You are on page 1of 230

UNIVERSITY OF CAPE TOWN

DEPARTMENT OF ELECTRICAL ENGINEERING


EEE3055F
2012
Module A: Electromagnetic Field Theory

Lecture Notes: Version 2.33 2012

Lecturer: Assoc. Prof. A.J.Wilkinson

Room: 6.09.1

Email: andrew.wilkinson@uct.ac.za

30th May 2012


Feel free to send me an email if you have questions regarding the course,
or would like to arrange a time to see me. This course module has a
web page where marks and notices will be posted (see courses link on
www.ee.uct.ac.za)
This part of EEE3055F consists of 24 lectures (two per week). Check
web site and notice board on 4th floor Menzies for venues.
Final mark for module A = exam (76%) + test1 (10%) + test2(10%)
+ lab (4%) + assignments (6%).
Two class tests, provisional dates to be announced.
There will be one laboratory practical on microwave antenna measure-
ments. You will be required to submit a short report that will count
4% of your total mark.
Computer simulation assignments will be set that will contribute up to
6% of the final mark.
The examination counts 70% of the mark awarded for this module.
There is a sub-minimum of 35% required for each module to pass the
course.
To qualify for a Duly Performed certificate, all assignments and lab
reports must be submitted by the due date. All submissions must
include a signed declaration, declaring that the work is ones own, and
has not been plagiarised.

Course notes will be handed out in class, supplemented by material on the


blackboard. There is no prescribed text, but the following books may be
consulted for supplementary reading:
Fundamentals of Physics, D. Halliday & R. Resnick
Introduction to Electrodynamics, D.J.Griffiths
Fields and Waves in Communication Electronics,
S. Ramo, J.R. Whinnery and T. Van Duzer.

2 AJW, EEE3055F, UCT 2012


Electromagnetism, I.S. Grant & W.R.Phillips
Electromagnetic Wave Propagation & Antennas, S. Cloude
Antenna Theory, C.A. Balanis

Lecture Topics
Lectures will be given on aspects of the following topics:
1. Review of basic static field theory: electrostatics; Gauss law; magne-
tostatics: Gauss law for magnetic fields; Amperes (Oersteds) law
2. Dynamic laws: Faradays law of electromagnetic induction
3. Maxwells equations & displacement current
4. Electromagnetic boundary conditions
5. Relationship to circuit theory an Kirchhoffs laws
6. Radiation and electromagnetic Waves; the radiation mechanism; the
wave equation and solutions
7. Sinusoidal EM waves
8. Plane waves in (i) free space (ii) non conducting dielectrics; polarization
9. Simulation of propagating waves using FDTD method
10. Power flow in electromagnetic waves; electromagnetic safety consider-
ations
11. Reflection and refraction at boundaries
12. Propagation in conducting media and the skin effect
13. Radiation from antennas: Hertzian dipole, wire antennas
14. Thermal Radiation from warm objects

3 AJW, EEE3055F, UCT 2012


Laboratory session
The laboratory session is on radiation from microwave horn antennas. It
covers:
experimentation with 10 GHz horn antennas.
exposure to some microwave components and instrumentation.
measurement of received power as a function of distance.
linear polarization and its effect on the orientation of the antenna.
measurement of the antenna beam patterns using a rotating turntable.
A formal report must be submitted, documenting the experiments and re-
sults obtained.
The laboratory venue is the Microwave Laboratory on the 7th floor of the
Menzies Building.

4 AJW, EEE3055F, UCT 2012


Contents

1 Introduction and course objectives 1-1


1.1 Course aims . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-1
1.2 Course outcomes . . . . . . . . . . . . . . . . . . . . . . . . 1-1
1.3 Examples of the applications of EM field and wave theory: . 1-1
1.4 Prerequisite Knowledge . . . . . . . . . . . . . . . . . . . . . 1-2
1.5 List of Symbols and Units . . . . . . . . . . . . . . . . . . . 1-2

2 The laws of Electromagnetics 2-1


2.1 Forces between charges (Coulombs Law) . . . . . . . . . . . 2-1
2.2 The Electric Field . . . . . . . . . . . . . . . . . . . . . . . . 2-2
2.3 Forces on charges from Magnetic Fields . . . . . . . . . . . . 2-3
2.4 Lorentz Force Law . . . . . . . . . . . . . . . . . . . . . . . 2-5
2.5 Biot-Savart Law . . . . . . . . . . . . . . . . . . . . . . . . . 2-5
2.6 Work and Potential . . . . . . . . . . . . . . . . . . . . . . . 2-7
2.7 Magnetic Vector Potential . . . . . . . . . . . . . . . . . . . 2-9
2.8 Energy stored in an Electrostatic Field . . . . . . . . . . . . 2-10
2.9 Energy stored in a Magnetostatic Field . . . . . . . . . . . . 2-11
2.10 Other useful Vector Quantities . . . . . . . . . . . . . . . . . 2-12
2.11 Fields in Linear Isotropic Media . . . . . . . . . . . . . . . . 2-12
2.12 Electric Fields in Dielectric Media . . . . . . . . . . . . . . . 2-12
2.13 Magnetic fields in Materials . . . . . . . . . . . . . . . . . . 2-14
2.14 Maxwells Equations governing Electromagnetic Behaviour . 2-18
2.15 Gauss Law for Electric Fields . . . . . . . . . . . . . . . . . 2-18
2.16 Gauss Law for Magnetic Fields . . . . . . . . . . . . . . . . 2-23
2.17 Faradays Law of Electromagnetic Induction . . . . . . . . . 2-24
2.18 Amperes Circuital Law . . . . . . . . . . . . . . . . . . . . 2-29
2.19 Maxwells Equations . . . . . . . . . . . . . . . . . . . . . . 2-32

5
3 The Differential Forms of Maxwells Equations 3-1
3.1 The Static EM Equations in Integral and Differential Form . 3-3
3.2 Electromagnetic Equations before Maxwell . . . . . . . . . . 3-10
3.3 Continuity of Charge . . . . . . . . . . . . . . . . . . . . . . 3-11
3.4 Fixing the Problem with Amperes Law . . . . . . . . . . . . 3-13
3.5 Maxwells Differential Equations . . . . . . . . . . . . . . . . 3-14
3.6 Summary Table of Maxwells Equations . . . . . . . . . . . . 3-15

4 Electromagnetic Boundary Conditions 4-1


4.1 Boundary Conditions for the Electric Field . . . . . . . . . . 4-1
4.2 Boundary Conditions for the Magnetic Field . . . . . . . . . 4-5
4.3 Examples of Boundary Conditions . . . . . . . . . . . . . . . 4-9

5 Relationship between Field Theory and Circuit Theory 5-1


5.1 Resistors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5-2
5.2 Capacitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5-5
5.3 Inductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5-6
5.4 Formulas for Practical Coils . . . . . . . . . . . . . . . . . . 5-11
5.5 Mutual Inductance . . . . . . . . . . . . . . . . . . . . . . . 5-15
5.6 Kirchhoffs Voltage Law . . . . . . . . . . . . . . . . . . . . 5-18
5.7 Kirchhoffs Current Law at a Node . . . . . . . . . . . . . . 5-22
5.8 The Relaxation Time of Conducting Materials . . . . . . . . 5-26
5.9 Shielding and The Faraday Cage . . . . . . . . . . . . . . . 5-29
5.10 Twisted Pair Cables . . . . . . . . . . . . . . . . . . . . . . 5-29

6 Electromagnetic Waves 6-1


6.1 Mathematical Description of Travelling Waves . . . . . . . . 6-2
6.2 Wave Equation derived from Maxwells Equations . . . . . . 6-6
6.3 Physical Interpretation of the Radiation Generation and Prop-
agation Mechanism . . . . . . . . . . . . . . . . . . . . . . . 6-9
6.4 Some Additional Notes on Wave Equations . . . . . . . . . . 6-12
6.5 Travelling Waves in a Plucked Guitar String . . . . . . . . . 6-14

7 Sinusoidal Waves 7-1


7.1 Signals as Sums of Sinusoidal Functions . . . . . . . . . . . . 7-1

6 AJW, EEE3055F, UCT 2012


7.2 Real Sinusoidal Travelling Waves . . . . . . . . . . . . . . . 7-2
7.3 Complex Phasor Representation . . . . . . . . . . . . . . . . 7-4

8 Plane Waves 8-1


8.1 Plane wave propagating in z direction . . . . . . . . . . . . . 8-2
8.2 Characteristic Impedance . . . . . . . . . . . . . . . . . . . . 8-4
8.3 Sinusoidal Representations . . . . . . . . . . . . . . . . . . . 8-5
8.4 Plane Wave Propagating in an Arbitrary Direction . . . . . 8-7
8.5 Polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . 8-9

9 Simulating Electromagnetic Waves using the Finite Difference


Time Domain (FDTD) Method [not covered in 2009] 9-1
9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 9-1
9.2 FDTD Solution to Maxwells equations in 1-D . . . . . . . . 9-3
9.3 FDTD Solution to Maxwells equations in 2-D Space . . . . 9-15
9.4 Solving Maxwells Equations in 3D . . . . . . . . . . . . . . 9-22
9.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9-22
9.6 MATLAB CODE . . . . . . . . . . . . . . . . . . . . . . . . 9-23

10 Power Considerations and the Poynting Vector 10-1


10.1 Power in a Sinusoidal Plane Wave . . . . . . . . . . . . . . . 10-3
10.2 Power density from Radiating Antennas . . . . . . . . . . . 10-5
10.3 Power Flow in a Simple Circuit . . . . . . . . . . . . . . . . 10-9
10.4 Electromagnetic safety considerations . . . . . . . . . . . . . 10-14

11 Reflection and Refraction at Boundaries 11-1


11.1 Reflection of Normally Incident Plane Waves from a Perfect
Conductor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11-1
11.2 Transmission Line Analogy . . . . . . . . . . . . . . . . . . . 11-4
11.3 Normally Incident Plane Wave on a Dielectric Interface . . . 11-5
11.4 Reflection and Transmission (refraction) at a dielectric inter-
face - arbitrary incident angle . . . . . . . . . . . . . . . . . 11-6

12 Propagation in Conducting Media and the Skin Depth 12-1


12.1 Wave propagation in a conducting medium . . . . . . . . . . 12-1

7 AJW, EEE3055F, UCT 2012


12.2 Skin Depth . . . . . . . . . . . . . . . . . . . . . . . . . . . 12-5
12.3 Skin Effect in Conducting Wires . . . . . . . . . . . . . . . . 12-7

13 Radiation from Antennas 13-1


13.1 Hertzian Dipole . . . . . . . . . . . . . . . . . . . . . . . . 13-1
13.2 Dipole Antennas . . . . . . . . . . . . . . . . . . . . . . . . 13-1
13.3 Phase array pattern . . . . . . . . . . . . . . . . . . . . . . 13-1
13.4 Aperture Antennas [simplified treatment; lab session] . . . . 13-1
13.5 Discussion on why circuits that are small compared to the
wavelength do not radiate . . . . . . . . . . . . . . . . . . . 13-1

14 Thermal Radiation from Warm Objects 14-1

8 AJW, EEE3055F, UCT 2012


1 Introduction and course objectives

1.1 Course aims


This course aims to provide students with an understanding of electromag-
netic field theory and wave propagation in the context of applications in
electrical engineering.

1.2 Course outcomes


Knowledge and understanding of Maxwells equations in integral and
differential form.
Understanding of the relationship between circuit theory and field the-
ory.
Understanding of electromagnetic wave phenomena: radiation, plane
wave propagation, reflection and refraction.
Exposure to tools and methods for simulating wave propagation
Basic antenna concepts and understanding radiation from wire anten-
nas.

1.3 Examples of the applications of EM field and wave


theory:
Understanding propagation of electromagnetic waves, and their inter-
action with matter.
Design of structures for efficient radiation of EM power in the form of
antennas for communication links and radar applications.

1-1
Predicting the behaviour of electronic circuits, particularly at higher
frequencies, where the dimensions of components approach the wave-
length.
Solving field problems, e.g. magnetic fields for the analysis of motors
and generators
In electrical impedance tomography, an image of object is reconstructed
from impedance measurements. These are obtained by injecting elec-
trical currents and measuring voltages around the boundary; the re-
construction algorithms require modelling the electric fields within the
medium under investigation.

1.4 Prerequisite Knowledge


Previous course covering electrostatics, magnetostatics.
Vector calculus to describe such fields, e.g. concepts such as div, curl,
Gauss theorem, Stokes theorem etc.

1.5 List of Symbols and Units


A - vector potential
B - the magnetic flux density vector in Wb m2 or Tesla T
Bt , Bn - tangential and normal B-field components at a surface
c - speed of propagation in ms1
C - used to label an integration contour
dl - vector used to define elemental section of a contour for contour integra-
tion
dS used to define a surface element for surface integra-
dS - vector dS n
tion
dV - volume element for volume integration

1-2 AJW, EEE3055F, UCT 2012


D - the electric flux density vector in Cm2
E - electric field strength in in NC1 or Vm1
Et , En - tangential and normal E-field components at a surface
E - the electric field intensity vector in NC1 or Vm1
F - force vector in N
H - the magnetic field intensity vector in Am1
J - current density vector in Am2
k - electric field constant k = 1/(4/0)
k - wavenumber in rad m1
n
- unit vector normal to a surface
P - Poynting vector in Wm2
S - used to label a surface for integration
U - energy in joules
UE energy in electric field in J
UM energy in magnetic field in J
I - current in A or Cs1
I - used occasionally to denote power density in Wm2
L - inductance in H.
M - mutual inductance in H.
t - time in s
f - frequency in Hz
G - antenna gain
r - vector denoting position in space

1-3 AJW, EEE3055F, UCT 2012


q, Q - electric charge in Coulombs C
T - period of a wave in s
V - voltage V or JC1
V - used to label a volume for volume integration
v - velocity vector of a particle in ms1
W - work in J
- charge density in Cm3
s - surface charge density in Cm2
- reflection coefficient
- conductivity in Sm1
- permittivity constant. In free space 0 = 8.854 1012 Fm1 (or
C2N1m2)
- permeability; in free space 0 = 4 107 in NA2
- impedance of a medium of propagation in
- wavelength in m
- potential difference or potential relative to zero reference in V or JC1
M - magnetic flux in webers Wb
E - electric flux in coulombs C
- frequency in rad s1

1-4 AJW, EEE3055F, UCT 2012


2 The laws of Electromagnetics
This section contains a review of the relationships covered in the static fields
course PHY210S. This leads on to the incorporation of the dynamic field
laws, including Faradays law and the modified form of Amperes law. These
laws are summarized in an elegant set of equations known as Maxwells
equations.
The fundamental electromagnetic (EM) field quantities are
E the electric field intensity vector in NC1 or Vm1
B the magnetic flux density vector in Wb m2 or Tesla T
These vector field quantities have been introduced to explain and pre-
dict the observed effects that charged particles have on one another in
static and dynamic situations.
EM fields are not directly visible to the eye; but their effects are.
Our eyes are of course sensitive to EM radiation in the optical band,
although we are not observing the field quantities E and B themselves.
Although field quantities are, in general, functions of position (x, y, z)
~
and time t, e.g. E E(x, y, z, t), for compactness, the arguments will
usually be omitted in the written notation. Bold symbols will be used
for vector quantities.

2.1 Forces between charges (Coulombs Law)


Two charged particles q1 and q2, separated by a distance r in free space
(vacuum), experience a force of either attraction or repulsion modelled by
q1 q2
F =k 2
R

2-1
1
where k = 40 is a constant. The permittivity constant 0 = 8.854 1012
Fm1
F Force on q2

R q1 q2
F= 40 r 2
R
R q2

q1

Figure 2.1: Force on charge q2 owing to q1 .

Note that the force is proportional to the charge values and inversely
proportional to r2.
The effect of such forces can be observed and measured.
We consider the force on q2 to be a result of placing q2 in the electric
field surrounding q1 .

If we place the points within a coordinate system at locations P~1 and P~2 ,
the force on charge q2 can be expressed in vector form as
q1 q2
F= R
40R2
where R = R is a unit vector in the direction of the vector R = P~1 + P~2 ,
|R|
pointing from q1 to q2 .

2.2 The Electric Field


The electric field arising from a charge q is defined in terms of the force it
exerts on a test charge q
F q
E= = R
q 40 r2
in units of N C 1.
The test charge q would generate its own field, which acts on q.

2-2 AJW, EEE3055F, UCT 2012


Superposition holds and so the total field at any point is the sum of
the components from all charges, i.e.
N
X
E= Ei
i=1

where Ei results from each of the N charges as illustrated in the fol-


lowing diagram.

Total Electric Field


E3 PN
E= i=1 Ei
E2
q1
P

E1

q2

q3

Figure 2.2: Electric field at point P is the vector sum of components all sources.

2.3 Forces on charges from Magnetic Fields


Magnetic fields were originally introduced to explain the effects of forces
between magnetized materials, and also current carrying conductors. Mag-
netic fields are created by moving charges. The magnetic fields produced by
permanent magnets can be explained in terms of the sum of fields arising
from the electrons orbiting atoms.
Recall, for a long straight wire conductor carrying current I, the sur-
rounding field at a radial distance r from the wire is given by B = 2r
0I
. To
determine the direction of the magnetic field:

2-3 AJW, EEE3055F, UCT 2012


Grasp the conductor with right hand
with thumb pointing in direction of current.
Fingers point in direction of magnetic field

B
r B
I
I
into page

Figure 2.3: Magnetic field surrounding a long straight current-carrying wire.

A charge moving through a magnetic field experiences a force, described


by the vector relationship
F = qv B
where v is the velocity vector of the particle. The directions of the field
quantities are depicted below (for a positive charge):

V (into page)

B
Magnetic force
on moving charge

F = qV B
(down)

Figure 2.4: Force on a charged particle moving through a magnetic field.

The units of B are deduced from the force per coulomb per unit of speed,
i.e. (NC1)/(ms1) = NC1m1s. For convenience the equivalent unit of
tesla T is commonly used (1 T = 1 NC1m1s).

2-4 AJW, EEE3055F, UCT 2012


2.4 Lorentz Force Law
The total electromagnetic force on a charged particle is the sum of forces
resulting from the electric and magnetic fields and is known as the Lorentz
law
F = qE + qv B
A conductor carrying a current experiences both electric and magnetic
forces from the EM fields in its environment. A stationary charged particle
can experience only an electric force, since the term qv B = 0 if v = 0.

2.5 Biot-Savart Law

The Biot-Savart law (1820) relates the current flowing in a wire to the
magnetic field resulting from it. Consider a linear homogeneous medium in
which a wire carries current.
Magnetic Field
(points out of page) B
P

I
r

dl

Figure 2.5: Magnetic field at point P owing to current I.

The contribution to the magnetic field at any point P resulting from an


elemental section of the wire dl carrying current I in the direction described

2-5 AJW, EEE3055F, UCT 2012


by the vector quantity dl is found experimentally to be
Idl sin
dB(r) =
4R2
where is the angle between vector R and dl , and is the permeability of
the medium. The magnetic field at P points out of the page. The field can
be expressed in vector form as

Idl R
dB(r) =
4R2

where R = (rr ) = R is a unit vector, given in terms of r being the position
|rr | |R|

of point P , and r being the position of the current element. The expression
is sometimes written as
Idl (r r )
dB(r) =
4 |r r |3
To calculate the total magnetic field resulting from a section (or entire
circuit loop), we must add contributions from all elemental sections by in-
tegrating along the length of the wire, usually a closed loop
I I
I dl R
B = dB =
4 R2

Homework

Use the Biot-Savart law to show by integration, that the field arising from an
I
infinitely long straight conductor carrying current I is B = 2R where R is
the perpendicular distance from the point of consideration to the conductor.

2-6 AJW, EEE3055F, UCT 2012


2.6 Work and Potential
The energy required to move a test chargeq from point A to point B along
a contour C is given by the contour integral
X
W = lim (Fi cos i )l
l0
i
Z B Z B Z B
= F dl = (qE + qv B) dl = qE dl
A A A

contour C
dl B
F

Figure 2.6: Contour C joining points A and B.

The dot product F dl = F l cos extracts the component of F tan-


gential to the contour C (in the direction of dl) and multiplies it by
the incremental distance l, to yield the incremental amount of work
dW = F dl.
It is interesting to see that no work is done against the magnetic field
because the magnetic force component of F is always perpendicular to
dl, i.e. qv B dl, and hence the dot product is zero (note that Fv
and vkdl hence Fdl). Work is only done against the electric field.
In electrostatics, it is often convenient to introduce the concept of poten-
tial. The potential difference between two points A and B is defined as the
work per Coulomb involved in moving a test charge q from A to B
Z B
W
BA = = E dl
q A

The units of potential difference are JC1 or volts V.


For static fields, this integral is path independent, i.e. electrostatic
fields are conservative.

2-7 AJW, EEE3055F, UCT 2012


It is often convenient to define a reference point (e.g. in circuit analysis)
as the zero potential, and specify potentials relative to that point.
For analysing problems involving point charges, the reference point is
sometimes set at infinity and thus the potential at any point P is
given by Z P
P = E dl

It can be shown (in many standard textbooks) that the potential at a
point P resulting from N point charges is given by
Z P N
X qi
P = E dl =
i=1
40 ri

where ri is the distance from charge qi to point P.


The electric field can be obtained from the gradient of an electrostatic
field,
x
E =
y
= grad =

z
This is a useful relationship, as the approach to finding the electric field
is often approached by first finding the (scalar) potential distribution1,
followed by calculation of the electric field from the gradient of the
potential function.

More generally, for solving dynamic fields problems, potential functions may also be defined. These
are known as retarded potentials and incorporate the wave propagating nature of electromagnetic
fields. Retarded potentials are functions of both position and time. It is often easier to solve a
problem (either analytically or numerically) by first solving for the potential functions, and then
subsequently calculating E and B field quantities. For example, the electric field is found from
A
E =
t
where (r, t) is the retarded electric potential function and A(r, t) is the retarded magnetic vector
potential.

2-8 AJW, EEE3055F, UCT 2012


2.7 Magnetic Vector Potential
In the same way that a scalar potential function is useful in solving electric
field problems, it is possible to define a special potential for solving magnetic
field problems.
Since B = 0, and (F) = 0 F, it follows that B can be expressed
as the curl of another vector, A, i.e.
B=A
As this is analogous to E = grad, the vector A is known as the magnetic
vector potential.
It is interesting to note that the flux through an arbitrary surface S is
related to A by a line integral around the perimeter:
Z Z I
M = B ds = A ds = A dl (by Stokes theorem)
S S

Although we shall not analyse this here, the magnetic vector potential
A can be determined at any point in space r by integrating contributions
from all currents or current densities that exist, i.e. for the case of a current
carrying conductor (see Figure 2.5),
Z
I(r )dl
A(r) =
4R
where dl is the elemental section of wire at position r on the contour, and
R = |r r |.
For a distributed current density within volume V ,
Z
J(r )dV
A(r) =
V 4R
Once A(r) has been obtained, the B field is calculated from:
B=A
The introduction of the magnetic vector potential is useful for solving both
static B field problems as well as for finding the fields surrounding an an-
tenna.

2-9 AJW, EEE3055F, UCT 2012


2.8 Energy stored in an Electrostatic Field

q1 P1

P2

q2
P3

q3

Figure 2.7: Collection of charges.

It can be shown [Griffiths and other texts] that the work required to assemble
a configuration of point charges by moving charges from infinitely far apart
into fixed position, is given by
N
1X
W = i qi
2 i=1
where i is the potential at point Pi resulting from all other charges, exclud-
ing the i th charge itself2. More generally, to assemble a continuous charge
distribution (x, y, z) with corresponding potential field (x, y, z), is found
by integrating elemental volume contributions
Z
1
W = dV
2 all space
An important result is that this expression can be re-written purely in terms
of the electric field [Griffiths] to yield
Z
1
W = 0 E 2dV
2 all space
2
Note: i qi is the work done to bring qi from far away into position Pi against the force from all other
charges, i.e. i qi counts work against qj (j 6= i), but j qj counts work against qi (j 6= i). Thus the
PN
quantity i=1 i qi counts the work in positioning qi against qj twice (for each pair). Hence we must
PN
divide by two, giving W = 12 i=1 i qi .

2-10 AJW, EEE3055F, UCT 2012


The term 12 0E 2 has units of Jm3 and can be thought of as the energy
density in the electrostatic field. This will have some significance in our
later discussion on radiating electromagnetic fields.

2.9 Energy stored in a Magnetostatic Field


To set up a current in a conducting wire coil requires work; the energy
involved can be recovered. The figure below shows an inductor carrying a
steady state current of I amperes. If the switch is opened, the inductor will
force the current through the resistor marked R, with a 1st order exponential
decay. The energy stored in the inductor is dissipated in the resistor.
Open

I L R

Figure 2.8: Energy is stored in the magnetic field of the inductor.

It can be shown [Griffiths and other texts], that the work required to
establish the current is W = 12 LI 2, which in turn, can be expressed purely
in terms of the surrounding magnetic field arising from the current as
Z
1
W = B 2dV
20 all space

and thus we can consider the term 21 0 B 2 as the energy density in Jm3 and
can be thought of as the energy density in the magnetic field. This will have
some significance in our later discussion on radiating electromagnetic fields.

2-11 AJW, EEE3055F, UCT 2012


2.10 Other useful Vector Quantities
For convenience, two additional EM field quantities are introduced:
D = E is called the electric flux density in Cm2. It is also known as the
displacement vector.
H= B

is called the magnetic field intensity in Am1
These quantities are useful for
writing compact expressions of the laws of electromagnetics
handling the case of fields in materials
aiding in conceptualising certain physical relationships e.g. Gauss law
in for electric fields; Amperes circuital law.
The D vector has a close analogy to the concept of fluid flow (hence the
nameflux density. The wordfluxliterally means therate of flow(which
would be in units of litres per second in the case of fluid flow).

2.11 Fields in Linear Isotropic Media


In the aforementioned equations, the effects of the medium in which a field
exists can be conveniently handled by introducing the concepts of permit-
tivity and permeability of a medium.
For the case of a linear isotropic medium, the parameters are related to
the free space values by constants, i.e.
= r 0 where r 1 is the relative permittivity
and similarly = r 0 where r is the relative permeability.

2.12 Electric Fields in Dielectric Media


When a dielectric (non-conductor) is subjected to an electric field, the
molecules deform and polarize forming tiny dipoles. The electric field
from the induced molecular dipoles is in the opposite direction to the ap-
plied electric field. The net effect of the polarization of the material is to

2-12 AJW, EEE3055F, UCT 2012


reduce the net electric field within the material. The polarization effect also
changes the direction of an electric field vector at the interface between two
different dielectric media (e.g. air and glass). This is studied later in the
chapter on boundary conditions.
Eapplied
air
E(z)

z z
Einduced
glass D(z)

air

Figure 2.9: Polarization of molecules caused by an applied electric field.

The E field (normal component) discontinuity arises from the polarization


effect in the dielectric which results in an induced layer of bound charge at
the interface between dielectrics. In the illustration, it can be seen that the
slight polarization of the molecules causes a thin positive layer on the upper
surface of the glass, and a thin negative layer on the bottom surface. An
electric field is induced within the material that is opposite to the applied
field. The total field within the dielectric is the vector sum of the two,
and is hence reduced. A plot of Ez (z) will show a step change at the
interfaces between glass and air. In a linear medium, Einduced Eapplied.
i.e. E = Eapplied kEapplied = (1 k)Eapplied = 1 Eapplied. In this situation,
the quantity D = E with be the same inside and outside of the medium -
hence the D vector is continuous (at least the normal component is) across
a boundary, analogous to fluid flowing.
The table below lists values of r for several materials.

2-13 AJW, EEE3055F, UCT 2012


Medium Dielectric constant r
Vacuum 1
Gases He,Ne,H,Ar,Ni <1.0007
Air (sea level) 1.00059
Air (pressurized 100 atm) 1.05
Water vapour 1.01
Polystyrene 2.5
Paper 3.5
Glass 4-7
Methyl Alcohol 35
Water 80.4

Table 2.1: Some permittivity values

2.13 Magnetic fields in Materials


For nearly all materials the permeability is very close to that of free space,
i.e. for practical problems, 0 , i.e. r 1.00. Iron (Fe) is one notable
exception, for which r is in the thousands or tens of thousands, depending
on the purity of the iron. Coils are wound on iron cores to increase the
magnetic flux density. It should also be noted that behaviour of iron is
nonlinear, and described by the BH curve.

Magnetic Domains
Insertion of a material into a magnetic field will modify the field. For
example, an electromagnet can be made by winding a coil around an iron
bar; the net effect of the presence of the iron bar is to boost the magnetic B
field strength, by a factor as much as 1000. To understand this phenomenon,
one must examine the microscopic structure of the material. The image in
Figure 2.10 [Ramo et al. 3rd ed. p 692] is a highly magnified picture of a
nickel crystal. The boundaries (revealed by an etching technique) shown are
local magnetic domains (typically 1010 to 1012 m3 in volume; or about 0.1-
0.5 mm in diameter) that have a net magnetic field. Under unmagnetized
conditions, the domains point in random directions. If an external magnetic
field is applied and increased with time, initially, the domains with directions
agreeing with the applied field enlarge, shifting the boundary walls of each

2-14 AJW, EEE3055F, UCT 2012


domain. Increasing the applied field causes all domains to rotate to align
with the applied field.

Figure 2.10: Ferromagnetic domains in nickel. [Ramo et al. 3rd ed. p 692]

Permanent Magnets
Subjecting an iron rod to a strong DC magnetic field causes the magnetic
domains to align with the applied field. If the applied field is reduced to
zero, the domains will tend to rotate back to their original positions, but
not entirely. The remaining or residual effect is a net magnetic polarization,
with a resulting magnetic field.
It should be noted that energy is required to rotate the domains. Some
of this energy is lost as heat, and some is stored in the magnetic field.
In a permanent magnet, the total magnetic field is the sum of contribu-
tions from magnetic fields generated by the electron current surrounding
individual atoms. The illustration below show two atoms within the mag-
netized material. The sum of many such current loops can be modelled as
a single current sheet circulating around the boundary of the material as
illustrated on the right (interior current contributions tend to cancel).

2-15 AJW, EEE3055F, UCT 2012


Figure 2.11: Permanent magnet modelled by an equivalent field-generating current sheet.

The B-H Curve


The figure below shows the B-H curve traced out for commercial iron [Grant
& Phillips]. On the x-axis is the applied H field to which the material is
subjected3; on the y-axis is the resulting B field. The H field arises from
current I in a coil of wire, i.e. H I and the resulting field is a non-linear
function of I (the B-H curve exhibits a hysteresis cycle). The permeability
dB
in the linear region (H < 200A/m) is = dH 10000. At H = 250 A/m,
the B field is a little over 1.5 Tesla. As H is increased beyond 250 A/m, the
dB
curve flattens and converges on a line of slope dH = 0 . This is explained
by the fact that once all domains are aligned with the applied H field,
increasing H has no further effect on material, and B = 0 H + constant.
Reducing H to zero leaves the so-called remanence magnetization Br . The
H required to reduce B back to zero is called the coercive force Hc .

3
The applied H field can also be seen as the applied B field divided by 0 , i.e. H = Bapplied /0 . Thus
the total B is Binduced by rotation of domains + 0 H = H where = r 0 .

2-16 AJW, EEE3055F, UCT 2012


B

Tesla
saturation
1.5 slope
0
remanence
1.0 magnetisation Br
0.5 H c
H

linear region
slope 0
saturation r

250 500 A/m

BH curve for commercial iron used to make mains transformers [Grant & Phillips]

Figure 2.12: B-H curve for iron.

In a 50 Hz AC machine or transformer, the iron is driven repeatedly


around the B-H curve 50 times per second. Each cycle results in an energy
loss in the core (and hence heating), proportional to the area within the
S-curve. The hysteresis power loss increases with frequency. If the core is
not driven close to saturation, the area is less, and so are the losses.

Example

Calculate the power loss in a 50 Hz transformer containing 1000 cm3 of iron,


operating along the hysteresis curve in the illustration above.

Solution

By integrating to obtain the area enclosed by the S-curve, the energy loss per
cubic metre is found to be approximately 350 Jm3 in one cycle. The time-
averaged power loss in 1000 cm3 is therefore 50 350 1000 (0.01)3 = 17.5
watts (which is quite high). In practice, the iron core is not driven so far
into saturation, and the area (and power loss) is about 1/6 of the B-H curve
shown above.

2-17 AJW, EEE3055F, UCT 2012


Magnetic Materials for Machines

Magnetic materials are divided according to their properties and applica-


tion.
Hard Magnetic Materials - have a wide B-H curve and can retain
a high degree of magnetization; used for making permanent magnets,
coating the surface of magnetic storage media like stiffy disks and audio
tapes.
e.g. Iron (bonded powder) Br = 0.6T and Hc = 0.0765Am1.
Soft Magnetic Materials - have a narrow B-H curve, low hysteresis
dB
losses, high initial permeability (r )0 dH at H = 0, high maximum
B
permeability (r )max = H , and high saturation Bsat.
e.g. Commercial Iron (99 Fe), (r )0 = 200, (r )max = 6000, Hc = 9000
Am1 and Bsat = 2.16 T.
Pure Iron (99.9 Fe),(r )0 = 25000, (r )max = 350 000, Hc = 100 Am1
and Bsat = 2.16 T.
Note that the above mentioned permeabilities (r ) are relative perme-
abilities i.e. = r 0 .

2.14 Maxwells Equations governing Electromagnetic


Behaviour
Maxwells equations are a set of equations that govern the behaviour of
electromagnetic fields for both static and dynamic situations. Some of these
equations you will have already met in PHY2010S covering electrostatics
and magnetostatics. In this section we shall describe the basic laws and
their physical interpretation.

2.15 Gauss Law for Electric Fields


Electric fields arise from charges. In free space, Gauss law relates a surface
integral of the quantity 0 E dS over a closed surface to the total charge

2-18 AJW, EEE3055F, UCT 2012


therein: I Z
0 E dS = dV
S V
In a mixed medium, the electric field is affected by the presence of dielectric
material, and Gauss law is written as
I Z
(r)E dS = dV
S V
where (r) is the position dependent permittivity. A useful analogy to fluid
flow can be created by defining the quantity D = E , which we call the
flux density arising from the free charge enclosed within S. Bound charge
(which may be found on the surface of a dielectric inteface) is not considered
- it is taken into account in the variation of (r).
Gauss law states that the total electric flux emanating from a closed
surface is equal to the (unbound) charge enclosed by that surface. It is
written as I Z
D dS = dV
S V

b
n b
Dn q
+
closed surface S surrouding charge q.

dS

Figure 2.13: Illustration of Gauss law applied to a single charge.

2-19 AJW, EEE3055F, UCT 2012


The quantity D dS D n b dS is the amount of flux emanating from
a small elemental surface area dS on S.
b is simply the normal component of D at the surface.
Dn
is the charge density4 in Cm3 that is integrated over the enclosed
volume to give the total enclosed charge Q.
The concept of flux from a charge is a bit like thinking of fluid from a
source. The total flow from a closed surface is numerically equal to
the charge contained within.
The law can also be expressed in terms of E by substituting D = E
where = r 0.
The advantage in working with D, apart from the slightly more com-
pact notation in some equations, is that the flux density/flow concept
holds in all types of media as well as free space. Boundary conditions
are specified in terms of the continuity of flux i.e. D dS is the same
on either side of a boundary. In keeping with the fluid flow analogy,
what flows in on one side of an interface between two materials must
flow out on the other. This does not hold for E dS on the interface
between two different dielectrics i.e. E dS is discontinuous on either
side of a dielectric interface. Shrinking dS to an infinitesimal area,
one concludes that the normal component of D is continuous across
a dielectric interface (i.e. at some point on the interface, the normal
component of D does not change) but the normal component of the
E vector does change. The E field (normal component) discontinuity
arises because of the polarization effect in a dielectric which results
in an induced layer of bound charge ob the boundary surface of each
dielectric.
Gauss law can be used to simplify calculations for fields in certain sym-
metrical situations e.g. determining the field surrounding a charged
4

Note: There is one subtlety here: in matter, refers only to free charge i.e. charge that is free to
move or form a current. It excludes the surface charge on dielectrics induced by an electric fields.

2-20 AJW, EEE3055F, UCT 2012


sphere, determining the field surrounding a long line of charge; deter-
mining the field from a flat (capacitor) plate of uniform charge.
Gauss law can in fact be derived from Coulombs force law and visa
versa. As an exercise, derive Coulombs law by applying Gauss law to
a sphere with a point charge located at its centre.

Example: Line of charge

The field surrounding a long line of charge of density l Cm1 can be esti-
mated by enclosing the line in an imaginary Gaussian cylinder, or radius r,
and length l. From symmetry,
H the electric field must point radially outwards.
Gauss law states: S D dS = Qenclosed. Ignoring the small contributions
of flux leaving the areas at the ends of the cylinder, Gauss law becomes
l
2rl Dr = l l from which we obtain Dr = 2r , or Er = 10 Dr = 02r
l
. Both
Dr and Er decay smoothly as a function of r.

Dr (r)

r
Er (r)

Figure 2.14: Application of Gauss law to a long line of charge.

Note: If we placed the line of charge inside a long hollow cylinder of glass
(along its axis), keeping cylindrical symmetry in the problem, then we can
still apply Gausss law to establish the field quantities. In this new situation,
Dr remains unchanged (compared to the previous case) - the D flux passes
through the glass unaffected (it is normal to the glass surface). The Er field
however abruptly decreases on entering the glass (i.e. is discontinuous), i.e.
Er = Dr /(r) has a step change as it enters and then leaves the glass owing
to the step change in (r). This modified example emphasises why is is
better to view the D field as a flowing fluid and not the E field.

2-21 AJW, EEE3055F, UCT 2012


0 Dr (r)
glass
0 r
Er (r)

Figure 2.15: Illustration of Gauss law applied to a line charge within a cycilinder.

Example: Parallel Plate Capacitor

The capacitance of a capacitor is defined as C = Q/V where Q is the


charge on either plate (+Q on one plate, and -Q on the other), and V is
the potential difference between plates. To estimate the capacitance of a
parallel plate structure, with surface area A and small gap d, we must relate
Q to V. Gauss law allows us to relate the electric field strength between the
plates, to the charge on the plates.
Consider first, a single flat plate, area A and containing surface charge Q.
Close to the plate, we can apply Gauss law to a small pillbox with upper
and lower surface of area a (though which the field lines penetrate), and
with sidewalls parallel to the field, as shown. Let Dn be the magnitude of
theH D field normal to the plate.
S D dS = Qenclosed => a Dn + a Dn s a => Dn = s /2 = (Q/A)/2 =
Q/(2A). If two plates are placed a small distance d apart, one charged to Q
and the other to Q, the total field between the plates is D = 2Dz = Q/A.
The field vectors in the gap add constructively, doubling the magnitude, i.e.
D = 2 Dn = Q/A. Above the upper plate (and below the lower plate), the
field vectors add destructively (being of equal magnitude Dn but opposit in
direction), and hence will be zero. The result is illustrated in Figure 2.16.
Thus the capacitance is
Q AD AE A
=C= = =
V Ed Ed d
which is a (should be) familiar result. Two plates, size 5cm 5cm, sepa-
rated by 5mm in air, have a capacitance of about 8.85E120.050.05
0.005 = 4.4pF.

2-22 AJW, EEE3055F, UCT 2012


Think of the physical size of a 5pF capacitor used for electronic circuits -
it is physically much smaller. This small size is made possible by reducing
the gap between plates, as well as the insertion of a dielectric material be-
tween the plates. Much larger capacitances in the micro-farad range can
be created in a compact package by sandwiching an electrolytic dielectric
between two thin metal sheets that are then rolled up like a swiss roll (so-
called electrolytic capacitors). More recently, supercapacitors have be-
come available, which make use an extremely thin dielectric layer to achieve
significantly higher (>1000 times) capacitances per unit volume than com-
mon electrolytic capacitors.
D
surface charge density Field from a
Charged Parallel s plate of positive charge
Plate Capacitor a

d
Gaussian surface
Area A

Field cancels Result of addition of fields from


positive and negative
plates

Field adds D = Q/A


Field cancels

Figure 2.16: Deriving the electric field for a charged parallel plate capacitor.

2.16 Gauss Law for Magnetic Fields


Gauss law for magnetic fields is
I
B dS = 0
S

2-23 AJW, EEE3055F, UCT 2012


The law reflects the fact that no point magnetic charges have been
found. Magnetic field lines arise from moving electric charges (e.g.
current in a piece of wire)
Magnetic flux lines form continuous loops; they do not begin and end
at fixed points in space (like electric field lines on point charges).
Any magnetic flux line that enters closed surface S at some point must
exit somewhere, as illustrated below:

Magnetic field lines


form continuous loops B

Gaussian surface
(closed surface)
North

South

Figure 2.17: Illustration to show that magnetic field lines form continuous loops; hence any
line that enters a close surface must exit it.

2.17 Faradays Law of Electromagnetic Induction


Faraday discovered that a changing magnetic field can induce an electric
effect. Faradays law is stated in integral form in terms of a relationship
between the line integral of the tangential component of the electric field

2-24 AJW, EEE3055F, UCT 2012


around a closed loop and the rate of change of magnetic flux threading the
loop. Stated in integral form
I Z
d dM
E dl = B dS =
dt S dt
R
where M = S B dS is the magnetic flux flowing through the open surface
S, bounded by the integration contour. The law may also be written as
I Z
B
E dl = dS
S t
In evaluating the line integral, the right hand rule must be Hobserved:
if the vector dS is defined to point up as shown above, then E dl is
made anticlockwise around the loop as illustrated below:

dS

Integration
direction

Figure 2.18: Illustration showing the direction of integration when applying Faradays law
to an open surface with surface normal dS. The right hand rule applies.

This implies that for the case where t > 0 and uniform, the electric
M

field vectors point clockwise around the loop as illustrated for the pole
face of an electromagnet shown below.

2-25 AJW, EEE3055F, UCT 2012


Increasing
B

TOP VIEW

E
Case of uniform (changing) E
magnetic field piercing
through flat circular area
e.g. on the pole face of M
an electromagnet. t >0
E
E(r)
Emax

r 1/r
i(t)
Increasing
r
rmax

Figure 2.19: Induced E-field vectors surrounding the pole face of an electromagnet for the
case where the magnetic field is increasing in the direction shown. Note:
although the B-lines must eventually curl back and form continuous loops, we
can regard the flux density for r > rmax as negligible.

H
In electrostatics, E dl = 0 indicating the conservative nature of
electrostatic fields and the fact that the net energy required to move a
test charge around a loop is zero. In general, electric fields are NOT
conservative.
The effect of the induced E field loop can be measured by inserting a
wire ring into a changing magnetic field and observing that a current
flows as a result of the induced electric field pushing the electrons
round the circuit. Note: current density J = E inside/on wire.

2-26 AJW, EEE3055F, UCT 2012


Increasing
B
Current flow
in wire ring

Recall
M
J = E t
>0
Vinduced
I= Rwire

Figure 2.20: Current flows in a wire ring placed on the pole face.

If the ring is broken by a small cut, the net effect is that the charges
pile up on either side of the cut, resulting in a local concentration of
electric field as illustrated below.

E=0
inside conductor
M
t
>0

Figure 2.21: Top view of a wire ring with an air gap. Note: the field inside the metal is zero.
Thus he field from the induced dipole must exactly cancel the Faraday-induced
electric field within the metal - try to draw the two fields.

If a voltmeter is connected as shown in the figure shown below left, the


voltmeter will read a value Vinduced = ddtM . The voltmeter wires extend
the definition of the loop i.e. the boundary of a surface through which
the magnetic flux passes.

2-27 AJW, EEE3055F, UCT 2012


Voltmeter Reading ?

B V

E=0
inside conductor
M
t
>0 M
>0
t

V voltmeter
M
Vmeasured = t

Figure 2.22: Wire ring with voltmeter placed in different positions. Are the readings the
same?

QUESTION: What happens if the voltmeter is moved to the opposite side


as depicted above on the right? What value will the voltmeter read?
ANSWER: virtually zero! Why? Ask yourself how much changing flux will
thread the loop.

2-28 AJW, EEE3055F, UCT 2012


2.18 Amperes Circuital Law
The dynamic form of Amperes law states that the line integral of the tan-
gential component of magnetic field intensity around a closed loop, bounding
an open surface S is equal to the sum of (i) the conduction current flowing
through S, and (ii) the rate of change of electric flux cutting through S,
I Z Z
D
H dl = J dS + dS
S S t
or I Z Z
d
H dl = J dS + D dS
S dt S
R
The term S J dS = Ic is the conduction current and refers to the
physical movement of charge.
In linear conducting media, the current density is proportional to the
electric field strength J = E where is the conductivity constant in
Sm1 (Siemens per metre).
R R dE
The term S D t dS = d
dt S D dS = dt = Id is known as the dis-
placement current as it has the same units of current Ic , although it
does not refer to the physical movement of charge. The displacement
current is the rate of change of electric flux on S.
R
The time varying term S D t
dS was added by Maxwell as a result
of a careful theoretical analysis that revealed that the static form of
Amperes law was incomplete.
D
The t is the displacement current density in units of Am2.
In the next section, we shall examine the dynamic form of Amperes law in
detail and establish a deeper understanding of displacement current.

2.18.1 Physical interpretation of Displacement Current


A classic example for gaining a physical understanding of the concept of
displacement current involves the consideration of Amperes circuital law in
a simple circuit involving a capacitor as depicted in Figure 2.23

2-29 AJW, EEE3055F, UCT 2012


S2 (balloon shaped surface passing
S1 (shaded) between plates)
+
H +
I +
+
+
+
+
C

Figure 2.23: Illustration showing how Amperes static law yields inconsistant results, de-
pending on the definition of the surface S.

Consider applying the magnetostatic form of Amperes circuital law to


the closed loop C, where the current passes through surface S1 in the plane
of the loop. Loosely speaking, the static form of Amperes law says that
the line integral of the tangential component of H around some contour C
equals the conduction current (i.e. physical movement of charge) passing
though a surface S bounded by loop C. For a long section of conductor,
from considerations of symmetry, one concludes that the H field lines are
circles centred on the wire. The magnitude of the circulating field can be
found by applying Amperes law to a circular contour, of radius r from the
wire. From Amperes static law
I
H dl = I = H2r

which implies
I
H= .
2r

2-30 AJW, EEE3055F, UCT 2012


If we consider an alternative choice of S = S2 , which is carefully chosen to
pass between the plates of the capacitor and not cut the wire as illustrated
in the sketch. Clearly the conduction current passing S2 is zero, implying
that the static form of Amperes law is either wrong, or is an incomplete
description.
Maxwell accounted for this apparent paradox, by introducing an addi-
tional term
H which he called the displacement current density. The line
integral H dl around closed loop C bounding R some arbitrary surface S
equals the sumR ofDthe conduction current Ic = S JdS plus the displacement
current Id = S t dS passing through S.
If the conduction current is Ic = I, then between the plates of the ca-
pacitor, an electric field will build up, resulting from the build up of a net
positive charge on one plate, and a negative charge on the other.
We wish to explore the displacement current term
Z Z
D d dE
Id = dS = D dS =
dS t dt S dt
dE
for the case of S = S2. We can apply Gausss law to determine dt and
relate it to other parameters.
Gausss law states that for a closed surface S,
I Z
D dS = dV = Qenclosed
S V
Although S2 is not closed (having an opening at its mouth), most of the
flux is highly concentrated between the plates, and hence the amount of
electric flux passing through surface S2 is approximately E Q where
Q is the charge on the left plate. Only a tiny amount of flux would pass
though the mouth at contour C.
Thus d dt
E
dQ
dt
= Ic, being the conduction current in the wire. If
we shrink the mouth of S2 to be a tiny hole then E Q and hence
Id (through S2 ) I. Thus the displacement current Id flowing through S2
is equal to the conduction current Ic flowing through S1 .
Maxwells form of Amperes law works for any definition of S, notably
for S = S1 , Ic = I and Id 0, which implies
I
H dl = Ic + Id I + 0

2-31 AJW, EEE3055F, UCT 2012


for S = S2 , Ic = 0 and Id dQ
dt = I, which implies
I
H dl = Ic + Id 0 + I

Figure 2.24: The magnetic field ring exists both in the region around the conducting wire,
and also around the displacement current in the gap between the plates.

2.19 Maxwells Equations


The set of four laws together are known as Maxwells equations:
I Z
D dS = dV
S V
I
B dS = 0
S
I Z
d
E dl = B dS
dt S
I Z Z
d
H dl = J dS + D dS
S dt S
In the following chapter, the differentials forms will be derived.

2-32 AJW, EEE3055F, UCT 2012


3 The Differential Forms of Maxwells
Equations
In the previous chapter, we examined the integral forms of Maxwells equa-
tions being:
1. Gauss law for electric fields
I Z
D dS = dV
S V

2. Gauss law for magnetic fields


I
B dS = 0
S

3. Faradays law I Z
d
E dl = B dS
dt S
4. Amperes (modified) law
I Z Z
d
H dl = J dS + D dS
S dt S
In this chapter, we shall show how a set of differential equations can be
derived from Maxwells four equations. These are known as the differential
formsand serve to describe field properties at a point in space and time; this
contrasts with the integral forms, which concern integrals along contours,
over surfaces and over volumes. Both the static laws and the dynamic laws
will be examined. Particulary, we shall also show how the displacement
current term introduced by Maxwell can be derived by consideration of the
differential forms.
The differential forms are found by applying two theorems from vector
calculus that apply to any vector field F = F(x, y, x, t) at any point in
space and time:

3-1
Gauss Divergence Theorem

For any (differentiable) vector field F, and any closed surface S defined in
space enclosing a volume V ,
I Z
F dS = F dV
S V
1
where F = div F is the divergence of the vector defined as
H
F dS
div F = limV 0
V
H
where the quantity F dS is the net flux leaving volume V .
The divergence theorem alows one to express a surface integral (over a
closed surface) as a volume integral.
The physical
R interpretation of the term on the right hand side of the
thereom V F dV is that we are summing up, for each elemental volume
P
dVi within V , the quantity H F dVi , i.e. we are calculating i F dVi .
The quantity F dVi = Si F dS is the flux leaving a volume element dVi
through its surrounding closed surface Si . The flux contributions through
the side walls of adjacent dV s cancel, leaving only the contributions through
theHside walls touching
H the outerRsurface S of the volume V , which we write
as S F dS. Hence S F dS = V F dV .

Stokes Curl Theorem

For any (differentiable) vector field F, and contour C bordering an open


surface S,
I Z
F dl = F dS
C S
where F = curl F is a vector2 that measures rotation in the vector
field. The curl F vector points along the axis or rotation, and has mag-

x Fx
Fy
1
In Cartesian coordinates, F =
y
Fy = Fx
x + y + Fz
z .
Fz
z
F x y
z    
x x 
Fy = x = Fz Fy Fy
2
and F = y
y

z y z x
+ Fx
z Fz
x y + x Fx
y z
Fz Fx Fy Fz
z

3-2 AJW, EEE3055F, UCT 2012


H
Fdl
nitude equal to limS0 S where S is an elemental area in the plane
perpendicular to the axis of rotation; the direction of integration along the
contour is defined by the right hand rule applied to vector dF.
If one imagines suspending a ball at some point in a fluid velocity vector
field, with non-zero curl at the point, then the ball will tend to rotate about
the curl F axis, and in a direction defined by the right hand rule applied to
vector curl F.
Stokes theorem relates a contour integral around the boundary of an open
surface to a surface integral over the surface. R
The physical interpretation of the right hand side S F dS, is that we
are summing up for each surface element dSi on S, the quantity H F dSi,
P
i.e we are calculating i F dSi . The quantity F dSi = Ci F dl
is the line integral on the contour Ci bordering elemental area dSi . The
contributions of the side edges of adjacent dSis cancel, leaving only the
contributions
H from the outer
H boundary R C bordering the surface S, which we
write as C F dl. Hence C F dl = S F dS.

3.1 The Static EM Equations in Integral and Differential


Form
The static forms of Maxwells equations in integral form are:
I Z
D dS = dV
S V
I
B dS = 0
S
I
E dl = 0
I Z
H dl = J dS
S
where D = E and B = H.
The term static refers to the fact that the field quantities are not
time varying, i.e. E D H B
t = 0, t = 0, t = 0 and t = 0.

3-3 AJW, EEE3055F, UCT 2012


The differential forms are derived by employing the divergence theorem and
Stokes theorem as follows:
H R
Gausss law for Helectric fields
R states S D dS = RV dV . From theR diver-
gence theorem, D dS = V D dV , hence V D dV = V dV .
Since this must be true for any closed surface, independent of its shape or
size, we conclude that D = .
H
Gausss law
H for magnetic
R fields states S B dS = 0. From the divergence
theorem, S B dS = V BdV . Again, since the closed surface can be
arbitrarily chosen, we conclude that B = 0.
H
The conservative property ofH electrostatic
R fields states RE dl = 0. Ap-
plying Stokes curl theorem, E dl = S E dS. Since S E dS = 0
for any surface, we conclude that E = 0.
H R
The static form of Amperes
H law Rstates that H dl = SJ
R dS. Applying
RStokes curl theorem, H dl = S H dS, we obtain S H dS =
S J dS for any surface, from which we conclude that H = J.

Thus the four static equations in differential form are

D=

B=0
E=0
H=J
These differential forms are used to solve for field distributions in static
situations. Like all differential equations, the solutions require the specifi-
cation of boundary conditions.

3-4 AJW, EEE3055F, UCT 2012


Examples of Static Field Problems
Example 1

In an electrostatic situation, we may wish to know how to calculate the field


quantities E(x, y, z) and (x, y, z) in space surrounding two metal plates at
fixed potentials relative to ground. As a more specific example, consider the
fields surrounding a charged capacitor. The potentials on the metal plates
(measured relative to a chosen reference ground) are the so called boundary
conditions.
A differential equation describing the surrounding electric field can be
found from Gauss law, D = , and observing that the density is zero
in the surrounding space (there may and will however be a thin layer of
surface charge on the plates themselves).
Substituting = 0 and D = E into Gauss law we get

(E) = = 0

The above relationship describes a relationship between the three compo-


nents of E. Since E = grad , we substitute into the above
equation to obtain a differential equation in terms of the single scalar func-
tion (x, y, z):
() = 0
or      

+ + =0
x x y y z z
This is a second order partial differential equation that may be solved using
numerical methods like the finite element method, together with the spec-
ified boundary conditions. An equation in terms of the potential is also
convenient as the boundary condition(s) are given in terms of the values of
the potential on the plates of the capacitor.
If the surrounding medium homogeneous, is a constant, the equation
reduces to a well-known equation called Laplaces equation

= 0

3-5 AJW, EEE3055F, UCT 2012


Laplaces equation is sometimes written in compact form as 2 = 0. Ex-
panding in Cartesian coordinates, the differential equation is

x x x x
=
y

y
(x, y, z) =
y

y


z z z z
or
2 2 2
+ 2 + 2 =0
x2 y z
Having obtained (x, y, z) using a numerical method, one can obtain E by
differentiation since
E = .

Example 2

In electrical resistance tomography3 (ERT), the goal is to reconstruct an im-


age of the electrical conductivity (x, y, z) within a tank from measurements
taken at electrodes placed around the circumference. ERT is applicable in
situation where the medium under investigation can be modelled as a resis-
tive medium. Applications include imaging inside mixing tanks and pipes
containing fluids. Any fluid containing dissolved salts is conductive.
Currents are injected between a pair of electrodes, and voltages are mea-
sured at all nodes. The reconstruction algorithm needs to be able to predict
the voltages at all electrodes around the boundary, given either specified po-
tentials or currents at the two driving electrodes. This relationship between
injected and measured quantities called the forward model. Reconstruc-
tion algorithms typically solve the so-called inverse problem by iteratively
finding a solution (i.e. the conductivity distribution (x, y, z)) for which
the voltages predicted by the forward model closely match the measured
voltages.

3
Tomography refers to a method by which an image of the interior of an object is obtained via measure-
ments taken around the boundary. Tomography can be implemented via the injection and measurement
of EM waves such as X-rays or microwaves, or using ultra-sound, or via low frequency impedance meth-
ods. The word tomography comes from the Greek words tomos which means part or section,
and the word graphein (to write). In some types of tomography (e.g. X-ray tomography), the image
is made up of slices taken at different angles and combined using an algorithm.

3-6 AJW, EEE3055F, UCT 2012


Electrical Resistance Tomography
v2

v1 v3

I J v4 Measure voltages
v4 v5
v8 between pairs

Inject
current
v7 v5
I
v6

Figure 3.1: Illustration of current injection in electrical resistance tomography (ERT). Volt-
age measurements are made between pairs of electrodes.

Stated mathematically, the goal is to find the distribution (x, y, z), de-
scribed by a set of parameters {} , which minimizes a cost function C({})
derived sum of the squares of the difference, i.e. minimise as a function of
{},
XN
C({}) = (Vpred. (n, {}) Vmeas.(n))2
n=1
where Vmeas. (n) is the nth voltage measurement, and Vpred. (n, {}) is the
voltage predicted by a simulator. In practice, a smoothness constraint on
(x, y, z) is usually applied as well. Minimizing C({}) is a multi-variable
optimization problem for which iterative solutions have been developed.
The forward model is a partial differential equation that must be solved
numerically to get . To construct the differential equation to be solved, we
start with the continuity equation4, J = 0, and substitute the relations
J = E and E = . The resulting differential equation is = 0.
For the special case where (x, y, z) = constant, the equation reduces to
Laplaces equation = 2 = 0.
The value of the (x, y, z) on the boundary electrodes provides the values
required to predict the measurements Vpred. (n, {}) in the cost function
given above.
4
Explained
H in aRlater section. J = 0 is the differential form of the continuity of charge relationship
S J dS = V t dV , which describes the flow of charge within a conducting medium.

3-7 AJW, EEE3055F, UCT 2012


Figure 3.2: Electrical resistance tomography performed on a tank of water containing two
non-conducting objects.

One can also extract other parameters of interest from the simulated po-
tential function, for example the electric field can be found by differentiation,
since E = . The current distribution can also be found via J = E.
This is all implemented using discrete approximations to the derivatives.

Example 3

In electrical capacitance tomography (ECT), the goal is to reconstruct an


image of the electrical permittivity (x, y, z) within a non-conducting object
(i.e. a dielectric medium) from capacitance measurements between electrode
plates placed around the circumference of the object.

The equation describing the potential distribution within the dielectric


medium is
= 0
This partial differential equation can be solved for a given permittivity dis-
tribution (x, y, z) and an applied boundary condition (being the potential
difference applied between two plates), using a numerical method. Tomo-

3-8 AJW, EEE3055F, UCT 2012


Figure 3.3: Electrical capacitance tomography - the image shown is of three objects inserted
into an (air-filled) container.

graphic algorithms are designed to find an estimate of (x, y, z) for which


simulated measurements best match the actual measured data. The im-
age shown is of three objects (a blackboard duster, and two plastic rods)
standing vertically inside in a plastic container.
A present research project at UCT concerns the condition monitoring of
wooden electricity distribution poles using electrical capacitance tomogra-
phy. An instrument has been built that allows one to detect decay and
termite damage within poles. A specially designed jacket of 12 electrodes is
attached to the base of a wooden pole, and an image of the interior can be
seen on a computer screen.
More generally, the method known as electrical impedance tomography

3-9 AJW, EEE3055F, UCT 2012


(EIT) is used in cases where both the conductivity and permittivity pa-
rameters are to imaged. Typically, a sinusoidal signal is injected (kHz or
MHz frequency range), and both the magnitude and phase at the sensing
electrodes are recorded. Pulse-based methods are also sometimes used.
Another tomography method involves stimulating the medium under in-
vestigation with a sinusoidal magnetic field (driving and sensing coils are
placed around the boundary). This method is known both as magnetic
induction tomography (MIT) or eddy current tomography. The applied
time-varying magnetic field creates eddy currents in/on conducting material
within the medium, which, in turn, creates a magnetic field which modifies
the magnetic field sensed by the coils.
A prototype MIT system is also being developed at UCT.

3.2 Electromagnetic Equations before Maxwell


Before Maxwell introduced the concept of displacement current, the four
equations describing the field relationships stood as
I Z
D dS = dV
S V
I
B dS = 0
S
I Z
B
E dl = dS
S t
I Z
H dl = J dS
S
and in differential form
D=
B=0
B
E=
t
H=J

3-10 AJW, EEE3055F, UCT 2012


The third differential equation (Faradays law) is obtained by applying
Stokes equation to E, i.e.
I Z
E dl = E dS
S
H R R R B
and comparing to E dl = S B t dS. Since S E dS = S t dS
B
must hold for all open surfaces, this implies, E = t .
A modern approach to justifying the addition of the displacement cur-
rent term is to show that the static form of Amperes law is inconsistent
with an relationship known as the continuity of charge equation, and that
the inclusion of the displacement current (density) term fixes the problem
(Maxwell may not have done it quite this way). For this approach, we need
the concept of the continuity of charge.

3.3 Continuity of Charge


If we consider a closed surface S, surrounding a volume V , the fact that
charge is neither created or destroyed leads us to the continuity requirement
I Z
d dQ
J dS = dV =
S dt V dt
which in words, simply states that the current leaving the volume equals
(minus) the rate of change of charge within the volume.
Note that if the net outflow is positive, the dQ
dt will be negative, and
hence the minus sign in the above equation.
R R
Note also, that dQdt = d
dt V dV = V t dV since differentiation and
integration are linear operations, which allows the order can be inter-
changed.
Thus we can write the continuity equation as
I Z

J dS = dV
S V t

3-11 AJW, EEE3055F, UCT 2012


The above integral form can be expressed as a differential form by applying
the divergence theorem to J, i.e.
I Z
J dS = JdV
S V

Thus Z Z

JdV = dV
V V t
Since this must hold for any chosen closed surface S containing volume V ,
we conclude that

J=
t
This is the differential form of the continuity of charge equation.

J
If dQ/dt < 0
then there is a
Charge
net current leaving
S
flowing out
dQ/dt<0

Figure 3.4: The continuity of charge equation implies that the rate at which charge leaves
a closed surface equals (minus) the rate of change of the total charge within
the enclosed volume.

An alternative, and more fundamental approach to obtain the continuity


of charge relationship in differential form is directly from the definition of
divergence:
H R
S J dS V t dV dV
J = lim = lim = lim t =
V 0 V V 0 V dV 0 dV t

3-12 AJW, EEE3055F, UCT 2012


3.4 Fixing the Problem with Amperes Law
A fundamental vector identity states that div(curlF) = ( F) = 0 for
any differentiable vector field F. If we take the divergence of the left and
right hand side of Amperes static form

H=J

we get
( H) = J
which would imply that J = 0. This clearly is not consistent with the
continuity equation J =
t , and so we must conclude that H 6= J
in general, although we know it holds for certain measurable cases (e.g.
magnetic field surrounding a conductor).
To fix the problem, we try adding a new term T, and see what we come
up with - i.e. let
H=J+T
Taking the divergence on both sides, and applying again our vector identity,
we get
0=J+T
Rearranging and substituting the continuity equation we get

T = J =
t
Noticing that we can substitute D = , we see
D D
T= = =
t t t
Mathematically,
D
+k T=
t
where k is a constant, satisfies the equality. The integration constant
does however not model anything physical5 and so, T = D t as introduced
5 D
For example, if k 6= 0, then it would suggest that H = k 6= 0 even when J = 0 and t = 0. Until
something is observed physically, the constant is zero.

3-13 AJW, EEE3055F, UCT 2012


by Maxwell, is sufficient to resolve the inconsistency in Amperes law. This
addition to Amperes magnetostatic law provides Maxwells fourth equation
D
H=J+
t
This simple fix makes Maxwells equations a mathematically consistent
set, and predicts the existence of an additional physical quantity called
the displacement current (density) Dt .

The units of D
t
must be the same as the conduction current J, being
Am , although D
2
t does not refer to any physical movement of charge.

The extended model has been shown to be accurate by experiment,


and predicts a variety of phenomena, some of which had not been
previously realised, including the fact that light is an electromagnetic
phenomenon, i.e. a propagating EM wave (discussed later).

We have derived Maxwells modified form of Amperes law in differential


form. To obtain the integral form, we again write down Stokes theorem for
H, and substitute H = J + D t
,
I Z Z Z Z
D D
H dl = H dS = J + dS = J dS + dS
S S t S S t

giving I Z Z
D
H dl = J dS + dS
S S t

3.5 Maxwells Differential Equations


The complete set of Maxwells Equations in differential form is
D=
B=0
B
E=
t
D
H=J+
t

3-14 AJW, EEE3055F, UCT 2012


where D = E and B = H. It is pointed out that this groups for equations,
together with the Lorentz force law
F = qE + qv B
summarize all the behaviour of classical electromagnetic field theory in both
matter and free space.
All other laws e.g. Coulombs force law, the Biot-Savart law and the
continuity equation can be derived from the above equations.
The only place where these differential forms do not hold are on bound-
aries between different materials where the derivatives are not finite.
At boundaries, we can derive a set of boundary conditions that relate
the field quantities immediately on either side of the boundary (see
section on boundary conditions).
The quantities and J refer to the free charge in the medium - i.e.
the charge that is free to move as current, as opposed to the charge
bound to atoms. (When a dielectric is polarized via exposure to an
electric field, a layer of surface charge forms on the polarized dielectric,
but this kind of bound charge is not included in the .)

3.6 Summary Table of Maxwells Equations


H R
Gauss law S D dS = V dV D =
H
Gauss law S B dS = 0 B=0
H R
Faraday E dl = dtd S B dS E = B
t
H R R
Ampere/Maxwell d
H dl = S J dS + dt S D dS H = J + D
t

Auxiliary equations:
F = qE + qv B
D = E
B = H
J = E

3-15 AJW, EEE3055F, UCT 2012


4 Electromagnetic Boundary Conditions
The integral forms of Maxwells equations describe the behaviour of elec-
tromagnetic field quantities in all geometric configurations. The differential
forms of Maxwells equations are only valid in regions where the parameters
of the media are constant or vary smoothly i.e. in regions where (x, y, z, t),
(x, y, z, t) and (x, y, z, t) do not change abruptly. In order for a differen-
tial form to exist, the partial derivatives must exist, and this requirement
breaks down at the boundaries between different materials.
For the special case of points along boundaries, we must derive the rela-
tionship between field quantities immediately on either side of the boundary
from the integral forms (as was done for the differential forms under differ-
entiable conditions).
Later, we shall apply these boundary conditions to examine the behaviour
of EM waves at interfaces between different materials - from them we can
derive the laws of reflection and refraction (Snells law).

4.1 Boundary Conditions for the Electric Field


Consider how the electric field E may change on either side of a boundary
between two different media as illustrated below.
The vector E1 refers to the electric field in medium 1, and E2 in medium 2.
One can further decompose vectors E1 and E2 into normal (perpendicular to
interface) and tangential (in the plane of the interface) components. These
components labelled En1, Et1 and En2, Et2 lie in the plane of vectors E1 and
E2 .
To derive boundary conditions for E, we must examine two of Maxwells
equations: I Z
B
E dl = dS
S t

4-1
Medium 1
E1 En1

Et2
Et1

En2 E2 Medium 2

Figure 4.1: Normal and tangential components of the electric field on either side of the
interface between two media.

and I Z
D dS = dV
V
which will allow us to relate the tangential and normal components of E on
either side of the boundary. Note that refers to the free, unbound change
within V , i.e. it excludes the charge bound to atoms.

Normal Component of D

The boundary condition for the normal component of the electric field can
be obtained by applying Gausss flux law
I Z
D dS = dV
V

to a small pill-box, positioned such that the boundary sits between its
upper and lower surfaces as shown in the illustration.

4-2 AJW, EEE3055F, UCT 2012


Medium 1
n
1 , 1 , l s
dS = n

111111
000000
s 000000
111111
000000
111111
000000
111111 h
000000
111111
Surface charge
s dS =
n s

Medium 2 2 , 2 , 2

Figure 4.2: Gaussian pill box straddling the interface between two media.

If we shrink the side wall h to zero (keeping the interface sandwiched


between the upper and lower surface) then all electric flux enters or leaves
the pill-box through the top and bottom surfaces, and
I
D dS D1 n s + D2 (
n) s = Dn1 s Dn2 s

where Dn1 and Dn2 are the normal components of the flux density vector
immediately on either side of the boundary in mediums 1 and 2, and s is
the elemental surface area.
The amount of charge enclosed as h 0 depends on whether there
exists a layer of charge on the surface (i.e. an infinitesimally thin layer of
charge)1. If a surface charge layer exists then
Z
dV = s s
V

and thus
Dn1 s Dn2s = s s
1
In perfect conductors, any excess free charge always resides on the surface of the conductor and is
denoted by s in units of Cm2 . Within the conductor, the charge density very rapidly goes to zero -
this is discussed in a later section on relaxation time.
It should also be noted that in the case of dielectrics subjected to an electric field, the material
polarizes, which does in fact result in a surface charge layer - however this charged layer is bound
charge caused by the polarization effect, and is not part of the quantity s .

4-3 AJW, EEE3055F, UCT 2012


from which we conclude
Dn1 Dn2 = s
For the case where s = 0,
Dn1 = Dn2
Or in terms of the electric field E,
1 En1 = 2 En2
or
2
En1 = En2
1

Tangential Component of E

We can derive the tangential component of E by applying Faradays law to


a small rectangular loop positioned across the boundary, and in the plane
of E1 and E2 , as illustrated in the diagram below.
E1

n
c
l a Medium 1
b
1 , 1 , l
h

2 , 2 , 2
d
c
Medium 2

E2

Figure 4.3: To determine the boundary condition on the tangential component of the E
field, Faradays law is applied to rectangular loop straddling the interface be-
tween two media.

Consider the limiting case where the sides h perpendicular to boundary


are allowed shrink to zero. In the limit as h 0 , the magnetic flux
threading the loop shrinks to zero, and thus
I Z b Z d
E dl E1 dl + E2 dl = 0
a c

4-4 AJW, EEE3055F, UCT 2012


E1 l + E2 (l) = 0
Writing the tangential components of E1 and E2 along the contour as Et1
and Et2 , we have
Et1 l Et2l = 0
from which we conclude that on either side of the boundary,

Et1 Et2 = 0
or
Et1 = Et2
i.e. tangential components immediately on either side of a boundary are
equal.

4.2 Boundary Conditions for the Magnetic Field


The derivation of boundary conditions for the magnetic field, follows similar
arguments to that of the electric field, but using equations
I
B dS = 0
I Z Z
D
H dl = dS
J dS +
S S t
Again we consider the normal and tangential components as illustrated be-
low.
Medium 1
B1 Bn1

Bt2
Bt1

Bn2
B2 Medium 2

Figure 4.4: Normal and tangential components the B field on either side of an interface.

4-5 AJW, EEE3055F, UCT 2012


Normal Component of B

The boundary condition for the normal component of the magnetic field
can be obtained by applying Gausss flux law
I
B dS = 0

to a small pill-box.
If we shrink the side wall h to zero, all magnetic flux leaves/enters the
pill-box through the top two surfaces,
I
B dS B1 n s + B2 (
n) s = Bn1s Bn2s

Equating to zero, we find


Bn1 Bn2 = 0
and hence the normal component of B is continuous at boundaries.

Tangential Component of H We can derive the tangential component of H


by applying Amperes law to a closed loop as illustrated below. Again, the
rectangular loop is in the plane of vectors H1 and H2 .
H1

n
c
l a Medium 1
b
1 , 1 , l
h

2 , 2 , 2
d
c
Medium 2

H2

Figure 4.5: To determine the boundary condition on the tangential component of the H
field, Amperes law is applied to rectangular loop straddling the interface be-
tween two media.

Amperes law states


I Z Z
D
H dl = J dS + dS
S S t

4-6 AJW, EEE3055F, UCT 2012


Consider the limiting case where the sides h perpendicular to the boundary
are allowed shrink to zero. The left hand side becomes
I Z b Z d
H dl H1 dl + H2 dl
a c
= H1 l + H2 (l)
R
On the right hand side, the displacement current term Id = S D t
dS shrinks
to zero. For physical media,R the conductivity is finite, and J is also finite.
Thus within the loop Ic = S J dS also shrinks to zero, and so
H1 l + H2 (l) = 0
which implies the tangential component of H does not change immediately
on either side of the boundary, i.e.
Ht1 = Ht2
For the special case of an idealised perfect conductor, , a surface
current may exist (i.e. current flowing within a vanishingly thin layer on
the surface). Some physical situations involving good conductors like metals
(e.g. skin effect and reflection of EM waves off metallic objects) may allow us
to treat currents concentrated on the surface as a surface current modelled
by a vector Js in units of Amps/m (NB not m2) flowing in an infinitesimally
thin layer. To get the correct orientation of the tangential component of Ht ,
we can orientate the rectangular loop until the line integral is a maximum.
This is achieved when Js flows perpendicular to our rectangular loop. The
current flowing through S is then
Z
Ic = J dS Js l
S
where Js is the magnitude of the surface current density.
We conclude that for the case where a surface current exists, the boundary
condition on the tangential component of H is therefore
Ht1 Ht2 = Js
and the tangential components, Ht1 and Ht2 , lie perpendicular to the surface
current vector Js . The direction of vectors Ht1 and Ht2 can be found by
applying the right hand rule with ones thumb pointing in the direction of
Js .

4-7 AJW, EEE3055F, UCT 2012


Summary of Boundary Conditions
Below are depicted the components on either side of the boundary in side
view.

B1 Medium 1
E1 En1 Bn1
Bt2 1 , 1 , l
Et2
Et1 Bt1
En2 Bn2 2 , 2 , 2

E2 B2 Medium 2

Figure 4.6: Normal and tangential components illustrated for the cases of the E field and
the B field.

The boundary conditions are summarised below.

Dn1 Dn2 = s

Et1 Et2 = 0

Bn1 Bn2 = 0

Ht1 Ht2 = Js
The boundary conditions can be expressed in vector form2 as:

D2 n
D1 n = s

E1 n
n E2 = 0
2
These vector forms require careful 3-D visualization. Taking the cross product between the surface
E1 extracts the tangential component (as a vector) .
normal and a vector, say n

4-8 AJW, EEE3055F, UCT 2012


B2 n
B1 n =0

(H1 H2) = Js
n
These general conditions can be further refined depending on the specific
media on either side of the interface. Some examples are given below.

4.3 Examples of Boundary Conditions


4.3.1 e.g. Dielectric - Dielectric Interface
Dielectrics are materials for which all electrons are bound to atoms, and are
non-conducting, i.e 0; no currents flow, and no unbound surface charge
exists unless explicitly put there (i.e. s = 0). Thus we have

Dn1 = Dn2 or 1En1 = 2En2

Et1 = Et2
Bn1 = Bn2
1 1
Ht1 = Ht2 or Bt1 = Bt2
1 2
An example of a dielectric-dielectric interface is the interface between
air and glass.
The above boundary conditions are applied when analysing the reflec-
tion and refraction of plane waves (studied in a later section).

For example, consider an air-glass interface where 1(air) = 0 and 2(air) =


50. Given the E1 vector in the air, how can we sketch the E2 vector?

4-9 AJW, EEE3055F, UCT 2012


E1
Medium 1
AIR
1 = 0

GLASS E2 =?
2 = 50, 2
Medium 2

Figure 4.7: E1 field at the the interface between two dielectrics.

We already know the tangential component must be sketched as Et2 = Et1 .


The normal component is related by En2 = 21 En1 = 15 En1 . The E2 can thus
be sketched as shown below:

E1
Medium 1
En1
AIR
1 = 0
En2
E2 Et2 = Et1
GLASS
2 = 50 , 2
Medium 2

Figure 4.8: E1 and E2 field at the the interface between two dielectrics.

4.3.2 e.g. Dielectric - Perfect Conductor


If one of the media is a dielectric (say medium 1 is air), and the other
medium (medium 2) is a perfect conductor 2 , then En2 = 0 and
Et2 = 0 inside the perfect conductor3.
3
The conductor is assumed to be stationary within the xyz frame of reference. If a conducting metal
object moves through a magnetic field, a non-zero E field can exist within the conductor. For example,
if a long thin metal rod moves through a uniform B field with velocity v, the electrons inside the rod

4-10 AJW, EEE3055F, UCT 2012


Since Dn1 Dn2 = s , we conclude that Dn1 = s
Since Et1 = Et2 and Et2 = 0, we conclude that Et1 = 0, i.e. there exists
no tangential component on the dielectric side of the interface.
In vector form we state the boundary conditions for the field in the dielectric
as
= s
D1 n
and
D1 = 0
n

The E field lines always meet a perfect conductor perpendicular to the


surface, and magnetic field lines parallel to the surface as is illustrated
in the figure below:

experience a force qv B, perpendicular to the direction of motion and perpendicular to B. Electrons


deplete on the one side and increase on the opposite side, causing an induced dipole.
Fields inside
B the rod

Rod moving v E
through uniform vB
magnetic field

As the dipole forms, an electric field builds up until the internal forces balance, i.e. qE = qv B,
and the electron current no longer flows. The internal field strength is E = v B. If the rod is
orientated in the direction of v B, then the potential difference between the ends of the moving rod
Rl
is = 0 E dl = vBl where l is the length of the rod. In cases where a metal object is stationary
within the frame of reference, the electrons will rearrange rapidly (if placed in an EM field) such that
the internal electric field goes to zero; the potential difference between any two points on the conductor
will also then be zero.

4-11 AJW, EEE3055F, UCT 2012


En1 =
dielectric e.g. air
H 1 = Js
n
E1 Medium 1
Bn1 = 0
Et1 = 0 B 1 , 1 , l

E2 = 0
B2 = 0 2 , 2 , 2 = inf
(AC fields) Medium 2
Perfect conductor

Figure 4.9: E field and the B field at the interface between air and a perfect conductor.

For AC fields, no time-varying magnetic field exists in a perfect conductor


- why? Recall that E = B t and since E = 0 in a perfect conductor,
E = 0 and hence B t
= 0. In other words, no changing magnetic field
can exist in a perfect conductor, and hence Bn2 = Bn1 = 0, i.e. the normal
component of the magnetic field is zero. A surface current can still exist,
implying a tangential component of B1 can exist. These two conditions can
be expressed in vector form as
=0
B1 n

H1 = Js
n
These boundary conditions are useful for establishing, for example, the
charge density or current distribution on the surface of a conductor,
when the field quantifies in the dielectric are known or specified.
These boundary conditions will be applied when analysing the reflec-
tion of an electromagnetic plane wave off the surface of a perfect con-
ductor.

4.3.3 e.g. Conductor-Conductor (steady state current)


For the case of two conductors under static field conditions (i.e. E
t = 0 and
B
t = 0), there can be no charge build up at the interface and hence

Jn1 = Jn2

4-12 AJW, EEE3055F, UCT 2012


Since Jn = En, we have an additional constraint on the normal component
of the electric field, i.e.
1 En1 = 2En2
For non-steady state conditions a more complicated boundary constraint
relates J1 and J2H, which can be Hderived by application of the continuity of

charge equation S J dS = t V dV at the boundary. The result is

s
+ t Js =
(J1 J2 ) n
t
where t is the two-variable divergence in the tangent plane applied to the
surface current Js , and
t is the rate of change of surface charge density in
s

Cm2s1 . See [Griffiths] for details.


Side View


Jn1 = J1 n
Medium 1

n
1 , 1 , l


Jn2 = J2 n
s 2 , 2 , 2
surface charge
Medium 2

Figure 4.10: Boundary condition for the normal component of J for conductors (for the
non-steady state case for which there may be charge build-up at the interface).

4-13 AJW, EEE3055F, UCT 2012


5 Relationship between Field Theory
and Circuit Theory
(ref: Ramo et al.)
At lower frequencies where physical circuit dimensions are small compared
to the wavelength1 of electromagnetic waves, the behaviour of circuits is ac-
curately modelled using lumped elementcomponent models, together with
Kirchhoffs laws. At higher frequencies where the distances between compo-
nents are a significant fraction of a wavelength and greater, the signals car-
rying information or power from one place in a circuit to another are treated
as waves. Signals must be routed from one point to another using trans-
mission lines, modelled using transmission line theory. If the component
dimensions be comparable to the wavelength then accurate understanding
and prediction of behaviour may require modelling using electromagnetic
field and wave theory.
In this section we examine the relationship between Maxwells equations
and circuit theory. Both Kirchhoffs voltage law, relating to voltage drops
Vi around a loop
XN
Vi = 0
i=1
and Kirchhoffs current law, relating relating currents Ii leaving a node
N
X
Ii = 0
i=1

can be explained in terms of field theory.


Consider for example, the circuit shown below shown firstly using standard
circuit symbols for R, L and C, and secondly as a physical representation
1 f 50 Hz 100 kHz 1 MHz 10 MHz 100 MHz 1 GHz 10 GHz 100 GHz
= c/f 6000 km 3 km 300 m 30 m 3m 30 cm 3 cm 3 mm

5-1
with wires of finite width, a resistor as a rod of carbon, an inductor as a
coiled wire, and the capacitor as a pair of metal plates. We shall examine
briefly each lumped component model from a field theory perspective.
1 2

I(t)
R 2
1
L
V(t)

3
0

0 3

C
1 2

I(t) R
1 L
2

V(t)
3
0 Coil

0 3

Figure 5.1: Circuit diagram showing component symbols (above) and a more physical de-
piction of the components (below).

5.1 Resistors
A resistor can be constructed from a resistive material of conductivity ,
length l and cross-sectional area A, as depicted below.

5-2 AJW, EEE3055F, UCT 2012


Area A E J = E
11
00
00
11
00
11
111
00
00
11 e 2
00
11
l

Figure 5.2: Resistor made from a cylinder of carbon. A current flows as a consequence of
the (axial component) of the electric field.

If the material is subjected to an electric field orientated along the length


of the cylinder, a current will flow, explained as follows:
Electrons move under the influence of the electric field to reach an
average drift velocity.
A classical model explains this as follows:
Electrons initially accelerate under the influence of the field, but repeat-
edly collide with bound atoms, and bounce off, resulting in deceler-
ation. The net result is a constant average velocity for the electrons.
This has some analogy to the terminal velocity reached by a falling
object as a results of the resistance from the air molecules.

Area A 1
0 E J = E
0
1
0
1
0
1
0
1 e Path of an accelerating
0
1 electron, which collides
Average drive velocity with atoms.
of the electrons

Figure 5.3: The electrons accelerate, but are impeded by the atomic structure, hence reach-
ing a finite (average) terminal velocity. An imagined path of a single electron
is shown.

Because of the high density of electrons, the average drift speed is sur-
prisingly slow. For example, Halliday, Resnick & Walker 6th Ed, do an
example calculation (Problem 27.3) in which the drift velocity with a cop-
per wire or radius 0.9 mm, carrying a current of 17 mA, is calculated to be
4.9 107m/s or 1.8 mm/hr.

5-3 AJW, EEE3055F, UCT 2012


The average current density in Am2 is proportional to the strength of
the electric field, i.e.
J = E
where has units [Sm1] and is a property of the medium.
It is noted that any particular electron continually accelerates and de-
celerates with each collision as illustrated.
Energy is dissipated as a result of the collisions (i.e. in the form of
heat).
The voltage developed across the resistor is found by integrating the electric
field through it from node 1 to node 2:
Z 2 Z 2 Z 2 Z 2
J I/A l
V21 = V2 V1 = Edl = Edl = dl = dl = I
1 1 1 1 A
l
The constant A
is identified as the resistance of the rod, i.e.
l
R=
A
In the labelled circuit loop,
V21 = V2 V1 = IR

5.1.1 Calculation of average drift velocity


To calculate the average drift velocity of the electrons in a cylindrical wire
conductor, one needs to know the current I, the charge on one electron
(qe = 1.6 1019 coulombs), the thickness of the wire, and the electron
density e [m3] (not to be confused with charge density). Copper contains
e = 8.5 1028 electrons per m3 - a very large number!
Imagine a slug of electrons moving along a wire rod in the x direction.
The electron drift velocity in metres per second (in the x direction) can be
expressed in terms of the current as
dx dx dQ 1
vdrif t = = = dQ (I)
dt dQ dt dx
where

5-4 AJW, EEE3055F, UCT 2012


dx
dt is the velocity of the leading edge of the slug as it passes some point
x0 ,
dQ
dx is the ratio of charge passing the point x0 per distance dx moved
in the x direction. This is identical the charge density in coulombs per
metre of wire.
dQ
dt is the charge per second passing point x0 per second. If the conven-
tional current is I amperes moving in the negative x direction, then
dQ
dt = I.

To determine dQ
dx
, consider a wire segment of length dx, and of thickness 2r.
The volume of the segment is r2 dx. The number of coulombs per metre is
dQ (charge/vol) vol (e qe )(r2dx)
= = = e qer2
dx length dx
and hence
I
vdrif t =
e qe r2
For copper wire of 1 mm thickness,
dQ
= e qe r2 = 8.5 1028 (1.6 1019) 3.14 (0.5 103)2 = 10676 Cm1
dx
If a current of I = 1 ampere flows in the wire, the electron drift velocity is
I 1 [Cs1 ] 5 1
vdrif t = 2
= 1 = 9.37 10 ms = 34 cm/hr.
e qe r 10676 [Cm ]
The number of electrons making up one coulomb is 1/ |qe | = 6.25 1018.
Thus for a 1 ampere current, 6.25 1018 electrons pass per second.

5.2 Capacitors
Consider a parallel plate capacitor. As the current flows through the wires,
a surface charge builds up on the inner sides of the capacitors plates.

5-5 AJW, EEE3055F, UCT 2012


Note that all excess charge will sit on the surface of the capacitor
plates, in a thin layer (not inside the metal). Recall that E = 0 inside
a perfect conductor, and since div D = or div E = this implies,
= 0 inside the conductor. All excess charge must therefore lie on the
surface, described by a surface charge density s in Cm2.
The charge Q on the plate onto which the conventional current flows is
found by integrating the current flowing onto the plate, i.e.
Z t
Q(t) = I(t)dt + Q0
t0

where Q0 is the initial charge at some starting time t0 . The other plate will
have a charge of Q(t).
A potential difference builds up between the places. The potential differ-
ence can be shown by careful argument [Griffiths] to be proportional to the
charge Q on the plates, i.e.
1
Vc (t) = Q(t)
C
where C is the constant known as the capacitance.
Substituting for Q(t), we get
Z
1 t Q0
Vc (t) = I(t)dt +
C t0 C
In the circuit loop, V03 = V0 V3 = Vc (t).

5.3 Inductors
Inductors are made by winding several turns of wire either in air, or around
some high permeability material (which boosts the inductance, requiring
fewer turns).
We shall explain the operation of an inductor by considering first a single
turn, and then a coil of several turns, in the context of the series circuit
under analysis.
As already discussed, we are interested in applying Faradays law around
the dashed loop shown in the physical circuit. For the inductor, we are

5-6 AJW, EEE3055F, UCT 2012


interested in the integral of the electric field through the air gap between
the terminals as indicated by the dotted line between nodes 2 and 3 in the
circuit.

5.3.1 Single-turn Inductor


Consider a single turn inductor made from a thin piece of wire and illus-
trated in Figure 5.4.
Integration
B contour C B
I
+
V

Figure 5.4: Single turn inductor.

We can apply Faradays law locally to a closed contour C that goes clockwise
around the inside of the wire and then across the air gap (in the shape of
the dotted path),
I Z Z Z
d dM
E dl = E dl + E dl = B dS =
C (air) (wire) dt S dt
where and dS points into the page, and M is the flux threading the inte-
gration loop (and cutting a chosen surface S, bounded by C).
Since E = 0 in the wire, the potential difference is then
Z Z
d
V = E dl = B dS
(air) dt S
The magnetic field B is linearly
R proportional to the current I flowing in the
2
wire , i.e. B I, and so is S B dS I. The constant of proportionality

2
The magnetic field vector owing to a short current segment can be computed using the Biot-Savart
law (reviewed in Section 2.5). The total field is found by integration of all contributions from current
elements in the wire.

5-7 AJW, EEE3055F, UCT 2012


is known as the inductance, i.e.
Z
B dS = LI
S
or R
B(t) dS M S
L= =
I(t) I
The units of inductance3 are henrys [H], The voltage across the inductor is
Z
d d[LI] dI(t)
V = B dS = =L
dt S dt dt
In the labelled series circuit,
dI(t)
V32 = V3 V2 = L
dt

5.3.2 Multi-turn Inductor


A multi-turn inductor is constructed by winding a coil of wire as depicted
in the figure below.
I
+
V3
V
V2

V1

Figure 5.5: Multi-turn inductor (N = 3 here).

The voltage drop across the terminals is the line integral along the dashed
line:
Z +
V = E dl
Z Z Z
= [ E dl + E dl + ... E dl]
gap1 gap2 gapN
= V1 + V2 + ...VN
3
The units of inductance are equivalently [H] = [Wb A1 ] = [VA1 s1 ] = [VCs2 ] = [NC1 mCs2 ] =
[Nms2 ].

5-8 AJW, EEE3055F, UCT 2012


If we further assume that the flux linking each turn is the same, then
d
V1 = = V2 = ... = VN
dt
and
d
V = N V1 = N
dt
Because there are N turns, the flux threading the coil will be N times
stronger than the contribution from a single turn, i.e.
= N 1turn
where 1turn is the flux contribution from a single turn. Substituting, we
obtain
d d d(L1turnI) dI
V = N V1 = N = N 2 1turn = N 2 = N 2 L1turn
dt dt dt dt
Thus the inductance for a tightly wound N -turn coil is
L = N 2 L1turn
where L1turn is the inductance of a single turn.

It is worth remembering that the inductance increases as a function of N 2 .


If one doubles the number of turns, the inductances increases by a factor of
four.

Alternative Explanation

Analysis of a multi-turn coil is similar to the case of a single turn coil, with
the added complication that the integration contour C is not a circle, but
rather made up of a spiral that follows the wire (and a short section in the
air gap between its terminals). As argued for a single turn case, the voltage
across the terminals is
Z Z
d
V = E dl = B dS
(air) dt S
where surface S is now a complicated-to-visualise surface that is bounded
by the contour C. It helps to imagine the surface within the coil as a smooth

5-9 AJW, EEE3055F, UCT 2012


spiral staircase winding around an imaginary centre line. The total flux
M passing through S is proportional to the current I in the wire, and is
given by Z
M = B dS = LI
S
where L is the inductance of the multi-turn coil. Thus
R
M B(t) dS
L= = S
I I(t)
where M must be carefully evaluated for the particular coil structure. For
a short, compact coil of N turns, it is not difficult to show that

L = N 2 L1turn

where L1turn is the inductance of a single turn coil of the same radius. To
see this, one must grasp two points:
the spiral surface S through with the flux lines pass consists of a stack
of N identical contributing flattish discs (the total surface area is
approximately N times larger than for a single turn)
the flux density on each component disc is N times stronger than the
flux density generated by a single turn (superposition of contributions
from N turns, each carrying current I)
Thus the total flux threading
Z the surface of the multi-turn contour is

M = B dS N N 1turn
S
where 1turn is equal to the flux generated by a single turn coil carrying
current I.
The inductance of the multi-turn coil is then
M N 2 1turn
L= = = N 2 L1turn
I I

5-10 AJW, EEE3055F, UCT 2012


5.4 Formulas for Practical Coils
In order to calculate the inductance accurately, we need to consider both the
field inside, and outside of the wire making up the coil. The total inductance
is usually computed in two parts:
L = Linternal + Lexternal
where Linternal is the contribution arising from the magnetic field within the
wire, and Lexternal is the contribution from the field outside the wire.

5.4.1 Inductance of a Circular Wire Ring

External field
in the surrounding
wire
air.
diameter wire 2a
r
2a
radius
of circle

Internal field
inside the metal

Figure 5.6: Circular wire ring, and its cross section.

The inductance of a wire ring can be found from


R
B dS M
L= S =
I I
where B can be found by integrating the field contributions (using the Biot-
Savart law, reviewed in Section 2.5) from each elemental current segment
around the ring.
Considering only the contribution from the flux outside the wire, it can
be shown, by integration, that M Ir[ln 8r
a 2] and hence the external
inductance (Ramo et al.) is
 
M 8r
Lext = 0 r[ln 2]
I a

5-11 AJW, EEE3055F, UCT 2012


for the case where a << r.
The internal inductance of a long straight wire (in henrys per metre of
wire), assuming uniform current density in the wire4, can be shown to be
independent of wire radius a, and for non-magnetic metal wire is given by
(ref Ramo et al.)
0
Lint = = 0.5 107 Hm1
8

Example Calculation

Calculate the inductance of a circular copper wire ring, radius r = 10 cm,


wire radius a = 0.5 mm. NOTE 0 in copper.
The external component is
   
8r 8 0.1
Lext 0 r[ln 2] = 4 107 0.1[ln 2] = 0.7 106 H
a 0.0005
The internal inductance in henrys per metre is
0 4 107
Lint = = = 0.05 106 Hm1
8 8
The total internal inductance

Lint Lint2r = 0.050 106 0.63 = 0.03 106 H

which is relatively small compared to Lext . The total inductance is

L = Lint + Lext = 0.73 106 H

4
At sufficiently low frequencies (kHz down to DC), the current density is fairly uniform across the cross
section of the wire. As the frequency increases from DC, the current tends to concentrate increasingly
towards the outside of the wire. This effect is known as the skin effect, and is discussed later in the
course.

5-12 AJW, EEE3055F, UCT 2012


5.4.2 Inductance of a Short Coil (short length to radius ratio)

Radius

Figure 5.7: Short coil.

For a short length-to-radius ratio coil such as that depicted in Figure 5.7,
the external inductance of an N turn coil is N 2 times that of a single turn,
i.e.  
2 8r
L N 0 r[ln 2]
a
When winding a coil, it is useful to remember that the inductance is
proportional to the square of the number of turns. The inductance may be
increased by winding the coil on an iron or ferrite rod or on a toroid, which
has a relative permeability of hundreds or thousands that of air.

5.4.3 Inductance of a Long Multi-turn Coil (long length to radius


ratio)
Figure 5.8 depicts a long coil of length l and coil-radius r l, containing N
closely-wound turns carrying current I amperes. Such a coil can be treated
as a wrapped sheet of current (the row of dots in the illustration), from
which H can easily be obtained by application of Amperes law
I Z
H dl = J dS
S

. Consider the dashed integration contour shown in the sketch, which


contains a horizontal length along the inside of the coil, and a length on the
outside. The vertical side contributions to the integral are negligibly small
because the flux lines are run perpendicular to the contour at the sides.

5-13 AJW, EEE3055F, UCT 2012


Integration contour

length l
I
B field

r
Radius

B ~ 0 outside
I

Figure 5.8: Long coil.

Outside the coil, the flux lines spread out, and H becomes negligibly small
compared to inside the coil. Consequently, we can also ignore the horizontal
segment of the integral on the outside of the coil. Thus
I
H dl H l

where H is the magnetic field inside


R the coil. The total current passing
through the integration contour is S J dS = N I.

Thus Amperes law implies


H l NI
from which
H N I/l
and
B 0 N I/l
An additional point to note is that this result is independent of the exact
position of the lower horizontal segment of the contour within the coil. This
implies that the field intensity is uniform on a cross section of a long coil, and
hence the flux through the cross section of such a coil is = Br2 = 0 Hr2

5-14 AJW, EEE3055F, UCT 2012


(i.e. linking one turn of the coil)

The inductance is the ratio of the total flux linking the coil to the current,
being
R
S B dS N N Br2 N (0 N I/l)r2 0 r2 N 2
L= = = =
I I I I l

5.4.4 Inductance of an Intermediate Length Multi-turn Coil


In cases where the coil can neither be considered very long or very short,
the following approximate formula is commonly used:
0 r2N 2
L
l + 0.9r
The formula incorporates an empirical correction factor (+0.9r) in the de-
nominator.

5.5 Mutual Inductance


If two wire coils are close to one another, flux resulting from current flowing
in one coil will thread the coil of the other. Thus a changing current in one
coil, will result in a changing flux in the other and hence induce a voltage
across its terminals. This is the basis of a transformer.
1 + 1f rom2 2 + 2f rom1

V1 (t) V2 (t)
I1 (t) I2 (t)

Figure 5.9: Coupled coils.

Consider two coils in close proximity, one containing N1 turns, and the
other containing N2 turns. Although not shown in the sketch, these coils
are parts of circuits and carry currents.

5-15 AJW, EEE3055F, UCT 2012


Let 1 (t) be the component of the flux threading coil 1, resulting purely
from the current I1(t) flowing in coil 1. Let the 1 f rom 2 be the flux threading
coil 1 arising from the current I2(t) in coil 2. From Faradays law, the voltage
across the terminals of coil 1 is
d (N11 + N11 f rom 2 ) d1 d1 f rom 2
V1 (t) = = N1 + N1
dt dt dt
where N11 is the total flux threading the multi-turn coiled surface. Since
1 is proportional to I1 and 1 f rom 2 is proportional to I2, we have,
dI1 dI2
+ M12
V1 (t) = L1
dt dt
where L1 is the self inductance constant of coil 1, and M12 is another con-
d f rom 2
stant. Since N1 1 dt = M12 dIdt2 ,
N11 f rom 2
M12 =
I2

Similarly, the voltage across the terminals of coil 2 is


d (N22 + N22 f rom 1 ) d2 d2 f rom 1
V2 (t) = = N2 + N2
dt dt dt
and
dI2 dI1
+ M21
V2 (t) = L2
dt dt
where L2 is the self inductance constant of coil 2, and M21 is a constant:
N22 f rom 1
M21 =
I1
It can be shown (consult more detailed texts), that regardless of the ge-
ometry, M12 = M21. The constant M = M12 = M21 must have the same
units of L1 and L2 being henrys, and is known as the mutual inductance
between the coils5.
5 N
One could in principle, calculate M12 = 1 1If2rom 2 for a given coil geometry by doing a surface inte-
R
gration of B(i2 ) dS to obtain 1 f rom 2 (I2 ) = S1 B(i2 ) dS. B(i2 ) can be obtained directly from the
Biot-Savart law (which requires a contour integration along coil 2). There is a better way to do it
(consult other texts for details).

5-16 AJW, EEE3055F, UCT 2012


Mutual Inductance for Tightly Coupled Coils
A special case is when the coils are tightly coupled, e.g. stacked on top
of one another such that 1 f rom 2 = 2 and 2 f rom 1 = 1 (or coils would
around a common toroid). For this case,
 
dI2 d1 f rom 2 d2 L2 dI2
M12 = N1 = N1 = N1
dt dt dt N2 dt
N1 N2
and hence M12 = N2 L2 . Similarly, M21 = N1 L1 .

Since M = M12 = M21 , the product M12M21 yields


N1 N2
M2 = L2 L1 = L1 L2
N2 N1
or p
M= L1 L2
for tightly coupled coils.
The voltage ratio is
p
V1 L1 dIdt1 + M dIdt2 N12 L1turn dIdt1 + N12 L1turnN22 L1turn dIdt2 N1
= dI2 = p =
V2 L2 dt + M dIdt1 N22 L1turn dIdt2 + N12 L1turnN22 L1turn dIdt1 N2

which is a well known result for tightly coupled coils.

5-17 AJW, EEE3055F, UCT 2012


5.6 Kirchhoffs Voltage Law
The relationship between Kirchhoffs law for a lumped element circuit model
and the physical component layout, is established by application of Fara-
days law.
Consider applying Faradays law to the closed contour indicated by the
dotted line in the following physical circuit representation. Define dS point-
ing into the page, which implies the integration direction is clockwise.

R
1 2

I(t)
V21 L
1 2
V32
V(t) V10
3
V03
0 Coil

0 3

Figure 5.10: Series circuit loop - Faradays law is applied along the dotted line to derive
Kirchhoffs voltage law.

Since E 0 in the wires, the voltage drops around the circuit occur
across the components. Thus, we can write
I Z 1 Z 2 Z 3 Z 0 Z
B dM
Edl = Edl Edl Edl Edl = dS =
0 1 2 3 S t dt
or
dM
V10 + V21 + V32 + V03 =
dt
The flux threading the loop can be split into three contributions:
M = applied + self + mutual
where

5-18 AJW, EEE3055F, UCT 2012


applied refers to any flux imposed on the circuit e.g. wave a bar magnet
past the circuit.
self refers to the flux generated by the current flowing in the circuit
loop itself (the circuit can be thought of as a single turn inductor).
self = Lself I where Lself is the self inductance of the loop, which
carries current I.
mutual refers any leakage flux from other parts of the circuit (notably
the inductive element) that threads the loop.
Substituting the lumped element relationships derived above,
Z
1 t dI dM
V (t) IR I(t)dt L =
C t0 dt dt
P
Kirchhoffs law N i=1 Vi = 0 describes the circuit model, and hence we
must introduce additional model component(s) into the circuit model
to account for the term ddtM .
It is noted that the term ddtM will modify the current flowing in the
circuit, and should be included for accurate prediction of the behaviour
of the circuit.
In practice however, this term is usually small compared to the other
terms, and is often neglected in practical circuit design.

For the complete model shown below, the equation in the form of Kirchhoffs
law is written as
Z
1 t dI dapplied dself dmutual
V (t) IR I(t)dt L =0
C t0 dt dt dt dt
or
N
X
Vi = 0
i=1

5-19 AJW, EEE3055F, UCT 2012


I
R

V (t) L

dS points into the page


dapplied (the direction of positive flux)
dt

Lself

Figure 5.11: Circuit modified to incorporate an additional series inductor Lself which mod-
els the series inductance of the loop, and an additional voltage source which
models unwanted external signals.

d
The term dtself = Lself dI
dt resulting from the current in the loop, is mod-
elled by a (small) series inductance Lself . A feeling for the magnitude of
this self inductance can be gained by considering a circuit arranged in a
circular loop of radius 10 cm. We previously calculated the self inductance
of a wire ring of radius 10 cm and wire radius of 0.5 mm to be 0.7 H.
At an operating frequency of say 1 kHz, the AC reactance of this term
is XL = 2f Lself 4 103 ohms, which is usually small enough to be
neglected from calculations. At higher frequencies, this term may become
significant.
d
The term applied
dt
arises if the circuit is exposed to some externally gen-
erated AC field, e.g. a nearby transmitter like a cell phone, or perhaps a
motor, or close-by transformer. Usually this term can be neglected. Of
course radio waves are ever present, but their contribution is usually in-

5-20 AJW, EEE3055F, UCT 2012


significant compared to the voltage signals of interest in the circuit. Higher
frequency fields are naturally suppressed by the series inductance of the cir-
cuit or intentional bandlimiting in the design of circuits with RF immunity.
Circuits can also be shielded from external sources by placing them in a
metal enclosure such as a Faraday cage 6.
The term dmutual
dt in this context arises from the leakage flux from the
inductor L, and is in practice usually small compared to the voltage drop
across the (multi-turn) inductor. The net effect may either be to increase
or decrease the current in the circuit, depending on the physical orientation
of the inductor. R
NOTE: The flux S B dS requires the direction of dS to be defined. Since
the integral of E was taken clockwise around the loop, the right hand rule
tells us that dS points into the page. The flux will be a positive quantity if
B (threading the loop) points into the page.

6
A Faraday cage will provide good shielding from electric fields. DC or slowly varying magnetic fields
however do penetrate metal enclosures. e.g. the earths magnetic field is still detected by a magnetic
compass within a Faraday cage. The degree of penetration of time-varying AC electromagnetic fields
is a function of a frequency dependent parameter of the metal known as the skin depth, which will
be studied later in this course. For good shielding at a particular frequency, the enclosure wall should
be considerably thicker that the skin depth (at that frequency).

5-21 AJW, EEE3055F, UCT 2012


5.7 Kirchhoffs Current Law at a Node
Consider the illustration in Figure 5.12 showing four wires connecting to a
node carrying currents I1, I2, I3 and I4. Kirchhoffs node current law states
that the sum of all currents leaving the node equals zero, i.e.
N
X
Ii = I1 + I2 + I3 + I4 = 0
i=1

I2

I1

I3

b
n
closed surface S1

dS

I4

closed surface S2

Displacement current "flows" between plates

Figure 5.12: The relationship between Kirchhoffs cuurent law at a node the continuity
equation.

Consider now the continuity of charge relationship


I Z
d dQ
J dS = dV =
S dt V dt

5-22 AJW, EEE3055F, UCT 2012


which states that the total conduction current leaving an arbitrary closed
surface S is equal to (minus) the rate of change of charge within the volume
V enclosed by S. One can re-express the continuity relationship in a form
that looks similar to Kirchhoffs law by moving the charge term to the left
hand side: I Z
d
J dS + dV = 0
S dt V
The 2nd termR can beHexpressed as a surface integral over S by substituting
Gauss law dV = S D dS,
I I
d
J dS + D dS = 0
S dt S
Moving the time derivative within the integral, the continuity equation be-
comes I I
D
J dS + dS = 0
S S t
which says that the sum of the conduction current Ic and the displacement
current Id leaving an arbitrary closed surface7 is zero. I.e. for any closed
surface S,
Ic + Id = 0
where I
Ic = J dS
S
and I
D
dS
Id =
S t
Thus we have derived a generalised form of Kirchhoffs current law, which
can be applied to an arbitrary closed surface.

For example, consider the closed surface S = S1 surrounding the node in


Figure 5.12. There are N = 4 wires piercing the surface and joining at the
node.

7
It is worth noting that the total displacement current flowing
H out of a closed surface is equal to the time
rate of change of charge enclosed by the surface, i.e. Id = S D
t dS = dQ/dt.

5-23 AJW, EEE3055F, UCT 2012


For S1 , the total conduction current leaving the surface is
I
Ic = J dS = I1 + I2 + I3 + I4
S1

The displacement current is typically insignificant (there is no significant


charge build up within S1 ), i.e.
I
D dQ
Id = dS = 0
S1 t dt
Thus we have
I1 + I2 + I3 + I4 0
If one shrinks surface S1 to a tiny surface surrounding the node, the dis-
placement current shrinks to zero and the relationship converges exactly to
Kirchhoffs law.

If however, we choose a surface S = S2 in such a way as to pass between the


plates of the capacitor as illustrated in Figure 5.12, then we have a slightly
more subtle situation.
As there is one less wire cutting the surface, the total conduction current is
I
Ic = J dS = I1 + I2 + I3
S2

There is however a significant charge build-up on the plate(s) of the capac-


itor as a result of current I4. The charge Qplate on the plate (within S2 )
dQ
builds up at a rate of dtplate = I4.

Thus the continuity relationship


I
dQ
J dS =
S dt
applied to surface S2 becomes
dQplate
I1 + I2 + I3 = I4
dt
Rearranging, we get
I1 + I2 + I3 + I4 0

5-24 AJW, EEE3055F, UCT 2012


which is consistent with the case where S = S1 and Kirchhoffs law.
The approximation ( ) is present in the above expression because a
small (and negligible) charge will exist on the surface8 of the conductors 1
to 4

Another way to look at the situation is to observe that the sum of all
currents, both conduction and displacement current, flowing out of a closed
surface is zero, i.e. for surface S2,

I1 + I2 + I3 + Id = 0
H
where Id = S D t dS is the displacement current leaving the surface. Id
is concentrated primarily between the plates of the capacitor (where the
electric field is strongest).
H D
Also, since we have shown that dQdt = S t dS = Id (for any closed surface),
dQ
and if the only significant dt within S2 is the charge build up on the inner
plate of the capacitor due to I4, then
dQ dQplate
Id = = I4
dt dt
which again for S2 implies

I1 + I2 + I3 + I4 0

8
Any excess charge must be the surface, because 0 very rapidly inside a metal conductor - see section
on relaxation time.

5-25 AJW, EEE3055F, UCT 2012


5.8 The Relaxation Time of Conducting Materials
The term conductor refers to a material that will carry current when sub-
jected to an electric field. In solid materials, like metals, electrons are free
to move, and the net movement of electrons constitutes a current. In liquids
(e.g. a salt solution), charged ions in solution are free to move allowing a
current to exist. Insulating materials, in contrast, are materials for which
the electrons are tightly bound to particular atoms, and hence no current
can flow.
A perfect conductor is one for which there is an unlimited abundance
of free electrons. The conductivity of a perfect conductor is infinite - an
infinitesimal electric field will create a large current. Metals can often be
approximated as perfect conductors in the analysis of their behaviour under
certain conditions.
If a conducting object is placed in a stationary position within an electric
field, the electrons will, given time, rearrange themselves such that:
E goes to zero inside the conductor (electrons quickly re-arrange them-
selves until the total E = 0 inside conductor). Note: in the steady state
situation, the net force on the electrons must go to zero - the electrons
will rearrange themselves to achieve this. Since the conductor is not
moving (stationary), in the steady state situation, the magnetic force
on the electrons will be zero, and hence the electric force must also be
zero9.
9
If a conductor is moving through a static magnetic field - then the E field inside the metal can be
non-zero - electrons will always rearrange themselves such that the sum of the magnetic and electric
forces equals zero. For example, a rod of length l moving at velocity v through a static magnetic field
B will experience a magnetic force on the electrons F = qv B. Electrons will re-arrange themselves
such that the total force on an electron of chage q is qv B + qE = 0, i.e. inside the metal, E = v B
once the electrons have rearranged themselves.
Fields inside
B the rod

Rod moving v E
through uniform vB
magnetic field

There will also exist a potential difference between the end points of the rod, i.e. b a =
Rb Rb
a E dl = a (v B) dl. If v is perpendicular to B and the rod is orientated such that its length is
perpendicular to v and B, then the potential difference between the ends will be vBl.

5-26 AJW, EEE3055F, UCT 2012


The charge density = 0 inside the conductor (since div D = and
E = 0, it means = 0) .
Any net charge (excess charge) resides on the surface in an infinitesi-
mally thin layer (we refer to this charge as a surface charge).
The potential (x, y, z) is constant throughout the conductor (since
the electric field is zero inside).
E is perpendicular at the boundary (i.e. no tangential component).
In practice, one might wonder just how long it would take for the electrons
to re-arrange themselves. Imagine setting up an arrangement of charge
(x, y, z) within a homogeneous conducting material and then releasing the
charge at some instant. The charge will redistribute itself such that the
electric field goes to zero at every point within the conductor10. This hap-
pens very rapidly in metals, so fast that it can be considered instantaneous
in many practical situations.

J J = E
charges move
(t)
0
(x, y, z, t)
(t) = 0 e t
J 0.370
t
=
J

Figure 5.13: The illustration shows an initial charge distribution within a homogeneous
conducting medium. The charge will rearrange as time progresses, with the
charge density at any monitored point decaying over time.

Consider the charge within the conducting body. The movement of charge
will be governed by the continuity equation, for which the differential form

J=
t

10
Here we are ignoring the granularity of electrons and treating the charge as a kind of fluid.

5-27 AJW, EEE3055F, UCT 2012


describes the relationship between current leaving a small volume element,
and the rate of change of charge within the element. If we substitute
J = E, we get

(E) =
t
and then eliminate E via Gauss law ( E = /) we obtain a first order
differential equation

+ =0
t
which has solution

(t) = 0 e t
where 0 is the initial charge density at time t = 0.
Thus the charge density at any point within the material will dissipate
to zero with an exponential decay. The decay curve is characterised by the
time constant = , also known as the relaxation time, which is the time
at which the charge density has reduced to e1 36.8% of its initial value.
After 5 the charge density will have decayed to less than 1% of the initial
value.
To see how quickly this happens in practice, the time constant may be
calculated for various materials. For example for a metal conductor like
copper ( = 5.8 107 Sm1, 0 = 8.85 1012 Fm1), the time
constant is = = 1.5 1019 s, which is extremely short compared to
say the period of a 100 GHz microwave sinusoid, being 1011 seconds. In
electronic circuits, the charge in the wires rearranges itself very quickly in
response to the dynamics of the circuits (i.e. to a very good approximation,
we consider E 0 and 0 inside the connecting copper wires - a small
component of E must however exist to drive the current).
For a weakly conducting liquid like tap water ( 102 Sm1, 810
Fm1), the relaxation time is about 70 109 s. For a good insulator like
glass (e.g. = 1014 Sm1, = 50), the relaxation time is calculated to
be about 4000 seconds (67 minutes).
Exercise: Calculate the relaxation time of iron ( = 0.9 107 Sm1).

5-28 AJW, EEE3055F, UCT 2012


5.9 Shielding and The Faraday Cage
Circuitry may be shielded from external electric fields by enclosing the cir-
cuit inside a metal box known as a Faraday cage. External electric fields
have no influence on the circuitry within a box made from a perfect con-
ductor - an electrically quiet zone exists within the box. On a larger scale,
Faraday cages can provide protection against lightning strikes. Low fre-
quency magnetic fields can however penetrate a real metal enclosure and
influence the circuitry inside it. Try for yourself to see if a permanent mag-
net is able to attract iron pieces through the walls of a metal box. DC
magnetic fields penetrate through metal enclosures. For example, a mag-
netic compass will still detect the earths magnetic field inside a Faraday
cage. Effective shielding from a magnetic field at 50 Hz requires a thick
wall (several mm), preferably made from a high permeability material. As
the frequency increases, a metal wall becomes more effective in attenuating
magnetic fields. In the MHz range and higher, metal enclosures are very
effective (if well sealed) for shielding circuits from external electromagnetic
fields, and also for preventing radiation leakage from the circuitry within
the enclosure. Electromagnetic waves reflect off the enclosure, and what
does enter the metal, decays exponentially with a decay constant called the
skin depth (covered later in the course).

5.10 Twisted Pair Cables


An interesting application of field theory concerns the understanding of how
twin-wire transmission lines are influenced by electric and magnetic fields.
Figure 5.14 depicts a parallel wire (non twisted) transmission line, which
could be used to carry a signal from one location to another. Such parallel
wire transmission lines are particularly susceptible to inductive coupling
of magnetic fields, especially when several signals need to be carried in the
same bundle. Changing magnetic flux d/dt within the circuit loop induces
an additional voltage which adds to the signal voltage in accordance with
Faradays law. A clever solution to minimising inductive coupling is to
reduce the net flux, by twisting the pair of wires as illustrated. The flux

5-29 AJW, EEE3055F, UCT 2012


contributions B dS in adjacent twists are opposite in polarity and will
tend to cancel, resulting in reduced d/dt and hence reduced magnetic
field interference. This type of cable is known as twisted pair and is very
commonly used for data networks.

V (t) dm /dt

V (t)

net dm /dt 0

Figure 5.14: Illustration comparing a straight wire transmission line with a twisted pair
transmission line.

In addition to the minimization of magnetic coupling, the twisting also


improves the immunity to capacitive coupling (an electric field effect). If,
for example, the cable lies close to a conductor that is varying in poten-
tial relative to ground, like the live wire wire 50 Hz mains supply, this 50
Hz signal will capacitively couple to the conductors (imagine small valued
capacitors (C1 and C2 in Figure 5.15 and Figure 5.16) between the 50 Hz
conductor and cables two wires). Twisting the cable, creates a more sym-
metrical coupling arrangement, independent of the of the orientation of the
pair, hence causing the effect to be common to both wires. A differential
amplifier at the receiver with a high common-mode rejection ratio extracts
the desired differential signal, and removes the common capacitively cou-
pled interference. It is also evident in Figure 5.15 that a balanced drive
creates a symmetrical arrangement in which the interfering signal will bet-
ter cancel at the receiver (if C1 C2 and the source resistors labeled R are
identical). Twisting the wires as in Figure 5.16, better matches the coupling

5-30 AJW, EEE3055F, UCT 2012


capacitance to each wire i.e. makes C1 C2.

Figure 5.15: Illustrations of capacitive coupling onto parallel wire transmission lines for the
case of balanced versus unbalanced driving circuitry.

Sometimes twisted pairs are also shielded (i.e. wrapped with an outer
braiding or foil sheath), offering increased immunity to electromagnetic in-
terference and noise. The shielding also further reduces radiation from the
cable itself. Several twisted pairs are sometimes bundled within the same
cable. The use of twisted pairs offers significantly lower cross talk between
data channels compared to non-twisted side-by-side wires within a cable.
Unshielded twisted pair (UTP) cable is now used for connecting standard
PCs in in-door local area networks (LANs). UTP network cables replaced
previously used 50 ohm coaxial cables for LANs because UTP cables are
cheaper to manufacture than coaxial cable, and offer adequate immunity
to electromagnetic interference. UTP network cable is typically used for

5-31 AJW, EEE3055F, UCT 2012


Figure 5.16: Illustrations of capacitive coupling onto wire transmission lines for the case
of straight-wire versus twisted pair cables. The twisting of the wires makes
the capacitive coupling between each wire and the interfering source more
equal. The interference is common to both wires and will be canceled at the
differential input of the receiver.

5-32 AJW, EEE3055F, UCT 2012


distances up 100m.
The Cat-5e series cable is the cable commonly used for PC LANs (for both
100 Mbit/s and gigabit ethernet networks), and is designed to carry frequen-
cies up to 100 MHz. PC LAN network cables contain four unshielded twisted
pairs, with RJ-45 connectors on each end. The characteristic impedance of
Cat-5e is 100 ohms.
Mechanical arrangement within the cable can further reduce coupling be-
tween pairs. For example the Power Cat-6 four pair cable sold by RS
Electronics contains four unshielded twisted pair (UTP) cables, with a cen-
tral separator, and is designed to support high speed data transmission
systems (frequencies up to 250 MHz).

Cat-5e Cat-6

RJ-45 connector

Figure 5.17: Photos from from RS website; connector from Intel website

5-33 AJW, EEE3055F, UCT 2012


6 Electromagnetic Waves
In this section we shall examine the generation and propagation charac-
teristics of electromagnetic radiation as solutions of Maxwells differential
equations.
One can divide the subject into six categories:
Generation of radiation (for which we design antennas)
Propagation through various media (e.g. free space and lossless media,
lossy conducting media)
Reflection at interfaces between different media
Refraction i.e. change of direction as a ray of light passes from one
medium to into another (consider a laser beam passing from air into a
prism)
Diffraction (e.g. bending around corners, and slit diffraction effects)
Scattering from objects (e.g. radars detect the energy scattered from
targets)

Maxwells Differential Equations


The complete set of Maxwells Equations in differential form is
D=
B=0
B
E=
t
D
H=J+
t

6-1
where D = E and B = H.
It is pointed out that this set, together with the Lorentz force law

F = qE + qV B

summarize all the behaviour of classical electromagnetic field theory in both


matter and free space.
All other laws e.g. Biot-Savart law and the continuity equation can be
derived from this set.
The only place where these differential forms do not hold are on bound-
aries between different materials where the derivatives are not finite.
At boundaries, we can derive a set of boundary conditions which re-
late the field quantities immediately on either side of the boundary (see
section on boundary conditions).

6.1 Mathematical Description of Travelling Waves


(ref S.Cloude 1995)
Consider the signal received at a distance z metres from a radiating source.
If the signal progates at speed of v metres per second then the time delay
will be z/v seconds.
In one dimension, a waveform propagating at a fixed speed v [ms1] in
the positive z-direction, can be described by a function of time and position
f (z, t) of the form
z
f1(t )
v
z
where v is the delay (time shift). A waveform travelling in the negative
z-direction can be expressed in the form
z
f2(t + )
v
See hand drawn illustrations of plots of f1(t vz ), plotted
1. as a function of time for different, fixed, values of z.
2. as a function of z at different time instants.

6-2 AJW, EEE3055F, UCT 2012


Figure 6.1: Illustration of wave propagation.

6-3 AJW, EEE3055F, UCT 2012


It is perhaps easiest to see why f1(t vz ) represents a forward propagating
wave by re-expressing the argument as
1
f1 ( (z vt))
v
and defining a new function w1 (.) such that
1
w1(z vt) = f1( (z vt))
v
The function w1(.) is a flipped and stretched version of f (.), i.e. w1(u) =
f ( uv ).
You should recognise that in w1(z vt), the quantity vt represents a
(time varying) shift along the z-axis. As time increases, the waveform w(z
vt) will therefore slide along the z axis in the positive z direction, at a
rate of v metres per second.
Similarly, the function f2(t + vz ) can be re-expressed as
1
f2 ( (z + vt)) = w2 (z + vt)
v
which reveals that waveform w2(.) will move in the negative z direction as
vt increases.

6.1.1 The Wave Equation


Waves travelling in the z directions are solutions to a 2nd order partial
differential equation known as the wave equation of the form
2f 1 2f
=
z 2 v 2 t2
or
2f 1 2f
=0
z 2 v 2 t2
A more compact notation uses subscripts to indicate the derivatives, i.e.
1
fzz = ftt
v2

6-4 AJW, EEE3055F, UCT 2012


This equation arises in many physical contexts e.g. propagation of sound
waves in air and fluids, electromagnetic waves in space, transmission line
theory.
The wave equation can be generalized to model propagation in 3-D space
and also to include a driving term that represents the source of the radiation.
The generalized form in Cartesian coordinates is
 2 
f 2f 2f 1 2f
+ + 2 2 = g(r, t)
x2 y 2 z 2 v t
where r =< x, y, z > represents position in space, and g(r, t) is the general
source term creating the waves. Away from the souce, g(r, t) = 0.
The source term g(r, t) models the underlying radiation generation mech-
anism e.g. a loudspeaker in acoustics, or an antenna in electromagnetic
radiation.
The 3-d wave equation may be more compactly written as

21 2f
f 2 2 = g(r, t)
v t
2 2 2
where 2 x2 + y 2 + z 2 is the Laplacian operator.

dAleberts solution to the 1-D wave equation

Solutions to the wave equation are propagating waves. It is easy to verify


that f1 (t zv ) and f2(t + vz ) (or equivalently w1 (z vt) and w2(z + vt)) are
solutions of the 1-D wave equation:

2f 1 2f
=0
z 2 v 2 t2
The solution known as dAlemberts solution is
z z
f (z, t) = f1 (t ) + f2 (t + )
v v
being the sum of forward and reverse propagating waves.
As an exercise, it is worth verifying the solution by expanding out the
2nd partial derivatives with respect to z and t and substiting into the wave
equation (do this yourself):

6-5 AJW, EEE3055F, UCT 2012


f 1 z 1 z
= f1 (t ) + f2(t + )
z v v v v
2f 1 z 1 z
= f 1 (t ) + f 2 (t + )
z 2 v2 v v2 v
and

f z z
= f1(t ) + f2(t + )
t v v
2f z z
= f 1 (t ) + f 2 (t + )
t2 v v
2 2
Clearly zf2 = v12 2ft and f (z, t) is a solution to the wave equation.
In any physical problem, the solutions are constrained by the so called
boundary conditions. In the case of electromagnetic radiation, these are
the constraints on E, D, B and H at interfaces between different media, and
the constraint on the field quantities at the source of the radiation.

6.2 Wave Equation derived from Maxwells Equations


In this section, we shall see how the wave equation may be derived directly
from the differential forms of Maxwells equations. We shall also see that
the source of electromagnetic radiation is the acceleration of electric charge.
We must show that the field quantities described by Maxwells equations
can be manipulated into a standard form of the wave equation
1 2f
2 f =g
v 2 t2
where g g(r, t), and f f (r, t).
Following the approach described in (S.Cloude 1995), we can find an ex-
pression for the electric field as follows:
Take the curl of Faradays law on both sides, i.e.
B
E=
t

6-6 AJW, EEE3055F, UCT 2012


becomes
B
( E) = ( )
t
Making use of the interchangeability of the partial derivatives

E= ( B)
t
To obtain an equation involving only E and not B, we can substitute
for B which we obtain from Amperes law,
D
H=J+
t
with D = E and B = H, i.e.
E
B = J +
t
Substituting for B we get
E 2E J
E = (J + ) = 2
t t t t
Note that the right hand side looks a bit like the wave equation.
To make further progress, we make use of a vector identity (proofs may
be found in texts on vector analysis),

E = ( E) 2E
2
Ex
2
2 Ex 2 Ex
where 2E E = 2Ey , and 2 Ex = xE2x + y 2 + z 2
2 Ez
Lastly, we use Gauss law

E =

to write the 1st term on the right hand side as

( E) = = grad( )

6-7 AJW, EEE3055F, UCT 2012


In a charge free region = 0 (e.g. a vacuum, or air, or a dielectric
material), grad( ) = 0, and the equation simplifies further to

2 2E J
E = 2
t t
or
2 2E J
E 2 =
t t
This is a vector form of the wave equation. For each Cartesian coordinate

2 2 Ex Jx
Ex 2 =
t t
2 2 Ey Jy
Ey 2 =
t t
2
Ez Jz
2Ez 2 =
t t
NOTE:
In free space, J = 0, and so all three components of E are of the
form
2 2 Ex
Ex 2 = 0
t
If we compare this equation to the standard form wave equation

2 1 2f
f 2 2 =0
v t
we identify v12 , and hence we conclude that electromagnetic waves
must propagate with velocity
1 1
v= = = 2.998 108 ms1
00 8.85 10 12 4 10 7

This is the same as the measured value for light, and hence sug-
gests that light is an electromagnetic wave - Maxwells great revelation
[around 1865]. At the time of Maxwell, this was an amazing result.
The constants 0 and 0 were purely the results of electrostatic and

6-8 AJW, EEE3055F, UCT 2012


magnetostatic theory and experiments, and had not been associated
with light. Maxwell also predicted the existance of EM waves at all
frequencies including radio waves, first demonstrated experimentally
by Hertz in 1888.
In other non-conducting media (e.g. dielectrics), the speed of
propagation is
1 1 1 1 c
v= = = =
r 0 r 0 r r 00 r r
Since, for most materials 0 , the speed of propagation is v cr .

The term n = r r is called the refractive index of the material.
Although we have derived the wave involving E, in exactly the same
way, we can derive a differential equation for B. In a non conducting
medium, the equation is
2B
2B =0
t2
identical in form to that of the electric field.

6.3 Physical Interpretation of the Radiation Generation


and Propagation Mechanism
By comparing the derived differential equation involving E (or its compo-
nents Ex , Ey or Ez ) with the standard form(s) of the wave equation, the
source of radiation is clearly identified as the term g(r, t) = J
t . Thus we
can conclude that the source of electromagnetic radiation is a time-varying
current (density). A constant current implies charge moving at a constant
speed. The time-derivative of a current corresponds to the acceleration of
charge1.

Thus radiation arises from accelerating charges.


1
The current density J resulting from current I passing through a small area S (S in a plane perpen-
I 2x
dicular to the direction of movement), is J = S x J
t . Thus t t2

6-9 AJW, EEE3055F, UCT 2012


6.3.1 Propagation Mechanism
To understand how the radiation propagates, examine the illustrations of
the electromagnetic wave propagating aways from a dipole radiating source.

SEE:
(1) HANDOUT OF ILLUSTRATIONS of EM Propagation [from book
on Antennas by J.D. Kraus]

(2) Simulations of EM propagating waves via links on the EEE3055F


course website.

Figure 6.2: a) Side view illustration of the E field propagating away from a dipole antenna
driven by a sinusoidal source. b) Top view illustration of the corresponding
B field propagating away from the dipole antanna. [images obtained from
simulation made available by Hsiu C Han of Iowa State University]

A short dipole anntena radiator can be constructed from two bits of wire.
Driving the dipole with a sinusoidal voltage source sets up time varying
currents in the two arms. The time varying current sets up a time-varying
electromagnetic field that propagates away from the source. A cross section
showing the E field at a particular time instant can be seen in Figure 6.2.

6-10 AJW, EEE3055F, UCT 2012


This disturbance propagates away from the accelerating charge governed
by two equations:
1. Faradays law:
B
Et =
t
or (by Stokess theorem)
I Z
B
Et dl = dS
t

2. Maxwells equation (from Amperes law):


E
Bt =
t
or (by Stokess theorem)
I Z
E
Bt dl = dS
t

The radiation pattern can be used to depict the power density (W/m2) in
a polar format in a particular direction. For an accelerating charge, a cross
section of the pattern is shown below - the distance from the origin at a
particular angle represents the power density radiated in that direction.
Radiation pattern from accelerating charge
J
t

P owerDensity()

Figure 6.3: Radiation pattern from an accelerating charge.

6-11 AJW, EEE3055F, UCT 2012


6.4 Some Additional Notes on Wave Equations
There are many forms of the wave equation, describing wave propagation
in 1, 2 and 3 dimensions, in Cartesian, cylindrical and spherical coordi-
nates, and in scalar as well as compact vector notations. The general form
describing some scalar field quantity is

2 1 2f
f 2 2 = g(r, t)
v t
where f = f ((r, t)) is a function of position in space and time. The function
g(r, t) is used to model a radiation source. The constant v is the velocity of
propagation.
Away from the source, g(r, t) = 0 and the wave equation is written as

21 2f
f 2 2 =0
v t
The term 2f is called the Laplacian of f.
2
In 1-D spatial dimension denoted by z, 2 f = zf2 and the equation is
written as
2f 1 2f
= g(z, t)
z 2 v 2 t2
It is easy to show (by differentiation) that a solution to the 1-D wave equa-
tion is of the form
z z
f (z, t) = f1 (t ) + f2 (t + )
v v
or alternatively written as
1 1
f (z, t) = f1 ( (z vt)) + f2( (z + vt))
v v
= w1(z vt) + w2(z + vt)

where w1(u) = f1 (vu) and w2 (u) = f2(vu).


The functions f1 and f2 (or alternatively w1 and w2) represent waves
travelling in the positive and negative z directions. This solution is known
as dAlemberts solution. It can be proved that this f (z, t) is a complete
solution to the 1-D wave equation.

6-12 AJW, EEE3055F, UCT 2012


The exact forms of the functions f1 and f2 are only known once a problem
has been more fully specified i.e. specification of (1) the initial conditions
and (2) the boundary conditions (an example of a 1-D problem involving a
plucked string follows).
In 3-D, the Laplacian can be expanded in Cartesian coordinates to
yield  2 
f 2f 2f 1 2f
+ + 2 2 = g(r, t)
x2 y 2 z 2 v t
which has solutions of the form

f (x, y, z, t) = f1 (ax + by + cz vt) + f2(ax + by + cz + vt)

which can be verified by differentiation. This form is used for modelling


problems with rectangular symmetry e.g. a plane wave propagating in space.
The values of the constants define the direction of propagation of the wave.
Waves can propagate simultaneously in several directions, i.e. f (x, y, z, t)
can be the superposition of one or more waves propagating in different
directions.
The wave equation can also be expressed in spherical coordinates,
which is convenient form modelling waves with spherical symmetry e.g.
spherical waves radiating away from a source. The Laplacian in spherical
coordinates is
   
2 1 2 f 1 f 1 2f
f= 2 r + 2 sin + 2 2
r r r r sin r sin 2
The wave wave equation becomes unwieldy to write, and so the compact
notation
2 1 2f
f 2 2 = g(r, t)
v t
is very convenient. If the wave amplitude depends only on the distance from

the source (e.g. radiation from a point source), 0 and 0 and the
Laplacian simplifies to
 
2 1 2 f 2f 2 f
f= 2 r +0+0= 2 +
r r r r r r

6-13 AJW, EEE3055F, UCT 2012


and the wave equation in spherical coordinates simplifies to
 2 
f 2 f 1 2f
+ =0
r2 r r v 2 t2
which has a general solution the form
1 1
f (r, t) = f1 (r vt) + f2(r + vt)
r r
where the first term represents a spherical wave travelling radially outward
from the source, with a decaying amplitude proportional to 1r ; the 2nd term
represents a wave converging on the origin. This solution can be verified by
substitution.
A 2-D example of a spherical wave, is the wave front propagating out on
the surface of water when a stone is dropped into onto the surface.
Sufficiently far from a radiating antenna, the amplitude of the electric and
magnetic fields decay as a function of 1r , like a spherical wave. The radiated
power density 2 decays proportional to r12 . Unlike a spherical wave, practical
antennas focus the radiated energy in particular directions, like the beam
of a torch.

6.5 Travelling Waves in a Plucked Guitar String


(references: Wylie & Barrett, and also Griffiths)
A simple example of the wave propagation is found in the analysis of a
plucked guitar string, clamped at x = 0 and x = L as shown in Figure 6.4.
f (x|t)
(transverse displacement)

x=0 x=L x

Figure 6.4: A plot of the displacement of the string as a function of position x (at some
given instant t in time).
2
As we shall see in a later section, the power density vector P = E H points in the direction of power
flow, and far from a source, the power density |P| r12 (in the so-called far field).

6-14 AJW, EEE3055F, UCT 2012


The function f (x, t) models the transverse displacement of the string as
a function of position x and time t, and turns out to be the solution of a
1-D wave equation. The wave equation is derived by considering the forces
acting on a small elemental section of the string shown in Figure 6.5.
Elemental section of string Horizontal and transverse
f (x|t) T2 components of T2
2
Displacement
T2
T2 sin2
Restoring 2
1 force T2 cos2
T1
x
x1
x

Figure 6.5: Forces on an element of the string; the magnitudes of the angles are exaggerated.

We shall make the following assumptions:


the transverse displacement of the string is small (typically mm) com-
pared to its length (normally 65 cm).
the resulting small increase in string length has negligible increase in
tension, i.e. tension T , in Newtons, is approximately constant i.e. T1 =
T2 = T . The tension in a real guitar string is in the range 70 N to 180 N.
the horizontal displacement of the elemental section is negligibly small.
the tension is high enough such that gravity can be neglected i.e. string
does not sag under its own weight.
air resistance is also negligible - although it would in practice contribute
to bringing a sting in motion to rest over time. Energy is lost as it is
converted into an audible sound wave.
the string has a uniformly distributed mass of w kg m1 (typically
0.4 103 to 8 103 kg m1).
for any segment considered, the angles 1 and 2 are small enough such
that we can make the small angle approximations sin1 tan1 1
and sin2 tan2 2.

6-15 AJW, EEE3055F, UCT 2012


The elemental section experiences a force T1 = T to the left directed along
the string, and a force T2 = T directed to the right along the string as
illustrated. The horizontal force components are approximately equal, i.e.
T2cos2 T1cos1 . The net transverse force on the elemental section is the
difference in the transverse components, i.e.

F = T 2 sin2 T1sin1 = T (sin2 T1sin1)



f f
T (tan2 tan1) = T ( )
x x x1 +x x1

For a positive lateral displacement (i.e. f (x1|t) > 0) the value of F will be
negative, indicating a restoring force in the direction of strings rest position
being the x-axis.
According to Newtons 2nd law, the force equals mass times acceleration.
For the small section of length l, its mass m = wl wx since 1 and
2 are small angles. Thus Newtons 2nd law implies

f f 2f
F = T( ) = wx 2
x x1 +x x x1 t
or
f
( f
x x +x x x )
1 1
w 2f
=
x T t2
2
f
Taking the limit x 0, the left hand side becomes x ( x ) = xf2 , being
the second partial derivative evaluated at position x = x1. Thus we have
derived the standard form of the 1-D wave equation
2f w 2f
=
x2 T t2
which describes the lateral displacement of the string at any position x and
time t between its clamped endpoints. The propagation
q speed of the wave
(in the +x or x direction) is identified as v = wT .
How do we obtain a solution? dAlemberts solution, we know is a com-
plete solution of the form

f (x, t) = f1 (x vt) + f2(x + vt)

6-16 AJW, EEE3055F, UCT 2012


where f1() and f2() to be determined from the initial and boundary condi-
tions. In this problem, the initial conditions are (i) the shape of the string
at time t = 0 when it is released, i.e. f (x, 0) = f (x, t)|t=0 = U0 (x), and (ii)
(x,t)
the initial transverse velocity ft = V0 (x) over the length.
t=0
From these initial conditions, we have two equations

U0 (x) = f1(x) + f2(x)

and
f (x, t)
V0(x) = = vf1 (x) + vf2 (t)
t t=0
Differentiating the first equation, U0 (x) = f1 (x) + f2 (x) and substituting
f2 (x) = U0 (x) f1 (x) into the second, we get

V0 (x) = vf1 (x) + v(U0 (x) f1 (x)) = 2vf1 (x) + vU0 (x)

Rearranging
1 1
f1 (x) = U0 (x) V0 (x)
2 2v
and by integration we get
Z x
1 1
f1 (x) = U0(x) V0 (u)du k
2 2v 0
where k is a constant, and
Z x
1 1
f2(x) = U0(x) f1(x) = U0(x) + V0 (u)du + k
2 2v 0

If we assume that the string is stationary at the time of release, V0 (x) = 0;


the integral terms are zero, and f (x, t) simplifies to

f (x, t) = f1(x vt) + f2 (x + vt)


1 1
= U0(x vt) + U0(x + vt)
2 2
The plucked string is the shown below as a triangular initial displacement3
at t = 0. The time-varying solution shows that f (x, t) is the sum of two

6-17 AJW, EEE3055F, UCT 2012


Wave solution to a plucked string

U0 (x)
U0 (x)/2 Initial displacement
at t=0

Waves moving apart


U0 (x + vt)/2 U0 (x vt)/2

No longer overlapping

Waves invert on reflection at clamped end points.

Figure 6.6: The illustration shows an initial condition U0 (x). Releasing the string at t = 0
sets up two propagating waves that reflect off the clamped ends of the string.
There is an inversion on reflection (similar to an voltage wave travelling down
a transmission line and reflecting off a short circuit termination).

6-18 AJW, EEE3055F, UCT 2012


waves, each being 21 U0( ), one propagating to the left and the other to the
right.
If a finite length string is considered and the boundary conditions are
applied (i.e. f (0, t) = 0 and f (L, t) = 0), then the full solution which can
be derived which shows that the waves will bounce back and forth off the
end points, reversing polarity at each bounce.
The time taken for the string to return to its initial state is the period
2L/v seconds - each pulse must reflect twice to return to its original state,
travelling a distance of 2L. The frequency of the lateral oscillation is there-
fore r
1 v 1 T
f= = =
period 2L 2L w
which is the fundamental frequency of sound that one hears. The vibration
of the string creates a compression wave in the air that propagates to ones
ears. Higher frequencies are also heard at integer multiples of the funda-
mental. If one plots the lateral displacement of a point on the string as a
function of time, it will be a periodic function. Fourier analysis will reveal
the amplitude of the harmonic components.

Example

= 2103kg m1
A guitar string has the following parameters: T = 100 N, w q
T
and L = 0.65 m. The speed of wave propagation is v = w = 224 ms1.
The fundamental frequency is
v 224
f= = = 172 Hz
2L 2 0.65
Tightning the string will increase speed of propagation and hence the fre-
quency.

3
One could imagine creating such an initial displacement using three drawing pins that are simultaneously
removed at time t = 0. A more realistic initial condition for a guitar string is a pluck made by a single
finger.

6-19 AJW, EEE3055F, UCT 2012


7 Sinusoidal Waves

7.1 Signals as Sums of Sinusoidal Functions


Analysis of a waveform can often be greatly simplified by representing the
waveform as the sum of functions that are more easily analysed that the
waveform itself. Examples of such representations for general signals are
the Fourier series (for periodic waveforms)

X
f (t) = Cn cos(n0 t + n )
n=1

the complex Fourier series (for periodic waveforms)



X
f (t) = Fn exp(jn0t)
n=

and the (inverse) Fourier transform for non-periodic waveforms, i.e.


Z
1
f (t) = F ()exp(jt)d
2
Well-known formulas exist for calculating the weightings of the sinusoids in
the above representations, i.e. Cn (and phase n), Fn and F ().
If, for example, one wishes to predict how a transmitted pulse propagates
to a receiver through a complex medium (the response of which may be
frequency dependent), one can carry out the analysis for a single frequency
sinusoid, and having done so, build the final output as the superposition of
all the sinusoidal components. One could in fact represent the relationship
between transmitter and receiver in terms of a frequency dependent transfer
function H(), for which
Vrec () = H()Vtran ()

7-1
where Vrec () and Vtran () are the Fourier transforms of the transmitted
signal Vtran (t) and received signal Vrec (t). The time domain waveform Vrec (t)
found by inverse transforming Vrec ().
This motivates the detailed study of sinusoidal solutions to Maxwells
equations.

7.2 Real Sinusoidal Travelling Waves


Any function f (t) may be transformed into a travelling wave moving in
the x direction at speed v by replacing the argument t by (t xv ), i.e.
f (t xv ). A forward travelling sinusoidal wave f (x, t) can be constructed
from cos(t + ) simply by replacing the variable t by (t xv ), i.e.
x
f (x, t) = A cos((t ) + )
v

= A cos( (x vt) + )
v
The function is also sometimes written as
x
f (x, t) = A cos(t + ) = A cos(t kx + )
v
where is the angular frequency in radians per second, and the constant
k = /v is called the wave number and is the spatial frequency in radians
per metre, and is a constant phase shift.
The waveform may be sketched as a function of x at a fixed time, or as a
function of t at a fixed position x.
Below are shown sketches of the function cos(t) and the travelling wave
kx
A cos(t kx + ) = A cos([t ])

as a function of time t at a fixed position x.

7-2 AJW, EEE3055F, UCT 2012


cos(t)
as a function of t

t
2
period T =

cos(t kx + ) shift
kx
as a function of t

Figure 7.1:

kx
It is pointed out that A cos(t) is shifted to the right by a time
=
x/v /.
The period T of the waveform is the time over which the phase changes
by 2, i.e. solving (t + T ) t = 2 yields
2
T =

A sketch of cos(kx) and the wave as a function of x at a fixed time t are


shown below:

7-3 AJW, EEE3055F, UCT 2012


cos(kx)
as a function of x

x
2
wavelength = k

cos(t kx + ) shift
t+
as a function of x k

Figure 7.2:

The wave A cos(t kx + ) = A cos(k[x + t+


k ]) which is A cos(kx)
t+
shifted to the right by a distance k = vt + /k
The wavelength is defined as the distance between consecutive points
of identical phase, i.e. the distance over which the phase changes by
2, i.e. solving k(x + ) kx = 2 yields
2 2 v v [ms1 ]
= = = =
k /v /2 f requency [Hz]
The complete sinusoidal solution to the 1-D wave equation is the sum of
forward and reverse travelling waves of the form:
f (x, t) = A1 cos(t kx + 1 ) + A2 cos(t + kx + 2 )
where the negative travelling wave is created by replacing x with x.

7.3 Complex Phasor Representation


Complex exponentials are often used in the analysis of linear system for a
number of reasons:
Complex exponential function are the basis functions used in the com-
plex Fourier representations of signals.

7-4 AJW, EEE3055F, UCT 2012


Mathematical notation and analysis can be simpler (e.g. in circuit
analysis)
The magnitude and phase shift changes that occur to a real sinusoid
when passed through a linear system, can be obtained directly from
analysis of the response to a complex sinusoid.
The real function
f (x, t) = A1 cos(t kx + 1 ) + A2 cos(t + kx + 2 )
is represented by the complex exponential form
f(x, t) = A1 exp [j(t kx + 1)] + A2 exp [j(t + kx + 2 )]
It is noted that
f (x, t) = Re{f(x, t)}
If we pass a real signal through a physical linear system, the output is a
convolution, i.e.
f (t) h(t)
where h(t) is the (real) impulse response of the system.
Passing the complex signal f(t) through the same linear system we get
h i

f (t) h(t) = Re{f (t)} + jIm{f (t)} h(t)
= Re{f(t)} h(t) + jIm{f(t)} h(t)
= f (t) h(t) + jIm{f(t)} h(t)
We see that
Re{f(t) h(t)} = f (t) h(t)
from which it is clear that we can perform the analysis using the complex
function f(t) and simply take the real part afterwards to obtain f (t) h(t).
The complex function f(x, t) = A1exp [j(t kx + 1 )] is sometimes
written more compactly as
f(x, t) = A1ej1 ejkx ejt = Ae
jkx ejt

where the factor A1ej1 has been replaced by a single complex constant
A = A1ej1 . It is also common in wave analysis to drop the sinusoidal

7-5 AJW, EEE3055F, UCT 2012


component ejt in the written notation, and write the forward and reverse
travelling waves as
A1ejkx + A2ejkx
The term A1ekx is called a complex phasor representation of the (forward)
wave.
The derivative of a complex sinusoid is
d
A1 exp [j(t kx + 1 )] = A1exp [j(t kx + 1 )] j
dt
Thus time derivatives of complex phasors are simply obtained by multi-
plying by j. Integration with respect to time is achieved by dividing by
j.

7.3.1 Maxwells Equations in Complex Phasor Notation


If all time varying quantities are sinusoidal, the set of equations

D=
B=0
B
E=
t
D
H=J+
t
can be written in phasor notation as
=
D

B =0
= j B
E
=J
H + j D

r, t) = E(~
where (~r, t) = (~r)ejt, E(~ r)ejt, B(~
r, t) = B(~
r)ejt, etc. Since
the ejt factors cancel on left and right hand sides, the phasor form of
Maxwells equations relate the non-time dependent portions of , E and B.

7-6 AJW, EEE3055F, UCT 2012


7.3.2 The Wave Equations in Complex Phasor Notation
The wave equations (derived for a non-conducting dielectric medium)

2 2E
E 2 = 0
t

2 2B
B 2 = 0
t
are written in phasor notation as
+ 2 E
2 E =0

+ 2 B
2 B =0
2
2
since t 2 (j)(j) = . It is noted that in phasor form, the wave
equations are purely a function of spatial coordinates and not time (the
ejt terms cancel). We shall use the phasor notation in studying wave
propagation in lossy conducting media.

7-7 AJW, EEE3055F, UCT 2012


8 Plane Waves
The 3-D wave equations for E and B in free space in vector notation are

2 2E
E 2 = 0
t

2 2B
B 2 = 0
t
Expanding the components of E in Cartesian coordinates,
2 Ex 2 Ex 2 Ex 2 Ex
+ + 2 = 0
x2 y 2 z 2 t
2 Ey 2 Ey 2 Ey 2 Ey
+ + 2 = 0
x2 y 2 z 2 t
2 Ez 2 Ez 2 Ez 2 Ez
+ + 2 = 0
x2 y 2 z 2 t
we see that each (scalar) component of the E vector obeys a 3-D wave
equation; and similarly for the components of B. Functions of the form
f (x, y, z, t) = fn (ax + by + cz vt) are solutions, and represent waves trav-
elling in a particular direction.
A plane wave is a propagating wave for which the field is uniform in any
plane perpendicular to the direction of propagation. This kind of field is an
idealization, as the fields radiated from sources spread out as a function of
distance with a curved wavefront (see handout of E field lines propagating
away from a dipole antenna). Far from the source, the wavefront may be
analysed locally as a plane wave (by locally, is meant a region over which
the curvature of the field lines is negligible).

8-1
8.1 Plane wave propagating in z direction
To simplify analysis and understanding, let us consider a wave front prop-
agating in the positive z-direction. Since the components of a plane wave
are by definition uniform within any plane perpendicular to the direction
of propagation, we conclude that E is only a function of z, independent of

x and y. Thus x = 0 and y = 0 and the 3-D wave equations in Cartesian
coordinates simplify to 1-D wave equations
2 Ex 2 Ex
2 = 0
z 2 t
2 Ey 2 Ey
2 = 0
z 2 t
2
Ez 2 Ez
2 = 0
z 2 t
Additionally, we shall assume that only an Ex component exists, and that
the Ey and Ez components of the E vector are zero.1i.e. E = hEx (z, t), 0, 0i.
Now we need only consider a single equation
2 Ex 2 Ex
2 = 0
z 2 t
dAlemberts complete solution is of the form

Ex (z, t) = f1 (z vt) + f2(z + vt),

with v = 1 .

1
The orientation of E field depends on the radiating source. If a wire dipole antenna is placed at the
origin and aligned with the x axis, then in the y z plane, the electric field will have only an Ex
component. Away from the y z plane, other components are non-zero, because of the curvature of
the E-field lines being closed loops.

8-2 AJW, EEE3055F, UCT 2012


Electromagnetic propagation of a short pulse
x
E(z, t) = hEx (z, t), 0, 0i

z
c
H(z, t)

Figure 8.1: Propagation of a short EM pulse.

What about the magnetic field?


The magnetic field is coupled to the electric field in Maxwells equations,
particularly,
B
E=
t
Expanding the curl of E in Cartesian coordinates,

x y z x y z

E = x = 0 0 = Ex y
y z z
E E E Ex 0 0 z
x y z

Thus
B Ex
= y
t z
from which we conclude that the time-varying magnetic field has only a y
component! i.e.  Z 
Ex
B= t y
z
It is pointed out that, mathematically, time-independent x, y and z com-
ponents could exist, however these DC components are not generated by
any source, and are therefore zero. Ex is a travelling wave of the general

8-3 AJW, EEE3055F, UCT 2012


form Ex (z, t) = f (z vt). Substituting, we get
Z  
Ex
By (z, t) = t
z
Z
= (f (z vt)) t
1
= f (z vt) + h(z) + C
v
where h(z) is some function of z and C is a constant. We have the additional
condition that By (z, t) is a travelling wave of the form By (z, t) = fn (z vt)
(the solution to the wave equation). The first term is of this form, thus
1
By (z, t) = Ex (z, t)
v

8.2 Characteristic Impedance


Often, in engineering, the H field is described; the Hy component is related
to the Ex component by
1 1 1
Hy (z, t) = By = Ex (z, t) = Ex (z, t)
v
p
where the constant = v = 1 = is called the characteristic
impedance of the medium and is the ratio of the electric to magnetic fields
Ex m1
i.e. = H y
. The units of are VAm 1 = V A
1
= ohms, hence of the use of
the term characteristic impedance.
For a plane wave propagating in free space, the characteristic impedance
is r r
0 4 107
0 = = 377 ohms (in free space)
0 8.85 1012
In air, 0 and 0 , and so the characteristic impedance of free space
is usually used for propagation in air.
In other media, for example a dielectric material like glass with 0
and = r 0 where r 5 say, the characteristic impedance is reduced to
r
r 0 0 377
= = 170 ohms (in glass)
r 0 r 5

8-4 AJW, EEE3055F, UCT 2012


8.3 Sinusoidal Representations
The sinusoidal representation of the forward travelling wave is then the EM
pair
Ex (z, t) = E1 cos(t kz + 1 )
1
Hy (z, t) = Ex (z, t) = H1 cos(t kz + 1 )


where k = /v = , and H1 = 1 E1.
The following illustration depicts the sinusoidal electric and magnetic fields
at a fixed instant in time:
x

E(z, t) = hEx (z, t), 0, 0i

H
E = Ex x

direction of propagation
n
z

E
1
H= Ex y

y

Figure 8.2: Sinusoidal EM wave propagating in the z direction.

NOTES:
The electric and magnetic components are perpendicular and are both
normal to the direction of propagation.
For the more general case of a plane wave propagating in an arbitrary
direction, involving Ex , Ey and Ez components , the associated H
vector is perpendicular, and concisely related by
1
H= nE

is a unit vector pointing in the direction of propagation.
where n
The cross product E H points in the direction of propagation.

8-5 AJW, EEE3055F, UCT 2012


Complex Phasor Representation
By assuming steady state complex exponential solutions, the standard wave
equation can be simplified and the solutions obtained from the simplified
equations. The 1-D wave equation representing the electric field component
of a plane wave propagating in the z direction is
2 Ex 2 Ex
2 = 0
z 2 t
We assume a solution of the form Ex (z, t) = Ex (z)ejt , where Ex (z) is a
(complex) function to be determined. Substituting into the above wave
equation, we get
2Ex jt
2
e + 2Ex ejt = 0
z
The term ejt term cancels yielding

2Ex
2
+ 2Ex = 0
z
which is now a simpler differential equation as it involves only functions
of z and not of time. The complete solution to this 2nd order ordinary
differential equation is known to be

Ex (z) = E1ejkz + E2ejkz

where E1 = E1ej1 and E2 = E2ej2 are constants (in general complex).


The real signal it represents is found by multiplying by ejt and taking the
real part, i.e.

Ex (z, t) = E1 cos(t kz + 1) + E2 cos(t + kz + 2 )

8-6 AJW, EEE3055F, UCT 2012


8.4 Plane Wave Propagating in an Arbitrary Direction
A plane wave propagating in an arbitrary direction in Cartesian coordinates
is related to the plane wave propagating in the z direction by a rotation
of the axes. The diagram illustrates a plane wave propagating in the k
direction.

l-axis

direction of propagation
k

r cos
k r=
l=0
l-axis x
r planes of constant
phase

z
y

Figure 8.3: Plane wave propagating in an arbitrary direction.

The phasor representation of an electromagnetic wave propagating in an


arbitrary direction can be compactly written as the pair
t) = E1 ej(tkr+) = E
E(r, 1 ejkr ejt

= 1k
H E


where

x
r = y specifies position in space.
z

8-7 AJW, EEE3055F, UCT 2012



kx
k = ky = k k is a vector that points in the direction of prop-
kz
q
agation. The wave number is given by k = kx2 + ky2 + kz2 , and the
2
wavelength = k .

The term k r accounts for the phase change in the direction of prop-
agation. To visualise this refer to the accompanying sketch. The term
r is just k times the component of r in the k
k r = kk direction, i.e.
k r = k r cos .
One can define an axis (labelled l-axis) pointing in the k direction,
as shown in the sketch. The quantity l = k r = r cos , is the distance
measured along the l-axis from its origin. Then factor

ejkr = ejkl

is easily recognised as spatially varying phase factor, along the axis of


propagation.
1 = E1 ej is a (complex) vector that is perpendicular to the direction
E
of propagation (i.e. E 1 k = 0), and lies in the plane of the plane
wave.
1 determines the orientation of the E field in the plane, i.e. in the ori-
E
entation of the linear polarization. E 1 determines both the amplitude,
phase shift, and polarization orientation of the plane wave. It can be
made complex to allow for an arbitrary phase shift .

8-8 AJW, EEE3055F, UCT 2012


8.5 Polarization
The polarization of a plane wave refers to the orientation of the E-field
vectors in the plane perpendicular to the direction of propagation. Up till
now, we have considered only the simplest case known as linear polarization,
in which the E field lines are orientated at a fixed angle in the plane.
More generally, a forward travelling sinusoidal EM wave travelling in the
+z direction can be represented as the sum of two orthogonal components,

E(z) = xE1ejkz + yE2ej ejkz

where E1 and E2 are the amplitudes of the x and y components, and


represents a possible relative phase shift between them. The associated H
field is given by

1 1
H(z) x E2ej ejkz + y E1ejkz
=

The physical sinusoidal electric and magnetic fields are modelled by the pair

E(z, t) = xE1 cos(t kz) + yE2 cos(t kz + )


1 1
H(z, t) = x E2 cos(t kz + ) + y E1 cos(t kz)

From the real forms can be derived the magnitude and orientation of the
E-field vector as a function of time. There are three classes of polarization,
which depend on the relative amplitudes of E1 and E2, and the phase shift
. These are described in detail in the following subsections.

8.5.1 Linear Polarization: Ex and Ey are in phase i.e. = 0


The electric vector is orientated at an angle determined by the relative values
of of amplitudes E1 and E2. The amplitude of the vector varies sinusoidally,
illustrated by a sequence of snapshots in a fixed plane (e.g. at z = 0), at
different fractions of the period T of the sinusoidal wave.

8-9 AJW, EEE3055F, UCT 2012


Linear Polarization - snapshots in time.

E
Ey

Ex

t=0 t = T /8 t = T /4 t = 3/8 t = T /2

Figure 8.4: Linear polarization.

The time varying magnitude of the resultant E vector is given by


q q q
Ex + Ey = E1 cos (t kz) + E2 cos (t kz) = E12 + E22 |cos(t kz)|
2 2 2 2 2 2

The orientation angle may be calculated from


Ey E2
tan = =
Ex E1
and jumps 180 degrees, when the resultant passes through zero. It is noted
that the magnitude of the resultant H vector is 1 times the magnitude of
the resultant E vector, as can be seen from
s 2  2
q E Ex 1
y
Hx2 + Hy2 = + = E

It is common in telecommunications and radar engineering to describe the
polarization by the orientation of the electric field vector: vertical polariza-
tion if the electric field is orientated vertically, and horizontal polarization
if the electric field is horizontal. The orientation of the polarization is de-
pendent on the orientation of the radiating antenna.

8-10 AJW, EEE3055F, UCT 2012


8.5.2 Circular Polarization: E1 = E2 and = 2
If the two orthogonal components are equal in amplitude, but with a rel-
ative phase shift of 90 degrees, the total electric field vector will rotate
as a function of time if viewed in a plane perpendicular to the direction of
propagation at a fixed location in space.
Imagine holding up a transparent sheet of paper through which the EM
wave passes. The electric field lines in the plane of the paper will rotate as
a function of time at a rate of rad/s, but the magnitude, indicated by the
line spacing, does not change as illustrated below.
Circular Polarization - snapshots in time.

t=0 t = T /8 t = T /4 t = 3/8 t = T /2

Figure 8.5: Circular polarization - snapshots at increasing times, in the same plane.

The E-vector at a point on the plane rotates as illustrated below, where


in this alternative representation, the magnitude is indicated via the length
of the vector.
Circular Polarization - snapshots in time (at a fixed position, z=0)
Wave propagates in +z direction out of page.

y E

x (t)

t=0 t = T /8 t = T /4 t = 3/8 t = T /2

Figure 8.6: Circular polarization - snapshots at increasing times, in the same plane.

The diagram below illustrates a side view snapshot along the axis of propa-
gation at a fixed instant in time - the vectors wind around like a corkscrew.

8-11 AJW, EEE3055F, UCT 2012


The corkscrew moves forwards in the z direction as time increases.

Figure 8.7: Circular polarization - E-vectors shown at a fixed instant in time. Note: this
is right-hand circular polarization.

To analyse the case of circular polarization, we set E1 = E2 = A and


= 2 . The magnitude of the resultant E-vector is given by
q q
Ex + Ey = A2 cos2 (t kz) + A2 sin2 (t kz) = A
2 2

which is independent of z and t. The orientation angle is however time-


varying, and may be calculated from
Ey A cos(t kz 2 ) A sin(t kz)
tan = = = = tan(t kz)
Ex A cos(t kz) A cos(t kz)
Thus the angle is

= (t kz) mod 2 if =
2
or

= (t kz) mod 2 if =
2
NOTE: At a fixed position z, the vector rotates at a rate of radians per
second in a circle in the plane. At a fixed time t, the vector rotates at k
radians per metre (with a cork-screw locus).

8-12 AJW, EEE3055F, UCT 2012


The IEEE defines two types of circular polarization, according to the
direction of rotation:
Left-hand circular: If the E vector of a circularly polarized plane wave
propagating out of the page rotates in the clockwise direction in the plane
of the page. You can imagine drawing the E vectors position on the page,
and watching it rotate clockwise. Another way to remember the definition
is to point the thumb of your left hand in the direction of propagation (i.e.
z-direction) with your fingers bent - your curled fingers will point in the
direction of rotation of the E vector in the x-y plane (i.e. the x-y plane is at
a fixed position, say z=0). It is noted that in our particular example, this
corresponds to the case where = 2 .

Right-hand circular if you use your right hand, with thumb pointing in
the direction of propagation (z direction), your curled fingers will point in
the direction of rotation of the E vector in the x-y plane (i.e. the x-y plane
is at a fixed position, say z=0). In our particular example, this corresponds
to the case where = 2 .

The illustrations below show the directions of rotation in an x-y plane for
the case of a plane wave travelling in the z-direction, being out of the page.
Left-Hand Circular Polarization Right-Hand Circular Polarization

y y
E E

x x
z z

(Propagation out of page in z-direction)

Figure 8.8: Left-hand versus right-hand polarization.

One can easily plot the E vector in the xy plane for t = 0, T /8, 2T /8, 3T /8, 4T /8
by considering the Ex and Ey components in our polarization model

E(z, t) = xE1 cos(t kz) + yE2 cos(t kz + )

8-13 AJW, EEE3055F, UCT 2012


For example, for the case of = 2 , the E field components at z = 0 are
Ex = A cos(t)

Ey = A cos(t )
2
The Ex and Ey values are plotted as functions of time below:
Ex = cos(t) y
T /4
T /4 T /2 3T /8 T /8
t
T /2 x
t=0
Ey = cos(t /2)
T /4
T /2
t
E vector rotates in z = 0 plane

Circular Polarization - snapshots in time (at a fixed position, z=0)


Wave propagates in +z direction out of page; hence this is right hand circular polarization

Figure 8.9: Circular polarization - Ex and Ey components plotted against time.


8.5.3 Elliptical Polarization: Either E1 6= E2 or 6= 0, 2 , 2
Elliptical polarization is the general case, in which the vector not only ro-
tates, but also varies in length, tracing out an ellipse in a plane at a fixed
position, as illustrated in the sequence below (which is left-hand elliptical
polarization):

Elliptical Polarization - snapshots in time.


y E

t=0 t = T /8 t = T /4 t = 3/8 t = T /2

Figure 8.10: Elliptical polarization - snapshots in time.

8-14 AJW, EEE3055F, UCT 2012


In all three classes of polarization, the H-field is always perpendicular to
the E-field, and in phase with it, as illustrated below:

Ey (t) E
E
H
H
Ex (t)

Linear Polarization Circular Polarization


- fixed orientation of resultant - rotating vector traces a circle
- amplitude varies sinusoidally.

E
H

Elliptical Polarization
- rotating vector traces an ellipse

Figure 8.11: All three classes of polarization, showing both E and H vectors, and the loci
of the tips of the vectors.

8.5.4 Applications of polarization


Communication Links

In broadcast transmitters e.g. radio, TV signal, linear polarization is most


commonly used. Wire antennas like dipoles, monopole and Yagi antennas
radiate linearly polarized radiation. Both horizontal and vertical polariza-
tion is used. The orientation of the transmitting antenna determines the
orientation of the polarization. The receiving antenna must be orientated
correctly to receive the maximum signal strength.
Consider for example, a dipole antenna receiving a vertically polarized EM
wave. The dipole must be orientated vertically to receive the signal - the

8-15 AJW, EEE3055F, UCT 2012


electrons rush vertically up and down the wires in response to the incoming
E-field, resulting in a voltage response across the centre terminals. If the
antenna is rotated 90 degrees to the electric field, it should be clear that the
electrons do not move up and down the rod as before, and so a potential
difference does not appear across the terminals. At an arbitrary angle, the
signal drops off as the cosine of the angle from the vertical (i.e. with the
component of the electric field in the direction of the dipole rod).

Figure 8.12: Vertical versus horizontal polarization from a radiating dipole.

Radar - Radio detection and ranging

Radar is a technique used to detect far away targets like aeroplanes ap-
proaching an airport, or ships out at sea. A pulse is transmitted in the
direction of interest, and the return echo is received and used to detect the
presence of targets. The range to the target is inferred from the time delay
of the echo, and the direction obtained from the pointing direction of the
antenna.
Circular polarization is sometimes used in radar applications as some scat-

8-16 AJW, EEE3055F, UCT 2012


tering structures only reflect electromagnetic energy of a particular polar-
ization. In circular polarization, the E-field vector rotates in a plane at the
reflector (being a fixed distance from the source), which ensures that there
will aways be some reflection from such structures.
For example, a vertically orientated thin wire rod will only reflect vertical
polarization, where as a horizontally orientated wire rod will only reflect
horizontal polarization. Circular polarization can be thought of as the sum
of two orthogonal, linearly polarized components. On reflection, only one
component, in line with the orientation of the reflecting rod will be re-
flected. An incident circularly polarized wave will be linearly polarized on
reflection. The received signal will be 3dB lower than one could obtain from
transmitting linear polarization, as only half the power is reflected.
Circular polarization can be generated by using two perpendicular dipoles
(crossed-dipoles), driven 90 degrees out of phase. Alternatively, the perpen-
dicular dipoles may be driven in phase, but physically separated by a quarter
wavelength in the propagation direction to achieve the required 90 degree
phase shift.

8-17 AJW, EEE3055F, UCT 2012


9 Simulating Electromagnetic Waves
using the Finite Difference Time
Domain (FDTD) Method [not
covered in 2009]
With the availability of fast computers with ever increasing memory and
computational performance, the simulation of wave propagation has been
made possible. In this section, a powerful method known as the finite dif-
ference time domain (FDTD) method for simulating the propagation of
electromagnetic waves will be introduced. To simplify the discussion, the
method will be described initially for wave propagation in 1-D space, fol-
lowed by the extension to 2-D, and 3-D1 . This chapter is meant to provide
only a brief overview.

9.1 Introduction
The finite difference time domain (FDTD) method is a full-wave, dynamic,
and powerful solution tool for solving Maxwells equations, introduced by
K.S. Yee in 1966 [Yee, 1966]. The algorithm involves direct discretizations of
Maxwells equations by writing the spatial and time derivatives in a central
finite difference form2.
The time-dependent Maxwells curl equations in a homogeous dielectric
medium ( = 0 r , = 0 , r = 1) are
1
The course coordinator wishes to thank doctoral student Pradip Mukhopadhyay for his assistance in
preparing the material in this section on FDTD simulation.
2
The central finite difference approximation for the derivative of the function f (x) at point P (x0 ) can be
written as
df (x0 ) f (x0 + x/2) f (x0 x/2)
= f (x0 )
=
dx x

9-1
E 1
= H (9.1)
t 0 r
H 1
= E (9.2)
t 0
E and H are vectors in three dimensions. The constants 0 and 0 are known
as the permittivity and permeability of free space and r is the relative
permittivity of the material.

9.1.1 Curl Equations in Cartesian Coordinates


Expanding
E 1
= H
t 0 r
in Cartesian coordinates3,
h     i
Ex Ey Hy Hy
x t +
y t + Ez
z t = 0 r x H
1
y
z
z + y Hx
z Hz
x + z x Hx
y

Equating the vector components, we obtain three equations, one for each
vector component
 
Ex 1 Hz Hy
=
t 0 r y z
 
Ey 1 Hx Hz
=
t 0 r z x
 
Ez 1 Hy Hx
=
t 0 r x y
Similarly expanding
H 1
= E
t 0
as


x
y z     
Hy Hy
3
H=
x

y

z
=x
Hz
y z + y Hx
z Hz
x + z x Hx
y
Hx Hy Hz

9-2 AJW, EEE3055F, UCT 2012


h     i
H Ey Ey
x H x
z H
y ty +
t + t
z
= 10 x E
y
z
z + y Ex
z Ez
x + z x Ex
y

we obtain three more equations,


 
Hx 1 Ey Ez
=
t 0 z y
 
Hy 1 Ez Ex
=
t 0 x z
 
Hz 1 Ex Ey
=
t 0 y x
In 1-D, we consider (i) exciting an Ex component, and assume Ey = 0

and Ez = 0 and (ii) no variation in the x-y plane, i.e. x = 0 and y = 0.
The equations reduce to
 
Hx 1 Ey Ez
= =0
t 0 z y
 
Hy 1 Ez Ex 1 Ex
= =
t 0 x z z
  0
Hz 1 Ex Ey
= =0
t 0 y x
 
Ex 1 Hz Hy 1 Hy
= =
t 0 r y z 0r z
Furthermore, if the source has no DC component, then Hx = 0 and Hz =
0, leaving only the Ex and Hy components.

9.2 FDTD Solution to Maxwells equations in 1-D


In a simple one-dimensional case, we will consider the case where only the
Ex and Hy components exist - consistent with modelling plane wave prop-
agation far away from an antenna. Equations 9.1 and 9.2 become
Ex 1 Hy
= (9.3)
t 0r z
Hy 1 Ex
= (9.4)
t 0 z

9-3 AJW, EEE3055F, UCT 2012


These equations represent a plane wave with the electric field orientated
in the x-direction and magnetic field oriented in the y-direction and trav-
elling in the z-direction. In the FDTD formulation, the central difference
approximations for both the temporal and spatial derivatives are obtained
at (z = kz, t = nt) for the first equation:

n+1/2 n1/2
Ex (k) Ex (k) 1 Hyn (k + 1/2) Hyn (k 1/2)
= (9.5)
t 0 r z
and at (z + z/2, t + t/2) for the second equation:
n+1/2 n+1/2
Hyn+1(k + 1/2) Hyn (k + 1/2) 1 Ex (k + 1) Ex (k)
= (9.6)
t 0 z
In the equations above, n is the time index and k is the spatial index, which
indexes times t = nt and positions z = kz, or positions t = (n 1/2)t
and positions z = (k 1/2)z. The time index is written as a superscript,
and the spatial index is within brackets.
Equations 9.5 and 9.6 can be rearranged as a pair of computer update
equations, which can be repeatedly updated in loop, to obtain the next time
n+1/2
values of Ex (k) and Hyn+1(k + 1/2), corresponding the Ex (t + t/2, z)
and Hy (t + t, z + z/2).
Figure 9.1 illustrates the interleaving of the E and H fields in space and
time in the FDTD formulation.

9-4 AJW, EEE3055F, UCT 2012


n1/2
Ex

k2 k1 k k+1 k+2

n
Hy

k1 1/2 k1/2 k+1/2 k+1 1/2 k+2 1/2

n+1/2
Ex

k2 k1 k k+1 k+2

Figure 9.1: Interleaving of the E and H fields in space and time in the FDTD formulation.
To calculate Hy (k + 1/2), for instance, the neighbouring values of Ex at k and
k + 1 are needed. Similarly, to calculate Ex (k + 1), the value of Hy at k + 1/2
and k + 1 21 are needed (Sullivan 2000).

In equations 9.5 and 9.6 0 and 0 differ by several orders of magnitude,


Ex and Hy will differ by several orders of magnitude. Numerical error is
minimised by making the following change of variables as
r
0
Ex = Ex (9.7)
0
which bring the field quantities to similar levels. Implementing the changing
of variables, equations 9.5 and 9.6 become
1 t  n 
Exn+1/2(k) = Exn1/2(k) Hy (k + 1/2) Hyn (k 1/2) (9.8)
r 0 0 z

1 t h n+1/2
i
Hyn+1(k + 1/2) = Hyn (k
+ 1/2) E (k + 1) Exn+1/2
(k)
0 0 z x
(9.9)
Stability and the FDTD method: For stability purposes, we need
to choose the cell size z to allow 10 to 15 points per wave length. In
free space, an electromagnetic wave travels a distance of one cell in time
t = z c0 , where c0 is the speed of light in free space. This limits the

9-5 AJW, EEE3055F, UCT 2012


maximum time step. In the case of a 2-D simulation, we have to allow for
the propagation in the diagonal direction, which brings the time requirement
to t = z
2c0
. Obviously, three-dimensional simulation requires t = z3c0
.
We will use in all our simulations a time step t of
z
t = , (9.10)
2 c0
where c0 is the velocity of light in free space, which satisfies the requirements
in 1-D, 2-D and 3-D for all media. The factor

1 t z/(2 c0 )
= c0 = 0.5 (9.11)
0 0 z z/2
Making use of equation 9.11 in equation 9.8 and 9.9, we obtain the following
computer update equations:

0.5
ex(k) = ex(k) + [hy(k 1) hy(k)] (9.12)
r (k)
hy(k) = hy(k) + 0.5 [ex(k) ex(k + 1)] (9.13)

which are used repeatedly in a loop to update the field quantities at every
position at all positions in space, as time progresses. Note that the n or
n + 1/2 or n 1/2 in the superscripts do not appear. In equation 9.12, the
ex on the right side of the equal sign is the previous value at n 1/2, and
the ex on the left side is the new value, n + 1/2, which is being calculated.
In case of the spatial index, k + 1/2 and k 1/2 are replaced by k and k 1
in order to specify an integer position in an array. It is understood from
the derivation, however, that the value stored in hy(k) is the H value at
position k + 1/2.
The value of ex(k) on the nth iteration, n+1/2(k), which is
q represents Ex
related to the electric field by Ex = Ex / 00 .

9.2.1 1-D FDTD code in Matlab


This section contains a code segment written in Matlab, in which the 1-D
FDTD algorithm is implemented. An electromagnetic pulse is radiated from

9-6 AJW, EEE3055F, UCT 2012


a source located in free space. The source waveform is a Gaussian shaped
pulse, injected at the centre of the array used to store the pulse in space.
This is achieved by repeatedly updating the E-field pulse value at the source
location; the values at the other locations are computed using the update
equations. Figure 9.2 shows snapshots of the simulation as time progresses.
The E-field pulse is seen to propagate away from the source to the left and
to the right. The corresponding H field is also calculated, but not plotted
here.
Note: In free space r = 1

1) Solution: The MATLAB file : fdtd 1d 1.m

KE = 201; % number of z samples.


kc = fix(KE/2) ; % centre of the grid
ex = zeros(1,KE); % initialize Ex field
hy = zeros(1,KE); % initialize Hy field
cb = zeros(1,KE);
cb(1,:) = 0.5;
% A Gaussian pulse parameters
t0 = 36.0 ; % location in time of the pulse
spread = 12 ; % Width of the pulse
NSTEPS = 190 ; % number of time step
for n=1:NSTEPS
for k=2:KE
ex(1,k) = ex(1,k) + cb(1,k).*(hy(1,k-1) -hy(1,k));
end
pulse = exp(-0.5*((t0-n)/spread)^2.0); %The Gaussian pulse
% adding the pulse at the centre of the grid
ex(1,kc) = pulse;
for k=1:KE-1
hy(1,k) = hy(1,k) + 0.5*(ex(1,k) -ex(1,k+1)) ;
end
plot(ex);
pause(0.05);
end

9-7 AJW, EEE3055F, UCT 2012


1D EM Propagation
1

0.8

Time step=35
0.6

0.4

0.2

0.2

0.4

0.6

0.8

1
20 40 60 80 100 120 140 160 180 200

1D EM Propagation
1

0.8

Time step=60
0.6

0.4

0.2

0.2

0.4

0.6

0.8

1
20 40 60 80 100 120 140 160 180 200

1D EM Propagation
1

0.8

0.6 Time step=120

0.4

0.2

0.2 Source location

0.4

0.6

0.8

1
20 40 60 80 100 120 140 160 180 200

9-8 of a Gaussian
Figure 9.2: 1-D Ez propagation. Propagation AJW,
pulseEEE3055F,
away fromUCT 2012
a source
located at the centre.
9.2.2 Wave hitting a dielectric medium (Matlab code)
We now consider the case where a plane wave travelling in free space
(medium 1) strikes a dielectric medium (medium 2), as is illustrated in Fig-
ure 9.3, which shows a source in free space on the left side, and a dielectric
slab on the right.

= 0 r
=
0

Source z

(or k)

kstart

Figure 9.3: EM wave hitting a dielectric surface

When the wave strikes the interface, a fraction of the incoming wave is
reflected, and a fraction is transmitted into the medium. The amplitude
of the reflected and transmitted waves, relative to the incident wave, are
described by the reflection coefficient and the transmission coefficient ,
which relate the amplitudes of the E field waves. From theoretical analysis,
these can be determined in terms of the characteristic impedances of the
media, as

Eref 2 1
= = (9.14)
Einc 2 + 1
Etrans 22
= = (9.15)
Einc 2 + 1
where 1 and 2 are the impedances of the respective media given by

9-9 AJW, EEE3055F, UCT 2012


r
0 r
= (9.16)
0 r
In case of a non-magnetic medium (r = 1), equations 9.14 and 9.15 become,


1 2
= (9.17)
1 + 2

2 1
= (9.18)
1 + 2
where 1 and 2 are the relative permittivities of medium 1 and 2 respec-
tively.
The 1-D FDTD code can be easily adapted to model propagation against
a dielectric interface, as shown below. Running the simulation should show
reflected and refracted waves as in the snapshots in Figure 9.4. Note here
that the reflected pulse inverts in sign, as indicated by a negative reflection
coefficient (check this by calculating it).

Modify the Matlab program as follows:

* Choose the dielectric constant: eps_r=4.0 (for example)


* Define were to start the dielec-
tric slab: k_start=kc+kc/2 (for example)
* Put the dielectric material in one side as
cb(1,k_start:KE)=0.5/eps_r

Modify the above program as follows:


.
.
eps_r = 4.0;
k_start = kc+kc/2;
cb = zeros(1,KE);
cb(1,:) = 0.5;
cb(1,k_start:KE) = 0.5/eps_r;
.

9-10 AJW, EEE3055F, UCT 2012


.
for n=1:NSTEPS
.
.
plot(ex);
pause(0.05);
end
%----------------------------------------------------

9-11 AJW, EEE3055F, UCT 2012


1D EM Propagation striking a dielectric slab on the right side
1

0.8

Time step=100
0.6

0.4

0.2

0.2

0.4

0.6

0.8

1
20 40 60 80 100 120 140 160 180 200

1D EM Propagation striking a dielectric slab on the right side


1

0.8 Transmitted
Time step=190
0.6

0.4

0.2

0.2

0.4
Reflected

0.6

0.8

1
20 40 60 80 100 120 140 160 180 200

Figure 9.4: Reflection and transmission at a dielectric (1-D simulation)

9-12 AJW, EEE3055F, UCT 2012


9.2.3 Removing the unwanted reflection at the boundary - The
Absorbing Boundary Condition (ABC)
In calculating the E field, we need to know the surrounding H values; this
is the fundamental assumption in FDTD. At the edge of the problem space
we will not have the value to one side, but we know there are no sources
outside the problem space. The wave travels z 2 (=c0 t) distance in one
time step, so it takes two time steps for a wave front to cross one cell.
Suppose we are looking for a boundary condition at the end where k=1.
Now if we write the E field at k=1 as

Exn(1) = Exn2(2)

then the fields at the edge will not reflect. This condition must be applied
at both ends.

Implementation

Modify the afore program as follows:

This is easy to implement, store the value of Ex (2) for two time steps and
then put it in Ex (1).
Modify the above program as follows:
.
.
ex_left_m1 = 0.0;
ex_left_m2 = 0.0;
ex_right_m1 = 0.0;
ex_right_m2 = 0.0;
cb = zeros(1,KE);
cb(1,:) = 0.5;
% You may choose cb(1,k_start:KE)=0.5/eps_r
.
.
for n=1:NSTEPS
.
.
% add the pulse at the centre of the grid
ex(1,kc) = pulse;

9-13 AJW, EEE3055F, UCT 2012


% add these additional lines after ex(1,kc)=pulse
ex(1,1) = ex_left_m2; % left boundary
ex_left_m2= ex_left_m1;
ex_left_m1 = ex(1,2);
ex(1,KE) = ex_right_m2; % right boundary
ex_right_m2= ex_right_m1;
ex_right_m1 = ex(1,KE-1);
.
.
plot(ex);
pause(0.05);
end
%-----------------------------------------------
Solution: The MATLAB file : fdtd_1d_3.m
% modified programme (use ABC)

KE = 201; % number of sampled in z direction


kc = fix(KE/2) ; % centre of the grid
ex = zeros(1,KE); % initialize the Ex field
hy = zeros(1,KE); % initialize the Hy field
cb = zeros(1,KE);
cb(1,:) = 0.5;
ex_left_m1 = 0.0;
ex_left_m2 = 0.0;
ex_right_m1 = 0.0;
ex_right_m2 = 0.0;
% A Gaussian pulse
t0 = 36.0 ;
spread = 12 ;
NSTEPS = 200 ; % number of time step
for n=1:NSTEPS
for k=2:KE
ex(1,k) = ex(1,k) + cb(1,k).*(hy(1,k-1) -hy(1,k));
end
%The Gaussian pulse
pulse = exp(-0.5*((t0-n)/spread)^2.0);
% adding the pulse at the centre of the grid
ex(1,kc) = pulse;
ex(1,1) = ex_left_m2; % left boundary
ex_left_m2= ex_left_m1;
ex_left_m1 = ex(1,2);
ex(1,KE) = ex_right_m2; % right boundary
ex_right_m2= ex_right_m1;
ex_right_m1 = ex(1,KE-1);

9-14 AJW, EEE3055F, UCT 2012


for k=1:KE-1
hy(1,k) = hy(1,k) + 0.5*(ex(1,k) -ex(1,k+1)) ;
end
plot(ex);
pause(0.05);
end
%---------------------------------------------------

9.2.4 Some Exercises


1. In free space, put the source at the middle and see how the wave travels.
See what happens when the pulse hits the boundary ?
2. Modify the program so that it has two sources 20 cells away from the
middle. What happens when the pulse meet each other.
3. Divide the problem space into two, one side being free space and the
other a dielectric material. What happens after the pulse hits the
boundary? Look at the relative amplitudes of the reflected and trans-
mitted pulses. Check the values with equation 9.17 and 9.18.

9.3 FDTD Solution to Maxwells equations in 2-D Space


In deriving 2-D FDTD formulation, we choose between one of two groups
of three vectors each:
1. The transverse magnetic (TM) mode, which is composed of Ez , Hx ,
and Hy
or
2. The transverse electric (TE) mode, which is composed of Ex , Ey , and
Hz
We will work with the TM mode wave propagation. Expanding the curl

equations4 9.1 and 9.2 with z = 0, Ex = 0, Ey = 0 and Hz = 0, we obtain

x
y z   
4
E=
x

y 0 = x E
y
z
+ y E
x
z
for the TM mode.
0 0 Ez

9-15 AJW, EEE3055F, UCT 2012


E H
for the non-zero components of t and t :

Ez 1 Hy Hx
= ( ) (9.19)
t 0 r x y
Hx 1 Ez
= (9.20)
t 0 y
Hy 1 Ez
= (9.21)
t 0 x
Now we can write the above three equations in the finite difference scheme
as

n+1/2 n1/2  n
Ez (i, j) Ez (i, j) 1 Hy (i + 1/2, j) Hyn (i 1/2, j)
=
t 0 r x

Hxn (i, j + 1/2, ) Hxn (i, j 1/2)
(9.22)
y
n+1/2 n+1/2
Hxn+1(i, j + 1/2) Hxn (i, j + 1/2) 1 Ez (i, j + 1) Ez (i, j)
=
t 0 y
(9.23)
n+1/2 n+1/2
Hyn+1 (i + 1/2, j) Hyn (i + 1/2, j) 1 Ez (i + 1, j) Ez (i, j)
=
t 0 x
(9.24)

Rearranging the above equations, we can write the final expression as

 n
n+1/2 n1/2 t Hy (i, j + 1/2, ) Hyn (i, j 1/2)
Ez (i, j) = Ez (i, j) +
0 r x

Hxn (i, j + 1/2, ) Hxn (i, j 1/2)
(9.25)
y

x
y z  
H =
x

y 0 = z H y
x
Hx
y
Hx Hy 0

9-16 AJW, EEE3055F, UCT 2012


1 t  n+1/2 
Hxn+1(i, j + 1/2) = Hxn (i, j + 1/2) Ez n+1/2
(i, j + 1) Ez (i, j)
0 y
(9.26)

1 t  n+1/2 
Hyn+1(i + 1/2, j)
= Hyn (i + 1/2, j) +
Ez n+1/2
(i + 1, j) Ez (i, j)
0 x
(9.27)
For computer implementation we can write the above equations update
equations as

ez(i,j)=ez(i,j)+{dt/(eps_0*eps_r)}
*[{Hy(i,j)-Hy(i-1,j)}/dx
-{Hx(i,j)-Hx(i,j-1)}/dy]
Hx(i,j) = Hx(i,j)-{dt/(mu_0*dy)}*{Ez(i,j+1)-Ez(i,j)}
Hy(i,j) = Hy(i,j)+{dt/(mu_0*dx)}*{Ez(i+1,j)-Ez(i,j)}

9.3.1 2-D FDTD code in Matlab


The code listing in this section simulated the case of 2-D EM wave propa-
gation in free space, source at the centre (r = 1).

Solution: The MATLAB file : fdtd_2d_TM_1.m

clear all;
cc=2.99792458e8; % Speed of light in free space
mu_0=4.0*pi*1.0e-7; % Permeability in free space
eps_0=1.0/(cc*cc*mu_0); % Permittivity in free space
eps_r = 1.0; % Dielectric constant in free space
ie = 201; % Grid pixels in Y direction
je = 201; % Grid pixels in X direction

9-17 AJW, EEE3055F, UCT 2012


is = floor(ie/2) ;
js = floor(je/2) ;
% Electric and Magnetic field
Ez = zeros(ie,je);
Hx = zeros(ie,je);
Hy = zeros(ie,je);
nmax = 150; % Number of time steps
ddx = 1.0e-3; % X grid size
ddy = ddx; % Y grid size
dt = 0.98/(cc*sqrt( (1/ddx)^2 + (1/ddy)^2 ));
%*********************************************************
tw = 26.53e-12;
t0 = 4.0*tw;
T = (0:1:nmax-1).*dt;
source = -2.0*((T-t0)./tw).*exp(-1.0*((T-t0)./tw).^2.0);
% Plot injected pulse figure
plot(source) title(Source pulse)
pause(1);
Emax = max(source);
Emin = min(source);
C1 = dt/(eps_0*eps_r);
C2 = dt/mu_0;
figure
for n = 1:nmax
%------------------- Ez -----------------
for jj = 2:je
for ii = 2:ie
Ez(ii,jj) = Ez(ii,jj) + C1*((Hy(ii,jj) - Hy(ii-
1,jj))./ddx -(Hx(ii,jj) - Hx(ii,jj-1))./ddy);
end
end
%----------------- Inject source --------
Ez(is,js) = source(n);
%------------------- Hx -----------------
for jj = 1:je-1
for ii = 1:ie
Hx(ii,jj) = Hx(ii,jj) - C2*((Ez(ii,jj+1) -
Ez(ii,jj))./ddy);
end
end
%------------------- Hy -----------------
for jj = 1:je

9-18 AJW, EEE3055F, UCT 2012


for ii = 1:ie-1
Hy(ii,jj) = Hy(ii,jj) + C2*((Ez(ii+1,jj) -
Ez(ii,jj))./ddx);
end
end
%plot the Ez component
imagesc(Ez, 0.4*[Emin Emax]);
title(Ez component propagating in free space)
colorbar;
colormap(hot);
pause(0.0002);
end
%--------------------------------------------

Source pulse
1

0.8

0.6

0.4

0.2

0.2

0.4

0.6

0.8

1
0 50 100 150

Figure 9.5: Source pulse [differential Gaussian]

9-19 AJW, EEE3055F, UCT 2012


Ez component propagating in free space

20

40

60

80

100

120

140

160

180

200
20 40 60 80 100 120 140 160 180 200

Figure 9.6: Propagation of a pulse away from a source. The pulse moves outwards as an
expanding circle, centred on the source.

9.3.2 Wave hitting a dielectric surface (right side), source at the


centre (2-D)
We will examine two examples:
1. Wave hitting a dielectric surface (right side), with the source at the
centre (code: fdtd 2d 2.m)
2. Wave hitting a dielectric surface (top right corner) (code: fdtd 2d 3.m)
To simulate 1 or 2 above, modify the previous programme (fdtd 2d 1.m) as
follows:

9-20 AJW, EEE3055F, UCT 2012


Specify the position of the dielectric surface (id=....., jd=.....)
Choose a value of eps r = ......
Modify the constant value C1 according to that.....
Run the program, you will see that on one side (free space) the wave
is propagating freely and on the other you will notice reflection and
transmission of the incident wave at the interface.
Figure 9.7 shows the result of case 1, where the pulse strikes the dielectric
interface. Notice that the wave propagates more slowly in the medium on
the right, and the pulse length is shorter.

2D EM propagation striking a dielectric slab on the right side

20

40

60

80

100

120

140

160

180

200
20 40 60 80 100 120 140 160 180 200

Figure 9.7: 2-D. Time snapshot of the radiating pulse interacting with a dielectric interface.
The reflected and transmitted waves are seen on either side of the interface.

9-21 AJW, EEE3055F, UCT 2012


9.3.3 Multiple source 2-D wave propagation
A last exercise it to add a second source in the 2D simulation. Run the code
(code: fdtd 2d TM multi source.m).
Radiation from an aperture antenna (e.g. a microwave horn antenna)
can be simulated as a line of (closely spaced) radiating sources (Huyguns
wavelets).

9.4 Solving Maxwells Equations in 3D


The FDTD concepts described for 1-D and 2-D space can be easily extended
to 3-D. The main difference is that all three components of the E and H fields
must be updated within the time loop. The equations are more complicated,
and the simulation is more computationally intensive.
Details of 3-D FDTD implementation can be found in the references.

9.5 References
Yee, K.S., Numerical solution of initial boundary value problems in-
volving Maxwells equations in isotropic media, 1966.
Taflove, A., Computational electrodynamics : the finite-difference time-
domain method, Boston, Mass. : Artech House,1995
Sullivan, D. M., Electromagnetic simulation using the FDTD method,
New York : IEEE Press, 2000.
Kunz, K.S. and Luebbers, R.J., The Finite Difference Time Domain
Method for Electromagnetics, Boca Raton, FL; CRC Press, 1993.

9-22 AJW, EEE3055F, UCT 2012


9.6 MATLAB CODE
These code segments will be placed on the course web page for download.

1-D code wave hitting a dielectric slab


MATLAB file : fdtd_1d_2.m

KE = 201; % number sample in z dirn


kc = fix(KE/2) ; % centre of the grid
ex = zeros(1,KE); % initialize the Ex field
hy = zeros(1,KE); % initialize the Hy field
k_start=kc+kc/2;
eps_r=4.0;
cb = zeros(1,KE);
cb(1,:) = 0.5;
cb(1,k_start:KE)=0.5/eps_r;
% A Gaussian pulse
t0 = 36.0 ;
spread = 12 ;
NSTEPS = 190; % number of time step
for n=1:NSTEPS
for k=2:KE
ex(1,k) = ex(1,k) + cb(1,k).*(hy(1,k-1) -hy(1,k));
end
%The Gaussian pulse
pulse = exp(-0.5*((t0-n)/spread)^2.0);
% adding the pulse at the centre of the grid
ex(1,kc) = pulse;
for k=1:KE-1
hy(1,k) = hy(1,k) + 0.5*(ex(1,k) -ex(1,k+1)) ;
end
plot(ex)
pause(0.05);
end
%---------------------------------------------

9-23 AJW, EEE3055F, UCT 2012


1-D code : Reflection at the boundary (Absorbing Boundary
Condition(ABC))
% modified programme (use ABC)

Solution: The MATLAB file : fdtd 1d 3.m


KE = 201; %number of grid in z direction
kc = fix(KE/2) ; % centre of the grid
ex = zeros(1,KE); % initialize the Ex field
hy = zeros(1,KE); % initialize the Hy field
cb = zeros(1,KE);
cb(1,:) = 0.5;
ex_left_m1 = 0.0;
ex_left_m2 = 0.0;
ex_right_m1 = 0.0;
ex_right_m2 = 0.0;
% A Gaussian pulse
t0 = 36.0 ;
spread = 12 ;
NSTEPS = 200 ; % number of time step
for n=1:NSTEPS
for k=2:KE
ex(1,k) = ex(1,k) + cb(1,k).*(hy(1,k-1) -hy(1,k));
end
%The Gaussian pulse
pulse = exp(-0.5*((t0-n)/spread)^2.0);
% adding the pulse at the centre of the grid
ex(1,kc) = pulse;
ex(1,1) = ex_left_m2; % left boundary
ex_left_m2= ex_left_m1;
ex_left_m1 = ex(1,2);
ex(1,KE) = ex_right_m2; % right boundary
ex_right_m2= ex_right_m1;
ex_right_m1 = ex(1,KE-1);
for k=1:KE-1
hy(1,k) = hy(1,k) + 0.5*(ex(1,k) -ex(1,k+1)) ;
end
plot(ex);
pause(0.05);
end
%------------------------------------------------

9-24 AJW, EEE3055F, UCT 2012


Wave hitting a dielectric surface (right side), source at the centre
Solution: The MATLAB file : fdtd 2d 2.m

% fdtd_2d_TM_2.m
% Excercise ....
% 2D FDTD free space propagation striking dielec-
tric slab on right side
% For EEE355F AJW
clear all;
cc=2.99792458e8; %speed of light in free space
mu_0=4.0*pi*1.0e-7; %permeability of free space
eps_0=1.0/(cc*cc*mu_0); %permittivity of free space
ie = 200; %Grid pixels in Y direction
je = 200; % Grid pixels in X direction
is = floor(ie/2) ;
js = floor(je/2) ;
% Electric and Magnetic field
Ez = zeros(ie,je);
Hx = zeros(ie,je);
Hy = zeros(ie,je);
eps_r = ones(ie,je);
jd = js+15; % Position of interface in X direction
eps_r(:,jd:je) = 10.0; % dielectric constant
nmax = 150; % number of time steps
ddx = 1.0e-3; % X grid size
ddy = ddx; % Y grid size
dt = 0.98/(cc*sqrt( (1/ddx)^2 + (1/ddy)^2 ));
%*********************************************
tw = 26.53e-12;
t0 = 4.0*tw;
T = (0:1:nmax-1).*dt;
source = -2.0*((T-t0)./tw).*exp(-1.0*((T-t0)./tw).^2.0);
[Emax] = max(source);
[Emin] = min(source);
C1 = dt./(eps_r.*eps_0);
C2 = dt/mu_0;
for n = 1:nmax
%---------------------update Ez--------------------
for jj = 2:je
for ii = 2:ie
Ez(ii,jj) = Ez(ii,jj) + C1(ii,jj).*((Hy(ii,jj) - Hy(ii-
1,jj))./ddx -(Hx(ii,jj) - Hx(ii,jj-1))./ddy);

9-25 AJW, EEE3055F, UCT 2012


end
end
%-----------------------source-----------------
Ez(is,js) = source(n);
%--------------------update Hx---------------------
for jj = 1:je-1
for ii = 1:ie
Hx(ii,jj) = Hx(ii,jj) - C2*((Ez(ii,jj+1) - Ez(ii,jj))./ddy);
end
end
%--------------------update Hy--------------------
for jj = 1:je
for ii = 1:ie-1
Hy(ii,jj) = Hy(ii,jj) + C2*((Ez(ii+1,jj) - Ez(ii,jj))./ddx);
end
end
imagesc(Ez, 0.4*[Emin Emax]);
patch([jd jd],[1 ie],w);
title(2D EM propagation --
striking a dielectric slab on the right side)
colormap(hot);
pause(0.0002);
end

9-26 AJW, EEE3055F, UCT 2012


Wave hitting a dielectric surface (top right corner)
Solution: The MATLAB file : fdtd 2d 3.m

% fdtd_2d_TM_3.m
% Excercise ....
% 2D FDTD free space propagation striking dielec-
tric slab on top right corner
% For EEE355F AJW
clear all;
cc=2.99792458e8; % speed of light in free space
mu_0=4.0*pi*1.0e-7; % permeability of free space
eps_0=1.0/(cc*cc*mu_0); % permittivity of free space
ie = 150; % Grid pixels in Y direction
je = 150; % Grid pixels in X direction
is = floor(ie/2);
js = floor(je/2);
% Electric and Magnetic field
Ez = zeros(ie,je);
Hx = zeros(ie,je);
Hy = zeros(ie,je);
eps_r = ones(ie,je);
id = is-15; % dielectric interface
jd = js+15;
eps_r(1:id,jd:je) = 40.0; % dielectric constant
nmax = 120; % number of time steps
ddx = 1.0e-3; % X grid size
ddy = ddx; % Y grid size
dt = 0.98/(cc*sqrt( (1/ddx)^2 + (1/ddy)^2 ));
%*****************************************************
tw = 26.53e-12;
t0 = 4.0*tw;
T = (0:1:nmax-1).*dt;
source = -2.0*((T-t0)./tw).*exp(-1.0*((T-t0)./tw).^2.0);
[Emax] = max(source); [Emin] = min(source);
C1 = dt./(eps_r.*eps_0);
C2 = dt/mu_0;
for n = 1:nmax
%-------------------update Ez-----------------
for jj = 2:je
for ii = 2:ie
Ez(ii,jj) = Ez(ii,jj) + C1(ii,jj).*((Hy(ii,jj) - Hy(ii-
1,jj))./ddx -(Hx(ii,jj) - Hx(ii,jj-1))./ddy);
end

9-27 AJW, EEE3055F, UCT 2012


end
%--------------------source---------------------
Ez(is,js) = source(n);
%--------------------update Hx------------------
for jj = 1:je-1
for ii = 1:ie
Hx(ii,jj) = Hx(ii,jj) - C2*((Ez(ii,jj+1) - Ez(ii,jj))./ddy);
end
end
%---------------------update Hy-----------------
for jj = 1:je
for ii = 1:ie-1
Hy(ii,jj) = Hy(ii,jj) + C2*((Ez(ii+1,jj) - Ez(ii,jj))./ddx);
end
end
imagesc(Ez, 1.0*[Emin Emax]);
patch([jd jd],[1 id],w);
patch([jd je],[id id],w);
title(2D EM propagation --
striking a dielectric slab on the right corner);
colormap(hot);
pause(0.0002);
end

9-28 AJW, EEE3055F, UCT 2012


Multiple source 2D wave propagation
Solution: fdtd 2d TM multi source.m

clear all;
cc=2.99792458e8; % Speed of light in free space
mu_0=4.0*pi*1.0e-7; % Permeability of free space
eps_0=1.0/(cc*cc*mu_0); % Permittivity of free space
eps_r = 1.0; % Dielectric constant in air
ie = 201; % Grid pixels in Y direction
je = 201; % Grid pixels in X direction
is = floor(ie/2) ;
js = floor(je/2) ;
% Electric and Magnetic field
Ez = zeros(ie,je);
Hx = zeros(ie,je);
Hy = zeros(ie,je);
nmax = 400; % Number of time steps
ddx = 1.0e-3; % X grid size
ddy = ddx; % Y grid size
dt = 0.98/(cc*sqrt( (1/ddx)^2 + (1/ddy)^2 ));
%***********************************************
T = (0:1:nmax-1).*dt;
f_int = 20.0e+9; %Frequency (20 GHz) of the incident pulse
source = sin(2.0*pi*T.*f_int); %Sine wave
% Plot injected pulse
figure
plot(source)
title(Source pulse: 20 GHz Sine wave)
pause(1);
[Emax] = max(source);
[Emin] = min(source);
C1 = dt/(eps_0*eps_r);
C2 = dt/mu_0;
figure
for n = 1:nmax
%------------------------------ Ez -----------------
for jj = 2:je
for ii = 2:ie
Ez(ii,jj) = Ez(ii,jj) + C1*((Hy(ii,jj) - Hy(ii-1,jj))./ddx -
(Hx(ii,jj) - Hx(ii,jj-1))./ddy);
end
end

9-29 AJW, EEE3055F, UCT 2012


%------------------ Inject source -----------------
for ii = is-20:7:is+30
Ez(ii,js) = source(n);
end
%--------------------- Hx -------------------------
for jj = 1:je-1
for ii = 1:ie
Hx(ii,jj) = Hx(ii,jj) - C2*((Ez(ii,jj+1) - Ez(ii,jj))./ddy);
end
end
%--------------------- Hy ---------------------
for jj = 1:je
for ii = 1:ie-1
Hy(ii,jj) = Hy(ii,jj) + C2*((Ez(ii+1,jj) -
Ez(ii,jj))./ddx);
end
end
imagesc(Ez, 0.5*[Emin Emax]);
title(2D EM wave propagation: Multiple source)
colorbar;
colormap(gray);
pause(0.0002);
end

9-30 AJW, EEE3055F, UCT 2012


10 Power Considerations and the
Poynting Vector
Electromagnetic waves carry energy (the capacity to do work) through
space. At any point in space, the flow of energy can be described by a
power density vector P, which specifies both the power density in watts per
square metre, and the direction of flow. The vector P is called the Poynting
vector, and is a simple cross product of electric and magnetic field vectors,

P=EH

The units of E are Vm1 and the units of H are Am1 and thus the units
of P are VAm2 or Wm2 or Js1 m2. The total power passing through a
surface S1 is obtained by integration over S1, i.e.
Z
W = P dS watts
S1

is
The power dP passing through an elemental surface dS with normal n

dS = P dS
dP = |P| cos dS = P n

To understand why P = E H represents power flow, we consider the


rate of energy loss though from a volume V enclosed by closed surface S,
and relate this to the electric and magnetic fields. If U is the total energy
in V , the rate of energy loss is
I Z
dU
= P dS = PdV (by Divergence theorem)
dt S V

We re-express the term P by substituting P = E H, and making use

10-1
of a vector identities1 and Maxwells curl equations:

P = (E H)
= ( E) H ( H) E
H E
= H (J + ) E
t t
1 (H H) 1 E E
= JE
2 t 2 t
1 1
= ( H 2 ) ( E 2) J E
t 2 t 2
Thus
I Z
P dS = PdV
S VZ Z Z
1 1
= ( H 2 )dV ( E 2)dV J E dV
VZt 2 VZ t 2
ZV
d 1 d 1
= ( H 2 )dV ( E 2)dV J E dV
dt V 2 dt V 2 V
Z
d d
= UM UE J E dV
dt dt V

where
R
UM = V ( 21 H 2 )dV is the energy in the magnetic field within the vol-
ume V . The quantity 12 H 2 is the magnetic energy density in Jm3.
R
UE = V ( 12 E 2)dV is the energy stored in the electric field within the
volume V . The quantity 12 E 2 is the electric energy density in Jm3.
R
V J E dV is a term that represents either power dissipated through
ohmic losses or power generated by a source inside V.
If representing power dissipated, the term J E can be re-expressed as
J E = E E = E 2 in Wm3.

1
Identity 1: (E H) = ( E) H ( H) E
1 (HH)
Identity 2: H
t H = 2 t
2 2 2
Hy 1 Hx 1 Hy 1 Hz
H
t H = Hx H
t + Hy
x
t + Hz H
t =
z
2 t + 2 t + 2 t

10-2 AJW, EEE3055F, UCT 2012


Thus
I
P dS = magnetic energy loss +electric energy loss +heat loss (in Js1)
S
H H
We therefore conclude that S P dS = E H dS represents the outflow
of power from a volume, and P = E H is the power density vector, in
Wm2.

Example of Electric Energy stored in a Capacitor


As an example of energy stored in a field, let us consider the energy stored
in the electric field of a parallel plate capacitor, of surface area A, gap d,
and charged to a voltage V. The capacitance is approximately C A d
if
the gap is small compared to the plate dimensions. The electric field is
concentrated in the gap between the plates, and has strength of E V /d.
The energy stored in the electric field is
Z
1
UE = ( E 2)dV
V 2
1 V2
2 Ad
2 d
1 A 2
V
2 d
1
CV2
2
a familiar result from circuit theory. The relationship is exact (i.e. UE =
1 2
2 C V ), and can be shown from first principles for a capacitor of any shape.

10.1 Power in a Sinusoidal Plane Wave


A sinusoidal electromagnetic plane wave propagating in the z direction is
described by the EM pair

E(z, t) = Ex (z, t)
x
H(z, t) = Hy (z, t)
y

10-3 AJW, EEE3055F, UCT 2012


where
Ex (z, t) = E1 cos(t kz + 1 )
Ex (z, t) E1
Hy (z, t) = = cos(t kz + 1 )

The power density is
E12 2
P = E H = Ex Hy sin z = cos (t kz + 1)
z
2
At observer or antenna at a stationary point would feel the power pulsating
at twice the frequency of the source. The time averaged power density is
E12 2 E12 1 1 1 E12
P = |E H| = cos ( ) = ( + cos(2t 2kz + 21) = Wm2
2 2 2

10.1.1 Poynting Vector for Complex Notation


Phasor notation is commonly used for treatment of steady state sinusoidal
signals. The phasor representation of the sinusoidal wave EM plane wave is
= Ex x
E = E1ej1 ejkz x

= H yy E1 j1 jkz
H = e e y

The Poynting vector for complex phasor representation is defined as
P = 1E H
2
is the time averaged power density, i.e. Re{P}
such that the real part of P =
P = E H. To see why the factor of 1/2 is required, we proceeding as
before, but substituting the phasor forms of Maxwells curl equations:
I Z
H
(E ) dS = H
(E ) dV
ZV h i
=
( E) H ( H ) E dV
V
Z h i
=
(jH) H (J jE ) E dV
ZV h i
=
jH H + jE E E J dV
V

10-4 AJW, EEE3055F, UCT 2012


In the non-phasor case, J E represents the instantaneous dissipated (or
generated) power density in Wm3 - in other words J E is a function of
time. Here E J
= E E equals twice the time averaged power density
is defined as P
(= 12 E 2 ). Hence P = 1E H in order that the real part
2
equals time averaged power flow in Wm2.
If we substitute the phasor representations for a plane wave,

= 1E
P H
2
1 1
= E1 ej1 ejkz E1ej1 ejkz sin( )z
2 2
2
1 E1
= z Wm2
2
which agrees with the result previously obtained with the real signal repre-
sentation.

10.2 Power density from Radiating Antennas


An isotropic antenna is an idealised radiating source that radiates power
uniformly in all directions. The power density at a distance r from the
source is
Pt
Pd = 2
Wm2
4r
where Pt is the power in watts flowing from the source and 4r2 is the
surface area of a spherical shell of radius r.

Isotropic
radiator

Figure 10.1: An isotropic radiator radiating in all directions.

10-5 AJW, EEE3055F, UCT 2012


A physical antenna is usually designed to focus the energy in a particular
direction, and hence the power density increases compared to an isotropic
radiator. The increase in power density is specified by a parameter called
the gain of the antenna, which is defined as the factor by which the power
density is increased above that of an isotropic radiator. The power density
in position (r, , ) is given by
Pt
Pd = G(, ) 2
Wm2
4r
where G(, ) is the gain pattern of the antenna, which is a function of
angle. A typical gain pattern for a directional antenna is shown in Figure
10.2. The radiated power is concentrated within the main lobe. Additional
side lobes are visible which are usually undesirable, and antenna designers
will often seek to minimise the side lobe levels.
The beam pattern can be understood as an interference pattern. The
signal radiated in a particular direction can be modelled as the sum of con-
tributions from elemental radiation sources covering the aperture (imagine
a regular grid of tiny sources). The radiation contributions sum coherently
(in phase) in the main-lobe direction (constructive interference), whereas as
the angle increases of the boresight, the relative differences in path length
results in destructive interferences, which reduces the total signal strength
in the side lobes.

10-6 AJW, EEE3055F, UCT 2012


Directional Beam
radiator

Pt
3dB

Typical antenna gain versus angle


G() Main lobe

Sidelobe


180 deg 3dB 180 deg
3dB Beamwidth (G drops to 1/2 of max)

Figure 10.2: A directional antenna and a typical antenna gain pattern.

The width of a beam is usually measured between the half power points
on the main lobe, i.e. where G drops to half its peak value, or equivalently
3 dB below the peak.
A well known result from antenna theory is that the beamwidth from a
radiating aperture (measured in a chosen plane) is approximately

3dB arcsin
D D
where D is the diameter of the aperture in the plane, and is the wave-
length. The approximation is valid for narrow beam antennas, for which
the diameter of the aperture (measured in the plane) of the is several wave-
lengths wide.

10-7 AJW, EEE3055F, UCT 2012


Example
An antenna radiates 100 watts. Calculate the electric field intensity and
the magnetic field intensity at a distance r = 1000 m from the antenna
assuming the antenna is:
(i) an isotropic radiator
(ii) a horn antenna with a peak gain of 10.
Assume that the polarization is linear and the field is locally approximated
as a plane wave.

Solution:

(i) For an isotropic radiator, the power density


100
Pd = 2
= 8.0 106 Wm2
41000
2
The power density for a plane wave is P = 12 E which implies
p p
E = 2P = 2 377 8.0 106 = 0.08 Vm1
E 0.08
H= = = 2.1 104 Am1
377

(ii) For the horn antenna with a gain of 10,


100
Pd = 10 2
= 8.0 105 Wm2
41000
p p
E= 2P = 2 377 8.0 105 = 0.25 Vm1
E 0.25
H= = = 6.5 104 Am1
377

10-8 AJW, EEE3055F, UCT 2012


10.3 Power Flow in a Simple Circuit
Consider the circuit shown below consisting of a resistor connected to a
battery via a pair of wires. The resistor is made of a cylinder of resistive
material, placed between two metal plates as illustrated.
H
Metal plate
I

H into page
Battery P J Resistor
V E E Length d
P P
Ez H

I
H

Figure 10.3: Power flow in a DC circuit.

The magnetic field lines circulate around the wires and resistor as illus-
trated (apply right hand rule for direction). The electric field lines extend
from the wire at the positive voltage down to negative voltage wire. The
power flow at any point in space is described by P = EH. Below is shown
a top view of the circuit. The Poynting vector points from left to right.

10-9 AJW, EEE3055F, UCT 2012


TOP VIEW P P
(plane of battery
and resistor) P
H
P=EH
r
Integration surface
E into page
for calculating
P total power entering load.

P P

Figure 10.4: Top view showing power flow in a DC circuit,

Consider the region local to the resistor. The electric field between the
plates is vertically directed (z-direction), and is given by
V IR
Ez = =
d d
and the magnetic field surrounding the resistor is
I
H =
2r
The power flow at any point in space is described by the cross product

P = EH
= Ez H sin 90
r
IR I
=
r
d 2r
I 2R
=
r
2rd
which is a vector that points radially inwards as illustrated2. From this we
observe that
the power flows from the outside inwards towards the centre axis of the
column.
2
You should verify the direction E H by considering the cross product at several points in space.
Note that
r points radially outwards from the centre axis.

10-10 AJW, EEE3055F, UCT 2012


the conductor will get warm; i.e. molecules vibrate more vigorously as
electromagnetic energy is converted into kinetic energy.
energy is re-radiated by the mechanisms of
thermal radiation, also known as black body radiation (so called
because objects that are dark in colour radiate energy effectively
in the optical frequency range; dark objects are also good absorbers
in the optical band).
convection, i.e. kinetic energy transfer to the molecules of the
surrounding air.
conduction - i.e. through objects that are physically in contact
with the resistor, e.g. metal plates and wires.
The total power entering the resistor can be calculated by integrating the
Poynting vector over a closed surface enclosing the resistor. Since E 0
in the top metal plates, we need only consider the contribution through the
side walls of a cylinder of radius r (see Top View figure), which has surface
area = circumf erence length = 2r d:
I
I 2R
P = P dS 2rd = I 2 R Watts
2rd
which is in agreement with circuit theory.

Question: How much power flows through the wires to the load?

Answer: For perfect wires, , and E 0 within the wires. Thus


P = E H 0, within the wires, and we conclude that NO power flows
through the wires. From a field theory perspective, the power flows through
the air. The power (i.e. the capacity to do work) is carried in the EM field,
which travels from source to load. Why then do we need wires? The pair of
wires serve to guide the wave from source to load. Transmission line theory
is used to predict the behaviour of such guided waves when the length of
the transmission line is comparable to the wavelength.

10-11 AJW, EEE3055F, UCT 2012


10.3.1 Transmission Lines for 50 Hz AC Mains Electricity Supply
Wires pairs are used for getting EM power from a power station to a home,
which may be many kilometres away. At 50 Hz, however, the wavelength
= c/f = 3 108/50 = 6 106 m = 6000km, which is considerably greater
than the transmission line lengths, and so circuit theory with appropriate
lumped element models is more applicable for describing power transfer -
even over considerable distances like across a country.
Below is depicted a cross section through a pair of wires: the conductor on
the left is carrying current I into the page (in the direction of the load) and
the conductor on the right is carrying the return current out of the page.
Note the directions of the electric and magnetic field lines at this instant in
time. The Poynting vector P = E H points into the page, in the direction
of the load. Since the current is alternating AC, what happens to the
Poynting vector when the current reverses direction? (Answer: E H still
points into the page - you should check this yourself).

H H
E

P=EH
points into the page
I
I E out of page
into page
H

Figure 10.5: Electromagnetic field surrounding a parallel wire transmission line. If the
polarity of the current changes, the E and H field lines reverse direction
(although P = E H points always points in the same direction - into the
page).

10-12 AJW, EEE3055F, UCT 2012


10.3.2 Transmission lines for Microwave Circuits
In microwave circuits, the wavelength (e.g. at 10 GHz, = 3 cm) is often
comparable to circuit board dimensions, and so wires carrying signals must
be treated as transmission lines. A common method of circuit construc-
tion is to use a ground plane on the underside of the printed circuit board
(PCB), and the signals are routed via tracks on the top side (see figure).
The wavelength of the guided wave and the characteristic impedance of the
guiding structure are functions of the dielectric constant of the material, the
width of the track, and the thickness of the PCB. For standard fibreglass
PCB, a transmission line with a 50 Ohm characteristic impedance can be
made by etching a track of about 2mm wide on the top side; the underside
is a ground plane. Increasing the width of the line reduces the characteristic
impedance; reducing the track width increases the characteristic impedance
of the line.

Copper track Propagation


Substrate material direction
fibreglass
or special
low loss H
microwave board
e.g. RT Duroid. E

Ground plane
(copper)

Figure 10.6: Micro-strip microwave transmission line. The EM wave propagates as a guided
wave towards a (matched) load.

10-13 AJW, EEE3055F, UCT 2012


10.4 Electromagnetic safety considerations
The safe exposure level of EM radiation to humans is a function of radiation
frequency, intensity and the particular organs exposed.
Detailed guidelines have been drawn up by panels of experts, which specify
maximum safe exposure levels under a wide range of conditions [1,2]. In this
section, we shall briefly summarise some of the key issues. The appropriate
regulations should be consulted for details.
EM Radiation is classified into two categories:
1) ionizing radiation (above visible frequencies f > 7.5 1014 Hz or
< 0.0004 mm, including ultraviolet 1016 Hz, X-rays 1018 Hz, gamma
rays 1023 Hz and higher).
2) non-ionising radiation (f < 7.5 1014 Hz or > 0.0004 mmfrom
i.e. from the visible frequencies range right down to DC, which includes
the infrared, microwave and RF ranges).
Ionizing radiation can break chemical bonds (creating ions - hence
the name), resulting in tissue damage and damage to DNA. Even low
level exposure can lead to severe health problems.
Non-ionizing radiation primarily causes heating in tissue, and is con-
siderably less harmful than ionizing radiation.
The safety to humans is a function of the B field and E field strengths
to which one is exposed.
Figure 10.7 contains an illustration of the frequency spectrum showing the
effects on human tissue.
In these lecture notes, we shall only discuss safety levels for non-ionizing
radiation in the frequency range from DC to 300 GHz (i.e. wavelengths
greater than 1 mm).

10-14 AJW, EEE3055F, UCT 2012


Figure 10.7: [http://www.epa.gov/radiation/understand/ionize nonionize.htm, from the
U.S. Environmental Protection Agency]. The border between ionizing and
non ionizing radiation is at f 7.5 1014 Hz or = 0.0004 mm.

10.4.1 Non-Ionizing Radiation


The International Commission for Non-Ionizing Radiation Protection (IC-
NIRP) [http://www.icnirp.de/] is an organisation that provides guidelines
[1] relating to exposure to non-ionising radiation.
In the far field from a radiating source, the ratio of E to B is a constant
(E/H = ). Hence it is sufficient to specify either the E field strength,
or the B field strength, or conveniently, the power density P = EH, as
a safety level. Usually the power density in Wm2 is specified.
Closer to a radiating source, in the so called near field, the ratio of
E/H is not constant, varying considerably as a function of position
in relation to the source, and the type of source itself (e.g. close to a
dipole antenna, the E field is comparatively stronger, whereas close to
a loop antenna, the B field is dominant. Thus specifying safety levels
is more complicated in the near field.
Electric fields inside a human body cause currents to flow: J = E

10-15 AJW, EEE3055F, UCT 2012


Parameter Units Frequency Range
Current density J Am2 DC to 10 MHz
Current I A up to 110 MHz
Specific Energy Absorption Rate (SAR) W or Wkg1 100 kHz to 10 GHz
Specific Energy Absorption (SA) Jkg1 for pulsed fields 300 MHz to 10 GHz
Power Density P (or symbol S) Wm2 10 GHz to 300 GHz

Table 10.1: Parameters used for specifying maximum levels for exposure to ionizing
radiation.

where is the conductivity in Sm1 in tissue and body fluids. Below


100 Hz, typical values of are 0.4 Sm1 in muscle, 0.7 Sm1 in blood,
and 0.04 Sm1 in fat. At higher frequencies, muscle in the tongue varies
from 0.4 Sm1 at 10 MHz, to 1.0 Sm1 at 1 GHz, and to 10 Sm1 at
10 GHz.
The degree of penetration of the field is also a function of frequency
(the skin depth phenomenon).
The ICNIRP guidelines [1] are specified in terms of maximum levels of one
or more of the parameters listed in Table 10.1. The guidelines cover the
frequency range from 300 GHz to DC.

10.4.2 Coupling mechanisms between EM Fields and the Human Body

There are three identified mechanisms:


1. Coupling of low frequency E fields:
E fields cause currents to flow in the body, as well as polarization of
bound charge, and reorientation of electric dipoles in the tissue.
2. Coupling of low frequency E fields:
Changing magnetic fields induce E fields and hence current flow.
3. Absorption of Energy from EM fields:

for frequencies < 100 kHz, the energy absorption is minimal


> 100 kHz can lead to significant energy absorption and temperatures
rises in the body.

10-16 AJW, EEE3055F, UCT 2012


Four frequency ranges are described (Durney et al. 1985, in [1]):
1. From 100 kHz to 20 MHz
2. From 20 MHz to 300 MHz: high absorption can occur in the whole
body. The adult human body resonance is between 70 MHz and 100
MHz depending on the height of a standing person, for vertical polar-
ization.
3. From 300 MHz to 10 GHz: significant local non-uniform absorption.
4. > 10 GHz: energy absorption is primarily at the body surface. (In this
range, it is better to specify the incident power density and not the
SAR.)

10.4.3 Guidelines for Limiting EM Exposure


Consult the ICNIRP guidelines [1] (extract distributed as a class handout).
In it you will find the following tables listing reference levels for limiting
EM exposure:
1. Table 4: Basic restrictions up to 10 GHz
2. Table 5: Basic restrictions for power density between 10GHz and 300
GHz.
3. Table 6: Reference levels for occupational exposure (i.e. people in a
working environment).
4. Table 7: Reference levels for the general public.
Plots of the E field and B field levels are shown in Figure 2 for the cases
of general public exposure, peak general public, occupational exposure
and peak occupational exposure.
A rule of thumb for microwave engineers working above 10GHz, is to avoid
exposure to radiation levels above about 10 W/m2. The reference level for
the range 10 GHz to 300 GHz for occupational exposure is 50 W/m2 and
for general public exposure is 10 W/m2.

10-17 AJW, EEE3055F, UCT 2012


To gain some feeling for the numbers, we can calculate the power density
at a distance of one metre from a 50 Watt light bulb, assuming isotropic
radiation. The radiation power density at a distance of 1m would be
Pt G/(4r2) = 50 1/(412) = 3.9 W/m2. This of the same order as the 10
W/m2 specified for microwave radiation (although it is noted that most of
the power emitted by light bulbs is in the optical to infrared region .

References on EM radiation Safety:


1. Guidelines for limiting exposure to time-varying electric, magnetic, and
electromagnetic fields (up to 300 GHz), published by the International
Commission for Non-Ionizing Radiation Protection (ICNIRP), reprinted in
Health Physics, April 1998, vol. 74, no. 4, pp. 494-522.
Downloadable from http://www.icnirp.de
2. A Local Government Officials Guide to Transmitting Antenna RF
Emission Safety: Rules, Procedures, and Practical Guidance, Federal Com-
munications Commission (FCC), June 2, 2000.
3. Compilation of the dielectric properties of body tissues at RF and
microwave frequencies, C. Gabriel and S.Gabriel, Technical Report 1996,
http://www.brooks.af.mil/AFRL/HED/hedr/reports/dielectric/Report/Report.html
4. Cellular Phone Antennas (Mobile Phone Base Stations) and Human
Health, J. Moulder, Professor of Radiation Oncology,
http://www.mcw.edu/gcrc/cop/cell-phone-health-FAQ/
5. U.S. Environmental Protection Agency Website:
[http://www.epa.gov/radiation/understand/ionize nonionize.htm].

10-18 AJW, EEE3055F, UCT 2012


11 Reflection and Refraction at
Boundaries
We shall consider the following cases:
1. Reflection from a perfect conductor: wave normally incident.
2. Reflection and transmission at a dielectric interface: wave normally
incident.
3. Reflection from a perfect conductor: wave incident at an arbitrary
angle.
4. Reflection and transmission (refraction) at a dielectric interface: wave
incident at an arbitrary incident angle.

11.1 Reflection of Normally Incident Plane Waves from a


Perfect Conductor
Consider a plane wave normally incident on a flat perfectly conducting plate
( = ). The pointing vector tells us that the energy does not pass through
the conductor - why? P = E H and E = 0 so P = 0. All energy must
therefore be reflected.
A good question is: Why should an incoming wave bounce off a conduct-
ing surface? What is physically going on on the surface? My explanation
is that the incoming wave does not bounce at all - rather the electrons on
the surface react to form an induced electric field of the opposite polarity
to that of the incident plane wave. Thus the reacting current sheet on the
surface re-radiates the so-called reflected wave (but with an inversion in
polarity).
The next question that arises is: Why does the incident wave not pass

11-1
straight through the metal? In a strict sense, yes it does on propagating
forward, but it is also accompanied by a forward propagating wave arising
from the surface currents. These two forward propagating waves are in-
verted, with respect to one another, and sum to zero. Thus in speech we
say that the incident wave bounces off the surface - but more correctly,
the phenomenon of reflection is a re-radiation effect.
x
Incident wave

Reflected
Perfect
conductor

Figure 11.1: Reflection from a perfect conductor.

We now analyse the steady state plane wave solution to the wave equation
subject to the appropriate boundary condition: tangential components of
E are zero at the boundary, i.e. Ex (0) = and Ey (0) = 0.
The phasor forms of the incident and reflected waves are

Ex = E+ ejkz + Ee+jkz
 
H y = 1 E+ejkz E e+jkz

Applying the boundary condition

0 = E+ejk0 + Ee+jk0

implies
E = E+
Thus the total field in terms of E+ is
 
Ex = E+ejkz E+ e+jkz = E+ ejkz e+jkz = E+2j sin kz

11-2 AJW, EEE3055F, UCT 2012


The instantaneous form is thus
n o

Ex (z, t) = Re Ex ejt

= Re {E+2j sin kz [cos t + j sin t]}


= 2E+ sin kz sin t
This sum of the forward and reflected wave does not contain a z vt
term and hence does not appear to move left or right - it is what is known
as a standing wave. The peak amplitude |Ex (z, t)|peak = |2E+ sin kz| (or
envelope) varies with distance from the interface, as plotted below.
2E+
Envelope of
electric field
z
2 z=0

2E+
0
Envelope of
magnetic field
z
3
4
4 z = 0

Figure 11.2: Standing wave resulting from reflection of a sinusoidal wave by a perfect con-
ducting surface.

The expression for the magnetic field is similarly obtained by substituting


E = E+ in the expression for H y yielding
   
y = 1 E+ejkz Ee+jkz = 1 E+ ejkz + e+jkz = 1 2E+ cos kz
H

with instantaneous form
n o  
yejt 1
Hy (z, t) = Re H = Re 2E+ cos kz [cos t + j sin t]

1
= 2E+ cos kz cos t

Notice, that at the metal interface, the peak value Hy (z = 0, t) is in fact a
maximum, as is shown in the plot of the resulting standing wave.

11-3 AJW, EEE3055F, UCT 2012


11.2 Transmission Line Analogy
The equations governing reflection of plane waves are identical in form to
those derived for transmission lines. The following table compares equiva-
lent relationships.

Fields Transmission Lines


Ex = E+ejkz + Ee+jkz Vx = V+ ejz + V e+jz
   
y = 1 E+ ejkz Ee+jkz Iy = 1 I+ ejz I e+jz
H Z0

k = = qLC
p
= Z0 = CL

All transmission line concepts and formulas are applicable to the reflection
of plane waves, e.g. SWR, reflection coefficient, Smith chart calculations etc.
The units of E are V/m, and the units of H are A/m, analogous to voltage
and current in the transmission line.
The wave impedance at a position z is defined as the ratio of total electric
to magnetic field in the plane
Ex (z)
(z) =
Hy (z)
Directly from the transmission line analogy, the impedance at a distance l
from a load plane of impedance L is
L cos kl + j sin kl
(l) =
cos kl + j sin kl
and the reflection and transmission coefficients are
E L
= =
E+ L +
and
E2 2L
= =
E+ L +
where E2 is the amplitude of the wave in medium two.

11-4 AJW, EEE3055F, UCT 2012


Example: Application of Transmission Line Analogy to Incidence on a
Perfect Conducting Plane
At the interface with a perfect conductor, the electrons on the surface move
in such a way as to force the electric field to zero. The load impedance as
seen by the plane wave is
Ex (0) 0
L = = =0
Hy (0) Hy (0)
Substituting, we get
= 1
and
=0
and
(l) = j tan kl
which should be familiar results from transmission line theory.

11.3 Normally Incident Plane Wave on a Dielectric


Interface
q
1
Again we apply the results of transmission line theory with 1 = 1 and
q
2 = 22 to obtain
E 2 1
= =
E+ 2 + 1
and
E2 22
= =
E+ 2 + 1
|E |2
The fraction of power that is reflected is |E+ |
2 = ||2 and the fraction of
power that is transmitted into the medium is therefore 1 ||2 .

11-5 AJW, EEE3055F, UCT 2012


Example Calculation
Calculate the fraction of transmitted and reflected power for the case of a
1mW laser beam normally incident from air ( = 0 and = 0 ) onto glass
(r = 5 = 0 ).

Solution
The impedances in air is
r r
0 4 107
1 377 Ohms
0 8.85 1012
and in the glass is
r
4 107
2 = 169 Ohms
5 8.85 1012
The reflection coefficient is
2 1 169 377
= = = 0.380
2 + 1 169 + 377

and hence ||2 = 0.145. The reflected power is ||2 Pinc = 0.145 mW and
the transmitted power is the balance being 0.855 mW.

11.4 Reflection and Transmission (refraction) at a


dielectric interface - arbitrary incident angle
(ref: Griffiths 1981, and Ramo et al. 1994)

In this section, the relationships between incident, reflected and refracted


waves are studied for a wave incident on a dielectric interface.

Consider a wave incident on a dielectric boundary on the plane z = 0 as


depicted below.

11-6 AJW, EEE3055F, UCT 2012


x

Reflected
Transmitted
k
k2
2
z
+

k+

Incident wave
Medium 1 Medium 2

Figure 11.3: Propagation directions of incident, reflected and refracted waves at an interface
between two media.

The problem is analysed as follows:


1. Boundary conditions are applied to the case of arbitrary linear polar-
ization to establish
a) the fact that the incident, reflected and refracted waves lie in the
same plane of propagation known as the plane of incidence; i.e.
+, k
the vectors k and k
2 all lie in the same plane.
b) Snells laws relating the angles of reflection and refraction to the
incident angle.
2. The amplitudes of the reflected and refracted waves are analysed by
resolving the incident wave into two orthogonal polarizations:
a) Case where the electric field is perpendicular to the plane of inci-
dence, known as transverse electric polarization (or TE polariza-
tion).
b) Case where the magnetic field is perpendicular to the plane of
incidence - referred to as transverse magnetic polarization (or TM
polarization).

11-7 AJW, EEE3055F, UCT 2012


11.4.1 Derivation of Snells Laws
We consider first, a wave of arbitrary linear polarization incident on the
z = 0 plane at an angle . We model the incident, reflected and refracted
waves in phasor notation as

+ (r, t) = E
E + ejk+ r ejt +(r, t) = 1 k
H + E
+ (r, t)
1

(r, t) = E
E ejk r ejt H (r, t) = 1 k E (r, t)
1
2 (r, t) = E
E 2 ejk2 r ejt 2 (r, t) = 1 k
H 2 E
2 (r, t)
1
NOTE: The term k+ r = k+ k r is just k+ (the wave number) times the
component of r in the k direction. Imagine defining an axis pointing in the
direction, and let the distance along this axis be . The term ejk+ kr
k is
jk+
equal to e .
The total electric field in medium 1 is the sum of incident and reflected
fields
1 (r, t) = E
E + ejk+ r ejt + E
ejk r ejt
The angles of incidence and reflection are found by imposing the boundary
conditions previously derived in Section 4, i.e.

(a) Dn1 = Dn2 Normal component of electric flux density is continu-


ous

(b) Et1 = Et2 Tangential components of E are equal on either side of


interface.

(c) Bn1 = Bn2 Normal component of magnetic flux density is contin-


uous.

(d) Ht1 = Ht2 Tangential components of H are equal on either side


of interface.

11-8 AJW, EEE3055F, UCT 2012


It can be shown that imposition of all boundary conditions results in equa-
tions of the form

( )ejk+ r + ( )ejk r = ( )ejk2 r at z = 0

where the bracketed quantities ( ) are either parallel, perpendicular or nor-


mal components of the E or H fields.

Consider, for example, the electric field. At the interface plane, the bound-
ary condition Et1 = Et2 becomes:
+ )x ejk+ r ejt + (E
(E )x ejk r ejt = (E
2)x ejk2 r ejt at z = 0

and
+ )y ejk+ r ejt + (E
(E )y ejk r ejt = (E
2)y ejk2 r ejt at z = 0

where (E +)x is notation for the x component of E


+ etc. The other boundary
conditions result in similar equations.

For the constraint

( )ejk+ r + ( )ejk r = ( )ejk2 r at z = 0

to hold true for all x, y in the z = 0 plane requires1 that

k+ r = k r = k2 r at z = 0

or
(k+)x x + (k+)y y = (k)x x + (k)y y = (k2)x x + (k2)y y
This must hold along the line y = 0 in the xy plane, and so

(k+)x = (k)x = (k2)x

and similarly it must hold along the line x = 0 in the xy plane, and so

(k+)y = (k)y = (k2)y


1
The only way this can hold true for all positions at z = 0 if the phase factors are equal.

11-9 AJW, EEE3055F, UCT 2012


Thus we can conclude that k+ , k and k2 lie in the same plane (known as
the plane of incidence).

If we orientate the axes such that the plane of incidence is the xz plane
as shown in the figure, then the y components of the k vectors are zero, and
the remaining boundary condition

(k+)x = (k)x = (k2)x

implies
k1 sin + = k1 sin = k2 sin 2
from which we obtain Snells laws of reflection and refraction:

k1 sin + = k1 sin + =

and
sin 2 k1 /v1 v2
k1 sin + = k2 sin 2 = = =
sin + k2 /v2 v1
Since for most dielectrics, 1 2 0 , the ratio

sin 2 v2 c/n2 n1 1 2
= = =
sin + v1 c/n1 n2 2 1
where n1 = c/v1 and n2 = c/v2 are the refractive indexes of the dielectrics.
To analyse the magnitudes of reflected and transmitted waves, we need
examine the two cases of a transverse electric (TE) incident wave, and a
transverse magnetic (TM) incident wave. An arbitrary polarization can be
split into the sum of these two orthogonal components.

11.4.2 Case Transverse Magnetic (TM) Incident Wave


(also known as parallel polarization)
If the electric field is parallel to the incident plane (and the magnetic field
perpendicular), there exist only Ex , Ez and Hy components as shown in
the illustration below.

11-10 AJW, EEE3055F, UCT 2012


x

E
CASE E2
Transverse k
H
Magnetic (into) k2 H2
(TM) 2
z
+

E+ k+
H+
(out)
Medium 1 Medium 2

Figure 11.4: Definitions of vector quantities on either side of an interface for the case of a
TM incident wave.

E+ jk+ r + = ( + )E+ejk+ r

H+ = y e E yk
1
= E = ( )Eejk r
H y ejk r E yk
1
E2 jk2 r 2 = ( 2 )E2ejk2 r
H2 = y
e E yk
2

Take note of the directions of the vectors indicated: these represent the di-
rections in which E and H point at the interface (i.e. at x = 0 and z = 0) at
time t = 0, assuming the constants E+ , E , E2 were all positive. Also, it is
noted that H is defined as pointing into the page, such that when + = 0,
and E
E + will point in the same direction (for positive E+ , E , E
2), to
allow comparison with the derivation in Section 9.1. The relationships be-
tween E+ , E and E2 will now be established by applying the boundary
conditions.

The boundary conditions imposed on the x, y and z components of the


electric and magnetic fields are (see diagram) are

11-11 AJW, EEE3055F, UCT 2012


(a) Dn1 = Dn2 1 (E+ sin + ) + 1 (E sin ) = 2E2 sin 2

(b) Et1 = Et2 E+ cos + + E cos = E2+ cos 2

(c) Bn1 = Bn2 nothing useful as Hz = 0

+ H
(d) Ht1 = Ht2 H = H
2 1 1 1
1 E+ 1 E = 2 E2

Thus (d) implies (so does (a) after substituting Snells law and some ma-
nipulation)
1
E+ E = E2
2
and (b) implies
cos 2
E+ + E = E2
cos +
We solve these two equations to obtain the reflected and transmitted am-
plitudes in terms of the incident amplitude:
2 cos 2 1 cos +
E = E+
2 cos 2 + 1 cos +
and
22 cos +
E2 = E+
2 cos 2 + 1 cos +
The reflection and transmission coefficients are defined as
E 2 cos 2 1 cos +
T M = =
E+ 2 cos 2 + 1 cos +
and
E2 22 cos +
T M = =
E+ 2 cos 2 + 1 cos +
For the special case of the incident angle being zero, i.e. + = = 2 = 0,
we obtain, as in Section 11.3,
2 1
T M =
2 + 1
and
22
T M =
2 + 1

11-12 AJW, EEE3055F, UCT 2012


The Brewster Angle
It is interesting2 to plot and as a function of incidence angle varying
from 0 degrees (normal incidence) to 90 degrees. Is is noted that at a
certain angle known as the Brewster angle, T M = 0 hence the no signal is
reflected. This angle occurs when the numerator q(2 cos 2 1 cos +) = 0,
which, after substituting Snells law and i = ii and some manipulation
we obtain 2 1
1 1 2
sin2 + = 1 2
1 ( 2 )
For the case where 1 2 0 , the relationship simplifies to
r
2 n2
tan +
1 n1
It is also noted that the sign of the reflection coefficient changes from nega-
tive to positive, implying that the phase of the reflected wave jumps by 180
degrees as the incidence angle crosses the Brewster angle.

11.4.3 Case Transverse Electric (TE) Incident Wave


(also known as perpendicular polarization)
If the electric field is perpendicular to the incident plane (and the magnetic
field parallel), there exist only Ey and Hx and Hz components as shown in
the illustration below.

2
See example plot for Example: Plane wave air into glass

11-13 AJW, EEE3055F, UCT 2012


x

H
CASE
Transverse k E2
E
Electric k2+
2 H2
(TE)
z
+
E+
(out) k+

H+
Medium 1 Medium 2

Figure 11.5: Definitions of vector quantities on either side of an interface for the case of a
TE incident wave.

+ = y E+ jk+ r
E E+ejk+ r
H+ = (k+ y ) e
1

= y y E
E Eejk r = (k
H ) ejk r
1

2 = y 2 y E2
E E2ejk2 r 2 = (k
H ) ejk2 r
2

The boundary conditions imposed on the x, y and z components of the


electric and magnetic fields are (see diagram) are

(a) Dn1 = Dn2 nothing useful as Ez = 0

(b) Et1 = Et2 E+ + E = E2

+ sin + + 1 H
(c) Bn1 = Bn2 1 H sin = 2 H
2 sin 2

1 1 2
1 E+ sin + + 1 E sin = 2 E2 sin 2

11-14 AJW, EEE3055F, UCT 2012


+ cos + + H
(d) Ht1 = Ht2 H cos = H
2 cos 2

11 E+ cos + + 1
E cos
1
= 12 E2 cos 2

Thus we have from (b) and (and(c) after substituting Snells law and some
manipulation)
E+ + E = E2
and from (d)
1 cos 2
E+ E = E2
2 cos +
Solving for E and E2 we obtain
2 cos + 1 cos 2
E = E+
2 cos + + 1 cos 2
and
22 cos +
E2 = E+
2 cos + + 1 cos 2
The reflection and transmission coefficients are
E 2 cos + 1 cos 2
T E = =
E+ 2 cos + + 1 cos 2
and
E2 22 cos +
T E = =
E+ 2 cos + + 1 cos 2

For TE polarization, it can be shown3 that for dielectrics with equal perme-
abilities, there exists no angle for which T E = 0. The Brewster angle only
occurs for TM polarization. Thus the Brewster angle is sometimes called
the polarization angle, as light with an arbitrary polarization incident at
the Brewster angle is polarized on reflection. The TM components are not
reflected, but the TE components are4.

3
Do this as an exercise by analysing the numerator of .
4
See example plot for Example: Plane wave air into glass; compare TM and TE plots.

11-15 AJW, EEE3055F, UCT 2012


11.4.4 Angle of Total Reflection
Total reflection occurs when || = 1. Analysis shows that for both TE and
TM polarization this occurs only for the case of an EM wave incident from
a medium of slower velocity to that of a higher velocity, i.e. for v2 > v1 .
The critical incidence angle + = c occurs at
r
v1 2
sin c =
v2 1
At angles greater than this complete internal reflection occurs. The interface
behaves as a perfect mirror.
Examples of the application of total reflection include: (i) use of prisms as
reflectors in optics (ii) light propagates down optic fibres, which are made
of glass, as a consequence of total internal reflection.

11-16 AJW, EEE3055F, UCT 2012


Example 1: Plane Wave from air into glass
Air: r = 1.0, r = 1.0 v1 = c = 3.0 108 ms1 1 = 377 ohms
Glass: r = 5, r = 1.0 v2 = cr = 1.35 108 ms1 2 = 169 ohms

The transmitted wave is refracted towards the normal, as illustrated in


the diagram below. The wavelength will be shorter in the glass than in air
by the factor
1 1
= = 0.46
r 5

Reflected
Transmitted

2
z
+

Incident wave
Medium 1 Medium 2
AIR GLASS

Figure 11.6: An incident wave going from air into glass.

11-17 AJW, EEE3055F, UCT 2012


The Brewster angle for TM polarization occurs at B = 65.9 degrees.
NOTE: Compare the reflection coefficient || as the incident angle changes
from 0 to 90 degrees. For TE polarization, |T E | increases smoothly from
about |T E | = 0.4 to |T E | = 1. For TM polarization, as increases from 0,
|T M | decreases smoothly and reaches |T M | = 0 at = B = 65.9 degrees;
increasing the angle further increases |T M | until |T M | = 1 at = 90
degrees.
Magnitude of Reflection coefficient
1

|rho| TM
0.8 |rho| TE
|T E |
0.6

0.4 Brewster
angle
0.2
|T M |
0
0 10 20 30 40 50 60 70 80 90
Incident angle in degrees

Magnitude of Transmission coefficient


0.7
|tau| TM
0.6
|tau| TE
0.5

0.4
|T M |
|T E |
0.3

0.2

0.1

0
0 10 20 30 40 50 60 70 80 90
Incident angle in degrees

CASE: Parameters for plane wave from AIR into GLASS.

Figure 11.7:

11-18 AJW, EEE3055F, UCT 2012


The set of figures below show T M , T M , T E , T E again for the air into
glass case, but the signed values; it can be seen that the phase of T M flips
by 180 degrees at the Brewster angle.
Angle of Refraction vs Angle of Incidence
30

25
2 versus +
Refraction angle (deg)

20

15

10
From Air into Glass
5

0
0 10 20 30 40 50 60 70 80 90
Incident angle in degrees
Transverse Magnetic Incident Wave
1
rho
0.8 tau
T M
0.6

0.4
Brewster
0.2
angle
0

0.2 T M
0.4
0 10 20 30 40 50 60 70 80 90
Incident angle in degrees
Transverse Electric Incident Wave
1
rho
T E tau
0.5

0.5 T E

1
0 10 20 30 40 50 60 70 80 90
Incident angle in degrees

Plots of various parameters for the case of AIR into GLASS


as a function of the incident angle.

Figure 11.8:

11-19 AJW, EEE3055F, UCT 2012


Example 2: Plane Wave from glass into air
The transmitted wave is refracted away from the normal as it enters the
air, as illustrated in the diagram
below. The wavelength will increase as the
wave enters by the factor 5 = 2.2, in the air.
x

Reflected
Transmitted

2
z
+

Incident wave
Medium 1 Medium 2
GLASS AIR

Figure 11.9: An incident wave going from glass into air.

NOTE:
Brewster angle for TM polarization occurs at B = 24.1 degrees.
Angle of total internal reflection occurs at c = 26.6 degrees.
Reflection coefficient becomes complex for > c and an angle-dependent
phase shift occurs (not plotted here).
Below are plotted and for 0 < < c.

11-20 AJW, EEE3055F, UCT 2012


Angle of Refraction vs Angle of Incidence
100

80
Refraction angle (deg)
60 2 versus +

40
From Glass into Air
20

0
0 10 20 30 40 50 60 70 80 90
Incident angle in degrees Brewster
Transverse Magnetic Incident Wave
4
rho
tau
3

2 T M
Brewster
1
T M angle
0

1
0 10 20 30 40 50 60 70 80 90
Incident angle in degrees
Transverse Electric Incident Wave
2
rho
T E tau
1.5

0.5
T E

0
0 10 20 30 40 50 60 70 80 90
Incident angle in degrees

Plots of various parameters for the case of GLASS into AIR


as a function of the incident angle.

Figure 11.10:

11-21 AJW, EEE3055F, UCT 2012


11.4.5 Power Considerations
For a wave at an oblique angle, the power density (referred to as the intensity
I) crossing the dielectric interface is given by the component of the Poynting
vector normal to the interface, i.e.

Wm2
I =Pn

=
where P is the Poynting vector and n z is the surface normal.
x

P P2
2
z
+
n

P+

Medium 1 Medium 2

Figure 11.11: Poynting vectors on either side of an interface.

For the incident wave


1 E+2
P+ = k+
2 1
and thus
1 E+2
=
I + = P+ n cos +
2 1

For the reflected wave


1 E2
P = k
2 1
and thus
1 E2
=
I = P n cos
2 1

11-22 AJW, EEE3055F, UCT 2012


For the transmitted wave
1 E22
P2 = k2
2 2
and thus
1 E22
=
I 2 = P2 n cos 2
2 2


By substituting E = || E+ and E2 = | | E+ it can be seen that

|I+ | |I | = |I2 |

which is consistent with the fact that the net energy per second entering a
surface from one side must balance with that leaving from the other side.

Note: It should be understood that

|P+ | |P | 6= |P2|

unless the incident angle is zero.

11-23 AJW, EEE3055F, UCT 2012


12 Propagation in Conducting Media
and the Skin Depth
In this chapter the following topics are covered
1. Wave Equation in conducting media
2. Skin depth
3. Skin effect in conducting wires carrying AC currents
References:
Electromagnetics by IS Grant and WR Phillips.
Introduction to Electrodynamics by DJ Griffiths

12.1 Wave propagation in a conducting medium


Wave propagation in a conducting medium results in an exponential decay
of the wave amplitude as it propagates. The following illustration shows
an exponentially decaying sinusoidal EM wave. Notice that there exists a
phase shift between E and B fields in such a medium.
In the treatment that follows, the particular case of a propagation in a
highly conductive material (known as a good conductor) is considered.

12.1.1 Conducting media (particularly good conductors)


In a conducting medium, free electrons will move under the influence of
the electric field i.e. Jc = E. A good conductor, is one for which the
conduction current Dis significantly greater than the displacement current,
i.e. |Jc | |Jd | = t , which allows us to neglect the displacement current
term in Maxwells equations.

12-1
Figure 12.1: Exponentially decaying plane wave in a conducting (lossy) medium. Note E
and H are not in phase in a lossy medium.


If all field quantities are sinusoidal of form ejt , then |Jc | D
t


E
|E| = |jE|
t

Thus for a physical material to be regarded as a good conductor, the fre-
quency satisfies



This requirement is satisfied for all metals from DC to 100 GHz (and con-
siderably higher in frequency).

12.1.2 Derivation of the Differential Equation


Consider Maxwells equations in a conducting medium for which 0
inside the medium and J = E Dt . The equations become

12-2 AJW, EEE3055F, UCT 2012


D = 0
B = 0
B
E =
t
D
H = J+ J
t
Let us derive an equation containing only E. We start by taking the curl of
Faradays law

E = ( B
t ) = B
t
into which we can substitute the good conductor approximation
B = J = E
Additionally, we can use the identity and the approximation E = 0
E = 2E + E 2E
Making these substitutions, we obtain a form of the wave equation governing
propagation in good conductors:
E
2E =
t

12.1.3 Plane wave solution inside the conducting medium



Consider a plane wave propagating in z direction. For this case, x = 0 and

y
= 0, and the wave equation simplifies to

2E E
=
z 2 t
which has a solution of the form
E = E0 ej(tz)ez
modelling a wave propagating in the forward direction. The factor ez
represents a decaying amplitude in the z direction.

12-3 AJW, EEE3055F, UCT 2012


To obtain expressions for the constants and , substitute this solution
back into the wave equation to obtain
 
E h j(tz) z
i
LHS = = E0 e e
z z z z
h j(tz) z
i
= ( j) E0 e e
z

= [( j) E]
z
E
= ( j)
z
2
= ( j) E
and
E
RHS =
t
= jE
Equating LHS = RHS implies
( j)2 = j
2 2 + 2j = j
Equating real parts gives
2 2 = 0
which implies = (or = ). Equating imaginary parts gives
2 =
substituting = implies
22 =
or
r

= =
2

The physical sinusoidal signal is


n o
j(tz) z
E = Re E0 e e
= E0 cos(t z) ez

12-4 AJW, EEE3055F, UCT 2012


Figure 12.2: Exponentially decaying plane E(z, t = 0) wave in a good conductor medium.
The E field envelope is E0 ez/ .

A plot of one of the components of the electric field is shown in the following
illustration.
Analysis of the H field shows that H is is 45 degrees out of phase with
the E field - see tutorial example.

12.2 Skin Depth


The exponential decay ez , may be characterised by the decay constant
1/, known as the skin depth
r
1 2
= =

The skin depth is the distance at which the wave in the medium has decayed
to 1/e 0.707 of its original amplitude. Note 1 i.e skin depth reduces
rapidly as frequency increases.

12-5 AJW, EEE3055F, UCT 2012


Figure 12.3:

Example
Consider a sinusoidal 50 Hz plane wave propagating inside copper (which
is a very good conductor) = 5.96 107 r = 1.00 at 300 K. The skin
depth at 50 Hz is
r
2
= = 0.0092 m 9.2 mm
4107 5.9 107 250

At f = 50 MHz (radio frequency)

105 m 0.01 mm

At f = 10 Gz (radar & microwave communications)

= 6.5 107 m

At RF and Microwave frequencies, the penetration into metal is small (less


than a micrometre). Significant currents flow on the surface of good con-
ductors (i.e. concentrated mainly within skin depth) - we talk of the current
as being a surface current.
The table shows the skin depth for copper ( = 5.96 107) for a range
of frequencies.

12-6 AJW, EEE3055F, UCT 2012


Frequency for Copper
50 Hz 9.2 mm
10 kHz 0.65 mm
100 kHz 0.21 mm
1 MHz 0.065 mm
10 MHz 0.021 mm
10 GHz 0.00065 mm

12.3 Skin Effect in Conducting Wires


At high frequencies, the AC current flowing in a conducting wire is concen-
trated in thin outer layer or skin of the wire, as is illustrated in the cross
sections below.

Figure 12.4:

This concentration of current density in a thin outer layer causes the


resistance (per metre) to increase as frequency increases. A plot showing
the variation in current density J across the diameter of a 1mm conductor
is shown below [Ramo, Whinnery & van Duzer].

12-7 AJW, EEE3055F, UCT 2012


Figure 12.5:

DC Resistance
The DC resistance of a piece of wire of length l and thickness 2r is

l l
R0 = A = []
r2
where A = r2 is the cross sectional area.

Figure 12.6:

12-8 AJW, EEE3055F, UCT 2012


High Frequency AC Resistance
The high frequency resistance of a length of wire of l metres can be crudely
estimated by pretending that the current is uniform within an effective
conducting area Askin = 2r. With this giant leap of faith, we obtain a
formula for the high frequency resistance
r
l l 1 0
RHF = = []
Askin 2r 2r 2
which turns out to be not far out in practice for the case where r.
The high frequency AC resistance can also be express as
l r
RHF = R0 []
2r 2

Example
Consider a the difference in resistance per metre at DC and at 50 MHz of
a copper wire of diameter 5 mm. r = 2.5 103 m

At DC,
R0
0.0008 m1
l

At 50 MHz,
  3
RHF R0 r 4 2.5 10
= 8 10 5
= 0.1 m1
l l 2 2 10
which is 125 times greater, than at DC.

12-9 AJW, EEE3055F, UCT 2012


13 Radiation from Antennas

13.1 Hertzian Dipole

13.2 Dipole Antennas

13.3 Phase array pattern


field pattern from a linear array.
mention SAR [not covered 2009]
explain the antenna pattern in terms of constructive and destructive
interference. [see handwritten handout issued 2009]

13.4 Aperture Antennas [simplified treatment; lab


session]
horn antenna

13.5 Discussion on why circuits that are small compared


to the wavelength do not radiate

13-1
14 Thermal Radiation from Warm
Objects
why warm objects radiate
frequency spectrum (watts/Hz)
good and bad radiators; black body radiators versus metalic objects.
noise power received by an antenna aimed at a warm surface. (ekTB)
demonstration of UCT radiomter (35 GHz or 10 GHz)

14-1

You might also like