You are on page 1of 22

X International Symposium on

Lightning Protection
9
th
-13
th
November, 2009 Curitiba, Brazil

A REVIEW OF FIELD-TO-TRANSMISSION LINE COUPLING
MODELS WITH PARTICULAR REFERENCE TO
LIGHTNING-INDUCED VOLTAGES

Farhad Rachidi
Swiss Federal Institute of Technology (EPFL)
Lausanne, Switzerland

E-mail: Farhad.Rachidi@epfl.ch


Abstract In this lecture, we discuss the Transmission Line (TL) theory and its application to the problem of lightning
electromagnetic field coupling to transmission lines. We start with the derivation of the general field-to-transmission line
coupling equations for the case of a single-wire line above a perfectly conducting ground. The derived equations are solely
based on the thin-wire approximation and they do take into account high frequency radiation effects. Under the transmission
line approximation, the general equations reduce to the Agrawal et al. field-to-transmission line coupling equations. After a
short discussion on the underlying assumptions of the TL theory, three seemingly different but completely equivalent
approaches that have been proposed to describe the coupling of electromagnetic fields to transmission lines are presented.
The derived equations are then extended to deal with the presence of losses and multiple conductors and expressions for the
line parameters, including the ground impedance and admittance are presented. The time-domain representation of the field-
to-transmission line coupling equations, which allows for a straightforward treatment of nonlinear phenomena as well as the
variation in the line topology, is also described. Solution methods in the frequency domain and in the time domain are given
and application examples with reference to lightning-induced voltages are presented and discussed. Specifically, the effects of
ground losses and corona are illustrated and discussed. When the travelling voltage and current waves are originated from
lumped excitation sources located at a specific location along a transmission line (direct lightning strike), both the corona
phenomenon and ground losses result in an attenuation and dispersion of propagating surges along transmission lines.
However, when distributed sources representing the action of the electromagnetic field from a nearby lightning illuminating
the line are present, ground losses and corona phenomenon could result in important enhancement of the induced voltage
magnitude.

1 INTRODUCTION

The problem of lightning-induced voltages on overhead and buried transmission lines has been seriously reconsidered
in recent years due to the increasing demand by customers for good quality in the power supply [1].
The evaluation of lightning-induced voltages requires the knowledge of the electromagnetic field change along the
considered line. This electromagnetic field is generally determined assuming that the lightning return stroke channel is a
straight vertical antenna above a conducting plane. Some studies have attempted to take into account the channel
tortuosity and inclination in the computation of electromagnetic fields (e.g. [2-6]). The spatial and temporal distribution
of the current along the channel is specified using a return stroke model.
The problem of return stroke modeling and electromagnetic field computation is beyond the scope of this lecture. It is
worth mentioning that
- Four classes of lightning return stroke models have been defined by Rakov and Uman [7]: (1) The gas dynamic
models, (2) electromagnetic models, (3) distributed-circuit models, and (4) engineering models. Outputs of the
electromagnetic, distributed-circuit, and engineering models can be directly used for the computation of electromagnetic
fields. A review on recent work on these three types of models can be found in [8].
- Three different approaches have been adopted to compute the electromagnetic fields, both above and below the earth
surface: (1) Numerical solution of the exact equations through dedicated algorithms, (2) numerical solution of the
Maxwells equations using numerical methods, such as the finite-difference time-domain (FDTD) technique or the
method of moments (MoM), and (3) use of simplified equations. These approaches have been recently reviewed by
Rakov and Rachidi [8].
In this lecture, we present the general theory describing the interaction of an impinging electromagnetic field with
transmission lines, with particular reference to lightning-induced voltages.
67

The lecture is organized as follows. In Section 2, we present the derivation of the general field-to-transmission line
coupling equations for the case of a single-wire line above a perfectly conducting ground [9]. The derived equations are
based on the thin-wire approximation and they do take into account high frequency radiation effects. In Section 3, we
show how the general equations reduce to the Agrawal et al. field-to-transmission line coupling equations derived under
the transmission line (TL) approximation [10]. We also discuss the underlying assumptions of the TL theory. In the
same section, we describe the equivalent approaches that have been proposed to describe the coupling of
electromagnetic fields to transmission lines [10-12]. The derived equations are then extended to deal with the presence
of losses and multiple conductors and expressions for the line parameters, including the ground impedance and
admittance are presented. Section 4 presents solution methods in the frequency domain and in the time domain.
Application examples of lightning-induced voltages are presented and discussed in Section 5, with special emphasis on
the effects of ground losses. General conclusions are given in Section 6.

2 FULL-WAVE FIELD-TO-TRANSMISSION LINE COUPLING EQUATIONS

To solve the coupling problem, i.e., the determination of voltages and currents induced by an external field on a system
of conducting wires, use could be made of the exact solutions of the Maxwell's equations using numerical methods,
such as Finite-Difference Time-Domain (FDTD) or the method of moments (MoM). However, due to the length of
typical overhead line installations, together with the need of modeling other components (e.g., in power transformers,
surge arresters, general line terminations), the use of such brute-force methods for the calculation of electromagnetic
field interaction with power lines is not straightforward and implies long computing times and considerable memory
requirements.
One way to cope with this problem is the elaboration of the so-called generalized or full-wave TL theory, which
incorporates high frequency radiation effects, while keeping the relative simplicity of TL equations [13]. Several
approaches have been proposed in the past decade or so (e.g. [13-18]).
In this Section, we will briefly describe the derivation of Tkachenko et al. [9] which refers to the geometry presented in
Fig. 1.
Consider a lossless current filament of finite length above a perfectly conducting ground. The line is in presence of an
external electromagnetic field.



Fig. 1 A single-wire line above a perfectly conducting ground in presence of an impinging electromagnetic field.


The boundary conditions on the wire surface implies that the total electric field tangential to the wire should be equal to
zero

( )
0
tot e s
x x
e E e E E = + =
r r r
r r
on the surface of the wire along the wire axis (1)

In (1)
e
E
r
is the exciting electric field obtained by the sum of the incident field
i
E
r
and the ground-reflected field
r
E
r
,
both determined in the absence of the wire;
s
E
r
is the scattered electric field which represents the reaction of the wire to
the excitation field.
The following development will be based on the thin-wire approximation, that is, the current and charge densities are
assumed to be distributed along the wire axis and the condition (1) is satisfied on the surface of the wire. The scattered
x
y
z
0
h
2a
L
e
E
r
e
B
r
68

electric field, produced by the linear charge and current densities and J
r
, can be expressed in terms of the retarded
scalar and vector potentials:

= A j E
s
r r
(2)
with
0
0
( ) ( ') ( , ) '
4
L
x
A r e I x g r r dx

=

r
r r r r
and
0
0
1
( ) ( ') ( , ) '
4
L
r x g r r dx

=

r r r
(3 a,b)
where ' x is the length variable along the wire axis,
z y x
e z e y e x r
r r r r
+ + = is the vector from the observation point to the
origin, r
r
is the vector from the source point to the origin, and ( ') I x and ( ') x are the current and charge density along
the wire. ) , ( r r g
r r
is the scalar Greens function given by

2 2 2 2 2 2
( ') ( ) ( ) ( ') ( ) ( )
2 2 2 2 2 2
( , )
( ') ( ) ( ) ( ') ( ) ( )
jk x x y y z h jk x x y y z h
e e
g r r
x x y y z h x x y y z h
+ + + + +
=
+ + + + +
r r
(4)

where the wave number k is related to the angular frequency by c k / = .

Defining the scattered voltage as in the standard transmission line theory [10]

0
( ) ( , 0, )
h
S s
z
V x E x z dz =

(5)

And after mathematical manipulations (see [9] for details), we obtain

0
0
( )
( ') ( ') ' ( , 0, )
4
s
L
e
x
dV x
j g x x I x dx E x h
d x

+ =

(6)
0
0
( ') ( ') ' 4 ( ) 0
L
s
d
g x x I x dx j V x
d x
+ =

(7)

in which the scalar Greens function ( ) g x is given by

2 2 2 2
4
2 2 2 2
( )
4
jk x a jk x h
e e
g x
x a x h
+ +
=
+ +
(8)

The system of equations (6) and (7) may be arranged and written in the following equivalent form

{ }
( )
( ) ( , 0, ) ( )
4
s
e
z
dV z
j L I x E x h j D I x
d x

+ = +
)
(9)
( )
{ }
( ) 1
( ) ( )
2ln 2
s
d I x d
j C V x D I x
d x h a d x
+ =
)
(10)

where the operator D
)
is defined as

{ } ( )
2 2 2 2
( ') ( ') 4
2 2 2 2
0
( ) 2ln 2 ( ) ( ') '
( ') ( ') 4
jk x x a jk x x h
L
e e
D I x h a I x I x dx
x x a x x h
+ +

=


+ +

)
(11)
and L and C are the per-unit-length inductance and capacitance of the horizontal infinite wire above a perfect
ground, as defined in the classical transmission line theory [19]. They are given by

69

( )
0
ln 2
2
L h a

=
( )
0
2
ln 2
C h a
h a

= >> (12 a,b)



Equations (9) and (10) are general, rigorous
1
equations describing the interaction of an external electromagnetic field
with the considered single-wire line above a perfectly conducting ground. The equivalent circuit representation of (9)
and (10) is shown in Fig. 2. This system (9)-(10) has a form similar to that of classical field-to-transmission line
coupling equations in the Agrawal et al. formulation [10]. However, additional distributed voltage and current sources
which depend on the unknown current are also present as additional source terms. These terms represent
electrodynamics corrections (including radiation) to the transmission line theory.


Fig. 2- Equivalent circuit corresponding to equations (9) and (10).



3 TRANSMISSION LINE THEORY

3.1 Transmission Line (TL) Approximation

Power networks extend, in general, over distances of several kilometres, much larger than the minimum wavelengths
associated with lightning electromagnetic pulse (LEMP). Indeed, significant portions of the frequency spectrum of
LEMP extend to frequencies up to of a few MHz, which corresponds to minimum wavelengths of about 100 m or less
(e.g. [20]). On the other hand, the transverse dimensions of typical overhead power lines are well below the minimum
wavelength associated to LEMP.
The main assumptions of the transmission line theory are [19]:

- Propagation occurs along the line axis.
If the cross-sectional dimensions of the line are electrically small, propagation can indeed be assumed to occur
essentially along the line axis only and the first assumption can be considered to be a good approximation.

- The sum of the line currents at any cross-section of the line is zero. In other words, the ground the reference
conductor is the return path for the currents in the n overhead conductors.
By assuming that the sum of all the currents is equal to zero, we are considering only transmission line mode currents
and neglecting the so-called antenna-mode currents [21]. If we wish to compute the load responses of the line, this
assumption is adequate, because the antenna mode current response is small near the ends of the line. Along the line,
however, and even for electrically small line cross sections, the presence of antenna-mode currents implies that the sum
of the currents at a cross section is not necessarily equal to zero [19, 21, 22]. However, the quasi-symmetry due to the
presence of the ground plane, if present, results in a very small contribution of antenna mode currents and,
consequently, the predominant mode on the line will be transmission line [21].

- The response of the line to the coupled electromagnetic fields is quasi transverse electromagnetic (quasi-TEM) or, in
other words, the electromagnetic field produced by the electric charges and currents along the line is confined in the
transverse plane and perpendicular to the line axis [23].
The condition that the response of the line is quasi-TEM is satisfied only up to a threshold frequency above which
higher-order modes begin to appear [19, 21]. For some cases, such as infinite parallel plates or coaxial lines, it is
possible to derive an exact expression for the cutoff frequency below which only the TEM mode exists [19]. For other

1
The only assumption in deriving (9) and (10) is the thin-wire approximation.
( , 0, )
e
x
E x h dx ' L dx
' C dx
( )
s
V x dx + ( )
s
V x
( ) I x ( ) I x dx +
x dx + x
+ +
{ } ( )
4
j D I x dx

)
( )
{ }
1
( )
2ln 2 /
d
D I x dx
h a dx
)
70

line structures (i.e. multiple conductors above a ground plane), the TEM mode response is generally satisfied as long as
the line cross section is electrically small [19].

For uniform transmission lines with electrically-small cross-sectional dimensions (not exceeding about one tenth of the
minimum significant wavelength of the exciting electromagnetic field), a number of theoretical and experimental
studies have shown a fairly good agreement between results obtained using the TL approximation and results obtained
either by means of antenna theory or experiments (see for example [24, 25]).


3.2 Derivation of the Field-to-Transmission Line Coupling Equations

Let us start with the general equations (6) and (7) and consider the integral
0
( ) ( ') ( ') '
L
C x g x x I x dx =

(13)
Assuming the TL approximation, the following conditions apply

1 << kh , h L 2 >> (14 a,b)

In this case, (13) can be written as
2 2 2 2
0
1 1
( ) ( ) '
( ') ( ') 4
L
C x I x dx
x x a x x h



+ +

(15)

Further, if the observation point x is sufficiently far from the wire ends, that is, if

2 2 h x L h << << (16)

then the integration limits of (15) can be taken from to + (since the integrals from to L and from L to +
are negligible) and the integral reduces to

( ) ( ) ( ) 2ln 2 / C x I x h a (17)

Introducing (17) in (6) and (7) results in the following equations

( )
( ) ( , 0, )
s
e
z
dV x
j L I x E x h
d x
+ = (18)
( )
( ) 0
s
d I x
j C V x
d x
+ = (19)

The above equations are the well-known field-to-transmission line equations of Agrawal et al. [10]. In their original
publication, Agrawal et al. [10] derived the same pair of equations following a different approach.
For termination impedances
A
Z and
B
Z (see Fig. 3), the boundary conditions in terms of the scattered voltage and the
total current as used in (9) and (10), are given by

0
(0) (0) (0, 0, )
h
s e
A z
V Z I E z dz = +

(20)
0
( ) ( ) ( , 0, )
h
s e
B z
V L Z I L E L z dz = +

(21)

71



Fig. 3- Field-illuminated single-wire line above ground terminated on impedances
A
Z and
B
Z .

The equivalent circuit representation of this model (equations (5), (18)-(21)) is shown in Fig. 4. For this model, the
forcing function (the exciting electric field tangential to the line conductor) is represented by distributed voltage sources
along the line. In agreement with boundary conditions (20) and (21), two lumped voltage sources are inserted at the line
terminations.


Fig. 4 Equivalent circuit of a lossless single-wire overhead line excited by an electromagnetic field according to the
Agrawal et al. model.


It is also interesting to note that this model involves only electric field components of the exciting field and the exciting
magnetic field does not appear explicitly as a source term in the coupling equations.

3.3 Equivalent Formulations

Taylor et al. model [11]

An equivalent formulation of field-to-transmission line coupling equations was proposed nearly two decades earlier
than was Agrawal et al.s, by Taylor, Satterwhite and Harrison in 1965 [11]. In the Taylor et al. formulation, the
coupling equations are expressed in terms of the total induced current and the total induced voltage. They are given by

0
( )
' ( ) ( , 0, )
h
e
y
dV x
j L I x j B x z dz
dx
+ =

(22)
0
( )
' ( ) ' ( , 0, )
h
e
z
dI x
j C V x j C E x z dz
dx
+ =

(23)

The boundary conditions for the load currents and voltages must be enforced. They are simply given by

) 0 ( ) 0 ( I Z V
A
= (24)
x
y
z
0
h
2a
L
e
E
r
e
B
r
A
Z B
Z
x x dx +
C
( , 0, )
e
x
E x h dx
' L dx
' C dx
( )
s
V x dx + ( )
s
V x
( ) I x ( ) I x dx +
A
Z B
Z
(0)
s
V ( )
s
V L
x dx + x
L 0
+
+ +
0
( , 0, )
h
e
z
E L z

0
(0, 0, )
h
e
z
E z

72


) ( ) ( L I Z L V
B
= (25)

Equations (22)-(25) can be represented using an equivalent circuit, as shown in Fig. 5. The forcing functions (source
terms) in (22) and (23) are included as a set of distributed series voltage and parallel current sources along the line.

Fig. 5 - Equivalent circuit of a lossless single-wire overhead line excited by an electromagnetic field.
Taylor et al. model.


Rachidi model [12]

Another form of the coupling equations, equivalent to the Agrawal et al. and to the Taylor et al. models, has been
derived by Rachidi [12]. In this model, only the exciting magnetic field components appear explicitly as forcing
functions in the equations:

0 ) ( '
) (
= + x I L j
dx
x dV
s
(26)
0
( ) 1
' ( ) ( , 0, )
'
e s
h
x
B dI x
j C V x x z dz
dx L y


+ =


(27)
in which ) (x I
s
is the so-called scattered current related to the total current by

) ( ) ( ) ( x
e
I x
s
I x I + = (28)
where the excitation current ) (x I
e
is defined as

0
1
( ) ( , 0, )
'
h
e e
y
I x B x z dz
L
=

(29)

The boundary conditions corresponding to this formulation are

0
(0) 1
(0) (0, 0, )
'
h
s e
y
A
V
I B z dz
Z L
= +

(30)
0
( ) 1
( ) ( , 0, )
'
h
s e
y
B
V L
I L B L z dz
Z L
= +

(31)

The equivalent circuit corresponding to the above equivalent set of coupling equations is shown in Fig. 6.

0
( , 0, )
h
e
y
j B x z dzdx

0
' ( , 0, )
h
e
z
j C E x z dzdx

' L dx
' C dx
( ) V x dx + ( ) V x
( ) I x ( ) I x dx +
A
Z
B
Z (0) V ( ) V L
x dx + x
L 0
+
73


Fig. 6 - Equivalent circuit of a lossless single-wire overhead line excited by an electromagnetic field. Rachidi model.


3.4 Remarks on the Three Equivalent Models

The three coupling models are different but fully equivalent approaches that predict identical results in terms of total
voltages and total currents, in spite of the fact that they take into account the electromagnetic coupling in different ways.
Nucci and Rachidi [26] have shown, on the basis of a specific numerical example that, as predicted theoretically, the
total induced voltage waveforms obtained using the three coupling models discussed earlier are identical. However, the
contribution of a given component of the exciting electromagnetic field to the total induced voltage and current varies
depending on the adopted coupling model.
The Agrawal et al. [10] model is the most widely used among the three models for essentially two reasons. First, the
distributed source term in the Agrawal et al. model is expressed very simply as the tangential electric field and its
evaluation does not require any differentiation or integration. Second, the source terms are expressed in the same way
when taking into account the finite ground conductivity
2
.
It is also worth noting that an earlier model was developed by Rusck [27] for the evaluation of lightning-induced
voltages on overhead lines. It has been shown by Nucci et al. [28] and by Cooray [29] that Ruscks equations, expressed
in terms of scalar potential, are absolutely equivalent to the above-discussed models, as far as the excited
electromagnetic field is originated by a straight vertical channel. The Rusck model and its extensions have been
extensively used in the literature (e.g. [30],[31]).

3.5 Inclusion of Losses

In the calculation of lightning-induced voltages, losses are, in principle, to be taken into account both in the wire and in
the ground. Losses due to the finite ground conductivity are the most important ones, and they affect both the
electromagnetic field and the surge propagation along the line [32].
Let us make reference to the same geometry of Fig. 3, and let us now take into account losses both in the wire and in the
ground plane. The wire conductivity and relative permittivity are
w
and
rw,
respectively, and the ground, assumed to be
homogeneous, is characterized by its conductivity
g
and its relative permittivity
rg
. The Agrawal et al. coupling
equations extended to the present case of a wire above an imperfectly conducting ground can be written as (for a step by
step derivation see [21])

( )
' ( ) ( , 0, )
s
e
x
dV x
Z I x E x h
dx
+ = (32)
0 ) ( '
) (
= + x V Y
dx
x dI
s
(33)

where Z' and Y' are the longitudinal and transverse per-unit-length impedance and admittance respectively, given by [21,
32]

g w
Z Z L j Z ' ' ' ' + + = (34)
( )
g
g
Y C j G
Y C j G
Y
' ' '
' ' '
'
+ +
+
= (35)

2
For example, in the Taylor et al. model, the ground finite conductivity results in an additional source term in the first equation given
by the exciting tangential electric field on the earths surface [21] .
0
1
( , 0, )
'
e
h
x
B
x z dzdx
L y


' L dx
' C dx
( ) V x dx +
( ) V x
( )
s
I x ( )
s
I x dx +
A
Z
B
Z
(0) V
( ) V L
x dx + x
L 0
0
1
( , 0, )
'
h
e
y
B L z dz
L

0
1
(0, 0, )
'
h
e
y
B z dz
L

74


in which
- L', C' and G' are the per-unit-length longitudinal inductance, transverse capacitance and transverse conductance,
respectively, calculated for a lossless wire above a perfectly conducting ground:

=

a
h
L
o 1
cosh
2
'

a
h
o
2
ln
2
for h >> a (36)

( ) a h
C
o
/ cosh
2
'
1

=
( ) a h
o
/ 2 ln
2
for h >> a (37)

' '
air
C G
o

= (38)

-
w
Z' is the per-unit-length internal impedance of the wire; assuming a round wire and an axial symmetry for the
current, the following expression can be derived for the wire internal impedance (e.g. [33]):

) ( I 2
) ( I
'
1
a a
a
Z
w w
w o w
w


= (39)

where ( )
rw o w o w
j j + = is the propagation constant in the wire and I
o
and I
1
are the modified Bessel
functions of zero and first order, respectively;

-
g
Z' is the per-unit-length ground impedance, which is defined as [34, 35]
'
) , (
' L j
I
dx z x B j
Z
h
s
y
g

=

(40)

where
s
y
B is the y-component of the scattered magnetic induction field.
Sunde [36] derived a general expression for the ground impedance which given by

dx
x x
j
Z
g
hx
o
g
+ +

=


0
2 2
2
e
' (41)

where ) (
rg o g o g
j j + = is the propagation constant in the ground.
The general expression (41) is not suitable for a numerical evaluation since it involves an integral over an infinitely long
interval. Several approximations for the ground impedance of a single-wire line have been proposed in the literature
(see [32] for a survey). One of the simplest and most accurate was proposed by Sunde himself and is given by the
following logarithmic function

h
h
j
Z
g
g
o
g
1
ln
2
' (42)

It has been shown [32] that the above logarithmic expression represents an excellent approximation to the general
expression (41) over the frequency range of interest.
Finally,
g
Y' is the so-called ground admittance, given by [21]
75

g
g
Z
Y
g
'
'
2

(43)

For typical overhead power lines, the effect of ground admittance and the wire impedance is negligible and can be
disregarded in the computation [1, 34]
3
.


3.6 Multiconductor Lines

The field-to-transmission line coupling equations for the case of a multi-wire system along the x-axis above an
imperfectly conducting ground and in the presence of an external electromagnetic excitation (Fig. 7) are given by [21,
24, 38]

[ ( )] [ ' ][ ( )] [ ' ][ ( )] [ ( , 0, )]
i
ij
s e
ij i g i x i
d
V x j L I x Z I x E x h
dx
+ + = (44)
[ ( )] [ ' ][ ( )] [ ' ][ ( )] [0]
i i
s s
i ij ij
d
I x G V x j C V x
dx
+ + = (45)

in which
- )] ( [ x V
s
i
and )] ( [ x I
i
are frequency-domain vectors of the scattered voltage and the current along the line;
- )] , ( [
i
e
x
h x E is the vector of the exciting electric field tangential to the line conductors;
- [0] is the zero-matrix (all elements are equal to zero);
- ] ' [
ij
L is the per-unit-length line inductance matrix. Assuming that the distances between conductors are much larger
than their radii, the general expression for the mutual inductance between two conductors i and j is given by [21]

+
+ +

=
2 2
2 2
) (
) (
ln
2
'
j i
ij
j i
ij
o
ij
h h r
h h r
L (46)
The self inductance for conductor i is given by

=
ii
i o
ii
r
h
L
2
ln
2
' (47)

- ] ' [
ij
C is the per-unit-length line capacitance matrix. It can be evaluated directly from the inductance matrix using the
following expression [21]

1
' '
ij o o ij
C L

=

(48)

- ] ' [
ij
G is the per-unit-length transverse conductance matrix. The transverse conductance matrix elements can be
evaluated starting either from the capacitance matrix or the inductance matrix using the following relations

1
' ' '
air
ij ij air o ij
o
G C L


= =

(49)

In most practical cases, the transverse conductance matrix elements
ij
G' are negligible in comparison with
ij
C j '
[19] and can therefore be neglected in the computation.

3
Note that for buried cables, the effect of the ground admittance is no longer negligible [37] .
76

- Finally, ] ' [
ij
g
Z is the ground impedance matrix. The general expression for the mutual ground impedance between
two conductors i and j derived by Sunde is given by [36]

dx x r
x x
j
Z
ij
g
x h h
o
g
j i
ij
) cos(
e
'
0
2 2
) (

+ +

=

+
(50)

In a similar way as for the case of a single-wire line, an accurate logarithmic approximation is proposed by Rachidi et
al. [38] which is given by

+
+

2 2
2 2
2 2
2
)
2
( 1
ln
4
'
ij
g
j i
g
ij
g
j i
g
o
g
r h h
r h h
j
Z
ij
(51)
Note that in (44) and (45), the terms corresponding to the wire impedance and the so-called ground admittance have
been neglected. This approximation is valid for typical overhead power lines [32].
The boundary conditions for the two line terminations are given by

0
[ (0)] [ ][ (0)] [ (0, 0, ) ]
i
i
h
s e
A i z
V Z I E z dz = +

(52)

0
[ ( )] [ ][ ( )] [ ( , 0, ) ]
i
i
h
s e
B i z
V L Z I L E L z dz = +

(53)

in which [Z
A
] and [Z
B
] are the impedance matrices at the two line terminations.



Fig. 7 - Cross-sectional geometry of a multiconductor line above a conducting ground plane in the presence of an
external electromagnetic field



e
E
r
e
B
r
i
h
j
h
ij
r
2
i
a
2
j
a
i
j
,
g rg

77

3.7 Coupling to Complex Networks

In order to take into account the presence of power system components, line-discontinuities, and complex system
topologies, the LEMP-to-transmission line coupling model has been linked with appropriate circuit solver software
taking advantage of the large available library of power system components (e.g [25, 39-49]). The developed models for
the calculation of LEMP-caused transients in overhead power lines have been experimentally validated using reduced
scale setups with LEMP and NEMP (nuclear electromagnetic pulse) simulators, and full-scale setups illuminated by
rocket-triggered lightning fields (see [25] for a review).


4 SOLUTION FOR FIELD-INDUCED CURRENTS AND VOLTAGES

4.1 Frequency-Domain Solutions

The field-to-transmission line coupling equations, together with the boundary conditions, can be solved using Greens
functions, which represent the solutions for line current and voltage due to a point voltage and/or current source [21]. In
this section, we will present the solutions, using the Agrawal at al. model for the case of a single-conductor line. Similar
solutions can be found for the case of a multiconductor line (see e.g. [21], [19]).
Considering a voltage source of unit amplitude at a location x
s
along the line, the Greens functions for the current and
the voltage along the line read, respectively [21],
( )
( ) ( )
( ; )
2 1
2
2 1
1 2
L
x x
e
x L x L
G x x e e e e
I s
L
Z e
c


> >



< <
=


(54)
( )
( ) ( )
( ; )
2 1
2
2 1
1 2
L
x x
e
x L x L
G x x e e e e
V s
L
e


> >



< <
= +


(55)
where
- x
<
represents the smaller of x or x
s
, and x
>
represents the larger of x or x
s
.
=1 for x > x
s
and = -1 for x < x
s
.
-
' ' Y Z = is the complex propagation constant along the transmission line,
- ' / ' Y Z Z
c
= is the lines characteristic impedance.
-
1
and
2
are the voltage reflection coefficients at the loads of the transmission line given by

c B
c B
c A
c A
Z Z
Z Z
Z Z
Z Z
+
-
+
-
2 1
= = (56)

The solutions in terms of the total line current and scattered voltage can be written as the following integrals of the
Greens functions [21]
0 0 0
( ) ( ; ) ( ; 0) (0, 0, ) ( ; ) ( , 0, )
L h h
e e
I s s s I z I z
I x G x x V dx G x E z dz G x L E L z dz = +

(57)
0 0 0
( ) ( ; ) ( ; 0) (0, 0, ) ( ; ) ( , 0, )
L h h
s e e
V s s s V z V z
V x G x x V dx G x E z dz G x L E L z dz = +

(58)
Note that the second and the third terms on the right hand side of (57) and (58) are due to the contribution of equivalent
lumped sources at the line ends (see Fig. 4).
The total voltage can be determined from the scattered voltage by adding the contribution from the exciting field as

0
( ) ( ) ( , 0, )
h
s e
z
V x V x E x z dz =

(59)

If we are interested in the transmission line response at its terminal loads, the solutions can be expressed in a compact
way by using the so-called BLT (Baum, Liu, Tesche) equations [21]

78


2
1
1
2
1
2
1
1 0
0 1
/ 1
) (
) 0 (
S
S
e
e
Z
L I
I
L
L
c
(60)

+
+
=


2
1
1
2
1
2
1
1 0
0 1
) (
) 0 (
S
S
e
e
L V
V
L
L
(61)

where the source vector is given by
1
2
0 0 0 1
2
( )
1
2
0 0 0
1
( , 0, ) (0, 0, ) ( , 0, )
2 2
1
( , 0, ) (0, 0, ) ( , 0, )
2 2
s
s
L
L h h
x e e e
x s s z z
L
L h h
L x e e e
x s s z z
e
e E x h dx E z dz E L z dz
S
S
e
e E x h dx E z dz E L z dz


+



=




+



(62)

Note that in the BLT equations, the solutions are directly given for the total voltage and not for the scattered voltage.
For an arbitrary excitation field, the integrals in Equation (62) cannot be carried out analytically. However, for the
special case of a plane wave excitation field, the integrations can be performed analytically and closed-form expressions
can be obtained for the load responses. General solutions for vertical and horizontal field polarizations are given in [21].



4.2 Time-Domain Solutions

A time domain representation of the field-to-transmission line coupling equations allows the straightforward treatment
of non linear phenomena as well as the variation in the line topology [24]. On the other hand, frequency-dependent
parameters, such as the ground impedance, need to be represented using convolution integrals.
The field-to-transmission line coupling equations (44) and (45) can be converted into time domain to obtain the
following expressions [23, 38]

x
|:

s
(x, t)] + |I'
]
]

t
|i

(x, t)] + |'


]
]@

t
|i

(x, t)] = |E
x
c
(x, y = u, z = b

, t)] (63)

x
|i

(x, t)] + |0'


]
]

t
|:

s
(x, t)] + |C'
]
]

t
|:

s
(x, t)] = u (64)

in which denotes convolution product and the matrix |'
]
] is called the transient ground resistance matrix; its
elements are defined as

j
i
]
[ = F
-1
||Z
i
g]
]]| (65)

The inverse Fourier transforms of the boundary conditions written, for simplicity, for resistive terminal loads read

|:

s
(x, t)] = -|R
A
]|i

(u, t)] + j] E
z
c
(x = u, y = u, z, t)
h
i
0
[ (66)

|:

s
(I, t)] = -|R
B
]|i

(I, t)] + j] E
z
c
(x = I, y = u, z, t)
h
i
0
[ (67)

where |R
A
] and |R
B
] are the matrices of the resistive loads at the two line terminals.

The general expression for the ground impedance matrix terms in the frequency domain does not have an analytical
inverse Fourier transform. Thus, the elements of the transient ground resistance matrix in time domain have to be, in
general, determined using a numerical inverse Fourier transform algorithm. However, analytical expressions have been
derived by Rachidi et al. [50] and Araneo and Celozzi [51] which have been shown to be reasonable approximations to
the numerical values obtained using an inverse FFT. More discussion on the validity of the approximate analytical
expressions can be found in [52].
One of the most popular approaches to solve the coupling equations in the time domain is the FDTD technique (e.g.,
79

[53]). Such a technique was already used by Agrawal et al. in [10] where partial time and space derivatives were
approximated using a 1
st
order FDTD scheme. In [54], instead, the use of a 2
nd
order FDTD scheme based on the Lax-
Wendroff algorithm [55, 56] was proposed. The 2
nd
order FDTD scheme shows much better stability compared to its 1
st

order counterpart, especially when analyzing complex systems involving nonlinearities [54, 57].
The 2
nd
order discretized solutions for the line current and scattered voltage are given by Paolone et al. [25]

2
1 1
1 1 1 1 1 1 1
2
2
1
2
' ' '
2 2 2
'
' '
2
n n n n n n n
i i xi xi i i i
n n k k k k k k k
i i ij ij ij
k k
gi
ij ij
i i E E v v v
t
v v t C L C
x x x
v
t
L C

+ + + +


+


= + +







+

1 1
'
2
n n
gi
k k
v
x
+





(68)

2
1 1
1 1 1 1 1
2
2
1
2
' ' ' '
2 2
' ' '
2
n n n n n
i i i i i n
n n n k k k k k
i i ij xi gi ij ij
k k k
k
xi
ij ij ij
v v i i i
t
i i t L E v C L
x x
E
t
C L C

+ + +


+


= + + + +





+

1
1 1
2
1
' '
' ' '
2 2
n n
n n
gi gi
xi
k k k k
ij ij ij
v v
E
t
C L C
t t






+







(69)

where

- x is the spatial integration step;
- t is the time integration step;
- k=0,1,2,, k
max
is the spatial discretization index (k
max
=(L/x)+1, where L is the line length);
- n= 0,1,2, , n
max
is the time discretization index;
-
1 n
i
k
v
+


is the vector of the scattered voltages corresponding to the spatial and time discretization indexes k and n+1,
respectively;
-
1 n
i
k
i
+


is the vector of the conductors currents corresponding to the spatial and time discretization indexes k and n+1,
respectively;
-
1 n
xi
k
E
+


is the vector of the exciting horizontal electric field along the wires corresponding to the spatial and time
discretization indexes k and n+1, respectively;
-
1
0
' '
k k
n i i n n h
n n
gi gij
k k
h
i i
v
t



=




=



.


5 APPLICATION TO LIGHTNING-INDUCED VOLTAGES

Ground losses are generally associated with attenuation and dispersion of propagating surges along transmission lines.
This is indeed the case when the travelling voltage and current waves are originated from lumped excitation sources
located at specific points along the line.
However, the situation is different when the traveling waves are originated from distributed voltage and/or current
sources along the line, representing the interaction of an external electromagnetic field with the line. Indeed, it has been
shown that line losses due to the ground finite conductivity (e.g. [32, 58-60]) or the corona phenomenon (e.g. [61])
could result in important enhancement of the induced voltages and currents.
The aim of this Section is to illustrate the complex effects of ground losses in field-to-transmission line interaction, and
to emphasize that such effects could result in important enhancement of voltages induced by external fields, as opposed
to direct overvoltages.



80

5.1 Effect of Ground Losses on Overvoltages due to a Direct Strike

Let us consider a 20-km long, 7.5-m high overhead line above a perfectly conducting ground. We shall present the
overvoltages due to a direct strike to the line, calculated at different observation points along the line, as shown in Fig.
8. The lightning current has a peak value of 4 kA and a maximum time derivative of 2 kA/s.










Fig. 8 - Direct lightning overvoltages along an overhead line.

Figure 9 shows the computed voltages calculated taking into account ground losses. The adopted ground parameters in
the simulations are
g
= 0.001 S/m and
r
= 10, and the expression for the ground transient resistance is the one
proposed in [50]. As expected, it can be seen that the resulting voltage wave experiences the typical dispersion as
traveling away from the strike point.


Fig. 9 - Effect of ground losses on traveling voltages along the line, due to a direct lightning strike (adapted from [62]).


It is important to mention that for direct lightning strikes, the effect of ground losses is far less significant than the effect
of corona. In Fig. 10, the voltages calculated taking into account the corona effect (and omitting the effect of ground
losses) are shown. The corona model, in other words the dependence of the dynamic capacitance as a function of the
voltage, is taken from Nucci et al. [61]. The effect of ground losses have been neglected in the simulation results shown
in Fig. 10.
It can be seen that the traveling waves exhibit the typical distortion and attenuation associated with corona effect. A
comparison between the results of Figs. 9 and 10 shows, additionally, that the overvoltages are more significantly
affected by the corona effect, rather than by the ground losses.

In the simulations presented in Figs. 9 and 10, we assumed the magnitude of the overvoltages are below the lightning
impulse withstand voltage of the line. In most of the cases, however, the surge propagating from the point of strike
along the line is altered by flashovers occurring between the strike location and the point of interest. Practically, all
flashovers to ground occur at the poles, as on overhead distribution lines, the weakest insulation is generally at a pole
structure rather than between conductors through air [63]. Fig. 11 shows a typical overvoltage (evaluated by
0
200
400
600
800
1000
0 5 10 15 20
Uo
2 km
4 km
6 km
V
o
l
t
a
g
e

(
k
V
)
Time (s)
2 km 2 km 2 km
U
3
U
2
U
1
U
O

R R
20 km
7 km 7 km
81

calculations)
Electromagn
eight spans (
as a transmis
(current-depe
calculated 60
overvoltage w


Fig. 1


Fig. 11 -

5.2 Effect of

We will now
radiated by a
detailed desc
The wire is
relative perm
kA/s (typica
Fig. 12 pres
terminal. Th
experimental
conducting; a
due to a direc
netic Transient
(9 poles) of 20
ssion line with
endent), with
00 m from th
which present
0 - Influence o
Example of a ty
f Ground Los
w consider a s
a nearby lightn
cription of mo
at a height of
mittivity
r
= 1
al of subseque
ents the volta
is configurati
l line in Mexi
and (ii) consid
ct lightning st
t Program (EM
00 m length e
h a characteris
30 at zero
he stroke loc
ts a few very s
f corona on trav
ypical lightning
sses on Induc
single-wire ov
ning return str
dels used in th
f 10 m above
10. The return
ent return stro
age induced a
ion has been
ico [65]. The
dering the fini
0
-80
-40
0
40
80
120
160
[kV]
trike of 30 kA
MTP) [64] for
each with a ch
stic impedance
o current. The
cation. The ex
short spikes, f
veling voltages
g overvoltage d
perform
ced Overvolta
verhead line m
roke. The com
he code).
e a ground pla
n stroke curren
oke).
along a 2.8-k
chosen becau
voltages repo
itely conductin
10
Voltage 600mf roms
current ampli
r a single-wir
haracteristic im
e of 300 . Th
e insulator fla
xample show
followed by an

along the line,
due to a direct st
med by T. Henr
ages
matched at bo
mputations are
ane characteri
nt has a peak v
km line for a
use it is simila
orted in Fig. 1
ng ground.
20
stroke location . Light
itude. The cal
e line with no
mpedance of 4
he footing DC
ashover voltag
ws the general
n impulse volt
due to a direct
trike to the MV
rikssen)
th ends, and i
carried out by
ized by a grou
value of 12 kA
return stroke
ar to an even
12 are calculat
30 40
tning current peak : 30
lculations hav
o ground wire
440 . Each p
C resistance w
ge was fixed
l characteristi
tage with a sm

lightning strike
V line. (Adapted
illuminated by
y means of the
und conductiv
A and a maxi
e located in t
nt recorded by
ted assuming
0 50
[us]
kA
e been perform
es. The line is
pole, 8-m hig
was assumed to
at 150 kV. T
ics of the dir
moother shape
e (adapted from

d from [63], com
y the electrom
e LIOV code
vity
g
= 0.00
imum time de
the vicinity o
y De La Rosa
(i) the groun
med using the
composed of
h, is modeled
o be nonlinear
The voltage is
rect lightning
[63].
m [61]).
mputations
magnetic field
(see [1]) for a
01 S/m and a
rivative of 40
f the left-end
a et al., on an
d as perfectly
e
f
d
r
s
g
d
a
a
0
d
n
y
82



Fig. 12 - Lightning-induced voltages along the line. Solid line: perfect ground; dashed line: lossy ground. Adapted from [62].


The computed results show that the voltages calculated taking into account ground losses exhibit significantly larger
amplitudes than those calculated assuming a perfectly conducting ground. Also, the voltages exhibit an inversion of
polarity as the observation point moves towards the far end of the line.
The reason why, for the examined case, ground losses can result in an amplitude enhancement has been discussed
thorougly in [1, 60]. In summary, it can be said that ground finite conductivity does not significantly affect the vertical
electric field amplitude and waveshape, but acts in modifying the horizontal electric field waveshape and in particular it
causes an inversion of polarity of it [32, 66, 67]. This inversion of polarity is more pronounced for larger distances and
for larger values of ground resistivity, as shown in Figure 13.


Fig. 13 - Radial electric field at four distances from the stroke location calculated with the Cooray-Rubinstein formula [66, 67].
Ground conductivity 0.001 S/m, ground relative permittivity 10. Observation point: at 10 m above ground. For illustrative purpose,
the values are multiplied by the square of the distance from the stroke location. (Adapted from [60])

It is important to bear in mind, however, that for different stroke locations and observation points along the line, ground
losses also could result in an attenuation of the induced voltages (see [1, 34]).
Figure 12 shows also that the voltage induced along the line exhibits larger amplitudes than that close to the stroke lo-
cation. Experimental data obtained by De La Rosa and co-workers [65] support this theoretical finding, since for an
-20
-16
-12
-8
-4
0
4
8
0 5 10 15 20
I
n
d
u
c
e
d

O
v
e
r
v
o
l
t
a
g
e

(
k
V
)
Time (s)
0 m
1 km
2 km
2.8 km
-100
-80
-60
-40
-20
0
20
0 2 4 6 8 10
100 m
500 m
1 km
2 km
E
x
.
r
2


(
V
/
m

*

k
m
2
)
Time (s)

2.8 km
730 m
4
3
0

m

83

event similar
the close one
Additional ex
were present
experimental
voltage mag
assuming a p
Fig. 14 -


As we have
effect. For th
for particular
this case, it h
opposed to th
retically, by
propagation
for the total
magnitudes (
As an examp
line, taking in

Fig. 15 - Vol
lines: taking i
r to the one pr
e.
xperimental d
ted by Ishii e
l reduced-scal
nitude for a g
perfectly-cond
Lightning-indu
seen in Sect.
he case of ind
rly severe exc
has been show
he typical atte
considering
speed of the v
induced volta
(see [61] for a
ple, figure 15
nto account th
ltage induced by
into account cor
resented in Fi
data supportin
et al. [58, 68]
le line is com
ground condu
ducting ground
uced voltage on
. 5.1, lightnin
duced voltages
citation condit
wn that the cor
enuation obser
that the incre
various surges
age - which re
a more detailed
shows the lig
he corona effe
y a nearby light
rona; dotted lin
line center an
g. 12, they ob
ng the enhance
. An example
mpared with co
uctivity of 0.0
d.
aan experimen
computati
ng overvoltage
s by a nearby
tions (very cl
rona effect wo
rved in the ca
ease of the li
s induced by l
esults from al
d explanation)
ghtning-induce
ect [61].
tning at three o
nes: disregarding
nd equidistant t
x
7.5m
1.5m
bserved line fl
ement of the i
e is shown in
omputation re
06 S/m is abo


ntal reduced sca
ions (Adapted f
es due to a d
lightning stri
lose impact po
ould result in a
ase of direct st
ine capacitanc
ightning. This
ll the contribu
).
ed voltages on
bservation poin
g corona. Groun
to the line termi
25m
V
lashover occur
induced volta
n Fig. 14 whe
esults. As it c
out twice as
ale line. Compar
from [68]).
direct strike ar
ike, corona ef
oint and/or lar
an enhanceme
trikes [61]. Th
ce produced
s reduction in
utions of the v
n a 1-km long
nts along a 1 km
nd: perfectly c
inations. (adapt
m
rring at the fa
age due to the
ere the induce
an be seen fr
large as the

rison between m
re essentially
ffect needs to
rge return stro
ent of the indu
his enhanceme
by corona re
the propagati
various induc
g, 7.5-m high,

m overhead line
onducting. Stro
ted from [61])
ar end of the l
finite ground
ed voltage me
om the figure
induced volta
measured wave
determined b
be taken into
oke current p
uced voltage m
ent can be exp
esults in a de
ion speed mak
ced surges - to
single-condu
e in presence of
oke location: at
ine and not at
d conductivity
easured on an
e, the induced
age computed
eforms and
by the corona
account only
eaks) [61]. In
magnitudes, as
plained, theo-
ecrease of the
kes it possible
o reach larger
ctor overhead
f corona. Solid
50 m from the
t
y
n
d
d
a
y
n
s
-
e
e
r
d
84


The simulations presented in Fig. 15 are obtained assuming a channel-base current with a peak value of 35 kA and a
maximum time derivative (di/dt)max =42 kA/s. The adopted lightning current is large enough to induce a voltage
exceeding the corona threshold at some points along the line, and low enough not to result in a direct strike to the line.
The stroke location is at 50 m from the line center and symmetrical to the line ends. The details of the computational
model can be found in [61]. The solid lines are the voltages calculated taking into account the corona effect, while the
dotted lines are the voltages calculated disregarding corona.

As of today, no experimental results are available to confirm the enhancement effect of corona when distributed sources
are present, although indications of using triggered lightning facilities are proposed in [61].


6 CONCLUSIONS

In this lecture, we discussed the Transmission Line (TL) theory and its application to the problem of lightning
electromagnetic field coupling to transmission lines.
We started with the derivation of the general field-to-transmission line coupling equations for the case of a single-wire
line above a perfectly conducting ground. The derived equations are based on the thin-wire approximation and they do
take into account high frequency radiation effects. It was shown that, under the transmission line approximation, the
general equations reduce to the Agrawal et al. field-to-transmission line coupling equations.
After a short discussion on the underlying assumptions of the TL theory, we described seemingly different but
completely equivalent approaches that have been proposed to describe the coupling of electromagnetic fields to
transmission lines.
The derived equations were then extended to deal with the presence of losses and multiple conductors and expressions
for the line parameters, including the ground impedance and admittance are presented. The time-domain representation
of the field-to-transmission line coupling equations, which allows for a straightforward treatment of nonlinear
phenomena as well as the variation in the line topology, was also described. Solution methods in the frequency domain
and in the time domain were given and application examples with reference to lightning-induced voltages were
presented and discussed.
Specifically, the effect of ground losses was illustrated and discussed. When the travelling voltage and current waves
are originated from lumped excitation sources located at a specific location along a transmission line (direct lightning
strike), both the corona phenomenon and ground losses result in an attenuation and dispersion of propagating surges
along transmission lines. However, when distributed sources representing the action of the electromagnetic field from a
nearby lightning illuminating the line are present, ground losses and corona phenomenon could result in important
enhancement of the induced voltage magnitude.


7 ACKNOWLEDGMENTS

The author expresses his gratitude to C.A. Nucci, M. Rubinstein and S. Tkachenko for their valuable contributions,
comments and suggestions.
This work was carried out within the framework of the European COST Action P18. Financial support from the Swiss
Office for Education and Research SER (Project No. C07.0037) and the Swiss National Science Foundation (Project
No. 200021-122457) are acknowledged.


8 REFERENCES


[1] C. Nucci and F. Rachidi, "Interaction of electromagnetic fields generated by lightning with overhead electrical
networks," in The Lightning Flash, V. Cooray, Ed. London: IEE, 2003, pp. 425-478.
[2] D. M. Le Vine and R. Meneghini, "Simulation of radiation from lightning return strokes: the effects of
tortuosity," Radio Science vol. 13, pp. 801809, 1978.
[3] D. M. Le Vine and R. Meneghini, "Electromagnetic fields radiated from a lightning return stroke: application
of an exact solution to Maxwell's equations," Journal of Geophysical Research, vol. 83, pp. 2377-84, 1978.
[4] A. Sakakibara, "Calculation of induced voltages on overhead lines caused by inclined lightning studies," IEEE
Transactions on Power Delivery, vol. 4, pp. 683693, 1989.
85

[5] S. C. Wu and W. T. Hsiao, "Characterization of induced voltages on overhead power lines caused by lightning
strokes with arbitrary configurations," in International Conference on Systems, Man, and Cybernetics, 1994,
pp. 27062710.
[6] R. Moini, S. H. H. Sadeghi, B. Kordi, and F. Rachidi, "An antenna-theory approach for modeling inclined
lightning return stroke channels," Electric Power Systems Research, vol. 76, pp. 945952, 2006.
[7] V. A. Rakov and M. A. Uman, "Review and evaluation of lightning return stroke models including some
aspects of their application," IEEE Transactions on Electromagnetic Compatibility, vol. 40, pp. 403-26, 1998.
[8] V. A. Rakov and F. Rachidi, "Overview of Recent Progress in Lightning Research and Lightning Protection,"
IEEE Transactions on Electromagnetic Compatibility, vol. 51, pp. 428-442, 2009.
[9] S. Tkachenko, F. Rachidi, and J. B. Nitsch, "High-frequency electromagnetic coupling to transmission lines:
Electrodynamics correction to the TL approximation," in Electromagnetic Field Interaction with Transmission
Lines: From Classical Theory to HF Radiation Effects, F. Rachidi and S. Tkachenko, Eds. Southampton: WIT
Press, 2008, pp. 123-158.
[10] A. K. Agrawal, H. J. Price, and S. H. Gurbaxani, "Transient response of multiconductor transmission lines
excited by a nonuniform electromagnetic field," IEEE Transactions on Electromagnetic Compatibility, vol. 22,
pp. 119-29, May 1980.
[11] C. D. Taylor, R. S. Satterwhite, and C. W. Harrison, "The response of a treminated two-wire transmission line
excited by a nonuniform electromagnetic field," IEEE Transactions on Antennas and Propagation, vol. AP-13,
pp. 987-989, 1965 1965.
[12] F. Rachidi, "Formulation of the field-to-transmission line coupling equations in terms of magnetic excitation
fields," IEEE Transactions on Electromagnetic Compatibility, vol. 35, pp. 404-407, 1993.
[13] F. Rachidi and S. Tkachenko, "Electromagnetic Field Interaction with Transmission Lines: From Classical
Theory to HF Radiation Effects," Southampton: WIT Press, 2008.
[14] S. Tkachenko and F. Rachidi, "Electromagnetic field coupling to a line of finite length: theory and fast iterative
solutions in frequency and time domains," IEEE Transactions on Electromagnetic Compatibility, vol. 37, pp.
509-518, 1995.
[15] S. Tkachenko, F. Rachidi, and M. Ianoz, "High-frequency electromagnetic field coupling to long terminated
lines," IEEE Transactions on Electromagnetic Compatibility, vol. 43, pp. 117-129 2001.
[16] H. Hasse, T. Steinmetz, and J. B. Nitsch, "New Propagation Models for Electromagnetic Waves Along
Uniform and Nonuniform Cables," IEEE Transactions on Electromagnetic Compatibility, vol. 46, pp. 345-
352, 2004.
[17] J. B. Nitsch and S. Tkachenko, "Complex-Valued Transmission-Line Parameters and Their Relation to the
Radiation Resistance," IEEE Transactions on Electromagnetic Compatibility, vol. 46, pp. 477-488, 2004.
[18] A. G. Chiariello, A. Maffucci, G. Miano, F. Villone, and W. Zamboni, "A Transmission-Line Model for Full-
Wave Analysis of Mixed-Mode Propagation," IEEE Transactions on Advances Packaging, vol. 31, pp. 275-
284, 2008.
[19] C. R. Paul, Analysis of multiconductor transmission lines. New York: John Wiley and Sons, 1994.
[20] V. Cooray, The Lightning Flash: IEE, 2003.
[21] F. M. Tesche, M. Ianoz, and T. Karlsson, EMC Ananlysis methods and computational models. New York:
Wiley Interscience, 1997.
[22] A. Vukicevic, F. Rachidi, M. Rubinstein, and S. Tkachenko, "On the Evaluation of Antenna-Mode Currents
along Transmission Lines," IEEE Transactions on Electromagnetic Compatibility, vol. 48, pp. 693-700, 2006.
[23] C. A. Nucci, F. Rachidi, and A. Rubinstein, "Derivation of telegrapher's equations and field-to-transmission
line interaction," in Electromagnetic Field Interaction with Transmission Lines: From Classical Theory to HF
Radiation Effects, F. Rachidi and S. Tkachenko, Eds. Southampton: WIT Press, 2008.
[24] C. A. Nucci and F. Rachidi, "Interaction of electromagnetic fields generated by lightning with overhead
electrical networks," in The Lightning Flash: IEE, 2003, pp. 425-478.
[25] M. Paolone, F. Rachidi, A. Borghetti, C. A. Nucci, M. Rubinstein, V. A. Rakov, and M. A. Uman, "Lightning
Electromagnetic Field Coupling to Overhead Lines: Theory, Numerical Simulations and Experimental
Validation," IEEE Transactions on Electromagnetic Compatibility, in press, vol. 51, pp. 532-547, 2009.
[26] C. A. Nucci and F. Rachidi, "On the contribution of the electromagnetic field components in field-to-
transmission lines interaction," IEEE Transactions on Electromagnetic Compatibility, vol. 37, pp. 505-508,
1995.
[27] S. Rusck, "Induced lightning overvoltages on power transmission lines with special reference to the
overvoltage protection of low voltage networks," Transactions of the Royal Institute of Technology, Stockholm,
vol. 120, 1958.
[28] C. A. Nucci, F. Rachidi, M. Ianoz, and C. Mazzetti, "Comparison of two coupling models for lightning-
induced overvoltage calculations," IEEE Transactions on Power Delivery, vol. 10, pp. 330-338, 1995.
86

[29] V. Cooray, "Calculating lightning-induced overvoltages in power lines: a comparison of two coupling models,"
IEEE Transactions on Electromagnetic Compatibility, vol. 36, pp. 179-182, 1994.
[30] A. Piantini and J. M. Janiszewski, "Induced voltages on distribution lines due to lightning discharges on nearby
metallic structures," IEEE Transactions on Magnetics, vol. 34, pp. 2799 - 2802 1998.
[31] A. Piantini and J. M. Janiszewski, "Lightning-Induced Voltages on Overhead Lines - Application of the
Extended Rusck Model," IEEE Transactions on Electromagnetic Compatibility, vol. 51, pp. 548-558, 2009.
[32] F. Rachidi, C. A. Nucci, I. M., and C. Mazzetti, "Influence of a lossy ground on lightning-induced voltages on
overhead lines," IEEE Transactions on Electromagnetic Compatibility, vol. 38, pp. 250-263, 1996.
[33] S. Ramo, J. R. Whinnery, and T. van Duzer, Fields and waves in communication electronics, 3 ed. New York:
Wiley, 1994.
[34] F. Rachidi, C. A. Nucci, M. Ianoz, and C. Mazzetti, "Importance of losses in the determination of lightning-
induced voltages on overhead lines," EMC '96 ROMA. International Symposium on Electromagnetic
Compatibility. Univ. Rome `La Sapienza', Rome, Italy, vol. 2, 1996.
[35] F. M. Tesche, "Comparison of the transmission line and scattering models for computing the HEMP response
of overhead cables," IEEE Transactions on Electromagnetic Compatibility, vol. 34, May 1992 1992.
[36] E. D. Sunde, Earth conduction effects in transmission systems. New York: Dover Publication, 1968.
[37] E. Petrache, F. Rachidi, M. Paolone, C. Nucci, V. A. Rakov, and M. A. Uman, "Lightning-Induced Voltages
on Buried Cables, Part I: Theory," IEEE Transactions on Electromagnetic Compatibility, vol. 47,, p. August
2005, 2005.
[38] F. Rachidi, C. A. Nucci, and M. Ianoz, "Transient analysis of multiconductor lines above a lossy ground,"
IEEE Trans. on PWDR, vol. 14, pp. 294-302, 1999.
[39] C. A. Nucci, V. Bardazzi, R. Iorio, A. Mansoldo, and A. Porrino, "A code for the calculation of lightning-
induced overvoltages and its interface with the Electromagnetic Transient program," in 22nd International
Conference on Lightning Protection (ICLP), Budapest, Hungary, 1994.
[40] D. Orzan, P. Baraton, M. Ianoz, and F. Rachidi, "Comparaison entre deux approches pour traiter le couplage
entre un champ EM et des rseaux de lignes," in 8me Colloque International sur la Compatibilit
Electromagntique, Lille, France, 1996.
[41] D. Orzan, "Couplage externe et interne entre un champ lectromagntique et un rseau de lignes multifilaires
(Ph.D. thesis)," Lausanne, Switzerland: Ecole Polytechnique Federale de Lausanne, 1998.
[42] H. K. Hoidalen, "Lightning-induced overvoltages in low-voltage systems (Ph.D. Thesis)," Norwegian
University of Science and Technology, 1997.
[43] H. K. Hoidalen, "Calculation of lightning-induced overvoltages using MODELS," International Conference on
Power Systems Transients. Tech. Univ. Budapest, Budapest, Hungary, 1999.
[44] H. K. Hoidalen, "Analytical formulation of lightning-induced voltages on multiconductor overhead lines above
lossy ground," IEEE Transactions on Electromagnetic Compatibility, vol. 45, pp. 92-100, 2003.
[45] H. K. Hoidalen, "Calculation of lightning-induced voltages in MODELS including lossy ground effects," in
International Conference on Power System Transient IPST 2003, New Orleans, USA, 2003.
[46] E. Perez, A. Delgadillo, D. Urrutia, and H. Torres, "Optimizing the Surge Arresters Location for Improving
Lightning Induced Voltage Performance of Distribution Network," in IEEE PES General Meeting, 2007.
[47] A. Borghetti, A. Gutierrez, C. A. Nucci, M. Paolone, E. Petrache, and F. Rachidi, "Lightning-induced voltages
on complex distribution systems: models, advanced software tools and experimental validation," Journal of
Electrostatics, vol. 60, pp. 163-174, 2004.
[48] F. Napolitano, A. Borghetti, C. A. Nucci, M. Paolone, F. Rachidi, and J. Mahserejian, "An advanced interface
between the liov code and the EMTP-RV," in 29th International Conference on Lightning Protection (ICLP),
Uppsala, Sweden, 2008.
[49] R. Montano, N. Theethayi, and V. Cooray, "An efficient implementation of the Agrawal et al. model for
lightning-induced voltage calculations using circuit simulation software," IEEE Transactions on Circuits and
Systems-I: Regular Papers, vol. 55, pp. 2959-2965, 2008.
[50] F. Rachidi, S. Loyka, C. A. Nucci, and M. Ianoz, "A New Expression For the Ground Transient Resistance
Matrix Elements of Multiconductor Overhead Transmission Lines," Electric Power System Research Journal,
vol. 65, pp. 41-46, 2003.
[51] R. Araneo and S. Cellozi, "Direct time domain analysis of transmission lines above a lossy ground," IEE Proc.
Science & Measurement Technology, vol. 148, pp. 73-79, 2001.
[52] N. Theethayi and R. Thottappillil, "Surge propagation and crosstalk in multiconductor transmission lines above
ground," in Electromagnetic Field Interaction with Transmission Lines. From Classical Theory to HF
Radiation Effects, F. Rachidi and S. Tkachenko, Eds.: WIT Press, 2008.
[53] A. Tafflove, Computational Electrodynamics: The Finite Difference Time Domain Method: Artech House,
1995.
87

[54] M. Paolone, C. A. Nucci, and F. Rachidi, "A New Finite Difference Time Domain Scheme for the Evaluation
of Lightning Induced Overvoltages on Multiconductor Overhead Lines," in 5th Int. Conf. on Power System
Transients, Rio de Janeiro, 2001.
[55] P. D. Lax and B. Wendroff, "System of conservations laws," Comm. Pure Apl. Math., vol. 13, pp. 217-237,
1960.
[56] S. R. Omick and S. P. Castillo, "A New Finite Difference Time-Domain Algorithm for the Accurate Modeling
of Wide-Band Electromagnetic Phenomena," IEEE Transactions on Electromagnetic Compatibility, vol. 35,
pp. 215-222, 1993.
[57] M. Paolone, "Modeling of lightning-induced voltages on distribution networks for the solution of power
quality problems, and relevant implementation in a transient program," in Department of Electrical
Engineering Bologna, Italy: University of Bologna, 2001, p. 117.
[58] M. Ishii, K. Michishita, Y. Hongo, and S. Ogume, "Lightning-induced voltage on an overhead wire dependent
on ground conductivity," IEEE Transactions on Power Delivery, vol. 9, pp. 109-118, 1994.
[59] V. F. Hermosillo and V. Cooray, "Calculation of fault rates of overhead power distribution lines due to
lightning induced voltages including the effect of ground conductivity," IEEE Trans on Electromagnetic
Compatibility, vol. 37, pp. 392-399, August 1995 1995.
[60] S. Guerrieri, C. A. Nucci, and F. Rachidi, "Influence of the Ground Resistivity on the Polarity and Intensity of
Lightning Induced Voltages," in 10th International Symposium on High Voltage Engineering, Montreal,
Canada, 1997.
[61] C. A. Nucci, S. Guerrieri, M. T. Correia de Barros, and F. Rachidi, "Influence of Corona on the Voltages
Induced by Nearby Lightning on Overhead Distribution Lines," IEEE Trans. on Power Delivery, vol. 15, pp.
1265-1273, 2000.
[62] F. Rachidi, C. A. Nucci, S. Guerrieri, and M. T. Correia de Barros, "On the Amplitude Enhancement of
Voltages Induced by External EM Fields on Transmission Lines due to Ground Losses and Corona
Phenomenon," in IEEE International Symposium on Electromagnteic Compatibility, Montreal, Canada, 2001.
[63] "CIGRE-CIRED JWG C4.4.02: Protection of MV and LV Networks against Lightning. Part I: Common
Topics," CIGRE Technical Brochure No 287, 2006.
[64] H. W. Dommel, "Electromagnetic Transient program Reference Manual (EMTP Theory Book)," Bonneville
Power Administration, Portland, OR August 1986 1986.
[65] F. De la Rosa, R. Valdivia, H. Prez, and J. Loza, "Discussion about the inducing effects of lightning in an
experimental power distribution line in Mexico," IEEE Transactions on Power Delivery, vol. 3, 1988.
[66] V. Cooray, "Lightning-induced overvoltages in power lines: validity of various approximations made in
overvoltage calculations," in 22nd International Conference on Lightning Protection, Budapest, Hungary,
1994.
[67] M. Rubinstein, "An approximate formula for the calculation of the horizontal electric field from lightning at
close, intermediate, and long range," IEEE Transactions on Electromagnetic Compatibility, vol. 38, pp. 531-5,
1996.
[68] M. Ishii, K. Michishita, and Y. Hongo, "Verification of coupling model for calculation of induced lightning
voltages," in CIGRE International Colloquium on Insulation Coordination, Toronto, Canada, 1997.


88

You might also like