Anda di halaman 1dari 57

Field, Force, Energy and Momentum in Classical Electrodynamics

Masud Mansuripur
College of Optical Sciences
The University of Arizona, Tucson































J ames Clerk Maxwell
13 J une 1831 5 November 1879







To Annegret, Kaveh, and Tobias

Contents
Preface ...................................................................................................................................... i
Keywords................................................................................................................................... iii
Chapter 1: Scalar and Vector Fields
1.1. Introduction.........................................................................................................3
1.2. Space and time....................................................................................................3
1.3. Scalar and vector fields.......................................................................................5
1.4. Gradient of a scalar field.....................................................................................6
1.5. Integration of fields over time and/or space.......................................................7
1.6. Divergence of a vector field................................................................................9
1.7. Theorem of Gauss...............................................................................................10
1.8. Curl of a vector field...........................................................................................10
1.9. Theorem of Stokes..............................................................................................11
1.10. Longitudinal and transverse vector plane-waves..............................................13
General References............................................................................................14
Problems.............................................................................................................15
Chapter 2: Foundations of the Classical Maxwell-Lorentz Theory of Electrodynamics
2.1. Introduction.........................................................................................................22
2.2. Definition: Permittivity
o
of free-space.............................................................23
2.3. Definition: Permeability
o
of free space...........................................................23
2.4. Speed of light c and impedance of free space Z
o
................................................23
2.5. Sources of electromagnetic fields.......................................................................23
2.6. Electric field E and magnetic field H..................................................................25
2.7. Electric displacement D and magnetic induction B............................................25
2.8. Rules of the game................................................................................................26
2.9. Rule 1: Maxwells first equation.........................................................................26
2.10. Rule 2: Maxwells second equation..................................................................27
2.11. Continuity equation of charge and current........................................................30
2.12. Rule 3: Maxwells third equation.....................................................................32
2.13. Rule 4: Maxwells fourth equation...................................................................33
2.14. Macroscopic versus microscopic equations......................................................34
2.15. Bound charge and bound current associated with polarization
and magnetization.............................................................................................34
2.16. Magnetic bound charge and bound current.......................................................35
2.17. Maxwells boundary conditions........................................................................36
2.18. Rule 5: Energy in electromagnetic systems......................................................38
2.19. Rule 6: Momentum density of the electromagnetic field..................................41
2.20. The Einstein-box gedanken experiment............................................................43
2.21. The thought experiment of Balazs....................................................................44
2.22. Rule 7: Angular momentum density of the electromagnetic field....................45
2.23. Rule 8: Force density exerted by electromagnetic fields on material media....46
2.24. Conservation of linear momentum...................................................................47
2.25. Rule 9: Torque density exerted by electromagnetic fields on material media..47

2.26. Conservation of angular momentum.................................................................48
General References............................................................................................49
Problems.............................................................................................................50
Chapter 3: Mathematical Preliminaries
3.1. Introduction.........................................................................................................62
3.2. Elementary special functions..............................................................................62
3.3. The Fourier transform operator...........................................................................66
3.4. The Fourier theorem...........................................................................................67
3.5. Fourier transformation in higher dimensions......................................................68
3.6. Bessel functions and their properties..................................................................71
General References.............................................................................................75
Problems..............................................................................................................76
Chapter 4: Solving Maxwells Equations
4.1. Introduction.........................................................................................................82
4.2. Plane-wave solutions of Maxwells equations....................................................83
4.3. Electric field produced by a stationary point-charge (electrostatics)..................86
4.4. Electric field of a line-charge (electrostatics).....................................................87
4.5. Electric field of a uniformly-charged plate (electrostatics) ................................87
4.6. Magnetic field of a long, thin wire carrying a constant current
(magnetostatics) .................................................................................................88
4.7. Magnetic field of a hollow cylinder carrying a constant current
(magnetostatics) .................................................................................................89
4.8. Electric field produced by a point-dipole (electrostatics)...................................90
4.9. Fields radiated by an oscillating point-dipole (electrodynamics).......................91
4.10. Radiation by an oscillating current sheet (electrodynamics)............................94
4.11. Radiation by an oscillating line-current (electrodynamics)..............................96
4.12. Radiation by a hollow cylinder carrying an oscillating current
(electrodynamics)..............................................................................................98
General References............................................................................................102
Problems.............................................................................................................103
Chapter 5: Solving Maxwells Equations in Space-time: The Wave Equation
5.1. Introduction.........................................................................................................113
5.2. Scalar potential ( , ) t r as the solution of a 2
nd
-order partial
differential equation............................................................................................113
5.3. Vector potential ( , ) t A r as the solution of a 2
nd
-order partial
differential equation............................................................................................114
5.4. Meaning of the Laplacian operator acting on a vector field...............................115
5.5. Relating scalar and vector potentials to their sources in the
space-time domain...............................................................................................116
5.5.1. Example: Oscillating point-dipole.............................................................118
5.5.2. Example: Infinitely-long, thin, current-carrying wire radiating
cylindrical waves......................................................................119

5.5.3. Example: Infinite sheet of oscillating current radiating plane-waves........120
General References.............................................................................................124
Problems..............................................................................................................125
Chapter 6: The Lorentz Oscillator Model
6.1. Introduction.........................................................................................................140
6.2. Mass-and-spring model of an atomic dipole.......................................................140
6.3. Generalization to the case of multi-electron atoms and molecules.....................142
6.4. Drude model of the conduction electrons...........................................................142
6.4.1. Example.....................................................................................................143
6.5. The Clausius-Mossotti relation...........................................................................144
6.6. Dependence of the real and imaginary parts of C() on frequency...................145
6.7. Phase and group velocities..................................................................................147
6.7.1. Example 1..................................................................................................148
6.7.2. Example 2..................................................................................................149
6.7.3. Example 3..................................................................................................150
6.7.4. Example 4..................................................................................................151
6.8. Step-response and Impulse-response..................................................................152
6.9. The Kramers-Kronig relations............................................................................154
General References.............................................................................................156
Problems..............................................................................................................157
Chapter 7: Plane Electromagnetic Waves in Isotropic, Homogeneous, Linear Media
7.1. Introduction.........................................................................................................163
7.2. Complex vector algebra of the electromagnetic field.........................................164
7.3. Plane electromagnetic waves and their properties..............................................166
7.4. Plane-waves in isotropic, homogeneous, linear media.......................................167
7.5. Energy flux and the Poynting vector ..................................................................168
7.6. Reflection and transmission of plane-waves at a flat interface between
adjacent media....................................................................................................169
7.6.1. Case of TM or p-polarized incident plane-wave at a flat interface
located at z =0........................................................................................171
7.6.2. Case of TE or s-polarized incident plane-wave at a flat interface
located at z =0........................................................................................172
7.7. Fresnel reflection and transmission coefficients in several cases
of practical interest..............................................................................................172
7.7.1. Special Case 1: normal incidence..............................................................173
7.7.2. Special Case 2: Brewsters angle...............................................................173
7.7.3. Special Case 3: total internal reflection.....................................................174
7.8. Concluding remarks............................................................................................174
General References.............................................................................................175
Problems..............................................................................................................176


11.2. Mass-and-spring model of polarization exhibiting spatial dispersion .............247
11.3. Dispersion relations .........................................................................................248
11.4. Case of s-polarized incident plane-wave .........................................................249
11.5. Case of p-polarized incident plane-wave .........................................................250
11.6. Mechanical energy density, energy loss rate, and a mechanical Poynting
vector.................................................................................................................252
General References...........................................................................................254
Chapter 12: The Reciprocity Theorem
12.1. Introduction.......................................................................................................255
12.2. Electromagnetic field radiated by an oscillating electric dipole.......................257
12.3. Electromagnetic field radiated by an oscillating magnetic dipole....................259
12.4. Reciprocity in a system containing electrically-polarizable media .................259
12.5. Reciprocity in systems containing both electric and magnetic media .............262
12.6. Reciprocity in the presence of spatial dispersion .............................................263
12.7. Comparison with standard proofs of reciprocity..............................................264
12.8. Summary and Concluding Remarks ................................................................267
References.........................................................................................................269
Problems...........................................................................................................270
Solutions to Selected Problems ..............................................................................................271
Appendix A: Vector Identities.................................................................................................319
Appendix B: Vector Operations in Cartesian, Cylindrical, and Spherical Coordinates..........320
Appendix C: Useful Integrals and Identities...........................................................................321
Index.........................................................................................................................................323
Preface
This book grew out of a graduate-level course in electrodynamics that I have taught at the University
of Arizonas College of Optical Sciences over the past six years. A typical student enrolled in the course
is a first year graduate student in Optical Sciences, Electrical Engineering, or Physics, who has had some
prior exposure to electromagnetic theory. The level of mathematics required for this subject is not
particularly advanced; students are expected to be familiar with calculus, vector algebra, complex
numbers, ordinary differential equations, and elementary aspects of the Fourier transform theory. Most of
the mathematical tools and techniques needed for developing the theory of electrodynamics are in fact
interwoven with the course material in the form of a section here, a chapter there, or a few problems at the
end of each chapter. The student is thus motivated to learn the required mathematics in the relevant
physical context whenever the need arises.
The approach of this book to classical electrodynamics is rather unconventional. It begins with a
minimum set of postulates that are considered fundamental in the sense that they cannot be derived from
each other or from other laws of classical physics. The set of postulates, of course, must be self-
consistent, as well as consistent with the conservation laws and with special relativity. These postulates
are described in their most general form at the outset, with no apologies for their sudden appearance and
no attempt to motivate them, say, by tracing the historical path that led to their discovery. The laws of
nature are what they are; it may have taken man a long and tortuous path to their discovery, but once the
laws are known, one should simply accept them and try to understand their consequences.
In this context, an analogy with a board game such as chess is constructive. Before one sets out to
play the game, one must learn the configuration of the board, the identity of the pieces, and the governing
set of rules in their detailed and complete form. For most practical purposes, it is irrelevant how the game
has evolved over the years, how the rules may have changed, and whether or not there is any justification
for the rules. The important thing is to learn the rules and play the game. In the case of physics, of course,
the postulates are justified because their consequences agree with observations. This, however, is
something that one will appreciate later, as one begins to understand the subject and learns how to deduce
the logical consequences that flow from the basic principles. The task before the student, therefore, is to
master the nomenclature and learn the basic rules of electrodynamics, then try to deduce their
consequences.
In the presence of known sources of radiation (i.e., sources whose spatio-temporal distributions are
given a priori) we will use the method of plane-wave decomposition and superposition to derive general
expressions for electromagnetic fields and potentials. This will enable us to examine several idealized
situations in Chapters 4 and 5, thereby gaining insight into the nature of electromagnetic fields and
radiation. In the course of this analysis, I find it useful to move back and forth between the space-time
domain, where the fields and their sources reside, and the Fourier domain, which is home to various
plane-waves whose superposition reproduces the fields and the sources. The mathematical methods used
in the two domains may differ, but the final results pertaining to physical observables of any given system
are invariably the same.
Throughout the book I have striven to be brief yet precise. Whenever possible, I develop a general
formalism to tackle a given class of problems, then specialize the solution to examine specific problems
within that class. For example, in dealing with plane-wave propagation in isotropic, homogeneous, linear
media (Chapter 7), Maxwells equations are solved in a way that is applicable to transparent as well as
absorptive media, encompassing both propagating and evanescent waves while accommodating arbitrary
states of polarization (i.e., linear, circular, elliptical). Once the general solution is at hand, a few specific
examples show its application to problems such as propagation in transparent or absorptive media, total
internal reflection, incidence at Brewsters angle, etc. The student is thus equipped with the tools needed
to tackle problems within a broad class, without having to learn each specific case as an isolated instance.
The theory of electrodynamics is too broad, and its applications too diverse, to allow coverage in a
brief textbook such as this one. My goal, therefore, is not to be comprehensive, but rather to build a
foundation upon which one could base future learning and further investigations. Throughout the book,
i

fundamental notions are laid out in their most general form and described in sufficient detail to give the
reader a firm grasp of their content. Examples are then used to bring out important logical consequences
of these concepts and to showcase certain practical applications. I have opted for idealized examples with
exact analytical solutions, as such solutions can be relied upon to provide valid answers to physical
questions even when one or more parameters are pushed to extreme limits. At the same time, each
idealized example corresponds to some physical setting in a recognizable limit (e.g., large diameter,
vanishing thickness, point-particle, uniform charge distribution, etc.), so that, in principle at least, the
predictions made in the context of an idealized situation could be subjected to experimental verification.
The exercise problems at the end of each chapter elaborate the concepts developed in the chapter,
providing the student with the means to test his/her understanding of the subject as well as extending the
methods and ideas in new directions. One gains broad insight into electrodynamics and appreciates the
richness of its various applications by solving these problems and trying to understand the physical
meaning behind each solution. In the case of problems marked with an asterisk, detailed solutions are
provided at the end of the book.
In a one-semester course, I have been able to cover the first seven chapters, with selected sections
from the remaining chapters (e.g., Chapters 8 and 10) assigned for self study. The more advanced topics
such as solving Maxwells equations in cylindrical coordinates (Chapter 9), plane-wave propagation in
spatially-dispersive media (Chapter 11), and the reciprocity theorem (Chapter 12) are better left for a
second semester. If the course is extended to a second semester (or third quarter), it would be desirable to
supplement the present text with other standard topics such as diffraction, spatial and temporal coherence,
wave propagation in dispersive media, reflection and refraction in the presence of birefringence and
optical activity, time-reversal symmetry, the Ewald-Oseen extinction theorem, and the Lorentz
transformation of electromagnetic fields and sources between inertial frames.
I thank many students who, in the course of the past few years, have asked penetrating questions,
corrected my mistakes, and generally motivated me by their enthusiasm for the subject. I am grateful to a
number of colleagues and associates who have guided me along the path of learning and provided
answers to my numerous questions. Special thanks are thus due to Brian Anderson, Stephen Barnett,
J ean-Pierre Delville, Poul J essen, Miroslav Kolesik, Henri Lezec, Rodney Loudon, J erome Moloney,
Miles Padgett, Pavel Polynkin, Din Ping Tsai, J ohn Weiner, Ewan Wright, and Armis Zakharian. I spent
the Spring semester of 2010 as a visiting professor at the Physics Department of the National Taiwan
University in Taipei, where I wrote several chapters of this book while teaching a group of bright and
motivated students. I take this opportunity to thank Taiwans National Science Council for supporting my
sabbatical leave, and also express my gratitude to Professor Din Ping Tsai for being a warm and gracious
host. Last but not least, I am grateful to my wife, Annegret, without her loving care and patient support
this book would not have become a reality.

Tucson, J uly 2011
Masud Mansuripur
ii

Field, Force, Energy and Momentum in Classical Electrodynamics
Masud Mansuripur

Keywords for Chapter 1:
Scalar Field, Vector Field, Divergence, Gradient, Curl, Gausss Theorem, Stokess Theorem,
Longitudinal Field, Transverse Field.

Keywords for Chapter 2:
Maxwells Equations, Electromagnetic Field, Displacement Field, Magnetic Induction, Free
Charge, Free Current, Bound Charge, Bound Current, Continuity Equation, Impedance of Free
Space, Electromagnetic Energy, Field Momentum, Field Angular Momentum.

Keywords for Chapter 3:
Fourier Transform, Fourier Operator, Fourier Theorem, Diracs Delta-Function, Sifting Property
of Delta-Function, Bessel Functions.

Keywords for Chapter 4:
Plane-wave Solutions, Maxwells Equations, Electromagnetic Radiation, Scalar Potential, Vector
Potential, Lorenz Gauge, Point Charge, Point Dipole, Line Current, Current Sheet, Oscillating
Electric Dipole, Oscillating Magnetic Dipole, Oscillating Hollow Cylinder.

Keywords for Chapter 5:
Solutions of Maxwells Equations, Vector Potential, Scalar Potential, Lorenz Gauge, Laplacian
Operator, Oscillating Point Dipole, Cylindrical Wave, Current-Carrying Wire, Oscillating Sheet
of Current.

Keywords for Chapter 6:
Lorentz Oscillator, Mass-and Spring Model, Single-electron Lorentz Model, Multi-electron
Lorentz Model, Drude Model, Conduction Electrons, Clausius-Mossotti Relation, Phase
Velocity, Group Velocity, Dispersion, Kramers-Kronig Relations.

Keywords for Chapter 7:
Plane-wave, Plane Electromagnetic Waves, Linear Media, Isotropic Media, Homogeneous
Media, Energy Flux, Poynting Vector, Reflection Coefficient, Transmission Coefficient, Fresnel
Coefficients, p-polarization, s-polarization, Normal Incidence, Oblique Incidence, Brewsters
Angle, Total Internal Reflection.

Keywords for Chapter 8:
Plane Electromagnetic Waves, Multilayer Stack, Reflection Coefficient, Transmission
Coefficient, Parallel-Plate Slab, Optical Cavity, Cavity Resonator, Perfectly Matched Layer,
Finite Difference Time Domain Method.

iii

Keywords for Chapter 9:
Solution of Maxwells Equations, Cylindrical Coordinates, Linear Media, Isotropic Media,
Homogeneous Media, Circular Symmetry, Cylindrical Symmetry, Bessel Functions, Plane-wave
Superposition, Hankel Functions, Guided Modes, Surface Plasmon Polariton, Energy Flux,
Poynting Vector.

Keywords for Chapter 10:
Electromagnetic Momentum, Electromagnetic Angular Momentum, Force, Torque, Radiation
Pressure, Momentum of a Light Pulse, Transparent Slab, Semi-Transparent Slab, Brewsters
Angle Incidence, Spherical Glass Bead, Optical Vortex, Circular Polarization, Momentum
Conservation.

Keywords for Chapter 11:
Plane-wave Propagation, Plane-wave Solutions of Maxwells Equations, Linear Media,
Homogeneous Media, Isotropic Media, Spatial Dispersion, Mass-and-Spring Model, Dispersion
Relations, s-Polarized Incidence, p-Polarized Incidence, Mechanical Energy Density, Mechanical
Poynting Vector.

Keywords for Chapter 12:
Reciprocity, Classical Electrodynamics, Electromagnetic Field, Oscillating Electric Dipole,
Oscillating Magnetic Dipole, Electrically-Polarizable Media, Magnetic Media, Spatial
Dispersion, Standard Proofs of Reciprocity.
i v










The tendency of modern physics is to resolve the whole material universe into waves,
and nothing but waves. These waves are of two kinds: bottled-up waves, which we
call matter, and unbottled waves, which we call radiation or light. If annihilation of
matter occurs, the process is merely that of unbottling imprisoned wave-energy and
setting it free to travel through space. These concepts reduce the whole universe to a
world of light, potential or existent, so that the whole story of its creation can be told
with perfect accuracy and completeness in the six words: God said, Let there be light.
Sir J ames J eans (1877-1946)
v
Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 3-21 3

CHAPTER 1
Scalar and Vector Fields
The aim of exact science is to reduce the problems of nature to the determination of quantities by
operations with numbers.
James Clerk Maxwell (1831-1879)
Abstract. The concepts of scalar and vector fields, which are central to the theory of
electrodynamics, are introduced. These fields are generally defined in 3-dimensional Euclidean
space as complex-valued functions of the space-time coordinates (x, y, z, t ). Integration and
differentiation in time and space, leading to such operations as gradient, divergence, and curl,
and subsequently to theorems of Gauss and Stokes, are developed. The intuitive approach taken
here avoids mathematical formalism in favor of physical understanding. Throughout the
chapter, examples based on complex-valued scalar and vector plane-waves help to illustrate the
various mathematical operations. The end-of-chapter problems should help refresh the readers
memory of elementary mathematical tools needed in this as well as in subsequent chapters.
1.1. Introduction. This chapter introduces the concepts of scalar and vector fields in flat space-
time, using Lorentzian coordinate systems. The fields are generally complex-valued, which is
convenient for algebraic manipulations. The physical fields, of course, are always real-valued
and, therefore, will be represented by the real parts of the complex entities that are used here to
describe the field strengths and their variations throughout space and time. There are various
ways to integrate and also to differentiate the fields in space-time, e.g., time integration and time
differentiation; spatial differentiation in the form of gradient, divergence, and curl operations;
and spatial line-, surface-, and volume-integrations. We will describe these operations in some
detail, prove the theorems of Gauss and Stokes, which pertain to vector fields and their spatial
derivatives and integrals, and provide examples in each case using a particularly useful field, the
plane-wave.
1.2. Space and time. Electromagnetic phenomena take place in space and time. An event in
space-time occurs at a point r in space at time t. We shall assume that all events occur in flat
space-time (i.e., in the absence of gravity and gravitational fields), and that all events are
observed by an inertial observer, namely, one whose motion is unaccelerated (relative to distant
stars). In addition, the observer uses a Lorentzian reference frame for all his observations. Thus
every event will be specified in an orthonormal coordinate system such as (r, t) = (x, y, z, t)
Cartesian or (, , z, t) cylindrical or (, , , t) spherical; see Fig. 1.
One may imagine that each point r in space has its own ideal (e.g., atomic) clock, which
runs at a fixed rate, and that the clocks at all locations r are coordinated and synchronized by the
inertial observer in whose reference frame all observations are made. Think of a clock located at
r as a compact, tightly-wound spiral curve, with a pointer moving inexorably
along the spiral in the same (forward) direction. Each point r has its own spiral
clock. When an event takes place at r, the pointer will be somewhere along the
spiral curve; the location of the pointer, as measured by the length of the spiral
from its starting point at t = 0, is the time associated with the event.
The reason we are using a tightly-wound spiral curve to represent the time axis, is that we
have run out of easily imaginable dimensions. In a 2-dimensional (2D) space, where events are
confined, for example, to a planar surface embedded within a 3D Euclidean space, one does not
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
4 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur

need this imaginary spiral, as the time-axis can be imagined to extend into the third spatial
dimension. Of course, one can always roll up the time axis, turn it into a compact spiral, then
place it at each point r = (x, y) in the 2D space, just as was done above in the case of 3D space.
















Fig. 1. Commonly used coordinate systems for specifying the location of a point r
o
in 3D
Euclidean space. From left to right: Cartesian, cylindrical, and spherical coordinate systems. In
Cartesian coordinates, the position is specified by the distances (x
o
, y
o
, z
o
) from the origin along the
three axes. In cylindrical coordinates, the radial distance

o
from the z-axis, the azimuthal angle

o
, measured from the x-axis in the counter-clockwise direction, and the vertical distance z
o
from
the xy-plane are used to specify the position of r
o
. In spherical coordinates, r
o
is specified by its
distance from the origin

o
, its azimuth
o
, and its polar angle
o
, measured from the positive
toward the negative z-axis.
The world-line of a point-particle moving through space-time is constructed from the
sequence of locations r visited by the particle at local time t. The continuous sequence of events
(r, t) visited by the particle thus defines its world-line; see Fig. 2. When one knows the world-line
of a particle, one knows everything about the movement of that particle through space-time; in
other words, at any desired time t, one knows the particles position, velocity, acceleration, etc.


















Fig. 2. World lines of particles residing in the 2D space of an xy-plane. The vertical axis is used to
specify the time at which the particle arrives at the corresponding point in the xy-plane. For a
stationary particle sitting motionless at (x
o
, y
o
) the world line is a straight vertical line, parallel to
the time-axis. For an arbitrarily moving particle, going from (x
o
, y
o
) at t
o
to (x
1
, y
1
) at t
1
, the world
line bends to the right, left, front, and back, but always with an upward orientation along the time
axis. The particle going around a circle in the xy-plane at constant angular velocity, has a world
line that resembles an upward-stretched spiral.
r
o


x
y
z

x
y
z
x
o

z
o

y
o

r
o

o

z
o

o


x
y
z
r
o

o

Stationary
particle

x
y
time
x
y
x
o

y
o

x
y
World line
Moving
particle
( x
o
, y
o
)
( x
1
, y
1
)

World line
t
o
t
1
time time
Rotating
particle
World line
Scalar and Vector Fields Field, Force, Energy and Momentum in Classical Electrodynamics 5

1.3. Scalar and vector fields. When a number, generally complex-valued, is associated with
each point or event (r, t) in space-time, we will have a complex function f (r, t), generally referred
to as a scalar field. The temperature T, specified at each point r in a room and at each instant of
time t is a good example of a real-valued scalar field T(r, t).
Similarly, when a vector V is associated with each point (r, t) in space-time, we have a
vector field V(r, t). In 3D space, vectors are generally specified by their 3 components along
specific directions. Thus, in a Cartesian coordinate system we have, at each point (r, t), three
numbers (V
x
, V
y
, V
z
) that specify the vector field. Similarly, in a cylindrical coordinate system, the
vector field is identified by (V

, V

, V
z
), while in spherical coordinates the components of the field
are (V

, V

, V

). Note that there are no good reasons to restrict the components of vectors to being
real-valued numbers. Thus, in general, a vector field assigns to each point (r, t) a complex-valued
vector, i.e., a vector whose three components in Euclidean space are complex-valued. One way
to visualize a complex-valued vector field is to imagine a pair of ordinary, real-valued vectors
(V , V) attached to each point in space, while the magnitudes and directions of both V and V
at each and every location r change arbitrarily with time. The complex-valued vector field is then
described by V(r, t) = V (r, t) +i V(r, t), where, in Cartesian coordinates,
V
x
(r, t) = V
x
(r, t) +i V
x
(r, t), (1a)
V
y
(r, t) = V
y
(r, t) +i V
y
(r, t), (1b)
V
z
(r, t) = V
z
(r, t) +i V
z
(r, t). (1c)
Similar expressions may be written for the components of a complex vector field in other
coordinate systems as well.
Of course, what distinguishes a vector field (real or complex) from a mere collection of
three scalar fields is vector algebra, namely, the rules of addition, subtraction, dot-multiplication,
and cross-multiplication of vectors. For concreteness, we shall describe these rules in Cartesian
coordinates only, although the same applies in any orthonormal coordinate system as well. For
addition and subtraction of vector fields we have
V
1
(r, t) V
2
(r, t) = [V
x1
(r, t) V
x2
(r, t)]x
^
+ [V
y1
(r, t) V
y2
(r, t)] y
^
+ [V
z1
(r, t) V
z2
(r, t)]z
^
. (2)
Note that the corresponding components of the two fields are simply added together or
subtracted from each other. The addition and subtraction rules for complex numbers being well
known, the above rule for addition and subtraction of vector fields clearly applies to real as well
as complex vector fields.
For dot- and cross-multiplication of two vector fields such as V
1
(r, t) and V
2
(r, t), one must
express each field as the sum of its three components, V
x
x
^
+V
y
y
^
+V
z
z
^
, then proceed to multiply
all the components of V
1
into all the components of V
2
using the following rules:
Dot: x
^
x
^
= y
^
y
^
= z
^
z
^
= 1; x
^
y
^
= y
^
x
^
= x
^
z
^
= z
^
x
^
= y
^
z
^
= z
^
y
^
= 0. (3)
Cross: x
^
x
^
= y
^
y
^
= z
^
z
^
= 0; x
^
y
^
= y
^
x
^
= zz
^
; y
^
z
^
= z
^
y
^
= x
^
; z
^
x
^
= x
^
z
^
= y
^
. (4)
Once again, the multiplication rule for complex numbers being well known, it makes no
difference whether the components of the vector fields being multiplied together are real- or
complex-valued.
A good example of a real-valued vector field is the wind velocity within a wind tunnel,
described as a function of time t for each point r in 3D space. To each point r is thus assigned a
22 Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 22-61
CHAPTER 2
Foundations of the Classical Maxwell-Lorentz Theory of Electrodynamics
" From a long view of the history of mankind, seen from, say, ten thousand years from now, there can
be little doubt that the most significant event of the 19
th
century will be judged as Maxwell's
discovery of the laws of electrodynamics. The American Civil War will pale into provincial
insignificance in comparison with this important scientific event of the same decade."
Richard P. Feynman (1918-1988)
Then came H. A. Lorentz's decisive simplification of the theory. He based his investigations with
unfaltering consistency upon the following hypotheses: The seat of the electromagnetic field is the
empty space. In it there are only one electric and one magnetic field vector. This field is generated by
atomistic electric charges upon which the field in turn exerts ponderomotive forces. The only
connection between the electromagnetic field and ponderable matter arises from the fact that
elementary electric charges are rigidly attached to atomistic particles of matter. For the latter
Newton's law of motion holds. Upon this simplified foundation Lorentz based a complete theory of
all electromagnetic phenomena known at the time, including those of the electrodynamics of moving
bodies. It is a work of such consistency, lucidity, and beauty as has only rarely been attained in an
empirical science.
Albert Einstein (1879-1955)
Abstract. The sources of electromagnetic fields are electric charge, electric current, polarization
and magnetization. The relationships among the fields and their sources, all of which
represented by functions of space and time, are described by Maxwells macroscopic equations.
The fields carry energy, whose rate-of-flow at each point in space at any instant of time is given
by the Poynting vector. At any location where one or more fields and one or more sources
reside simultaneously, there could occur an exchange of energy between the fields and the
sources. The time-rates of such exchanges are uniquely specified by the Poynting theorem,
which is a direct consequence of Maxwells macroscopic equations in conjunction with the
definition of the Poynting vector. Electromagnetic fields also carry momentum and angular
momentum, whose densities at all points in space-time are simple functions of the local
Poynting vector. A generalized version of the Lorentz law of force dictates the time-rate of
exchange of momentum between the fields and the sources in regions of space-time where they
overlap. There also exists a simple expression for the torque exerted by the fields on the
sources, which defines the time-rate of exchange of angular momentum between them. This
chapter is devoted to a precise and detailed description of the relations among the fields and
their sources, as well as their interactions involving electromagnetic force, torque, energy,
momentum, and angular momentum.
2.1. Introduction. Throughout this chapter we shall treat the classical theory of electrodynamics
as a game of chess. The board on which the game unfolds is the three-dimensional Euclidean
space; in other words, we assume a flat space-time, in which no gravitational deformations of
space-time geometry are allowed. We choose an inertial (i.e., unaccelerated) observer, and use a
Lorentzian reference frame to assign coordinates to each and every point in space-time.
Locations on our cosmic chessboard are thus uniquely identified (within the given Lorentzian
frame of reference) by their four-dimensional space-time coordinates (r, t). We shall identify the
pieces that reside within the 3D space (i.e., the chessboard) and move around through time.
Also specified will be the rules of the game, according to which the pieces interact with each
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
Foundations of the Classical Maxwell-Lorentz Field, Force, Energy and Momentum in Classical Electrodynamics 23

other and evolve in space and time. Throughout the chapter, the system of units will be MKSA
(meter, kilogram, second, ampere).

2.2. Definition: Permittivity
o
of free-space. The permittivity of free-space (i.e., vacuum,
empty space) is
o
= 8.854

10
12
farad/meter. In what follows we will clarify the meaning of
permittivity as well as the relationship between farad and the basic units of the MKSA system.
2.3. Definition: Permeability
o
of free space. The permeability of free space is
o
= 4 10
7

henry/meter. In what follows we will clarify the meaning of permeability as well as the
relationship between henry and the basic units of the MKSA system.
2.4. Speed of light c and impedance of free space Z
o
: The speed of light in vacuum, c, can be
derived from Maxwells equations. The exact relation between c and the permittivity and
permeability of free space will be seen to be c = 1/
o

o
. Given that, in principle, c is precisely
measurable (c = 2.99792458

10
8
m/s), and that the exact value of
o
is 4 10
7
henry/meter,
it is common practice to express the precise value of
o
as 1/(
o
c
2
). Equating the units on both
sides of the equation yields: farad henry = sec
2
.
Also, as will be seen later, the impedance of free space is defined as Z
o
=
o
/
o
377ohm.
(The symbol for ohm, the unit of electrical resistance, is .) Equating the units on the two sides
of the equation yields: henry/farad =
2
. Combining this with the previously obtained relation
between farad and henry, we find: farad = sec/ and henry = sec.
The special combinations of
o
and
o
in the expressions for c and Z
o
appear quite naturally
in electrodynamics equations. Therefore, whenever possible, we will use c and Z
o
to simplify
expressions that contain various combinations of
o
and
o
.
2.5. Sources of electromagnetic fields: There exist four material sources in electromagnetic
(EM) systems:

free
(r, t), J
free
(r, t), P(r, t), and M(r, t). These are continuous and differentiable
functions of space r and time t. The density of free charge,

free
, is a scalar; its MKSA units are
coulomb/m
3
. (Since current I is the time-rate of flow of charge Q, namely, I = dQ/dt, the units of
charge are: coulomb = ampere sec.) The other three sources, namely, density of free current J
free

(units = ampere/m
2
), polarization P (units = coulomb/m
2
), and magnetization M (units =
henry ampere/m
2
= weber/m
2
), are vector functions of r and t. Loosely speaking, these four
sources furnish some of the pieces of the aforementioned chess set: Charge, Current,
Polarization, and Magnetization reside in 3D space (and change in time) in ways that are roughly
similar to the way in which pawns, bishops and knights occupy positions on the chessboard and
move around. The analogy is not perfect, of course, and one should not push it too far. For
instance, the various sources of EM fields can overlap in the same region of space; their
magnitudes at each point in space can vary continuously with time; they produce the EM fields,
but also are influenced by these fields; the sources can convert from one form to another, for
example, current can contribute to charge density and vice versa; etc.
Digression: The current density J could arise from the motion of

. So, for example, J(r, t) =

(r, t)V(r, t), where V(r, t) is the velocity of charge at the point (r, t) in space-time. (Check the
consistency of the units of J with those of the product

V.) However, there could also exist J


without a net

. For example, when, in a typical copper wire, the positive charges (copper ions)
24 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur

are stationary, while the negative charges (conduction electrons) move with some velocity V
along the length of the wire (under the influence of an electric field), assuming equal densities
for the two types of charge, we have a situation in which the net charge density

is zero, yet
there exists a non-zero current density J.
Charge and current densities are intimately related via the continuity equation
J
free
(r, t) +

free
(r, t)/t = 0. Thus any net current flowing in or out of a given volume must
change the total charge content of that volume.














Fig. 1. (a) An electric dipole consists of a pair of equal and opposite charges, q, separated by a
small distance d. The dipole moment is defined as p=qd, with d always pointing from the negative
to the positive charge. (b) The total strength of the dipole moments p
n
(t) within a small volume V
surrounding a given point r, when normalized by V yields the polarization P(r, t) of the material
medium at location r and time t.
P(r, t) is the density of atomic electric dipole moments. Each atomic dipole is characterized
by its equal and opposite charges q, and the small separation d between these charges. By
definition, the direction of d is from negative to positive charge. The dipole moment is defined as
p=qd (units of p = coulomb meter). To determine the polarization P(r, t), take a small volume
V centered on r, then find the vector sum of all the individual dipole moments p
n
within that
volume. Normalizing the total dipole moment within V by the volume V yields the local
polarization P(r, t); units of P = coulomb/ m
2
.















Fig. 2. (a, b) Magnetic dipoles are produced by spinning elementary particles, by the orbital
motion of charged particles, and also by various combinations of spin and orbital magnetic
moments within atoms and molecules. In the case of a small, flat loop of area A carrying an
electric current I, the magnitude and direction of the dipole moment are given by m=
o
IA z
^
, with
z
^
being the surface normal in the direction determined by the right-hand rule in conjunction with
the sense of circulation of the current I. (c) The total strength of the dipole moments m
n
(t) within a
small volume V surrounding a given point r, when normalized by V yields the magnetization
M(r, t) of the material medium at location r and time t.

+
p=qd
q
-q
Small volume V
(centered at r)
P(r, t) = (1/V)

p
n
(t)
p
1
p
2
p
3
p
n
d
Electric dipole moment
(a)
(b)
m=
o
IAz
^
Small volume V
(centered at r)
M(r, t) = (1/V)

m
n
(t)
m
1
m
2
m
3 m
n
Orbital magnetic
moment
Spin magnetic
moment
m
(a) (b) (c)
62 Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 62-81
CHAPTER 3
Mathematical Preliminaries
" the enormous usefulness of mathematics in the natural sciences is something bordering on the
mysterious and that there is no rational explanation for it."
Eugene Wigner (1902-1995), in "The Unreasonable Effectiveness of Mathematics in the Natural Sciences."
1

Abstract. In preparation for a Fourier analysis of Maxwells equations in the following chapter,
we describe here the mathematics of Fourier transformation, exploring certain properties of the
forward and reverse Fourier operators. Several special functions are also discussed notable
among them, Diracs delta-function and various Bessel functions which appear frequently in
Fourier analysis and elsewhere. Simple charge- and current-density distributions serve as
exemplary electromagnetic systems that can be readily transformed into the Fourier domain.
3.1. Introduction. This chapter provides a brief overview of the mathematical tools and
techniques needed for the analysis of Maxwells equations by means of Fourier transformation.
We begin by introducing a few elementary functions that are specially useful in the context of
Fourier transform theory, explore their properties, and proceed to rely on them when describing
the properties of the Fourier operator. Another class of special functions, which appear
frequently in Fourier transform theory and elsewhere, are Bessel functions discussed at the end
of the chapter.
The forward and reverse Fourier integrals are initially defined and analyzed for complex-
valued functions of a single real-valued variable; these will be referred to as one-dimensional
(1D) Fourier transforms. Subsequently, we generalize the concept to higher-dimensional spaces,
where complex functions of two or more real variables are transformed back and forth between a
multi-dimensional space and its corresponding Fourier domain. With regard to the solution of
Maxwells equations, which generally reside in Lorentzian space-time, the functions of interest
are usually four-dimensional (4D), as they depend on the space-time coordinates (r, t) = (x, y, z, t).
The corresponding Fourier domain in this case will also be a 4D space whose coordinates, often
referred to as the spatio-temporal frequencies, are denoted by (k, ) = (k
x
, k
y
, k
z
, ).
3.2. Elementary special functions. Certain special functions play an important role in the theory
of Fourier transforms. In this section we provide a brief review of the various properties of these
functions, and demonstrate their usefulness in problems involving Fourier transforms.
a) The unit-step function Step(x) is equal to 0.0 when x <0 and equal to 1.0 when x >0. The
value of the function at x = 0 could be assigned arbitrarily, as it does not affect the properties of
the function. For the sake of completeness, however, we define Step(x) = when x = 0. A plot of
this function appears in Fig. 1(a).
b) The unit rectangular pulse function Rect(x) is equal to 1.0 when |x| <, and equal to 0.0 when
|x| >. The value of the function at x = could be assigned arbitrarily, as it does not affect the
properties of the function. For the sake of completeness, however, we define Rect(x) = when
x = . A plot of this function appears in Fig. 1(b). Note that the area under Rect(x) is unity.
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
Mathematical Preliminaries Field, Force, Energy and Momentum in Classical Electrodynamics 63
c) The unit triangular pulse function Tri (x) is equal to 1|x| when |x| <1, and equal to 0.0 when
|x| >1, as shown in Fig. 1(c). Like the rectangular pulse, the area under the triangular pulse is
equal to 1.0.
d) The sinc function sinc(x) is defined as sin(x)/(x) over the entire x-axis. A plot of this
function appears in Fig. 1(d). Note that the value of the function at x =0 is 1.0. Also, the area
under the function can be shown to be unity, that is,

sinc(x)dx =1.0.
e) The Dirac delta-function (x) does not have a simple definition, and cannot be easily
visualized in a unique and unambiguous way. A good way to describe it would be as a very tall
and very narrow function of x, centered at x =0, symmetric around this central point, and with an
area equal to 1.0, that is,

(x)dx =1.0. Thus


1
Rect(x/), where is a small, real-valued,
positive constant would be the simplest representation of (x); see Fig. 2(a). Similarly,

1
Tri(x/) approaches a delta-function in the limit when 0; see Fig. 2(b). For sufficiently
small , there is essentially no difference between
1
Rect(x/) and
1
Tri(x/), except for the
latter function being continuous and readily differentiable which would be useful if one were
interested in the first derivative (x) of (x).






















Fig. 1. Plots of several elementary functions. (a) Unit-step function. (b) Unit rectangular pulse. (c)
Unit triangular pulse. (d) The sinc function.
Another embodiment of Diracs delta-function is
1
sinc(x/) in the limit when 0. The
fact that sinc(x) has an infinite number of oscillations is of no consequence, so long as the chosen
value of is small enough to cram a large number of these oscillations into a small
neighborhood of the origin (x =0), thus ensuring that the area under the function in that
neighborhood is as close to unity as is desired.
The various properties of (x) can be easily understood with the aid of the above definitions
and analogies. For example, the fact that (x) = (x) is a direct consequence of the requirement
of symmetry around x =0. Or, (2x) = (x) is clearly true given that, for sufficiently small ,
the function
1
Rect(2x/), is even, tall, narrow, and has an area equal to , as can be seen in
Fig. 2(c). By the same token, ( x) =(1/| | ) (x) for any real-valued and ( x +) =
(1/||) [x+( / )]. The latter delta-function is centered at x= /, having an area equal to 1/|| .
x
Step(x)


x
Rect(x)
1.0

(a) (b)
(c)
Tri(x)

(d)
sinc(x)
x
1.0
1.0
1.0
1.0
1.0
1.0
2.0
3.0
x
64 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur

Figure 2(d) shows that a unit-step function, when smoothed-out over the interval (, )
and differentiated with respect to x, yields the delta-function (x) in the limit when 0.
The most important property of (x), which is always true irrespective of the functional
form used to visualize the delta-function, is its sifting property. If an arbitrary function f (x)
happens to be continuous at x =0, then the sifting property of (x) is stated as follows:

f (x) (x)dx = f (0). (1a)


More generally, when f (x) is discontinuous at x =0, we will have

f (x) (x)dx = [ f (0
+
) +f (0

)]. (1b)






























Fig. 2. Visualizing Diracs delta-function (a) as a tall, narrow rectangular pulse, and (b) as a tall,
narrow triangular pulse. (c) Representing (2x) as
1
Rect (2x/) for sufficiently small . (d) A
unit-step function whose transition from 0 to 1 has been softened and extended over the small
interval (, ) yields the rectangular-pulse approximation to a delta-function when
differentiated with respect to x.
The sifting property can be explained by super-imposing the graph of f (x) on any one of the
visualizations of (x), then multiplying the two functions together and integrating over any
region of the x-axis that contains a small neighborhood of the origin (x =0). It should also be
obvious that the sifting property remains valid when (x) is shifted along the x-axis by an
arbitrary amount. For instance, if (x) is shifted to x =x
0
, at which point f (x) happens to be
continuous, we will have

f (x) (x x
0
)dx = f (x
0
). (1c)
x

1
Rect(x/)
1/
(a)
(d)
Step

(x)
/2
1.0
x
/2
x

1
Tri(x/)
1/
(b)

1
Rect(2x/)
1/
(c)
/4 /4
/2 /2
x
d
1/
/2 /2
Step

(x)
dx
82 Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 82-112
CHAPTER 4
Solving Maxwells Equations
Lorentz proclaimed the very radical thesis which had never before been asserted with such
definiteness: The ether is at rest in absolute space. In principle this identifies the ether with absolute
space. Absolute space is no vacuum, but something with definite properties whose state is described
with the help of two directed quantities, the electrical field E and the magnetic field H, and, as such
is called the ether.
Max Born (1882-1970)
I cannot but regard the ether, which can be the seat of an electromagnetic field with its energy and
its vibrations, as endowed with a certain degree of substantiality, however different it may be from
all ordinary matter.
Hendrik Antoon Lorentz (1853-1928)
Abstract. We solve Maxwells macroscopic equations under the assumption that the sources of
the electromagnetic fields are fully specified throughout space and time. Charge, current,
polarization, and magnetization are thus assumed to have predetermined distributions as
functions of the space-time coordinates (r, t). In this type of analysis, any action by the fields on
the sources will be irrelevant, in the same way that the action on the sources by any other force
be it mechanical, chemical, nuclear, or gravitational need not be taken into consideration. It
is true, of course, that one or more of the above forces could be responsible for the presumed
behavior of the sources. However, insofar as the fields are concerned, since the spatio-temporal
profiles of the sources are already specified, knowledge of the forces would not provide any
additional information. In this chapter, we use Fourier transformation to express each source as
a superposition of plane-waves. Maxwells equations then associate each plane-wave with other
plane-waves representing the electromagnetic fields. Inverse Fourier transformation then
enables us to express the electric and magnetic fields as functions of the space-time coordinates.
4.1. Introduction. Electromagnetic (EM) fields originate from the sources
free
(r, t), J
free
(r, t),
P(r, t), and M(r, t). When these sources are fully specified throughout space and time, the
resulting EM fields can be uniquely and unambiguously derived from Maxwells macroscopic
equations.
A simple yet powerful method of solving Maxwells equations under such circumstances
involves shuttling back and forth between the space-time domain (r, t) and the Fourier domain
(k, ). Since the sources are specified everywhere in space for all time, they can be Fourier
transformed and, therefore, expressed as superpositions of plane-waves. Through Maxwells
equations, these plane-waves may be related to other plane-waves that describe the scalar and
vector potentials (r, t), A(r, t), and also the EM fields E(r, t), D(r, t), H(r, t), and B(r, t). Once
the amplitudes of the plane-waves representing the fields and/or the potentials for all values of
(k, ) are determined, the linearity of Maxwells equations allows one to compute the exact
distributions of the fields and/or the potentials by superposing these plane-waves via inverse
Fourier transformation.
In this chapter we employ the mathematical tools and techniques developed in Chapter 3 to
calculate the EM field distributions for several systems of general as well as practical interest.
The Fourier domain relations that connect the fields and the potentials to their sources are
derived in the next section. The utility of these relations are subsequently demonstrated through
several examples.
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
Solving Maxwells Equations Field, Force, Energy and Momentum in Classical Electrodynamics 83

4.2. Plane-wave solutions of Maxwells equations. The scalar and vector potential functions in
the form of plane-waves in vacuum are given by

o
( , ) exp[i( )], t t = r k r (1a)

o
( , ) exp[i( )]. t t = A r A k r (1b)
In general, k and A
o
are complex vectors,
o
is a complex scalar, and is a complex-valued
constant.
One can always associate the curl of the vector potential with the B-field, for the simple
reason that the divergence of the curl of any vector field is identically zero, which, in the present
case, means that Maxwells 4
th
equation, ( , ) 0, t = B r is automatically satisfied. We thus write

o
( , ) ( , ) i exp[i( )]. t t t = = B r A r k A k r (2)
Clearly, the complex B-field amplitude is
o o
i , = B k A and the divergence of ( , ) t B r
vanishes everywhere because
2
o o
i ( ) ( ) 0. = = k k A k k A It might be helpful to think
momentarily of k and A
o
as real-valued vectors, to realize that B
o
is related only to the transverse
component
o
A of A
o
, namely, the projection of A
o
in the plane perpendicular to the k-vector.
The longitudinal component
o
A of A
o
is in no way constrained by our association of the B-field
with the vector potential in accordance with Eq. (2); therefore,
o
A remains free and available for
later adjustments. As will be seen below, this degree of freedom associated with the longitudinal
component of A
o
is intimately related to the so-called freedom to choose the gauge.
Having eliminated Maxwells 4
th
equation, we now turn to his 3
rd
equation and rewrite it as
follows:
( , ) ( , ) / [ ( , )]/ t t t t t = = E r B r A r ( , ) ( , ) / 0. [ ] t t t + = E r A r (3)
The above equation informs us that the vector field / t + E A is curl-free. It must,
therefore, be equal to the gradient of some scalar field, because the gradient of any scalar field is
curl-free as well. For instance, the gradient of (r, t) given by Eq. (1a) is
o
i exp[i( )], t k k r
whose curl
2
o
i exp[i( )] t k k k r is identically zero due to the fact that kk =0. By
convention, the scalar potential of electrodynamics is defined to be the field whose gradient is
( / ). t + E A We thus have
( , ) ( , ) ( , )/ . t t t t = E r r A r (4)
For the plane-wave solutions of Maxwells equations, the above identity yields

o o o
( , ) exp[i( )] ( i i ) exp[i( )]. t t t = = + E r E k r k A k r (5)
Equations (2) and (5) thus relate the electromagnetic E- and B-fields to the scalar and vector
potentials of Eq. (1) in such a way as to automatically satisfy Maxwells 3
rd
and 4
th
equations.
Instead of having to find the six components of E
o
and B
o
, one only needs to determine the scalar
entity
o
and the three components of A
o
. It should also be borne in mind that the longitudinal
component
o
A of A
o
is as yet unconstrained.
84 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur

Next, we turn to Maxwells 1
st
equation, assume that the contributions of free charge density
and polarization are combined in the total bound charge density
total
( , ), t

r and that the total


charge density present in the entire space-time is given by
total o
( , ) exp[i( )]. t t

= r k r With
the aid of Eq. (5) we obtain

o total o o o o o o o
( , ) ( , ) i ( ) . t t

= = = E r r k E k k k A (6)
Equation (6) relates the scalar- and vector-potential amplitudes
o
and A
o
to the charge-
density amplitude
o
. Taking advantage of the fact that
o o
k A =

k A is as yet unconstrained, we
simplify Eq. (6) by setting k A
o
equal to
o
/c
2
; the reason for this choice becomes clear
shortly. The above choice of the longitudinal component of the vector potential, generally
referred to as working in the Lorenz gauge, is equivalent to forcing A(r, t) and (r, t) into the
following relationship:

2
( , ) (1/ ) ( , )/ 0. t c t t + = A r r (7)
With the Lorenz gauge thus set, Eq. (6) becomes an equation for
o
in terms of
o
, which
may readily be solved as follows:

o
o 2 2
o
( , )
.
( , )
( / ) [ ] k c

k
k (8)
In the above equation, the fact that
o
and
o
depend on the plane-waves k-vector as well as
its frequency is explicitly emphasized. We also have written k k as k
2
to highlight its scalar
nature. In general, of course, k =k+i k is a complex vector, resulting in k
2
=(k
2
k
2
) +2i k k
being a complex scalar. Equation (8) clearly indicates that, in the Lorenz gauge, the scalar
potential is intimately related to the total charge-density distribution
total
( , ), t

r but not in any


explicit way to the current density distribution
total
( , ), t J r which is the only other source of
electromagnetic fields and radiation.
The remaining Maxwell equation,
o total o o
( , ) ( , ) ( , )/ , t t t t = + B r J r E r may also be
written in terms of the scalar and vector potentials. Writing
total o
( , ) exp [i( )], t t = J r J k r and
substituting for ( , ) t B r from Eq. (2) and for ( , ) t E r from Eq. (5), we arrive at

2
o o o o o o o
i ( ) i ( i i ) = + k k A J k A

2
o o o o o o o o o o
( ) ( ) + k k A k A k = J k A

2 2
2
o o o o o
( / ) ( / ) . [ ] k c c + A = k A k J (9)
The vector identity ( ) ( ) ( ) = A B C A C B A B C has been used in the above derivation.
Using the fact that k A
o
=
o
/c
2
(Lorenz gauge), Eq. (9) is simplified, yielding the relation
between the vector potential and the current density as follows:

o o
o
2 2
( , )
.
( , )
( / ) k c

J k
A k (10)
Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 113-139 113
CHAPTER 5
Solving Maxwells Equations in Space-time: The Wave Equation
A real field is then a set of numbers we specify in such a way that what happens at a point
depends only on the numbers at that point. We do not need to know any more about whats going on
at other places. It is in this sense that we will discuss whether the vector potential is a real field.
Richard P. Feynman (1964), arguing in favor of according vector potential the
status of physical reality on the basis of the Aharonov-Bohm effect.
1

For the rest of my life I will reflect on what light is.
Albert Einstein (1917), as quoted in S. Perkowitz, Empire of Light.
Abstract. The problem addressed in the present chapter is the same problem as discussed in the
preceding one, namely, the determination of fields for given distributions of charge, current,
polarization and magnetization. Here, however, we will derive expressions for the scalar and
vector potentials as functions of space and time coordinates for the given source distributions,
which are also specified in space-time. Once the potentials are obtained, the electromagnetic
fields will be calculated by straightforward differentiation. The integrals will look very different
from those encountered in Chapter 4, but the final results will be exactly the same.
5.1. Introduction. The first and third of Maxwells equations, namely,
free
( , ) t = D r and
( , ) ( , )/ , t t t = E r B r become equations that relate the E-field to its sources,
free
( , ) t r and
bound
( , ) ( , ), t t = r P r whenever the B-field happens to be time-independent, that is,
( , )/ 0. t t = B r The first and third equations thus form a complete basis for electrostatics in the
absence of time-varying magnetic fields. Similarly, the second and fourth equations,
free
( , ) ( , ) ( , )/ , t t t t = + H r J r D r and ( , ) 0, t = B r form a complete basis for magnetostatics
whenever the displacement field happens to be time-independent, i.e., ( , )/ 0. t t = D r In the
absence of time-variation, therefore, the four equations split into two independent sets, each
containing only two equations.
Such splitting of the equations becomes impossible, however, when the fields are functions
of time, as the sources of the E- and B-fields become intertwined. Nevertheless, by working with
potentials rather than with the fields, and by staying in the Lorenz gauge, it is possible to arrive
at two 2
nd
-order partial differential equations, one for the scalar potential ( , ), t r the other for
the vector potential ( , ), t A r with the scalar potential depending only on total charge distribution
total
( , ), t r and the vector potential being a function only of the total current distribution
total
( , ). t J r These equations, generally referred to as wave equations, will be derived and analyzed
in the following sections.

5.2. Scalar potential ( , ) t r as the solution of a 2
nd
-order partial differential equation. From
Maxwells first equation,
free
( , ) , t = D r we have

o free total
( , ) ( , ) ( , ) ( , ) t t t t = = E r r P r r (1)
Using the relation between the E-field and the potentials, namely, / , t = E A we may
rewrite Eq. (1) as follows:
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
114 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur


o o total
.
t

= A (2)
Now, in the Lorenz gauge we have
2
(1/ ) / 0. c t + = A Therefore,

2
2 total
2 2
o
1
( , ) ( , )
( , ) .
t t
t
c
t


=
r r
r (3)
In the above equation,
2
( , ) t r is the divergence of the gradient of the scalar potential, an
operation commonly referred to as Laplacian. Substituting in Eq. (3) for ( , ) t r and
total
( , ) t r
their inverse Fourier integrals, we find

2
2
2
4 1
(2 ) ( , ) exp[i( )] d d ( )
t
t
c

k k r k

o total
4
(1/ )(2 ) ( , ) exp[i( )]d d . t

k k r k (4)
The operators on the left-hand-side of the above equation act under the integral sign, with the
Laplacian bringing out the factor (i k) (i k), and the 2
nd
order differentiation with respect to time
bringing out (i )
2
. The resulting expressions on both sides of the equation are Fourier integrals.
Since the two inverse Fourier transforms are equal, the functions being transformed must
themselves be equal. We thus have

2 2
o total
[ ( / ) ] ( , ) (1/ ) ( , ), k c = k k
or, equivalently

total
o
2 2
( , )
( , ) (1/ ) .
( / ) k c

k
k (5)
It is thus clear that Eq. (5) is the Fourier domain version of Eq. (3), namely, the wave
equation in the space-time domain. In principle, one can always move back-and-forth between
the two domains, as the need arises.

5.3. Vector potential ( , ) t A r as the solution of a 2
nd
-order partial differential equation.
From Maxwells 2
nd
equation,
2
o total
( , ) ( , ) (1/ ) ( , )/ , t t c t t = + B r J r E r using the relations
between the fields and the potentials in the Lorenz gauge, we have

2
o total
( , ) ( , ) (1/ )( / ) ( , ) ( , )/ [ ] [ ] t t c t t t t = + A r J r r A r

2 2 2 2 2
o total
( ) [(1/ ) / ] (1/ ) / c t c t = A A J A

2 2 2 2 2
o total
(1/ ) / (1// ) / [ ] c t c t + + = A A A J



2 2 2 2
o total
( , ) (1// ) ( , )/ ( , ). t c t t t = A r A r J r (6)

( Lorenz gauge) = 0
The vector identity (A) =(A)

2
A may be verified in Cartesian
coordinates, where, by definition,

2
A=(
2
A
x
)x
^
+(
2
A
y
) y
^
+(
2
A
z
)z
^
;
see Chapter 1, Problem 35(c).
Solving Maxwells Equations in Space-time Field, Force, Energy and Momentum in Classical Electrodynamics 115

The last equation is the wave equation for the vector potential, relating ( , ) t A r to
total
( , ). t J r
Applying the Fourier transform operation to both sides of Eq. (6) yields

total o
2 2
( , )
( , ) .
( / ) k c

k J
A k (7)
Once again, one can easily go back and forth between Eq. (6) in the space-time domain and
its counterpart in the Fourier domain, Eq. (7), as the need arises.

5.4. Meaning of the Laplacian operator acting on a vector field. If we expand the well-
defined operation ( , ) [ ] t A r in Cartesian coordinates, we find it to be equal to another
well-defined operation, namely,
2
( ) , A A where
2 2 2 2
( ) ( ) ( ) .
x y z
A A A = + + A x y z
Since A
x
, A
y
, and A
z
are scalar functions of r and t, the meaning of the above terms should be
perfectly clear. Thus the wave equation for ( , ), t A r Eq. (6), is in fact a compact formula that
contains three equations, one for each Cartesian component of ( , ). t A r The problem with this
definition of the Laplacian operator is that it is valid only in Cartesian coordinates; one cannot
obtain
2
( , ) t A r in cylindrical or spherical coordinates, for example, by applying the scalar
Laplacian operator to the components ( , , )
z
A A A

or ( , , ) A A A

of the vector field, then
combining the results to create a vector Laplacian field. Needless to say, one can always use the
operational definition,
2
( , ) ( ) ( ), t = A r A A to find the Laplacian of a vector field in
any coordinate system, but the algebra quickly becomes tedious.
A good way to understand the essence of the Laplacian operation on a vector field is to
analyze it in the Fourier domain, where we have

4
( , ) (2 ) ( , ) exp[i( )]d d [ ] { } t t

A r A k k r k

4 2
(2 ) i ( , ) exp[i( )]d d [ ] t

k k A k k r k

4 2
(2 ) ( , ) exp[i( )]d d . k t


A k k r k (8)
Note that ( ) A extracts the longitudinal component of ( , ) A k before multiplying it with k
2
,
thereby throwing out the information contained in the transverse component, ( , ),

A k of the
field. In contrast, the operation ( ) A retains the entire vector field in the Fourier domain,
but it adds an undesirable element, as follows:

4 2
( , ) (2 ) i ( , ) exp[i( )]d d . [ ] [ ] t t

A r k k A k k r k (9)
Using the vector identity ( ) ( ) ( ) , = A B C A C B A B C we now write Eq. (9) as follows:

4 2
( , ) (2 ) ( , ) ( , ) exp[i( )]d d . [ ] [ ] t k t


A r A k A k k r k (10)
The undesirable term in Eq.(10) turns out to be the same expression as appears on the right-hand-
side of Eq.(8). Subtracting this term then yields
140 Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 140-162
CHAPTER 6
The Lorentz Oscillator Model
Everything should be made as simple as possible, but not any simpler!
Albert Einstein (1879-1955)
Abstract. The electric and magnetic dipoles, which produce the polarization and magnetization
of material media, are frequently induced by the action of electromagnetic fields on the atomic
and molecular constituents of these media. The electric field, for example, may draw the
electron cloud surrounding the nucleus of an atom slightly to one side or the other, thus creating
a small separation between the centers of positive and negative atomic charges. Or the magnetic
field might speed up or slow down the motion of electrons circling a nucleus, thus producing a
change in the magnetic dipole moment of individual atoms. Numerous other possibilities exist
as well, such as the reorientation of permanent dipoles in the presence of electromagnetic fields.
As it turns out, a simple mass-and-spring model, first proposed by the Dutch physicist H. A.
Lorentz, captures the essence of induced polarization and magnetization in many situations of
practical interest. In this chapter, we describe the classical Lorentz oscillator model and explore
its fascinating properties.
6.1. Introduction. Material media are composed of atoms and/or molecules which have massive
nuclei and are surrounded by electrons with a relatively small mass. These electrons can often be
excited by an applied electric field E(r, t). The E-field usually displaces the electrons away from
their host nuclei, thus inducing electric dipoles, which collectively give rise to the polarization
P(r, t) of the medium. A simple yet powerful mass-and-spring model, first proposed by the Dutch
physicist H. A. Lorentz, yields a remarkably accurate picture of the E-field dependence of
polarization for a wide class of materials. The Lorentz oscillator model has been applied to
material media ranging from dilute to dense gases, and to transparent as well as absorptive
liquids and solids including dielectrics, semiconductors, and metals. This chapter presents the
classical Lorentz oscillator model and describes its properties under diverse circumstances as
well as its application to a broad class of practical problems.
6.2. Mass-and-spring model of an atomic dipole.
With reference to Fig. 1, consider a particle of
mass m and charge q, attached to a massive
particle of equal but opposite charge +q via a
spring whose real-valued, positive spring-constant
is . Furthermore, assume that this mass-and-
spring system is subject to dynamic friction having
a real-valued, positive friction coefficient . In the
absence of external excitation, the equilibrium
position of the q charge coincides with the center
of the +q charge, resulting in a net total charge of
zero, and an electric dipole moment p =p
x
x
^
=0.
When an electric field E(t) =E
xo
cos(t )x
^
is
applied along the x-axis, it drives the negative
charge away from its equilibrium position.
Denoting the displacement of the negative charge
m, q
M, +q
x
E
xo
cos(t )
z
Spring constant >0
Dynamic friction coefficient >0
p(t ) =qx(t) x
^
Fig. 1. The Lorentz oscillator model of a single
electron of mass m and charge q attached via a
spring to a heavy nucleus of mass M and charge +q.
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
The Lorentz Oscillator Model Field, Force, Energy and Momentum in Classical Electrodynamics 141

along the x-axis by x(t), the force exerted on the charge will consist of the external force qE(t),
the restoring spring force x(t)x
^
, and the dissipative friction force V(t)x
^
, where the velocity
of the particle along the x-axis is V(t) =dx(t)/dt. The total force must then be equal to mass times
acceleration according to Newtons law of motion, namely,
md
2
x(t)/dt
2
= qE(t) x(t) dx(t)/dt. (1a)
Defining two new parameters,
o
= /m and = /m, we rewrite the above equation as follows:
d
2
x(t)/dt
2
+ dx(t)/dt +
o
2
x(t) = (q/m)E(t). (1b)
For reasons that will become clear later,
o
is usually referred to as the resonance frequency and
as the damping coefficient. Using complex notation, we write E(t) =Re[E
xo
exp(it )] and
x(t) =Re[x
o
exp(it )], where, in general, x
o
=| x
o
| exp(i
o
) is a complex constant. We then
rewrite Eq. (1b) as

2
x
o
i x
o
+
o
2
x
o
= (q/m)E
xo
, (2a)
which yields




Denoting by
p
(t) the electric dipole moment of the oscillating mass-and-spring system, namely,

p
(t)= qx(t) x
^
=Re[qx
o
exp(it ) x
^
] =Re[
p
o
exp(it )], (3)
we will have




Let there be N such dipoles in a unit volume of space at a given location r, and define the
polarization as P(r, t) =Re[N
p
o
exp(it )]. Defining the dimensionless function C() such that
P(r, t ) =Re[
o
C()E
xo
exp(it )x
^
], we will have




Using a new parameter
p
= Nq
2
/(
o
m), called plasma frequency, we rewrite Eq. (5) as follows:



In the MKSA system of units,
o
has units of farad/ meter while the units of the E-field are
volt/ meter, yielding for the polarization P(r, t) the units of farad volt /m
2
, which is the same as
coulomb/m
2
. Also, one may readily confirm that
p
has units of s
1
.
Generalizing the above result, we now state that the polarization P(r, t) at a given location r
in an isotropic and linearly-polarizable medium excited by the monochromatic E-field
E(r, t) =Re[E(r) exp(it )] is obtained via the complex-valued proportionality constant
o
C(),
namely,
P(r, t ) =Re[
o
C()E(r) exp(it )]. (7)
The complex-valued function of frequency C() depends on three effective material parameters:
the plasma frequency
p
, the resonance frequency
o
, and the damping coefficient .

o
2
+i
x
o
=
(q/m)E
xo
.
(2b)

o
2
+i
p
o
=
(q
2
/m)E
xo
.
(4)

o
2

2
i
C() =
(Nq
2
/
o
m)
.
(5)

o
2

2
i
C()
=

p
2
.
(6)
142 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur
6.3. Generalization to the case of multi-electron atoms and molecules. A typical material is
made up of atoms or molecules that have many electrons associated with their individual atomic
nuclei. Each electron may therefore be represented by a mass-and-spring system similar to that
depicted in Fig. 1. Electrons occupying different orbitals, of course, must have different
parameters
p
,
o
, and . The polarizability coefficient C() of a medium whose constituent
atoms/molecules each have K different electrons may thus be expressed as





Note in Eq. (8) that we have kept
p
independent of k, while introducing the new parameter f
k
,
called the k
th
oscillator strength, in order to account for differences in the effective mass m
k
and
effective charge q
k
of the various oscillators, as well as possible differences in their number
densities N
k
. The parameters (
p
, f
k
,
ok
,
k
) of the Lorentz oscillator model where k ranges
over all the electrons in distinct orbitals associated with the constituent atoms/molecules of the
medium must be obtained either from quantum-mechanical calculations or through
experimental measurements of the material properties.
6.4. Drude model of the conduction electrons. The Lorentz oscillator model may be used to
describe the conduction electrons in a metal or semiconducting medium, provided that the spring
constant or, equivalently, the resonance frequency
ok
for these electrons is set equal to zero.
In other words, since the conduction electrons are not bound to any particular atom/molecule,
they should not be subject to the restoring force of a spring. However, when 0, the
conduction electron may still be imagined to respond to an oscillating E-field by vibrating
around its equilibrium position (whatever that term may mean for a delocalized electron), subject
only to the force exerted by the E-field and the dynamic friction force represented by the
damping coefficient . The proportionality constant (actually not a constant but a function of )
between the E-field and the polarization P(r, t ) of conduction electrons is now represented by

e
(), rather than C(), and referred to as the electric susceptibility of the conduction electrons.
Setting
o
= 0 in Eq. (6), we find



The above expression is known as the Drude model of the conduction electrons, even though
it is only a special case of the Lorentz oscillator model. At high frequencies where >>, Eq. (9)
may be further simplified by ignoring the term i in the denominator. We then have

e
()

(
p
/)
2
, which is known as plasma susceptibility. The electric permittivity of the
plasma is thus given by () =1+

e
() 1(
p
/)
2
.
Note: In isotropic linear media, the displacement field D(r,) =
o
E(r,) +P(r,) may be expressed as

o
[1+

e
()]E(r,). The dimensionless proportionality coefficient relating D(r,) to E(r,) is generally known as
the electric permittivity () of the medium, and related to the electric susceptibility via () =1+

e
(). Since the
actual proportionality coefficient relating D(r,) to E(r,) is
o
(), it is not uncommon to see () referred to as
the relative permittivity. It is also common practice to call () the dielectric constant of the material medium,
although a better name would be dielectric function, since permittivity is not a constant but a function of .
Similarly, the magnetic induction B(r,) =
o
H(r,) +M(r,) may be written
o
[1+

m
()]H(r,). The
dimensionless proportionality coefficient relating B(r,) to H(r,) is known as the magnetic permeability () of

ok
2

2
i
k

C
K
()
=


f
k

p
2
.
(8)
k =1
K

2
+i

e
()
=


p
2
.
(9)
Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 163-195 163
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
CHAPTER 7
Plane Electromagnetic Waves in Isotropic, Homogeneous, Linear Media
What, then, is light according to the electromagnetic theory? It consists of alternate and opposite
rapidly recurring transverse magnetic disturbances, accompanied with electric displacements, the
direction of the electric displacement being at the right angles to the magnetic disturbance, and both
at right angles to the direction of the ray.
James Clerk Maxwell (1831-1879)
The velocity of light is one of the most important of the fundamental constants of Nature. Its
measurement by Foucault and Fizeau gave as the result a speed greater in air than in water, thus
deciding in favor of the undulatory and against the corpuscular theory. Again, the comparison of the
electrostatic and the electromagnetic units gives as an experimental result a value remarkably close to
the velocity of light a result which justified Maxwell in concluding that light is the propagation of
an electromagnetic disturbance. Finally, the principle of relativity gives the velocity of light a still
greater importance, since one of its fundamental postulates is the constancy of this velocity under all
possible conditions.
Albert Abraham Michelson (1852-1931)
Abstract. Material media typically react to electromagnetic fields by becoming polarized or
magnetized, or by developing charge- and current-density distributions within their volumes or
on their surfaces. The response of a material medium to the fields could be complicated, as
would be the case, for instance, when the relation between induced polarization and the electric
field is non-local, non-linear, or history-dependent, or when the induced magnetization is an
anisotropic function of the local magnetic and/or electric fields. In many cases of practical
interest, however, the media are homogeneous, isotropic, and linear, with the electric dipoles
responding only to the local E-field (and magnetic dipoles responding only to the local H-field)
in accordance with the Lorentz oscillator model of the preceding chapter. Irrespective of the
manner in which the charge-carriers or the dipoles of the medium respond to the fields, there is
always an additional complication that the fields are not merely those imposed on the medium
from the outside. The motion of the charges and/or the oscillation of the dipoles in response to
the fields give rise to new electromagnetic fields, which must then be added to the external
fields before the induced charge, current, polarization, or magnetization can be computed. In
other words, the entire system of interacting fields and sources, whether originating outside or
induced within the media, must be treated self-consistently. This chapter provides a detailed
analysis of plane-wave propagation within the simplest kind of material media, namely, those
that are homogeneous and isotropic, whose induced electric dipoles are linear functions of the
local E-field, and whose induced magnetic dipoles are linear functions of the local H-field.
7.1. Introduction. A plane electromagnetic wave is specified by its temporal frequency , a
propagation vector k, an electric-field amplitude E
o
, and a magnetic-field amplitude H
o
. In
general, electromagnetic fields are real-valued vector-fields throughout space and time, that is,
the E- and H-fields, E(r, t) and H(r, t), are real-valued vector functions of the spatial and temporal
coordinates, r =(x, y, z) and t. It turns out, however, that complex-valued functions of (r, t) in
which k, E
o
and H
o
are allowed to be complex vectors may be used to mathematically
represent the E- and H-fields, so long as the real parts of these complex functions are recognized
as corresponding to the actual fields. In this way, the physical fields always end up being real-
valued, while their complex-valued representations simplify mathematical operations. Section 2
provides a detailed description of the complex vector algebra used throughout the chapter.
164 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur

While a traditional, real-valued k-vector is used to represent a homogeneous plane-wave in
free space or within a transparent medium, the use of a complex k-vector enables one to discuss
inhomogeneous plane-waves such as evanescent waves in transparent media, exponentially
decaying (or attenuating) plane-waves in absorptive media, and also exponentially growing (or
amplifying) plane-waves within gain media. Moreover, the use of complex-valued field
amplitudes E
o
and H
o
enables one to consider arbitrary states of polarization (i.e., linear, circular,
elliptical) within the same mathematical formalism. A general discussion of complex-valued
plane-waves and their properties is given in section 3.
Isotropic, homogeneous, linear media are characterized by the uniformity and isotropy of
their electromagnetic properties, and by the fact that their dielectric susceptibility
o

e
() relates
the polarization P(r, t) to the local E-field E(r, t), while their magnetic susceptibility
o

m
()
relates the magnetization M(r, t) to the local H-field H(r, t). Plane-wave solutions to Maxwells
equations in such media are discussed in section 4, where relationships among the frequency ,
the k-vector, the field amplitudes (E
o
, H
o
), and material parameters () =1+

e
() and
() =1+

m
() are derived from Maxwells macroscopic equations.
The rate of flow of electromagnetic energy (per unit area per unit time) in an optical or
electromagnetic field is given by the Poynting vector S(r, t) =E(r, t) H(r, t). A general
expression for the time-averaged Poynting vector of a monochromatic plane-wave is derived in
section 5. This result is applicable to all sorts of plane-waves, whether propagating in a
transparent medium, exponentially decaying within an absorptive medium, exponentially
growing inside a gain medium, or evanescent. Using the results of section 5, one can calculate
the energy flux of any plane-wave residing in an isotropic, homogeneous, linear medium, and
confirm the conservation of energy under various circumstances.
When an electromagnetic wave arrives at the interface between two (physically distinct)
media, a part of the wave is reflected at the interface, while the remaining part enters from the
first (incidence) medium into the second. Assuming the two media are isotropic, homogeneous,
and linear each specified by its dielectric permittivity () and magnetic permeability ()
and also assuming that the two media occupy adjacent, semi-infinite, half-spaces separated at a
flat interface, one can readily calculate the reflection and transmission coefficients for arbitrary
plane-waves that arrive at the interface. These so-called Fresnel reflection and transmission
coefficients are derived in section 6. The results of section 6 are completely general, and may be
used in conjunction with transparent or absorptive media. They encompass reflection and
transmission of arbitrarily-polarized plane-waves directed at the interface between two arbitrary
media at an arbitrary angle of incidence. The media under consideration may be metals, plasmas,
or dielectrics with positive or negative refractive indices. The results of section 6 are applicable
to ordinary reflection and transmission, total internal reflection, and reflection from plasmas. We
discuss several cases of practical interest in section 7.
7.2. Complex vector algebra of the electromagnetic field. This chapter describes the
dependence of plane electromagnetic waves on space and time coordinates using a powerful
complex notation. A complex number such as c=a+i b is specified by its real and imaginary
parts a and b, respectively. The sum and difference of two complex numbers are defined as
c
1
c
2
=(a
1
a
2
)+i (b
1
b
2
). The product of two complex numbers is c
1
c
2
=(a
1
a
2
b
1
b
2
)+
i (a
1
b
2
+a
2
b
1
). Division is the inverse of multiplication, in the sense that if c
3
=c
1
/c
2
, then c
1
=c
2
c
3
.
The complex-conjugate of c=a+i b is defined as c*=a i b. The product cc*=a
2
+b
2
is always
real and non-negative. It is often useful to write c
1
/c
2
as (c
1
c
2
*)/(c
2
c
2
*), which has a real-valued
Plane Electromagnetic Waves in Isotropic Field, Force, Energy and Momentum in Classical Electrodynamics 165

denominator; complex division is thus reduced to complex multiplication of c
1
and c
2
*, followed
by normalization by the real-valued c
2
c
2
*.
In polar representation, a complex number is written as c =|c|exp(i
c
). The magnitude (or
modulus) |c| of the complex-number c is a non-negative real number; its phase
c
is in the range
<
c
(modulo 2). The Euler identity exp(i ) =cos +i sin is generally used to convert c
from one representation to the other, namely, c = |c|exp(i
c
) =|c| cos
c
+i |c| sin
c
=a+ib.
A vector V in 3D-space is written in Cartesian coordinates as V=V
x
x
^
+ V
y
y
^
+V
z
z
^
. The sum
and difference of two vectors are defined as V
1
V
2
=(V
x1
V
x2
)x
^
+(V
y1
V
y2
)y
^
+(V
z1
V
z2
)z
^
. As
for vector multiplication, two types of products are defined in vector algebra. The rule for dot
multiplying V
1
=V
x1
x
^
+V
y1
y
^
+V
z1
z
^
and V
2
=V
x2
x
^
+V
y2
y
^
+V
z2
z
^
, denoted by V
1
V
2
, is that the
three terms of V
1
are multiplied into the three terms of V
2
using the following identities:
x
^
x
^
= y
^
y
^
= z
^
z
^
=1; x
^
y
^
= x
^
z
^
= y
^
x
^
= y
^
z
^
= z
^
x
^
= z
^
y
^
=0. (1)
The other type of vector multiplication, denoted V
1
V
2
and referred to as cross-product, requires
that the three terms of V
1
be multiplied into the three terms of V
2
using the alternative set of rules
listed below:
x
^
x
^
=0, y
^
y
^
=0, z
^
z
^
=0, (2a)
x
^
y
^
= z
^
, y
^
z
^
= x
^
, z
^
x
^
= y
^
, (2b)
y
^
x
^
= z
^
, z
^
y
^
=x
^
, x
^
z
^
=y
^
. (2c)
In the complex notation used in this chapter, we expand the domain of vector algebra by
allowing the three components V
x
, V
y
, V
z
of an arbitrary vector V to be complex numbers. Note
that we do not change the rules for adding, subtracting, dot-multiplying, or cross-multiplying
pairs of vectors; all we require is that the rules of complex-number addition and multiplication be
followed when adding and multiplying the components V
x1
, V
y1
, V
z1
of one vector with the
components V
x2
, V
y2
, V
z2
of another, while retaining the rules of vector multiplication for the unit
vectors x
^
, y
^
, z
^
listed in Eqs. (1) and (2). A complex-valued vector may thus be written
V=V
x
x
^
+V
y
y
^
+V
z
z
^
=(V
x
+iV
x
)x
^
+(V
y
+iV
y
)y
^
+(V
z
+iV
z
)z
^
. (3a)
Combining the real parts of the various components of V into one real-valued vector V , and the
imaginary parts into another (also real-valued) vector V , we rewrite Eq. (3a) as
V=(V
x
x
^
+V
y
y
^
+V
z
z
^
) +i(V
x
x
^
+V
y
y
^
+V
z
z
^
) =V +iV . (3b)
Each complex-valued vector V may therefore be considered a combination of two real-valued
vectors V and V in accordance with Eq. (3b).
The reader should be warned that, although complex-numbers may, on occasion, be
represented as 2D-vectors in the complex-plane, there must be no confusing such vectors with
the 3D-vectors that are the subject of the present discussion. For our purposes, therefore,
numbers in the complex-plane are scalars, not vectors. The only vectors we shall encounter are
those in the conventional 3D-space; they are written as V=V
x
x
^
+V
y
y
^
+V
z
z
^
, their components
V
x
, V
y
, V
z
being either real- or complex-valued. For example, a point r in 3D-space will be
denoted in Cartesian coordinates as r =xx
^
+y y
^
+z z
^
, a 3D-vector with real-valued components
x, y, z. On the other hand, the propagation vector k =k
x
x
^
+k
y
y
^
+k
z
z
^
will be treated as a 3D-vector
with complex components k
x
, k
y
, k
z
.
196 Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 196-210
CHAPTER 8
Simple Applications Involving Plane Electromagnetic Waves
Science is built up with facts, as a house is with stones. But a collection of facts is no more a
science than a heap of stones is a house.
Jules Henri Poincar (1854-1912)
The mere formulation of a problem is often far more essential than its solution, which may be merely
a matter of mathematical or experimental skills. To raise new questions, new possibilities, to regard
old problems from a new angle requires creative imagination and marks real advances in science.
Albert Einstein (1879-1955)
Abstract. Plane-waves are the building blocks of arbitrary electromagnetic waves residing in
homogeneous, linear media. A basic understanding of plane-wave properties and their behavior
upon arrival at a flat interface between two such media is sufficient for the analysis of a number
of interesting electrodynamic problems. This chapter presents several exemplary problems that
are easy to set up and to describe, yet the understanding and appreciation of their full impact
requires subtle arguments involving the properties of plane-waves and those of the Fresnel
reflection and transmission coefficients. These examples reveal certain interesting as well as
useful features of optical and electromagnetic systems that are frequently encountered in
practical applications.
8.1. Introduction. This chapter contains a number of examples that require only an elementary
knowledge of the properties of plane-waves, yet the conclusions reached in each case are far
from trivial or obvious. The examples cover a range of problems of practical interest, including
the reflection and transmission properties of multilayer stacks, characteristics of a Fabry-Perot-
like optical resonator, and the intricacies of the perfectly-matched boundary layer used in certain
numerical solutions of Maxwells equations.

8.2. Transmission through a multilayer stack. A multilayer stack consists of N layers of
isotropic, homogeneous, linear materials, each specified in terms of its complex-valued
n
()
and
n
() as well as a thickness d
n
. The stack is surrounded by vacuum on all sides, as shown in
Fig. 1. For a plane-wave normally incident on this multilayer, we show that the complex-valued
Fresnel transmission coefficient does not depend on which side of the stack faces the light
source. This result can also be extended to oblique incidence at an arbitrary incidence angle .
Let us split off the first layer from the rest of the stack to create a small gap between the two
parts, as shown in Fig. 2. Note that the first layer by itself, surrounded by vacuum, is symmetric,
and, therefore,
1
=
1
and
1
=
1
. Let the amplitude of the plane-wave inside the gap and
traveling from left to right be A
1
. Denoting by
1
and
2
the reflection coefficients from the two
exposed surfaces within the gap, and by
1
the transmission coefficient of the first layer, we have
A
1
=
1
E
(i)
+
1

2
A
1
exp(i 2). (1)
Here 2 =4d/
o
is the round-trip phase acquired by the beam within the gap, which will vanish
in the limit when d0. From Eq. (1) we find
A
1
=
1
E
(i)
/ [1
1

2
exp(i 2)]. (2)
Denoting by
2
the transmission coefficient of the rest of the stack, we see that the total
transmission coefficient is
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
Simple Applications Involving Plane Field, Force, Energy and Momentum in Classical Electrodynamics 197

=
1

2
exp(i)/ [1
1

2
exp(i)], (3)
which, in the limit when d0, approaches =
1

2
/ (1
1

2
).
























Fig. 1. A plane-wave having amplitude E
(i )
is normally incident on an arbitrary multilayer stack
surrounded by vacuum. In general, there is no a priori reason to believe that the reflection and
transmission coefficients (, ) for incidence from the left-hand side should be the same as (, )
for incidence from the right-hand side. As it turns out, will be equal to under all
circumstances, whether or not the stack is absorptive. The reflection coefficients from the opposite
sides, however, are not necessarily the same. In the absence of absorption, one can readily show
that and must have equal magnitudes, that is, | | =| |, although the phase angles of and
will, in general, be different.























Fig. 2. (a) The first layer is split off and placed at a short distance d from the rest of the stack. The
complex amplitude of the E-field propagating from left to right inside the gap is A
1
, that
propagating from right to left is A
2
. The relations between the various field amplitudes are:
A
2
=
2
A
1
exp(i 2d/
o
), A
1
=E
(i)

1
+
1
A
2
exp(i 2d/
o
), E
(t)
=
2
A
1
exp(i 2d/
o
). The first layer
being symmetric, its reflection and transmission coefficients from the two sides must be identical,
that is,
1
=
1
and
1
=
1
. (b) Similar to (a), except that the incident beam now arrives from the
right-hand side.
E
(i)

E
(r)
=E
(i)

E
(t)
= E
(i)
E
(i)
E
(r)
=E
(i)
E
(t)
= E
(i)
Partially transmitting mirror
(incidence from the left-hand side)
Vacuum
Vacuum
Partially transmitting mirror
(incidence from the right-hand side)
(b)
(a)
E
(i)

E
(r)
E
(i)

E
(t)
E
(i)
Gap
A
1

A
2

1
,
1
,
1
,
1

2
,
2
,
2
,
2

d
E
(i)
E
(r)
E
(i)
E
(t)
E
(i)
B
1
B
2

1
,
1
,
1
,
1


2
,
2
,
2
,
2

d
(b) (a)
198 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur

Next, suppose the direction of incidence is reversed. Everything in the above expression for
will remain the same except for
2
, which becomes
2
. Now, if the reduced stack (i.e., the one
with N1 layers) happens to be symmetric, we will have
2
=
2
, resulting in the entire stack
having the same transmission coefficient from both sides. If the reduced stack is not
symmetric, we can continue to split from it one layer at a time and complete the proof by
induction.
The case of oblique incidence at an arbitrary angle is similar to that of normal-incidence
discussed above, except that the two polarization states, p and s, must be treated separately. Also,
the round-trip phase-shift inside the gap will be 2 =4dcos /
o
, but the cos factor will not
affect the final result, as will vanish in the limit when d0, irrespective of the value of .

8.3. Reflection and transmission coefficients of a non-absorbing slab. A slab made of
unknown but lossless materials (perhaps a multilayer made of transparent dielectric layers) has
the property that its Fresnel reflection and transmission coefficients from one side are (, ),
while the corresponding coefficients from the opposite side are (, ). Let =|| exp(i

),
=| | exp(i

), and so on for the other coefficients. Because the slab is lossless, we must have
||
2
+| |
2
=1 and ||
2
+| |
2
=1. Place the slab in front of a perfect reflector, with a vacuum gap of
arbitrary thickness between the two, as shown in Fig. 3. Shine a plane-wave on the slab at normal
incidence from the left-hand side, and observe that the reflection coefficient of the entire
structure (slab + gap + perfect mirror) must have unit amplitude but, in general, a non-zero
phase. Denote the reflection coefficient of the entire structure with exp(i), where depends on
the slabs reflection and transmission coefficients as well as the size of the gap d and the
wavelength
o
. We will show that

= (

), and also that || =|| and | |=| |. It is


impossible to prove with this technique that

, but that is something we already know from


the preceding section.
















Fig. 3. A lossless slab, having reflection and transmission coefficients (, ) from the left-hand
side and (, ) from the right-hand side, is placed at a distance d from a perfect reflector. Inside
the gap, the E-field amplitude immediately after the slab is A
1
, that immediately after the perfect
mirror is A
2
. Since the entire structure is lossless, its reflection coefficient is denoted by exp(i),
which is a complex number of unit amplitude and arbitrary phase.
E
(i)

E
(r)
=exp(i)E
(i)
Gap
A
1
A
2
d
Partially-transmitting
lossless mirror
(, , , )
Perfect
reflector
( =1)
Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 211-221 211
CHAPTER 9
Maxwells Equations in Cylindrical Coordinates
The news of the Nobel prize reached us in the middle of the night at 3 am in California, through a
telephone call from Stockholm (then in their morning) no doubt carried on optical fibers;
congratulations came literally minutes later from friends in Asia (for whom it was evening), again
through messages carried on optical fibers. Too much information is not always a good thing: we had
to take the phone off the hook that night in order to get some sleep!
From Charles K. Kaos December 8, 2009 Nobel lecture. Kao shared the Nobel Prize in Physics "for
groundbreaking achievements concerning the transmission of light in fibers for optical communication."

Abstract. We describe the solution of Maxwells equations in systems of homogeneous,
isotropic, and linear media that exhibit cylindrical symmetry around a given axis. The solutions
are generally expressed in terms of Bessel functions of various kinds and integer orders. The
properties of these solutions are discussed in some detail throughout the chapter, where several
practical applications are also pointed out.
9.1. Introduction. In many systems of practical interest, electromagnetic waves propagate
within linear, isotropic, homogeneous, circularly symmetric media. Examples include hollow-
tube microwave wave-guides, optical fibers, fiber lasers, nano-wires, and nano-rods. In this
chapter we discuss the solution of Maxwells equations for systems that exhibit circular
symmetry around a given axis. We then proceed to examine the fundamental characteristics of
these solutions, which are generally referred to as the modes (or eigenmodes) of the system.
9.2. Solving Maxwells equations in linear, isotropic, homogeneous, circularly symmetric
media. An electromagnetic wave in a system with cylindrical symmetry has eigenfunctions of
the form E(r, , z, t) = [E
r
(r)r

+ E

(r)

+ E
z
(r)z

]exp(im)exp(ik
o

z
z)exp(it), with a similar
expression for the magnetic field H(r, , z, t). Here k
o
=/c=2/
o
, where
o
is the vacuum
wavelength; the integer m is the azimuthal mode number; and the complex-valued
z
is the
propagation constant along the z-axis. In general, the beam resides within an environment having
complex permittivity and permeability
o
() and
o
(). As usual, the speed of light in
vacuum c = 1/
o

o
and the impedance of free space Z
o
=
o
/
o
. In the absence of free charges
and free currents, Maxwells equations are written
E
r
+ rE
r
/r + imE

+ ik
o
r
z
E
z
= 0, (1)
E
r
=
z
Z
o
H

(mZ
o
/k
o
r)H
z
, (2a)
E

=
z
Z
o
H
r
(iZ
o
/k
o
)H
z
/r, (2b)
E
z
= (mZ
o
/k
o
r)H
r
+ (iZ
o
/k
o
r)H

+ (iZ
o
/k
o
)H

/r, (2c)
Z
o
H
r
= (m/k
o
r)E
z

z
E

, (3a)
Z
o
H

=
z
E
r
+ (i/k
o
)E
z
/r, (3b)
Z
o
H
z
= (m/k
o
r)E
r
(i/k
o
r)E

(i/k
o
)E

/r, (3c)
H
r
+ rH
r
/r + imH

+ ik
o
r
z
H
z
= 0. (4)
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
212 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur

Substituting for E
r
and E

from Eqs. (2a) and (2b) into Eqs. (3a) and (3b) yields
(
z
2
)H
r
= (m /Z
o
k
o
r)E
z
+i (
z
/k
o
)H
z
/r, (5)
(
z
2
)H

= (m
z
/k
o
r)H
z
+i (/k
o
Z
o
)E
z
/r. (6)
Placing the above expressions for H
r
and H

in Eq. (2c) results in the following 2


nd
order
differential equation for E
z
:

2
E
z
/r
2
+ (1/r)E
z
/r +[k
o
2
(
z
2
) (m/r)
2
]E
z
= 0. (7)
We define the radial propagation constant
r
=
z
2
, so that the radial coordinate r may
be written in normalized form as = k
o

r
r. [For given values of , , and
z
, we shall always
choose the value of
r
such that Im(
r
) 0.] Equation (7) may now be written as follows:

2
E
z
/
2
+ (1/)E
z
/ +(1m
2
/
2
)E
z
= 0. (8)
In similar fashion, we substitute for H
r
and H

from Eqs. (3a) and (3b) into Eqs. (2a) and (2b)
to obtain
(
z
2
)E
r
= i (
z
/k
o
)E
z
/r (mZ
o
/k
o
r)H
z
, (9)
(
z
2
)E

= (m
z
/k
o
r)E
z
i(Z
o
/k
o
)H
z
/r. (10)
Placing the above expressions for E
r
and E

in Eq. (3c) results in the following 2


nd
order
differential equation for H
z
:

2
H
z
/r
2
+ (1/r)H
z
/r +[k
o
2
(
z
2
) (m/r)
2
]H
z
= 0. (11)
Once again, use of the normalized radial coordinate = k
o

r
r leads to the standard Bessel
equation for H
z
(), that is,

2
H
z
/
2
+ (1/)H
z
/ +(1m
2
/
2
)H
z
= 0. (12)
We mention in passing that Eq. (12) could also have been obtained by substituting H
r
and H


of Eqs. (5) and (6) into Eq. (4). Similarly, Eq. (8) could have been obtained by substituting E
r
and
E

of Eqs. (9) and (10) into Eq. (1). This should not be surprising, considering that Maxwells
equations are interdependent.
The solutions of Eqs. (8) and (12) may now be expressed in terms of Bessel functions of the
radial distance r from the z-axis. In general, any linear combination of Bessel functions of the
first and second kind, order m, namely, J
m
() and Y
m
(), satisfies the above equations. Thus the
general form of the cylinder function appearing in the solution to Eqs. (8) and (12) will be
C
m
() = AJ
m
()+BY
m
(), where A and B are arbitrary complex constants. The derivative of the
cylinder function with respect to its complex argument , denoted by C
m
(), will appear in the
expressions for H
r
, H

, E
r
, and E

, as can be readily observed by examining Eqs. (5), (6), (9), and


(10). The following identities will be useful:

2
C
m
() + C
m
() + (
2
m
2
)C
m
() = 0, (13a)
C
m
() = (/2m)[C
m1
() + C
m+1
()], (13b)
C
m
() = C
m1
() (m/)C
m
(), (13c)
C
m
() = (m/)C
m
() C
m+1
(), (13d)
Maxwells Equations in Cylindrical Coordinates Field, Force, Energy and Momentum in Classical Electrodynamics 213

C
m
() = (1)
m
C
m
(), (13e)
C
m
() = (1)
m
C
m
(). (13f)
We divide the solutions to Eqs. (1-4) into two sets; for one set, we let E
z
= 0, in which case
the solution will be referred to as Transverse Electric (TE); for the other, we let H
z
= 0, and call
the solution Transverse Magnetic (TM). Occasionally, it will be possible to express the complete
solution to a given problem in terms of either TE or TM modes; this occurs, for instance, when

z
= 0 or when m = 0. Often, however, the boundary conditions cannot be matched with just one
set of solutions; in such cases both TE and TM modes appear in the solution to Maxwells
equations, and the resulting mode will be referred to as EH or HE, depending on whether E
z
or
H
z
dominates the longitudinal field component. The TE mode wavefunctions are written below:

TE: E
m
(r) = (m/k
o
r
r
2
)C
m
(k
o

r
r)r

(i/
r
)C
m
(k
o

r
r)

, (14a)
Z
o
H
m
(r) = (i
z
/
r
)C
m
(k
o

r
r)r

(m
z
/k
o
r
r
2
)C
m
(k
o

r
r)

+ C
m
(k
o

r
r)z

. (14b)

Similarly, the TM mode wavefunctions are given by:

TM: E
m
(r) = (i
z
/
r
)C
m
(k
o

r
r)r

(m
z
/k
o
r
r
2
)C
m
(k
o

r
r)

+ C
m
(k
o

r
r)z

, (15a)
Z
o
H
m
(r) = (m /k
o
r
r
2
)C
m
(k
o

r
r)r

+ (i /
r
)C
m
(k
o

r
r)

. (15b)

The case of m = 0 may be simplified as follows:

TE (m=0): E
0
(r, , z) = (i/
r
)C
1
(k
o

r
r)exp(i k
o

z
z)

, (16a)
Z
o
H
0
(r, , z) = [(i
z
/
r
)C
1
(k
o

r
r)r

+ C
0
(k
o

r
r)z

]exp(ik
o

z
z). (16b)

TM (m=0): E
0
(r, , z) = [(i
z
/
r
)C
1
(k
o

r
r)r

+ C
0
(k
o

r
r)z

]exp(i k
o

z
z), (17a)
Z
o
H
0
(r, , z) = (i/
r
)C
1
(k
o

r
r)exp(ik
o

z
z)

. (17b)

In general, the field components in the xy-plane, E
m||
and H
m||
, may be expressed in terms of a
superposition of right- and left-circularly polarized states, as follows:

TE: E
m||
(r, , z) = (/2
r
){C
m1
(k
o

r
r)exp[i(m1)](x

+i y

)+C
m+1
(k
o

r
r)exp[i(m+1)](x

i y

)}
exp(ik
o

z
z), (18a)
Z
o
H
m
(r, , z)={(i
z
/2
r
)[C
m1
(k
o

r
r)exp[i(m1)](x

+i y

)C
m+1
(k
o

r
r)exp[i(m+1)](x

i y

)]
+ C
m
(k
o

r
r)exp(im)z

}exp(ik
o

z
z). (18b)

TM: E
m
(r, , z)= {(i
z
/2
r
)[C
m1
(k
o

r
r)exp[i(m1)](x

+i y

)C
m+1
(k
o

r
r)exp[i(m+1)](x

i y

)]
+ C
m
(k
o

r
r)exp(im)z

}exp(ik
o

z
z), (19a)
Z
o
H
m||
(r, , z) = (/2
r
){C
m1
(k
o

r
r)exp[i(m1)](x

+i y

)+C
m+1
(k
o

r
r)exp[i(m+1)](x

i y

)}
exp(ik
o

z
z). (19b)
222 Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 222-245
CHAPTER 10
Electromagnetic Momentum, Angular Momentum, Force and Torque
The experimental investigation by which Ampere established the law of the mechanical action
between electric currents is one of the most brilliant achievements in science. The whole theory and
experiment, seems as if it had leaped, full grown and full armed, from the brain of the ' Newton of
Electricity'. It is perfect in form, and unassailable in accuracy, and it is summed up in a formula
from which all the phenomena may be deduced, and which must always remain the cardinal formula
of electrodynamics.
James Clerk Maxwell (1831-1879)
Abstract. Electromagnetic fields are capable of exerting force and torque on material media.
This is the mechanism by which linear and angular momenta are exchanged between the fields
and the media. The fundamental principles that govern such exchanges were described in some
detail in Chapter 2. The present chapter provides several examples that demonstrate the
application of these principles in certain situations of practical interest.
10.1. Introduction. Electromagnetic waves carry linear and angular momenta, whether they
reside in free space or within material media. As mentioned briefly in Chapter 2, the linear
momentum density of an EM field is always given by
2
EM
( , ) ( , )/ , t t c = p r S r where S is the
Poynting vector at point r in space and instant t in time. Similarly, the angular momentum
density of the EM field at (r, t ) relative to an arbitrary point r
o
is always given by
2
EM o
( , ) ( ) ( , )/ . t t c = L r r r S r It makes no difference whether the field is in vacuum or inside a
material medium, whether the field is of the propagating-type or evanescent, whether the
medium is linear or nonlinear, absorbing or transparent, birefringent or optically active, etc.
Once the E- and H-fields at (r, t ) are known, the local Poynting vector may be calculated using
the standard formula ( , ) ( , ) ( , ), t t t = S r E r H r from which the linear and angular momentum
densities of the EM field can then be obtained in accordance with the above prescription.
Whenever the E- and/or the H-field happen to be co-located with a material medium at a
point (r, t ) in space-time, there could occur an exchange of linear and/or angular momenta
between the EM field and the matter. Such exchanges take place through the action of
electromagnetic force and torque on material media, which media present themselves to the EM
field as
free
(r, t), J
free
(r, t), P(r, t ), M(r, t ), or a combination thereof. The expressions of
electromagnetic force and torque densities were given in Chapter 2, Eqs. (35) and (37), where it
was emphasized that any gain or loss of field momentum will be accompanied by a
corresponding loss or gain of material momentum, in such a way that the total linear and angular
momenta of a closed system are always conserved.
The exchange of momentum between electromagnetic fields and material media can go both
ways. For example, a light pulse entering a transparent glass slab from free space loses a fraction
of its EM momentum to the slab. Subsequently, the slab acquires a mechanical momentum and
begins to move forward, in the direction of propagation of the light pulse. This motion will be
extremely slow if the mass of the slab happens to be substantial. The slabs momentum,
however, is generally a significant fraction of the initial momentum of the light pulse; in other
words, the product of a large mass and a small velocity is not necessarily small. In contrast, the
same light pulse, when leaving the slab through the exit facet, will take back the momentum it
had initially given the slab. In the process of exiting the glass slab and re-entering free space, the
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
Electromagnetic Momentum, Angular Momentum Field, Force, Energy and Momentum in Classical Electrodynamics 223

light pulse recovers its initial momentum by putting a brake on the slab; the slab thus transfers all
its mechanical momentum to the light pulse and comes to a halt once the pulse is fully outside.
(To avoid complications arising from reflections at the entrance and exit facets, we have
assumed in the above discussion that the glass slab is anti-reflection coated on both sides.)
Although energy exchange between EM fields and material media is a commonplace
occurrence, it is not necessary for momentum transfers to be accompanied by exchanges of
energy. The reflection of a light pulse from a perfectly conducting mirror, for example, results in
a reversal of the momentum of the light pulse. Conservation of linear momentum then dictates
that the mirror must pick up twice the momentum of the incident pulse. If the mirror is massive
(e.g., fastened to the Earth), its kinetic energy will be negligible, even when its momentum is
substantial. If, however, the kinetic energy of the mirror needs to be accounted for, it can be
found in the reduced energy of the reflected light manifest in its Doppler-shifted frequency. As
another example, consider the case of a perfectly absorbing medium illuminated by a light pulse.
The absorption of the pulse transfers to the material medium both the energy and the momentum
of the light pulse. The acquired momentum causes the absorber to move forward, with a velocity
that may be large or small, depending on the mass of the absorber. Whatever kinetic energy the
material medium may need to maintain its momentum must come from the initial EM energy of
the light pulse; the more massive the absorber, of course, the smaller will be its kinetic energy.
Needless to say, the remaining energy of the pulse will be converted to heat upon its absorption.
This chapter provides a framework for the analysis of certain problems involving
momentum exchange between EM fields and material media. We confine our attention to simple
examples that require only an application of the conservation laws. In principle, of course, many
complex problems can be solved by first determining the solution of Maxwells equations,
namely, E(r, t) and H(r, t ), given the source distributions
free
(r, t ), J
free
(r, t), P(r, t ), and M(r, t),
then computing the resulting force and torque densities using Eqs. (35) and (37) of Chapter 2. In
practice, this recipe is not always easy to implement because the fields and the sources are
usually interdependent. For example, the sources may be induced by the fields, and the
governing constitutive relations may be nonlinear, non-local, hysteretic, etc. Moreover, the EM
force and torque exerted on the material media may produce motion (e.g., translation, rotation,
deformation), which will result in changes to the spatio-temporal distribution of the sources.
These changes will, in turn, modify the distribution of fields and forces. Also, absorption of EM
energy could give rise to thermal effects, which modify the optical, thermal, and mechanical
properties of the host media. All in all, the solution of electromagnetic problems cannot, in
general, be reduced to solving Maxwells equations, but may entail the simultaneous solution of
equations involving mechanical motion, thermodynamics, elasticity, and hydrodynamics as well.
In the present chapter, it is not our intent to delve into such complex problems. Nevertheless, by
providing simple examples of situations in which the mechanical effects of light play an
important role, we hope to inspire the reader to pursue this fascinating subject to greater depth.
The references listed at the end of the chapter should provide adequate entry points into the rich
and often controversial history of efforts, both theoretical and experimental, directed at
unraveling the nature of electromagnetic momentum in its interactions with material media.
10.2. Brief review of classical mechanics. In Newtonian mechanics a point-particle of mass m
and velocity V has linear momentum p = mV and kinetic energy E =mV
2
. Einsteins special
theory of relativity modifies these formulas by assigning to the particle the relativistic
momentum p =mV/ 1V
2
/c
2
and a total energy (that is, rest energy plus kinetic energy)
224 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur

E =mc
2
/ 1V
2
/c
2
. A Taylor series expansion in powers of V/c of the latter expression yields, to
lowest order in V/c, the approximate formula E mc
2
+mV
2
. In other words, at non-relativistic
velocities, where the higher-order terms in V/c can safely be ignored, the energy of a particle of
mass m and velocity V is the sum of its rest energy, mc
2
, and its Newtonian kinetic energy,
mV
2
. To maintain the generality of the argument in the following discussions, we shall derive
various formulas using the relativistic expressions of p and E , with the understanding that non-
relativistic variants can be substituted in the formulas whenever the circumstances permit.
Newtons second law, F(t) =dp(t)/dt, relates the force exerted on a point-particle to the
time-rate-of-change of its momentum. Within any given inertial reference frame, the formula is
valid under all circumstances provided, of course, that the relativistic expression is used for the
momentum p of the particle. For a system of N particles, the total momentum is the sum of
individual momenta, while the total force on the system is the sum of all the forces experienced
by each and every individual particle, that is,

1 1 1
d ( ) d ( )
.
( ) ( ); ( ) ( ) ( )
d d
N N N
n
n n
n n n
t t
t t t t t
t t
= = =
= = = =

p p
F F p p F (1)
Newtons third law, sometimes referred to as the law of equal action and reaction, states that
two particles exert equal and opposite forces on each other. Thus, summing up the forces on all
the particles within a given system results in the cancellation of all the internal forces that the
various particles exert upon each other. The total force F(t), therefore, is due only to external
influences on the system. For an isolated system with no external forces acting on the particles
that comprise the system, the total force vanishes and, in accordance with Eq. (1), the total
momentum p(t) of the system becomes time-independent. Under such circumstances, the total
momentum of the system is said to be a conserved entity.
Equality of action and reaction, however, could break down in some situations, for instance,
when the particles are far apart and the finite speed of light is taken into consideration. Suppose,
for example, that two charged particles have been held stationary for a long time, so that the
Coulomb force between them has reached a steady state. If now one of the particles is suddenly
set in motion, that particle will experience a time-varying force due to the fact that the electric
field it sees from the stationary particle is changing with its position. The force experienced by
the immobile particle, however, cannot possibly change until it has received the news of the
movement of the first particle. Conservation of momentum is thus seen to break down in this
example if one insists on counting only the momenta of the particles. It turns out, however, that
the electromagnetic field carries momentum as well. Adding the field momentum to the particle
momenta, while replacing action-at-a-distance between remote particles with local interactions
between the field and the particles, will restore the principle of conservation of momentum for
the complete system of fields-plus-particles.
In an analogous fashion, the principle of conservation of energy must be applied not to the
system of particles alone, but to an isolated system that contains both the particles and the EM
field. It is in their interaction with the EM field that the particles generally gain or lose energy.
Once the concept of instantaneous action-at-a-distance between remote particles has been
abandoned in favor of local interactions between the field and the particles, it becomes necessary
to add the energy of the field to that of the particles in order to restore the principle of
conservation of energy for isolated systems.
Angular momentum is another important property of mechanical systems that can be
generalized to encompass systems containing particles as well as EM fields. Let a point-particle
246 Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 246-254
CHAPTER 11
Plane-wave Propagation in Linear, Homogeneous, Isotropic Media
Exhibiting Temporal as well as Spatial Dispersion
The theory I propose may therefore be called a theory of the Electromagnetic Field because it has to
do with the space in the neighbourhood of the electric or magnetic bodies, and it may be called a
Dynamical Theory, because it assumes that in the space there is matter in motion, by which the
observed electromagnetic phenomena are produced.
James Clerk Maxwell (1831-1879)
Abstract. The reaction of a material body to electromagnetic fields depends not only on the
nature of its individual atoms and molecules which are the seats of electric and magnetic
dipoles but also on the classical as well as quantum-mechanical interactions among its various
atomic and molecular constituents. The interatomic interactions may be short-ranged, such as
those associated with exchange-coupling among the near neighbors, or they could be medium-
range or even long-range, such as those mediated by conduction electrons. As useful and
powerful as the Lorentz oscillator model of Chapter 6 has been in many practical applications,
its main shortcoming is that it represents dipoles that respond to nothing but local
electromagnetic fields. When the states of adjacent dipoles, or those of dipoles that are further
removed, happen to have a direct impact on the behavior of a given electric or magnetic dipole,
the Lorentz oscillator model must be modified to account for such interactions. In this chapter
we explore the consequences of nearest-neighbor couplings among the electric dipoles of a
homogeneous, linear, isotropic medium. It is these interactions that lead to spatial dispersion.
11.1. Introduction. In a linear, homogeneous, isotropic medium, when the electric susceptibility

e
and/or the magnetic susceptibility

m
happen to be functions of the frequency , the medium
is said to be temporally dispersive. The Lorentz oscillator model provides a good foundation for
the electromagnetic analysis of temporally dispersive materials. The Lorentz oscillators are
essentially mechanical mass-and-spring systems that interact with the local electromagnetic
field(s), carry kinetic energy in their moving mass(es), store potential energy in their spring(s),
and lose energy to their environment by means of an internal friction mechanism. Nothing in the
Lorentz oscillator model, however, provides for the possibility of mechanical interactions among
various dipoles of the medium, be these dipoles adjacent to each other or further removed from
one another.
Spatial dispersion arises from mechanical interactions among dipoles residing in different
locations within the medium. In the presence of both types of dispersion (i.e., spatial and
temporal), the susceptibilities becomes functions of the k-vector as well as the temporal
frequency , written generally as

e
(k, ) and

m
(k, ).
In the simplest possible example of a medium exhibiting spatial dispersion, an electric
dipole p
0
(t) =p
x0
x
^
+p
y0
y
^
+p
z0
z
^
=q[x
0
(t)x
^
+y
0
(t)y
^
+z
0
(t)z
^
], sitting at the origin of a Cartesian
coordinate system, interacts with its nearest-neighbor dipoles located at x
^
, y
^
, and z
^
,
where is the separation distance between adjacent dipoles; see Fig. 1. The dipole qx
0
(t)x
^
then
experiences a force [x
1
(t) x
0
(t)]x
^
from its neighbor to the right ( is the coupling constant), a
similar force [x
2
(t) x
0
(t)]x
^
from its neighbor to the left, another force [x
3
(t) x
0
(t)]x
^
from the
neighbor above, and so on. The sum of these forces, all exerted on p
x0
(t)x
^
, is readily seen to be
expressible as (
2
/q)[
2
p
x
(r, t)]x
^
. Now, the volume occupied by each dipole is
3
, which is
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
Plane-wave Propagation in Linear, Homogeneous Field, Force, Energy and Momentum in Classical Electrodynamics 247

also equal to 1/N, the inverse number density of the dipoles. Since p
x
=P
x

3
, we see that the force
experienced by an x-oriented dipole at (r, t) is (
2
/Nq)[
2
P
x
(r, t)]x
^
. Similar forces act on y- and
z-oriented dipoles as well. To simplify the polarization dynamics equations, it is customary to
introduce a new parameter, =
2
/m, where m is the moving mass of each dipole. The force of
the adjacent neighbors acting on a given dipole is thus given by (m/Nq)
2
P(r, t); the units of
are m
2
/s
2
.

























Fig. 1. Mass-and-spring dipole oscillators (mass =m, charge =q) experience a force from their adjacent dipoles.
In the following sections we restrict attention to linear, homogeneous, isotropic, and non-
magnetic media for which

m
(k, ) =0. We begin in Sec. 2 by deriving the functional dependence
on k and of the electric susceptibility

e
for a medium whose mechanical dipole-dipole
interactions are limited to nearest neighbors, as described in the preceding paragraph. For plane-
wave propagation in such media the dispersion relations are derived in Sec. 3. We proceed to
examine the case of semi-infinite media exhibiting temporal as well as spatial dispersion in the
presence of an obliquely-incident plane-wave arriving from the free space. The case of an s-
polarized incident plane-wave on the flat surface of a semi-infinite, dispersive medium is treated
in Sec. 4, followed by the case of p-polarized incident plane-wave in Sec. 5. Section 6 is devoted
to a discussion of mechanical energy and its storage, propagation, and dissipation in dispersive
media. In particular, we describe a mechanical Poynting vector that governs the propagation of
mechanical energy within such media.
11.2. Mass-and-spring model of polarization exhibiting spatial dispersion. The following
differential equation describes the dynamics of electric dipole oscillators within a linear,
homogeneous, isotropic medium where each dipole, having charges q, behaves as a mass m
attached to a spring with spring constant , has a dynamic friction coefficient , and is coupled
to its nearest neighbors with a coupling coefficient . (The normalized parameters used in the
equation are
o
=/m, =/m, and =
2
/m.)

2
P(r, t)/ t
2
+ P(r, t)/ t +
o
2
P(r, t)
2
P(r, t)= (Nq
2
/m)E(r, t). (1)
z
y
Spring constant >0
Dynamic friction coefficient >0
Near-neighbor interaction coefficient >0
Distance between adjacent dipoles >0
Moving particles mass and charge m>0, q
p
n
(t ) =q[x
n
(t) x
^
+y
n
(t) y
^
+z
n
(t) z
^
]
x
z
0
(t)
2
248 Field, Force, Energy, and Momentum in Classical Electrodynamics Masud Mansuripur

In the above equation, P(r, t) is the polarization density, N is the number density of the dipoles,
and E(r, t) is the local electric field. If happens to be positive, then when an adjacent neighbor
is stretched more than the local dipole, it tends to pull on the local dipole in an attempt to make it
larger as well; when is negative, the neighboring dipoles will have the opposite influence on
each other.
Substituting in Eq. (1) the plane-wave expressions E(r, t) = E
o
exp[i(k r t)] and P(r, t) =
P
o
exp[i(k r t)], then solving the resulting algebraic equation, one finds
P
o
=
o
[(Nq
2
/m
o
)/(k
2
+
o
2

2
i)]E
o
. (2)
Defining the plasma frequency as
p
= Nq
2
/m
o
, the relative susceptibility of the medium is
found to be

e
(k, ) = P
o
/(
o
E
o
) =
p
2
/(k
2
+
o
2

2
i). (3)
With the exception of the new term k
2
appearing in the denominator, the dielectric
susceptibility of Eq. (3) is very much the same as that obtained from the Lorentz oscillator model
of Chapter 6. The new term, however, will turn out to have profound effects on the solutions of
Maxwells equations within spatially dispersive media.
11.3. Dispersion relations. Setting B(r, t) =
o
H(r, t) and solving Maxwells equations for plane-
wave propagation within the homogeneous, isotropic, and linear medium described by the
relative permittivity and permeability functions (k, ) =1+

e
(k, ) and (k, ) =1, leads to the
dispersion relation
k
2
= (/c)
2
[1+

e
(k, )]. (4)
With reference to Fig. 2, This 4
th
order polynomial equation in k can be solved by setting k
x
= 0,
k
y
= (/c)sin, as these components of k are determined by the boundary conditions, when a
plane-wave is incident from the free space onto the surface of the medium at an angle of
incidence within the yz-plane of incidence. We denote the four solutions of Eq. (4) by k
z1
and
k
z2
, where the signs are chosen such that both k
z1
and k
z2
correspond to downward propagating
plane-waves.



















A careful examination of Maxwells equations, however, reveals the possibility of a third
plane-wave co-existing, within the spatially dispersive medium, with the other two obtained from
z
Fig. 2. A plane-wave, having frequency and k-
vector k
i
in the yz-plane, is incident from the
free-space onto a semi-infinite, linear, isotropic,
homogeneous medium that is both temporally
and spatially dispersive. In general, Maxwells
equations allow the possibility of three plane-
waves co-existing within the dispersive medium.
The mode associated with k
0
t
is a longitudinal
electric mode, that is, the magnetic field
associated with this plane-wave identically
vanishes everywhere. Since the only allowed E-
field components of the longitudinal mode are E
y

and E
z
, this mode is not excited with an s-
polarized incident beam.
k
i
k
r

k
0
t

k
1
t

k
2
t

Free-space
(k, ) =1+

e
(k, )
(k, ) =1
y

Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 255-270 255
CHAPTER 12
The Reciprocity Theorem


"God does not care about our mathematical difficulties; He integrates empirically."
Albert Einstein (1879-1955)
Abstract. This chapter provides a simple physical proof of the reciprocity theorem of classical
electrodynamics in the general case of material media that contain linearly polarizable as well as
linearly magnetizable substances. The excitation source is taken to be a point-dipole, either
electric or magnetic, and the measured field at the observation point can be electric or magnetic,
regardless of the nature of the source dipole. The electric and magnetic susceptibility tensors of
the material system may vary from point to point in space, but they cannot be functions of time.
In the case of spatially non-dispersive media, the only other constraint on the local susceptibility
tensors is that they be symmetric at each and every point. The proof is readily extended to
media that exhibit spatial dispersion: For reciprocity to hold, the electric susceptibility tensor

E_ mn
that relates the complex-valued magnitude of the electric dipole at location r
m
to the
strength of the electric field at r
n
must be the transpose of
E_nm
. Similarly, the necessary and
sufficient condition for the magnetic susceptibility tensor is
M_mn
=

T
M_nm
.
12.1. Introduction. The principle of reciprocity in acoustic as well as electromagnetic (EM)
systems was first enunciated by Lord Rayleigh [1]. Soon afterward, H. A. Lorentz and
J. R. Carson extended the concept and provided sound physical and mathematical arguments that
underlie the rigorous proof of the reciprocity theorem [2, 3]. Over the years, the theorem has been
embellished and extended to cover a broader range of possibilities, and to apply with fewer
constraints [4-11]. The basic concept and its proof based on Maxwells macroscopic equations
are discussed in standard textbooks on electromagnetism [12, 13]. For a recent review of
reciprocity in optics, the reader is referred to the comprehensive article by Potton [14].
Roughly speaking, the idea of reciprocity can be stated as follows: In a linear, time-invariant
electromagnetic system subject to certain restrictions, if the source of radiation is placed in
region A and the resulting EM field is monitored in region B, then switching the locations of the
source and the observer will result in the field monitored at A to be intimately related to that
previously measured at B. Examples include:
i) The radiation pattern of an antenna is closely related to its reception pattern.
ii) Thin-film multilayer stacks of metals and dielectrics may have different reflectivities when
illuminated from opposite directions (i.e., front illumination versus rear illumination), but
they always exhibit precisely the same transmissivity at any given angle of incidence,
irrespective of which side of the stack is illuminated [14, 15].
iii) When a diffraction grating is illuminated at an arbitrary angle of incidence, several
diffraction orders usually emerge, each propagating in a different direction and each having a
specific diffraction efficiency. If now the direction of incidence is made to coincide with the
path taken by one of the emergent orders, say, the one coming off at an angle
m
and with a
diffraction efficiency
m
, then one of the newly emerging diffraction orders will follow the
previous path of incidence (in the reverse direction, of course), and will have the same
efficiency
m
[16].
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
256 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur

This chapter presents a simple yet general proof of the reciprocity theorem, which brings out
the essential physics of the phenomenon and clarifies the need for the restrictions under which
the theorem applies. The only mathematical fact needed in our proof is that the product of any
number of matrices, say, M
1
, M
2
, , M
n
, when transposed, will be equal to the product of the
transposed matrices in reverse order, that is,

1 2 3 1 1 3 2 1
( ... ) ... .
T T T T T T
n n n n
M M M M M M M M M M

= (1)
Generally speaking, reciprocity in electrodynamic systems applies when the material media
that surround the source of radiation are linearly polarizable and/or magnetizable. The electric
and magnetic susceptibility tensors of the material system can vary from point to point in space,
but they must be time-independent. When the media are spatially non-dispersive, the only other
constraint on the susceptibility tensors is the requirement of symmetry at each and every point. In
other words, if the polarization density P(r,
o
) and magnetization density M(r,
o
) at a given
point r in space and at a fixed frequency
o
are related to the local electric and magnetic fields,
E(r,
o
) and H(r,
o
), through the following linear relations:
P(r,
o
) =
o

E
(r,
o
)E(r,
o
), (2a)
M(r,
o
) =
o

M
(r,
o
)H(r,
o
), (2b)
then we must have

(r,
o
) =

T
(r,
o
) for electric polarization (subscript E) as well as
magnetization (subscript M). In the above equations,
o
and
o
are the permittivity and
permeability of free space, the system of units employed is MKSA, and both
E
and
M
are
dimensionless entities. We shall impose no other restrictions on
E
and
M
, allowing their
components to be arbitrary complex-valued functions of r and
o
.
The excitation source will be taken to be a stationary, monochromatic point-dipole, either
electric or magnetic, located at r
s
. The measured field at the observation point r
o
will be either
electric or magnetic, regardless of the nature of the source. When the source and the observed
field are of the same type, i.e., both electric or both magnetic, one must use in the reverse path
the same type of source and measure the same type of field as in the forward path. In contrast,
when the source and the observed field are of different types, then, upon reversing the path, one
must switch both the source type and the measured field. For example, if the source at r
s
is an
electric point-dipole, p
o
exp(i
o
t), while the measured field at r
o
is the H-field, then, in the
reverse path, the source placed at r
o
must be a magnetic point-dipole, m
o
exp(i
o
t), and the
field measured at r
s
must be the E-field. (Note that the subscript o used in conjunction with the
observation point r
o
is italicized. This should not be confused with the subscript o used with
the frequency
o
of the oscillations, with the amplitudes p
o
and m
o
of the dipoles, and with the
EM field amplitudes E
o
and H
o
.)
The chapter is organized as follows. In Section 2 we describe the radiation field of an
electric point-dipole in the surrounding free space, expressing the radiated EM fields in Cartesian
coordinates for an arbitrary electric point-dipole p
o
exp(i
o
t) with components along the x-, y-,
and z-axes. The corresponding formulas for the EM fields radiated by a magnetic point-dipole
m
o
exp(i
o
t) are given in Section 3. In Section 4 we prove the reciprocity theorem in the simple
case where the media surrounding the source are electrically polarizable, having a local,
symmetric, time-independent electric susceptibility
E
(r,
o
). The restriction to electrically
polarizable media will be lifted in Section 5, where the media surrounding the source are allowed
The Reciprocity Theorem Field, Force, Energy and Momentum in Classical Electrodynamics 257

to have an electric as well as a magnetic susceptibility,
M
(r,
o
). Section 6 generalizes the results
to spatially-dispersive media, where both susceptibilities will be delocalized. In Section 7 we
remark on the differences between our approach to reciprocity and the conventional methods of
proving the theorem. Section 8 summarizes the results of the chapter and concludes with an
observation regarding the Feld-Tai reciprocity lemma [8, 9].

12.2. Electromagnetic field radiated by an oscillating electric dipole. With reference to Fig. 1,
consider the electric point-dipole
o o
( , ) ( ) exp( i ) ,
z s
t p t = p r r r z located at the fixed point r
s

and oscillating along the z-axis with constant amplitude p
zo
and at fixed frequency
o
. At another
point, say r
o
, the radiated fields given by an exact solution of Maxwells equations are [17, 18]:

o o o o
( , ) ( ) exp( i ) ( /4 )
o o s z
t t p r = = E r E r r (3a)

2 2 2
o o o o

2cos [(1/ ) i( / )] sin [(1/ ) i( / ) ( / ) ] exp[ i ( / )], { } r cr r cr t r c c +



2
o o o o o o

( , ) ( ) exp( i ) ( /4 ) sin [( / ) i ( / )] exp[ i ( / )]. { }


o o s z
t t cp r c cr t r c = = + H r H r r (3b)
In the above equations, c =1/
o

o
is the speed of light in vacuum (with
o
and
o
being the
permeability and permittivity of free-space),
o s
= r r r is the separation vector between the
source dipole and the observation point, | | r = r is the length of r, and (, ) are the polar and
azimuthal coordinates of the observation point as seen from the location r
s
of the dipole.

















Fig. 1. An electric point-dipole located at r
s
and oriented along the z-axis oscillates with frequency

o
and (complex) amplitude p
zo
. The electric and magnetic fields radiated by the dipole are
observed at r
o
, whose position relative to r
s
is given by r =r
o
r
s
. The spherical coordinates of the
separation r between the source and the observer are (r, , ).
Defining
,
o s
x x x = (4a)
,
o s
y y y = (4b)
,
o s
z z z = (4c)
cos ( )/ , z r = (4d)

1
2
2 2
sin [( ) ( ) ] / , x y r = + (4e)
x
y
z
r
s
r
o

(p
zo
,
o
)


Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 271-318 271
Solutions to Selected Problems

Chapter 1
1-7) For all integer values of n we have: i exp[i(2 /2)]. n = + Therefore,
( )
2
i
i
i exp[i(2 /2)] exp[i (2 /2)] exp[ (2 /2)]. n n n = + = + = +

1-8) a) | | exp(i ) | | cos i| | sin . A A A A = = + Therefore,
exp(i ) | | (cos i sin )(cos i sin ) A A = + +
| | [(cos cos sin sin ) i(cos sin sin cos )] A = + +
| | [cos( ) i sin( )] | | exp[i( )]. A A = + + + = +
b)
| | (cos i sin ) | | (cos i sin )(cos i sin )
exp(i ) cos i sin (cos i sin )(cos i sin )
A A A

+ +
= =
+ +


2 2
| | [(cos cos sin sin ) i(sin cos cos sin )]
cos sin
A

+ +
=
+

| | [cos( ) i sin( )] | | exp[i( )]. A A = + =

1-13) a) The cross-product AB has a magnitude equal to the area of the base of the parallele-
piped, and is perpendicular to this base. Dot-multiplication with C then yields the product of the
base area with the projection of C on the normal to the base (i.e., the height of the parallele-
piped). The product of the height and the base area is thus equal to the volume.
b) (AB) C=A (BC) =B (CA) because all three combinations represent the volume of the
same parallelepiped. Note that the volume calculated in this way may have a minus sign, and,
therefore, the order of cross-multiplication is important.

1-17) Let . = C D E Then, from Problem 13, ( ) ( ). A B E = A B E Using Problem 15, we
will have ( ) ( ) ( ) . = = B E B C D B D C B C D Therefore,
( ) ( ) [( ) ( ) ] ( )( ) ( )( ). = = A B C D A B D C B C D A C B D A D B C

1-18) Method 1: Define C=BA. Then from Problem 13, we will have
[ ( )] ( ) ( ) ( ) ( ) ( ) ( ). = = = = A B A B A C B = C B A C A B C A B A B A B
Method 2: Use Problem 15 to write
[ ( )] [( ) ( ) ] ( )( ) ( )( ) = = A B A B A A B A B A B A A B B A B A B

2 2 2 2 2 2 2 2 2
cos sin ( ) ( ). | | | | | | | | | | | | | | = = = = A B A B A B A B A B A B

Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
272 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur
1-25) In Cartesian coordinates we have .
x x y y z z
C S C S C S = + + C S Now, by definition of
the curl operator,
x x
C S is the integral of A(r) around the triangle in the yz-plane. Similarly,
y y
C S is the integral of A(r) around the triangle in the xz-plane, while
z z
C S is the integral of
A(r) around the triangle in the xy-plane. When these three integrals are added together, the
contributions to the integrals by those sides of the triangles which are shared between adjacent
triangles (i.e., the sides of the triangles on the x, y, and z axes) cancel out. What is left then is the
contribution by the sides of the triangle at the base of the pyramid. Therefore, C S is the
integral of A(r) around the base of the pyramid. Note that, for the contributions of adjacent sides
to cancel out, the right-hand rule must be observed for each triangle.
This exercise confirms that the curl at any given point need only be evaluated along three
orthogonal directions. Once C
x
, C
y
, C
z
are known, the loop integral of A(r) around any other
triangle (e.g., the base triangle) can be obtained from the dot-product . C S



Chapter 2
2-17) ( ) [ ( ) ( )] ( ) ( ) ( ) ( ). = = S r E r H r H r E r E r H r
In static situation ( ) ( )/ 0. t = = E r B r
Also,
free free
( ) ( ) ( )/ ( ) t = + = H r J r D r J r in magnetostatics.
Therefore,
free
( ) ( ) ( ) 0. = = S r E r J r

2-21) a) J(r, t) = qV(t)(x)(y) [z
0
t
V(t )dt] z
^
.
b) E (r, t)/ t = E(r, t) J(r, t) = E
o
qV(t)(x)(y) [z
0
t
V(t)dt ].
Integrating over the energy density, let E (t) =

E (r, t)dxdydz denote the energy of the


particle at time t. We will have
dE (t)/dt =(d/dt)

E (r, t)dxdydz =

[E (r, t)/ t]dxdydz = qE


o
V(t).
The change in the kinetic energy of the particle as it moves from z =0 to z =d is thus given by
E =
to
t
1
qE
o
V(t)dt =qE
o

to
t
1
V(t )dt =qE
o
d.
c) The E-field of the system is the sum of the constant E-field between the plates and that of the
point particle, namely, E(r, t) = E
o
Rect(z/d)z
^
+(q/4
o
)[r r
o
(t)]/ | r r
o
(t)|
3
. The E-field energy
density is therefore given by
E (r, t) =
o
E
2
(r, t) =
o
E(r, t) E(r, t)
=
o
{E
o
2
Rect(z/d)+(q/4
o
)
2
/ | r r
o
(t)|
4
+2(qE
o
/4
o
)Rect (z/d)[r r
o
(t)] z
^
/ | r r
o
(t)|
3
}
=
o
E
o
2
Rect (z/d)+(q
2
/32
2

o
){x
2
+y
2
+[z z
o
(t)]
2
}
2

+(qE
o
/4)Rect (z/d)[z z
o
(t)]/ {x
2
+y
2
+[z z
o
(t)]
2
}
3/2
.
Solutions to Selected Problems Field, Force, Energy and Momentum in Classical Electrodynamics 273
The total E-field energy of the system at time t is obtained by integrating the above energy
density over the entire space, that is,
E (t) =

E (r, t)dxdydz =
o
E
o
2

Rect (z/d)dxdydz
+(q
2
/32
2

o
)

(x
2
+y
2
+[z z
o
(t)]
2
)
2
dxdydz
+(qE
o
/4)

Rect (z/d)[z z
o
(t)](x
2
+y
2
+[z z
o
(t)]
2
)
3/2
dxdydz
=
o
E
o
2
Ad +(q
2
/32
2

o
)

(x
2
+y
2
+z
2
)
2
dxdydz
+(qE
o
/4)

Rect (z/d)[z z
o
(t)]dz
0

(r
2
+[z z
o
(t)]
2
)
3/2
2r dr.
The first term in the above expression is the E-field energy of the plates in the absence of the
point particle, which is a constant, independent of the position of the particle in the system. The
second term, although infinite in magnitude because of the zero diameter of the point particle
is constant nonetheless, as it does not depend on the particles position within the system. The
only relevant term, therefore, is the last term, which may be evaluated as follows:
E (t) =constant +qE
o

Rect (z/d)[z z
o
(t)]{(r
2
+[z z
o
(t)]
2
)
1/2
}|
0

dz
=constant +qE
o

Rect(z/d)[z z
o
(t)] | z z
o
(t)|
1
dz
=constant +qE
o
[d 2z
o
(t)].
Thus, as the particle moves from z
o
(t
o
) =0 to z
o
(t
1
) =d, the total E-field energy E (t) of the system,
aside from the additive constant, drops from +qE
o
d to qE
o
d, for a total decline of qE
o
d.
This, of course, is precisely the kinetic energy gained by the particle when it moves from the
bottom plate to the top plate, as evaluated in part (b).

2-22) a) dE (t)/dt = E d
p
(t)/dt = E
o
d
p
z
(t)/dt = E
o
d[
p
(t)cos (t)]/dt
= E
o
cos(t)[d
p
(t)/dt] E
o
p
(t)sin (t)[d (t)/dt]
= E
o
cos(t)[d
p
(t)/dt] +|
p
(t)E| [d(t)/dt]
= E
o
cos(t)[d
p
(t)/dt] +I
o
[d
2
(t)/dt
2
][d(t)/dt]
= E
o
cos(t)[d
p
(t)/dt] +(d/dt){I
o
[d(t)/dt]
2
}
The time-rate-of-change dE (t)/dt of the electromagnetic energy given to the dipole thus has two
components. The first, associated with the changing dipole magnitude, d
p
(t)/dt, appears as the
first term on the right-hand-side of the above equation. The second component, associated with
the increase or decrease in the kinetic energy I
o
[d(t)/dt]
2
of the rotating dipole, constitutes the
second term.
b) The electromagnetic energy exchanged with the dipole in consequence of its changing
orientation angle (t) appears as either an increase or a decrease in the dipoles rotational kinetic
energy. The energy exchanged with the field in the process of changing the dipole strength p(t),
however, causes either an increase or a decrease in the internal energy of the dipole. This energy
A = area of plates
self energy of
point charge =
T
x
(t), component of
torque along x-axis
Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 319-319 319
Appendix A
Vector Identities
= A B B A
( ) + = + A B C A B A C
= A B B A
( ) + = + A B C A B A C
( ) + = + A B C A C B C
( ) ( ) ( ) = = A B C B C A C A B
( ) ( ) ( ) = A B C A C B A B C
( ) ( ) ( )( ) ( )( ) = A B C D A C B D A D B C
( ) 0 = A
( ) 0 =
( ) ( ) = + A A A
( ) ( ) = + A A A
( ) ( ) ( ) = A B B A A B
2
( ) ( ) = A A A
( ) ( ) ( ) ( ) ( ) = + + + A B A B B A A B B A
( ) ( ) ( ) ( ) ( ) = + A B A B B A B A A B
If r is a point in 3-dimensional Euclidean space, then
3
(1/| |) /| | = r r r
3
( /| | ) 4 ( ) = r r r
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 320-320 320
Appendix B
Vector Operations in Cartesian, Cylindrical, and Spherical Coordinates

Cartesian (x, y, z):
x y z


= + + x y z
y
x z


x y z



= + +

y y
x x z z


y z z x x y





= + +



x y z
2 2 2
2 2 2
2
x y z


= + +
Cylindrical (,, z):


= + + z
( )
z


z




= + +
1


( )
z z


z z





= + +


z
2 2
2 2 2
2
/ ( )
z


= + +
Spherical (r,,):
1

sin r r r


= + + r
2
2
sin 1 1
sin sin
( ) ( )
r

r
r r r r




= + +
(sin ) ( )
( ) 1 1 1

sin sin
r r
r
r
r r r r r r






= + +


r
2 2
2 2 2 2 2
2
/ 1 sin / 1
sin sin
( ) ( ) r r
r r r r


= + +
2 2
2 2
( / ) ( )
Notethat
r r r
r r r r



=




Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 321-322 321
Appendix C
Useful Integrals and Identities
1.

dx/(x
2
+a
2
)=/a; Re(a)>0; Gradshteyn & Ryzhik 3.112-2
2.
0

[cos(ax)/(x
2
+b
2
)]dx =[/(2b)]exp(ab); a >0, Re(b)>0; G&R 3.7232
3.
0

{[x sin(ax)]/(x
2
+b
2
)}dx =(/2)exp(ab); a >0, Re(b)>0; G&R 3.7233
/2; a >b0
4.
0

x
1
sin(ax)cos(bx)dx = /4; a =b0 G&R 3.7412
0; b>a 0

a/2; b a >0
5.
0

x
2
sin(ax)sin(bx)dx = G&R 3.7413
b/2; a b>0

6.
0

dx/[cosh(x)+cos(a)]=a/sin(a); G&R 2.4442


7.
0

x
2
dx/[cosh(x)+cos(a)]=a(
2
a
2
)/(3sina); 0<a <; G&R 3.5313
8.
0

x
4
dx/[cosh(x)+cos(a)]=a(
2
a
2
)(7
2
3a
2
)/(15sina); 0<a <; G&R 3.5314
9.
0

exp(ax)cos(bx)dx =a/(a
2
+b
2
); Re(a)>0; G&R 3.8932
10.
0

exp(ax
2
)cos(bx)dx =/a exp[b
2
/(4a)]; Re(a)>0; G&R 3.8964
11.
0

r
1
exp(r)sin(kr)dr =arctan(k/); Re()>0; G&R 3.9481
12.

exp(i x
2
)dx =/ exp(i/4); >0. G&R 3.3223
13.
0

sinx cosx exp(i cosx)dx =2i [sin() cos()]/


2
G&R 3.71511
14.

{[exp(i px)]/(x
2
+
2
)}dx =(/ )exp(|p| ); Re( )>0; G&R 3.3895
15.

{exp(x)/[b+exp(x)]}dx =b
1
/sin(); 0<Re()<1, |arg(b)|<; G&R 3.3119
16.
0

x exp(p
2
x
2
)sin(ax)dx =[a/(4p
3
)]exp[a
2
/(4p
2
)]; G&R 3.9521
17.
0
2
exp(i cosx)dx =2J
0
( ); [J
0
(): Bessel function of 1
st
kind, 0
th
order] G&R 3.9152
18.
0
2
cos(x)exp(i cosx)dx =i2J
1
( ); [J
1
(): Bessel function of 1
st
kind, 1
st
order] G&R 3.9152
19.
0

cos(nx)exp(icosx)dx =i
n
J
n
( ); [J
n
(): Bessel function of 1
st
kind, n
th
order] G&R 3.9152
20.

x
n+1
J
n
(x)dx =x
n+1
J
n+1
(x) G&R 5.521
21.

xJ
0
(x)dx =xJ
1
(x) G&R 5.562
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
Field, Force, Energy and Momentum in Classical Electrodynamics, 2011, 323-339 323


Index

Absorber 42, 43, 44, 223, 242
Absorbing medium (media) 223, 260
Absorption 43, 101, 197, 203, 205, 209, 223, 229, 244
Absorptive media 164
Acceleration 4, 39, 141, 162
Action at a distance 224, 226
Aluminum 242
Ampre, Andr-Marie 27, 222
Ampere 23, 143
Ampres law 27, 37, 53, 60, 89, 90, 129, 130, 131, 135, 293
Ampere-Maxwell law 27, 29
Amplification 101
Angle of incidence 164, 179, 183, 194, 226, 229, 230, 238, 248, 255, 306
Angular momentum 22, 26, 45, 46, 55, 70, 133, 222, 226, 231, 236, 238, 244
density 45, 222, 234, 236
spin 71
Angular velocity 4, 52, 70, 104, 106, 109, 128, 132, 234, 237
Antenna 255
electric-dipole 134
Anti-reflection coating 21, 180, 188, 232, 233
Ashkin, Arthur 239
Atom(s) 24, 28, 140, 142, 246
Atomic 24, 25, 46, 47, 140, 142, 145, 246, 249
Attenuation 215
Azimuthal 29, 132, 234, 235, 257, 275, 293
angle 4,
direction 130
mode number 211
symmetry 90, 91

B-field 32, 34, 35, 38, 54, 57, 59, 83, 85, 103, 113, 120, 122, 125, 132, 228, 259, 292, 302, 312
B-line(s) 33
Balazs thought experiment 44, 45
Battery 58
Berenger, J . P. 205, 208
Bessel function 62, 71, 72, 74, 80, 110, 112, 135, 211, 212, 214, 221, 279
modified 73, 97, 99, 322
of the first (1
st
) kind 71, 73, 81, 87, 96, 99, 119, 137, 212, 279, 280, 288, 321
of the second (2
nd
) kind 71, 97, 99, 119, 212, 288
of the third (3
rd
) kind 71, 97, 99
Biot-Savart law 131, 132, 292, 293
Bohr magneton 71
Born, Max 82
Boundary condition(s) 36, 54, 60, 90, 95, 100, 111, 126, 137, 169, 171, 180, 187, 191, 213, 244, 248, 250, 254, 304
Bound charge 34, 35, 41, 46, 79, 84, 90, 91, 95, 109, 144, 178, 195
electric 35, 46, 103, 109
magnetic 35, 46, 79
Bound current 34, 35, 41, 46, 79, 109, 178, 195
electric 28, 35, 40, 46, 80, 106
magnetic 35, 40, 46
Brewster(s) angle 173, 178, 179, 186, 187, 230, 231, 297, 304, 306
Masud Mansuripur
All rights reserved 2011 Bentham Science Publishers
324 Field, Force, Energy and Momentum in Classical Electrodynamics Masud Mansuripur
Brewster(s) incidence 174, 189
Bullet 42, 43
Canister 137
Capacitance 25, 30, 58, 61, 159,
Capacitor 25, 29, 30, 58, 60, 112, 158, 159, 289, 290
Carrier(s) 39, 40, 101, 102, 144, 148, 163, 181
Carson, J . R. 255, 267, 268, 269, 270
Cartesian coordinates 4, 5, 9, 50, 115, 116, 165, 256, 258, 272
Cavity 107, 111, 137, 144, 193, 203, 205, 209, 210, 293, 294, 310, 311, 312
Center of mass 43, 44, 158, 161, 225, 295, 296
Chess 23, 25
game of 22
Chessboard 22, 23, 25
Circ function 52, 65, 66, 74, 79, 104, 280
Circuit 58, 59, 124, 137
Circular loop 29, 30, 104, 131, 132, 281, 293
Circular polarization 45, 46, 235, 237, 238
Circularly polarized 45, 184, 190, 193, 194, 234, 237, 238
left 45, 213, 221, 237, 238
right 45, 213, 221, 237, 238
Clausius-Mossotti relation 144, 145, 152, 157, 160, 294
Clock 3
Closed loop 8, 12, 20, 28, 32, 33, 34
Closed surface 8, 9, 10, 11, 12, 20, 27, 30, 33, 55, 270,
Clockwise 13, 16, 160, 234
Coaxial cable 61
Coil 130, 293
doughnut-shaped 130
toroidal 293
Collision 42, 62
elastic 42, 62
Complex-amplitude 9, 162, 197, 257, 260
distribution 215
Complex-conjugate 91, 94, 164, 174, 214, 313
Complex vector 5, 6, 7, 9, 10, 11, 13, 14, 16, 83, 84, 163, 164, 166
Conducting (or Conductive) 57, 59, 60, 94, 111, 126, 132, 133, 137, 193, 223
cylinder 61
mirror 193, 223
sphere 127
wire 31, 57, 59, 132, 134
Conductor 53, 103, 127, 136, 137, 180, 181, 182, 191, 221, 228, 238, 243, 244, 268, 300, 315
cylindrical 100, 101, 221
metallic 143, 191
Conduction electron(s) 24, 31, 142, 143, 144, 149, 157, 246
Conductivity 59, 143, 270
tensor 268, 270
Continuous wave (cw) 228, 229
Convolution 77, 117, 155
theorem 77, 155
Conservation 41, 43, 45
of angular momentum 26, 48, 55, 226, 235
of energy 26, 42, 129, 164, 180, 224, 241, 274, 299, 307
of momentum 26, 42, 47, 223, 224, 227, 228, 229, 231, 232, 241, 274
Conservation law(s) 41, 223
Core 136, 191, 203, 204
Dielectric 203, 204, 205, 209, 210, 310, 311
Index Field, Force, Energy and Momentum in Classical Electrodynamics 325
Coupled field 204
Classical Maxwell-Lorentz theory 22, 26
Classical theory 22, 33, 40, 48
Coefficient matrix 259, 261, 296
Continuity equation 24, 30, 38, 50, 94, 103, 109, 117, 128, 138, 170, 177, 180, 191, 217, 221, 253, 288, 292, 301
Copper 23, 143, 144
ion(s) 31
wire 23, 31
Coulomb 23, 25, 143
force 224, 226
gauge 103, 111, 125
Coulombs law 27, 86, 291
Counter-clockwise 4, 8, 13, 16, 234
Coupled oscillator(s) 161, 297
Cross-multiplication 5, 271
Cross-product 165
Curl 3, 10, 11, 18, 20, 27, 28, 30, 32, 35, 80, 83, 131, 264, 272, 312
equation 264, 265
operation 3
operator 10, 11, 18, 20, 33, 272, 293
Current 22, 23, 27, 41, 50, 82, 101, 113, 130, 143, 222, 259, 301
bound 34, 40, 178, 242
carrier 39, 101
constant 28, 33, 88
density 9, 23, 27, 35, 46, 53, 62, 70, 79, 89, 105, 118, 135, 157, 163, 191, 221, 267, 312
free 30, 34, 46, 103, 167, 185, 211
loop 71, 282
oscillating 94, 98, 120
sheet 53, 94, 95, 120, 123, 138
sinusoidal 31
slowly-varying 29, 30, 60
source 97, 119, 124
time-varying 29
uniform 137
Current-carrying 9
sheet 129
wire 29, 117, 119
Current sheet 53, 94, 95, 120, 122, 123, 124, 138, 139
Current-voltage relation 25, 30
Cylinder 31, 32, 55, 61, 78, 89, 90, 98, 100, 105, 110, 130, 136, 217, 218, 221, 236, 237, 289
axis 31, 135, 214
function 212, 214, 217
hollow 55, 89, 98, 104, 110, 111, 130, 136
Cylindrical coordinate(s) 4, 5, 6, 7, 52, 60, 79, 87, 89, 104, 110, 111, 128, 130, 211, 218, 293
Cylindrical wave 101, 110, 119, 120, 215, 220

D-field 26, 27, 29, 36, 54, 126, 189, 209, 291, 294, 301, 309
Damping
coefficient 105, 141, 150, 157, 297
force 161, 162
mechanism 252
parameter 148, 149, 190
Delocalized 144, 257
electron 142
Delta-function 50, 62, 63, 64, 66, 69, 76, 78, 86, 106, 109, 125, 152, 177, 254
Derivative 3, 6, 17, 21, 32, 50, 63, 65, 73, 76, 133, 145, 148, 152, 161, 212, 221, 243, 253, 277

Anda mungkin juga menyukai