Вы находитесь на странице: 1из 182

LEVEL 5 ELECTROMAGNETIC THEORY

AND RFID APPLICATIONS


Part 1: Electromagnetic theory - A simple
introduction
Peter H. Cole
July 11, 2010

Contents
1 INTRODUCTION
1.1 Introduction . . . . . . . . . . . .
1.2 Course Aims . . . . . . . . . . . .
1.3 Administrative Details . . . . . . .
1.4 Relation to Other Physical Theories
1.4.1 Relativity . . . . . . . . . .
1.4.2 Quantum mechanics . . . .
1.5 Subject Development . . . . . . .

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

2 MATHEMATICAL BACKGROUND
2.1 Notation . . . . . . . . . . . . . . . .
2.1.1 Scalar quantities . . . . . . .
2.1.2 Vector quantities . . . . . . .
2.1.3 Calligraphic characters . . . .
2.2 Co-ordinate systems . . . . . . . . .
2.2.1 Scope . . . . . . . . . . . . .
2.2.2 Cylindrical polar co-ordinates
2.2.3 Spherical polar co-ordinates .
2.3 The Field Concept . . . . . . . . . .
2.4 Source and Vortex Fields . . . . . . .
2.4.1 Source-type fields . . . . . . .
2.4.2 Vortex-type fields . . . . . . .
2.5 Names and Units for Variables . . .
2.6 A Modest Proposal . . . . . . . . . .
2.7 Vector Algebra . . . . . . . . . . . .
2.7.1 Scope . . . . . . . . . . . . .
2.7.2 Scalar product . . . . . . . .
2.7.3 Vector product . . . . . . . .
2.8 Vector Integrals . . . . . . . . . . . .
2.8.1 Line integrals . . . . . . . . .
2.8.2 Surface integrals . . . . . . .
2.8.3 Contour integrals . . . . . . .
2.8.4 Closed surface integrals . . . .
2.9 Representation of Fields . . . . . . .
2.10 Inverse Square Central Fields . . . .
2.11 Properties of Conservative Fields . .
2.12 Flux and Circulation . . . . . . . . .
i

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

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

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

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

.
.
.
.
.
.
.

1
1
1
2
2
2
3
3

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

5
5
5
6
7
8
8
8
8
9
10
10
11
11
12
12
12
12
13
14
14
14
16
16
16
18
18
20

ii

CONTENTS
2.12.1 Flux of a vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.12.2 Circulation of a vector . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.13 A Sneak Preview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3 ELECTROSTATICS IN EMPTY SPACE


3.1 Introduction . . . . . . . . . . . . . . . . . . . . .
3.2 Experimental Observations . . . . . . . . . . . .
3.2.1 Static electricity . . . . . . . . . . . . . .
3.2.2 Conservation of charge . . . . . . . . . . .
3.2.3 Coulombs Law . . . . . . . . . . . . . . .
3.3 The Electrostatic Field . . . . . . . . . . . . . .
3.4 Superposition . . . . . . . . . . . . . . . . . . . .
3.5 Representation of Electric Fields . . . . . . . . .
3.6 Illustration of Some Fields . . . . . . . . . . . . .
3.6.1 An isolated charge . . . . . . . . . . . . .
3.6.2 Two equal charges . . . . . . . . . . . . .
3.6.3 An electric dipole . . . . . . . . . . . . .
3.7 The Electric Dipole . . . . . . . . . . . . . . . .
3.7.1 Electric field pattern . . . . . . . . . . . .
3.7.2 Parameter definition . . . . . . . . . . . .
3.7.3 Dipole field analysis . . . . . . . . . . . .
3.7.4 Torque on a dipole . . . . . . . . . . . . .
3.7.5 Induced dipoles . . . . . . . . . . . . . . .
3.7.6 Force on a dipole . . . . . . . . . . . . . .
3.8 Electric Flux and Flux Density . . . . . . . . . .
3.8.1 Introduction . . . . . . . . . . . . . . . .
3.8.2 Definition of flux density . . . . . . . . .
3.8.3 Definition of electric flux . . . . . . . . .
3.9 Gauss Law for the Electric Flux . . . . . . . . .
3.9.1 Objective . . . . . . . . . . . . . . . . . .
3.9.2 Point charge at the centre of a sphere . .
3.9.3 A more general surface . . . . . . . . . .
3.9.4 Case of several charges . . . . . . . . . .
3.9.5 Applications of Gauss theorem . . . . . .
3.9.6 Definition of a Gaussian surface . . . . .
3.9.7 How we proceed . . . . . . . . . . . . . .
3.9.8 Flux from a line charge . . . . . . . . . .
3.9.9 Flux from a surface charge on an insulator
3.9.10 Flux from a surface charge on a metal . .
3.9.11 Solid metallic conductor . . . . . . . . . .
3.9.12 Hollow metallic conductor . . . . . . . . .
3.10 Electrostatic Potential . . . . . . . . . . . . . . .
3.10.1 General observations . . . . . . . . . . . .
3.10.2 Definition . . . . . . . . . . . . . . . . . .
3.10.3 Potential of an isolated charge . . . . . . .
3.10.4 Equipotential surfaces . . . . . . . . . . .
3.11 Capacitors and Capacitance . . . . . . . . . . . .

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

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

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

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

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

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

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

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

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

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

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

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

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

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

23
23
23
23
25
26
27
28
28
28
29
29
29
31
31
31
31
32
32
33
34
34
34
34
35
35
35
35
37
37
38
38
38
39
40
41
41
42
42
43
44
44
46

CONTENTS
3.11.1 Concept of a capacitor . . . . .
3.11.2 Definition of capacitance . . .
3.11.3 Calculation of capacitance . . .
3.11.4 Energy stored in a capacitor .
3.11.5 Capacitors in parallel . . . . .
3.11.6 Capacitors in series . . . . . .
3.12 Energy Stored in an Electrostatic Field
3.13 Static Electricity Machines . . . . . .
3.13.1 Introduction . . . . . . . . . . .
3.13.2 Electrical discharge from points
3.13.3 Van der Graa generator . . .

iii
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

4 ELECTROSTATICS IN MATERIAL BODIES


4.1 Introduction . . . . . . . . . . . . . . . . . . .
4.2 Experimental Observations . . . . . . . . . . . .
4.3 Model of a Dielectric . . . . . . . . . . . . . . .
4.4 Charge Movement in Dielectrics . . . . . . . .
4.4.1 Classification of charges . . . . . . . . .
4.4.2 Definition of polarisation . . . . . . . .
4.4.3 Induced charge crossing a plane . . . . .
4.4.4 Non-polar atoms . . . . . . . . . . . . .
4.4.5 Polar atoms . . . . . . . . . . . . . . . .
4.4.6 Induced charge entering a closed volume
4.4.7 Induced charge at a boundary . . . . .
4.5 Flux and Flux Density . . . . . . . . . . . . . .
4.5.1 Flux density in free space . . . . . . . .
4.5.2 Flux density in a dielectric . . . . . . .
4.5.3 Flux in free space . . . . . . . . . . . .
4.5.4 Flux in a dielectric . . . . . . . . . . . .
4.6 Gauss Law With Dielectrics . . . . . . . . . .
4.6.1 Classification of charges . . . . . . . . .
4.6.2 Gauss law in free space . . . . . . . . .
4.6.3 Continued applicability . . . . . . . . .
4.6.4 Conduction charge form . . . . . . . . .
4.6.5 Traps for the unwary . . . . . . . . . .
4.7 Characterisation of Dielectric Media . . . . . .
4.7.1 General discussion . . . . . . . . . . . .
4.7.2 A word of warning . . . . . . . . . . . .
4.7.3 Dielectric susceptibility . . . . . . . . .
4.7.4 Dielectric permittivity . . . . . . . . . .
4.7.5 Dielectric constant . . . . . . . . . . . .
4.8 Analysis of Capacitor . . . . . . . . . . . . . .
4.9 Stored Energy in Dielectrics . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

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

.
.
.
.
.
.
.
.
.
.
.

46
49
49
49
50
50
51
52
52
52
54

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

55
55
56
56
58
58
59
59
60
61
61
62
62
62
62
62
63
63
63
64
64
64
65
65
65
66
66
66
67
67
68

iv
5 CURRENT AND RESISTANCE
5.1 Introduction . . . . . . . . . . . .
5.2 Charge Transport in Conductors .
5.2.1 Introduction . . . . . . . .
5.2.2 Electron drift . . . . . . .
5.2.3 Electron mobility . . . . .
5.2.4 Volume current density . .
5.2.5 Conductivity . . . . . . .
5.2.6 Resistivity . . . . . . . . .
5.2.7 Ohms law . . . . . . . . .
5.3 Lumped Circuit Concepts . . . .
5.3.1 Introduction . . . . . . . .
5.3.2 Total current . . . . . . .
5.3.3 Potential dierence . . . .
5.3.4 Ohms law again . . . . .
5.4 Power Dissipation . . . . . . . . .
5.4.1 Lumped circuit form . . .
5.4.2 Field form . . . . . . . . .
5.5 Resistors in Combination . . . . .
5.5.1 Resistors in series . . . . .
5.5.2 Resistors in parallel . . . .

CONTENTS
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

71
71
71
71
72
72
72
74
74
74
74
74
74
75
75
75
75
76
76
76
77

6 MAGNETOSTATICS IN EMPTY SPACE


6.1 Introduction . . . . . . . . . . . . . . . . . .
6.2 Experimental Observations . . . . . . . . . .
6.2.1 Early history . . . . . . . . . . . . .
6.2.2 Forces between poles . . . . . . . . .
6.2.3 Exploring the Field . . . . . . . . . .
6.2.4 Dipolar behaviour . . . . . . . . . . .
6.2.5 Iron filings experiment . . . . . . . .
6.2.6 Magnetic eects of a current . . . . .
6.2.7 Field classification . . . . . . . . . .
6.2.8 Other eects . . . . . . . . . . . . . .
6.3 Measurement of Magnetic Field . . . . . . .
6.3.1 Torque on compass needle . . . . . .
6.3.2 Torque on a small loop . . . . . . . .
6.3.3 Electron ballistics . . . . . . . . . . .
6.3.4 Hall eect . . . . . . . . . . . . . . .
6.3.5 Faraday induced emf . . . . . . . . .
6.3.6 Summary . . . . . . . . . . . . . . .
6.4 Fields Produced by Currents . . . . . . . .
6.4.1 Introduction . . . . . . . . . . . . . .
6.4.2 Field of a long straight wire . . . . .
6.4.3 The law of Biot and Savart . . . . .
6.4.4 Field of a Complete Circuit . . . . .
6.4.5 Caution . . . . . . . . . . . . . . . .
6.5 Amperes Law of Magnetostatics . . . . . .

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

79
79
79
79
80
80
81
82
82
83
84
84
85
85
85
85
85
86
86
86
86
87
88
88
88

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

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

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

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

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

CONTENTS

6.6

6.7

6.8

6.9

6.5.1 Introduction . . . . . . . . . . .
6.5.2 The simple case . . . . . . . . .
6.5.3 The general case . . . . . . . .
Magnetic Flux and Flux Density . . . .
6.6.1 Magnetic flux density . . . . . .
6.6.2 Magnetic flux . . . . . . . . . .
Magnetic Forces . . . . . . . . . . . . .
6.7.1 Force on a moving charge . . .
6.7.2 Force on a current element . . .
6.7.3 Small loop in uniform field . . .
6.7.4 Force between parallel wires . .
Some Magnetic Field Distributions . .
6.8.1 Interior of current carrying wire
6.8.2 A long solenoid . . . . . . . . .
6.8.3 Toroidal coil . . . . . . . . . . .
6.8.4 Centre of circular loop . . . . .
6.8.5 Axis of circular loop . . . . . .
Eects of Magnetic Forces . . . . . . .
6.9.1 Introduction . . . . . . . . . . .
6.9.2 Hall eect . . . . . . . . . . . .
6.9.3 Circulating charges . . . . . . .
6.9.4 The cyclotron . . . . . . . . . .

v
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

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

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

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

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

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

7 MAGNETOSTATICS IN MATERIAL BODIES


7.1 Introduction . . . . . . . . . . . . . . . . . . . . .
7.2 Comparison of Dielectric and Magnetic Media . .
7.2.1 Points of similarity . . . . . . . . . . . . .
7.2.2 Points of dierence . . . . . . . . . . . . .
7.3 The Origins of Magnetism . . . . . . . . . . . . .
7.4 Summary . . . . . . . . . . . . . . . . . . . . . .
7.5 Magentisation . . . . . . . . . . . . . . . . . . . .
7.6 Magnetic Field Produced by Magnetisation . . . .
7.6.1 Illustration . . . . . . . . . . . . . . . . .
7.7 Magnetic Flux and Flux Density . . . . . . . . .
7.7.1 Flux density in free space . . . . . . . . .
7.7.2 Flux density in a magnetic medium . . . .
7.7.3 Magnetic flux in free space . . . . . . . .
7.7.4 Magnetic flux in a magnetic medium . . .
7.7.5 Reasons for definition . . . . . . . . . . . .
7.8 Characterisation of Magnetic Media . . . . . . .
7.8.1 General discussion . . . . . . . . . . . . .
7.8.2 A word of warning . . . . . . . . . . . . .
7.8.3 Magnetic susceptibility . . . . . . . . . .
7.8.4 Magnetic permeability . . . . . . . . . . .
7.9 Large Signal Behaviour of Ferromagnetic Media .
7.9.1 Linear media . . . . . . . . . . . . . . . .
7.9.2 Ferromagnetic media . . . . . . . . . . . .

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

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

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

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

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

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

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

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

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

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

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

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

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

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

88
89
89
90
90
91
91
91
91
91
92
93
93
95
96
96
97
97
97
97
98
99

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

101
. 101
. 102
. 102
. 102
. 102
. 103
. 104
. 104
. 105
. 107
. 107
. 107
. 107
. 107
. 107
. 108
. 108
. 108
. 108
. 109
. 109
. 109
. 109

vi

CONTENTS
7.9.3 Measurement of M - H curves . . . . . . . . . . . . . . . . . . . . 110
7.10 The Source Free Nature of B . . . . . . . . . . . . . . . . . . . . . . . . . . 110

8 ELECTRODYNAMICS
8.1 Introduction . . . . . . . . . . . . . . . . . . . .
8.2 Electromagnetic Induction . . . . . . . . . . . .
8.3 Conductor Moving in a Field . . . . . . . . . . .
8.4 A further use of Lenzs law . . . . . . . . . . . .
8.5 Inductance . . . . . . . . . . . . . . . . . . . . .
8.5.1 Introduction . . . . . . . . . . . . . . . .
8.5.2 Mutual inductance . . . . . . . . . . . .
8.5.3 Expression in terms of Faradays law . .
8.5.4 The sign from Lenzs law . . . . . . . . .
8.5.5 Self inductance . . . . . . . . . . . . . .
8.5.6 The induced voltage . . . . . . . . . . .
8.5.7 Circuit expressions of inductance . . . .
8.6 Stored Energy in an Inductance . . . . . . . . .
8.6.1 Introduction . . . . . . . . . . . . . . . .
8.6.2 Context . . . . . . . . . . . . . . . . . .
8.6.3 Establishing an inductor current . . . . .
8.6.4 Power input . . . . . . . . . . . . . . . .
8.6.5 Stored energy in field form . . . . . . . .
8.6.6 Generalisation . . . . . . . . . . . . . . .
8.7 Measurments on Magnetic Materials . . . . . .
8.7.1 Method . . . . . . . . . . . . . . . . . .
8.7.2 Consequences for electromagnetic theory
8.8 Depolarising and Demagnetising Factors . . . .
8.9 Maxwells Contribution . . . . . . . . . . . . . .
8.9.1 Introduction . . . . . . . . . . . . . . . .
8.9.2 Displacement Current Concept . . . . .
8.9.3 Resolution . . . . . . . . . . . . . . . . .
8.10 The Complete Laws . . . . . . . . . . . . . . . .
8.10.1 Faradays law . . . . . . . . . . . . . . .
8.10.2 Amperes law as modified by Maxwell . .
8.10.3 Gauss law for the electric flux . . . . . .
8.10.4 Gauss Law for the magnetic flux . . . .

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

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

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

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

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

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

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

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

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

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

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

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

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

A STANDARD INTERNATIONAL TERMINOLOGY AND UNITS


A.1 Objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A.2 Informal discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A.3 The Base Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A.4 Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A.5 The Fundamental Vectors . . . . . . . . . . . . . . . . . . . . . . . . .
A.6 List of Standard Symbols . . . . . . . . . . . . . . . . . . . . . . . . . .

113
113
113
115
117
118
118
118
120
120
120
121
121
123
123
123
124
124
125
126
127
127
128
129
129
129
129
130
131
131
131
131
131

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

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

.
.
.
.
.
.

133
. 133
. 133
. 134
. 136
. 136
. 137

CONTENTS

vii

B USEFUL PHYSICAL CONSTANTS


B.1 Physical Constants . . . . . . . . . .
B.2 Material Conductivities . . . . . . . .
B.3 Relative Dielectric Permittivities . .
B.4 Relative Magnetic Permeabilities . .
C REFERENCES
D ADVICE ON STUDY FOR
D.1 Use of these lecture notes
D.2 Advice on units . . . . . .
D.3 Other Advice . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

139
139
139
140
140
141

EXAMINATIONS
143
. . . . . . . . . . . . . . . . . . . . . . . . . . . 143
. . . . . . . . . . . . . . . . . . . . . . . . . . . 143
. . . . . . . . . . . . . . . . . . . . . . . . . . . 143

E EXERCISE SET 1

145

F EXERCISE SET 2

149

G EXERCISE SET 3

153

H EXERCISE SET 4

157

EXERCISE SET 5

161

J EXERCISE SET 6

165

K SOLUTIONS TO EXERCISE SET 1

169

L SOLUTIONS TO EXERCISE SET 2

177

M SOLUTIONS TO EXERCISE SET 3

183

N SOLUTIONS TO EXERCISE SET 4

193

O SOLUTIONS TO EXERCISE SET 5

203

P SOLUTIONS TO EXERCISE SET 6

209

viii

CONTENTS

List of Figures
2.1
2.2
2.3
2.4
2.5
2.6
2.7
2.8
2.9
2.10
2.11
2.12
2.13
2.14
2.15
2.16

Polar representation of a complex number. . . . . . . . .


Cylindrical polar co-ordinate system. . . . . . . . . . . .
Spherical polar co-ordinate system. . . . . . . . . . . . .
Illustration of a source type field. . . . . . . . . . . . . .
Illustration of a vortex type field. . . . . . . . . . . . . .
Illustration of a scalar product. . . . . . . . . . . . . . .
Illustration of a vector product. . . . . . . . . . . . . . .
A line integral for a vector field. . . . . . . . . . . . . .
A surface integral for a vector field. . . . . . . . . . . .
A closed contour integral. . . . . . . . . . . . . . . . . .
A closed surface integral. . . . . . . . . . . . . . . . . .
Two methods of representing a field. . . . . . . . . . . .
Vector diagram to determine a resultant field. . . . . . .
Three dimensional aspect of the Faraday representation.
Flux passing through a surface. . . . . . . . . . . . . . .
Circulation around a contour. . . . . . . . . . . . . . . .

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

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

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

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

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

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

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

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

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

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

6
8
9
10
11
13
13
15
15
16
17
19
19
20
21
21

3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
3.9
3.10
3.11
3.12
3.13
3.14
3.15
3.16
3.17
3.18
3.19
3.20

Pith ball repelled by a charged rod. . . . . . . . . . . .


Pith balls in attraction and repulsion. . . . . . . . . . .
Charging by induction. . . . . . . . . . . . . . . . . . .
Point charges in free space . . . . . . . . . . . . . . . .
Application of superposition in electric field calculations.
Electric field of an isolated charge . . . . . . . . . . . . .
Electric field of two equal charges . . . . . . . . . . . . .
Electric field of a dipole . . . . . . . . . . . . . . . . . .
Parameters of an electric dipole . . . . . . . . . . . . . .
Analysis of a dipole field . . . . . . . . . . . . . . . . . .
Calculation of torque on a dipole . . . . . . . . . . . . .
A naturally created dipole . . . . . . . . . . . . . . . . .
Point charge at centre of spherical surface. . . . . . . .
A general closed surface surrounding a point charge. . .
Equality of flux over two surface elements. . . . . . . . .
Flux from a line charge. . . . . . . . . . . . . . . . . . .
Flux from a surface charge on an insulator. . . . . . . . .
Flux from a surface charge on a metal. . . . . . . . . . .
Solid metallic conductor. . . . . . . . . . . . . . . . . . .
Hollow metallic conductor. . . . . . . . . . . . . . . . . .

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

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

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

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

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

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

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

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

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

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

24
24
25
26
28
29
30
30
31
32
33
33
35
36
36
38
39
40
41
42

ix

LIST OF FIGURES
3.21
3.22
3.23
3.24
3.25
3.26
3.27
3.28
3.29
3.30
3.31
3.32

Path of integration for a simple calculation


Equipotentials for an isolated charge. . . .
Equipotentials for a dipole field. . . . . . .
Equipotentials for another field. . . . . . .
Equipotentials for a uniform field. . . . .
Concept of a capacitor. . . . . . . . . . . .
A parallel plate capacitor. . . . . . . . . .
Capacitors in parallel. . . . . . . . . . . .
Capacitors in series. . . . . . . . . . . . .
Electrical discharge from a sharp point. .
Modelling the eect of a sharp point. . .
Van der Graa Generator. . . . . . . . .

of the potential.
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .

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

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

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

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

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

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

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

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

44
45
46
46
47
47
48
50
51
53
53
54

4.1
4.2
4.3
4.4
4.5
4.6
4.7
4.8
4.9
4.10

Capacitor with dielectric . . . . . . . . . . . . . . . . . . . .


A non-polar atom . . . . . . . . . . . . . . . . . . . . . . . .
Non-polar atom behaviour in an internal field . . . . . . . .
An elementary dipole . . . . . . . . . . . . . . . . . . . . . .
Polar atom behaviour in an internal field . . . . . . . . . . .
Plane in polarised dielectric . . . . . . . . . . . . . . . . . .
A rectangular volume in a non-uniformly polarised dielectric
A general volume in a non-uniformly polarised dielectric . .
Induced charge at dielectric surface . . . . . . . . . . . . . .
Fields and charges in a capacitor with dielectric . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

56
57
57
57
58
60
61
61
62
67

5.1
5.2
5.3

Conducting bar with an electric field . . . . . . . . . . . . . . . . . . . . . 73


Resistors in series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
Resistors in parallel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

6.1
6.2
6.3
6.4
6.5
6.6
6.7
6.8
6.9
6.10
6.11
6.12
6.13
6.14
6.15
6.16
6.17
6.18
6.19

A bar magnet. . . . . . . . . . . . . . . . . . . . . . . . . . .
Field lines near a bar magnet. . . . . . . . . . . . . . . . . .
Division of a bar magnet. . . . . . . . . . . . . . . . . . . . .
Iron filings experiment. . . . . . . . . . . . . . . . . . . . . .
Exploring field lines near a long straight wire. . . . . . . . .
Map of field lines near a long straight wire. . . . . . . . . . .
Co-ordinates for Biot and Savart Law. . . . . . . . . . . . .
Magnetic field lines surrounding a long straight wire. . . . .
Small current carrying loop in a magnetic flux density. . . .
Long straight parallel wires. . . . . . . . . . . . . . . . . . .
Wire with uniformly distributed current. . . . . . . . . . . .
Magnetic field variation with distance from wire centre. . . .
Calculation of interior field of a long solenoid. . . . . . . . .
Coil on non-magnetic toroidal core. . . . . . . . . . . . . . .
A circular current carrying loop. . . . . . . . . . . . . . . . .
Position on the axis of a circular current carrying loop. . . .
Deflection of charge carriers in a magnetic flux. . . . . . . .
Circular motion of charge in a uniform magnetic flux density.
Construction of the cyclotron. . . . . . . . . . . . . . . . . .

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

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

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

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

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

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

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

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

80
81
81
82
83
83
87
89
92
93
94
94
95
96
96
97
98
98
99

LIST OF FIGURES

xi

7.1 Depolarising and demagnetising fields in dielectric and magnetic media . . 105
7.2 Magnetisation vs magnetic field inside a ferromagnetic medium . . . . . . . 110
7.3 A north pole in a magnetic material . . . . . . . . . . . . . . . . . . . . . . 111
8.1
8.2
8.3
8.4
8.5
8.6
8.7
8.8
8.9
8.10
8.11
8.12
8.13
8.14
8.15

Electromagnetic induction by changing currents . . . . .


Electromagnetic induction by moving magnetic material
Conductor moving in a magnetic field . . . . . . . . . . .
Induced source in series with bar. . . . . . . . . . . . . .
Complete circuit with a moving conductor. . . . . . . . .
Illustration of the use of Lenzs law. . . . . . . . . . . . .
Illustration of mutual inductance concept . . . . . . . . .
Illustration of self inductance concept . . . . . . . . . . .
Expressions of Faradays and Lenzs laws. . . . . . . . . .
Current voltage relations in R, L and C. . . . . . . . . .
Circuit for establishing an inductor current. . . . . . . .
A toroidal inductor. . . . . . . . . . . . . . . . . . . . . .
Apparatus for investigation of M - H curves. . . . . . .
Triangular current waveform. . . . . . . . . . . . . . . .
Displacement current in a capacitor. . . . . . . . . . . .

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

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

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

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

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

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

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

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

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

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

114
114
115
116
117
118
119
121
122
123
124
126
127
128
129

E.1 Representation of field lines. . . . . . . . . . . . . . . . . . . . . . . . . . . 146


E.2 Representation by array of vectors. . . . . . . . . . . . . . . . . . . . . . . 146
E.3 Five Gaussian surfaces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
F.1
F.2
F.3
F.4
F.5
F.6

Four closed surfaces enclosing charge.


Long uniformly charged conductor. .
Charged slab of insulator. . . . . . .
Charged conducting slab. . . . . . . .
Pair of conducting plates. . . . . . .
Charged hollow conductor. . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

149
151
151
151
151
152

G.1
G.2
G.3
G.4

Simplification of a resistor network. . . . . . . . . . .


Resistor network for exercise with an example above.
Network with two voltage sources. . . . . . . . . . . .
An electric dipole. . . . . . . . . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

154
154
155
156

H.1
H.2
H.3
H.4
H.5
H.6

An RC circuit. . . . . . .
Three resistor circuit. . . .
Potential divider circuit. .
Integration contours. . . .
Crossed field configuration.
Bridge circuit. . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

157
158
158
159
160
160

I.1
I.2
I.3
I.4

Air filled toroidal coil. . . . . . . .


Wire antenna in space. . . . . . . .
conducting loop in magnetic field. .
Conducting strip in magnetic field.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

161
162
162
163

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

J.1 An R L circuit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

xii

LIST OF FIGURES
J.2 An R L C circuit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
K.1
K.2
K.3
K.4
K.5

Two charged masses on a string. . . . . . . .


Calculation of a field using the superposition
Angle between normal and electric field. . .
A pyramid. . . . . . . . . . . . . . . . . . .
An electric field illustration. . . . . . . . . .

. . . . . .
principle.
. . . . . .
. . . . . .
. . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

170
171
172
173
175

L.1
L.2
L.3
L.4
L.5
L.6
L.7

Point P from segment l. . . . . . .


Electric field at 0. . . . . . . . . . . .
Work done moving from A to B. . . .
First Gaussian surface. . . . . . . . .
Second Gaussian surface. . . . . . . .
Third Gaussian surface. . . . . . . .
Gaussian surface inside a conductor.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

178
180
181
181
181
182
182

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

Position of points A and B relative to


Calculation of equivalent resistance. .
Conservation of charge. . . . . . . . .
Charge sharing. . . . . . . . . . . . .
Resistance of a wire. . . . . . . . . .
Kirchhos junction rule. . . . . . . .
Potentials at nodes. . . . . . . . . . .
Connection of a second earth. . . . .
A simple resistor circuit. . . . . . . .

the sphere.
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

184
185
186
186
187
188
189
190
190

N.1 Reference directions for current and voltage variables. . . . .


N.2 Charging a capacitor through a resistor. . . . . . . . . . . .
N.3 Voltage across resistor vs time. . . . . . . . . . . . . . . . .
N.4 Three resistor circuit and its equivalent. . . . . . . . . . . .
N.5 Alternative positions of meters. . . . . . . . . . . . . . . . .
N.6 Potential divider circuit. . . . . . . . . . . . . . . . . . . . .
N.7 Electrons in a uniform magnetic flux density. . . . . . . . . .
N.8 Conversion of galvanometer to a voltmeter and an ammeter.
N.9 Bridge circuit with capacitor. . . . . . . . . . . . . . . . . .
N.10 Bridge circuit with battery removed. . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

193
194
195
196
196
197
199
200
202
202

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

O.1 Air filled toroidal coil. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203


O.2 Boat approaching bridge. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
O.3 Boat moving away from bridge. . . . . . . . . . . . . . . . . . . . . . . . . 207
P.1
P.2
P.3
P.4
P.5
P.6
P.7

Expressions of Faradays and Lenzs Law. .


Current voltage relations in R, L and C. .
The simple RC circuit. . . . . . . . . . . .
The RLC circuit. . . . . . . . . . . . . . .
Excitation of inductor with resistance. . .
Voltages across L and R vs time. . . . . .
Plot of v 2 . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

209
211
211
212
213
213
214

LIST OF FIGURES
P.8
P.9
P.10
P.11
P.12
P.13
P.14

Current and capacitor charge of a LC circuit. . .


Energy in capacitor and inductor of an LC circuit.
Vector addition. . . . . . . . . . . . . . . . . . . .
Voltages in an RL circuit. . . . . . . . . . . . . .
Voltages in a series LRC circuit. . . . . . . . . . .
Power factor from a vector diagram. . . . . . . . .
Vector diagram of voltages. . . . . . . . . . . . .

xiii
.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

215
216
217
217
218
219
219

xiv

LIST OF FIGURES

Chapter 1
INTRODUCTION
1.1

Introduction

Electric and magnetic eects seen in our everyday lives include: static electricity and
lightning; the behaviour of magnetic materials as shown for example in a compass or in
an electrical relay or an actuator; all the benefits of domestic electricity including lighting,
heating, and running electric motors; the communications industry providing radio and
television entertainment and sometimes some information; the telephone and other data
communications networks; and the now ubiquitous computer technology. These eects
are supported by abundant (if you pay) supplies of electricity from electric generators and
certain widely advertised brands of batteries. Our task is to begin an understanding of
all these phenomena.
The phenomena are so diverse that we will concentrate in this simple introduction on
the fundamental laws, and leave to other parts of the course the understanding of complex
devices and systems which can be derived from those fundamentals. To understand this
introduction to the subject we will need to:
take note of observations of physical phenomena;
formulate some new concepts;
assemble appropriate equations describing those concepts; and
maintain scrutiny of the eectiveness of those equations in describing increasingly
diverse and complex phenomena.
What we have just described is the essence of the scientific method.

1.2

Course Aims

Among the things we hope to understand over a several year program of development of
electrodynamic theory are:
How things can push or pull one another without intervening materials.
How signals can come to us from outer space.
1

CHAPTER 1. INTRODUCTION
How we can broadcast from mountain tops or indeed from valleys.
How we can create moving pictures on the screen of a television set.
How we can remotely power devices through focussed electromagnetic energy beams.
The nature of electric and magnetic fields and the distinction between them.
How a magnetic field can cause an electric field and an electric field can cause a
magnetic field.
How we can construct and utilise Radio Frequency Identification (RFID) devices.

Finally we would like to know the complete set of principles of electrodynamics, so


that we may analyse any new electrodynamic situation we encounter, and so that we may
be in a position to create new electrodynamic technologies.
We also hope to learn something about the pioneer scientists who first discovered the
laws of electrodynamics, and the progressive way and the order in which those laws were
first understood.

1.3

Administrative Details

The lecturer will give details of:


The number and location of lectures and tutorials.
Laboratory arrangements, if any.
Lecturer location and availability.
Text and reference books.
Assumed knowledge.
Examination.

1.4
1.4.1

Relation to Other Physical Theories


Relativity

The modern electrodynamic theory is consistent with the theory of relativity. Indeed the
results of the fully developed electrodynamic theory embodied in Maxwells equations were
found, shortly after they were discovered, to be at variance with other aspects of physical
theory. Resolving the conflict led to the realisation that previous physical theories had
made an unwarranted assumption about the nature of time and led also to the theory of
relativity, but the equations of electrodynamics were found to consistent with that theory.

1.5. SUBJECT DEVELOPMENT

1.4.2

Quantum mechanics

When quantities become small it is necessary to take into account the eects described
by the quantum theory. While we will for completeness recognise the discreteness of
electronic charge, we will be dealing in our applications with systems of charge containing
many electrons, and quantum eects will not be evident. Thus we will present a theory
of macroscopic electrodynamics in which quantum eects are for the most part absent.
There is a corresponding theory of quantum electrodynamics which, because of its
complexity, is never used where it is not required.
Just occasionally we will bring in to our macroscopic theory some results of the quantum theory but in a form which fits comfortably into the macroscopic theory.

1.5

Subject Development

The most succinct description of the laws of electrodynamics is by means of vector dierential equations. We may not have dealt with this branch of mathematics yet, so we are
denied for the moment the benefit of that formalism.
However, just as in the analysis of ordinary functions there is both a dierential and
an integral version of the calculus, so too in vector calculus we have both a dierential
and integral form. Although the dierential form is too complex for study this year, the
concepts of the integral form are simpler than the dierential form, so we will be able to
state the fully general forms of the laws of electrodynamics in the integral form. It is one
of the principle objectives of this course to do so.
There are four fundamental laws, each expressed by an equation. Two of them are
associated with the name of Gauss, that is there are two dierent Gauss laws, one known
as Gauss law for the electric flux density, and one known as Gauss law for the magnetic
flux density. One is associated with the name of Ampere, and is known as Amperes law,
and one with the name of Faraday, and is not surprisingly known as Faradays law.
The equations expressing these fundamental laws are together known as Maxwells
equations. The versions of the laws thought to be true prior to the work of Maxwell were
self-contradictory. Maxwell was the first to make a correction to one of the equations to
bring them into harmony, and to make the prediction of the existence of electromagnetic
waves by combining the corrected equations with one another.
On the way to uncovering the four fundamental laws, other laws of considerable utility
will be introduced. Many of them were discovered early in the development of the subject,
and are still of great practical use. They are not considered to be in the fundamental group
of four, as they are really derivable from the fundamental laws, although the process of
doing so is largely an exercise in mathematics and does not add much to the understanding
of the subject.
The subject lends itself to study at varying levels of complexity. Electrostatics is the
study of the eects of stationary electric charges. Magnetostatics is the study of the
eects of unvarying electric currents or stationary magnetised material. In the level of
study known as electrodynamics we relax the restriction that charges, currents or material
be unvarying or stationary, and allow arbitrary time variation or movement. We will
encounter the fascinating result that time varying electric fields can create magnetic fields,
and time varying magnetic fields can create electric fields.

CHAPTER 1. INTRODUCTION

It is suitable to introduce the laws of electrostatics and magnetostatics first as these


are simpler than the fully general laws. But we must be careful to realise that while some
of the equations of electrostatics and magnetostatics continue to apply in the fully general
context of electrodynamics, others of those equations will require modification before they
apply in the more general context. It is the intention of the lecturer to make clear which
equations are fully general and which apply only in a limited context.

Chapter 2
MATHEMATICAL
BACKGROUND
2.1
2.1.1

Notation
Scalar quantities

Most of the time in this course we will be dealing with variables which directly express
the values of the physical quantities, such as for example, voltage or current, and if those
physical quantities have a time variation, so do the variables of our equations.
In such a case, when the quantities represented are scalars, as in the example of
voltage or current just mentioned, we use lower case Roman of sometimes Greek letters.
Sometimes the time variation is shown, and sometimes it is not, as for example in the
equation
v = v(t).

(2.1)

In some courses and occasionally later in this course, it will be convenient to restrict
the time variation of all physical quantities to be either constant or sinusoidal, or more
explicitly to be of cosine form.
In such cases, the behaviour of each time varying quantity is known for all time if we
know the frequency, the amplitude and the phase of the cosine function of time. In a single
context, all such quantities are assumed to have the same frequency, which is stated once
as a fixed part of that context, but the dierent variables representing dierent quantities
can have various amplitudes and phases.
In this situation, it is convenient to introduce a variable called a complex phasor which
while not itself being a function of time, does represent a time varying quantity. What we
do is to introduce a complex number, constructed so that the magnitude of the complex
number is the amplitude of the cosine waveform, and the angle of the complex number
in its polar representation, as shown for example in Figure 2.1, is the phase angle of the
cosine waveform.
Thus for the sinusoidally varying quantity
v = Vm cos(t + )

(2.2)

which has, when expressed as a cosine wave as has been done above, an amplitude Vm
and a phase angle , we construct the time invariant complex phasor V given by
5

CHAPTER 2. MATHEMATICAL BACKGROUND

Figure 2.1: Polar representation of a complex number.


V = Vm expj

(2.3)

It may be noted that the relation between the time invariant complex phasor V and
the time varying variable v(t) which it represents is
v(t) =

Vejt

(2.4)

In words, this relation says that to recover the time function from the complex phasor,
we multiply the complex phasor by ejt and take the real part.
A graphical interpretation of the mathematical operation just defined is that to recover
the time function from the complex phasor, we can first represent the phasor in the Argand
diagram, and take its projection on the horizontal axis as the expression of the value of
the time function at the time t = 0. To visualise the behaviour of the physical quantity as
a function of time, we must take the complex phasor, and rotate it in a counter-clockwise
direction on the Argand diagram at an angular frequency , starting at time t = 0 at the
position illustrated in Figure 2.1, and watch the values of the projection on the horizontal
axis of the rotating arm which results.
Notice in the above exposition that we have not said that the phasor rotates. To do
so would contradict the definition of the phasor as a time invariant quantity. We have
given a dierent name, namely rotating arm, to the thing that rotates.
In establishing the notation described above, we have been able, because it is available, to use dierent calligraphy, namely v and V, to distinguish the real time varying
variables directly representing the physical quantities, and the time invariant complex
phasors indirectly representing them. The dierence in notation is helpful in avoiding
misunderstandings.

2.1.2

Vector quantities

When we come to the representation of a sinusoidally varying vector quantity, we use


as is common bold face characters to represent vectors. It is traditional to represent
electromagnetic field quantities by upper case letters. If we are to distinguish clearly
between the real time-varying variables and the time invariant complex phasors which
in the particular case of sinusoidal time variation may be used to represent physical
quantities, we are in need of two bold face character sets of dierent appearance.

2.1. NOTATION

This need is satisfied by the use of so called calligraphic characters, for example E, D,
H, and B, for the time varying vectors directly representing time varying physical vector
quantities, and upright Roman letters, such as E, D, H, and B, for the time invariant
complex vectors which can be used to represent sinusoidally varying vector quantities.
In both cases the term vector indicates that the physical quantity has three Cartesian
components, each of which is a scalar quantity, and each of which may have any of: no
time variation, a general non-sinusoidal time variation, or perhaps a sinusoidal variation.
In all three cases representation by the calligraphic letters is appropriate, but only in the
last case is representation in the phasor notation also appropriate.
The field vectors can in the general (non-sinusoidal) case have, in addition to their time
variation, a spatial variation. In setting out equations, we may for emphasis explicitly
show the time or space variation, or we may for compactness just write the symbol for
the variable with the functional variation understood. Both of these things are done in
the equation for an electric field vector
E = E(x, y, z, t).

(2.5)

Just as we have for scalar quantities the possibility of representing a sinusoidally


varying quantity by a time invariant complex phasor, so we can for a sinusoidally varying
vector quantity E introduce a time invariant complex vector phasor E (really just three
phasors representing the components along the three co-ordinate axes), with the relation
between the two being
E(x, y, z, t) =

E(x, y, z)ejt

(2.6)

which shows in the case of sinusoidal time variation of a vector the relation between
the vector of time varying functions providing a direct representation of the field E and
the vector of time invariant complex phasors providing the indirect representation of that
field.

2.1.3

Calligraphic characters

As some of the calligraphic characters are a little unusual in appearance, we provide here
a table showing the most often used letters paired with their Roman counterparts. The
table also, for future reference, gives the names and the Standard International Units of
the physical quantities most commonly represented by those variables.
E
H
D
B
J

E
H
D
B
J

Electric field intensity


Magnetic field intensity
Electric flux density
Magnetic flux density
Volume current density

V m1
Am1
Cm2
W bm2
Am2

Table 2.1: Names and units of common field vectors

2.2
2.2.1

CHAPTER 2. MATHEMATICAL BACKGROUND

Co-ordinate systems
Scope

In practical situations we often have to deal with objects of round or spherical shape, and
an algebraically simple description of the object is not available in the familiar rectangular cartesian co-ordinate system. For this reason we will sometimes use the cylindrical
polar co-ordinate system or the spherical polar co-ordinate system defined in the following
sections.

2.2.2

Cylindrical polar co-ordinates

Figure 2.2 shows the standard cylindrical polar co-ordinate system, with co-ordinates
(r, , z) and an associated rectangular cartesian co-ordinate system with co-ordinates
(x, y, z). The relations between co-ordinates of the two systems are
x = r cos
y = r sin
z = z

(2.7)

Figure 2.2: Cylindrical polar co-ordinate system.

2.2.3

Spherical polar co-ordinates

Figure 2.3 shows the standard spherical polar co-ordinate system, with co-ordinaes (r, , ),
and an associated rectangular cartesian co-ordinate system with co-ordinates (x, y, z). The
relations between co-ordinates of the two systems are
x = r sin cos
y = r sin sin
z = r cos

(2.8)

2.3. THE FIELD CONCEPT

Figure 2.3: Spherical polar co-ordinate system.

2.3

The Field Concept

The word field has many meanings in the English language. There is the familiar meaning in agriculture, a quite dierent meaning (a commutative ring with a multiplicative
identity) in group and number theory, and a particular meaning in physical science and
engineering which we must now define. For our purposes, a field is some physical variable
which has a value (generally varying) at all points of a region of three-dimensional space.
A field may be a scalar field, ie the mapping of a variable which has a magnitude
but no direction. Examples of interest to weather forecasters are the temperature or air
pressure variables which can be measured over a region of the earths surface and the
atmosphere above it. A field may be a vector field, the most familiar example being the
gravitational field, this field having both a magnitude and a direction.
At first sight the gravitational field appears to be uniform, having the same value here
as next door, but if we jump really high, we find that the gravitational field is non-uniform,
weakening as we get higher, and also changing in direction, pointing always towards the
earths centre even if we move to other hemispheres. In fact we find, when we explore
it, that the earths gravitational field belongs to the class of inverse square central force
fields, which have many interesting properties which we will explore later.
An important aspect of a field is that to understand its practical significance it is
necessary to develop a capacity to visualise it, ie to create in the mind a mental picture
of some representation of the field.
Another important example of a vector field is provided in the study of fluid dynamics,
in which we trace the movement of a particle within a fluid as it moves throughout a three
dimensional space. In this the velocity field of such a particle has many mathematical
properties similar to the quite dierent physical fields we will study in electrodynamics,
and in consequence much of the terminology originally developed for the mathematical
study of fluid dynamics has been adopted in the mathematical formulations used for the
study of the various forms of electrodynamics including electrostatics and magnetostatics.

10

CHAPTER 2. MATHEMATICAL BACKGROUND

2.4

Source and Vortex Fields

In our study of electric and magnetic fields we will be greatly and in many ways aided by
the concepts of sources and vortices, originally developed, as the names suggest, in the
study of fluid dynamics. The concepts themselves will be explained in the next section,
but we will remark first that:
An ability to form in the mind electromagnetic field pictures appears to be an
essential skill required for mastering electromagnetic theory.
The source and vortex concepts provide a basis for picturing the electromagnetic
field created in a wide range of situations.
In a theorem first proved by Helmholtz, it may be shown that any vector field may
be uniquely decomposed as the sum of two vector fields, one of which is purely
source type, and one of which is purely vortex type.
The sources and vortices of a field are described in a mathematical sense by derivatives of the vector calculus known as the divergence and curl derivatives, but these
are too complex for study this year, and will be defined in future years.
Maxwells equations (which are the fundamental laws of electrodynamics) are direct
statements about the source and vortex properties of electromagnetic fields.

2.4.1

Source-type fields

Within a small region, a field is considered to be source-type if there is, as illustrated in


Figure 2.4, a nett flux of the field which emerges from that region.

Figure 2.4: Illustration of a source type field.


The archetypical example of a source-type field is the electrostatic field which is caused
by a distribution of electric charge within the field region.

2.5. NAMES AND UNITS FOR VARIABLES

2.4.2

11

Vortex-type fields

Within a small region, a field is considered to be vortex-type if there is, as illustrated


in Figure 2.5, a nett circulation of the field around a contour within that region. The

Figure 2.5: Illustration of a vortex type field.


archetypical example of a vortex-type field is the magnetostatic field which is caused by
a distribution of electric current within the region.

2.5

Names and Units for Variables

Several disparate terminologies and a considerable number of sets of units have been (and
sometimes are still) in use by scientists, text book writers and lecturers in discussing the
subject of electrodynamics. This can and does cause considerable confusion for students.
Fortunately a set of International Standards have been defined and will be followed
throughout this course. Probably you have already been exposed to terminology dierent
from that in use in this course, and experience has shown that confusion will arise unless
a deliberate attempt is made by the student to master the dierences (and hopefully to
adopt the International Standard terminology). The names and units will be given again
as the corresponding concepts are elucidated, but in order to achieve clariy at the earliest
possible point they are summarised below.
1. The names, usual symbols, and Standard International units for the four vectors
used to describe electromagnetic fileds in a general medium are
E
electric field intensity
Vm1
H magnetic field intensity Am1
D
electric flux density
Cm2
B
magnetic flux density Wbm2
The units of magnetic flux density B have the alternative name of Tesla, for which
the abbreviation is T.

12

CHAPTER 2. MATHEMATICAL BACKGROUND


2. The names, usual symbols, and Standard International units for the vectors used
to describe the state of a dielectric medium and the state of a magnetic medium
respectively are
P
polarisation Cm2
M magnetisation Am1
3. The fully general relations between the above six vectors are
(i)
(ii)

D = 0E + P
B = 0 (H + M)

by definition
by definition

4. The values of the magnetic permeability and the dielectric permittivity of free space
are in standard international units
(i) 0
(ii) 0

4 107
8.854 1012

Hm1
Fm1

by definition
approximately

In the most recent definition of the standard international system of units, both of
the above values have in eect become defined values, but we will have to learn more
electromagnetic theory before we can see why this is so. This matter is discussed
further in Appendix A.

2.6

A Modest Proposal

It is sugested that at the earliest practicable time students commit to memory the material
in the preceding section.

2.7
2.7.1

Vector Algebra
Scope

It is assumed in this course that students will have a familiarity with the basic concepts
of vector algebra, and a complete treatment is not attempted here. We will merely for
safety refresh the two concepts of the scalar product and the vector product between two
vectors.

2.7.2

Scalar product

As shown in Figure 2.6, the scalar product of two vectors v1 and v2 is equal to the product
of the magnitudes of the two vectors times the cosine of the angle between them. If the
vectors happen to be parallel, the scalar product is the product of the magnitudes. If the
vectors happen to be orthogonal, the scalar product is zero.

2.7. VECTOR ALGEBRA

13

Figure 2.6: Illustration of a scalar product.

2.7.3

Vector product

As shown in Figure 2.7, the vector product of two vectors v1 and v2 is another vector,
and thus has a magnitude and a direction. The magnitude of the vector product is equal
to the product of the magnitudes of the two vectors times the sine of the angle between
them. The direction of the vector product is the direction perpendicular to the plane
defined by the two vectors v1 and v2 , with the sense (i.e. the choice between the two
directions which are perpendicular to the plane) determined by the right hand rule.
If the vectors happen to be orthogonal, the magnitude of the vector product is the
product of the magnitudes. If the vectors happen to be parallel, the vector product is
zero.

Figure 2.7: Illustration of a vector product.

14

2.8

CHAPTER 2. MATHEMATICAL BACKGROUND

Vector Integrals

In order to be able to state fully general laws of electric and magnetic fields we are going
to have to use the concept of the integral of a vector field. As in any integral, for example
the familiar integral
8 b

f(x)dx

we need to specify the function which is being integrated, the variable with respect to
which it is being integrated, and the domain, here specified by the limits a and b, over
which the integral is being performed.
Although it would be possible to create many dierent types of vector integrals, we
will be concerned just with the two discussed below. Although both of them are integrals
of a vector field F, the results of both of the particular integrals defined here are scalars,
because we form in the process of integration the scalar product between the vector
function F and an incremental vector element of the domain over which the variable is
being integrated. There are two types of such integrals, namely line integrals, where the
domain is a line (not necessarily a straight one) and a surface integral, where the domain
in a surface (not necessarily a flat one). These concepts are made clearer in the following
two sections.

2.8.1

Line integrals

An example of the line integral


8 B
A

F dr

(2.9)

is provided in Figure 2.8. The integral is performed over a specified pathway leading
from the point A to the point B. To perform the integral, we perform a dissection of the
pathway into short segments, each denoted by a short vector distance dr. At each such
element of the path we form the scalar product Fdr between the local value of the vector
field F and the path element dr, then form the sum of these elemental scalar products,
and then take the limit as the dissection is made more fine such that the length of the
longest element of the path tends to zero.

2.8.2

Surface integrals

An example of the surface integral

F ds

is provided in Figure 2.8. The integral is performed over a specified surface which is the
shaded region shown in the diagram. To perform the integral, we perform a dissection of
the area into small patches, each denoted by a vector area ds. For each such patch we
form the scalar product Fds, between the local value of the vector field F and the patch
element ds, form the sum of these elemental scalar products, and take the limit as the
dissection is made more fine such that the area of the largest patch tends to zero.

2.8. VECTOR INTEGRALS

Figure 2.8: A line integral for a vector field.

Figure 2.9: A surface integral for a vector field.

15

16

2.8.3

CHAPTER 2. MATHEMATICAL BACKGROUND

Contour integrals

In addition to the line integral defined above we sometimes need to integrate over a closed
contour shown in Figure 2.10. This is the same as the line integral defined above except
that the end point B has been moved to coincide with the start point A. The value of the
integral is then the same for any start point on the contour, so we no longer specify the
start and finish points, but merely name the contour, in this case adopting the name C,
and write the integral as
C

F dr

Note that a special integration symbol dierent from that used for a line integral has
been used to indicate that the line has closed upon itself.

Figure 2.10: A closed contour integral.

2.8.4

Closed surface integrals

In addition to the surface integral defined above we sometimes need to integrate over a
closed surface shown in Figure 2.11. This is the same as the surface integral defined above
except that the surface no longer has a boundary, but instead encloses a volume.
The integral integral over a closed surface is denoted by
-

F ds

Note that a special integration symbol dierent from that used for a line integral has been
used to indicate that the surface S is closed.

2.9

Representation of Fields

Two dierent methods of representing a vector field are illustrated in Figure 2.12. Both are
representations, with various degrees of completeness, of the same vector field. Although
it is not pertinent to the general points to be made in discussing representations of vector
fields, you may be interested to know that the field represented is that produced by two
small charged metallic spheres, the one at the left carrying a positive charge and the one
at the right carrying a negative charge of lesser magnitude.

2.9. REPRESENTATION OF FIELDS

17

Figure 2.11: A closed surface integral.


In the representation at the top of the figure, the dierent values of the field at the
limited number of points denoted by P, Q, R, S, T, U, and V are each represented as
a vector with of course direction indicating the field direction and length indicating the
field magnitude. In this representation, while we have a clear indication of the field at
that limited number of points, we can only guess at the values of the field in between.
The representation in the lower part of the diagram is that devised by Faraday, and
makes use of what he called lines or tubes of force. In this representation the field lines
are not straight and are not vectors. They do have the important properties listed below.
At any point through which a field line passes the direction of the field is indicated
by the tangent to the field line.
The magnitude of the field at such a point is indicated by the number of field lines
per unit area that will intercept an area perpendicular to the direction of the field
line. Thus where the field lines are crowded together, the field is strongest, and
where the field lines are more spaced, the field has become weaker.
Of course because we have drawn the field lines on the plane of a sheet of paper, the
planes through which the field lines are imagined to pass and which are associated with
the definition and the magnitude of the field are perpendicular to the paper. It is not
practical to draw them in this diagram, so they must be imagined.
The two representations have very significantly dierent properties, but both have
their uses. The vector representation shown at the top of the figure is useful if we wish
to consider at one point various contributions to a field which might come from dierent
sources, and sum those contributions in a vector diagram to determine the resultant field.
This process is illustrated in Figure 2.13.
On the other hand, the method shown lower on the page is useful for representing the
entire field within the area occupied by the field lines. Examination of the diagram will
suggest that there does not seem to be any position at which both the direction and the
magnitude of the field is not clearly indicated.
It might be noted that in this representation the field lines only start and stop at
positions where charges exist. We have simply, as a matter of convenience, terminated
many of the field lines at some distance from the charge because we have only a finite

18

CHAPTER 2. MATHEMATICAL BACKGROUND

amount of paper. In this representation the field lines have this special property, not as
a general facet of the Faraday representation, but as a particular consequence of the fact
that the field that we have chosen to represent here is an electrostatic field caused by a
charge distribution placed at particular points.
In the above discussion it has been pointed out that in the Faraday representation the
magnitude of the field is represented by the density per unit area of field lines intercepting
a plane perpendicular to the local direction of a field line. This requirement implies two
things. Firstly that the fields we represent are in a three-dimensional space because three
dimensions are needed to contain both the field line and the plane perpendicular to it.
Secondly there is a degree of imagination needed to create a visualisation of the field
lines existing in this three-dimensional space. Rarely in a text book is a diagram to
represent this three dimensional nature of the Faraday representation attempted. Such a
rare example however is provided in Figure 2.14.
Although we are concerned with techniques of field representation rather than representation of a particular field, it may be of interest to know that the field represented in
that diagram is the electrostatic field created by an isolated point charge located some
distance above a metallic conducting plane of infinite extent, on the surface of which is
spread out a negative charge of magnitude equal to the positive point charge. The shading
of the ground plane is intended to indicate the regions in which the density of negative
charge per unit area is greatest.

2.10

Inverse Square Central Fields

An important sub-class of vector fields is the class known as inverse square central fields.
These are fields which, at any point, have a direction either away from or towards a fixed
centre (and hence the term central) and which vary in their magnitude as the inverse
square of the distance from that central point (and hence the term inverse square).
A field we are already familiar with which has this property is the gravitational field.
We will see later that the same properties are present in the electrostatic field produced
by a point charge as embodied in Coulombs Law.

2.11

Properties of Conservative Fields

Many of the vector fields which we will study are associated with forces on objects which
we may move around. A good example is provided by the gravitational field which represents the force per unit mass on an object, or the electrostatic field which we will come to
recognise represents the force per unit charge on an object. Both of these objects are of
the type that we can grab hold of and move around. In the process we will have to exert
on the object a force which is the opposite of that exerted by the field. In the process of
moving the object we will do work.
An important sub-class of the force fields are those known as conservative fields. These
fields have the two properties, which are entirely equivalent, listed below.
The work done in moving the object around a closed path, so that we end at the
same point at which we started, is zero.

2.11. PROPERTIES OF CONSERVATIVE FIELDS

Figure 2.12: Two methods of representing a field.

Figure 2.13: Vector diagram to determine a resultant field.

19

20

CHAPTER 2. MATHEMATICAL BACKGROUND

Figure 2.14: Three dimensional aspect of the Faraday representation.


The work done in moving the object from a particular start point to a particular
finish point is independent of the path we take.
It may be shown, and this is a purely mathematical rather than physical result, that
the class of inverse square central fields is conservative.

2.12

Flux and Circulation

Closely related to the concepts of sources and vortices which we discussed earlier are two
mathematical concepts of flux and circulation which are discussed in the next section and
which serve as a measure of the source strength or vortex strength of a particular field. As
the names of those concepts might imply (flux is the Latin word for flow) the concepts
originally were defined in the context of the study of fluid dynamics. Although they play
an absolutely fundamental role in establishing the laws of electrodynamics, it should be
emphasised that in that latter context nothing is really flowing. It is just that the field
lines of electric and magnetic fields merely resemble the stream lines of a flowing fluid.

2.12.1

Flux of a vector

When a vector field intersects a possibly non-planar $surface S of which ds is an element


of the surface area as shown in Figure 2.15, we call S F ds the flux of the vector F
flowing through the surface S in the direction defined by the vector ds. In the definition
it is essential to define a positive direction for the vector ds.
The positive direction may in general be arbitrarily chosen although it must always
be stated. In the particular context of expressing the laws of electrodynamics, however,
there are some conventions which are required to be observed in that choice.
The first of these applies to a non-closed surface such as that illustrated in Figure 2.15.
Such a surface is always bounded by a contour. There is usually a theorem which relates
the surface integral of some vector to the contour integral of another vector. Since both
of integrals can change sign if the sense chosen for the positive direction of the surface is
changed or the contour integral is performed in the opposite direction, it is necessary, in
order to sustain a theorem, for there to be a relationship between the sense chosen for
the area and the sense chosen for the contour.

2.12. FLUX AND CIRCULATION

21

Figure 2.15: Flux passing through a surface.


The relationship which is always chosen is one defined by the right hand rule. In this
rule if one directs the thumb of the right hand in the positive direction chosen for the
surface the fingers of the right hand should indicate the positive direction for traversing
the contour.
Of course not all surface integrals are of the type illustrated in Figure 2.15. In Figure 2.11 we encountered a closed surface integral. When the surface is a closed surface
the sense which must always be chosen for the positive direction of each surface element
is the outward direction.

2.12.2

Circulation of a vector

In the above discussion we encountered the concept of the integral of a vector around a
closed path. That integral has in fact the ocial name of circulation.
When a closed path C is defined in a vector field F as shown Figure 2.16 we may
define the circulation of F around the contour C as
=

F dr

(2.10)

Figure 2.16: Circulation around a contour.


It is hopefully clear that the value of the circulation will depend upon the sense with

22

CHAPTER 2. MATHEMATICAL BACKGROUND

which we traverse the contour and will change sign if, in the integral, the contour is
traversed in the opposite direction.

2.13

A Sneak Preview

As it has been stated that an objective of this course is to introduce the fully general laws
of electrodynamics, it is of interest to see how these laws may be succinctly expressed in
terms of the mathematical concepts just defined.
In the usual notation, the vector E represents the electric field intensity measured
in V/m and the vector H represents the magnetic field intensity measured in Am1 .
Two additional vectors P representing the polarisation of a dielectric medium and M
representing the magnetisation of a magnetic medium allow the definition of an electric
flux density vector D = 0 E + P measured in Cm2 , and a magnetic flux density vector
B = 0 (H + M) measured in Wbm2 . In the below statements, the term circulation
refers to the integral with respect to distance around a stated closed path of the scalar
product between a named field vector and a vector element of distance around that path,
while the term flux refers to the integral with respect to area over a stated area of the
scalar product between a named field vector and a vector element of that area. In terms
of that terminology and the four vectors E, H, D and B the four fundamental laws are:
Faradays law: The circulation of the electric field vector E around a closed contour
is equal to minus the time rate of change of magnetic flux through a surface bounded by
that contour, the positive direction of the surface being related to the positive direction
of the contour by the right hand rule.
Amperes law as modified by Maxwell: The circulation of the magnetic field
vector H around a closed contour is equal to the sum of the conduction current and the
displacement current (defined as the integral over the surface of the time rate of change
of the D vector passing through a surface bounded by that contour, with again the right
hand rule relating the senses of the contour and the surface.
Gauss law for electric flux: The total electric flux (defined in terms of the D
vector) emerging from a closed surface is equal to the total conduction charge contained
within the volume bounded by that surface.
Gauss Law for magnetic flux: The total magnetic flux (defined in terms of the B
vector) emerging from any closed surface is zero.
The principal objective of this course is to establish within students a clear and permanent understanding of these statements.

Chapter 3
ELECTROSTATICS IN EMPTY
SPACE
3.1

Introduction

With our mathematical background behind us we now begin our study of electromagnetic
fields which the study of electrostatics in empty space. This is the study of the properties
of stationary charges. The study may be seen to be important for the reasons that:
It is the beginning of the study of the broad subject of electrodynamics which is the
basis of a very large amount of electrical and electronic technology which we enjoy
in our lives.
The electrostatic force is the principal force at work in chemical bonding which is
of course not only enjoyable but essential for our lives.

3.2
3.2.1

Experimental Observations
Static electricity

We are hopefully all familiar with the elementary phenomena which are observable in the
study of static electricity.
In a systematic study of these phenomena we may begin with lightweight spheres,
perhaps made with pith, suspended on lightweight insulating threads, perhaps made with
silk, such as are illustrated in the figures which follow.
But, you may ask, where does this pith come from? Where can I get some? Well,
these days it most commonly occurs in text books, but if you want to handle some you
can use the soft lining of bamboo, dried out, or the lightweight interior parts of other
growing things. What is good about it is that when dried it is light in weight for the
volume it occupies, but it has just enough electrical conductivity left for charges which
are deposited at one point on it to spread over its surface. It also provides, for people old
enough, a nostalgic recall of the early days in which the subject was first learned.
The first experimental observation that we might make is that, as shown in Figure 3.1,
if we first rub a glass rod with silk, and touch that rod onto a pith ball suspended as shown
in Figure 3.1 the ball will show a tendency to be repelled from the glass rod. A similar
23

24

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

observation might be made if a perspex rod is rubbed with fur. In either case the pith
ball touched with the rod experiences a repulsion from that rod.

Figure 3.1: Pith ball repelled by a charged rod.


A curious observation, illustrated in Figure 3.2, may be made. Pith balls rubbed with
the same type of rod appear to repel one another, however, a pith ball touched by the
glass rod is attracted to the pith ball touched with the perspex rod.

Figure 3.2: Pith balls in attraction and repulsion.


Experiments along these lines eventually lead us to the following conclusions.
That the rubbing of the rods with the named substances produces on them an
electric charge.
That there are two types of electric charge which we call positive and negative.
That like charges repel while unlike charges attract.

3.2. EXPERIMENTAL OBSERVATIONS

25

On the atomic theory the explanation we adopt is that matter contains charged particles, some of them negatively charged and called electrons and others positively charged
called protons, these charges having the same magnitude but opposite signs. An electron
is sometimes detachable from the remainder of the atom. A positive charge occurs when
electrons have been detached from a substance, and a negative charge occurs on the body
to which the electrons have been transferred.
We can make the further observation that a plastic ruler rubbed with wool can attract
small pieces of paper without there having to be any touching. We will for the moment
have to defer our explanation of this phenomenon until a later time.

Figure 3.3: Charging by induction.


Among the elementary phenomena which we might study is that of charging by induction. Essential steps of this process are illustrated in Figure 3.3. In this drawing an
uncharged conductor, perhaps made of copper, is mounted on an insulated stand perhaps
made of glass and a positively charged glass rod is brought near but does touch as shown
in part (a). The result will be, as shown in that part of the figure, an attraction of electrons to the end near to the glass rod with there being a dearth of electrons, that is a
positive charge, at the other end of the conductor.
The conductor can then be touched with a finger at the end remote from the rod.
Touching will establish a conducting path up from the earth to that end of the rod,
and electrons will be drawn up through that path so that that particular end of the rod
becomes electrically neutral. The conducting path may then be removed as shown in part
(c) of the diagram with the result that one end only, that closer to the nearby glass rod,
contains an imbalance of charge.
Finally the charged glass rod, which has lost none of its charge, can be removed to
a suitable distance, upon which the charges on the conductor redistribute themselves as
shown in (d). We observe in this process that we have achieved a charge on the insulated
conductor which is of opposite sign to the charge on the glass rod, and that we have lost
none of the charge on the glass rod in the process. We may, however, have done work in
moving the glass rod around.

3.2.2

Conservation of charge

In the above explanations we have tacitly assumed what turns out to be one of the most
enduring laws of physics, namely that of conservation of charge. When we put charge

26

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

from the rod onto the pith ball we assumed it came from the rod, and when we charged
the conducting body in Figure 3.3 we assumed that the negative charge came from the
earth. This idea of conservation of charge is in fact a very fundamental concept which has
endured through all the changes in physical theory which have been seen in this century.
We will give this charge conservation concept a mathematical expression later, but for now
we will take it as our first and most very basic concept in our study of electrodynamics.
Although we have yet to define a method of measuring charge, let us state now that in
the Standard International system of units it is measured in coulombs, but we will delay
the definition of the coulomb until more of the phenomena of electrodynamics have been
revealed.

3.2.3

Coulombs Law

Figure 3.4: Point charges in free space


The first quantitative laws dealing with the phenomena discussed above were determined by Coulomb with the experimental apparatus not unlike that illustrated above but
augmented by methods of measuring small forces, principally through the use of a delicate
torsion balance. Coulomb determined two important empirical results, the first of which
is always quoted and the second of which is often overlooked. The first of these results is
that the force exerted by charge q1 on charge q2 distant r12 from charge q1 in free space
as shown in Figure 3.4 is
q1 q2r12
(3.1)
2
4 0 r12
Putting this formula into words we can say that the force of mutual repulsion between
two electronic charges is proportional to the magnitudes of those charges and inversely proportional to the distance between them. The formula contains a physical constant 0 which
has in the Standard International system of units the approximate value 8.8541012 and
the units of Farads per metre. Why the units are given this name cannot be made clear
until more of the subject has been expounded.
In the formula we may note that the inverse square part of the law resembles the
inverse square part of the law of gravity, but that the electrostatic formula is unlike
F 12 =

3.3. THE ELECTROSTATIC FIELD

27

gravity in that like charges repel and unlike charges attract, whereas in gravity masses
always attract.
We might also note the symmetry of the formula, and the action-reaction principle
implied thereby. Finally we might note that this form of the formula emphasises the
concept of action at a distance. With no intervening medium, this Coulomb force is able
to act across empty space.
The second major discovery of Coulomb, and one mentioned above as generally overlooked, is that the force between two charges is not aected by the introduction of a third
charge. Put another way, the concept of superposition applies. If we have three charges,
we can calculate the force on the third charge by separately calculating the forces on that
charge exerted by the first two, and adding those forces. Although the linearity evident
in Equation 3.1 suggests that this may be so, it should be regarded as an empirically
established fact, and not a consequence of the neatness of any mathematical description
which encompasses the behaviour of the forces between only two charges.

3.3

The Electrostatic Field

If we rewrite the formula for the Coulomb force to focus attention on the force on just
one of those charges we might locate charge q1 at the origin and charge q2 may move to
various points r of space so r12 used before becomes just r. The Coulomb formula thus
becomes
F(r) =

q1r
q2
4 0 r2

(3.2)

Rewriting the formula in this way leads us to the concept of the electrostatic field
which is defined for any point in space as the force per unit charge on a test charge placed
at that point. The Equation 3.2 above also invites a subtle change in viewpoint wherein
the first part the equation is taken to indicate that the presence of the charge q1 at the
origin creates at each point r in space a property which is experienced by the second
charge q2 in the form of a force which is proportional to that second charge. We are thus
associating the force on the test charge more with the point in space than with the other
charge which we have placed at the origin.
This viewpoint was strongly supported by Faraday and greatly assisted in his creative
thinking about electromagnetic fields. We thus make a more formal separation of the
parts of the equation by defining a concept of electrostatic field. The electrostatic field at
a point in space in thus interpreted as the force per unit charge on a test charge which
has been placed at that point.
The electric field E(r) caused at a point r of space by an electric charge q1 situated at
the origin is then
E(r) =

q1r
4 0 r2

(3.3)

An expression for the force on charge q at position (r) is then


F = qE

(3.4)

28

3.4

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

Superposition

Because we have found, empirically, that the superposition principle applies to Coulomb
forces it will also apply to the electric field just defined. Expressed in words this will say
that the electric field at one point caused by a number of charges is the sum of the electric
fields which might be caused by those other charges individually. Mathematically we may
say that the expression for the field E(r0 ) at point r0 due to charges q1 to qn at points r1
to rn is given by
E(r0 ) =

q1r10
q2r20
qnrn0
+
+ +
2
2
2
4 0 r10 4 0 r20
4 0 rn0

(3.5)

where rn0 is a unit vector in the direction from the point rn to the point r0 , and
rn0 is the scalar distance between these points. An illustration of the application of
superposition to electric field calculation is given in Figure 3.5.

Figure 3.5: Application of superposition in electric field calculations.

3.5

Representation of Electric Fields

We have already discussed the representation of vector fields in Section 2.9. It may be
worth giving some support to the Faraday representation, as it is a belief of the author
that this is the one which best supports visualisation of an entire field distribution, and it
is a further belief that skill in such visualisation is indispensable to a clear understanding
of the subject.

3.6

Illustration of Some Fields

We take the opportunity to show in Figures 3.6 to 3.8 the electric fields of some elementary
change distributions.

3.6. ILLUSTRATION OF SOME FIELDS

3.6.1

29

An isolated charge

The electric field of an isolated charge is shown in a representation resembling the Faraday
representation Figure 3.6. Notable features are the spherical symmetry and the spreading
of the lines as we proceed away from the change. The paradox that the field lines in the
diagram separate as the first power of distance, whereas in the Faraday representation
they should separate as the second power of distance may be resolved by realising that
this planar diagram really only hints at a three-dimensional field line structure in which
the correct behaviour occurs.

Figure 3.6: Electric field of an isolated charge

3.6.2

Two equal charges

The electric field distribution of a pair of equal positive charges is illustrated in Figure 3.7.
From a distance the field distribution resembles that of a single charge, but in the region of
the charges there is considerable dierence. Notable is the neutral point mid-way between
the charges. At this point the electric field is zero.

3.6.3

An electric dipole

The electric field of two charges of equal magnitude but opposite sign is shown in Figure 3.8. Such an arrangement of charges is known as an electric dipole.
The electric dipole charge distribution is important in many ways, and deserves it own
section which follows.

30

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

Figure 3.7: Electric field of two equal charges

Figure 3.8: Electric field of a dipole

3.7. THE ELECTRIC DIPOLE

3.7
3.7.1

31

The Electric Dipole


Electric field pattern

The electric field pattern has already been given in Figure 3.8.

3.7.2

Parameter definition

In this section we will define an important parameter known as the dipole moment. As
shown in Figure 3.9 the dipole consists of charges q and q separated by a vector distance
2d, the vector being directed from the negative charge toward the positive charge. This
choice of direction is important, and often mis-remembered by students.
We will find in below sections that many of the eects of a dipole are proportional
to the product of the charge and the vector separation as defined above. Thus we define
in Equation 3.6 below for the electric dipole a parameter p known as the electric dipole
moment.
p = 2qd

(3.6)

The electric dipole moment is therefore seen to be the product of the positive charge
and its vector distance from the corresponding negative charge.

Figure 3.9: Parameters of an electric dipole

3.7.3

Dipole field analysis

We recall from Coulombs law that the field of an isolated charge diminishes as the inverse
square of distance from the charge. The mathematical description of the electrostatic field
of a dipole is much more complex, and it is certainly not a simple central field like that
of an isolated charge. It does however have the interesting property that once we get far
enough from the dipole the field diminishes in magnitude as the inverse third power of the
distance. Providing a general proof of this result for all directions is a greater burden than
we wish at this moment to undertake, so we will content ourselves here with providing an
analysis just for the case when we move from the mid-point of the dipole in a direction
at right angles to the direction of the dipole. The geometry for this analysis is given in
Figure 3.10.

32

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

E+



Er

_q


q
d

Figure 3.10: Analysis of a dipole field

where
and

|E + | = |E | =
cos =

q
4 0 (r2 + d2 )

(3.7)

d
1
(r2 + d2 ) 2

(3.8)

At P, the y components of E + , E cancel while the x components E + cos , E cos


add.
So

E = |E| = 2E cos =

If

p
3

4 0 (r2 + d2 ) 2
1 p
d, then E =
4 0 r3

(3.9)
(3.10)

We see that because of field cancellations the field drops o more rapidly than for a
point charge.
It may be shown that the r3 dependence holds in all directions, not just in the
perpendicular direcetion, although we wont prove this.

3.7.4

Torque on a dipole

It is a simple matter and should be taken as an exercise to show that when a dipole of
strength p is placed in a uniform electric field E it experiences a torque
T =pE

(3.11)

The calculation is illustrated in Figure 3.11

3.7.5

Induced dipoles

If the electrons in a solid body have some freedom of movement then it is easy for such a
body when it is placed in an electric field to develop a dipole moment, even if no charge

3.7. THE ELECTRIC DIPOLE

33

Figure 3.11: Calculation of torque on a dipole


is transferred to the body. If some of the electrons can move a significant distance from
their parent atoms, they will move in a direction opposite to that of the field, leaving as
shown in Figure 3.12 the two ends of the body oppositely charged, so that the body has
a dipole moment. A body which is long and thin in the direction of the field shows this
eect more strongly than others. You should think about why.

Figure 3.12: A naturally created dipole

3.7.6

Force on a dipole

As we have seen in Figure 3.11, in a uniform electrostatic field the two forces on the two
charges of a dipole are equal and opposite and hence there is no nett force on a dipole.
This result is quite well known. Less well known, however, is that in a non-uniform
electrostatic field a dipole can experience a nett force. Referring again to Figure 3.11 we
can see that if the field is non-uniform in an appropriate way it may well be that the force
on one of the charges of the dipole does not exactly cancel the force on the other, and
there is a non-zero resultant. It should be easy to appreciate that the force is proportional
to (although not parallel to) the strength p of the dipole, and the force is stronger when
the field is more rapidly spatially varying.
We are now in a position to explain a phenomenon first noted in Section 3.2 but not
there explained. It was noted that a plastic ruler rubbed with wool will attract small
pieces of paper. To explain this, we will first note that the field near the charged end of
the ruler is spatially non-uniform as it gets stronger close to the ruler. Secondly we note
that charges can, as illustrated in Figure 3.12, move from one end of the paper to the
other, either by the mechanism of conduction or through the mechanism of polarisation
to be discussed in Chapter 4. Once we have recognised that we have a dipole in a non-

34

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

uniform field it is easy to see that we can have a net force produced by the mechanism
discussed earlier in this section.

3.8
3.8.1

Electric Flux and Flux Density


Introduction

We now come to the formal definition of an important mathematical concept relating to


an electrostatic field distribution, i.e. that of electric flux.
As remarked in Chapter 2, the term flux means flow, and has originated in the study of
fluid dynamics. If the vector E representing an electrostatic field were instead the velocity
v resembling the flow of a fluid in a fluid dynamics context, then the scalar product of v
and a vector area A would represent the amount of that fluid per unit time which flows
through the area A.
As we study electrostatics further, we will find that there is a physical significance
to the scalar product of the quantity 0 E and an area A, and we will, in memory of the
fluid dynamics analogy, give the name flux to that product. Once again however we must
emphasise that in electrostatics that nothing is really flowing we are just employing a
rather useful analogy as an aid to visualisation of the field.

3.8.2

Definition of flux density

We will now formally define in free space the flux density vector D by the equation
D=

0E

(3.12)

We should emphasise that this definition applies to free space only, and that in the
presence of material media, as discussed in Chapter 4, a new definition, which reduces to
the definition given above, will apply.
We have given this new variable the name flux density both in honour of international
standards and so that it will bear a logical relation to the concept of electric flux to be
introduced in the next section.

3.8.3

Definition of electric flux

We define the electric flux which crosses a surface S as the integral of the flux density
over that surface, i.e.
=

D ds

(3.13)

This is the same definition of flux as was introduced in Chapter 2, except that this
time it is applied to the D vector instead of a general vector F.
The reason we have bothered to define the concept of electric flux is that it participates
in a very general theorem known as Gauss law for the electric flux, and which will be
discussed in the next section.

3.9. GAUSS LAW FOR THE ELECTRIC FLUX

3.9
3.9.1

35

Gauss Law for the Electric Flux


Objective

Our objective in this section is to establish a general law of electrostatics known as Gauss
law for the electric flux, which states in words that for any closed surface of any shape
enclosing a distribution of charge, the total electric flux emerging from that surface is
equal to the total electric charge enclosed. We will approach the proof of the theorem by
a series of steps.

3.9.2

Point charge at the centre of a sphere

The theorem is easy to prove for the case of a point charge at the centre of a spherical
surface such as is shown in Figure 3.13.

Figure 3.13: Point charge at centre of spherical surface.


We note that for an element of area ds of the surface the direction of the surface
element and the direction of the flux density vector D are parallel, both being radially
outward, and the radial component of the D vector is given by
Dr =

0 Er

0Q

(3.14)
4 0 r2
Since this value is constant everywhere on the surface, all we have to do to obtain the
flux is to multiply it by the total surface area 4r2 . Thus
=

3.9.3

0 Q4r
2
0 4r

=Q

(3.15)

A more general surface

A quite arbitrary closed surface enclosing a point charge Q is shown in Figure 3.14.
In this case the charge is not at the centre of the surface because the surface is too
irregular to have one.
We can however make sure the point charge is at the centre of a spherical surface by
drawing as shown in Figure 3.15 such a surface, of radius a, centred on the charge.

36

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

Figure 3.14: A general closed surface surrounding a point charge.

Figure 3.15: Equality of flux over two surface elements.

3.9. GAUSS LAW FOR THE ELECTRIC FLUX

37

In the diagram we have also shown an element of area ds on the irregular surface, and
the projection of that element of area on the spherical surface.
The essence of the proof is that the amount of flux which emerges from the element
of area on the irregular surface is the same as the amount of flux which emerges from the
corresponding area on the spherical surface.
In comparing these two amounts of flux, we make the following observations.
The ratio of the magnitude of the flux density at the irregular surface to the flux
density at the spherical surface is (a/r)2 .
The flux density vector and the area vector at the irregular surface are not parallel,
but are inclined at an angle . If the area at the irregular surface were not inclined
at that angle , then the ratio of its area to the projected area at the spherical
surface would be (r/a)2 .
If the area at the irregular surface, inclined at the angle , were rotated so that it
kept the same projection on the spherical surface while the inclination angle became
zero, that area on the irregular surface would shrink by a factor cos .
All these factors combine to assure the equality of fluxes asserted above. Thus we conclude
for the irregular surface as well as for the spherical surface
=Q

3.9.4

(3.16)

Case of several charges

So far we have proven the theorem for a single point charge within an irregular but closed
surface. Because it did not matter where the charge was situated (other than that it
was inside) and because the field (and hence the flux) from a number of charges is the
superposition of the fields (and the fluxes) from the charges separately, we may conclude
the theorem provides that the flux from a number of point charges inside a closed surface
is equal to the sum of these charges, ie the total charge enclosed. Finally if we had
instead of a number of point charges, a continuous distribution of charge of a volume
charge density v within the closed surface we would, by modelling this distribution as a
number of small point charges, conclude that the theorem still applies.
i.e.
where

D ds = Q

Q=

v dv

(3.17)
(3.18)

Gauss law may be regarded as an alternative statement of the inverse-square centralforce property of electrostatic fields, expressed previously by Coulombs law, and which
is so much a part of the Faraday line of force concept discussed above.

3.9.5

Applications of Gauss theorem

The theorem we have just proven if of significance to us in two major ways. Firstly it
occupies a central role in the theory of electrodynamics, and is regarded in fact as one
of the four principal laws. Secondly it is of great practical use for determining, with
little eort, field configurations for a number of important charge distributions. The four
sections below will illustrate how this is done.

38

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

3.9.6

Definition of a Gaussian surface

A Gaussian surface is an imaginary closed surface which we introduce into a region for
the purpose of applying Gauss law and thereby drawing a conclusion about the nature of
the field on some part of that surface.

3.9.7

How we proceed

The basic method by which this is done involves the following steps.
We make a plausible guess as to the qualitative nature of the field distribution. For
example we may assume, based upon symmetry considerations, that it is in a particular direction, but we do not know the functional dependence on the coordinates.
We construct an appropriate Gaussian surface in the form of a closed surface which
encloses either all of the charge or at least a known portion of it.
Generally, the Gaussian surface is such that under the assumed field distribution,
flux will penetrate some of its surfaces, while flux will not penetrate other surfaces
because the assumed field lines are parallel to them.
We apply Gauss theorem so that we may obtain a quantitative relationship between
the field at some particular point and the charge enclosed.
We have thus obtained the missing part of the information describing the field
distribution, that is its functional form.

3.9.8

Flux from a line charge

We will apply this general method to the determination of the field distribution surrounding a uniform distribution of charge on a long thin straight conductor such as is shown in
Figure 3.16.

Figure 3.16: Flux from a line charge.

3.9. GAUSS LAW FOR THE ELECTRIC FLUX

39

In the diagram we have a line charge density of Cm1 . Surrounding a length l of


this line we construct a Gaussian surface in the form of a cylinder of radius r and closed
ends. We make the assumption that the electric field is purely radial, originating on the
charged conductor at the centre and passing through the curved surface. None of the field
passes through the disk shaped end surfaces as it is parallel to them. If we employ Gauss
law to equate the flux penetrating the closed surface with the total charge enclosed within
that closed surface we obtain
Dr 2rl = l

(3.19)

0 Er 2rl

(3.20)

which may be rewritten as


= l

and finally simplified to


Er =

3.9.9

2 0 r

(3.21)

Flux from a surface charge on an insulator

Our task in this section is to investigate the field distribution produced by a uniform
surface charge of Cm2 which might for example lie on a thin insulating membrane.
The situation we study is shown in Figure 3.17.

Figure 3.17: Flux from a surface charge on an insulator.


For this calculation we construct a Gaussian surface in the form of a pill box, in this
case a short cylinder of which each circular end cap has an area A and each such end lies
on the opposite side of the charge distribution. The height of the cylinder is unimportant
except that it is intended to be small in relation to the diameter.
Our assumption about the nature of the field distribution, in this case is that it
is directed on the top side vertically upward from the charge distribution, and on the
bottom side vertically downward (as we might expect from the symmetry of the charge
distribution).
We again notice that the surface is a closed surface consisting of a top surface a bottom
surface and a cylindrical side surface.
Under the assumption made for the field distribution, no flux will penetrate the side
surface but equal amounts of flux will penetrate the top and bottom surfaces. Equating

40

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

the flux to the total charge enclosed produces


2Dn A = A

(3.22)

2 0 En = A

(3.23)

which can be re-written as


and can be simplified to
En =

3.9.10

2 0

(3.24)

Flux from a surface charge on a metal

Our task in this section is to investigate the field distribution produced by a uniform
surface charge density of Cm2 which is lying on a matallic surface. The situation we
study is shown in Figure 3.18.

Figure 3.18: Flux from a surface charge on a metal.


For this calculation we construct a Gaussian surface again in the form of a pill box, in
this case again a short cylinder of which each circular end cap has an area A. One of the
end caps lies within the metal and one of the end caps lies just outside the metal in such
a way that the charge on the metallic surface lies within the pill box. The height of the
cylinder is unimportant except that it is intended to be small in relation to the diameter,
and the end cap outside the metal is intended to be close to the metallic surface.
Our assumption about the nature of the field distribution in this case is that, at the
end cap outside the metal, the electric field and flux density are directed perpendicularly
to and away from the metal surface, and at the bottom end cap there is no electric field
or flux density because that surface is within the metal.
Under the assumption made for the field distribution no flux will penetrate the side
surface because the direction of the field is tangential to that surface, and no flux will
penetrate the bottom surface because there is no field there. Equating the flux emerging
from the top surface to the total charge enclosed produces
Dn A = A

(3.25)

which can be re-written as


0 En

(3.26)

3.9. GAUSS LAW FOR THE ELECTRIC FLUX

41

and can be simplified to


En =

3.9.11

(3.27)

Solid metallic conductor

Our intention now is to apply Gauss theorem to answer the question of when we apply
a charge to a conducting body, where does it reside?
Figure 3.19 shows a somewhat egg-shaped solid metallic conductor on which a positive
charge has been deposited. The dotted line within the shaded region is intended to
indicate the boundary of a Gaussian surface which we have drawn within the conducting
solid object.
We first make use of the rule that in electrostatics, which is the study of the eects
of charges which are not moving, there will be no electric field, and hence no electric flux
density, within the conductor. This assumption can be made because when an electric
field does exist within a conductor it causes the motion of charge and we are then no longer
in an electrostatic context. Eventually the motion of charge will produce a rearrangement
of the charges to weaken the field and that field will in fact decay to zero, but such a
conclusion lies within the realm of circuit theory, not electrostatics.
As we can say there is no electric field or electric flux density within the conductor it
is very clear that the normal component of those variables on any Gaussian surface will
be zero. Gauss theorem then indicates that there can be no charge within that surface.
We are at liberty to expand the Gaussian surface until it reaches (but does not cross) the
boundary. We can therefore conclude that if the conductor does carry any charge, it must
all reside on the surface.

Figure 3.19: Solid metallic conductor.

3.9.12

Hollow metallic conductor

We now turn to the case of considering for a hollow metallic conductor the question of
where the charge must lie. In Figure 3.20 the shaded section represents the conductor and
the white section in the centre is a void. We have just established that in electrostatics
there can be no charge within the solid sections, but we could have charge somewhere on
the surfaces. Although we will later show that it cannot happen, let us for the moment
assume that there can be charge on the inner surface as shown in the diagram.
On the inner surface near the point A we have shown, for purpose of argument, some
positive charge. That positive charge will give rise to electric field lines flowing out from
that surface into the void. Field lines cannot just stop. It is a property of electric field

42

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

lines that they always start on a positive charge and end on a negative charge. Since
it is assumed that there are no charges floating around inside the void, and it is known
that the field lines cannot penetrate the metal surface, and it is further known that they
cannot form closed circles, they must terminate on negative charge elsewhere on the inner
surface, perhaps in the region near to the point B as shown.
To show that this situation cannot occur we really have to draw upon the insights of
the next section which discusses electrostatic potential. Therein it will be shown that the
line integral of the electric field between any two points must be independent of the path.
Between the points A and B we can construct two dierent paths. One is a path through
the void where the field lines are shown, and the other is a path through the metal where
we already know that in electrostatics there is no field. Since the line integral along the
first path is clearly non-zero and the line integral along the second path is clearly zero, we
have arrived at a contradiction which we can only resolve by saying that within the void,
and contradictory to the figure, there is in fact no electrostatic field. If there is then no
field within the void then the charges on the inner surface which we originally introduced
to explain the possibility of a field within the void must be absent. Thus we obtain the
important result that in a charged but hollow conductor all of the charge must reside on
the outside surface.

Figure 3.20: Hollow metallic conductor.

3.10

Electrostatic Potential

3.10.1

General observations

It is generally true that when we have a conservative force field we can introduce the
concept of a potential. For example in the conservative force field produced by gravity we
can define a concept called gravitational potential which is the work done per unit mass
in moving a mass from some reference point to some other point in space.
The electrostatic field has a similar property. It is, because it is also of the form of an
inverse square central force field, conservative. At any point in space we can define the
electrostatic potential as the work we must do, per unit charge, in moving a charge from
some reference point to the point in space at which we are defining the potential. The
reference point is considered to have a potential zero. In this definition we do not need
to specify the path for use in moving the charge, because the work done is independent
of the path.
In the above definition the concept of a reference point has appeared. The potential
at the reference point is by definition zero, but there is the question where we will place
the reference point in our definition. Two common conventions are in use.

3.10. ELECTROSTATIC POTENTIAL

43

In practical problems we often select the reference point to be somewhere on a large


sheet of metal which is connected to the earths surface. Again remembering that we are
in the context of electrostatics, we note that there are no electric fields within that metal
or within the earth or in the connecting wire, so all of those points will be at the zero of
potential if one of them is.
However for a theoretical development, it is not desirable to disturb the symmetry of
a charge distribution we may be studying by intruding into the space we are studying
with a conductor. In such cases therefore we employ another convention which elects the
reference point to be a point at infinity.
Now, you may ask, which point at infinity? There are lots of them! Too many in
fact! Well the answer is that it does not matter which one, because they are so far away
that the electric fields from any finite distribution of charge we are studying will have
decayed to negligible value there and even then any residual field is directed orthogonally
to any path we might follow in going from one point at infinity to another. So there is no
work done in getting from one point at infinity to another, and all of them will do as the
reference point.
A certain amount of care is needed in the introduction of the electrostatic potential, as
potential is a concept well developed in the abstract mathematical description of vector
fields, and the definition used in that abstract mathematical description is dierent from
the definition used in the electrodynamic context in an important way.
In the abstract mathematical theory, attention is focused on the forces exerted by the
field on an object and the work done by those forces when the object is moved. In the
theory of electrodynamics we take a more personal view. To move something we have to
exert a force. The force we have to exert is the opposite of the force which is exerted by
the field on the object. Since we are doing the work (often with very little pay) we feel
entitled to base the definition of potential on the work we do, and the forces we exert,
rather than on any forces exerted by the field. Of course the two forces we are discussing
are simply equal and opposite, so that the dierence between the two conventions is one
of a sign change, but it is necessary that we get it right.

3.10.2

Definition

In the light of the above discussion we define formally the electrostatic potential V at a
point r of space to be the work done by ourselves per unit charge in bringing a charge
from infinity to that point
V (r) =

8 r

qE(r) dr
q

(3.28)

Dividing by the common factor q gives


V (r) =

8 r

E(r) dr

(3.29)

It is hopefully understandable that people also give the definition of potential as the
work which we do in moving a unit (and therefore positive) charge from the reference
point normally at infinity to the point r in space.

44

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

3.10.3

Potential of an isolated charge

For a charge q, situated for convenience at the origin, the electric field is given by
E=

qr
4 0 r2

(3.30)

Choosing our zero of potential to be at a point at infinity, the potential at a point r


distant r from the charge is
8

r qr dr
(3.31)

4 0 r 2
The simplest calculation is obtained if we make our path along a straight line parallel
to r from a point at r = to the point r as shown in Figure 3.21. Then the equation for
the potential at the point r becomes
V (r) =

E(r) dr =

Figure 3.21: Path of integration for a simple calculation of the potential.

8 r

qdr
4 0 r 2
}
]
q r
=
4 0 r
q
=
4 0 r

V (r) =

(3.32)
(3.33)
(3.34)

This is the potential created by a charge q at a distance r from that charge.


A more elaborate calculation in which the starting point can be in any direction from
the origin, and the path can be any curve, produces the same result. You may care to
reflect upon the method for performing that calculation and upon that result.

3.10.4

Equipotential surfaces

A very important concept in electrostatics is that of equipotential surfaces. These are


simply surfaces in space which are orthogonal to the field lines. It is clear that if we move
around on an equipotential surface the force that we must exert on the charge (to keep

3.10. ELECTROSTATIC POTENTIAL

45

it from flying away out of our grasp) is orthogonal to the direction we are moving so
ocially no work is done. All points on that surface are therefore at the same potential
and deserve the name equipotential. For an isolated charge a set of equipotential surfaces
is shown in Figure 3.22.
This diagram seeks to illustrate some important points about equipotential surfaces.
These are
The surfaces are orthogonal to the field lines.
Where the field is strongest the equipotential surfaces for equal potential intervals
become closer together.
Potential goes up when we are pushing against the direction of the electric field. If
you let the charge move in the direction of the field, the potential is going down.

Figure 3.22: Equipotentials for an isolated charge.


So that we get practice in viewing the relation between equipotential surfaces and
field lines we have shown the equipotential surfaces and the field lines for a dipole in
Figure 3.23. Again we may note the crowding together of the potential surfaces where
the field is strongest.
As a further illustration of equipotential surfaces we show in Figure 3.24 the equipotential surfaces for the substantially uniform field which can be formed between two charged
conducting plates which we see in edge view in that figure. Here the substantially uniform
spacing of the equipotential surfaces in a uniform field is illustrated.
The equipotential potential surfaces for a completely uniform field are shown in Figure 3.25. Again we see only the edges of the equipotential surfaces which form planes
perpendicular to the field direction. Again we may note that in the uniform field the

46

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

Figure 3.23: Equipotentials for a dipole field.


spacing between equipotential surfaces for equal potential increments is constant. Again
we note that the potential goes up if you move against the field. Finally we note that the
spacing s between equipotential surfaces for a potential increment V is given by
|s| = |

V
|
E

(3.35)

Figure 3.24: Equipotentials for another field.

3.11

Capacitors and Capacitance

3.11.1

Concept of a capacitor

The general concept of a capacitor is illustrated in Figure 3.26. A more normal structure
is illustrated in Figure 3.27.
This structure has the benefit for analysis of providing a simple charge and electric
field distribution.

3.11. CAPACITORS AND CAPACITANCE

Figure 3.25: Equipotentials for a uniform field.

Figure 3.26: Concept of a capacitor.

47

48

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

Figure 3.27: A parallel plate capacitor.

3.11. CAPACITORS AND CAPACITANCE

3.11.2

49

Definition of capacitance

The equations relating for a fixed geometry the electric field and electric flux to the
charges which cause them are linear in those variables, we expect that for any capacitor
the charge stored on the electrodes will be proportional to the potential dierence between
them. Thus
QV

(3.36)

We introduce the capacitance C to turn this relation into the equation


Q = CV

3.11.3

(3.37)

Calculation of capacitance

In order to develop a formula for the capacitance C we will assume that the electric field
between the plates has the value E, and we will determine in terms of this variable both
the potential dierence between the plates and the charge on each of them.
The potential dierence of the positively charged plate relative to the negatively
charged plate is easily seen from the definition of potential to be given by V = Ed.
If the electric field between the plates is E, the associated electric flux density will be
D = 0 E. We have earlier shown in Section 3.9 that the relation between charge per unit
area on a metal surface and the adjacent outwardly directed electric flux density is
s = Dn =

0 En

(3.38)

For plates of area A the total charge on the positive plate is therefore
Q = As =

0 AEn

(3.39)

while the voltage between the plates may be expressed as


V = En d

(3.40)

Dividing the charge by the voltage gives the expression for the capacitance C
C=

3.11.4

Q
0A
=
V
d

(3.41)

Energy stored in a capacitor

In order to charge a capacitor from its initially uncharged state we will have to do work in
transferring a charge from the negative plate to the positive plate. We will leave for the
moment unspecified just how we get the charge to detach itself from the negative plate,
but it is true that that can be done without the performance of work. The real work is
done in moving the charge from one plate to the other, and it does not matter, because
we are in an electrostatic field, what path we take. We will consider the work to be done
incrementally, that is when the capacitor is partially charged with the voltage V we are
going to move some extra charge of amount q to increase the charge on one plate. The

50

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

amount of work done in moving that extra charge is V q and the total work done We in
establishing the final charge Q is therefore
We =

8 Q

V dq

(3.42)

If we take note that in a charging process the voltage V is always proportional to the
charge already deposited, ie V = q/C we have
8 Q

q dq
Q2
=
(3.43)
C
2C
0
The formula which we normally employ is easily obtained from this by the substitution
Q = CV and is
W =

1
We = CV 2
2

3.11.5

(3.44)

Capacitors in parallel

Quite often in practical circuits two capacitors are connected in parallel, as is shown for the
three capacitors in Figure 3.28. Clearly the capacitors have the same potential dierence
V between the negative and positive plates, and clearly for n capacitors the total charge
Q stored on the combination is the sum of the individual charges Q1 , Q2 , . . . Qn stored
on the n capacitors. Thus the capacitance C of the single capacitor which is equivalent
to the combination is given by
C = C1 + C2 + + Cn

(3.45)

This is the formula for combining capacitors in parallel.

Figure 3.28: Capacitors in parallel.

3.11.6

Capacitors in series

A dierent arrangement of connections between three capacitors is shown in Figure 3.29.


The capacitors are said to be connected in series. When n capacitors are connected in
series and the individual capacitor voltages denoted by V1 , . . . , Vn it is hopefully clear
from the diagram that the potential dierence V across the entire combination is the sum
of the individual potential dierences listed above. What the diagram also implies, but

3.12. ENERGY STORED IN AN ELECTROSTATIC FIELD

51

will require some justification, is that each capacitor carries the same charge Q, and also
that the two plates of each capacitor carry equal and opposite charges.
The justification for the second assertion can be found in the structure normally
adopted for the construction of practical capacitors, which have plates of separation very
small in relation to their transverse dimensions. It is a consequence of this that the electric flux which originates on one plate of the capacitor is confined almost entirely to the
interior of the capacitor, and in consequence almost all of the flux which originates on the
positive plate (and is equal to the total charge on that plate) terminates on the negative
plate, and being equal (apart from sign) to the total charge on that plate ensures that
equal and opposite charges reside on the two plates of a single capacitor.
The justification of the first assertion, namely that each capacitor carries the same
charge Q derives from the assumption that all capacitors were initially uncharged. As the
diagram shows, and the conservation of charge principle requires, the charge Q driven out
of the right hand plate of the first capacitor becomes the charge Q on the left hand plate
of the second capacitor, and so on.
If C is the capacitance of the single capacitor which can replace the series combination
of n capacitors, then we have Q = CV from which we may write
1
1
1
1
=
+
+ +
C
C1 C2
Cn

(3.46)

Figure 3.29: Capacitors in series.

3.12

Energy Stored in an Electrostatic Field

This section commences the study of the subject of energy storage in electromagnetic
fields. The context in which we are beginning our study is limited in two ways. Firstly,
we are in the electrostatic context, and secondly we are in free space, and not yet in a
dielectric medium as we will be in Chapter 4. Thus while the formulae we derive here are
valid in the presently limited context of electrostatics in free space, we will have to revise
the formulae when we enter the broader contexts of later chapters.
We have seen earlier in Section 3.11.4 that we must do an amount of work equal to
1
CV 2 to charge a capacitor of capacitance C to a potential dierence V . As we can by
2
discharging in the capacitor recover the work done, we regard that work as having created

52

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

stored energy of that amount. It is usual to regard the energy as stored in the electric
field within the capacitor. This view, which may seem at first somewhat arbitrary, will
become more compelling as the subject evolves. Let us now derive an expression in field
terms for the stored energy. We begin with the expression
1
1
(3.47)
We = CV 2 = QV
2
2
In this last formula we can relate Q to the electric flux density D and V to the
magnitude of the electric field E via the equations Q = AD and V = Ed. The formula for
stored energy then becomes
1
We = EDAd
(3.48)
2
To translate this to the concept of stored energy per unit volume Ue of the electrostatic
field we divide by the volume Ad of the space between the capacitor plates, which is the
volume of space containing the field, and obtain any of the three equivalent forms
1
2
0E
2

(3.49)

1
Ue = E D
2

(3.50)

Ue =

Ue =

D2
20

3.13

Static Electricity Machines

3.13.1

Introduction

(3.51)

A number of interesting machines designed for the development of very high potentials
make use of the phenomenon of discharge of ions under the influence of the intense electric
fields existing near a sharp point of a conductor at a high potential, and the subsequent
physical transport of the charges carried by those ions on moving insulated media.
Although in these machines charge is moved around, it is done so slowly that the eects
of charge movement covered in later chapters on magnetostatics and electrodynamics are
not evident, any more than they were in the movement of charge discussed at the beginning
of this chapter, and we will provide an analysis based on electrostatic equations.
The operating principles of these machines, when identified, also illustrate much of the
theory developed so far in this chapter. Time will permit us to discuss in Section 3.13.3
only the van der Graa generator.

3.13.2

Electrical discharge from points

The phenomena of the discharge of electricity from the sharp point on a conductor charged
to a high potential is illustrated in Figure 3.30. As will be shown in the section below, the
electric field near such a sharp point is very intense. The result is that molecules of air
close to the sharp point can become ionised, with the positive and negative ions moving in

3.13. STATIC ELECTRICITY MACHINES

53

dierent directions, this movement constituting a flow of electric current through the air.
In everyday life this eect may be viewed as a corona discharge from lightning conductors
or high voltage power lines. A diagram which is of assistance in modelling the eect of
electric fields near a sharp point is provided in Figure 3.31.

Figure 3.30: Electrical discharge from a sharp point.

Figure 3.31: Modelling the eect of a sharp point.


At the left of the diagram is shown a conductor with a radius of curvature at one end
much less than at the other. For the purpose of analysis, we will model the structure
as two spheres of respective radii r1 and r2 carrying respective charges q1 and q2 . The
spheres are assumed to be suciently far apart that they each have the uniform surface
charge density which each would have in the absence of the other, but they are connected
by a fine wire so that they are at the same potential. In expression of this last fact we
may write
q2
q1
=
4 0 r1
4 0 r2
from which we determine the charges are in the ratio
q1
r1
=
q2
r2
Now the total charges are related to the surface charge densities by
q1 = 4r12 1 ; q2 = 4r22 2

(3.52)

(3.53)

(3.54)

and substituting these results into the equation above we obtain


q1
r1
r2 1
=
= 12
(3.55)
q2
r2
r2 2
from which we can deduce the surface charge densities are related to the radii by

54

CHAPTER 3. ELECTROSTATICS IN EMPTY SPACE

1
r2
= .
(3.56)
2
r1
Noting that the electric field at the surface is proportional to the surface charge density
we have

Thus if r1
r2 then E1
a sharp point is justified.

3.13.3

E1
1
r2
=
=
(3.57)
E2
2
r1
E2 , and our assertion that the electric field is large close to

Van der Graa generator

The structure of the Van der Graa generator is illustrated in Figure 3.32, and its operation will be discussed in lectures.

Figure 3.32: Van der Graa Generator.

Chapter 4
ELECTROSTATICS IN MATERIAL
BODIES
4.1

Introduction

In studying the behaviour of various materials with respect to electric fields we find it
useful to classify the materials into the groups characterised below.
Conductors, of which the best conductors are generally metals, in which some of the
electrons are not tightly bound to their parent atoms but can move about under the
influence of electric fields and can travel far from their origin. We call these mobile
charges conduction charges.
Insulators, in which the atoms are tightly bound to their parent atoms, and while
those electrons may be pulled a little to one side of their rest position in an atom
or molecule, are not free to travel to distant locations. We will call any charges
involved in this slight distortion process bound charges.
In the Chapter 3 we noted that in conductors there can be no electrostatic field, as
any internal field temporarily established therein will cause current to flow to the surfaces
of the medium, and thereby bring to those surfaces conduction charges which make an
additional contribution to the internal electric field which eventually cancels the original
internal field, and the transport of conduction charge ceases.
In an insulator, this process of continuous charge transport for as long as there is an
internal electric field cannot occur, and it is possible for electrostatic fields to continue
to exist inside an insulator for an indefinitely large time. We use the alternative name
dielectric to signify that the material can sustain an internal electrostatic field.
In the above discussion, we have been referring to the internal electric field in such
a material. This may be a good time to emphasise that when the dielectric material
is of finite extent, there is generally also an electric field external to the material, and
although the internal and external fields are somewhat related, it is generally true that the
magnitude of the internal field is not the same as the magnitude of the nearby external
electric field.
Most dielectric materials behave in a particular way, characterised (when we finally
get to them) by linear equations. Such media are called simple dielectrics, and produce
results to be described in the following section. We should realise, however, that not all
55

56

CHAPTER 4. ELECTROSTATICS IN MATERIAL BODIES

dielectrics behave as simple dielectrics, and some surprising and non-linear phenomena are
possible. Although we will not have time to discuss these alternative forms of behaviour
in this course, we should not be bemused into the belief that all dielectrics are simple
dielectrics.

4.2

Experimental Observations

The easiest experimental observation which we can make about simple dielectric materials
is that if we insert as shown in Figure 4.1 such a material between the plates of a capacitor,
the capacitance of that capacitor is increased by a factor which is characteristic of that
material.
We will later, after analysing this situation, introduce a parameter known as the
dielectric constant of the material, and will show that the ratio of the capacitance with
the dielectric material present to the capacitance of the same capacitor when the dielectric
material is absent, is equal to that dielectric constant.

Figure 4.1: Capacitor with dielectric


As indicated in the section above, this observation is restricted to the case of simple
dielectrics, but those are what we study this year.

4.3

Model of a Dielectric

The molecules which can be found in simple dielectrics can be classified as either polar or
non-polar. The latter are simpler and will be discussed first. A simple representation of
a non-polar molecule, generally an atom, is shown in Figure 4.2. In this figure is shown a
small positive charge distribution surrounded by a larger and more diuse negative charge
distribution, of equal magnitude, but both are spherical and have the same centre. In
consequence there is no electric field produced externally to this charge distribution.
If this atom is placed in an electric field, for example one directed to the right, the
positive charge distribution will move a small distance in that direction, and the negative
charge distribution will move a small amount in the opposite direction, until an equilibrium is established with the force of attraction between the two charge distributions
being balanced by the forces exerted by the external field. The situation will then be as
depicted in Figure 4.3.
The essential characteristic of the situation is that the separation of charge centres
has produced for each molecule an elementary dipole as shown in Figure 4.4.
Many molecules, however, have a dipole moment even in the absence of an external
field. Such molecules are termed polar molecules. However thermal agitation ensures that
in the absence of an electric field the dipole moments are randomly oriented, as shown

4.3. MODEL OF A DIELECTRIC

Figure 4.2: A non-polar atom

Figure 4.3: Non-polar atom behaviour in an internal field

Figure 4.4: An elementary dipole

57

58

CHAPTER 4. ELECTROSTATICS IN MATERIAL BODIES

at the left in Figure 4.5, and on a macroscopic scale no nett electric dipole moment is
evident.

Figure 4.5: Polar atom behaviour in an internal field


When, however, an internal electric field is present in the material, there is a tendency
for the randomly oriented dipoles to align themselves, to some extent, with the internal
electric field, as shown on the right of Figure 4.5. We will regard this situation, on a
macroscopic scale, as equivalent to that for non-polar atoms, in that we will say that
each of the randomly oriented dipoles acquires an increment in its dipole moment, and,
relative to the situation present before the internal electric field is introduced, develops a
dipole moment again of the type illustrated in Figure 4.4.

4.4
4.4.1

Charge Movement in Dielectrics


Classification of charges

In our earlier studies of electrodynamics in free space we considered all charges to be of


the same type (although not of sign). When we begin to consider the eects of electric
fields in material media, it is useful to recognise that various groups of electrons have
dierent freedom to move. We might recognise three situations for an electron.
In free space outside of any medium an electron is free to move under the eect of
electric fields, and can move a large distance.
Inside a metallic conductor, an electron may be a valence electron and may be only
very loosely bound to a particular atom and may be substantially free to move large
distances throughout the metal.
Inside an insulator, an electron may be tightly bound to a parent atom, and while
it may move a little under the influence of an electric field internal to the material,
it does not leave the atom.
In understanding the behaviour of both conducting and insulating media, we will
find it convenient to classify electrons according to whether they are free to move large
distances or are bound to parent atoms. The charges of the first group, those free to move

4.4. CHARGE MOVEMENT IN DIELECTRICS

59

large distances, will be called conduction charges, and their density per unit volume will
be denoted by qvc , while a total charge of this kind in a definite volume will be denoted by
Qc . The charges of the second group, those bound to atoms, will called bound charges,
and their density per unit volume will be denoted by qvi , while a total charge of this kind
in a definite volume will be denoted by Qi .

4.4.2

Definition of polarisation

We begin our study of dielectric media with a reminder of our model of an elementary
dipole shown in Figure 4.4 as a pair of charges q and q which are separated by a vector
distance 2d, the vector being directed from the negative charge to-ward the positive
charge.
In material media we consider that when an internal electric field E is not present,
the positive and negative charges within an atom have a spherical distribution with the
same centre, and are equal in magnitude, and so produce no external field. When an
electric field is introduced into the material, there is a tendency for the positive charge to
move, to a limited extent, in the direction of the field, and the negative charge to move,
to a limited extent, in the opposite direction, and thus the centres of the positive and
negative charge distributions may become separated so that each atom becomes a dipole
of the form shown in Figure 4.4 and which is aligned along the direction of the field. The
strength p of each dipole is given by
p = 2qd

(4.1)

The units of p are clearly Cm. If in a polarised medium there are N such dipoles each
of strength p created per unit volume, we say the medium has a polarisation P given by
The units of P are clearly Cm2 .

4.4.3

P = Np

(4.2)

Induced charge crossing a plane

In the following section we are going to prove a number of important results about charge
movement within dielectrics under the influence of internal electric fields.
The first and most basic of these is that when an field is introduced into a dielectric,
it is accompanied by a consequential limited movement of charge across any plane which
the electric field intersects. The movement is not a current; when the field has stopped
changing, the charge movement comes to an end with charges in new positions but no
longer moving. This result will be established for non-polar atoms in Section 4.4.4, and
for polar atoms in Section 4.4.5.
Two further results, each depending on this basic result, will be established in Section 4.4.6 and Section 4.4.7. The first of these derived results is that when the polarisation
is non-uniform within a dielectric, there is in any closed volume of the material a nett
induced charge inside that volume caused by the fact that more bound charges may enter
some parts of the bounding surface of that volume than leave through other parts of the
bounding surface of that volume. The second of these derived results is that when the
dielectric material is of finite extent, there will be, on the surfaces of the medium, an

60

CHAPTER 4. ELECTROSTATICS IN MATERIAL BODIES

induced charge per unit area of the surface, this induced surface charge being related to
the magnitude and direction of the polarisation of the medium inside the surface.

4.4.4

Non-polar atoms

Figure 4.6: Plane in polarised dielectric


When a material becomes polarised, positive charges move in one direction and negative charges move in the opposite direction, with the result, as will be explained below,
that there is a net movement of charge across any plane in the material perpendicular to
the direction of polarisation.
To determine the net charge which crosses an area A, we assume that both the area
and the electric field are directed rightwards, and we consider the number of polarised
atoms which lie as shown in Figure 4.6 in a slice of thickness d and area A just to the left
of the plane, and also the number of polarised atoms which lie within a slice of thickness
d and area A just to the right of the plane.
Any positive charges which were in the left hand slice before the electric field was
applied will be found after the electric field has been applied to have crossed the plane
in a rightwards direction. Any negative charges which were in the right hand slice before
the electric field was applied will be found after the electric field has been applied to have
crossed the plane in a leftwards direction. Thus as each of these atoms will contribute a
nett amount q to the movement of charge rightwards across the area A. In this way we
conclude that the total charge Q which crosses the area A is given by
Q = (qdA + qdA)N

(4.3)

Recalling our definition of P as 2qNd we see that this may be written


Q = PA

(4.4)

Considering the more general case when the normal vector of the plane A is not
parallel to the direction of polarisation, we arrive at the result for the charge crossing the
vector area A of any plane
Q=P A

(4.5)

4.4. CHARGE MOVEMENT IN DIELECTRICS

4.4.5

61

Polar atoms

In the light of our previous discussion of polar atoms we will assusme that the same result
holds.

4.4.6

Induced charge entering a closed volume

Referring to Figure 4.7 the rectangular box shown has only two faces, each of area A,
through which induced charge passes, as the polarisation is parallel to the other four
faces. We conclude that the total induced charge which enters the box is
Qi = (P 1 P 2 ) A

(4.6)

Figure 4.7: A rectangular volume in a non-uniformly polarised dielectric


In Figure 4.8 we have an irregular closed surface S and a spatially non-uniform polarisation. The total induced charge which enters the closed surface S is given by
i

Q =

P ds

Figure 4.8: A general volume in a non-uniformly polarised dielectric

(4.7)

62

CHAPTER 4. ELECTROSTATICS IN MATERIAL BODIES

4.4.7

Induced charge at a boundary

When the polarised medium comes to an abrupt end as shown in Figure 4.9, we consider
the charge crossing a plane placed as close as we please to the end of the medium but
still inside it to conclude that at the surface of the medium (or just inside it) there is an
induced surface charge density

Figure 4.9: Induced charge at dielectric surface

is = P n

(4.8)

is a unit vector normal to the surface and pointing outward.


where n

4.5

Flux and Flux Density

4.5.1

Flux density in free space

In free space we have defined in Section 3.8.2 the flux density to be


D=

4.5.2

0E

(4.9)

Flux density in a dielectric

Now we generalise the above definition of electric flux density, which applied to the context
of free space, to be when dielectric media are present
D = ( 0 E + P)

(4.10)

Since in free space P = 0, this definition is consistent with the free space definition
given above, and we can therefore now use this definition as applying in all contexts.

4.5.3

Flux in free space

In free space we had the definition of electric flux


=

D ds =

0E

ds

(4.11)

4.6. GAUSS LAW WITH DIELECTRICS

4.5.4

63

Flux in a dielectric

In the presence of dielectric media we will follow international standards and retain
=

D ds

(4.12)

as the definition of electric flux. It will in consequence bear the expected relation to
electric flux density D. Because however the simple equation D = 0 E no longer applies,
there will be interesting consequences to the various ways in which Gauss law in the
presence of dielectrics is expressed. These will be discussed in the next section, which
should be studied with care, as it sometimes presents diculty. The diculty arises in
part because it is an unfortunate but common practice to use a single identifier to denote
various dierent quantities, the task of deciding which one is intended at each usage being
left to the reader. (Poor reader!)

4.6
4.6.1

Gauss Law With Dielectrics


Classification of charges

We have already discussed this matter in Section 4.4.1, and recognised that the charges
in a dielectric can be classified as either conduction charges or bound charges (sometimes
called induced charges).
We now introduce notation for the dierent types of charges and charge densities just
mentioned, for use on occasions when we wish to remove the ambiguities discussed above.
In our notation we denote the type of charge by a superscript. Thus the conduction
charge is denoted by Qc , and the induced (i.e. the bound) charge is denoted by Qi . The
sum of these charges is called the total charge and is denoted by Qt . Naturally
Qt = Qc + Qi

(4.13)

In the same way, when we are concerned with charge densities per unit volume, the
conduction charge density per unit volume is denoted by c , and the induced (i.e. the
bound) charge density per unit volume is denoted by i . The sum of these charge densities
is called the total charge density per unit volume and is denoted by t . Naturally
t = c + i

(4.14)

The importance of this classification is that we as engineers have dierent types of


control over the dierent types of charge. For example, we can put conduction charge on
to a body, as we did in the generation of static electricity, and move that body about at
will if we want to by applying forces. On the other hand, induced charges move strictly
in response the local electric field. Certainly we can take steps to influence the electric
field, and in so doing influence the induced charges, but our control over those charges
is significantly more indirect than over the conduction charges. Thus as engineers we
might wish to express our fundamental laws in a form in which conduction charges are
prominent. In particular we will investigate in the next sections whether there is a form
of Gauss law involving only conduction charges.

64

CHAPTER 4. ELECTROSTATICS IN MATERIAL BODIES

4.6.2

Gauss law in free space

We recall from Chapter 3 that Gauss law, expressed in terms of the electric field rather
than the flux density, takes the form
-

ds = Q =

0E

dv

(4.15)

It is entirely appropriate that the law be expressed in terms of electric field, as Gauss
law may be though of as a re-writing of Coulombs law, and Coulombs law is the basic
physical law relating electric field and charge.

4.6.3

Continued applicability

When we consider what happens inside a dielectric medium we say that the same law
applies, provide we interpret the right hand side as the total charge Qt , as all of the
charges are equally eective in producing electric field. Introducing this clarification, we
write Gauss law inside a medium as
-

ds = Q =

0E

4.6.4

t dv

(4.16)

Conduction charge form

Now suppose that we wish to have only the conduction charges appear explicitly in the
equation. We can achieve this by subtracting the induced charges from each side to obtain
-

0E

ds Qi = Qt Qi = Qc

(4.17)

On the left hand side we have a term Qi , but we recall that we have earlier in
Section 4.4.6 in considering movement of bound charges within a dielectric established
that the induced charges in a closed volume are related to the polarisation P by the
equation
Qi =

P ds

(4.18)

Substituting this expression for Qi on the left hand side of Equation 4.17 we obtain
-

0E

ds +

P ds = Qc

(4.19)

Now we recall that the combination 0 E + P has been defined in the general case as
the electric flux density D, so the above equation simplifies to
-

D ds = Qc

(4.20)

This equation may be called the engineering form of Gauss law, in contrast with
equation
-

0E

ds = Qt

(4.21)

4.7. CHARACTERISATION OF DIELECTRIC MEDIA

65

which may be called the physical form of Gauss law. It is important to realise that
the two forms represent the same law, just written in dierent terms suited to dierent
purposes. There is no new physical fact implied by our having two forms of the law.

4.6.5

Traps for the unwary

We now have to recognise that in may contexts, the careful superscript notation introduced
above to clearly distinguish what groups of charges are referred to in the various forms
of Gauss law is not always employed. It is in text books very common to omit the
superscripts, and leave it to the reader to determine from the context and his or her clear
understanding of the subject just which group of charges in referred to by the unadorned
symbol Q or . Thus you may find Gauss law written in the form
-

0E

ds = Q

(4.22)

and you are supposed to realise that the charge Q on the right hand side is the total
charge. Alternatively you may find Gauss law written in the form
-

D ds = Q

(4.23)

and you are supposed to realise that the charge Q on the right hand side is only the
conduction charge.
While it is highly unfortunate that this confusing and contradictory usage of symbols
occurs, if is well established, and there seems to be no cure for it other than to understand the subject so completely that you may instantly make a correct interpretation of
the symbols Q or whenever they occur as representing either Qt or Qc or t or c as
appropriate.
Sorry, but thats the way it is! I didnt do it. I cant kill the people who started it, as
they are already dead.

4.7
4.7.1

Characterisation of Dielectric Media


General discussion

From our discussion of polarisation given in Section 4.4.2 it appears reasonable to suppose
that in the type of medium described the polarisation vector P will be in the direction of
the internal electric field E and will have its magnitude proportional thereto.
While this is certainly a good approximation to the behaviour of a wide range of
materials, is should come as no surprise that when atoms are packed as closely are they
are to form a solid, the interactions between them are much more complex than was
contemplated in that simple description given earlier, and a wider range of behaviour
than simple linear response of the polarisation to the internal electric field is possible.
It must be realised that the internal structure of solids varies from a single crystal, to
polycrystalline to amorphous. In the first two of these structures molecules may contain
large numbers of atoms with alternative sites for some of them. The dipole moments of
atoms may have a preferred direction along the crystal lattice, or atoms or electrons may
have a limited number of stable positions within the crystal lattice. Atoms or individual

66

CHAPTER 4. ELECTROSTATICS IN MATERIAL BODIES

electrons may switch between alternative sites under the influence of an electric field,
but it may take a particular field strength before such movement occurs. Thus while
simple elastic and reversible distortion of the atomic structure of the type discussed in
Section 4.4.2 may indeed occur, we must be prepared to encounter materials where the
response of the polarisation to electric field is of a much more complex nature, including
anisotropic, non-linear and irreversible behaviours.

4.7.2

A word of warning

In the next section we will discuss the behaviour of so-called linear media, wherein the
polarisation vector P is simply proportional to the internal elenctic field E.
However the discussion above should have made us aware that the resulting linear
model is by no means universal for dielectric media, and that there are important materials
for which other forms of behaviour can and desirably do occur.

4.7.3

Dielectric susceptibility

In the case when the polarisation P is proportional to and aligned with the electric field
E we write
P = e 0 E

(4.24)

as a way of introducing the dielectric susceptibility constant e for the material. The
insertion of the physical constant 0 in the above equation makes e a dimensionless
constant, and is done so that this will occur. Values of e for various common materials
range from of the order of one or less to about ten.

4.7.4

Dielectric permittivity

The linear relation between P and E shown for simple dielectrics in the equation above
leads for such materials (but not for all) to a simple linear relation between P and E.
Recalling that we have defined D in all media as
D=

0E

+P

(4.25)

+ e 0 E

(4.26)

we can write for simple media


D=

0E

.
Defining for simple dielectrics
r

= 1 + e

(4.27)

we can write for simple dielectrics


D=

r 0E

(4.28)

.
Finally we introduce the dielectric permittivity

of the material to be

4.8. ANALYSIS OF CAPACITOR

67
=

r 0

(4.29)

and we can then write


D= E

(4.30)

We should note that whereas the relative dielectric permittivity r is a dimensionless


number, the units of dielectric permittivity are F/m, the same as those of 0 .

4.7.5

Dielectric constant

An alternative and very common name for the relative dielectric permittivity is the dielectric constant.

4.8

Analysis of Capacitor

We now return to the analysis of the situation with which we began this chapter, namely
to behaviour of a capacitor when a dielectric medium rather than free space separates the
plates.

Figure 4.10: Fields and charges in a capacitor with dielectric


The capacitor is shown again in Figure 4.10, but this time we have augmented the
drawing to show not only conduction charges Qc and Qc but induced charges Qi and
Qi on the adjacent surfaces of the dielectric. The capacitor plates still have an area
A and separation d. The magnitudes of the electric field and the electric flux density
between the plates are denoted by E and D. The total conduction charge on the left plate
is Q, and the potential of that plate relative to the right plate is V .
The relation between the electric field and the voltage is unaected by the presence
of the dielectric so we have
V = Ed

(4.31)

However, the electric field within the dielectric now produces a rightward polarisation

68

CHAPTER 4. ELECTROSTATICS IN MATERIAL BODIES


P = e 0 E

(4.32)

which produces on the surfaces of the dielectric the induced surface charge densities
of minus and plus is given in magnitude by
is = P = e 0 E

(4.33)

There is also a rightward electric flux density D given by


D=

r 0E

(4.34)

This flux density must be equal to the surface density of electric conduction charge on
the left plate. Multiplying by the plate area A we obtain the charge Q on left top plate
as
Q=

r 0 EA

(4.35)

If we now substitute for the voltage V from Equation 4.31 and the charge Q from
Equation 4.35 in the defining equation
Q
V

C=

(4.36)

for the capacitance we obtain


C=

r 0A

(4.37)
d
We see that the capacitance has been increased by a factor r relative to the case when
no dielectric material is used.
It is probably reasonable to attribute this increase in capacitance to the eect of
induced surface charges on the dielectric surfaces. These induced surface charges are of
opposite sign to the conduction charges on the adjacent plates, and serve (because they are
of opposite sign to the adjacent conduction charges) to weaken the internal electric field,
relative to the case when there is no dielectric but the plates carry the same conduction
charge. The weaker electric field reduces the voltage for a given charge, thus increasing
the capacitance.

4.9

Stored Energy in Dielectrics

As was foreshadowed in Section 3.12 our formulae for energy stored in a capacitor and
in an electrostatic field may have to be revised when dielectric media are present. It is
our task now to generate the more general formulae. It is our intention to cast the new
formulae in a form which automatically simplify to the previously obtained formulae when
dielectric media are not present.
The previously derived formulae
We =
and

8 Q
0

V dq

(4.38)

4.9. STORED ENERGY IN DIELECTRICS

69

1
We = CV 2
(4.39)
2
for the energy stored in a capacitor need no revision as the arguments used in their
derivation still apply. However when we translate the latter formula, by substituting for
the capacitance in terms of the dimensions and material within the capacitor, to terms
which involve fields within the capacitor, we find that we now have
1
2
(4.40)
r 0 E Ad
2
Noting that the volume of the capacitor is still Ad, we interpret this equation as saying
that the energy stored per unit volume of the field is
We =

Ue =

1
2

r 0E

(4.41)

An alternative version of this formula is


1
(4.42)
Ue = E D
2
Either of the last two formulae can be used for the case when a dielectric material is
either present or absent, as they simplify in the absence of such a medium to the formulae
derived for the free space case. Thus these are the formulae to learn.

70

CHAPTER 4. ELECTROSTATICS IN MATERIAL BODIES

Chapter 5
CURRENT AND RESISTANCE
5.1

Introduction

In previous chapters we found that an electric field places forces on charges, and in some
materials called conductors those charges are free to move. In this chapter we will study
some of the eects of charges moving under the influence of electric fields. The principal
eects we study here are the capacity of electric field to produce a current, and the capacity
of that current and the electric field together to produce conversion of energy per unit
time into heat in the material. The study of the other principal eect of a current, ie
its capacity to produce a magnetic field, will be deferred to a later chapter. In the work
of this chapter we will give further illustration to a concept already encountered, namely
that of volume current density.
In certain materials (but not in all) we will find that the amount of current flow
involved in a volume current density is proportional to the electric field. We will call such
media linear resistive media, and for those media take note of that linearity in a law which
is known as Ohms Law. We will also in that process define some new concepts called
conductivity, resistivity, conductance and resistance. We will also take note of a circuit
element called a resistor and deduce the formulae for combining resistors in series and
resistors in parallel.

5.2
5.2.1

Charge Transport in Conductors


Introduction

We have noted that in materials which we call conductors some of the electrons are not
tightly bound to the parent atoms but are free to move about. In that movement they
are subject to substantial thermal agitation, with a random distribution of velocities
which are quite quite high compared with the velocities which we might calculate in the
phenomena to be studied in this chapter, and capable of wandering far from the parent
atoms. In that thermal agitation the electrons make frequent collisions with the crystalline
or polycrystalline lattice containing the parent atoms. In the process there is a transfer
of energy and momentum between the mobile charge carriers and the fixed atoms of the
lattice.
Mostly the charge carriers are electrons, that is they carry a negative charge. If they
were positively charged they would be pushed in the direction of the field, but as they
71

72

CHAPTER 5. CURRENT AND RESISTANCE

are negatively charged the field will exert a force in the opposite direction, and they will
tend to move in that opposite direction.

5.2.2

Electron drift

When an electron in a vacuum is in the region of an electric field it experiences a force


as noted above in the opposite direction to that of the field, and hence an acceleration,
and thus a steady change per unit time in its velocity, in that direction.
When, however, an electron in a conductor is in the region of an electric field, its
acceleration does not produce much of a change in velocity before a collision with the
lattice occurs. For clarity we will assume that the electric field is directed to the right. In
that collision there is an exchange of momentum and energy between the charge carrier
and the atoms of the lattice. We cannot here go into detail, but the result is that the
electron loses the velocity increment which it had received as a result of its acceleration
for the limited time between collisions under the influence of the electric field, and is
brought back eectively to the unaccelerated condition. After the collision it begins again
to acquire a rightward motion, only to lose that motion soon after in a further collision.
The result is that instead of acquiring a steady acceleration as an electron in free space
would do, an electron in a conductor acquires, as an average over time containing many
collisions, a steady driftvelocity vd .

5.2.3

Electron mobility

In many (but not all) media the drift velocity is proportional to the electric field. In that
case we can introduce a constant of proportionality between the electron drift velocity
and the electric field E and write
vd = e E

(5.1)

where we have inserted a minus sign to take note of the fact that vd and E will be in
opposite directions, and we want our constant of proportionality to have a positive value.
The constant of proportionality e is called the electron mobility.

5.2.4

Volume current density

We now consider the situation depicted in Figure 5.1 in which a bar of conducting material of length L and cross-sectional area A has an internal electric field E directed for
definitness to the left. We will suppose that there are n electrons per cubic metre. The
bar will then contain a number of electrons
N = nAL

(5.2)

Each of which carries a charge q where q is the magnitude of the charge on an


electron. Under the influence of the electric field the electrons will acquire a rightward
drift velocity of
vd = e E
and after a time T given by

(5.3)

5.2. CHARGE TRANSPORT IN CONDUCTORS

73

Figure 5.1: Conducting bar with an electric field

T =

L
vd

(5.4)

where vd is the magnitude of the velocity vd , all of these electrons will have passed
out of the right face of the bar. In so doing they will have carried a charge Q = qN =
qN AL rightward through that face. The same charge will cross any cross sectional plane
in the bar. Thus the amount of charge per unit time crossing any cross sectional plane in
a leftward direction is
Q = qN = qnAL

(5.5)

Now this charge per unit time flowing in a leftward direction is called a current I in
that direction so we write the above equation in the form
I=

Q
qnAL
=
= qnAvd
T
L/vd

(5.6)

Now the current will be uniformly distributed over all parts of the area A, so the
amount of current per unit area, which we call the volume current density J , is given by
J = qnvd

(5.7)

We notice that by now the dimensions of the bar have dropped out of the equation,
and we have an equation which is characteristic of the material and not the dimensions
of any particular specimen of it. To make progress we substitute for the magnitude vd of
the drift velocity and obtain
J = (qne )E

(5.8)

We observe that in this sort of material, in which the drift velocity is assumed proportional to the electric field, we have deduced that the volume current density is in the
direction of and is proportional to the magnitude of the electric field. The constant of proportionality in brackets in the equation above is dependent upon the fundamental physical
constant q and the material properties n and e .

74

CHAPTER 5. CURRENT AND RESISTANCE

5.2.5

Conductivity

We give the name conductivity and the symbol to this constant and write
J = E

(5.9)

= qne

(5.10)

where

We note that and has units of S/m.

5.2.6

Resistivity

Sometimes we choose to rewrite the equation above so that the field is defined in terms
of the volume current density, ie in the form
E = J

(5.11)

where
1
1
=

qne
is known as the material resistivity and has units of ohm metre.
=

5.2.7

(5.12)

Ohms law

The relation in equation 5.9 we have just deduced is known as the field form of Ohms
Law. It is simply an assertion of the proportionality between the volume current density
and the electric field which causes it.

5.3
5.3.1

Lumped Circuit Concepts


Introduction

Frequently we are not concerned with the details of the charge transport within the
material, but merely with the total current flowing in the bar, and the potential dierence
between the ends of that bar. We will investigate these matters below. Again we assume
that the field is directed to the left, and the electrons are moving to the right, producing
a transport of negative charge to the right which has the same eect as the transport of
positive charge to the left. We notice that the field and the transport of positive charge
are both in the same direction.

5.3.2

Total current

The total current I flowing, in the leftward direction, is just the product of the volume
current density and the cross sectional area A, i.e.
I = JA

(5.13)

5.4. POWER DISSIPATION

75

or
I = AE

5.3.3

(5.14)

Potential dierence

If the current flow is to the left and the field is directed to the left, then the right hand
end of the bar will be at a higher potential than the left. The potential dierence V is,
for a uniform field, the product of the field and the distance so
V
L
If we substitute this last equation into the one above we obtain
E=

(5.15)

V
A
I = A
=
V
(5.16)
L
L
In this last equation the quantity in brackets is called the conductance G of the bar
and we write
I = GV

(5.17)

where the conductance G is give by


A
(5.18)
L
The units of conductance are Siemens, abbreviated S. Alternatively we may write the
voltage in terms of the current in the form
G=

V = IR

(5.19)

where
1
L
L
=
=
(5.20)
G
A
A
is called the resistance R of the bar. Units of resistance are Ohms, abbreviated .
R=

5.3.4

Ohms law again

Either of the equations 5.16 or 5.19 is called the lumped circuit form of Ohms Law.

5.4

Power Dissipation

5.4.1

Lumped circuit form

As the charge carriers move in the field, the force exerted by the field gives them additional
kinetic energy, which is subsequently dissipated in the form of heat with the frequent
collisions with the crystal lattice. When a charge Q has moved from the right end of the

76

CHAPTER 5. CURRENT AND RESISTANCE

bar to the left end of the bar in time T through the potential dierence V the energy
given by the field to the charge in subsequently converted to heat is
W = QV

(5.21)

The power dissipated P is the rate W/T at which energy is converted to heat and is
thus given by
QV
W
=
= VI
(5.22)
T
T
Since Ohms Law allows us to express, for linear resistive media, the voltage in terms
of the current and vice versa, we can express the power dissipated in any of the alternative
forms
P =

P = I 2R = V 2G

5.4.2

(5.23)

Field form

The power is dissipated uniformly throughout the volume of the material. We can therefore obtain an expression for the power dissipated per unit volume Pv of material by
dividing by the volume. Thus
w

Ww

P
VI
V
I
=
=
Pv =
(5.24)
AL
AL
L
A
In the above formula we recognise the first factor as the magnitude of the electric field
E and the second factor as the magnitude of the volume current density J . Thus
Pv = J E

(5.25)

Pv = J 2 = E 2

(5.26)

We notice that in this form the dimensions of the bar have dropped out of the formula.
We assume therefore that this formula is appropriate for calculating the power dissipated
per unit volume of the material of any size and shape.
Because Ohms Law in field form allows us to express the volume current density J
in terms of electric field E and vice versa, we may re write the above result in any of the
alternative forms

These expressions are known as the field forms of power dissipated per unit volume.

5.5
5.5.1

Resistors in Combination
Resistors in series

It will be shown in lectures that when a number n of resistors are placed in series as
shown in Figure 5.2, the value of the single resistor equivalent to the combination is given
by
R = R1 + R2 + + Rn

(5.27)

5.5. RESISTORS IN COMBINATION

77

Figure 5.2: Resistors in series

5.5.2

Resistors in parallel

Figure 5.3: Resistors in parallel


It would be shown in lectures that when a number n of resistors are placed in parallel
as shown in Figure 5.3 the value of the single resistor equivalent to the combination is
given by
1
1
1
1
=
+
+ +
(5.28)
R
R1 R2
Rn
This expression may be re-written in terms which involve the conductances rather
than the resistances of the resistors. In that form the relation is
G = G1 + G2 + + Gn

(5.29)

The formula for resistors in series provides a clear expression of the fact that there is
a common current and an overall potential dierence which is the sum of the individual
potential dierences.
The last formula for resistors in parallel provides a clear expression of the fact that
there is a common potential dierence and an overall current which is the sum of the
individual currents.
The similarity between the first and third formulae above is an illustration of the
principle of duality which often occurs in physical theory and occurs here in a simple way
in lumped circuit theory.

78

CHAPTER 5. CURRENT AND RESISTANCE

Chapter 6
MAGNETOSTATICS IN EMPTY
SPACE
6.1

Introduction

We will take note of the fact that some naturally ocurring materials, which we call
magnetic materials can exert across empty space a force on one one another which
is dierent from the electrostatic force in ways which we will identify.
We will attribute this force to a new type of field which we call the magnetic field
H.
We will also find that magnetic fields can be caused by electric currents, and that
they can exert forces on wires carrying electric currents.
We define the study of magnetstatics in empty space as the study of the forces which
cross empty space between magnetic materials and between wires carrying steady
electric currents.
Just as we did with electrostatics, we place a description of what is happening inside
magnetic materials in another and later chapter.
In our study, we will find it convenient to define a new variable, associated with the
magnetic field H, and we will call that new quantity the magnetic flux density, and
will denote it by B.
We should be wary of calling B the magnetic field. Although some text books do
this, it is at variance with Standard International terminology, and causes confusion
to students.

6.2
6.2.1

Experimental Observations
Early history

Small magnets exist. You can get them from the mineral lodestone, or from the
hardware store.
79

80

CHAPTER 6. MAGNETOSTATICS IN EMPTY SPACE

Figure 6.1: A bar magnet.


An example of a bar magnet is shown in Figure 6.1.
You can use such materials to make compass needles.
Compass needles show alignment in: the field of another magnet, in the field of the
earth, or in the field of a steady current.
These compass needles are not aected by electrostatic fields, but, although responding to a dierent type of field, behave like small dipoles in that they experience
torques which tend to align themselves in a particular direction in that new type of
field which is called a magnetic field.
In a bar magnet, the field is strong near the ends.
We call those end regions poles and find that there is a north-seeking pole, which
we call a north pole and a south-seeking pole, which we call a south pole.
It appears that the earth itself produces a magnetic field which is responsible for the
alignement of suitably suspended magnetic material in a compass so that the north
pole of the magnet points roughly to-wards the north geographic pole of the earth,
and the south pole of the magnet points roughly to-wards the south geographic pole.

6.2.2

Forces between poles

Like poles repel and unlike poles attract.

6.2.3

Exploring the Field

Figure 6.2 shows the field lines near a magnet. We cannot on our present knowledge know
the field lines within the magnet, as we have as yet no way of penetrating it. Although it
is true that the fields which originate from the poles are of dipolar type, we will have to
defer proving this until we have assembled more of the laws, and illustrated more of the
phenomena.
Small compass needles can be used to map the field pattern. They in an obvious
way give the direction, and if we want to become quantitative, we can measure the
torques to determine the relative strength of the field at dierent points.
People have studied the force eects near the poles separately, by making the magnets long.

6.2. EXPERIMENTAL OBSERVATIONS

81

Figure 6.2: Field lines near a bar magnet.


Away from the cause of the field, the field lines seem to provide in the Faraday field
representation the same expression of the field in magnitude and direction as do the
electric field lines do of the electric field, i.e. they spread out in the same way.
It follows that away from the cause of the field once we know the way the field
modifies itself in direction, we can determine the way it changes in strength.
Iron filings can help get an idea of the field picture. This will be illustated in a
forthcoming section.

6.2.4

Dipolar behaviour

Figure 6.3: Division of a bar magnet.

Exploring the field shows that, at least at a distance, the magnetic field of a small
magnet it is the same shape as an electric field of a small electric dipole.

82

CHAPTER 6. MAGNETOSTATICS IN EMPTY SPACE


We find other eects reminiscent of dipolar behaviour - inverse square law near to
a pole, and inverse cube law when we are far away.
The dipolar nature of magnets is reinforced by the fact that if cut a magnet in an
eort to separate the poles we just get two more dipoles as in Figure 6.3.
We note that no-one has made a magnetic field which resembles the electric field of
an isolated charge.
It is also found that a small loop of wire carrying a steady current has at a sucient
distance a magnetic field of the same shape, viz dipolar.

6.2.5

Iron filings experiment

Figure 6.4: Iron filings experiment.


Describe the experiment.
Give some explanation.

6.2.6

Magnetic eects of a current

After Volta had invented the battery, generation of significant currents became possible,
and the eect that an electric current can cause a magnetic field was discovered by
Oersted. The eect was explored experimentally and quantitively by Ampere, and also
by Biot and Savart.
Figure 6.5 shows the discovery of Oerstead. In the figure is shown a long straight wire
carrying a current, with the field created by the wire being explored by a series of small
compass needles. Although the conservation of charge requires that the current in the
wire be provided with a return path, in which the current flows in the opposite direction,
the use of long wires allows the return path to be placed at some distance, and we are

6.2. EXPERIMENTAL OBSERVATIONS

83

Figure 6.5: Exploring field lines near a long straight wire.


prepared to believe that the field close to the wire is caused by the currents in the nearby
wire.

Figure 6.6: Map of field lines near a long straight wire.


An illustration of the field as mapped by Ampere using the Faraday representation,
but of course really requiring a three dimensional visualisation, is shown in Figure 6.6

6.2.7

Field classification

It will hopefully be recognised that this is an entirely new type of field. Whereas in
Chapter 3 we saw that electrostatic fields caused by charges are of source type, as was
illustrated in Figure 2.4 and we have just asserted above (and we will later prove it)

84

CHAPTER 6. MAGNETOSTATICS IN EMPTY SPACE

that the magnetostatic fields created by magnetic material are also entirely source type,
the magnetic fields created by electric currents are of vortex type, as was illustrated in
Figure 2.5. Thus, at the risk of repeating ourselves, we will summarise these results below.
The magnetic field H caused solely by magnetic materials is source type.
This means that when the field is represented by field lines in the Faraday representation, the field lines always begin in regions called sources and end in places called
sinks, and that none of the field lines form closed curves.
The magnetic field H caused solely by electric currents (whether they be conduction
currents, induced currents or total currents) is entirely vortex type.
This means that when the fields is represented by field lines in the Faraday representation, there are no regions where the field lines start or stop, and that all of
the field lines form closed loops.
The magnetic field H which is caused partly by magnetised material and partly by
electric currents has both source and vortex properties. This means that when the
field is represented in the Faraday representation, some of the field lines can start
or stop at sources, and some of the field lines can form closed loops.
.

6.2.8

Other eects

Shortly after the discovery that electric currents can produce a magnetic field it was also
discovered that a current carrying wire (or a moving electric charge) can experience a
force in a magnetic field. The direction of the force is orthogonal to the direction of the
current (or the moving charges) and is also orthogonal to the direction of the magnetic
field. The mathematical formulation of this eect will be given in a later section.
Even later, it was discovered by Faraday that a magnetic field changing in time, or the
action of moving a conductor in a magnetic field, can produce an electric field, and what
is called an induced voltage around a circuit containing that moving conductor. Because
this eect is concerned with magnetic fields or conductors which are changing in time, it
is outside of the context of magnetostatics, which is the subject of the present chapter,
and apart from a brief mention in the following section, no further discussion of this eect
will occur until Chapter 8.

6.3

Measurement of Magnetic Field

Before setting out in the following sections the quantitative laws of magnetostatics it is
appropriate that we follow our empirical tradition by at least mentioning some of the
ways in which magnetic field can be measured, and thus the ways in which the laws to be
stated below can be verified.

6.3. MEASUREMENT OF MAGNETIC FIELD

6.3.1

85

Torque on compass needle

If we were living long ago, and had only a piece of lodestone as a magnetic material, we
could fashion a compass needle, make it a component of a torsion balance, and measure
the strength of various magnetic fields in terms of the torque which they exerted on this
small magnet. Of course such measurements would be only comparative, as we would not
know the strength of our naturally occurring magnet, and not yet have any theoretical or
practical background against which we could determine it.

6.3.2

Torque on a small loop

Following the discovery that a current carrying wire experiences in a magnetic field a
force perpendicular to the wire, we may easily conclude that a small current carrying loop
will experience in a magnetic field a torque, and again using a torsion balance to measure
magnetic fields in terms of their eect on a loop of wire of known size carrying a known
current.

6.3.3

Electron ballistics

An alternative method of measurement would be to explore not the force produced by a


magnetic field on a current carrying conductor, but instead the force on a charged particle
moving in a vacuum by seeing to what extent the trajectory of such a particle is influenced
by a magnetic field. Of course in that method of measurement we would not be measuring
the magnetic field in a small region, but rather some sort of average of the magnetic field
over the entire trajectory of the particle.

6.3.4

Hall eect

Yet another method of measurement of magnetic field, again depending on the eect that a
charged particle moving in a magnetic field experiences a force orthogonal to the direction
of the current, is to make use of the Hall eect, and which derives from the fact that the
charge carriers which provide the current in a piece of wire immersed in a magnetic field
are pushed to one side by that transverse force, and so establish a measurable potential
dierence between the two sides of the wire. This phenomenon allows the construction of
compact magnetic field probes which can explore the magnetic field over a small region,
will be discussed in detail in Section 6.9.2

6.3.5

Faraday induced emf

Probably the most common method of measurement of magnetic fields makes use of the
phenomenon of electromagnetic induction discovered by Faraday, the essence of which is
that a conductor which is moving within a magnetic field (or is stationary in a magnetic
field which is changing with respect to time) will experience an induced voltage around
a circuit of which that conductor forms a part. This phenomenon which will be discussed in detail in Chapter 8 allows the construction of compact magnetic field exploring
instruments.

86

CHAPTER 6. MAGNETOSTATICS IN EMPTY SPACE

6.3.6

Summary

We have in the preceding sections discussed a number of magnetic field measurement


principles, these being based on physical phenomena which range from those which fit into
the magnetostatic context we are discussing in this chapter to those which are based on
electrodynamic laws not yet fully expounded. At least some of these provide a respectable
basis for the empirical confirmation of the relationship between current and magnetic field
to be provided in the next section.

6.4
6.4.1

Fields Produced by Currents


Introduction

In this section we produce two formulae for the magnetic field H produced by current.
Before we do this however we must note the several things listed below.
These formulae apply only to that part of a magnetic field which is produced by
current, which as we have stated before is vortex type.
They do not apply to any magnetic field H caused by magnetised material. We
will have to use other formulae to be introduced later (probably in Chapter 7) to
calculate a magnetic field produced by magnetised material, which we have stated
earlier to be source type.
Steady electric currents must, in view of the conservation of charge law, flow in
closed circuits. It is really not possible to produce an empirically verifiable formula
for the magnetic field produced by an incomplete circuit, as such a circuit cannot
be constructed in nature. If such a formula appears in the below sections to have
been produced, it is subject to one or another of the restrictions below.
In the case of a formula to be given below for the magnetic field produced by a
current in a long straight wire we are assuming that the length of the wire is very
great compared with the transverse distance from the wire at which we calculating
the field, and that the return path for the current is at a similarly large distance.
In the case of the law of Biot and Savart, to be given below, the formula for the field
produced by a current element is really only meaningful after it has been integrated
over the complete circuit of which the current element forms a part.

6.4.2

Field of a long straight wire

The experiments of Ampere showed that the magnetic field H produced by a long straight
wire forms closed circles centred on and surrounding the wire, and is inversely proportional
to the radial distance from the wire. Such a field distribution was shown in Figure 6.6.
Amperes discovery can then be written as
H

I
r

(6.1)

6.4. FIELDS PRODUCED BY CURRENTS

87

The Standard International system of units has been devised and has been defined in
such a way that the constant of proportionality implied in the above equation is 1/(2)).
Thus in the S.I. system of units
I
(6.2)
2r
We can see from this equation that in the S.I. system the units for magnetic field are
Am1 .
H =

6.4.3

The law of Biot and Savart

The above formula serves well to illustrate the relation between current and magnetic field
for a simple geometry, but does not tell us how to calculate the magnetic field distribution
produced by a current I flowing in a circuit of more complex shape. Biot and Savart,
working at about the time of Ampere, devised a formula which does this.

Figure 6.7: Co-ordinates for Biot and Savart Law.


To explain the law, we use as shown in Figure 6.7, the position vector r1 to denote
a point on a closed circuit C and dr1 to denote a vector element of length along that
contour, and we use the position vector r2 to denote a point in space at which we wish
to calculate the magnetic field. Then the law states that the contribution d H(r2 ) to the
magnetic field at r2 due to the current element composed of I and dr1 is
dH(r2 ) =

Idr1 r12
2
4r12

(6.3)

r12 is a unit vector (unit


In the above equation, r12 is the vector r2 r1 from r1 to r2 ,
magnitude, no dimensions) in that direction, and r12 is the magnitude of r12 .
In the figure, the plane of the paper is the plane containing dr1 and r12 . The vector
dH(r2 ) is perpendicular to that plane, varies in magnitude as sin , and inversely as the
square of the distance r12 .
The law indicates that the contribution to the field is inversely proportional to the
square of the scalar distance between the current element and the field point, and is

88

CHAPTER 6. MAGNETOSTATICS IN EMPTY SPACE

directly proportional to the sine of the angle between them. Thus the eect of the current
element is greatest when its direction is orthogonal to the direction between itself and
the field point. As shown in Figure 6.7 the right hand rule, with the thumb along the
direction of the current, allows the fingers to indicate the direction of the contribution to
the magnetic field.
Although it is good for us to get to know the above formula in its current element
form, we should remember, as we were warned in the previous section, that this formula
does not of itself apply to a physically realisable situation. The current I must flow in
the complete circuit C, and it is only the sum of all the contributions from all the current
elements which is physically meaningful. This sum is obtained from the integral below.

6.4.4

Field of a Complete Circuit

For a closed circuit following a contour C on which r1 is a position vector, the total
magnetic field at a point r2 is
H(r2 ) =

6.4.5

Idr1 r12
2
4r12

(6.4)

Caution

The physical meaning of the Biot-Savart law as expressed in equation 6.3 is restricted
to the notion that when it is integrated over a complete circuit, as in equation 6.4, it
correctly gives the magnetic field of that circuit.
As is not possible to obtain a current element in isolation, we should be cautious in
attaching a physical interpretation to equation 6.3 alone. If such an attempt is made, for
example, by using the formula and the law of force (to be introduced later) to calculate
the forces exerted between a pair of such current elements, it is found that the forces
calculated do not obey the reaction principle, whereas the forces between two charges
calculated by Coulombs law do. On the other hand, if a calculation of the forces between
two complete circuits is made, the result does obey the reaction principle.
We must therefore regard equation 6.4 as describing physical reality, while equation 6.3
can be considered as merely providing an expression for the kernel of the integral in the
complete result as expressed by equation 6.4.

6.5
6.5.1

Amperes Law of Magnetostatics


Introduction

We are now going to produce a very general law relating magnetic fields and currents
which produce them.
In many ways this law has a similarity to Gauss Law for the electric flux, in that
just as we were able to use that law to deduce electric fields in useful practical cases, so
we will be able to use the forthcoming law to deduce magnetic fields in useful practical
cases. There is however the considerable contrast that whereas Gauss law for the electric
flux dealt with the flux emerging from electric charges, and involved a surface integral,
Amperes law deals with the circulation around a current, and involves a contour integral.

6.5. AMPERES LAW OF MAGNETOSTATICS

6.5.2

89

The simple case

Figure 6.8: Magnetic field lines surrounding a long straight wire.


We return to the case of the magnetic field produced by a long straight wire which is
illustrated again in Figure 6.8 and given by equation
I
(6.5)
2r
In that figure we have selected the positive direction for the current to be upward. In
that figure we have also introduced a contour C in the form of a circle of radius r and
sensed in relation to the positive direction of the current by the right hand rule, that is if
the thumb points in the positive direction for the current, the fingers point in the positive
direction for the contour.
Let us now calculate the circulation of the magnetic field around the contour. The
result is clearly
H =

H dr = I

(6.6)

Where we are using dr to denote an element of the countour and not some change in
radius.

6.5.3

The general case

We are still considering only fields due to current.


We will establish that the circulation of the magnetic field H around a closed cntour
is the current enclosed within that contour.
The proof has been easy for a circular contour centred on a long striaght wire.
The result however has in fact great generality as described below.

90

CHAPTER 6. MAGNETOSTATICS IN EMPTY SPACE


Any closed curve will do, and the wire does not have to be striaght
The circulation of H is found to be constant and equal to the total current enclosed
The proof can be taken as empirical, or can be derived from the Biot and Savart
law, already discussed above, itself taken as an empirical law, by the methods of the
vector calculus, but that is too much eort at this stage of the course, so we take it
for now on trust, to be established more formally in the future.
The general result is called is Amperes law. It is one of the four fundamental laws
of electrodynamcs called Maxwells equations.
In the form just given it is valid only in the magnetostatic context.
In the electrodynamic context it will have to be modified. It will then be called
Amperes law as modified by Maxwell.
Beware of substitutes involving B.

6.6
6.6.1

Magnetic Flux and Flux Density


Magnetic flux density

In free space we define the magnetic flux denisty B associated with a magnetic field H as
B = 0 H

(6.7)

where 0 is a physical constant known as the magnetic permeability of free space and
has the value in the S.I. system of units 4 107 , and the units of henry per metre. (It
is given this value by definition, the definition having an eect on the size of the S.I. unit
of current.)
We may see some similarity between this definition and the definition of the electric
flux density
D=

0E

(6.8)

appearing in Chapter 3. There is the further similarity that, just as we needed to


revise the definition of electric flux density (but in a consistent way) when we encourered dielectric media, we will revise the definition of the magnetic flux density (but in
a consistent way) when we encounter magnetic media, but that will happen in the next
chapter.
There are two dierent names for the S.I. unit of magnetic flux density. The first is
weber per square metre, and has the advantage of exhibiting an obvious parallel with the
units of electric flux density, which are coulombs per square metere, and also has a clear
relation to the unis of magnetic flux, to be introduced in the next section, but which we
state here to be Webers. The alternative name for the units of magnetic flux density is
Tesla, and has been introduced to celebrate the name of Nikola Tesla.

6.7. MAGNETIC FORCES

6.6.2

91

Magnetic flux

If we have defined a surface S, either closed or not closed, and have a rule for deciding
the sense of each vector element of area of that surface, we define the magnetic flux
linking the surface as the integral over that surface of the scalar product between the flux
density vector B and each vector element of surface area, i.e.
=

B ds

(6.9)

We will find that magnetic flux and flux density are very important concepts in both
megnetostatics and electrodynamics, being associated both with forces on conductors and
with electromagnetically induced electric fields.

6.7
6.7.1

Magnetic Forces
Force on a moving charge

It is found experimentally that a force on a charge q moving with velocity v in a magnetic


flux density B is given by
F = qvB

(6.10)

When an electric field is also present the force is


F = q (E + v B)

6.7.2

(6.11)

Force on a current element

If we do not have a point charge q but a dierential charge element dq = dv in a volume


element dv, that charge moving with velocity v, we may write (when E is zero),
dF = dv v B

(6.12)

We may have seen earlier that v is just the volume current density J . Thus
dF = J dv B

(6.13)

Now if the volume element has small cross section, the filamentary current approximation in which J dv is equivalent to I dr, where v, J and dr are all in the same direction,
may be used. Thus for a filamentary current element I dr
dF = I dr B

6.7.3

(6.14)

Small loop in uniform field

Consider the small loop shown in Figure 6.9 carrying a current I in a substantially spatially
uniform magnetic flux density B. Since the loop is small, the flux density B can be
considered uniform over its extent and the total force is zero.

92

CHAPTER 6. MAGNETOSTATICS IN EMPTY SPACE

Figure 6.9: Small current carrying loop in a magnetic flux density.


It is easy to show, however, that the forces on the dierent sections of the loop, while
not producing a significant force, do produce a torque given by
T = I ds B

(6.15)

In this relation, ds is the vector area of the loop, sensed in relation to the reference
direction for the current, using the right hand rule. Intorducing the definition for the
magnetic moment m = I ds of the loop, the result above becomes
T = mB

(6.16)

This result may be compared with the similar result for the torque on a small electric
dipole in an electric field.
T =pE

(6.17)

Notice that despite the similarity, in one equation a field is involved while in the other
equation a flux density is involved. This lack of symmetry between the equations is further
reflected in a lack of symmetry between the units of p which are Cm, and m which are
Am2 rather than Wbm.

6.7.4

Force between parallel wires

It is a further experiemental observation, or can be deduced from the equations


above, that two long straight parallel wires distant s from one another and carrying
currents of magnitudes I1 and I2 experience a force of attraction per unit length
of one wire

6.8. SOME MAGNETIC FIELD DISTRIBUTIONS

93

Figure 6.10: Long straight parallel wires.


Force per unit
0 I1 I2
=
length of one wire
2s

(6.18)

The force is attractive if the currents are in the same direction and repulsive if the
currents are in the opposite direction.
This phenomenon is in fact used in the standard Internatioal System of units to
define the unit of current.

6.8
6.8.1

Some Magnetic Field Distributions


Interior of current carrying wire

A wire of radius a and carrying a current I directed out of the page and uniformly
distributed over its cross section is shown in Figure 6.11.
Please show as an exercise that insde the wire at a radius r the circumferential component of the magnetic field H is given by
H =

Ir1
2a2

(6.19)

I
2r2

(6.20)

while outside the wire it is given by


H =

A graph of this result is given in Figure 6.12.

94

CHAPTER 6. MAGNETOSTATICS IN EMPTY SPACE

Figure 6.11: Wire with uniformly distributed current.

Figure 6.12: Magnetic field variation with distance from wire centre.

6.8. SOME MAGNETIC FIELD DISTRIBUTIONS

6.8.2

95

A long solenoid

A solenoidal coil of wire of n turns per metre, each carrying the same current I, is shown
in Figure 6.13. Inside the solenoid the magnetic field H is approximately uniform and is
direced to the right. To calculate the value of the magnetic field H we apply Amperes
law to the contour ABCD shown in the Figure. Thus

Figure 6.13: Calculation of interior field of a long solenoid.


-

H dr = nIL

(6.21)

Dividing this contour into the four straight segments shown, and taking note of the
orthogonality of the field direction and the path direction on two of the segments, and
the fact that at the distant segment the field is expected to be negligible, gives
-

H dr =

8 Q
P

H dr +

8 R
Q

H dr +

= 0 + HL + 0 + 0
= HL

8 S
R

H dr +

8 P
S

H dr

(6.22)
(6.23)
(6.24)

Substituting this result into equation 6.21 above gives


H = nI

(6.25)

We note that this result has the correct units of A/m as n was the number of turns
per unit length of the solenoid and thus the units of n are m1 .

96

CHAPTER 6. MAGNETOSTATICS IN EMPTY SPACE

Figure 6.14: Coil on non-magnetic toroidal core.

6.8.3

Toroidal coil

A coil of N turns each carrying the same curent I and wound around a non-magnetic
toriodal core is shown in Figure 6.14.
Please use Amperes law to show as an exercise that the circumferential component
of the magnetic field H at a radius r within the toroid is given by
H =

6.8.4

NI
2r

(6.26)

Centre of circular loop

A circular loop of wire of radius a and carrying a current I in a counter-clockwise direction


is shown in Figure 6.15.

Figure 6.15: A circular current carrying loop.


Please use the Biot-Savart law to show as an exercise that at the centre of the loop
the component of magnetic field directed out of the page is given by
H=

I
2a

(6.27)

6.9. EFFECTS OF MAGNETIC FORCES

6.8.5

97

Axis of circular loop

A dierent view of the same single turn circular loop of wire of radius a and carrying a
current I is shown in Figure 6.16.

Figure 6.16: Position on the axis of a circular current carrying loop.


This time we are interested in the axial component of magnetic field at a point on the
axis of the coil, and a distance z from its centre. It may be shown as a more dicult
exercise that the axial component of magnetic field at that position is given by
Hz (z) =

6.9
6.9.1

I a2
2 (z 2 + a2 )3/2

(6.28)

Eects of Magnetic Forces


Introduction

We will study in this section some natural phenomena and some pieces of technology, all
depending for their existence upon the magnetic force eects described in Section 6.7.

6.9.2

Hall eect

Figure 6.9.2 shows a conducting strip, immersed in a magnetic flux density B directed into
the page, and carrying a current I direced vertially downward, the curent being provided
by charge carriers q each with a downward drift velocity vd .
As a result of the force exerted by the magnetic flux density on the moving charge
carriers, they will be deflected sideways in the diagram. The charge carriers experience a
force
F = qvd B

(6.29)

The analysis below shows that, for the directions of magnetic flux density and current
shown, the deflection of the charge carriers will always be to the right, independently of
the sign of the charge carriers.

98

CHAPTER 6. MAGNETOSTATICS IN EMPTY SPACE

Figure 6.17: Deflection of charge carriers in a magnetic flux.


Thus if q is positive, then vd is downward, and if q is negative then vd is upward. In
both cases F is to the right and the charge carriers are forced rightwards.
The right hand side of the material will take up a positive potential relative to the
left hand side if the charge on the charge carriers has a positive sign, and will take up
a negative potential relative to the left hand side if the sign of the charge on the charge
carriers is negative.
This phenomenonis known as the Hall efect, and is useful in showing that in most materials the sign of the charge carriers is negative, ie it is the electrons which are providing
the conduction mechanism.

6.9.3

Circulating charges

Figure 6.18 shows a charge carrier q with velocity v in the plane of the paper has entered
a region in which there is a uniform magnetic flux density B perpendicular to the plane
of the paper, and directed inward.

Figure 6.18: Circular motion of charge in a uniform magnetic flux density.


As a result of the interaction between the charge, its velocity and the magnetic flux
density, the charge carrier will experience a force orthogonal it its direction, and will

6.9. EFFECTS OF MAGNETIC FORCES

99

therefore travel in a circular trajectory in the plane of the paper. Equating the electromagmagnetically induced force with the inertial force required to produce the acceleration
in a circular path give the equations
mv 2
r
from which we may calculate the radius of the circular path to be
qvB = ma =

(6.30)

mv
qB

(6.31)

r=

The cyclotron resonant frequency f is given by


f=

m
2r
=
v
qB

(6.32)

We observe that this frequency is independent of both v and r. This fact has a
hopefully obvious practical importance.

6.9.4

The cyclotron

A sketch of the well known cyclotron particle accelerator which depends in part upon the
principle just outlined is provided in Figure 6.19.

Figure 6.19: Construction of the cyclotron.


The operation of the structure will be outlined in lectures. Significant points are

100

CHAPTER 6. MAGNETOSTATICS IN EMPTY SPACE

The device is intended to accelerate bunches of particles to hign energies.


It contains two hollow evacuated D-shaped electrodes, with a uniform magnetic flux
density B directed perpendicular to the plane of the diagram.
A high frequency alternating electric field accelerates groups of charged perticles
across the gap. The frequency of the electric field matches the cyclotron frequency
of the circular motion of the charges.
The magnetic flux density B keeps the particles in circular oribits.
The radius increases as the particles accelarate, but the clyclotron frequency remains
fixed.
The structure can achieve particle energies up to 20 MeV.

Chapter 7
MAGNETOSTATICS IN
MATERIAL BODIES
7.1

Introduction

While we have already considered in the last chapter some of the eects produced by
magnetic media in the free space surrounding such media, it is now time to consider
eects internal to the magnetic media themselves. Presentation of these ideas presents
considerable challenge because
Students may have been previously exposed to elementary treatments in which
terminology contradictory to that defined in international standards has been used.
Such exposure often leads to incorrect beliefs, the most common of which is that
magnetic fields are source-free, ie they are purely vortex type. Such an incorrect
belief is for an engineer a considerable handicap. The truth is that the magnetic
field H is partly source type and partly vortex type.
In our exposition we will be aided, to an extent, by having observed in our study of
the properties of dielectric media, some concepts which are parallel to those required for
an understanding of the present material.
In that treatment of dielectric eects, we were able to honour our empirical tradition
by giving, at the beginning, a definition of the concept of the dielectric constant in terms
of observations made externally to the medium. Following that definition we provided a
description of what is happening inside the dielectric medium.
Because of the order of chapters in these notes, we cannot in the case of magnetic media proceed in an exactly similar way. The external description of dielectric eects which
provided us with an empirical basis for measurement of dielectric properties depended
upon the concept of capacitance with which we were already familiar. The corresponding
external measurements on magnetic materials which throw light on their internal behaviour depend upon the concept of inductance which we will not meet until Chapter 8. In
this situation what we will do is:
provide a description of the internal behaviour of magnetic media which parallels
that given for dielectric media, and
leave until Chapter 8 the oering of experimental evidence that this description is
correct.
101

102

CHAPTER 7. MAGNETOSTATICS IN MATERIAL BODIES

7.2

Comparison of Dielectric and Magnetic Media

We will find that the behaviour of dielectric and magnetic media exhibit a striking number
of parallel features, as well as a number of important dierences. To better enable an
understanding of these similariteis and dierences, we make comparisons in the sections
below between these behaviours.

7.2.1

Points of similarity

Just as we defined the polarisation P to be the density per unit volume of electric
dipoles inside a dielectric medium, we will define a quantity magnetisation M to
be density per unit volume of magnetic dipoles inside a magnetic medium.
Just as we combined E and P to make the electric flux density
D=

0E

+P

(7.1)

we will combine H and M to make the magnetic flux density


B = 0 (H + M)

(7.2)

Just as the normal component of P at a material surface produces an induced surface


charge density that acts as a source or sink of electric field E, a normal component of
magnetisation M at a material surface creates an induced surface density of north
or south magnetic pole which acts as a source or sink of the magnetic field H.
Corresponding to the definitions of dielectric susceptibility e and relative dielectric
permittivity r we will have a definition of magnetic susceptibility m and relative
magnetic permittivity r .

7.2.2

Points of dierence

In the two definitions D = 0 E + P and B = 0 (H + M) we notice the appearance


of brackets in only one of these expressions.
The microscopic mechanisms underlying magnetism, at least in the strongly magnetic materials of iron, cobalt and nickel, are considerably dierent from these which
are responsible for dielectric behaviour, in ways to be discussed in the next section.

7.3

The Origins of Magnetism

The origins of magnetic eects lie in the special property of an electron that is it has, in
addition an electric charge, a magnetic dipole moment, of a fixed size. Thus an electron can
both create a magnetic field external to itself, and can experience a torque in a magnetic
field.
In many materials these magnetic eects are not evident, because the electrons tend to
pair up, either within an atom or within a molecule, so that their magnetic moments cancel

7.4. SUMMARY

103

and there is thus no external eect. In some special materials, however, the magnetic
moment of one atom tends to align itself with that of a neighbouring atom, and magnetic
eects become evident.
Despite this tendency for neighbours to become aligned, there are other, longer range,
forces at work inside a magnetic material, tending to produce misalignment. The result
is generally that groups of atoms of one hundred or so become mutually aligned, but
dierent groups adopt dierent orientations, with the result that on the average over
thousands of atoms, there is no coherent alignment.
When, however, a magnetic field is introduced inside the material, and that magnetic
field is steadily increased in strength, more and more of the magnetic dipoles become
aligned with that magnetic field, and the material develops a dipole moment per unit
volume which increases with the internal magnetic field, just as a dielectric material
develops an increasing internal electric dipole moment for unit volume as the internal
electric field increases.
There is however an important practical dierence. In dielectric media the electric
dipole moments of individual atoms are not fixed, and also it is usually impossible to
achieve in dielectric media complete alignment of the separate dipole moments of individual atoms. Thus as the internal electric field increases, so does the electric dipole moment
per unit volume, and no limit to the polarisation is reached.
However, in magnetic media, the magnetic dipole moments of those electrons which
participate in the production of magnetic eects are fixed, and in addition it is practicable
to produce inside the material a magnetic field of sucient strength to align them all.
When this has occurred, the magnetic moment per unit volume of the medium has reached
an upper limit called the saturation magnetisation, and cannot increase further. This
eect will be discussed further in a later section.

7.4

Summary

The simple fact is that the magnetic fields caused by currents are vortex type, and
the magnetic fields caused by magnetised materials are source type. When both
currents and magnetic material are present, the field is partly vortex type and partly
source type.
We are familiar from the previous chapter with the vortex type part of the field, so
we concentrate in this chapter with the source type part caused by the magnetic
material.
It is well known that there seem to be no isolated magnetic charges, but in the
interior of a magnetic medium there is a density per unit volume of magnetic dipoles.
The situation is similar to that in a dielectric medium in which there is a density
per unit volume of electric dipoles.
The origin of the magnetic dipoles is in the electrons in the incomplete shells of the
electronic structure of the crystalline lattice forming a magnetic material. It is a
basic property of electrons to have an intrinsic magnetic moment. In completed electron shells, the moment is not externally very evident, as the electrons in completed

104

CHAPTER 7. MAGNETOSTATICS IN MATERIAL BODIES


shells are arranged in pairs in which the magnetic moments are equal in magnitude
and oppositely directed.

In an incomplete shell, however there is an uncompensated electron magnetic moment, and the atom behaves as it is has a net magnetic moment which we will call
m. In the Standard International system of units, the units of magnetic moment
are Am2

7.5

Magentisation

It will be recalled that in dielectric media we constructed a concept called the electric
polarisation by multiplying the number N per cubic metre of electric dipoles, each of
strength p, by the strength of each dipole. Thus
P = Np

(7.3)

In a similar way we construct in magnetic media a concept called the magnetisation


M by forming a similar product
M = Nm

(7.4)

where N is now the number per cubic meter of magnetic dipoles each of strength m.

7.6

Magnetic Field Produced by Magnetisation

The calculation of the magnetic field caused by the magnetisation can follow, at least in
the context of a uniformly magnetised medium, the procedure already described in the
context of a dielectric medium. In that context we were able at each surface to identify
an induced surface charge density equal to P
n. In a small element of area ds on the
surface, there is an element of induced charge P
nds coulomb.
The electric field at a point r2 caused by the various elements of charge situated at
points r1 is then given by Coulombs law
E(r2 ) =

r12 ds
P n
2
4 0 r12

(7.5)

In the magnetic medium context we can identify an induced surface magnetic pole
density of M
n. The magnetic field H(r2 ) at a point r2 caused by the various elements
of magnetic pole situated at points r1 is the given by a corresponding formula
H(r2 =

r12 ds
Mn
2
4r12

(7.6)

We notice some dissimilarity between the two equations 7.5 and 7.6 in that the former
contains a denominator factor 0 , whereas the latter does not contain a corresponding
factor 0 .

7.6. MAGNETIC FIELD PRODUCED BY MAGNETISATION

7.6.1

105

Illustration

The process of production of electric or magnetic field by induced surface charges or


induced surface poles is illustrated in Figure 7.1.

Figure 7.1: Depolarising and demagnetising fields in dielectric and magnetic media
In the left of the diagram is shown a sample of dielectric material which is subject
to what we call an applied electric field E a caused by some agency, for convenience not
shown, but which could be conduction charges on a pair of plates placed one above and
one below the sample, and carrying equal and opposite electric surface charge densities.
The significance of the concept of applied electric field is that we placed the charges
which caused it. In the case illustrated in Figure 7.1 the applied electric field points
upward. The applied electric field is by no means the total electric field present, either
inside or outside the sample. It is merely one of the several possible contributions (two
in this case) to that total field, which we are going to denote by E t .
The other contribution to the total electric field is a field caused by induced surface
charges appearing as shown at the surfaces of the sample as a result of the sample becoming polarised with a polarisation P which we assume is also in an upward direction.
We could call this contribution to the total field the induced field E i , in recognition of its
having derived from induced charges, but it is more common to call it the depolarising field
E d , in recognition of the fact that inside the material it points in the direction opposite to
the polarisation, in this case downward, although outside the material it points upward.
The total electric field E t is of course the sum
Et = Ea + Ed

(7.7)

106

CHAPTER 7. MAGNETOSTATICS IN MATERIAL BODIES

both inside and outside the material. We note that inside the material the depolarising
field is in the opposite direction to the applied field and has the eect of weakening the
internal total field. Outside the material, the depolarising field points in the same direction
as the applied field, and has the eect of re-inforcing the applied field to produce the total
field.
In the right of the diagram is shown a sample of magnetic material which is subject
to what we call an applied magnetic field Ha caused by some agency, for convenience not
shown, but which could be a coil with a vertial axis and carrying a current I surrounding
the sample of magnetic material.
The significance of the concept of applied field is that we placed the currents which
caused it. In our case the applied magnetic field points upward. The applied magnetic
field is by no means the total magnetic field present, either inside or outside the sample.
It is merely one of the several possible contributions (two in this case) to that total field,
which we are going to denote by Ht .
The other contribution to the total magnetic field is a field caused by induced magnetic
poles appearing as shown at the surfaces of the sample as a result of the sample becoming
magnetised with a magnetisation M, which we assume is also in an upward direction.
We could call this contribution to the total field the induced field Hi , in recognition of
its having derived from induced magnetic poles, but it is more common to call it the
demagnetising field Hd , in recognition of the fact that inside the material it points in the
direction opposite to that of the magnetisation, in this case downward, although outside
the material it points upward.
The total field Ht is of course the sum
Ht = Ha + Hd

(7.8)

both inside and outside the material. We note that inside the material the demagnetising field is in the opposite direction to the applied field and has the eect of weakening
the internal total field. Outside the material, the demagnetising field points in the same
direction as the applied field, and has the eect of re-inforcing the applied field to produce
the total field.
It may be noted that the depolarising field points away from the induced postive
charges, so that in the figure it points, above the top surface, in an upward direction, and
below that surface it points in a downward direction.
In a corresponding way, when magnetic poles form on the surface of a magnetic material, as is illustrated in Figure 7.1, the magnetic field created by the north pole points,
above the top surface of the medium, in the upward direction, while the magnetic field
created by the north pole points, below the top surface of the medium, in the downward
direction.
We also note that in both cases the field created by the induced charges or induced
poles is normally not as strong as the original applied field, so that although the total
field inside the material is less in magnitude than the original applied field, it is still in
the original (upward in this case) direction, and the polarisation P or magnetisation M
which results from the total internal field E t or Ht is also in that direction.

7.7. MAGNETIC FLUX AND FLUX DENSITY

7.7
7.7.1

107

Magnetic Flux and Flux Density


Flux density in free space

In free space we have defined in Section 6.6.1 the magnetic flux density to be
B = 0 H

7.7.2

(7.9)

Flux density in a magnetic medium

Now we generalise the above definition of magnetic flux density, which applied to the
context of free space, to be, when magnetic media are present,
B = 0 (H + M).

(7.10)

Since in free space M = 0, this definition is consistent with the free space definition
given above, and we can therefore now use this definition as applying in all contexts.

7.7.3

Magnetic flux in free space

In free space we had the definition of magnetic flux


=

7.7.4

B ds

(7.11)

Magnetic flux in a magnetic medium

In the presence of magnetic media we will follow international standards and retain
=

B ds =

0 (H + M) ds

(7.12)

as the definition of magnetic flux. It will in consequence bear the expected relation to
magnetic flux density B.

7.7.5

Reasons for definition

The reasons for adopting in Chapter 4 the defnition


D=

0E

+P

(7.13)

for the electric flux density vector were not there explicitly stated, but might be stated
now to be so that we would have an electric vector which bore a direct relation to the
conduction charge density rather than the total charge density.
The reasons for adopting the definition
B = 0 (H + M)

(7.14)

for the magnetic flux density vector are not entirely parallel. They are set out below.

108

CHAPTER 7. MAGNETOSTATICS IN MATERIAL BODIES

Although the forces on charge carriers moving in a vacuum can be calculated either
from the flux density vector B or the equivalent quantity 0 H, when we determine,
by means of the Hall eect, the forces on charge carriers flowing inside a magnetic
medium, we find that it is the magnetic flux density vector B which is no longer
equal to 0 H, which gives the correct experimental result.
In the next chapter we will be introduced to Faradays law of electromagnetic induction. It has been established experimentally that changing magnetic field H and
changing magnetisation M are equally eective in inducing electric fields. So it is
by using the combination
B = 0 (H + M)

(7.15)

that we achieve the most succinct expression of Faradays law of electromagnetic


induction.

7.8
7.8.1

Characterisation of Magnetic Media


General discussion

From our discussion of the origins of magnetism given in Section 7.3 it appears reasonable
to suppose that in some magnetic media the magnetisation vector M will be in the
direction of the internal magnetic field H and will, at least until saturation beigns to
occur, have its magnitude proportional thereto.

7.8.2

A word of warning

In the next section we will discuss the behaviour of so-called linear media, wherein the
magnetisation vector M is in the direction of and is simply proportional to the internal
magnetic field H. However the discussion in Section 7.3 of magnetic saturation should
have made us aware that the resulting linear model of magnetic media will only apply for
small internal magnetic fields, i.e. those producing a magnetic moment per unit volume
whose value is well below the saturation magnetisation value.

7.8.3

Magnetic susceptibility

In the case when the magetisation M is proportional to and aligned with the magnetic
field H we write
M = m H

(7.16)

as a way of introducing the magnetic susceptibility constant m for the material. We


note that as M and H have the same units, this constant is dimensionless. Values of m
for various common materials range from almost zero to several thousand. The dierence
between this behaviour and the corresponding behaviour for dielecric media should be
noted.

7.9. LARGE SIGNAL BEHAVIOUR OF FERROMAGNETIC MEDIA

7.8.4

109

Magnetic permeability

The linear relation between M and H shown for simple magnetic media in the equation
above leads for such materials (but not for all) to a simple relation between M and B.
Recalling that we have defined B in all media as
B = 0 (H + M)

(7.17)

B = 0 (H + m H)

(7.18)

we can write for simple media

.
Defining for simple magnetic media
r = 1 + m

(7.19)

we can write for simple magnetic media


B = r 0 H

(7.20)

.
Finally we introduce the magnetic permeability of the material to be
= r 0

(7.21)

B = H

(7.22)

and we can then write

We should note that whereas the relative magnetic permeabilty r is a dimensionless


number, the units of magnetic permeability are henry per metre, the same as those of
0 .

7.9
7.9.1

Large Signal Behaviour of Ferromagnetic Media


Linear media

In previous sections we introduced a simple linear model for the relation between magnetic
field and magetisation in linear magnetic media. The most important magnetic media,
the so-called ferromagnetic media, however, do not exhibit this simple behaviour.

7.9.2

Ferromagnetic media

An illustration of the relation between the magnetisation and the internal magnetic field
for such a medium is given in Figure 7.2. Notable features of the graph are listed below.
The high slope for small fields.
The eventual saturation in which the magnetisation reaches a constant value when
all of the dipole moments of the individual atoms are fully aligned.

110

CHAPTER 7. MAGNETOSTATICS IN MATERIAL BODIES

Figure 7.2: Magnetisation vs magnetic field inside a ferromagnetic medium


The fact that the curve followed for increasing field diers from the curve for decreasing field.

7.9.3

Measurement of M - H curves

The measurementof M - H curves such as are shown in Figure 7.2 is based on the capacity
of a changing magnetic flux to induce a voltage in a coil which surrounds the material
in which the flux change is taking place. This phenomenon is known as Faradays law of
electromagnetic induction, and is one of two basic elelctromagnetic laws to be discussed in
the following chapter. Further discussion of this eect is logically postponed until then.

7.10

The Source Free Nature of B

We intend to establish in this section the important result that the magnetic flux density
B has no sources, i.e. it has only vortices. This result is considered to be one of the
fundamental prinicples of electromagnetic theory.
In the study of electrostatics we found that the polarisation and the induced charge
densities had the relation
-

P ds = Qi

(7.23)

and that this equation, when added to Gauss Law in form involving all sources, viz

7.10. THE SOURCE FREE NATURE OF B


-

0E

ds = Qt

111
(7.24)

gave a form of Gauss Law in which the only sources are those of conduction charge,
viz
-

( 0 E + P) ds = Qc

(7.25)

In magnetostatics a similar result ensues when we combine 0 H and 0 M to form


the flux density vector B. If we look at the vector M from a source point of view, we find
that the north poles occur in regions where there is, as shown in Figure 7.3, a net inflow
of the vector M just below the top surface which is not balanced by a corresponding
outflow of M above the top surface.

Figure 7.3: A north pole in a magnetic material


Thus in the region of the poles there is a net inflow of the vector M. Expressed
dierently, there is a net outflow of the vector -M. But a net outflow of a vector from a
region indicates that there are sources of the vector within that region. Thus we conclude
that the sources of the vector -M are the magnetic poles, just as the sources of the vector
- P were the induced charges.
But we have already stated that the sources of the magnetic field vector H are the
magnetic poles. Taking note then that the vectors H and -M have the same sources, we
can conclude that if we add the vectors M and H we should have no sources, i.e. the
vector B = 0 ( H + M) is source-free. The situation is similar to that which occurred
in the study of electrostatics, in which we concluded that by adding the vectors 0 E and
P we obtained a vector in which the induced sources were eliminated, leaving only the
conduction charges as sources. In the magnetostatic case since no-one has found any free
magnetic charges, the only sources of H are the poles, and if we eliminate them in a
vector by combining H with M, the resulting vector has no sources.
The statement that the magnetic flux density vector B is source-free may be embodied
in the equation

112

CHAPTER 7. MAGNETOSTATICS IN MATERIAL BODIES


-

B ds = 0

(7.26)

In words, the equation states that the net outflow of magnetic flux from a closed surface
is zero. This equation is known as Gauss Law for the magnetic flux, and qualifies as one of
the four Maxwell equations which between them embody the principles of electrodynamics.

Chapter 8
ELECTRODYNAMICS
8.1

Introduction

To date we have made separate studies of the topics of electrostatics and of magnetostatics,
both within and outside material media. It is now time to take note of the discoveries of
Faraday and Maxwell which deal with time varying fields.
The histories of these discoveries are interestingly dierent. Faradays discovery was
experimental, and was made after years of painstaking search for evidence of an eect
which he intuitively believed should exist. Maxwells discovery was of a purely mathematical nature, and derived from his examination of the laws of electrodynamics as they
were then believed, and noting that they were inconsistent with basic mathematical principles. Having noticed the inconsistency, Maxwell hypothesised a correction to the basic
equations, and then went on to pursue the further mathematical consequences of that correction. These consequences included the prediction of electromagnetic waves, a prediction
soon confirmed experimentally by Hertz, and the era of electromagnetic communication
was born.

8.2

Electromagnetic Induction

After a search covering o and on more than a decade, Michael Faraday was able to
confirm experimentally his belief that, not only could electric currents produce magnetic
fields, but magnetic fields produced either by electric currents or by magnetised materials
could be used to produce electric currents. What Faraday discovered was that an essential
ingredient of the process was that the magnetic field had to be changing with time. The
two discoveries of Faraday are illustrated in Figure 8.1 and Figure 8.2.
In Figure 8.1 the switch S is used to allow the passage of, or to interrupt, a current i
in coil one. A galvanometer connected in series with coil two shows, when the switch S
has just been closed, a momentary deflection in one direction, and when later the switch
S is opened, a momentary deflection in the opposite direction.
In Figure 8.2 the galvanometer shows no eect while the magnet is stationary, but is
deflected in one direction while the magnet is being brought towards the coil, and in the
opposite direction while the magnet is being moved backward away from the coil.
Exploring these things further, Faraday established that
113

114

CHAPTER 8. ELECTRODYNAMICS

Figure 8.1: Electromagnetic induction by changing currents

Figure 8.2: Electromagnetic induction by moving magnetic material

8.3. CONDUCTOR MOVING IN A FIELD

115

The potential is induced by a change of either the magnetic field H or the magnetisation M in the combination of the magnetic flux density B = 0 (H + M).
The potential induced around a closed circuit is proportional to the time rate of
change of the magnetic flux linked by that circuit.
When motion is involved, it is relative motion between the coil in which the currents
are being induced and the agency producing the magnetic field which induces the
electric field or potential which can produce the eect.
The question of in what direction the emf is induced is obviously of interest. The
clearest statement of how to determine that direction was given by Lenz, who determined
that when an emf is induced in a circuit by a change in the flux linked by that circuit,
the emf will be induced in the direction such that any current which it causes to flow in
that circuit will be in a direction so as to produce a contribution to the flux linked by
that circuit which is in a direction so as to oppose the original change. This statement is
known as Lenzs Law.

8.3

Conductor Moving in a Field

The law of electromagnetic induction can in some cases be seen to have a strong relation
with laws described in Section 6.7 giving the forces on charge carriers moving in a magnetic
field. This matter is illustrated first in Figure 8.3, in which a conductor (shown shaded)
is moving leftwards in a uniform magnetic flux density B directed into the page.

Figure 8.3: Conductor moving in a magnetic field


The charge carriers are electrons, of which one is labelled as e , and experience as a
result of their movement in the magnetic flux density an upward force which results in an
accumulation of negative charge at the top of the bar, and a positive charge at the bottom

116

CHAPTER 8. ELECTRODYNAMICS

end. It does not actually matter whether the charge carriers are negative or positive, as
little thought will show that if they were positive, the top of the bar would still acquire a
negative charge and the bottom would acquire a positive charge.
This process is equivalent to placing a source of emf Vi in series with the bar as shown
in the diagram of Figure 8.4.

Figure 8.4: Induced source in series with bar.


That induced emf can cause a current to flow if the bar is now made part of a complete
circuit as shown in Figure 8.5. In that figure the bar, still moving leftward, and still
immersed in a magnetic flux density B directed into the page, makes sliding contact with
a fixed conductor bent into a U shape as shown. The upward movement of electrons in
the bar produces a counter-clockwise current shown as I in the diagram.
It is instructive to see the way in which this situation follows both Faradays and
Lenzs Laws. In the bar the magnitude of the eective electric field is B, and over the
length L of the bar the induced emf is vBL. If we replace v by dx/dt, where x indicates
the position of the bar, we have
dx
dt

(8.1)

d
(BLx)
dt

(8.2)

Vi = BL
=

As both B and L are constant,the product BLx in the numerator is the flux linked
by the circuit, so we have
d
(8.3)
dt
ie the induced voltage is in magnitude the time rate of change of the flux linked by
the circuit, in accord with Faradays law.
Vi =

8.4. A FURTHER USE OF LENZS LAW

117

Figure 8.5: Complete circuit with a moving conductor.


As far as the direction is concerned, we note that when the bar is moving to the left
as shown in Figure 8.5, the flux linked by the circuit is increasing, whereas the current I
flowing in the counter-clockwise direction will produce within the circuit a flux which flows
out of the page, ie in a direction to oppose the change of flux linked which is occurring
as a consequence of the motion of the bar. Thus the direction of the current, which we
deduced from the law of force on the charge carriers, is in accord with Lenzs law.

8.4

A further use of Lenzs law

We illustrate the use of Lenzs Law to find the sense of the current which can flow as a
result of an induced emf in Figure 8.6
In this figure a bar magnet is being moved in a leftward direction so that the flux
emerging from the north pole begins to link a circular coil with a horizontal axis in the
plane of the paper. In the diagram we show that part of the coil which lies below the
plane of the paper, and the places where the coil intersects that plane are shown shaded.
As the bar magnet moves leftward, the flux (sensed in a leftward direction) linked by the
coil increases. The current I induced in the coil must by Lenzs Law be in a direction so
that its contribution to the flux linked by the coil opposes that change. The right hand
rule applied to detect the direction of the magnetic field set up by the current I, sensed
as shown, will tell us that this current will be positive when the north pole is moving
leftward to approach the coil.
You should consider the question of what will happen if the leftward movement of the
magnet is continued so that it passes right through the coil, and proceeds to a considerable
distance leftward of it.

118

CHAPTER 8. ELECTRODYNAMICS

Figure 8.6: Illustration of the use of Lenzs law.

8.5

Inductance

8.5.1

Introduction

Inductance is a concept which can be defined when we have a coil carrying a current i
which sets up a magnetic flux density B so that an amount of magnetic flux associated
with B is linked by either another coil or perhaps by the same coil. When two coils are
involved, the concept defined is that of mutual inductance. When the coil which links
the flux is the same coil as carries the current i the concept defined is that of self
inductance.

8.5.2

Mutual inductance

The concept of mutual inductance is illustrated in Figure 8.7. It is appropriate that we


describe the notation used in this figure in detail, as that notation has bearing upon the
mathematical expression of the inductance concept and of its relation to Faradays and
Lenzs Laws as embodied in the series of equations below.
Firstly the diagram contains two coils, labelled as coil one and coil two, which are
placed in proximity with one another. Coil one and coil two are popular names for coils
these days. It is not necessary that the proximity be of the nested nature shown in the
diagram: the coils merely need be close enough so that one is in the magnetic flux density
caused by currents in the other. It is likewise not necessary that the coils take the form
of solenoids: any shape at all will do. What is necessary is that each of the coils has
established a well defined complete circuit which may however have a pair of terminals
suciently close together that when they are joined each coil forms the boundary of an
unambiguously defined surface over which we may integrate the flux density to produce
a definition for the flux linked by the coil.
In this definition it is necessary to establish for each coil a reference direction which is
shown in the diagram by an arrow on the wire of each coil. That reference direction plays

8.5. INDUCTANCE

119

Figure 8.7: Illustration of mutual inductance concept


a part in the definition of the flux, as we must use the right hand rule to establish, from
the reference direction for traversing the coil, the positive direction for the vector area of
which the coil forms a boundary.
In addition we will need symbols for the currents which flow, or may flow, in the coils.
In the diagram the currents which flow or may flow in coil one and coil two are known as i1
and i2 respectively. Note that to be meaningful, these currents must also have a reference
direction. These reference directions should be the same as were chosen as described in
the paragraph above.
Finally we note that in addition to each coils having the same reference direction
used for the definition of both the current and the flux, we should have a relation between
the separate reference directions for the two coils. Although the first coil may have its
reference direction assigned arbitrarily, the reference direction for the second coil must be
chosen so that a positive current in coil two produces a flux linking coil two of the same
sign as the dierent flux which links coil two as a result of a positive current flowing in
coil one. This relation between reference directions will have an eect upon the signs
which appear in the equations below.
We note also from the figure that coil one has a terminal pair, marked as A, to which
we can connect a voltage generator or some other agency which can cause a current i1 to
flow in coil one. Coil two has a terminal pair, marked as B, which may be bridged by a
resistor, a short circuit, or some other circuit element and so may allow a current i2 to
flow in coil two.
Finally we note that coil two has in series a voltage generator Vi , and that the voltage
generator has a polarity (indicated by the + and - signs) such that if the terminal pair B
is bridged by a resistor, a current i2 of the same sign as the voltage Vi will flow.
The voltage generator Vi is not a real piece of hardware. It is a circuit symbol which
has been introduced to show the eect of the induced voltage which can occur in coil two
in accord with Faradays and Lenzs Laws.
We are now in a position to define the concept of mutual inductance. If the current

120

CHAPTER 8. ELECTRODYNAMICS

i1 in coil one produces a flux 2 which links coil two, with the reference directions all as
defined above, then the mutual inductance M between the two coils is defined as
2
(8.4)
i1
Thus mutual inductance is flux linked per unit current, the flux being linked by coil
two, and the current being in coil one, there being for the moment no current in coil
two. We note that as 2 is proportional to i1 , what we have just defined is a constant
independent of i1 . It depends just upon the geometry of the coils, and fundamental
physical constants. In the SI system of units the units of mutual inductance are henry
(H).
M=

8.5.3

Expression in terms of Faradays law

From Faradays Law we can state that the magnitude of the induced voltage is given by
d2
|
dt
d
= | (M i1 )|
dt

|Vi | = |

(8.5)
(8.6)

Since M is a constant, we may write this in the form


di1
|
(8.7)
dt
To determine a relation between Vi and i1 without the modulus signs, we make use,
as explained in the Section below, of Lenzs law.
|Vi | = M |

8.5.4

The sign from Lenzs law

We now take note that according to Lenzs Law, if we were to allow by placing a resistor
across the terminal pair B a current i2 to flow in coil two in response to the induced voltage
Vi , it must flow in such a direction as to oppose the change in flux 2 which results from
the charge in i1 . Taking note of all the reference directions defined earlier, this means
that when i1 is increasing, i2 must be negative and hence Vi must also be negative. We
conclude that the equation for Vi which embodies both Faradays and Lenzs Laws is
Vi = M

8.5.5

di1
dt

(8.8)

Self inductance

Figure 8.8 illustrates the context of and establishes the variables and reference directions
used to define the concept of self inductance.
In this figure a single coil is carrying a current i which has a reference direction which
is used, in addition to giving meaning to the variable i, to establish, via the right hand
rule, the reference direction for the flux linked by the coil. The coil and its close together
terminal pair, provide a well-defined boundary to the area over which the flux density B,

8.5. INDUCTANCE

121

Figure 8.8: Illustration of self inductance concept


established by the coil, is integrated to define the flux linked by the coil. The voltage
generator Vi has a polarity (defined by the + and - signs) so that if the terminal pair A
is bridged by a resistor R, the current i which is caused to flow is of the same sign as Vi .
The self inductance L for the coil is defined as the flux linkage per unit current, ie

(8.9)
i
We note that because of the relation between the reference direction for the flux and
the reference direction for the current, the self inductance is always positive.
L=

8.5.6

The induced voltage

When the current i is changing, a voltage Vi (proportional to the rate of charge of flux) is
induced in the coil in a direction so as to oppose the change. By reasoning parallel to that
used in the case of mutual inductance, we determine that the correctly signed equation
for the induced voltage is
di
(8.10)
dt
The minus sign in the equation above is of course a consequence of Lenzs Law, and
of the way in which reference directions were defined.
Vi = L

8.5.7

Circuit expressions of inductance

There are two ways in which induced emf in an inductance, as defined in magnitude by
Faradays law and direction by Lenzs law, may be given expression in a schematic circuit.
Both are illustrated in Figure 8.9.
In the left part of the diagram accessible nodes of the physical inductor are labelled as
A and B. The wriggly line drawn between nodes A and C does not represent all aspects
of the inductor, but merely serves as a reminder that an inductor is present, and that
one of its terminals is connected to node A. The induced emf aspect of the inductor is
represented by the voltage generator vi appearing between nodes C and B. We note that

122

CHAPTER 8. ELECTRODYNAMICS

Figure 8.9: Expressions of Faradays and Lenzs laws.


the current i and the voltage generator vi are sensed in the same direction, downward in
this case. The negative sign in the equation
vi = L

di
dt

is an expression of Lenzs law. It should be noted that node C in not a physically


accessible node, but is introduced so that we can show the induced voltage separately
from the part of the symbol which serves as a reminder that an inductor is present.
We note also that in this representation there is supposed to be no potential dierence
between the ends of the wiggly line, i.e. between nodes A and C.
A more compact schematic circuit expression of the same inductor is shown in the
right hand part of Figure 8.9. Here the wiggly line does represent the entire inductor, in
that firstly it is drawn between nodes A and B to which the ends of the physical inductor
are connected, and secondly in that it has between its ends the same voltage vL as would
appear between nodes A and B in the left hand diagram, interpreted as described above.
We should note that in the right hand diagram the current i and the inductor voltage vL
are sensed in opposite directions, ie. the terminal of vL which is marked + is the node at
which current enters the inductor.
Comparing the two diagrams, we see that the appropriate relation between i and vL
which honours both Faradays and Lenzs laws is
di
(8.11)
dt
We note specifically here the absence of a negative sign. Lenzs law has been acommodated by suitable choices of the reference directions of i and vL .
We see slso that the significance of the wiggly line in the two circuits is quite dierent;
in one case there is no voltage between its ends and in the other case the inductor voltage
vL , appropriately sensed, appears between its ends.
In the second circuit the expression given for vL should not be thought of as defining a
voltage generator in parallel with the inductor, but instead as making a statement about
the relation between the current through the inductor and the potential dierence across
it.
In most electrical circuit diagrams the second convention is the one normally chosen
to represent inductors. This choice is made for the reasons firstly that it is more compact
vL = L

8.6. STORED ENERGY IN AN INDUCTANCE

123

than the other, and secondly that because when compared, as in Figure 8.10, with the
representations of the two other basic elements of electric circuits, i.e. the resistor and
the capacitor, there is the similarity that when the currents and voltages are sensed in
the same relative directions in the three diagrams, none of the accompanying equations
contain a negative sign.

Figure 8.10: Current voltage relations in R, L and C.

8.6
8.6.1

Stored Energy in an Inductance


Introduction

Just as we found in our study of electrostatics that placing a charge on a capacitor requires
that work be done, and results in stored energy, we will also find that establishing a current
in an inductor requires that work be done, and that also results in stored energy. It is
the business of this section to investigate these matters. We will first establish a formula
for the stored energy in terms of circuit parameters, such as the inductance and the
current. We will later manipulate those formulae so that they are in terms of of purley
field quantities. A result of this manipulation will be that we will develop a concept
of stored electric energy per unit volume in a magnetic field to accompany our existing
concept of stored magnetic energy per unit volume in an electric field.

8.6.2

Context

In the analysis to follow, we will assume the possible presence of magnetic material, so
our results will not be limited to free space but will cover all contexts, subject to the
qualification below.
Our definition of inductance as flux linkage per unit current necessarily assumes that
the flux linkage in an inductor is proportional to the current flowing therein. If magnetic
media are present, the relation between the internal magnetic field and the magnetisation
must be a linear one. This requires that firstly the magnitude of the internal magnetic
field be low enough for the saturation shown in Figure 7.2 not to occur, and that material

124

CHAPTER 8. ELECTRODYNAMICS

be of a nature that the hysteresis shown in Figure 7.2 be not significant. If the inductor
contains no magnetic material, these restrictions do not apply; the flux-current relation
is a linear one for all values of current.

8.6.3

Establishing an inductor current

A circuit diagram which will enable us to perform the required calculations is shown in
Figure 8.11. In that diagram, a variable voltage source v under our control is connected
to the inductor terminals through which a current i is flowing. From Faradays law for
the inductor, we know that the relation between the voltage source and the current is,
when the reference directions are as shown in Figure 8.11,
v=

d
di
=L
dt
dt

(8.12)

Figure 8.11: Circuit for establishing an inductor current.


It is important to realise that without the voltage source, we cannot establish an
inductor current from a zero initial current condition. If the voltage source is not provided,
there is no path for the current to flow. If the voltage source is provided, but has a zero
value, we can see from equation 8.12 that the derivative of the current remains zero, and
the current remains at its initial value, which has been assumed to be zero. So we see
that the voltage source plays an indispensable part in establishing the current.

8.6.4

Power input

The interacton between the voltage source and the current shows that while the current
is changing, and in consequence the voltage source is non-zero, the voltage source is
supplying a power P and that power is being delivered to the inductor. The power P is
of course given by
P = vi

(8.13)

In the above expresion the voltage and current are related through equation 8.12 so
we may write the power input as

8.6. STORED ENERGY IN AN INDUCTANCE

125

di
(8.14)
dt
In the process of establishing over a time extending from 0 to T , from an initial zero
current a final current I in the inductor, the total work done Wm is obtained by integration
of the above power. Thus
P = vi = Li

Wm =
=

8 T

P dt

(8.15)

vidt

(8.16)

8 T
0

8 T

di
dt
dt
0
8 T
di
= L
i dt
dt
0
=

= L

Li

8 I

idi

(8.17)
(8.18)
(8.19)

1 2
LI
2

(8.20)

Thus our formula for the energy stored in an inductor with current I is
1
Wm = LI 2
(8.21)
2
We may notice the similarity to the formula for the energy stored in a capacitor with
voltage V , viz.
1
We = CV 2
2

8.6.5

(8.22)

Stored energy in field form

Using the relation = Li we first change the stored energy equation


1
Wm = LI 2
2

(8.23)

to the equivalent form


1
Wm = I
(8.24)
2
This is a apropriate change, bringing the formula more into line with the physical
processs which supplied the energy. We recall that the energy was suplied as power,
being the product of a voltage and a current, that product being integrated over time.
In the above formula, we have one factor of current and one factor of magnetic flux, the
derivative of which is the voltage applied to the inductor.
To express this result in the form of the various magnetic field variables, we will have
to assume a particular geometrical form for the inductor. The form which most suits our
purpose is the toroid shown in Figure 8.12. One reason for assuming this form is that

126

CHAPTER 8. ELECTRODYNAMICS

Figure 8.12: A toroidal inductor.


the current I has a simple relation to the internal magnetic field H within the material,
more so than with other forms of inductor.
Another reason for the selection of the toroid for performing this analysis is that both
the magnetic field and magnetic flux are confined entirely to the interior of the winding,
so we are in no doubt about the extent of the space occupied by the field. A third reason
for favoring this structure is that in the interior, the magnitudes of both the magnetic
field and the magnetic flux density are (at least approximately) spatially uniform, so no
complications from spatial non-uniformity arise.
In our analysis we will assume that the radius of the toriod is a and its cross section is
A. The volume occupied by the magnetic field and the flux is therfore 2aA. In pursuit
of a concept of stored energy per unit volume, let us now take the equation 8.24 for the
stored energy in terms of the current and the flux and divide by the volume 2aA to
obtain the stored energy per unit volume Um
Um =

Wm
1 I
=
2aA
2 2aA

(8.25)

On the right hand side we may group the factors as


w Ww

1
Um =
2

I
2a

Ww

(8.26)

The second and third factors are recognisable as the magnitude of the magnetic field
H and the magnitude of the magnetic flux density B, both of which are parallel. We can
therefore write the formula for stored magnetic energy per unit volume as
1
Um = H B
2

8.6.6

(8.27)

Generalisation

We make the inductive generalisation that for all magntetic field distributions involving
linear media, this formula correctly describes the energy stored per unit volume in the
magnetic field. That generalisation then acquires validity through its coninued success in
making experimental predictions.

8.7. MEASURMENTS ON MAGNETIC MATERIALS

8.7
8.7.1

127

Measurments on Magnetic Materials


Method

Faradays law of electomagnetic induction provides a basis for measuring changes in the
magetisation internal to a magnetic material. Figure 8.13 shows how the relation between
a magnetic field H applied to two dierently shaped samples of magnetic material and
the magnetisation which is produced therein may be investigated.

Figure 8.13: Apparatus for investigation of M - H curves.


In the left of the figure the sample of material takes the shape of a circular cylinder,
while on the right the material is formed into a toroid. In each case a magnetic field H
is applied by surrounding the material by a uniform winding of n turns per unit distance
(either length or circumfernce), each carrying a current i. It is easily determined from
Amperes law that the resulting magnetic field produced by that coil (either longitudinally
or circumferentially) is given by
H = ni

(8.28)

We agan note that as n is the number of turns per metre, the formula will give the
correct units for H, i.e. Am1 .
For the determination of flux changes in the material, a second winding of N turns
uniformly covering the sample is applied, and the induced voltage which occurs in response
to changes in the flux linked by the second winding is observed. If the sample has cross
sectional area A and an internal flux density B, which we may assume is perpendicular to
and uniform over the cross section, then the flux linked by each turn is BA, and the total
flux linked by N turns is N BA. The induced voltage may be observed on an oscilloscope,
which should have a high enough input impedance for only a negligible current to flow in
the second winding.
In operation the current i producing the driving field H is made time varying, perhaps
oscillating with a triangular wave form as shown in Figure 8.14. We can determine the
wavform of the applied magnetic field H by making measurements of the current i.
The voltage indicated on the oscilloscope will be, in accord with Faradays law,

128

CHAPTER 8. ELECTRODYNAMICS

Figure 8.14: Triangular current waveform.

v(t) =

d
dt

(8.29)

.
This voltage is not directly indicative of the flux within the sample, but of its time
derivative. With modern computer techniques the waveform may be recorded and integrated with respect to time to produce the wavform of flux and then the flux density
B itself. By dividing B by 0 and subtracting H, we may determine for each value of H
the magnetisation M.

8.7.2

Consequences for electromagnetic theory

When the sample is in the form of a toroid, there is no doubt that the internal field within
the material is the applied field H produced by the current i, as the field, and the resulting magnetisation are all in a circular direction inside the material, and are as a result
parallel to the material surface. There is thus no point at which the magnetisation meets
the surface at right angles, and so no magnetic poles are produced. The relation between
the applied field H and the magnetisation M seen in the experiment is the true relation between the internal magnetic field and the magnetisation, and for a ferromagnetic
material takes the form seen aleady in Figure 7.2.
When however we examine the results for the cylindirical sample, we find that to
achieve a particular value of magetisation, we must apply through the current i a larger
value of applied field H than was the case with the toroidal sample. The explanation
is that when the cylinder becomes magnetised, poles appear on the end surfaces, and
produce their own contribution to the magnetic field in the cylinder, which contribution
opposes the applied field, so that to achieve a particular internal field within the cylinder,
a greater applied field is required.
This experiment provides the evidence which was promised in Sections 7.1 and 7.9.3
for the existence of the demagnetising eect of the magnetic poles which can occur on the
surfaces of a magnetised material.

8.8. DEPOLARISING AND DEMAGNETISING FACTORS

129

Figure 8.15: Displacement current in a capacitor.

8.8

Depolarising and Demagnetising Factors

There will hopefully be a short discussion during lectures of the concept and utility of
depolarising and demagnetising factors.

8.9
8.9.1

Maxwells Contribution
Introduction

It is now time to describe the final element of the progressive uncovering of the complete
laws of electromagnetic theory which occurred with the discovery by Maxwell in 1864
of the role of what is known as displacement current in contributing to the vortices of
the magnetic field H. Maxwells contribution lay firstly in recognising that the laws of
electrodynamics, as they were thought to exist at that time, were mutually incosistent,
and required modification to render then self-consistent, secondly in discovering the appropriate modification, and thirdly in making experimentally verifiable predictions of the
consequences of the correction he had introduced.

8.9.2

Displacement Current Concept

The inconsistency is made evident by examining Amperes law of magnetostatics as shown


in equation 8.30 below
-

H dr = I =

J ds

(8.30)

in which the current I is the total current enclosed by a contour C around which the
integral of the magnetic field H takes place, in the light of Figure 8.15.
In this figure we see an ac circuit, in which the current i in part of the circuit is carried
along wires while in another part of the circuit the same current flows through the series
capacitor shown.

130

CHAPTER 8. ELECTRODYNAMICS

The determination of the issue of whether the contour encloses the current is decided
by constucting any surface bounded by the contour, and seeing whether the current travels
through that surface.
As is shown in Figure 8.15, it is possible for two dierent surfaces, one cutting the
wire and the other passing between the capacitor plates, to be drawn such that these two
surfaces have the same boundary in the form of the contour C as shown.
For each of these surfaces, the surface integral on the right hand side of equation 8.30
is supposed to be equal to the line integral on the left hand side of the same equation along
the common contour C. For the integral over the surface cut by the wire, the value of the
integral is the current. For the integral over the surface passing between the capacitor
plates there is no current, and hence the surface integral has a zero value.
What physically has happened is that the current has led to charges being distributed
over the capacitor plates, and these charges have led to an electric flux density between
the plates. When the fields are time varying, this time-varying flux density eectively
continues the conduction current from one plate to the other in the form of a displacement
current.
d
dt

D ds

(8.31)

.
As it is a consequence of Gauss law for the electric flux density that the normal
component of the surface charge density is equal to the normal component of the electric
flux density D, the displacement current just defined is equal to the conduction current
I in the wires leading to and from the capacitor.

8.9.3

Resolution

Maxwell realised that the diculty that the right hand side of equation 8.30 depends on
where surface is placed could be removed if the term
d 8
D ds
dt S

(8.32)

involving the displacement current could be added to the right hand side of that
equation to give the result
-

d 8
H dr = I +
D ds
dt S
C

(8.33)

as a new version of Amperes law suitable for time varying fields. It is hopefully
clear that when the fields are static, the added term is zero, so the amended equation is
consistent with the already establised laws for magnetostatic fields.
The result of this action is the completion of the set of electrodynamic equations which
are set out in Section 8.10 below. These equations have remained unchanged since their
formulation by Maxwell in 1864.

8.10. THE COMPLETE LAWS

8.10

131

The Complete Laws

We conclude this course by summarising the complete laws of electrodynamics which have
been described over the present and preceding chapters. The results can be compared
with those which were quoted in Section 2.13.

8.10.1

Faradays law

The circulation of the electric field vector E around a closed contour is equal to minus
the time rate of change of magnetic flux through a surface bounded by that contour, the
positive direction of the surface being related to the positive direction of the contour by
the right hand rule. In a mathematical formula this law takes the form
-

8.10.2

E dr =

d
dt

B ds

(8.34)

Amperes law as modified by Maxwell

The circulation of the magnetic field vector H around a closed contour is equal to the
sum of the conduction current and the displacement current passing through a surface
bounded by that contour, with again the right hand rule relating the senses of the contour
and the surface. In a mathematical formula this law takes the form
-

d 8
H dr = J ds +
D ds
dt S
C
S

8.10.3

(8.35)

Gauss law for the electric flux

The total electric flux (defined in terms of the D vector) emerging from a closed surface
is equal to the total conduction charge contained within the volume bounded by that
surface. In a mathematical formula this law takes the form
-

D ds =

dv

(8.36)

In the above equation we are expected to remember that the charge density appppearing on the right hand side contains only the conduction charge density c .

8.10.4

Gauss Law for the magnetic flux

The total magnetic flux (defined in terms of the B vector) emerging from any closed
surface is zero. In a mathematical formula this law takes the form
-

B ds = 0

(8.37)

132

CHAPTER 8. ELECTRODYNAMICS

Appendix A
STANDARD INTERNATIONAL
TERMINOLOGY AND UNITS
A.1

Objective

It is the objective of this Appendix to make available to students up to date information


on those parts of the Standard International system of units which should be used for
study in the area of electodynamics.

A.2

Informal discussion

Students are hopefully familiar with the usage of Standard International units in the
subjects of mechanics and dynamics, although perhps not with recent changes in the way
the fundamental quantities of mass, length and time are defined. Although it might appear
that those fundamental quantities should be definable in a way which is independent of
electrodynamic theory, the modern approach to those definitions requires for its proper
undersanding a knowlege of the fundamenal principles of electodynamics, and in particular
of the properties of electromagnetic waves and the relation of their velocity to fundamenal
physical constants.
During the course we have encountered the fundamental physical constants 0 which
appeared in the law of force between electric charges and 0 which appeared in the law
of force between current carrying wires. What we did not have the opportunity to do,
because of the limited scope of the course, was to establish that the values of these
constants are not indepedent, but form part of a basic relation which links themselves and
the value of the velocity of light.
In the further development of the subject of electrodynamics it may be shown that
the magnetic permeability of free space, the dielectric permittivity of free space and the
velocity of light in free space are related by the equation
c2 0

=1

(A.1)

In the past, 0 was always given the defined value 4 107 H/m in the process
of setting up the Standard International system of units. On the assumption that the
units of mass, length and time were already established, and hence the unit of force
was established, the action of giving 0 this defined value in eect established the unit
133

134 APPENDIX A. STANDARD INTERNATIONAL TERMINOLOGY AND UNITS


of current. The velocy of light was then an experimentally determined value, and the
dielectric permittiviy of free space became a calulated value based in part on a defined
value and in part on an experimentally determined value, the calculation being based
upon equation A.1 above.
In the modern approach, the unit of time (the second) is defined in terms of the
frequency of oscillation of an atomic clock based on a particular form of radiation. Instead,
however, of defining, as was done in the past, the unit of distance (the metre) as the
distance between two marks on the internationl prototype metre maintained in a controlled
environment, the metre has been re-defined in terms of the distance traveled by light in a
particular time. Such a defintion requires that the velocity of light be known. It becomes
known by giving it the defined value of 299,979,245.8 m/s. What has happened is that
the length of the meter has been adjusted, (hopefully not by much), so that the velocity
of light now has exactly this value.
Among the electrodynamic constants, the value of 0 is still given the defined value
4 107 H/m. Equation A.1 is regarded as being firmly established by fundamental
physical law as providing a relation between 0 , 0 and the velocity of light c. As both c
and 0 have now received defined values, so now also has 0 . Its value is exactly
0

107

1
(299, 979, 245.8)2

(A.2)

Its value is quoted to a convenient number of decimal places later in this Appendix.
In the modern system, the unit of mass (the kg) is still defined as the mass of the
international prototype kilogram maintained in a controlled environment, and from which
other standard kilogram masses can be copied and distributed.
As the units of mass, length and time are now established, then the unit of force is
establised through the hopefully well known equations of dynamics.
The unit of current is then established through the equation 6.18 drawn from Chapter 6
and reproducd for convenience below.
Force per unit
0 I1 I2
=
length of one wire
2s

(A.3)

and hence the unit of charge (the coulomb) is established as the charge which a current
of one ampere would deliver in one second.
Although for economy of expression and to celebrate the work of eminent scientists,
particular names have been given to the remainder of the electrodynamc quantities discussed in this course, they can all be reduced to alternative terms involving the fundamental units of mass, length, time and current which are defined, among others, as belonging
in the set of base units listed in Table A.1 below, and hence the meaning of all such
quantites is clearly established.

A.3

The Base Units

We provide in Table A.1 the formal definitions of the base units of the SI system.

A.3. THE BASE UNITS

135

BASE UNITS
SYMBOL DEFINITION OF UNIT
The metre is the length of the path travelled
by light in vacuum during a time interval of
1/299 792 458 of a second. (What this does
length
metre
m
is give the velocity of light a defined value of
299 979 245.8 m/s.)
The kilogram is the unit of mass; it is equal
mass
kilogram kg
to the mass of the international protype of
the kilogram.
The second is the duration of 9 192 631 770
periods of the radiation corresponding to the
second
s
time
transition between the two hyperfine levels of
the ground state of the caesium-133 atom.
The ampere is that constant current which,
if maintained in two straight parallel conductors of infinite length, of negligible circular
ampere
A
cross-section, and placed one metre apart in
electric current
a vacuum, would produce between these conductors a force equal to 2 107 newton per
metre of length.
The kelvin unit of thermodynamic temperathermodynamic
ture, is the fraction 1/273.16 of the thermokelvin
K
temperature
dynamic temperature of the triple point of
water.
The mole is the amount of substance which
contains as many elementary entities as there
are atoms in 0.012 kilograms of carbon-12.
amount of submole
mol
When the mole is used, the elementary enstance
tities must be specified and may be atoms,
molecules, ions, electrons, other particles, or
specified groups of such particles.
The candela is the luminous intensity, in a
given direction, of a source that emits monoluminous intenchromatic radiation of frequency 540 1012
candela
cd
sity
hertz and that has a radiant intensity in that
direction of 1/683 watt per steradian.
QUANTITY

NAME

Table A.1: The base units of the SI system.

136 APPENDIX A. STANDARD INTERNATIONAL TERMINOLOGY AND UNITS

A.4

Abbreviations

We provide in Table A.2 a summary of the names and standard abbreviations for commonly used electrodynamic variables.
Full name
ampere
coulomb
farad
henry
joule
metre
newton
ohm
siemen
tesla
volt
watt
weber

Abbreviation
A
C
F
H
J
m
N

S
T
V
W
Wb

Table A.2: Full and abbreviated forms of SI units used for electrodynamic variables.

A.5

The Fundamental Vectors

1. The names, usual symbols, and standard international units for the four vectors
used to describe electromagnetic fileds in a general medium are
E
electric field intensity
Vm1
H magnetic field intensity Am1
D
electric flux density
Cm2
B
magnetic flux density Wbm2
The units of magnetic flux density B have the alternative name of Tesla, for which
the abbreviation is T.
2. The names, usual symbols, and standard international units for the vectors used
to describe the state of a dielectric medium and the state of a magnetic medium
respectively are
P
polarisation Cm2
M magnetisation Am1
3. The fully general relations between the above six vectors are
(i)
(ii)

D = 0E + P
B = 0 (H + M)

by definition
by definition

A.6. LIST OF STANDARD SYMBOLS

137

4. The values of the magnetic permeability and the dielectric permittivity of free space
are in standard international units
(i) 0
(ii) 0

4 107
8.854 1012

Hm1
Fm1

by definition
approximately

In the most recent definition of the standard international system of units, both of
the above values have become, in eect, as explained above, defined values.

A.6

List of Standard Symbols

We provide in Table A.3 a summary of the names and SI units used for the variables
employed in these notes.

138 APPENDIX A. STANDARD INTERNATIONAL TERMINOLOGY AND UNITS

VARIABLE
Area
Capacitance
Charge
Conductance
Conductivity
Current
Dielectric permittivity
Dielectric susceptibility
Distance
Electric dipole moment
Electric field intensity
Electric flux
Electric flux density
Energy
Force
Inductance
Linear charge density
Magnetic field intensity
Magnetic flux
Magnetic flux density
Magnetic moment
Magnetic permeability
Magnetic susceptibility
Magnetisation
Mobility
Polarisation
Potential
Power
Resistance
Surface charge density
Surface current density
Torque
Velocity
Volume
Volume current density
Volume charge density
Work

SYMBOL
A
C
Q
G

I
e
l, r
p
E

D
W
F
L
ql
H

B
m

m
M

P
V
P
R
qs
J s or K
T
v
v
J v
qv or v
W

SI UNITS
m2
F
C
S
S/m
A
F/m
None
m
Cm
V/m
C
Cm2
J
N
H
Cm1
A/m
Wb
Wbm2 or T
Am2
H/m
None
A/m
m2 Vs
Cm2
V
W

Cm2
Am1
Nm
m/s
m3
Am2
Cm3
J

Table A.3: Names symbols and units for electrodynamic variables.

Appendix B
USEFUL PHYSICAL CONSTANTS
B.1

Physical Constants
QUANTITY
Electronic charge
Electron mass
Permittivity of free space
Permeability of free space
Velocity of light

SYMBOL VALUE
e
1.60217733 1019
m
9.1093897 1031
8.854187817 1012
0
0
4 1012
c
2.99792458 108

Table B.1: Physical constants.

B.2

Material Conductivities
MATERIAL
Silver
Copper
Gold
Aluminium
Iron
Nichrome
Graphite
Sea water
Distilled water
Quartz

CONDUCTIVITY
6.17 107 S/m
5.80 107 S/m
4.10 107 S/m
3.82 107 S/m
1.03 107 S/m
0.10 107 S/m
7 104 S/m
5 S/m
1 104 S/m
1 1017 S/m

Table B.2: Material conductivities.

139

UNIT
C
kg
F/m
H/m
m/s

140

B.3

APPENDIX B. USEFUL PHYSICAL CONSTANTS

Relative Dielectric Permittivities


MATERIAL
Air
Polyethylene
Fused quartz
Bakelite
Glass
Aluminium oxide
Silicon
Distilled water
Titanium dioxide
Barium titanate

DIELECTRIC CONSTANT
1.0005
2.26
3.8
4.74
4 to 7
8.8
11.8
80
100
1200

Table B.3: Relative dielectric permittivities.

B.4

Relative Magnetic Permeabilities

It should be noted that for a ferromagnetic medium the concept of a relative permeability
implies a linear relation between M and H. This assumption is only an approximation,
and applies with varying degrees of correctness to various materials, and is only valid well
below magnetic saturation.
MATERIAL
Nickel
Cobalt
Powdered iron
Ferrite (higly variable)
Pure iron
Mumetal

RELATIVE MAGNETIC PERMEABILITY


50
60
100
1,000
4,000
20,000

Table B.4: Relative magnetic permeabilities.

Appendix C
REFERENCES
1. SI units and recommendations for the use of their multiples and of certain other
units, International Standard ISO 1000 (1992), International Organisation for Standardisation, Case postale 56, CH-1211, Genevre 20, Switzerland.

141

142

APPENDIX C. REFERENCES

Appendix D
ADVICE ON STUDY FOR
EXAMINATIONS
D.1

Use of these lecture notes

You are advised to print your own copy of these lecture notes, and to study them carefully.

D.2

Advice on units

1. Students will be expected to know the Standard International units for all of the
electrodynamic variables mentioned in the lecture notes, and to correctly use and
state those units in providing an answer to any examination question.
2. The use of equivalent but unexpected combination of units in place of the accepted
standards is not welcome. A good example is the use of N/C in place of V/m for
the units of the electric field.

D.3

Other Advice

1. Where an explanation of a phenomenon or principle is sought, it is hoped that the


answer will be presented in well formed English sentences rather than in the form
of ungrammatical notes.
2. Students should pay particular attention to correctly understanding the matter below in which our previous experience has shown that their prior teaching in electromagnetic theory often leads to erroneous views. This is
The relative directions of E, P, and D in a linear dielectric.
3. The lecturer has tried to present the concepts of sources and vortices as indispensible tools needed in gaining a good understanding of electromagnetic theory. It is
suggested that students obtain a good mastery of these concepts in the following
respects.
The appearance of simple source type and simple vortex type fields.
143

144

APPENDIX D. ADVICE ON STUDY FOR EXAMINATIONS


Which of the fundamental electromagnetic field vectors have sources and which
of them have vortices.
How the four fundamental laws can be expressed as statements about sources
or vortices of the fundamental field vectors.

Appendix E
EXERCISE SET 1
1. (a) State the names, usual symbols and standard international units, as defined
in ISO 1000 (1992), for four vector fields used to describe electromagnetic
phenomena in a general medium.
(b) State the names, usual symbols and standard international units, as defined in
ISO 1000 (1992), for vectors describing the states of both a dielectric medium
and a magnetic medium.
(c) Quote the fully general relations between the vectors:
i. D, E and P; and
ii. B, H and M

(d) In the SI system of units, what are the values of:


i. the magnetic permeability 0 of free space?
ii. the electric permittivity 0 of free space?
2. When a glass rod is charged by rubbing it with silk, where does the charge come
from?
3. When defining the electric field strength, why is it necessary to specify that the
magnitude of the test charge is very small?
[Begin by defining the electric field strength at a point.]
4. Two similar conducting balls each of mass m are hung from a common point on silk
threads of length l and carry equal charges q. Show that when the balls are at rest,
the distance x between them is
x=

q2l
2 0 mg

~1/3

(Assume that the angle of each thread to the vertical is so small that cos can be
taken to be equal to one).
5. Figure ?? shows the electric field lines for two charged spheres separated by a small
distance.
(a) Estimate the ratio q1 /q2 .
145

146

APPENDIX E. EXERCISE SET 1

q2

q1

Figure E.1: Representation of field lines.

q2

q1

Figure E.2: Representation by array of vectors.


(b) What are the signs of q1 and q2 ?
6. The electric field can be represented either by electric field lines as in Figure E.1,
or by an array of vectors, each of which has a length and direction indicating the
strength and direction of the electric field at that point, as in Figure E.2.
(a) Comment on the advantages and disadvantages of each representation.
(b) State the principle of superpostion, explain how it is used in determining the
electric field strength, and use it to draw vectors representing the electric field
strength at the points P and Q in Figure ??.
(c) Explain why electric field lines never cross. [Hint: Consider the electric field if
the field lines were to cross.]
7. State whether each of the following statements about electric field lines is true or
false, and explain your reasons.
(a) Where there is no electric field line, the electric field stength is zero.
(b) Along a given field line, the force on a test charge is constant.

147
(c) The force on a test charge at a point is tangential to the field line through that
point.
(d) A charged particle in an electric field always travels travels along a field line.
8. (a) Explain how in free space the electric flux through an area is related to the
electric field strength.
(b) If you had to represent the direction of an area using a single vector, which
direction would you choose?
(c) A flat surface having an area of 3.2 m2 is rotated in a uniform electric field of
intensity E = 6.2 105 Vm1 . Calculate the electric flux through this area
when the electric field:
i. is perpendicular to the surface;
ii. is parallel to the surface;
iii. makes an angle of 75 with the plane of the surface.
9. (a) Explain what is meant by a gaussian surface, and why it is useful.

S4
S5
+2Q
+Q
S1
-3Q
S2

S3

Figure E.3: Five Gaussian surfaces.


(b) Figure E.3 indicates five closed gaussian surfaces that surround various charges
as indicated. Determine the total electric flux through each surface.
10. If the net electric flux through a gaussian surface in free space is zero, which of the
following statements are true?
(a) There are no charges inside the surface.
(b) The net charge inside the surface is zero.
(c) The electric field is zero everywhere on the surface.
(d) The number of electric field lines entering the surface equals the number leaving
the surface.

148

APPENDIX E. EXERCISE SET 1

11. Challenge question [Think of an easy way to do it]:


A pyramid with a 6 m square base and height of 4 m is placed in a vertical electric
field of 52 Vm1 . Calculate the total electric flux through the pyramids four
slanted surfaces. No conducting or dielectric material is present in this problem.

Appendix F
EXERCISE SET 2
1. (a) State Gauss law for electrostatics in free space.
(b) The diagram of Figure F.1shows four closed surfaces that surround as indicated
various charges in free space.

S4
S2

-2Q

+Q
S3

+Q

S1

Figure F.1: Four closed surfaces enclosing charge.


i. Determine the electric flux through each surface.
ii. Can Gauss law be used to find E at any point on any of these surfaces?
Explain.
(c) On a clear day, the electric field near the earths surface is 100 Vm1 pointing
radially inward. If the same electric field existed everywhere on the earths
surface, what would the total net charge stored in the earth be?
2. A long line of positive charge of linear density Cm1 is surrounded in a vacuum
by a concentric conducting cylinder of inner radius ao and thickness t, and which
carries no net charge.
(a) Without carrying out any calculations, explain how you would find the electric
field at a point P distance r < ao from the axis, if you could not use Gauss
law.
(b) Use Gauss law to find:
i. the electric field at a distance r < ao from the axis;
ii. the electric field at a distance r > (ao + t) from the axis;
149

150

APPENDIX F. EXERCISE SET 2


iii. the charge per unit length on the inner and outer walls of the cylinder.

3. (a) Describe the Gaussian surface used in lectures to derive the electric field near
a charged flat plate in a vacuum. Describe two other Gaussian surfaces which
could be used instead. Be careful in specifying their position and orientation.
(b) In response to the question: Use Gauss law to find the electric field just outside
a conducting surface carrying in a vacuum a charge per unit area Cm2 , the
following student answer was given.
=
=
o EA =
E =

q
A
A
/ o

Comment on whether this answer is satisfactory, and make any improvements


you think are necessary.
4. (a) What is the distinction between electric potential and electric potential energy?
(b) An -particle (Z = 2) travels directly toward a gold nucleus (Z = 79). When
it is far from the nucleus, the -particle has kinetic energy K. The particle
approaches the nucleus (assumed spherical), and is repelled by the electrostatic
force.
i. Calculate K assuming that the centres of the -particle and gold are 5.0
1015 m apart at the distance of closest approach, and that they are not
touching at this point. Express your answer in Joules and in MeV.
ii. The actual -particle energy Rutherford and his collaborators used in their
experiment was only 5.0 MeV. Compare their experiment with that described above.
5. (a) Define the electric potential at a point in an electric field.
(b) What is the value of the potential at the centre of a spherical conducting shell
of radius 100 mm if it carries a charge of 2 107 C?

(c) Two point positive charges each of value q are separated by a distance 2a.
Determine the electric field E and potential V , at a point midway between the
charges.

6. In Figure F.2 points A and B are located at distances a1 and a2 respectively from a
long conductor of length l carrying positive charge of linear density Cm1 . Find
the potential dierence VBA = VB VA between the two points. Use the value of
electric field derived in lectures.
7. Using carefully chosen Gaussian surfaces, find the electric field in air at the point
P, shown in Figures F.3,F.4, and F.5, in each of the following situations:
(a) Charged slab of insulator, with area charge density .
(b) Charged conducting slab, with surface charge density on each surface.

151

Figure F.2: Long uniformly charged conductor.

P
+++++++++++++++

Figure F.3: Charged slab of insulator.

P
+++++++++++++++
+++++++++++++++

Figure F.4: Charged conducting slab.

-----------------P
++++++++++++++++

Figure F.5: Pair of conducting plates.


(c) Pair of conducting plates, with surface charge densities + and .
8. Challenge question:
The diagram of Figure F.6 shows a hollow conductor carrying positive charge on
its outside surface, and equal amounts of positive and negative charge on its inner
surface.
(a) Is this arrangement of charge consistent with Gauss law?
(b) Does this arrangement violate any other laws of physics, ie could it happen in
practice?
[Hint: Think in terms of conservation of energy, and work done by an electric
field.]

152

APPENDIX F. EXERCISE SET 2

+
+
+

+
+

---

++
+
++

+
+

+
+

Figure F.6: Charged hollow conductor.

Appendix G
EXERCISE SET 3
1. (a) Consider the following expressions for capacitance, C:
C = Q/V

and C = A/d.

Define the symbols in each expression, and explain what each expression refers
to.
(b) Find the capacitance of a parallel plate capacitor with plates of area 1.2 103
m2 , separated by a distance of 2.0 mm (in vacuum).
(c) Deduce the charge on each plate of the capacitor, and the electric field in the
region between the plates, when the potential dierence between the plates is
50 V.
2. An isolated conducting sphere of radius a carries a positive charge Q.
(a) Where can you consider the equal and opposite charge Q to be?

(b) Beginning with an expression for the electric field due to the sphere, find the
potential of the sphere, and its capacitance.
3. Consider whether each of the following statements is true or false, and explain your
reasoning.
(a) The potential across a resistor in a circuit means the same as the potential
dierence between the ends of the resistor.
(b) The capacitance of a parallel plate capacitor is proportional to the charge on
its plates.
(c) There can be no electric field within a conductor.
(d) When there is a potential dierence between the ends of a uniform conductor,
the electrons experience a uniform electric field, so they experience constant
acceleration.
(e) If the electric potential is zero at some point in space the electric field must
also be zero at that point.
(f) If you are caught in the open in an electrical storm, the safest thing to do is
to lie flat on the ground.
153

154

APPENDIX G. EXERCISE SET 3

4. Often a complicated circuit can be redrawn to look much simpler and to make the
calculation of its eective or equivalent value much easier. Using the principles
shown in the example of Figure ?? redraw the circuit shown in the exercise of Figure G.2 so that it appears as some resistors connected between two input terminals
a and b (as done in the example), and then if possible, calculate the single equivalent
resistance Req . Take each resistance to be 2 .
R2

R1

a
R3

R4
a

R1

R4

R2
R3

b
a

Req

Figure G.1: Simplification of a resistor network.


R5

R1

a
R6

R4

R2

R3

Figure G.2: Resistor network for exercise with an example above.


5. (a) A 6.0 F capacitor is connected in series with a 4.0 F capacitor and a potential
dierence of 200 V is applied across the pair.
i. What is the charge on each capacitor?
ii. What is the potential dierence across each capacitor?
(b) A 100 pF capacitor is charged to a potential dierence of 50 V and the charging
battery is then disconnected. The charged capacitor is connected in parallel
with a second (initially uncharged) capacitor. If the potential dierence across
each is now 35 V, what is the capacitance of this second capacitor?

155
6. Two wires of Al and Cu having equal lengths of 40 m and cross-sectional area 1
mm2 (standard lighting cable) are connected in series. A potential dierence of 17.8
V is applied between the ends of the connected wires. The resistance of the Al wire
is 1.10 and the resistance of the Cu wire is 0.68 . Determine:
(a) the resistivities of Al and Cu;
(b) the potential dierence across each wire;
(c) the electric field in each wire.
(d) What happens to your answers if both wires are carefully stretched so that they
double in length? (Assume that the cross-sectional area reduces but remains
uniform along the length of the wire to keep the volume of material constant.)
7. (a) Determine the current through each resistor in the circuit shown in Figure G.3.
(b) Taking the zero of potential at the negative terminal of the 3 V battery, find
the potential at each of the labelled points.

4W

2W

+
3V _

5W

C
+
_ 1V

Figure G.3: Network with two voltage sources.


(c) Discuss what happens if a cathode ray oscilloscope (CRO) is used to measure
the voltage across the 2.0 resistor, connecting the earthed terminal to the
point C.
8. Challenge question:
Two resistors, R1 and R2 , are connected in series across a battery of emf E and of
negligible internal resistance.
If E = 25 V and R1 = 50 find the values of R2 such that the power dissipated in
R2 is 2 W.
9. Challenge question:
(a) The diagram of Figure G.4 shows an electric dipole consisting of charges +q
and q located at positions +a and a respectively on the z axis. (What is
the direction of the x axis?) Show that the potential at a point o-axis at a
great distance r from the origin is given approximately by:

156

APPENDIX G. EXERCISE SET 3

V =

2qa cos
p cos
pz
=
=
2
2
4 0 r
4 0 r
4 0 r3

1
1
[Hint: Show that r+
r
r/r2 where r = r r+ 2a cos .]

r+

+q

r
r-

2a

-q

y
Dr

Figure G.4: An electric dipole.


(b) Find the x,y and z components of the electric field at a point o-axis but in
the yz plane.
[Hint: it is easiest to use partial dierentiation of the potential.]

Appendix H
EXERCISE SET 4
1. (a) Use Kirchhos loop rule to obtain a dierential equation for the circuit shown
in Figure H.1 in terms of the charge q on the capacitor at time t after the
switch is closed.

Figure H.1: An RC circuit.


(b) Solve the equation for q, and compare your solution with the result given in
lectures.
[Hint: Rearrange the dierential equation so that you can integrate over q on
one side of the equation, and over t on the other side.]
What values of q does your solution give for t = 0 and t ?
(c) What is meant by the time constant of the circuit?

(d) Find the initial charging rate of the capacitor.


2. A 10,000 resistor and a capacitor are connected in series and a 10 V potential
dierence is suddenly applied across the combination.
(a) Draw a circuit diagram describing this situation and indicate the direction
of flow of conventional current through the circuit. Write down an equation
describing the potential dierence across the resistor as a function of time and
illustrate this equation with a sketch graph; indicate the potential dierence
when t is equal to the time constant.
At a time t1 = 1.0 1.06 s (1.0 s), the potential across the capacitor is 5.0
V.
(b) What is the current through the resistor at time t1 ?
(c) Find the time constant of the circuit and hence find the capacitance of the
capacitor C shown in Figure H.2.
157

158

APPENDIX H. EXERCISE SET 4

3. (a) Redraw the circuit, of Figure H.2, showing an ammeter and voltmeter connected to measure the current through the 6 resistor and the potential difference across the 2 resistor.

Figure H.2: Three resistor circuit.


What are the readings on the two meters? (Assume the meters are ideal.)
(b) How would you accurately determine the resistance of the 6 resistor by
measuring both the potential across and the current through it?
4. (a) For the potential divider circuit shown in Figure H.3, derive an expression for
V2 in terms of V1 , R1 and R2 .

Figure H.3: Potential divider circuit.


(b) A voltmeter with resistance Rv is connected between terminals a and b ofthat
circuit. Derive an expression for V2 in this new situation.
(c) To assess the eect of Rv on the circuit, calculate V2 in terms of V1 and R2 for
the cases where:
i. R1 = R2 and Rv = 1000R1 ;
ii. R1 = R2 = Rv .
Summarise your findings.
(d) Draw a potentiometer circuit, and show how it should be connected to measure
the potential V2 . Does it have any advantages over a voltmeter?

159

5. Calculate the value of Hdr (sometimes called the circulation) around the paths
indicated in Figure H.4. In parts (b) and (c) of the figure the current I is directed
out of the page.
I
[Use the relation H = 2r
for the magnitude of the magnetic field H a distance r
from a long straight wire carrying current I.]

What has this question to do with Amperes law?

Figure H.4: Integration contours.


6. Suppose that the electrons in the beam of a television tube have an energy of 12,000
eV. The tube is oriented so that the electrons move horizontally from south to north.
The vertical component of the earths magnetic field H points up and has the value
of 44 Am1 .
(a) In what direction will the beam be deflected?
(b) What is the velocity of a given electron?
(c) How far will the beam be deflected in moving 200 mm through the television
tube?
[You may assume the deflecting force is constant in direction.]
(The ratio e/m for an electron is 1.76 1011 C/kg)
7. In Figure H.5 a uniform electric field of strength E directed down the page is provided by a pair of charged metal plates as shown in the diagram. A magnetic field
H (indicated by crosses) is applied in vacuo into the page. Positively charged particles of velocity v are injected midway between the plates. Assume the force on the
particles due to gravity is small compared to those due to the electric and magnetic
fields.
(a) For a charged particle that has just entered the B fields midway between the
plates in the direction indicated, write down the force (and direction) due to
the magnetic flux density B; and due to the electric field E. Under what
conditions will this particle emerge through the slit?
(b) Particles emerging from the slit are found to have velocity vs . What will happen
to particles that enter the system with a velocity v < vs ? What about v > vs ?

160

APPENDIX H. EXERCISE SET 4

metal plates

+q

XB

slit

Figure H.5: Crossed field configuration.


(c) If particles have velocity vs but a range of masses when they enter the system,
will they all emerge through the slit?
8. Suppose that a galvanometer with a sensitivity of 50 A for full scale deflection
(F.S.D.) and internal resistance of 1000 is available to you.
(a) Draw circuits that would allow this meter to be used as:
i. a voltmeter;
ii. and an ammeter.
(b) Calculate the resistance needed if the potential dierence between the voltmeter
terminals is 10 V when the current through it is 50 A.
(c) Design a circuit that allows the meter to measure currents up to 1A F.S.D.
9. Challenge Question: In relation to Figure H.6:

Figure H.6: Bridge circuit.


(a) What is the voltage across the capacitor after the battery has been connected
for a long time?
(b) If the battery is suddenly disconnected, how long does it take for the capacitor
to discharge to 1/e of its initial voltage?

Appendix I
EXERCISE SET 5
1. Consider an air-filled toroidal coil such as the one shown in Figure I.1. It is wound
on a doughnut shaped core which has an inner radius of 40 mm, an outer radius of
60 mm, has 400 turns, and carrier a current of 0.5 A.

Figure I.1: Air filled toroidal coil.


(a) Describe the magnetic field H inside the core.

(b) Use Amperes Law to calculate the value of H at a point midway between the
inner and outer walls of the core.
(c) What is the value of H outside the toroidal coil? Compare this with the value
of H outside a long solenoid.
(d) What can be said about the values of B inside and outside the core?
2. Recently, the space shuttle attempted an experiment shown in Figure I.2 that involved the deployment of a 20 km wire in space with the idea of using it as a power
supply.
(a) For the situation above where the wire is perpendicular to the magnetic field
lines throughout its orbit, calculate the emf produced if the wire is 20 km long,
moves with a velocity of 8 km/s and the magnetic flux density is is 0.33 104
T. (These are realistic parameters for this case.)
(b) What would happen to the potential dierence between the wires ends if the
orbit were reversed?
(c) Discuss the practical aspects of using this extremely long power supply to drive
a circuit.
161

162

APPENDIX I. EXERCISE SET 5

wire

Figure I.2: Wire antenna in space.


3. A conducting loop of fixed size and shown in Figure I.3 is in position P within the
confines of a uniform magnetic field in air and of finite extent. Specify the direction
of the current induced in the loop (if any) when it is moved from P to:
B Uniform
R

Figure I.3: conducting loop in magnetic field.


(a) position Q;
(b) position R;
(c) position S.
(d) While the loop is moving through position R, indicate the direction of any
forces on the loop.
4. (a) A flexible wire of length l is joined at the ends to form a single-turn circular
loop, and placed on a horizontal wooden table. Find the area of the loop in
terms of l, and hence find the flux through the loop if the vertical component
of the flux density arising from the Earths magnetic field is B.
(b) Suppose that the loop is collapsed during a time interval t, so that the area
enclosed in negligible. Find an expression for the average current through the
loop, and show that the amount of charge to pass any point in the wire is given
by Bl2 /(4R) where R is the resistance of the wire.
(c) Estimate the average current that would flow if l = 1 m, R = 0.1 , B =
2 105 T and the collapse occurs in 2 seconds.
5. Unknown to Captain Jolly, a passenger on Popeye is holding a powerful bar
magnet such that its North Pole is closest to the front of the boat and its South
Pole is closest to the back.

163
(a) Consider the continuous conducting loop formed by the river and the University
footbridge. Give the direction in which the electric current in this loop will
flow:
i. as Popeye approaches the footbridge;
ii. as the passenger is directly underneath the footbridge;
iii. as Popeye moves away from the footbridge.
(b) The electric current in the footbridge/River Torrens loop dissipates some energy due to the resistance of the loop. Where does this energy come from?
(c) Is it possible for Captain Jolly to prevent a current flowing in the loop by
reseating the passenger (and reorienting the magnet)?
6. Consider whether each of the following statements is true or false, and explain your
answer.
(a) If a charged particle is moving in a straight line, it must be in a region of zero
magnetic field.
(b) A conventional current I flows in a conducting strip in a magnetic field, with
the field direction out of the page. The resulting Hall potential, shown in the
diagram of Figure I.4, proves that the charge carriers must be positive.

- - - - - - - I

B out

+ + + + + + + +

Figure I.4: Conducting strip in magnetic field.


(c) The induced emf in a circuit is proportional to the magnetic flux through the
circuit.

(d) Since s B dA = 0 (where S is a closed surface), magnetic flux through any
area must always be zero.
7. (a) Consider the following expressions for inductance.
N
0 N 2 A
L=
and
L=
I
l
Define the symbols in each expression, and explain what each expression refers
to.
(b) In electronic equipment, wires carrying equal and opposite currents are often
twisted together. Why?
(c) A resistor is made by winding a piece of resistance wire into a coil. How can
this be done so that there is no inductance?
8. Challenge Question.
Suppose you are given a length of wire from which to construct a solenoid with
maximum possible self inductance. Demonstrate whether it would be better to
have a long solenoid with small radius or a short solenoid with large radius.

164

APPENDIX I. EXERCISE SET 5

Appendix J
EXERCISE SET 6
Useful information

cos(A + B) = cos A cos B sin A sin B


cos( ) = cos
cos 2A = cos2 A sin2 A
1. (a) A circuit with a battery, a coil and a resistor in series carries a steady current.
Does the coil aect the current in the circuit? Does the coil have an inductance?
(b) In each of the circuits of Figures ?? and J.2 introduce symbols and reference
directions which will allow specification of the voltages across all of the circut
elements in relation to other circuit variables, such as charge or current, and
use Kirchhos loop theorem to write down an equation for the total voltage
change around the loop.

L
+

Figure J.1: An R L circuit.


In the circuit of Figure ?? express your result in terms of a loop current i, and
in the circuit of Figure ?? express your result in terms of the capacitor charge
q.
2. At time t = 0, a coil with an inductance of 2.0 H and resistance of 10 is suddenly
connected to a battery of emf E = 100 V and negligible internal resistance.
165

166

APPENDIX J. EXERCISE SET 6

L
i

R
C

-q

vC

Figure J.2: An R L C circuit.


(a) Treat the coil as an ideal resistor and ideal inductor in series, draw a schematic
circuit diagram for the situation, and introduce therein appropriate notation
and reference directions for the loop current and the voltages across the inductor and resistor.
(b) Sketch graphs of the potential dierence across each component as a function
of time.
(c) At 0.10 s after the connection is made, find the rates at which:
i. energy is being stored in the magnetic field;
ii. thermal energy is being dissipated in the resistance; and
iii. energy is being supplied by the battery.
Are your answers consistent with energy conservation?
3. (a) For a
sinusoidal voltage v = Vmax cos t, show that the r.m.s. value of v is
Vmax / 2.
(b) A 2.4 kW electric heater operates at mains voltage of 240 V r.m.s. Find the
values of the r.m.s. current, the peak current, the peak voltage and the heater
resistance.
4. A 5 F capacitor is charged to 20 V and then connected in series with an inductor
so that the current oscillates sinusoidally.
(a) Sketch graphs of the charge on the capacitor and the current as a function
of time. Hence sketch graphs of the energy stored in the capacitor and the
inductor as a function of time.
(b) Briefly describe the eect of including a resistance in the circuit.
5. A capacitor acts as a short circuit at high frequencies and as an open circuit at
low frequencies.
Explain this statement remembering that a short circuit means there is no impedance and an open circuit means there is infinite impedance.
(Hint: Write down an expression for the reactance of a capacitor as a function of
frequency.)

167
6. Sinusoidal signals with frequency f, but with dierent amplitude and phase, combine to give a resultant sinusoidal signal of frequency f. [Alternating voltages in a
circuit, and interfering light waves are examples.] The addition may be carried out
graphically, using trigonometry or using phasors.
(a) Use trigonometric identities to show that the sum of the voltage signals v1 =
A cos t and v2 = B cos(t + ) can be expressed in the form:
v = v1 + v2 = D cos(t + )
and find D in terms of A, B and .
(b) Compare your value of D with the length of the third side of a triangle, the
other two sides of which make an angle and have lengths A and B.
(c) Write down an expression for the r.m.s. value of the resultant voltage v in
terms of r.m.s. values of v1 and v2 .

7. A 50 Hz sinusoidal emf of peak value 200 V is applied to a resistor of 300 in series


with an inductor of 2 H (assume that the inductor has no resistance, which is an
ideal case).
Carefully construct a phasor diagram and read o the angle by which the current
lags the applied voltage.
8. (a) A 50 Hz 240 V r.m.s. supply is connected to a series LRC circuit. The circuit is then tuned i.e. adjusted to give a maximum current; the value of the
capacitance is found to be 1.3 F, and the current is 4 A r.m.s.
i. With the aid of a phasor diagram, show that a resonant circuit behaves
like a pure resistance.
ii. What is the value of the resistance?
iii. What is the value of the r.m.s. voltage across the inductance?
(b) The circuit is then de-tuned so that the net reactance of L and C together is
80 . What are the values of the current, the power factor and the dissipated
power?
9. Challenge question.
A circuit contains a coil (containing inductance and resistance) in series with the a
non-inductive resistance. A 50 Hz sinusoidal voltage of 60 V r.m.s. applied to the
circuit results in voltage drops of 20 V r.m.s. across the resistor and 45 V r.m.s.
across the coil.
How can the sum of the voltage drops be greater than the emf?
Show the voltages on a phasor diagram drawn to scale and determine the voltage
drops across the resistive part and the inductive part of the coil. If the current
flowing is 1.2 A r.m.s., determine the resistance and the inductance of the coil.
What is the power dissipated in the circuit?

Вам также может понравиться