You are on page 1of 324

The University of Queensland

Department of Physics
2008

Lecture notes and tutorial problems of the


undergraduate course
PHYS3050/PHYS7051
ELECTROMAGNETIC THEORY III

Lecturer: Zbigniew Ficek


Physics Annexe(6): Rm 436
Ph: 3365 2331
email: ficek@physics.uq.edu.au
http://www.physics.uq.edu.au/people/ficek/

Consultation Hours: Wednesday 2pm – 4pm


Preface
This lecture notes covers the principal elements of classical electromagnetic
theory embodying Maxwell’s equations with applications mainly to situations
where electric charge can be treated as a continuous fluid. The intention is
to introduce students to the background of classical field theory and the
applications of the electromagnetic theory to solid state physics, classical
optics, radiation theory and telecommunication.
The goal of this course is to provide a compact logical exposition of the
fundamentals of the electromagnetic theory and the applications to various
areas of physics and engineering. The treatment is quantitative throughout
and an attempt has been made to imbue students with a sound understanding
of the Maxwell’s equations and with the ability to apply them to modern
problems in physics.
The organization of the lectures is fairly standard and includes vector
analysis, electrostatic, magnetostatic, mathematical techniques in the solu-
tion of the Maxwell’s equations and the Laplace’s equation, time varying
fields and applications of the solution of the Maxwell’s equations. The ma-
terial on vector analysis gives greater emphasis to the relationship between
fields and their sources.
A number of revision questions have been included at the end of each
chapter. These questions have been designed not only to point out to the
student the essential material of the chapter but also to test the students’
understanding of the material presented in the chapter. In addition, tutorial
problems have been included at the end of the chapters that contain the
most important elements of the electromagnetic theory. These problems
will be discussed and solved in details during the tutorial sessions, but it
is hoped that the student will attack these problems before attending the
tutorial session.

2
Contents
1 The Classical Theory of the Electromagnetic Field 13
1.1 Elementary Aspects of Electromagnetism . . . . . . . . . . . . 13
1.2 Macroscopic Charges and Currents . . . . . . . . . . . . . . . 16
1.2.1 Charge Density . . . . . . . . . . . . . . . . . . . . . . 16
1.2.2 Current density . . . . . . . . . . . . . . . . . . . . . . 18
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2 Mathematical Description of Vector Fields 20


2.1 Gradient of a Scalar Function . . . . . . . . . . . . . . . . . . 20
2.2 Divergence Function . . . . . . . . . . . . . . . . . . . . . . . 22
2.3 The Flux of a Vector Field . . . . . . . . . . . . . . . . . . . . 24
2.4 Gauss’s Divergence Theorem . . . . . . . . . . . . . . . . . . . 24
2.5 The Continuity Equation for Electric Current . . . . . . . . . 25
2.6 Curl (Rotation) Function . . . . . . . . . . . . . . . . . . . . . 26
2.7 Stokes’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.8 Successive Application of ∇ . . . . . . . . . . . . . . . . . . . 28
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3 Vectorial Equations in Electromagnetism: Scalar and Vector


Potentials 34
3.1 Maxwell’s Equations, an Example of Vectorial Equations . . . 34
3.2 Scalar Potential . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.3 Vector Potential . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

4 The Experimental Basis of the Development of Electromag-


netic Theory 41
4.1 Coulomb’s Law – Force between Static Charges . . . . . . . . 41
4.2 Gauss’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.3 Biot-Savart Law – Force between Static Currents . . . . . . . 49
4.4 Current Element and Charge Element . . . . . . . . . . . . . . 51
4.5 The Lorentz Force . . . . . . . . . . . . . . . . . . . . . . . . 52
4.6 Amperes Circuit Law . . . . . . . . . . . . . . . . . . . . . . . 53
4.7 Faraday’s Law of Electromagnetic Induction . . . . . . . . . . 55

3
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

5 Differential Equations for the EM Field and Maxwell’s The-


ory 61
5.1 Differential Equations for the EM Field . . . . . . . . . . . . . 61
5.1.1 Divergence of E ~ . . . . . . . . . . . . . . . . . . . . . . 62
5.1.2 Curl of E~ . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.1.3 Divergence of B ~ . . . . . . . . . . . . . . . . . . . . . . 63
5.1.4 Curl of B~ . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.2 The Maxwell’s Theory . . . . . . . . . . . . . . . . . . . . . . 64
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

6 Maxwell’s Equations and Prediction


of Electromagnetic Waves 69
6.1 The Wave Equation for EM Waves in Vacuum . . . . . . . . . 70
6.2 Plane Wave Solution to the Wave Equation . . . . . . . . . . . 71
6.3 Harmonic Waves . . . . . . . . . . . . . . . . . . . . . . . . . 73
6.4 The Transverse Nature of Plane Waves in Vacuum . . . . . . . 74
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

7 EM Theory and Einstein’s Special Theory of Relativity 80


7.1 Lorentz Transformation Equations for Space and Time . . . . 81
7.2 Force Transformation Equations . . . . . . . . . . . . . . . . . 82
7.2.1 The Force between Two Charges Moving with Con-
stant Velocities . . . . . . . . . . . . . . . . . . . . . . 83
7.3 Electric Field Lines of a Moving Charge . . . . . . . . . . . . 87
7.4 Magnetic Field Lines of a Moving Charge . . . . . . . . . . . . 88
7.5 Invariance of the Maxwell equations under the Lorentz trans-
formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
7.5.1 Remark on the Electromagnetic Induction . . . . . . . 95
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

8 Energy in the Electromagnetic Field:


Poyntings’ Theorem 99

4
8.1 Rate of Doing Work by the Field on the Current – Ohmic
Heating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
8.2 Electrostatic Field Energy Density . . . . . . . . . . . . . . . 102
8.3 Magnetostatic Field Energy Density . . . . . . . . . . . . . . . 103
8.4 Poynting Vector . . . . . . . . . . . . . . . . . . . . . . . . . . 103
8.5 Phase Relationships in EM Waves . . . . . . . . . . . . . . . . 105
8.6 Momentum Flux . . . . . . . . . . . . . . . . . . . . . . . . . 105
8.7 EM Energy Flow: Circuit versus Field Theory . . . . . . . . . 106
8.7.1 Energy Flow in a Resistive Wire . . . . . . . . . . . . . 106
8.7.2 Energy Flow out of Battery . . . . . . . . . . . . . . . 108
8.7.3 Propagation of an Electromagnetic Wave along a Wire 109
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

9 Solution of Laplace’s Equation and Boundary Value Problem112


9.1 Uniqueness of the Solution of Laplace’s Equation . . . . . . . 113
9.1.1 Dirichlet Theorem . . . . . . . . . . . . . . . . . . . . 113
9.2 Solutions of Laplace’s Equation . . . . . . . . . . . . . . . . . 114
9.2.1 Method of Separation of Variables . . . . . . . . . . . . 115
9.2.2 Solution of the Laplace Equation in Spherical Coordi-
nates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

10 General Solution of the Maxwell’s Equations 138


10.1 Difficulty of the Direct Solution of Maxwell’s Equations . . . . 138
10.2 Scalar and Vector Potentials . . . . . . . . . . . . . . . . . . . 140
10.2.1 Lorenz Gauge . . . . . . . . . . . . . . . . . . . . . . . 142
10.2.2 Coulomb Gauge . . . . . . . . . . . . . . . . . . . . . . 143
10.3 Solution of the Inhomogeneous Wave Equations . . . . . . . . 145
10.4 Rigorous Solution: Green Functions Method . . . . . . . . . . 148
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

11 Electromagnetic Antennas: Hertzian Dipole 152


11.1 Generation of electromagnetic waves . . . . . . . . . . . . . . 154
11.1.1 Field of an Element of Alternating Current . . . . . . . 154
11.2 Power Radiated from the Current Element . . . . . . . . . . . 161
11.3 Gain of the Dipole Antenna . . . . . . . . . . . . . . . . . . . 163

5
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

12 Electromagnetic Theory of Polarizable


Materials 166
12.1 Potential and Electric Field of a Single Dipole . . . . . . . . . 167
12.2 Polarization Vector . . . . . . . . . . . . . . . . . . . . . . . . 168
12.3 Maxwell’s Equation for ∇ · E ~ in a Dielectric . . . . . . . . . . 171
12.4 Macroscopic Effects of the Polarizability . . . . . . . . . . . . 173
12.5 Dense Dielectrics: The Clausius-Mossotti Relation . . . . . . . 176
12.6 Dielectric in a Time Dependent Field . . . . . . . . . . . . . . 178
12.7 The Complex Susceptibility and Permitivity . . . . . . . . . . 182
12.8 The Loss Tangent . . . . . . . . . . . . . . . . . . . . . . . . . 185
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

13 Electromagnetic Theory of Magnetizable Materials 190


13.1 Magnetic Polarization Currents . . . . . . . . . . . . . . . . . 190
13.2 The Magnetic Intensity Vector H ~ . . . . . . . . . . . . . . . . 195
13.2.1 Static Magnetic Fields . . . . . . . . . . . . . . . . . . 196
13.3 Linear Magnetic Materials . . . . . . . . . . . . . . . . . . . . 197
13.4 Non-Linear Magnetic Materials . . . . . . . . . . . . . . . . . 200
13.4.1 Work done in magnetization of iron . . . . . . . . . . . 201
13.5 Permanent Magnetic Materials: Ferromagnets . . . . . . . . . 202
13.5.1 B~ and H~ fields of a ferromagnet . . . . . . . . . . . . . 203
13.5.2 Magnetic Poles . . . . . . . . . . . . . . . . . . . . . . 205
13.6 Time Dependent Magnetic Fields and Energy Loss . . . . . . 207
13.7 The complex magnetic susceptibility . . . . . . . . . . . . . . 208
13.8 Maxwell’s Equations in Dielectric and Magnetic Materials . . 209
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212

14 Poynting’s Theorem Revisited 214


14.1 Poynting Vector in Terms of E ~ and H ~ . . . . . . . . . . . . . 214
14.2 Poynting Vector for Complex Sinusoidal Fields . . . . . . . . . 216
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 219

6
15 Plane Wave Propagation in Dielectric and Magnetic Media 220
15.1 Dispersive Equation and Complex Propagation Number . . . . 222
15.2 Wave Refraction and Attenuation . . . . . . . . . . . . . . . . 225
15.2.1 Propagation of an EM wave in a low loss dielectric . . 225
15.2.2 Propagation of an EM Wave in a Good Conductor . . 227
15.2.3 Skin Effect . . . . . . . . . . . . . . . . . . . . . . . . . 229
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231

16 Transitions Across Boundaries for Electromagnetic Fields 234


16.1 Normal Components of the Fields . . . . . . . . . . . . . . . . 234
16.1.1 Normal Component of B ~ . . . . . . . . . . . . . . . . . 234
16.1.2 Normal Component of H ~ . . . . . . . . . . . . . . . . . 236
~
16.1.3 Normal Component of D and E ~ . . . . . . . . . . . . . 237
16.2 Tangential Components of the Fields . . . . . . . . . . . . . . 237
16.2.1 Tangential Component of E ~ . . . . . . . . . . . . . . . 238
16.2.2 Tangential Component of H ~ . . . . . . . . . . . . . . . 240
16.2.3 Tangential component of B ~ . . . . . . . . . . . . . . . 240
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242

17 Propagation of an EM Wave Across a Boundary 244


17.1 Representation of Plane Waves in Different Directions . . . . . 245
17.1.1 Representation of B~ in terms of E ~ . . . . . . . . . . . . 246
17.2 Directions of Reflected and Transmitted Waves . . . . . . . . 247
17.3 Angle of Reflection and Snell’s Law of Refraction . . . . . . . 252
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 253

18 Fresnel’s Equations 254


~ i normal to plane of incidence . . .
18.1 E . . . . . . . . . . . . . . 255
~ i in the plane of incidence . . . . .
18.2 E . . . . . . . . . . . . . . 257
18.3 Fresnel Equations for dielectric media . . . . . . . . . . . . . . 258
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261

19 Applications of the Boundary Conditions and the Fresnel


Equations 262
19.1 Applications in dielectrics . . . . . . . . . . . . . . . . . . . . 262

7
19.1.1 Polarization by reflection . . . . . . . . . . . . . . . . . 262
19.1.2 Total internal reflection . . . . . . . . . . . . . . . . . . 264
19.2 Transmission and Reflection at a Conducting
Surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
19.2.1 Field vectors at normal incidence . . . . . . . . . . . . 269
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272

20 Propagation of an EM Wave in a Rectangular Waveguide 273


20.1 Transverse Electric (TE) Modes . . . . . . . . . . . . . . . . . 275
20.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . 278
20.3 TE Modes in a Lossless Waveguide . . . . . . . . . . . . . . . 280
20.4 Phase and Group Velocities of Mode Propagation . . . . . . . 283
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 286

21 Relativistic Transformation of the Electromagnetic Field 287


21.1 The Principle of Relativity . . . . . . . . . . . . . . . . . . . . 287
21.2 Transformation of Electric and Magnetic Field Components . 293
21.3 Transformation Rules in Terms of Parallel and Normal Com-
ponents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294
21.3.1 Rules for Parallel Components . . . . . . . . . . . . . . 295
21.3.2 Rules for Normal Components . . . . . . . . . . . . . . 295
21.4 Transformation of the Components
of a Plane EM Wave . . . . . . . . . . . . . . . . . . . . . . . 297
21.5 Doppler Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
21.6 Transformation of Energy of a Plane EM Wave . . . . . . . . 301
Revision questions . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
Tutorial problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304

Revision Questions for the Final Examination 306

Appendix A: Proof of the Amperes law 312

Appendix B: Proof of the vector theorem, Eq. (73) 314

Appendix C: PHYS3050 Facts and Formulae 315

8
Literature
1. J.D. Jackson, Classical Electrodynamics, 3rd ed. Wiley 1999.
The course is aimed at this level of treatment.

2. R.K. Wangsness, Electromagnetic Fields, 2nd ed. Wiley 1986.


This is an alternative textbook for the course.

3. R. Plonsey and R.E. Collin, Principles and Applications of Elec-


tromagnetic Fields, McGraw Hill 1961.
This is a good general text on electromagnetic fields.

4. B.I. Bleaney and B. Bleaney, Electricity and Magnetism, 3rd ed.


Oxford U.P. 1983.
A very readable book ranging over circuit theory, electronics, electric
properties of matter and field theory.

5. J.A. Stratton, Electromagnetic Theory, McGraw Hill, 1941.


This is a very comprehensive reference on electromagnetic theory.

6. W.K.H. Panofsky and M. Phillips, Classical Electricity and Mag-


netism, 2nd ed. Addison-Wesley 1962.
This is a good text on the relativistic representation of electromag-
netism.

7. R.P. Feynman, R.B. Leighton, and M. Sands, The Feynaman


Lectures on Physics, Vol. 2, Addison Wesley, 1964.
A good reference to the physics of the emission of radiation.

9
”To anyone who is motivated by anything
beyond the most narrowly practical,
it is worthwhile to understand Maxwell’s equations
simply for the good of his soul”
− J.R. Pierce

11
1 The Classical Theory of the Electromag-
netic Field
Classical theory of the electromagnetic field or Classical Electrodynamics,
formulated by Maxwell more than 100 years ago, is now a well established
theory with many applications in different areas of physics, chemistry and en-
gineering. In this context ‘classical’ means ‘non-quantum’, but we would
like to point out that the basic equations of electromagnetism, the Maxwell’s
equations, hold equally in quantum and classical field theory.

We begin the study of the electromagnetic theory with a brief review of the
elementary aspects of electromagnetism that the student learnt in the first
year PHYS1002 course. These aspects are in fact about the microscopic
nature of the electromagnetic theory. To indicate the classical approach to
the electromagnetic theory and the macroscopic nature of electromagnetic
phenomena we are going to explore in this lecture notes, we first introduce
the concept of macroscopic distributions of charges and currents, i.e. we
define volume, surface and linear charge densities. Next, we will introduce
the concept of the current density and a total (macroscopic) current.

1.1 Elementary Aspects of Electromagnetism


In many textbooks on electromagnetism, the elementary aspects of electro-
magnetism are introduced as postulates. In fact, they are based on experi-
mental observations, and we will explore this fact in full details in one of the
lectures. At this introductory lecture, let us review briefly the elementary
aspects of electromagnetism that are:

1. Electromagnetic interactions are ONE of FOUR fundamental types:1


Type of interaction Relative Strength
Strong interaction (nuclear) 1
Electromagnetic 10−2
Weak interaction (e.g. β decay) 10−12
Gravitation 10−40

1
At our level of discussion there is no relation between these four types of interaction,
i.e. they cannot be considered as different manifestations of a single FORCE.

13
2. Electromagnetic (EM) forces or interactions are due to ELECTRIC
CHARGE, which is NOT in turn explicable in terms of anything else.

3. Charges are of two kinds called positive and negative. In the static
limit like charges repel and unlike attract.

4. Charges are quantized in units of e ' 1.6 × 10−19 [Coulombs].

5. In the static limit the inverse square (Coulomb) law of force holds:
1 q2 q 1 r̂
F~q2 = r̂ , q1 −→ q2 ,
4πε0 r2

that the charge q1 acts on the charge q2 with the force F~q2 . The force
acts along the distance r and the parameter ε0 determines the property
of the medium and is called the electric permittivity.

The Coulomb’s law holds for microscopic charges, i.e. it holds


only for charges whose the spatial dimensions are small com-
pared with the distance separating them.

The Coulomb’s law does not tell us how the first charge knows the
other one is present. One usually assumes that the charge produces an
electric field which then interacts with the other charges. To express
this explicitly, the Coulomb force is often written as

F~q2 = q2 E
~ q1 ,

~ q1 is the electric field produced by the charge q1 .


where E

The Coulomb law can be tested to great accuracy indirectly by showing


that no charge rests on the inside of a statically charged hollow con-
ductor.2

2
S.J. Plimpton and W.E. Lawton, Phys. Rev. 50, 1066 (1936).

14
If the exponent in the Coulomb’s law for point charges
1
E∼
rn
is not n = 2 but n = 2(1 ± ), the potential inside the hollow conductor
would be large. Since the potential inside the hollow conductor found
by the experiments was less in magnitude than a small detectable po-
tential, then  ' 10−9 , the level of sensitivity of the detector.

6. Electric charge is conserved (algebraically)


X
q = constant
whole universe

7. In the NON-STATIC case, i.e. the case of moving charges, the force is
no longer given by Coulomb’s Law. It is given by the Lorentz equation

F~q2 = q2 (E
~ + ~v × B)
~ ,

where E~ and B ~ are the electric and magnetic fields due to q1 , and ~v is
the velocity of the charge q2 .

8. In the electromagnetic theory, we assume that the fields E ~ and B~ de-


pend on the frame of reference of the observer, that E, ~ B ~ and the
~
force F must follow the required relativistic transformation law.

However, we do not think that q depends on the frame of reference.


i.e. q does not depend on its velocity with respect to an observer. In
terms of the relativistic theory, the charge is invariant under the Lorentz
transformation.3

The student has noticed that the Lorentz force involves both, the electric and
magnetic fields. Why there must be a B~ and how E ~ and B ~ are computed for
3
This is because in ordinary matter electrons move much faster than ions, their speeds
depend on temperature, and electric fields are not observed to arise from changes in
temperature.

15
arbitrary motion of charges is the substance of electromagnetic theory.

We now have the following picture of the basis of the electromagnetic theory:
ELECT ROM AGN ET IC F ORCE ON
1. CHARGE 1 =⇒ F IELD =⇒ CHARGE 2

2. Fields are generated by charges - NOT by other fields.

1.2 Macroscopic Charges and Currents


We know from the Millikan experiment that electric charge is quantized. The
electron is a point charge on the smallest scale measurable. We may then
speak, on a subatomic scale, of a microscopic theory of electromagnetism. On
a subatomic scale there must be very strong and rapidly varying electric and
magnetic fields on spatial scales ∼ 10−8 m and temporal scales ∼ 10−10 s.
When we measure the fields around a macroscopic circuit, clearly we are
not looking at these fields. We are measuring fields on distance scales much
larger than 10−8 m and time scales much longer than 10−10 s.

In the macroscopic context, we do not distinguish individual charges. It is


convenient and justifiable to regard the charge as a continuous ”fluid” dis-
tributed over the volume or surface of a charged material.

Thus, we may introduce the concept of macroscopic charges and currents to


indicate the macroscopic nature of electromagnetic phenomena.

1.2.1 Charge Density


Charges may be distributed throughout the volume of a material, or on the
surface of the material, or may be distributed along one dimensional wires.
Thus, there are different spatial distributions of charges considered in the
electromagnetic theory that result in three types of charge densities: volume
charge density, surface charge density and linear charge density.

16
When we encounter a large number of point charges in a finite volume, it
is convenient to describe the source in terms of a volume charge density,
defined as
Σq
ρ = lim ,
∆V
where Σq is the algebraic sum of the charge in the volume ∆V .

The limit is not to zero, but to a ∆V much larger than atomic scale size,
which is still very small on the laboratory scale.

If the volume charge density is represented by a continuous function ρ(~r),


the total charge Q in a volume V is given by
Z
Q= ρ(~r) dV .
V

Note that the charge density is a function of the position which varies smoothly
inside a charged material. An exception is a boundary between two charged
materials where ρ may change discontinuously due to the presence of surface
charges of a non-zero density. Thus, we may introduce the concept of surface
charge density.

A surface charge density is defined, analogously to the volume charge den-


sity, as
Σq
σ = lim ,
∆S
where Σq is the algebraic sum of the charge on the surface ∆S.

Then, the total charge continously distributed on a surface S is


Z
QS = σ(~r) dS .
S

In the same fasion, we may define a linear charge density on a length ∆L


Σq
γ = lim ,
∆L
and then the total linear charge continously distributed on a length L is
Z
QL = γdL .
L

17
1.2.2 Current density
For many purposes in the electromagnetic theory, it will be necessary to
introduce the concept of current density.
It measures the amount of current flowing
through an area normal to the direction
of the current.
We define the current density in the fol-
lowing way. Let IA = ∂q/∂t is a current
through the area A, as shown in Figure 1.
Then the current density, normal to the
area A is defined by
I δq
J~ = lim n̂ = lim n̂
Figure 1: Illustration of the current A δt A
flow through an area A and the evalua- ρ A v δt
tion of the current density through the = lim n̂ = ρ~v ,
area A.
δt A
where ~v = vn̂, and n̂ is the unit vector
normal to the area A, and the limit is
taken in the same sense as for ρ.

Total current through an arbitrary surface area

For a surface of an arbitrary shape, the current may not be normal to the
surface at all points on the surface.

However, if the current density J~ is


known at all very small areas, that
can be approximated by flat surfaces of
a macroscopic surface A, we still can
obtain the total current through the
area A.

Consider a current density J~ through


a macroscopic area A, as shown in Figure 2: Illustration of evaluation of the
Figure 2. To find the total current current density through an area A. The
through the area, we first divide the vector n̂ is a unit vector normal to the sur-
area A into a lot of small flat areas dA, face element dA.

18
and find the current through dA as4

δI = J~ · n̂ dA = J cos(θ)dA .

Then the current through the total area A is the sum of the contributions
from all the small elements of the area:
Z Z Z
IA = δI = J~ · n̂ dA = J~ · dA
~,
A A

~ = dAn̂ is a vector representing the element dA of the surface A.


where dA

Note: In vector analysis it is common to represent a surface by a vector


whose length corresponds to the magnitude of the surface area and whose
the direction is specified by the unit vector n̂ normal to the surface.

In summary, Electromagnetic theory formulated in terms of space depen-


dent quantities, charge densities ρ, σ or γ, and current density J~ is regarded
as a MACROSCOPIC THEORY.

Revision questions

Question 1. What are the elementary aspects of electromagnetism?

Question 2. How do we define volume, surface and linear charge densities?

Question 3. How do we define and evaluate the current density through an area?

Question 4. If the current density through an area is known, how does the
current through the area is evaluated?

4
In literature, very often small areas that a macroscopic area is divided to are called
elementary areas or elementary surfaces.

19
2 Mathematical Description of Vector Fields

The study of electromagnetic theory requires considerable knowledge of vec-


tor analysis. In this lecture, we will introduce vector operations such as
gradient, divergence and curl that we will need for our study of electromag-
netic theory. As we shall see, these vector operations are very convenient to
determine the properties of electromagnetic field, also considerably simplify
the formulation of electromagnetic theory and allow get a better inside into
electromagnetic phenomena.

2.1 Gradient of a Scalar Function

Let us first define a vector operation: Gradient of a scalar function. Assume


that a function Φ represents a scalar field and that Φ is a single valued, con-
tinuous, and differentiable function of position. Let dΦ represents a change
of Φ with a distance ds.
The gradient of the scalar function Φ is
defined as:
∂Φ
grad Φ ≡ ∇Φ = n̂ ,
∂s
where n̂ is a unit vector in the direction
the rate ∂Φ/∂s has its maximum value. In
other words, gradient tells us in which
direction the change in Φ is maxi-
mal.

~ a change of Φ can
For some other direction dX,
~
Figure 3: Illustration of the be found by projection of gradient of Φ on dX:
evaluation of gradient of a scalar
function Φ.
~ = ∂Φ ~ = ∂Φ cos θ dX .
dΦ = ∇Φ · dX n̂ · dX
∂s ∂s
We know from the vector analysis that it is convenient to represent a vec-
tor in a reference (coordinate) frame. Commonly used are the rectangular
(cartesian) coordinates, in which we can easily find that the x component of

20
the gradient is
∂Φ ∂Φ
(∇Φ)x = ∇Φ · î = n̂ · î = cos θ
∂s ∂s
δΦ δΦ ∂Φ
= lim = lim = .
δ(s/ cos θ) δx ∂x
Similarly, the y and z components of the gradient are
∂Φ ∂Φ
(∇Φ)y = , (∇Φ)z = .
∂y ∂z
Hence, in cartesian coordinates, the gradient of a scalar function Φ can be
written as5
∂Φ ∂Φ ∂Φ
∇Φ = î + ĵ + k̂ .
∂x ∂y ∂z
The gradient is analogous to multiplication of a vector by a scalar. The re-
sult, of course, is a vector.6

Remember: Mathematically, gradient of an arbitrary scalar function is a


vector that is pointing in the direction of maximal changes of the function.

Exercise in class:
Consider a scalar function Φ = x. Calculate the gradient
of Φ = x.

Solution: Since
∂Φ ∂Φ ∂Φ
=1 and = =0,
∂x ∂y ∂z
we find that
∂x ∂x ∂x
∇Φ = î + ĵ + k̂ = î .
∂x ∂y ∂z
This exercise explicitly shows that gradient is a vector pointing in the direc-
tion of maximal changes in Φ.

5
Expressions for the gradient in other coordinate systems are given in the Appendix C.
6
We do not usually take a gradient of a vector, the result would be a tensor.

21
The student should be able now to give a quick answer to simple questions
as ”Is gradient a scalar or a vector?” or ”In which direction does gradient of
a given scalar function point?” A practical test: Try to give the answer to
the following question without any calculations: What is the direction of the
gradient of a scalar function Φ = x + y?

2.2 Divergence Function


We have already defined the nabla operator and showed how it evaluates a
certain property of a scalar function. We now define the divergence function,
which involves the nabla operator and, as we shall see, tells us about the flow
of a vector field and its sources.

Definition of divergence:

The divergence is the scalar function which results from operation


of ∇ upon a vector F~ in a fashion analogous to the dot product of
two vectors. The result is a scalar function.

In Cartesian coordinates, the divergence function is written as7


∂Fx ∂Fy ∂Fz
div F~ ≡ ∇ · F~ = + + .
∂x ∂y ∂z
Properties: A positive divergence means that there is a source of a vector
field, and a negative divergence means the presence of a sink of the field.
When ∇ · F~ = 0 everywhere, the field F~ is called solenoidal, since no start-
ing points or sources can be assigned to the lines describing the field. In
other words, it has no sources or sinks.

Things to remember: Mathematically, divergence of an arbitrary vector


is a scalar. Physically, a non-zero ∇ · F~ implies a source (if positive) or a
sink (if negative), and if ∇ · F~ = 0, there is no source or sink – the field lines
have no beginnings or ends.
7
The forms that the divergence takes in other coordinate systems are given in the
Appendix C.

22
Exercise in class:
Consider two vectorial fields F~1 = xî and F~2 = y î. Which of
the fields is solenoidal?

Solution: A vector field is solenoidal when ∇ · F~ = 0. Thus,


we calculate the divergence of the fields
∂x ∂x ∂x
div F~1 = (î · î) + (ĵ · î) + (k̂ · î) = 1 .
∂x ∂y ∂z
and
∂y ∂y ∂y
div F~2 = (î · î) + (ĵ · î) + (k̂ · î) = 0 ,
∂x ∂y ∂z

from which we see that the field F~2 = y î is solenoidal.

This exercise shows that divergence of a given vector field is different from
zero only if the field amplitude changes in the direction of the field. So now it
is easy to visualize the divergence of a vector field. The divergence is related
to how the field changes as you move in the direction of the field.

One more example for a better understanding of the meaning of divergence.


Consider a vector ~r = xî + y ĵ + z k̂ that represents the position of an arbitrary
point in the Cartesian coordinates. The student can easily find that div~r = 3.
What does the number 3 mean? It means that the field changes by one unit in
each direction. However, this statement may not be generally true. For example,
divergence of a vector F~ = 2xî + y ĵ is also equal to 3, but in this case the field
changes by two units in the x direction and one unit in the y direction.

23
2.3 The Flux of a Vector Field
A vector field propagating in space may cross some surfaces not necessary
normal to the field direction.
In this case, we may speak about a flux of the
field through the surface. The flux is measured
by the number of field lines crossing the sur-
face.

Mathematically, the flux Ψ of a field F~ through


a surface S, not necessary a closed S, as shown
in Figure 4, is defined as:
Figure 4: A vector field F~ lines Z Z
crossing a surface S. Ψ= F~ · dS
~= F~ · n̂ dS ,
S S

where n̂ is the unit vector normal to the surface at the point where the ele-
ment of area dS is located. Thus, we calculate the flux through a surface S of
arbitrary shape by dividing it up into a lot of small surfaces dS and calculate
the flux through each of these small surfaces. next, we add (integrate) the
fluxes to obtain the total flux through the surface S.

2.4 Gauss’s Divergence Theorem


We now introduce an important identity in vector analysis, the Gauss’s di-
vergence theorem or Gauss’s law, which we will use frequently in our study
of the theory of the electromagnetic fields.

Consider a vector field F~ crossing a closed surface bounding a volume V , as


shown in Figure 5.

Gauss’s law, or sometimes called as the Gauss’s divergence theorem, says that
Z I
∇ · F~ dV = F~ · n̂ dS . (1)
V S

The Gauss’s law states that the volume integral of the divergence
of a vector field over a volume V is equal to the closed surface in-
tegral of the vector over the surface bounding the volume V .

24
Mathematically, the Gauss’s divergence the-
orem converts a volume integral of the di-
vergence of a vector to a closed surface inte-
gral of the vector, and vice versa. Physically,
the Gauss’s divergence theorem says that the
number of the field lines flowing through
the surface S is equal to the ”strength”
of the field source contained inside the vol-
ume V .

Figure 5: Illustration of the eval- Remember that n̂ is the unit outward


uation of the Gauss’s law for a vec- normal over S, as shown in Figure 5,
tor field F~ crossing a closed sur- and S is a closed surface bounding a vol-
face bounding a volume V .
ume V .

Associated with this is the pictorial representation of fields by lines of force,


the direction of F~ given by the tangent to a line and the strength of F~ given
by the line density per unit area.

Note that from the divergence theorem


1 I ~ Ψ
   
~
∇ · F = lim F · n̂dS = lim ,
dV →0 dV dV →0 dV

i.e. the divergence of a vector field is the emanating flux per unit volume.

2.5 The Continuity Equation for Electric Current


The Gauss’s divergence theorem allows us to derive the continuity equation
for the electric current known as the equation of conservation of the charge.

Suppose we have some charge of density ρ in a volume V enclosed by a sur-


face S, as shown in Figure 6. Let ~v is a macroscopic velocity of the charge.
Then, the rate of decrease of the total charge in the volume V is equal to the
rate of transport of the charge out through the surface S.

Mathematically, we can express this statement as


∂ Z I
− ρ dV = ρ~v · n̂ dS .
∂t
V S

25
Using the Gauss’ divergence theorem, we
can write
I Z
ρ~v · n̂ dS = ∇ · (ρ~v ) dV .
S V

Thus
Z
∂ρ Z
dV = − ∇ · (ρ~v ) dV .
∂t
V V

Figure 6: Illustration of a flow of However, this relation holds for arbitrary V ,


a charge through a surface S closing
a volume V .
so we can drop the integrals and obtain
∂ρ
= −∇ · (ρ~v ) .
∂t
Since

ρ~v = J~ ,

we finally obtain
∂ρ
+ ∇ · J~ = 0 ,
∂t
which is well known as the continuity equation, or the equation of conser-
vation of charge.

When stationary currents are involved, then ∂ρ/∂t = 0. In this case ∇·J~ = 0,
that is for stationary currents the current density is solenoidal.

2.6 Curl (Rotation) Function


Another very important property of vector fields is curl or rotation or cir-
culation. Curl is the operation of ∇ operator upon a vector in a fashion
analogous to the cross product of two vectors. The result is a vector that in
Cartesian coordinates is written as8
∂ F~ ∂ F~ ∂ F~
curl F~ ≡ ∇ × F~ = î × + ĵ × + k̂ × ,
∂x ∂y ∂z
8
Expressions for the curl in other coordinate systems are given in the Appendix C.

26
or
! ! !
∂Fz ∂Fy ∂Fx ∂Fz ∂Fy ∂Fx
∇ × F~ = − î + − ĵ + − k̂
∂y ∂z ∂z ∂x ∂x ∂y

î ĵ k̂

∂ ∂ ∂
= ∂x ∂y ∂z

.

F Fy Fz
x

Properties: Curl is nonzero when the field increases (or decreases) in a dif-
ferent direction that the field pointed. If the field is pointed in the same
direction as that in which is increased, the curl is zero. So the curl is related
to how the field changes as you move across the field.

When ∇ × F~ = 0 everywhere, the field F~ is called irrotational.

Remember: Mathematically, curl of an arbitrary vector is a vector.

Exercise in class:
Find which of the following vectorial fields is irrotational.

F~1 = 2 , F~2 = 2~r , F~3 = ∇r3 ,
r
where r = |~r| 6= 0 and r̂ is the unit vector in the direction
of ~r.

2.7 Stokes’s Theorem


The curl function allows us to introduce an another important identity in
the theory of vectorial fields, the Stokes’s theorem that is formulated as
I Z
F~ · d~l = ∇ × F~ · n̂dS ,
l S

~ is the infinitesimal vector length tangent to a closed path of length l


where dl
bounding a surface area S.

27
The Stokes’s theorem states that
the closed line integral of a
vector field F~ along the con-
tour bounding an open sur-
face S is equal to the sur-
face integral of the curl of
the vector field over the sur-
face.

We may write the Stokes’s theorem in


the form
Figure 7: Illustration of the evaluation of
1 I ~ ~
 
the Stokes’s Theorem. ∇ × F~ · n̂ = lim F · dl .
dS→0 dS
This gives an intuitive meaning to any component of ∇ × F~ in terms of the
line integral around a small element of surface, as shown in Figure 7. Simply,
∇ × F~ is a measure of the vorticity of the field.

2.8 Successive Application of ∇


We can introduce scalar and vector products in which the operator ∇ ap-
pears more than once. For example, since the gradient of an arbitrary scalar
function Φ is a vector, we can take the divergence of the gradient
∂ 2Φ ∂ 2Φ ∂ 2Φ
!
∂Φ ∂Φ ∂Φ
∇ · ∇Φ = ∇ · î + ĵ + k̂ = + + 2 .
∂x ∂y ∂z ∂x2 ∂y 2 ∂z
The same result is obtained if we take ∇ · ∇ as a new operator ∇2 with
properties
∂2 ∂2 ∂2
∇2 ≡ ∇ · ∇ = + + .
∂x2 ∂y 2 ∂z 2
The operator ∇2 is called the Laplacian and is a scalar.

The Laplacian may also be applied to a vector, with the result

∂ 2 F~ ∂ 2 F~ ∂ 2 F~
∇2 F~ = + + .
∂x2 ∂y 2 ∂z 2

28
It is also possible to form the curl of the gradient, which is identically zero

∇ × ∇Φ = 0 . (2)

The divergence of the curl of a vector is also identically zero9

∇ · ∇ × F~ = 0 . (3)

These two properties are very useful in the vector field theory, in particular in
the electromagnetic theory. The relation (2) shows that an irrotational field
can always be expressed as gradient of an arbitrary scalar field. Similarly,
relation (3) shows that any solenoidal field can always be expressed as a curl
of an arbitrary vector field. Further discussion of the successful application
of ∇ is left to the tutorial problems.

9
The proof of this and the above property is left to the student as a tutorial problem.
Note that each of these properties involves two vector operations, but in writing them
we usually do not put brackets that would indicate in which order the two operations
should be applied. Is it clear to the student in which order the vector operations should
be applied?

29
Revision questions

Question 1. What is the physical meaning of the gradient of a scalar function?

Question 2. Derive the explicit form of ∇ in the Cartesian coordinates.

Question 3. What is the physical meaning of the divergence of a vector?

Question 4. What is the physical meaning of the curl of a vector?

Question 5. Write the curl of an arbitrary vector in the Cartesian coordinates.

Question 6. Explain, without using any mathematics, why ∇ · ∇ × F~ = 0.

Tutorial problems

Problem 2.1 Direction of a vector

Given a vector A~ = −î + 2ĵ − 2k̂ in Cartesian coordinates, find


~
the expression for the unit vector  in the direction of A.

Problem 2.2 Relation between two vectors

Show that, if A~·B ~ = A ~·C~ and A~×B ~ = A


~ × C,
~ where A
~ is
~ ~ ~ ~
not a null vector (A 6= 0), then B = C.

Problem 2.3 Multiple applications of · and ×

~ B,
Consider arbitrary vectors A, ~ C
~ and D.
~

~ × (B
(a) Is A ~ × C)
~ equal to (C
~ × B)
~ × A?
~ Explain.

30
~ · (B
(b) Is A ~ × C)
~ equal to (C
~ × B)
~ · A?
~ Explain.

~×B
(c) Does A ~ =A
~×C
~ implies B
~ = C?
~ Explain.

~ =∇×B
(d) Does ∇ × A ~ implies A
~ = B?
~ Explain.

(e) Which of the following expressions do not make sense? Ex-


plain.
~ × B/|
A ~ B|,
~ ~ · D/(
C ~ A ~ × B),
~ A~ × B/(
~ C ~ · D),
~
~×B
A ~ × C,
~ ~
∇ · (∇A), ~ · B),
∇2 (B ~
~ × B),
∇2 (A ~ ∇ × (A~ · B)
~ .

(f ) Which of the following expressions are vectors and which are


scalars?
~
∇(∇ · A), ~ · ∇)B,
(A ~ ∇(A ~ · B),
~ ~ × B),
∇2 (A ~
~ · (B
∇A ~ × C),
~ ~ × ∇[B
A ~ · (C
~ × D)],
~ ~ · B),
∇2 (A ~
~ × B).
∇ × (A ~

Problem 2.4 Vector components and directions

(a) Find the relative position vector R~ of the point P (2, −2, 3)
~
with respect to Q(−3, 1, 4). What are the direction angles of R?
~
(The direction angles are the angles between R and the x, y and z
axes repectively).

~ = î + 2ĵ + 3k̂ and B


(b) Given the two vectors A ~ = 4î − 5ĵ + 6k̂,
find the angle between them. Find the component of A ~ in the
~
direction of B.

~ = 2î + 3ĵ − 4k̂ and B


(c) Given the vectors A ~ = −6î − 4ĵ + k̂ find
~ ~
the component of A × B along the direction of C ~ = î − ĵ + k̂.

31
Problem 2.5 Flux of a vector field

(a) Given the vector field A ~ =


xy î + yz ĵ + zxk̂, evaluate directly
from the definition the flux of A ~
through the surface of a rectan-
gular parallelepiped of sides a, b, c
with the origin at one corner and
edges along the positive directions
of the rectangular axes, see Fig-
ure 8.

~
R
(b) Also evaluate ∇ · AdV directly
Figure 8: Contour for evaluation of the flux and show that it is equal to the cal-
of a vector field. culated flux as predicted by Gauss’s
Theorem.

Problem 2.6 Component in the direction of a gradient

(a) A family of hyperbolas in the xy plane is given by u = xy.


Find ∇ u.

(b) Given the vector A ~ = 3î + 2ĵ + 4k̂, find the component of A
~
in the direction of ∇ u at the point on the curve for which u = 3
and x = 2.

Problem 2.7 Unit normal to ellipsoids

The equation giving a family of ellipsoids is:

x2 y 2 z 2
u= + 2 + 2 .
a2 b c
Find the unit vector normal to each point of the surface of these
ellipsoids.

Problem 2.8 Divergence of a radial field

For fields of the form rn r̂, (r 6= 0), find for which values of n
the divergence is zero.

32
Problem 2.9 Field of a cylindrical form

For fields of the cylindrical form ρn φ̂, (ρ 6= 0), find for which
values of n the curl is zero.

Problem 2.10 Line integral of a vector

The vector field


~ = x2 y î + xy 2 ĵ + a3 e−βy cos α xk̂ ,
A

where a, α, β are constants.

~
Evaluate directly the line integral of A
around the closed path in the xy plane
as shown in the Figure 9. The straight
portions are parallel to the axes and
the curved portion is the parabola
y 2 = kx, where k = constant. Eval-
~ over
uate the surface integral of ∇ × A
the area S enclosed by C. Show that
the two are related as expected by
Stokes’ Theorem.

Figure 9: Contour for evaluation of the curl


of a vector.

33
3 Vectorial Equations in Electromagnetism:
Scalar and Vector Potentials

In this lecture, we will illustrate the application of the Gauss’s and Stockes’s
theorems to vectorial equations involving ∇· and ∇× operations. These type
of operations are involved in the Maxwell’s equations, the basic equations for
the electromagnetic theory. We shall discuss only the vectorial nature of
the equations, deferring a detailed physical interpretation of the equations
to the next chapter when we consider the experimental basis for electromag-
netism. Furthermore, we illustrate the Helmholtz Theorem for the unique
determination of a given vector and discuss some properties of the successive
application of ∇, which will allow us to introduce the concept of vector and
scalar potentials to the electromagnetic field theory.

3.1 Maxwell’s Equations, an Example of Vectorial Equa-


tions
Consider the Maxwell’s equations in the differential form that are an example
of differential vectorial equations involving ∇· and ∇× operations. These
~ and B
differential equations are for the vector fields E ~ and are of the form

~ =ρ
∇·E , (4)
ε0
~ = 0,
∇·B (5)
~
~ = − ∂B ,
∇×E (6)
∂t
~
~ = µ0 J~ + 1 ∂ E ,
∇×B (7)
c2 ∂t

These four equations determine ∇· and ∇× of the two vectors E ~ and B,


~
and we need both operations for each of the EM field vectors to completely
determine E~ and B.
~
An obvious question arises: Why do we need both ∇· and ∇× equations for
~ and B?
each of the vectors to completely determine E ~

34
The answer is in the Helmholtz Theorem, which says that an arbitrary
vector F~ can always be written as

1 Z ∇ · F~ 1 Z
∇ × F~
F~ = − ∇ dV + ∇× dV
4π V r 4π V r

= F~l + F~t .

In other words, if the divergence and curl of a vector field are known every-
where in a finite region, then the vector field can be found uniquely. Thus,
specification of ∇ · F~ and ∇ × F~ is necessary and sufficient to determine F~ .
Hence, we need four equations of ∇· and ∇× type to determine E ~ and B.
~
~
Since F can be found from the knowledge of ∇· and ∇×, the divergence and
curl are often called the sources of the field.

3.2 Scalar Potential


The electromagnetic field equations, the Maxwell’s equations, are coupled dif-
ferential equations whose the solution is not in general simple and straight-
forward. Even the calculation of electric field produced by a macroscopic
charge from the Coulombs law is usually a difficult task. Often the solutions
are aided by the use of potentials.

What do we mean by a potential and how helpful it is in the calculation of


the fields?

The idea is that the expression of the fields in terms of potentials simpli-
fies the equations to a form that can be readily solved. In particular, the
Maxwell’s equations are simplified to a form that can be solved in a reason-
ably simple way.

A potential is a quantity from which a vector field can be derived by some


process of differentiation. To show this, consider a field with zero rotation
or curl, ∇ × F~ = 0. Since ∇ × ∇Φ ≡ 0, then the equation ∇ × F~ = 0 will
have an integral of the form:

F~ = ∇Φ .

35
Thus, the field with zero curl may be derived from the gradient of the scalar
function Φ. In the field theory, the function Φ is called a scalar potential.

The scalar potential function is very often used in the electromagnetic field
~ =
theory. For example, the electrostatic (Coulomb) field has zero curl, ∇× E
10
0. Hence, we can always write the electrostatic field as
~ = −∇Φ .
E

The minus sign is inserted in the definition of Φ to agree with the definition
of Φ as a potential ENERGY. The negative sign can also be understood phys-
~ is in the direction that a positive charge moves,
ically from the fact that E
hence in the direction of decreasing potential.

The scalar potential is useful in the calcu-


lations of work done on charges moving in
an electric field.
The work done by the field in moving the
unit charge q from a point A to a point B
~ as illustrated in Fig-
in an electric field E,
ure 10, is given by
ZB ZB
W ~ · d~l =
= − E ∇Φ · d~l
q
A A
Figure 10: Illustration of a path ZB B
∂Φ ~
Z
along which a unit charge is moved = r̂ · dl = dΦ
from a point A to a point B in an ∂r
A A
~
electric field E. = ΦB − ΦA = ∆ΦAB .

We see that the work done on moving the


charge from the point A to the point B
is equal to the potential difference between
the two points.

The work is independent of the path, it depends only on the position of the
start and end points. If the charge gains energy in moving from A to B, we
10 ~ 6= 0 in general in electromagnetism,
As one can see from the Maxwell’s equations, ∇×E
~
so it is not in general possible to write E = −∇Φ.

36
say that the point B is at higher potential than the point A, and vice versa.

~ = 0.
Note an another interesting property of the field with ∇ × E
From the Stokes’s law we have that
I Z
~ · d~l =
E ~ · n̂dS = 0 .
∇×E

It says that no work is done by the field


on a test charge if it is moved along
an arbitrary closed path in the field.
Consequently, the work along two ar-
bitrary paths from A to B, as illus-
trated in Figure 11, is
Figure 11: Two arbitrary paths leading
∆Φ(AB)1 = −∆Φ(BA)2 = ∆Φ(AB)2 . from a point A to a point B.

In summary, the work done by the field is independent of the path chosen.
In other words, if a charge is moved around any closed path, no net energy
is requred. Thus, a field F~ with ∇ × F~ = 0 is a conservative field of force.

3.3 Vector Potential

An another useful potential in the electromagnetic theory is the vector po-


tential. To illustrate the idea of the vector potential and how it is defined,
consider a field with zero divergence, ∇ · F~ = 0. Since for an arbitrary
~
vector A:
~≡0,
∇·∇×A

we have that the vector F~ can always be written as

F~ = ∇ × A
~.

Thus, the field with zero divergence may be derived from the curl of the
~ In the field theory, the vector A
vector field A. ~ is called a vector potential.

~ = 0 always in electromagnetism so there will always


We shall see that ∇ · B
~ such that B
be a vector field A ~ = ∇×A ~ and such an A
~ is referred to as

37
the vector potential in electromagnetism (though there may be other elec-
tromagnetic field functions with zero divergence).

Note also that just writing ∇ × A ~=B ~ does not completely specify A~ even
~ is known everywhere. According to the Helmholtz Theorem, one needs
if B
~ as well to completely determine A.
to specify ∇ · A ~ Equivalently, we can say
that defining ∇ × A ~=B ~ still leaves us free to define ∇ · A.
~

Exercise in class: Applications of ∇· and ∇×.

(a) For a given scalar function Φ and vectors A,~ B,


~ C,
~ in-
dicate successive steps you would follow in the calculation
of the following expressions:
~;
∇×∇·A ~ · ∇B
A ~ ; ~ ·B
A∇ ~ ; ~×∇·B
A ~ ;

~×B
∇A ~ ·C
~ ; ~×B
∇Φ · A ~ ; ~·B
A ~ × ∇Φ ;

~·B
∇×A ~ ×C
~ ; ~ · ∇Φ ;
A ~ × ∇B
∇·A ~ ·C
~ .

(b) Which of the expressions in part (a) are vectors?

38
Revision questions

Question 1. Explain why in electromagnetism we need two divergence and two


~ and magnetic B
curl equations for the electric E ~ fields to com-
~ and B?
pletely determine the fields E ~

Question 2. What is the Helmholtz’s theorem telling us about a vector F~ ?

Question 3. Why are the scalar and vector potentials useful in electromagnetism?

Question 4. Explain, which vectors can be expressed in terms of a scalar poten-


tial and which can be expresses in terms of a vector potential.

Question 5. Explain, when a field F~ is a conservative field of force?

39
Tutorial problems

Problem 3.1 If the potential Φ satisfies the Laplace equation, ∇2 Φ = 0, show


that the vector ∇Φ is both solenoidal and irrotational.

Problem 3.2 Show from the form of Maxwell’s equations that it should be
~ and Φ such that
possible to define electromagnetic potentials A
the fields can be calculated from
~
~ = −∇Φ − ∂ A ,
E
∂t
~ ~
B = ∇×A .

Problem 3.3 In an application of the electromagnetic theory to nuclear physics,


~ B,
the so-called vectorial mesons are described by three vectorial fields E, ~ A
~
and a real scalar field Φ. The fields satisfy the following equations
~ = −µ2 Φ ,
∇·E
~
~ = ∂ E − µ2 A
∇×B ~,
∂t
~
~ = − ∂ A − ∇Φ ,
E
∂t
where µ2 is a positive constant.

Show that the above equations can be converted into two differen-
tial equations

~+ ∂Φ
(i) ∇·A =0,
∂t
∂2Φ
(ii) ∇ 2 Φ − µ2 Φ − 2 = 0 .
∂t

40
4 The Experimental Basis of the Develop-
ment of Electromagnetic Theory
In this lecture, we will analyse experiments that led to the development of
the electromagnetic theory. We will show that the Coulomb Law for the elec-
trostatic force between two point charges is the experimental basis for the
development of the electric theory and the Biot-Savart Law for the magnetic
force between two static currents is the experimental basis for the develop-
ment of the magnetic theory. We also illustrate two useful laws of electro-
magnetism, the Gauss’s and Amperes Circuit laws, that are very helpful in
calculations of the electric and magnetic fields produced by some symmetric
distributions of charges and currents.

4.1 Coulomb’s Law – Force between Static Charges


In 1785, Coulomb investigated the nature of the force between charged bod-
ies, and the results of his experiments can be formulated mathematically in
what is known as Coulomb’s law
1 q q0
F~2 = r̂ .
4πε0 r2
We can write Coulomb’s law in terms of the electric field produced by the
charge q that acts on the charge q 0 :

F~q0 = q 0 E
~ ,

where

~ = 1 q
E r̂
4πε0 r2
is the electric field produced by the charge q. The electric field is an example
of a vector field. In principle, we can always calculate an electric field using
Coulomb’s law. However, there is an alternative way we can find the electric
field. In particular, the field may be represented by means of the flux con-
cept. The total flux of E ~ from a point charge q may be readily calculated by
integrating E ~ · dS~ over a surface enclosing q.

41
4.2 Gauss’s Law
Consider a macroscopic charge q closed by a surface S, as shown in Figure 12.
We will show that the flux of the electric field produced by the charge q is
proportional to the charge q, and is independent of the shape of the surface
closing the charge.
Consider first the flux through a spherical sur-
face, for which we find after a simple algebra,
that the flux is
I
~
I
q
ΨS1 = E · n̂dS = r̂ · n̂dS
4πε0 r02
S1 S1
q I q q
= 2
dS = 2
4πr02 = .
4πε0 r0 4πε0 r0 ε0
S1

Thus, we see that the flux through the


spherical surface is proportional to the
charge enclosed by the surface. An obvi-
Figure 12: Illustration of the ous question arises: Is this relation valid
calculation of the Gauss’s law. only for the case of spherical surfaces or
maybe it is valid for an arbitrary sur-
face?

In order to answer this question, consider the flux through an arbitrary sur-
face
I I
q
ΨS2 = dΨS2 = r̂ · n̂dS ,
4πε0 r2
S2 S2

where S2 is an arbitrary surface enclosing the charge q.

However, from the inverse square law and some geometry, we find
q q dS cos θ q
dΨS2 = 2
r̂ · n̂dS = 2
= dΩ ,
4πε0 r 4πε0 r 4πε0
where dΩ is the solid angle subtended by dS at q.

We see from Figure 12 that the element of solid angle dΩ is independent of


where we cut the bundle of electric lines of force. Thus
I I
q
dΨS2 = dΨS1 and then dΨS2 = dΨS1 = .
ε0

42
Hence
I
q
ΨS2 = dΨS2 = . (8)
ε0
Equation (8) is the statement of the Gauss’s Law. It says that the flux
~ through a closed surface equals 1/ε0 times the total
of the elcetric field E
charge contained within the surface.

Furthermore, the Gauss’s Law applies not


only to a single charge contained within
a surface S, but also for some arbitrary
number of point charges q1 , q2 , q3 , · · · , qn or
a continuous distribution of charges within
the surface S.

We can write this statement in the following


way
Figure 13: Illustration of the I I
Gauss’s law for a number of point dΨ = ~1 + E
(E ~2 + E
~ 3 + · · ·) · n̂ dS
charges enclosed by a surface S.
= Ψ 1 + Ψ2 + Ψ3 + · · ·
q1 q2 q 3 1 X
= + + + ··· = q,
ε0 ε0 ε0 ε0
P
where q is the algebraic sum of all charges
within the surface S. For a contiuous distribution of charges inside the
surface S
X Z
q= ρdV ,
V

where ρ is the charge density in the volume V enclosed by the surface S.

This general property is what one could expect, the Gauss’s law applies to an
arbitrary number of charges due to the additive nature of the fields produced
by each charge separately.

Gauss’s Law for a source field charge outside S

43
If the source charge q is outside S, as illustrated in Figure 14, the surface in-
tegral vanishes since the total solid angle subtended at q by the surface is zero.

Proof:

Consider first the flux of the electric field through a surface element dS1 seen
from the charge q under a solid angle dΩ:
q r̂ · n̂1 dS1
dΨ1 = .
4πε0 r12
However
r̂ · n̂1 dS1 dS1
2
= − 2 = −dΩ ,
r1 r1
where minus sign is from the fact that at the surface dS1 the angle between r̂
and n̂1 is larger than 90◦ .
Similarly, the flux of the electric field
through a surface element dS2 seen from
the position of the charge q under the
same solid angle dΩ is
q r̂ · n̂2 dS2 q
dΨ2 = 2
= dΩ ,
4πε0 r2 4πε0
where now we have positive sign since at
the surface dS2 the angle between r̂ and
n̂2 is smaller than 90◦ . Thus, we find that
Figure 14: Illustration of the calcula-
tion of the Gauss’s law for the source
dΨ1 + dΨ2 = 0 ,
charge outside the surface S.
and integrating over all S, we obtain that
the total flux through the surface S is

I
Ψ= dΨ = 0 .
S

The physical interpretation of this result is that field lines originating from
an external charge and entering the surface S must also leave this surface.

44
In summary, the Gauss’s law says that the total electric flux through a
surface S enclosing a charge q is:

q X
Ψ= , where q = charges INSIDE S .
ε0
Using the definition of the flux, we often write the Gauss’s law as
~= q .
I
~ · dS
E
S ε0
The power of the Gauss’s law lies in the fact that we are free to apply it to any
closed surface whose shape can be chosen arbitrary such that the evaluation
of the surface integral becomes a simple straightforward task. The Gauss’s
law is particularly useful in the calculation of the electric field produced by
certain symmetrical charge distributions. If the distribution of the charges
does not correspond to any simple symmetry, the Gauss’s law is not much
helpful in the calculations. We illustrate this in the following example.

Example of an application of the Gauss’ Law

An infinitely long line is positively and uniformly charged with a constant


linear charge density ρl . Use (a) Coulomb law, (b) Gauss’s law to find the
electric field about the line.

(a) If we have to calculate the field due to a static macroscopic distribution


of charge using the Coulomb law, we divide the macroscopic charge into
~ The field
infinitesimal (point) charges dq which produce an electric field dE.
dE~ of the point charge dq is given by the Coulomb field

~ = 1 dq
dE r̂ .
4πε0 r2
Then the total electric field is found by vector addition
Z
~ =
E ~ .
dE

Since we are adding vectors, a caution must be employed. We use the fol-
lowing procedure, which is general and can be employed to any system:

45
~ produced by the infinites-
1. Write the expression for the electric field dE
imal (point) charge dq.
~ into components
2. Choose a reference frame and resolve the vector dE
dEx , dEy , and dEz .
3. Calculate
R
each component of E separately by integration, e.g. calculate
Ex = dEx .
~ from its components
4. Find the resultant E
~ = Ex î + Ey ĵ + Ez k̂ .
E

Return now to our example of the charged infinitely long line and lets try to
solve the example problem following the above procedure.
Take a small element dl of the line contain-
ing a point charge dq. Electric field pro-
duced by the point charge at A distance r
from dl is given by the Coulomb field

~ = 1 dq
dE r̂ .
4πε0 r2
We see from the figure that
h
r= , l = h cot θ .
sin θ
Since the density of the charges on the line is
Figure 15: Illustration of an ap- constant, the charge on the line element dl is
plication of the Coulomb law to the
calculation of the electric field of an
h
dq = ρl dl , where dl = − dθ .
infinitelt long line charge. sin2 θ
Then
sin2 θ 
!
~ = ρl h 
dE − 2 dθ cos θ î + sin θ ĵ
4πε0 sin θ h2
ρl  
= − cos θî + sin θĵ dθ ,
4πε0 h
where we have decomposed the unit vector r̂ into two (x, y) components
r̂ = cos θî + sin θĵ .

46
Integrating over θ from θ = π to θ = 0, as the line is infinite and we go in the
direction of increasing l (from x = ∞, θ = π to x = −∞, θ = 0), we obtain

~ = ρl h i
E (− sin 0 + sin π) î + (cos 0 − cos π) ĵ
4πε0 h
ρl
= ĵ .
2πε0 h
Thus, the electric field produced by the charged line depends inversely on the
distance from the line and points in the direction perpendicular to the line.

(b) Let us now calculate the field by the direct application of the Gauss’s law.

The electric field near the uniformly charged line must be radially directed
because of the symmetry of the problem. The field must have cylindrical
symmetry because the problem is unchanged by rotating the line about its
axis. The field must also be independent of position along the line because
the distance to either end is infinite.

This is the ideal situation for the


application of Gauss’s Law. Due
to the cylindrical symmetry of the
field, we can apply a cylinder sur-
face of radius h and length L cen-
tered about the line of charge, as shown
in Figure 16. Such a surface is of-
ten referred to as a Gaussian sur-
face.

According to the Gauss’s law


Figure 16: Application of the Gauss’s
law to an infinitely long line charge.
~= q ,
I
~ · dS
E
S ε0
where q = ρl L is the charge on the line closed by the cylinder surface.

The flux through the cylinder surface splits into three fluxes
I Z Z Z
E ~ =
~ · dS E ~ A+
~ · dS ~ B+
~ · dS
E ~ C .
~ · dS
E
S A B C

47
Since E ~ A, E
~ ⊥ dS ~ C, E
~ ⊥ dS ~ B , and the magnitude of E
~ k dS ~ is constant
along the surface B, the flux through the cylinder reduces to that over the
surface B only
Z Z
E ~ =
~ · dS EdSB = 2πhLE ,
S B

and since the cylinder is symmetrically positioned about the line of charge,
the magnitude of E is constant over the surface B. Then, according to the
Gauss’s law
ρl L
2πhLE = ,
ε0
which gives
ρl
E= .
2πε0 h
Note how simple are the calculations of the electric field using the Gauss’s
law. However, we were able to solve this problem because we knew the
direction of the field at any point around the line.

48
4.3 Biot-Savart Law – Force between Static Currents

There is a force not only between electric charges but also between electric
currents. This force has a different nature than that one due to electric
charges, it is due to magnetic fields produced by the currents.

It has been first noticed by Oer-


sted in 1819, and few years later,
in 1827 by Ampere who showed that
quantitatively the magnetic forces
in macroscopic circuits can be ac-
counted for by what has come
to be known as the Biot-Savart
Law.

To illustrate the presence of mag-


netic forces between electric currents,
consider an experiment involving two
Figure 17: Example of an experiment long parallel wires carrying currents I
1
illustrating the existance of the magnetic
force between two current carrying wires. and I2 and separated by a distance d,
as shown in the Figure 17.

Experimental observations:

• If I1 kI2 then the force F is attractive.


• If I1 antik I2 then F is repulsive.
• If one of the wires is rotated through 90◦ then F = 0.
• The force is proportional to the currents I1 and I2 .

All the above observations can be conbined into a single equation for the
force acting on the current I2 :

µ0 I2 d~`2 × (I1 d~`1 × r̂)


dF~2 = ,
4π r2
where µ0 = 4π × 10−7 [H/m] in SI units, is the permeability of the vacuum.

49
We can write the force as

dF~2 = I2 d~`2 × dB
~ ,

where

~ = µ0 I1 d~`1 × r̂
dB ,
4π r2
which is known as the Biot-Savart law for magnetic field produced by the
current element I1 d~`1 .

The Biot-Savart law allows to compute magnetic field produced by an arbi-


trary current distribution Id~`

~ µ0 I Z d~` × r̂
B= . (9)
4π l r2
The method requires integration of small current elements. Note, all three
quantities appearing under the integral change during the integration, which
complicates the evaluation of the integral.

~ by using the following procedure.


We can simplify the calculations of B

If we replace r̂/r2 by −∇(1/r), the integrand becomes

−d~` × ∇(1/r) .

Next, using a vector identity we find that


 
d~` 1 1 1
   
∇×  ~ ~
= ∇ × d` − d` × ∇ ~
= −d` × ∇ ,
r r r r

since ∇ × d~` = 0. Hence, we can write

~ µ0 I Z d~`
B =∇× . (10)
4π l r
Thus, the magnetic field can be expressed as
~ =∇×A
B ~.

50
~ = 0 always. Thus, in this case we can first calculate A:
We see that ∇ · B ~

~ µ0 I Z d~`
A= , (11)
4π l r

which involves only two variables d~` and r, and then using (10), we find B.
~

~ is easier to calculate than the original expression (9) for B.


The integral for A ~
Since the curl operation is readily performed, we may use (11) as an inter-
mediate step for finding B ~ in a simpler way.

As we have already mentioned, the vector A ~ is called a vector potential,


~ in electromag-
and will see later in the course many useful applications of A
netic theory.

4.4 Current Element and Charge Element

The Biot-Savart law shows that the magnetic field is produced by a current
element. Here, we will show that in fact the magnetic field is produced by
moving charges.
A charge dq moving with a velocity v
is equivalent to an element of cur-
rent Id`:
d`
dq = Idt = I ,
v
where dt is the time for all the charge
in d` to pass out of the volume, as il-
lustrated in Figure 18.
Figure 18: Current element of a charge Hence
moving with a velocity ~v .

dq~v = Id~` .

The force between the two current elements can then be written as
µ0 (dq1~v1 × r̂)
dF~2 = dq2~v2 × ,
4π r2

51
or

dF~2 = dq2 ~v2 × dB


~1 ,

with

~1 = µ0 ~v1 × r̂
dB dq1 2 .
4π r
This shows that magnetic field is produced by a moving electric charge.
Moreover, it shows that the magnetic field can act with a force only on
moving charges. No force of a magnetic field on stationary charges.

4.5 The Lorentz Force

When a charge is moving in electric and magnetic fields, the force acting on
the charge is no longer the Coulomb force. It is the Lorentz force that is
obtained putting Coulomb’s Law F~E = q E~ and the Biot-Savart Law F~M =
~ together:
q~v × B

F~EM = F~E + F~M = q(E


~ + ~v × B)
~ .

Thus, a motion of electric charges is modified by both the electric and mag-
~ if it
netic forces. If the charge is stationary, the force depends only on E,
moves, there is an additional force proportional to ~v .

52
4.6 Amperes Circuit Law

Amperes circuit law is a useful relation between currents and magnetic fields.
This law allows us to calculate magnetic field produced by certain current
distributions in a very effective way.
The Amperes law says that for an arbi-
trary closed path around a current car-
rying conductor, as shown in Figure 19,
the component of magnetic field tangent
to the path is proportional to the net cur-
rent passing through the surface bounded
by the path
I
~ · d~` = µ0 I ,
B
where the integration is over a closed loop
Figure 19: Contour for evaluation of
the Amperes circuit law.
and I is the total current through the
loop.11

The Amperes law can be applied in highly symmetric situations to find the
magnetic field more easily than by computing with the Biot-Savart law. In
either case, the result is the same. In case that lack the proper symmetry,
the Amperes law is not easily applied. The following example illustrates the
use of Ampere’s law.

Worked Example: An application of the Amperes Law

An infinitely long wire caries a constant current I. Use (a)


Biot-Savart law, (b) Amperes law to find the magnetic field
about the wire.

(a) First, we divide the current into the small current el-
ements Id~`. The element of magnetic field dB ~ due to an
~
element I d` at a point A distance ~r = rr̂ is found from the
Biot-Savart formula:
dB~ = µ0 I d~` × r̂ ,
4π r2
11
The prove of the Amperes law is complicated and will not be presented at the lecture.
The student, if interested in details of the prove is referred to the Appendix A.

53
where we note that all dB ~ are in the same φ̂ direction normal
to the direction of the current. So we see from this symme-
try that the field lines are circles concentric with the current.
Furthermore, along any such circular path the field is con-
stant in magnitude.

Let us calculate the magnitude of the magnetic field. Since


h
d~` × r̂ = dl sin θφ̂ , r= , l = h cot θ ,
sin θ
we have
h
dl = − dθ
sin2 θ
and then
2
~ = µ0 I sin θ − h µ0 I
 
dB 2 sin θdθ φ̂ = − sin θdθφ̂ .
4π h2 sin θ 4πh
Integrating the above equation over the length of the wire,
we obtain
0
~ = − µ0 I µ0 I
Z
B sin θdθφ̂ = (cos 0 − cos π) φ̂
4πh π 4πh
µ0 I
= φ̂ .
2πh

(b) Let us now calculate the field using the Amperes law.

Since the field lines are circles concentric with the current,

54
and along any such circular path the field is constant in mag-
nitude, this is the ideal situation for the application of Am-
peres law:

~ · d~` = µ0 I ⇒ 2πhB = µ0 I ⇒ B = µ0 I .
I
B
2πh

Note how much simpler are the calculations of the magnetic


field using the Amperes law.

4.7 Faraday’s Law of Electromagnetic Induction

Faraday discovered electromagnetic induction by changing magnetic field. If


we consider a closed stationary circuit located in a varying magnetic field,
as shown in Figure 20, the induced electromotive force around this circuit is
equal to the negative time rate of change of the magnetic flux through the
circuit

E =− ,
dt
where Φ is the total magnetic flux through the circuit.12
From the definition of the flux
Z
Φ= B ~ ,
~ · dS
S

and from the fact that the emf force E is


equal to the work done per unit charge,
we have
E I
~ · d~` .
~ + ~v × B)
= (E
q
`

Figure 20: Illustration of a closed cir-


cuit located in a magnetic field.
In a stationary circuit ~v = 0 (and any-
way ~v × B~ ⊥ d~` since ~v k d~`). Thus
12
The electromotive force is induced by varying the flux of the magnetic field that is
proportional to the magnetic field and the area of the loop. Thus, the electromotive force
can also be induced in a constant magnetic field by varying the area S of the loop.

55
~ · d~` = 0.
~v × B

Hence
I
E= ~ · d~`
E
`

and finally

~ · d~` = − ∂
I Z
E ~ · dS
B ~ ,
∂t
` S

or
I Z ~
∂B
~ · d~` = −
E ~ .
· dS
∂t
` S

This is the Faraday’s law written in the integral form.

The student can easily show, using the Stokes’s theorem, that the Faraday’s
law can be written as
~
~ = − ∂B ,
∇×E
∂t
which is called the differential form of the Faraday’s law.

In summary: The Faraday’s law tells us that time-varying magnetic fields


give rise to electric fields. This shows that the fields are related to each other,
and we then must speak of electromagnetic fields, rather than separate
electric and magnetic fields.

56
Revision questions

Question 1. State the Gauss’s Law and explain under what conditions the
Gauss’s law is useful in calculating the electric field produced by a
macroscopic charge.

Question 2. State the Biot-Savart Law and show that magnetic field is produced
by moving charges.

Question 3. State the Ampere’s Law and explain under what conditions the
Ampere’s law can be successfully applied for the calculation of the
magnetic field produced by a macroscopic currents.

Question 4. What is the Faraday’s law of electromagnetic induction and derive


its differential form.

57
Tutorial problems

Problem 4.1 Field of a capacitor

(a) An infinite flat sheet has a uniform charge density of σ [Cm−2 ].


Use Gauss’ Theorem plus an argument from symmetry to find the
electric field strength everywhere.

(b) A capacitor that has plates (area A) of diameter very much


greater than the spacing d between them has charge densities +σ
and −σ on the plates, respectively. Use the result in (a) to find
the total electric field strength due to the charges on the plates,
within and outside the capacitor.

(c) Calculate the force per unit area on one plate due to its at-
traction by charges of opposite sign on the other plate.

Problem 4.2 Properties of a radial vector field

If F~ is the vector rn r̂, (r 6= 0) show that:


∇ × F~ = 0,
∇ · F~ = (n + 2)rn−1 .
Problem 4.3 The Poisson equation

Note from the above tutorial problem 4.2 that the divergence of
the radial field is zero for n = −2. The student may immedi-
ately comment that this result means that the Coulomb field of
this symmetry has no point source. However, we know that the
Coulomb field is produced by a point charge. How then the diver-
gence is zero if there is a point source of the field?

To resolve this dilemma, show that for all r including r = 0



∇· = 4πδ(~r) ,
r2

where δ(~r) is the three-dimensional Dirac delta function.

58
Problem 4.4 Field of a non-uniformly charged sphere

A sphere of radius a has a charge density increasing linearly with


radius from zero at its centre to ρ0 at r = a. Use the Gauss’s law
to find the electrostatic field inside as well as outside the sphere.
Use the results to confirm that
∇·E ~ = ρ and ∇ × E ~ =0,
ε0
everywhere.

Problem 4.5 Field of charged coaxial cylinders

Two infinitely long coaxial cylinders have


radii a and b with b > a, as shown in
Figure 21. The region between them is
filled with charge of volume density given
in cylindrical coordinates as
ρ = A rn ,
where A and n are constants. The charge
density is zero everywhere else. Find the
~ everywhere, i.e. inside,
electric field E
between and outside the cylinders.
Figure 21:

Problem 4.6 Field of a uniformly charged cylinder

An infinitely long cylinder has a circular cross section of radius a.


It is filled with a charge of constant volume density ρ. Find the
electric field of this charge for all points both inside and outside
the cylinder.

Problem 4.7 Charge density for a cylindrically symmetric electric field

~ =
A certain electric field is given in cylindrical coordinates by E
3
E0 (r/a) r̂ for 0 < r < a and E ~ = 0 otherwise. Find the volume
charge density.

59
Problem 4.8 Charge in a uniform magnetic field

(a) A charge q enters a region of uniform magnetic field B~ moving


at right angles to the field. Show that the charge will undergo cir-
cular motion in the field and find an expression for the frequency
of the circular motion.

(b) Describe the motion of the charge if it is not moving at right


angles to the magnetic field.

Problem 4.9 Magnetic field of an infinite current sheet

Consider an infinite plane sheet containing an electric current


which is in the same direction everywhere (the current can be con-
sidered to close at infinity if we want to justify use of closed circuit
theorems). The strength of the current would be described by a
current per unit length (with the ‘length’ drawn at right angles
to the direction of the current). So the units of ‘surface current
density’ would be amps/metre. Note that if the current is entirely
confined to a plane the ordinary current density J in amps/metre2
must be considered as infinite.

(a) Use an argument from symmetry plus Ampères circuital law


to find the magnetic field due to an infinite plane with surface
current density Js amps/metre. It may be helpful to think of the
surface current as the limit of a large number of parallel wires.

(b) Using the result of part (a), find the force per unit area be-
tween two parallel current sheets containing surface current den-
sities I1 and I2 if the currents in the two sheets are both in the
same direction.

60
5 Differential Equations for the EM Field and
Maxwell’s Theory

We know now that electromagnetic forces are carried by electromagnetic


fields that propagate at speed c ' 3 × 108 ms−1 . Because of the finite propa-
gation speed we are forced to assign energy and momentum to the fields i.e.
we must think of them as real physical entities as against mere mathematical
conveniences (as is the case for static fields). An electromagnetic system then
qualifies as static only if all the charges have been at rest longer than the
time taken to traverse the system at speed c.

In the 1830’s Michael Faraday carried out experiments to measure a finite


electromagnetic propagation speed. He was unsuccessful due to lack of time
resolution in his apparatus. Faraday would have had no reason to think that
the electromagnetic speed was the same as the speed of light (then known).
In the 1860’s, James Clerk Maxwell, seeking to advance Faraday’s ideas about
electromagnetic fields, by a brilliant process of intuition worked out how to
generalize certain differential equations deduced from static experiments. He
produced a set of field equations known by his name today. Maxwell also
had a theory i.e. a set of qualitative ideas underpinning his equations. The
theory, unlike the equations, has not stood the test of time.

5.1 Differential Equations for the EM Field

Let us take as the source of the electromagnetic field a continuous macro-


scopic charge and current distribution represented in terms of the macro-
scopic source quantities, charge and current densities
Z Z
q= ρ dV , I= J~ · ~n dS .

Using the Gauss’s and Stokes’s Theorems, we will find differential vector
equations for the fields E~ and B~ from the integral forms of observational
results discussed in the previous chapter.

61
5.1.1 ~
Divergence of E
Consider the Coulomb’s Law that can be written in the form of the Gauss’s
law as
~= q = ρ
I Z
~ · dS
E dV .
S ε0 V ε0

Applying the Gauss’s Theorem, Eq. (1), to the left-hand side of the above
equation, we obtain
Z
~ dV =
Z
ρ
∇·E dV .
V V ε0
Since this relation must hold for arbitrary V , no matter how small, we can
drop the integrals, and obtain

~ = ρ .
∇·E
ε0
This equation is called the differential form of the Coulomb’s (Gauss’s) Law.
From this equation it follows that the divergence of the electric field is zero
in all regions where there is no electric charge.

5.1.2 ~
Curl of E
Consider the Faraday’s flux cutting rule
I Z ~
∂B
~ · d~` = −
E · n̂ dS .
` S ∂t
Applying Stokes’s Theorem to the left-hand side of the above equation, we get
Z Z ~
∂B
~ · n̂ dS = −
∇×E · n̂ dS .
S S ∂t
Since this relation must hold for arbitrary S and n̂, it implies that
~
~ = − ∂B .
∇×E
∂t
This equation is called the differential form of the Faraday’s law.

62
5.1.3 ~
Divergence of B
~ from the Biot-Savart law
We calculate the divergence of B

~ = µ0 I d~` × r̂ .
B
4π r2
By taking ∇· of both sides of the Biot-Savart law and applying the vector
identity
~ × B)
∇ · (A ~ =B
~ ·∇×A
~−A
~·∇×B
~ ,

we obtain
!
~ = µ0 I
∇·B
r̂ r̂
· ∇ × d~` − d~` · ∇ × 2 . (12)
4π r 2 r

However, ∇ × d~` = 0, since d~` is a constant vector with respect to the


differentiation over x, y and z, and

∇× ≡ 0 for any n .
rn
Hence, the right-hand side of Eq. (12) is always zero, so that always

~ =0.
∇·B

This equation, sometimes called Gauss’s law for magnetism, states that the
~ is zero through any closed surface.
flux of the magnetic field B

5.1.4 ~
Curl of B
~ from the Ampere’s circuit law
We find curl of B
I Z
~ · d~` = µ0 I = µ0
B J~ · n̂ dS .
` S

Simply, by applying Stokes’s Theorem to the Ampere’s law, we find that the
following relation
Z Z
~ · n̂ dS = µ0
∇×B J~ · n̂ dS
S S

63
holds for arbitrary S. Thus
~ = µ0 J~ .
∇×B (13)

This equation is called the differential form of the Ampere’s circuit law.

However, Maxwell realized that unlike the previous three differential equa-
tions, this one, Eq. (13), could not be generally true. To see this, take its
divergence and remember that ∇ · ∇ × F~ ≡ 0 for any vector function F~ , so
that
~ = 0 = µ0 ∇ · J~ .
∇·∇×B

Thus

∇ · J~ ≡ 0 ,

which means that the Ampere’s law does not include source currents. This
result is in contradiction with the continuity equation of J~ varying with time.

We have already seen that conservation of electric charge requires


∂ρ
∇ · J~ = − .
∂t
Thus ∇ · J~ ≡ 0 implies that
∂ρ
≡0 i.e. ρ = const.
∂t
Hence, the Ampere law may be applied only for constant currents, i.e. ac-
cepting the Ampere law as general, we could never charge or discharge a
capacitor.

No wonder Maxwell was confused!

5.2 The Maxwell’s Theory


~ as follows. Since
Maxwell guessed the right form of ∇ × B
∂ρ
∇ · J~ + =0
∂t
64
and using for ρ the first Maxwell equation
~ ,
ρ = ε0 ∇ · E

we have
 
~
∂ε0 ∇ · E ~
∂E
∇ · J~ + =0, or ∇ · J~ + ε0 =0.
∂t ∂t

If then we, after Maxwell, write:


 
~
~ = µ0 J~ + ε0 ∂ E  ,
∇×B
∂t

instead of
~ = µ0 J~ ,
∇×B

we obtain
 
~
~ ≡ 0 → ∇ · J~ + ε0 ∂ E  = 0 ,
∇·∇×B
∂t

which is in accordance with conservation and motion of charge.

Note from above that the term that Maxwell added


~
∂E
ε0 ,
∂t
has the dimensions of current density. Maxwell called it the displacement
current density.

Note that the displacement current density does not explicitly involve charges,
they do not appear explicitly in the definition of the displacement current
density. Why then we call it as a ”current”? Maxwell had a theory under-
pinning his equations in which the displacement current was a real physical
current - due to ‘polarization of the electromagnetic ether’. This theory has
not survived. Nevertheless the above term is still referred to as the ‘displace-
ment current’.

65
With the modification of the current density, we then write the fourth Maxwell’s
equation as
~
~ = µ0 J~ + ε0 µ0 ∂ E ,
∇×B
∂t
which contains all the physical processes involved in the generation of mag-
netic fields.

In summary:

The differential form of the Maxwell’s equations is easier to inter-


pret physically and is also useful in deriving the boundary condi-
tions that the field vectors must satisfy.

The Maxwell’s equations are self-consistent and no experimental


evidence for requiring any further modifications has been found.
The equations are used to explain and predict all physical phenom-
ena that involve charges and/or currents.

Exercise in class: Fields within a capacitor

A plane parallel capacitor is being charged with a current I.


Show that the displacement current between the plates of
the capacitor is equal to the conduction current I in the ex-
ternal charging circuit. Remember that the displacement
~
current density is, by definition, ε0 ∂∂tE so the displacement
current through a surface S is:
Z ~
∂E
ID = ε0 · n̂ dS .
∂t
S

Assume that the external wires are perfect conductors so


that E~ is zero in them. Assume the space between the
plates is a perfect insulator so no conduction current flows
within the capacitor.

66
Can you see any curious consequence in this case if the
displacement current is assumed to be a real physical cur-
rent (flow of charges)?

67
Revision questions

Question 1. Derive the differential forms for the Coulomb, Ampere and Faraday
laws.

Question 2. Derive the so-called Gauss’s law for magnetism.

Question 3. Explain, how Maxwell resolved the difficulty with general applica-
tions of the Ampere’s law.

Question 4. What is the displacement current and explain its significance in


the EM theory.

Question 5. Are the Maxwell’s equations independet of each other? Explain.

Tutorial problems

Problem 5.1 Coulomb field

Demonstrate that the Coulomb field for stationary point charge

~ = 1 q
E r̂
4πεo r2
follows from Maxwell’s equations.

Problem 5.2 Magnetic field of an infinite straight wire

Magnetic field at distance r from an infinite straight wire carrying


constant current I in the z direction is given by

~ = µo I φ̂ ,
B
2πr
~ in the xy plane.
where φ̂ is the unit vector in the direction of B
Show, using only the circuit, surface and volume integrations, and
the Gauss’ and Stokes’ Theorems that the magnetic field satisfies
Maxwell’s equations.

68
6 Maxwell’s Equations and Prediction
of Electromagnetic Waves

In the previous lecture, we have derived from the experimental laws vectorial
~
expressions for both the divergence and curl of the two basic field vectors E
~ These expressions put together completely determine the electro-
and B.
magnetic fields and are termed the Maxwell’s equations

~ = ρ
I. ∇·E ,
ε0
II. ~ =0,
∇·B
~
III. ∇×E~ = − ∂B ,
∂t
~
∂E
IV. ~ = µ0 J~ + ε0 µ0
∇×B .
∂t
As we have shown in the previous lecture, the first equation, which we shall
call the Maxwell’s equation number I, is the differential form of the Gauss’s
law, the second equation, which we shall call the Maxwell’s equation num-
ber II, tells us about the non-existence of magnetic charges, the third equa-
tion, which we shall call the Maxwell’s equation number III, is the differen-
tial form of the Faraday’s law, and the final equation, which we shall call
the Maxwell’s equation number IV, is the differential form of the modified
Ampere’s law.

Maxwell’s immediate triumph with the modification of the Ampere’s law


was to predict the existence of electromagnetic waves and their propagation
speed. The calculated speed came (within experimental error) to be equal
to the measured speed of light. This prediction obviously led to the con-
clusion that light was electromagnetic in nature. Thus arose a synthesis of
electromagnetism and optics. In this lecture, we will show how the Maxwell’s
equations predict the existence of the electromagnetic waves and will analyse
their properties when the fields propagate in vacuum. Much of our discus-
sion, however, will be about ‘how to solve the Maxwell’s equations’ for fields
propagating in vacuum.

69
6.1 The Wave Equation for EM Waves in Vacuum
In a vacuum there are no sources, i.e. ρ = 0 and J~ = 0. Hence, the Maxwell’s
equations reduce to the following differential equations
~ = 0,
∇·E (14)
~ = 0,
∇·B (15)
~
~ = − ∂B ,
∇×E (16)
∂t
~
~ = ε0 µ 0 ∂ E .
∇×B (17)
∂t
These equations are readily to solve. The procedure is to eliminate ~ or B
E ~
between equations (16) and (17) to obtain differential equations for ~ or B
E ~
alone, using where required Eqs. (14) and (15).

Method:

Think of ∇× and ∂/∂t as linear (differential) operators. By analogy with


methods of solving linear algebraic equations, applying ∇× into (16) and
∂/∂t into (17), we obtain
  ∂∇ × B ~
~
∇× ∇×E = − ,
∂t
2~
∂ ~ = ε0 µ 0 ∂ E .
∇×B
∂t ∂t2
Hence
  ~
∂ 2E
~
∇× ∇×E = −ε0 µ0 ,
∂t2
and using a vector identity
 
~ = ∇∇ · E
∇× ∇×E ~ − ∇2 E
~ ,

we obtain
2~
~ = −ε0 µ0 ∂ E .
~ − ∇2 E
∇∇ · E
∂t2
70
~ = 0 in the vacuum, it follows at once that
Because ∇ · E
~
1 ∂ 2E
~ =
∇2 E , (18)
c2 ∂t2
where
1
c2 = .
ε0 µ 0
The parameter c has the dimensions of velocity and is numerically equal
to 3 × 108 ms−1 .

Equation (18) is the standard form of a three-dimensional vector wave equa-


tion. Following the same procedure, we can easily show that the magnetic
~ satisfies the same equation. The proof is left to the student.
field B

6.2 Plane Wave Solution to the Wave Equation


The wave equation in a vacuum is
~
1 ∂ 2X
~ =
∇2 X
c2 ∂t2
~ ≡ E,
for X ~ B.
~

We will solve the wave equation in one dimension assuming that the wave
propagates in the z direction.
In this case, ∂/∂x = ∂/∂y ≡ 0, and then

∂2 ∂2 ∂2 ∂2
∇2 = + + = .
∂x2 ∂y 2 ∂z 2 ∂z 2
~ and B
The differential equations for E ~ both have the same form:

~
∂ 2X ~
1 ∂ 2X
= .
∂z 2 c2 ∂t2
Such an equation has solutions of the form

X = f (z ± ct) , (19)

71
where f is an arbitrary function with the argument z ± ct.

This solution represents a signal propagating with speed c as can be seen


from the following discussion:

Let X0 = f (z0 − ct0 ) i.e. it represents the solution X at t = t0 and z = z0 .


Now examine X at time ∆t later and distance ∆z further along in z. Since
a harmonic wave does not change in vacuum, we have

X1 = f (z0 + ∆z − c(t0 + ∆t))


= f (z0 − ct0 ) = X0 when ∆z = c∆t ,

i.e. the signal propagates a distance ∆z = c∆t in time ∆t i.e. it propagates


with speed c.

Proof of solution (19):

Let f represent f (z − ct). Then

∂f ∂f ∂(z − ct) ∂f
= = = f0 ,
∂z ∂(z − ct) ∂z ∂(z − ct)

where ”0 ” means differentiation wrt z − ct. Similarly

∂ 2f ∂f 0 ∂f 0 ∂(z − ct)
= = = f 00 ,
∂z 2 ∂z ∂(z − ct) ∂z

and
∂f ∂f ∂(z − ct)
= = −cf 0 .
∂t ∂(z − ct) ∂t

Next
∂ 2f
= (−c)(−c)f 00 ,
∂t2
and consequently

∂ 2f 1 ∂ 2f
= .
∂z 2 c2 ∂t2

72
Thus
1 1
c2 = →c= √ ,
ε0 µ 0 ε0 µ 0

which with the numerical values of the parameters

ε0 ' 8.85 × 10−12 Fm−1 , µ0 = 4π × 10−7 Hm−1

gives

c ' 3 × 108 ms−1 .

The fact that the numerical value of c is equal to the velocity of light in
vacuum led Maxwell to propose an electromagnetic theory of light, one of
the brilliant contributions to physics in the nineteenth century.

6.3 Harmonic Waves


In vacuum, we chose a plane wave representation for the propagating EM
wave. For a harmonic wave
ω 2π
c = νλ = , ω = 2πν, k= ,
k λ
where ν = frequency (Hz), λ = wavelength (m), ω = radian frequency (ra-
dians s−1 ), and k = propagation constant (m−1 ).
ω
f (z − ct) = f (z − t) = f1 (ωt − kz) .
k
A plane wave is represented by the electric field
~ =E
E ~ 0 ei(ωt−kz) , (20)

whose the propagation is characterized by the frequency ω and the wave


vector ~k. Since the phase of the wave described by Eq. (20) is by definition
the argument of the exponent, we see that the surfaces of the constant phase
are planes whose normal is the z-axis:
z
 
i(ωt − kz) = iω t −
c

73
given at any definite time by z = const. These surfaces of constant phase
travel with a constant velocity c often referred as the phase velocity of the
wave.13

Therefore, the waves described by Eq. (20) are called plane waves.

6.4 The Transverse Nature of Plane Waves in Vacuum

We now investigate the relations between the amplitudes and phases of the
electric and magnetic fields in a plane harmonic wave. While it is true that
the magnetic field satisfies the same wave equation as the electric field, it
is not independent of the latter, since one must satisfy the Maxwell equa-
tions III and IV.

~ = 0 always in electromagnetism, and


Since ∇ · B

~ = ∂Bx ∂By ∂Bz ∂Bz


∇·B + + =0+0+
∂x ∂y ∂z ∂z
for a plane wave propagating in the z direction, we have
∂Bz
=0.
∂z
However, for a plane wave
∂Bz
= −ikBz .
∂z
Hence, the rhs must be zero, which means that either k = 0 (zero frequency)
or Bz = 0 (transverse wave).
~
For a propagating wave k 6= 0, so the wave is transverse in B.

In a vacuum ∇ · E ~ = 0 and then by the same argument we conclude that


~ In other cases in electromagnetism
the plane wave is also transverse in E.
13
Note that some textbooks on electromagnetism, for engineers in particular, use the
letter j instead of i for the imaginary number. We will use i throughout this lecture notes.

74
(e.g. for a plasma in a magnetic field) a plane wave may not be purely trans-
~
verse in E.

~ ⊥B
In addition: E ~ for a plane EM wave in a vacuum.

Proof:

Consider a harmonic wave propagating parallel to the z axis. In this case


the field components are of the forms
~ =E
E ~ 0 ei(ωt−kz) , ~ =B
B ~ 0 ei(ωt−kz) .

In order to prove the statement, we will use the Maxwell’s equation III
~
~ = − ∂B ,
∇×E
∂t
and expand it in Cartesian coordinates remembering that ∂/∂t ≡ iω

î ĵ k̂

∂ ∂ ∂
∂x ∂y ∂z

= −iω(Bx î + By ĵ + Bz k̂) . (21)

E Ey Ez
x

For electromagnetic plane waves propagating along the z axis in a vacuum:


∂ ∂ ∂
Ez = Bz = 0, = = 0, = −ik .
∂x ∂y ∂z
Hence, Eq. (21) reduces to


î ĵ k̂
0 0 −ik = −iω(Bx î + By ĵ) .



Ex Ey 0

Expanding the determinant, and comparing the left and right-hand sides,
we find
k
x component : ikEy = −iωBx ⇒ Bx = − Ey ,
ω
k
y component : −ikEx = −iωBy ⇒ By = Ex .
ω
75
~ and B?
What does it tell us about the relation between E ~

~ · B,
If we consider a scalar product E ~ and use the above result, we find

~ ·B
E ~ = (îEx + ĵEy ) · (îBx + ĵBy )
−k k
= (îEx + ĵEy ) · (î Ey + ĵ Ex )
ω ω
k k
= − Ex Ey + Ex Ey = 0 .
ω ω
~ ⊥ B.
This means that E ~

In addition, note that


E ω
= =c.
B k
Thus, we may conclude that in the electromagnetic theory when E and B
are related, their ratio is always a velocity characteristic of the problem in
hand.

This is as far as Maxwell took the subject. It was for others like Heinrich
Hertz 1884 to show how to solve Maxwell’s equations with source terms ρ, J~
included (i.e. the generation of electromagnetic waves). We will consider
this later.

You should be aware that we have not derived Maxwell’s equations from the
static limits like Coulomb’s Law and the Biot-Savart Law. The solutions to
Maxwell’s equations include the static limits as special cases but many more.
Maxwell’s equations have the status of postulates suggested by experimental
results.

In summary, we have the following important results for related electric


and magnetic fields propagating in vacuum:
1. The electric and magnetic fields propagate in a form of plane waves,
so-called electromagnetic (EM) waves.

2. The plane EM wave is transverse in E ~ and B,~ i.e. both fields are
perpendicular to the direction of propagation.

76
3. The electric and magnetic fields are perpendicular to each other.

4. The ratio E/B is constant and equal to the velocity of the wave, that
is equal to the speed of light in vacuum.

Exercise in class: Electric and magnetic fields of a plane wave

For a given electric field polarized in x direction and prop-


agating in z direction
~ = E0 sin (ωt − kz) î ,
E

calculate B~ from III and show that the fields represent


a plane electromagnetic wave propagating in vacuum, i.e.
they satisfy the Maxwell’s equations (14)−(17).

77
Revision questions

Question 1. Show that the Maxwell’s equations for the EM fields propagating
in vacuum can be reduced to two independent second order differ-
ential equations, the so-called the wave equations.

Question 2. State the properties of an EM wave propagating in vacuum.

Question 3. How do we define phase velocity of an EM wave?

Question 4. The wave equations for the electric and magnetic fields propagating
in vacuum are independent of each other. How then the electric
and magnetic fields are related to each other? Explain.

Tutorial problems

Problem 6.1 Fields of a plane wave

A plane electromagnetic wave propagates in the z direction in


a vacuum. Its electric field is polarized in the x direction and so
has the form:
~ = E0 sin(ωt − kz)î
E

(a) If the wave amplitude is 10 volts m−1 and its frequency is 1


MHz, find numerical values for E0 , ω, k and the wavelength.

~ directly from the above expression for E.


(b) Calculate ∇ × E ~
~ ~
Plot E and ∇ × E as functions of z for t = 0 (on the same plot).

(c) Hence calculate the magnetic field of the wave using the Maxwell
equation:

~
∂B ~ .
= −∇ × E
∂t

78
~ varies in phase with E.
(i) Show that B ~

(ii) Calculate the ratio E/B.

(d) Show that in the stated circumstances, the remaining Maxwell’s


equations should be:

~
~ = 0,
∇·E ~ = 0,
∇·B and ~ = ε0 µ0 ∂ E .
∇×B
∂t

~ and B
(e) Write expressions for E ~ of a right-circularly polarized
plane wave having the same amplitude and frequency as the above
one.

Problem 6.2 EM wave in free space

An EM wave traveling in free space (ρ = 0, J~ = 0) in the pos-


itive z direction is described by

~ = E0 cos (ωt − kz) î,


E
~ = B0 cos (ωt − kz) ĵ,
B

where k, E0 , and B0 are constants. Under what circumstances do


~ and B
these E ~ fields satisfy all of Maxwell’s equations?

79
7 EM Theory and Einstein’s Special Theory
of Relativity
The purpose of this lecture is to test the laws of electromagnetism against
the principle of special relativity. We will demonstrate how one might infer
the law of Biot-Savart and in general relate the magnetic field to electric
field from application of special relativity to Coulomb law. Special relativ-
ity, formulated in 1905, grew out of Einstein’s meditation on electromagnetic
theory and the properties of space and time. Historically, the insights of
Einstein’s theory follow after electromagnetism. Logically however, special
relativity contains more general statements about nature than electromag-
netism. Electromagnetic field theory is just one of a possible set of field
theories that are compatible with the Einstein theory of space and time. It
is evident that relativistic effects are important if we would have to calculate
the field of a charge moving with a speed comparable to that of light. What
is not so obvious is that special relativity offers insights into aspects of elec-
tromagnetic theory even in the case of the low speed charges we consider in
this course.

Two such aspects are:

(1) The unity of the electromagnetic field i.e. the field is a single entity
with 6 components (represented by two vectors E ~ and B,~ each with three
components).

(2) Understanding the nature of causal relationships in electromagnetic the-


ory, e.g.
~
∂B
~ =−
∇×E .
∂t
~ causes a spatially changing E?
Does this mean that a time changing B ~

80
7.1 Lorentz Transformation Equations for Space and
Time
We will analyse how the fundamental laws of electromagnetism change when
charges move with a constant velocity. In particular, we will show that the
Maxwell’s equations are derived from the Coulomb’s law. Thus, the student
will learnt that the whole electromagnetism follows naturally from electro-
statics. In our calculations, we will follow the principle of special relativity.

The principle of special relativity

1. The laws of physics are the same in all inertial reference frames.

2. The speed of light in vacuum is independent of the uniform motion of the


observer or source.

The constancy of the velocity of light, in-


dependent of the motion of the source,
gives rise to the relations between space
and time coordinates in different inertial
reference frames known as Lorentz trans-
formations.

Consider a stationary reference frame S


and a inertial frame S 0 moving with a ve-
locity ~v parallel to the x axis, as shown in
Figure 22. Let x, y, z are coordinates in
the S frame and x0 , y 0 , z 0 are the coordi-
nates in the S 0 frame. The time and space
Figure 22: Two inertial reference coordinates in S 0 are related to those in S
frames in relative motion. A stationary
frame S and a frame S 0 moving with a by the 1D Lorentz transformations:
velocity ~v parallel to the x axis.
x0 = γ(x − vt) ,
y0 = y ,
z0 = z ,
vx
 
0
t = γ t− 2 ,
c

81
where
!− 1
v2 2
γ = 1− 2
c
is the Lorentz factor.

The above transformation corresponds to a situation of ~v parallel to the x


axis. Later in the course, Chap. 21, we will consider the general case of the
velocity ~v of the frame S 0 in an arbitrary direction.

7.2 Force Transformation Equations


Let us investigate the force between two charges as viewed by observers in S
and S 0 . We will evaluate the force from the point of view of the observer in S 0 .

Consider a particle which is moving with velocity ~u = ux î + uy ĵ + uz k̂ in


the S frame and is acted on by a force with components Fx , Fy and Fz . Then,
according to the special theory of relativity, the force in the S 0 frame is
vuy vuz
Fx0 = Fx − 2 vux Fy − 2 Fz ,
c (1 − c2 ) c (1 − vu
c2
x
)
2 1
(1 − uc2 ) 2 Fy
Fy0 = vux Fy = ,
1 − c2 γ(1 − vu
c2
x
)
2 1
(1 − vc2 ) 2 Fz
Fz0 = vux Fz = .
1 − c2 γ(1 − vu
c2
x
)
Suppose Fx , Fy , Fz represents the velocity independent Coulomb force. Then
in the S 0 frame (source of the field now moving) the force is no longer veloc-
ity independent. Thus, the observers disagree about the magnitude of the
force. They, in fact, disagree not only about the magnitude but also about
the origin of the force. In electromagnetic theory we say that there is now a
magnetic force and we define a magnetic field B ~ that determines the mag-
netic force.

We now present a detailed calculation that illustrates how the form of the
Maxwell’s equations is determined by nature obeying Einstein’s special the-
ory of relativity.

82
7.2.1 The Force between Two Charges Moving with Constant Ve-
locities

Invariance of electric charge

In ordinary matter, electrons move with much greater speeds than protons
yet there is no associated electric field. This implies that electric charge is
independent of velocity unless electromagnetic laws modified in some compli-
cated way (see discussion by King in Physical Review Letters 5, 562 (1960)).

Consider charges q1 and q2 moving


with velocities ~u and ~v in an iner-
tial frame S. No loss of generality
occurs if ~v is taken parallel to the x
axis.

Now consider a frame S 0 moving


with velocity ~v along x axis, that q2
is stationary in S 0 . Assume that at
time t = 0, the frames S and S 0
overlap.
Figure 23: Two moving charges seen in two
different reference frames. From the principle of relativity, in
the S 0 frame Coulombs law holds.
The force on q1 seen in S 0 is therefore
1 q 1 q2 0
F~ 0 = ~r .
4πε0 r0 3
We shall transform this expression to find the force observed in the S frame
in which the source of the field (q2 ) is moving.

We shall see that what we normally call the MAGNETIC FIELD arises
as a natural consequence of relativistic invariance with no extra assumptions.

We start with the x component of the force, that is


1 q1 q2 0
Fx0 = x ,
4πε0 r0 3

83
and similarly for y and z.

We need transformations of x0 to x and r0 to r. The transformation of x0 to x


is found from the Lorentz transformations that at t = 0 are

x0 = γx ,
y0 = y ,
z0 = z ,
vx
t0 = −γ 2 .
c
To do the complete transformation of the force component, we also need
transformation from r0 to r. It can be found as follows. It is seen from
Figure 23 that there is an “axial symmetry”, so that the transformation
could be expressed in terms of the angle θ. Thus, we can write

r02 = x02 + y 02 + z 02 = γ 2 x2 + y 2 + z 2
y2 + z2 v2  2
! " ! #

2 2 2 2 2
= γ x + =γ x + 1− 2 y +z
γ2 c
v2  2
" #

2 2 2 2 2
= γ x +y +z − 2 y +z
c
2
v2
! !
2 2 v 2 2 2 2 2
= γ r − 2 r sin θ = γ r 1 − 2 sin θ ,
c c
q
where sin θ = (y 2 + z 2 )/r. Hence
!1
0 v2 2
r = γr 1 − 2 sin2 θ .
c

Thus, substituting the transformations into Fx0 , we obtain


1 q1 q2 γx
Fx0 = 3 = q1 gx ,
4πε0 
v2
γ 3 r3 1 − c2
sin2 θ 2

where
1 q2
g= 3 .
4πε0 
v2 2 2
γ 2 r3 1− c2
sin θ

84
Similarly for the remaining two components, we find
q1 gy q1 gz
Fy0 = , Fz0 = .
γ γ
In summary, the force transformations are

a) x component
vuy vuz
Fx0 = Fx − 
vux
 Fy − 
vux
 Fz .
c2 1 − c2
c2 1 − c2

Thus
vuy vuz
q1 gx = Fx − 
vux
 Fy − 
vux
 Fz .
c2 1 − c2
c2 1 − c2

b) y component
Fy
Fy0 = 
vux
 ,
γ 1− c2

from which we find


vux vux q1 gy
   
Fy = γ 1 − 2
Fy0 = γ 1 − 2 ,
c c γ
which we can write as
vux
 
Fy = q1 gy 1 − 2 .
c
c) z component
vux
 
Fz = q1 gz 1 − 2 .
c
Substituting for Fy,z in x equation
vuy vuz
Fx = q1 gx + q1 gy 2
+ q1 gz 2 .
c c
Note: Here is the germ of the magnetic field. The last two terms are typical
second order relativistic terms ≈ v 2 /c2 . In a nonrelativistic calculation we

85
would have Fx = Fx0 .

We now combine results a), b) and c) into a single vector equation for the
force F~ . First, note that
vux = ~u · ~v .
Next, we can write the x component as
vux vux vuy vuz
 
Fx = q1 gx 1 − 2 + q1 gx 2 + q1 gy 2 + q1 gz 2
c c c c
vux v

= q1 g 1 − 2 x + q1 g 2 (~u · ~r)
c c
and with the y and z components
vux
 
Fy = q1 g 1 − 2 y ,
c 
vux

Fz = q1 g 1 − 2 z ,
c
these three components combine into
!
~v · ~u ~v
F~ = q1 g 1 − 2 ~r + q1 g 2 (~u · ~r)
c c
q1 g
= q1 g~r + 2 [~v (~u · ~r) − ~r (~u · ~v )] ,
c
which can be written as
q1 g
F~ = q1 g~r + 2 ~u × (~v × ~r) .
c
We can write this equation in the form of the Lorentz force
 
F~ = q1 E
~ + ~u × B
~ ,
where
~ = g~r = 1 q2
E 3 ~r ,
4πε0 
v2 2 2
γ 2 r3 1− c2
sin θ

and
~
~ = ~v × g~r = ~v × E .
B (22)
c2 c2
86
Thus, a force which we see in the moving frame S 0 as the Coulomb force
appears as the Lorentz force in the stationary frame S.

The relation (22) shows that B ~ and E ~ have no independent existence. A


pure electric field in one coordinate frame appears as a mixture of electric
and magnetic fields in another coordinate frame.

Note that as v → 0, γ → 1, and then

~ → 1 q2
E ~r .
4πε0 r3
Moreover, the ration of magnitudes of magnetic to electric term in the force
equation is uv/c2 , i.e. magnetic forces are second order relativistic effects.

7.3 Electric Field Lines of a Moving Charge


Let β = v/c. Then, we can write the electric field as

~ = 1 q(1 − β 2 )
E 3 r̂ .
4πε0 
r2 1 − β 2 sin2 θ 2

For a given θ, the electric field E


still varies as 1/r2 , but the field
lines are crowded in the direction per-
pendicular to ~v , as shown in Fig-
ure 24.

In the forward direction θ = 0, and


then
1 q(1 − β 2 )
E= < Es ,
4πε0 r2
Figure 24: Electric field lines of the sta- where Es is the electric field of the sta-
tionary charge and the charge moving with
uniform velocity ~v . tionary charge, i.e. the static electric
field (at v = 0). Thus, the field ampli-
tude is lower in the direction of motion
relative to the static field amplitude.

87
In the perpendicular direction θ = π/2, and then
1 q   1
2 −2
E= 1 − β > Es .
4πε0 r2
Thus, the field amplitude is larger in the transverse directions relative to the
to the static field amplitude.14

In summary, the electric field lines radiate from the present position of the
charge and are crowded in the direction perpendicular to the direction of
motion of the charge.

7.4 Magnetic Field Lines of a Moving Charge


From the relation between electric and magnetic fields, we find
~ q(1 − β 2 ) ~v × ~r
~ = ~v × E = 1
B 3 .
c2 4πε0 
2 2
c2 r 3 1− β2 sin θ

In order to find the direction of the


magnetic field, we refer to the spheri-
cal polar coordinates, and find that
~ = Br r̂ + Bθ θ̂ + Bφ φ̂ = Bφ φ̂ ,
B

~ ∼ ~v × ~r, and r̂, θ̂, φ̂ are unit


since B
vectors.

Thus, the magnetic field lines form


concentric rings about ~v , and there is
symmetry about the plane θ = π/2. In
Figure 25: Magnetic field lines of the
charge moving with uniform velocity ~v .
the non-relativistic case of v  1 the
factor β → 0, and then the magnetic
field reduces to

~ = 1 q ~v × ~r
B ,
4πε0 c2 r3
14
An interesting question: What is the electric field distribution of a charge moving with
velocity v = c?

88
which is the Biot-Savart law. Applied to a continuous line current I:
~ × ~r
1 I dl
~ =
B .
4πε0 c2 r3
The constant 1/(ε0 c2 ) is normally written µ0 - the magnetic permeability of
free space.

7.5 Invariance of the Maxwell equations under the Lorentz


transformation
According to the principles of relativity, the laws of physics are invariant un-
der the Lorentz transformation. In this lecture, we will show that the basic
equations for the electromagnetic theory, the Maxwell’s equations, are invari-
ant under the Lorentz transformation, i.e. they represent laws of physics.

(1) Equation for the total electric flux

Consider the total electric flux through a surface S closing a moving charge q:
I
ΨE = ~ · n̂ dS .
E
S

We will use the axial symmetry and


break sphere up into rings, as shown
in Figure 26, lying between θ and θ +
dθ. The radius of a ring seen
from the center under the angle θ
is r sin θ, and the width of the ring
is rdθ.

~ has a radial symmetry, E


Since E ~ k n̂
everywhere, and then the flux dΨE is
~ · n̂ dS = EdS .
dΨE = E (23)
Figure 26: Evaluation of the total electric
flux through a surface S closing a moving However, the area of the ring (stripe)
charge. between θ and θ + dθ is equal to
dS = 2π(r sin θ) rdθ = 2πr2 sin θdθ .

89
Hence, substituting this result into Eq. (23) and the explicit form of E, we
find that the flux through the ring is

1 q(1 − β 2 )
dΨE = 3 2πr2 sin θdθ
4πε0 
2 2
r2 1− β2 sin θ
q(1 − β 2 ) sin θdθ
= 3 .
2ε0 
2 2
1− β2 sin θ

This gives the total flux through the surface S:


π
Z
q(1 − β 2 ) Z sin θdθ
ΨE = dΨE = 3 .
2ε0 
θ=0 1 − β 2 sin2 θ 2

To calculate the integral, put cos θ = x, so that sin θdθ = −dx, and then
Z
dx 1 Z dx
I = − =− 3  2
(1 − β 2 + β 2 x2 )3/2
3/2
β 1−β
+ x2 β2
1 Z
dx
= − ,
β3 (a2 + x2 )3/2

where a = 1 − β 2 /β. Performing the integration, we obtain
1 x
I=− ,
a2 β 3 (a2 + x2 )1/2

and finally including the limits of the integral, we get


Z−1
dx
I = −
1
+ β 2 x2 )3/2
(1 − β2
1 −1 1 1 2
= − 2 3√ 2 + 2 3√ 2 = .
a β a +1 a β a +1 1 − β2
Thus, the total flux through the surface is

q(1 − β 2 ) 2 q
ΨE = 2
= ,
2ε0 1−β ε0

90
which is the same as for a stationary charge.

Hence, the electric field produced by a moving charge satisfies the Gauss’s law.
If we apply the Gauss’ theorem to ΨE , we get

∇·E ~ = ρ .
ε0
Thus, we conclude that the Gauss’ law (Maxwell’s equation I) is invariant
under the Lorentz transformation.

(2) Magnetic flux through a closed surface

Consider the magnetic flux through a closed surface


I
ΨM = ~ · n̂ dS .
B

~ is
If we choose a sphere centered on q to calculate the integral, we find that B
perpendicular to n̂ everywhere on the sphere, i.e. is tangential to the surface
of the sphere. Thus, the flux through the sphere ΨM = 0.

~ = 0 always. It is easy to see that


Similarly, we can show that ∇ · B

~ = 1 
~

∇·B ∇ · ~
v × E
c2
1 ~ 
~ =0,
= 2 E · ∇ × ~v − ~v · ∇ × E
c
~
since ∇ × ~v = 0 as ~v is constant, and ~v is perpendicular to ∇ × E.

Thus, we conclude that the Maxwell’s equation II is invariant under the


Lorentz transformation.

~ derivative related to temporal B


(3) Spatial E ~ derivative.

We shall show that for a point charge under uniform velocity


~
~ = − ∂B .
∇×E
∂t
91
~ in spherical polar coordinates
To do it, we write ∇ × E
" #
~ = r̂ ∂ (sin θEφ ) ∂Eθ
∇×E −
r sin θ ∂θ ∂φ
" #
θ̂ 1 ∂Er ∂ (rEφ )
+ −
r sin θ ∂φ ∂r
" #
φ̂ ∂ (rEθ ) ∂Er
+ − .
r ∂r ∂θ

Since the electric field has a radial symmetry, we have

Eθ = Eφ = 0 ,

and because the electric field amplitude dependes only on r and θ, we get

~ =− 1 ∂Er
∇×E φ̂ ,
r ∂θ
where
1 q(1 − β 2 ) K
Er = 3 = 3 .
4πε0 
2 2 2 2
r2 1− β2 sin θ 1− β2 sin θ

Calculate ∂Er /∂θ:

∂Er 3 − 5 
   
= − K 1 − β 2 sin2 θ 2 −2β 2 sin θ cos θ
∂θ 2
3Kβ 2 sin 2θ
=  5 .
2 1 − β 2 sin2 θ 2

Hence
2
~ = − 1 3q(1 − β )  β 2 sin 2θ
∇×E 5 φ̂ .
4πε0 2r3
1 − β 2 sin2 θ 2

~
To calculate ∂ B/∂t, we use the theorem of partial derivatives. If

y = f (x, t)

92
then from the maximum change of y
∂y ∂y
dy = dx + dt = 0 ,
∂x ∂t
we obtain
∂y ∂y ∂x
=− .
∂t ∂x ∂t
Thus
~
∂B ~
∂B
= −v .
∂t ∂x
Alternatively, to see this “physically”, remember that the field pattern moves
with constant velocity ~v . Let a stationary observer measure the change in the
~ in a time interval dt. This change is the same as he would observed
field B
at a fixed time by moving a distance dx = −vdt, i.e.
~
dB ~
in dt ≡ dB in dx = −vdt .

Hence
~
∂B ~
∂B
=− ,
∂x v∂t
and then
~
∂B ~
∂B
= −v .
∂t ∂x
Now

~ = 1 q(1 − β 2 ) v sin θ
B 3 φ̂
4πε0 
2 2
c2 r 2 1− β2 sin θ

and
~
∂B ∂Bφ ∂Bφ ∂r
= φ̂ = φ̂ .
∂x ∂x ∂r ∂x
Since
√ 2
y + z2 a
sin θ = = ,
r r
93
we can write Bφ as
Ka Ka
Bφ = 2 3/2
= ,
(r2 − β 2 a2 )3/2
 
β2a
r3 1− r2

where
1 q(1 − β 2 )v q
K= and a = y2 + z2 .
4πε0 c2
Next
1/2
∂r ∂ (x2 + y 2 + z 2 ) 1 2 −1/2 x
= = x + y2 + z2 2x = = cos θ ,
∂x ∂x 2 r
∂Bφ 3  2 3Ka r
  −5/2
= Ka − r − β 2 a2 2r = − .
∂r 2 (r2 − β 2 a2 )5/2
Hence
∂Bφ 1 3q(1 − β 2 ) vr2 sin θ cos θ
= − 5
∂x 4πε0 
c2 r5 1 − β 2 sin2 θ 2
1 3q(1 − β 2 ) v sin 2θ
= − 5 ,
4πε0 
2 2
2c2 r3 1− β2 sin θ
and then
~
∂B ∂Bφ 1 3q(1 − β 2 ) v 2 sin 2θ
= −v φ̂ =  5 φ̂
∂t ∂x 4πε0 2 3  2
2c r 1 − β sin θ 2
2

1 3q(1 − β 2 )β 2 sin 2θ
=  5 φ̂ .
4πε0 
2
2r 1 − β sin θ 2
3 2

~ we see that
Comparing with ∇ × E,
~
~ = − ∂B .
∇×E
∂t
Thus, we conclude that the Faraday’s Law (Maxwell’s equation III) is invari-
ant under the Lorentz transformation.

94
7.5.1 Remark on the Electromagnetic Induction
It has been known since about 1831 when Faraday first waved a magnet
near an electric circuit and played with transformers that when the magnetic
flux through a circuit changes, an electromotive force (emf) E appears in it.
Faraday gave the rule
∂ΦM
E =− .
∂t
The nature of this “phenomenon” is, however often misinterpreted.

Consider a point charge q moving


near a closed circuit as shown in
Figure 27. Because of the θ de-
pendence of E, the electric field
on side (1) is larger than that
on side (2). Thus, there is a
net driving force round the cir-
cuit.

Calculate the resulting electromotive


force in the circuit, which is equal to
the work done on a charge q 0 in the
Figure 27: Contour for evaluation of an circuit
electromotive force.
Z ~
F~ · dl
E = Wq 0 =
q0
Z
= ~ + ~u × B)
(E ~ ,
~ · dl

where u is the velocity of the charge q 0 in the circuit.

~ is perpendicular to both ~u and B,


Now, since ~u × B ~ it is also perpendicular
~ and then
to dl,
Z  
~ =0.
~ · dl
~u × B

Hence
Z Z Z
E= E ~ =
~ · dl ~ · n̂ dS =
∇×E ~ .
~ · dS
∇×E

95
If we now take the conclusion that where is a spatially varying electric field
there is also a time varying magnetic field, we obtain
~
∂B ~ =−∂ B ~ = − ∂ΨM .
Z Z
E =− · dS ~ · dS
∂t ∂t ∂t
However, it is obvious in this example that the changing magnetic flux is not
the CAUSE of the emf. The changing magnetic field and the electric field
have a common CAUSE through the charge q.

We can conclude that: Electric and magnetic fields do not produce


each other - they are both due to electric charges.

It is often thought in the textbooks, however that e.g. in a transformer the


changing ΨM produces E. It happens because the flux cutting rule is an
extremely powerful one for calculating the integrated electric field of electric
currents.

The Faraday’s rule E = −∂ΦM /∂t, which arises from ∇ × E ~ = −∂ B/∂t


~
should not be thought of as a casual relationship. What it means is that if
a charge moving with a constant velocity produces a time varying magnetic
field then that charge also produces a spatially varying electric field.

~ and temporal variation


(4) Relation between spatial variation of B
~
of E.

Since
~
~ = ~v × E ,
B
c2
and
     
~
∇ × ~v × E = ~ · ∇ ~v − (~v · ∇) E
E ~ + ~v ∇ · E
~ −E
~ (∇ · ~v ) ,

with ~v constant, we obtain


 
∇ · ~v = 0 , ~ · ∇ ~v = 0 .
E

96
Then
~ = 1 ∇ × ~v × E
~ = 1 − (~v · ∇) E
  n  o
∇×B ~ + ~v ∇ · E
~ .
c2 c2
Thus,
~ ~
!
~ = vx ∂ + vy ∂ + vz ∂ E
(~v · ∇) E ~ = v ∂E = − ∂E
∂x ∂y ∂z ∂x ∂t
as
∂ ∂
vx = v vy = vz = 0 and = −v .
∂t ∂x
Hence
 
~
~ = 1 ~v ∇ · E
∇×B ~ + ∂E  .
c2 ∂t
If we recognize that
~ = ρ
∇·E
ε0
and next that ~v ρ = J~ and 1/c2 ε0 = µ0 , we get the Maxwell’s equation IV.

Thus, we may conclude that the Maxwell’s equation IV is invariant under


the Lorentz transformation.

In summary:

Maxwell’s equations for a point charge moving with uniform velocity are
∇·E~ = ρ ,
ε0
∇·B~ = 0,
~
∇×E ~ = − ∂B ,
∂t
~
∇×B ~ = µ0 J~ + 1 ∂ E .
c2 ∂t
97
These equations arise from the necessity for the correct relativistic trans-
formations between frames in uniform relative motion. This shows that
the Maxwell’s equations are invariant under the Lorentz transfor-
mation. If the postulates of relativity are correct and Coulomb’s law gives
the field of a stationary charge, these equations follow, and the force on a
charge is
 
F~ = q E
~ + ~v × B
~ .

Revision questions

Question 1. State the principle of special relativity.

Question 2. Is the Coulomb’s law for moving charges equivalent to the Lorentz
force?

Question 3. Are the electric and magnetic fields invariant under the Lorentz
transformation? Explain.

Question 4. Are the Maxwell’s equations invariant under the Lorentz transfor-
mation?

Question 5. Why the Faraday’s law is often misinterpreted? What is the correct
interpretation of the the Faraday’s law?

98
8 Energy in the Electromagnetic Field:
Poyntings’ Theorem

Energy may be transported through space by means of electromagnetic waves.


We expect energy flow in the direction of propagation of the wave, E ~ × B,
~ as
illustrated in Figure 28. In this lecture, we will derive an expression for the
rate of the energy flow with a traveling EM wave. We will use the Maxwell
equations to derive this expression, which at the same time will show that
the Maxwell equations are consistent with the law of conservation of energy.

We know from the circuit theory that


the power (energy) flow is related to the
product of voltage and current

P =VI .

We will show that in the field theory,


where everything is expressed in terms
of the electric and magnetic field ampli-
tudes, the power flow across an element
~
Figure 28: Direction of the E~ × B~ vec- of area dS is given by
tor of the field components of a plane ~ ×B
~ · dS
~ ,
wave. c2 ε 0 E

where S is a closed surface bounding a volume V containing a source of


an EM field.

~ B,
If the field propagates in the direction determined by the cross product E× ~
consider the source of the field, i.e. consider the expression
~ × B)
∇ · (E ~ . (24)

If we employ the vector identity


~ × B)
∇ · (E ~ =B
~ ·∇×E
~ −E
~ ·∇×B
~

and use the Maxwell equations III and IV:


~ ~
~ = − ∂B
∇×E and ~ = µ0 J~ + ε0 µ0 ∂ E ,
∇×B
∂t ∂t
99
we then may write Eq. (24) as

~ ~
~ × B)
∇ · (E ~ · ∂ B − µ0 E
~ = −B ~ · ∂E ,
~ · J~ − ε0 µ0 E
∂t ∂t
or in the form
1 B2
!
1 ~ × B) ~ · J~ − ∂
~ = −E 1
∇ · (E ε0 E 2 + .
µ0 ∂t 2 2 µ0

On the left-hand side, we put 1/µ0 = ε0 c2 and integrate the equation over
some closed surface S enclosing a volume V . Then, we obtain

1 B2
!
~ · J~ dV − ∂ 1
Z Z Z
2 ~ × B)
ε0 c ∇ · (E ~ dV = − E ε0 E 2 + dV . (25)
∂t 2 2 µ0
V V V

Now we apply Gauss’s theorem to the left-hand side of the above equation,
to convert the volume integral into the closed surface integral, and find
I
~ × B)
ε0 c2 (E ~ · n̂ dS ←− Energy flux
S
Z
=− ~ · J~ dV
E ←− Rate of doing work by field on the current
V
1 B2
!
∂ Z 1
− ε0 E 2 + dV ←− Field energy. (26)
∂t 2 2 µ0
V

In words, this equation states that the instantaneous flow of power


across a closed surface S is equal to the time rate of decrease of the
energy stored in the field in the interior of S plus the power loss due
to the work on current J,~ conducting charges, within the volume V .

The above interpretation of the three terms in this equation can be justified
in the following way.

100
8.1 Rate of Doing Work by the Field on the Current
– Ohmic Heating

Consider first the second term: − E


R
~ · J~ dV . We will prove that this term
V
is equal to the rate of doing work by the electric field on the current that,
in other words, it is the ohmic heating. Simply, it shows that the power
produced by a surce contained inside the volume can be converted into the
ohmic power dissipated inside the volume.

We start our calculations from the circuit theory where we express all the
circuit variables in terms of the field variables.

From the circuit theory, we know that


in a resistive medium, as shown in Fig-
ure 29, the Ohm’s Law is: V = IR and
R = R`/A, where R is the resistivity
of the medium, I is the current flowing
through the medium and A is the area
through which the current is flowing.
Thus, we can write the current as
I = V /R = (V A)/(R`) .
Figure 29: Illustration of a current I
flowing through a resistive medium of a ~ = −∇ V , so that E = V /` and
Since E
length ` and cross section A.
then we can find the current density
through the area A, and write it in terms of the field variable E as
1
J = I/A = E.
R
Defining the conductivity σ = 1/R, the current density can be written as
I/A = J = σE .
Having the current density expressed in terms of the field variable E, we
can calculate the rate of conversion (dissipation) of the electromagnetic field
energy into heat
V2 E 2 `2
H = IV = = = σE 2 `A = σE 2 V ,
R R A`

101
where V = `A is the volume of the resistive medium. Hence, the energy
dissipated per unit volume is
H ~ · J~ .
= σE 2 = E J = E
V
~ · J~ is the rate of heating per unit volume in this case.
Thus, E

8.2 Electrostatic Field Energy Density


R 1 2
We now show that the term V 2 ε0 E dV represents energy contained in the
electric field E, so that
∂ Z 1
− ε0 E 2 dV
∂t V 2
is the time rate of decrease of the energy stored in the electric field.

Again, we start from the circuit theory from which we know that capacitors
are devices where electric field energy can be stored. Thus, consider the work
required to charge a capacitor of a capacitance C to a voltage V
1
W = CV 2 .
2
According to the field theory of electromagnetism this work done corresponds
to conversion from some other form of energy into electrostatic field energy
1 ε0 A
E = W = CV 2 and C = .
2 d
Since the electric field in the capacitor is given by E = V /d, we obtain
1 ε0 A 2 2 1
E= E d = ε0 E 2 V ,
2 d 2
where V = Ad is the volume of the capacitor. Hence
E 1
= ε0 E 2
V 2
is the energy per unit volume contained in or carried out by the electric
field E.

102
8.3 Magnetostatic Field Energy Density

B 2 /(2µ0 )dV represents energy contained in


R
We now show that the term V
the magnetic field B, so that

∂ Z 1 B2
− dV
∂t V 2 µ0
is the time rate of decrease of the energy stored in the magnetic field.

As before, we refer to the circuit theory from which it is well known that a
solenoid is a device where the magnetic field energy can be stored. Thus,
the work required to energize an inductor of inductance L to a current I is
E = 12 LI 2 and that the magnetic field within a long solenoid of self-inductance
L = µ0 n2 A` is B = µ0 nI. According to the field theory of electromagnetism
this work done corresponds to transformation of some other form of energy
to magnetic field energy.

Thus, for a solenoid of length ` and area of cross section A we have energy

1 1 1 1 1 B2
E = W = LI 2 = µ0 n2 A`I 2 = (µ0 nI)2 A` = V ,
2 2 2 µ0 2 µ0
oand then energy per unit volume is

E 1 B2
= .
V 2 µ0
This is the energy per unit volume contained in or carried out by the mag-
netic field B.

8.4 Poynting Vector

Since the right-hand side of Eq. (26) represents the rate of increase of the
electric and magnetic energies stored in the EM field and substracted by
the ohmic power dissipated as heat, the left-hand side must be equal (in

103
consistency with the law of conservation of energy) to the power flowing into
the volume through its surface. Thus, the expression
~ = ε 0 c2 ( E
N ~ × B)
~

is a vector representing the power flow per unit area, and is referred to as
the Poynting vector.

It represents the energy flux in the electromagnetic field, i.e. the energy flow
per unit area (measured normal to the flow) per unit time.

The quantity N has dimension of power per square meter. It is easy to see:
E has the dimension of volts per meter, ε0 c2 B has dimension of amper per
meter. Thus, the Poynting vector has dimension volts × amper/(square me-
ter) = Watts/(square meter).

The energy flow equation can be converted into the form of a differential
continuity equation or energy conservation law. From Eq. (25), we have

∂U ~ = −J~ · E
~ ,
+∇·N
∂t
where
1 1 B2
U = ε0 E 2 + (27)
2 2 µ0
is the energy density of the EM field.

The physical meaning of the differential continuity equation is that the time
rate of change of electromagnetic energy within a certain volume, plus the
energy flowing out through the boundary surface of the volume is equal to
the negative of the total work done by the fields on the source inside the vol-
ume. Thus. J~ · E~ is a conversion of electromagnetic energy into heat energy.

Equation (27) shows that we consider the energy stored in electric and mag-
netic fields as distributed throughout the region of space where these fields
are present with densities ε0 E 2 /2 and B 2 /2µ0 , respectively.

104
8.5 Phase Relationships in EM Waves

Here, we will show that only the in-phase components of E ~ and B~ contribute
to net energy flow averaged over a whole cycle of the radiation. If somehow,
the fields would oscillate with different phases, no energy can be transported.

Consider orthogonal fields E~ and B ~ oscillating in phase. Thus, if E


~ =
~
E0 cos(ωt)î and B = B0 cos(ωt)ĵ then:
~ = ε 0 c2 E
N ~ ×B
~ = ε0 c2 E0 B0 cos2 (ωt) k̂ .

Since, cos2 (ωt) = 12 , we obtain


1
N = ε0 c2 E0 B0 = ε0 c2 Erms Brms ,
2
√ √
where Erms = E0 / 2 and Brms = B0 / 2.

Hence, the average Pointing vector is different from zero indicating that the
~ and B
orthogonal fields E ~ oscillating in phase can transport energy.

~ = E0 cos(ωt)î and B
If, however, E ~ = B0 sin(ωt)ĵ then:

~ = ε 0 c2 E
N ~ ×B
~ = ε0 c2 E0 B0 cos(ωt) sin(ωt) k̂ .

Since, cos(ωt) sin(ωt) = 0, we have N = 0.

Hence, the average Pointing vector is zero indicating that the orthogonal
~ and B
fields E ~ oscillating out of phase phase cannot transport energy.

~ and B
In summary, electromagnetic waves in which the E ~ fields oscillate
in phase can transport energy.

8.6 Momentum Flux


It should be noted here, that EM waves can transport not only energy but
also momentum.

105
To obtain an expression for the momentum carried by the electromagnetic
field, we may employ the relativistic energy-momentum relationship

E 2 = p2 c2 + m20 c4 .

Since for electromagnetic radiation m0 = 0, we obtain


E
p= .
c
Thus the momentum flux of the electromagnetic field is:
~ ×B
ε 0 c2 E ~
M= ~ ×B
= ε 0 cE ~ ,
c
which leads to the conclusion that electromagnetic waves transport not only
energy, whose the flow is given by the Poynting vector, but also momentum.

8.7 EM Energy Flow: Circuit versus Field Theory


We now consider simple examples illustrating applications of the Poynting
vector to some familiar circuit problems to show how the field theory provides
an alternative (surprising!) way of viewing the energy transfer through the
circuits.

8.7.1 Energy Flow in a Resistive Wire


As a first example, consider a wire (resistor) of length `, carrying a current I,
as shown in Figure 30.
Let V is a potential difference ap-
plied along a resistive wire. We cal-
culate the power dissipated in the
wire using the circuit theory, and next
will compare the formalism with that
used in the field theory. We are
particularly interested in the predic-
Figure 30: A resistive wire of length ` tion of the two theories of the direc-
carrying a current I. tion of energy flow through the cir-
cuit.

106
Circuit theory calculation:

According to circuit theory the power dissipated in the wire is


P = V I = I 2R ,
where I is the current flowing through the wire and R is the resistance of
the wire. Since R is proportional to the length of the wire, a more power
(energy) is dissipated for a longer wire. Thus, according to the circuit theory,
energy flows along the wire.

Field theory calculation:

Let us now look at the same problem from the point of view of the field
theory. According to electromagnetic field theory, the direction of the flow
of energy is described by the Poynting vector N ~ × B.
~ = ε 0 c2 E ~

In order to find the direction of the


Poynting vector, we calculate the di-
rections of propagation of the electric
and magnetic fields produced by the
current. The electric field propagates
along the wire, as shown in Figure 31,
and is given by

Figure 31: Directions of the electric and ~ = −∇ V (z) = V (z) ẑ .


E
magnetic fields produced by a wire and the `
direction of the resulting Poynting vector.
From the Amperes line integral the-
orem, we find that the the magnetic
field propagates around the wire, and at points distant a from the wire is

~ = µ0 I φ̂ .
B
2πa
Hence, the Poynting vector is

~ = ε 0 c2 E
N ~ = ε0 c2 V µ0 I r̂ = V I r̂ ,
~ ×B
` 2πa 2π`a
where we have used the relations ẑ × φ̂ = r̂ and ε0 µ0 c2 = 1.

107
Thus, the field theory predicts that energy flows into the wire from the air
not along the wire. The energy is in the fields, the wire provides boundary
conditions and guides the fields.

The student may argue that the EM field carries only a small part of the
energy dissipated in the wire. However, the total rate at which field energy
flows into the wire by crossing the surface of the wire is given by
I I I
~ · dS
N ~ = ~ · r̂ dS =
N N dSside
S
I
VI
= N dSside = 2π`a = V I ,
2π`a
~ ⊥ dS
(N ~ on the ends of the cylinder), which is in agreement with the re-
sult of the circuit theory. This result demonstrates quantitatively that all the
power which heats the resistor enters through the sides not through the wires.

If the energy is in the fields, it means that the electromagnetic energy goes
out of a battery into the air, and then goes into the wire from the air. This
is exactly the case we will show in the following example.

8.7.2 Energy Flow out of Battery


In the above example, we have shown that according to the field theory, the
energy enters the resistor from the air.
Then, a question arises: If the en-
ergy enters the resistor from the air,
how does the energy get out to the
air from a source of energy (bat-
tery)?
To answer this question, consider a bat-
tery which provides energy to the resis-
tor. As illustrated in Figure 32, inside
the battery J~ and E~ are in opposite di-
rections. The magnetic field circulates
~
Figure 32: Directions of the current J around the battery, so we see that the
~ in a battery.
and the electric field E Poynting vector N~ points out into the air,
not along the wire.

108
Thus, the battery sends energy out into the air, not along the wire.

8.7.3 Propagation of an Electromagnetic Wave along a Wire

One of the above examples showed that field energy flows into a wire so that
it can be dissipated as heat. Consider now a different situation that one
would like to transmit an electromagnetic wave through a resistive wire.

It is well known from experiments,


that an electromagnetic wave can be
transmitted along the wire with very
little loss of its energy. Why does it
happen?

This effect has a simple expla-


nation in terms of the EM the-
ory.

If we consider the case of a perfect


Figure 33: Directions of the electric and conductor we find that there will be a
magnetic fields around a wire transmitting
an EM wave of wavelength λ. current wave along the wire with sur-
face charges induced producing a ra-
dial electric field. Then E ~ ×B ~ is parallel to the wire and the field descrip-
tion of energy transmission is that it is transmitted in the space around the
wire. In the space around the wire E ~ and B ~ are in phase and N
~ is always in
the same direction. Within a perfect conductor we will show later in Chap-
ter 12.6 that E ~ and B~ are π/2 out of phase so the mean N averaged over a
cycle is zero. Thus, there is no net energy transmission within the perfectly
conducting wire.

Exercise in class: Energizing of a capacitor

Consider the energizing of a plane parallel capacitor with


circular plates. Show that circuit and field calculations
agree as to the rate of energizing of the capacitor, i.e.
Pc = Pf where:

109
Pc = V I = rate of doing work (by current I and voltage V
between the plates) in charging the capacitor according to
circuit theory.

~ ×B
ε0 c2 E ~ · dS
~ = rate of energy flow into the surface
R
Pf =
S
of the capacitor according to field theory.

From which direction does the energy flow in to the ca-


pacitor according to field theory?

We will keep the calculation simple assuming the plates


of the capacitor to be uniformly charged. Under what con-
ditions is this assumption likely to be true?

Exercise in class: No Fluxes from Static Fields

~ and magnetostatic
Consider a source of electrostatic field E
~ ~
field B. If it is arranged so that E ⊥ B, ~ one may ask:
~ × B?
Should we expect to see an energy flux of ε0 c2 E ~

110
Revision questions

Question 1. Derive formula for the electrostatic field energy density.

Question 2. Derive formula for the magnetostatic field energy density.

Question 3. State the definition of the Poynting vector.

Question 4. In a circle containing a resistor, what is the direction of the energy


flow according to the circuit theory and the field theory?

Tutorial problems

Problem 8.1 Poynting vector on the surface of a long wire

Find the direction and the magnitude of the Poynting vector on


the surface of an infinitely long, straight conducting wire of ra-
dius r and conductivity σ, carrying a current I.

Problem 8.2 Energy flow in a plane wave

A plane electromagnetic wave propagates in the z direction in


a vacuum. Its electric field is polarized in the x direction and so
has the form:
~ = E0 sin(ωt − kz)î
E

(a) Find an expression for the vector energy flux in this wave.

(b) On the same time scale plot the electric field strength and
energy flux magnitude versus time at position z = 0 for 2 cycles
of the wave.

(c) What is the mean energy flux averaged over cycles?

111
9 Solution of Laplace’s Equation and Bound-
ary Value Problem
In one of the previous lectures, we discussed an advantage of using the scalar
and vector potentials in the calculation of the electric and magnetic fields. In
this lecture, we will illustrate applications of the scalar potential to physical
problems involving bounded fields.

There is a class of problems in electromagnetism in which a field can be


derived without involvement of the complete set of the Maxwell’s equations.
This can be done with the help of the gradient of a scalar potential which
satisfies Laplace’s equation

∇2 Φ = 0 .

Proof:

The condition for this to happen is that a vector field F~ has vanishing diver-
gence and curl

∇ · F~ = 0 and ∇ × F~ = 0 . (28)

Since, ∇× F~ = 0, we can always write F~ as F~ = ∇Φ, where Φ is an arbitrary


scalar function. Then, ∇ · F~ = 0 means that

∇ · (∇Φ) = ∇2 Φ = 0 .

Thus, the scalar potential Φ contains all the necessary information to com-
pletely specify the field of the properties (28).

Example 1: Electrostatic problems involving Laplace’s equation

~ = ρ/ε0 and ∇× E
Since in general ∇· E ~ =−∂B ~ we see that the requirement
∂t
for Laplace’s equation to be relevant is that ρ = 0 and ∂/∂t = 0, i.e. a source-
free region and static conditions. Of course there must be a source of charges
somewhere or there would be no field anywhere. The typical situation where
solution of Laplace’s equation is relevant in electrostatic is where we have
source-free non-conducting regions between statically charged conductors.

112
Example 2: Magnetostatic problems involving Laplace’s equation

~ = 0 and ∇ × B
Since in general ∇ · B ~ = µ0 J~ + 12 ∂ E,
~ we see that the require-
c ∂t
ment for Laplace’s equation to be relevant is that J~ = 0 and ∂/∂t = 0, i.e.
a current free region and static conditions. Again, there must be currents
somewhere or there would be no fields anywhere. The typical situation is
to be calculating the magnetic field in the non-conducting region between
constant currents.

9.1 Uniqueness of the Solution of Laplace’s Equation


As we shall see below, general solutions of Laplace’s equation are in terms of
some constants which are usually found from boundary (initial) conditions
for a given problem. A question arises:
What boundary conditions are appropriate for the Laplace equa-
tion to ensure that a unique and well-behaved (physically reason-
able) solution will exist inside the bounded region?

Our experience leads to believe that specification of the potential on a closed


surface defines a unique potential problem. This is called Dirichlet theorem
or Dirichlet boundary conditions.

9.1.1 Dirichlet Theorem

Consider a volume V completely bounded by a closed surface S. Within S


there is a potential Φ satisfying ∇2 Φ = 0. The Dirichlet theorem says that
the value of Φ inside the volume is uniquely determined if the value of the
potential is specified everywhere on the whole boundary.

Proof:

Suppose, to the contrary, that there exist two solutions Φ1 and Φ2 satisfying
the same boundary condition, i.e. ∇2 Φ1 = 0 and ∇2 Φ2 = 0 within S, but
Φ1 = Φ2 on S.

113
Let U = Φ1 − Φ2 is the difference between the solutions. Since Φ1 and Φ2
are known to be solutions of the Laplace equation, then from the linearity
of the ∇2 operator ∇2 U = 0, i.e. U is also a solution of the Laplace equation.

We will prove that U = 0 inside the volume. To show this, we introduce a


vector F~ = U ∇U . Then using the vector property that

∇ · F~ = ∇ · U ∇U = U ∇ · (∇U ) + ∇U · ∇U ,

and the Gauss’ Divergence Theorem, we get


Z Z Z
∇ · F~ dV = U ∇ · ∇U dV + ∇U · ∇U dV
V V V
Z Z
= U ∇2 U dV + (∇U )2 dV
V V
I
= ~ .
U ∇U · dS
S

Now, the right-hand side of the above equation is equal to zero because U = 0
over S. Also, the integral
Z
U ∇2 U dV = 0 ,
V

because Φ1 and Φ2 both satisfy the Laplace equation throughout V . Hence


Z
(∇U )2 dV = 0 .
V

Since the integral of a positive function is always positive, ∇U must be zero


for the integral to be zero. Thus ∇U = 0 and consequently, inside the vol-
ume V , the function U is constant. Since U = 0 on S, so that inside V , we
have then that Φ1 = Φ2 everywhere, as required.

9.2 Solutions of Laplace’s Equation


There are different methods of solving the Laplace equation

• Method of Images

114
• Green functions method

• Variational method

• Method of lattices

• Numerical Monte-Carlo simulations method

• Method of separation of variables

• Solution in spherical coordinates


We will illustrate last two methods which can be applied to a large class of
problems in electromagnetism. The other methods can be applied to spe-
cific problems. For these methods it is necessary that the boundaries over
which the potential is specified coincide with the constant bounding surfaces.

9.2.1 Method of Separation of Variables

In cartesian coordinates the Laplace equation for the scalar potential can be
written as
∂ 2Φ ∂ 2Φ ∂ 2Φ
∇2 Φ = + + 2 =0. (29)
∂x2 ∂y 2 ∂z
Since x, y, z are independent variables, the solution of the Laplace equation
is of the form

Φ(x, y, z) = X(x)Y (y)Z(z) .

Substituting this into the Laplace equation and dividing both sides by XY Z,
we obtain
1 d2 X 1 d2 Y 1 d2 Z
+ + =0.
X dx2 Y dy 2 Z dz 2
This equation can be separated into three independent equations. To show
this, we write this equation as
1 d2 X 1 d2 Y 1 d2 Z
= − − .
X dx2 Y dy 2 Z dz 2

115
Both sides of the above equation depend on different (independent) variables,
thus are equal to a constant, say −α2 :
1 d2 X
= −α2 ,
X dx2
1 d2 Y 1 d2 Z
− − = −α2 .
Y dy 2 Z dz 2
The second equation can be written as
1 d2 Y 2 1 d2 Z
= α − .
Y dy 2 Z dz 2
Again, both sides depend on different variables, thus are equal to a constant,
say −β 2 :
1 d2 Y
= −β 2 ,
Y dy 2
1 d2 Z
α2 − = −β 2 .
Z dz 2
Hence, after the separation of the variables, we get three independent ordi-
nary differential equations
1 d2 X
+ α2 = 0 ,
X dx2
1 d2 Y
+ β2 = 0 ,
Y dy 2
1 d2 Z
− (α2 + β 2 ) = 0 .
Z dz 2
The solutions of these equations depend on whether α2 and β 2 are positive
or negative constants. If we choose α2 and β 2 to be positive, the solutions
of the differential equations for X and Y are in the form of trigonometric
functions whereas the solution for Z is in the form of a hyperbolic function
X 
X(x) = Ak eiαx + Bk e−iαx ,
k
X 
Y (y) = Cl eiβy + Dl e−iβy ,
l
X √ √ 
α2 +β 2 z − α2 +β 2 z
Z(z) = Ep e + Fp e .
p

116
If α2 and β 2 had been chosen as negative constants, the hyperbolic and
trigonometric solutions would be interchanged.

The solutions can be equally well written in the form


X
X(x) = [Ak sin(αx) + Bk cos(αx)] ,
k
X
Y (y) = [Cl sin(βy) + Dl cos(βy)] ,
l
X q  q 
Z(z) = Ep sinh α2 + β 2 z + Fp cosh α2 + β 2 z .
p

The above solutions of the Laplace’s equation are in a general form valid for
an arbitrary problem. The constants α, β, Ak , Bk , Cl , Dl , Ep and Fp are found
from specific boundary conditions.

Consider two examples of an application of the general solutions:


1. We have a solution with known boundary conditions, find the problem.
2. We have a problem with specific boundary conditions, find the solution.

Example 1.

Consider the following two-dimensional solution of the Laplace’s equation


Φ(x, z) = X(x)Z(z) = V0 sin(αx) sinh(αz) (30)
with the lower boundary Φmin = 0 and the upper boundary Φmax = V0 .

In what circumstance is the above the solution?

Consider Φ(x, z) in some limits. Note that X = 0 for x = 0 or αx = π, i.e.


x = π/α.

The solution thus satisfies the boundary conditions along the ”vertical” lines
for α = π/b.

Since sinh(αz) = 0 for z = 0, the boundary condition along the lower bound-
ary is satisfied.

117
For the solution to satisfy the upper
boundary condition, the shape of the up-
per boundary must be such that
V0 sin(αx) sinh(αz) = V0 ,
for all points x, z on the line, i.e.
πx πz
   
sin sinh =1.
b b
 
Since sin πxb
→ 0 at the edges, it is
Figure 34: greatest (= 1) at the center x = b/2.
 
πz
Hence, sinh b must be equal to one at x = b/2. This happens when
πz
= arc sinh(1) ≈ 0.885 ,
b
from which we find
0.885b
z= = 0.282b .
π
Figure 34 shows the shape of the box inside which the potential is of the
form (30).

Example 2.

We usually have reverse problems to the above that we have a set of elec-
trodes which constitute equipotential lines or surfaces, and need to find the
appropriate solution of the Laplace equation. In this example we will illus-
trate this situation and we will try to find potential inside a rectangular box
whose three sides have potential equal to zero, and the remaining side has a
potential V0 .

Consider a two-dimensional problem with boundary conditions shown in Fig-


ure 35. This two-dimensional problem has a general solution
X
Φ(x, z) = [An sin(αx) + Bn cos(αx)]
n
× [En sinh(αz) + Fn cosh(αz)] .

118
The boundary condition of Φ = 0
at the side x = 0 can be sat-
isfied by Bn = 0. The bound-
ary condition of Φ = 0 at the
side z = 0 can be satisfied by Fn =
0. In order to have Φ = 0
at x = b, we must have αb =
nπ.

Figure 35: Hence with the boundary conditions


along three sides of Φ = 0, the general
solution reduces to

nπx nπz
X    
Φ(x, z) = Kn sin sinh , (31)
n=1 b b
where Kn = An En .

To find Kn we apply the remaining boundary condition that Φ = V0 at z = a:



nπx nπa
X    
Φ(x, a) = V0 = Kn sin sinh . (32)
n=1 b b
This is a Fourier series in x and in the usual way we use the orthogonality
properties of sine functions to calculate Kn :
Z 2π (
0 for m 6= n
sin(mφ) sin(nφ) dφ =
0 π for m = n
Z 2π (
0 for m 6= n
cos(mφ) cos(nφ) dφ =
0 π for m = n
Z 2π
sin(mφ) cos(nφ) dφ = 0 for all m and n .
0

Thus, multiplying both sides of Eq. (32) by sin(mπx/b) and integrating over
x = 0 → b, we get
Z b ∞
mπx nπa
  X  
V0 sin dx = Kn sinh
0 b n=1 b
Z b
mπx nπx
   
× sin sin dx .
0 b b

119
From the orthogonality of the sine functions, we find that all integrals on the
right-hand side are equal to zero except for m = n. Thus, after performing
the integration on the left-hand side, we are left with the term
b
nπx b nπa Z b 2 nπx
     
V0 − cos = Kn sinh sin dx ,
b 0 nπ b 0 b
which we can write as
V0 b nπa
 
[1 − cos(nπ)] = Kn sinh
nπ b
Z b 
1 2nπx
 
× 1 − cos dx .
0 2 b
The cos (2nπx/b) integrates to zero over the range 0 → b, so that the above
equation becomes
V0 b nπa b
 
[1 − cos(nπ)] = Kn sinh ,
nπ b 2
from which we find the constant Kn :
2V0 1 − cos(nπ)
Kn =   .
nπ sinh nπa
b

The solution for Kn can be further simplified:

Namely, if n is an even number, cos(nπ) = 1 and then Kn = 0.

If n is an odd number, cos(nπ) = −1, and then 1 − cos(nπ) = 2. Hence, we


finally arrive at
4V0 1
Kn =   , for odd n
nπ sinh nπa
b

and Kn = 0 for even n.

Substituting Kn into Eq. (31), we find that the potential inside the box has
the form
 
nπz
4V0 sinh b nπx
X  
Φ(x, z) =   sin .
odd n nπ sinh nπa b
b

120
As an exercise, check the correctness of this solution by showing that it
satisfies the Laplace’s equation.

Exercise in class: Reduction of the solution of a three dimesional problem


to a two dimensional problem

Why for a two dimensional problem we simplify the general


three-dimensional solution of the Laplace equations into

Φ = X(x)Z(z) ,
or
Φ = Y (y)Z(z) ,

but never

Φ = X(x)Y (y) .

121
9.2.2 Solution of the Laplace Equation in Spherical Coordinates

In this lecture, we continue the discussion of boundary-value problems and


will illustrate solution of the Laplace equation for general problems of spher-
ical symmetry. In this case, we shall work in spherical coordinates (r, θ, φ),
in which the Laplace equation takes the form

∂ 2Φ
! !
2 1 ∂ ∂Φ 1 ∂ ∂Φ 1
∇ Φ= 2 r2 + 2 sin θ + 2 2 =0.
r ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂φ2

The procedure we follow in the solution of the above equation is as follows:


Multiplying both sides by r2 , the Laplace equation can be written as a sum
of two separate parts

1 ∂ 2Φ
! !
∂ ∂Φ 1 ∂ ∂Φ
r2 + sin θ + =0.
∂r ∂r sin θ ∂θ ∂θ sin2 θ ∂φ2

The first part depends solely on r, whereas the second part depends solely on
the angles θ, φ. Thus, the solution of this equation is of the separable form

Φ = R(r)Y (θ, φ) .

Hence, substituting Φ = R(r)Y (θ, φ) and dividing both sides by R(r)Y (θ, φ),
we obtain
1 ∂ 2Y
! " ! #
1 d dR 1 1 ∂ ∂Y
r2 =− sin θ + .
R dr dr Y sin θ ∂θ ∂θ sin2 θ ∂φ2

Note, both sides of the above equation depend on different variables. It


follows that the both sides must be equal to the same constant, say −α:
!
d dR
r2 + αR = 0 ,
dr dr
1 ∂ 2Y
!
1 ∂ ∂Y
sin θ + − αY = 0 .
sin θ ∂θ ∂θ sin2 θ ∂φ2

Thus, the Laplace equation splits into two independent differential equations.
We will call them the radial part and angular part, respectively.

122
Angular part: Equation for Y .

Multiplying both sides by sin2 θ, we get

∂ 2Y
!
∂ ∂Y
sin θ sin θ − α sin2 θY + =0.
∂θ ∂θ ∂φ2

This equation contains two separate parts, one dependent only on θ and the
other dependent only on φ. Therefore, the solution will be of the form

Y (θ, φ) = X(θ)Ψ(φ) .

Hence, we get

1 d2 Ψ
!
1 d dX
sin θ sin θ − α sin2 θ = − .
X dθ dθ Ψ dφ2

As before, both sides must be equal to a constant, say m2 :


!
1 d dX
sin θ sin θ − α sin2 θ = m2 ,
X dθ dθ
2
1dΨ
= −m2 .
Ψ dφ2
First, we will solve the equation for Ψ, which we can write as

d2 Ψ
= −m2 Ψ .
dφ2
It is the familiar differential equation for a harmonic motion. The solution
of this equation is of simple exponent form:

Ψ(φ) = A exp(imφ) ,

where A is a constant.

To determine the constant m note that in rotation, φ and φ + 2π correspond


to the same position in space: Ψ(φ) = Ψ(φ + 2π), which is satisfied when

exp(imφ) = exp[im(φ + 2π)] .

123
This leads to

exp(i2πm) = 1 ,

that is satisfied when m = 0, ±1, ±2, . . ..


Hence, the constant m2 is not an arbitrary number, is an integer.

Now, we will find X(θ) to complete the solution for Y .

Using the solution for Ψ, the differential equation for X(θ) can be written as

m2
! !
1 d dX
sin θ − α+ X=0.
sin θ dθ dθ sin2 θ

Introducing a new variable z = cos θ, we can rewrite this equation as


 d2 X m2
!

2 dX
1−z − 2z − α + X=0,
dz 2 dz 1 − z2

or

m2
" # !
d   dX
1 − z2 − α+ X=0. (33)
dz dz 1 − z2

The above equation is known as the generalized Legendry differential equa-


tion, and its solutions are the associated Legendry polynomials. For m = 0,
the equation is called the ordinary Legendry differential equation whose so-
lution is given by the Legendry polynomials.

Lets look into the solution procedure of the above equation. This will allow
us to find α and X(θ).

We assume that the whole range of z (cos θ), including the north and south
poles (z = ±1), is in the region of interest. The desired solution should be
single valued, finite, and continuous on the interval −1 ≤ z ≤ 1 in order to
represent a physical potential.

124
The differential equation for X has poles at z = ±1. In order to find the
solution of this equation, we first check what solution could be continuous
near the poles.

Lets check a possible solution near z = 1. Substituting x = 1 − z, then


dx = −dz and in terms of x the equation takes a form

m2
" # !
d dX
x (2 − x) − α+ X=0.
dx dx x(2 − x)

We look for a solution in the trial form of power series in x



X(x) = xs an x n .
X

n=0

Substituting this into the differential equation for X, we get

2s2 a0 xs−1 + (s + 1)(2sa1 − sa0 + 2a1 )xs + . . .


m2
!
− α+ (a0 + a1 x + . . .)xs = 0 .
x(2 − x)

Near x ≈ 0, we can replace x(2 − x) by 2x, and obtain

m2
!
2
2s a0 − a0 xs−1 + (. . .)xs . . . = 0 .
2

This equation is satisfied for all x only if the coefficients at xs , xs±1 , . . . are
zero. From this, we find that
1
s = ± |m| .
2
We take only s = + 12 |m| as for s = − 21 |m| the solution for X(x) at x = 0
would go to infinity. We require the solution to be finite at any point x.

Thus, the solution that is continuous near x = 0 is of the form



1
X(x) = x 2 |m| an x n ,
X

n=0

125
or in terms of z
1

|m|
a0n z n .
X
X(z) = (1 − z) 2

n=0

Using the same procedure, we can show that near the pole z = −1, the
continuous solution is
1

X(z) = (1 + z) 2 |m| a00n z n .
X

n=0

Hence, we will try to find the solution in the form


 ∞
 1 |m| X
X(z) = 1 − z 2 2
bn z n . (34)
n=0

What left is to determine the coefficiets bn .

Substituting Eq. (34) into the differential equation for X(z), Eq. (33), and
collecting all terms at the same powers of z n , we obtain
X
{(n + 1)(n + 2)bn+2 − n(n − 1)bn
n
o
− 2(|m| + 1)nbn − (α + |m| + m2 )bn z n = 0 ,

from which we find a recurrence relation for the coefficients bn


(n + |m|)(n + |m| + 1) + α
bn+2 = bn .
(n + 1)(n + 2)

We have two separate solutions for even and odd n. For b0 6= 0, we put
b1 = 0, and the solution is given in terms of even n. For b0 = 0, we put
b1 6= 0, and the solution is given in terms of odd n.
We cannot accept both the even and odd solutions at the same time, because
in this case the solution X(z) would not be a single valued function that is
would not be accepted as a physical potential.
For example, for b0 6= 0, we have α = −|m| − m2 , but for b1 6= 0, we have
α = −2 − 3|m| − m2 . If we would accept both of the solutions at the same
time, the potential would have two different values.

126
We check now whether the series is converting when n → ∞ which would
ensure that the potential is finite.

Since bn+2 > bn , the series diverges for z = ±1. Therefore, in order to get
the potential finite everywhere in the space, we have to terminate the series
at some n = n0 . In other words, we assume that bn0 +1 = bn0 +2 = . . . = 0.

The series terminating at n = n0 indicates that

(n0 + |m|)(n0 + |m| + 1) + α = 0 .

Introducing

l = n0 + |m| ,

we see that l ≥ |m|, and

α = −l(l + 1) , l = 0, 1, 2, . . .

Thus, the solution for X(z) is

  1 |m| l−|m|
Xlm (z) = 1 − z 2 bn z n ,
2
X

where the sum is over even n when l − |m| is an even number, and over odd n
when l − |m| is an odd number.

First few solutions

X00 (z) = b0 = b0 P00 (z) ,


X10 (z) = b1 z = b1 P10 (z) ,

X11 (z) = b0 1 − z 2 = b0 P11 (z) ,

where P00 (z) = 1, P10 (z) = z, P11 (z) = 1 − z 2 , . . . are the associate Legendry
polynomials of the order l.

Useful examples of the associate Legendry polynomials, written in terms


of θ, (z = cos θ):

P00 (cos θ) = 1 ,

127
P10 (cos θ) = cos θ ,
P11 (cos θ) = sin θ ,
1
P20 (cos θ) = [3 cos(2θ) + 1] ,
4
3
P21 (cos θ) = sin(2θ) ,
2
3
P22 (cos θ) = [1 − cos(2θ)] .
2
An important property of the Legendry polynomials is orthogonality that
Z 1
Plm (cos θ)Pkn (cos θ) d(cos θ) = 0 for m 6= n and l 6= k ,
−1
Z 1
2 (l + m)!
[Plm (cos θ)]2 d(cos θ) = for m = n and l = k .
−1 2l + 1 (l − m)!
Finally, with the above notation, the solution for the angular part of the
Laplace equation Y (θ, φ) is of the form

Alm Plm (cos θ)eimφ .


XX
Y (θ, φ) =
l m

Radial part: Equation for R(r).

With α = −l(l + 1), the differential equation for R takes the form
!
d dR
r2 − l(l + 1)R = 0 .
dr dr

Dividing by r and introducing a new function U (r) = rR(r), we obtain

d2 U l(l + 1)
− U =0.
dr2 r2
Lets first check the asymptotic solution for r  1. In this case we can ignore
the second term in the differential equation, and find that the asymptotic
equation has a solution U (r  1) = Cr, where C is a constant.

Following this asymptotic behavior, we will try the general solution of a form

U (r) = rs .

128
Substituting this into the differential equation, we obtain

[s(s − 1) − l(l + 1)] rs−2 = 0 .

This equation is satisfied for any r when

s = (l + 1) or s = −l .

Thus, the general solution for U that satisfies the asymptotic solution is of
the form

U (r) = C1 rl+1 + C2 r−l ,

and then

R(r) = C1 rl + C2 r−(l+1) .

Hence, the general solution of the Laplace equation in spherical polar coor-
dinates is of the form:
XX 
Φ(r, θ, φ) = C1l rl + C2l r−(l+1) Alm Plm (cos θ)eimφ .
l m

The solution can be written as


X X n 
Φ(r, θ, φ) = C1l rl + C2l r−(l+1)
l m
× [alm cos(mφ) + blm sin(mφ)] Plm (cos θ)} .

This is the final solution for the potential of an arbitrary spherically sym-
metric problem. The potential is finite, determined and is continuous at each
point (r, θ, φ).

Example 1: Potential outside a conducting sphere

Consider an example of boundary-value problem with azimuthal symmetry:


A conducting sphere of a radius a in an uniform electric field, as shown
in Figure 36.

129
In order to find the potential, we first have to determine boundary conditions
for the potential.

There are two boundaries: One of the boundaries is the surface of the sphere
and the other is the region at r → ∞.

The conducting sphere is an equipotential volume (else there would be electric


fields driving currents till it became so). We take, with no loss of generality,
the zero potential on the sphere.

Thus, the boundary conditions to be sat-


isfied are:

1. The potential on the surface of the


sphere Φ(a, θ, φ) = 0.

2. The potential at infinity is the uni-


form field potential (no effect of the
sphere at infinity), so Φ = −E ~ · ~r =
−Er cos θ at infinity.

Since the applied potential is indepen-


Figure 36: A dielectric sphere in a dent of the angle φ, the induced poten-
uniform electric field. tial will also be independent of φ. This
is a general property of the potential in-
side a bounded area that we shall ex-
plain in details at one of the tutorial se-
sions.

Thus, using this property, we can set m = 0 in the general solution and get
Xh i
Φ(r, θ) = C1l rl + C2l r−(l+1) Pl0 (cos θ) .
l

At infinity (r → ∞), the boundary condition is satisfied for all constants


C1l = 0 except for l = 1 (remember P1 = cos θ). Thus
X C2l Pl (cos θ)
Φ(r, θ) = C11 rP1 (cos θ) + .
l rl+1

130
We first find the explicit form of C11 . As r → ∞, the potential Φ(r, θ) →
C11 r cos θ = −Er cos θ. Therefore C11 = −E.

The remaining coefficients C2l we find from the other boundary condition
that Φ(a, θ) = 0 on the surface of the sphere. Thus, at r = a
X C2l Pl (cos θ)
Φ(a, θ) = 0 = −EaP1 (cos θ) + .
l al+1

We can determine the coefficients C2l using the orthogonality properties of


the Legendry polynomials. Multiplying the above equation by Pk (cos θ) sin θ
and integrating over sin θdθ = d(cos θ), we obtain
Z 1
0 = −aE P1 (cos θ)Pk (cos θ) d(cos θ)
−1
X C2l Z 1
+ Pl (cos θ)Pk (cos θ) d(cos θ) .
l al+1 −1

For k 6= 1 the first term vanishes by orthogonality of the Legendry polyno-


mials. Moreover, all terms in the summation vanish except that for l = k.
Thus, for k 6= 1
C2k Z 1
0 = k+1 Pk (cos θ)Pk (cos θ) d(cos θ) .
a −1

Since the integral is nonzero, then C2k = 0.

For k = 1 and using the orthogonality property of the Legendry polynomials


Z 1
2 (l + m)!
Plm (cos θ)Plm (cos θ) d(cos θ) = ,
−1 2l + 1 (l − m)!

we find that for l = 1 and m = 0, the integral is equal to 2/3. Thus

2 C21 2
0 = −aE + 2 ,
3 a 3
from which, we find C21 = Ea3 . Hence
cos θ
Φ(r, θ) = −Er cos θ + Ea3 .
r2

131
The first term is just the potential of a uniform field E. The second term
is the potential due to the induced surface charges or, equivalently, is the
potential of the induced dipole moment p = 4πε0 Ea3
p cos θ
Φdip = .
4πε0 r2
This result shows that the sphere placed in an uniform electric field behaves
as a dipole because of the effect of the charge distribution induced on its
surface.

Exercise in class: Potential inside a sphere

Let us suppose that the student would try to find Φ in-


side a sphere using the general solution for Φ. We have
found before, using the Gauss’s law, that Φ = const. in-
side the sphere.

The student immediately notice that the solutions appear


to be mathematically different since that found by using the
Gauss’s law is in a closed form while the general one is in
the form of a series. The question to be settled is whether
the two solutions are identical or whether, perhaps, only
one of the solutions is correct?

132
Revision questions

Question 1. What conditions are imposed on the field that is completely deter-
mined by a scalar potential?

Question 2. State the Dirichlet theorem.

Question 3. Explain the method of separation of variables on example of the


Laplace equation in cartesian coordinates.

Question 4. Explain the method of solving the generalized Legendry differential


equation that satisfies the conditions imposed on the scalar poten-
tial.

133
Tutorial problems

Problem 9.1 Electrostatic potential inside a rectangular box with symmetrical


boundary conditions

Find a series solution in rectangular harmonics for the electrostatic


potential inside a 2-dimensional box in which the sides at y = 0
and y = b are at zero potential and the sides at z = a and z = −a
are at potential V0 .

Problem 9.2 Potential inside a rectangular box

Consider a two-dimensional region with boundaries at x = 0, b


and z = 0, a, as shown in Figure 37. The boundary conditions are

Figure 37:

∂Φ
= 0 at z=0,
∂z
Φ = 0 at x = 0, b ,
Φ = V0 at z=a.

Find the potential at any point inside the two-dimensional region.

134
Problem 9.3 Potential inside an open rectangular box

Find the potential everywhere inside a two dimensional region


bounded from three sides by a finite plane at z = 0 and by semi-
infinite planes located at x = 0 and x = b, as shown in Figure 38.

The boundary conditions are

Φ = 0 at x=0 and x = b ,
Φ = V0 at z = 0 ,
Φ = 0 at z→∞.

Figure 38:

Problem 9.4 Potential inside a 3-dimensional box

A 3-dimensional conducting box is defined by x = 0 to a, y =


0 to b and z = 0 to c. All the sides of the box are connected
together and grounded except that side defined by z = c, which
is insulated from the rest and held at potential V0 . Show that the
potential within the hollow box can be written as the series
mπx
 nπy 
X X 16 V0 sin a sin b
Φ(x, y, z) = q  
nmπ 2 mπ 2
 nπ 2

n m
odd odd
sinh a + b c
s  
2 2
mπ nπ
 
× sinh  +  z .
a b

135
Problem 9.5 Potential due to a thin circular ring

(a) Show that for a system with axial symmetry such as a ring
of charge, the electrostatic potential can be written as a series in
spherical polar coordinates:
X 
Φ(r, θ) = A` r` + B` r−(`+1) P` (cos θ) .
`

To determine the coefficients A` and B` consider the solution along


the z-axis, where θ = 0. It is useful to know that P` (1) = 1 for
any value of `. Thus, along the axis:
X 
Φ(z) = A` z ` + B` z −(`+1) .
`

(b) Calculate Φ(z) directly from the Coulomb potential. Expand


this expression as a power series and hence find the A` and B` by
comparison.

Hence, show that for a ring of radius a and charge per unit length λ
the potential everywhere can be written as the series:
   
`
λ  X (−1) 1.3.5 . . . (` − 1) 
2
Φ(r, θ) = 1 + r` P` (cos θ) ,
 
`
 
2ε0 ` 2 2
`
! a`
even 2

where ` = 2, 4, 6, . . . (even).

The Legendre functions can be computed from the recurrence re-


lation:
(` + 1) P`+1 (x) = (2` + 1) x P` (x) − ` P`−1 (x)
and knowing that P0 (x) = 1 and P1 (x) = x.
However, the functions become rather tedious for hand calcula-
tions for large `, e.g.
1 n
P10 (x) = −252 + 13860x2 − 120120x4 + 360360x6
1024 o
−437580x8 + 184756x10 .

136
However, there is Matlab etc.

(c) Calculate also the electric field components of the circular


~ = −∇Φ.
ring using the relation E

(There is a useful recurrence relation for the derivative of the Leg-


endre function:

d
(x2 − 1) P` (x) = `x P` (x) − ` P`−1 (x) .
dx

See e.g. Abramowitz & Stegun ‘Handbook of Mathematical Func-


tions’).

137
10 General Solution of the Maxwell’s Equa-
tions

In Chap. 6, we presented a method of solving the Maxwell’s equations for


the EM fields propagating in vacuum. We were able to reduce the Maxwell’s
~ and B
equations into two differential equations for E ~ alone. The equations
were in the form of the wave equations whose the solutions are in the form
of plane and transverse waves. In this lecture, we will present a method for
general solution of the Maxwell’s equations in the presence of the sources,
charges and currents that depend on position ~r and time t. The concept of
vector and scalar potentials and gauges will be useful in the general solution.

10.1 Difficulty of the Direct Solution of Maxwell’s Equa-


tions
Consider the Maxwell’s equations for the fields in the presence of charges and
currents
I. ~ = ρ/ε0 ,
∇·E (35)
II. ~ =0,
∇·B (36)
III. ∇×E~ =−∂B ~ , (37)
∂t
IV. ∇×B~ = µ0 J~ + 1 ∂ E ~ . (38)
c2 ∂t
~ and B
In the Maxwell’s equations, the fields E ~ depend on (~r, t), the charge
and current densities also depend on (~r, t). It is not explicitly stated in the
above equations, but we shall remember about this dependence in the fol-
lowing calculations.

~ and B
Let us try to solve the Maxwell’s equations to find the fields E ~ pro-
~
duced by the source charges ρ and currents J.

The Maxwell’s equations involve two fields that satisfy coupled differential
equations. First, we will try to separate the Maxwell’s equations into a dif-
~ alone or B
ferential equation for E ~ alone.

138
Assuming in the usual way that space and time operators commute, we act

with c12 ∂t on III and ∇× on IV, and obtain from III:
2
1 ∂ ~+ 1 ∂ B ~ =0,
∇ × E
c2 ∂t c2 ∂t2
and from IV:
1 ∂ ~ − ∇ × (∇ × B)
~ = −µ0 ∇ × J~ .
∇×E
c2 ∂t
~ by subtraction of the two above equations, we get
Eliminating E
2
~ + 1 ∂ B
∇ × (∇ × B) ~ = µ0 ∇ × J~ .
c2 ∂t2
Using the vector identity for double × product, and from II:
~ = −∇2 B
∇ × (∇ × B) ~ + ∇(∇ · B)
~ = −∇2 B
~ ,

we arrive to a wave equation

~− 1 ∂2 ~
∇2 B B = −µ0 ∇ × J~ . (39)
c2 ∂t2
~ gives
Similarly, elimination of B

~− 1 ∂2 ~ 1 ∂
∇2 E 2 2
E = ∇ρ + µ0 J~ . (40)
c ∂t ε0 ∂t
Equations (39) and (40) are in the form of coupled wave equations known as
inhomogeneous Helmholtz equations. We see that the current density J~ en-
ters into both of these equations and enters in a relatively complicated way.
These equations are coupled through J, ~ and for this reason these equations
and are not readily soluble in general.

In the absence of currents and charges, J~ = 0, ρ = 0, and then the above


equations describe a free EM field, and can be solved separately, as it was
illustrated in Chap. 6. The general solution of the wave equations, in the
presence of space and time varying currents and charges is complicated and
is more readily attained via the electromagnetic potentials.

139
10.2 Scalar and Vector Potentials

~ and B
Generally, we do not find fields E ~ directly by integration of Eqs. (39)
and (40). We rather first compute scalar and vector potentials from which
the fields may be found. We will illustrate here the advantage of working
with the potentials, i.e. with the the scalar and vector potentials.

~ defined such that the Maxwell’s equation II


Introduce the vector potential A
remains unchanged. The field B ~ always has zero divergence, ∇ · B
~ = 0, and
hence we can always write
~ =∇×A
B ~, (41)

since ∇ · ∇ × A~ is identically zero. Substitute this relation to the Maxwell’s


equation III, we obtain
 
~
~ =−∂ ∇×A~ = ∇ × − ∂ A  .
 
∇×E (42)
∂t ∂t

The two curl’s are equal, but it does not mean that the vectors under the
curls are equal. Since always ∇ × ∇Φ = 0, where Φ is an arbitrary scalar
function, the two vectors are equal with the accuracy to ∇Φ:

~ =−∂A
E ~ − ∇Φ . (43)
∂t
Scalar potential is a quantity from which a field can be derived by a process
of differentiation, e.g. in electrostatics
~ = −∇Φ ,
E

where Φ is the electrostatic potential.

~
In the static limit of ∂ A/∂t = 0, the scalar function Φ reduces to the familiar
electrostatic potential.

Equation (43) shows that the electric field depends on the specific choice of
the potentials. We can change A~ and Φ and still get the same E.~ One can
~ so Eqs. (41) and (43) are
object that Eq. (41) ensure a fixed value for A,

140
satisfied for fixed A~ and Φ. However, we can define new potentials without
changing E ~ and B ~
~0 = A
A ~ + ∇Λ ,

Φ0 = Φ − Λ . (44)
∂t
Proof:

From Eq. (43), we find that

~0 = − ∂ A
E ~ 0 − ∇Φ0
∂t !
∂ ~  ∂
= − A + ∇Λ − ∇ Φ − Λ
∂t ∂t
∂ ~ ~ .
= − A − ∇Φ = E
∂t
Similarly, fro Eq.(41), we find that
~0 = ∇ × A
B ~0 = ∇ × A
~ + ∇ × (∇Λ) = ∇ × A
~=B
~ .

as required.

The transformation (44) is called a gauge transformation, and the invariance


of the fields under such transformations is called gauge invariance.

~
In this case, how do we completely determine A?

From the Helmholtz theorem we know that the definition B ~ = ∇×A ~ does
not completely define A ~ despite the fact that B ~ is completely defined. The
~
vector potential A is arbitrary to the extent that the gradient of some scalar
function can be added. Thus, infinite set of possible potentials Φ corresponds
to an infinite set of possible vector potentials.

Recall the Helmholtz Theorem which says that any vector field can be
written as a sum two terms

~ 1 Z ∇ · F~ 1 Z
∇ × F~
F = − ∇ dV + ∇× dV
4π V r 4π V r

= F~l + F~t ,

141
where F~l is called the longitudinal part of the field and it has ∇ × F~l = 0,
while F~t is called the transverse part as it has ∇ · F~t = 0.

We see that ∇ · F~ and ∇ × F~ together determine F~ but neither do alone.


~ we complete the definition of A.
Thus, if we define ∇ · A, ~ This is called
”choosing the gauge of the potential”. The above is an excellent illustration
of the power of the Helmholtz theorem. This theorem enables us to recog-
nize basic common properties of vector fields independent of their individual
physical properties.

10.2.1 Lorenz Gauge

~
How do we define ∇ · A?

It is done as follows. We take ∇· of Eq. (43) and obtain

∇·E~ =−∂ ∇·A ~ − ∇2 Φ . (45)


∂t
~ will satisfy the Maxwell’s equation I when
Thus, the electric field E
∂ ~ − ∇2 Φ = ρ/ε0 .
− ∇·A (46)
∂t
~ = ∇ × A,
From the Maxwell’s equation IV and B ~ we have
!
~ = µ0 J~ + 1 ∂ − ∂ A
∇ × (∇ × A) ~ − ∇Φ ,
c2 ∂t ∂t
which can be written as
2
~ = µ0 J~ − 1 ∂ A
~ + ∇(∇ · A)
−∇2 A ~− 1∇∂Φ,
2
c ∂t 2 c2 ∂t
or
2
!
~− 1 ∂ A
∇A2 ~+ 1 ∂Φ
~ = −µ0 J~ + ∇ ∇ · A . (47)
2
c ∂t 2 c2 ∂t

The freedom of choosing A~ and Φ means that we can choose a set of potentials
to satisfy the condition

~+ 1 ∂Φ=0.
∇·A
c2 ∂t
142
~ This equation is sometimes
This is called the Lorenz gauge and defines ∇ · A.
called the Lorenz equation.

Under the Lorenz gauge, Eq. (47) reduces to

~− 1 ∂2 ~
∇2 A A = −µ0 J~ ,
c2 ∂t2
and applying the Lorenz gauge to Eq. (46), we get

2 1 ∂ 2Φ
∇ Φ − 2 2 = −ρ/ε0 .
c ∂t
We see the advantage of using the vector and scalar potentials. In terms
of the potentials, the the Maxwell equations reduce to two uncoupled wave
equations that can be solved separately. The equations also show that the
Lorenz gauge is consistent with our experience on the sources of the EM
fields. The source of the vector potential that is related to the magnetic field
is a current density, and the source of the scalar potential that is related to
charges is a charge density.

10.2.2 Coulomb Gauge


Another useful gauge of the potentials is the Coulomb gauge or transverse
gauge
~=0.
∇·A

The origin of the name ”Coulomb gauge” is in equation (45) that under the
~ = 0 reduces to the Poisson equation
condition ∇ · A

∇2 Φ = −ρ/ε0 ,

that determines the Coulomb potential due to the charge density ρ.

Where the name transverse gauge came from?

Before we give the answer to this question, we first show that the solution of
the Poisson equation is of the form
1 Z ρ
Φ(r) = dV .
4πε0 r

143
It can be proved in the following way.

From the Coulomb’s law


~ = 1 Z ρr̂
E dV .
4πε0 r2
and using the relation
1 r̂
−∇ = 2 ,
r r
we can write the electric field as
~ 1 Z
ρ
E=− ∇ dV = −∇Φ ,
4πε0 r
where
1 Z ρ
Φ= dV . (48)
4πε0 r
Since the electric field satisfies the Maxwell’s equation I, we find

∇·E ~ = −∇2 Φ = ρ ,
ε0
as required.

~ under the Coulomb gauge


Now, we can find the wave equation for A

~− 1 ∂2 ~ 1 ∂Φ
∇2 A 2 2
A = −µ0 J~ + 2 ∇ . (49)
c ∂t c ∂t
In this equation, the term involving the scalar potential is called ”longitudi-
nal” as it has vanishing ∇×. This suggests that it may cancel the longitudinal
part of the current density J.~

Let us check if the longitudinal part of the current density can be expressed
in terms of the scalar potential. According to the Helmholtz Theorem, the
current density can be written as

~ 1 Z ∇ · J~ 1 Z
∇ × J~
J = − ∇ dV + ∇× dV
4π V r 4π V r

= J~l + J~t .

144
Using the continuity equation
∂ρ
+ ∇ · J~ = 0
∂t
and the solution of the Poisson equation, we find the longitudinal part of the
current density

1 Z ∇ · J~ 1 Z − ∂ρ
J~l = − ∇ dV = − ∇ ∂t
dV
4π V r 4π V r
1 ∂ Z ρ ∂
= ∇ dV = ε0 ∇Φ .
4π ∂t V r ∂t
Then
∂ 1 ∂Φ
µ0 J~l = µ0 ε0 ∇Φ = 2 ∇ ,
∂t c ∂t
which is equal to the second term on the right hand side of Eq. (49).

Hence, the inhomogeneous term in the wave equation (49) can be expressed
~
entirely in terms of the transverse current and then the wave equation for A
reduces to
1 ∂2 ~
~
∇ A − 2 2 A = −µ0 J~t .
2
c ∂t
This explains the origin of the name ”transverse gauge”.15

10.3 Solution of the Inhomogeneous Wave Equations


We have shown that the Maxwell equations can be reduced to two indepen-
dent wave equations for the potentials A ~ and Φ. In fact, we have four scalar
equations for (Ax , Ay , Az , Φ). Each of these equations has the same form.
Therefore, it is enough to solve one of the four equations.

15
The transverse gauge is often used in atomic physics to calculate the EM fields pro-
duced by orbiting electrons. In this case, the orbiting electrons produce a current that is
solenoidal.

145
We will illustrate the solution on the equation for Φ:

2 1 ∂ 2Φ
∇ Φ − 2 2 = −ρ/ε0 . (50)
c ∂t
A general solution of the above equation may be found by considering two
limiting cases:

(a) Electrostatic limit: ∂/∂t ≡ 0

In this limit the wave equation for Φ reduces to the Poisson equation whose
the solution is
1 Z ρ(r0 )
Φ(r) = dV 0 .
4πε0 r0

(b) Source free limit: ρ=0

In this case, the wave equation (50) reduces to the homogeneous equation
1 ∂ 2Φ
∇2 Φ − =0.
c2 ∂t2
This equation has a spherically symmetric solution of the form
f (t − r/c)
Φ(r, t) = ,
r
where f (t − r/c) is an arbitrary function of the retarded time t − r/c. The
retardation r/c is equal to the time needed for the electromagnetic wave to
pass the distance from the source to a given point in space.

Proof:

If there are no charged (boundary) surfaces in the space, the potential can
depend only on r, and must in fact be spherically symmetric. Thus, in
spherical coordinates only the radial part of the Laplacian will contribute to
the wave equation
" #
1 ∂
2 ∂Φ
∇ Φ= 2 r2 .
r ∂r ∂r

146
Since
∂f ∂f ∂tr 1
= = − f0 ,
∂r ∂tr ∂r c
where tr = t − r/c and f 0 = ∂f /∂tr , we have

f0 1 ∂ rf 0
" !# !
2 1 ∂ f
∇Φ = − 2 r2 + 2 =− 2 +f
r ∂r cr r r ∂r c
1 f0 r
"  #
1 00 1 0 1
  
= − 2 + − f + − f = 2 f 00 ,
r c c c c rc

where f 00 = ∂ 2 f /∂t2r .

Moreover, ∂/∂t = ∂/∂tr , and then

1 ∂ 2Φ 1
2 2
= 2 f 00 .
c ∂t cr
Consequently, we obtain

1 ∂ 2Φ 1 1
∇2 Φ = 2 2
= 2 f 00 − 2 f 00 = 0 ,
c ∂t cr cr
as required.

Summarizing the above analysis, we can construct a general solution of the


wave equation by noting that it must represent a spherical wave outside the
source and reduce to the appropriate static limit. This solution is
1 Z ρ(t − r/c)
Φ(r, t) = dV ,
4πε0 V r
where r is the distance coordinate from the source (from the charge ρdV ) at
the time when the potential wave left it. This exhibits the causal behavior
associated with the wave disturbance. The argument of ρ shows that an
effect observed at the point r at time t is caused by the action of the source
a distant r away at an earlier or retarded time tr = t − r/c. The time r/c is
the time of propagation of the disturbance from the source to the point r.
Thus, the Maxwell’s equations satisfy the causality principle.

147
10.4 Rigorous Solution: Green Functions Method

The wave equations all have the basic structure


1 ∂ 2 Φ(r, t)
∇2 Φ(r, t) − = −4πf (r, t) ,
c2 ∂t2
where f (r, t) is a known (source distribution) function.

To solve this equation, we will introduce the Green function of the equation
and solve it as an inhomogeneous Helmholtz equation.

Suppose that Φ(r, t) and f (r, t) have the Fourier integrals


Z ∞
Φ(r, ω) = Φ(r, t)eiωt dt ,
−∞
Z ∞
f (r, ω) = f (r, t)eiωt dt .
−∞

When we insert it into the wave equation, we find that the Fourier transform
Φ(r, ω) satisfies the inhomogeneous Helmholtz wave equation
 
∇2 + k 2 Φ(r, ω) = −4πf (r, ω) ,
where k = ω/c is the wave number.

The advantage of working in Fourier components is to remove the derivative


over time, and consequently to reduce the wave equation to a differential
equation involving the space variables only.

Green function

For a unit point source the potential satisfies the Poisson equation
1
∇2
= −4πδ(r) .
r
The function 1/r = G(r) is called a Green function of the above differential
equation.
In analogy, we can define the Green function of the wave equation
1 ∂2
!
2
∇ − 2 2 G(r, t − t0 ) = −4πδ(r)δ(t − t0 ) .
c ∂t

148
The Fourier transform gives
 
∇2 + k 2 Gk = −4πδ(r)eiωt0 .

where Gk is the Fourier transform of the Green function G(r, t − t0 ), which


we are trying to find.

If there are no boundary surfaces, the Green function depends only on r,


and then the Laplacian operator in spherical coordinates depends only on r
giving

1 d2  −iωt0

rG k e + k 2 Gk e−iωt0 = −4πδ(r) .
r dr2
Everywhere except r = 0, the function rGk e−iωt0 satisfies the homogeneous
equation

d2  −iωt0

rG k e + k 2 (rGk e−iωt0 ) = 0 ,
dr2
whose the solution is

rGk e−iωt0 = Aeikr + Be−ikr .

In this general solution for the Green function we could equally well choose
the exponential form

e±ikr eiωt0
Gk = .
r
Using the inverse Fourier transform, we find

1 Z ∞ e±ikr −iωτ
G(r, τ ) = e dω ,
2π −∞ r
where τ = t − t0 .

The integral
1 Z ∞ −iω(τ ∓r/c)
e dω
2π −∞

149
is the delta function δ(τ ∓ r/c). Thus
1
G(r, τ ) = δ(τ ∓ r/c) .
r
The Green function is a casual response function, and has the same property
as the scalar potential of a point source.

In summary of the general solution of the Maxwell’s equations

The general (retarded) solutions of the Maxwell’s equations are given in terms
of the vector and scalar potentials
~
~ =∇×A
B ~ , ~ = −∇Φ − ∂ A ,
E (51)
∂t
with
1 Z ρ(t − r/c)
Φ(r, t) = dV , (52)
4πε0 r
~ − r/c)
1 Z J(t
~
A(r, t) = dV , (53)
4πε0 c2 r
and the potentials satisfy the Lorenz gauge.

Thus, we do not find the fields by a direct integration of the Maxwell’s equa-
tions. If the charge and current distributions are known, we first calculate
the scalar and vector potentials from Eqs. (52) and (53), and then find the
electric and magnetic fields from Eqs. (51).

Revision questions

Question 1. Do we find the fields by a direct integration of the Maxwell’s equa-


tions? Explain.

150
Question 2. Explain the usefulness of the scalar and vector potentials in the
solution of the Maxwell’s equations.

Question 3. Why do we use gauges in the solution of the Maxwell’s equations?

Question 4. Explain how do we solve inhomogeneous wave equations.

Question 5. What is the Green function of a given differential equation?

151
11 Electromagnetic Antennas: Hertzian Dipole

The electromagnetic fields of charges in uniform motion are effectively bound


to the charges. The fields of accelerated (oscillating) charges on the other
hand can propagate as electromagnetic (EM) waves at the speed c and can
have a life of their own (until absorbed by some other charges).

In this lecture we will show how electromagnetic waves are generated by oscil-
lating charges. This will also illustrate an application of the general solution
of the Maxwell’s equations in calculations of the electric and magnetic fields
produced by a source system containing time varying charges and currents.
First, we will show that this problem can be solved with the help of only
the vector potential A.~ Next, we will apply this concept to the problem of
generation of electromagnetic waves.

Consider the retarded solutions of the Maxwell’s equations


~
~ =∇×A
B ~ , ~ = −∇Φ − ∂ A ,
E (54)
∂t
with
1 Z ρ(t − r/c)
Φ ≡ Φ(r, t) = dV ,
4πε0 r
~ − r/c)
1 Z J(t
~ ≡ A(r,
A ~ t) = dV .
4πε0 c2 r
The above solution holds for the Lorenz gauge in which

~=− 1 ∂Φ
∇·A .
c2 ∂t
Assume that the position and time variations of the charges and currents can
be separated, so that the charges and currents vary sinusoidally in time

ρ(r, t) = ρ(r)eiωt ~ t) = J(r)e


and J(r, ~ iωt
.

In this case, the Lorenz gauge takes the form

~ = − iω Φ ,
∇·A
c2
152
which gives
c2 ~.
Φ=− ∇·A (55)

~
Thus, the scalar potential can be expressed in terms of the vector potential A.
In other words, the scalar potential can be eliminated from the field equations
leaving only the dependence on A.~ As a result, we can express both E ~ and B~
~
in terms of the vector potential A alone. Substituting Eq. (55) into Eq. (54),
we obtain
2 ~
~ = c ∇(∇ · A)
E ~ − ∂A , B~ =∇×A ~. (56)
iω ∂t

Hence, both the electric and magnetic fields can be found from the vector
~
potential A.

This result may seem rather strange at first, since normally we should expect
to need both the scalar and vector potentials in order to completely determine
the EM fields. The explanation and in fact an another way of saying the same
thing is that time varying charge must satisfy the continuity equation
∂ρ
∇ · J~ = − = −iωρ ,
∂t
so that
∇ · J~
ρ=− .

We see that the time varying charge density can be expressed in terms of the
current density. Then the scalar potential becomes
1 Z ρ(t − r/c) 1 Z ∇ · J~
Φ(r, t) = dV = − dV
4πε0 r 4πε0 iω r
Z ~
1 J c2 ~.
= − ∇· dV = − ∇ · A
4πε0 iω r iω
Thus, specification of J~ alone is sufficient to completely determine all sources
of moving (oscillating) charges, and hence a solution for A ~ in terms of J~ con-
tains all the necessary information to completely specify the time-varying
EM fields.

153
11.1 Generation of electromagnetic waves
We have learnt that the introduction of the concept of displacement current
by Maxwell led to the prediction of electromagnetic waves in vacuum. Now,
we inquire into the sources of electromagnetic waves, i.e. how to generate EM
waves of different wavelengths and different properties, how to control these
properties and how to propagate these waves in desired directions.

11.1.1 Field of an Element of Alternating Current


Consider a linear element ∆l carrying an alternating current I = I0 exp(iωt),
as shown in Figure 39. The current element may be viewed as two charges Q
and −Q oscillating back and forth and can be served as an antenna, i.e. a
source of electromagnetic waves.
Assume that ∆l is much smaller than
the wavelength λ = 2πc/ω correspond-
ing to the frequency of the oscillation.
In this case, we can ignore the phase
variation of the current along ∆l. Of
course, this is an approximation that
may not hold for many realistic an-
tennas. However, an understanding
of the properties of such an antenna
is of great interest since, in principle,
all radiating structures can be consid-
ered as a sum of small radiating ele-
Figure 39: A linear alternating cur- ments. Moreover, many practical an-
rent element ∆l oriented in the z di-
rection. The generated fields are calcu- tennas working at low frequencies are
lated in the ~r direction. very short compared with the wave-
length.

For the fields around a small current element, ∆l  λ, there are three spatial
regions (zones) of interest:

• The near field (static) zone: ∆l  r  λ

• The intermediate field (induction) zone: ∆l  r ∼ λ

• The far field (radiation) zone: ∆l  λ  r

154
We will see that the fields have different properties in the different zones.
In the near zone the fields have the character of static fields, with a strong
dependence on the properties of the source. In the far field zone, the fields
are transverse to the radius vector and fall of as r−1 , typical of radiation
fields, and are independent of the source.

~ and B
Let us calculate the E ~ fields around the current element. We start by
considering the retarded current element that cab be written as
~ − r/c)dV = I(t
J(t ~ ,
~ − r/c)dl = I0 ei(ωt−kr) dl

where k = ω/c.

Since the current is the same at any point of the antenna (∆l  λ), it
~ and then the vector potential
appears as a constant for the integartion in A,
becomes

~= I0 ei(ωt−kr) ~
A ∆l ,
4πε0 c2 r
~ = R dl.
where ∆l ~

In the near zone, where r  λ (or kr  1), the exponent exp(−ikr) can be
replaced by unity. In the far field zone kr  1, the exponential oscillates
rapidly, and in this region it is sufficient to approximate exp(−ikr) ≈ 1−ikr.
In the intermediate zone, all powers of kr must be retained.

We now calculate directions and magnitudes of the electric and magnetic


fields produced by the current element. The fields are most easily evaluated
in spherical polar coordinates, if we choose the direction of the current ele-
ment ∆l~ along the z-axis, i.e. ∆l
~ = ∆l~k. We then have that in the spherical
coordinate system, the vector potential A~ has components

I0 ∆l cos θ ei(ωt−kr)
Ar = Az cos θ = ,
4πε0 c2 r
I0 ∆l sin θ ei(ωt−kr)
Aθ = −Az sin θ = − , (57)
4πε0 c2 r

Aφ = 0 .

155
~ and B,
According to Eq. (56), to find the fields E ~ we have to calculate ∇ · A
~
~ that in spherical polar coordinates are given by
and ∇ × A,

~= 1 ∂ (r2 Ar ) 1 ∂ (Aθ sin θ) 1 ∂Aφ


∇·A 2
+ + ,
r ∂r r sin θ ∂θ r sin θ ∂φ
and
" # " #
~ = r̂ ∂ (Aφ sin θ) ∂Aθ θ̂ 1 ∂Ar ∂ (rAφ )
∇×A − + −
r sin θ ∂θ ∂φ r sin θ ∂φ ∂r
" #
φ̂ ∂ (rAθ ) ∂Ar
+ − .
r ∂r ∂θ
Since Aφ = 0 and there is no φ dependence of Ar and Aθ , i.e. ∂Ar,θ /∂φ = 0,
the above equations reduce to

~ = 1 ∂ (r2 Ar ) 1 ∂ (Aθ sin θ)


∇·A 2
+ , (58)
r ∂r r sin θ ∂θ
" #
~ = ∂ (rAθ ) ∂Ar φ̂
∇×A − . (59)
∂r ∂θ r

~
Magnetic field B

Calculate first direction and magnitude of the magnetic field produced by the
antenna. Since, B ~ = ∇ × A,~ we easily find from Eq. (59) that the magnetic
field of the current element is
~ =∇×A
B ~ = Bφ φ̂ ,

where
" #
1 ∂ (rAθ ) ∂Ar
Bφ = − , (60)
r ∂r ∂θ
and Br = Bθ = 0.

Conclusion: The magnetic field produced by the antenna is perpen-


dicular to the radius vector at all distances.

156
Calculate the magnitude Bφ . Substituting Eq. (57) into Eq. (60), we obtain
   
i(ωt−kr)
−I0 ∆l  ∂ sin θe ei(ωt−kr) ∂ cos θ 
Bφ = +
4πε0 c2 r ∂r r ∂θ
ei(ωt−kr)
" #
I0 ∆l i(ωt−kr)
= ik sin θe + sin θ .
4πε0 c2 r r

Hence
" #
I0 ∆l ik 1
Bφ = + sin θei(ωt−kr) .
4πε0 c2 r r2

A number of interesting general conclusions follow from this equation. In the


first place, we note that the magnetic field is composed of two terms: the
near zone term ∼ 1/r2 and the far zone term ∼ 1/r. Secondly, we note that
in the limit of ω → 0, the magnetic field is composed of only the near zone
term that is the familiar Biot-Savart formula. Finally, the most important
is that the far zone term is only present for an oscillating field (ω 6= 0) and
therefore it represents a radiation field arising from accelerated (oscillating)
charge.

~
Electric field E

We now calculate direction and magnitude of the electric field produced by


the antenna, which as we have shown before can be found from the vector
potential
2 ~
~ = c ∇(∇ · A)
E ~ − ∂A ,
iω ∂t
where
~
∂A  
~ = iω Ar r̂ + Aθ θ̂ .
= iω A
∂t
~ Substituting Eq. (57) into Eq. (58), we get
We first calculate ∇ · A.

~ = 1 ∂ (r2 Ar ) 1 ∂ (sin θAθ )


∇·A 2
+
r ∂r r sin θ ∂θ
157
   
i(ωt−kr)
1 ∂ r cos θe i(ωt−kr) 
!
I0 ∆l  1 ∂ 2 e
= − sin θ
4πε0 c2  r2 ∂r r sin θ ∂θ r 

2 sin θ cos θei(ωt−kr)


( )
I0 ∆l cos θ h i(ωt−kr) i(ωt−kr)
i
= e − ikre −
4πε0 c2 r2 r2 sin θ
" #
I0 ∆l 1 ik
= − + cos θei(ωt−kr) .
4πε0 c2 r2 r

~ Since in spherical coordinates


Next, we take gradient of ∇ · A.

∂ 1 ∂ φ̂ ∂
∇ = r̂ + θ̂ + ,
∂r r ∂θ r sin θ ∂φ
we obtain for the components of the gradient:

ik k 2 ik i(ωt−kr)
" #
h 
~
i I0 ∆l cos θ 2
∇ ∇·A = − − − + − 2 e
r 4πε0 c2 r3 r2 r r
2ik k 2 i(ωt−kr)
" #
I0 ∆l cos θ 2
= + − e ,
4πε0 c2 r3 r2 r
" #
h 
~
i I0 ∆l 1 ik
∇ ∇·A = 2 3
+ 2 sin θ ei(ωt−kr) ,
θ 4πε0 c r r
h  i
~
∇ ∇·A = 0.
φ

Hence, the radial part of the electric field, Er , is


 
c2 h ~
~ −  ∂A 
i
Er = ∇(∇ · A)
iω r ∂t
r
2ik k 2 i(ωt−kr) I0 ∆l cos θ ei(ωt−kr)
" #
I0 ∆l cos θ 2
= + − e − iω
4πε0 iω r3 r2 r 4πε0 c2 r
" #
I0 ∆l cos θ 2 2 ik iω i(ωt−kr)
= 3
+ 2+ − e .
4πε0 c ikr r r cr

Since k = ω/c, the 1/r terms cancel and then Er simplifies to

I0 ∆l cos θ 2 2
 
Er = 3
+ 2 ei(ωt−kr) . (61)
4πε0 c ikr r

158
Similarly, we find the polar component of E as
 
c2 h ~
~ −  ∂A 
i
Eθ = ∇(∇ · A)
iω θ ∂t
θ
I0 ∆l sin θ ei(ωt−kr)
" #
I0 ∆l sin θ 1 ik i(ωt−kr)
= + e + iω
4πε0 iω r3 r2 4πε0 c2 r
" #
I0 ∆l sin θ 1 1 ik i(ωt−kr)
= 3
+ 2+ e . (62)
4πε0 c ikr r r

The azimuthal component of the electric field Eφ = 0.

Note that the radial part of the electric field, Eq. (61), contributes only to
the near and intermediate zones, whereas the angular polar part, Eq. (62),
contributes to all of the zones.

Theorem:

The near-zone 1/r3 part is the Coulomb type contribution. It is similar in


nature to a static field surrounding a small linear-current element and an
electric dipole.

Proof:

The Coulomb or static field is for ω → 0. In this limit the 1/r3 contribution is

I0 ∆l sin θ 1 i(ωt−kr)
Eθ = e
4πε0 c ikr3
I0 ∆l sin θ 1 1
 
2
= 1 − ikr + (−ikr) + · · ·
4πε0 c ikr3 2
I0 ∆l sin θ 1
= − ,
4πε0 c r2
where we have taken only the real (physical) part of the field.

Since I0 = ∆q/∆t and ∆l/∆t = c, we get

∆q sin θ 1
Eθ = − ,
4πε0 r2

159
as required.

Electric and magnetic fields in near and far field zones

Consider first the near field zone (r  λ). In this limit the magnetic and
electric fields are
B~ near = I0 ∆l 1 sin θei(ωt−kr) φ̂ ,
4πε0 c2 r2
~ near = −i I0 ∆l 2 ei(ωt−kr) cos θr̂ + sin θθ̂ .
 
E
4πε0 c kr3
Since the magnetic field is real and the electric field is imaginary, the Pointing
vector involving the near-zone field components is a pure imaginary quantity.
It does not represent any flow of energy. This imaginary quantity represents
energy that oscillates back and forth between the source and the region of
space surrounding the source.

Consider now the far zone or radiation components of the magnetic and
electric fields.
~ rad = EθR θ̂ , I0 ∆l sin θ ik i(ωt−kr)
E EθR = e ,
4πε0 c r
~ rad = BφR φ̂ , I0 ∆l sin θ ik i(ωt−kr)
B BφR = e . (63)
4πε0 c2 r
Note the following properties of the radiation components:
1. The electric and magnetic fields oscillate in phase.
2. The electric and magnetic fields are orthogonal to each other.
EθR
3. The ratio BφR = c, the value for plane waves in free space.

4. The electric and magnetic fields are transverse to the radius vector at
all distances.
5. The Poynting vector N ~ = c2 ε 0 E
~ rad × B
~ rad is a real quantity and is in
the direction of the radius vector, indicating that the energy of the field
propagates away from the current element.
These properties show that in the far zone the EM field is in a form of plane
waves propagating with the speed of light c.

160
11.2 Power Radiated from the Current Element

Consider now the radiation power emitted by the antenna. By the radiation
power we mean that part of energy which is carried by the radiation com-
ponents of the field. We will show that the radiation power is equal to the
radiation losses, i.e. energy carried by the plane electromagnetic wave that
propagates on its own independent of the source.

The power flux at any point is given by the Poynting vector


~ = c2 ε 0 E
N ~ ×B
~ .

Then, the total power radiated across a sphere of radius r is


I
W = N ~ ,
~ · dS
S

where
dS = r2 sin θ dθdφ .
Only those partial products in E ~ ×B~ which vary as 1/r2 will have net radi-
ated power. The other partial products are small as they fall off more rapidly
than 1/r2 . Thus, the only part of the fields entering into the expression for
the radiated power is the far field zone part (radiation component) consisting
of the terms varying as 1/r.

The radiation components of E ~ and B~ oscillate in phase as sine or cosine


functions. Thus, an average of their product over time is
1  R  R
E R¯B R = E B ,
2 θ 0 φ 0
   
where EθR and BφR are amplitudes of EθR and BφR .
0 0

On substituting from Eq. (63), we find that the time averaged Poynting
vector is
1 2 I 2 ∆l2 k 2
N̄ = c ε0 0 2 2 3 2 sin2 θ
2 16π ε0 c r
#2
I02 ∆l2 4π 2 I02 ∆l sin2 θ
"
2
= sin θ = .
32π 2 ε0 c λ2 r2 8ε0 c λ r2

161
Hence, integrating over all directions, we get the total power emitted by the
antenna
Z π Z 2π
W = r2 N̄ sin θ dθdφ
0 0
#2
I02 ∆l
Z 2π " Z π
= dφ sin3 θdθ .
0 8ε0 c λ 0

Performing the integration over φ that is equal to 2π, and over θ that is equal
to 4/3, we get
#2 #2
I 2 ∆l πI02 ∆l
" "
4
W = 2π 0 = .
8ε0 c λ 3 3ε0 c λ

We can write the total power radiated in terms of the power absorbed in an
equivalent resistance, called the radiation resistance, as
" #2
1 2π ∆l 1
W = I02 = RI02 .
2 3ε0 c λ 2

where
" #2
2π ∆l
R=
3ε0 c λ

is the radiation resistance.


q
Since 1/(ε0 c) = µ0 /ε0 = 377, or 120π, we obtain for R:
" #2
2 ∆l
R = 80π .
λ

For example, if ∆l/λ ≈ 0.1, then R = 0.8π 2 ≈ 8 ohms.


This example shows that for a current element which is 10% of the wave-
length long, the resistance is very small.

Thus, if ∆l/λ  1, the radiation losses are negligible, that the radiated power
is very small. In terms of the emitted radiation, the emitted EM waves are
very weak in power.

162
With a short linear current element, an appreciable power would be radiated
only if the current amplitude I0 were very large. A large current, on the other
hand, would lead to large amounts of power dissipation in the conductor, and
hence a very low efficiency.

We can conclude that current carrying systems that have linear dimensions
small compared with the wavelength radiate negligible power. An efficient
antenna should have dimensions comparable to or greater than the wave-
length.

11.3 Gain of the Dipole Antenna

A further property of the dipole antenna that is worthy of consideration is


the directional property of power radiated in different directions.

The gain or directivity function of a transmitting antenna is the ratio of the


Poynting flux to the flux due to an isotropic radiator emitting the same total
power W :

N (θ, φ)
gT (θ, φ) = ,
Niso
where
W
Niso =
4πr2
is the energy flux uniform in all directions.

Since for the infinitesimal dipole:


#2
I 2 ∆l sin2 θ
"
N (θ, φ) = 0 ,
8ε0 c λ r2

and
#2
πI02 ∆l
"
W = ,
3ε0 c λ

163
we obtain
3 2
gT (θ, φ) = gT (θ) = sin θ .
2
The directivity function gT (θ, φ) defines a three-dimensional surface called
the polar radiation pattern of the antenna.
The function varies as sin2 θ, and
hence the radiation is most intense
in the θ = π/2 direction (perpen-
dicular to the axis of the dipole)
and zero in the directions θ = 0, π
(on the axis of the dipole). The
maximum gain then is 1.5 for di-
rections defined by θ = π/2, in
Figure 40: The polar radiation pattern of the equatorial plane of the dipole.
the dipole antenna.
The gain function is independent
of φ.

The spatial distribution of the radiated power can be shown on a polar di-
agram, such as in Figure 40, which gives the relative values of the radiated
power at different positions on the surface of a sphere centered on the dipole.

We can conclude, that the directivity function gT (θ, φ) is a measure of how


effective the antenna is in concentrating the radiated power in a given direc-
tion.

164
Revision questions

Question 1. Under what condition the fields can be expressed solely by the vec-
tor potential?

Question 2. What do we mean by ”near zone” and ”far field zone”?

Question 3. What are the properties of the EM field in the far field zone?

Question 4. Show that the power radiated by an antenna is equal to the radia-
tion losses.

Question 5. Why emitted waves of a short antenna are very weak?

Question 6. What is the spatial distribution of the power emitted by a short


antenna?

165
12 Electromagnetic Theory of Polarizable
Materials

Condense matter physics and electromagnetism theories used to be regarded


as almost separate subjects. However, in next few lectures, we will illustrate
how matter and electromagnetism are integrated and how the electromag-
netic theory may be applied to condense matter physics to explain the prop-
erties of conducting materials and insulators, or dielectrics as they are often
called. Materials such as conductors or insulators are composed of atoms
that contain free and bounded electric charges. The charges can be redis-
tributed by an application of external fields. We will focus on ”new” fields
induced by redistributed electric charges and will show how these problems
can be solved with the help of only the scalar potential Φ.

We proceed our analysis by introducing the microscopic (atomic) model of


materials and for this purpose simplified models of the atom will be used.
In this simplified model, the atom consists of a positively charged nucleus
surrounded by a spherically symmetric cloud of electrons. In the absence of
an external electric field, the electrons and nucleus are in a stable equilib-
rium. What happens if the atom is placed in an external electric field? If
an uncharged dielectric (insulator) is placed in an electric field, the field will
redistribute the charges within the dielectric. An effective non-zero charge is
induced by rearrangement of bound charges within the atoms or molecules
of the dielectric. In dielectrics these charges are a set of molecular dipoles
that, in turn, will set up a secondary (induced) field so that the net field will
be modified from its original (external) value. We will show how the modifi-
cation of the field led to the concept of an another field vector; the dielectric
displacement vector D. ~ The advantage of using the dielectric displacement
vector and its physical properties will be explored.

In materials, electric dipoles may exist permanently or may be induced by


the external field. Dielectrics with permanent dipole moments are usually
electrically neutral due to random orientation, in the absence of electric fields,
of the dipole moments. An example is H2 O. The application of an external
electric field causes all molecules composing the dielectric to align themselves
with the external field. This produces a macroscopic dipole moment. We will

166
study electromagnetic theory of polarizable materials in terms of microscopic
objects, electric dipoles, which we will treat as building blocks of dielectric
materials. As we shall see, despite this simplicity, the model satisfactorily
predicts the macroscopic behavior of dielectric materials.

12.1 Potential and Electric Field of a Single Dipole


Mathematically, for the calculations of the field of distributed charges in-
side a dielectric material, it is convenient to deal with a separate object, the
dipole, that is composed of two bounded charges of oposite signs separated
by a small distance, not as just a pair of individual plus and minus charges.

Suppose that two opposite sign point


charges ±q are separated by a dis-
tance d. We will find the potential Φ
~ at a point A dis-
and the electric field E
tance r and angle θ under the assump-
tion that the distance r is much larger
than the separation between the charges,
i.e. r  d. The potential Φ pro-
duced by two separated charges of op-
posite sign is called the dipole poten-
Figure 41: A schematic model for tial.
the calculation of the potential pro-
duced by an electric dipole.
We define electric dipole moment as the
product of the charge times the separation.
It is a vector that points from the negative charge to positive charge
p~ = q d~ .
Since r2 − r1 ≈ d cos θ at distances r  d and d cos θ = p~ · r̂, we get for the
potential at the point A:
q q
Φ = −
4πε0 r1 4πε0 r2
q r2 − r1 q d cos θ 1 p~ · r̂
= = ≈ . (64)
4πε0 r1 r2 4πε0 r1 r2 4πε0 r2
Hence, we can find electric field of the dipole using the relation E~ = −∇Φ.
Since the potential of the dipole depends on r and θ, it is convenient to work

167
in the spherical coordinates in which the electric field is given by
~ = Er r̂ + Eθ θ̂ + Eφ φ̂ ,
E

where the components are


∂Φ 1 2p cos θ
Er = − = ,
∂r 4πε0 r3
1 ∂Φ 1 p sin θ
Eθ = − = ,
r ∂θ 4πε0 r3
1 ∂Φ
Eφ = − =0.
r sin θ ∂φ
Figure 42: The electric field lines of a
dipole moment. Hence
~ = p  
E 2 cos θ r̂ + sin θ θ̂ .
4πε0 r3
One can note that the field has az-
imuthal symmetry. The field is pro-
portional to (1/r)3 so that it falls
rapidly with distance. Moreover, in the direction perpendicular to the dipole
moment, θ = 90◦ , the field points in the opposite direction to the dipole
moment at all distances. Figure 42 shows a sketch of the electric field lines
of an electric dipole moment.

12.2 Polarization Vector


If there are N dipoles per unit volume of a dielectric, the total dipole mo-
ment is:
N
P~ =
X
p~i ,
i=1

and is called the polarization.

The electric potential set up in the space by a dielectric material with an


arbitrary volume distribution of electric dipoles can be calculated by using
the potential produced by a single dipole, Eq. (64), and the above definition

168
of P~ . The electric potential at an arbitrary point A distance r from a volume
element dV containing such dipoles is

1 P~ · r̂
dΦ = dV ,
4πε0 r2
where r̂ is the unit vector from dV towards A, and we have assumed that r
is much larger than the extent of the volume element dV .

Let r̂ be the unit vector from A to-


wards dV . (We want to integrate over V
with the position of A fixed). In this case,
we change r̂ → −r̂, and obtain
!
1 ~ r̂
dΦ = P· − 2 dV ,
4πε0 r

and then the potential measured in the


direction towards the volume V is
!
1 Z ~ r̂
Figure 43: Schematic diagram for cal- Φ= P · − 2 dV .
culations of the scalar potential pro-
4πε0 r
V
duced at point A by a macroscopic ma-
terial of the polarization P~ . It is difficult to interprete this form of
the potential as it is not in the form of
a potential known from the electrostat-
ics.

However, the result can be transformed into a form that can be interpreted
with the help of the general form of the static potentials.

First, noting that


r̂ 1
− 2
=∇ ,
r r
we have that
!
r̂ 1
P~ · − 2 = P~ · ∇ .
r r

169
~ = Φ∇ · A
Next, applying a vector identity ∇ · (ΦA) ~+A
~ · ∇Φ, we can write
the above expression as
 
1 P~ ∇ · P~
 
P~ · ∇ =∇· − .
r r r

Hence, we can write the potential as a sum of two terms both involving
the 1/r function
   
1 Z P~ 1 Z  ∇ · P~ 
Φ= ∇ ·   dV + − dV .
4πε0 r 4πε0 r
V V

Note that the second term varies with r explicitly as the 1/r function, but
in the first term, the 1/r function is under the nabla operator. However, we
can use the Gauss’s divergence theorem to transform the first term into the
surface integration. In this case, the first term becomes a simple function
of 1/r, so that the potential takes the form
 
1 I
P~ · n̂ 1 Z  ∇ · P~ 
Φ= dS + − dV , (65)
4πε0 r 4πε0 r
S V

where n̂ is the unit vector normal to the surface of the material.

Figure 44: Surface (left picture) and volume charges (right picture).

On comparing with the general form of the static potential, Eq. (48), the ex-
pression (65) can be interpreted as follows. The first term on the right-hand

170
side, a surface integral, is a potential equivalent to that of a surface charge
density σs = P~ · n̂. The second term is a potential equivalent to that of a
volume charge density σV = −∇ · P~ .

Where did this interpretation came from? It can be explained by using a


simple picture shown in Figure 44. Surface charges exist because there are
no neighboring charges at the end surfaces of the material to cancel them
out. Volume charges exist because the number of dipoles per unit volume
changes, that there is an incomplete cancellation of charge density from ad-
jacent dipoles.

Materials which have a non-zero volume charge density are called inhomo-
geneous materials. Thus, a sufficient condition for a material to be homoge-
neous is that the polarization of the material have a zero divergence.

12.3 ~ in a Dielectric
Maxwell’s Equation for ∇ · E

We have just seen that a non-zero potential outside a dielectric material is


produced by surface and volume charges. A question then arises: What is
the electric field inside and outside the dielectric when extra polarization
charges are present?

In general, the electric field in the dielectric can be found from the Maxwell’s
equation I:

~ = ρ
∇·E ,
ε0

where ρ is the total volume charge density inside the dielectric.16

In the dielectric it is convenient to (mentally) separate the polarization


charges from whatever other charges might be there also. The other charges
are usually referred to as the free charges or conducting charges to distin-
guish them from the bound charges in the dielectric.

16
Why the surface charge density is not taken into account in the Maxwell’s equation I?

171
We may write the Maxwell’s equation I as
~ = ρf + ρp ,
ε0 ∇ · E

where ρf is the ”free” charge density and ρp is the polarization charge density
throughout the volume. If we express ρp in terms of P~ , i.e. if we write
ρp = −∇ · P~ , we obtain

~ = ρf − ∇ · P~ .
ε0 ∇ · E

Hence
~ + P~ ) = ρf .
∇ · (ε0 E (66)

Now it is common practice to drop the subscript f , but one must remember
that ρ now stands for the charge density not counting the polarization charges.

Next, we can define a new vector


~ = ε0 E
D ~ + P~ , (67)

called the dielectric displacement field,17 and write Eq. (66) as

~ =ρ.
∇·D
~ inside a dielectric is
We can read this equation that the source of the field D
the free charge density ρ.

17
The reason for the name ”dielectric displacement” can be easily understood if we refer
to the Maxwell’s theory. Take time derivative of both sides of Eq. (67):

~
∂D ~
∂E ∂ P~
= ε0 + .
∂t ∂t ∂t
We know from the Maxwell’s theory that the first term on the rhs of the above equation
represents displacement current density, and the second term is the polarization current
~
density. Therefore, ∂ D/∂t can be called a generalization of the displacement current
density, and then D~ can be regarded as the dielectric displacement.

172
12.4 Macroscopic Effects of the Polarizability

We turn now to a consideration of the macroscopic effects of the polarizabil-


ity of dielectric materials. We will consider only ideal dielectrics.

Ideal dielectrics can be divided into following categories:


1. Homogeneous – properties independent of the position.

2. Isotropic – properties independent of direction.


~
3. Linear – polarization proportional to E.

4. Stationary – properties independent of time.

Case of simple isotropic and linear dielectrics

Ordinary dielectrics (glass, teflon, plastics etc.) are linear in polarization for
fields not strong enough to cause dielectric breakdown i.e. P~ ∝ E. ~
For these materials, we can write

P~ = N αE
~ = χε0 E
~ ,

where α is the polarizability of a single atom (molecule), and χ is the dielec-


tric susceptibility. Then, we can write the dielectric displacement as
~ = ε0 E
D ~ + P~ = ε0 (1 + χ)E
~ = ε0 εr E
~ = εE
~ ,

where εr is the relative permittivity or dielectric constant, and ε is the per-


mittivity. Hence
~ = ∇ · (εE)
∇·D ~ =ρ

or
~ = ρ
∇·E ,
ε
if ε is independent of position i.e. if ε is a permittivity of a homogeneous
dielectric.

173
The same result would be obtained by replacing ε0 in Coulomb’s law by ε.
1 q1 q2
F~ = r̂ .
4πε r2
The ratio ε/ε0 then represents the relative shielding of q1 from q2 by the
polarization charges induced in the medium.

How do we determine the permittivity of a given material?

The theory of the molecular structure of a material will yield an estimate


of χ and ε. The dipole moment p~ of a molecule will be proportional to
~ so we can define a molecular susceptibility χm such
the local electric field E
that p = χm ε0 E. Then

P~ = ~ ,
X
p~i = N χm ε0 E
i

where N is the number of molecules per unit volume. Thus

ε = ε0 εr = ε0 (1 + N χm )

and ε is a quantity directly measurable from the measurement of capacitance


εA ε0 A
C= = εr = εr C 0 ,
d d
where C0 is the capacitance without the dielectric.

We can summarize: Filling a capacitor with dielectric multiplies its ca-


pacitance by εr .

Exercise in class: Capacitor filled with a homogeneous dielectric

A plane parallel capacitor has charges +σ and −σ per unit


area on its plates. The capacitor is filled with a homoge-
neous and linear dielectric of dielectric constant εr = 1 + χ.
Show that:

174
(a) The electric field within the dielectric is:
σ
E= .
εr ε0

(b) The polarization charge per unit area on the surface


of the dielectric adjacent to the surface of the negatively
charged plate is:
χσ
σs = .
1+χ

(c) The capacitance of the capacitor is C = εr C0 where C0


is the capacitance of the same capacitor without the dielec-
tric (i.e. a vacuum or air between the plates).

175
12.5 Dense Dielectrics: The Clausius-Mossotti Rela-
tion
In the standard calculations of the polarization of macroscopic materials, it is
~ is the same at any point of the material. How-
often assumed that the field E
ever, for an extended dense materials the field E~ may vary with the position
and the calculation of the dielectric constant of the material may not agree
with an experimental measurement. This is what really happens. Therefore
the approach of a constant field through the whole area of the material must
be modified.

It was Lorentz, who proposed an approach that resolved this problem. The
Lorentz theory of polarizability of dense dielectric materials distinguishes be-
tween the mean electric field E~ and the local electric field E ~ loc as seen by a
typical dipole. The typical dipole is considered to be at the center of a small
sphere that has been excavated from the dielectric. E ~ loc is thus the mean
~ ~ ~
field E minus Eplug where Eplug is the field of the spherical volume excavated.

Let us see what would be the field at this area if the dipole is not there.

Let ~ be the mean field throughout the dielectric.


E
Let ~
Eplug be the field due to the spherical plug alone.
Let ~ loc be the field in the spherical hole.
E
~
~ =E
E ~ loc + E
~ plug and ~ plug = − P .
E
3ε0

Thus, the field acting to polarize the molecule is equal to E~ minus the con-
tribution to the total average field from the molecule itself. Hence
~
~+ P .
~ loc = E
E
3ε0
The argument now is that each molecule is at the centre of a small hole
~ loc . If α is the molecular
and the field acting on the molecule is thus E
polarizability, its induced dipole is thus:
~ loc
p~ = αE

176
If there are N molecules per unit volume then:
 
~
P~ = N p~ = N αE
~ loc ~+ P  .
= N α E
3ε0

By definition: P~ = χε0 E
~ = (εr − 1)ε0 E.
~ Substituting for P~ in the above
equation, we obtain
" #
(εr − 1)ε0 E ~ + (εr − 1)ε0 E
~ = Nα E ~ .
3ε0
Hence
εr − 1 Nα
 
(εr − 1)ε0 = N α 1 + = (εr + 2) ,
3 3
and finally
εr − 1 Nα
= .
εr + 2 3ε0
This is known as the Clausius-Mossotti relation for a dense dielectric. Assum-
ing that α is known, we can compute εr - dielectric constant of the material.

The Clausius-Mossotti relation works pretty well for most of dense materials.
However, it has a weak point. To illustrate this, we solve the Clausius-
Mossotti equation for εr . Since

εr − 1 = (εr + 2) ,
3ε0
we obtain for εr :
2N α Nα
   
εr = 1 + / 1− .
3ε0 3ε0
Note that
2N α
1+ 3ε0 Nα
εr = Nα → ∞ as →1.
1 − 3ε0 3ε0

We see that one can adjust the number of the dipoles to get an infinite di-
electric constant of a finite material. This is called the Clausius-Mossotti

177
catastrophe.

This effect can be explained as follows: Removal of the plug leaves polariza-
tion charges, whose field tends to line up the dipole parallel to the field. The
system is self-polarizing (Clausius-Mossotti catastrophe) if N α
3ε0
→ 1, i.e. the
system induces P~ without an external field.

12.6 Dielectric in a Time Dependent Field


Let us consider what will happen if a dielectric is introduced into an alter-
nating electric field. The student immediately conclude that the alternating
electric field will create oscillating dipole moments inside the material, or
equivalently, an oscillating polarization of the material. This is true, but
then an another question arises: How does the polarization follow the oscil-
lating external field?

To answer this question, let us


consider an experiment, illustrated
in Figure 45, involving a plane
plate capacitor filled with a di-
electric and connected to an os-
cillating AC current. The alter-
nating current will create an oscil-
lating electric field inside the ca-
pacitor which will cause the elec-
tric dipoles of the dielectric to os-
cillate in time. On the other
Figure 45: A plane plate capacitor filled hand, the oscillating electric field
with a dielectric and connected to an alter-
nating current. will induce oscillating polarization
charges.

What are the oscillating polarization charges equivalent to? The induced
polarization charges do not produce any currents inside the dielectric. There
is no DC current in response to a DC electric field, but if P~ is changing with

178
~ is changing with time) there will be an AC current density:
time (because E

∂σs ∂ P~
J~ = n̂ = .
∂t ∂t
Thus, ∂ P~ /∂t plays the role of polarization current density.

With the alternating electric field present, the polarization P~ may lag in
~ This means there is internal friction and
phase behind the driving field E.
heat dissipation. If it happen, the capacitor will exhibit resistive as well as
capacitive properties.

Since resistivity leads to a dissipation of the energy, we proceed by considering


the work done in charging the capacitor. It is defined as
dW dQ
= V I = Ed ,
dt dt
where V is the voltage.

Thus, we have to find how quickly the charge on the plates is changing in
time.

Since we are interested in the time variation of the electric field and the
polarization, we express the charge in terms of the field quantities
εA
Q = CV = CEd = Ed = AεE = AD ,
d
from which, we obtain

Q = A(ε0 E + P ) .

Then work in terms of the field variables is


dW d d
= Ed A(ε0 E + P ) = E dA (ε0 E + P ) .
dt dt dt
Since dA = V is the volume of the capacitor, we can write
" #
dW Z
d dP
= E (ε0 E) + E dV ,
dt dt dt
V

179
or
dW d 1 dP
Z  Z 
2
= ε0 E dV + E dV ,
dt dt 2 dt
V V

where we took into account a possibility that the electric field and polariza-
tion can vary across the capacitor’s plates.

~
The first term in the above equation is the rate of doing work building up E
field.

~
The second term is the rate of doing work on the dipoles by E.

Thus, the supplied energy to the capacitor is used to build up the electric
field inside the capacitor and to polarize the dielectric. Consider separately
both terms.

First term:

If E = E0 cos(ωt), the first term takes the form


Z
(−E0 cos(ωt) ε0 ωE0 sin(ωt)) dV .
V

The element of work done per unit volume and unit time is
dW
= −ε0 ωE02 cos(ωt) sin(ωt) .
dV
Averaging over a cycle, we get
2π/ω
dW 2
Z
= −ε0 ωE0 cos(ωt) sin(ωt) dt = 0 .
dV
t=0

No energy has been used to build up the electric field inside the capacitor.
Work is done building up the field in one part of the cycle but the stored
energy is given back in another part.

Second term:

180
If P = χε0 E = χε0 E0 cos(ωt) the same zero net energy conversion averaged
over a cycle will happen with this term. If there is internal friction there will
be a phase difference between P and E, that P does not follow the changes
in E. If we write
P = χε0 E0 cos(ωt + φ) ,
where φ represents a phase difference between P and E, we get for the
polarization
P = χε0 E0 cos φ cos(ωt) − χε0 E0 sin φ sin(ωt) .
Taking the time derivative, we get
dP
= −ωχε0 E0 cos φ sin(ωt) − ωχε0 E0 sin φ cos(ωt) .
dt
Hence, the work done per unit volume per cycle will be
2π/ω
dW Z
= − ωχε0 E02 cos φ cos(ωt) sin(ωt) dt
dV
0
2π/ω
Z
− ωχε0 E02 sin φ cos2 (ωt) dt .
0

Since the first integral is zero, we get


2π/ω
dW 2
Z
= −ωχε0 E0 sin φ cos2 (ωt) dt . (68)
dV
0

The integral on the rhs of the above equation is positive and dW/dV must
be positive corresponding to energy dissipation (or the dielectric would keep
getting energy from its interior and building up the field with it).

Thus, sin φ must be negative, so −π < φ < 0. This means that the polariza-
tion lags in phase the electric field.
P = χε0 E0 cos(ωt − φ) .
In summary: Equation (68) shows that an energy is lost in each cycle of the
oscillations. In practice, it is dissipated as a heat in the material. The energy
loss is caused by the work required to change the polarization of the material.

181
12.7 The Complex Susceptibility and Permitivity

We have learnt that the polarization of a linear dielectric is proportional to


the field and the proportionality is determined by the dielectric constant or
susceptibility. In the previous lecture, we have modeled the fact that in real
dielectrics, the polarization may lag in phase behind the driving field E~ by
introducing the phase difference φ between P~ and E.
~ We may now ask: If P~
lags in phase behind E, how does then the relation between P~ and E
~ ~ is af-
fected?

To answer this question, it is convenient to use complex exponentials to


represent amplitude and phase of an oscillating quantity. We write
~ =E
E ~ 0 eiωt and P~ = P~0 ei(ωt−φ) , (φ is positive) .
We can write the complex polarization in different forms
P~ = P~0 e−iφ eiωt = (P~0 cos φ − iP~0 sin φ)eiωt ,
 
P~0 P~0
P~ =  cos φ − i sin φ E0 eiωt ,
E0 E0

P~ = ε0 (χ0 − iχ00 )E
~ 0 eiωt = ε0 χc E0 eiωt = ε0 χc E
~ .

The relation between P~ and E


~ is the same as before for a constant external
field. However, now χc = χ0 − iχ00 is a complex quantity called a complex
susceptibility.

Thus, the phase difference between the polarization and the exter-
nal field leads to a complex susceptibility of a dielectric material.
In other words, the internal friction of the material results in a
complex susceptibility.

With the complex polarization, the dielectric displacement takes the form
~ = ε0 E
D ~ + P~ = ε0 E
~ + ε0 χ c E
~ = ε0 (1 + χc )E
~ ,

or
~ = ε0 (1 + χ0 − iχ00 )E
D ~ = ε0 εr E
~ = εc E
~ ,

182
where εc is a complex permittivity, and εr is a complex relative permittivity
or dielectric constant

εc = ε0 (1 + χ0 ) − iε0 χ00 .

How do we in practice estimate losses in a given material? In other words,


how do we calculate χ00 for a given material?

The imaginary part of εr , χ00 , can be determined from the following experi-
ment involving a capacitor filled with a dielectric.
As we have just shown, the imag-
inary part of the dielectric suscep-
tibility represents losses, i.e. cor-
responds to net energy dissipation.
In the circuit theory language the
imaginary component of the dielec-
tric susceptibility adds a resistive
component to the capacitor. The
material filling the capacitor could
Figure 46: also have some ordinary ohmic con-
ductivity (due to the presence of
‘free’ charges as well as ‘bound’ charges in the material). Let us calculate
the magnitude of the total resistance of the dielectric. Simply, we will con-
sider the Ohm’s law for the capacitor and will find the relation between an
external current supplied to the capacitor and voltage. The relation will give
us an information about the resistance of the capacitor.

Let I be the oscillating current in the external circuit, and

~ = A dP
Z
Ip = J~p · dA
dt
be the polarization current in the dielectric. Let

~ = AσE = Aσ V
Z
Ic = J~c · dA
d
be the conduction current in the dielectric due to its finite conductivity σ.

183
If σs is the charge density on the plates, then the effective charge on the
plates is
Q = σ s A = qI − qp − qc ,
where qI is the charge supplied by I, qp is the charge removed by the polar-
ization current Ip , and qc is the charge removed by the conduction current Ic .

We can write the effective charge as


Z Z Z
Q= I dt − Ip dt − Ic dt .

Since the electric field inside the capacitor is


σs V
E= = ,
ε0 d
we can find the surface charge density
V
σs = ε0 .
d
and write the effective charge as
V Z Z Z
Q = σs A = ε0 A = Idt − Ip dt − Ic dt .
d
By taking a derivative in time of the both sides of the above equation, and
substituting for P the relation, P = χc ε0 E = χc ε0 Vd , we get
A dV χc ε0 dV σV
ε0 = I − Ip − Ic = I − A −A . (69)
d dt d dt d
Putting
dV
= iωV , χc = χ0 − i χ00 ,
dt
and solving Eq. (69) for I, we get
ε0 A σ
 
I= (1 + χ0 − iχ00 )iω + V .
d ε0
Separating real and imaginary parts and putting C0 = ε0 A/d (the capaci-
tance there would be if the dielectric were lossless), we find
σ
   
I = C0 + ωχ00 + iωC0 (1 + χ0 ) V .
ε0

184
Since the capacitor transmits some
charges through the internal dielec-
tric, in the circuit theory this system
is equivalent to a parallel circuit, as
shown in Figure 47. For the parallel
circuit
1
 
I= + iωC V .
R
Comparing with the above result for
Figure 47: current flow in the lossy capacitor
we see that the effective capacitance
0
is C0 (1 + χ ) and the effective resistance is
1
R= 
σ
 ,
C0 ε0
+ ωχ00

where we remember that C0 is the capacitance in the absence of losses.

12.8 The Loss Tangent

The properties of a dielectric material are usually specified by giving its di-
electric constant K, and its loss tangent tan δ.

We can write the general Ohm’s law as


σ
  
0 00
I = iωC0 (1 + χ ) − i χ + V .
ε0 ω
The quantity in [ ] brackets can be defined as a complex relative permittiv-
ity εr :
σ
 
0 00
εr = 1 + χ − i χ + .
ε0 ω
This is a generalization on the previous definition of complex relative per-
mittivity to include the effects of ohmic conductivity. Then we can define a

185
generalized permittivity and dielectric constant ε = ε0 εr .

Now if we write

εr = K 0 e−iδ = K 0 (cos δ − i sin δ)

= K 0 cos δ(1 − i tan δ) = K(1 − i tan δ) .

Then the standard form for permittivity is

ε = ε0 K(1 − i tan δ) ,

where tan δ is the ‘loss tangent’.18

Since K 0 cos δ = (1 + χ0 ) and K 0 sin δ = (χ00 + σ


ε0 ω
) it follows that:
h i
χ00 + σ
ε0 ω
tan δ = .
[1 + χ0 ]

In this equation, tan δ includes the effects of finite conductivity and the effects
of polarization damping force.

18
In practice, it is read out on some AC bridges as an alternative to reading out the
resistive property of a lossy capacitor.

186
Revision questions

Question 1. How do we calculate the electric field of a microscopic dipole mo-


ment?

Question 2. What are the surface and volume charge densities?

Question 3. Explain the advantage of introducing the concept of dielectric dis-


placement field.

Question 4. How do we calculate the polarization vector of a dense dielectric?

Question 5. In real dielectrics, does the polarization vector follow the changes
of an external electric field?

Question 6. How are losses in a dielectric represented in the permitivity of the


dielectric?

Question 7. What it meant by the loss tangent of a medium?

Tutorial problems

Problem 12.1 Capacitor with 2 different dielectrics

Consider a capacitor in which there are two different dielectrics


of thickness d1 and d2 having dielectric constants ε1 and ε2 . The
separation of the plates is then d1 + d2 .

(a) Calculate the polarization charge distributions everywhere.


Show that at the junction of the two dielectrics there is a surface
charge density:
ε1 − ε2
σs = σ.
ε1 ε2

(b) Show that the capacity of the capacitor, if the area of the

187
plates is A, is:
ε1 ε2 ε0 A
C= .
ε2 d1 + ε1 d2

Problem 12.2 Capacitor filled with an inhomogeneous dielectric

A parallel plate capacitor has charges +σ and −σ per unit area


on its plates which are separated by a distance d. The volume be-
tween the plates is filled with an inhomogeneous dielectric. The
dielectric susceptibility is zero at the positively charged plate and
increases linearly with distance, reaching its maximum value of
unity at the negatively charged plate.

If x is the coordinate measured from the positive plate toward


the negative plate and α is the gradient of dielectric susceptibil-
ity, i.e. χ = αx, show that:

(a) The electric field through the dielectric varies like:


σ
E= 0≤x≤d.
ε0 (1 + αx)

(b) There is a volume charge distribution throughout the dielec-


tric given by:
ασ
ρ=− ,
(1 + αx)2
together with a surface charge distribution on the dielectric adja-
cent to the negative plate given by:
σ
σ1 = .
2

(c) The capacitance of the capacitor is:


ε0 αA
C= .
ln 2

Hint :
You can see there will be a volume charge distribution within the
dielectric because:
ρ = −∇ · P~ = −∇ · (χε0 E)
~ 6= 0 , because χ = αx .

188
Also since this is the only charge in the region the polarization
charge density ρ also satisfies the general Maxwell equation:
∇·E ~ = ρ .
ε0

From these two equations you can develope a differential equation


describing the variation of the electric field across the capacitor.
Notice that it is ε0 we use in the equations (not ε) because we are
treating the polarization charges explicitly.
After solving for the electric field E(x) one can calculate the
charge distribution from:
ρ = −∇ · P~ = −∇ · (χε0 E)~ .

Problem 12.3 Equivalent circuit of a lossy capacitor

So-called electrolytic capacitors are made by electrolytic deposit


of a thin layer of dielectric on aluminium. Consider a 1000 µF
capacitor that might be used in smoothing the ripple in a DC
voltage supply generated by rectifying the mains (50 Hz) supply.
The dielectric has a complex dielectric constant:
εr = 1 + χ0 − iχ00 = 5 − i 10−4
and a conductivity of σ = 2 × 10−13 [mho m−1 ].

What is the equivalent parallel resistance of the capacitor at fre-


quency 50 Hz? and at 1 kHz?

Problem 12.4 Field at the centre of a uniformly polarized sphere

Consider a dielectric sphere with uniform polarization P~ through-


out. The equivalent polarization charges will be a volume charge
density −∇ · P~ which will be zero because of the uniform charac-
ter of P~ together with a surface charge density P~ · n̂. Show then
that the polarization charges will give rise to an electric field at
the centre of the sphere given by:
~
~ = −1 P .
E
3 ε0

189
13 Electromagnetic Theory of Magnetizable
Materials
In first lecture on electromagnetic theory of polarizable materials, we have
discussed how the polarization of dielectrics by an externally applied electric
field is equivalent to creation of volume and surface distributions of charges.
Analogously, a magnetic field can act on molecular scale current loops exist-
ing in the building blocks of materials, atoms, to produce macroscopic effects.
It was Ampère who first suggested that the magnetism of matter was due to
the cooperative effects of currents circulating in atoms and not, as previously
thought, due to a separate magnetic charge called poles. Thus, the magnetic
properties of materials can be considered from an atomic viewpoint of elec-
tron’s currents, in which case a fundamental understanding of sources of the
magnetic field can be developed.

13.1 Magnetic Polarization Currents

Let us illustrate the microscopic (atomic) theory of magetism from which we


will then formulate the macroscopic theory.
According to the atomic theory, the elec-
trons rotating in orbital paths are equiva-
lent to circulating currents on an atomic
scale. We can define the magnetic mo-
ment of an electron current loop as

µ
~ = IAn̂ ,

which is equal to the product of the area


Figure 48: A schematic diagram of of the plane loop closed by the electron’s
a current loop created by orbiting elec- orbit and the magnitude of the circulat-
tron. ing current. The vector direction n̂ of the
moment is perpendicular to the plane of
the loop and along the direction set by the right-hand rule, as shown in
the Figure 48.

190
A macroscopic material body contains a lot of current loops, so we can de-
fine a macroscopic dipole moment per unit volume of the material, called
magnetization
~ =
X
M µ
~i .
i

The theory of magnetism is based on the above definition of magnetization,


and is formulated by a simple theorem:

Theorem: If a volume V enclosed by surface S has a magnetic dipole mo-


ment per unit volume M ~ , which may be a function of position,
the macroscopic magnetic fields so produced are equivalent to
those of:
• A volume current density J~V = ∇ × M ~.
• A surface current density J~S = M
~ × n̂.

Proof of the Theorem:

The theorem is proved by showing that the vector potential due to the dipole
distribution in a volume V closed by a surface S can be written in the form

1 Z ∇×M ~ 1 I M ~ × n̂
~=
A dV + dS ,
4πε0 c2 R 4πε0 c2 R
V S

where R is a distance from the centre of the volume dV closed by a surface S.

Consider a current loop of radius a, as shown


in Figure 49. We will find the vector po-
tential produced by the current loop at the
point X.
We start from the solution of the Maxwell’s
equations, which gives the (static) vector
potential of a current loop

1 Z J~ I I dl~
~=
A dV = . (70)
4πε0 c2 r 4πε0 c2 r
Due to the radial symmetry of the loop, it
Figure 49: A geometry for the cal- is convenient to work in the polar spherical
culations of the vector potential pro-
duced at point X by a current loop 191
of radius a.
~
coordinates, in which the current element dl
can be written as
~ = adφ φ̂ = −a sin φ dφ î + a cos φ dφ ĵ ,
dl
and the distance from dl to X is given by
h i1/2
r = (x − a cos φ)2 + (y − a sin φ)2 + z 2 ,
which for a  R can be written as
 1/2
r = x2 + y 2 + z 2 − 2ax cos φ + a2 − 2ay sin φ
 1/2
≈ R2 − 2ax cos φ − 2ay sin φ
!
ax cos φ ay sin φ
≈ R 1− − .
R2 R2
Hence
!
−1 −1 ax cos φ ay sin φ
r =R 1+ + ,
R2 R2
where we have used the Taylor expansion of 1/(1 − x) = 1 + x + . . ..

Thus, the vector potential (70) takes the form


Z 2π !
~ I ax cos φ ay sin φ
A = 2
1+ +
4πε0 c R 0 R2 R2
 
× −a sin φ dφ î + a cos φ dφ ĵ .
Since
Z 2π Z 2π Z 2π
sin φ dφ = cos φ dφ = sin φ cos φ dφ = 0 ,
0 0 0

~ simplifies to
the formula for A
Z 2π 
~= I 2 2 2 2

A −a y sin φ î + a x cos φ ĵ dφ ,
4πε0 c2 R3 0
which can be written as
Ia2 Z 2π Z 2π
 
~=
A −y î 2
sin φ dφ + xĵ cos2 φ dφ .
2
4πε0 c R 3 0 0

192
Next, since
Z 2π Z 2π
2
sin φ dφ = cos2 φ dφ = π ,
0 0

we obtain

~= Ia2 π h i
A −y î + x ĵ . (71)
4πε0 c2 R3
Using the relation
x y z x y
 
k̂ × R̂ = k̂ × î + ĵ + k̂ = ĵ − î ,
R R R R R
we can write Eq. (71) as

~ = Ia2 π µ R̂
A 2 2
k̂ × R̂ = 2
k̂ × 2
4πε0 c R 4πε0 c R
1 1
= − ~ ×∇ ,
µ
4πε0 c2 R
where we have used the result

R̂ 1
2
= −∇ ,
R R
and the fact that µ
~ is in the direction normal to the loop, i.e. in the z direc-
tion.

Summarizing what we have obtained so far: We have obtained a formula for


the vector potential produced by a single magnetic dipole moment µ
~.

If we consider a set of dipole moments


and change, for a convenience, the direc-
tion of R̂ into −R̂, as shown in Figure 50,
we can write that the vector potential dA~
produced by a set of dipole moments con-
tained in a volume element dV is

~= 1 ~ × ∇ 1 dV ,
dA M
4πε0 c2 R

193
Figure 50:
where M~ is the magnetic dipole moment
per unit volume.

Then, the vector potential produced by


the dipole moments contained in the total volume V is given by

~= 1 Z
~ × ∇ 1 dV .
A 2
M
4πε0 c R
Using a vector identity
~ = ∇Φ × A
∇ × (ΦA) ~ + Φ∇ × A
~,

we can write the integral function as


~ ~
M~ × ∇ 1 = −∇ 1 × M ~ = ∇×M −∇× M ,
R R R R
and then the vector potential takes the form
1 Z
∇×M~ 1 Z ~
M
~=
A dV − ∇ × dV . (72)
4πε0 c2 R 4πε0 c2 R
The first term is in the form easy to interpret. Refering to the definition of
the vector potential, Eq. (53), we see that ∇ × M~ can be interpreted as the
volume current density. Thus, the first term represents the vector potential
produced by the volume currents existing inside the material.

We still have to do something with the second term as it is in an unfamiliar


form.

In order to transform the second term into a familiar form, we apply a the-
orem that:19
Z ~
M Z ~
M × n̂
− ∇× dV = dS . (73)
V R S R
Thus, the application of the theorem to the second term in Eq. (72) leads to
the vector potential of the form
1 Z
∇×M~ 1 Z ~
M × n̂
~=
A dV + dS ,
2
4πε0 c V R 2
4πε0 c S R
19
Proof of the theorem is given in Appendix B.

194
or
Z ~ Z ~
~= 1 JV 1 JS
A dV + dS .
4πε0 c2 V R 4πε0 c2 S R
We see by refering to the definition of the vector potential that the effective
(Ampere) currents associated with a macroscopic dipole moment M ~ per unit
volume are:

(i) A current density J~V = ∇ × M


~ throughout the volume.

(ii) A surface currents J~S = M


~ × n̂.

Figure 51 illustrates the source of the macroscopic surface current.

Figure 51: Source of the macroscopic surface current.

13.2 ~
The Magnetic Intensity Vector H

When we were dealing with dielectric materials in the presence of electric


fields, it was convenient to introduce a new vector, the displacement vector
~ in order to eliminate the necessity of taking the electric dipole polariza-
D,
tion P~ of the material into account explicitly.

A similar procedure is used for the magnetic materials, where a new magnetic
~ is introduced to eliminate the magnetization M
intensity vector H ~ . We will

195
illustrate this idea for both static and time-varying fields.

13.2.1 Static Magnetic Fields


For a static current distribution, the Maxwell’s equation IV reduces to
~ = µ0 J~ .
∇×B

In a medium where there are magnetic polarization currents as well as con-


duction currents we can write

J~ = J~c + J~m ,

where J~c is the conducting current, and J~m is the magnetization current.
Thus

J~ = J~c + ∇ × M
~ ,

and then
~ = µ0 (J~c + ∇ × M
∇×B ~).

This equation can be written as


~ − µ0 M
∇ × (B ~ ) = µ0 J~c ,

or in the form
 
~
B
∇× −M~  = J~c .
µ0

This shows that the vector B/µ ~ 0−M ~ has as its source only the conduction
current J~c . Therefore, to eliminate the necessity of dealing directly with the
magnetization current J~m , we can define a new vector
~
~ = B −M
H ~ ,
µ0
which is called the magnetic (field) intensity vector.

196
~ the Maxwell’s equation IV for static fields in magnetic mate-
In terms of H,
rials takes the form
~ = J~c .
∇×H

In dealing with magnetic materials we often know J~c but not M ~ (well not
directly anyway.) Think e.g. of an inductor filled with some magnetizable
~ becomes useful. It is a way of avoiding a detailed
material like iron. Then H
calculation of the polarization currents. The magnetic intensity H ~ is the
~
magnetic analogue of the dielectric displacement D in the electric case. We
may drop the subscript c and write
~ = J~ ,
∇×H

but we should remember that J~ is now not the total electric current density
everywhere.

In summary: The use of the vector H ~ enables us to write the Maxwell’s


equation IV in terms of only the conducting current density in any magnetic
material. There is no need to deal with the magnetization M~.

13.3 Linear Magnetic Materials

For most materials (excluding ferromagnetics) the magnetization M~ is lin-


~ and then because of the
early proportional to the applied external field B,
linear relation
~
~ = B −M
H ~ ,
µ0
~ Thus, for linear magnetic ma-
the magnetization is also proportional to H.
terials we write
~ = χm H
M ~ ,

~ M
and then at any point the vectors B, ~ , and H
~ will be in the same direction,
~
and we get the following relation between B and H~

~ = µ0 H
B ~ + µ0 M
~ = µ0 (1 + χm )H
~ = µ0 µr H
~ = µH
~ .

197
The parameter µr = (1 + χm ) is called the relative permeability, χm is the
magnetic susceptibility, and µ = µ0 (1 + χm ) is called the magnetic permeabil-
ity.

Since
~ = µ0 (H
B ~ +M
~),

~ such that
and we have defined M
~ ~
~ = µ0 χ m B = χ m B ,
~ = µ0 χ m H
µ0 M
µ 1 + χm
we get

B~
~ =
H .
µ0 (1 + χm )
~
Thus, if we know the material we use, we can find H.

Example: A solenoid filled with magnetizable material

In this example, we illustrate some interesting properties and re-


lations between magnetic field vectors produced by a solenoid filled
~ is mod-
with a magnetizable material. We will show how the field B
ified due to the presence of the material and what is the source of
the H~ vector.
Consider the expression for the magnetic field produced by a long
solenoid

B = µ0 N I .

In the presence of the material, we should include the Ampere surface


currents as well as the conduction currents in the wire. Since

B = µ0 I 0 ,

where I 0 is the total current per unit length, we have


χm B
 
B = µ0 (N I + M ) = µ0 (N I + χm H) = µ0 NI + ,
µ

198
which can be written as
µ0 χm
 
B 1− = µ0 N I ,
µ
or
µ0 N I µ0 N I
B= µ0 χm = χm .
1− µ 1 − 1+χ m

Note that if χm is positive then B is greater than it would have


been in the absence of the magnetizable material. Evidently in this
case the macroscopic Ampere current is in the same sense as the
conduction current in the solenoid.
Since
χm 1 1
1− = = ,
1 + χm 1 + χm µr
we have for the magnetic field

B = µr µ0 N I = µN I . (74)

Thus, the introduction of a magnetizable material into the solenoid


is equivalent to replace µ0 by µ.
We may go further and consider the following question: What is the
source of the H~ vector?
It is easy to answer this question. Since

B = µH ,

we then have from Eq. (74) that

H = NI .

Thus, H~ depends only on the parameters of the solenoid.


We see that the effect of filling the solenoid with a material of rela-
tive permeability µr is to multiply B by a factor µr (assuming the
current in the wire remains the same).

Example:
~ depends only on the parameters of
To illustrate further that H
the solenoid, consider an another example which shows that in a
solenoid, the magnetic intensity H is independent of the presence

199
or absence of the magnetic material.

As in the preceding example, we consider a solenoid filled with


a magnetic material. By the definition
B
H= −M .
µ0
When we remove the magnetic material, M = 0, and then
B µ0 N I
H= = = NI .
µ0 µ0
With the material present
B µN I
H= −M = − χm H = µr N I − χm H .
µ0 µ0
Hence

H = (1 + χm )N I − χm H ,

which can be written as

H(1 + χm ) = (1 + χm )N I .

Thus

H = NI ,

as before.

This gives rise to the notion of H as an inducing field and B as a resultant


field. This concept is much used in the study of magnetic properties of ma-
~ is the field introduced to hide the properties of
terials. The reason is that H
~
the material. Thus, H is the same independent of the presence or absence of
the material.

13.4 Non-Linear Magnetic Materials


There are materials, i.e. iron, whose the magnetic properties are not de-
~ = µH.
scribed by the linear formula B ~ They are called ferromagnetic mate-
rials and have a specific property that placed in an external magnetic field

200
their magnetization undergoes a saturation. This is because all the internal
current loops are lined up, which breaks the linear property. The reason for
this unusual behavior is that ferromagnetic materials do not have a unique
value of magnetic susceptibility because of strong magnetic nonlinearities.
~ and H,
For this reason, it is difficult to provide a relation between B ~ and
it is usually presented graphically in terms of the so-called hysteresis. An
example of the hysteresis is shown in Figure 52.

Figure 52: Hysteresis loop of a ferromagnetic material.

13.4.1 Work done in magnetization of iron


Think of the case of the solenoid of length `, cross-section area A, and filled
with magnetic material. The applied voltage V is:
dΨ d
V = −E = = `N BA ,
dt dt
where N ` is the number of turns.

The rate of doing work to power the solenoid is


dB dB
P = V I = A`N I =V H ,
dt dt
where V = Al is the volume of the solenoid.

201
In the time dt that it takes to change B by dB, then the work done is
dW = P dt = V H dB. Thus, the work done per unit volume is
dW = H dB .
Hence, the work done per unit volume in one cycle of the hysteresis loop of
a non-linear material is given by the area of the loop.

Figure 53: An example of hysteresis loops of a ’hard’ (iron) and a ’soft’ (ferrite) fer-
romagnetic materials. Hard ferromagnetics retain some magnetization in the absence of
external fields. This property of hard ferromagnetics makes them useful for permanent
magnets.

It is sometimes useful to write


Z Z Z Z
H dB = µ0 H d(H + M ) = µ0 H dH + µ0 H dM
1 Z
= µ0 H 2 + µ0 H dM ,
2
where the first term on the rhs is the work to establish magnetic field, and
the second terms is the work by the field H to establish magnetization dM .

Which part really uses the energy: 12 µ0 H 2 or µ0 H dM ? To check it, con-


R

sider properties of magnetic materials in a time varying magnetic field.

13.5 Permanent Magnetic Materials: Ferromagnets

We have so far considered magnetic materials that are called diamagnetics


and paramagnetics. In these materials the magnetization exists only in the

202
~ is a function of the external
presence of an external magnetic field, i.e. M
~ ∼ B).
field, (M ~ However, there is a class of materials, called ferromagnetics
or permanent magnets in which macroscopic magnetization exists even in the
absence of the external field.

Consider first an exercise, which will illustarte several points that are of in-
terest in designing of ferromagnetic materials.

Exercise in class. Magnetic field of a solenoid

Consider a cylindrical current sheet on the surface of a cylinder of radius a,


called a solenoid. The lines of current flow are circles round the surface of
the cylinder.
As in the case of the plane sheet we
measure the strength of the current
by a current per unit length with the
length measured along the surface of
the cylinder parallel to the axis and
at right angles to the direction of cur-
rent flow.

(a) Assuming the magnetic field within


the cylinder is uniform and parallel to the
axis and that the magnetic field is zero
Figure 54: Ampère’s loop to calcu- outside the cylinder show that within the
late magnetic field inside and outside a ~ = µ0 I¯ where I¯ is
solenoid.
cylinder the field is B
the current per unit length.

(b) Can you think of the arguments from Ampère’s theorem for the field
being uniform within the cylinder and zero outside?

13.5.1 ~ and H
B ~ fields of a ferromagnet

We can now easily extend the arguments from the above exercise to analyse
~
the magnetic field a long homogeneous ferromagnetic material. In this case B

203
~:
inside the material is due solely to the M

J~s = M
~ × n̂ ,

or in terms of the magnitudes Js = M .

~ and H
Let us find B ~ inside and outside the material.

Using Ampère’s law


H
~ · d~` = µ0 I, where I = Js ` = M `, we obtain
B

B` = µ0 M ` i.e. B = µ0 M ,

where we have used the fact that B is zero outside the magnet.

Thus
B µ0 M
H= −M = −M =0 ,
µ0 µ0
in the region where M 6= 0, i.e. inside the magnet.

Outside the magnet


B B
H= −0= ,
µ0 µ0
~ is just a scaled replica of B.
i.e. H ~

204
13.5.2 Magnetic Poles
~ = 0 always and the lines of B
Since ∇ · B ~ form closed loops, we have
 
~
~ = ∇·B −M
∇·H ~  = −∇ · M
~ .
µ0

Thus, H~ has a source (field lines start and stop) where M


~ varies i.e. at the
ends of the magnet (magnetic poles).

For a ferromagnet, we can write


~ = −∇ · M
∇·H ~ = ρm .

Thus, we can think of ρm as a volume density of ‘magnetic charge’ giving


~ field. It must be stressed that this equivalence is purely math-
rise to the H
ematical, and does not prove a physical existence of magnetic charges.

Moreover, for the ferromagnet


~ = J~c = 0 .
∇×H
~ here to those of the E
Note the similar mathematical properties of H ~ field
in electrostatics
~ = ρ ~ =0.
∇·E and ∇ × E
ε0
Historically, magnetostatics of permanently magnetized materials (ferromag-
nets) developed via use of the H~ field and ’magnetic charges’ or ’poles’. One
obtains a law analogous to Coulomb’s Law for the force between magnetic
poles.

205
Exercise in class: Plane magnetized material

An infinite plane surface divides the universe into a vac-


uum on one side and a magnetic material on the other.
Within the magnetic material there exists a uniform
magnetic moment per unit volume M ~ which is parallel
to the surface.

(a) Show that while the direction of the magnetic in-


duction vector B~ is different on the two sides of the
surface, its magnitude is given everywhere by:
M
B= .
2ε0 c2

(b) Find the magnitude and direction of the magnetic


~ everywhere due to an infinite plane parallel slab
field B
of material of thickness d which is permanently uni-
formly magnetized with dipole moment per unit vol-
ume M ~ lying parallel to the bounding surfaces.

~ everywhere.
(c) Find the magnetic intensity H

206
13.6 Time Dependent Magnetic Fields and Energy Loss

~ may not stay in


In a time-dependent magnetic field the magnetization M
~ This corresponds to internal friction and
phase with the driving field H.
heat dissipation.

Let
H = H0 cos(ωt) .
Then the magnetisation in a lossy material varies as
M = M0 cos(ωt + φ) ,
where M0 and φ represent the amplitude and phase of the magnetization
response to the magnetizing field. Thus
M = M0 cos φ cos(ωt) − M0 sin φ sin(ωt) .
Now the work done in magnetization per cycle of the AC current producing
the magnetizing field is
2π/ω
Z Z
dM
W = µ0 H dM = µ0 H dt .
dt
t=0

Since
dM
= −ωM0 cos φ sin(ωt) − ωM0 sin φ cos(ωt) ,
dt
we have W = W1 + W2 . Consider the term W1 :
2π/ω
Z
W1 = −µ0 H0 M0 ω cos φ cos(ωt) sin(ωt) dt .
t=0

This term is zero - it represents reversible energy conversion to and from H.

Consider now the term W2 :


2π/ω
Z
W2 = −µ0 H0 M0 ω sin φ cos2 (ωt) dt .
0

207
This term represents work done against internal friction during magnetizing
and demagnetizing the material.

Since cos2 (ωt) dt is positive, sin φ must be negative so that work is done on
R

the material i.e. φ is negative.

13.7 The complex magnetic susceptibility


Let
H = H0 eiωt ,
which for a physical field can be written as
 
H = H0 cos(ωt) = Re H0 eiωt .
Then
M = M0 ei(ωt−φ) = M0 e−iφ eiωt
or
M = (M0 cos φ − iM0 sin φ)eiωt

M0 M0
 
= cos φ − i sin φ H0 eiωt .
H0 H0
This result can be written in terms of real and imaginary susceptibility
M = (χ0 − iχ00 )H0 eiωt = (χ0 − iχ00 )H .
Using this complex number notation, we find
B = µ0 (H + M ) = µ0 H + µ0 (χ0 − iχ00 )H
= [µ0 (1 + χ0 ) − iµ0 χ00 ]H
and then
0 00
B = (µ − iµ )H = µH ,
0 00
where µ = µ0 (1 + χ0 ), µ = µ0 χ00 , and µ is the complex permeability.

208
13.8 Maxwell’s Equations in Dielectric and Magnetic
Materials

Let us complete the lectures on dielectric and magnetic properties of mate-


rials by incorporating our findings into the Maxwell’s equations, to see how
the basic equations of electromagnetism are modified in the presence of the
material.

The Maxwell equation IV contains a current density term


~
~ = µ0 J~ + µ0 ε0 ∂ E .
∇×B
∂t

We think this is always true provided J~ is the total electric current den-
sity. Applying this in a region where there may be electric and magnetic
polarization effects we can write the current density as

∂ P~
J~ = J~c + J~E + J~M = J~ + ~ .
+∇×M
∂t
Thus, the Maxwell’s equation IV takes the form
~ ~
~ = µ0 J~c + µ0 ∂ P + µ0 ∇ × M
∇×B ~ + µ 0 ε0 ∂ E ,
∂t ∂t
which can be written as
 
~
B
∇× −M~  = J~c + ∂ (ε0 E
~ + P~ ) .
µ0 ∂t

~ and H,
With the introduction of the new vectorial fields D ~ this equation
takes the form
~
~ = J~ + ∂ D ,
∇×H
∂t

where we must remember that J~ represents the conduction current only.

209
In summary: The Maxwell’s equations for the electromagnetic field prop-
agating in a material are of the following form
~ = ρ,
∇·D
~ = ∇·H
∇·B ~ =0,
~
∇×E~ = −µ ∂ H ,
∂t
~
∇×H~ = J~ + ∂ D ,
∂t
but in general
~ = −∇ · M
∇·H ~ .

These equations are supplemented by appropriate constitutive relations, which


~ and the magnetic induction B
connect the electric field E ~ with the displace-
~ and the magnetic field H
ment field D ~

~ = εE
D ~ and ~ = µH
B ~ .

These relations carry information about the material.

Revision questions

Question 1. State the definition of magnetization.

Question 2. How do we define a volume current density and a surface current


density?

~ instead of B
Question 3. Explain why it is useful to use the H ~ for a magnetic
field in a magnetic material.

Question 4. What is the form of the Maxwell’s equation IV for static fields in
magnetic materials?

210
Question 5. What it is a hysteresis loop, and what does it correspond to?

Question 6. What is meant by a homogeneous and isotropic material?

Question 7. How do we distinguish between soft and hard ferromagnetics?

Question 8. What are the modifications to the Maxwell’s equations of an elec-


tromagnetic field propagating in a material?

211
Tutorial problems

Problem 13.1 Field of a uniformly magnetized sphere

~ per unit
A magnetized sphere has a uniform dipole moment M
volume.

(a) Find and sketch a diagram of the magnetic polarization cur-


rents in and on the sphere.

(b) Apply the Biot-Savart law to this current distribution to find


the magnetic flux density B~ at the centre of the sphere. Show
that:
~
B~ = 2M .
3ε0 c2

Problem 13.2 Fields in an inhomogeneous magnetic material

A long solenoid of radius a and having N turns per metre car-


ries current I. It is filled with an inhomogeneous material with a
gradient in magnetic susceptibility. The susceptibility is zero at
the centre and increases linearly with radial distance r from the
centre of the solenoid, attaining a value of 2 at r = a.

~ M
(a) Find the vectors B, ~,H ~ throughout the solenoid.
(Hint: Study H~ first by integrating the Maxwell equation
~ = J~ ,
∇×H
and use the symmetries of the situation. Show that H ~ is un-
changed by the presence of the magnetizable material).

(b) Show that the self-inductance of the solenoid is the same


as would be obtained by replacing the above-mentioned magnatic
material with a homogeneous material having a magnetic suscep-
tibility of 34 .

(c) What are the magnetic polarization currents on the surface


and within the solenoid? Calculate them in terms of I.

212
Problem 13.3 Fields in a dense magnetic material

Using a method similar to that used in the case of electric fields


within dense dielectrics, show that if a spherical cavity is exca-
vated in a uniformly magnetized material, the magnetic field B ~
~
and the magnetic intensity H in the hole and in the material are
related by

~
~ material − 2M ,
~ hole = B
B
3ε0 c2
~
~ material + M .
~ hole = H
H
3

213
14 Poynting’s Theorem Revisited
We have seen in Chapter 8 how energy of the electromagnetic field may
be transported through vacuum (empty space) by means of electromagnetic
waves. We have shown that the direction of propagation of energy is de-
termined by the Poynting vector. In this lecture, we will reconsider the
Poynting theorem taking into account propagation of the electromagnetic
field in magnetizable materials. A question we will try to answer: How the
energy is propagated inside a magnetizable material?

14.1 ~ and H
Poynting Vector in Terms of E ~
~
We have shown that it is useful to represent the magnetic field in terms of H
~ vector when the field propagates inside a magnetizable
vector rather then B
material.

~ vector for the field in vacuum


Using the concept of the H
~
~ = B = ε 0 c2 B
H ~ ,
µ0
the Poynting vector of the EM field in vacuum takes the form
~ = ε 0 c2 E
N ~ ×B
~ =E
~ ×H
~ .
~ ×H
The cross product E ~ also turns out to be the correct expression for the
Poynting vector when magnetizable materials are involved and is the expres-
sion most commonly quoted for it.

In order to show this, we consider, as before in Section 8, a flow of the energy


through a closed surface S. Using the Gauss’s theorem and a vector identity
~ × B)
∇ · (A ~ =B
~ · (∇ × A)
~ −A
~ · (∇ × B)
~ ,

we can write
I Z
~ × H)
(E ~ · dS
~= ~ × H)
∇ · (E ~ dV
SZ V Z
= ~ · (∇ × E)
H ~ dV − ~ · (∇ × H)
E ~ dV .
V V

214
Now, substitute from the Maxwell equations

~ = − ∂B
∇×E ~ ,
∂t
~ = J~ + ∂ D
∇×H ~ ,
∂t
and obtain
~ ~
~ · ∂ B dV − E ~ · ∂ D dV − E
I Z Z Z
~ × H)
(E ~ · dS
~=− H ~ · J~ dV .
S V ∂t V ∂t V

The individual terms in this equation have the following interpre-


tation:
1. The lhs is the rate of flow of field energy out of the volume V enclosed
by the surface S.
2. First term on the rhs is the rate of work in establishing the magnetic
field in V .
3. Second term is the rate of doing work in establishing the electric field
in V .
4. Third term is the rate of doing work on the currents in V .

We go further with the Poynting’s Theorem.

~ = µH,
If we substitute for B ~ we then get
Z ~
∂B Z
∂ 1
 
∂ Z 1 B2
~
H· dV = µH 2 dV = dV ,
∂t ∂t 2 ∂t 2 µ
and
~
~ · ∂ D dV = ∂ 1 2 ∂ Z 1 2
Z Z  
E εE dV = εE dV .
∂t ∂t 2 ∂t 2
Hence
1 2 1 B2
!
~=−∂
I Z Z
~ × H)
(E ~ · dS εE + dV − ~ · J~ dV .
E
S ∂t 2 2 µ

215
Thus, the rate of flow of field energy out of volume V is equal to the rate of
changing energy of the EM field plus the arte of doing work on the currents
in V .

14.2 Poynting Vector for Complex Sinusoidal Fields

It is well known that the electromagnetic field (e.g. light) is a real physical
quantity (observable). However, in the electromagnetic theory it is advanta-
geous to represent the real electromagnetic field by complex sinusoidal quan-
tities because of its mathematical simplicity. In addition, what we usually
~ ∗ · Ei,
measure is the average intensity of the field, hE ~ which is a real quantity.
For time varying fields, we usually write
~ =E
E ~ 0 eiωt and ~ =H
H ~ 0 eiωt ,

where E~ 0 and H
~ 0 are complex quantities including both amplitude and phase
information. We understand that the electric and magnetic fields are given
by the REAL PARTS of E ~ and H.
~

The power of the complex exponential scheme lies in the fact that for oper-
ations such as summation, subtraction, integration etc., we take real parts
AFTER the operation. For example:
 
~ 1 + ReE
ReE ~ 2 = Re E
~1 + E
~2 ,
dE~1 dE~2 d ~ 
Re + Re = Re ~2 .
E1 + E
dt dt dt
However, some care has to be taken in evaluating the Poynting vector.

~ =E
The Poynting vector is given by N ~ × H,
~ but if we write complex expo-
~ × H,
nential expressions for E ~ we note that
 
~ = ReE
N ~ c × ReH
~ c 6= Re E
~c × H
~c ,
where we put a subscript c to indicate that we are writing the fields using
complex exponentials.

Proof:

216
We can write the complex electric field as
 
~c = E
E ~ 0 eiωt = E
~ r + iE
~ i (cos ωt + i sin ωt)
   
= ~ r cos ωt − E
E ~ i sin ωt + i E
~ r sin ωt + E
~ i cos ωt .

Similarly, for the magnetic field we can write


 
~c = H
H ~ 0 eiωt = H
~ r + iH
~ i (cos ωt + i sin ωt)
   
= ~ r cos ωt − H
H ~ i sin ωt + i H
~ r sin ωt + H
~ i cos ωt ,

~ r, E
where E ~ i, H
~ r and H
~ i are real vectors.
However, from the above, we see that
~c = E
ReE ~ r cos ωt − E
~ i sin ωt ,
~c = H
ReH ~ r cos ωt − H
~ i sin ωt .

Clearly, if we calculate Re(E~c × H


~ c ) we get extra terms in addition to those
in the expression for ReE~ c × ReH~ c , and then

~ c × ReH
ReE ~ c 6= Re(E
~c × H
~ c) ,

as required.
¯~
~ , but rather N
In experiments, we do not measure N , the mean Poynting
vector averaged over the cycles of oscillations.

There is a useful expression for the MEAN POYNTING VECTOR in


terms of the complex exponential E ~ c and H
~ c:

¯~ 1 ~ ~∗
 1 ~∗ ~  ~ ¯ ~
N = Re E c × Hc = Re Ec × Hc = ReEc × ReHc ,
2 2

where the bar over N ~ indicates an average over the whole cycle of the sinu-
soidal field to eliminate the rapid temporal oscillations at frequency ω.

Proof:

217
Calculate the Poynting vector involving measurable (real) parts of the com-
plex fields
   
~ = ReE
N ~ c × ReH
~c = E
~ r cos ωt − E
~ i sin ωt × H
~ r cos ωt − H
~ i sin ωt
~r × H
= E ~ r cos2 ωt + E
~i × H
~ i sin2 ωt − (E
~r × H
~i + E
~i × H
~ r ) cos ωt sin ωt .

Since
1ZT 1ZT 1
cos2 ωt dt = sin2 ωt dt = ,
T 0 T 0 2
1ZT
and cos ωt sin ωt dt = 0 ,
T 0
we obtain for the average Poynting vector
¯~ 1 ~ ~ ~ ~
N = (E r × Hr + Ei × Hi ) .
2
On the other hand, take
 
~ c∗ =
H ~ r − iH
H ~ i e−iωt ,
 
~c =
E ~ r + iE
E ~ i eiωt ,

and then
~c × H
(E ~ ∗ ) = (E
~r × H
~r + E
~i × H
~ i ) + i(E
~i × H
~r − E
~r × H
~ i) .
c

Hence
1 ~ ~ c∗ = 1 (E

~r × H
~r + E
~i × H
~ i) ,
Re Ec × H
2 2
¯~
which is equal to N , as required.

In summary: The average Poynting vector N ~ of complex exponential fields


satisfies the relation
¯~ ¯ ~ c = 1 Re E
 
N ~c ×
= ReE ReH ~c × H~ c∗ ,
2
and this is the expression for calculation of the Poynting vector of complex
fields.

218
Revision questions

~ and H
Question 1. What is the form of the Poynting vector in terms of the E ~
vectors?

Question 2. How do we write the electric and magnetic field vectors for oscillat-
ing fields?

Question 3. What is the useful form of the Poynting vector for complex sinusoidal
fields?

219
15 Plane Wave Propagation in Dielectric and
Magnetic Media

In this lecture, we shall examine in some details how existing radiation field is
modified by the material it passes through. We will consider the propagation
in a linear and homogeneous material and as we shall see, the conductivity
is the most significant parameter for modifications, rather than the dielectric
and magnetic constants.

We have learnt that the properties of a lossy linear and homogeneous dielec-
tric material can be described using a complex permittivity and similarly,
the properties of a lossy magnetic material are described by a complex per-
meability. Thus, for a lossy linear and homogeneous material the Maxwell’s
equations describing the EM field inside the material are of the form
~ = ρ/ε ,
∇·E
∇·H~ = 0,
~
∇×E~ = −µ ∂ H ,
∂t
~
∇×H~ = J~ + ε ∂ E , (75)
∂t
where ε, µ are complex quantities that characterize the material, ρ is a free
charge in the material, and J~ is conduction current only, i.e. J~ = σ E.
~

Note from Eq. (75) that an EM field propagating in a linear and homoge-
~ and H
neous material is solely described by the E ~ fields, and its properties
depend only on the material constants ε and µ. Thus, the only factors that
can modify the wave from that propagating in vacuum are the constants ε
and µ.

We remember that in the vacuum the EM field propagates in the form of


plane transverse waves. Questions we will try to answer are: What are the
properties of the EM waves propagating inside a material? Does an EM wave
retain its transverse nature?

220
To answer these questions, consider a plane wave propagating in one dimen-
sion, the z direction
~ =E
E ~ 0 ei(ωt−kz) . (76)
~ to check whether the wave retaines its transverse prop-
First, we shall find H
erties when propagating through the lossy material. Thus, for the electric
field (76), the Maxwell’s equation IV takes the form
~ = σE
∇×H ~ + iωεE
~ = (σ + iωε)E
~ .

Since for a plane wave propagating in the z direction the derivatives ∂/∂x
and ∂/∂y of E~ and H~ are zero, we obtain


î ĵ k̂


0 ∂
0 ∂z

~ .
= (σ + iωε)E

Hx Hy Hz

From the LHS, we see that


~ z = 0 → (σ + iωε)Ez = 0 .
(∇ × H)
~ ⊥ k̂.
Hence Ez = 0 unless (σ + iωε) = 0. Thus, E

~ = 0. Also
Since Ez = 0, we have that ∇ · E
~
∂H
~ = −µ
∇×E
∂t
~ ⊥ H.
can be used to show that E ~

In addition, we can find from the above equation that Hz = 0, which means
~ ⊥ ~k.
that H

We can summarize our thusfar findings that the wave propagating


in a lossy material retains its transverse nature.

However, the losses are expected to modify somehow the wave. What kind
of modifications are made by the losess?

221
~ = 0, the quantity ε occurs only in the equation for ∇ × H,
Since ∇ · E ~ and
it is common to proceed as follows:
~
~ = J~ + ε ∂ E = σ E
∇×H ~ + iωεE~ ,
∂t
~ = iω σ + ε E ~ = iω ε − i σ E
   
∇×H ~ .
iω ω
Now
σ σ σ
 
ε−i = ε0 − iε00 − i = ε0 − i ε00 + = ε̄ ,
ω ω ω
which gives
~ = iω ε̄E
∇×H ~ .

Physically what has been done is to lump together the conduction and the
lossy dielectric constant term. To an external observer they are inseparable.
Only using some theory of the internal structure of the dielectric, they can
be separated.

We can summarize that the electric and magnetic fields of the propagating
wave in a lossy conducting material satisfy the following equations
~ = 0 ,
∇·E (77)
~ = 0 ,
∇·B (78)
~
~ = −µ ∂ H = −iωµH
∇×E ~ , (79)
∂t
~ = iωεE
∇×H ~ , (80)
where we have left the bar off ε.

15.1 Dispersive Equation and Complex Propagation


Number
We now proceed to solve the propagation equations (77)–(80). We will see
that the fact that µ and ε are complex numbers, the solution leads to a dis-
persive equation.

222
Thus, we look for plane wave solutions, and will try to find how the propa-
gation number k behaves.

The procedure is as follows: Taking ∇× of (79) and using (80), we obtain


~ = ω 2 µεE
∇ × (∇ × E) ~ ,

which, after applying the double × vector identity, simplifies to


~ = ω 2 µεE
−∇2 E ~ .

~ of the transverse EM wave is independent of x and y, we have


Since E
~
∂ 2E
− ~ .
= ω 2 µεE
∂z 2

~ given in Eq. (76), is


On the other hand, the second derivative of E,
~
∂ 2E ~ .
= − k2E
∂z 2

Hence, we obtain a dispersion equation

k 2 = ω 2 µε .

This dispersion equation is not as simple as it looks. We cannot just say that
phase velocity is
ω 1
vp = =√ ,
k µε
as we usually do for EM waves propagating in vacuum, because ε and µ are
complex quantities and then k is a complex number.

What does an imaginary value of k mean?

To give the answer to this question, we write the complex number k as

k = α − iβ ,

where α and β are real. Then

E = E0 ei(ωt−kz) = E0 ei(ωt−αz+iβz) = E0 e−βz ei(ωt−αz) .

223
Clearly from this, the phase velocity is
ω
vp = ,
α
i.e. the real part of k plays the role of the propagation number, and β =
− Im(k) is the attenuation (losses) coefficient.

In practice, the losses come from (1) Conduction currents, (2) Lossy dielec-
tric, and (3) Lossy magnetic material, which are lumped together in ε and µ.

We can find the explicit forms of α and β.

To do this, we write the complex number k as


1
σ
  
1 2
0 00 0 00
k = ω (µε) = ω
2 ε −i ε + (µ − iµ ) ,
ω
which can be written as
 1
σ σ
     
2
k=ω ε0 µ0 − µ00 ε00 + − i µ0 ε00 + + ε0 µ00 .
ω ω
Let us introduce a short notation for the real and imaginary parts of k:
σ
 
p = ε0 µ0 − µ00 ε00 + ,
ω
σ
 
q = µ0 ε00 + + ε0 µ00 .
ω
Then
1
k = ω (p − iq) 2 ,

which can be written as


 1  12
2 2 2 −iarctan(q/p)
k=ω p +q e ,

or
 1
k = ω p2 + q 2 4
e−iθ ,

224
where θ = 12 arctan (q/p). Hence
 1
α = Re(k) = ω p2 + q 2 4
cos θ ,
 1
β = −Im(k) = ω p2 + q 2 4
sin θ ,

which in general are quite complicated and difficult to interpret. The inter-
pretation becomes more transparent when we make some simplifications that
are discussed in the following two examples.

15.2 Wave Refraction and Attenuation

We now consider two examples of the propagation of an EM wave in different


materials: Low loss dielectrics and conducting materials.

15.2.1 Propagation of an EM wave in a low loss dielectric


In the first example, we illustrate the propagation of an EM wave in a low
loss dielectric with no magnetic effects.

For dielectrics, σ is negligible and also we disregard magnetic properties.


Thus, for a low loss dielectric
σ
ε00 +  ε0 ,
ω
and
1
µ = µ0 = .
ε 0 c2
Hence, the propagation number simplifies to
 1
ω σ
 
2
k = √ ε0 − i ε00 +
ε0 c ω
 12
2 ! 21

ω  0 2 σ

= √ (ε ) + ε00 + e−iΘ  ,
ε0 c ω

225
where
ε00 + σ
ε00 + σ
!
ω ω
Θ = arctan ≈ ,
ε0 ε0

as the losses are small. Then


√ s  00 
ω ε0 −iΘ/2 ω ε0 −i ε +σ/ω
2ε0
k= √ e = e .
c ε0 c ε0

Hence, the real part of k takes the form


s √
ε00 + σ/ω
!
ω ε0 ω ε0 ω√
α= cos ≈ √ = εr .
c ε0 2ε0 c ε0 c

We can then find phase velocity of the wave and refractive index of the
dielectric material
ω ε0
r
vp = =c 0 <c for ε0 > ε0 ,
α s ε
c ε0 √
n = = = εr > 1 .
vp ε0

Thus, we may conclude that inside the dielectric, the EM wave


propagates with a phase velocity vp < c and can be refracted from
the original direction on the entry to the dielectric.

Similarly, we can find β


s
ε00 + σ/ω
!
ω ε0
β = −Im(k) = sin .
c ε0 2ε0

Since for small θ, sin θ ≈ θ, we get


s
ε00 + σ/ω ω ε00 + σ/ω
!
ω ε0
β≈ = √ .
c ε0 2ε0 c 2 ε0 ε0

Thus, losses (absorption of the wave) are small.

226
We can summarize, that the theory predicts that the refractive index for a
lossless dielectric is given by

n = εr .

Table below compares theoretical values of n with that obtained experimen-


tally. An excellent agreement is observed, except for polar molecules. For

example, water has a refractive index of 1.33, whereas εr = 9. The reason
for this discrepancy lies, not in the inadequacy of the laws of electromag-
netism, but rather in the tacit assumption that the dielectric constant is a
constant independent of the frequency. The refractive index of the water
exhibits a strong dependence on ω as it is composed of polar molecules that
the polarization depends strongly on ω.

εr n (experimental)
Air 1.00029 1.00029
Argon 1.00028 1.00028
CO2 gas 1.00047 1.00045
Benzene 1.49 1.48
Ethanol 5.3 1.36
NaCl 2.47 1.54
Water 9.0 1.33

15.2.2 Propagation of an EM Wave in a Good Conductor


As a second example consider a propagation of the EM field in a good con-
ductor with no dielectric and magnetic losses: ε00 = µ00 = 0.

Examples of good conductors: Cu, Ag.

Consider the Maxwell equation IV, which for an EM wave propagating in a


conducting material can be written as
~
~ = J~ + ε ∂ E = σ E
∇×H ~ + iωεE
~ . (81)
∂t
227
Definition of a Good Conductor

Good conductor is when the conduction term on the right-hand


side of the Maxwell equation (81) dominates over the displace-
ment current, i.e. when σ  ωε. On physical grounds, for a good
conductor the conductivity current is much larger than the polar-
ization current.

For example, for copper at ω = 1 MHz = 2π × 106 rad

σ 5.8 × 107
= = 1.0 × 1012 .
ωε0 2π × 106 × 8.85 × 10−12
Consider general expression for k:
 1
σ σ
     
2
0 0 00 00 0 00
k = ω εµ −µ ε + −i µ ε + + ε0 µ00 ,
ω ω
that for a conductor with no dielectric losess reduces to
 1

 
2
0 0
k = ω (ε µ − 0) − i µ +0 ,
ω
and finally
1


2
0 0
k = ω ε µ − iµ .
ω
Now we might as well drop the dashes on ε, µ understanding that they are
real quantities, and obtain
1 1
σ √ σ
 
2 2
k = ω εµ − iµ = ω εµ 1 − i ,
ω εω
that can be written as
1
2 # 12
"
√ 

σ σ
2
k = ω εµ 1 + e−i arctan εω .
 εω 

228
σ
Remembering that for a good conductor εω
 1. Thus, we can drop ”1” in
the [ ] brackets, and get
1
√ σ −i π 2 √ σ −i π
 r
k = ω εµ e 2 = ω εµ e 4 .
εω εω

Since cos(π/4) = sin(π/4) = 1/ 2, we obtain
!
√ 1 1 ωµσ
r
k = ωµσ √ − i √ = (1 − i) .
2 2 2

Then
ωµσ
r
α = Re(k) = = β = −Im(k) .
2
Knowing α, we can find the phase velocity of the wave
s

vp = ,
µσ

which shows that the velocity of propagation in a good conductor depends


on the frequency, even if µ and σ are assumed independent of frequency.

15.2.3 Skin Effect


Since the propagation number is a complex number and for a good conduc-
tor β = α, there is a very heavy attenuation of a propagating wave. For a
complex propagation number, we can write the electric field as

E = E0 e−βz ei(ωt−αz) .

We see, that the amplitude of the electric (and also magnetic) field decreases
exponentially as z increases.

Since we can write



α=β= ,
λ

229
1
the distance for attenuation ”e” fold (i.e. amplitude falls to a factor e
of its
original value in a distance β1 ) is

1 λ
z= = ,
β 2π
i.e.
1 λ λ
δ= = ≈ !!
β 2π 6
where δ is called the skin depth in the conductor. It is the distance the wave
must propagate in order to decay by an amount e−1 . This effect is sometimes
called ”skin effect” as with an increasing σ the current flows in a narrower
and narrower layer, until in the limit of σ → ∞ a true current exists only on
the surface of the conductor.

230
Revision questions

Question 1. Show that the propagation number of an electromagnetic field in


dielectric and magnetic materials is a complex number.

Question 2. What are the physical consequences of the propagation number of


an electromagnetic field being a complex number?

Question 3. What is the definition of a good conductor?

Question 4. Why the theoretical and experimental values of the refractive index
of water are significantly different?

Question 5. What is meant by the skin effect of a propagating wave?

Tutorial problems

Problem 15.1 Relations between the electric and magnetic field vectors in dielec-
tric and conductor

(a) Show that for a plane wave propagating in a dielectric along


the x-direction, the electric and magnetic field vectors satisfy the
relations
∂Ey ∂Bz ∂Ez ∂By
= − , = ,
∂x ∂t ∂x ∂t
∂Hz ∂Dy ∂Hy ∂Dz
= − , = .
∂x ∂t ∂x ∂t

(b) Show that for the propagation in a conductor


∂Ey ∂Bz
= − ,
∂x ∂t
∂Hz ∂Dy
= − − σEy ,
∂x ∂t

where σ is the conductivity of the conductor.

231
(c) Show that Ey and Hz each satisfies a second order differential
equation of the form

∂ 2 Ey ∂ 2 Ey ∂Ey
2
= εµ 2
+ µσ .
∂x ∂t ∂t

Problem 15.2 Energy flow in a conductor

Using the relations of Problem 15.1, part (b), show that in a


conductor the rate of decrease of electromagnetic field energy in a
volume element dV is equal to the net rate at which energy flows
out of this volume element plus the rate of Joule heating in the
element.

Problem 15.3 Practical example on propagation and attenuation

If a field propagates as a plane wave in the form:

E = E0 ei(ωt−kz) ,

where ω is the (real) radian frequency and k is a complex propa-


gation constant that can include the effects of losses due to con-
duction, dielectric loss and magnetic loss, in lectures we derived
a dispersion equation
 1
σ σ
     
2
k=ω ε0 µ0 − µ00 ε00 + − i µ0 ε00 + + ε0 µ00 .
ω ω

Consider propagation at a frequency of 1 MHz in a medium with


the following properties:
1. Conductivity σ = 10−9 mho m−1 .
2. Dielectric susceptibility χe = 3 − i10−4 .
3. Magnetic susceptibility χm = 2 − i10−3 .

(a) Calculate the phase velocity and wavelength at 1 MHz.

(b) Compare these values with those for free space propagation
at 1 MHz.

232
(c) Calculate the attenuation coefficient at 1 MHz. How far will
the wave propagate before its amplitude falls to one tenth of its
initial value?

(d) Is this material a good conductor at 1 MHz? Over what


frequency range would it be classified as a good conductor ?

233
16 Transitions Across Boundaries for Elec-
tromagnetic Fields
In the lectures on electric and magnetic properties of materials we have
confined our attention principally to the case of a single material medium
completely occupying the spatial region where electric and magnetic fields
existed. We now investigate the relations between the fields which hold at
a boundary between two different materials. They are of great help in the
calculations of propagation problems in optics, where light can propagate
between different material media.

Across boundaries between different materials there are sharp changes in


electrical properties of ε, µ, σ. On a macroscopic scale the fields may have to
be regarded as varying discontinuously across such boundaries. The source
of such discontinuities will be the surface polarization charges P~ · n̂ and cur-
rents M~ × n̂ discussed previously.

The question we will address in this lecture is: How the electromagnetic fields
transfer through a boundary between two materials? In the analysis of the
transfer properties of the electromagnetic field across a boundary between
two different materials, it is convenient to decompose the fields into two
components, normal and tangential to the boundary between the materials
and study how each of the components transfers through the boundary.

16.1 Normal Components of the Fields


We will use the Maxwell’s divergence equations I and II to investigate the
transition of normal field components at a boundary between two materials.

16.1.1 ~
Normal Component of B

Consider two materials with different constants ε and µ, as shown in the


Figure 55. Suppose that a magnetic field B ~ propagates from material (1)
to (2). We expect that B ~ will be different in different materials due to the
presence of different surface currents.

234
~ = 0, to find
To check this, we first apply the Maxwell’s equation II, ∇ · B
how the normal component of B ~ transfers through the boundary between
the two materials.

Using the Gauss’ divergence theo-


rem, the Maxwell’s equation II can
be written in the integral form as
I
~ · n̂ dS = 0 ,
B (82)
S

where S is an arbitrary surface clos-


ing some area on the boundary
plane.
Figure 55: A Gauss’s surface for the deriva-
tion of the boundary condition on the normal
In order to evaluate the integral,
component of B.~
we consider a thin cylinder of
area δS and thickness δ` including
the boundary, as shown in Figure 55. In this case, we can distinguish three
surfaces that we represent by unit vectors n̂1 , n̂2 and n̂3 , and therefore the
surface integral in Eq. (82) splits into three terms
Z
(B~2 · n̂1 ) δS + (B~1 · n̂2 ) δS + ~ · n̂s dS = 0 ,
B
sides

where the first term is the flux of the magnetic field through the top surface,
the second term is the flux through the bottom surface, and the third term
is the flux through the side surface of the cylinder.

Since B~ is finite everywhere and we are interested in the transformation of


the field at the boundary, we make the height δ` tend to zero, δ` → 0. Then,
the integral
Z
~ · n̂ dS → 0 as δ` → 0 .
B
sides

~ over the surface has


Thus, we find that the total flux of the magnetic field B
contributions from the top and bottom ends only

B~2 · n̂1 + B~1 · n̂2 = 0 . (83)

235
Now, if we introduce the notation
n̂1 = n̂ then n̂2 = −n̂ ,
and we find that Eq. (83) becomes
n̂ · (B~2 − B~1 ) = 0 ,
or equivalently, we may write that
B2⊥ = B1⊥ ,
~ in material (1) normal to the
where B1⊥ is the component of the field B
boundary and B2⊥ is the component in material (2) normal to the boundary.

Thus, the normal component of B~ is continuous at all points across


a boundary separating two different materials.

On physical grounds, we can understand this result by noting that the mag-
~ 1 × n̂1 and M
netic fields of polarization currents M ~ 2 × n̂2 are parallel to
the boundary and so do not affect the normal component of B, ~ as shown
in Figure 56.

16.1.2 ~
Normal Component of H

~ = µH,
Since B ~ we have
B1⊥ = µ1 H1⊥ = µ2 H2⊥ = B2⊥ .
Thus, we obtain that
µ2
H1⊥ = H2⊥ ,
µ1
which shows that the nor-
mal component of H ~ is not
continuous across a boundary
between two different materi-
als since µ1 and µ2 are not
equal.
Figure 56: Direction of magnetizations and re-
On physical grounds, this result sulting polarization currents at the boundary be-
for the normal component of H ~ tween two different materials.
is the consequence of different magnetizations of the materials, when µ1 6= µ2 .

236
16.1.3 ~ and E
Normal Component of D ~

~ = 0, an identical argument to that


Since in nonconducting dielectrics ∇ · D
~ will show that the normal component of D
applied the above for B ~ is
continuous across a boundary.

~ = εE.
In the case of dielectrics we write D ~ Since the normal component of D
~
is continuous across a boundary, we have

ε1 E1⊥ = ε2 E2⊥ ,

which shows that the normal component of E ~ is not continuous across the
boundary.
On the physical grounds, we can
understand it as follows. The
electric field of the dielectric sur-
face charge P~ · n̂ is normal to
the boundary, as shown in Fig-
ure 57. Since the polarizations
of the dielectrics are different, so
the fields of the dielectrics are
different. This difference causes
the discontinuity in the incident E~
Figure 57: Polarization charges at the field.
boundary between two different materials.
The discontinuity of the normal
component of the electric field leads to a change of the direction of prop-
agation of the field, as it is illustrated in Figure 58. The change of the
direction depends on the ratio ε2 /ε1 . When ε2 /ε1 > 1, θ2 > θ1 .

16.2 Tangential Components of the Fields

Different boundary conditions apply to the components of the fields parallel


(tangential) to a boundary between two different materials. We will use the
Maxwell’s curl equations III and IV to investigate the transition of tangential
field components at a boundary between different materials.

237
Figure 58: Boundry between two dielectrics showing change of the direction of propa-
gation of an electric field.

16.2.1 ~
Tangential Component of E
~ we will apply the Faraday induction law,
For the tangential component of E,
the Maxwell’s equation III, to a closed path such as shown in the Figure 59.

We choose a closed loop in which


the sides perpendicular to the
boundary are made infinitely short
compared to the parallel sides L.

Consider the Maxwell’s equation III


~
~ = − ∂B .
∇×E
∂t
Figure 59: A closed loop for the derivation
of boundary conditions for the tangential com- Integrating both sides of this equa-
~
ponent of E. tion over the surface a and applying
the Stokes’s theorem, we obtain
I Z ~
∂B
~ · d~l = −
E · n̂0 da .
l a ∂t
Hence
  Z Z ~
∂B
~ 2 · t̂2 + E
E ~ 1 · t̂1 L + ~ · d~l = −
E · n̂0 da , (84)
ends ∂t

238
where n̂o is the unit vector normal to the surface a, and t̂1 and t̂2 are unit
vectors along the paths L on the side (1) and (2), respectively.20

As dl → 0 the right-hand side and the second term on left-hand side of


~ and B
Eq. (84) go to zero, since E ~ are finite everywhere. In this limit, the
area enclosed by the path approaches zero.

Since, t̂2 = −t̂1 = t̂, we then have


 
~2 − E
E ~ 1 · t̂ = 0 .

But E~ · ~t is the component of E~ tangential to the surface. Since this is true


for any t̂, we obtain E1k = E2k , where E1k and E2k are the components of E ~1
and E~ 2 parallel to the boundary. Thus, we may conclude that:

The tangential component of E ~ is continuous across a boundary


between two different dielectric materials.

The continuity of a tangential com-


ponent can be written in an equiv-
alent form as
 
~2 − E
n̂ × E ~1 = 0 .

Explanation

From Figure 60, we see that the


~ can be writ-
cross product n̂ × E
Figure 60: Illustration of the direction of the ten as
~
vector n̂ × E.

~
n̂ × E = E sin(π/2 − θ) = E cos θ .

In other words, n̂ × E~ is equal to the projection of the vector E ~ on the


~
boundry. Thus, n̂ × E is tangential to the boundary, i.e. it is the tangential
~
component of E.

20
Do not mix the unit vector n̂o normal to the surface a with the unit vector n̂ normal
to the boundary. Actually, n̂o ⊥ n̂.

239
16.2.2 ~
Tangential Component of H

Consider the Maxwell’s equation IV:


~
~ = J~c + ε ∂ E ,
∇×H
∂t

where J~c is the conduction current, i.e. not counting polarization currents.

The application of the Stokes’s theorem then gives


I Z Z ~
∂E
~ · d~l =
H J~c · n̂da + ε · n̂da .
l a a ∂t

Both terms on the right-hand side go to zero as δl → 0 because J~c and ∂ E/∂t
~
are finite. 21 Hence
 
~2 − H
H ~ 1 · t̂ = 0 ,

or
 
~2 − H
n̂ × H ~1 = 0 .

~ is continuous
Thus, we conclude that the tangential component of H
at all points across a boundary between two different materials.

16.2.3 ~
Tangential component of B

We may easily deduce that the tangential component of B ~ is not in general


continuous across a boundary because of the presence of the magnetic polar-
ization surface currents M~ × n̂, which do not have a finite current density as
they flow in an infinitely thin surface layer.

~ the term
Thus, if we examine the corresponding Maxwell equation for ∇× B,
in the integral involving J~ may stay finite as δj → 0.

21
Infinite current density can occur in a material which has infinite electrical conduc-
tivity, but here we ignore this special case.

240
This conclusion can be justified as follows. Since
~ = ε 0 c2 B
H ~ −M
~ ,

we have
   
~2 − H
H ~ 1 = ε 0 c2 B
~2 − B
~1 − M
~2 −M
~1 .

Take scalar product of both sides


with t̂, and using the fact that the tan-
gential component of H ~ is continuous
across a boundary
 
~2 − H
H ~ 1 · t̂ = 0 ,

we obtain

~ 1 · t̂ = 1 M
   
~2 − B
B ~2 −M
~ 1 · t̂ .
ε 0 c2
~ 1 6= M
It is seen that if M ~ 2 , the differ-
Figure 61: Directions of magnetizations ence between the two magnetizations
and resulting magnetic fields produced at
generates a discontinuity in the B ~
a boundary between two materials.
field, as shown in Figure 61.

Summarizing:
Field components that are continuous across a boundary between
two materials:
~
• The normal component of D.
~
• The tangential component of E.
~
• The normal component of B.
~
• The tangential component of H.

241
Revision questions

Question 1. Which of the Maxwell’s equations are used to analyse the transmis-
sion of the normal components of the EM field through a boundary
between two materials?

Question 2. Which of the Maxwell’s equations are used to analyse the trans-
mission of the tangential components of the EM field through a
boundary between two materials?

Question 3. Explain, using physical grounds, why the normal component of the
~ is continuous across a boundary between two ma-
magnetic field B
terials.

Question 4. Explain, using physical grounds, why the tangential component of


~ is continuous across a boundary between two
the electric field E
materials.

Tutorial problems

Problem 16.1 Change in direction of a field line

A (static) electric field line makes an angle θ1 to the normal to


the boundary between two dielectrics. The dielectric constant on
the ‘incident’ side is ε1 and on the other side is ε2 . There are no
local charges other than the polarization charges in and on the
dielectrics. Show that the angle the electric field makes to the
normal to the boundary in the second dielectric is
ε2
 
θ2 = arctan tan θ1 .
ε1

Problem 16.2 Angles and polarization charges

An infinite slab having a dielectric constant of 4 is placed in a


vacuum which is permeated by an electric field E~ 0.

242
(a) Find the angle that E~ 0 should make to the boundary normal
~
so that E inside the slab makes an angle π/4 with the boundary
normal.

(b) Find the surface density of polarization charge on the slab


in terms of E0 .

Problem 16.3 Wave reflection from a dielectric at normal incidence

Consider waves normally incident on a boundary between two di-


electrics having dielectric constants 4 (on the incident side) and 9.

~
(a) Draw a diagram showing the directions and magnitudes of E
~
and B in the incident, reflected and transmitted waves. Apply
~ and B
the general boundary conditions for E ~ directly. Remem-
ber that B = E/vp .

(b) Show that the magnitude of the amplitude reflection coef-


ficient is 0.2.

243
17 Propagation of an EM Wave Across a Bound-
ary

In the preceding lecture, we have derived relations for transmission of electric


and magnetic fields through a boundary between two non-conducting mate-
rials. We have treated each field separately. We now proceed to consider a
propagation of an EM wave across a boundary between two materials. We
shall find that for the propagation of an EM wave the situation is different
that we cannot treat the fields separately. The boundary conditions must
be satisfied simultaneously for both the electric and magnetic fields. This
leads to a modification that at a boundary between two different materials,
the general boundary conditions cannot be satisfied by a transmitted wave
alone. There has to be a reflected wave also. What happens is that the inci-
dent radiation is absorbed by charges in the boundary and reradiated in all
directions. The waves interfere destructively except in two directions along
those of the reflected and transmitted waves. The directions, amplitudes
and phases of the reflected and transmitted waves can be derived from the
general boundary conditions already obtained. This is an explanation often
presented in texbooks on classical optics.

In this lecture, we shall attack these problems from the standpoint of the elec-
tromagnetic theory of light and will show how the propagation phenomenon
can be understood with the help of the Maxwell’s equations. More precisely,
in this and the following lecture, we consider a number of well-known op-
tical phenomena and will show that they can be explained in terms of the
electromagnetic theory of light.

244
17.1 Representation of Plane Waves in Different Di-
rections
Let us first set up the formalism we shall use in the study of propagation
of an EM wave through a boundary between two materials. The question
we will try to answer is: What is the convenient method to represent plane
waves propagating in different directions?
Suppose that a plane wave prop-
agates in the z direction. Then
~ =E
E ~ 0 exp[i (ωt − kz)] .

Let n̂p is the unit normal to


the phasefront (wavefront) of the
propagating wave, and ~r is a po-
sition vector that is independent
of n̂p , as shown in Figure 62. The
vector ~r determines position of
a wave-front as seen by an ob-
Figure 62: Illustration of the wave front mon-
itored by an observed O distant ~r from the
server O.
point A.
We see from Figure 62 that the
distance z the wave propagated in time t is

z = n̂p · ~r .

Hence, we can write


~ =E
E ~ 0 exp[i (ωt − n̂p · ~r k)] .

Thus, if at the point A appear few plane waves propagating in different di-
rections, the observer can distinguish them by different n̂’s. In other words,
the vector ~r sets a reference frame for the observation of different waves at
the point A.

Remember, that an EM wave is composed of the electric and magnetic fields,


so we have to consider the propagation of both fields. However, there is
a useful relation between B~ and E
~ of a plane wave that we can find the
magnetic field from the knowledge of the electric field.

245
17.1.1 ~ in terms of E
Representation of B ~

The magnetic field of a plane wave propagating in the direction determined


by the unit vector n̂p can be found from the relation

~ = k n̂p × E
B ~ . (85)
ω
Proof:

We shall show that the relation (85) arises from the Maxwell equation III:
~
~ = − ∂B .
∇×E
∂t
To prove it, we expand both sides of this equation, and obtain


î ĵ k̂ ~
∂B ∂Bx ∂By

0 0 =− = −î − ĵ ,

∂z

∂t ∂t ∂t
Ex Ey 0

where we have used the fact that the wave is transverse and propagates in
the z direction.

Comparing the coefficients standing at the same unit vectors, we find that
the x component gives
∂Ey ∂Bx
− =−
∂z ∂t
from which, we obtain
Ey ω
ikEy = −iωBx or =− .
Bx k
The y component gives:
∂Ex ∂By
=−
∂z ∂t
from which, we obtain
Ex ω
−ikEx = −iωBy or = .
By k

246
Thus
 1 ω 2 1 ω
E = Ex2 + Ey2 2
= Bx + By2 2 = B .
k k
~ B,
Since E, ~ n̂p , are mutually orthogonal, E~ ×B~ gives the direction of n̂p
~ gives the direc-
(Poynting result) of the propagation direction. Then n̂p × E
~ Hence, written vectorially
tion of B.
~
~ = B = k n̂p × E
H ~ . (86)
µ ωµ

If ε and µ are real, k = ω εµ, and then we have
s
~ = ε ~ .
H n̂p × E (87)
µ

In the next few lectures, we will use the continuity conditions for E~ and H
~ to
analyse different optical phenomena. With the relations (86) or (87), we will
be able to limit the analysis to the electric field alone, as knowing properties
~ we can infer from Eq. (86) properties of H.
of E, ~

17.2 Directions of Reflected and Transmitted Waves


Armed with the formalism to study propagation of an EM wave between
two materials, we now consider some familiar elementary results of classical
optics on reflection and transmission to show that they can be derived from
the Maxwell’s equations.
The most useful results in this
connection concern the continu-
ity of tangential components of
the E~ and H ~ fields. How-
ever, most of the results can
be derived from the fact that
the tangential component of E ~
is continuous across a bound-
ary.

Figure 63: Propagation vectors for the in-


cident, reflected and refracted beams at a
boundary between two materials. 247
Consider properties of an EM wave
(radiation beam) at a boundary be-
tween two materials, as illustrated
in Figure 63.

Let the origin of position coordinates ~r be located on the boundary S. Then


the electric fields for incident, reflected and transmitted waves are
~i = E
E ~ 0 exp [i (ωt − n̂i · ~r k1 )] ,
~r = E
E ~ 1 exp [i (ωt − n̂r · ~r k1 )] ,
~t = E
E ~ 2 exp [i (ωt − n̂t · ~r k2 )] .

These equations determine directions of the three electric fields relative to


the direction of observation ~r. Note that we have assumed the presence of
the reflected wave without any particular reasons.

Why do we need the presence of the reflected wave in the propa-


gation of an incident wave through the boundary?

The answer to this question is provided by the requirement that the tangen-
~ and H
tial components of E ~ must be continuous through the boundary.

Suppose that E ~ is polarized in the plane of incidence, i.e plane containing n̂


and n̂i . Then, the following two equations result from the continuity condi-
tions

Ei cos θi − Er cos θr = Et cos θt , (88)

and

Hi + Hr = Ht . (89)

Since
k
H= E,
ωµ
the relation (89) takes the form
k2
Ei + Er = Et , (90)
k1

248
as for a dielectric µ1 = µ2 = µ0 .

Now we see that without Er , Eq. (88) gives

cos θt
Ei = Et ,
cos θi
while Eq. (90) gives

k2
Ei = Et .
k1
Thus, without Er we would get two different values for Ei or Et , which we
cannot accept as both continuity conditions Eqs. (88) and (89) must be satis-
fied at the same moment. Hence, we conclude that the continuity conditions
~ and H
for E ~ will be satisfied only if Er 6= 0.

The same argument applies to the need of the transmitted wave. Without Et ,
we get two equations
cos θi
Er = Ei and Er = −Ei ,
cos θr
which evidently cannot be in general satisfied simultaneously.

Now, we will try to answer the following question:

What are the relative directions of the three waves represented


here by the unit vectors n̂i , n̂r and n̂t ?

To answer this question, we use the boundary condition for the tangential
~
component of E:
 
~i + E
n̂ × E ~ r = n̂ × E
~t ,

i.e.
n o
~ 0 exp [i (ωt − n̂i · ~r k1 )] + E
n̂ × E ~ 1 exp [i (ωt − n̂r · ~r k1 )]
n o
~ 2 exp [i (ωt − n̂t · ~r k2 )]
= n̂ × E .

249
This relation must hold over the whole surface S for all ~r (subject to n̂·~r = 0).
Thus, the exponential phase factors must all be the same. Otherwise, if it
was true for one ~r it would not be true for other ~r’s, but we have a freedom
of choosing ~r. Hence

k1 n̂i · ~r = k1 n̂r · ~r = k2 n̂t · ~r ,

from which we have

n̂i · ~r = n̂r · ~r , (91)

and
k2
n̂i · ~r = n̂t · ~r .
k1
These relations will help us to prove that:

Incident, reflected and transmitted waves are coplanar.

In other words, n̂i , n̂r , n̂t are coplanar, the property observed in experiments.

To show this, we first define planes determined by pairs of the unit vectors
(n̂i , n̂), (n̂r , n̂) and (n̂t , n̂), and next will show that all the pairs of the unit
vectors form the same plane.

It is well known from the vector analysis that a cross product between two
vectors determines a plane in which these two vectors are. Thus, we will
determine the cross products n̂i × n̂, n̂r × n̂ and n̂t × n̂, and then will show
that the cross products are equal.

In order to find the cross products, we use the following relation valid for
arbitrary ~r and n̂, such that n̂ · ~r = 0.

~r = −n̂ × (n̂ × ~r) . (92)

Proof:

Since

n̂ × (n̂ × ~r) = (n̂ · ~r)n̂ − (n̂ · n̂)~r ,

250
and n̂ · ~r = 0 as the vector ~r lies on the plane S, we obtain
n̂ × (n̂ × ~r) = −~r ,
as required.
Using Eq. (92), we can then write
Eq. (91) as
n̂i · [n̂ × (n̂ × ~r)] = n̂r · [n̂ × (n̂ × ~r)] .
Interchanging (·) and (×) products, we
find
(n̂i × n̂) · (n̂ × ~r) = (n̂r × n̂) · (n̂ × ~r) .
This must be true for all ~r in
Figure 64: Illustration that the inci- plane S. Thus:
dent and reflected beams are coplanar.
n̂i × n̂ = n̂r × n̂ .
This implies that n̂r is in the ”plane of incidence”, i.e. the plane containing n̂
and n̂i , as shown in Figure 64.

Similarly, the relation


k1 n̂i · ~r = k2 n̂t · ~r
implies that
k1 n̂i · n̂ × (n̂ × ~r) = k2 n̂t · n̂ × (n̂ × ~r) .
As before, by interchanging (·) and (×) products, we find that
k1 (n̂i × n̂) · (n̂ × ~r) = k2 (n̂t × n̂) · (n̂ × ~r) .
This relation is valid for all ~r in S. Thus
k1 n̂i × n̂ = k2 n̂t × n̂ .
This means that n̂t is in the plane of incidence. Thus, we may conclude that
n̂i , n̂, and n̂t are coplanar.

The coplanar property of the waves is observed in any experiment. Note that
what we have just shown is an another example of a remarkable triumph of
the Maxwell’s electromagnetic theory.

251
17.3 Angle of Reflection and Snell’s Law of Refraction

We now illustrate that other familiar laws of elementary optics can also be
derived from the Maxwell’s equations.

We have already shown that


n̂i × n̂ = n̂r × n̂ ,
from which and from the definition of the cross product, we have that
sin θi = sin θr .
Hence
θi = θr .
Thus, the angle of incidence equals the angle of reflection, another
familiar law of elementary optics.

Moreover, we have shown that


k1 n̂i × n̂ = k2 n̂t × n̂ ,
from which we have
k1 sin θi = k2 sin θt ,
and then
sin θi k2 n2
= = n12 = .
sin θt k1 n1
This is the well-known law of refraction in optics, called the Snell’s law.

In the case where k1 and k2 are purely real (e.g. in dielectrics), the refractive
index has a simple physical interpretation
k2 ω/k1 v1
n12 = = = ,
k1 ω/k2 v2
i.e. the refractive index in equal to the ratio of phase velocities.

252
Revision questions

Question 1. Why there are three waves, incident, reflected and refracted in a
propagation of an EM wave between two dielectric materials?

Question 2. What is a plane of incidence?

Question 3. Incident, reflected and transmitted waves are coplanar: True or


false?

Question 4. State the Snell’s law, and briefly explain how it is derived from the
Maxwell’s equations.

Question 5. How does the refractive index depend on phase velocities of an EM


wave propagating between two dielectric materials?

253
18 Fresnel’s Equations
In the propagation of an EM wave between two different materials, it is im-
portant to know how much of the energy of the incident beam is reflected
and transmitted. In this lecture, we will find relations between the ampli-
tudes of the incident, reflected and refracted beams. More precisely, we will
express the amplitudes of the reflected and transmitted beams in terms of
the amplitude of the incident beam. We shall see that the relations depend
on the angle of propagation of the incident beam and the material constants.

The boundary condition on tangential E ~ does not give sufficient information


~ r and E
to calculate E ~ t in terms of E~ i . For a given E ~ i there are still two
unknowns in the equation for continuity of tangential E ~ viz E
~ r and E
~ t . We
need a second relation between E ~ i, E
~ r and E~ t . This can be obtained from
the continuity of the magnetic field.

We know that tangential component of B ~ is not continuous across a boundary


~
because of the presence of M × n̂ surface currents in magnetized materials.
To allow for such possibilities we can use the more general condition that
tangential component of H~ is continuous across a boundary, i.e.
 
~i + H
n̂ × H ~ r = n̂ × H
~t .

Since
~
~ = B = k n̂p × E
H ~ ,
µ ωµ
where n̂p is a unit ray vector in the direction of propagation, we have an
equation
! !
k1 ~ i + k1 n̂r × E
~ r = n̂ × k2 n̂t × E
~t
n̂ × n̂i × E , (93)
ωµ1 ωµ1 ωµ2
which together with
 
~i + E
n̂ × E ~ r = n̂ × E
~t , (94)

~ r and E
contains sufficient information to determine E ~ t in terms of E
~ i.

254
Since these two equations involve vectors of an arbitrary polarization, the
solution of these two equations is greatly facilitated by decomposing each
of the vectors into two components: electric field components parallel and
normal to the plane of incidence. Then, we can solve these two cases sep-
arately that are necessary and sufficient to determine the relations between
the amplitudes valid for an arbitrary polarization since a superposition of
these two cases gives the solution for an arbitrary polarized incident wave.

Equations (93) and (94) also provide a simple explanation of why we need
reflected and transmitted fields at the boundary to obtain the correct results
for the field amplitudes.

18.1 ~ i normal to plane of incidence


E
~ i of an EM wave acting on a boundary is
Suppose that the electric field E
normal to the plane of incidence, as shown in Figure 65.
In this case, the incident electric
~ i is purely tangential to the
field E
boundary. Since the materials are
~ r and E
isotropic, the induced fields E ~t
will also be tangential to the the
boundary. Thus the condition
 
~i + E
n̂ × E ~ r = n̂ × E
~t

gives
~0 + E
E ~1 = E
~2 . (95)
Figure 65: Incident beam with the elec-
tric field normal to the plane of incidence.

Note then that


~ 0 = n̂ · E
n̂ · E ~ 1 = n̂ · E
~2 = 0 .

If we want to express E1 and E2 in terms of E0 , we need an another equation


~ 0, E
for E ~ 1 and E~ 2 , which comes from the continuity condition of H
~ through
the relation
! !
k1 ~ k1 ~ k2 ~
n̂ × n̂i × Ei + n̂r × Er = n̂ × n̂t × Et .
ωµ1 ωµ1 ωµ2

255
~ α exp i(ωt − n̂α · ~r k) are the same, we obtain
Since the phase factors in E
k1 h   
~ 1 = k2 n̂ × n̂t × E
~ 0 + n̂ × n̂r × E
i  
~2 .
n̂ × n̂i × E
ωµ1 ωµ2

Using a vector identity A ~ × (B~ × C)


~ = (A ~ · C)
~ B~ − (A
~ · B)
~ C,
~ and the fact
~ ~ ~
that n̂ · E0 = n̂ · E1 = n̂ · E2 = 0, we obtain
k1  ~ 1 = k2 n̂ · n̂t E
~ 0 + n̂ · n̂r E

~2 .
n̂ · n̂i E
ωµ1 ωµ2
However

n̂ · n̂i = cos(π − θi ) = − cos θi ,


n̂ · n̂r = cos θr = cos θi ,
n̂ · n̂t = cos(π − θt ) = − cos θt ,

and then, we obtain


k1  ~ ~ 1 cos θi = k2 E

~ 2 cos θt .
E0 cos θi − E (96)
ωµ1 ωµ2
~ are all in the same direction, we might as well drop the vector
Since E’s
signs. Thus, Eqs. (95) and (96) take the form

E0 + E1 = E2 , (97)
k2 µ1
E0 cos θi − E1 cos θi = E2 cos θt . (98)
k1 µ2
Eliminating E1 using Eq. (97), that E1 = E2 − E0 , we get
k2 µ1
E0 cos θi − (E2 − E0 ) cos θi = E2 cos θt ,
k1 µ2
which can be written as
!
k2 µ1
2E0 cos θi = cos θi + cos θt E2 ,
k1 µ2
or

2k1 µ2 E0 cos θi = (k1 µ2 cos θi + k2 µ1 cos θt ) E2 .

256
Thus
2k1 µ2 cos θi
E2 = E0 . (99)
k1 µ2 cos θi + k2 µ1 cos θt
Using the Snell’s law

k1 sin θi = k2 sin θt ,

we can eliminate θt . This will allow us to predict the amplitude of the


reflected wave from only knowing materials properties (k1 , k2 , µ1 , µ2 ) and the
angle of incidence. Thus, from the Snell’s law, we have
q
k2 cos θt = k22 − k12 sin2 θi

and then substituting to Eq. (99), we get


2k1 µ2 cos θi
E2 = q E0 . (100)
k1 µ2 cos θi + µ1 k22 − k12 sin2 θi

Similarly, eliminating E2 = E0 + E1 from Eq. (98) above, we obtain


q
k1 µ2 cos θi − µ1 k22 − k12 sin2 θi
E1 = q E0 . (101)
k1 µ2 cos θi + µ1 k22 − k12 sin2 θi

Equations (100) and (101) are called Fresnel equations for the electric field
amplitudes.

~ fields are not parallel to each other, but their relative


The corresponding H
magnitudes can be deducted from equations of the form

~ = k n̂ × E
H ~ i.e. H=
kE
.
µω µω

18.2 ~ i in the plane of incidence


E

Suppose now that E~ i is in the plane of incidence. In this case H


~ i is tangential
to the boundary plane and then H ~ r and H~ t are tangential too. Thus
 
~i + H
n̂ × H ~ r = n̂ × H
~t

257
becomes
~0 + H
H ~1 = H
~2 .
~ is given by
The continuity of tangential E
 
~0 + E
n̂ × E ~ 1 = n̂ × E
~2 .

We have two simultaneous equations. In contrast to the previous case, it is


~
now simpler to work in terms of H.

Thus, we express E ~ in terms of H ~

E~ = − ω n̂ × B ~ = − µω n̂ × H
~ .
k k
Hence
µ1 ~ 0 + µ1 n̂r × H
~ 1 = n̂ × µ2 n̂t × H
   
n̂ × n̂i × H ~2 .
k1 k1 k2
~ i normal to plane of
Continuing with a procedure similar to the case of E
incidence, we obtain
q
k22 µ2 cos θi − µ1 k1 k22 − k12 sin2 θi
H1 = q H0 , (102)
k22 µ1 cos θi + µ2 k1 k22 − k12 sin2 θi
2k22 µ2 cos θi
H2 = q H0 . (103)
k22 µ1 cos θi + µ2 k1 k22 − k12 sin2 θi
Equations (102) and (103) are called Fresnel equations for the magnetic
field amplitudes.

~ are not parallel, but their relative amplitudes can be de-


This time the E’s
ducted from the relation E = (µω/k)H.

18.3 Fresnel Equations for dielectric media

Consider now two specific examples of propagation of an EM wave between


two dielectric materials. We shall show that in this case, the Fresnel equa-
tions can be simplified to forms containing only geometrical factors.

258
In a dielectric: conductivities σ1 = σ2 = 0, µ1 = µ2 = µ0 , k = 2π/λ =
real, and
ω 1
vp = =√ .
k εµ0

Since k ∼ 1/λ ∼ 1/vp ∼ ε, the Snell’s law (k1 sin θi = k2 sin θt ) becomes
√ √
ε1 sin θi = ε2 sin θt ,

and then
s
sin θi ε2 v1
= = = n12 .
sin θt ε1 v2

Consider two examples:

Example 1: In the first example we assume that E is normal to the plane


of incidence. From Eq. (100), we have that in general

E2 2k1 µ2 cos θi
= .
E0 k1 µ2 cos θi + k2 µ1 cos θt
Since for a dielectric
k2 sin θi
µ1 = µ2 = µ0 and = ,
k1 sin θt
we obtain
E2 2 cos θi 2 cos θi
= k2 = sin θi
E0 cos θi + k1 cos θt cos θi + sin θt
cos θt
2 cos θi sin θt
= .
cos θi sin θt + sin θi cos θt
Hence
E2 2 cos θi sin θt
= .
E0 sin (θt + θi )

259
Similarly, we can readily show that

E1 sin(θt − θi )
= . (104)
E0 sin (θt + θi )

Example 2: Consider now the case of E in the plane of incidence. Following


the same procedure as above, we can show that
E2 2 cos θi cos θt
= ,
E0 sin (θt + θi ) cos (θi − θt )

and similarly

E1 tan(θi − θt )
= .
E0 tan (θt + θi )

It is seen from the above examples that the reflection and refraction ampli-
tudes are different for linearly polarized waves with the electric field vector
oscillation in the direction normal to the plane of incidence and for waves
with the electric field vector oscillating in the plane of incidence. However,
the above examples show an interesting feature that for normal incidence,
(θi = 0), the reflected and refracted amplitudes are independent of the po-
larization.

Revision questions

Question 1. Explain in general, what are the Fresnel’s equations?

Question 2. Under which conditions the ratios of the reflected to the incident
and transmitted to the incident beams amplitudes depend solely on
the angles involved?

Question 3. Does the amplitude of the reflected beam depend on polarization of


the incident beam?

260
Tutorial problems

Problem 18.1 Reflected and transmitted fields at a boundary between two non-
conducting materials

Consider a propagation of an electromagnetic wave between two


different non-conducting materials. Prove the following:

~ i in the plane of incidence


(a) For E

E1 sin 2θt − sin 2θi E2 4 sin θt cos θi


= and = .
E0 sin 2θt + sin 2θi E0 sin 2θt + sin 2θi

~ i normal to the plane of incidence


(b) For E

E1 sin(θt − θi ) E2 2 sin θt cos θi


= and = .
E0 sin(θt + θi ) E0 sin(θt + θi )

261
19 Applications of the Boundary Conditions
and the Fresnel Equations

In this lecture, we examine some of the consequences of the Snell’s law. There
are two cases possible: n2 > n1 and n1 > n2 . In the first case, an optical
wave travels from an optically ”rarer” to optically ”denser” medium. In the
second case, we have the inverse situation. We will consider these two cases
separately for dielectrics and for conductors.

19.1 Applications in dielectrics


In propagation between two dielectrics there are interesting polarization
dependent effects that the two components of the EM wave (parallel and
perpendicular to the plane of incidence) are not transmitted and reflected
equally.

19.1.1 Polarization by reflection


~ in the plane of incidence, as shown in Figure 66.
Consider the case of E
From the Fresnel equations, we have
that the ratio of reflected to incident
amplitude is

E1 tan(θi − θt )
= .
E0 tan(θi + θt )

Note from the ratio, if θi + θt = π2 then


tan(θi + θt ) = ∞ and consequently

E1 = 0 .

However, according to Eq. (104),


at the same time E1 will not be
Figure 66: Geometry of reflection and
~ field polarized zero for the electric field compo-
refraction for the incident E
in the plane of incidence. nent normal to the plane of inci-
dence.

262
~ i has arbitrary polarization then E
Thus, if E ~ r will be plane polarized with E
~r
normal to the plane of incidence.

π π
If θi + θt = 2
then θt = 2
− θi . Thus
s
sin θi ε2 sin θi
= n21 = = = tan θi .
sin θt ε1 sin( π2 − θi )

Hence, the angle of incidence for total linear polarization of the reflected
wave is
s
ε2
θi = arctan .
ε1

It is known in the literature as the Brewster’s angle.

~ is in the plane of incidence,


There is an equivalent alternative proof that if E
then E1 = 0 (no reflected field polarized in the plane of incidence).

Alternatively, it can be proved using the continuity conditions for the tan-
gential components of E~ and H,
~ from which we have

E0 cos θ − E1 cos θ = E2 cos θt , (105)

and

H0 + H1 = H2 . (106)

Since
k
H= E, (107)
ωµ
π
and θt = 2
− θ, we get

E0 − E1 = E2 tan θ ,
E0 + E1 = n12 E2 , (108)

where n12 = k2 /k1 .

263
However, from the Snell’s law we have that
sin θ sin θ sin θ
=   = = tan θ = n12 .
sin θt sin π2 − θ cos θ

Thus, Eqs. (108) will be satisfied simultaneously only if E1 = 0, i.e. there is


no reflected field in the plane of incidence.

~ i polarized normal
We now prove that the above conclusion is not true for E
to the plane of incidence. In this case
H0 cos θ − H1 cos θ = H2 cos θt , (109)
and
E0 + E1 = E2 . (110)
From Eq. (109) we find, after applying Eq. (107), that
cos θt
E0 − E1 = E2 = n12 E2 .
cos θ
Thus, E1 must be present, otherwise E0 or E2 would have two different values.

19.1.2 Total internal reflection

We now illustrate an another interesting effect that has many practical ap-
plications: Total internal reflection at a boundary between two dielectrics.
It is well known from experiments that in the case of propagation from an
optically more dense to optically less dense medium, e.g. from water in to
air, the incident beam can be completely reflected at the boundary with no
transmission. Does it mean that there is no refracted beam? If this is the
case, one can easily find from the conditions of continuity of the tangental
components of E ~ and H~ that the refracted beam has to be present to satisfy
those two conditions.

It looks that the EM theory is in a trouble, but no panic, lets see how we
can handle this problem in terms of the EM theory.

264
Consider the Snell’s law
s
sin θi ε2
sin θt = , where n21 = .
n21 ε1

One can see that in the case of n21 < 1, that happens when ε2 < ε1 , and
when the wave is going from an optically more dense to optically less dense
medium, real angles θt are obtained only for sin θi ≤ n21 .

Since θt increases with θi , an interesting situation arises when θt = π/2,


at which angle the refracted wave glaze along the boundary. The angle of
incidence at which θt = π/2 or equivalently at which

sin θi = n21

is called a critical angle for propagation.

For greater θi , sin θt > 1, and then the angle of refraction θt becomes imagi-
nary. In this case, there is no real refracted wave, only a reflected wave.

What does it mean ”no real” refracted wave?


The answer to this question is provided by calculating the cosine of θt , which
in the case of the total internal reflection is an imaginary number
v
sin2 θi
q u
u
cos θt = 1 − sin2 θt = t1 −
n221
v
u sin2 θi
u
= ±it − 1 = iβ ,
n221

We see that although sin θt is still real, cos θt becomes imaginary when
sin θt > 1.

What are the consequences of this property?

Consider the propagation of the transmitted wave in the less optically dense
medium, as illustrated in Figure 67, for which

Et = E2 e−i(ωt−n̂t ·~rk) ,

265
with the propagation distance

n̂t · ~r = z cos θt + x sin θt = iβz + αx .

Then
0 0 0 0
Et = E2 e−i(ωt−iβ z−α x) = E2 e−β z e−i(ωt−α x) .

Here β 0 = kβ and α0 = kα. One can see that there is attenuation of the field
amplitude in the z direction but no phase propagation. Phase propagation
occurs in the x direction along the boundary.
Thus, for sin θt > 1, an evanes-
cent wave exists along the bound-
ary (in the x direction) which is at-
tenuated exponentially in the second
medium in the normal z direction. We
may say that the field enters into the
second material, but the wave does
not.

This illustrates a general method of ap-


plying the Fresnel equations. For only a
limited range of circumstances will all the
Figure 67: A schematic diagram for angles θi , θr , θt be real. We can how-
calculation of the propagation distance ever always apply a generalized Snell’s
of the transmitted wave under the total Law k sin θ = k sin θ to find sin θ and
2 t 1 i t
reflection.
cos θt and proceed as above.
Note that the planes of constant phase
are normal to the boundary (i.e. they have their normals tangential to the
boundary). The phase of the transmitted wave below the boundary must
match the incident wave above. The wavelength in the second medium is
2π 2π 2π λ0 n21
λ0 = 0
= = 2π sin θi
= ,
α kα λ0 n21
sin θi

where λ0 is the wavelength of freely propagating waves in this medium.


The planes of constant amplitude in the transmitted medium are parallel to
the boundary.

266
19.2 Transmission and Reflection at a Conducting
Surface
Propagation of EM fields in conductors (metals) is more complicated phe-
nomenon than in dielectrics. Consider a propagation of an EM wave in a
medium composed of a dielectric and a conductor, and assume that the in-
cident wave originates in the dielectric.

From previous work, we know that in a dielectric the propagation number is


k12 = ε1 µ1 ω 2
and in a conductor

 
k22 2 2
= ε2 µ2 ω − iωσµ2 = ε2 µ2 ω 1 − .
ε2 ω
A characteristic property of propagation in the conductor is that in the limit
of of a good conductor, σ/(ε2 ω)  1, the propagation number k2 → ∞. As
we shall see this is a major factor for the propagation in the conductor to be
completely different than in the dielectric.

Consider the Snell’s law which can be


written as
k1
sin θt = sin θi .
k2
Assume that the metal is a good con-
ductor. Since in a good conductor k2 →
∞, we see from the above equation that
in the propagation from a dielectric to
a good conductor, sin θt → 0 indepen-
dent of θi . Thus, the direction of the
transmitted wave is normal to the sur-
Figure 68: Propagation of an EM face independent of the angle of inci-
wave from a dielectric to a conductor.
dence θi .

~ and H
As a consequence, the field vectors E ~ in the conductor lie tangential
to the boundary and so the normal components of these vectors on the con-
ductor side of the boundary are zero.

267
It follows that:

• Since the normal component of B ~ (or H)


~ is continuous across the
boundary, the normal component of H ~ or B~ is zero on the dielectric
side also.
Thus the normal component of B ~ of the reflected wave must be equal
and opposite to that of the incident wave.

• In a good conductor

k2


σ
2
2 = K 1 − →∞ as 1,
k1 εω εω

and then we have from the Fresnel equations

E2 → 0 and E1 → −E0 .

This means that the electric field in the conductor (which is tangential to
the boundary) → 0.
Since Ek is continuous across the boundary, we have that Ek is zero also in
the dielectric at the boundary.

Thus, the tangential component of E~ of the reflected wave must be equal and
opposite to that of the incident wave.

In summary, we have two useful special boundary conditions at the surface


between a dielectric and a perfect conductor:
~ = 0.
1. The tangential component of E
~ or H
2. The normal component of B ~ = 0.

268
19.2.1 Field vectors at normal incidence
We now consider the special case of normal incidence at a boundary, i.e. when
the wave propagation vector coincides with the normal to the boundary.
Figure 69 shows how the field vectors
must look in the plane of the surface be-
tween dielectric and conductor. Since we
already know that for a good conductor
for which
σ
→∞,
εω
the tangential component of the trans-
mitted electric field Et = 0, and the tan-
gential component of the reflected mag-
Figure 69: The field vectors in the netic field Hr = Hi , we find that the re-
plane of the surface between dielectric
lations between the field components take
and conductor.
the form

Et = Ei − Er → 0,
Ht = Hi + Hr → 2Hi .

In this case, the power reflection coefficient becomes

Er Hr E2 σ
αp = = r2 → 1 as →∞.
Ei Hi Ei εω

Thus, we have the total reflection of the energy carried by the EM wave.

We see an advantage of using dielectric-conductor boundary to propagate


an EM wave in a material.

1. The energy of the wave is completely reflected, and in fact the total
reflection is independent of the angle of incidence.

2. The polarization remains constant in the propagation.

Before concluding the lecture, let us clarify a problem regarding the reflec-
tion at the normal incidence. Namely, one could think that under the normal
incidence, there is only the incident and transmitted wave with no reflected

269
wave. Here, we prove the necessity of assuming the existence of the reflected
wave.

Assume that E~ is normal to the plane of incidence. Then, from the continuity
of the tangential components at the boundary, we have

Ei − Er = Et , (111)
Hi + Hr = Ht . (112)

However
s s s s
ε ε ε0 ε0
H= E= E=n E.
µ0 ε0 µ0 µ0

Thus, Eq. (112) can be written as

n1 Ei + n1 Er = n2 Et ,

and then we get two equations for the amplitudes of the electric field

Ei − Er = Et ,
n2
Ei + Er = Et .
n1
If Er is missing, we could not simultaneously satisfy both equations. Thus,
there always is a reflected wave in the normal incidence.

270
Revision questions

Question 1. What is a Brewster’s angle and why one could call it a polarizing
angle?

Question 2. Under what conditions will the reflected and transmitted amplitudes
for perpendicular polarization be the same as those for parallel po-
larization?

Question 3. Under which conditions an incident wave will be completely reflected


at a boundary between two dielectrics?

Question 4. What is meant by an evanescent wave?

Question 5. What are the boundary conditions at a surface between a dielectric


and a perfect conductor?

Question 6. In the propagation of an EM wave between a dielectric and a perfect


conductor, the direction of the transmitted wave is normal to the
boundary surface independent of the angle of incidence: True or
false?

Question 7. In the propagation between a dielectric and a perfect conductor


what happens to the electric field at the boundary?

Question 8. Is there a reflected beam at the normal incidence to a boundary


between two materials?

271
Tutorial problems

Problem 19.1 Brewster’s angle for different polarizations

Consider a propagation of an electromagnetic wave between two


non-conducting materials.

(a) For the polarization of the E ~ i in the plane of incidence, find


the relation between the critical angle of incidence and the Brew-
ster’s angle. Note, at the critical angle of incidence θt = π/2, and
at the Brewster’s angle of incidence θi + θt = π/2.

(b) Show that under the condition of no reflection at a boundary


between two non-conducting materials, the sum of the Brewster’s
angle and the angle of refraction is π/2 for

~ i normal to the plane of incidence, ε1 = ε2 = ε0 and µ1 6= µ2 .


(i) E

~ i in the plane of incidence, µ1 = µ2 = µ0 and ε1 6= ε2 .


(ii) E

Problem 19.2 Total internal reflection independent of polarization

Show, using the Fresnel’s equations that under the total inter-
nal reflection, energy of an incident wave is completely reflected
independent of polarization of the wave.

272
20 Propagation of an EM Wave in a Rectan-
gular Waveguide

In many practical problems we want to send a signal (EM wave) to differ-


ent places in an efficient way. Sending it in an uncontrolled way through
the free space in air is not a good choice. Due to scattering of the wave on
different objects there will be significant losses and uncontrolled changes of
the direction of propagation. Therefore, we need some device which would
help us to send the signal in more controlled way and where it would be pro-
tected from degradation. Conducting rectangular waveguides provide this:
the waves can be propagated and delivered to the specific points (receivers)
without significany losses.

In this lecture, we discuss the propagation of bounded EM waves by consid-


ering the propagation of radiation through a waveguide where the radiation
is fully confined in the transverse plane. We will consider the case where
the bounding walls are made of a conductor, are planar and cross section is
rectangular. Figure 70 shows an example of the rectangular waveguide.
An EM wave propagating along
the z direction may undergo a
multiple reflection from the walls.
Thus, we are likely to get stand-
ing waves in the x direction due
to reflections between the walls
x = 0 and x = a. We also
get standing waves in the y di-
rection due to reflections between
the walls y = 0 and y =
b.
Figure 70: A schematic diagram of a rect-
angular waveguide.
The principles behind the propaga-
tion of an EM wave along the rect-
angular waveguide are analysed by solving the Maxwell’s equations for the
fields inside the waveguide subject to the good conductor boundary condi-
tions being satisfied at x = 0, x = a, y = 0 and y = b. We write the z

273
dependence of any field component in the form

e−γz Thus ≡ −γ ,
∂z
where γ describes the propagation conditions, e.g. γ purely imaginary de-
scribes a wave propagating without loss.

~ H,
We describe the electromagnetic field by the vector pair E, ~ and we use
the Maxwell’s equations in the form

∇·E ~ = ρ =0,
ε
~
∇ · B = 0 or ∇ · H ~ =0,
~ ~
∇×E ~ = − ∂ B = −µ ∂ H ,
∂t ∂t
~ ~
∇×H ~ + ε ∂E ,
~ = J~ + ε ∂ E = σ E
∂t ∂t
where σ is the conductivity of the interior of the waveguide, not the bounding
metallic surfaces. We assume the interior if filled with a dielectric, but allow
a possible non-zero conductivity of the dielectric.

As we shall see, certain characteristic modes of propagation are found.


In general a characteristic mode is one which propagates with constant po-
larization.

In a rectangular waveguide, one may have TE (transverse electric) waves, or


TM (transverse magnetic) wave, but not TEM wave. If both E and B fields
are transverse, the wave would be going straight down the guide. However,
such a wave would not satisfy various boundary conditions and could not be
transmitted through the waveguide.

~ is transverse to the direction in which the


For a TE wave, the electric field E
wave is propagating that is, down the waveguide. The magnetic field B ~ is
perpendicular to E~ and therefore there is a substantial component of B ~ that
points in the direction of the waveguide.

274
20.1 Transverse Electric (TE) Modes

As we have already mentioned is possible to propagate a wave with the elec-


tric field polarized in the xy plane, a TE wave with E ~ transverse to the
waveguide axis, by lifting the restriction that the magnetic field is transverse
to the direction of propagation.

To show this, we will look for the solution of the Maxwell’s equations with
Ez = 0 that satisfies the good conductor boundary conditions.

The outline of the procedure is as follows:


First, using the Maxwell’s equations, we find relations between the compo-
nents of the electric and magnetic fields. Next, we convert these relations
into a single differential (wave) equation for Hz , the longitudinal component
~ Then we find the solution of the wave equation that satisfies the good
of H.
conductor boundary conditions.

We proceed as follows: Since the waveguide is of a rectangular shape, we


consider the Maxwell’s equations in Cartesian coordinates.
~ + µ ∂ H~ = 0, we have
• From III, ∇ × E ∂t

î ĵ k̂


∂x ∂
−γ ~
+ iωµH

=0.
∂y
E Ey 0
x

Hence

x component : γEy + iωµHx = 0 , (113)


y component : − γEx + iωµHy = 0 , (114)
∂Ey ∂Ex
z component : − + iωµHz = 0 . (115)
∂x ∂y

~ = 0, we have
• From II, ∇ · H
∂Hx ∂Hy ∂Hz
+ + =0,
∂x ∂y ∂z

275
and from the fact that ∂/∂z = −γ, we get

∂Hx ∂Hy
+ − γHz = 0 . (116)
∂x ∂y

~ + ε ∂ E~ ) = 0, we have
~ − (σ E
• From IV, ∇ × H ∂t

î ĵ k̂


∂x ∂
−γ − (σ
~ =0,
+ jωε)E
∂y
H Hy Hz
x

which in terms of the components gives


∂Hz
+ γHy − (σ + iωε)Ex = 0 , (117)
∂y
∂Hz
−γHx − − (σ + iωε)Ey = 0 , (118)
∂x
∂Hy ∂Hx
− =0 (Ez = 0) . (119)
∂x ∂y
~ = ρ/ε = 0, we have
• From I, ∇ · E

∂Ex ∂Ey
+ =0 (Ez = 0) , (120)
∂x ∂y

where we have used the TE condition Ez = 0.

We shall now express Ex , Ey , Hx and Hy in terms of Hz to get a wave equa-


tion for Hz .

From Eqs. (113) and (114), we get

Ex Ey iωµ
=− = . (121)
Hy Hx γ

Using this relation, we can then express Ex and Ey in terms of Hy and Hx


and then show that Eq. (115) becomes identical to Eq. (116).

276
Thus, we use Eq. (121) in Eqs. (117) and (118). In Eq. (117) substitute
for Ex :
∂Hz iωµ
+ γHy − (σ + iωε) Hy = 0 .
∂y γ
Hence
−1 ∂Hz
Hy = iωµ
γ − (σ + iωε) γ ∂y
−γ ∂Hz γ ∂Hz
= =− 2 , (122)
γ2 − iωµ(σ + iωε) ∂y k ∂y

where k 2 = γ 2 − iωµ(σ + iωε).

In Eq. (118) substitute for Ey and find

−γ ∂Hz γ ∂Hz
Hx = = − , (123)
γ 2 − iωµ(σ + iωε) ∂x k 2 ∂x

Using Eqs. (121), (122), and (123), we find that Eqs. (119) and (120) are
automatically satisfied.

Substituting Eqs. (122) and (123) into Eq. (116), we obtain two equations

−γ ∂ 2 Hz
,
γ 2 − iωµ(σ + iωε) ∂x2
γ ∂ 2 Hz
− − γHz = 0 ,
γ 2 − iωµ(σ + iωε) ∂y 2
which can be written as
∂ 2 Hz ∂ 2 Hz
+ + k 2 Hz = 0 , (124)
∂x2 ∂y 2

where k 2 = γ 2 − iωµ(σ + iωε).

We have obtained the wave equation for Hz which we will solve assuming the
good conductor boundary conditions.

277
Now, we proceed to solve Eq. (124) for Hz with the boundary conditions for
the field components at the surface of a good conductor. Having Hz , then we
can solve Eqs. (122) and (123) for Hx and Hy . With these solutions, we then
will be able to solve Eq. (121) for Ex and Ey . After that, we will know all
the field components. In other words, we will know how the fields propagate
through the rectangular waveguide.

20.2 Boundary Conditions

Let us proceed with the steps mentioned above. First, we solve the wave
equation (124) with the known boundary conditions at the surface of a good
conductor:
~ normal to boundary (in xy plane) = 0.
1. H
~ tangential to boundary (in xy plane) = 0.
2. E

According to Figure 71: We must have at the boundaries along the x-axis,
x = 0 and x = a, the field components Hx = 0 and Ey = 0. Looking at
Eq. (118), we see that this means that the derivative ∂Hz /∂x must be equal
to zero at x = 0 and x = a.

Similarly, at the boundaries along


the y-axis, at y = 0 and y =
b, the field components Hy and Ex
must vanish, Hy = 0 and Ex =
0. Looking at Eq. (117), we see
that this means that the derivative
∂Hz /∂y must be equal to zero at
the boundaries y = 0 and y =
b.

Summarizing, we see that the solu-


Figure 71: Electric and magnetic field
components at the surface of the waveguide. tion of Eq. (124), which satisfies these
boundary conditions is of the form

Hz = H0 cos(kx x) cos(ky y) eiωt−γz ,

278
with kx x and ky y satisfying the standing wave conditions:
kx a = mπ and ky b = nπ .
Hence, possible values for Hz inside the waveguide are
mπx nπy
   
Hz = H0 cos cos eiωt−γz , (125)
a b
where m, n = 0, 1, 2 . . ..

An (m, n) combination of the integer numbers represents a possible TE mode


of propagation. The modes are designated in the form TEmn .

What left is to determine the coefficient γ. With the solution (125), the wave
equation (124) gives a condition for k:
" 2 2 #
mπ nπ
 
2
− − +k Hz = 0 .
a b
For a non-trivial solution, Hz 6= 0, and then
2 2
mπ nπ
 
2
k = + .
a b
Since that on the other hand
k 2 = γ 2 − iµω(σ + iεω) ,
we find that
2 2
mπ nπ
 
2
γ = + + iµω(σ + iεω) . (126)
a b
We see that in general the propagation constant γ is a complex number. We
can write γ = β + iα. Then the solution (125) takes the form
mπx nπy
   
Hz = H0 cos cos e−βz ei(ωt−αz) .
a b
Thus, α = 2π/λg , where λg is the wavelength in the waveguide at frequency ω.
The phase velocity in the waveguide is vp = ω/α.

The parameter β is the attenuation coefficient describing losses in the waveg-


uide. Energy loss may be due to:

279
• Ohmic resistivity (σ of the medium finite).

• Dielectric losses (described by ε having an imaginary component)

• Magnetic losses (described by µ having an imaginary component)

20.3 TE Modes in a Lossless Waveguide


Suppose that the interior of the waveguide is filled with a perfect dielectric,
i.e. we assume a lossless propagation for which we have σ = 0 and ε and µ
both purely real. In this case, we find from Eq. (126) that
2 2
mπ nπ
 
2 2
γ = −εµω + + ,
a b
or introducing the phase velocity, we obtain

ω2
2 2
mπ nπ
 
2
γ =− 2 + +
v0 a b

where v0 = 1/ εµ is the phase velocity of propagation of waves in an
infinite (unbounded) medium of the type filling the waveguide.

It is convenient to express γ 2 in terms of the wavelength of the input wave.


Since
ω 2πf 2π
= = ,
v0 f λ0 λ0
where λ0 is wavelength of the input wave, or equivalently, the infinite size
medium wavelength, we can write the expression for γ 2 as
2 2 2
2π mπ nπ
  
2
γ =− + + .
λ0 a b

One can see that the parameter γ 2 can be negative or positive even after
all the assumptions of a lossless propagation. The nature of the propaga-
tion depends on the size of the cross section of the waveguide, and changes
according as:

280
1. Assume that γ 2 negative. In this case, γ is purely imaginary, and we
have a propagating wave
mπx nπy
   
Hz = H0 cos cos ei(ωt−kg z) ,
a b
where we have put γ = ikg , with kg the guide propagation constant.
The wave then propagates with a guide wavelength λg given by
s 2 2 2
2π 2π mπ nπ
 
= kg = − − .
λg λ0 a b
There is a maximum wavelength λ0 (= λc say) such that kg is real
2 2
1 m n
 
2
= + .
λc 2a 2b
Thus, there is a minimum frequency fmn such that the TEmn mode will
propagate down the waveguide
s
2 2
m n
 
fmn = vm + .
2a 2b
This can be derived from the cut-off condition fmn λc = vm .
Thus, the waveguide acts as a high-pass filter for any (m, n) mode.

2. Consider γ 2 positive. Then γ = β say is purely real and


mπx nπy
   
Hz = H0 cos cos e−βz eiωt .
a b
There is no phase propagation but amplitude attenuation. Note there
is no energy loss mechanism available. This is an evanescent mode
at frequencies less that fmn analogous to the case of total internal
reflection.
Field components in the TE modes.

Since
mπx nπy
   
Hz ∼ cos cos eiωt−γz ,
a b

281
we find from Eq. (122) the Hy component of the field
!
∂Hz mπx nπy
   
Hy ∼ ∼ cos sin eiωt−γz .
∂y a b

Then from Eq. (123), we find the Hx component


!
∂Hz mπx nπy
   
Hx ∼ ∼ sin cos eiωt−γz .
∂x a b

Next, from Eq. (121), we find the components of the electric field
mπx nπy
   
Ex ∼ Hy ∼ cos sin eiωt−γz ,
a b
and
mπx nπy
   
Ey ∼ Hx ∼ sin cos eiωt−γz .
a b
We see from the above equations that the mode m = n = 0 is not allowed to
propagate through the waveguide, since Ex , Ey , Hx and Hy all contain sine
terms so all the field components vanish for m = n = 0. All other TEmn
modes are allowed.22

TM (transverse magnetic) modes

Put Hz = 0 and go through the whole procedure again. Ez 6= 0 now. Equa-


tions analogous to Eqs. (122), (123), and (124) appear for components of
the E~ vector this time. Consequently the previous discussion about modes
and their cut-off frequencies for TE modes is also true for TM modes. The
only difference is that more modes are not allowed, i.e. in contrast to the TE
propagation, TM modes with either m = 0 or n = 0 are not allowed.23

22
The mode m = n = 0 is never possible in a transmission line consisting of a single
closed conductor like a rectangular waveguide. It is possible in 2-conductor lines e.g. the
coaxial line or the twin wire transmission line.
23
This is an interesting difference between properties the TE and TM modes, and the
student is encourage to analyze, as a tutorial problem, where the difference is coming from.

282
Special properties of the TE10 mode

We now proceed to discuss interesting properties of some of the propagating


modes. For example, If the transverse dimensions of the rectangular waveg-
uide are different (a 6= b) there is a finite range of frequencies over which the
TE10 mode is the only allowed mode.
This means that a waveguide can be
designed to allow propagation in one
mode only. In other words, it can work
as a frequency filter transmitting non-
degenerate modes. To show this, con-
sider the frequency of the TEmn mode
s
2 2
m n
 
fmn = vm + ,
2a 2b
from which we find that frequencies
of two neighbouring modes TE10 and
Figure 72: Electric and magnetic fields TE01 are
of the TE10 mode.
vm vm
f10 = < f01 = ,
2a 2b
where we adopt the convention that a > b. Thus, in the frequency range
from f10 → f01 , the TE10 mode is the only mode allowed.

20.4 Phase and Group Velocities of Mode Propagation

Consider the velocity with which the wave propagates inside the waveguide.
We look into the phase velocity, the velocity the wave front propagates
ω ω
vp = f λg = = r  .
kg 2π
2 

2 

2
λ0
− a
− b

Thus
v0
vp = r  2  2 ,
mλ0 nλ0
1− 2a
− 2b

283
where v0 is the phase velocity in the unbounded medium.

We see that the phase velocity of the wave inside the waveguide is greater
than in an unbounded medium, and so may be greater than the speed of
light in vacuum.

Alternately, we could write


2 2 2
mπ nπ 2π
  
+ = = kc2 ,
a b λc
where λc is the infinite medium wavelength at the cut-off frequency for
the m, n mode. Then
ω ω v0
vp = q = r 2 =r 2 , (127)
k02 − kc2
 
kc ωc
k0 1 − k0
1− ω

provided ω > ωc i.e. in the pass-band for that mode.

Note that for ω → ωc , the phase velocity vp → ∞. Moreover, if a vacuum or


air fills the waveguide then v0 = c and vp > c.

The group velocity

Since vp > c typically, we see that the phase velocity vp is not the velocity
of propagation of energy or information down the waveguide. It is prop-
agated with the group velocity which is smaller than the phase velocity if
the medium is dispersive. When dispersion is not present, phase and group
velocities are equal. According to Eq. (127), the waveguide is a dispersive
medium, since the phase velocity depends on frequency, so the energy or
information is propagated with the group velocity vg = dω/dkg that differs
from the phase velocity vp = ω/kg .

Just a brief explanation how do we define the group velocity. The group
velocity is the velocity of propagation of some modulation of multi-frequency
wave that carries information. A single frequency harmonic wave carries no
information. It is just there. A finite bandwidth is required to carry infor-
mation.

284
We can illustrate the concept of group velocity by considering a sum of two
cosine waves of slightly different frequencies, ω and ω + dω:

cos(ωt − kz) + cos[(ω + dω)t − (k + dk)z]


! " #
dω dk 2ω + dω 2k + dk
= 2 cos t − z cos t− z
2 2 2 2
!
dω dk
= 2 cos t − z cos(ωt − kz) ,
2 2
where we have assumed that dω  2ω and dk  2k.

The velocity of the amplitude modulation is



2 dω
vg = dk = .
2
dk

In the rectangular waveguide


q
2π q ω 2 − ωc2
kg = = k02 − kc2 = ,
λg vm
where k0 = ω/vm and kc = ωc /vm .

Remember, vm is the infinite medium phase velocity and ωc is the cut-off


(angular) frequency of the (m, n) mode. Then differentiating
dkg 1 1 2 1 ω
= (ω − ωc2 )− 2 2ω = q .
dω vm 2 vm ω 2 − ωc2

Finally
q s
dω vm ω 2 − ωc2  2
ωc
vg = = = vm 1 − < vm .
dkg ω ω
Thus, the group velocity of a wave propagating inside the waveguide is smaller
than the phase velocity. Moreover
s
2
ωc vm

2
vg vp = vm 1 − r = vm .
ω 
ωc
2
1− ω

285
In a vacuum-filled waveguide vg vp = c2 . Thus, relativity is still all right.

Revision questions

Question 1. Explain what is meant by TE and TM waves.

Question 2. Explain briefly the procedure of finding the components of the elec-
tric and magnetic fields transmitted through a waveguide.

Question 3. Why a TEM wave cannot be transmitted through a rectangular


waveguide?

Question 4. Explain the notation, what does it mean TEmn ?

Question 5. Explain, how a waveguide can work as a frequency filter.

Question 6. Define the phase and group velocities and the relation between them.

Question 7. In a vacuum waveguide the phase velocity of an EM wave is greater


than speed of light: True or false?

286
21 Relativistic Transformation of the Elec-
tromagnetic Field

The final part of the course is devoted to relativistic effects in the EM theory.
Using the argument that the Maxwell’s equations, as physical observables,
are invariant under the Lorentz transformation, we shall show how the fields
transform according to the relativistic rules. Then, we will illustrate on few
examples, how the fields change when one goes between different inertial
frames and how this affects propagation of an EM wave. Finally, we show
how frequency and energy transform according to the transformation rules
and point out the evidence of the dependence of the energy on frequency.

21.1 The Principle of Relativity


We begin from the recollection of the principles of relativity.

1. The laws of physics are the same in all inertial reference frames.

2. The speed of light in vacuum is independent of the uniform motion of


the observer or source.

The constancy of the velocity of light, independent of the motion of the


source, gives rise to the relations between space and time coordinates in dif-
ferent inertial reference frames known as Lorentz transformations.

Consider a stationary reference frame S and a inertial frame S 0 moving with a


velocity ~u parallel to the x axis, i.e. ~u = uî. The time and space coordinates
in S 0 are related to those in S by the Lorentz transformations

x0 = γ(x − βct) ,
y0 = y,
z0 = z,
ct0 = γ(ct − βx) ,
−1/2
where γ = (1 − β 2 ) is the Lorentz factor, and β = u/c.

287
The above transformation corresponds to a situation of ~u parallel to the x
axis. If the axis in S and S 0 remain parallel, but the velocity ~u of the frame S 0
is in an arbitrary direction, the generalization of the above transformations is
~ β~
(~r · β)
~r 0 = ~r + (γ − 1) ~ ,
− γ βct
β2
 
ct0 = γ ct − β~ · ~r ,

where β~ = ~u/c.

Proof:

Decompose the vector ~r into two components: parallel and normal to ~u

~r = ~rk + ~r⊥ .

Then, using the one dimensional Lorentz transformations, we have


 
~
~rk0 = γ ~rk − βct ,
~r⊥0 = ~r⊥ .

However, we can write the parallel and normal components as


~ β~
(~r · β)
~rk =
β2
~r⊥ = ~r − ~rk .

Hence
 
~
~r 0 = ~rk0 + ~r⊥0 = γ ~rk − βct + ~r − ~rk
~ β~
(~r · β)
= ~r + (γ − 1) ~ .
− γ βct
β2
The transformation of time can be proved in the similar way. Since
 
ct0 = γ ct − βrk ,

and
~
(~r · β)
βrk = β~ · ~rk = β~ · β~ = ~r · β~ ,
β2

288
we obtain
 
ct0 = γ ct − β~ · ~r ,

as required.

We will need the inverse Lorentz transformations, which are


~ β~
(~r 0 · β)
~r = ~r 0 + (γ − 1) ~ 0,
+ γ βct
β2
 
ct = γ ct0 + β~ · ~r 0 .

Note that ~r is a function of ~r0 and t0 .

The principle of relativity indicates, and we have showed it explicitly in


Chap. 7, that the Maxwell equations are invariant under the Lorentz
transformation. The same rule applies to the continuity equation.

Thus, the Maxwell’s equations together with the continuity equation have
the same form in two different inertial frames.

If in the frame S:

∇·D ~ = ρ, ~ =0,
∇·B
~ ~
∇×E ~ = − ∂B , ∇×H~ = J~ + ∂ D ,
∂t ∂t
∂ρ
∇ · J~ = − ,
∂t
then in the frame S 0 :
~ 0 = ρ0 ,
∇0 · D ~0 = 0 ,
∇0 · B
~0 ~0
∇0 × E ~ 0 = − ∂B , ∇ 0
× ~ 0 = J~ 0 + ∂ D ,
H
∂t0 ∂t0
0
∂ρ
∇0 · J~ 0 = − 0 ,
∂t
where the prime variables are functions of the transformed variables, t0 and ~r 0 .

289
Of course, the EM fields together with the current and charge densities in
the S 0 frame must be different from that in the S frame in order to match the
same Maxwell equations. From this, an interesting question arises: What are
the relations between the EM fields, charge and current densities in the S 0
and S frames?

In order to answer this question, we have to find the transformation of a


time derivative of an arbitrary scalar function Ψ, and the transformation of
divergence ∇ · F~ .

Consider the transformation of a time derivative ∂Ψ/∂(ct)


∂Ψ ∂Ψ ∂ct0 ∂Ψ ∂x0 ∂Ψ ∂y 0
= + +
∂(ct) ∂(ct0 ) ∂(ct) ∂x0 ∂(ct) ∂y 0 ∂(ct)
∂Ψ ∂z 0 ∂Ψ ∂ct0 0 ∂~r 0
+ = + ∇ Ψ · .
∂z 0 ∂(ct) ∂(ct0 ) ∂(ct) ∂(ct)
However
∂ct0 ∂~r 0
=γ , = −γ β~ ,
∂(ct) ∂(ct)
which gives
!
∂Ψ ∂
=γ − β~ · ∇0 Ψ .
∂(ct) ∂(ct0 )
Consider now the divergence
∂Fx ∂Fy ∂Fz
∇ · F~ = + + .
∂x ∂y ∂z
Since
∂Fx ∂Fx ∂ct0 ∂Fx ∂x0
= +
∂x ∂(ct0 ) ∂x ∂x0 ∂x
∂Fx ∂y 0 ∂Fx ∂z 0
+ + ,
∂y 0 ∂x ∂z 0 ∂x
and
∂ct0
= −γβx ,
∂x
290
∂x0 β2
= 1 + (γ − 1) x2 ,
∂x β
0 0
∂y ∂z
= =0,
∂x ∂x
we obtain
∂Fx ∂Fx ∂Fx
= −γβx 0
+ αx 0 ,
∂x ∂(ct ) ∂x
where

αx = 1 + (γ − 1)βx2 /β 2 .

Similarly, for Fy and Fz , and finally we get

∂ F~
∇ · F~ = ᾱ ∗ (∇0 · F~ ) − γ β~ · ,
∂(ct0 )
where ᾱ is a 3 × 3 diagonal matrix
2


1 + (γ − 1) ββx2 0 0

β2

ᾱ = 0 1 + (γ − 1) βy2 0 ,


2

0 0 1 + (γ − 1) ββz2

as required.

Using the above transformations, we can derive transformations for the cur-
rent density J~ and the charge density ρ.
In order to do it, we consider the continuity equation, that can be written as
∂cρ
∇ · J~ = − .
∂(ct)
Hence
~
!
0~ − γ β~ · ∂ J = −γ
ᾱ ∗ (∇ · J)

− β~ · ∇0 cρ ,
∂(ct0 ) ∂(ct0 )
or

~ − γ β~ · ∇0 (cρ) = −γ ∂  
~ · J~ .
∇0 · (ᾱ ∗ J) cρ − β
∂(ct0 )

291
Since

β~ · ∇0 (cρ) = ∇0 · (cρβ)
~ ,

we obtain
  ∂ h  i
∇0 · ᾱ ∗ J~ − γcρβ~ = − γ cρ − ~ · J~ .
β
∂(ct0 )
Thus, the continuity equation will be invariant under the Lorentz transfor-
mation if
 
cρ0 = γ cρ − β~ · J~ ,
J~ 0 = ᾱ ∗ J~ − γcρβ~ . (128)

In order to understand the physical meaning of these equations, consider the


following example.

Example

Assume that in the S frame there is a stationary volume charge of density


ρ 6= 0. Since ρ is stationary, there are no currents in the S frame (J~ = 0).
What are the charge and current densities as seen in the S 0 frame?

In the S frame

J~ = 0 , ρ 6= 0 .

According to Eq. (128), in the S 0 frame

J~ 0 = −γcρβ~ , ρ0 = γρ .

Thus, there is a current in the S 0 frame. As seen from S 0 a given part of the
charge is length contracted in the direction of motion so the charge density
is correspondingly increased by the factor γ > 1. The length contracted
charge density appears from S 0 to move in the opposite direction. We can
understand this result: The stationary charge in the S frame moves with
velocity −~u in the S 0 frame. This effect is predicted by the Galileo transfor-
mation.

292
Less obvious and more interesting is the following situation. Let’s in the S
frame one observes

J~ 6= 0 , ρ=0.

Then, according to Eq. (128), someone will see a non-zero charge density
ρ0 6= 0 in the S 0 frame.
This is a pure relativistic effect, which cannot be predicted by the Galileo
transformation. Note, the charge ρ0 is proportional to u/c2 .

21.2 Transformation of Electric and Magnetic Field


Components

To find the transformation rules for electric and magnetic field components
we will use the transformations of the time and space derivatives derived
above.

Consider two of the Maxwell equations that in the S frame are


~ = ρ,
∇·D
~
~ = J~ + ∂ D .
∇×H
∂t
These equations should be equivalent of two equations

~ 0 = ρ0 ,
∇0 · D
~0
~ 0 = J~ 0 + ∂ D ,
∇0 × H
∂t0
in the S 0 frame.

Using the transformations of the time and space derivatives, we have


" # !
∂ ∂
ᾱ ∗ ∇ − γ β~
0
0
~ −γ
×H 0
− β~ · ∇0 cD ~ = J~ ,
∂(ct ) ∂(ct )
" #
0 ~ ∂ ~ = cρ .
ᾱ ∗ ∇ − γ β · cD
∂(ct0 )

293
Substituting the transformations of J~ and ρ, we find that the D
~ and H
~
vectors transform as
 
cD ~ + β~ × H
~ 0 = γ ᾱ−1 ∗ cD ~ ,
 
~ 0 = γ −β~ × cD
H ~ + ᾱ−1 ∗ H
~ , (129)

where ᾱ−1 is the inverse of the matrix ᾱ


2


1 + ( γ1 − 1) ββx2 0 0

β2

ᾱ−1 = 0 1 + ( γ1 − 1) βy2 0 .


2

0 0 1 + ( γ − 1) ββz2
1

From the Maxwell equations


~ = 0,
∇·B
~
~ = − ∂B ,
∇×E
∂t
~ and B
we find that the E ~ vectors transform as
 
E ~ + β~ × cB
~ 0 = γ ᾱ−1 ∗ E ~
 
~ 0 = γ −β~ × E
cB ~ + ᾱ−1 ∗ cB
~ . (130)

21.3 Transformation Rules in Terms of Parallel and


Normal Components

Suppose that the frame S 0 is moving with speed u in the direction paral-
lel to the z axis. In this case, βx = βy = 0, βz = β 6= 0, and then the
transformations take the form
~ 0 = γcDx î + γcDy ĵ + cDz k̂ + γβ k̂ × H
cD ~ ,
~ 0 = −γβck̂ × D
H ~ + γHx î + γHy ĵ + Hz k̂ ,
~ 0 = γEx î + γEy ĵ + Ez k̂ + γcβ k̂ × B
E ~ ,
~ 0 = −γβ k̂ × E
cB ~ + γcBx î + γcBy ĵ + cBz k̂ . (131)

294
It is useful to rephrase the transformation rules in terms of components
parallel and normal to ~u. The parallel components are the z components
and the normal components lie in the xy plane. For example
 
~ =E
E ~⊥ + E
~ k = Ex î + Ey ĵ + Ez k̂ ,

~ H
and the same for D, ~ and B.
~

21.3.1 Rules for Parallel Components

From Eq. (131), we readily find that


~ k0 = E
E ~ z0 k̂ = E
~ z k̂ = E
~k ,

and similarly
~ k0 = B
B ~k , ~ k0 = H
H ~k , ~ k0 = D
D ~k .

Thus, the components parallel to the direction of ~u are invariant under the
transformations (129) and (130).

21.3.2 Rules for Normal Components

Consider now the components perpendicular to ~u:


~ 0 = cD 0 î + cD 0 ĵ
cD ⊥ x y

= γcDx î + γcDy ĵ + γβHx ĵ − γβHy î


= γ (cDx − βHy ) î + γ (cDy + βHx ) ĵ ,

~ 0 = H 0 î + H 0 ĵ
H⊥ x y

= γ (Hx + βcDy ) î + γ (Hy − βcDx ) ĵ ,

~ ⊥0 = Ex0 î + Ey0 ĵ
E
= γ (Ex − βcBy ) î + γ (Ey + βcBx ) ĵ ,

295
~ 0 = cB 0 î + cB 0 ĵ
cB⊥ x y

= γ (cBx + βEy ) î + γ (cBy − βEx ) ĵ ,

which can be written as


 
cD ~ ⊥ + β~ × H
~ 0 = γ cD ~⊥

 
H ~ ⊥ − β~ × cD
~0 = γ H ~⊥

 
E ~ ⊥ + β~ × cB
~0 = γ E ~⊥

 
cB ~ ⊥ − β~ × E
~ 0 = γ cB ~⊥ . (132)

We shall illustrate the transformation rules on the following examples:

Example 1 - purely electric field in S

~ 6= 0 but B
Suppose that in S, one observes E ~ = 0.
0
Then from the transformation rules, in S :
~ k0 = E
E ~k , ~ ⊥0 = γ E
E ~⊥ ,

~ k0 = 0 ,
B ~ ⊥0 = − γ ~u × E
B ~⊥ .
c2
Thus
~ ~
~ ⊥0 = − ~u × E⊥ = − ~u × E ,
~0 =B
B
c2 c2
~ k = 0.
since ~u × E

Thus what appears to be purely an electric field to one observer is seen as


both an electric and a magnetic field to a second observer moving with re-
spect to the first.

Example 2 - purely magnetic field in S

296
~ = 0, but B
Now suppose that in S, one observes E ~ =
6 0.
0
Then using the transformation rules, in S :
~ k0 = B
B ~k , ~ ⊥0 = γ B
B ~⊥ ,

~ k0 = 0 ,
E ~ ⊥0 = γ~u × B
E ~⊥ .
Thus
~0 =E
E ~ 0 = ~u × B
~ ⊥ = ~u × B
~ .

We see that what appears to be a purely magnetic field for one observer will
appear to be both an electric and a magnetic field to a relatively moving ob-
server.

This result could be used to calculate the emf in an electric dynamo from
the point of view of an observer watching the conductor move in a magnetic
field or from the point of view of an observer moving with the conductor.

21.4 Transformation of the Components


of a Plane EM Wave

In this final lecture of the course, we illustrate the transformation rules of


the components of a plane EM wave. We illustrate this on two examples.

Example 1

Suppose that a plane wave propagates in vacuum along the z axis. Then the
electric and magnetic fields of the wave are
E~ = îEei(ωt−kz) = îE0 = îEx ,
~ = ĵBei(ωt−kz) = ĵB0 = ĵBy .
B
Hence from the transformation rules (132), in the S 0 frame moving in the
same direction:
 
~ 0 = γ E0 î − βcB0 î = γ (E0 − uB0 ) î ,
E
 
~ 0 = γ cB0 ĵ − βE0 ĵ = γ (cB0 − βE0 ) ĵ .
cB

297
Since in vacuum

cB0 = E0 ,

we obtain
s
u
~0 =γ 1− u 1− c−u
 
c
E E0 î = r E0 î = E0 î ,
c  2
u c+u
1− c

and
s
~ = γ 1 − u cB0 ĵ = c−u
 
0
cB cB0 ĵ .
c c+u

Thus, we observe that the amplitudes of the fields are reduced, but the ratio
~ 0 |/|B
|E ~ 0 | = |E0 |/|B0 | is constant and independent of u, as it should be, since
speed of light is the same in all inertial frames.

This is consistent with the principle of relativity that speed of light is inde-
pendent of the motion of the observer.

Example 2

Suppose that an observer S 0 moves in the direction of the electric field of


an EM wave, i.e. ~u = uî, as shown in Figure 73.
In this case

~ 0 = E0 î + γβcB0 k̂ = î + γ u k̂ E0 ,
 
E
c
and
~ 0 = γcB0 ĵ .
cB

Thus, the magnetic field remains un-


changed, but the electric field turns to-
wards the direction of propagation of the
wave.
Figure 73: Text

298
An interesting situation occurs when the
velocity u → c. In this case, we have
u 1 u
γ =r →∞,
c 
u
 2 c
1− c

and then
~ 0 = E k̂ ,
E
~ 0 becomes perpendicular to ~u.
i.e. the direction of E

However, the Poynting vector of the wave is still in the direction of ~u:

~ = γ u E0 γB0 k̂ × ĵ = −γ 2 u E0 B0 î .
~0×H
E
c c

21.5 Doppler Effect

Consider a plane wave propagating in vacuum


~
E(t) ~ 0 ei(ωt−~k·~r) .
=E

In the moving frame S 0 this wave will have a different frequency ω 0 and the
wave vector ~k 0 , but the phase of the wave will remain unchanged as it is
invariant under the transformation, i.e.

φ = ωt − ~k · ~r = ω 0 t0 − ~k 0 · ~r 0 .

Using the inverse Lorentz transformation


~ 0,
~r = ᾱ ∗ ~r 0 + γ βct 
ct = γ ct0 + β~ · ~r 0 ,

we will find the relations between ω 0 , ω and k 0 , k that ensure the invariance
of the phase of the wave.

299
Thus, with the above transformations, we find
ω 0 t0 − ~k 0 · ~r 0 = ωt − ~k · ~r
ωγ
= ct0 + β~ · ~r 0 − ᾱ ∗ ~k · ~r 0 − γ~k · βct
~ 0
c
ωγ ~
 
0 0~ ~
= ωγt − γt k · ~u + β − ᾱ ∗ k · ~r 0
c
ωγ ~
   
~
= γ ω − k · ~u t − 0 ~
β − ᾱ ∗ k · ~r 0 .
c
Hence
 
ω 0 = γ ω − ~k · ~u ,

~k 0 = ᾱ ∗ ~k − ωγ ~u .
c2
Consider two special cases. In the first case, assume that the wave propagates
ˆ
in vacuum, ~k = (ω/c)~k. Then
u
 
ω 0 = γω 1 − cos θ ,
c
where θ is the angle between the direction of propagation of the wave and
the direction of the motion ~u.
For the propagation direction θ = 0
s
c−u
ω0 = ω .
c+u
The frequency in S 0 is smaller than that
in S, and ω 0 → 0 as u → c. On the
other hand, the frequency ω 0 → ∞ when
u → −c.

If the wave propagates in a material body,


where vf < c, we have
Figure 74: Text ~k = (ω/vf )k̂ ,

and then the frequency in S 0 is


!
0 u
ω = γω 1 − cos θ .
vf

300
For the case of θ = 0
u
1− vf
ω0 = ω r  2 .
u
1− c

When u = vf , the frequency ω 0 = 0, but for c > u > vf , we obtain that the
frequency ω 0 < 0.

21.6 Transformation of Energy of a Plane EM Wave

Consider an EM wave propagating in the ~k direction and an observer moving


in the z direction, as shown in the Figure 75.
Let E~ = E î. Then, in the frame S 0 :

E ~ + γ β~ × cB
~ 0 = γE ~⊥
~ + γ cos φ β~ × cB
= γE ~ .

~ = B ĵ + Bk k̂ (the wave propagates in the plane yz), and cB = E,


Since B
we obtain
~ 0 = γE
E ~ − γ cos φ βcB î
~ .
= γ (1 − β cos φ) E

Consider now energy of an electric field


of a plane EM wave confined in a vol-
ume ∆V . In the S frame
1
We = ε0 E02 ∆V .
4
In the S 0 frame
1
We0 = ε0 (E00 )2 ∆V 0 .
4
However

Figure 75: An EM wave is propagat- ∆V 0 = ∆x0 ∆y 0 ∆z 0 .


ing in the ~k direction and an observer
is moving in the z direction.

301
Since, we have assumed that the observer (S 0 frame) moves in the direction
of the z axis, and the wave propagates in the direction ~k · ~u = u cos φ, we
obtain

∆V 0 = ∆x∆y∆z 0 ,

where
∆z
∆z 0 =   .
γ 1 − uc cos φ

Hence
2
1 u ∆V

We0 = ε0 E02 γ 2 1 − cos φ  
4 c γ 1 − uc cos φ
u
 
= We γ 1 − cos φ .
c
It is interesting to compare the transformation of energy with the transfor-
mation of frequency. Since
u
 
We0 = We γ 1 − cos φ ,
c
and
u
 
0
ω = ωγ 1 − cos φ ,
c
we see that the energy and frequency transform in the similar way!

This indicates that We ∼ ω. Note, that a similar proportionality was pre-


dicted in quantum physics, We = h̄ω.

In the electromagnetic theory, the proportionality forms backgrounds of the


so called quantum electrodynamics.

302
Revision questions

Question 1. Are the Maxwell’s equations invariant under the Lorentz transfor-
mation? Explain.

Question 2. What are the rules for transformation of the electric and magnetic
field components?

Question 3. What are the rules for transformation of the components of an EM


wave?

Question 4. What is Doppler effect?

Question 5. How does the energy of an EM wave transform in the relativistic


case?

303
Tutorial problems

Problem 21.1 A relativistic linear dynamo

A linear dynamo is constructed using a hor-


izontal U-shaped or rectangular section of
wire with one side moving at speed v = dxdt
~ as shown in
in a vertical magnetic field B
Figure 76.

The rectangle has width `. The dynamo is


relativistic in that v may be comparable to c.
The general definition of emf round a cir-
Figure 76: cuit is:
I ~
F
I
E =W = · d~` = ~ · d~` .
~ + ~v × B)
(E
q

(a) Calculate the emf induced along the moving arm:

(i) In the frame of reference in which the U-shaped section is


at rest.

(ii) In the frame of reference in which the slide-wire is at rest.

Do you arrive at the same value of emf in each frame? If not,


does that bother you?

(b) In practice, emf ’s are usually calculated using Faraday’s Flux


Cutting Rule because it is the simplest way mathematically. The
flux cutting rule E = − dΦdt may be thought of as an integral of the
Maxwell equation ∇ × E ~ = − ∂ B~ . Now Maxwell’s equations are
∂t
supposed be ‘relativistically correct’, i.e. they take the same form
in all inertial frames of reference. Show that the flux cutting rule
holds in both reference frames mentioned above.

Problem 21.2 Plane wave propagating in a material

Suppose that a plane wave propagates in a material with ε 6= ε0


and µ 6= µ0 . Show that in the frame S 0 , moving with velocity u

304
in the direction of propagation of the wave (z-axis), the ratio
~ 0 |/|H
|E ~ 0 | depends on the velocity u.

Problem 21.3 Electric and magnetic fields in a moving frame

Suppose that in a stationary frame S one has measured an uniform


~ = E0 î and an uniform magnetic field B
electric field E ~ = B0 ĵ,
such that cB0 > E0 .

Is it possible to find a frame S 0 moving along the z-axis in which


one could not measure either E ~ 0 or B
~0 ?

If yes, find the amplitude of the measured field and the veloc-
ity u of the frame S 0 .

305
Revision Questions for the Final Examination
Question 4.1 Prove that the total electric flux through a closed surface S is
proportional to the total charge inside the surface.

Question 4.2 Prove the Amperes circuit law.

Question 4.3 Derive the integral form of the Faraday’s law and then transform
it into the differential form
~
∂B
~ =−
∇×E .
∂t

Question 5.1 Starting from the Maxwell’s equations derive the continuity equa-
tion, i.e. show that conservation of charge is built into the Maxwell’s
equations.

Question 6.1 ~ satisfies the same


Using the Maxwell’s equations show that B
~
wave equation as E.

Question 6.2 Show, using the proof of solution of the wave equation, that
f (z + ct) represents a signal propagating in the negative z direc-
tion with speed c.

Question 7.1 Show that magnetic and electric fields of a charge moving with a
constant velocity ~v are related by
~
~ = ~v × E .
B
c2
Question 7.2 Show that the magnetic field lines produced by a charge moving
with a constant velocity ~v form concentric rings about ~v .

Question 7.3 Show that the electric field produced by a moving charge satisfies
the Gauss’s law.

Question 7.4 Explain the statement: Electric and magnetic fields do not pro-
duce each other - they are both due to electric charges.

306
Question 8.1 Using the Maxwell’s equations derive the continuity equation for
the Poynting vector.

Question 8.2 Show, using the field theory calculation, that the power dissipated
along a resistive wire is P = V I, the same predicted by the circuit
theory.

Question 9.1 (a) Obtain a series solution of the Laplace equation in the cartesian
coordinates for the electrostatic potential inside a closed three di-
mensional area.

(b) What would be the solution of the Laplace equation for a


closed two dimensional problem.

Question 9.2 Consider a conducting sphere of radius R. The surface of the


sphere is kept at a potential

Φ(R, θ, φ) = V0 sin θ sin φ .

Using the above as a boundary condition, find the potential at


any point inside the sphere.

Question 10.1 Show that in general the Maxwell’s equations cannot be simpli-
~ and B
fied to two separate differential equations for the E ~ fields.
Then show that by introducing the concept of vector and scalar
potentials one can arrive to differential equations for A~ and Φ
that can be separated from each other under the Lorenz gauge.

Question 10.2 Explain, why the Coulomb gauge is often called ”Transverse
gauge”.

Question 10.3 Prove that the homogeneous wave equation

1 ∂2Φ
∇2 Φ − =0
c2 ∂t2
has the solution of the form
f (t − r/c)
Φ(r, t) = ,
r

307
where f (t − r/c) is an arbitrary function of the retarded time
t − r/c.

Question 11.1 Show that in spherical polar coordinates, the magnetic field of a
~ = ∆lI
short current element I ∆l ~ 0 exp(iωt) has only an azimuthal
component of the form

~ = I0 ∆l ik + 1 sin θ ei(ωt−kr) φ̂ .
 
B
4πε0 c2 r r2
Question 11.2 Show that in the far field zone of a radiating short current ele-
ment, the electric and magnetic fields oscillate in phase and are
orthogonal to each other.

Question 11.3 (a) Given the expressions for the EM field of a Hertzian dipole,
show that the total radiated power from the dipole is
2
πI02 ∆l

W = .
3ε0 c λ

(b) Show that the emitted power is equivalent to the power lost
on a resistor of resistance
2
∆l

R = 80π 2 .
λ
Question 11.4 Show that the time averaged Poynting vector of the EM field
emitted by a short current element is maximal in the equatorial
plane of the element.

Question 12.1 Show that the electrostatic potential due to a distribution of elec-
tric dipoles of moment per unit volume P~ throughout a volume V
enclosed by a surface S is that of a volume charge density −∇ · P~
together with a surface charge density P~ · n̂.

Question 12.2 Illustrate the Lorentz theory of polarizability of dense dielec-


tric materials and derive the Caussius-Mossotti relationship for a
dense dielectric.

Question 12.3 Show that the polarization of a dielectric driven by a time varying
electric field lags in phase the driving field.

308
Then show that the phase difference between the polarization and
a time-varying electric field results in a complex permittivity of
the dielectric.

Question 13.1 Explain the concept and advantage of introducing the magnetic
~
intensity vector H.

Question 13.2 Show that inside a ferromagnet H = 0 and explain the physical
meaning of this result.

Question 13.3 Derive the Maxwell’s equations for the EM fields in electric and
magnetic materials.

Question 14.1 (a) Derive the special form of Poynting’s Theorem applicable in
certain material media
~ ~
~ · ∂ B dV − ~ · ∂ D dV −
I Z Z Z
~ × H)
(E ~ · dS
~=− H E ~ · J~ dV .
E
S ∂t ∂t
(b) Interpret the above equation in terms of energy storage and
energy flow etc.

(c) State qualitative meaning of each term in the equation.

Question 14.2 Prove that the useful formula for the mean Poynting vector for
sinusoidal fields is
¯~ 1 ~ 
~∗ ,
N = Re E c × Hc
2
~c = E
where E ~ 0 exp(iωt) and H
~ c∗ = H
~ ∗ exp(−iωt).
0

Question 15.1 Show that the amplitude of a plane wave propagating in a non-
conducting material is damped with the rate β which arises from
the imaginary parts of the complex permittivity and permeability.

Question 15.2 Derive the expressions for the attenuation coefficient and the
phase velocity of an EM wave propagating in a conducting medium.
What are the values of the quantities for a propagation inside a

309
good conductor?

Question 15.3 Show that in a good conductor an EM wave propagates on the


surface of the conductor.

Question 16.1 Prove the following boundary conditions at a bounding surface


between two dielectrics:

~ is continuous across the bound-


(a) The normal component of B
ary.

~ is continuous across the bound-


(b) The tangential component of E
ary.

~ is continuous across the bound-


(c) The tangential component of H
ary.

Question 17.1 Show, using the Maxwell’s equations that the electric and mag-
netic vectors of an EM wave are related by
~
~ = B = k n̂p × E
H ~ ,
µ ωµ
where n̂p is the unit vector in the direction of propagation of the
wave.

Question 17.2 Prove that in the reflection and refraction at a bounding surface,
the direction of incident, reflected and refracted waves are copla-
nar.

Question 17.3 Derive the familiar laws of elementary optics:

(a) Angle of reflection equals to the angle of incidence.

(b) Snell’s law of refraction.

Question 17.4 Show, using the continuity conditions for E ~ and H ~ that both
reflection and refraction takes place in the incidence of light on a
boundary between two dielectrics.

310
Question 19.1 Show that under the Brewster’s angle of incidence there is no
reflected electric field in the plane of the incidence.

Question 20.2 Show that in a vacuum-filled rectangular waveguide vp > c, and


vg vp = c2 .

Question 21.1 Find the condition under which the continuity equation for ρ
and J~ is invariant under the Lorentz transformation.

Question 21.2 Show that electric charge Q is invariant under the Lorentz trans-
formation.

311
Appendix A: Proof of the Amperes Law

Consider a long wire of radius a carrying current I. Let P is a point on the


integration path, as shown in Figure 77.

Figure 77: The source circuit and the integration path to prove the Ampere law.

The magnetic field at P is

~ µ0 I I d~s × (−r̂)
B= .
4π s r2

Moving P by d~` is equivalent to moving the current circuit by −d~`.


The solid angle subtended by −d~`, d~s at P is:

(−d~` × d~s) · r̂ −d~` · d~s × r̂ d~s × −r̂


2
= 2
= d~` · .
r r r2

(The element of area normal to r̂ is −d~` × ~s · r̂)

Thus due to the path element d~`, the change in solid angle subtended at P
by the circuit is:
I
d~s × (−r̂)
dΩ = d~` · .
s
r2

312
Hence
4π I ~ 4π ~ ~
dΩ = d~` · dB = B · d` ,
µ0 I s µ0 I

~ at some
where integration is around the circuit s giving the magnetic field B
point P as shown in the diagram.

Now integrating round the closed path:

~ · d~` = µ0 I dΩ .
I I
B

` `

If P moves round a closed path (returning to its original position but not
circulating through the current loop:
I
dΩ = 0 .
`

But if P circulates through the loop:


I
dΩ = 4π ,
`

and then
~ · d~` = µ0 I 4π = µ0 I .
I
B

`

We conclude that the line integral of the magnetic field round a closed loop
path is equal to µ0 I, where I is the current passing through the loop.

313
Appendix B
Proof of the vector theorem, Eq. (73):
Z ~
M Z ~
M × n̂
− ∇× dV = dS .
V R S R
This is an application of the more general theorem
Z Z
− ∇ × F~ dV = F~ × n̂ dS .
V S

~ be a constant vector. Then


Let C
∇ · (F~ × C)
~ = (∇ × F~ ) · C
~ − (∇ × C)
~ · F~ = C
~ · (∇ × F~ ) . (133)
We will prove the general theorem by using the divergence theorem
Z Z
∇ · (F~ × C)
~ dV = (F~ × C)
~ · n̂ dS .
V S

~ and using (133), we get


For an arbitrary constant vector C,
Z Z

C ∇ × F~ dV = ~ · ∇ × F~ dV
C
V ZV Z
= ∇ · (F~ × C)
~ dV = F~ × C
~ · n̂ dS .
V S

Hence
Z Z

C ∇ × F~ dV = − ~ × F~ · n̂ dS .
C
V S

However
~ × F~ · n̂ = C
C ~ · F~ × n̂

and then we obtain


Z Z

C ∇ × F~ dV = −C
~· F~ × n̂ dS .
V S

~ we finally have
Since this is true for arbitrary C,
Z Z
∇ × F~ dV = − F~ × n̂ dS ,
V S

as required.

314
Appendix C: PHYS3050 Facts and Formulae
I Z
Gauss0 Divergence Theorem : F~ · n̂ dS = ∇ · F~ dV
S V
I Z
Stokes0 s Theorem : F~ · d~` = ∇ × F~ · n̂ dS
` S

Numerical values in SI units:


ε0 = 8.85 × 10−12 , µ0 = 4π × 10−7 , c = 3 × 108 [ms−1 ] .
−19
For the electron: e = 1.6 × 10 [C] , m = 9.11 × 10−31 [kg]

The Lorentz force law : F~ = q(E


~ + ~v × B)
~

1 q1 q2
Coulomb0 s Law : F~ = r̂
4πε0 r2

~ = µ0 I d~` × r̂
Biot − Savart Law : dB
4π r2
~ · n̂ dS = Q
I
Gauss0 Law : E
ε0
I
0
Ampère s Circuital Law : ~ · d~` = µ0 I
B

Maxwell’s Equations in vacuum:

∇·E ~ = ρ , ∇·B ~ =0,


ε0
~ ~
∇×E ~ = − ∂B , ∇×B ~ = µ0 J~ + 1 ∂ E .
∂t c2 ∂t
Maxwell’s Equations in material bodies:

~ =ρ,
∇·D ~ =0,
∇·H
~ ~
~ = − ∂B ,
∇×E ~ = J~ + ∂ D .
∇×H
∂t ∂t
315
Poynting vector:
~ = ε 0 c2 E
N ~ ×B
~

Poynting’s Theorem:

1 B2
!
~ · J~ dV − ∂ 1
I Z Z
2~ × B)
ε0 c (E ~ · n̂ dS = − E ε0 E 2 + dV
∂t 2 2 µ0
S V V

~ = ε0 E
but in polarizable materials where it is convenient to define D ~ + P~
~
~ = B −M ~:
and H µ0

~ ~
~ · ∂ B dV − E~ · ∂ D dV − E
I Z Z Z
~ ×H
E ~ · n̂ dS = − H ~ · J~ dV
∂t ∂t
S V V V

A theorem on the calculation of the mean Poynting vector from complex


fields:
¯~ 1 h
~c × H
~ c∗
i
N = Re E
2

The rate of doing work in magnetization


dW dB
=H
dt dt
and B = µ0 (H + M ).

Hemholtz Theorem: An arbitrary vector F~ can be written as:

~ 1 Z ∇ · F~ 1 Z
∇ × F~
F = − ∇ dV + ∇× dV
4π V r 4π V r

= F~l + F~t

Fields and potentials:

~
~ = −∇Φ − ∂ A ,
E ~ =∇×A
B ~
∂t

316
Differential equation for the vector potential:
2~
!
~ − 1 ∂ A = −µ0 J~ + ∇ ∇ · A
∇A2 ~ + 1 ∂Φ
2
c ∂t 2 c2 ∂t
In the Lorentz gauge:

∇·A ~ = − 1 ∂Φ
c2 ∂t
the differential equations for the electromagnetic potentials are:

1 ∂ 2Φ ρ
∇2 Φ − 2 2
=−
c ∂t ε0
2~
∇2 A~ − 1 ∂ A = −µ0 J~
c2 ∂t2
and these have solutions of the form:
1 Z ρ
Φ= dV
4πε0 r

~= 1 Z J~
A dV
4πε0 c2 r

The field of a Hertzian dipole:


I0 ∆` cos θ 2 2
 
Er = 3
+ 2 ei(ωt−kr)
4πε0 c ikr r
" #
I0 ∆` sin θ 1 1 ik i(ωt−kr)
Eθ = + + e
4πε0 c ikr3 r2 r
" #
I0 ∆` sin θ ik 1
Bφ = 2
+ 2 ei(ωt−kr)
4πε0 c r r
The mean energy flux from the Hertzian dipole:
#2
I 2 ∆` sin2 θ
"
N̄ = 0
8ε0 c λ r2

317
A series solution in 2 dimensions to Laplace’s equation in Cartesian coordi-
nates:
X
Φ(x, z) = [Ak sin(αx) + Bk cos(αx)][Ck sinh(αz) + Dk cosh(αz)]
k

A series solution in 3 dimensions to Laplace’s equation in spherical polar


coordinates:
( " #)
  X
` −(`+1)
blm sin(mφ)] P`m
X
Φ(r, θ, φ) = C1` r + C2` r [alm cos(mφ) + (cos θ)
` m

Useful properties of trigonometrical functions:

sin(α ± β) = sin α cos β ± sin β cos α


cos(α ± β) = cos α cos β ∓ sin α sin β
1
sin2 α = (1 − cos 2α)
2
1
cos2 α = (1 + cos 2α)
2
Z π
4
sin3 θ dθ =
0 3
Z 2π (
0 for m 6= n
sin(mφ) sin(nφ) dφ =
0 π for m = n
Z 2π (
0 for m 6= n
cos(mφ) cos(nφ) dφ =
0 π for m = n
Z 2π
sin(mφ) cos(nφ) dφ = 0 for all m and n
0

318
Properties of Legendry polynomials:
Z1
Plm (cos θ) Pkn (cos θ) d(cos θ) = 0 unless m = n and l = k
−1

Z1
2 (l + m)!
[Plm (cos θ)]2 d(cos θ) =
2l + 1 (l − m)!
−1

1
P0 = 1 , P10 = cos θ , P11 = sin θ , P20 = (3 cos(2θ) + 1) ,
4
3 3
P21 = sin(2θ) , P22 = (1 − cos(2θ)) ,
2 2
Pl (1) = 1 , for all l .

A theorem on the electrostatic potential due to a distribution of electric


dipoles of moment per unit volume P~ :

1 I P~ · n̂ 1 Z ∇ · P~
Φ= dS + − dV
4πε0 r 4πε0 r
S V

A theorem on the vector potential due to a distribution of magnetic dipoles


of moment per unit volume M~:

1 Z ∇×M ~ 1 I M ~ × n̂
~=
A dV + dS
4πε0 c2 r 4πε0 c2 r
V S

A dispersion equation:
 1
σ σ
     
0 0 00 00 0 00
0 00 2
k=ω εµ −µ ε + −i µ ε + +εµ
ω ω

319
The skin depth in a good conductor:
s
2
δ=
ωµσ

General boundary conditions:


~ is continuous across a boundary.
• The normal component of B
~ is continuous across a boundary.
• The normal component of D
~ is continuous across a boundary.
• The tangential component of E
~ is continuous across a boundary.
• The tangential component of H

Special boundary conditions at the surface between a dielectric and


a perfect conductor:
~ = 0.
• The tangential component of E
~ or H
• The normal component of B ~ = 0.

The Fresnel equations:

Case 1: E~ normal to the plane of incidence.


Reflection:
q
k1 µ2 cos θi − µ1 k22 − k12 sin2 θi
E1 = q E0
k1 µ2 cos θi + µ1 k22 − k12 sin2 θi
Transmission:
2k1 µ2 cos θi
E2 = q E0
k1 µ2 cos θi + µ1 k22 − k12 sin2 θi

Case 2: E~ in the plane of incidence.


Reflection : q
µ1 k2 cos θi − µ2 k1 k22 − k12 sin2 θi
2
H1 = q H0
µ1 k22 cos θi + µ2 k1 k22 − k12 sin2 θi

320
Transmission:
2µ1 k22 cos θi
H2 = q H0
µ1 k22 cos θi + µ2 k1 k22 − k12 sin2 θi

In dielectric media the Fresnel equations become:

~ normal to the plane of incidence:


Case 1: E

E1 sin(θi − θt ) E2 2 cos θi sin θt


= , =
E0 sin(θi + θt ) E0 sin(θi + θt )

~ in the plane of incidence:


Case 2: E

E1 tan(θi − θt ) E2 2 cos θi cos θt


= , =
E0 tan(θi + θt ) E0 sin(θi + θt ) cos(θi − θt )

Rectangular waveguides:

In the propagation direction z, the fields vary as e−γz .

~ satisfies:
For TE modes, the longitudinal component of H

∂ 2 Hz ∂ 2 Hz
2
+ 2
+ k 2 Hz = 0 ,
∂x ∂y
where
k 2 = γ 2 − iωµ(σ + iωε)
Then satisfying the boundary conditions (assuming the walls are perfect con-
ductors) requires:
2 2
mπ nπ
 
γ2 = + + iµω(σ + iεω)
a b
For the lossless waveguide:

321
Cut-off frequency for the mn mode:
s
2 2
m n
 
fmn = vm +
2a 2b
Phase velocity:
vm
vp = r
fc 2
 
1− f

Group velocity:
v
u !2
u fc
vg = vm t
1−
f

VECTOR FORMULAS

∇(Φ + Ψ) = ∇Φ + ∇Ψ
~ + B)
∇ · (A ~ = ∇·A
~+∇·B~
∇ × (A~ + B)
~ = ∇×A~+∇×B ~
∇(ΦΨ) = Φ∇Ψ + Ψ∇Φ
~ = A
∇ · (ΦA) ~ · ∇Φ + Φ∇ · A
~
~ × B)
∇ · (A ~ = B
~ · (∇ × A)
~ −A ~ · (∇ × B)
~
∇ × (ΦA)~ = ∇Φ × A ~ + Φ∇ × A~
∇ × (A~ × B)
~ = A∇
~ ·B ~ − B∇
~ ·A ~ + (B~ · ∇)A
~ − (A
~ · ∇)B
~
∇ · ∇Φ = ∇2 Φ
~ = 0
∇ · (∇ × A)
∇ × ∇Φ = 0
∇ × (∇ × A)~ = ∇(∇ · A) ~
~ − ∇2 A
~ · (B
A ~ × C)
~ = B
~ · (C
~ × A)
~ =C ~ · (A
~ × B)
~
~ × (B
A ~ × C)
~ = B(
~ A~ · C)
~ − C(
~ A~ · B)
~

322
FORMS OF VECTOR OPERATIONS IN CYLINDRICAL
COORDINATES

∂Φ φ̂ ∂Φ ∂Φ
∇Φ = ρ̂ + + ẑ
∂ρ ρ ∂φ ∂z

~ = 1 ∂(ρAρ ) + 1 ∂Aφ + ∂Az


∇·A
ρ ∂ρ ρ ∂φ ∂z

! ! !
~ = ρ̂ 1 ∂Az − ∂Aφ
∇×A + φ̂
∂Aρ ∂Az
− + ẑ
1 ∂(ρAφ ) ∂Aρ

ρ ∂φ ∂z ∂z ∂ρ ρ ∂ρ ∂φ

1 ∂ 2Φ ∂ 2Φ
!
2 1 ∂ ∂Φ
∇Φ = ρ + 2 2 + 2
ρ ∂ρ ∂ρ ρ ∂φ ∂z

VECTOR AND DIFFERENTIAL OPERATIONS IN


SPHERICAL COORDINATES

~ = (Ax sin θ cos φ + Ay sin θ sin φ + Az cos θ) r̂


A
+ (Ax cos θ cos φ + Ay cos θ sin φ − Az sin θ) θ̂
+ (−Ax sin φ + Ay cos φ) φ̂
= Ar r̂ + Aθ θ̂ + Aφ φ̂

∂Φ 1 ∂Φ 1 ∂Φ
∇Φ = r̂ + θ̂ + φ̂
∂r r ∂θ r sin θ ∂φ

323
~ = 1 ∂ (r2 Ar ) 1 ∂ (sin θAθ ) 1 ∂Aφ
∇·A + +
r2 ∂r r sin θ ∂θ r sin θ ∂φ

" # " #
~ = r̂ ∂ (sin θAφ ) ∂Aθ θ̂ 1 ∂Ar ∂ (rAφ )
∇×A − + −
r sin θ ∂θ ∂φ r sin θ ∂φ ∂r
" #
φ̂ ∂ (rAθ ) ∂Ar
+ −
r ∂r ∂θ

∂ 2Φ
! !
2 1 ∂ ∂Φ 1 ∂ ∂Φ 1
∇Φ = 2 r2 + 2 sin θ + 2 2
r ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂φ2

324

You might also like