Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

2

Electromagnetic Fields

2.1 Maxwell’s Equations


This chapter provides a short introduction of the classical electromagnetic
theory. More detailed representations can be found in the literature, e.g. [6]
[28] [37] [49] [51].
Maxwell’s equations describe all macroscopic electromagnetic phenomena.
The time-dependent differential form is given by
∂D
∇×H = J+ (2.1)
∂t
∂B
∇×E = − (2.2)
∂t
∇·D = ρ (2.3)
∇·B = 0 (2.4)
with the following physical quantities:
H magnetic field strength
E electric field strength
D electric flux density
B magnetic flux density
J conduction current density
ρ volume charge density
Equation 2.1 is also known as Ampere’s law. Equation 2.2 is Faraday’s
law of induction and Equations 2.3 and 2.4 are Gauss’ laws for the electric
and magnetic field, respectively. Historically, the displacement current den-
sity ∂D/∂t was introduced by Maxwell and is essential for the description of
propagating electromagnetic waves. The term J +∂D/∂t is referred to as total
current density.
In Maxwell’s equations the current density and the charges are the sources
of the electric and magnetic fields. In source-free regions with non-conductive
material (e.g., free space) the equations become more simple:
6 2 Electromagnetic Fields

∂D
∇×H = (2.5)
∂t
∂B
∇×E = − (2.6)
∂t
∇·D = 0 (2.7)
∇·B = 0 (2.8)

Maxwell’s equations may also be written in their time-dependent integral


form:
   
∂D
H · ds = J+ · dA (2.9)
∂t
C(A) A
 
∂B
E · ds = − · dA (2.10)
∂t
C(A) A
 
 D · dA = ρ dv (2.11)
A(V ) V

 B · dA = 0 (2.12)
A(V )

where A and V are an arbitrary surface and volume, respectively. C(A) de-
notes the closed contour line of the surface A, and A(V ) denotes the closed
surface that encloses the volume V . The physical phenomena described by
Maxwell’s equations are more illustrative in their integral form. Figure 2.1a
shows a visualization of Ampere’s law (Equation 2.9) in integral form. Inte-
grating the magnetic field along the closed contour line C(A) yields the total
current through the area A. Figure 2.1b shows a visualization of Faraday’s
law (Equation 2.10) in integral form. Integrating the electric field along the
closed contour line C(A) yields the time derivative of the magnetic flux density
integrated over the area A.

2.2 Material Equations


The electric and magnetic field strengths are related to the current and flux
densities by the following material equations

B = µ0 µr H (2.13)
D = ε0 ε r E (2.14)
J = σE . (2.15)

where σ is the electrical conductivity, µ0 = 4π · 10−7 Vs/(Am) is the perme-


ability of free space and ε0 = 8.854 · 10−12 As/(Vm) is the permittivity of free
2.2 Material Equations 7

Fig. 2.1. Visualization of (a) Ampere’s law and (b) Faraday’s law in integral form

space. The relative permeability µr and the relative permittivity εr charac-


terize magnetic and dielectric properties of different materials. If the material
is
linear − material properties are independent of the field
isotropic − material properties are independent of direction
non-dispersive − material properties are independent of frequency
µr and εr become scalar values greater than or equal to one.
In technical disciplines signals with harmonic time-dependence are widely
used. More general signals with non-harmonic time-dependence can be de-
scribed as linear combinations of mono-frequent signals by using the Fourier-
Transform. If we assume a linear system, all physical quantities have a time
dependence cos(ωt + ϕ) with different phase angle ϕ but the same angular
frequency ω = 2πf . The advantage of this concept is that the time-dependent
physical quantities can be replaced by complex-valued phasors. The follow-
ing equation shows the relation between the physical real-valued electric field
strength E(r, t) and the complex-valued phasor E(r)
 
E(r, t) = Re E(r)ejωt , (2.16)

where j is the imaginary unit −1. Since the time dependency of the signals
is given by ejωt , the time-derivatives ∂/∂t can be replaced by a factor of jω.
Consequently, Maxwell’s equations in harmonic, differential form are given by

∇ × H = J + jωD (2.17)
∇ × E = −jωB (2.18)
∇·D = ρ (2.19)
∇·B = 0 (2.20)

where the E, H, D, B and J are complex-valued phasors. The real-valued phys-


ical quantities can be calculated analogous to Equation (2.16). The right-hand
side of Equation 2.17 is often rewritten in the following form:
8 2 Electromagnetic Fields

J + jωD = (σ + jωε0 εr ) E (2.21)


 
σ
= jωε0 εr − j E (2.22)
ωε0
= jωε0 (ε − jε ) E . (2.23)

where ε − jε presents a complex permittivity. The ratio between the imag-
inary part ε and the real part ε of the complex permittivity is called loss
tangent. It is a measure how lossy a material is:

ε σ
tan δ = 
= . (2.24)
ε ωε0 εr

2.3 Boundary Conditions

Fig. 2.2. Visualization of the source-free boundary conditions for the electric field
strength E and the electric flux density D. E1 and D1 indicate the field in medium 1
immediately before the interface and E2 and D2 indicate the field in medium 2
immediately behind the interface. The tangential component of the electric field and
the normal component of the electric flux density are constant at the interface

At the interface between two materials the electric and magnetic field com-
ponents have to fulfill special boundary conditions. In the absence of surface
charges ρs and surface current densities Js on the interface, the tangential
components of the electric and magnetic field strength are constant at the
interface. Furthermore, the normal components of the electric and magnetic
flux densities are constant. In the more general case including surface charges
and surface current densities we can write:
2.4 Waves in Free Space 9

E1t − E2t = 0 (2.25)



0 without surface charge density
D1n − D2n = (2.26)
ρs in case of surface charge density
B1n − B2n = 0 (2.27)

0 without surface current density
H1t − H2t = (2.28)
Js in the case of surface current density

where the letter t denotes the tangential component and the letter n denotes
the normal component of the fields. Figure 2.2 illustrates the behavior of the
electric field strength and the displacement current density at the boundary
between two materials with different permittivities. Equivalent figures can be
drawn for the magnetic field H and the magnetic induction B.

2.4 Waves in Free Space

In free space Maxwell’s equations can be transformed into the following wave
equations for the electric and magnetic field strength

∂2E
∆E − εµ =0 (2.29)
∂t2
∂2H
∆H − εµ 2 = 0 . (2.30)
∂t
where ∆ is the Laplacian operator. For harmonic fields the wave equations
assume the following form

∆E + k02 E = 0 (2.31)
∆H + k02 H = 0 . (2.32)

where E and H are phasors and k0 is the wave number of free space

k0 = ω ε0 µ0 . (2.33)

The simplest solution that fulfills the wave equations in free space is a
sinusoidally time-varying uniform plane wave. As an example Fig. 2.3 shows
a constant-time (t = T /4) plot of the magnetic and electric field strength
of a uniform plane wave travelling in positive z-direction. The electric and
magnetic field vectors are given by

E(z, t) = E0 cos(ωt − k0 z)ex and (2.34)


H(z, t) = H0 cos(ωt − k0 z)ey . (2.35)
10 2 Electromagnetic Fields

Surfaces of constant phase (ωt − k0 z = const.) are planes with surface


normal vector in the direction of propagation (z-direction). The spatial period
of the plane wave is the wavelength λ. The relation between the wavenumber
k0 , the free space wavelength λ0 and the phase velocity in free space c0 is
given by
c0 2π
λ0 = = . (2.36)
f k0
In free space the phase velocity is
1 m
c0 = √ ≈ 2.9979 · 108 . (2.37)
ε0 µ0 s

Fig. 2.3. Uniform plane wave propagating in positive z-direction

The electric field vector, the magnetic field vector, and the direction of
propagation are perpendicular to each other. The ratio between the ampli-
tude E0 of the electric field strength and amplitude H0 of the magnetic field
strength is given by the characteristic impedance of free space ZF0 .
2.5 Polarization 11

µ0 E0
ZF0 = ≈ 377 Ω = (2.38)
ε0 H0

An energy flow is associated with the propagation of the plane wave. The
Poynting-vector S describes the rate of energy transport per unit area.

E02
S=E×H= · ez = ZF0 · H02 · ez (2.39)
ZF0

2.5 Polarization

By convention, the direction of the electric field vector determines the po-
larization of the plane wave. In Fig. 2.3 the plane wave is linearly polarized
along the x-direction. If we superimpose two uniform plane waves with the
same angular frequency ω, but different amplitudes and phase angles ϕ0 , we
get elliptical or circular polarization.
In order to understand technically relevant polarizations, the superposition
of the following two uniform plane waves in free space is investigated. The
vector of the electric field strength E1 of the first plane wave is oriented
in x-direction and the vector of the electric field strength E2 of the second
plane wave is oriented in y-direction. Both waves are propagating in positive
z-direction.

E1 (z, t) = E1,0 cos(ωt − k0 z + ϕ1,0 )ex and (2.40)


E2 (z, t) = E2,0 cos(ωt − k0 z + ϕ2,0 )ey . (2.41)

If we look at the polarization of the resulting wave E = E1 + E2 we can


define the following different polarizations:
Linear polarization: If the phase angles are equal (ϕ1,0 = ϕ2,0 ) the direc-
tion of the vector of the resulting electric field strength E is independent
of time (Fig. 2.4a).
Circular polarization: If the amplitudes are equal (E1,0 = E2,0 ) and the
phase angles differ by π/2 (ϕ1,0 = ϕ2,0 ± π/2) the resulting electric
field vector rotates on a circular curve in the xy−plane. The angular
frequency of the rotation is ω. Depending on the sign of the phase dif-
ference the vector rotates clockwise or counterclockwise. The wave is said
to be right-handed circularly polarized (RHCP) or left-handed circularly
polarized (LHCP) (Fig. 2.4b).
Elliptical polarization: If the amplitudes of the waves are not equal (E1,0 =
E2,0 ) and the phase difference is still π/2 (ϕ1,0 = ϕ2,0 ± π/2) the resulting
electric field vector rotates again with angular frequency ω. In the xy-plane
the tip of the vector follows an ellipse with the main axes directed into
the x- and y-direction. If the amplitudes of the waves differ (E1,0 = E2,0 )
and the phase angle difference is neither 0 nor π/2 (ϕ1,0 = ϕ2,0 + ϕ with
12 2 Electromagnetic Fields

(a) (b)

(c) (d)

Fig. 2.4. Polarizations of a wave propagating in positive z-direction: (a) linear


polarization, (b) right-handed circular polarization (RHCP), (c)(d) right-handed
elliptical polarization (RHEP)

arbitrary ϕ) the resulting electric field vector follows an elliptical curve in


the xy−plane but now the main axes of the ellipse are rotated with respect
to the x- and y-direction. Depending on the sign of the phase difference the
vector rotates clockwise or counterclockwise. The wave is said to be right-
handed elliptically polarized (RHEP) or left-handed elliptically polarized
(LHEP) (Fig. 2.4c-d).

2.6 Waves in Lossy Media

In lossy media the wave equations take a more complex form than in the case
of free space. Due to the conductivity an additional term is introduced into
the wave equations
2.6 Waves in Lossy Media 13

∂E ∂2E
∆E − µσ − εµ 2 = 0 (2.42)
∂t ∂t
∂H ∂2H
∆H − µσ − εµ 2 = 0 . (2.43)
∂t ∂t
For harmonic fields the wave equations assume the following form

∆E − jωµσE + k 2 E = 0 (2.44)
∆H − jωµσH + k 2 H = 0 . (2.45)

These equations can be written as

∆E − γ 2 E = 0 (2.46)
∆H − γ 2 H = 0 (2.47)

by defining a propagation constant γ



γ = jωµσ − ω 2 εµ (2.48)

= jωµ(σ + jωε) . (2.49)

The propagation constant is a complex number. The real part α is the


attenuation constant and the imaginary part β is the phase constant.

γ = α + jβ (2.50)

The attenuation constant and the phase constant are given by



 
εµ σ
α=ω 1+ −1 (2.51)
2 ωε

 
εµ σ
β=ω 1+ +1 . (2.52)
2 ωε

Furthermore, the phase velocity c, the wavelength λ, and the intrinsic


impedance ZF become
ω
c= and (2.53)
β

λ= (2.54)
β
jωµ
ZF = . (2.55)
γ
For good conductors (σ >> ωε) the attenuation coefficient is approxi-
mately the reciprocal value of the skin depth δ

1 2
≈δ= . (2.56)
α ωµσ
14 2 Electromagnetic Fields

The simplest solution that fulfills the wave equations in lossy media is again
a sinusoidally time-varying uniform plane wave. To illustrate the change in
the solution we return to the example of the previous section. In the equation
of the uniform plane wave travelling in positive z-direction an additional term
e−αz is introduced. Furthermore, the wave number k0 is replaced by the phase
constant β. Consequently, the amplitude of the plane wave decreases as the
wave propagates in positive z-direction (see Fig. 2.5):

E(z, t) = E0 e−αz cos(ωt − βz)ex and (2.57)


−αz
H(z, t) = H0 e cos(ωt − βz)ey . (2.58)

Fig. 2.5. Attenuation of a uniform plane wave in a lossy material

2.7 Energy Conservation

In electric and magnetic fields energy is stored. Electric energy may be trans-
formed into magnetic energy and vice versa. The energy density we in the
electric field and the energy density wm in the magnetic field are given by
1 1 2
we = D·E= ε |E| (2.59)
2 2
1 1 2
wm = B·H= µ |H| (2.60)
2 2
In the presence of lossy materials energy may be transformed into heat.
The power dissipated into heat per unit volume is
2.8 Electromagnetic Potentials 15

dP 2
= J · E = σ |E| (2.61)
dV
Considering harmonic time-dependence the time-averaged energy density
and power loss per unit volume can be expressed using the phasors of the
electric and magnetic field strength by
 
1 1 ∗
we = Re εE · E (2.62)
2 2
 
1 1
wm = Re µH · H∗ (2.63)
2 2
dP 1
= Re {J · E∗ } , (2.64)
dV 2
where E ∗ and H ∗ denote the complex conjugate electric and magnetic field
phasor, respectively.
With the above relations we can write the law of electromagnetic energy
conservation in the following form.
  
d
(we + wm ) dv = −  S · dA − JE dv (2.65)
dt V V
A(V )

where S is the Poynting-vector defined in Equation 2.39. We can interpret


this important equation of energy conservation in the electromagnetic field in
the following form: A decrease in energy inside the volume V is a result of (a)
power leaving the volume or (b) the transformation of energy into heat.

2.8 Electromagnetic Potentials


In the previous sections we have described electromagnetic phenomena by us-
ing electric and magnetic field functions E and H. There exists an alternative
formulation in terms of potential functions. These potential functions are the
scalar electric potential φ and the magnetic vector potential A. The advan-
tage with this concept of potentials is that they can be determined quite easily
from the sources of the electromagnetic fields, i.e., the current density J and
the charge density ρ.
The magnetic induction B can be given in terms of the magnetic vector
potential A. The magnetic vector potential A is defined by convention as

B = ∇×A . (2.66)

Furthermore, the electric field E can be given in terms of the scalar electric
potential φ and the derivative of the magnetic vector potential.
16 2 Electromagnetic Fields

∂A
E = −∇φ − (2.67)
∂t
In order to derive an explicit solution of the potentials from the differential
Eqs. 2.66 and 2.67 the Lorentz condition in Eq. 2.68 is normally used:
∂φ
∇ · A = −εµ . (2.68)
∂t

Fig. 2.6. Retarded potentials A and φ at observation point r originating from the
conduction current density J and charge density ρ at source point r . Time delay
t0 = R/c due to finite velocity of electromagnetic wave propagation

Using the definitions given in Equations 2.66 and 2.67 and the Lorentz
condition the following wave equations for the electromagnetic potentials can
be derived from Maxwell’s equations
∂2A
∆A − εµ = −µJ (2.69)
∂t2
∂2φ ρ
∆φ − εµ 2 = − . (2.70)
∂t ε
Solutions of these time-dependent wave equations are the so-called retarded
potentials
  
ρ r , t − Rc
φ(r, t) = dv  (2.71)
4πεR

V
 
µJ r , t − Rc
A(r, t) = dv  , (2.72)
V 4πR
2.9 Green’s Function 17

where the term

R = |r − r | (2.73)

denotes the distance between the source point r and the observation point
r. The term t = t − R/c takes into account the delayed response of the
potential at the observation point r due to the finite propagation velocity c
of an electromagnetic wave: a change of charge density ρ or current density
J at the source point r does not alter the potential in the spatial domain
instantaneously. It takes the wave the time t0 = R/c to travel from the source
point to the observation point.

2.9 Green’s Function


If we consider time harmonic signals, the wave equations of the potential
functions (Equations 2.69 and 2.70) are called Helmholtz equations for the
electromagnetic potentials

∆A + k 2 A = −µJ (2.74)
ρ
∆φ + k 2 φ = − , (2.75)
ε
where k = ω/c is the wave number. The solutions of these equations are given
by

1
φ(r) = ρ(r )G(r, r )dv  (2.76)
ε
 V

A(r) = µ J(r )G(r, r )dv  , (2.77)


V

where G is Green’s function of free space and the volume integral covers all
sources in the entire volume. Green’s function of free space is determined by
solving the scalar Helmholtz equation with a Dirac source function

∆G(r, r ) + k 2 G(r, r ) = −δ(r − r ) . (2.78)

The solution is given by

1 e−jk|r−r |

e−jkR
G(r, r ) = 
= with R = |r − r | . (2.79)
4π |r − r | 4πR

Green’s function describes the potential at the observation point r that is


generated by a point source at r . In free space Green’s function describes a
spherical wave originating at the source point r .
Now we insert Equation 2.79 into Equations 2.76 and 2.77 and obtain the
harmonic solutions of the potentials for free space as
18 2 Electromagnetic Fields

ρ (r ) ejkR 
φ(r) = dv (2.80)
V 4πεR

µJ (r ) ejkR 
A(r) = dv . (2.81)
V 4πR

From the potentials A and φ the electric and magnetic fields can be cal-
culated using Equations 2.66 and 2.67. With the use of the Lorentz condition
in Equation 2.68 the fields can be expressed in terms of the vector potential
only.

B = ∇×A (2.82)
1
E= ∇ (∇ · A) − jωA . (2.83)
jωµε

2.10 Waves on Transmission Lines

In Sect. 2.4 we discussed the propagation of uniform plane waves in free space.
These waves fill the entire space uniformly and propagate in one direction
without the support of any material. Waves can be guided by so-called trans-
mission lines that transports energy from a source to a load. There is a vari-
ety of structures that can guide waves. We will restrict our discussion here to
transmission lines made of two conductors for this chapter.

Fig. 2.7. Examples of two conductor transmission lines: (a) a pair of parallel wires
(b) a coaxial cable

Figure 2.7 shows two common types of such transmission lines: a pair of
parallel wires and a coaxial cable. The latter consists of an inner and an outer
conductor. These lines are uniform, i.e., the cross section of the line remains
constant over the length of the structure. Furthermore, let us assume that the
waves travelling on this lines propagate in an homogeneous lossless medium
and that the conductors are perfect (σ → ∞). In the following section we will
consider waves travelling in vacuum. Under these conditions a fundamental
mode wave called transversal electric magnetic (TEM) wave can propagate
along the line. In a TEM wave the vectors of the electric and magnetic field
strengths are perpendicular to each other and perpendicular to the direction
2.10 Waves on Transmission Lines 19

of propagation. There is no field component in the direction of propagation.


Figure 2.8 shows the electric and magnetic field orientations in a cross section
for a two-wire line and a coaxial transmission line. The electric field is directed
from one conductor to the other and the magnetic field is circulating around
the conductors.

Fig. 2.8. Electric and magnetic field distribution for a TEM wave on (a) a pair
of parallel wires and (b) a coaxial transmission line. Definition of integration paths
CE and CH for the calculation of voltage U and current I on (c) a pair of parallel
wires and (d) a coaxial transmission line

Within such a structure it is possible to define a voltage U (z, t) between


the conductors and a current I(z, t) flowing in the conductors evaluating the
following line integrals:

U (z, t) = − E ds (2.84)
CE

I(z, t) = H ds . (2.85)
CH

In Equation 2.85 the current I is the current in one of the conductors. The
other conductor carries the same current but with opposite sign. In Fig. 2.8
this is indicated by the symbol I + for the current in one conductor and I −
for the current in the other conductor.
20 2 Electromagnetic Fields

In terms of an equivalent circuit model the conductor with a current flow-


ing in it and a magnetic field around it represents an inductance. The pair of
conductors with the electric field in between represents a capacitance. There-
fore the line can be represented by an equivalent circuit as shown in Fig. 2.9,
where L is the inductance per unit length and C  is the capacitance per unit
length.

Fig. 2.9. Equivalent circuit (per unit length) of a lossless transmission line

From this equivalent circuit the transmission line equations can be derived
∂U ∂I
= −L (2.86)
∂z ∂t
∂I ∂U
= −C  . (2.87)
∂z ∂t
These equations can be transformed into the following equations for the volt-
age U and the current I:
∂2U  ∂ U
2
− L C =0 (2.88)
∂z 2 ∂t2
∂2I ∂2I
2
− L C  2 = 0 (2.89)
∂z ∂t
Equations 2.88 and 2.89 are wave equations of the same form as given in
Equations 2.29 and 2.30 for the electric and magnetic field strength. Therefore
we can use the solution from the plane wave as obtain in Sect. 2.4. This leads
to the propagation of voltage and current waves along the transmission line.
We can give the phasors of voltage and current waves travelling in positive
z-direction in the following form:

U (z) = U0 e−jkz (2.90)


−jkz
I(z) = I0 e (2.91)

where k is the wave number. When evaluating the wave number k it turns out
that k is the wave number of free space
2.10 Waves on Transmission Lines 21
√ √ ω
k = ω L C  = ω µ0 ε0 = . (2.92)
c0
The ratio of the forward-moving voltage wave U (z) and the forward-
moving current wave I(z) is independent of the longitudinal coordinate z
and called line impedance Z0 or characteristic impedance of the line:

U (z) U0 L
Z0 = = = . (2.93)
I(z) I0 C

The line impedance of technically important two-wire transmission lines


is commonly Z0 = 50 Ω.

You might also like