You are on page 1of 106

SPIE PRESS

Maxwells Equations of Electrodynamics: An Explanation


is a concise discussion of Maxwell's four equations of
electrodynamicsthe fundamental theory of electricity,
magnetism, and light. It guides readers step-by-step through
the vector calculus and development of each equation.
Pictures and diagrams illustrate what the equations mean in
basic terms. The book not only provides a fundamental
description of our universe but also explains how these
equations predict the fact that light is better described as
"electromagnetic radiation."

P.O. Box 10
Bellingham, WA 98227-0010
ISBN: 9780819494528
SPIE Vol. No.: PM232

Bellingham, Washington USA

Library of Congress Cataloging-in-Publication Data


Ball, David W. (David Warren), 1962Maxwells equations of electrodynamics : an explanation / David W. Ball.
pages cm
Includes bibliographical references and index.
ISBN 978-0-8194-9452-8
1. Maxwell equations. 2. Electromagnetic theory. I. Title.
QC670.B27 2012
530.1401dc23
2012040779
Published by
SPIE
P.O. Box 10
Bellingham, Washington 98227-0010 USA
Phone: +1 360.676.3290
Fax: +1 360.647.1445
Email: Books@spie.org
Web: http://spie.org
c 2012 Society of Photo-Optical Instrumentation Engineers
Copyright
(SPIE)
All rights reserved. No part of this publication may be reproduced or
distributed in any form or by any means without written permission of
the publisher.
The content of this book reflects the work and thoughts of the author(s).
Every effort has been made to publish reliable and accurate information
herein, but the publisher is not responsible for the validity of the information or for any outcomes resulting from reliance thereon.
Printed in the United States of America.
First printing

Dedication
This book is dedicated to the following cadets whom I had the honor
of teaching while serving as Distinguished Visiting Faculty at the US
Air Force Academy in Colorado Springs, Colorado, during the 201112
academic year:
Jessica Abbott, Andrew Alderman, Bentley Alsup, Ryan Anderson,
Austin Barnes, Daniel Barringer, Anthony Bizzaco, Erin Bleyl,
Nicholas Boardman, Natasha Boozell, Matthew Bowersox, Patrick
Boyle, Andrew Burns, Spencer Cavanagh, Kyle Cousino, Erin
Crow, Michael Curran, Chad Demers, Nicholas Fitzgerald, Kyle
Gartrell, James Gehring, Nicholas Gibson, Ahmed Groce, Kassie
Gurnell, Deion Hardy, Trevor Haydel, Aaron Henrichs, Clayton
Higginson, Anthony Hopf, Christopher Hu, Vania Hudson,
Alexander Humphrey, Spencer Jacobson, Stephen Joiner, Fedor
Kalinkin, Matthew Kelly, Ye Kim, Lauren Linscott, Patrick Lobo,
Shaun Lovett, James Lydiard, Ryan Lynch, Aaron Macy, Dylan
Mason, Ryan Mavity, Payden McBee, Blake Morgan, Andrew
Munoz, Patrick Murphy, David Myers, Kathrina Orozco, Nathan
Orrill, Anthony Paglialonga, Adam Pearson, Emerald Peoples,
Esteban Perez, Charles Perkins, Hannah Peterson, Olivia Prosseda,
Victoria Rathbone, Anthony Rosati, Sofia Schmidt, Craig Stan,
James Stofel, Rachele Szall, Kevin Tanous, David Tyree, Joseph
Uhle, Tatsuki Watts, Nathanael Webb, Max Wilkinson, Kamryn
Williams, Samantha Wilson, Trevor Woodward, and Aaron Wurster.
May fortune favor them as they serve their country.

Contents
Preface .................................................................................................... ix
Chapter 1

1.1
1.2
1.3
Chapter 2

2.1
2.2
2.3
2.4
2.5
Chapter 3

3.1
3.2
3.3
Chapter 4

4.1
4.2
4.3
4.4

History.............................................................................

History (Ancient) ............................................................................. 2


History (More Recent).................................................................... 7
Faraday ................................................................................................ 11
First Equation of Electrodynamics ............................ 17

Enter Stage Left: Maxwell ............................................................


A Calculus Primer............................................................................
More Advanced Stuff ......................................................................
A Better Way .....................................................................................
Maxwells First Equation ..............................................................

17
19
26
35
40

Second Equation of Electrodynamics ....................... 47

Introduction ........................................................................................ 47
Faradays Lines of Force ............................................................... 48
Maxwells Second Equation ......................................................... 50
Third Equation of Electrodynamics ........................... 55

Beginnings ..........................................................................................
Work in an Electrostatic Field......................................................
Introducing the Curl ........................................................................
Faradays Law ...................................................................................
vii

55
59
67
70

viii

Contents

Chapter 5

5.1
5.2
5.3
Afterword

A.1
A.2
A.3

Fourth Equation of Electrodynamics ........................ 75

Ampres Law ................................................................................... 75


Maxwells Displacement Current ............................................... 78
Conclusion .......................................................................................... 82
Whence Light? ............................................................... 83

Recap: The Four Equations .......................................................... 83


Whence Light? .................................................................................. 84
Concluding Remarks....................................................................... 87

Bibliography .......................................................................................... 89
Index ....................................................................................................... 91

Preface
As the contributing editor of The Baseline column in Spectroscopy
magazine, I get a lot of leeway from my editor regarding the topics I cover
in the column. For example, I once did a column on clocks, only to end
with the fact that atomic clocks, currently our most accurate, are based on
spectroscopy. But most of my topics are more obviously related to the title
of the publication.
In late 2010 or so, I had an idea to do a column on Maxwells equations
of electrodynamics, since our understanding of light is based on them.
It did not take much research to realize that a discussion of Maxwells
equations was more than a 2000-word column could handleindeed,
whole books are written on them! (Insert premonitional music here.) What
I proceeded to do was write about them in seven sequential installments
over an almost two-year series of issues of the magazine. Ive seldom had
so much fun with, or learned so much from, one of my ideas for a column.
Not long into writing it (and after getting a better understanding of how
long the series would be), I thought that the columns might be collected
together, revised as needed, and published as a book. There is personal
precedent for this: In the early 2000s, SPIE Press published a collection of
my Spectroscopy columns in a book titled The Basics of Spectroscopy,
which is still in print. So I contacted then-acquisitions-editor at SPIE
Press, Tim Lamkins, with the idea of a book on Maxwells equations.
He responded in less than two hours. . . with a contract. (Note to budding
authors: thats a good sign.)
Writing took a little longer than expected, what with having to split
columns and a year-long professional sojourn to Colorado, but here it is. I
hope the readers enjoy it. If, by any chance, you can think of a better way
to explain Maxwells equations, let me knowthe hope is that this will be
one of the premiere explanations of Maxwells equations available.
Thanks to Tim Lamkins of SPIE Press for showing such faith in
my idea, and for all the help in the process; also to his colleagues,
ix

Preface

Dara Burrows and Scott McNeill, who did a great job of converting
manuscript to book. Thanks to the editor of Spectroscopy, Laura Bush,
for letting me venture on such a long series of columns on a single topic.
My gratitude goes to the College of Sciences and Health Professions,
Cleveland State University (CSU), for granting me a leave of absence so I
could spend a year at the US Air Force Academy, where much of the first
draft was written, as well as to the staff in the Department of Chemistry,
CSU, for helping me manage certain unrelinquishable tasks while I was
gone. Thanks to Bhimsen Shivamoggi and Yakov Soskind for reading over
the manuscript and making some useful suggestions; any remaining errors
are the responsibility of the author. Never-ending appreciation goes to my
familyGail, Stuart, and Caseyfor supporting me in my professional
activities. Finally, thanks to John Q. Buquoi of the US Air Force Academy,
Department of Chemistry, for his enthusiastic support and encouragement
throughout the writing process.
David W. Ball
Cleveland, Ohio
November 2012

Chapter 1

History
Maxwells equations for electromagnetism are the
fundamental understanding of how electric fields,
magnetic fields, and even light behave. There are various
versions, depending on whether there is vacuum, a charge
present, matter present, or the system is relativistic or
quantum, or is written in terms of differential or integral
calculus. Here, we will discuss a little bit of historical
development as a prelude to the introduction of the laws
themselves.

One of the pinnacles of classical science was the development of an


understanding of electricity and magnetism, which saw its culmination in
the announcement of a set of mathematical relationships by James Clerk
Maxwell in the early 1860s. The impact of these relationships would
be difficult to minimize if one were to try. They provided a theoretical
description of light as an electromagnetic wave, implying a wave medium
(the ether). This description inspired Michelson and Morley, whose failure
inspired Einstein, who was also inspired by Planck and who (among
others) ushered in quantum mechanics as a replacement for classical
mechanics. Again, its difficult to minimize the impact of these equations,
which in modern form are known as Maxwells equations.
In this chapter, we will discuss the history of electricity and magnetism,
going through early 19th-century discoveries that led to Maxwells work.
This historical review will set the stage for the actual presentation of
Maxwells equations.
1

Chapter 1

1.1 History (Ancient)


Electricity has been experienced since the dawn of humanity, and from
two rather disparate sources: an aquatic animal and lightning. Ancient
Greeks, Romans, Arabs, and Egyptians recorded the effect that electric
eels caused when touched, and who could ignore the fantastic light shows
that accompany certain storms (Fig. 1.1)? Unfortunately, although the

Figure 1.1 One of the first pictures of a lightning strike, taken by William N.
Jennings around 1882.

History

ancients could describe the events, they could not explain exactly what
was happening.
The ancient Greeks noticed that if a sample of amber (Greek elektron)
was rubbed with animal fur, it would attract small, light objects such as
feathers. Some, like Thales of Miletos, related this property to that of
lodestone, a natural magnetic material that attracted small pieces of iron.
(A relationship between electricity and magnetism would be re-awakened
over 2000 years later.) Since the samples of this rock came from the nearby
town of Magnesia (currently located in southern Thessaly, in central
Greece), these stones ultimately were called magnets. Note that it appears
that magnetism was also recognized by the same time as static electricity
effects (Thales lived in the 7th and 6th century BCE [before common era]),
so both phenomena were known, just not understood. Chinese fortune
tellers were also utilizing lodestones as early as 100 BCE. For the most
part, however, the properties of rubbed amber and this particular rock
remained novelties.
By 1100 CE (common era), the Chinese were using spoon-shaped
pieces of lodestone as rudimentary compasses, a practice that spread
quickly to the Arabs and to Europe. In 1269, Frenchman Petrus Peregrinus
used the word pole for the first time to describe the ends of a lodestone
that point in particular directions. Christopher Columbus apparently had a
simple form of compass in his voyages to the New World, as did Vasco de
Gama and Magellan, so Europeans had compasses by the 1500sindeed,
the great ocean voyages by the explorers would likely have been extremely
difficult without a working compass (Fig. 1.2). In books published in 1550
and 1557, the Italian mathematician Gerolamo Cardano argued that the
attractions caused by lodestones and the attractions of small objects to
rubbed amber were caused by different phenomena, possibly the first time
this was explicitly pointed out. (His proposed mechanisms, however, were
typical of much 16th-century physical science: wrong.)
In 1600, English physician and natural philosopher William Gilbert
published De Magnete, Magneticisque Corporibus, et de Magno Magnete
Tellure (On the Magnet and Magnetic Bodies, and on the Great Magnet
the Earth) in which he became the first to make the distinction between
attraction due to magnetism and attraction due to rubbed elektron. For
example, Gilbert argued that elektron lost its attracting ability with heat,
but lodestone did not. (Here is another example of a person being correct,
but for the wrong reason.) He proposed that rubbing removed a substance
he termed effluvium, and it was the return of the lost effluvium to the

Chapter 1

Figure 1.2 Compasses like the one drawn here were used by 16th century
maritime explorers to sail around the world. These compasses were most likely
the first applied use of magnetism.

object that caused attraction. Gilbert also introduced the Latinized word
electricus as meaning like amber in its ability to attract. Note also what
the title of Gilberts book impliesthe Earth is a great magnet, an idea
that only arguably originated with him, as the book was poorly referenced.
The word electricity itself was first used in 1646 in a book by English
author Thomas Browne.
In the late 1600s and early 1700s, Englishman Stephen Gray carried out
some experiments with electricity and was among the first to distinguish
what we now call conductors and insulators. In the course of his
experiments, Gray was apparently the first to logically suggest that the
sparks he was generating were the same thing as lightning, but on a much
smaller scale. Interestingly, announcements of Grays discoveries were
stymied by none other than Isaac Newton, who was having a dispute with
other scientists (with whom Gray was associated, so Gray was guilty by
association) and, in his position of president of the Royal Society, impeded
these scientists abilities to publish their work. (Curiously, despite his

History

advances in other areas, Newton himself did not make any major and
lasting contributions to the understanding of electricity or magnetism.
Grays work was found documented in letters sent to the Royal Society
after Newton had died, so his work has been historically documented.)
Inspired by Grays experiments, French chemist Charles du Fay also
experimented and found that objects other than amber could be altered.
However, du Fay noted that in some cases, some altered objects attract,
while other altered objects repel. He concluded that there were two types
of effluvium, which he termed vitreous (because rubbed glass exhibited
one behavior) and resinous (because rubbed amber exhibited the other
behavior). This was the first inkling that electricity comes as two kinds.
French mathematician and scientist Ren Descartes weighed in with
his Principia Philosophiae, published in 1644. His ideas were strictly
mechanical: magnetic effects were caused by the passage of tiny
particles emanating from the magnetic material and passing through
the luminiferous ether. Unfortunately, Descartes proposed a variety of
mechanisms for how magnets worked, rather than a single one, which
made his ideas arguable and inconsistent.
The Leyden jar (Fig. 1.3) was invented simultaneously in 1745 by
Ewald von Kleist and Pieter van Musschenbrk. Its name derives from
the University of Leyden (also spelled Leiden) in the Netherlands, where
Musschenbrk worked. It was the first modern example of a condenser or
capacitor. The Leyden jar allowed for a significant (to that date) amount of
charge to be stored for long periods of time, letting researchers experiment
with electricity and see its effects more clearly. Advances were quick,
given the ready source of what we now know as static electricity. Among
other discoveries was the demonstration by Alessandro Volta that materials
could be electrified by induction in addition to direct contact. This, among
other experiments, gave rise to the idea that electricity was some sort of
fluid that could pass from one object to another, in a similar way that
light or heat (caloric in those days) could transfer from one object to
another. The fact that many conductors of heat were also good conductors
of electricity seemed to support the notion that both effects were caused by
a fluid of some sort. Charles du Fays idea of vitreous and resinous
substances, mentioned above, echoed the ideas that were current at the
time.
Enter Benjamin Franklin, polymath. It would take (and has taken) whole
books to chronicle Franklins contributions to a variety of topics, but here
we will focus on his work on electricity. Based on his own experiments,

Chapter 1

Figure 1.3 Example of an early Leyden jar, which was composed of a glass jar
coated with metal foil on the outside (labeled A) and the inside (B).

History

Franklin rejected du Fays two fluid idea of electricity and instead


proposed a one fluid idea, arbitrarily defining one object negative if it
lost that fluid and positive if it gained fluid. (We now recognize his idea
as prescient but backwards.) As part of this idea, Franklin enunciated an
early concept of the law of conservation of charge. In noting that pointed
conductors lose electricity faster than blunt ones, Franklin invented the
lightning rod in 1749. Although metal had been used in the past to adorn
the tops of buildings, Franklins invention appears to the first time such a
construction was used explicitly to draw a lightning strike and spare the
building itself. [Curiously, there was some ecclesiastical resistance to the
lightning rod, as lightning was deemed by many to be a sign of divine
intervention, and the use of a lightning rod was considered by some to be
an attempt to control a deity! The danger of such attitudes was illustrated
in 1769 when the rodless Church of San Nazaro in Brescia, Italy (in
Lombardy in the northern part of the country) was struck by lightning,
igniting approximately 100 tons of gunpowder stored in the church. The
resulting explosion was said to have killed 3000 people and destroyed
one-sixth of the city. Ecclesiastical opposition weakened significantly after
that.]
In 1752, Franklin performed his apocryphal experiment with a kite, a
key, and a thunderstorm. The word apocryphal is used intentionally:
different sources have disputed accounts of what really happened. (Modern
analyses suggest that performing the elementary-school portrayal of the
experiment would definitely have killed Franklin.) What is generally
not in dispute, though, is the result: lightning was just electricity, not
some divine sign. As mundane as it seems to us today, Franklins
demonstration that lightning was a natural phenomenon was likely as
important as Copernicus heliocentric ideas were in demonstrating the lack
of privileged, heaven-mandated position mankind has in the universe.

1.2 History (More Recent)


Up to this point, most demonstrations of electric and magnetic effects were
qualitative, not quantitative. This began to change in 1779 when French
physicist Charles-Augustin de Coulomb published Thorie des Machines
Simples (Theory of Simple Machines). To refine this work, through the
early 1780s Coulomb constructed a very fine torsion balance with which
he could measure the forces due to electrical charges. Ultimately, Coulomb
proposed that electricity behaves as if it were tiny particles of matter, acting

Chapter 1

in a way similar to Newtons law of gravity, which had been known since
1687. We now know this idea as Coulombs law. If the charge on one body
is q1 , and the charge on another body is q2 , and the bodies are a distance r
apart, then the force F between the bodies is given by the equation
F=k

q1 q2
,
r

(1.1)

where k is a constant. In 1785 Coulomb also proposed a similar expression


for magnetism but, curiously, did not relate the two phenomena. As
seminal and correct as this advance was, Coulomb still adhered to
a two fluid idea of electricityand magnetismas it turned out.
(Interestingly, about 100 years later, James Clerk Maxwell, about whom
we will have more to say, published papers from reclusive scientist Henry
Cavendish, demonstrating that Cavendish had performed similar electrical
experiments at about the same time but never published them. If he had,
we might be referring to this equation as Cavendishs law.) Nonetheless,
through the 1780s, Coulombs work established all of the rules of static
(that is, nonmoving) electrical charge.
The next advances had to do with moving charges, but it was a while
before people realized that movement was involved. Luigi Galvani was an
Italian doctor who studied animal anatomy at the University of Bologna. In
1771, while dissecting a frog as part of an experiment in static electricity,
he noticed a small spark between the scalpel and the frog leg he was cutting
into, and the immediate jerking of the leg itself. Over the next few decades,
Galvani performed a host of experiments trying to pin down the conditions
and reasons for this muscle action. The use of electricity from a generating
electric machine or a Leyden jar caused a frog leg to jerk, as did contact
with two different metals at once, like iron and copper. Galvani eventually
proposed that the muscles in the leg were acting like a fleshy Leyden jar
that contained some residual electrical fluid he called animal electricity. In
the 1790s, he printed pamphlets and passed them around to his colleagues,
describing his experiments and his interpretation.
One of the recipients of Galvanis pamphlets was Alessandro Volta, a
professor at the University of Pavia, about 200 km northwest of Bologna.
Volta was skeptical of Galvanis work but successfully repeated the frogleg experiments. Volta took the work further, positioning metallic probes
exclusively on nerve tissue and bypassing the muscle. He made the
crucial observation that some sort of induced activity was only produced
if two different metals were in contact with the animal tissue. From

History

these observations, Volta proposed that the metals were the source of the
electrical action, not the animal tissue itself; the muscle was serving as the
detector of electrical effects, not the source.
In 1800, Volta produced a stack of two different metals soaked in
brine that supplied a steady flow of electricity (Fig. 1.4). The construction
became known as a voltaic pile, but we know it better as a battery. Instead
of a sudden spark, which was how electricity was produced in the past
from Leyden jars, here was a construction that provided a steady flow,
or current, of electricity. The development of the voltaic pile made new
experiments in electricity possible.
Advances came swiftly with this new, easily constructed pile. Water
was electrolyzed, not for the first time but for the first time systematically,
by Nicholson and Carlisle in England. English chemist Humphrey Davy,
already a well-known chemist for his discoveries, proposed that if
electricity were truly caused by chemical reactions (as many chemists at
the time had perhaps chauvinistically thought), then perhaps electricity
can cause chemical reactions in return. He was correct, and in 1807 he
produced elemental potassium and sodium electrochemically for the first
time. The electrical nature of chemistry was first realized then, and its
ramifications continue today. Davy also isolated elemental chlorine in
1810, following up with isolation of the elements magnesium, calcium,
strontium, and barium, all for the first time and all by electricity. (Davys
greatest discovery, though, was not an element; arguably, Davys greatest
discovery was Michael Faraday. But more on Faraday later.)
Much of the historical development so far has focused on electrical
phenomena. Where has magnetism been? Actually, its been here all
along, but not much new has been developed. This changed in 1820.
Hans Christian rsted was a Danish physicist who was a disciple of
the metaphysics of Emmanual Kant. Given the recently demonstrated
connection between electricity and chemistry, rsted was certain that there
were other fundamental physical connections in nature.
In the course of a lecture in April 1820, rsted passed an electrical
current from a voltaic pile through a wire that was placed parallel to a
compass needle. The needle was deflected. Later experiments (especially
by Ampre) demonstrated that with a strong enough current, a magnetized
compass needle orients itself perpendicular to the direction of the current.
If the current is reversed, the needle points in the opposite direction.
Here was the first definitive demonstration that electricity and magnetism
affected each other: electromagnetism.

10

Chapter 1

Figure 1.4 Diagram of Voltas first pile that produced a steady supply of electricity.
We know it now as a battery.

History

11

Detailed follow-up studies by Ampre in the 1820s quantified the


relationship a bit. Ampre demonstrated that the magnetic effect was
circular about a current-carrying wire and, more importantly, that two
wires with current moving in them attracted and repelled each other as
if they were magnets. Magnets, Ampre argued, are nothing more than
moving electricity. The development of electromagnets at this time freed
scientists from finding rare lodestones and allowed them to study magnetic
effects using only coiled wires as the source of magnetism. Around this
time, Biot and Savarts studies led to the law named after them, relating
the intensity of the magnetic effect to the square of the distance from the
electromagnet: a parallel to Coulombs law.
In 1827, Georg Ohm announced what became known as Ohms law,
a relationship between the voltage of a pile and the current passing
through the system. As simple as Ohms law is (V = IR, in modern
usage), it was hotly contested in his native Germany, probably for political
or philosophical reasons. Ohms law became a cornerstone of electrical
circuitry. Its uncertain how the term current became applied to the
movement of electricity. Doubtless, the term arose during the time that
electricity was thought to be a fluid, but it has been difficult to find the
initial usage of the term. In any case, the term was well established by the
time Sir William Robert Grove wrote The Correlation of Physical Forces
in 1846, which was a summary of what was known about the phenomenon
of electricity to date.

1.3 Faraday
Probably no single person set the stage more for a mathematical treatment
of electricity and magnetism than Michael Faraday (Fig. 1.5). Inspired by
rsteds work, in 1821 Faraday constructed a simple device that allowed
a wire with a current running through it to turn around a permanent
magnet, and conversely a permanent magnet to turn around a wire that
had a current running through it. Faraday had, in fact, constructed the first
rudimentary motor. What Faradays motor demonstrated experimentally is
that the forces of interaction between the current-containing wire and the
magnet were circular in nature, not radial like gravity.
By now it was clear that electrical current could generate magnetism.
But what about the other way around: could magnetism generate
electricity? An initial attempt to accomplish this was performed in 1824
by French scientist Franois Arago, who used a spinning copper disk to

12

Chapter 1

Figure 1.5 Michael Faraday was not well versed in mathematics, but he was a
first-rate experimentalist who made major advances in physics and chemistry.

deflect a magnetized needle. In 1825, Faraday tried to repeat Aragos work,


using an iron ring instead of a metal disk and wrapping coils of wire on
either side, one coil attached to a battery and one to a galvanometer. In
doing so, Faraday constructed the first transformer, but the galvanometer
did not register the presence of a current. However, Faraday did note
a slight jump of the galvanometers needle. In 1831, Faraday passed
a magnet through a coil of wire attached to a galvanometer, and the
galvanometer registered, but only if the magnet was moving. When the

History

13

magnet was halted, even when it was inside the coil, the galvanometer
registered zero.
Faraday understood that it wasnt the presence of a magnetic field that
caused an electrical current, it was a change in the magnetic field that
caused a current. It did not matter what moved; a moving magnet can
induce current in a stationary wire, or a moving wire can pick up a current
induced by moving it across a stationary magnet. Faraday also realized that
the wire wasnt necessary. Taking a page from Arago, Faraday constructed
a generator of electricity using a copper disk that rotated through the
poles of a permanent magnet (Fig. 1.6). Faraday used this to generate a
continuous source of electricity, essentially converting mechanical motion
to electrical motion. Faraday invented the dynamo, the basis of the electric
industry even today.
After dealing with some health issues during the 1830s, in the mid1840s Faraday was back at work. Twenty years earlier, Faraday had
investigated the effect of magnetic fields on light but got nowhere.
Now, with stronger electromagnets available, Faraday went back to that
investigation on the advice of William Thomson, Lord Kelvin. This time,
with better equipment, Faraday noticed that the plane of plane-polarized
light rotated when a magnetic field was applied to a piece of flint glass with
the light passing through it. The magnetic field had to be oriented along the
direction of the lights propagation. This effect, known now as the Faraday
effect or Faraday rotation, convinced Faraday of several things. First, light

Figure 1.6 Diagram of Faradays original dynamo, which generated electricity


from magnetism.

14

Chapter 1

and magnetism are related. Second, magnetic effects are universal and not
just confined to permanent or electromagnets; after all, light is associated
with all matter, so why not magnetism? In late 1845, Faraday coined
the term diamagnetism to describe the behavior of materials that are
not attracted and are actually slightly repelled by a magnetic field. (The
amount of repulsion toward a magnetic field is typically significantly less
than the magnetic attraction by materials, now known as paramagnetism.
This is one reason that magnetic repulsion was not widely recognized until
then.)
Finally, this finding reinforced in Faraday the concept of magnetic fields
and their importance. In physics, a field is nothing more than a physical
property whose value depends on its position in three-dimensional space.
The temperature of a sample of matter, for example, is a field, as is the
pressure of a gas. Temperature and pressure are examples of scalar fields,
fields that have magnitude but no direction; vector fields, like magnetic
fields, have magnitude and direction. This was first demonstrated in the
13th century, when Petrus Peregrinus used small needles to map out the
lines of force around a magnet and was able to place the needles in curved
lines that terminated in the ends of the magnet. That was how he devised
the concept of pole that was mentioned above. (See Fig. 1.7 for an

Figure 1.7 An early photo of iron filings on a piece of paper that is placed over
a bar magnet, showing the lines of force that make up the magnetic field. The
magnetic field itself is not visible; the iron filings are simply lining up along the lines
of force that define the field.

History

15

example of such a mapping.) An attempt to describe a magnetic field in


terms of magnetic charges analogous to electrical charges was made by
the French mathematician/physicist Simeon-Denis Poisson in 1824, and
while successful in some aspects, it was based on the invalid concept of
such magnetic charges. However, development of the BiotSavart law
and some work by Ampre demonstrated that the concept of a field was
useful in describing magnetic effects. Faradays work on magnetism in
the 1840s reinforced his idea that the magnetic field was the important
quantity, not the materialwhether natural or electronicthat produced
the field of forces that could act on other objects and produce electricity.
As crucial as Faradays advances were for understanding electricity
and magnetism, Faraday himself was not a mathematical man. He had
a relatively poor grasp of the theory that modern science requires to
explain observed phenomena. Thus, while he made crucial experimental
advancements in the understanding of electricity and magnetism, he made
little contribution to the theoretical understanding of these phenomena.
That chore awaited others.
This was the state of electricity and magnetism in the mid-1850s.
What remained to emerge was the presence of someone who did have the
right knowledge to be able to express these experimental observations as
mathematical models.

Chapter 2

First Equation of
Electrodynamics E =

In this chapter, we introduce the first equation of


electrodynamics. A note to the reader: This is going to
get a bit mathematical. It cant be helped. Models of the
physical universe, like Newtons second law F = ma, are
based in math. So are Maxwells equations.

2.1 Enter Stage Left: Maxwell


James Clerk Maxwell (Fig. 2.1) was born in 1831 in Edinburgh, Scotland.
His unusual middle name derives from his uncle, who was the 6th Baronet
Clerk of Penicuik (pronounced penny-cook), a town not far from
Edinburgh. Clerk was, in fact, the original family name; Maxwells father,
John Clerk, adopted the surname Maxwell after receiving a substantial
inheritance from a family named Maxwell. By most accounts, James Clerk
Maxwell (hereafter referred to as simply Maxwell) was an intelligent but
relatively unaccomplished student.
However, Maxwell began blossoming intellectually in his early teens,
becoming interested in mathematics (especially geometry). He eventually
attended the University of Edinburgh and, later, Cambridge University,
where he graduated in 1854 with a degree in mathematics. He stayed
on for a few years as a Fellow, then moved to Marischal College in
Aberdeen. When Marischal merged with another college to form the
University of Aberdeen in 1860, Maxwell was laid off (an action for which
the University of Aberdeen should still be kicking themselves, but who
can foretell the future?), and he found another position at Kings College
London (later the University of London). He returned to Scotland in 1865,
17

18

Chapter 2

Figure 2.1

James Clerk Maxwell as a young man and an older man.

only to go back to Cambridge in 1871 as the first Cavendish Professor of


Physics. He died of abdominal cancer in November 1879 at the relatively
young age of 48; curiously, his mother died of the same ailment and at the
same age, in 1839.
Though he had a relatively short career, Maxwell was very productive.
He made contributions to color theory and optics (indeed, the first photo in
Fig. 2.1 shows Maxwell holding a color wheel of his own invention), and
actually produced the first true color photograph as a composite of three
images. The MaxwellBoltzmann distribution is named partially after him,
as he made major contributions to the development of the kinetic molecular
theory of gases. He also made major contributions to thermodynamics,
deriving the relations that are named after him and devising a thought
experiment about entropy that was eventually called Maxwells demon.
He demonstrated mathematically that the rings of Saturn could not be
solid, but must instead be composed of relatively tiny (relative to Saturn, of
course) particlesa hypothesis that was supported spectroscopically in the
late 1800s but finally directly observed for the first time when the Pioneer
11 and Voyager 1 spacecrafts passed through the Saturnian system in the
early 1980s (Fig. 2.2).
Maxwell also made seminal contributions to the understanding
of electricity and magnetism, concisely summarizing their behaviors
with four mathematical expressions known as Maxwells equations of

First Equation of Electrodynamics

19

Figure 2.2 Maxwell proved mathematically that the rings of Saturn couldnt be
solid objects but were likely an agglomeration of smaller bodies. This image of a
back-lit Saturn is a composite of several images taken by the Cassini spacecraft in
2006. Depending on the reproduction, you may be able to make out a tiny dot in
the 10 oclock position just inside the second outermost diffuse ringthats Earth.
If you cant see it, look for high-resolution pictures of pale blue dot on the Internet.

electromagnetism. He was strongly influenced by Faradays experimental


work, believing that any theoretical description of a phenomenon must be
grounded in experimental observations. Maxwells equations essentially
summarize everything about classical electrodynamics, magnetism, and
optics, and were only supplanted when relativity and quantum mechanics
revised our understanding of the natural universe at certain of its limits.
Far away from those limits, in the realm of classical physics, Maxwells
equations still rule just as Newtons equations of motion rule under normal
conditions.

2.2 A Calculus Primer


Maxwells laws are written in the language of calculus. Before we move
forward with an explicit discussion of the first law, here we deviate to a
review of calculus and its symbols.
Calculus is the mathematical study of change. Its modern form was
developed independently by Isaac Newton and German mathematician
Gottfried Leibnitz in the late 1600s. Although Newtons version was
used heavily in his influential Principia Mathematica (in which Newton
used calculus to express a number of fundamental laws of nature), it is

20

Chapter 2

Leibnitzs notations that are commonly used today. An understanding of


calculus is fundamental to most scientific and engineering disciplines.
Consider a car moving at constant velocity. Its distance from an initial
point (arbitrarily set as a position of 0) can be plotted as a graph of distance
from zero versus time elapsed. Commonly, the elapsed time is called the
independent variable and is plotted on the x axis of a graph (called the
abscissa), while distance traveled from the initial position is plotted on the
y axis of the graph (called the ordinate). Such a graph is plotted in Fig. 2.3.
The slope of the line is a measure of how much the ordinate changes as the
abscissa changes; that is, slope m is defined as
m=

y
.
x

(2.1)

For the straight line shown in Fig. 2.3, the slope is constant, so m has
a single value for the entire plot. This concept gives rise to the general
formula for any straight line in two dimensions, which is
y = mx + b,

(2.2)

where y is the value of the ordinate, x is the value of the abscissa, m is the
slope, and b is the y intercept, which is where the plot would intersect with
the y axis. Figure 2.3 shows a plot that has a positive value of m. In a plot
with a negative value of m, the slope would be going down, not up, moving

Figure 2.3 A plot of a straight line, which has a constant slope m, given by
y/x.

First Equation of Electrodynamics

21

from left to right. A horizontal line has a value of 0 for m; a vertical line
has a slope of infinity.
Many lines are not straight. Rather, they are curves. Figure 2.4 gives an
example of a plot that is curved. The slope of a curved line is more difficult
to define than that of a straight line because the slope is changing. That is,
the value of the slope depends on the point (x, y) where you are on the
curve. The slope of a curve is the same as the slope of the straight line that
is tangent to the curve at that point (x, y). Figure 2.4 shows the slopes at
two different points. Because the slopes of the straight lines tangent to the
curve at different points are different, the slopes of the curve itself at those
two points are different.
Calculus provides ways of determining the slope of a curve in any
number of dimensions [Figure 2.4 is a two-dimensional plot, but we
recognize that functions can be functions of more than one variable, so
plots can have more dimensions (a.k.a. variables) than two]. We have
already seen that the slope of a curve varies with position. That means
that the slope of a curve is not a constant; rather, it is a function itself.
We are not concerned about the methods of determining the functions for
the slopes of curves here; that information can be found in a calculus text.
Here, we are concerned with how they are represented.
The word that calculus uses for the slope of a function is derivative. The
derivative of a straight line is simply m, its constant slope. Recall that we
mathematically defined the slope m above using symbols, where is the
Greek capital letter delta. is used generally to represent change, as in T

Figure 2.4 A plot of a curve showing (with the thinner lines) the different slopes
at two different points. Calculus helps us determine the slopes of curved lines.

22

Chapter 2

(change in temperature) or y (change in y coordinate). For straight lines


and other simple changes, the change is definite; in other words, it has a
specific value.
In a curve, the change y is different for any given x because the
slope of the curve is constantly changing. Thus, it is not proper to refer
to a definite change because (to overuse a word) the definite change
changes during the course of the change. What we need here is a thought
experiment: Imagine that the change is infinitesimally small over both
the x and y coordinates. This way, the actual change is confined to an
infinitesimally small portion of the curve: a point, not a distance. The
point involved is the point at which the straight line is tangent to the curve
(Fig. 2.4).
Rather than using to represent an infinitesimal change, calculus starts
by using d. Rather than using m to represent the slope, calculus puts a
prime on the dependent variable as a way to represent a slope (which,
remember, is a function and not a constant). Thus, the definition of the
slope y0 of a curve is
y0 =

dy
.
dx

(2.3)

We hinted earlier that functions may depend on more than one variable.
If that is the case, how do we define the slope? First, we define a partial
derivative as the derivative of a multivariable function with respect to only
one of its variables. We assume that the other variables are held constant.
Instead of using d to indicate a partial derivative, we use a symbol based
on the lowercase Greek delta known as Jacobis delta. It is also common
to explicitly list the variables being held constant as subscripts to the
derivative, although this can be omitted because it is understood that a
partial derivative is a one-dimensional derivative. Thus, we have
f x0

f
=
x

y,z,...

f
,
x

(2.4)

spoken as the partial derivative of the function f (x, y, z, . . .) with respect to


x. Graphically, this corresponds to the slope of the multivariable function
f in the x dimension, as shown in Fig. 2.5.
The total differential of a function, df, is the sum of the partial
derivatives in each dimension; that is, with respect to each variable

First Equation of Electrodynamics

23

Figure 2.5 For a function of several variables, a partial derivative is a derivative


in only one variable. The line represents the slope in the x direction.

individually. For a function of three variables, f (x, y, z), the total


differential is written as
df =

f
f
f
dx +
dy +
dz,
x
y
z

(2.5)

where each partial derivative is the slope with respect to each individual
variable, and dx, dy, and dz are the finite changes in the x, y, and z
directions. The total differential has as many terms as the overall function
has variables. If a function is based in three-dimensional space, as is
commonly the case for physical observables, then there are three variables
and so three terms in the total differential.
When a function typically generates a single numerical value that is
dependent on all of its variables, it is called a scalar function. An example
of a scalar function might be
F(x, y) = 2x y2 .

(2.6)

According to this definition, F(4, 2) = 2 4 22 = 8 4 = 4. The final


value of F(x, y), 4, is a scalar: it has magnitude but no direction.
A vector function is a function that determines a vector, which
is a quantity that has magnitude and direction. Vector functions can
be easily expressed using unit vectors, which are vectors of length 1
along each dimension of the space involved. It is customary to use the

24

Chapter 2

representations i, j, and k to represent the unit vectors in the x, y, and z


dimensions, respectively (Fig. 2.6). Vectors are typically represented in
print as boldfaced letters. Any random vector can be expressed as, or
decomposed into, a certain number of i vectors, j vectors, and k vectors
as is demonstrated in Fig. 2.6. A vector function in two dimensions might
be as simple as
F = xi + yj,

(2.7)

as illustrated in Fig. 2.7 for a few discrete points. Although only a few
discrete points are shown in the figure, understand that the vector function
is continuous. That is, it has a value at every point in the graph.
One of the functions of a vector that we will evaluate is called a dot
product. The dot product between two vectors a and b is represented and

Figure 2.6 The definition of the unit vectors i, j, and k, and an example of how
any vector can be expressed in terms of the number of each unit vector.

First Equation of Electrodynamics

25

Figure 2.7 An example of a vector function F = xi + yj. Each point in two


dimensions defines a vector. Although only 12 individual values are illustrated here,
in reality, this vector function is a continuous, smooth function on both dimensions.

defined as
a b = |a||b| cos ,

(2.8)

where |a| represents the magnitude (that is, length) of a, |b| is the magnitude
of b, and cos is the cosine of the angle between the two vectors. The dot
product is sometimes called the scalar product because the value is a scalar,
not a vector. The dot product can be thought of physically as how much one
vector contributes to the direction of the other vector, as shown in Fig. 2.8.
A fundamental definition that uses the dot product is that for work w, which
is defined in terms of the force vector F and the displacement vector of a
moving object s, and the angle between these two vectors:
w = F s = |F||s| cos .

(2.9)

Thus, if the two vectors are parallel ( = 0 deg, so cos = 1) the


work is maximized, but if the two vectors are perpendicular to each other

26

Chapter 2

Figure 2.8 Graphical representation of the dot product of two vectors. The dot
product gives the amount of one vector that contributes to the other vector. An
equivalent graphical representation would have the b vector projected into the a
vector. In both cases, the overall scalar results are the same.

( = 90 deg, so cos = 0), the object does not move, and no work is done
(Fig. 2.9).

2.3 More Advanced Stuff


Besides taking the derivative of a function, the other fundamental operation
in calculus is integration, whose representation is called an integral, which
is represented as
Z

f (x) dx,

(2.10)

R
where the symbol
is called the integral sign and represents the
integration operation, f (x) is called the integrand and is the function to
be integrated, dx is the infinitesimal of the dimension of the function, and
a and b are the limits between which the integral is numerically evaluated,
if it is to be numerically evaluated. [If the integral sign looks like an
elongated s, it should be numerically evaluated. Leibniz, one of the
cofounders of calculus (with Newton), adopted it in 1675 to represent sum,
since an integral is a limit of a sum.] A statement called the fundamental

First Equation of Electrodynamics

27

Figure 2.9 Work is defined as a dot product of a force vector and a displacement
vector. (a) If the two vectors are parallel, they reinforce, and work is performed. (b)
If the two vectors are perpendicular, no work is performed.

theorem of calculus establishes that integration and differentiation are the


opposites of each other, a concept that allows us to calculate the numerical
value of an integral. For details of the fundamental theorem of calculus,
consult a calculus text. For our purposes, all we need to know is that the
two are related and calculable.
The most simple geometric representation of an integral is that it
represents the area under the curve given by f (x) between the limits a
and b and bound by the x axis. Look, for example, at Fig. 2.10(a). It is a
graph of the line y = x or, in more general terms, f (x) = x. What is the
area under this function but above the x axis, shaded gray in Fig. 2.10(a)?
Simple geometry indicates that the area is 1/2 units; the box defined by
x = 1 and y = 1 is 1 unit (1 1), and the right triangle that is shaded
gray is one-half of that total area, or 1/2 unit in area. Integration of the
function f (x) = x gives us the same answer. The rules of integration will
not be discussed here; it is assumed that the reader can perform simple
integration:
Z

f (x) dx =


1
1 2 1 1 2 1 2 1
x dx = x = (1) (0) = 0 = . (2.11)
2 0 2
2
2
2

28

Chapter 2

Figure 2.10 The geometric interpretation of a simple integral is the area under
a function and bounded on the bottom by the x axis (that is, y = 0). (a) For
the function f (x) = x, the areas as calculated by geometry and integration are
equal. (b) For the function f (x) = x2 , the approximation from geometry is not a
good value for the area under the function. A series of rectangles can be used
to approximate the area under the curve, but in the limit of an infinite number of
infinitesimally narrow rectangles, the area is equal to the integral.

It is a bit messier if the function is more complicated. But, as first


demonstrated by Reimann in the 1850s, the area can be calculated
geometrically for any function in one variable (most easy to visualize, but
in theory this can be extended to any number of dimensions) by using
rectangles of progressively narrower widths, until the area becomes a
limiting value as the number of rectangles goes to infinity and the width
of each rectangle becomes infinitely narrow. This is one reason a good
calculus course begins with a study of infinite sums and limits! But, I
digress. For the function in Fig. 2.10(b), which is f (x) = x2 , the area under
the curve, now poorly approximated by the shaded triangle, is calculated
exactly with an integral:
Z

f (x) dx =


1
1 3 1 1
x dx = x = 0 = .
3 0 3
3
2

(2.12)

As with differentiation, integration can also be extended to functions of


more than one variable. The issue to understand is that when visualizing
functions, the space you need to use has one more dimension than variables
because the function needs to be plotted in its own dimension. Thus, a plot
of a one-variable function requires two dimensions, one to represent the
variable and one to represent the value of the function. Figures 2.3 and
2.4, thus, are two-dimensional plots. A two-variable function needs to be

First Equation of Electrodynamics

29

Figure 2.11 A multivariable function f (x, y) with a line paralleling the y axis. The
equation of the line is represented by P.

plotted or visualized in three dimensions, like Figs. 2.5 or 2.6. Looking at


the two-variable function in Fig. 2.11, we see a line across the functions
values, with its projection in the (x, y) plane. The line on the surface is
parallel to the y axis, so it is showing the trend of the function only as the
variable x changes. If we were to integrate this multivariable function with
respect only to (in this case) x, we would be evaluating the integral only
along this line. Such an integral is called a line integral (also called a path
integral). One interpretation of this integral would be that it is simply the
part of the volume under the overall surface that is beneath the given
line; that is, it is the area under the line.
If the surface represented in Fig. 2.11 represents a field (either scalar
or vector), then the line integral represents the total effect of that field
along the given line. The formula for calculating the total effect might
be unusual, but it makes sense if we start from the beginning. Consider a
path whose position is defined by an equation P, which is a function of one
or more variables. What is the distance of the path? One way of calculating
the distance s is velocity v times time t, or
s = v t.

(2.13)

But velocity is the derivative of position P with respect to time, or dP/dt.


Let us represent this derivative as P0 . Our equation now becomes
s = P0 t.

(2.14)

30

Chapter 2

This is for finite values of distance and time, and for that matter, for
constant P0 . (Example: total distance at 2.0 m/s for 4.0 s = 2.0 m/s 4.0
s = 8.0 m. In this example, P0 is 2.0 m/s and t is 4.0 s.) For infinitesimal
values of distance and time, and for a path whose value may be a function
of the variable of interest (in this case, time), the infinitesimal form is
ds = P0 dt.

(2.15)

To find the total distance, we integrate between the limits of the initial
position a and the final position b:
s=

P0 dt.

(2.16)

The point is, its not the path P that we need to determine the line integral;
its the change in P, denoted as P0 . This seems counterintuitive at first,
but hopefully the above example makes the point. Its also a bit overkill
when one remembers that derivatives and integrals are opposites of each
other; the above analysis has us determine a derivative and then take the
integral, undoing our original operation, to obtain the answer. One might
have simply kept the original equation and determined the answer from
there. Well address this issue shortly. One more point: it doesnt need to
be a change with respect to time. The derivative involved can be a change
with respect to a spatial variable. This allows us to determine line integrals
with respect to space as well as time.
Suppose that the function for the path P is a vector. For example,
consider a circle C in the (x, y) plane having radius r. Its vector function
is C = r cos i + r sin j + 0k (see Fig. 2.12), which is a function of the
variable , the angle from the positive x axis. What is the circumference of
the circle?; that is, what is the path length as goes from 0 to 2, the radian
measure of the central angle of a circle? According to our formulation
above, we need to determine the derivative of our function. But for a vector,
if we want the total length of the path, we care only about the magnitude
of the vector and not its direction. Thus, well need to derive the change
in the magnitude of the vector. We start by defining the magnitude: the
magnitude |m| of a three- (or lesser-) magnitude vector is the Pythagorean
combination of its components:
|m| =

x2 + y2 + z2 .

(2.17)

First Equation of Electrodynamics

31

Figure 2.12 How far is the path around the circle? A line integral can tell us, and
this agrees with what basic geometry predicts (2r).

For the derivative of the path/magnitude with respect to time, which is the
velocity, we have
|m0 | =

(x0 )2 + (y0 )2 + (z0 )2 .

(2.18)

For our circle, we have the magnitude as simply the i, j, and/or k terms of
the vector. These individual terms are also functions of . We have
d(r cos i)
= r sin i,
d
d(r sin j)
= r cos j,
y0 =
d
z0 = 0.
x0 =

(2.19)

From this we have


(x0 )2 = r2 sin2 i2 ,
(y0 ) = r2 cos2 j2 ,

(2.20)

32

Chapter 2

and we will ignore the z part, since its just zero. For the squares of the unit
vectors, we have i2 = j2 = i i = j j = 1. Thus, we have
s=

P dt =

p

r2

sin +

r2

cos2


d.

(2.21)

We can factor out the r2 term from each term and then factor out the r2
term from the square root to obtain
s=

p
r sin2 + cos2 d.

(2.22)

Since, from elementary trigonometry, sin2 + cos2 = 1, we have


s=

(r 1) d =

2

r d = r = r(2 0) = 2r. (2.23)
0

This seems like an awful lot of work to show what we all know, that the
circumference of a circle is 2r. But hopefully it will convince you of the
propriety of this particular mathematical formulation.
Back to total effect. For a line integral involving a field, there
are two expressions we need to consider: the definition of the field
F[x(q), y(q), z(q)] and the definition of the vector path p(q), where q
represents the coordinate along the path. (Note that, at least initially, the
field F is not necessarily a vector.) In that case, the total effect s of the field
along the line is given by
s=

F[x(q), y(q), z(q)] |p0 (q)| dq.

(2.24)

The integration is over the path p, which needs to be determined by the


physical nature of the system of interest. Note that in the integrand, the
two functions F and |p0 | are multiplied together.
If F is a vector field over the vector path p(q), denoted F[p(q)], then the
line integral is defined similarly:
s=

F[p(q)] p0 (q) dq.

(2.25)

First Equation of Electrodynamics

33

Here, we need to take the dot product of the F and p0 vectors.


A line integral is an integral over one dimension that gives, effectively,
the area under the function. We can perform a two-dimensional integral
over the surface of a multidimensional function, as pictured in Fig. 2.13.
That is, we want to evaluate the integral
Z

g(x, y, z) dS ,

(2.26)

where g(x, y, z) is some scalar function on a surface S . Technically, this


expression is a double integral over two variables. This integral is generally
called a surface integral.
The mathematical tactic for evaluating the surface integral is to project
the functional value into the perpendicular plane, accounting for the
variation of the functions angle with respect to the projected plane (the
region R in Fig. 2.13). The proper variation is the cosine function, which
gives a relative contribution of 1 if the function and the plane are parallel
[i.e., cos(0 deg) = 1] and a relative contribution of 0 if the function and the
plane are perpendicular [i.e., cos(90 deg) = 0]. This automatically makes
us think of a dot product. If the space S is being projected into the (x, y)
plane, then the dot product will involve the unit vector in the z direction,
or k. (If the space is projected into other planes, other unit vectors are
involved, but the concept is the same.) If n(x, y, z) is the unit vector that

Figure 2.13 A surface S over which a function f (x, y) will be integrated. R


represents the projection of the surface S in the (x, y) plane.

34

Chapter 2

defines the line perpendicular to the plane marked out by g(x, y, z) (called
the normal vector), then the value of the surface integral is given by
"
R

g(x, y, z)
dx dy,
n(x, y, z) k

(2.27)

where the denominator contains a dot product, and the integration is over
the x and y limits of the region R in the (x, y) plane of Fig. 2.13. The dot
product in the denominator is actually fairly easy to generalize. When that
happens, the surface integral becomes
"
R

g(x, y, z)

f
1+
x

!2

f
+
y

!2

dx dy,

(2.28)

where f represents the function of the surface, and g represents the


function you are integrating over. Typically, to make g a function of only
two variables, you let z = f (x, y) and substitute the expression for z into
the function g, if z appears in the function g.
If, instead of a scalar function g we had a vector function F, the above
equation becomes a bit more complicated. In particular, we are interested
in the effect that is normal to the surface of the vector function. Since we
previously defined n as the vector normal to the surface, well use it again;
we want the surface integral involving F n, or
Z

F n dS .

(2.29)

For a vector function F = F x i + Fy j + Fz k and a surface given by the


expression f (x, y) z, this surface integral is
Z

#
" "
f
f
Fy
+ Fz dx dy.
F n dS =
F x
x
y
R

(2.30)

This is a bit of a mess! Is there a better, easier, more concise way of


representing this?

First Equation of Electrodynamics

35

2.4 A Better Way


There is a better way to represent this last integral, but we need to back up
a bit. What exactly is F n? Actually, its just a dot product, but the integral
Z

F n dS

(2.31)

is called the flux of F. The word flux comes from the Latin word fluxus,
meaning flow. For example, suppose you have some water flowing
through the end of a tube, as represented in Fig. 2.14(a). If the tube is
cut straight, the flow is easy to calculate from the velocity of the water
(given by F) and the geometry of the tube. If you want to express the
flow in terms of the mass of water flowing, you can use the density
of the water as a conversion. But what if the tube is not cut straight,
as shown in Fig. 2.14(b)? In this case, we need to use some morecomplicated geometryvector geometryto determine the flux. In fact,
the flux is calculated using the last integral in the previous section. So, flux
is calculable.
Consider an ideal cubic surface with the sides parallel to the axes (as
shown in Fig. 2.15) that surround the point (x, y, z). This cube represents
our function F, and we want to determine the flux of F. Ideally, the flux
at any point can be determined by shrinking the cube until it arrives at a
single point. We will start by determining the flux for a finite-sized side,
then take the limit of the flux as the size of the size goes to zero. If we look
at the top surface, which is parallel to the (x, y) plane, it should be obvious

Figure 2.14 Flux is another word for amount of flow. (a) In a tube that is cut
straight, the flux can be determined from simple geometry. (b) In a tube cut at an
angle, some vector mathematics is needed to determine flux.

36

Figure 2.15
small?

Chapter 2

What is the surface integral of a cube as the cube gets infinitely

that the normal vector is the same as the k vector. For this surface by itself,
the flux is then
Z
F k dS .
(2.32)
S

If F is a vector function, its dot product with k eliminates the i and j parts
(since i k = j k = 0; recall that the dot product a b = |a||b| cos , where
|a| represents the magnitude of vector a) and only the z component of F
remains. Thus, the integral above is simply
Z

Fz dS .

(2.33)

If we assume that the function Fz has some average value on that top
surface, then the flux is simply that average value times the area of the

First Equation of Electrodynamics

37

surface, which we will propose is equal to x y. We need to note,


though, that the top surface is not located at z (the center of the cube),
but at z + z/2. Thus, we have for the flux at the top surface:
!
z
xy,
top flux Fz x, y, z +
2

(2.34)

where the symbol means approximately equal to. It will become


equal to when the surface area shrinks to zero.
The flux of F on the bottom side is exactly the same except for two small
changes. First, the normal vector is now k, so there is a negative sign on
the expression. Second, the bottom surface is lower than the center point,
so the function is evaluated at z z/2. Thus, we have
bottom flux Fz x, y, z

!
z
xy.
2

(2.35)

The total flux through these two parallel planes is the sum of the two
expressions:
!
!
z
z
xy Fz x, y, z
xy.
flux Fz x, y, z +
2
2

(2.36)

We can factor the xy out of both expressions. Now, if we multiply this


expression by z/z (which equals 1), we have
"

!
!#
z
z
z
flux Fz x, y, z +
Fz x, y, z
xy .
2
2
z

(2.37)

We rearrange as follows:
flux


Fz x, y, z +

z
2


Fz x, y, z

z
2

i

xyz,

(2.38)

and recognize that xyz is the change in volume of the cube V:


flux


Fz x, y, z +

z
2


Fz x, y, z

z
2

i

V.

(2.39)

38

Chapter 2

As the cube shrinks, z approaches zero. In the limit of infinitesimal


change in z, the first term in the product above is simply the definition
of the derivative of Fz with respect to z! Of course, its a partial derivative
because F depends on all three variables, but we can write the flux more
simply as
flux =

Fz
V.
z

(2.40)

A similar analysis can be performed for the two sets of parallel planes;
only the dimension labels will change. We ultimately obtain
Fy
Fx
Fz
V +
V +
V
x
y
z
!
F x Fy Fz
V.
+
+
=
x
y
z

total flux =

(2.41)

(Of course, as x, y, and z go to zero, so does V, but this does not


affect our end result.) The expression in the parentheses above is so useful
that it is defined as the divergence of the vector function F:
divergence of F

F x Fy Fz
+
+
(where F = F x i + Fy j + Fz k). (2.42)
x
y
z

Because divergence of a function is defined at a point, and the flux (two


equations above) is defined in terms of a finite volume, we can also define
the divergence as the limit as volume goes to zero of the flux density
(defined as flux divided by volume):
1
F x Fy Fz
+
+
= lim
(total flux)
V0 V
x
y
z
Z
1
= lim
F ndS .
(2.43)
V0 V S

divergence of F =

There are two abbreviations to indicate the divergence of a vector


function. One is to simply use the abbreviation div to represent

First Equation of Electrodynamics

39

divergence:
div F =

F x Fy Fz
+
+
.
x
y
z

(2.44)

The other way to represent the divergence is with a special function. The
function (called del) is defined as
=i

+j +k .
x
y
z

(2.45)

If we were to take the dot product between and F, we would obtain the
following result:
!



F x i + Fy j + Fz k
F= i +j +k
x
y
z
F x Fy Fz
=
+
+
,
x
y
z

(2.46)

which is the divergence! Note that, although we expect to obtain nine terms
in the dot product above, cross terms between the unit vectors (such as i k
or k j) all equal zero and cancel out, while like terms (that is, j j) all
equal 1 because the angle between a vector and itself is zero and cos 0 = 1.
As such, our nine-term expansion collapses to only three nonzero terms.
Alternately, one can think of the dot product in terms of its other definition,
a b = ai bi = a1 b1 + a2 b2 + a3 b3 ,

(2.47)

where a1 , a2 , etc., are the scalar magnitudes in the x, y, etc., directions. So,
the divergence of a vector function F is indicated by
divergence of F = F.

(2.48)

What does the divergence of a function mean? First, note that the
divergence is a scalar, not a vector, field. No unit vectors remain in the
expression for the divergence. This is not to imply that the divergence is a
constant; it may in fact be a mathematical expression whose value varies

40

Chapter 2

in space. For example, for the field


F = x3 i,

(2.49)

F = 3x2 ,

(2.50)

the divergence is

which is a scalar function. Thus, the divergence changes with position.


Divergence is an indication of how quickly a vector field spreads out at
any given point; that is, how fast it diverges. Consider the vector field
F = xi + yj,

(2.51)

which we originally showed in Fig. 2.7 and are reshowing in Fig. 2.16.
It has a constant divergence of 2 (easily verified), indicating a constant
spreading out over the plane. However, for the field
F = x2 i,

(2.52)

whose divergence is 2x, the vectors grow farther and farther apart as x
increases (see Fig. 2.17).

2.5 Maxwells First Equation


If two electric charges were placed in space near each other, as shown in
Fig. 2.18, there would be a force of attraction between the two charges; the
charge on the left would exert a force on the charge on the right, and vice
versa. That experimental fact is modeled mathematically by Coulombs
law, which in vector form is
F=

q1 q2
r,
r2

(2.53)

where q1 and q2 are the magnitudes of the charges (in elementary units,
where the elementary unit is equal to the charge on the electron), and
r is the scalar distance between the two charges. The unit vector r
represents the line between the two charges q1 and q2 . The modern version
of Coulombs law includes a conversion factor between charge units

First Equation of Electrodynamics

41

Figure 2.16 The divergence of the vector field F = xi + yj is 2, indicating a


constant divergence and, thus, a constant spreading out, of the field at any point
in the (x, y) plane.

Figure 2.17 A nonconstant divergence is illustrated by this one-dimensional field


F = x2 i whose divergence is equal to 2x. The arrowheads, which give the
divergence and not the vector field itself, represent lengths of the vector field at
values of x = 1, 2, 3, 4, etc. The greater the value of x, the farther apart the
vectors become; that is, the greater the divergence.

42

Chapter 2

Figure 2.18 It is an experimental fact that charges exert forces on each other.
That fact is modeled by Coulombs law.

(coulombs, C) and force units (newtons, N), and is written as


F=

q1 q2
r,
40 r2

(2.54)

where 0 is called the permittivity of free space and has an approximate


value of 8.854 . . . 1012 C2 /N m2 .
How does a charge cause a force to be felt by another charge? Michael
Faraday suggested that a charge has an effect in the surrounding space
called an electric field, a vector field, labeled E. The electric field is defined
as the Coulombic force felt by another charge divided by the magnitude of
the original charge, which we will choose to be q2 :
E=

q1
F
=
r,
q2 40 r2

(2.55)

where in the second expression we have substituted the expression for F.


Note that E is a vector field (as indicated by the bold-faced letter) and
is dependent on the distance from the original charge. E, too, has a unit
vector that is defined as the line between the two charges involved, but,
in this case, the second charge has yet to be positioned, so, in general, E
can be thought of as a spherical field about the charge q1 . The unit for an
electric field is newton per coulomb, or N/C.
Since E is a field, we can pretend it has flux; that is, something is
flowing through any surface that encloses the original charge. What is
flowing? It doesnt matter; all that matters is that we can define the flux
mathematically. In fact, we can use the definition of flux given earlier. The
electric flux is given by
Z
=
E n dS ,
(2.56)
S

which is perfectly analogous to our previous definition of flux.

First Equation of Electrodynamics

43

Let us consider a spherical surface around our original charge that has
some constant radius r. The normal unit vector n is simply r, the radius
unit vector, since the radius unit vector is perpendicular to the spherical
surface at any of its points (Fig. 2.19). Since we know the definition of
E from Coulombs law, we can substitute into the expression for electric
flux:
=

q1
r r dS .
40 r2

(2.57)

Figure 2.19 A charge in the center of a spherical shell with radius r has a normal
unit vector equal to r in the radial direction and with unit length at any point on the
surface of the sphere.

44

Chapter 2

The dot product r r is simply 1, so this becomes


=

q1
dS .
40 r2

(2.58)

If the charge q1 is constant, 4 is constant, is constant, the radius r is


constant, and the permittivity of free space is constant, these can all be
removed from the integral to obtain
q1
=
40 r2

dS .

(2.59)

What is this integral? Well, we defined our system as a sphere, so the


surface integral above is the surface area of a sphere. The surface area
of a sphere is known: 4r2 . Thus, we have
=

q1
4r2 .
40 r2

(2.60)

The 4, the , and the r2 terms cancel. What remains is


=

q1
.
0

(2.61)

Recall, however, that we previously defined the divergence of a vector


function as
1
F x Fy Fz
+
+
= lim
(total flux)
V0 V
x
y
z
Z
1
F ndS .
= lim
V0 V S

div F =

(2.62)

Note that the integral in the definition has exactly the same form as the
electric field flux . Therefore, in terms of the divergence, we have for E:
1
div E = lim
V0 V

1
1 q1
= lim
, (2.63)
V0 V
V0 V 0

E ndS = lim

First Equation of Electrodynamics

45

where we have made the appropriate substitutions to obtain the final


expression. We rewrite this last expression as
div E =

lim q1
V0 V
0

(2.64)

The expression q1 /V is simply the charge density at a point, which we


will define as . This last expression becomes simply
div E =

.
0

(2.65)

This equation is Maxwells first equation. It is also written as


E=

.
0

(2.66)

Maxwells first equation is also called Gauss law, after Carl Friedrich
Gauss, the German polymath who first determined it but did not publish
it. [It was finally published in 1867 (after Gauss death) by his colleague
William Weber; Gauss had a habit of not publishing much of his work,
and his many contributions to science and mathematics were realized only
posthumously.]
What does this first equation mean? It means that a charge puts out a
field whose divergence is constant and depends on the charge density and
a universal constant. In words, it says that the electric flux (the left side
of the equation) is proportional to the charge inside the closed surface.
This may not seem like much of a statement, but then, were only getting
started.

Chapter 3

Second Equation of
Electrodynamics B = 0
Maxwells first equation dealt only with electric fields,
saying nothing about magnetism. That changes with the
introduction of Maxwells second equation. A review of
the mathematical development of the divergence, covered
in detail in Chapter 2, may be helpful.

3.1 Introduction
A magnet is any object that produces a magnetic field. Thats a rather
circular definition (and saying such is a bit of a pun, when you understand
Maxwells equations!), but it is a functional one; a magnet is most simply
defined by how it functions.
As mentioned in previous chapters, technically speaking, all matter
is affected by magnets. Its just that some objects are affected more
than others, and we tend to define magnetism in terms of the more
obvious behavior. An object is magnetic if it attracts certain metals such
as iron, nickel, or cobalt, and if it attracts and repels (depending on its
orientation) other magnets. The earliest magnets occurred naturally and
were called lodestones, a name that apparently comes from the Middle
English leading stone, suggesting an early recognition of the rocks
ability to point in a certain direction when suspended freely. Lodestone, by
the way, is simply a magnetic form of magnetite, an ore whose name comes
from the Magnesia region of Greece, which is itself a part of Thessaly in
central eastern Greece bordering the Aegean Sea. Magnetites chemical
formula is Fe3 O4 , and the ore is actually a mixed FeO-Fe2 O3 mineral.
Magnetite itself is not uncommon, although the permanently magnetized
47

48

Chapter 3

form is, and how it becomes permanently magnetized is still an open


question. (The chemists among us also recognize Magnesia as giving its
name to the element magnesium. Ironically, the magnetic properties of Mg
are about 1/5000th that of Fe.)
Magnets work by setting up a magnetic field. What actually is a
magnetic field? To be honest, it may be difficult to put into words, but
effects of a magnet can be measured all around the object. It turns out that
these effects exert forces that have magnitude and direction; that is, the
magnetic field is a vector field. These forces are most easily demonstrated
by objects that have either a magnetic field themselves or are moving and
have an electrical charge on them, as the exerted force accelerates (changes
the magnitude and/or direction of the velocity of) the charge. The magnetic
field of a magnet is represented as B, and again is a vector field. (The
symbol H is also used to represent a magnetic field, although in some
circumstances there are some subtle differences between the definitions of
the B field and the of the H field. Here we will use B.)

3.2 Faradays Lines of Force


When Michael Faraday (see Chapter 1) was investigating magnets starting
in the early 1830s, he invented a description that was used to visualize
magnets actions: lines of force. There is some disagreement as to whether
Faraday thought of these lines as being caused by the emanation of discrete
particles or not, but no matter. The lines of force are what is visualized
when fine iron filings are sprinkled over a sheet of paper that is placed over
a bar magnet, as shown in Fig. 3.1. The filings show some distinct lines
about which the iron pieces collect, although this is more a physical effect
than it is a representation of a magnetic field. There are several features
that can be noted from the positions of the iron filings in Fig. 3.1. First, the
field seems to emanate from two points in the figure, where the iron filings
are most concentrated. These points represent poles of the magnet. Second,
the field lines exist not only between the poles but arc above and below the
poles in the plane of the figure. If this figure extended to infinity in any
direction, you would still see evidence (albeit less and less as you proceed
farther away from the magnet) of the magnetic field. Third, the strength of
the field is indicated by the density of lines in any given space; the field
is stronger near the poles and directly between the poles, and the field is
weaker farther away from the poles. Finally, we note that the magnetic
field is three-dimensional. Although most of the figure shows iron filings

Second Equation of Electrodynamics

49

Figure 3.1 Photographic representation of magnetic lines of force. Here, a


magnetic stir bar was placed under a sheet of paper, and fine iron filings were
carefully sprinkled onto the paper. While the concept of lines of force is a useful
one, magnetic fields are continuous and are not broken down into discrete lines
as is pictured here. (Photo by author, with assistance from Dr. Royce W. Beal,
Mr. Randy G. Ramsden, and Dr. James Rohrbough of the US Air Force Academy
Department of Chemistry.)

on a flat plane, around the two poles the iron filings are definitely out of
the plane of the figure, pointing up. (The force of gravity is keeping the
filings from piling too high, but the visual effect is obvious.) For the sake
of convention, the lines are thought of as coming out of the north pole
of a magnet and going into the south pole of the magnet, although in
Fig. 3.1 the poles are not labeled.
Faraday was able to use the concept of lines of force to explain attraction
and repulsion by two different magnets. He argued that when the lines of
force from opposite poles of two magnets interact, they join together in
such a way as to try to force the poles together, accounting for attraction of
opposites [Fig. 3.2(a)]. However, if lines of force from similar poles of two
magnets interact, they interfere with each other in such a way as to repel
[Fig. 3.2(b)]. Thus, the lines of force were useful constructs to describe the
known behavior of magnets.
Faraday could also use the lines-of-force concept to explain why
some materials were attracted by magnets (paramagnetic materials,
or in their extreme, ferromagnetic materials) or repelled by magnets
(diamagnetic materials). Figure 3.3 illustrates that materials attracted by
a magnetic field concentrate the lines of force inside the material, while

50

Chapter 3

Figure 3.2 Faraday used the concept of magnetic lines of force to describe
attraction and repulsion. (a) When opposite poles of two magnets interact, the
lines of force combine to force the two poles together, causing attraction. (b) When
like poles of two magnets interact, the lines of force resist each other, causing
repulsion.

materials repelled by a magnetic field exclude the lines of force from the
material.
As useful as these descriptions were, Faraday was not a theorist. He was
a very phenomenological scientist who mastered experiments but had little
mathematical training with which to model his results. Others did that
others in Germany and Francebut none more so than in his own Great
Britain.

3.3 Maxwells Second Equation


Two British scientists contributed to a better theoretical understanding of
magnetism: William Thomson (also known as Lord Kelvin) and James
Clerk Maxwell. However, it was Maxwell who did the more complete job.
Maxwell was apparently impressed with the concept of Faradays lines
of force. In fact, the series of four papers in which he described what was
to become Maxwells equations were titled On Physical Lines of Force.
Maxwell was a very geometry-oriented person; he felt that the behavior of
the natural universe could, at the very least, be represented by a drawing
or picture.
So, consider the lines of force pictured in Fig. 3.1. Figure 3.4 shows
one ideal line of force for a bar magnet in two dimensions. Recall that this
is a thought experimenta magnetic field is not composed of individual
lines; rather, it is a continuous vector field. And it is a vector field, so
the field lines have some direction as well as magnitude. By convention,
the magnetic field vectors are thought of as emerging from the north pole

Second Equation of Electrodynamics

51

Figure 3.3 Faraday used the lines-of-force concept to explain how objects
behave in a magnetic field. (a) Most substances (such as glass, water, or elemental
bismuth) actually slightly repel a magnetic field; Faraday explained that they
excluded the magnetic lines of force from themselves. (b) Some substances (such
as aluminum) are slightly attracted to a magnetic field; Faraday suggested that
they include magnetic lines of force within themselves. (c) Some substances (such
as iron) are very strongly attracted to a magnetic field, including (according to
Faraday) a large density of lines of force. Such materials can be turned into
magnets themselves under the proper conditions.

52

Chapter 3

Figure 3.4 Hypothetical line of force about a magnet. Compare this to the photo
in Fig. 2.19.

of the magnet and entering the south pole of the magnet. This vector
scheme allows us to apply the right-hand rule when describing the effects
of the magnetic field on other objects, such as charged particles and other
magnetic phenomena.
Consider any box around the line of force. In Fig. 3.4, the box is shown
by the dotted rectangle. What is the net change of the magnetic field
through the box? By focusing on the single line of force drawn, we can
conclude that the net change is zero: there is one line entering the box on
its left side, and one line leaving the box on its right side. This is easily seen
in Fig. 3.4 for one line of force and in two dimensions, but now lets expand
our mental picture to include all lines of force and all three dimensions;
there will always be the same number of lines of force going into any
arbitrary volume about the magnet as there are coming out. There is no net
change in the magnetic field in any given volume. This concept holds no
matter how strong the magnetic field and no matter what size the volume
considered.
How do we express this mathematically? Why, using vector calculus,
of course. In the previous discussion of Maxwells first law, we introduced
the divergence of a vector function F as
divergence of F

F x Fy Fz
+
+
,
x
y
z

(3.1)

where F = F x i + Fy j + Fz k. Note what the divergence really is: it is the


change in the x-dimensional value of the function F across the x dimension
plus the change in the y-dimensional value of the function F across the y
dimension plus the change in the z-dimensional value of the function F
across the z dimension. But we have already argued form our lines-of-force
illustration that the magnetic field coming in a volume equals the magnetic

Second Equation of Electrodynamics

53

field going out of the volume, so that there is no net change. Thus, using B
to represent our magnetic field,
Bx By Bz
=
=
= 0.
x
y
z

(3.2)

That means that the divergence of B can be written as


div B =

Bx By Bz
=
=
=0
x
y
z

(3.3)

or simply
div B = 0.

(3.4)

This is Maxwells second equation of electromagnetism. It is sometimes


called Gauss law for magnetism. Since we can also write the divergence
as the dot product of the del operator () with the vector field, Maxwells
second equation is
B = 0.

(3.5)

What does Maxwells second equation mean? Because the divergence


is an indicator of the presence of a source (a generator) or a sink (a
destroyer) of a vector field, it implies that a magnetic field has no
separate generator or destroyer points in any definable volume. Contrast
this with an electric field. Electric fields are generated by two different
particles, positively charged particles and negatively charged particles. By
convention, electric fields begin at positive charges and end at negative
charges. Since electric fields have explicit generators (positively charged
particles) and destroyers (negatively charged particles), the divergence
of an electric field is nonzero. Indeed, by Maxwells first equation, the
divergence of an electric field E is
E=

,
0

(3.6)

which is zero only if the charge density is zero; if the charge density is
not zero, then the divergence of the electric field is also not zero. Further,

54

Chapter 3

the divergence can be positive or negative depending on whether the charge


density is a source or a sink.
For magnetic fields, however, the divergence is exactly zero, which
implies that there is no discrete source (positive magnetic particle)
or sink (negative magnetic particle). One implication of this is that
magnetic field sources are always dipoles; there is no such thing as a
magnetic monopole. This mirrors our experience when we break a
magnet in half, as shown in Fig. 3.5. We do not end up with two separated
poles of the original magnet. Rather, we have two separate magnets,
complete with north and south poles.
In the next chapter, we will continue our discussion of Maxwells
equations and see how E and B are related to each other. The first two of
Maxwells equations deal with E and B separately; we will see, however,
that they are anything but separate.

Figure 3.5 If you break a magnet, you dont get two separate magnetic poles
(monopoles, top), but instead you get two magnets, each having north and south
poles (bottom). This is consistent with Maxwells second law of electromagnetism.

Chapter 4

Third Equation of
Electrodynamics E = t B
Maxwells equations are expressed in the language of
vector calculus, so a significant portion of the previous
chapters has been devoted to explaining vector calculus,
not Maxwells equations. For better or worse, thats par
for the course, and its going to happen again in this
chapter. The old adage the truth will set you free might
be better stated, for our purposes, as the math will set
you free. And thats the truth.

4.1 Beginnings
In mid-1820, Danish physicist Hans Christian rsted discovered that
a current in a wire can affect the magnetic needle of a compass.
His experiments were quickly confirmed by Franois Arago and, more
exhaustively, by Andr Marie Ampre. Ampres work demonstrated
that the effects generated by the current, which defined a so-called
magnetic field (labeled B in Fig. 4.1), were centered on the wire, were
perpendicular to the wire, and were circularly symmetric about the wire.
By convention, the vector component of the field had a direction given by
the right-hand rule: if the thumb of the right hand were pointing in the
direction of the current, the curve of the fingers on the right hand gives the
direction of the vector field.
Other careful experiments by Jean-Baptiste Biot and Flix Savart
established that the strength of the magnetic field was directly related to
the current I in the wire and inversely related to the radial distance from
55

56

Figure 4.1
through it.

Chapter 4

The shape of a magnetic field about a wire with a current running

the wire r. Thus, we have


1
B ,
r

(4.1)

where means proportional to. To make a proportionality into an


equality, we introduce a proportionality constant. However, because of
the axial symmetry of the field, we typically include a factor of 2 (the
radian angle of a circle) in the denominator of any arbitrary proportionality
constant. As such, our new equation is
B=

I
,
2 r

(4.2)

where the constant is our proportionality constant and is called the


permeability of the medium that the magnetic field is in. In a vacuum,
the permeability is labeled 0 and, because of how the units of B and I are
defined, is equal to exactly 4 107 T m/A (tesla meters per ampere).
Not long after the initial demonstrations, Ampre had another idea:
curve the wire into a circle. Sure enough, inside the circle, the magnetic
field increased in strength as the concentric circles of the magnetic field
overlapped on the inside (Fig. 4.2). Biot and Savart found that the magnetic
field B created by the loop was related to the current I in the loop and the
radius of the loop R by the following:
B=

I
.
2R

(4.3)

Multiple loops can be joined in sequence to increase B, and from 1824


to 1825 English inventor William Sturgeon wrapped loops around a piece

58

Chapter 4

Figure 4.3 (a) A magnet inside a coil of wire does not generate a current. (b) A
magnet moving through a coil of wire does generate a current.

Actually, this is not far from the truth (it would have become another of
Maxwells equations if it were the truth), but the more complete truth is
expressed in a different, more applicable form.

Third Equation of Electrodynamics

59

4.2 Work in an Electrostatic Field


The simple physical definition of work w is force F times displacement
s:
w = F s.

(4.5)

This is fine for straight-line motion, but what if the motion occurs on a
curve (Fig. 4.4 in two dimensions) with perhaps a varying force? Then
calculating the work is not as straightforward, especially since force and
displacement are both vectors. However, it can be easily justified that the
work is the integral, from initial point to final point, of the dot product
force vector F with the unit vector tangent to the curve, which we will
label t:
Z final point
w=
F t ds.
(4.6)
initial point

Because of the dot product, only the force component in the direction of
the tangent to the curve contributes to the work. This makes sense if you

Figure 4.4 If the force F is not parallel to the displacement s (shown here as
variable, but F can be vary, too), then the work performed is not as straightforward
to calculate.

60

Chapter 4

remember the definition of the dot product, a b = |a||b| cos ; if the force
is parallel to the displacement, work is maximized [because the cosine of
the angle between the two vectors is cos(0 deg) = 1], while if the force is
perpendicular to the displacement, work is zero [because now the cosine
is cos(90 deg) = 0].
Now consider two random points inside of an electrostatic field E
(Fig. 4.5). Keep in mind that we have defined E as static; that is, not
moving or changing. Imagine that an electric particle with charge q were to
travel from P1 to P2 and back again along the paths s1 and s2 , as indicated.
Since the force F on the particle is given by qE (from Coulombs law), we
have for an imagined two-step process
w=

P2
P1

qE t ds1 +

P1

qE t ds2 .

(4.7)

P2

Each integral covers one pathway, but eventually you end up where you
started.

Figure 4.5 Two arbitrary points in an electric field. The relative strength of the
field is indicated by the darkness of the color.

Third Equation of Electrodynamics

61

This last statement is a crucial one: eventually you end up where you
started. According to Coulombs law, the only variable that the force or
electric field between the two particles depends on is the radial distance r.
This further implies that the work w depends only on the radial distance
between any two points in the electric field. Even further still, this implies
that if you start and end at the same point, as in our example, the overall
work is zero because you are starting and stopping at the same radial point
r. Thus, the equation above must be equal to zero:
Z

P2

qE t ds1 +

P1

P1

qE t ds2 = 0.

(4.8)

P2

Since we are starting and stopping at the same point, the combined paths
s1 and s2 are termed a closed path. Notice, too, that, other than being
closed, we have not imposed any requirement on the overall path itself;
it can be any path. We say that this integral, which must equal zero, is path
independent.
H
The symbol for an integral over a closed path is . Thus, we have
I

qE t ds = 0.

(4.9)

We can divide by the constant q to obtain something slightly more simple:


I

E t ds = 0.

(4.10)

This is one characteristic of an electrostatic field: the path-independent


integral over any closed path in an electrostatic field is exactly zero.
The key word in the above statement is any. You can select any
random closed path in an electric field, and the integral of E t over that
path is exactly zero. How can we generalize this for any closed path?
Let us start with a closed path in one plane, as shown by Fig. 4.6. The
complete closed path has four parts, labeled T, B, L, and R for top, bottom,
left, and right sides, respectively, and it surrounds a point at some given
coordinates (x, y, z). T and B are parallel to the x axis, while L and R are
parallel to the y axis. The dimensions of the path are x by y (these will
be useful shortly). Right now the area enclosed by the path is arbitrary, but
later on we will want to shrink the closed path down so that the area goes

62

Chapter 4

Figure 4.6

A closed, two-dimensional path around a point.

to zero. Finally, the path is immersed in a three-dimensional field F whose


components are F x , Fy , and Fz . That is,
F = iF x + jFy + kFz

(4.11)

in terms of the three unit vectors i, j, and k in the x, y, and z dimension,


respectively.
Let us evaluate the work of each straight segment of the path separately,
starting with path B. The work is
wB =

Z
F t ds.

(4.12)

The tangent vector t is simply the unit vector i, since path B points along
the positive x axis. When you take the dot product of i with F [see
Eq. (4.12)], the result is simply F x . (Can you verify this?) Finally, since
the displacement s is along the x axis, ds is simply dx. Thus, we have
wB =

Z
F x dx.
B

(4.13)

Third Equation of Electrodynamics

63

Although the value of F x can vary as you move across path Bin fact, it
is better labeled as F x (x, y, z)let us assume some average value of F x as
indicated by its value at a y-axis position of y y/2, which is the y value
that is one-half of the height of the box below the point in the center. Thus,
we have
!
y
, Z x,
x, y
z

wB =

F x dx F x
B

(4.14)

where we have replaced the infinitesimal dx with the finite x.


We can do the same for the work at the top of the box, which is path T.
There are only two differences: first, the tangent vector is i because the
path is moving in the negative direction, and second, the average value of
F x is judged at y + y/2, which is one-half of the height of the box above
the center point. Hence we can simply write
wT F x

!
y
x, y +
, z x.
2

(4.15)

The sum of the work on the top and bottom are thus
!
!
y
y
x, y
, z x F x x, y +
, z x.
2
2

WT+B F x

(4.16)

Rearranging Eq. (4.16) so that it is in the form top minus bottom and
factoring out x, this becomes
"
WT+B F x

!
!#
y
y
, z F x x, y
, z x.
x, y +
2
2

(4.17)

Let us multiply this expression by 1, in the form of y/y. We now have


h
WT+B


F x x, y +

y
2 ,z


F x x, y

y
2 ,z

i
xy.

(4.18)

Recall that this work is actually a sum of two integrals involving,


originally, the integrand F t. With this in mind, Eq. (4.18) can be written

64

Chapter 4

as
WT+B =

Z
F t ds


F x x, y +

y
2 ,z


F x x, y

y
2 ,z

i

T+B

xy. (4.19)

The term xy is the area A of the path. Dividing by the area on the left
side of the equation, we have
1
A

Z
F t ds


F x x, y +

y
2 ,z


F x x, y

T+B

y
2 ,z

i
.

(4.20)

Suppose that we take the limit of this expression as x = y = A 0.


What we would have is the amount of work done over any infinitesimal
area defined by any random path; the only restriction is that the path is in
the (x, y) plane. Equation (4.20) above becomes
"
lim

A0

1
A

Z
T+B

h 
#
F x x, y +
F t ds = lim
y0

y
2 ,z


F x x, y

y
2 ,z

i

(4.21)

Looking at the second limit in Eq. (4.21) and recalling our basic calculus,
that limit defines a derivative with respect to y! But because F x is a function
of three variables, this is better defined as the partial derivative with respect
to y. Thus, we have
" Z
#
1
Fx
F t ds =
.
A0 A T+B
y
lim

(4.22)

Note the retention of the minus sign.


We can do the same for paths L and R. The analysis is exactly the same;
only the variables that are affected change. What we obtain is (you are
welcome to verify the derivation)
"

1
lim
A0 A

#
Fy
F t ds =
.
x
L+R

(4.23)

Third Equation of Electrodynamics

65

Now, we combine the two parts. The work done over an infinitesimally
small closed path in the (x, y) plane is given by
" I
#
Fy F x
1
w
F t ds =

.
lim = lim
A0 A
A0 A
x
y

(4.24)

Now isnt that a rather simple result?


Let us see an example of this result so we can understand what it means.
Consider a two-dimensional sink in the (x, y) plane, as diagrammed in
Fig. 4.7. A thin film of water is going down the central drain, and in
this case it is spinning in a counter-clockwise direction at some constant
angular velocity. The vector field for the velocity of the spinning water is
v=i

x
y
+j .
t
t

(4.25)

In terms of the angular velocity , this can be written as


v = (iy + jx) = iy + jx.

(4.26)

Figure 4.7 A two-dimensional sink with a film of water rotating counterclockwise


as it goes down the drain.

66

Chapter 4

(A conversion to polar coordinates was necessary to go to this second


expression for v, in case you need to do the math yourself.) In this vector
field, F x = y, and Fy = x. To determine the limit of the work per unit
area, we evaluate the expression

(x) (y).
x
y

(4.27)

= () = 2.

(4.28)

This is easy to evaluate:

Suppose that we stand up a piece of cardboard on the sink, centered at


the drain. Experience suggests to us that the cardboard piece will start
to rotate, with the axis of rotation perpendicular to the flat sink. In this
particular case, the axis of rotation will be in the z dimension, and in order
to be consistent with the right-hand rule, we submit that in this case the axis
points in the positive z direction. If this axis is considered to be a vector,
then the unit vector in this case is (positive) k. Thus, vectorally speaking,
the infinitesimal work per unit area is actually
"

#
Fy F x
k.

x
y

(4.29)

Thus, the closed loop in the (x, y) plane is related to a vector in the z
direction. In the case of a vector field, the integral over the closed path
is referred to as the circulation of the vector field.
As a counterexample, suppose that water in our two-dimensional sink
is flowing from left to right at a constant velocity, as shown in Fig. 4.8. In
this case, the vector function is
v = Ki,

(4.30)

where K is a constant. If we put a piece of cardboard in this sink, centered


on the drain, does the cardboard rotate? No, it does not. If we evaluate the
partial-derivative expression from above (in this case, F x = K and Fy = 0):
"

(K) (0) k = 0k = 0.
x
y

(4.31)

Third Equation of Electrodynamics

Figure 4.8
velocity.

67

Water flowing in a two-dimensional sink with a constant left-to-right

(Recall that the derivative of a constant is zero.) This answer implies that
no rotation is induced by the closed loop.

4.3 Introducing the Curl


For a vector function F = F x i + Fy j + Fz k, we define the function
"

#
Fy F x
k

x
y

(4.32)

as the one-dimensional curl of F. Possibly improperly, it is designated one


dimensional because the result is a vector in one dimension, in this case
the z dimension. The analysis we performed in the earlier section (defining
a closed path in a single plane and taking the limit of the path integral) can
be performed for the (x, z) and (y, z) planes. When we do that, we obtain
the following analogous results:
#
F x Fz
j;

(x, z) plane :
z
x
"

68

Chapter 4

#
Fz Fy
i.

(y, z) plane :
y
z
"

(4.33)

The combination of all three expressions in Eqs. (4.32) and (4.33) gives us
a general expression for the curl of F:
"
curl F =

#
"
#
"
#
Fy F x
F x Fz
Fz Fy
i+
j+
k. (4.34)

y
z
z
x
x
y

This expression allows us to determine


"
lim

A0

1
A

#
F t ds

(4.35)

for any vector function F in any plane.


But what does the curl of a vector function mean? One way of thinking
about it is that it is a variation in the vector function F that causes a
rotational effect about a perpendicular axis. [Indeed, curl F is sometimes
still designated rot F, and a vector function whose curl equals zero (see
Fig. 4.8) is termed irrotational.] Also, a vector function with a nonzero
curl can be thought of as curving around a particular axis, with that axis
being normal to the plane of the curve. Thus, the rotating water in Fig. 4.7
has a nonzero curl, while the linearly flowing water in Fig. 4.8 has a zero
curl.
A mnemonic (that is, a memory aid) for the general expression for curl
F takes advantage of the structure of a 3 3 determinant:

i j

curl F =
x y
F x Fy


k

.
z
Fz

(4.36)

Understand that curl F is NOT a determinant; a determinant is a number


that is a characteristic of a square matrix of numerical values. However,
the expression for curl F can be constructed by performing the same
operations on the expressions as one would perform with numbers to
determine the value of a 3 3 determinant: constructing the diagonals,
adding the right-downward diagonals, and subtracting the left-upwards

Third Equation of Electrodynamics

69

diagonals. In case you have forgotten how to do this, Fig. 4.9 shows how
to determine the expression for the curl.
The determinental form of the curl can be expressed in terms of the del
operator . Recall from Chapter 2 that the del operator is
i

+j +k .
x
y
z

(4.37)

Recall from vector calculus that the cross product of two vectors A
iA x + jAy + kAz and B defined analogously is written A B and is given
by the expression


i j k
A B = A x Ay Az .


Bx By Bz

(4.38)

By comparing this expression to the determinental form of the curl, it


should be easy to see that the curl of a vector function F can be written
as a cross product:
curl F F.

(4.39)

Figure 4.9 To determine the expression using a determinant, multiply the three
terms on each arrow and apply the positive or negative sign to that product, as
indicated. Combining all terms yields the proper expression for the curl of a vector
function F having components F x , Fy , and Fz .

70

Chapter 4

Like the fact that curl is not technically a determinant, the curl of a function
is technically not a cross product, as del is an operator, not a vector. The
parallels, however, make it easy to gloss over this technicality and use the
del cross F symbolism to represent the curl of a vector function.
Because the work integral over a closed path through an electrostatic
field E is zero, it is a short logical step to state that, therefore,
E = 0.

(4.40)

This is one more property of an electrostatic field: the field is not rotating
about any point in space. Rather, an electrostatic field is a purely radial
field, with all field lines going from a point in space straight to the
electric charge.

4.4 Faradays Law


An electrostatic field caused by a charged particle is thought of as
beginning at a positive charge and ending at a negative charge. Since
overall, matter is electrically neutral, every electric field emanating from a
positive charge eventually ends at a negative charge. This is illustrated in
Fig. 4.10(a).
However, when a changing magnetic field creates a current in a
conductor, this current is the product of an induced electric field. In the
case of a bar-type magnet, the magnetic field is axially symmetric about
the length of the bar, so the induced electric field is axially symmetric as
well; that is, it is circular. This is illustrated in Fig. 4.10(b).
The induced, circular electric field caused by a moving magnet causes
charges to move in that circle. The circulation of the induced electric field
vector can be constructed from our definition of circulation above; it is
I
circulation =
E t ds,
(4.41)
s

where E is the induced electric field, t is the tangent vector along the path,
and s is the infinitesimal amount of path. In this case, the circulation is
defined as the electromotive force, or EMF. What is this force doing?
Why, causing charges to move, of course! As such, it is doing work, and
our arguments using work in the sections above are all valid here.
What Faraday found experimentally is that a changing magnetic field
induced an electric field (which then forced a current). If you imagine that

Third Equation of Electrodynamics

71

Figure 4.10 (a) In an electrostatic field, the field lines go from the positive charge
to the negative charge. (b) A moving magnetic field induces an electric field, but in
this case the electric field is in a circle, following the axial nature of the magnetic
field lines.

a magnetic field is composed of discrete field lines, what is happening is


that as the number of magnetic field lines in a given area changes with
time, an electric field is induced. Figure 4.11 illustrates this. Consider the
loop of area outlined by the black line. As the magnet is moved to the right,
the number of magnetic field lines that intersect the loop changes. It is this
change that induces the electric field.
The number of field lines per area is called the magnetic flux. In
Chapter 2, we presented how to determine the flux of a vector field. For
a changing vector field F having a unit vector perpendicular (or normal) to

72

Chapter 4

Figure 4.11 As the magnet is moved farther from the loop, the number of
imaginary magnetic field lines intersect the loop changes [here, from (a) seven
lines to (b) three lines]. It is this change that induces an electric field in the loop.

its direction of motion n over some surface S, the flux is defined as


Z
F n dS .
(4.42)
flux =
S

For our magnetic field B, this becomes


magnetic flux =

Z
B n dS .
S

(4.43)

Third Equation of Electrodynamics

73

However, the induced electric field is related to the change in magnetic


flux with time. Thus, we are actually interested in the time derivative of
the magnetic flux:

Z
B n dS .

(4.44)

At this stage, we bring everything together by citing the experimental


facts as determined by Faraday and others: the electromotive force is equal
to the change in the magnetic flux over time. That is,
I
S

E t ds =
t

Z
B n dS .

(4.45)

Let us divide each side of this equation by the area A of the circular path
of the induced current. This area also corresponds to the surface S that the
magnetic field flux is measured over, so we divide one side by A and one
side by S to obtain
1
A

E t ds =
s

1
S t

Z
B n dS .

(4.46)

Suppose that we want to consider the limit of this expression as the area of
the paths shrink to zero size; that is, as A 0. We would have
1
A0 A

1
S 0 S t

E t ds = lim

lim

Z
B n dS .

(4.47)

The left side of Eq. (4.47) is, by definition, the curl of E. What about
the right side? Rather than proving it mathematically, lets consider the
following argument. As the surface S goes to zero, the limit of the
magnetic flux ultimately becomes one magnetic flux line. This single line
will be perpendicular to the infinitesimal surface. Look at the rendering of
the magnetic field lines in Fig. 4.1 if you need to convince yourself of this.
Thus, the dot product Bn is simply B, and the infinite sum of infinitesimal
pieces (which is what an integral is) degenerates to a single value of B. We
therefore argue that

lim
S 0 t

Z
S

B ndS =

B.
t

(4.48)

74

Chapter 4

So what we now have is


E=

B.
t

(4.49)

We are almost done. The law of conservation of energy must be satisfied.


Although it appears that we are getting an induced current from nowhere,
understand that this induced current also generates a magnetic field. In
order for the law of conservation of energy to be satisfied, the new
magnetic flux must oppose the original magnetic flux (this concept
is known as Lenzs law after Henrich Lenz, a Russian physicist who
discovered it). To represent this mathematically, a negative sign must be
included in Eq. (4.49). By convention, the minus sign is put on the right
side, so our final equation is
E=

B.
t

(4.50)

This expression is known as Faradays law of induction, given that Michael


Faraday discovered (or rather, first announced) magnetic induction of
current. It is considered the third of Maxwells equations: a changing
magnetic flux induces an electromotive force, which in a conductor will
promote a current.
Not meaning to minimize the importance of Maxwells other equations,
but the impact of what this equation embodies is huge. Electric motors,
electrical generators, and transformers are all direct applications of a
changing magnetic field being related to an electromotive force. Given the
electrified nature of modern society and the machines that make it that way,
we realize that Maxwells equations have a huge impact on our everyday
lives.

Chapter 5

Fourth Equation of
Electrodynamics

B = 0(J + 0 E
t )
5.1 Ampres Law
One of the giants in the development of the modern understanding of
electricity and magnetism was the French scientist Andr-Marie Ampre
(17751836). He was one of the first to demonstrate conclusively that
electrical current generates magnetic fields. For a straight wire, Ampre
demonstrated that (1) the magnetic fields effects were centered on the
wire carrying the current, (2) were perpendicular to the wire, and (3) were
symmetric about the wire. Though illustrated in an earlier chapter, this is
illustrated again in Fig. 5.1.
This figure is strangely reminiscent of Fig. 4.7, reproduced here
as Fig. 5.2, which depicts water circulating around a drain in a

Figure 5.1
through it.

The shape of a magnetic field about a wire with a current running


75

76

Chapter 5

Figure 5.2

Water going in a circular path about a drain (center).

counterclockwise fashion. But now, lets put a paddle wheel in the drain,
with its axis sticking in the drain as shown in Fig. 5.3. We see that the
paddle wheel will rotate about an axis that is perpendicular to the plane
of flow of the water. Rotate our water-and-paddle-wheel figure by 90 deg
counterclockwise, and you have an exact analogy to Fig. 5.1.
We argued in the previous chapter that water circulating in a sink as
shown in Figs. 5.1 and 5.2 represents a function that has a nonzero curl.
Recall that the curl of a vector function F designated curl F or F
is defined as
#
"
#
"
#
"
Fy F x
Fz Fy
F x Fz
curl F = F =
i+
j+
k, (5.1)

y
z
z
x
x
x
where F x , Fy , and F x are the x, y, and z magnitudes of F, and i, j, and k
are the unit vectors in the x, y, and z dimensions, respectively. In admitting
the similarity between Figs. 5.1 and 5.2, we suggest that the curl of the
magnetic field B is related to the current in the straight wire. There is a
formal way to derive this. Recall from the last chapter that the curl of a

Fourth Equation of Electrodynamics

77

Figure 5.3 The paddle wheel rotates about a perpendicular axis when placed in
circularly flowing water. We say that the water has a nonzero curl. The axis of the
paddle wheel is consistent with the right-hand rule, as shown by the inset.

function F as
1
curl F = F = lim
A0 A

E t ds.

(5.2)

For the curl of the magnetic field, we thus have


1
B = lim
A0 A

B t ds.

(5.3)

Recall that S is the surface about which the line s is tracing, t is the tangent
vector on the field line, and A is the area of the surface.
This simplifies easily when one remembers that we have a formula for
B in terms of the distance from the wire r; it was presented in Chapter 3
and is
B=

I
,
2r

(5.4)

where I is the current, r is the radial distance from the wire, and is the
constant known as the permeability of the medium; for vacuum, the symbol
0 is used, and its value is defined as 4 107 T m/A (tesla meters
per ampere). We also know that the magnetic field paths are circles; thus,
as we integrate about the surface, the integral over ds becomes simply the
circumference of a surface, 2r. Substituting these expressions into the

78

Chapter 5

curl of B, we obtain
B = lim

A0

1 I
(2r).
A 2r

(5.5)

The 2r terms cancel, and we are left with


B = lim

A0

I
.
A

(5.6)

The constant can be taken out of the limit. What we have left to
interpret is
lim

A0

I
.
A

(5.7)

This is the limit of the current I flowing through an area A of the wire
as the area grows smaller and smaller, ultimately approaching zero. This
infinitesimal current per area is called the current density and is designated
by J; it has units of coulombs per square meter, or C/m2 . Since current is
technically a vector, so is current density J. Thus, we have
B = J.

(5.8)

This expression is known as Ampres circuital law. In a vacuum, the


expression becomes
B = 0 J.

(5.9)

This is not one of Maxwells equations; it is incomplete. It turns out that


there is another source of a magnetic field.

5.2 Maxwells Displacement Current


The basis of Ampres circuital law was discovered in 1826 (although
its modern mathematical formulation came more than 35 years later). By
the time Maxwell was studying electromagnetic phenomena in the 1860s,
something new had been discovered: magnetic fields from capacitors.
Heres how to think of this new development. A capacitor is a device
that stores electrical charge. The earliest form of a capacitor was the

Fourth Equation of Electrodynamics

79

Leyden jar, described in Chapter 1, and a picture of which is shown in


Fig. 5.4. Although the engineering of modern capacitors can vary, a simple
capacitor can be thought of as two parallel metal plates separated by a
vacuum or some other nonconductor, called a dielectric. Figure 5.5 is a
diagram of a parallel-plate capacitor.

Figure 5.4 A series of four Leyden jars in a museum in Leiden, The Netherlands.
This type of jar was to be filled with water. The apparatus on the bottom side is a
simple electrometer, meant to give an indication of how much charge was stored
in these ancient capacitors.

80

Chapter 5

Figure 5.5

Diagram of a simple parallel-plate capacitor.

A capacitor works because of the gap between the plates; in an electrical


circuit, current enters a plate on one side of the capacitor. However,
because of the gap between the plates, the current builds up on one side,
ultimately causing an electric field to exist between the plates. We know
now that current is electrons, so in modern terms, electrons build up on
one side of the plate. However, electrons have a negative charge, which
repel other electrons residing on the other plate. These electrons get forced
away, resulting in the other plate building up an overall positive charge.
These electrons that are forced away represent a current on the other side
of the plate that continues until the maximum charge has built up on the
other plate. This process is illustrated in Fig. 5.6.
Although electrons are not flowing from one side of the capacitor to
the other, during the course of charging the capacitor, a current flows and
generates a magnetic field. This magnetic field, caused by the changing
electric field, is not accounted for by Ampres circuital law because it is
the result of a changing electric field, not a constant current.
Maxwell was concerned about this new source of a magnetic field.
He called this new type of current displacement current and set about
integrating it into Ampres circuital law. Because this magnetic field was
proportional to the development of an electric field; that is, the change in

Fourth Equation of Electrodynamics

81

Figure 5.6 Charging a capacitor: (1) Current enters one plate. (2) Electrons build
up on the plate. (3) Electrons on the other plate are repelled, causing (4) a shortlived current to leave the other plate.

E with respect to time, we have


B

E
,
t

(5.10)

The proportionality constant needed to make this an equality is the


permittivity of free space, symbolized as 0 . This fundamental constant
has a value of about 8.854 1012 C2 /J m. Since an electric field has
units of volts per meter (V/m), the combined terms 0 (E/t) have units
of C2 /(J m) V/(m s), where the s unit comes from the t in the
derivative. A volt is equal to a joule/coulomb, and a coulomb/second is
equal to an ampere, so the combined units reduce to A/m2 , which is a
unit of current density! Thus, we can add the displacement current term
0 (E/t) to the original current density J, yielding Maxwells fourth
equation of electrodynamics:
B = 0 J + 0

!
E
.
t

This equation is sometimes called the AmpreMaxwell law.

(5.11)

82

Chapter 5

5.3 Conclusion
The presentation of Maxwells equations, in their modern differential
form and in a vacuum, is complete. Collectively, they summarize all
of the properties of electric and magnetic fields. However, there is one
more startling conclusion to Maxwells equations that we will defer to an
Afterword.

Afterword
Whence Light?
A.1 Recap: The Four Equations
In a vacuum, Maxwells four equations of electrodynamics, as expressed
in the previous four chapters, are:

Gauss law,
0
B = 0 Gauss law of electromagnetism,
B
E=
Faradays law, and
t
!
E
B = 0 J + 0
AmpreMaxwell law.
t
E=

(A.1)
(A.2)
(A.3)
(A.4)

A few comments are in order. First, Maxwell did not actually present the
four laws in this form in his original discourse. His original work, detailed
in a four-part series of papers titled On Physical Lines of Force contained
dozens of equations. It remained to others, especially English scientist
Oliver Heaviside, to reformulate Maxwells derivations into four concise
equations using modern terminology and symbolism. We owe almost as
much a debt to the scientists who took over after Maxwells untimely death
in 1879 as we do to Maxwell himself for these equations.
Second, note that the four equations have been expressed in differential
forms. (Recall that the divergence and curl operations, and
respectively, are defined in terms of derivatives.) There are other forms
of Maxwells equations, including forms for inside matter (as opposed to
a vacuum, which is what is considered exclusively in this book), integral
forms, so-called macroscopic forms, relativistic forms, even forms that
83

Afterword

84

assume the existence of magnetic monopoles (likely only of interest to


theoretical physicists and science fiction writers). The specific form you
might want to use depends on the quantities you know, the boundary
conditions of the problem, and what you want to predict. Persons interested
in these other forms of Maxwells equations are urged to consult the
technical literature.

A.2 Whence Light?


We began Chapter 1 by claiming that light itself is explained by Maxwells
equations. How? Actually, it comes from an analysis of Faradays law and
the AmpreMaxwell law, as these are the two of Maxwells equations
that involve both E and B.
Among the theorems of vector calculus is the proof (not given here)
that the curl of a curl of a function is related to the divergence. For a given
vector function F, the curl of the curl of F is given by
( F) = ( F) 2 F.

(A.5)

Hopefully, you already recognize F as the divergence of the vector


function F. There is one other type of function present on the right side of
the equation, the simple by itself (without a dot or a cross). Unadorned
by the dot or cross, represents something called the gradient, which is
nothing more than the three-dimensional slope of a vector function, itself
expressed as a vector in terms of the unit vectors i, j, and k:
F = i

F
F
F
+j
+k .
x
y
z

(A.6)

The first term on the right of a curl of a curl [Eq. (A.5)], then, is the
gradient of the divergence of F. The gradient can also be applied twice.
(The gradient is the last term on the right-hand side, as seen by the 2 .)
When this happens, what initially seems complicated simplifies quite a
bit:
F
F
F
+j
+k
F= i
x
y
z
2

!
F
F
F
i
,
+j
+k
x
y
z

(A.7)

Whence Light?

85

which simplifies to
2 F =

2F 2F 2F
+
+
.
2 x 2y 2z

(A.8)

Note that there are now no i, j, or k vectors in these terms, and that there
are no cross terms between x, y, and z. This is because the i, j, and k vectors
are orthonormal: n1 n2 = 1 if n1 and n2 are the same (that is, both are i
or both are j), while n1 n2 = 0 if n1 and n2 are different (for example, n1
represents i and n2 represents k).
What we do is take the curl of both sides of Faradays law:
( E) =

!
B
.
t

(A.9)

Because the curl is simply a group of spatial derivatives, it can be brought


inside the derivative with respect to time on the right side of the equation:
!
( B)
,
( E) =
t

(A.10)

and we can substitute the expression for what the curl of a curl is on the
left side:
!
( B)
2
( E) E =
.
(A.11)
t
The expression B is defined by the AmpreMaxwell law (Maxwells
fourth equation), so we can substitute for B:
i
h 

0 J + 0 E
t
.
( E) E =

t
2

(A.12)

Faradays law (or Maxwells first equation) tells us what E is: it equals
/0 . We substitute this into the first term on the left side:
i
h 
!

0 J + 0 E

t
.

2 E =

0
t

(A.13)

Afterword

86

Now we will rewrite the right side by separating the two terms to obtain
two derivatives with respect to time. Note that the second term becomes a
second derivative with respect to time, and that 0 , the permeability of a
vacuum, distributes through to both terms. We obtain

!
2E
(0 J)

0 0 2 .
2 E =
0
t
t

(A.14)

In the absence of charge, = 0, and in the absence of a current, J = 0.


Under these conditions, the first terms on both sides are zero, and the
negative signs on the remaining terms cancel. What remains is
2 E = 0 0

2E
.
t2

(A.15)

This is a second-order differential equation that relates an electric field that


varies in space and time. That is, it describes a wave, and this differential
equation is known in physics as the wave equation. The general form of
the wave equation is
2 F =

1 2F
,
v2 t2

(A.16)

where v is the velocity of the wave. The function F can be expressed in


terms of sine and cosine functions or as an imaginary exponential function;
the exact expression for an E wave (light) depends on the boundary
conditions and the initial value at some point in space.
The wave equation implies that
1
vlight =
.
0 0

(A.17)

This can be easily demonstrated:


1
vlight = q


12 C2
(4 107 Tm
A ) 8.8541878 10
Jm
= 2.9979246 . . . 108 m/s.

(A.18)

Whence Light?

87

(You need to decompose the tesla unit T into its fundamental units kg/As2
to see how the units work out. Remember also that J = kg m2 /s2 , and that
A = C/s, and everything works out naturally, as it should with units.) Even
by the early 1860s, experimental determinations of the speed of light were
around that value, leading Maxwell to conclude that light was a wave of
an electric field that had a velocity of (1/0 0 )1/2 .
Light is also a magnetic wave. How do we know? Because we can take
the curl of the AmpreMaxwell law and perform similar substitutions.
This exercise is left to the reader, but the conclusion is not. Ultimately, you
will obtain
2 B = 0 0

2B
.
t2

(A.19)

This is the same form of the wave equation, so we have the same
conclusions: light is a wave of a magnetic field having a velocity
of (1/0 0 )1/2 . However, because of Faradays law (Maxwells third
equation), the electric field and the magnetic field associated with a light
wave are perpendicular to each other. A modern depiction of what we now
call electromagnetic waves is shown in Fig. A.1.

A.3 Concluding Remarks


Along with the theory of gravity, laws of motion, and atomic theory,
Maxwells equations were triumphs of classical science. Although

Figure A.1
light.

A modern depiction of the electromagnetic waves that we know as

88

Afterword

ultimately updated by Planck and Einsteins quantum theory of light,


Maxwells equations are still indispensable when dealing with everyday
phenomena involving electricity and magnetism and, yes, light. They help
us understand the natural universe betterafter all, isnt that what good
scientific models should do?

Bibliography
Baigrie, B., Electricity and Magnetism, Greenwood Press, Westport, CT
(2007).
Darrigol, O., Electrodynamics from Ampere to Einstein, Oxford University
Press, Oxford, UK (2000).
Halliday, D., R. Resnick, and J. Walker, Fundamentals of Physics, 6th
edition, John Wiley and Sons, New York (2001).
Hecht, E., Physics, Brooks-Cole Publishing Co., Pacific Grove, CA (1994).
Marsden, J. E. and A. J. Tromba, Vector Calculus 2nd edition, W. H.
Freeman and Company, New York (1981).
Reitz, J. R., F. J. Milford, and R. W. Christy, Foundations of
Electromagnetic Theory, Addison-Wesley Publishing Co., Reading,
MA (1979).
Schey, H. M., Div, Grad, Curl, and All That: An Informal Text on Vector
Calculus, 4th edition, W. W. Norton and Co., New York (2005).

89

Index
A
abscissa, 20
amber, 3
Ampre, 9, 15
Ampres circuital law, 78
Ampre, Andr Marie, 55, 75
AmpreMaxwell law, 81, 84, 85,
87
Arago, Franois, 11, 55

color theory, 18
Columbus, Christopher, 3
compasses, 3
Copernicus, 7
Coulombs law, 8, 40, 61
curl, 67, 68, 73, 76, 83
D
Davy, Humphrey, 9
de Coulomb, Charles-Augustin, 7
de Gama, Vasco, 3
del, 39, 53, 69
delta, 21, 22
derivative, 21, 38, 64
Descartes, Ren, 5
determinant, 68
diamagnetic, 49
diamagnetism, 14
dielectric, 79
displacement current, 80
divergence, 38, 39, 83
dot product, 24, 36, 73
du Fay, Charles, 5
dynamo, 13

B
battery, 9
Biot and Savart, 11
Biot, Jean-Baptiste, 55
BiotSavart law, 15
Browne, Thomas, 4
C
calculus, 19
Cambridge, 18
capacitors, 78
Cardano, Gerolamo, 3
Cavendish, Henry, 8
changing electric field, 80
charge density, 45
Charles du Fay, 5
Church of San Nazaro, 7
circulation, 66, 70
closed path, 61

E
Earth, 4
effluvium, 3, 5
electric eels, 2
91

92

electric field, 42, 80


electric motors, 74
electricity, 2, 4
electrons, 80
elektron, 3
F
Faraday effect, 13
Faradays law, 74, 84, 85, 87
Faraday, Michael, 9, 11, 48, 57,
70, 74
ferromagnetic, 49
flux, 35, 42
Franklin, Benjamin, 5
G
Galvani, Luigi, 8
Gauss law, 45
Gauss, Carl Friedrich, 45
generators, 74
Gilbert, William, 3
gradient, 84
Gray, Stephen, 4
Grove, William Robert, 11
H
Heaviside, Oliver, 83
Henry, Joseph, 57
I
induced electric field, 70
infinitesimal, 26
integral, 26
integral sign, 26
integrand, 26
integration, 26, 28
K
kinetic molecular theory of gases,
18

Index

Kings College London, 17


L
law of conservation of energy, 74
Leibnitz, Gottfried, 19, 26
Lenzs law, 74
Lenz, Henrich, 74
Leyden jar, 5, 9, 79
lightning rod, 7
line integral, 29, 33
lines of force, 48, 49
lodestone, 3, 47
luminiferous ether, 5
M
Magellan, 3
Magnesia, 3, 47
magnesium, 48
magnet, 47
magnetic field, 14, 47, 48, 55, 75,
76
magnetic flux, 71
magnetic monopole, 54
magnetite, 47
magnitude, 30
Marischal College, 17
Maxwells demon, 18
Maxwells equations of
electrodynamics, 19
Maxwells first equation of
electromagnetism, 45
Maxwells fourth equation of
electrodynamics, 81
Maxwells second equation of
electrodynamics, 53
Maxwells third equation of
electrodynamics, 74
Maxwell, James Clerk, 8, 17, 50,
78, 83

93

Index

MaxwellBoltzmann distribution,
18
N
Newton, Isaac, 4, 19, 26
normal vector, 34
O
Ohms law, 11
Ohm, Georg, 11
ordinate, 20
rsted, Hans Christian, 9, 55
other forms of Maxwells
equations, 83
P
paramagnetic, 49
paramagnetism, 14
partial derivative, 22
path independent, 61
path integral, 29
Peregrinus, Petrus, 3, 14
permeability, 56
permeability of a vacuum, 86
permittivity of free space, 42, 81
Poisson, Simeon-Denis, 15
pole, 3, 48

scalar, 39
Savart, Flix, 55
slope, 20
speed of light, 87
Sturgeon, William, 56
surface integral, 33
T
Thales of Miletos, 3
thermodynamics, 18
Thomson, William, 50
Thomson, William, Lord Kelvin,
13
total differential, 22
transformers, 74
U
unit vectors, 24, 32, 76
University of Aberdeen, 17
University of Leyden, 5

Q
quantum mechanics, 19

V
van Musschenbrk, Pieter, 5
vector, 48
vector function, 23
volt, 81
Volta, Alessandro, 5, 8
voltaic pile, 9
volume, 37
von Kleist, Ewald, 5

R
Reimann, Bernhard, 28
relativity, 19
right-hand rule, 55
Royal Society, 4

W
Water, 9
wave equation, 86, 87
Weber, William, 45
work, 59

S
Saturn, 18

Y
y intercept, 20

David W. Ball is a Professor in the Department of


Chemistry at Cleveland State University (CSU) in
Ohio. He received a Ph.D. in chemistry from Rice
University in 1987 and, after post-doctoral research
at Rice University and at Lawrence Berkeley
Laboratory in Berkeley, California, joined the
faculty at CSU in 1990, rising to the rank of
Professor in 2002. He has authored more than
200 publications, equally split between research
papers and works of a more educational bent,
including eight books currently in print. Dr. Balls
research interests include low-temperature infrared
spectroscopy and computational chemistry. His wife, Gail, is an engineer
who keeps him on his toes. His two sons, Stuart and Casey, also keep him
on his toes.
Professor Ball has a doppelgnger, David W. Ball, who writes historical
fiction and lives in the Rocky Mountains. David (the chemistry professor)
has read some of Davids (the real authors) books and has enjoyed them.
There is no word if the favor has been returned.

SPIE PRESS

Maxwells Equations of Electrodynamics: An Explanation


is a concise discussion of Maxwell's four equations of
electrodynamicsthe fundamental theory of electricity,
magnetism, and light. It guides readers step-by-step through
the vector calculus and development of each equation.
Pictures and diagrams illustrate what the equations mean in
basic terms. The book not only provides a fundamental
description of our universe but also explains how these
equations predict the fact that light is better described as
"electromagnetic radiation."

P.O. Box 10
Bellingham, WA 98227-0010
ISBN: 9780819494528
SPIE Vol. No.: PM232

You might also like