You are on page 1of 74

1 Introduction to Electromagnetics

Electromagnetic fields are caused by electric charges at rest and in


motion. Positive and negative electric charges are sources of the
electric fields and moving electric charges yielding a current is the
source of magnetic fields. Time-varying electric and magnetic fields are
coupled in an electromagnetic field radiating from the source.

Figure 1 Positive and negative electric charges are sources of an electric field

Electromagnetic fields are divided into four different quantities:

the magnetic flux density B with the unit T (Tesla or volt-second


per square meter)
the magnetic field intensity H with the unit A/m (Ampere per
meter)
the electric field intensity E with the unit V/m (Volt per meter)
the electric flux density D with the unit C/m2 (Coulomb per square
meter)

A time-varying E and D will give rise to B and H, and vice versa where
the relation depends on the properties of the medium. Far enough from
the source the magnetic field, H, will be perpendicular to the electric
field, E, and both are normal to the direction of propagation, as shown
in the following figure:

Figure 2 A time-varying electric field, E, will give rise to a perpendicular magnetic field, H, and vice
versa. Far enough from the source it will become a uniform plane wave and the ratio between E and
H will be the intrinsic impedance of the medium.

Far enough from the source, the wave-front, which will become almost
spherical, can be seen as an almost plane front if the sphere is large
enough. Then we have a uniform plane wave where the ratio between
the electric field and the magnetic field, called the wave impedance:

Equation 1

is a constant named the intrinsic impedance of the medium, . The


electromagnetic theory is also based on three universal constants:

the velocity of an electromagnetic wave in free space c (the


speed of light 3.108 m/s)
the permittivity of free space 0 ( 8.854.10-12 F/m)
the permeability of free space 0 (= 4 .10-7 H/m)

These constants are related by [8]:

Equation 2

The permittivity is a proportionality constant between the electric flux


density D and the electric field intensity E, in free space as:

Equation 3

and the permeability is the proportionality constant between the


magnetic flux density B and the magnetic field intensity H, in free
space as:

Equation 4

From these constants, the intrinsic impedance of free space can be


calculated as [8]:

Equation 5

[ ]

2 Maxwells Equations
James Clerk Maxwell (1837-1879) gathered all
prior knowledge in electromagnetics and
summoned the whole theory of electromagnetics
in four equations, called the Maxwells equations.
To evolve the Maxwells equations we start with
the fundamental postulates of electrostatics and
magnetostatics. These fundamental relations are
considered laws of nature from which we can
build the whole electromagnetic theory.
According to Helmholtzs theorem, a vector field
is determined to within an additive constant if both its divergence and
its curl are specified everywhere [8]. From this an electrostatic model
and a magnetostatic model are derived only by defining two
fundamental vectors, the electric field intensity E and the magnetic flux
density B, and then specifying their divergence and their curls as
postulates. Written in their differential form we have for the
electrostatic model the following two relations' [8]:

Equation 6

Equation 7

where is the volume charge density:

[C/m3]

Equation 8

These are based on the electric field intensity vector, E, as the only
fundamental field quantity in free space. Then to account for the effect
of polarisation in a medium the electric flux density, D, is defined by
the constitutive relation:

Equation 9

where the permittivity is a scalar (if the medium is linear and


isotropic). Similarly for the magnetostatic model we have the following
two relations, based on the magnetic flux density vector, B, as the
fundamental field quantity:

Equation 10
Equation 11

where J is the current density. To account for the material here as well,
we define another fundamental field quantity, the magnetic field
intensity, H, and we get the following constitutive relation:

Equation 12

where is the permeability of the medium. Using the constitutive


relations we can rewrite the postulates and the relations derived is
gathered in the following table:

Table 1 Fundamental Relations for Electrostatic and Magnetostatic Models (The Governing
Equations)

The Governing Equations


Electrostatic Model
Equation 13

Equation 14

Magnetostatic Model
Equation 15

Equation 16

These equations must, however, be revised for calculation of time


varying fields. The electrostatic model must be modified due to the
observed fact that a time varying magnetic field gives rise to an
electric field and vice versa and the magnetostatic model must be
modified in order to be consistent with the equation of continuity.
The complete model for electromagnetic fields (Maxwells equations) is
gathered in the following table (Table 2), where the integral forms of
the equations are added [8]:
Table 2 Maxwell's Equations, both in differential and integral form

Maxwells Equations
Faradays law
Equation 17

Equation 18

Ampres circuital law


Equation 19

Equation 20

Gausss law
Equation 21

Equation 22

No isolated magnetic charge


Equation 23

Equation 24

We can see in Equation 17 that the electric field intensity vector


(Equation 13) is replaced with
according to
Faradays law of electromagnetic induction [8]. In Equation 19, the
term
is called displacement current density and its introduction was
one of the major contributions by Maxwell. The displacement current
density is necessary in order to make the equations consistent with the
principle of conservation of charge in the time varying case.
There are many ways of solving and using these equations. One
technique to make the solution of Maxwells equations easier, which
we will use later, is to use potential functions. It is known that if a
vector field is divergenceless, then it can be expressed as the curl of
another vector field. For instance since the divergence of B is zero,
, then B can be expressed as the curl of the vector field A:

Equation 25

where A is called the vector magnetic potential and it can be


determined from the current distribution J:

Equation 26

where:

is the wave number


r is the distance between the radiating source and the point of
observation

Using the constitutive relations, Equation 9 and Equation 12, the


magnetic field intensity H can then be calculated as:

Equation 27

The electric field intensity E can also be calculated using potential


functions by introduction of the scalar electric potential V:

Equation 28

When both the vector magnetic potential A and the scalar electric
potential V are known, the electric field intensity E is derived by:

Equation 29

It is however not necessary to calculate both the magnetic field


intensity H and the electric field intensity E since they are related by
the equation:

Equation 30

3 Near-Field and Far-Field


In this chapter we will solve the Maxwells equations for a radiating
wire and by analysing the solution we will define the near-field and the
far-field. The electromagnetic field radiating from a wire can be
calculated by solving Maxwells equations for a short current element
and then placing the current elements end-to-end.

Figure 3 A small current element or an electric dipole.

To be able to derive the field of a conducting element acting as an


electrical dipole, the current in the element is defined to flow along an
axis z. This gives that the magnetic vector potential can be expressed
as (see Equation 26):

Equation 31

where:

Equation 32

It is wise to use spherical coordinates when deriving the field of a small


current element that can be approximated with a point source

Equation 33

If A is then expressed in the form

Equation 34

it can apparently be divided into the spherical components:

Equation 35

Equation 36

Equation 37

The intensity ofq he magnetic field is then (using Equation 27):

Equation 38

According to Equation 30 above E can be determined when H is known:

Equation 39

Which results in:

Equation 40

Equation 41

Equation 42

By rewriting those equations we can identify the wavelength and get


the following equations in Table 3. A summation of the theory above
states that:
If the length of one element is much less than a wavelength (dl << )
and the element is considered as an oscillating dipole the electric and
magnetic fields for one element can, with the use of some algebra, be
expressed as:
Table 3 The electric and magnetic field equations for a small oscillating electric dipole

Electric and Magnetic Field Equations for an Electric Dipole


Equation 43

Equation 44

Equation 45

where:

dl is the length of the current element


is a short for (2 f/ )- t
is the radian frequency of the signal
t is the time (=1/f)
c is the speed of light (
m/s)
Z0 is the free space impedance ( 377 )
I is the current in the element
is the zenith angle to radial distance r
is the wavelength of the signal
r is the distance from the element to the point of observation

Analysing these equations (Equation 43 to Equation 45) we can divide


the terms into three different basic terms:

one inverse distance term proportional to r-1 which is called the


radiation term. This term represents the flow of energy away
from the wire.
one inverse square distance term proportional to r-2 which is
called the induction term and represents energy stored in the
field during one quarter of a cycle and then returned to the
antenna in the next.
and finally one inverse cube distance term proportional to r-3 that
is called the quasi stationary term, or the electrostatic field term,
and results from the accumulation of charge at the ends of the
element.

Figure 4 The significance of the different terms for the electric field strength

Notice that all these terms with the coefficients

and

will all be equal (=1) at the distance


. This distance is in
this report said to be the boundary between near-field and far-field and
the contributions from the radiation, induction and the electrostatic
term are all of the same magnitude.

When r <<
only the first term in each equation is significant and will
in this case mean that the wave impedance will be:

Equation 46

that is much greater than the free space impedance Z0 i.e. we will have
a high E-field and a low H-field. If the current element had been a
current loop with a low circuit impedance instead of the high circuitimpedance of the current element, the first term, or the electrostatic
term, in the first two equations (Equation 43 and Equation 44) would
disappear and a similar equation would appear in Equation 45. In this
case the wave impedance would be:

Equation 47

representing a high H-field and a low E-field in the near-field.

When r >>
the last term proportional to r-1 in Equation 43 and
Equation 45 will dominate and the wave impedance will approach the
free space impedance Z0 = 377 . This is called the far-field or
radiation field. The E and the H fields will then be in phase and
orthogonal to each other producing plane waves. This is illustrated in
Figure 5 below:

Figure 5 Wave impedance at different distances from either an electric source or a magnetic source

When such current elements are placed end-to-end to produce a model


of a radiating wire, the charge at the ends of the elements will cancel
and the term due to the electrostatic field (the one proportional to r-3)
will disappear. This is however only true with a constant current
distribution on the line. With a varying current distribution the
electrostatic fields will not cancel entirely. However, if the wire is
divided into a sufficient number of segments per wavelength then this
error will be small.
The length of telecommunication lines and the size of
telecommunication systems are often greater than one wavelength, D
. If the dimension of the field source D is greater than a wavelength
the near-field/far-field boundary is said to be at the distance [5]:

Equation 48

At a shorter distance maxima and minima would appear due to


interference caused by different distances to different parts of the
source.

4 Transmission Line
4.1 Lossy Transmission Line Model
A transmission line can be approximated by a distributed-parameter
network with the circuit parameters distributed throughout the line.
One line segment with the length z can be approximated with an
electric circuit as in the following figure:

Figure 6 Equivalent circuit of an element of a transmission line with a length of z

Note that this model is actually for an unbalanced line but the model
for a balanced line is the same except for R and L which is divided
symmetrically on both conductors with R/2 and L/2 both on the upper
conductor and the lower conductor in Figure 6. This lossy transmission
line model is described by four lumped parameters, which will be
derived later in this section:

R is the resistance in both conductors per unit length in /m


L is the inductance in both conductors per unit length in H/m
G is the conductance of the dielectric media per unit length in
S/m
C is the capacitance between the conductors per unit length in
F/m

where R and G is zero under lossless conditions. We will also derive the
propagation constant and the characteristic impedance in this section.

To derive the parameters we begin by using Kirchhoffs voltage law on


the circuit in Figure 6:

Equation 49

which can be written as:

Equation 50

and then letting z 0 we get:

Equation 51

Then we have one equation containing R and L. To get another


equation relating G and C we apply Kirchhoffs current law on the
circuit and get:

Equation 52

and letting z 0 in this equation also we get:

Equation 53

The first order partial differential equations, Equation 51 and Equation


53, are called the general transmission line equations. These equations
can be simplified if the voltage v(z,t) and the current i(z,t) are timeharmonic cosine functions:

Equation 54
Equation 55

Using Equation 54 and Equation 55, the general transmission line


equations in Equation 51 and Equation 53 can be written as:

Equation 56

Equation 57

These equations are called the time-harmonic transmission line


equations. These equations (Equation 56 and Equation 57) can be used
to derive the propagation constant and the characteristic impedance of
the line. By combining Equation 56 and Equation 57 we get:

Equation 58

Equation 59

where is the propagation constant:

Equation 60

The real part, , of the propagation constant is called the attenuation


constant in Np/m and the imaginary part, , is called the phase
constant of the line in rad/m. In a transmission line the series
resistance of the conductors are so low that it can be neglected in the
computation of the propagation constant. Equation 60 can then be
simplified to:

Equation 61

The solution of Equation 58 and Equation 59 is:

Equation 62
Equation 63

where the two terms in each equation denotes travelling waves in


positive and negative direction along z respectively. The characteristic
impedance of a transmission line is the ratio between the voltage and
the current for an infinitely long line. For an infinitely long line with the
source at the left end there are no reflecting waves so only the waves
travelling in the positive z direction exist (only V+ and I+ in Equation 62
and Equation 63). Using Equation 56 and Equation 57 the
characteristic impedance can be written as:

Equation 64

[ ]
Note that it is independent of z.
If the distance D between the two wires is much greater than the
radius, a, of a conductor, the capacitance per unit length, C, can be
written as:

Equation 65

[F/m]
And then by comparing Equation 61 with the propagation constant for
a transverse electromagnetic wave in a medium with the constitutive
parameters , and written as:

Equation 66
-1

[m ]
and assuming that the following relationship is known:

Equation 67

we get:

Equation 68

Using Equation 67 and Equation 68 we can get L and G for the


transmission line from Equation 65:

Equation 69

[H/m]
and

Equation 70

[S/m]
To derive the series resistance R we look at the ohmic power dissipated
per unit length of both conductors. Assuming the surface current Js to
flow in a very thin surface layer and to be uniform over the
circumference of both conductors the current in each conductor is
then:

Equation 71

Since copper, and other materials used in transmission lines, are good
conductors (i.e. c >> c) and the radius a is much larger than the
skin depth the above assumptions will hold for transmission lines and
then the surface impedance can then be written as the intrinsic
impedance, c, of a good conductor:

Equation 72

[ ]
The dissipated power per unit length, p , assuming the presence of a
non-vanishing axial electric field at the surface, will then be:

[W/m2]

Equation 73

Using Equation 71 the total ohmic power dissipated in a unit length of


the line (in both conductors) will then be:

[W/m]

Equation 74

The series resistance per unit length is then:

Equation 75

[ /m]
where the parameters c and c is that of the material in the
conductors. The real part of the surface impedance, Rs, is calculated
as:

Equation 76

[ ]
Having the distance, D, between the conductors and the radius, a, of
the conductors, the characteristic impedance from Equation 64 can be
simplified using Equation 65, Equation 69, Equation 70 and Equation
75 into:

Equation 77

[ ]

It can also be shown that for a balanced pair of wires near ground the
characteristic impedance will also depend on the height over ground, h,
as:

Equation 78

[ ]
The derived equations is collected in the following table:
Table 4 Basic transmission line equations

Description

Paramet
er

Equation

Unit

Resistance
R

/m

H/m

S/m

F/m

Z0

Z0

Z0

m-1

Inductance

Conductance

Capacitance

Characteristic
Impedance

Characteristic
Impedance
(simplified)
Characteristic
Impedance
(at height h above
ground)
Propagation
Constant

4.2 Termination and Reflections


In an infinitely long line there are only forward travelling waves and no
reflected waves. The second term in Equation 62 and Equation 63 will
be zero. This is however also true for a line terminated with its
characteristic impedance. A line is called a matched line when the load
impedance is equal to the characteristic impedance. If we consider a
line with the characteristic impedance Z0, a propagation constant and
with the length l terminated with a load impedance ZL connected to a
sinusoidal voltage source, and then the voltage and current distribution
on the line can be calculated as:
[V]

[A]

Equation 79
Equation 80

where z = l-z is the distance measured backward from the load. These
equations are derived from Equation 62 and Equation 63 by looking at
z=0 and z=l. These equations are solved using the fact that:

Equation 81

[ ]

and then using the hyperbolic functions to simplify. Having the voltage
and current distribution the input impedance, Zi, can be calculated,
which is the impedance that the source sees at z = 0 or z= l:

Equation 82

[ ]
Notice that when ZL equals Z0 the input impedance will be equal to the
characteristic impedance. So the voltage source only sees an
impedance Zi and the input current, Ii. The voltage, Vi, can then easily

be calculated from the source voltage, Vg, and the internal impedance
of the source Zg as:

Equation 83

Equation 84

Note that if the line is matched, Equation 79 and Equation 80 reduces


to:

Equation 85
Equation 86

This means that the voltage and current distribution on a matched line
are exactly the same as though the line has been extended to infinity.
The average input power delivered by the source can then be
calculated as:

Equation 87

where * denotes the complex conjugate. The average power delivered


to the load is calculated as:

Equation 88

In the area of telecommunication the absolute power is often measured


in dBm. That is the power in decibel, with the reference power one
milliwatt.

Equation 89

Where P is the power in milliwatt, V is the circuit voltage in volts and R


is the circuit impedance in across which V is measured. When
describing a voltage in decibel, measured in dBu, the reference is said
to be the power of 1 mW over 600 which represents a voltage of
0.775 V. Note that dBm and dBu will be the same if measured over 600
. If the impedance, Z, is not 600 , however, the absolute power in
dBm can be computed from the measured voltage, U, as:

Equation 90

4.3 Balanced and Unbalanced Transmission


A good balance of the telecommunication line is important for reducing
common-mode signals. Common-mode signals are the main cause of
the electromagnetic field around an aerial wire since the distance
between the conductors in the wire is much less than the distance
between the wire and ground. For differential-mode signals the phase
difference is 180 and the amplitudes are equal, therefore fields
generated by the two conductors will tend to cancel each other.

Figure 7 Radiation from common-mode signals and radiation from differential-mode signals.

Figure 7 illustrates simulated radiation from a wire, carrying a commonmode current to the left and a differential-mode current to the right.

The magnitudes of the currents were the same and the differences in
field strength between the two lines are also the same in both figures.
The existing radiation from the line to the right is due to the distance
between the conductors.
If the wire is not perfectly balanced a common-mode signal will appear
due to introduced differences in amplitude and phase. The effect of this
phenomenon can be calculated with the use of the longitudinal
conversion loss (LCL). In recommendation G.117 from ITU-T a
standardised method is described for determination of LCL [4]. First EL
is applied and VT is measured according to the following figure:

Figure 8 Method for measurement of LCL [4]

Then LCL can be calculated as [4]:

Equation 91

[dB]
To see the relation between the longitudinal conversion loss (LCL), as
measured in Figure 8, the unbalance of the line and the conversion
from differential signals to common-mode signals, we will study a
model of a telecommunication line as in the following figure:

Figure 9 Equivalent circuit of equipment and line

where ZC is the source impedance to ground and Z1 + Z2 is the


differential mode impedance (normally the termination which is
customarily the same as the characteristic impedance of the line). The
common-mode impedance is the impedance for the common mode
signals to ground, (Z1 in parallel with Z2) + ZL. If the measurement setup for the LCL-measurements shown in Figure 8 is applied to the line in
Figure 9, we get the following equivalent circuit:

Figure 10 Equivalent circuit of the LCL measurement set-up (Figure 8) applied to the
telecommunication line in Figure 9.

Then the common-mode current, I, in Figure 10, is given by:

Equation 92

Furthermore the currents, I1 and I2, in each conductor of the line are
given by:

Equation 93

Equation 94

The differential mode voltage, VT, can then be expressed as

Equation 95

Combining Equation 92, Equation 93, Equation 94 and Equation 95 and


substituting Equation 91 leads to:

Equation 96

Because the differential mode impedance is fixed and the source is


perfectly balanced the following assumptions can be made:

Equation 97
Equation 98

We can express the cable unbalance as the difference between the


differential impedances Z1 and Z2:

Equation 99

Substituting Equation 97, Equation 98 and Equation 99 in Equation 96


gives the following relation for the longitudinal conversion loss:

Equation 100

Finally by assuming a reasonable balance on the line,


the approximation:

, give

Equation 101

In the next step of the study this expression is more practical to use
than Equation 91. From Equation 101 it can easily be observed that
LCL is inversely dependent of the unbalance Z. In addition larger load
impedance to ground, ZL, will increase LCL and thus less current will be
converted from differential-mode to common-mode.

To be able to investigate the relationship between differential mode


signals and common mode signals a general model in Figure 9 is
refined below.

Figure 11 Equivalent circuit of Figure 9 showing differential mode and common mode currents

Note that the signal source is primarily producing a differential signal in


this case. By applying Kirchhoffs voltage law on the equivalent circuit
in Figure 11, the following two equations are obtained:

Equation 102
Equation 103

Combining these equations will result in the following common-mode


current:

Equation 104

Substituting Equation 97, Equation 98 and Equation 99 in Equation 104


leads to:

Equation 105

Assuming that we have a reasonable balanced line,


, results in:

Equation 106

or if we instead write this in decibel:

Equation 107

where the factor Q is:

Equation 108

But we want an expression containing the LCL so we rewrite Equation


108 and then use Equation 101 to insert the LCL. Q can then be
rewritten as:

Equation 109

Substituting the CM-impedance, which can be written as


, gives:

Equation 110

For the CM-current in Equation 107 this leads to:

Equation 111

However, if it is the common mode voltage (VC) that is the wanted


variable,
is simply subtracted from ICM in Equation 111 and
we get VC in dB V.
In most cases this relationship can under some assumptions be
simplified, for instance assuming that the line is approximately
balanced, i.e. ES in Equation 111 is equal to the logarithmic differentialmode voltage (VD).
Having the longitudinal conversion loss, the common-mode voltage
(VC) can then be calculated from the differential-mode voltage (VD) [1]:

[dB V]

Equation 112

Where ZCM is the common-mode impedance and Z0 is the differentialmode impedance. This relation between the differential-mode voltage
(VD) and the common-mode voltage (VC) is the most commonly used for
calculations on balanced transmission lines.
When the common-mode voltage is known it is possible to estimate the
interference field strength (E) by the use of a conversion factor (FAV)
from common-mode voltage to an electric field determined at a
specified distance from the cable [1].
[dB V/m]

Equation 113

or if we use the relationship in Equation 112 and replace the commonmode voltage in Equation 113 we get:

[dB V/m]

Equation 114

The relation between LCL and the radiated electric field strength have
also been measured [1] (in Figure 12 below) and it have been shown
that the radiated field strength is approximately inversely proportional
to LCL.

Figure 12 Relationship between LCL of the telecommunication line and the radiated field strength

The conversion from an unbalanced signal from the source to a


balanced signal on the line and back to an unbalanced signal again is
done with balanced to unbalanced transformers (baluns). There are
mainly three kinds of broadband baluns, shown in Figure 13 below.

Figure 13 Three different kinds of baluns with their equivalent schematic to the right

These can be divided into two distinct categories. The first two baluns
are voltage baluns causing equal and opposite voltages to appear at
the two output terminals. The third balun is a current balun, or choke
balun, forcing equal and opposite currents to flow on the line. If the line
and the terminating load is perfectly symmetrical, the voltage baluns
will force the voltages to be equal and opposite and thus the currents
flowing from the balun output terminals will be equal and opposite and
there will be no CM-current on the line. If, however, the line is not
perfectly symmetrical unequal currents, resulting in CM-currents, will
appear leading to line radiation.
A solution, that is good in the aspect of reducing line radiation, would
be to use a current balun, which makes it possible for opposite currents
to flow and rejects CM-currents. Another solution is to use a voltage
balun in combination with a current balun. Notice that even if the
voltage is perfectly balanced at the ends of the line, some CM-currents
will still appear due to non-symmetrical line and load. When using a
current balun on a balanced line it will function as a CM-suppressor.

The wire is wrapped around a ferrite core as in Figure 13. Then the
differential currents of opposite phase will produce magnetic fluxes of
opposite phase in the core that cancel each other and there will be no
remaining reactance in the core for differential-mode signals. The
common-mode current will however produce a magnetic flux that
appears as a reactance in series with the line.

Figure 14 Impedance in different ferrite beads

Ferrite cores are well suited as core material but the characteristic will
vary a lot with the frequency and with the ferrite material as shown in
Figure 14 above.

5 Models
Following are models derived to, in theory, determine the important
factors affecting the coupling between an aerial telecommunication
line and an antenna. We will also try to find the main contributions to
the radiated electromagnetic field in different situations. Since these
models are going to be used in the study of xDSL signals we will focus
on frequencies up to 30MHz. The models are divided into eight
different coupling paths. They are initially divided into four near-field
models and four far-field models. The near-field models, based on the
mutual impedance, will have more significant contributions to the
coupling when the distance between the emitting line and the
receiving antenna r is much less than
and the far-field models,
based on the radiated field, will dominate when the distance is much

greater than
(see Equation 43 to Equation 45). At distances
between the near-field and the far-field the contribution will be a
combination between the far-field models and the near-field models.

Figure 15 The major contribution to the coupling at frequencies up to 30MHz (FAR-FIELD =


radiated, NEAR-FIELD = mutual impedance)

Both the near-field models and the far-field models are divided into
looking at the magnetic coupling or fields and the electric coupling or
fields separately. This is because of the different characteristics of the
two. All cases are finally divided into one model for the common-mode
signals and one model for the differential-mode signals. For the
common-mode signals, the signals on a two-wire line are approximated
with one signal on a single conductor. In almost all cases the
differential-mode models show much less tendency to radiate than the
common-mode model, but since the differential-mode signals usually
are of a much greater magnitude than the common-mode signals in a
balanced transmission system, the contributions from the differentialmode signals can not always be neglected.

Notice that the far-field is said to start at the distance


if the
size of the source, in this case the length of the line D, is greater than a
wavelength. This has nothing to do with the different types of coupling,
i.e. if the induction terms or the radiation terms will dominate the
coupling (see Equation 43 to Equation 45). The distance is, however,
often used as the boundary between the near-field and the far-field
because at a shorter distance maxima and minima would appear due

to interference caused by different distances to different parts of the


source. That means that it is preferred to do discrete measurements
beyond this distance to get repeatable results. The distance, r, is
calculated for a couple of line-lengths in Figure 16 below.

Figure 16 Distances to a far-field without fluctuations using a long line compared to the wavelength

The coupling between the telecommunication line radiating


electromagnetic energy and a receiving antenna can be modelled with
an equivalent two-port network as in the following figure:

Figure 17 Equivalent two-port network of the coupling between the telecommunication line and an
antenna, with the antenna to the right and the coupling impedance Z12 (which is assumed to be equal
to Z21)

The voltage on the line V1 principally represents the common-mode


voltage of the line but could also be made up by differences in distance
to the differential-mode voltages. I1 is the current in the
telecommunication line and I2 is the current in the antenna. V2 is the

voltage induced across the load impedance of the antenna. These


voltages and currents are linearly related as:

Equation 115
Equation 116

where Z11, Z12, Z21 and Z22 are open circuit impedance coefficients. When
the medium between the line and the antenna is bilateral, governing
reciprocity relations, the coupling impedances Z12 and Z21 are equal.
This equivalent network is used in the following theory only by the
near-field models but it will hold for the far-field also since at large
distances, r, between the line and the antenna, the coupling
impedances will become very small:

Equation 117

Then the impedances Z11 and Z22 will be nearly equal to the input
impedances of the line and the antenna ZA and ZB. That means that
with weak coupling Equation 115 could be written as:

Equation 118

If we apply Thvenins theorem to the left of the load impedance of the


antenna, ZL, we can determine the open-circuit voltage VOC and internal
impedance Zg as:

Equation 119

Equation 120

Also here we can see that if we are in the far-field and have a weak
coupling, the internal impedance, Zg, for the antenna will be
approximately equal to its input impedance as:

Equation 121

We can write the produced voltage and current in the antenna due to
the current in the telecommunication line as:

Equation 122

Equation 123

These relations will be used in the following near-field models, where


we look at the magnetic coupling considering the coupling impedance
to be purely inductive and looking at the electric coupling considering
the coupling to be purely capacitive. We will also look at commonmode signals and differential-mode signals separately.
The mutual impedance, Z12 or Z21, between two arbitrary positioned
conductors above ground can be calculated according to the equations
in Appendix C. These equations are however quite difficult to solve.
Therefor we have decided to look at different situations to get
simplified but restricted solutions.

5.1 Near-Field Models


In the near-field, the main coupling between a telecommunication line
and an antenna is the mutual impedance, mainly consisting of mutual
inductance from magnetic field coupling and mutual capacitance from
the electric field coupling. The models for the near-field are divided in
magnetic or electric coupling in which both common-mode coupling
and differential-mode coupling are studied. The models are limited to
looking at aerial telecommunication lines since they radiate more than
buried cables, and represent a worst case situation. The results from
each model could then be superimposed to determine the overall
coupling. This can however produce errors when the length of the
cable is much less than the wavelength. In our case this is almost
always true. The purpose with the models will then be to find the main
cause of radiation and the characteristic of the radiation in a particular
case. This can be used to derive the important factors for this
particular kind of radiation. Knowing these factors is very important

when building realistic set-ups to measure the radiation and also when
looking at different reduction techniques.
Notice that the results and reduction techniques can also be applied to
cable-to-cable crosstalk problems or antenna to cable disturbances.
5.1.1 Magnetic Coupling
5.1.1.1 Common-mode Coupling

In this chapter we will study the magnetic coupling of the commonmode current to an antenna due to mutual inductance, M, shown in the
following figure:

Figure 18 Common-mode magnetic (mutual inductance) coupling between a telecommunication line


above and an antenna below. The common-mode current on the two wire line is approximated with
one current on a single conductor.

The distance between the conductors is considered very small


compared to the distance to the antenna. Therefore the two
conductors carrying the common-mode current are considered as a
single conductor.
The common-mode current in an aerial telecommunication line will
follow a closed loop, denoted C1. The currents will follow the copper

leads and then be coupled to ground via the common-mode


impedance, which can be purely capacitive, and return in the groundplane back to the source producing a closed loop bounding the surface
S1. The common-mode current I1 flowing around the circumference C1
of S1, will produce a magnetic field B1. A part of the magnetic flux
caused by B1 will link with the antenna, which bounds the surface S2.
The mutual flux is then:

Equation 124

If we take the surface integral of the fundamental postulate for


electromagnetic induction (Equation 17) we get:

Equation 125

where the left side is the induced electromotive force in the antenna,
U2. Notice that C2 does not have to be a physical closed loop. The right
side can be rewritten using Equation 124 and the result is the
Faradays law of electromagnetic induction, which states that the
electromotive force induced in a stationary closed loop circuit is equal
to the negative rate of inverse of the magnetic flux linking the circuit
[8]:

Equation 126

Since the emf U2 is induced by a time-varying magnetic field it is called


a transformer emf. The negative sign states that the induced emf will
cause a current, which in turn will produce an opposite magnetic flux.
This is known as Lenzs law.
Then by using Biot-Savarts law, which is obtained by taking the curl of
the vector magnetic potential A (as in Equation 25 and Equation 26):

Equation 127

where aR is the unit vector from the source point to the field point and
R is the distance from the wire element dl to the field point. We can see
that B1 is directly proportional to the common-mode current I1 and from

Equation 124 we can see that the mutual flux 12 is also proportional
to I1. If we define that proportionality factor as the mutual inductance,
M, between the telecommunication line and the antenna we have:

Equation 128

which states that the mutual inductance is the magnetic flux linkage
with one circuit per unit current in the other. In a similar way the self
inductance is defined as the magnetic flux linkage per unit current in
the loop itself as:

Equation 129

For a linear medium, the self inductance does not depend on the
current in the loop and exists regardless of whether the loop is open or
closed. The inductances depend on the geometrical shape of the
elements constituting the current and on the permeability of the
medium.
If we consider the wire as one long conductor and a parallel dipole
antenna as another long conductor where both conductors have the
same radius, a, for simplicity reasons. These are separated by a
distance d, which is much larger than a. Both conductors are in free
space. To calculate the mutual inductance we first combine Equation
124 and Equation 128 into:

Equation 130

where we use the subscript 1 to denote the telecommunication line


and the subscript 2 for the antenna. Since B1 can be calculated from
the curl of the vector magnetic potential A1 we can write (Equation 25):

Equation 131

where the vector magnetic potential for a thin wire, using Equation 26,
is:

Equation 132

where R is the distance to the point of observation, in this case the


distance between the telecommunication line and the antenna.
Combining Equation 131 and Equation 132 we get the mutual
inductance as:

Equation 133

where we can see that in this situation the mutual inductance would
vary inversely with the distance R between the antenna and the
telecommunication line. We can also see that for a linear medium, it is
proportional to the permeability and independent of the currents in the
circuits. The contour integrals over C1 and C2 is however hard to
calculate since the contour of a dipole antenna is non-obvious.
Interchanging the subscripts would not change the value of the double
integral which means that the reciprocity relations hold as discussed
earlier, Z12 = Z21. Equation 133 is called the Neumann formula for
mutual inductance.
The mutual inductance M as calculated in Equation 128 or Equation
133 would represent the value of a coupling inductor Z12 in Figure 17.
The self-inductances, as calculated in Equation 129, would be part of
the impedances Z11 and Z22 in the same figure. If Z12 is purely inductive
then -Z12 in the upper impedances in Figure 17 would represent pure
capacitances and the schematic in Figure 17 could be illustrated as in
the following figure:

Figure 19 Magnetic coupling with the mutual inductance M

From Figure 19 we can clearly see the origin of the term mutual
inductance, since both the current I1 and I2 are going through the same
inductance M. From Equation 122 we get the induced current in the
antenna as:

Equation 134

The mutual inductance as illustrated in Figure 19 can also be illustrated


as a transformer coupling using a lumped parameter circuit as in the
following figure:

Figure 20 Equivalent circuit of the magnetic coupling between a telecommunication line above and
an antenna below, showing the mutual inductance M and the self inductances L1 and L2.

U2 is induced emf in the antenna due to the common-mode current I1.


The mutual inductance can be written as:

Equation 135

where L1 is the self inductance of the telecommunication line, L2 is the


self inductance of the antenna and the constant k is called the
coefficient of coupling. This coefficient will depend on how much of the
magnetic flux from the common-mode current that is coupled to the
antenna. If there is no leakage flux this coefficient will be one, but in
our case it will be much less than one. Using the circuit in Figure 20 the
induced emf could be written as:

Equation 136

which is directly proportional to the common-mode current. It also


depends on the geometry by k and the self inductances of the
telecommunication line and of the antenna.
We also know from Equation 133 that the mutual inductance vary
inversely with the distance between the line and the antenna, R. We
could extract the distance from the coupling coefficient, k, and define a
new constant which will depend on the geometry of the loops only as
K. Then we can rewrite Equation 136 as:

Equation 137

5.1.1.2 Reduction of Common-mode Magnetic Coupling

The most obvious reduction techniques like reducing the loop area (i.e.
reducing the height and the length of the telecommunication line) or
increasing the distance, R, between the interfering and the interfered
object is however hard to do in practice.
Reducing the common-mode current in the line will of cause have a
direct effect on induced voltage in the antenna, as seen in Equation
137. Increasing the balance of the cable would reduce the commonmode current (see Equation 91). Increasing the common-mode
impedance and the use of different common-mode rejection
techniques could also be utilised to reduce the common-mode current.
The common-mode current is also reduced by reducing the differentialmode current since that is the source of the common-mode current.
Operating at lower frequencies, if possible, would decrease the
reactance of M.

5.1.1.3 Differential-mode Coupling

With differential-mode currents on the line, the two wires will cause
almost equal and opposite magnetic fluxes that will tend to cancel
each other. We denote the mutual inductance from one wire to the
antenna M13 and the mutual inductance from the other wire to the
antenna M23, as in the following figure:

Figure 21 The magnetic (mutual inductance) coupling between a two wire telecommunication line
carrying differential-mode currents above and an antenna below.

The differential-mode currents are equal and have opposite direction.


We can then write the induced current by using Equation 134 and the
law of superposition:

Equation 138

or with the emf-method as in Equation 136 (superposition):

Equation 139

or with the distance extracted as in Equation 137:

Equation 140

where we can see that if the mutual inductances are equal, there
would not be an induced emf in the antenna. Due to slight differences
in the distances between the antenna and the two lines and
differences in the self inductances in the lines there will however be
some coupling.
5.1.1.4 Reduction of Differential-mode Magnetic Coupling

The reduction techniques for reducing differential-mode magnetic


coupling is based on getting the mutual inductance from the two wires
to be as equal as possible. This is done by twisting the wire, which is
equal to reducing the loop area between the wires. It is important to
keep the wires as close as possible.
Also here we can see that the induced voltage in the antenna is
directly proportional to the differential-mode current and varies
inversely with the distance between the antenna and the line, R.
5.1.2 Electric Coupling
5.1.2.1 Common-mode Coupling

In this chapter we will study the electric coupling of the common-mode


current to an antenna due to mutual capacitance, C, shown in the
following figure:

Figure 22 Common-mode electric (mutual capacitance) coupling between a telecommunication line


above and an antenna below. The two wire line carrying the common-mode signal is approximated
with a single conductor.

As in the case of magnetic coupling, the distance between the


conductors are considered very small compared to the distance to the
antenna. Therefore the two conductors carrying the common-mode
current are considered as a single conductor.
The capacitance between the telecommunication line and the antenna
is a physical property. It depends on the geometry of the line and the
antenna and of the permittivity of the medium between them. The
capacitance of an isolated conducting body is the electric charge that
must be added to the body per unit of increase in its electrical
potential. The capacitance was defined from the observation that the
ratio between the charge, Q, and the voltage, V, is a proportionality
constant which remains constant.

Equation 141

where the unit is coulomb per volt or farad, F. Recall that for a parallel
plate capacitor of area S the capacitance C is expressed as:

Equation 142

where d and are the distance between the plates and the permittivity
of the dielectric that space.
Consider an infinitely long line charge with a charge density
will cause a cylindrical electric field with intensity E at the
perpendicular distance r from the line charge.

[C/m]. It

Equation 143

This relationship can be used to approximate the electric field intensity


of both the dipole antenna and the transmission line.
At a distance r from the line charge an electric potential can be
calculated by integrating the electric field intensity E over the distance
from the line charge to the point where the potential is to be
calculated.

Figure 23 Cross section of a line charge, in P, and its image in a parallel conductor

If the ekvivalent diameter of the transmission line is approximately


equal to the diameter of the dipole antenna wire, both diameters can
be written as the same variable a.

Figure 24 Cross section of a telecommunication cable in parallel with the conductor of a dipole
antenna and equivalent line charges

The potential in the dipole antenna at a distance d from a line charge in


the transmission line is:

Equation 144

and the corresponding potential caused by a line charge in the dipole


antenna is:

Equation 145

If l is the length for which the transmission line and the dipole antenna
are parallel, the mutual capacitance can simply be expressed as (using
Equation 141):

Equation 146

To use the mutual capacitance in the coupling impedance concept as in


Figure 17, we have to rewrite the mutual capacitance coupling as a
mutual inductance by assuming a very large resistance to ground as in
the following figure:

Figure 25 Equivalent circuit of mutual capacitance, using mutual inductance instead

That the left side is the same as the right side in Figure 25 is realised
when calculating the voltage out on the right, V2, from applying a
voltage, V1, to the left (voltage dividing), first on the left circuit (in
Figure 25):

Equation 147

and then the same with the circuit to the right:

Equation 148

The equivalent circuit will then be the same as in Figure 19 with the
capacitance C as the value of M. The resistance 1/R will be neglectible
compared to Z11 in Figure 19 since R is very large. Then we get the
following relation of the current in the antenna due to the commonmode current in the telecommunication line (as in Equation 134):

Equation 149

which is directly proportional to the common-mode voltage in the


telecommunication line and to the capacitance between them. Notice
that this current has the same sign as that from mutual inductance in

Equation 134 and thus these two contributions will be additive and not
cancel each other.
5.1.2.2 Reduction of Common-mode Electric Coupling

Since we can not control the antennas ground impedance, the only
factors we can change are either the common-mode voltage or the
capacitance between the line and the antenna. The capacitance is hard
to reduce since we can not separate the line and the antenna or
reduce the radius of the wires. A possibility is ofcourse to operate at
lower frequencies to increase the reactance caused by C. Reducing the
common-mode voltage is done by increasing the balance of the cable
(see Equation 91), increasing the common-mode impedance and the
use of different common-mode rejection techniques (these methods
also increases LCL as seen in Equation 101). The common-mode
voltage is also reduced by reducing the differential-mode voltage since
that is the source of the common-mode voltages.
5.1.2.3 Differential-mode Coupling

With differential-mode voltages on the line, the two wires will cause
almost equal and opposite currents in the antenna which will tend to
cancel each other. We denote the mutual capacitance from one wire to
the antenna C13 and the mutual capacitance from the other wire to the
antenna C23, as in the following figure

Figure 26 The electric (capacitive) coupling between a two wire telecommunication line carrying
differential-mode signals above and an antenna below.

The current in the antenna due to the differential-mode voltages will


then be:

Equation 150

where I1 is the current in one conductor in the line and I2 is the current
in the other conductor. But since these are differential currents only, I2
will have the same magnitude as I1 but opposite direction:

Equation 151

Then we can write Equation 150 as:

Equation 152

which is directly proportional to the differential-mode currents on the


line and to the differences in mutual capacitance to the two conductors
in the line. We can see that if the mutual capacitances are equal, the
induced currents in the antenna will cancel each other. Due to slight
differences in the distances between the antenna and the two wires
and differences in the geometry between the two wires, there will
however be some nonzero capacitive coupling even here.
5.1.2.4 Reduction of Differential-mode Magnetic Coupling

The most common reduction technique is to twist the wires to keep


them as close as possible to each other. Reducing the differential-mode
currents on the line will have a direct effect as seen in Equation 152.
Equal diameters of the wires are important. The frequencies should be
kept as low as possible. Separating the antenna and the line is very
favourable, but hard to accomplish.

5.2 Far-Field Models


In the far-field the main cause of coupling between a
telecommunication line and an antenna is the radiation term as
derived in chapter 3. We chose here to look at the electromagnetic
fields generated instead of the inductive and capacitive coupling as in
the near-field models. The models are divided into magnetic fields from
current loops and electric fields from time-varying voltages. These
models are divided into common-mode radiation and differential-mode
radiation. The results could, with care, be superimposed, but the
purpose is to locate the main cause of radiation in a particular case as
with the near-field models. Also here it is very important to know the
main factors in each type of radiation to be able to build a realistic
case for measuring, or to know which factor to focus on to reduce a
specific kind of radiation.
In the models for both the magnetic and the electric radiation we will
focus on a straight aerial telecommunication line as shown in the
following figure:

Figure 27 The main radiation direction and polarisations for a rectangular loop

5.2.1 Magnetic Fields


5.2.1.1 Common-mode Radiation

In this chapter we will study the magnetic radiation due to commonmode current flowing as shown in the following figure:

Figure 28 The common-mode current loop causing magnetic fields, with aerial telecommunication
lines. The common-mode signals on the two wire line are approximated with one signal on a single
conductor.

As before the distance between the conductors are considered very


small compared to the distance to the point of observation. Therefore
the two conductors carrying the common-mode current are considered
as a single conductor carrying a current.
Consider a small loop of radius b carrying a uniform time-harmonic
current i(t)=Icos t around its circumference as in the following figure:

Figure 29 A magnetic dipole or current loop.

This is called an elemental magnetic dipole, which has a vector


magnetic moment, m, as:

[Am2]

Equation 153

To determine the electromagnetic field from this current loop, we need


the vector magnetic potential, A, since the magnetic flux density, B, is
the curl of the vector magnetic potential. Assuming that we have a thin
wire and the current is flowing entirely along the wire we have:

Equation 154

This integral is however hard to calculate exactly, because R1 will


change with the location of dl on the loop. If we assume that we have
a small loop we can solve the vector magnetic potential as:

Equation 155

Then the electric and magnetic field intensities, E and H, can be solved
by deriving the magnetic flux from the vector magnetic potential, A,
for the magnetic field and then the electric field can be calculated from
the curl of the magnetic field intensity as:

Equation 156

Equation 157

From this we get the electric and magnetic field intensities as:

Equation 158

Equation 159

Equation 160

Notice the similarity with the equations for the electric dipole, derived
in Equation 43, Equation 44 and Equation 45, and that the nature of
the near and far-field discussed earlier also applies to these equations.
For the far-field (R >> /2 ) these equations will simplify to:

Equation 161

[V/m]
Equation 162

[A/m]
where =2 c/ . We can see that the far-field intensities vary
inversely as R and their ratio E /H equals the intrinsic impedance of
free space, 0. We can also see that the maximum fields are produced
in the same plane as the current loop, where is /2.
The vector magnetic moment, m, is the current, I, times the area of the
loop which we denote S. That means that the electric (and the
magnetic) field intensity vary linearly with the current in the loop and
the area of the loop. If we look at the electric field intensity in a point in
the same plane as the loop, in the x-y plane, where we have the
maximum field intensity, we could write the electric field intensity as a
constant depending only on the frequency and the distance R times the
current and the area of the loop (in free space):

Equation 163

where the area, S, is in m2 and the current, I, is the peak amplitude


current in amperes at appropriate frequency, then the constant K is
derived as [5]:

Equation 164

The current loop can be either a small loop or a large loop which will
affect the input impedance of the loop and therefor the current flowing
in the loop. We have earlier derived the input impedance for a
transmission line (Equation 82) repeated here:

Equation 165

If we consider an ideal transmission line and assume negligible line


losses it can be shown that the input impedance can be written as:

Equation 166

where Z0 is the characteristic impedance of the line, ZL is the load


impedance, l is the length of the line and is the wavelength as in the
following figure:

Figure 30 The input impedance for a line of length l

From Equation 166 we can see that for small loops with l << (often
written as l < /10) we will have the input impedance, Zi, almost equal
to the load impedance, ZL. If we have the line as illustrated in Figure 30
we will have a current in the circuit as:

Equation 167

At low frequencies or short lines where we have a small loop, the


current is essentially uniformly distributed around the loop. When the
length become a significant fraction or multiple of the wavelength, the
loop becomes a large loop. Since the input impedance vary with the
frequency and the length of the line there will be a corresponding
variation in the amplitude of the loop current. The input impedance will
vary between maximum and minimum values depending on the
relationship between the wavelength and the length of the cable or the
circumference of the loop. The current distribution along the
telecommunication line is generally in the form of standing waves with
current maxima and minima along the line resulting in directional
radiation field patterns. The amplitude of the standing waves is
determined by the match between the characteristic impedance of the
line and the load impedance.
Notice that in the common-mode situation, the load impedance is the
common-mode impedance, which could be reactive (capacitive
coupling to ground). If the load impedance is purely reactive, ZL=jXL,
then the input impedance can be written as:

Equation 168

and we will have a purely reactive input impedance with poles and
zeros at different frequencies depending on the relation between the
line length, load impedance and the characteristic impedance of the
line.
If we look at the common-mode current loop as illustrated in Figure 28
we know that the electric and magnetic field intensity in the far-field is
proportional to the loop area which is the medium height over ground
times the length of the line and to the common-mode current flowing
in the loop. Notice that in the aerial telecommunication line case, the
loop will often become a large loop.
5.2.1.2 Reduction of Common-mode Magnetic Radiation

The most obvious way would be to reduce the loop area by reducing
the height over ground or reducing the length of the cable, but since
this would be to expensive, the reduction techniques are focusing on

reducing the common-mode current amplitude. This is mainly done by


reducing the ground loop coupling. The ground loop coupling can be
reduced by on or more of the following techniques:

The use of ferrite beads on signal lines (common-mode choke)


Float (ungrounded) box shields inside equipment enclosures
Install RF chokes in the case-to-ground path
Use an isolation transformer
Use an optical isolator
Using a balanced system, with a balanced transmitter and
receiver
Float (unground) either or both the transmitter and receiver

Shielding the loop area by gorunding it is, at present, quite expensive.


We can see that a buried cable would not have as large loop area as an
aerial telecommunication cable. So if burying the cable is too
expensive, the only way to reduce the radiation is to have as good
balance as possible to reduce the amount of common-mode current
and then to make it as hard as possible for the common-mode current
to flow around the loop.
5.2.1.3 Differential-mode Radiation

The theory presented above for common-mode signals will hold in this
case too except for the fact that the loop area will be much less than in
the case of common-mode current in aerial cables. The loop area for
the differential mode signal will be the area between the conductors of
the cable, as illustrated in the following figure:

Figure 31 The differential-mode current loop causing magnetic fields.

Note that this area will be even less in a real cable since the cables are
twisted, intentionally or unintentionally. The different parts of a twisted
cable will not radiate in the same direction and will in some cases tend
to cancel each other. Also what was said about the input impedance,
see Figure 30, depending on the length of the cable can be applied to
this situation, see Equation 166 and about the maxima and minima in
Equation 168.
5.2.1.4 Reduction of Differential-mode Magnetic Coupling

Also here the radiation will be proportional to the current in the loop,
see Equation 163, which means that a reduction of the differentialmode current would have a direct effect on the amount of radiated
fields, by the same amount. It is also direct proportional to the loop
area, Equation 163, and inversely proportional to the distance, R, to the
point of observation (Equation 161 and Equation 162).
The most common reduction technique is the use of twisted pairs. The
reduction of the differential mode coupling that is offered by twisted
pair can be written as [8]:

Equation 169

[dB]
where n is the number of twists per meter, l is the length of the wire in
meters and is the wavelength in meters. The reduction of differentialmode coupling offered by twisted pair is plotted in the following figure:

Figure 32 Differential-mode coupling reduction offered by twisting wire pair

The x-axis is the total number of twists along the whole wire and each
curve corresponds to the product of the number of twists per meter, n,
and the wavelength, .
5.2.2 Electric Fields
5.2.2.1 Common-mode Radiation

A two wire communication line carrying a common mode signal can be


approximated as a single conductor carrying a signal. To be able to
derive the far field radiation from such a line, it is modelled as a
composition of many short conducting elements. The length of each
element should be only a fraction of the wavelength of the signal in the
conductor so that the current distribution within each element can be
assumed to be uniform. We can now use Equation 43, Equation 44 and
Equation 45 to calculate the radiation from each element. However
since we in this part are only interested in the far field radiation
(R>> /2 ) the contribution to the far field radiation from each
element can be calculated according to the following simplified
equations:

Equation 170

[A/m]
Equation 171

[V/m]
By examining the relationships in these equations we see that the
magnitude of the fields in the far field is proportional to the current
through the element I, the length of the element dl and inversely
dependent of the distance R and that the ratio between E and H
equals the intrinsic impedance of free space, 0. Another important
observation is that the magnitude of the field is maximised
perpendicular to the conducting element where = /2, and minimised
along the axis of the conducting element where =0, see Figure 3.
Consider a typical aerial telecommunication line with horizontally and
vertically oriented segments according to Figure 33. According to the
theory above, the common mode current will produce an
electromagnetic field. The current distribution will also give rise to a
charge distribution that also will produce an electromagnetic field. The
value of the common-mode impedance will decide whether the

common-mode current or the common-mode voltage will be the main


cause of radiation.

Figure 33 The model for the electric fields generated by common-mode signals on the line. The
common-mode signals on the two wire line are approximated with one signal on a single conductor.

The distribution of currents along a telecommunication line above a


perfectly conducting ground has a corresponding image of an opposite
current distribution equidistant to, but at the opposite side of the
ground plane. This is called method of images. The vertical segments
(downleads) in Figure 33 represent the common mode current path to
ground. This path may consist of a capacitive coupling as well as a
conductor and a common-mode impedance.
If the common-mode impedance is relatively small then the resulting
radiation is mainly due to currents. Radiation caused by common-mode
currents in the horizontal oriented segment of the line and its image
will tend to cancel each other at some point in the far-field where the
distance to the line and its image is equal or differs by a multiple of a
wavelength. Since we are only interested in the fields at almost ground
level the distance to the line and its image will be almost equal and the
fields will cancel, see Figure 34 below.

Figure 34 Simulated electric fields from a telecommunication line above ground using the method of
images

If the horizontal telecommunication line is several wavelength long and


properly terminated it behaves as a travelling wave antenna with the
current distribution:

Equation 172

At points where the distance to the line and its image differs by an
uneven fractions of a wavelength, radiation will occur.

Figure 35 Radiation pattern of a travelling wave antenna ( is the elevation angle above earth)

Again, this is not the case at ground level in the far field.
If the common-mode impedance instead is relatively large then the
resulting radiation is mainly due to the charge distribution. In this case
the radiation caused by the potential of the horizontal oriented
segment of the line and its image will form a electrical dipole radiating
a vertical polarised electric field.

Figure 36 A telecommunication line over ground containing a charge distribution, and its
corresponding image.

Instead of expressing the electric field in terms of current distribution,


as in Equation 171, a time varying field can be expressed in terms of
charge distribution as:

Equation 173

where

Equation 174
Equation 175

To be able to derive the electric field intensity in the far field we need
to integrate Equation 173 to get:

Equation 176

The capacitance per unit length between the telecommunication line


and earth can be calculated as:

Equation 177

where D is the average ekvivalent cable diameter of the


telecommunication line and h is the heigth of the line above earth. This
results in that an aerial telecommunication line of length L will have a
total capacitance to earth of:

Equation 178

The fundamental expression for a charge

Equation 179

can by substitution, using Equation 178, be rewritten as:

Equation 180

By examining Equation 176 we can identify a dependence of the


derivative of the charge. The derivative of the charge can be written
as:

Equation 181

This is illustrated in the following figure:

Figure 37 A telecommunication line over ground acting as a vertical dipole

It can can also be expressed in terms of the potential V and the


capacitance per unit length of the telecommunication line.

Equation 182

This can be used to express the radiation as:

Equation 183

A quick analyse of Equation 183 confirms that the electric field


intensity in the far field will increase when:

the potential of the telecommunication line increases


the length of the telecommunication line increases
the capacitance per unit length between the line and earth
increases
the height of the telecommunication line above earth increases
the distance between the line and the observation point
decreases
the angle between a normal to the ground plane and a line
from the telecommunication line to the observation point
increases towards 90

The resulting magnitude of the radiation will at ground level also


depend on the azimuth angle to the communication line. Perpendicular
to the line the magnitude will be maximised and in parallel with the
line the magnitude will depend on the distance between the downleads
in relation to the wavelength as in the situation of a two element
antenna array.
The dependence of the distance between the communication line and
the observation point can also be calculated as the free space
propagation loss. By assuming that far-field propagation and a
spherical wavefront applies and neglecting effects on the propagation
of polarisation, antenna heights, curvature of earth, atmospheric
conditions, reflections and interfering objects along the path, the free
space propagation loss can be approximated with the following
formula, also shown in Figure 38.
[dB]

Equation 184

where FSPL is the Free Space Propagation Loss, R is the distance to the
point of observation and f is the frequency in MHz.

Figure 38 The Free Space Propagation Loss

Notice that this is intended only to be used as a rule of thumb when


evaluating the propagation loss at a distance from the source.
5.2.2.2 Reduction of Common-mode Electric Radiation

The signal source will ideally only produce differential-mode signals,


common-mode signals will then arise from differential-mode signals
due to longitudinal conversion loss (Equation 91). Hence, a way to
reduce the common-mode electric radiation is to increase LCL
(increase balance) to prevent conversion from differential-mode signals
to common-mode signals. There will however always be some
remaining common-mode signal that will cause radiation. To minimise
the electric radiation we then have to reduce the common mode
impedance, an undesired consequence of this is that the magnetic
radiation will increase.
5.2.2.3 Differential-mode Radiation

The electric field radiation due to differential-mode signals is caused by


opposite charges in the two wires of the telecommunication line.

Figure 39 The model for the electric fields generated by differential-mode signals on the line.

The theory for common-mode radiation above applies to this case too,
if we consider having the opposite charges located in the two wires
instead of in the line and its image. This will cause the distance
between the opposite charges to be much smaller in this case for an
aerial communication line. The capacitance between the two wires
(Equation 65) is larger than the capacitance between the line and
earth. As in the case of differential-mode magnetic radiation in the far
field the magnitude of the radiation is also dependent on how the
telecommunication line is twisted. As in illustrated in Figure 7 the
electrical fields generated from the two conductors will tend to cansel
each other.
5.2.2.4 Reduction of Differential-mode Electric Radiation

Because of the similarities with the case of differential-mode magnetic


field radiation in the far field, the same reduction techniques are
applicable. Twisting the cable is the best reduction technique (Equation
169).

Appendix A - Symbols and Units


Table 6 Fundamental SI Units

Quantity

Symbol

Length

Unit

Abbreviation

meter
m

l
Mass

kilogram
kg

m
Time

second
s

t
Current

ampere
A

I, i
Table 7 Derived Quantities

Quantity

Symbol

Unit

Abbreviatio
n

Admittance

siemens
S

Y
Angular frequency
Attenuation constant

Capacitance

radian/second
rad/s
neper/meter
Np/m
farad
F

C
Charge

coulomb
C

Q, q
Charge density
(linear)
Charge density
(surface)
Charge density
(volume)

l
s

Conductance

coulomb/meter
C/m
coulomb/meter2
coulomb/meter3

Current density
(surface)
Current density
(volume)
Dielectric constant

C/m3

siemens
S

G
Conductivity

C/m2

siemens/meter
S/m
siemens/meter
A/m

Js
siemens/meter2
J

A/m2

r
coulomb/meter2

Electric flux density


D
Electric field intensity

C/m2

volt/meter
V/m

E
Electric potential

volt
V

V
Electromotive force

volt
V

Energy

joule
J

W
joule/meter3

Energy density
w
Frequency

hertz
Hz

f
Impedance

Z,

Inductance

ohm

henry
H

L
ampere-meter2

Magnetic dipole
moment

Magnetic field
intensity

A/m
weber
Wb

Magnetic flux density

tesla
T

B
Magnetic potential

weber/meter
Wb/m

Permittivity
Phase
Phase constant

0
0

Power

henry/meter
H/m
farad/meter
F/m
radian
rad
radian/meter
rad/m
watt
W

P
Propagation constant

A.m2

ampere/meter

Magnetic flux

Permeability

J/m3

Reactance

meter-1

m-1

ohm
X

Relative permeability
Relative permittivity

r
r

Resistance

ohm

R
Voltage

volt
V

V
Wavelength

Wavenumber

meter
m
radian/meter
rad/m

k
Work

joule
J

Appendix B - Constants
Table 8 Constants of Free Space

Constant

Symbol

Velocity of light
c
Permittivity
Permeability
Intrinsic impedance

Value
3.108 [m/s]

[H/m]

0
0

120 or 377 [ ]

Table 9 Physical Constants

Constant

Symbol

[F/m]

Value

9.107.10-31 [kg]

Rest mass of electron


me

-1.602.10-19 [C]

Charge of electron
-e

-1.759.1011 [C/kg]

Charge-to-mass ratio of electron


-e/me

2.81.10-15 [m]

Radius of electron
Re

1.673.10-27 [kg]

Rest mass of proton


mp

Table 10 Relative Permittivities (average low-frequency values at room temperature)

Material

Value

Material

Value

Air

1.0

Plexiglas

3.4

Backelite

5.0

Polyethylene

2.3

Glass

4-10

Polystyrene

2.6

Mica

6.0

Porcelain

5.7

Oil

2.3

Rubber

Paper

2-4

Soil

3-4

Paraffin wax

2.2

Teflon

2.1

2.3-4.0

Table 11 Relative Permeabilities (average low-frequency values at room temperature)

Material

Value

Material

Ferromagnetic
:

Value

Diamagnetic:

Nickel

Bismuth
250

Cobalt

0.99983
Gold

600
Iron

0.99996
Silver

4000

0.99998

Mumetal

Copper
100000

0.99999

Paramagnetic:
Aluminium
1.000021
Magnesium
1.000012
Palladium
1.00082
Titanium
1.00018
Table 12 Conductivities (average low-frequency values at room temperature)

Material

Value [S/m]

Material

Value [S/m]

Silver

6.17.107

Distilled water

2.10-4

Copper

5.80.107

Dry soil

10-5

Gold

4.10.107

Transformer oil

10-11

Aluminium

3.54.107

Glass

10-12

Brass

1.57.107

Porcelain

2.10-13

Bronze

107

Rubber

10-15

Iron

107

Fused quartz

10-17

Table 13 Ground Conductivity and Dielectric Constants

Surface Type

Dielectric Conductivi Commen


Constant ty [S/m]
t

Salt water
81

5.0

80

0.001

Excellent
Ground

Fresh water
Pastoral, low hills, rich soil type

Very good
20

0.03

Flat country, marchy, densely


wooded

13

0.007

Pastoral, medium hills, few


trees

13

0.006

Pastoral, heavy clay soil


Rocky soil, steep hills,
mountainous

Average
13

0.005

12

0.002

10

0.002

Poor

Sandy, dry, flat


Cities
Cities, high buildings, industrial
areas

Very poor
5

0.001

0.001

You might also like