EM Lec Notes
EM Lec Notes
1
Updated August 23, 2023
ii Electromagnetic Field Theory
Contents
Preface v
Acknowledgements vi
iii
iv Electromagnetic Field Theory
Preface
This set of lecture notes is from my teaching of ECE 604, Electromagnetic Field Theory, at
ECE, Purdue University, West Lafayette. It is intended for entry level graduate students.
Because different universities have different undergraduate requirements in electromagnetic
field theory, this is a course intended to “level the playing field”. From this point onward,
hopefully, all students will have the fundamental background in electromagnetic field theory
needed to take advance level courses and do research at Purdue.
In developing this course, I have drawn heavily upon knowledge of our predecessors in this
area. Many of the textbooks and papers used, I have listed them in the reference list. Being
a practitioner in this field for over 40 years, I have seen electromagnetic theory impacting
modern technology development unabated. Despite its age, the set of Maxwell’s equations
has endured and continued to be important, from statics to optics, from classical to quantum,
and from nanometer lengthscales to galactic lengthscales. The applications of electromagnetic
technologies have also been tremendous and wide-ranging: from geophysical exploration, re-
mote sensing, bio-sensing, electrical machinery, renewable and clean energy, biomedical en-
gineering, optics and photonics, computer chip design, computer system, quantum computer
designs, quantum communication and many more. Electromagnetic field theory is not every-
thing, but it remains an important component of modern technology developments.
The challenge in teaching this course is on how to teach over 150 years of knowledge in
one semester: Of course this is mission impossible! To do this, we use the traditional wisdom
of engineering education: Distill the knowledge, make it as simple as possible, and teach the
fundamental big ideas in one short semester. Because of this, you may find the flow of the
lectures erratic. Some times, I feel the need to touch on certain big ideas before moving on,
resulting in the choppiness of the curriculum.
Also, in this course, I exploit mathematical “homomorphism” as much as possible to
simplify the teaching. After years of practising in this area, I find that some complex and
advanced concepts become simpler if mathematical homomorphism is exploited between the
advanced concepts and simpler ones. An example of this is on waves in layered media. The
problem is homomorphic to the transmission line problem: Hence, using transmission line
theory, one can simplify the derivations of some complicated formulas.
A large part of modern electromagnetic technologies is based on heuristics. This is some-
thing difficult to teach, as it relies on physical insight and experience. Modern commercial
software has reshaped this landscape: The field of math-physics modeling through numeri-
cal simulations, known as computational electromagnetic (CEM), has made rapid advances
in recent years. Many cut-and-try laboratory experiments, based on heuristics, have been
v
vi Lectures on Electromagnetic Field Theory
Acknowledgements
I like to thank Dan Jiao for sharing her lecture notes in this course, as well as Andy Weiner
for sharing his experience in teaching this course in the beginning. Mark Lundstrom gave me
useful feedback on Chapters 38 and 39 on the quantum theory of light.
I like to thank Dan Jiao for sharing her recent stellar contributions to fast algorithms in
computational electromagnetics. Thanks also to Andy Weiner and Mahdi Hosseini for sharing
their fascinating advances in quantum optics from their research group. I like to thank Erhan
Kudeki of Illinois who always takes an interest on my writing on this subject matter.
Also, I am thankful to Dong-Yeop Na for helping teach part of this course as well as
collaborating on advances in quantum electromagnetics when he was at Purdue. Thanks also
are due the other members of our research team, Boyuan Zhang, Jie Zhu, Ivan Okhmatovskii,
Akila Murugesan, and Sina Vaezi for supporting this course and also to Robert Hsueh-Yung
Chao who took time to read the lecture notes and gave me some very useful feedback. Re-
cently, Chris Ryu and Thomas Roth also gave useful feedback on the last two chapters of
these notes.
Thomas Roth, Dong-Yeop Na, and I recently taught short courses on quantum electromag-
netics at AP-S/URSI, Singapore 2021, and Denver 2022. Some of the materials for the short
courses are factored into the last two chapters of the lecture notes. We have also collaborated
on research in this exciting emerging topic.
Part I
1
Chapter 1
Introduction, Maxwell’s
Equations
In the beginning, this field is either known as the field of electricity and magnetism or the
field of optics. But later, as we shall discuss, these two fields are found to be based on the
same set equations, the Maxwell’s equations. Maxwell’s equations unified these two fields,
and it is common to call the study of electromagnetic theory based on Maxwell’s equations
electromagnetics. It has wide-ranging applications from statics to ultra-violet light in the
present world with impact on many different technologies.
Maxwell’s equations are valid over a vast length scale from subatomic dimensions to
galactic dimensions. Hence, these equations are valid over a vast range of wavelengths:
3
4 Electromagnetic Field Theory
Maxwell’s equations are relativistic invariant in the parlance of special relativity [1]. In
fact, Einstein was motivated with the theory of special relativity in 1905 by Maxwell’s
equations [2]. These equations look the same, irrespective of what inertial reference
frame2 one is in.
Maxwell’s equations are valid in the quantum regime, as it was demonstrated by Paul
Dirac in 1927 [3]. Hence, many methods of calculating the response of a medium to
classical field can be applied in the quantum regime also. When electromagnetic theory
is combined with quantum theory, the field of quantum optics came about. Roy Glauber
won a Nobel prize in 2005 because of his work in this area [4]. Alain Aspect, John F.
Clausser, and Anton Zeilinger won the Nobel prize in 2022 for their works in quantum
optics.
Maxwell’s equations and the pertinent gauge theory has inspired Yang-Mills theory
(1954) [5], which is also known as a generalized electromagnetic theory. Yang-Mills
theory is motivated by differential forms in differential geometry [6]. To quote from
Misner, Thorne, and Wheeler, “Differential forms illuminate electromagnetic theory,
and electromagnetic theory illuminates differential forms.” [7, 8]
Maxwell’s equations are some of the most accurate physical equations that have been
validated by experiments. In 1985, Richard Feynman wrote that electromagnetic theory
had been validated to one part in a billion.3 Now, it has been validated to one part in
a trillion (Aoyama et al, Styer, 2012).4
1 Current lithography process is working with using deep ultra-violet light with a wavelength of 193 nm
Figure 1.1: The impact of electromagnetics in many technologies. The areas in blue are
prevalent areas impacted by electromagnetics some 20 years ago [9], and the areas in
brown are modern emerging areas impacted by electromagnetics.
Figure 1.2: Knowledge grows like a tree. Engineering knowledge and real-world applica-
tions are driven by fundamental knowledge from math and the sciences. At a university,
we do science-based engineering research that can impact wide-ranging real-world ap-
plications. But everyone is equally important in transforming our society. Just like the
parts of the human body: no one can claim that one is more important than the others
(1 Corinthians 12).
6 Electromagnetic Field Theory
Figure 1.3: The electromagnetics spectrum goes from statics to UV. Deep UV and ex-
treme UV (EUV) are now used in nano-lithography, while statics is used in circuit theory
and geophysical exploration.
Figure 1.2 shows how knowledge are driven by basic math and science knowledge. Its
growth is like a tree. Due to the vast ocean of knowledge that we are immersed in, it
is important that we collaborate, especially between engineering and science, to develop
technologies that can transform this world.
laws as opposed to those for optics. This is understandable as the physics of electricity and
magnetism is quite different of the physics of optics as they were known to humans then.
For example, lode stone was known to the ancient Greek and Chinese around 600 BC to
400 BC. Compass was used in China since 200 BC. Static electricity was reported by the
Greek as early as 400 BC. But these curiosities did not make an impact until the age of
telegraphy. The coming about of telegraphy was due to the invention of the voltaic cell or
the galvanic cell in the late 1700’s, by Luigi Galvani and Alesandro Volta [10]. It was soon
discovered that two pieces of wire, connected to a voltaic cell, can transmit information at a
distance.
So by the early 1800’s this possibility had spurred the development of telegraphy. Both
André-Marie Ampére (1823) [11, 12] and Michael Faraday (1838) [13] did experiments to
better understand the properties of electricity and magnetism. And hence, Ampere’s law and
Faraday law were named after them. Kirchhoff voltage and current laws were also developed
in 1845 to help better understand telegraphy [14, 15]. Despite these laws, the technology of
telegraphy was poorly understood. For instance, it was not known as to why the telegraphy
signal was distorted. Ideally, the signal should be a digital signal switching between one’s
and zero’s, but the digital signal lost its shape rapidly along a telegraphy line.5
It was not until 1865 that James Clerk Maxwell [17] put in the missing term in Am-
pere’s law, the displacement current term, only then the mathematical theory for electricity
and magnetism was complete. Ampere’s law is now known as generalized Ampere’s law.
The complete set of equations are now named Maxwell’s equations in honor of James Clerk
Maxwell.6
The rousing success of Maxwell’s theory was that it predicted wave phenomena, as they
have been observed along telegraphy lines. But it was not until 23 years later that Hein-
rich Hertz in 1888 [19] did experiments to prove that electromagnetic field can propagate
through space across a room. This illustrates the difficulty of slow knowledge dissemination
then when new knowledge was discovered. Moreover, from experimental measurement of the
permittivity and permeability of matter, it was decided that electromagnetic wave moves at a
tremendous speed. But the velocity of light has been known for a long while from astronomi-
cal observations (Roemer, 1676) [20]. The interference phenomena in light has been observed
in Newton’s ring (1704) [21]. When these pieces of information were combined together, it
was decided that electricity and magnetism, and optics, are actually governed by the same
physical law or Maxwell’s equations. And optics and electromagnetics are unified into one
field!
5 As a side note, in 1837, Morse invented the Morse code for telegraphy [16]. There were cross pollination
of ideas across the Atlantic ocean despite the distance. In fact, Benjamin Franklin associated lightning with
electricity in the latter part of the 18-th century. Also, notice that electrical machinery was invented in 1832
even though electromagnetic theory was not fully understood.
6 However, it was Oliver Heaviside (1850-1925) who distilled Maxwell’s equations into four equations that
Figure 1.4: A brief history of electromagnetics and optics as depicted in this figure.
In the early days, it was thought that optics is a different discipline from electricity
and magnetism. Then after 1865, the two fields are unified and governed by Maxwell’s
equations.
In Figure 1.4, a brief history of electromagnetics and optics is depicted. In the beginning,
it was thought that electricity and magnetism, and optics were governed by different physical
laws. Low frequency electromagnetics was governed by the understanding of fields and their
interaction with media. Optical phenomena were governed by ray optics, reflection and
refraction of light. But the advent of Maxwell’s equations in 1865 revealed that they can be
unified under electromagnetic theory. Then solving Maxwell’s equations becomes a rewarding
mathematical endeavor.
The photoelectric effect [22, 23], and Planck radiation law [24] point to the fact that
electromagnetic energy is manifested in terms of packets of energy, indicating the corpuscular
nature of light. Each unit of this energy is now known as the photon. A photon carries an
energy packet equal to ℏω, where ω is the angular frequency of the photon and the Planck
constant ℏ = 6.626×10−34 J s, which is a very small constant. Hence, the higher the frequency,
the easier it is to detect this packet of energy, or feel the graininess of electromagnetic energy.
Eventually, in 1927 [3], quantum theory was incorporated into electromagnetics, and the
quantum nature of light gives rise to the field of quantum optics. Recently, even microwave
photons have been measured [25,26]. They are difficult to detect because of the low frequency
of microwave (109 Hz) compared to optics (1015 Hz): a microwave photon carries a packet of
energy about a million times smaller than that of an optical photon.
The progress in nano-fabrication [27] allows one to make optical components that are sub-
Introduction, Maxwell’s Equations 9
optical wavelength as the wavelength of blue light is about 450 nm.7 As a result, interaction
of light with nano-scale optical components requires the solution of Maxwell’s equations in its
full glory, whereas traditionally, ray optics were used to describe many optical phenomena.
In the early days of quantum theory, there were two prevailing theories of quantum in-
terpretation. Quantum measurements were found to be random. In order to explain the
probabilistic nature of quantum measurements, Einstein posited that a random hidden vari-
able caused the random outcome of an experiment. On the other hand, the Copenhagen school
of interpretation led by Niels Bohr, asserted that the outcome of a quantum measurement is
not known until after a measurement [28].
In 1960s, Bell’s theorem (by John Steward Bell) [29] said that an inequality should be
satisfied if Einstein’s hidden variable theory was correct. Otherwise, the violation of the
inequality implies that the Copenhagen school of interpretation should prevail. However,
experimental measurement showed that the inequality was violated, favoring the Copenhagen
school of quantum interpretation [28]. This interpretation says that a quantum state is in a
linear superposition of states before a measurement. But after a measurement, a quantum
state “collapses” to the state that is measured. This implies that quantum information can
be hidden incognito in a quantum state. Hence, for a quantum particle, such as a photon,
its quantum state is unknown until after its measurement. In other words, quantum theory
is “spooky” or “weird”. This also has the profound and beautiful implication that “our
karma is not written on our forehead when we were born, our future is in our own hands!”.
This leads to growing interest in quantum information and quantum communication using
photons. Quantum technology with the use of photons, an electromagnetic quantum particle,
is a subject of growing interest.
d
E · dl = − B · dS Faraday’s Law (1838) (1.2.1)
C
d
dt
S
E: V/m H: A/m
D: C/m2 B: W/m2
I: A Q: C
The first equation above, which is the static form of Faraday’s law also gives rise to Kirchhoff
voltage law. The second equation is the original form of Ampere’s law where displacement
current was ignored. The third and the fourth equations remain unchanged compared to the
time-varying (dynamic) form of Maxwell’s equations.
Figure 1.5: The force between two charges q1 and q2 . The force is repulsive if the two
charges have the same sign.
r2 − r1
r̂12 = , r = |r2 − r1 | (1.3.6)
|r2 − r1 |
Using the definition for unit vector, the force between two charges can also be rewritten as
q1 q2 (r2 − r1 )
f1→2 = , (r1 , r2 are position vectors) (1.3.7)
4πε|r2 − r1 |3
q∆q
f= r̂ (1.3.8)
4πεr2
where r̂ is a unit vector pointing from charge q to the incremental charge ∆q. Then the
electric field E, which is the force per unit charge, is given by
f q
E= = r̂, (V/m) (1.3.9)
△q 4πεr2
Therefore, in general, the electric field E(r) at location r from a point charge q at r′ , using
the definition for r̂, is given by
q(r − r′ )
E(r) = (1.3.10)
4πε|r − r′ |3
12 Electromagnetic Field Theory
Figure 1.6: Emanating E field from an electric point charge as depicted by (1.3.9) and
(1.3.10). The physical meaning is that for a positive charge q
If one knows E due to a point charge, one will know E due to any charge distribution
because any charge distribution can be decomposed into sum of point charges. For instance, if
there are N point charges each with amplitude qi , then by the principle of linear superposition
assuming that linearity holds, the total field produced by these N charges is
N
X qi (r − ri )
E(r) = (1.3.11)
i=1
4πε|r − ri |3
where qi = ϱ(ri )∆Vi is the incremental charge at ri enclosed in the volume ∆Vi . In the
continuum limit, the above becomes
ϱ(r′ )(r − r′ )
E(r) = dV (1.3.12)
V 4πε|r − r′ |3
In other words, the total field, by the principle of linear superposition, is the integral sum-
mation of the contributions from the distributed charge density ϱ(r).
Introduction, Maxwell’s Equations 13
where D is electric flux density with unit C/m2 and D = εE, dS is an incremental surface
at the point on S given by dS n̂ where n̂ is the unit normal pointing outward away from the
surface, and Q is total charge enclosed by the surface S.
Figure 1.7: Electric flux through an incremental surface dS where n̂ is the unit normal,
and D is the electric flux density passing through the incremental surface.
The left-hand side of (1.3.13) represents a surface integral over a closed surface S. To
understand it, one can break the surface into a sum of incremental surfaces ∆Si , with a
local unit normal n̂i associated with it. The surface integral can then be approximated by a
summation
X X
D · dS ≈ Di · n̂i ∆Si = Di · ∆Si (1.3.14)
S i i
where one has defined the incremental surface ∆Si = n̂i ∆Si . In the limit when ∆Si becomes
infinitesimally small, the summation becomes a surface integral.
14 Electromagnetic Field Theory
flux D that it produces on a spherical surface satisfies,
Figure 1.8: Electric flux from a point charge satisfies Gauss’s law.
Even when the shape of the spherical surface S is distorted from a sphere to an arbitrary
shape surface S, it can be shown that the total flux through S is still q. In other words, the
total flux through sufaces S1 and S2 in Figure 1.9 are the same.
This can be appreciated by taking a sliver of the angular sector as shown in Figure 1.10.
Here, ∆S1 and ∆S2 are two incremental surfaces intercepted by this sliver of angular sector.
The amount of flux passing through this incremental surface is given by dS · D = n̂ · D∆S =
n̂ · r̂Dr ∆S. Here, D = r̂Dr is pointing in the r̂ direction. In ∆S1 , n̂ is pointing in the r̂
direction. But in ∆S2 , the incremental area has been enlarged by that n̂ not aligned with
D. But this enlargement is compensated by n̂ · r̂. Also, ∆S2 has grown bigger, but the flux
at ∆S2 has grown smaller by the ratio of (r2 /r1 )2 . Finally, the two fluxes are equal in the
limit that the sliver of angular sector becomes infinitesimally small. This proves the assertion
that the total fluxes through S1 and S2 are equal. Since the total flux from a point charge q
through a closed surface is independent of its shape, but always equal to q, then if we have a
total charge Q which can be expressed as the sum of point charges, namely.
X
Q= qi (1.3.17)
i
Introduction, Maxwell’s Equations 15
Figure 1.9: Same amount of electric flux from a point charge passes through two surfaces
S1 and S2 . This allows Gauss’s law for electric flux to be derivable from Coulomb’s law
for statics.
Then the total flux through a closed surface equals the total charge enclosed by it, which is
the statement of Gauss’s law or Coulomb’s law.
Figure 1.10: When a sliver of angular sector is taken, same amount of electric flux from a
point charge passes through two incremental surfaces ∆S1 and ∆S2 at different distances
from the point charge.
16 Electromagnetic Field Theory
Problem 1-1:
(i) Explain why the electric flux going through ∆S1 and ∆S2 are the same in Figure 1.10.
(ii) Find the field due to a ring of charges with line charge density ϱ C/m as shown in the
figure (courtesy of Ramo, Whinnery, and Van Duzer). Hint: Use symmetry.
Figure 1.11: Electric field of a ring of charge (courtesy of Ramo, Whinnery, and Van
Duzer [33]).
(iii) What is the electric field between coaxial cylinders of unit length in a coaxial cable?
Hint: Use symmetry and cylindrical coordinates to express E = ρ̂Eρ and apply Gauss’s
law.
Introduction, Maxwell’s Equations 17
(iv) The figure shows a sphere of uniform charge density. Find the electric field E as a
function of distance r from the center of the sphere. Hint: Again, use symmetry and
spherical coordinates to express E = r̂Er and apply Gauss’s law.
Figure 1.13: Figure for Problem 1-1 for a sphere with uniform charge density.
(v) Given an infinitely long cylindrical circular wire carrying a DC current I, find the
magnetic field around the wire using symmetry argument, and Ampere’s law.
18 Electromagnetic Field Theory
Chapter 2
Maxwell’s Equations,
Differential Operator Form
Maxwell’s equations were originally written in integral form as has been shown in the previous
lecture. Integral forms have nice physical meaning and can be easily related to experimental
measurements. However, the differential operator form1 can be easily converted to differential
equations or partial differential equations where a whole sleuth of mathematical methods and
numerical methods can be deployed. Therefore, it is prudent to derive the differential operator
form of Maxwell’s equations.
The right-hand side of the above is the total electric flux D that comes out of the surface S.
In the above, assuming that V → ∆V , an infinitesimal volume, then ∇ · D is defined as
D · dS
∆S
∇ · D = lim (2.1.2)
∆V →0 ∆V
The above can be used to derive the explicit for of ∇ · D as shall be shown later. The above
implies that the divergence of the electric flux D, or ∇ · D is given by first computing the
flux coming (or oozing) out of a small volume ∆V surrounded by a small surface ∆S and
1 We caution ourselves not to use the term “differential forms” which has a different meaning used in
19
20 Electromagnetic Field Theory
taking their ratio as shown on the right-hand side of the above. As shall be shown, the ratio
has a finite limit and eventually, we will find a simplified expression for it. We know that if
∆V ≈ 0 or small, then the above implies that,
∆V ∇ · D ≈ D · dS (2.1.3)
∆S
First, we assume that a volume V has been discretized3 into a sum of small cuboids, where
the i-th cuboid has a volume of ∆Vi as shown in Figure 2.1. Then
N
X
V ≈ ∆Vi (2.1.4)
i=1
Figure 2.1: The discretization of a volume V into a sum of small volumes ∆Vi each of
which is a small cuboid. Stair-casing error occurs near the boundary of the volume V
but the error diminishes as ∆Vi → 0.
Figure 2.2: Fluxes from adjacent cuboids cancel each other leaving only the fluxes at the
boundary that remain uncancelled. Please imagine that there is a third dimension of the
cuboids in this picture where it comes out of the paper.
By summing the above over all the cuboids, or over i, one gets
X X
∆Vi ∇ · Di ≈ Di · dSi ≈ D · dS (2.1.6)
i i ∆Si S
The last approximation follows, because it is easily seen that the fluxes out of the inner surfaces
of the cuboids cancel each other, leaving only fluxes flowing out of the cuboids at the edge of
the volume V as explained in Figure 2.2. The right-hand side of the above equation (2.1.6)
becomes a surface integral over the surface S except for the stair-casing approximation (see
Figure 2.1). However, this approximation becomes increasingly good as ∆Vi → 0. Moreover,
the left-hand side of (2.1.6) becomes a volume integral, and we have
dV∇ · D = D · dS (2.1.7)
V S
Figure 2.3: Figure to illustrate the calculation of fluxes from a small cuboid where a
corner of the cuboid is located at (x0 , y0 , z0 ). There is a third z dimension of the cuboid
not shown, and coming out of the paper. Hence, this cuboid, unlike that shown in the
figure, has six faces.
Accounting for the fluxes going through all the six faces, assigning the appropriate signs
in accordance with the fluxes leaving and entering the cuboid, one arrives at the following six
terms
D · dS ≈ −Dx (x0 , y0 , z0 )∆y∆z + Dx (x0 + ∆x, y0 , z0 )∆y∆z
∆S
−Dy (x0 , y0 , z0 )∆x∆z + Dy (x0 , y0 + ∆y, z0 )∆x∆z
−Dz (x0 , y0 , z0 )∆x∆y + Dz (x0 , y0 , z0 + ∆z)∆x∆y (2.1.8)
Factoring out the volume of the cuboid ∆V = ∆x∆y∆z in the above, one gets
D · dS ≈ ∆V {[Dx (x0 + ∆x, . . .) − Dx (x0 , . . .)] /∆x
∆S
+ [Dy (. . . , y0 + ∆y, . . .) − Dy (. . . , y0 , . . .)] /∆y
+ [Dz (. . . , z0 + ∆z) − Dz (. . . , z0 )] /∆z} (2.1.9)
Or that
D · dS ∂Dx ∂Dy ∂Dz
≈ + + (2.1.10)
∆V ∂x ∂y ∂z
In the limit when ∆V → 0, then
D · dS ∂Dx ∂Dy ∂Dz
lim = + + =∇·D (2.1.11)
∆V →0 ∆V ∂x ∂y ∂z
Maxwell’s Equations, Differential Operator Form 23
where (2.1.2) has been used for the above definition of ∇ · D. Furthermore,
∂ ∂ ∂
∇ = x̂ + ŷ + ẑ (2.1.12)
∂x ∂y ∂z
D = x̂Dx + ŷDy + ẑDz (2.1.13)
The above is the definition of the divergence operator in Cartesian coordinates. The diver-
gence operator ∇· has its complicated representations in cylindrical and spherical coordinates,
a subject that we would not delve into in this course. But they can be derived, and are best
looked up at the back of some textbooks on electromagnetics [37].
Consequently, one obtains the Gauss’s divergence theorem given by
dV ∇ · D = D · dS (2.1.14)
V S
The physical meaning of divergence is that if ∇·D ̸= 0 at a point in space, it implies that there
are fluxes oozing or exuding from that point in space [38]. On the other hand, if ∇ · D = 0,
it implies no net flux oozing out from that point in space. In other words, whatever flux that
goes into the point must come out of it. The flux is then termed divergence free. Thus, ∇ · D
is a measure of how much sources or sinks exist for the flux at a point. The sum of these
sources or sinks gives the amount of flux leaving or entering the surface that surrounds the
sources or sinks.
Moreover, if one were to integrate a divergence-free flux over a volume V , and invoking
Gauss’s divergence theorem, one gets
D · dS = 0 (2.1.15)
S
In such a scenerio, whatever flux that enters the surface S must leave it. In other words, what
comes in must go out of the volume V , or that flux is conserved. This is true of incompressible
fluid flow [39], electric flux flow in a source free region, as well as magnetic flux flow, where
the flux is conserved.
24 Electromagnetic Field Theory
Figure 2.4: In an incompressible flux flow, flux is conserved: whatever flux that enters a
volume V must leave the volume V .
∇·D=ϱ (2.1.18)
4 Named after George Gabriel Stokes who lived between 1819 to 1903. In this course, we will use Stokes’
In the above, the contour C is a closed contour, whereas the surface S is not closed.5
From the above definition, we can derive explicit form for ∇ × E. First, applying Stokes’s
theorem to a small surface ∆S, we define a curl operator6 ∇× at a point to be measured as
E · dl
∆C
(∇ × E) · n̂ = lim (2.2.2)
∆S→0 ∆S
In the above, E is a force per unit charge, and ∇ × E is a vector. The above can be viewed as
the definition of ∇ × E. Taking ∆C E · dl as a measure of the torque or rotation of the field
E around a small loop ∆C, the ratio of this rotation to the area of the loop ∆S has a limit
when ∆S becomes infinitesimally small. This ratio is related to (∇ × E) · n̂ where n̂ is a unit
normal to the surface ∆S. As in angular momentum, the direction of the torque is along the
rotation axis of the force.
First, the flat surface S enclosed by C is tessellated (also called meshed, gridded, or
discretized) into sum of small rects (rectangles) as shown in Figure 2.5. Then, (2.2.2), is
applied to one of these small rects to arrive at
where one defines ∆Si = n̂∆S. Next, we sum the above equation over i or over all the small
5 In other words, C has no boundary whereas S has boundary. A closed surface S has no boundary like
rects to arrive at
X X
Ei · dli = ∇ × Ei · ∆Si (2.2.4)
i ∆Ci i
Again, on the left-hand side of the above, all the contour integrals over the small rects cancel
each other internal to S save for those on the boundary. In the limit when ∆Si → 0, the
left-hand side becomes a contour integral over the larger contour C, and the right-hand side
becomes a surface integral over S. One arrives at Stokes’s theorem, which is
E · dl = (∇ × E) · dS (2.2.5)
C S
Figure 2.6: We approximate the integration over a small rect using this figure. There
are four edges to this small rect.
As of this point, ∇ × E is not defined explicitly. Hence, we next need to prove the details
of definition (2.2.2) using Figure 2.6. Performing the integral over the small rect, one gets
We have picked the normal to the incremental surface ∆S to be ẑ in the above example,
and hence, the above gives rise to the identity that
∆S
E · dl ∂ ∂
lim = Ey − Ex = ẑ · ∇ × E (2.2.7)
∆S→0 ∆S ∂x ∂y
Picking different ∆S with different orientations and normals n̂ where n̂ = x̂ or n̂ = ŷ, one
gets
∂ ∂
Ez − Ey = x̂ · ∇ × E (2.2.8)
∂y ∂z
∂ ∂
Ex − Ez = ŷ · ∇ × E (2.2.9)
∂z ∂x
∂ ∂ ∂ ∂
∇ × E = x̂ Ez − Ey + ŷ Ex − Ez
∂y ∂z ∂z ∂x
∂ ∂
+ẑ Ey − Ex (2.2.10)
∂x ∂y
where
∂ ∂ ∂
∇ = x̂ + ŷ + ẑ (2.2.11)
∂x ∂y ∂z
The curl operator ∇× is a measure of the rotation, the torque, or the circulation of a field at
a point in space.7 On the other hand, ∆C E · dl is a measure of the circulation of the field E
around the loop formed by C. To see if a field has a non-zero curl, one can imagine a paddle
wheel placed in such a field. If the field is uniform, the paddle wheel will not rotate implying
that the curl of a uniform field is zero. However, if the field is varying in the z direction, then
the paddle wheel will rotate, implying that a non-uniform field has non-zero curl.
7 In many old textbook, the notation “rot” is still used for the curl or ∇× operator.
28 Electromagnetic Field Theory
Figure 2.7: A paddle wheel can be used to test if a field has zero or non-zero curl as
explained in the text.
Again, the curl operator has its complicated representations in other coordinate systems
like cylindrical or spherical coordinates, a subject that will not be discussed in detail here [37].
It is to be noted that our proof of the Stokes’s theorem is for a flat open surface S, and
not for a general curved open surface. Since all curved surfaces can be tessellated into a union
of flat triangular surfaces according to the tiling theory of simplices,8 the generalization of
the above proof to curved surface is straightforward. An example of such a triangulation of
a curved surface into a union of flat triangular surfaces is shown in Figure 2.8.
Figure 2.8: An arbitrary curved surface can be triangulated with flat triangular patches,
called simplices. The triangulation can be made arbitrarily accurate by making the
patches arbitrarily small.
8 It says that any curve in 1D can be approximated by union of line segments, a 2D surface can be
approximated by union of triangles, while a 3D volume can be approximated by union of tetrahedrons. Line
segments, triangles, and tetrahedrons are simplices in 1D, 2D, and 3D.
Maxwell’s Equations, Differential Operator Form 29
Assuming that the surface S is not time varying, one can take the time derivative into the
integrand and rewrite the above as
∂
E · dl = − B · dS (2.2.13)
C S ∂t
In the above, d/dt becomes ∂/∂t inside the integrand since B = B(r, t) is a multivariable
function. One can replace the left-hand side with the use of Stokes’ theorem to arrive at
∂
∇ × E · dS = − B · dS (2.2.14)
S S ∂t
In the above, dS is an arbitraty elemental surface which can be made very small. Then the
integral can be removed, and one has
∂
∇ × E(r, t) = − B(r, t) (2.2.15)
∂t
The above is Faraday’s law in differential operator form.
In the static limit, ∂B
∂t = 0, giving
∇×E=0 (2.2.16)
∂B
∇ × E(r, t) = − (r, t) (2.3.1)
∂t
∂D
∇ × H(r, t) = (r, t) + J(r, t) (2.3.2)
∂t
∇ · D(r, t) = ϱ(r, t) (2.3.3)
∇ · B(r, t) = 0 (2.3.4)
These equations are point-wise relations as they relate the left-hand side and right-hand side
field values at a given point in space. Moreover, they are not independent of each other. For
9 Faraday’s law is experimentally motivated. Michael Faraday (1791-1867) was an extraordinary exper-
imentalist who documented this law with meticulous care. It was only decades later that a mathematical
description of this law was arrived at.
30 Electromagnetic Field Theory
instance, one can take the divergence of the first equation (2.3.1), making use of the vector
identity that ∇ · (∇ × E) = 0, one gets
∂∇ · B
− = 0 → ∇ · B = constant (2.3.5)
∂t
This constant corresponds to magnetic charges, and since they have not been experimentally
observed, one can set the constant to zero. Thus the fourth of Maxwell’s equations, (2.3.4),
follows from the first (2.3.1).
Similarly, by taking the divergence of the second equation (2.3.2), and making use of the
current continuity equation that
∂ϱ
∇·J+ =0 (2.3.6)
∂t
one can obtain the second last equation (2.3.3). Notice that in (2.3.3), the charge density ϱ
can be time-varying, whereas in the previous lecture, we have “derived” this equation from
Coulomb’s law using electrostatic theory.
The above logic follows if ∂/∂t ̸= 0, and is not valid for static case. In other words, for
statics, the third and the fourth equations are not derivable from the first two. Hence all
four Maxwell’s equations are needed for static problems. For electrodynamic problems, only
solving the first two suffices.
Something is amiss in the above. If J is known, then solving the first two equations implies
solving for four vector unknowns, E, H, B, D, which has 12 scalar unknowns. But there are
only two vector equations or 6 scalar equations in the first two equations. Thus, one needs
more equations. These are provided by the constitutive relations that we shall discuss next.
It is to be noted that when James Clerk Maxwell first wrote his equations down, it was in
many equations and very difficult to digest [17,41,42]. It was an eccentric English genius
Oliver Heaviside, an electrical engineer by training, together with the Maxwellians [43],
who distilled those equations into their present form found in textbooks. Putatively,
most cannot read Maxwell’s treatise [41] beyond the first 50 pages [18].
Maxwell wrote many poems in his short lifespan (1831-1879) and they can be found
at [44].
Also, the ancestor of James Clerk Maxwell married from the Clerk family into the
Maxwell family. One of the conditions of marriage was that all the descendants of the
Clerk family should adopt the family name Clerk Maxwell. That was why Maxwell was
addressed as Professor Clerk Maxwell in his papers.
Maxwell’s Equations, Differential Operator Form 31
Problem 2-1:
(i) By going through proper flux counting, show that (2.1.9) is valid.
(ii) By going through the math carefully, starting from (2.2.6), show that (2.2.10) is correct.
(iii) Explain why Stokes’ theorem can be generalized to curved surfaces.
(iv) In Section 2.3 of Chapter 2, show that for the four Maxwell’s equations, equations
(2.3.3) and (2.3.4) are derivable from the first two Maxwell’s equations.
(v) Explain why this derivation is not valid for static electromagnetic fields.
(vi) By converting the current continuity equation into integral form, explain why it is the
same as charge conservation.
32 Electromagnetic Field Theory
Bibliography
33
34 Electromagnetic Field Theory
[13] B. Jones and M. Faraday, The life and letters of Faraday. Cambridge University Press,
2010, vol. 2.
[14] G. Kirchhoff, “Ueber die auflösung der gleichungen, auf welche man bei der untersuchung
der linearen vertheilung galvanischer ströme geführt wird,” Annalen der Physik, vol. 148,
no. 12, pp. 497–508, 1847.
[15] L. Weinberg, “Kirchhoff’s’ third and fourth laws’,” IRE Transactions on Circuit Theory,
vol. 5, no. 1, pp. 8–30, 1958.
[16] T. Standage, The Victorian Internet: The remarkable story of the telegraph and the
nineteenth century’s online pioneers. Phoenix, 1998.
[18] P. J. Nahin, Oliver Heaviside: sage in solitude. IEEE Press New York, 1987.
[19] H. Hertz, “On the finite velocity of propagation of electromagnetic actions,” Electric
Waves, vol. 110, 1888.
[20] M. Romer and I. B. Cohen, “Roemer and the first determination of the velocity of light
(1676),” Isis, vol. 31, no. 2, pp. 327–379, 1940.
[22] A. Arons and M. Peppard, “Einstein’s proposal of the photon concept–a translation of
the Annalen der Physik paper of 1905,” American Journal of Physics, vol. 33, no. 5, pp.
367–374, 1965.
[23] A. Pais, “Einstein and the quantum theory,” Reviews of Modern Physics, vol. 51, no. 4,
p. 863, 1979.
[24] M. Planck, “On the law of distribution of energy in the normal spectrum,” Annalen der
physik, vol. 4, no. 553, p. 1, 1901.
[30] C. Pickover, Archimedes to Hawking: Laws of science and the great minds behind them.
Oxford University Press, 2008.
[31] R. Resnick, J. Walker, and D. Halliday, Fundamentals of physics. John Wiley, 1988.
[32] J. L. De Lagrange, “Recherches d’arithmétique,” Nouveaux Mémoires de l’Académie de
Berlin, 1773.
[33] S. Ramo, J. R. Whinnery, and T. Duzer van, Fields and waves in communication elec-
tronics, Third Edition. John Wiley & Sons, Inc., 1995, also 1965, 1984.
[34] J. A. Kong, Electromagnetic Wave Theory. EMW Publishing, 2008, also 1985.
[35] H. M. Schey, Div, grad, curl, and all that: an informal text on vector calculus. WW
Norton New York, 2005.
[36] R. P. Feynman, R. B. Leighton, and M. Sands, The Feynman lectures on physics, Vols.
I, II, & III: The new millennium edition. Basic books, 2011, also 1963, 2006, vol. 1,2,3.
[40] V. J. Katz, “The history of Stokes’ theorem,” Mathematics Magazine, vol. 52, no. 3, pp.
146–156, 1979.
[41] J. C. Maxwell, A Treatise on Electricity and magnetism. Dover New York, 1954, first
published in 1873, vol. 1 and 2.
[42] A. D. Yaghjian, “Reflections on Maxwell’s treatise,” Progress In Electromagnetics Re-
search, vol. 149, pp. 217–249, 2014.
[43] B. J. Hunt, The maxwellians. Cornell University Press, 2005.
[44] J. C. Maxwell, “Poems of James Clerk Maxwell,” https://mypoeticside.com/poets/
james-clerk-maxwell-poems.
36 Electromagnetic Field Theory
Index
37
38 Electromagnetic Field Theory