Cil
i2

and then we find


X = Ma
which has solution
a = M1 X
Suppose that hundreds of identical experiments have given us the opportunity to compute the
set of parameters a many times, and we obtain a set of average values for these parameters

a, What is the standard deviation in the values of these fit parameters?


(am a
m ) =

X
k

k ) =
(M 1 )mk (Xk X

(M 1 )mk

kj

Cjk
(xj xj )
j2

however if we assume that the measurements resulting in each data point are uncorrelated,
then
< (xj xj )(xi xi ) >= i2 ij
and so
< (am am )(al a
l ) >=
=

(M 1 )mk

ijkp

Cjk
Cip
(M 1 )lp 2 i2 ij
2
j
i

(M 1 )mk (M 1 )lp Mpk = (M 1 )ml

kp

and so the standard deviation in am is


am = q

1
(M 1 )mm

So by finding the matrix M we essentially find the curve of best fit and its tolerances.
Example
Consider the following data table

185

x
0

f(x)
1.1

0.1

2.9

0.1

8.9

0.1

19.1 0.1

We propose a theoretical form


f (x) = a0 + a1 x2
The theoretical values of f(x) should be then
1 = a 0
2 = a 0 + a 1
3 = a0 + 4a1
and
4 = a0 + 9a1
We find the matrix Cim is

Cim =
the data vector elements are
X1 =

1
1
1
1

0
1
4
9

2.9
8.9
19.1
1.1
+
+
+
= 3200
0.01 0.01 0.01 0.01

X2 =

2.9
8.9
19.1
+4
+9
= 21040
0.01
0.01
0.01

and the error matrix is

Mlm =

This has inverse


(M

0.005
0.0007143
0.0007143 0.0002041

100 100 100 100


0 100 400 900

0
1
4
9

)lm =

400 1400
1400 9800

186

1
1
1
1

We arrive at the following parameters


a1
a2

0.005
0.0007143
0.0007143 0.0002041

3200
21040

0.9711
2.0085

and so the function


f (x) = 0.9711 + 2.0085x2
best fits our lab data. Furthermore we know

a1 = 0.005 = 0.071, a2 = 0.000204 = 0.01428


are the errors in the determination of these best fit parameters.

18.3

Appendix III. The Plotting of Lab Data with GNU plotutils

This suite of graphics utilities can be used to build publication quality graphs from raw data.
It consists of the command-line programs graph, plot, spline, ode and several others. The
use of graph is the subject of this section.
Graph can output is work into a variety of formats and devices, chosen with the commandline option -T. For example
graph -T X data
will output the resulting figure into xplot, an X windows primitive graphing canvas. To
output to pnm, xfig, Postscript, or GIF, one just passes this choice to graph at invocation
and redirects the resulting data into a file, for example
graph -T ps data > data.ps
Read through the voluminous /usr/local/info/plotutils.info file for more command line options, such as those needed to limit the abcissa and ordinates xmin x xmax, ymin
y ymax
graph -T ps -x xmin xmax -y ymin ymax data > data.ps
or plotting several data files on the same figure
graph -T ps data1 data2 data3

> data.ps

or perhaps using a logarithmic x-axis within domain 100 x 104


graph -T ps -l x -x 0 4 data > data.ps
One can add labels to the x and y axes as well as a banner label
graph -T ps -X x, m -Y Volts -L Voltage per meter data > data.ps
The labels can have Greek letters or math symbols, which are given later in a table in this
document
187

graph -T ps -L \*Q\sp*\ep(x,t)\*Q(x,t) for t=0.1 s data > data.ps


will print (x, t)(x, t)f ort = 0.1s along the top of the figure.
Graph has the capacity to produce graphs within graphs, or to make multigraphs consisting
of many repositioned plots. For example the command
graph -T ps --reposition 0.0 0.6 0.3 -L "a=0.07" 07 --reposition 0.3 0.6 0.3
-L "a=0.09" 09 --reposition 0.6 0.6 0.3 -L "a=0.11" 11 --reposition 0.0 0.3
0.3 -L "a=0.13" 13 --reposition 0.3 0.3 0.3 -L "a=0.15" 15 --reposition 0.6
0.3 0.3 -L "a=0.17" 17 --reposition 0.0 0.0 0.3 -L "a=0.19" 19 --reposition
0.3 0.0 0.3 -L "a=0.21" 21 --reposition 0.6 0.0 0.3 -L "a=0.23" 23 >page1.ps
will create the following multiplot
a=0.07

a=0.09

700

600

600

500

500

a=0.11
500

400

400
300

400
300
300

200
200

200

0
15

100

100

100

10

10

15

0
15

10

a=0.13

10

15

0
10

a=0.15

500

500

400

400

300

300

200

200

100

100

10

a=0.17
400

300

200

100

0
8

0
8

a=0.19

0
6

a=0.21

500

a=0.23

400

400

300

300

200

200

100

100

400

300

200

100

0
6

0
6

0
6

The fact that these programs are command-line means that they can be used in shell-scripts
or batch files to perform bulk processing of graphics files.
For plotting discrete data points the switch S allows you to use a large variety of symbols
for individual data points
188

Symbol number
1
2
3
4
5

Symbol

+
*

with 31 different symbols in all. For example


graph -T ps -S 4 0.01 data > data.ps
uses circles whose size is 0.01 times the size of the box within which the entire plot is drawn.
Plots can be rotated by 90, 180, 270 degrees with rotation
graph -T ps --rotation 90 > data.ps
Lines connecting data points in plots can be manipulated with -m, line mode 0 is no connecting line. Line style, width, and even color can be manipulated. For example we use
-m to change linemode from solid to dotted or dashed, but adding -C changes this to color
linemode rather than breaking the line in different ways. We can place a plot within a plot
by repositioning the second plot within the first and shrinking it;
graph -T ps -C -m 2 data1 --reposition 0.5 0.5 0.3 data2>data.ps
results in
1.0
5

0.5

0.0

0.5

1.0

or we can try
graph -T ps -C -m 3 data1 --reposition 0.5 0.5 0.1 data2>data.ps
which produces
189

1.0

0.5
5

0.0

0.5

1.0

The reposition is relative to the point (0.0, 0.0) being the lower left corner of the physical
display, and (1.0, 1.0) being the upper right corner. In the first example a miniature virtual
display of size 0.3 will be place with its lower left corner at (0.5, 0.5).
These few examples are enough to get started with in using graph, and you really should
read through the info file to learn more since this barely scratches the surface.
One of the most useful features of libplot is the generous collection of font symbols that
it recognizes for drawing strings on plots. These include Greek letters and mathematical
symbols. For example
graph -T X -L "P\sb2\eb(cos\*h)"
will result in the string P2 (cos ) being drawn on the plot title or label. A partial table of
useable symbols is given below
symbol

6
=

escape seq. symbol


\a

\D

\m

\H

\U

\z

\h

\q

\+

\cu

\li
h

\es

\! =

escape seq.
\b
\e
\n
\R
\W
\d
\p
\if
\pt
\gr
\hb
\c+
\ >=

symbol

190

escape seq. symbol escape seq.


\G

\x
\F

\L
\y

\P
\S

\t
\C

\Q
\g

\s
\f

\l
\<

\ >
\pd

\ca
R
\SU
\is
\dg

\wp
\c

\Ah

\ <=
\sr

Some of these symbols are only available if pl fontname has been used to set the font to
Hershey or HersheySans. Subscripts and superscripts can be created using the delimiters
illustrated by the following example;
graph -T X -X"\*Q\sp*\ep\\sbn,l,m\eb(r,\*f,\\*h)"
writes n,l,m (r, , ) to the x-axis label,
graph -T ps -L" L\sp\*a\ep\sb n\eb(r)" > out.ps
writes Ln (r) to the label. This PostScript output is being redirected to a file.

18.4

Appendix IV. Preparing a lab report with LaTeX

LaTeX is very easy to learn and produces documents of superior quality. Nearly every scientific textbook and journal is typeset in TeX or LaTeX. LaTeX has the further advantage
of being platform independent. In addition it is free software available in binary format for
all existing operating systems. It is installed on several of our lab computers. The support
software for this manual contains templates for all of the labs, and the data analysis programs will output LaTeX tables and formatted equations.
LaTeX is designed to typeset equations and tables, and it excels at these tasks. The basic
math equations are typeset in math mode in which offset, centered equations are surrounded with double dollar signs, the math mode delimiters. The language is best learned
by example, and so below you will see the LaTeX commands and corresponding typeset
equations for a wide variety of mathematics.
$$ \int_{-a}^{\infty} f(x) dx= F(a)$$
Z

f (x)dx = F (a)

$$ \oint_C {f(z) \over z-z_0 } dz = 2 \pi i f(z_0) $$


I

f (z)
dz = 2if (z0 )
z z0

$$ \vec{A} \times \vec{B} = \vec{C}

$$

~B
~ =C
~
A
$$\vec{\nabla} \cdot \vec{E} = {\rho \over \epsilon_0}
~ E
~ =

0
191

$$

$$ \gamma \rightarrow \beta \bar{\beta} $$



$$ \oint \vec{p} \cdor d\vec{q}= 2 \pi n \hbar $$
I

p~ d~q = 2n
h

$$ \sum_{n=0}^{\infty} {1 \over n^s} = \zeta{s} $$

1
= s
s
n=0 n
$$ \prod_{n=1}^{\infty} (1-{x^2 \over n^2 \pi^2})={\sin x \over x} $$

n=1

(1

x2
sin x
)=
2
2
n
x

$$ \lim_{n \rightarrow \infty} (1-{x \over n})^n = e^{-x}$$


lim (1

x n
) = ex
n

$$ P(x)={1 \over \sqrt{2 \pi \sigma^2}}e^{-{(x-\bar{x})^2 \over 2\sigma^2}}$$

P (x) =

(x
x)2
1
e 22
2 2

$$ \ln |x| = \int_{1}^{x} {dy \over y}$$

ln |x| =
$$ \alpha

\beta

\gamma

\delta

\pi

x
1

dy
y

\lambda

\epsilon

\sigma

\omega $$


$$ 1={2 \over \pi}[\sin x +{1 \over 3}\sin 3x +{1 \over 5}\sin 5x +\cdots] $$

192

1
1
2
[sin x + sin 3x + sin 5x + ]

3
5
This shows a pretty full range of LaTeX math typesetting commands.
1=

Below is the basic template for the LaTeX documentstyle article. Type this into a file
to use as a template for all of your lab reports.

\documentclass[12pt]{article}
\pagenumbering{arabic}
\usepackage[dvips]{graphics}
\begin{document}
\title{ Your Title Here }
\author{
author 1 \\
Department of Physics,
900 Wood Road, Kenosha
\and
author 2 \\
Department of Physics,
900 Wood Road, Kenosha
}

University of Wisconsin-Parkside \\
WI \\

University of Wisconsin-Parkside \\
WI \\

\begin{abstract}
Here is the very short description of your experiment.
\end{abstract}
\maketitle
\noindent
{\bf Section 1.}\\
\\
Describe the equipment used, draw a diagram of the apparatus using
xfig and convert it to a postscript figure. It can be included in
the document with the code line \\
\rotatebox{270}{\scalebox{0.5}[0.5]{\includegraphics{apparatus.ps}}}\\
\noindent
{\bf Procedure.}\\
\\
Describe the experimental procedure.
193

Be to the point, brief and go step by step.

\noindent
{\bf Data.}\\
\\
Your data must be in tabular form. There is a handout on using awk
and sed to process your data files and put them into LaTeX table format.
Here is a typical 4 column table including some multicolumn rows.
\begin{tabular}{|c|c|c|c|}\hline
\multicolumn{2}{|c|}{\bf Trial } & \multicolumn{2}{|c|}{\bf Trial 2}\\
\hline
Bin & Number & Bin & Number \\
\hline
1 & 3 & 5 & 9 \\
\hline
2 & 4 & 6 & 3 \\
\hline
3 & 7 & 7 & 1 \\
\hline
4 & 1 & 8 & 2 \\
\hline
\end{tabular}
Notice that an ampersand is used to separate
the table fields and a LaTeX newline \\
ends each line in the table.
\noindent
{\bf Analysis.}\\
\\
This is the section in which you perform sample computations
and analyse your data. Any graphs generated can be included
in this section with
\rotatebox{270}{\scalebox{0.5}[0.5]{\includegraphics{../ps/graph.ps}}}\\
It may not be necessary to include the rotate box command here.
This is the place where your error analysis and propagation is performed
and your results compared to the accepted (correct) values.

\noindent
{\bf Conclusions.}\\
194

\\
In this section draw your conclusions.
\end{document}
To format your lab report, type it up using this model and call it report.tex. Run
LaTeX on it from the command line in the directory the file lives in
latex report.tex
If there are no errors, you will be left with a new file called report.dvi. Convert this to
postscript with
dvips -o report.ps report.dvi
The result of this is report.ps which can be printed. If there are errors the LaTeX engine
will stop on the offending line and point out the error. It wiil prompt you, respond by typing
X
which exits LaTeX. Open your report in the editor and fix the error. Usual LaTeX errors
are unpaired dollar signs and other simple things, easily remedied.
It is well worth your while to become adept with LaTeX if you plan on pursuing Physics
or any technical field. An excellent book is A Guide to LaT eX2 by Helmut Kopka and
Patrick W. Daly, Addison-Wesley, second edition 1995.
Incidentally, what does the table example in the lab template look like?
Bin
1
2
3
4

Trial
Number
3
4
7
1

18.5

Trial 2
Bin Number
5
9
6
3
7
1
8
2

Appendix V. The Oscilloscope

This is a device used to analyse electronic waveforms and signals. To measure a signal one
does the following.
1. Apply the signal to the Input terminals of the vertical deflection amplifier.
2. Horizontal display is controlled by the triggering sections on the face of the scope. For
External triggering, sweep is initiated whenever a preset voltage level is reached, in this case
by an externally supplied source, applied to the Trigger in terminals. For Internal triggering,
Trigger signal is the same as the input vertical signal, the wave being observed triggers itself.
Line setting triggers from the 60 Hz power line that powers the scope. Auto mode means
that the scopes internal sawtooth generator is used to initiate the sweep. As this voltage
195

increases, the horizontal trace moves from left to right on the screen until a preset value is
reached, then the sweep starts over.
In all but Auto modes, the voltage at which the sweep is initiated is set by the Stability
level control, which can select the voltage with a positive or negative slope, and an AC or
DC portion of the input signal.
3. Horizontal sweep rate can be set with the Sweep time
control
cm
4. The y(x) relationship between two applied voltages y(t), x(t) can be determined by setting Horizontal sweep to External and applying x(t) to the Horizontal input terminal and
y(t) to the Vertical input terminal. This bypasses the sawtooth generator and horizontal
deflection is controlled by x(t) If the two input signals are harmonic with frequencies in a
rational ratio, a Lissajous figure will be displayed on the screen.
Trigger Mode
Line
Int
Auto

Ext
Trigger in

Dc

Stability
+

AC

Focus

Intensity

Horizontal
sweep time/cm
0.5 0.2
1
0.1

Scale illum.
msec
2
AC off

Vertical
V/cm
Input
1 0.5

0.1

sec
Vert. Pos.

50
20

5
10

18.6

5
10

10
20

50

1 5

mV

0.1 Ext

V/cm

2
5
10

20
10
5 sec

1
0.5
0.2
0.1

input

Horiz. Pos.

Appendix VI. Support software

The program below chi2 table gen.c will generate 2 in tabular format extremely quickly
using the methods of Gaussian quadrature. It is limited to no more than about 25 degrees
of freedom.

196

#include<stdio.h>
#include<math.h>
#include<stdlib.h>
#include<plot.h>

float gammaf(float x);


float chi2function( float aa, int DOF);
int DF,n,number;
float high, low,dchi,chi,P;
main(int argc, char *argv[]){
if(argc < 5 || argc > 5){
printf("chi2_table_gen low high number degree_of_freedom \n");
printf("low is lower bound, high is upper bound (floats).\n");
printf("number is how many table entries to generate\n");
printf("between low and high.\n");
printf("degree_of_freedom is an integer <= 20.\n");
exit(0);
}
low=atof(argv[1]);
high=atoi(argv[2]);
number=atoi(argv[3]);
DF=atoi(argv[4]);

if(low<=0.0){
printf("Invalid lower bound\n");
exit(0);}
if(high < low){
printf("Invalid upper bound\n");
exit(0);}
if(DF<1 || DF > 20){
printf("Invalid number of degrees of freedom\n");
exit(0);}
dchi=(high-low)/(float)number;
printf("Generating $\\chi^2$ probabilities $\\{1\\over 2^{{%d\\over 2}} \\Gamma({%d\\ov
printf("for $%f\\le x \\le %f$\n\n", low,high);
printf("\\begin{tabular}{|c|c|}\\hline\n");
printf(" $\\chi^2_d$ & $\\wp(\\chi^2\\ge\\chi^2_d)$\\\\\n");
printf("\\hline\n");
for(n=0;n<=number;n++){
chi=low+(float)n*dchi;
P=chi2function(chi,DF);
printf("%f & %f\\\\\n",chi,P);
197

}
printf("\\end{tabular}\n");
}
The program requires the following (called chi2.c ) subroutine to compute the perform
the quadrature.

/* computation of chi^2 statistic by Gaussian */


/* quadrature. Uses Laguerre polynomial roots */
/* accuracy is limited to 4-5 decimal places */
/* use GMP lib to improve the accuracy
*/
#include<stdio.h>
#include<math.h>
#define PI 3.1415926
float gammaf(float x);
float chi2function( float aa, int DOF){
/* a=experimental chi^2, N= degrees of freedom */
/* good up to N=14 but is best well below that point */
int m,p;
float sum,denominator,exponent;
float w[15];
float y[15];
/* the abscissa points */
y[0]=9.3307812017e-2;
y[1]=4.926917403e-1;
y[2]=1.215595412;
y[3]=2.26994953;
y[4]=3.66762272;
y[5]=5.42533663;
y[6]=7.56591623;
y[7]=10.12022856;
y[8]=13.13028248;
y[9]=16.65440771;
y[10]=20.77647890;
y[11]=25.62389423;
y[12]=31.40751917;
y[13]=38.53068331;
y[14]=48.02608557;
/* the weight factors */
w[0]=2.18234886e-1;
w[1]=3.42210178e-1;
w[2]=2.63027578e-1;
198

w[3]=1.26425818e-1;
w[4]=4.02068649e-2;
w[5]=8.56387780e-3;
w[6]=1.21243614e-3;
w[7]=1.11674392e-4;
w[8]=6.45992676e-6;
w[9]=2.22631691e-7;
w[10]=4.22743038e-9;
w[11]=3.92189727e-11;
w[12]=1.45651526e-13;
w[13]=1.48302705e-16;
w[14]=1.60059491e-20;
exponent=(float)DOF/2.0-1.0;
denominator=pow(2.0, exponent)*gammaf((float)DOF/2.0);
sum=0.0;
for(m=0;m<15;m++){
sum=sum+w[m]*pow(2.0*y[m]+aa, exponent );
}
return(exp(-aa/2.0)*sum/denominator);
}
Also required is the following rather inefficient program for the gamma function gammaf.c

#include<math.h>
#define PI 3.1415926
float gammaf(float x)
{
/* The Gamma function */
float sum, prefactor;
float b[9],arg[9];
int m;
b[0]=1.00;
b[1]=-0.577191652;
b[2]=0.988205891;
b[3]=-0.897056937;
b[4]=0.918206857;
b[5]=-0.756704078;
b[6]=0.482199394;
b[7]=-0.193527818;
b[8]=0.035868343;
199

sum=1.0;
arg[0]=1.0;
prefactor=1.0;
if(x<0.0)
/* this is the only recursive call (1) */
return(PI/(sin(PI*x)*gammaf(1-x)));
if(x>2.0){
do{
x=x-1.0;
prefactor=prefactor*x;}
while(x>2.0);}
if(x>0.0 && x<1.0){
prefactor=1.0/x;
x=x+1.0;}
for( m=1;m<=8;m++){
arg[m]=(x-1)*arg[m-1];
sum=sum+b[m]*arg[m];}
return(prefactor*sum);
}

everything can be compiled with the command


gcc chi2.c gammaf.c chi2_table_gen.c -o chi2_table_gen

200

-lm

Index
accelerating voltage, 53
accept-reject, 13, 175
activity, 22
adiabatic process, 40
alpha particle, 20
Aluminum crystal, 62
angular correlation, 145
angular momentum, 36
Awk script, 24

fcc structure, 63
Fermi Golden Rule, 145
Fermis Golden Rule, 143
fig, 187
four-momentum, 126
Franck-Hertz experiment, 83
Geiger counter, 20
Geiger tube, 6, 20
GIF, 187
gnuplot, 25
graph, 187

background radiation correction, 24


Bargmann-Fock space, 143
bcc structure, 63
beam-splitter, 37
Bequeral, 22
beta decay, 20
Binomial theorem, 41
binsort.c, 16
Bohr levels, 89
Bragg condition, 55, 63
Bragg scattering, 53
Bravais lattice, 61
bulk modulus, 41

half-life, 5, 21
Helium-Neon laser, 36
Hershey fonts, 191
high pressure gas tank, 49
hypergeometric function, 28
interpolaion, 122
ionization potential, 83
ionizing radiation, 6
isotopes, 5
K capture, 36

characteristic curve, 6
chi squared gen.c, 16
Cobalt-60, 148
coincidence circuit, 148
Compton edge, 127
Compton scattering, 126
conventional cell, 63
Coulomb energy, 36
Curie, 22

lattice spacing, 57, 62


Laue condition, 63
linemode, 189
linetype, 189
lissojou figure, 43
longitudinal waves, 40
markers, 188
Miller indices, 63
molecular weight, 42

de Broglie relation, 126


deBroglie wavelength, 56
degree of freedom, 47
density of states, 111
dipole selection rule, 145

Neon, 36
neutrino, 20
Newtons method, 80
Normal distribution, 13
nuclear spin, 36
null measurement, 16

EM wave, 111
equipartition theorem, 47, 112
escape sequences, 190

oscilloscope, 43, 85
201

photoelectric effect, 69
Planck apparatus, 69
Planck radiation law, 114
pnm, 187
Poisson distribution, 17
Poissonian distribution, 8
polarization, 145
Postscript, 187
primitive cell, 62
radiation, 5
radio-isotope generator, 23
Rayleigh-Jeans law, 112
reposition, 188
rotational degrees of freedom, 45
Rydberg constant, 89
scaler-counter, 5
scaler-timer, 20
scattering planes, 53
selection rules, 145
simulation, 13, 65, 150, 175
Sodium, 36
space lattice, 61
specific heat, 41
speed of sound, 41
Stefan-Boltzmann law, 114
stopping voltage, 71
symbols, 190
T versus R program, 122
translation vectors, 61
unbiased estimator, 17
wave quation, 111
Weins law, 114
work function, 70

202

You might also like