(Eduard M. Bazelyan, Yuri P. Raizer) Lightning
(Eduard M. Bazelyan, Yuri P. Raizer) Lightning
(Eduard M. Bazelyan, Yuri P. Raizer) Lightning
Lightning Protection
E M Bazelyan
and
Yu P Raker
IOP
Institute of Physics Publishing
Bristol and Philadelphia
A catalogue record for this book is available from the British Library.
Preface ix
1 Introduction: lightning, its destructive effects and protection 1
1.1 Types of lightning discharge 1
1.2 Lightning discharge components 5
1.3 Basic stages of a lightning spark 6
1.4 Continuous and stepwise leaders 9
1.5 Lightning stroke frequency 11
1.5.1 Strokes at terrestrial objects 11
1.5.2 Human hazard 11
1.6 Lightning hazards 12
1.6.1 A direct lightning stroke 12
1.6.2 Induced overvoltage 17
1.6.3 Electrostatic induction 18
1.6.4 High potential infection 19
1.6.5 Current inrush from a spark creeping along the
earth’s surface 20
1.6.6 Are lightning protectors reliable? 21
1.7 Lightning as a power supply 23
1.8 To those intending to read on 24
References 26
Vlll
Today, we know sufficiently much about lightning to feel free from the mystic
fears of primitive people. We have learned to create protection technologies
and to make power transmission lines, skyscrapers, ships, aircraft, and space-
craft less vulnerable to lightning. Yes, the danger is getting less but it still
exists! With every step of the technical progress, lightning arms itself with
a new weapon to continue the war by its own rules against the self-confident
engineer. As we improve our machines and stuff them with electronics in an
attempt to replace human beings, lightning acts in an ever refined manner. It
takes us by surprise where we do not expect it, making us feel helpless again
for some time.
We do not intend to present in this book a set of universal lightning
protection rules. Such a task would be as futile as advertising a universal anti-
biotic lethal to every harmful microbe. The world is changeable, and today’s
panacea often becomes a useless pill even before the advertising sheet fades.
Technical progress has so far failed to take lightning unawares. Improvement
and miniaturization of devices increase our concern about the refined
destructive behaviour of lightning, but no prophet is able to foresee all of
its destructive effects.
We d o not plan to discuss in detail all available information on light-
ning. There are already some excellent books providing all sort of reference
data, among them the two volumes of Lightning edited by R H Golde and
Lightning Discharge by M Uman. Our aim is different. We think it important
to give the reader some clear, up-to-date physical concepts of lightning
development, which cannot be found in the books referred to. These will
serve as a basis for the researcher and engineer to judge the properties of
this tremendous gas discharge phenomenon. Then we shall discuss the
nature of various hazardous manifestations of lightning, focusing on the
physical mechanisms of interaction between lightning and an affected
construction. The results of this consideration will further be used to estimate
ix
If you want to observe lightning, the best thing to do is to visit a special light-
ning laboratory. Such laboratories exist in all parts of the globe except the
Antarctic. But you can save on the travel if you just climb onto the roof of
your own house to give a good field of vision. Better, fetch your camera.
Even an ordinary picture can show details the unaided human eye often
misses. You might as well sit back in your favourite armchair, having
pulled it up to a window, preferably one overlooking an open space. The
camera can be fixed on the window sill, There is nothing else to do but
wait for a stormy night.
There is enough time for the preparations to be made because the storm
will be approaching slowly. At first, the air will grow still, and it will get
much darker than it normally is on a summer night. The cloud is not yet visible,
but its approach can be anticipated from the soundless flashes at the horizon.
They gradually pull closer, and the brightest of them can already be heard as
delayed and yet amiable roaring. T h s may go on for a long time. It may
seem that the cloud has stopped still or turned away, but suddenly the sky is
ripped open by a fire blade. This is accompanied by a deafening crash, quite
different from a cannon shot because it takes a much longer time. The first
lightning discharge is followed by many others falling out of the ripped
cloud. Some strike the ground while others keep on crossing the sky, competing
with the first discharge in beauty and spark length. This is the right time to start
observations: remove the camera shutter and try to take a few pictures.
no other bright light source should be present within the vision field of the
camera lens. The film can then be exposed for many minutes until a spark
finds its way into the frame. After this, the lens should be closed with the shut-
ter and the camera should be set ready for another shot. Experience has shown
that at least one third of pictures taken during a good night thunderstorm
prove successful.
All lightning discharges can be classified, even without photography,
into two groups - intercloud discharges and ground strikes. The frequency
of the former is two or three times higher than that of the latter. An inter-
cloud spark is never a straight line, but rather has numerous bends and
branchings. Normally, the spark channel is as long as several kilometres,
sometimes dozens of kilometres.
The length of a lightning spark that strikes the ground can be defined
more exactly. The average cloud altitude in Europe is close to three kilo-
metres. Spark channels have about the same average length. Of course,
this parameter is statistically variable, because a discharge from a charged
cloud centre may start at any altitude up to 10 km and because of a large
number of spark bends. The latter are observable even with the unaided
eye. In a photograph, they may look strikingly fanciful (figure 1.1). A photo-
graph can show another important feature inaccessible to the naked eye - the
main bright spark reaching the ground has numerous branches which have
stopped their development at various altitudes. A single branch may have
a length comparable with that of the principal spark channel (figure 1.2).
Branches can be conveniently used to define the direction of lightning
propagation. Like a tree, a lightning spark branches in the direction of
storm cloud [4]. When the probe rises to 200-300m above the earth, an
ascending spark is induced from it. A discharge artificially induced in the
atmosphere is often referred to as triggered lightning. To raise the chances
for a successful experiment, the electric field induced by the storm charges at
the ground surface are measured prior to the launch. The probe is triggered
when the field strength becomes close to 200 V/cm, which provides spark
ignition in 60-70% of launches [5].
The value 200 V/cm is two orders of magnitude smaller than the thresh-
old value of E = 30 kV/cm, at which a short air gap with a uniform field is
broken down under normal atmospheric conditions. Clearly, no spark
ignition would be possible without the local field enhancement by electric
charges induced on the probe and the wire. Below, we shall discuss the
triggered discharge mechanism in more detail.
A field detector on the Earth’s surface (it might as well be placed on the
window of your own room) can easily determine the polarity of the charge
transported by a lightning spark to the ground. The polarity of the spark
is defined by that of the charge. About 90% of descending sparks occurring
in Europe during summer storms carry a negative charge, so these are known
as negative descending sparks. The other descending sparks are positive. The
even at a relatively low voltage. Imagine now that the needle is falling down
on to the earth, pulling the wire behind it. The strong field region, in which
the air molecules become ionized, will move down together with the needle.
A lightning spark has no wire at its disposal. The function of a conductor
connecting the leader tip to the starting point of the discharge is performed by
the leader plasma channel. It takes a fairly long time for a leader to develop -
up to 0.01 s, which is eternity in the time scale of fast processes involving an
electric impulse discharge. During this period of time, the leader plasma
must be maintained highly conductive, and this may become possible only if
the gas is heated up to an electric arc temperature, i.e. above 5000-6000K.
The problem of the channel energy balance necessary for the heating and com-
pensation for losses is a key one in leader theory. It is discussed in chapters 2
and 4, as applied to various kinds of lightning discharge.
A leader is an indispensable element of any spark. The initial and all
subsequent components of a flash begin with a leader process. Although its
mechanism may vary with the spark polarization, propagation direction
and the serial number of the component, the process remains essentially
the same. This is the formation of a highly conductive plasma channel due
to the local enhancement of the electric field in the leader tip region.
A return stroke is produced at the moment of contact of a leader with
the ground or a grounded object. Most often, this is an indirect contact: a
counterpropagating leader, commonly termed a counterleader, may start
from an object to meet the first leader channel. The moment of their contact
initiates a return stroke. During the travel from the cloud to the ground, the
lightning leader tip carries a high potential comparable with that of the cloud
at the spark start, the potential difference being equal to the voltage drop in
the leader channel. After the contact, the tip receives the ground potential
and its charge flows down to the earth. The same thing happens with the
other parts of the channel possessing a high potential. This ‘unloading’ pro-
cess occurs via a charge neutralization wave propagating from the earth up
through the channel. The wave velocity is comparable with the velocity of
light and is about 10’ mjs. A high current flows along the channel from the
wave front towards the earth, carrying away the charge of the unloading
channel sites. The current amplitude depends on the initial potential distribu-
tion along the channel and is, on average, about 30 kA, reaching 200-250 kA
for powerful lightning sparks. The transport of such a high current is accom-
panied by an intense energy release. Due to this, the channel gas is rapidly
heated and begins to expand, producing a shock wave. A peal of thunder
is one of its manifestations.
The return stroke is the most powerful stage of a lightning discharge
characterized by a fast current change. The current rise can exceed
10’ A/s, producing a powerful electromagnetic radiation affecting the
performance of radio and TV sets. This effect is still appreciable at a distance
of several dozens of kilometres from the lightning discharge.
This introductory chapter contains no theory, and this makes the discussion
of leader details a very complicated task. So we shall mention only its
principal features which can be registered by a continuously moving film.
Continuous streak photographs show lightning development in time. One
needs, however, a certain skill and experience to be able to interpret them
adequately. Suppose a small light source moves perpendicularly to the
earth at a constant velocity. It may be a luminant bomb descending with a
parachute. A film moving horizontally, i.e. in the transverse direction, at a
constant speed will show a sloping line (figure lS(u)). Given the film speed
(the display rate), one can easily calculate the light source velocity from
the line slope. A uniformly propagating vertical channel will leave on a
film a sloping wedge (figure 1.5(b)) rather than a line. From its slope, too,
one can find the channel velocity, or its propagation rate. The higher the
rate of the process in question, the higher must be the display rate in
streak photography. The highest display rates can be obtained using an
electron-optical converter, in which an image is converted to an electron
beam scanned across the screen by an electric field. A conventional photo-
camera registers the displayed electronic image from the screen onto an
immobile film. Electron-optical converters have provided much information
on long sparks, but their application in lightning observations has been
limited. The main results here have been obtained using mechanical streak
cameras. We described this technique and analysed streak pictures in our
book on long sparks [8].
Figure 1.6(a) shows the leader of an ascending lightning spark going
up from the top of a grounded tower in the electric field of a negatively
charged cloud cell. The leader carries a positive space charge and, therefore,
it should be referred to as a positive leader. One can clearly see the bright
trace of the channel tip, which looks like a nearly continuous line. This
kind of leader is known in literature as a continuous leader. The changing
trace slope suggests that the leader velocity changes during its propagation.
These changes are, however, quite smooth, not interrupting the tip travel up
to the cloud.
0 t o t
a b
Figure 1.6. A schematic streak picture of a positive ascending (a) and a negative
descending (b) lightning leader.
a distance r 6 Re, are attracted by it, the others missing the object. This
primitive pattern of lightning attraction generally leads to a correct result.
For estimations, one can use Re! RZ 3h and borrow the stroke frequency
per unit unperturbed area per unit time, nl, from meteorological observa-
tions. The latter are used to make up lightning intensity charts. For example,
the lightning intensity in Europe is nl < 1 per 1 km2 per year for the tundra,
2-5 for flat areas, and up to 10 for some mountainous regions such as the
Caucasus. A tower of h = lOOm is characterized by Re, = 0.3 km with
NI = n17rR&M 1 stroke per year at the average value of nl = 3.5 kmP2year-’.
This estimation is meaningful for a flat country and only for not very high
objects, h < 150m, which do not generate ascending lightnings.
is as small as 5-6m and the attraction area is less than lop4 km2. If a man
had stopped alone in the middle of a large field two thousand years ago,
he might have expected to attract a direct lightning stroke only by the end
of the third, coming millennium. In actual reality, however, the number of
lightning victims is large, and direct strokes have nothing to do with this.
It is known from experience that one should not stay in a forest or hide
under a high tree in an open space during a thunderstorm. A tree is about
10 times higher than a man, and a lightning strikes it 100 times more fre-
quently. When under the tree crown, a man has a real chance to be within
the zone of the lightning current spread, which is hazardous.
After a lightning strikes the tree top, its current ZM runs down along its
stem and roots to spread through the soil. The root network acts as a natural
grounding electrode. The current induces in the soil an electric field E = pj,
where p is the soil resistivity and j is the current density. Suppose the current
spreads through the soil strictly symmetrically. Then the equipotentials will
represent hemispheres with the diagonal plane on the earth’s surface. The
current density at distance r from the tree stem is j = IM/(27rr2)the field is
IMp/(27rr2)and the potential difference between close points r and r + Ar
+
is equal to A U = (ZMp/27r)[r-’ - (Y AY)-’] x E(r)Ar.If a person is stand-
ing, with his side to the tree, at distance r = 1 m from the tree stem centre and
the distance between his feet is Ar x 0.3 m, the voltage difference on the soil
with resistivity p = 200 f2/m will be A U x 220 kV for a moderate lightning of
ZM = 30 kA. This voltage is applied to the shoe soles and, after a nearly inevi-
table and fast breakdown, to the person’s body. There is no doubt that the
person will suffer or, more likely, will be killed - the applied voltage is too
high. Note that this voltage is proportional to Ar. This means that it is
more dangerous to stand with one’s feet widely apart than with one’s feet
pressed tightly together. It is still more dangerous to lie down along the
radius from the tree, because the distance between the extreme points
contacting the soil becomes equal to the person’s height. It would be much
safer to stand still on one foot, like a stork. But it is, of course, easier to give
advice than to follow it. Incidentally, lightning strikes large animals more
frequently than humans, also because the distance between their limbs is larger.
If you have a cottage equipped with a lightning protector, take care that
no people could approach the grounding rod during a thunderstorm. The
situation here is similar to the one just described.
Figure 1.7. Traces of lightning strokes at the steel tip of the Ostankino Television
Tower in Moscow.
The lightning current will flow down the metallic tower to the ground
electrodes to be dissipated in the earth. Let us take point A at the height
of the insulator string connection. Due to the lightning current i(t), the
potential at this point, pA,will differ from the zero potential of the earth
by the voltage drop in the grounding resistance R, and in the tower induc-
tance L, between the tower base up to the point A:
di
p = R,i+ L,-
dt
However, the power wire potential will practically remain the same (in this
qualitative description, we ignore all inductances between the power wires,
tower and grounded wire). The power wire potential is due to the operating
voltage source of the power line: qw = Uop.Then, the insulator string voltage
will be
di
U = pw - (FA = Uop - R,i - Ls-,
dt
Note that the lightning current and operating voltage may have different
polarities. As a result, the overvoltage U may prove to be the sum of the
three terms in equation (1.2).
The inductance component of the overvoltage, L,di/dt, has a short
lifetime: it acts about as long as the lightning current rises. For a current
impulse with an average amplitude IM 30kA and an average rise time
tf = 5 p , the inductance voltage at L, 50pH will be about 300kV. The
resistance component U , at a typical grounding resistance R, = 10R will
have about the same value but will act an order of magnitude longer, i.e.,
as long as the lightning current flows. For this reason, this component
makes the principal contribution to the insulation flashover.
The emergency situation just described is not as bad as a direct stroke at
a power wire when the same lightning current can induce an order of magni-
tude higher voltage. The insulation of a ultrahigh voltage line can withstand
short overvoltages up to 1000- 1500kV and seldom suffers from lightning
strokes at a tower or a lightning protection wire. To produce a harmful
effect, the lightning current must be 3-5 times the average value. Lightning
strokes of this power do not occur frequently, making up less than 1% of
all strokes. Quite different is the effect of a direct stroke at a power network
with an operating voltage of 35 kV and lower. The insulation system will
suffer equally from a stroke at a power wire or a tower. It is no use protecting
such a line with grounded wire.
Insulation flashover due to the tower potential rise is referred to as
reverse flashover. This name does not imply the definite direction of the
discharge development but only indicates the direction from which the
potential rises, i.e., the grounded end of the insulator string rather than
the power wire.
When the length of the current conductor inducing the magnetic field is much
longer than the distance to the circuit, rc, and when the width of the circuit
normal to the magnetic field is much smaller than rc, we have
poS di
EemfM --
27rrc dt
where po = 47r x lop7Him is vacuum magnetic permeability. The order of
magnitude of the induction emf amplitude is defined as
where A,, is the maximum rate of the current impulse rise equal to 10" A/s
for the subsequent components of a powerful lightning flash. A circuit of area
S = 1 m2 located at a distance r, = 10 m from a lightning current conductor
may become the site of induced overvoltage with an amplitude up to 20 kV.
This value is only an arbitrary guideline, because induced overvoltage may
vary with the circuit area, its orientation and distance from the lightning cur-
rent. Circuits with an area of hundreds and thousands of square metres may
be created by large industrial metallic constructions and power transmission
lines. The distance between the circuit and the current flow may also vary
greatly. For such diverse parameters of a system, the problem will be more com-
plicated in the case of fast current variations along the spark and in time. It
Figure 1.9. Schematic input of high potential from remote lightning strokes.
however, becomes loaded by additional current, which finds its way there
from a remote lightning stroke, using the service line as a conductor.
When the earth connection resistance is low and the service line goes through
the ground with a high resistivity so that the current leakage through the side
surface is not large, nearly all of the lightning current arrives at the con-
nection from the stroke site. This situation appears to be somewhat similar
to a direct lightning stroke. Sometimes, special measures must be taken to
restrict the infection current. A detailed treatment of the problem of current
and potential infections will be offered in chapter 6.
1.6.5 Current inrush from a spark creeping along the earth’s surface
This phenomenon is familiar to all communications men who have to repair
communications cables damaged by lightning. The damaged site can be
detected easily, because it is indicated by a furrow in the ground extending
far away from the stroke site. A furrow may be as long as several dozens
of metres, or 100-200m in a high resistivity ground. Such a long gap can
be bridged by a spark because of the electric field created by the spark current
spreading out through the ground. The mechanism of spark formation along
a conducting surface differs from that of a ‘classical’ leader propagating
through air. A creeping spark can develop in low fields and have a very
high velocity.
Underground cables are not the only objects suffering from creeping
spark current. Similarly, it can find its way to underground service lines
and to the earth connections of constructions well equipped by lightning
protectors. But a protector palisade cannot stop lightning. When the conven-
tional way from the earth surface is blocked, it breaks through from beneath,
making a bypass in the ground. Lightning thus behaves very much like a
clever general in ancient times, who ordered his soldiers to make a secret
underground passageway instead of attacking openly the impregnable
castle walls. It is reasonable to suggest that the contact of a creeping spark
with combustible materials is as frequent a cause of a fire as a direct lightning
stroke.
The details of the creeping discharge mechanism have been unknown
until quite recently. They are analysed in chapter 6.
Figure 1.10. This lightning has missed the teletower tip by over 200m.
Figure 1.11. Distributions of breakdown voltages in air gaps of various lengths with a
sharply non-uniform electric field.
gaps have a common high voltage electrode (the leader that has descended to
a certain altitude), while the zero potential electrodes are formed by the
earth’s surface with grounded objects and protectors distributed on it. The
problem of protector effectiveness thus reduces to the calculation of
breakdown probabilities for the elementary gaps in a multielectrode
system. The general formulation of this problem is very complex, since the
spark development in the elementary gap in real conditions cannot be
taken to be independent. The discharge processes affect one another by
redistributing their electric fields, which eliminates straightforward use of
statistical relations describing independent processes.
We cannot say that the spark discharge theory for a multielectrode
system has been brought to any stage of completion. But what has been
done, theoretically and experimentally, allows the formulation of certain
concepts of the lightning orientation mechanism and the development of
engineering approaches to estimate the effectiveness of protectors of various
heights (see chapter 5).
Investigation of multielectrode systems is also important from another
point of view: we must find ways of affecting lightning actively. It would
be reasonable to leave the discussion of this issue for specialized chapters
of this book, but they will, however, attract the attention of professionals
only, or of those intending to become professionals. It is not professionals
but amateurs who, most often, try to invent lightning protectors. They
have at their disposal a complete set of up-to-date means: lasers, plasma
jets, corona-forming electrodes for cloud charge exchange, radioactive
There will be no more popularized stories about lightning in this book. Nor
shall we mention ball lightning here. The next chapter will contain a
thorough analysis of available data and theoretical treatments of the long
spark, because we believe that without these preliminaries the lightning
mechanism may not become clear to the reader. Nature has eagerly employed
standard solutions to its problems, so lightning is quite likely to represent the
limiting case of the long spark. It would be useful for readers to familiarize
themselves with our previous book S p a r k Discharge, because it is totally
concerned with this phenomenon. But even without it, they will be able to
find here basic information on long sparks. We have tried to describe their
general mechanisms and to give predictions as to their extension to air
gaps of extrema1 length. Even for this reason alone, the next chapter is not
a summary of the previous book. Lightning is as complicated a phenomenon
as the long spark and is definitely more diverse. It is a multicomponent
process. Since its subsequent components sometimes take the path of an
earlier component, we must consider the effects of temperature and residual
conductivity in the spark channel on the behaviour of new ionization
waves.
References
[l] Bazelyan E M, Gorin B N and Levitov V I 1978 Physical and Engineering Funda-
mentals of Lightning Protection (Leningrad: Gidrometeoizdat) p 223 (in Russian)
[2] McEachron K 1938 Electr. Engin. 57 493
[3] Berger K and Vogrlsanger E 1966 Bull. SEV 57 No 13 1
[4] Newman M M, Stahmann J R, Robb J D, Lewis E A et a1 1967 J . Geophys. Res.
72 4761
[5] Uman M A 1987 The Lightning Discharge (New York: Academic Press) p 377
[6] Berger K, Anderson R B and Kroninger H 1975 Electra 41 23
[7] Schonland B, Malan D and Collens H 1935 Proc. Roy. Soc. London Ser A 152 595
[8] Bazelyan E M and Raizer Yu P 1997 Spark Discharge (Boca Raton: CRC Press)
p 294
191 Schonland B 1956 The Lightning Discharge. Handbuch der Physik 22 (Berlin:
Springer) p 576
This chapter will deal with the spark discharge in a long air gap. We have
already mentioned in chapter 1 that this material should not be ignored by
the reader. But for the long spark, specialists would know much less about
lightning. Today, high voltage laboratories are able to produce and study
long sparks of several tens and even hundreds of metres long [l-31. Many
of the long spark parameters and properties lie close to the lower boundary
of respective lightning values. Most effects observable in a lightning
discharge were, sooner or later, reproduced in the laboratory. One exception
is a multicomponent discharge, but the obstacles lie in the technology rather
than in the nature of the phenomenon. It would be very costly to instal and
synchronize several high voltage power generators, making them discharge
consecutively into the same air gap.
As for the fine structure of gas-discharge elements, long spark research-
ers are far ahead of lightning observers. This could not be otherwise, since a
laboratory discharge can be reproduced as often as necessary, by starting the
generator at the right moment, within a microsecond fraction accuracy, and
strictly timing the switching of all fast response detectors. But with lightning,
the situation is different. It strikes every square kilometre of the earth’s sur-
face in Europe approximately 2 to 4 times a year. So, even such a high con-
struction as the Ostankino Television Tower (540m) is struck by lightning
only 25-30 times a year. Of these, only 2-3 discharges are descending,
while the others go up to a cloud. Normally, lightning observations have
to be made from afar, so that many details of the process are lost. The
gaps in the study of its fine structure must, of necessity, be filled in laboratory
conditions.
The long spark theory is far from being completed, and there is no
adequate computer model of the process. Still, there has lately been some
progress, primarily owing to laboratory investigations. It would be unwise
to discard these data and not to try to use them for the description of
21
The key point is how a spark channel develops in a weak electric field, by 1-2
orders of magnitude lower than what is necessary to increase the electron
density in air (Ei x 30 kV/cm under normal conditions). Naturally, we
speak of a discharge in a sharply non-uniform field. Near an electrode
with a small curvature radius (suppose this is a spherical anode of radius
r, x 1-lOcm), the field is Ea(ra) Ea > E, at the voltage U x 50-500kV.
This is the site of initiation of a discharge channel. At a distance r = lor,
from the electrode centre, the channel tip enters the outer gap region,
where the initial value of E = Ea(r,/r)* is one hundredth of that on the
electrode. This weak field is incapable of supporting ionization. Nevertheless,
the channel moves on, changing the neutral gas to a well-ionized plasma.
There is no other reasonable explanation of this fact except for a local
enhancement of the electric field at the tip of the developing channel. The
enhancement is due to the action of the channel’s own charge. Indeed, a con-
ductive channel having a contact with the anode tends to be charged as much
as its potential U , relative to the grounded cathode. Current arises in the
channel, which transports the positive electric charge from the anode (more
exactly, from the high voltage source, to which the anode is connected). (In
reality, electrons moving through the channel toward the anode expose low
mobility positive ions.) Such would be exactly the mechanism of charging a
metallic rod if it could be pulled out of the anode like a telescopic antenna.
Then the strongest field region would move through the gap together with
the rod tip. We can say that a strong electric field wave is propagating through
a gap, in which ionization occurs and produces a new portion of the plasma
channel. We can also name it as an ionization wave, and t h s term is commonly
used.
The wave mechanism of spark formation was suggested as far back as
the 1930s by L Loeb, J Meek, and H Raether. The channel thus formed
was termed a streamer (figure 2.1). Experiments showed that the streamer
velocity could be as high as 107m/s. In lightning, this velocity is demon-
strated by the dart leader of a subsequent component. Even the mere fact
that these velocities are comparable justifies our interest in the streamer
mechanism. It is important to know what determines the streamer velocity
and how it changes with the tip potential. For this, we have to analyse
grounded
cathode
processes taking place in the streamer tip region where ionization occurs. It is
necessary to find out how the processes of charged particle production are
related to electron motion in the electric field, due to which the charged
region travels through the gap like the crest of a sea wave.
The specific nature of spark breakdown is not restricted to the ionization
wave front, because its crucial parameter is the channel tip potential U,. Its
value may be much smaller than the potential U , of the electrode, from which
the streamer has started, since the channel conductivity is always finite and
the voltage drops across it. Therefore, the analysis of streamer propagation
for a large distance will require a knowledge of the electron density behind
the wave front and the current along the channel in order to eventually
calculate the electric field in the travelling streamer and to derive from it
the voltage drop on the channel. Incidentally, the field and the current
preset the power losses in the channel. It will become clear below how
important this parameter is for spark theory.
The streamer creates a fairly dense plasma. Without this, it would be
unable to transport an appreciable charge into the gap. A quantitative
description of the ionization wave propagation provides the initial electron
density in the channel and defines its initial radius. Behind the wave front,
the streamer continues to live its own life. A streamer channel may
expand, through ionization, in the radial electric field of its intrinsic
charge, provided that the latter grows. The cross section of the current
flow then becomes larger. The channel continuously loses the majority cur-
rent carriers - electrons. The rates of electron attachment to electronegative
particles and electron-ion recombination strongly affect the fate of the
discharge as a whole. If the air through which a streamer propagates is
cold and the power input into the channel is unable to increase its tempera-
ture considerably (by several thousands of degrees), the process of electron
loss is very fast, since the attachment alone limits the electron average lifetime
to lop7 s. This is a very small value not only at the scale of lightning but
should be called the ionization wave front. The streamer tip charge is primarily
concentrated in the region behind the wave front. The field there becomes low,
dropping to a value E, in the channel, small as compared with E,. The lines of
force going radially away from the tip in front of it become straight lines inside
the tip and align axially along the streamer channel.
Let us mentally subdivide the continuous process of streamer develop-
ment into stages. The strong field region in front of the tip is the site of
ionization of air molecules by electron impact. The initial seed electrons
necessary for this process are generated by the streamer in advance. Their
production is due to the emission of quanta, accompanying the ionization
process because of electronic excitation of molecules. In our case, highly
excited N2 molecules are active so that the quanta emitted by them ionize
the O2 molecules, whose ionization potential is lower than that of N2. The
radiation is actively absorbed, but its intensity is high enough to provide
an initial electron density M~ of about 105-106~m-3at a distance of 0.1-
0.2cm from the tip. Each of these electrons gains energy from the strong
field, giving rise to an electron avalanche. Since the number of avalanches
developing simultaneously is very large, they fill up the space in front of
the streamer tip to form a new plasma region. Owing to the electron outflow
towards the channel body, the positive space charge of the plasma becomes
exposed. Simultaneously, electrons that have advanced from the ahead
region neutralize the positive charge of the ‘old’ tip which turns to a new
channel portion, thereby elongating the streamer.
The gas in the wave front region must be highly ionized for the electron-
ion separation to produce an appreciable charge capable of creating a strong
ionizing field in front of the newly formed tip. For this reason, the region of
a0
= exp
s vidt =exp
Figure 2.4. Ionization frequency of air molecules by electron impact under normal
conditions (from the data on ionization coefficient cy and electron drift velocity V,
in [l 11).
This type of formula was first suggested by Loeb [5] and has been used since
that time, in this or modified form, in all streamer theories [6-lo]. The
velocity of a fast streamer is weakly related to the initial no and final n,
electron densities and is determined only by the maximum field E, and the
tip radius r,.
The quantities E, and r , which determine V , are not independent. They
are interrelated by the tip potential U,. For an isolated conductive sphere with a
uniformly distributed surface charge Q', we have U = r,E, = Q ' / ~ T T T E ~ Y , ,
where E~ = ( 3 6 x~ lo")-' FZ 8.85 x 10-12F/m is vacuum permittivity. A
streamer looks more like a cylinder with a hemispherical rounded end (see
figure 2.3). We can show [4] that in a long perfect conductor of this shape,
one half of the potential at the hemisphere centre is created by charges
concentrated on the hemisphere surface and the other half by those on the
cylinder surface, so that the tip charge is Q = ~ T T T E ~The Y , Ufield
~ . at the tip
front point is, to good approximation, only one half of that in an isolated
vim
It leaves behind it a positive charge of the same surface density q. It is this
charge that creates the field E,. We shall see soon that the effective thickness
of a positively charged layer is Ax << r,. In electrostatics, the field of such a
layer at the conductor surface is equal to E, M q / q ( t h s equality is
absolutely exact for the surface charge of a perfect conductor). By substituting
q from (2.4),we get an estimate for the density behind the ionization wave:
n, M Eovim/epCL,.
The plasma density n, is not related directly to the streamer velocity and is
essentially determined by the maximum field value which defines the ioniza-
tion frequency.
Let us make sure that the tip charge is indeed concentrated in the thin layer.
The effective time for the charge to be formed in the layer approximately is
the transition from an order of magnitude to a specific value have been deliber-
ately ignored. This is justified because we have simplified all initial conditions
and the derivation procedure in order to reveal the physics of the phenomenon
in question. The significance of a formula will increase if it is ‘equipped’ with
numerical factors, even approximate ones. Since we know the origin of these
factors, we can partly judge about the theory validity and meaningfully
compare the analytical results with computations and measurements.
A more rigorous treatment of a fast ionization wave, using one-
dimensional equations consistent with the streamer physical model [4,9],
yields the expressions
qmrm nm
v,= n, = nC
- = ln- (2.6)
(2k - 1) ln(n,/no) ’ kePe ’ nm no
where k is the power index from the approximate formula vi(E) N Ek and n, is
the electron density in the wave front at the point of maximum field (the
density is an order of magnitude smaller than the maximum achievable
density n,). In the field range characteristic of an air streamer, k = 2.5.
The issue of the streamer tip radius or maximum field represents the
most complicated and least convincing point in streamer theory. It is likely
that their values are established under the action of a self-regulation mechan-
ism related to the proportionality V , = q(E,) and to the rapidly increasing
(at first) and then slowly growing dependence of q on E (figure 2.4). If, at
constant tip potential, the tip radius turns out to be too small and the field
E, respectively too high, corresponding to the slow growth of vi(E), the
channel front end will not only move forward quickly but it will expand as
fast under the action of a strong transverse field. The value of r , will rise
while that of E,, according to (2.3), will fall.
Suppose, on the contrary, that the radius ,Y is too large and the field E,
is too low, corresponding to a rapid growth of vi(E).Any slight plasma pro-
trusion in the tip front will locally enhance the field. The ionization rate will
greatly increase there, and the protrusion will run forward as a channel of a
smaller radius. Some qualitative considerations of this kind were suggested in
an old work of Cravath and Loeb [12], but these authors discussed a lightning
channel obeying other mechanisms, because lightning develops via the leader
process. These ideas were used in [6,7] to formulate an approximate streamer
theory. A semi-qualitative criterion was suggested for choosing the maxi-
mum field feasible in the streamer tip. According to [6,7], E, corresponds
to the saturation point or bending in the function vi(E). This criterion was
refined in [13] by establishing a quantitative relation of E, to the slope of
the q ( E ) function and to the charge and normal field distributions over
the streamer tip surface. It was shown that the field E, at the tip front
point is established such that the normal field on its lateral surface
corresponds to the point of transition from the rapid to the slow growth of
the ? ( E ) function (figure 2.4).
wave front decreases, all electrons are drawn out of the tip, and space charge
fills it up. However, the final plasma density behind the ionization wave does
not decrease. The tip radius becomes very small, and the streamer stops at
low U,. This result agrees with experiments: no streamers with a velocity
less than (1.5-2) x lo5mjs have ever been observed in air under normal
conditions.
(2.11)
The current at the tip is defined mainly by the tip potential and streamer
velocity. At the anode, the current is
/ 1
1,. = dQ
dt, Q= J’ ~ ( xdx) = s o C1 [ U ( x )- Uo(x)]dx
0
(2.12)
anode to the channel tip. In this case, the anode current is close to the end
current defined by (2.11) including potential U, which may be much lower
than U,. Many experiments have shown that the average channel field
must exceed a certain minimum value of about 5 kV/cm for air in normal
conditions (see sections 2.2.6 and 2.2.7) to be able to support a long positive
streamer. For instance, if U , = 600 kV at the anode and the streamer length
is 1 1 m, nearly all voltage drops in the channel and U, << U,, but currents
i, and il still do not differ much.
To get a general idea about the orders of magnitudes, let us consider a
variant which seems quite realistic and is manifested by some calculations
(section 2.2.6). This is the variant with constant applied voltage and
slowly varying average field in the channel, when the current along it
changes little and, therefore, can be evaluated from (2.1 1). For example, at
I = 1 m, r = 0.1 cm, V , = 1.7 x lo6 m/s, and Ul = U, 34 kV, as in the
illustration in section 2.2.2 (with U, > 500 kV), we have ln(l/r) = 6.9,
C1 = 8 x 10-12F/m, r1= 2.7 x lOP7C/m, and i, = i, = 0.46A. Streamer
currents of this order of magnitude (as well as much higher or much lower
currents) have been registered in many experiments. These values can also
be obtained from calculations with the account of possible streamer velocities
from lo5 to lo7 m/s in air, which have been found in some experiments to be
even higher [4].
In a simple model of potential and current evolution in a developing
streamer channel; the latter can be represented as a line with distributed
parameters: the capacitance C1 and resistivity R1 = (xr2epene)-' per unit
length. The electron density n,(x. t ) should be calculated in terms of the
plasma decay kinetics (see section 2.2.5). Estimations show that self-induction
effects are not essential in streamer development [4]. Then, the process is
described by the following equations for current and voltage balance:
d r di dU .
-+--0, -- = zR1, 7 = C1(U - Uo). (2.13)
at dx dX
A boundary condition in the set of equations (2.13) at x = 1 is the equality
4 = c1[U1 - UO(4l Vs (2.14)
equivalent to (2.1 1). Formula (2.12) automatically follows from (2.13) and
(2.14). Another boundary condition may be the setting of anode potential,
since U ( 0 ,t ) = U a ( t ) .Equations (2.13) and (2.14) will be used in the next
section to evaluate the heating of a streamer channel. Illustrations of
streamer development calculations will be given in sections 2.2.6 and 2.2.7
after a discussion of the plasma decay mechanism. A complete set of
equations for a long line, generalized by taking self-induction into account,
will be applied in section 4.4 to the treatment of a lightning return stroke.
Equality (2.11) allows evaluation of longitudinal field E, in the channel
behind the streamer tip, where the electron density is still as high as that
s2 s
passage is
W = a E dt = aE2dx/Vs (2.16)
where the integrals are formally taken from --x to +xbut actually over the
ionization wave region. The principal contribution to energy release is made
by a thin layer behind the wave front where the electron density and field are
high. The integral of (2.16) was found rigorously to be ~ ~ E i / using 2,
equations for this wave region [4]. This value has the physical meaning of
electrical energy density at maximum field. The contribution of the region
before the wave front is In (nm/no)times, or an order of magnitude, smaller
than this value. Although the field there is as high as that behind the front,
the electron density is of the order of n, and the conductivity 0 is
In (nm/no)times smaller (section 2.2.2). Therefore,
W x e o E k / 2 x 2.6 x J/cm3 (2.17)
where the numerical value corresponds to E, = 170 kV.
The fact that the density of energy release in a gas is of the same order of
magnitude as the energy density of the electric field is quite consistent with
electricity theory. When a capacitor with capacitance C is charged through
resistance R to voltage U of a constant voltage supply, half of the work
QU = CU2 done by the supply is stored by the capacitor as electrical
energy, and the other half is dissipated due to resistance, irrespective of its
value. The value of R determines only the characteristic time of the capacitor
charging, RC. Something like this is valid for the case in question but, of
course, without both energies being rigorously equal to each other, because
this situation is much more complicated. Indeed, according to the results
of section 2.2.2, the tip capacitance is C, = Q / U , % 27re0rm, volume
V, x 4rrLI3, and field E , E Ut/rm, so that the energy dissipation per unit
tip volume is W FZ CtU:/2Vt z eOEk (we have ignored the unessential
term Uo(l)).
Joule heat is released directly in a current carrier gas, or an electron gas.
Then electrons give off their energy to molecules in collisions. An appreciable
portion of electron energy (even most of it in a certain range of E I N ) is used
for the excitation of slowly relaxing vibrations of nitrogen molecules. Some
energy is used for ionization and electron excitation of molecules, about
U’ = 100 eV per pair of charged particles produced, i.e., n , ~ ’= J/cm3 at
n, FZ 1014cmP3.But even without the account of these ‘losses’, the gas tempera-
ture rise in the wave front region appears to be negligible: AT < Wlcv = 3 K.
Here, cv = qkBN = 8.6 x loP4J/(cm3/K) is the heat capacity of cold air and
kB is the Boltzmann constant.
Let us see what subsequent gas heating can provide by the moment it
is somewhere in the middle of a long streamer channel. We multiply the
second equation of (2.13) by i and integrate over the whole channel length,
assuming, for simplicity, that constant voltage U, is applied to the anode.
After taking a by-part integral in the left side of the equation, we substitute
ailax from (2.13) and i(1) i, from (2.14), followed by simple transforma-
tions. As a result: we have
U,i, =
So
1
I
i RI dx + - -d x
dt
+ [y - Ci U I U O ( / )Vs
] (2.18)
which describes the power balance in the system; here, Uo(1) is unperturbed
potential of the external field at the streamer tip point x = 1. The input power
Uai, into a discharge gap is used for Joule heat release in the channel (the first
term on the right), for increasing the electric energy stored in its capacitance
(the second term), and for the creation of new capacitance due to the channel
elongation (the third term). Joule heat associated with the ionization wave is
not represented here. The field burst and the tip impulse current that make up
W calculated above are absent from equations (2.13) and (2.18). Having
integrated equality (2.18) over the period of time from the moment of
channel initiation to the moment t the channel has acquired length I , we
get the equation for the energy balance in the system at the moment t :
where charge Q is given by (2.12). The energy input into the channel, U,Q, is
used to create capacity (the last term on the right), partly stored in this
capacity (the integral) and partly dissipated (&Is). The braces ( ) t indicate
the time averaging of the process. In case of a long channel, much of the
applied voltage drops across its length, so the tip potential U , is small
most of the time, as compared with average channel potential U,, of about
U,. Then, the last term in (2.19) can be neglected.
If we compare the left side of (2.19) with the substituted expression for Q
from (2.12) and the integral in the right side of (2.19), we can conclude that
the difference between these values cannot be much smaller than their own
values but rather have the same order of magnitude. Therefore, the energy
KdlSdissipated in the channel is equal, in order of magnitude, to the gained
electrical energy, which is in agreement with a similar situation discussed
above.
The average energy dissipated per unit channel length is W,, x CI Uiv/2
and the average energy contributed per unit channel volume is
(2.20)
where rav is the average channel radius. With the formation of every new
portion of the channel, its radius was approximately proportional to the
tip potential owing to the fact that the maximum tip field remained approxi-
mately constant. So we have Uav/r,, M E,. Substituting this expression and
(2.8) into (2.20), we find
One can see that subsequent heating of the channel gas adds little to the initial
heating by an ionization wave passing through the particular channel site.
To conclude, gas heating due to streamer development is negligible if the
gap voltage remains constant. Higher voltage does not change the situation
because the energy dissipated in the channel grows in proportion with the
channel cross section and the air volume to be heated. Specific heating
remains unchanged, since it is determined by a more or less fixed volume
density of electric energy.
3
dt
= -vane - pn,2 (2.30)
(2.31)
where the time is counted from the moment the streamer tip passes through a
particular point of space.
Accordingto(2.15), wehaveE x 4.2kV/cmandE/N x 1 . 7 10-16V/cm ~
for a streamer channel just behind the tip at p = 1 atm. The electron
attachment frequency from (2.21) is va x 1.2 x lo7 sC1 and the characteristic
attachment time is 7, = v;’ x 0.8 x Over this time, most simple 0;
ions in dry air turn to complex 0; ions. Electrons recombine with them
with the coefficient ,8 x 2.2 x lop7cm3/s corresponding to electron tempera-
ture Te x 1 eV = 1.16 x lo4K at the above value of E/N. The initial electron
density n, x 1014cmP3 is so high that the parameter @nc/vax 2 determining
the relative contributions of recombination and attachment is larger than
unity. This means that at an early decay stage with t < ra x lop7s, electrons
are lost primarily due to recombination, with attachment playing a lesser role.
Later, at t > 2ra x 2 x lO-’s, the electron density decreases exponentially, as
is inherent in attachment, but as if starting from a lower initial value
nl = nJ(1 + @ n c ~ ax) 0.3n,; ne x nl exp (-vat).
The plasma conductivity decreases by two orders of magnitude, as
compared with the initial value, over t x 3 x s. At the streamer velocity
V , x lo6 m/s, this occurs at a distance of 30 cm behind the tip. A micro-
second later, the conductivity drops by six orders of magnitude. The streamer
plasma in cold humid air decays still faster because of a several times higher
rate of recombination with hydrated ions and due to the appearance of an
additional attachment source involving water molecules. These estimations
indicate a low streamer viability. It is only very fast streamers supported
by megavolt voltages that are capable of elongating to 1 x 1 m in cold air
without losing much of their galvanic connection with the original electrode.
This is supported by experiments with a single streamer and a powerful
streamer corona [4].
Note that a streamer plasma has a longer lifetime in inert gases, where
attachment is absent and recombination is much slower. This makes it
possible to heat the plasma channel by flowing current for a longer time
after the streamer bridges the gap (the estimations of section 2.2.4 do not
extend to these conditions). Such a process sometimes leads to a streamer
(leader-free) gap breakdown [ 171. Still, the formulation of the streamer
breakdown problem is justified for hot air and is related to lightning (see
section 4.8 about dart leader).
A U , = U, << U,, nearly all applied voltage drops in the channel. Therefore,
E,, can be treated as the lowest field limit, at which a streamer is still capable
of propagating.?
This interpretation remains valid when a streamer does not bridge a gap
but stops somewhere on the way. Indeed, according to (2.6) and (2.10), the
streamer velocity, and hence its ability to move on, is determined only by
the tip potential excess over the external potential, A U , = U , - Uo(I),and
is independent of the latter. No matter where a long streamer stops, we
shall have A U , << U,, though the external potential value at this point,
Uo(lmax),may be high. Generally the average field limit in the channel and
the streamer length at the moment it stops, ,I are interrelated as
In order to be able to use these relations in practice, we must know not only
the easily registered gap voltage U , but also potential Uo(Imax)inaccessible to
measurement. In most cases, it is hard to estimate even by calculation. For a
particular streamer, the external field is determined, in addition to the anode
charge, by the whole combination of charges that have emerged in the gap
and its vicinity. Especially important is the charge of all other streamers
that were formed together with the one under study. Consequently, the
field U o ( x ) ,in which the streamer is moving, represents a self-consistent
field. An outburst of hundreds of streamer branches is characteristic of air;
they fill up a space comparable with l,,,. It is this maximum length, rather
than the small anode radius, that will determine the external field fall
along the gap length. Estimations of a self-consistent field U o ( x )involve
considerable difficulties and errors (we shall come back to this when
evaluating the size of the streamer zone of a leader). So in reality, critical
field E,, can be evaluated only from experimental data that relate to a
situation with streamers bridging a discharge gap. Then, the potential
Uo(I,,,) is known reliably because it coincides with the potential of the
electrode, usually the grounded one: Uo(lmax)= U o ( d )= 0. But if it is
known that Uo(lmax)<< U,, as in the case of a long streamer moving in a
sharply non-uniform field, the criterion of (2.32) for a definite streamer
length will be extremely simplified: I, x U,/Ecr.
The existence of critical field E,, has a rather clear physical meaning. The
reason for the appearance of a minimum average field in a channel is its finite
t The average quantities E,, and E,, describe, to some extent, the actual field strengths in the
channel even when the external field is extremely non-uniform, changing by several orders of
magnitude along the gap far from the conductor. For example, if we close the gap with such a
thin wire that short-circuiting current does not change the electrode voltage, a short time
later, after the current along the wire is equalized, the actual gap field will become constant
along its length and exactly equal to Eav.
E,,/N does not remain constant and the value of E,, decreases more rapidly
than density N [19]. When a streamer propagates through heated air, the
critical field becomes lower not only due to a lower density but as a result
of a direct temperature action. This was found from measurements at various
p and T , up to 900 K [ 19,201. The reason is clear: on gas heating, the action of
attachment and recombination becomes weaker (section 2.2.5). In electro-
positive gases, in which there is no attachment, the value of E,, is lower
than in cold air, other things being equal. For instance, in nonpurified
nitrogen with an oxygen admixture up to 2%, the field is E,, M 1.5kV/cm
at p = 1 atm. In inert gases where the attachment is absent and the recombi-
nation has a lower rate than in molecular gases, E,, is much lower, about
0.5 kV/cm [21,22].
The channel field does not vary much in time along its length because of
the compensation due to countereffects. On the one hand, the conductivity in
an old channel portion is lower than in a new one because of the plasma
decay. On the other, that old portion was produced by a faster ionization
wave at a higher tip potential corresponding to a larger channel cross section.
As a result, the resistivity per unit length R1 = ( ~ r ~ e p ~ ndoes , ) - ~not vary
much along the channel. Of course, it grows in time because of electron
loss, but at the same time, the streamer velocity decreases together with the
channel current. For this reason, the time variation of the channel field
E ( x ,t ) = R I (x,t ) i ( x .t ) is much slower than that of any of the cofactors.
t The numerical simulation was made in cooperation with M N Shneider. This type of equation,
but with a constant channel radius, was solved in [23].
5M)
400 2 - 23.37 ns
3- 53.5ns
4- IOOns
2 300 5- 310ns
-
.-m
s
L
2 200
v
'E
0
100
0
20 40 60 80 1
x , cm x , cm
U -t 4
0.1
0 20 40 60 80 1 0
x , cm
Figure 2.5. Streamer propagation in air from a spherical anode of 5cm radius at
500 kV. The distributions of potential U , current I , field E , and electron density ne
along the channel at various moments of time until the streamer stops.
t The calculation using a simplified formula without the account of Uo(lmax)at l,, = 0.39m
gives an error: 5.7 instead of 4.9 kV cm.
Figure 2.6. The potential distribution along a conductive rod in a uniform electric
field. Broken line, the potential in the absence of a conductor.
space. The dipole charge enhances the field (E,,, > Eo) at the ends of the
polarized body (figure 2.6).
The problem of field redistribution by polarization charge can be solved
rigorously by numerical methods for any geometry, but simplified evalua-
tions are also possible. In the close vicinity of a charged dipole ‘tip’ of
radius Y << 1, the longitudinal field varies nearly in the same way as the
field of a sphere of identical radius. Therefore, the external field perturbation
by polarization charges is attenuated at a distance Y from each of the two
conductor ends. Let us take the conductor middle point to be the coordinate
origin. The end potentials of a polarized conductor differ from that of
an unperturbed one, U. = -Eox at the same points by AU x Eol/2. The
absolute strengths of the total field at the conductor ends rise to
E, x A U / r x Eol/2r, and the field increases with increasing l/r. This
estimate fits fairly well the numerical evaluation in [24].
At 1 >> r, ionization processes and streamers may arise at the ends of a
polarized conductor even at a relatively low external field Eo (figure 2.7(a)).
Ionization waves run in both directions, leaving plasma channels behind, in
much the same way as with a streamer starting from a high-voltage electrode.
If their velocity V , appreciably exceeds the electron drift velocity, the condi-
tions of streamer travel from the positive and negative ends will not differ
much. The total charge of developing streamers is zero at any moment of
time. This could not be otherwise, because none of them is connected with
the electrode and, through it, with the high-voltage source. Charges do not
escape the gap but are only redistributed by the streamer current. A streamer
Figure 2.7. Excitation of streamers of both signs from the ends of a conductor in a
uniform field. The charge distributions per unit channel length are shown schemati-
cally: ( a ) with active plasma, (b) with plasma decay in the older channel portions.
preserved conductivity. Here, A U also represents the excess of the streamer tip
potential over the external one. In particular, as the length I increases and the
field at the conductor ends becomes as high as E, M 150-170kV/cm for
normal density air (section 2.2.2), the growth of E, ceases. As I and A U
increase further, the streamer tip radius Y,, not E,, increases because
A U z E,Y,. This is accounted for by the self-regulation mechanism discussed
in section 2.2.2.
For the understanding of the streamer process in the streamer zone of
a leader, where the field is nearly uniform and the leader channel acts as a
high voltage ‘electrode’, it is essential that the ‘electrode’ and ‘electrodeless’
situations should be strictly equivalent, provided that the positive and the
negative streamers are identical and the external field is uniform. Let us
mentally cut, at the centre, a plasma conductor developing in both directions
from this centre and discard, say, the negatively charged half. Let us now
replace it by a plane anode under zero potential and assume a negative
potential to be applied to a remote plane cathode. A cathode-directed
streamer produced at the anode by a local inhomogeneity, whose field
initially supported ionization, will be identical to a positive streamer in the
electrodeless case. Indeed, in both cases, the conductor potential U coincides
with the external potential, U ( 0 ) = Uo(0). The charge pumped from the
negative half into the positive one will now be supplied by the source current
from the anode. Here, the principle of mirror reflection in a perfectly
conducting plane, well-known from electrodynamics, reveals itself in every
detail. According to this principle, the distributions of charge, current and
field in half-space do not change if the plane is replaced by the mirror
reflection of half-space charges.
These considerations were used in the calculations and representation of
results on streamer development in a uniform field U, = -Eox from the
point x = 0 towards lower potential (figures 2.8 and 2.9).t The solution
was derived from the same set of equations (2.13), (2.14), (2.6), (2.10),
(2.31) and the same plasma decay characteristics as in section 2.2.6. The
calculations show that a streamer does not develop if the external field is
lower than a certain minimum value. The values of Eommdo not differ
much from the critical channel field E,, which determines the streamer
length in a non-uniform field calculated with (2.32): Eo M 7.7 kV/cm. One
may probably use for estimations the experimental value E,, x 5 kV/cm as a
realistic Eo (section 2.4.1). If the uniform external field slightly exceeds
the minimum, the excess tip potential is small, the streamer velocity is low,
and the channel field is close to the unperturbed external field Eomn.This
situation is illustrated in figure 2.8.
If Eo is appreciably higher than Eo however, the tip potential is much
higher than the external potential, and the streamer develops a high velocity
' 2
0 0 5 10 15 20 25 30 35 0
x, cm x, cm
Figure 2.8. An air streamer in a uniform field Eo =7.7 kV/cm, slightly exceeding the
critical minimum, with calculated distributions of potential U , current I , field E and
electron density ne. Dashed line, applied field potential counted from the streamer
origin. The oppositely charged streamer running in the opposite direction is not
shown.
(figure 2.9). The current in well-conducting portions a t the tip is high but
decreases towards the channel centre. Owing to the high current, much
positive charge is pumped into the tip region even at an early stage; in the
tail, however, where the conductivity has decreased, the current is low. As
a result, positive charge pumped out of it is not reconstructed; moreover,
--_ ---I_. h
5 .4
-800 3 - 482.1 ns \
4- 576.6ns
5 - 653.1ns
-1000
0 20 40 60 80 100
x, cm x, cm
2.5 7
x, cm x, cm
Figure 2.9. An air streamer in a uniform field Eo =10 kV/cm with the charge distribu-
tion 7 instead of the ne curves similar to those in Figure 2.8.
This section is a key one for the understanding of a long spark and the first
lightning component. We shall try to answer the question why a simple struc-
tureless plasma channel has no chance to acquire a considerable length in cold
air of atmospheric pressure. The reader will see what is necessary for a spark to
become long and have a long lifetime and how Nature realizes this possibility.
but the streamer cross section 7rrk also increases as U 2 .So the released energy
density proportional to ( U / r m ) 2 , which defines the heating, remains low. To
increase the channel temperature considerably, it is necessary to accumulate
a much higher energy in a much narrower plasma column. For this, the func-
tional relation providing the low U / r m ratio must be violated. This is
impossible in a primary ionization wave but becomes possible in a differently
organized channel development. Let us try to approach this problem by
considering the final result and estimate voltages and plasma channel radii,
at which the gas temperature would become sufficiently high. This can be
done in terms of the general energy considerations discussed in section 2.2.4.
Suppose there is a charged space with characteristic size R in the front
region of a developing plasma channel. Its capacitance is C x m o R with
C1 x x0per unit length along the channel axis. If the tip potential is U ,
the energy dissipated per unit length of a new portion of the system including
the channel and the space being charged is C1U 2 / 2 ,provided the spark devel-
ops steadily. Since the capacitance per unit length of a new system portion is
independent of its radius R or of any other geometrical dimension, we are
free to assume any nature, size, and volume of this charged space. (The
specific capacitance of the central portion of a long system does vary with
total length and radius but only logarithmically, as is clear from formula
(2.8)) The dissipated energy includes all expenditures for the creation of a
new channel portion and space charge. Attribution of this energy to the
various expenditures is a special problem which requires details of the
process to be specified. But we can estimate the upper limit of air mass
that can be heated to the necessary temperature, say, to T = 5000K. For
this, let us assume that all energy has been used heating an air column of
initial radius ro. This will be an estimate of the upper radius limit. This
temperature will lead to considerable thermal gas expansion, because a hot
channel, as will be shown later, develops much more slowly than a cold
streamer channel. Current must have enough time to heat the gas, because
it is eventually the released Joule heat of current that does the heating. If
the heating rate is not high enough, pressure in the gas space is equalized,
so that the gas of a thin channel becomes less dense. The air heat capacity
does not remain constant within a wide temperature range, so energy calcu-
lation should be made in terms of specific enthalpy h( T , p ) .Therefore, the
expression to a maximum radius rOmaxof a cold air column that can be
heated to temperature T is
7rrim,,p0h(T) x T & ~ U ~ / ~ . (2.34)
Here, h( T ) is specific enthalpy for air at p = 1 atm and po is its density at
p = 1 atm and To = 300K.
With tip potential U, = 1 MV and T = 5000 K, when h ( 5 ) = 12 kJ/g, an
air column that can be heated must have an initial radius less than
yomax = 0.054cm. The maximum radius due to thermal expansion will be
less than rmax = ro [ p o / p (5 ) ]' I 2 M 0.26 cm, where p( 5) is the air density at
T = 5000 K and pressure 1 atm. A channel of this thickness has been observed
in laboratory spark leaders. At U, M 100 MV, characteristic of very powerful
lightning, the radius estimated from formula (2.34) must be two orders of
-
magnitude larger. A lightning leader, however, has a temperature higher
than 5000K and h T 2 approximately (h(10) = 48kJ/g), so that the
radius does not grow as much as U, and remains as small as several centi-
metres. It may seem surprising, but a leader channel is thinner than a streamer
channel at the same tip potential (their radii are established due to different
reasons: a leader radius follows the heating conditions, while a streamer
radius is such that the lateral field is too low for intensive ionization).
away from the leader tip. We already know (sections 2.2.6 and 2.2.7) that the
average field necessary for streamer development in atmospheric air must be at
least E,, x 5 kV/cm. Since streamers stop at the end of the streamer zone, the
voltage drop along the zone length R is about AU, x E,,R (cf. formula (2.32)).
About as high voltage drops outside the streamer zone, because the field there,
i.e., in a zero-charge space, drops from about E,, to zero, as for a solitary
sphere of radius R. Hence, we have U x 2AUs and R = U/2Ec,. At
U x 1 MV, the streamer zone radius is R x 1 m, in agreement with laboratory
measurements.
It follows from both calculations and measurements that the current,
field, electron density, and conductivity of a heated leader channel are
generally comparable with respective parameters of a fast streamer. If they
are somewhat larger, the difference is not orders of magnitude. So the heating
time to achieve a much higher gas temperature must be much longer. This
explains why a leader propagates much more slowly than a fast ionization
wave.
The capacitance per unit length of a leader system (the channel plus a
charged cover) will be described by the same formula (2.8) if I is substituted
by leader length L and the conducting channel radius Y by cover radius R , the
actual radius of a charged volume. This follows directly from electrostatics.
Similarly, the current iL at the leader channel front is related to the tip
potential U and leader velocity VL by the same expression (2.1 l)t
(2.35)
t Here and below, the external field potential Uo(x)is omitted for brevity. It is indeed small in
laboratory leaders normally observed in a sharply non-uniform field.
single streamer. Part of the streamer zone appears to be behind the tip,
transforming to a new cover for the newly born leader portion. But this
does not decrease the streamer zone length, because meanwhile the zone
has moved forwards together with the tip. Note only that if there are
many streamers they are very close to one another, and they travel in a
self-consistent field close to the critical field (sections 2.2.6 and 2.2.7). Such
streamers are slow and have a low current [4], so that the leader current is,
indeed, a sum of numerous low streamer currents.
larger than K x 0.3 cm2/s. If a non-uniformity takes less time for dispersion
than for development, i.e., if Tamb is smaller than the instability lifetime qns,
the latter is suppressed at its origin.
The scale for qns is the characteristic time of, say, gas temperature
doubling in a perturbed plasma column, as compared with initial tempera-
ture To. This time is pocpTo/jE,where j E is the power of Joule heat release
and cp is specific heat at constant pressure with the account of thermal
expansion. But this is not all. The higher the instability development rate
-
is, the greater is the steepness of the ionization frequency dependence on
reduced field E / N E T , i.e., on gas temperature. For instance, if a 10%
increase in T raises the ionization rate by 20%, the instability will, generally,
double its rate, as compared with a 10% increase in the ionization rate. This
circumstance brings the factor Ci d In vild In ( E / N ) into the theoretical
formula for qns[26], which characterizes the q ( E / N ) function steepness.
This yields the following expression to be used for estimations:
(2.37)
We have shown above that a streamer zone plays the key role in a leader
process. It is here that a space charge cover is formed which stabilizes the
leader channel, preventing its ionization expansion which would otherwise
exclude plasma heating. A streamer zone is the site of current generation
for heating the leader, providing its long life. In this section, we shall deal,
in some detail, with processes occurring in the streamer zone and leader
cover, defining the priorities in the causative relationships among leader
parameters. We shall show how the process of streamer generation from a
leader tip becomes automatic.
laboratory sparks and -10 MV or, probably, more for lightning. Streamers
are continuously produced in a leader tip, which means that the field at its
surface Et exceeds the ionization threshold Eiz 30 kV/cm (under normal
conditions). This excess cannot be very large, otherwise a streamer flux would
become too intensive. The excessive charge of the same sign as U , introduced
into the space would create a much stronger reverse field which would reduce
Et to a level close to E,. Therefore, the field E, does not exceed much E, and
has the same order of magnitude. An automatic field stabilization is inherent
in any continuous threshold process of charge generation by an electrode, for
example, in a steady-state corona. Measurements have shown that the field
near a corona-forming electrode is stabilized with high accuracy and does
not respond to voltage rise across the gap; what changes is the corona
intensity, i.e., its current. A leader tip, too, is the site of corona formation,
with an intensity high enough to support a quasi-stationary state in the tip
and streamer zone, corresponding to potential U,. The field at the corona
electrode E,, is shown by stationary corona experiments to be by a factor
of 1.5 higher than E,, if the electrode radius is about the leader tip radius
r - lcm.
At E, = E,, = 50 kV/cm and Y, = 1 cm, the leader tip charge q, =
4mO~:Et % 5 x lO-'C is capable of creating only a small portion of
E,r, = 50kV of an actually megavolt potential U,. The main potential
source is, therefore, the space charge of the streamer zone and cover
surrounding the tip. But the value of U, is primarily determined by character-
istics external relative to the tip. This is the electrode (anode) potential minus
the voltage drop across the leader channel. Consequently, the charge Q, and
the size R of a streamer zone, as well as respective cover parameters, are
established such that they correspond to the proper potential U,. The
mechanism by which a leader 'chooses' the values of Q, and R are directly
related to streamer properties. There are many streamers present in the
zone at every moment of time. They are emitted by the tip at a high frequency
(see below), have different lengths at any given moment and are at different
stages of evolution, with their charges filling up the zone space. Every single
streamer moves in a self-consistent field created by the whole combination of
streamers. The contribution of the leader tip itself (or of its channel) to the
total field has just been shown to be small. One exception is the region
around the tip with a size of its radius.
There are experimental and theoretical grounds to believe that the field
strength in the streamer zone, except for the tip vicinity, is more or less
constant and close to the minimum at which streamers can grow. This is
indicated by measurements of streamer velocity, which does not vary along
the streamer zone. (Attempts to measure a single streamer in the tip
region, where the streamer density is high, have so far failed.) Experiments
show that until the streamer zone of a laboratory leader touches the opposite
electrode, streamers move slowly, at a nearly limit velocity of about lo5 mjs.
This is possible only in a uniform field close to Eo mn (section 2.2.7).Streamers
can travel for such a distance R,at whch the field Eo mn still exists, but they
stop on entering the region with E < E,, ml,.
Suppose, for simplicity, that the streamer zone is a hemisphere with the
centre in the leader tip. The hemisphere changes to a cylindrical cover of the
same radius R with the same order of the space-charge density. A thin con-
ducting leader channel goes along the cylinder axis as far as the hemisphere
centre. When evaluating the zone parameters, one should take into account
the cover charge at the leader end, which also affects the zone field. We can
do this simply by connecting the hemisphere, simulating a streamer zone, to
another hemisphere by mentally cutting it out of the cover space. Let us
assume that there is a uniform radial field E = Eomi, in the sphere. As was
mentioned in section 2.2.7, the theoretical limit of Eo is close to the experi-
mental critical average field in the streamer channel, below which a streamer
cannot propagate. For air, therefore, we have E x E,, x 5 kV/cm. A uni-
form field in sphere geometry corresponds to the space charge density
p = 2 ~ & / r . If the leader tip is far from the earth and grounded electrodes,
its potential is
(2.39)
(2.40)
UR = U, - E R = Ut/2.
Figure 2.12. Streak photograph of the positive leader channel bottom in an air rod-
plane gap of 12 m in length. The streamer zone is seen to elongate when approaching
the plane cathode.
(2.41)
which was not taken into account in the calculation or (2.39), it is by one
order less than that distributed along the channel. At N , M lo5, the average
interstreamer space is R/N;I3,i.e., about several centimetres. With this large
separation of streamer tips, the streamers can really be considered solitary
and propagating in an average self-consistent field.
When a streamer reaches the end of the streamer zone, it stops because it
enters a field lower than Eo ,in. Since the streamer zone approaches this field
fairly slowly, at leader velocity VL an order of magnitude lower than
streamer velocity V , (for laboratory streamers), the streamer loses its
conductivity entirely. The ions of its space charge are gradually repelled
(and diffuse), reducing the field near the charge trace, so that the streamer
trace becomes lifeless 'forever'. The streamer zone still passes by for a time
tL = R / V , N lop4s, after which the immobile charged trace, which is now
behind the leader tip, becomes a cover component. Viable streamers fly
-
across the streamer zone over time t , = R / V s m i n lop5s, an order of
magnitude shorter. Therefore, if the frequency of streamer production is
-
v,, the number of viable streamers in a streamer zone is N I v,t,, while
--
the number of non-viable traces, practically coinciding with the total
- -
number of charged streamer portions, is N v,tL. Hence, the streamer
generation frequency is v, N / t L N V L / R lo9 s-'.
touched its surface [29]. The counts were integrated over the area. Measure-
ments made at different currents showed that the relation
iL/v, = qs RZ 5 x 10-”C remained constant. It is consistent with measure-
ments and the estimations of average streamer charge presented here.
The formula for leader tip current can be given in a conventional form of
the type i = 7rr2 eneve = T ~ Vwhen
~ , current is expressed as the number of
charge carriers (electrons) per unit current column length T, and electron
velocity V,. This formula can also be changed to the phenomenological
expression (2.36) describing the result of the current process without
indicating the nature of carriers. Although the current carriers in this case
are electrons, we can also speak of ‘macroscopic’ carrier-moving charged
streamer sections. With what we mentioned at the end of section 2.3.2 and
formula (2.36), we have
(2.42)
electrode remains the same, U,, whereas the length L, becomes shorter. The
average field in the streamer zone E = U,/L, rises, together with streamer
velocity (section 2.2.7) and leader current it V,. The final jump current
N
negatively (if the leader is positive), the field around it will become negative, its
module exceeding Ei,. A negative corona will be excited and introduce into the
cover as much charge as is necessary to reduce IE,I to Ein.This situation
becomes feasible when the gap voltage is constant or decreases, since then
the channel ‘enters’ the cover with a too-large charge. Indeed, suppose the
charge of a streamer hemisphere of radius R is distributed as p = 2 ~ ~ E / r
(section 2.4.1), creating potential U, = 2ER (2.39) in its centre. The cover
inherits a charge of the same radial density distribution p = 2 ~ ~ E=/ r
~ E ~ U ,and/ R amount
~ r’ = 27reoUt per unit length. At point x far enough
from the channel ends, L >> R , this charge will create potential
U’(x)
=-
1
YTEO
5”J’ 2 ~ r dr
p dz
o o [r2+ ( z - x) ]
1,2 M
2
r’
-In-
~
Le
= U, In-
~R 0
Le
R
(2.43)
when the channel potential becomes equal to the earth’s zero potential (see
section 4.4).
All ionization processes responsible for the leader development are localized in
the streamer zone, leader tip and a short channel section behind it. In the latter,
gas heating is completed and a quasi-stationary state characteristic of a long
spark is established. In t h s sense, the rest of the channel plays a minor role,
simply connecting the operating part of the leader to a high-voltage source.
High potential and current vital to the ionization and energy supply are
transmitted through the channel. But how much voltage reaches the leader
tip depends on the channel conductivity which, in turn, is determined by the
channel state. For this reason, what is going on in a developing channel is
as important to the leader process as the mechanisms described above.
bending, which may increase the length by 20-30%. Finally, the accuracy of
E,, is not as high as is necessary for such a delicate operation. Experimental
researchers know that the streamer zone field varies with air pressure,
humidity, and temperature but they do not know the respective corrections.
Nevertheless, the data in table 2.1 demonstrate a decrease in the average field
with increasing channel length. This is also evident from experiments with
superlong sparks. A voltage of 3-5 MV is sufficient to create a spark 100 m
long or longer. The tip potential necessary for the development of a streamer
zone of several metres in length is 1-2 MV, as is clear from table 2.1, there-
fore the average field in such a long channel will be as low as 200-250 V/cm.
These values are more applicable to older, remote channel sections which
have acquired a quasi-stationary state, but the fields close to the tip are much
higher. This follows from many experiments indicating a regular increase in
average field with decreasing leader length. More explicitly this was shown
by supershort spark experiments, when the channel length was only a few
dozens of centimetres [31]. The field in a supershort leader and, therefore,
the field at the respective distance from the tip of a long spark may be 2-
4 kV/cm. At the site of ionization-thermal instability, where current is
accumulated within a thin column, the field was found (section 2.3) to be
20 kV/cm [4]. But far from the tip, it is nearly two orders of magnitude lower.
At a typical experimental leader current of i x 1 A, its velocity is
V , M 1.5-2cm/p, and the lifetime of a leader section at a distance of 3 m
from the tip is at least 150 ps. This time is long enough for the relaxation
processes in the channel to be nearly completed and for a nearly steady-
state to be established. The thermal expansion of the channel, very fast at
the beginning, is also completed by that time. Measurements made in a
10m gap between a cone anode and a grounded plane [28] (voltage 1.6-
1.8 MV, average current about 1 A, and average leader velocity 2 cm/ps)
showed that the channel expansion rate was l00m/s at first but loops
later the rate dropped to 2m/s. Measurements made by the shadow tech-
nique showed the average thermal expansion radius to be rL = 0.1 cm.
According to spectroscopic measurements, the temperature of a channel
which has reached the gap middle is 5000-6000 K. Some other experimental
data on laboratory leaders can be found in [4, 27, 28, 31, 321.
The air ionization mechanism changes radically at temperatures T M 3000-
6000K and relatively low reduced fields (for example, at E = 450V/cm,
T = 5000K, and p = 1 atm, we have E / N = 3 x V/cm2).t In cold air,
t We should like to warn against the commonly used postulate that the field in a leader channel
has a constant value of E / N 8 x 10-'6VcmZ. Some authors use it for the calculation of
breakdown voltages in air gaps, including long ones. At T = SO00 K, we have E = 1.15kV/cm,
which disagrees with experimental data for more or less long leaders (table 2.1) and contradicts
the physics. The underlying implicit suggestion is that the only consequence of air heating is a
change in its density. We shall show that this is not the case.
1o4
0 1 2 3 4 . 5 6
Temperature, kK
Figure 2.13. Parameters of the initial leader channel right behind the tip as a function
.
of the gas temperature (model calculations of [34]; 1 Td = lo-’’ V cm2).
.T
Figure 2.14. Schematic arc channel with nearly the same distributions of T and o at
the axis of the ‘old’ portion of a leader channel.
but not very rapidly. But the conductivity 0 and current i = .irr2uE would
grow very quickly if the field remained constant. As fast would be the
growth of energy release P I = iE,which would set the system out of heat
balance. The balance is maintained because the field drops with rising
current, while the power and temperature do not change much. In an ideal
case with T, = const, which is close to the low temperature conditions for
a leader and low current arc, we have E x i-' from (2.47). The arc CVC is
indeed a descending curve, though its goes down somewhat more slowly
because T, and P1 slightly rise with current [26].
A similarity between the states and CVC curves for quasi-stationary
leaders and arcs was established in model experiments [4]. Sparks 7cm long
were generated in air between rod electrodes. The circuit parameters were
chosen such that the channel current was stabilized at a level characteristic
of a long laboratory spark at the moment of gap bridging. The stabilizing
mode lasted several milliseconds. During this time, a quasi-stationary state
was established under energy supply conditions close to those of the leader
process. The CVC thus measured is approximated by the expression
E = 32 + 52/i V/cm, i [A]. (2.48)
The obtained field appears to be lower than in a leader (84 V/cm at 1 A) and
closer to the arc field.
and gas excitation in the streamer zone, this value will increase by, at least, a
factor of 1.5 [4].Therefore, the tip potential in the initial leader stage will be
several hundreds of kilovolts even under favourable conditions.
The voltage U. applied to the gap drops across the leader channel and is
partly transported to the tip. The general formula is
U0 = EL + U, (2.49)
where E is the average field in a leader channel of length L. We showed in
section 2.5 that in a long channel, most of which is in a quasi-stationary
state, E is a more or less definite value varying with current i. The channel
field decreases with increasing current. But current growth requires that
the tip potential determining the leader velocity and current i = C1U, VL
should be raised. At a fixed length L , the Uo(i, L ) function has a minimum,
since it is the sum of a falling component and a component rising with i.
Minimum voltage U 0 ~ * ( Lcorresponds
) to current iOpt(L)optimal for a
leader of length L. It is hardly possible, in the present state of the art, to
find the Uomin(L) function theoretically. We shall try to define its character
using semi-empirical data.
Many experimental physicists have measured the leader velocity varia-
tion with applied voltage U,. Much work has been done on short leaders
because one can neglect the voltage drop across the channel, assuming
U, = U,. With the account of this approximate equality, Bazelyan and
Razhansky [35] suggested an empirical formula: VL z uU:’~, where
a E 1.5 x lo3V-1/2 cm s - l . Physically, the velocity increase with voltage
looks quite natural (though this variation is not strong). We also know
that the tip current is defined by (2.36). This gives the relation
U, = Ai2l3(VL i 1 / 3 )with A = ( C ~ U ) =
N - ~( /2~7 r ~ ~ a ) -Let
~ / ~us. use the
analogy between a well developed leader and an arc and take the CVC
E = b / i typical for a low current arc. Let us put b = 300 V A/cm for numerical
calculations and ignore the difference between the tip and channel currents.
+
We shall then get U, = Lb/i AiZi3and after differentiation
iopt= (3Lb/2A)3i5, = $ U, opt
U. mln = A3/5(3bL/2)215 (2.50)
If a leader develops under optimal conditions, the applied voltage is
shared by the tip and the channel in comparable proportions. The mode
with a low tip potential close to the limit admissible from the energy criteria,
is unprofitable for a long leader, because it corresponds to low current lead-
ing to a considerable voltage drop across the channel. The long spark param-
eters in table 2.3 found from (2.50)with semi-empirical constants are quite
reasonable: these orders of magnitude for current, voltage, and velocity
meet the requirements on the optimal experimental conditions for long
leader development. Besides, the experiment requires a nonlinear, slow
dependence of minimum breakdown voltage on the gap length. It is generally
known that increasing the length of a multi-metre gap is not a particularly
effective way of raising its electrical strength. This is a key challenge to those
working in high-voltage technology.
The results of extrapolation of formula (2.50) to a lightning leader
( L = 3 km) also lie within reasonable limits.
What is the rate of gap voltage rise necessary for the optimal mode of
spark development? Clearly, the gap voltage must be raised as the spark
length becomes longer according to (2.50), where L is an instantaneous
leader length. The existence of an optimal mode of spark development has
been confirmed experimentally [36-381. It has been shown that for a
breakdown to occur at minimum voltage, the pulse risetime t f must increase
with the gap length d. The authors of [39] recommend the following empirical
formula for the evaluation of an optimal risetime:
lfopt 50d [PSI, d [ml (2.51)
Generally, optimal voltage impulses have a fairly slow risetime. Their values
vary between 100 and 250ps in modern power transmission lines with the
insulator string length of 2-5m. We shall return to this issue in chapter 3,
when considering the diversity of time parameters of lightning current
impulses. The minimum electric strength of an air gap with a sharply non-
uniform field can be found from the formulas [4]
3400
= -[kV], d < 15m
u50% min
1 + 8/d (2.52)
U,,,, min = 1440 + 55d [kV], 15 < d < 30 m
Most lightnings carry a negative charge to the earth because they are ‘anode-
directed’ discharges. It is always more difficult to break down a medium-
length gap between a negative electrode and a grounded plane. A negative
leader requires a higher voltage. The difference between leaders of different
polarities is due to the streamer zone structure, while their channels and
voltage drop across them are quite similar. Indeed, a gap of about lOOm
long, in which an appreciable part of voltage drops across the channel, is
bridged by leaders of both signs at about the same voltages [2,3].
poorly understood. In the 1930s, when Schonland started his famous studies
of lightning [41], a negative leader was found to have a discrete character of
elongation, so it was termed stepwise. Streak photographs exhibit a series of
flashes, indicating that the leader propagates in a stepwise manner. Later, a
similar process was found in a long negative leader produced in laboratory
conditions 142,431. With every step, a negative leader elongates by dozens
of centimetres, or by several metres in superlong gaps [3]; steps of a hundred
metres have been registered in negative lightning discharges. Every step of a
laboratory leader is accompanied by a detectable current overshoot which
quickly vanishes during the time between two steps.
Without going into theoretical explanations of this mechanism, based
on an unverified hypothesis, let us see what information can be derived
from streak photographs of the process, made during laboratory experiments
[44]. These are naturally more informative than streak photographs of a step-
wise lightning leader. It is seen from figures 2.16 and 2.17 that in the intervals
between the steps, the tip of a negative leader slowly and continuously moves
on together with its streamer zone made up of anode-directed streamers. The
main events occur near the external boundary of the negative streamer zone.
It seems that a plasma body elongated along the field arises there and is
polarized by the field (compare with the discussion in section 2.2.7). The
positive plasma dipole end directed towards the main leader tip serves as a
starting point for cathode-directed streamers. They move towards the tip,
thus elongating the conducting portion of the channel and enhancing the
negative field at its end directed to the anode. Almost at the same time, the
plasma body generates an anode-directed streamer. T h s nearly mystic
picture of streamer production in the gap space is clearly seen in a streak
photograph in figure 2.18. Nothing like this has ever been observed with a
positive leader.
Figure 2.16. A schematic streak picture of a negative stepped laboratory leader: (1,2)
secondary cathode- and anode-directed streamers from the gap interior; (3) secondary
volume leader channel; (4)main negative leader channel; (5) its tip; (6) plasma body;
(7), (8) tip of secondary positive and negative leader (9) leader flash concluding step
development.
Figure 2.17. A streak photograph of the initial stage in a negative laboratory leader.
Marking numbers correspond to figure 2.16.
The polarized plasma section becomes the starting point not only of
streamers but also of secondary leaders which follow them. They are
known as volume leaders. A positive cathode-directed volume leader grows
intensively. Normally, its streamer zone almost immediately reaches the
main negative leader, so it looks as if the secondary positive leader develops
Figure 2.18. The origin of anode-directed (1) and cathode-directed (2) streamers from
the gap interior; (3) initial flash of a negative corona (static photograph) which trigger
a streak photograph regime; (4)arisen negative leader.
in the final jump mode, i.e., very quickly. The negative volume leader moves
towards the anode somewhat more slowly. When the tips of the main
negative and of the positive volume leaders come into contact, they form a
common conducting channel, giving rise to the process of partial charge
neutralization and redistribution. As a result, the former volume leader
acquires a potential close that of the main negative leader tip. This process
looks like a miniature return stroke of lightning, accompanied by a rapidly
rising and just as rapidly falling current impulse in the channel and external
circuit. The intensity of the channel emission increases for a short time. It is
hard to say what exactly stimulates this increase - the short temperature rise
or the ionization in the channel cover: which changes the cover charge,
thereby getting ready for a potential redistribution along the channel (see
section 2.4.4). The negative portion of the plasma dipole turns to a new
negative tip of the main leader. This is the mechanism of step formation
and stepwise elongation of the main channel. Then the story is repeated.
The picture just described gives no ground to draw the conclusion about
a stepwise character of negative leader development. The motion of a nega-
tive leader is continuous, but secondary positive volume leaders, also contin-
uous, produce a stepwise effect. Discrete is the final result of their ‘secret
activity’, but only if the observer is equipped with imperfect optical instru-
ments. In other words, what is generally known as a step is an instant
result of a long continuous leader process. As for gap bridging by a main
negative leader, one should bear in mind that most of the channel is created
by auxiliary agents - by a succession of positive volume leaders.
This picture has been reconstructed from streak photographs. But we
still do not know how polarized plasma dipoles are formed far ahead of
the main leader tip. Their appearance is hardly a result of our imagination.
Steps can be produced deliberately by making a volume leader start from a
desired site in the gap. For this, it suffices to place there a metallic rod several
centimetres long (figure 2.19). A series of rods placed in different sites of a
gap will create a regular sequence of volume leaders. The work [45] describes
an experiment with a negative leader 200 m long. Its perfectly straight trajec-
tory was predetermined by seed rods suspended by insulation threads at a
distance of 2-3m from each other. A volume leader started from a rod
when it was approached by the negative streamer zone boundary of the
main leader. Clearly, the rods are polarized by the streamer zone field to
serve as seed dipoles instead of natural (hypothetical) plasma dipoles.
There are many hypotheses concerning the stepwise leader mechanism,
but they are so imperfect, lacking strength, and, sometimes, even absurd that
we shall not discuss them here. We are not ready today to suggest an alter-
native model either. Additional special-purpose experiments could
certainly stimulate the theory of this complicated and challenging phenom-
enon. It would be desirable to take shot-by-shot pictures of a negative
leader tip region with a short exposure. A sequence of such pictures would
Figure 2.19. An artificially induced step: (1) initial flash of a negative corona from a
spherical cathode; (2, 3) cathode- and anode-directed leaders from a metallic rod
2.5 cm long, placed in the gap interior; (4) leader flash concluding the step develop-
ment; (5) new streamer corona flash from the tip of the elongating channel.
References
of different scales, but it is also clear that both have a common nature. For
this reason, we shall often compare the parameters of lightning with those of
a long spark. We should like to emphasize that this will be a comparison
rather than a direct extrapolation, because there is no complete analogy
between the two phenomena.
km
6-
4-
2-
0 I
Figure 3.1. The ‘dipole’ model of the charge distribution in a storm cloud.
space between the cloud and the earth should be scanned, and this must be
done for some fractions of a second, just before a lightning discharge, in
the vicinity of its anticipated trajectory. Unfortunately, such attempts have
not been quite successful. More successful were measurements made at
points on the earth’s surface separated at a distance of hundreds and
thousands of metres [17-191. These have been used to reconstruct the
charge distribution within a storm cloud, invoking the results of direct
cloud probing. The reconstruction procedure and its possible errors are
discussed in [20]. Generally, with simultaneous field measurements made at
n points, one can write a closed set of equations for the same number of
parameters of charged regions. Its solution provides the parameters, for
example, the average space charge densities in pre-delineated regions.
Most often, the number of points is too small, so the results obtained only
permit the construction of simplified models with point charges. Very common
is the dipole model with a negative charge at 3-5 km above the earth with the
same value of the positive charge raised at a double altitude. Sometimes, a
small positive point charge is added to them, which is placed at a distance
by 1-2 km closer to the earth than the negative charge. All point charges are
assumed to be located along the same vertical line (figure 3.1).
The concept of a cloud filled by charged layers of different signs is
based on probe measurements of charge polarity in hydrometeors. In this
respect, this model raises no doubt. But as for the field distribution, the
measurement error is too large, especially for the space in the cloud between
two point charges. Luckily, the descending lightning trajectory lies mostly
outside of the cloud, in the air free from charged particles. For this part of
the trajectory, the average field evaluation in terms of a simple model
makes sense.
2q 4
El = E2 = - q + (3.1)
47r&o(h+ Y)2 ' 47reor2 47r&o(2h+ r)2 '
The factor 2 in El and the second term in E2 are due to the action of charge
induced in the earth's conducting plane (mirror reflection). Since in reality
El << E2, it is natural to suggest that r << h. Then, with the second term in
E2 ignored, we find
leadp channel
L- return
E stroke
b- 1.5 ms -A
,
Figure 3.2. Schematic streak picture of a positive descending lightning leader regis-
tered on the San Salvatore Mount in Switzerland [22].
Figure 3.3. Schematic streak picture [22] of a positive ascending leader from a 70-m
tower on the San Salvatore Mount.
zone. With the data of section 2.4.1, we can calculate the leader tip potential
U,, at which the streamer zone will exceed at least twice the length of about
loom, a threshold value for the measuring equipment. For the average
streamer zone field E, = 5 kVjcm, the potential derived from (2.39) is
U, % 100 MV. Such a high value cannot be typical of lightning with average
parameters. There is another circumstance preventing streamer zone registra-
tions - different radiation wavelengths of a hot leader channel and a cold
streamer zone. Violet and ultraviolet radiation from streamers is dissipated
by water vapour and rain droplets in the air much more than long wavelength
radiation characteristic of a mature channel. At a distance of about a kilo-
metre between the lightning and the registration site (closer distances are
practically unfeasible), a streamer zone may become quite invisible to the
observer's equipment. Note that this is totally true of optical registrations
of an ascending positive leader.
A schematic streak picture of an ascending positive leader, based on 18
successful registrations [22],is shown in figure 3.3. All lightnings started from
a 70-m tower on the San Salvatore Mount near the Lake Lugano. The leader
does not exhibit specific features that would distinguish it from a long labora-
tory spark. On the whole, it developed in a continuous mode with irregular
short-term enhancements of the channel intensity. Normally, they did not
accelerate the leader development. Something like this has been observed
in a long laboratory spark. The streak picture in figure 3.4 demonstrates
this with reference to a positive leader in a sphere-plane gap 9 m long. But
this phenomenon has nothing to do with a stepwise elongation of a negative
leader channel.
The velocity of an ascending positive leader near the starting point is
close to that of a laboratory spark, about 2 x 104m/s. From some data
[22], it was in the range of (4-8) x 104m/s for a channel length of 40-
100m; but when the leader tip was at a height of 500-1 150m, it increased
by nearly an order of magnitude, to 105-106 mjs.
Figure 3.4. A streak photograph of the initial stage of a positive leader in a 9-m rod-
plane gap, displaying short flashes of the channel.
I I I I I 1
b
0 5 10 15
20 25 Yp
Figure 3.6. Typical integral velocity distributions for descending leaders of the CY- and
P-types (from [24, 251).
length varied between 10 and 200 m with the average value being 30 m. The
duration of a step is likely to be within several microseconds. The available
streak photographs are not good enough to identify the details of a step.
In any case, it is hard to decide whether it is similar to a step of the long
negative spark described in section 2.7.
A stepwise negative leader approaches the earth at an average velocity
of 105-106m/s. Two descending leader types can be identified in terms
of their velocity: slow a-leaders and fast P-leaders. The former travel
at their step-averaged velocity; it varies with the discharge in the range of
(1-8) x105m/s with the average value of 3 x 105m/s. The respective P-
leader parameters are 3-4 times higher. This can be seen in figure 3.6 showing
the integral velocity distributions described in [24,25]. Usually, P-leaders are
more branched and their steps are longer. They abruptly slow down when
they approach to the earth, after which they behave as a-leaders.
An ascending negative leader also has characteristic steps. Most of the
13 registered leaders ascending from a 70-m tower on the San Salvatore
Mount [22] were identified as a-leaders. They have relatively short steps
(5-18 m) and a velocity of (1.1-4.5) x lo5mjs. Two of the discharges were
referred to @-leaders because their velocity was (0.8-2.2) x lo6 m/s and the
step length up to 130m. On the whole, ascending and descending stepwise
leaders do not show significant differences.
The registrations of ascending discharges from the San Salvatore Mount
provide direct evidence for the existence of a streamer zone in a lightning
leader. Registrations made at a sufficiently close distance, which became
Figure 3.7. Schematic diagram of the initial development of a leader ascending from
a 70-m tower on the San Salvatore Mount, as viewed from close distance streak
photographs.
possible due to the tower top being the only starting point, show streamer
flashes arising at the moment a new step begins. Streamers were initiated
not only from the tip of the main channel but also from its branches
(figure 3.7).
Leaders of lightning components following the first one are known as dart
leaders because of the absence of branches. The streak photograph in
figure 3.8 shows the trace of only one bright tip looking like a sketch of an
arrow or dart. A dart leader follows the channel of the previous lightning
component with a velocity up to 4 x lo7mjs. Averaging over many registra-
tions gives the value (1-2) x 107m/s, with the minimum values being an
order of magnitude less than the maximum one [23,25]. The dart leader
velocity does not vary much on the way from the cloud to the earth.
We can only guess about the values of descending leader currents or estimate
them from indirect data. We shall make such estimations in section 3.5, using
leader charge data, also obtained indirectly. Ascending leader currents are
not difficult to measure, and there have been many measurements of this
kind. Normally, a current detector is mounted on top of a tower dominating
the locality [28-301. The current impulse of an ascending leader, registered on
an oscillogram, lasts for about 0.1 s, corresponding to the time of ascending
leader development. The current nearly always rises in time (figure 3.11). The
current supplies an elongating leader with charges. Physically, these charges
are induced by the electric field of a cloud. When a leader approaches a cloud,
going through an increasingly higher field, the linear density of induced
charge T increases. Besides, the leader goes up with an increasing velocity
V,, reducing the time for the charge supply. A combination of these factors
raises the current i = 7VL. At the moment an ascending leader starts its
travel, its current is lower than 10A, whereas at the end of the travel, it
may rise to 200-600 A, with an average value of about 100 A. Sometimes,
just before the leader begins its continuous elongation, impulses with an
amplitude of several amperes may arise against the background of a milli-
ampere corona current.
Current oscillograms of an ascending leader triggered from a thin wire
elevated by a small rocket to 100-300m [13,31] give a similar picture.
They show the same slowly rising impulse with an amplitude of 100-200A
and duration 50-looms. It has no overshoots at the front, even if the
leader goes up in a stepwise mode.
There are no reasons to suggest any principal difference between average
currents of ascending and descending leaders. In both cases, the leader is
supplied by charges induced by the electric field of a storm cloud, and the
leader lifetimes are comparable because they move at approximately the
same velocity.
Qualitatively, the current variation of the first component leader is
similar to that of a laboratory spark. When the gap voltage is raised
slowly, one can observe initial leader flashes at the high-voltage electrode,
followed by distinct current impulses [32]. As for long spark steps, they
practically do not change the current at the leader base. It has been shown
The subdivision of experimental data between this and the previous section is
somewhat arbitrary. Electric field measurements provide information about
leader charge, while charge and current are related by leader velocity. On the
whole, this is a general problem. If the observations were arranged properly
and the data analysis was made carefully, relatively simple field measure-
ments can add much to our knowledge of electrical parameters of lightning.
The knowledge of the field itself is rarely of importance, probably, except in
some applied problems of lightning protection of low voltage circuits. For
this reason, it is not the measurements but, rather, methodological
approaches to their treatment which are significant. So we shall begin with
these approaches and the general principles underlying a treatment of most
lightning stages.
-
Figure 3.12. Measurement of fast variations in the electric field during the lightning
development.
Suppose the electric field at an observation point near the earth’s surface
is Eo(0) at the moment of the lightning start. When the discharge is com-
pleted, the field takes a new value Eo(tm).The field has changed by the
value AEo = Eo(tm)- Eo(0)over the time t,. When the measurement time
is relatively short, it is more convenient to register the field change rather
than to measure its values. This is usually done with electrostatic antennas,
i.e., metallic conductors (normally, flat) grounded through a reservoir capa-
citor C (figure 3.12). If the capacitor and the measurement circuits connected
to it have an infinitely high leakage resistance RI,the capacitor voltage, at
any moment of time, is
(3.3)
where Sa is the area of a flat antenna and qc is the charge induced on it. If RI
is finite (which is always the case due to the input resistance of the circuit
taking voltage readings from the capacitor), the use of (3.3 ) requires the con-
dition RIC >> t,, which can be easily met for the lightning duration ~ 1 0 - s*
but becomes problematic for a time interval of several minutes between two
flashes.
An accurate measurement of the field change AEo(t,) requires the
necessary time constant of the measurement circuit RIC and the account of
effects of external field variation in the atmosphere, Eo, by making allowance
for local effects. (The antenna may be raised above the earth, say, mounted
on a building roof, so that the field there will be higher than on the earth. On
the other hand, a nearby high construction may reduce the field, acting as an
electrostatic screen.) The field value obtained is not particularly informative.
In order to get information about the lightning discharge, we have to make
certain assumptions concerning the distribution of charges which have
changed the field.
Let us begin with a simple illustration. Suppose a lightning leader
passing from a spherical volume has changed the charge of only one sign
in a storm cloud cell. If there are other changes in the sphere charge during
the leader travel, they are assumed to have been completely neutralized later,
at the return stroke stage. If this assumption is correct, the measured value of
A E o ( t m )can give an idea about the quantity of charge transported by the
leader from the cloud to the earth:
+
2 7 r ~ o ( H ~R2)3/2AEo(tm)
AQM = H (3.4)
Here, H is the altitude of the charged storm centre and R is its radial dis-
placement relative to the registration point. Both parameters should be
measured by an independent method or simultaneous field registrations at
two more points should be made at given distances from the first one. This
will provide additional equations for the unknown values of H and R.
Such an unambiguous treatment results from the simple model we have
chosen, which contains no geometrical parameter except for the distance to
the charge. However, a slightly more complicated, dipole model deprives the
measurement treatment of this advantage. Still, electric field measurements
have always been attractive to lightning researchers owing to their
simplicity. Interest in such measurements increased with the application of
lightning triggering by small rockets raising a grounded wire to 150-300 m
above the earth’s surface (triggered lightning). The first component of such
lightning is genuinely artificial, but then the first trace channel is used by
practically natural dart leaders travelling to the earth. Their point of contact
with the earth is predetermined, so field detectors can be placed at any dis-
tance from the leader. This registration system is quite sensitive and capable
of responding to the linear charge density not far from the leader tip when it
approaches the earth.
To illustrate our analysis, we shall use the field measurements described
in [34,35]. The authors of this work kindly made them available to us after
their discussion at the IXth International Conference on Atmospheric Elec-
tricity, held in St. Petersburg in 1992. The files contained detailed records of
electric fields, taken during the flight of dart leaders, and of their return
stroke currents. Detectors were placed at the distance of R = 500m and
30m from the contact point. Regretfully, the recordings at these distances
were not simultaneous but made in different years. Their comparison is
still possible because the fields were recorded at the same time as the
return stroke currents. By sorting out identical current oscillograms, one
can select lightning discharges with about the same leader tip potentials.
This provides close values of leader velocity and linear charge density in
the charge cover not too far from the leader tip. Some representative oscillo-
grams of AE(t)/AEmax normalized by their amplitudes are shown in
figure 3.13. They correspond to discharges with really close currents in the
return strokes (IM = 6 kA at point R = 500 m and 7 kA at point 30 m). The
amplitude values of field variation AE,,, over the time of the dart leader
1.0-
I
0.8 -
0.6 -
w
0.4-
10 20 30 40
Time, ps
Figure 3.13. Oscillograms taken in Florida, USA [35], from the vertical field com-
ponent during the development of the dart leader in the subsequent component of
a triggered lightning. The detectors were positioned at 30 and 500 m from the point
of strike; the pulses are related to their maximum amplitudes.
travel were 6.9 V/cm and 120 V/cm, respectively. Note that the measure-
ments in [35] result in AEm,,/IM M const at every point. There is no
geometrical similarity between the pulses AE( t)/AEmax at the different
points. On the contrary, there is a sharp difference in the rates of strength
rise, as a dart leader was approaching the earth. The field increase in the
range (0.5-1.0)AEm,, took Atl12 = 76ps for point R = 500m and only
5 ps for point R = 30 m.
These data will be treated in terms of a simple model, in which a dart
leader is represented as a uniformly charged axis with linear charge density
rL. Naturally, the real cover radius R, can be ignored in the field calculation
at a distance R.We shall show below that field calculations can only take into
account the charge distribution along a relatively short length behind the tip,
comparable with R. This will justify the assumption of rL being constant,
because it actually refers to a short length of about R near the tip. Therefore,
the field change due to the leader charge at point R at the earth, with the
allowance for its mirror reflection by the earth, is described as
where h is the height of the leader tip from the earth at the moment of regis-
tration and H is the height of the leader base. The field change is maximum
when the tip contacts the earth, and for R << H this gives
It indeed follows from (3.6) that only the charge distribution along a short
length comparable with R is important for field evaluation. (For example,
at H > 5R, the error of the model with rL= const will be less than 20%
for any charge distribution, unless TL grows rapidly from the tip toward
the base; but there is no reason for this, because the channel field E, is
weak and the cloud potential does not vary much.)
Formula (3.6) allows charge density evaluation with a good accuracy,
since the leader is strictly vertical at the earth - it reproduces the path of
the rocket taking up the wire which has evaporated. The value calculated
from the measurements at point R = 30m appears to be unexpectedly
small: rLx 2 x lop5Cjm. Nearly as much charge is transported by long
laboratory sparks (section 2.4). The potential of a lightning leader tip,
U,, does not seem to be much larger than that of a laboratory spark.
According to (2.8) and (2.35), the linear leader capacitance is
C1 FZ 2mo/ln ( H / R L )x (2-10) x F/m, even with indefinite leader
radius R L . From this, we have U, x rL/C1x 2-10MV.
The velocity of a dart leader proves to be very high. For its evaluation, we
shall use the measured value of A t l p ,which is 5 ps for R = 30 m. Formula (3.5)
gives A E = AE,,,/2 at h = J?;R. Hence, the average velocity along a path of
length h FZ 50m at the earth’s surface is VL FZ f i R / A t 1 / 2= lo7 m/s, quite
consistent with direct measurements. It should be emphasized that this velocity
refers to the perfectly vertical path at the earth’s surface, so it is the true velocity.
Similar evaluations can be made with the measurements at the far point
R = 500 m but with a lower reliability, since the parameter averaging is to be
made over a leader length of about lo3m with an unknown path. Neverthe-
less, the values of T~ = 2.3 x Cjm and VL = 1.15 x lo7 mjs are found to
be close to those above. It will be shown in the next section that an indefinite
trajectory may produce an error much larger than the obtained difference in
the values of TL and VL. So the dart leader of triggered lightning with the
definite path at the earth is a lucky exception.
Another illustration of A E ( t ) ,cited in [35], characterizes a more power-
ful dart leader. The current amplitude in the return stroke was as high as
40 kA. The maximum field change was found to be AE,,, = SlOVjcm,
i.e., a little more than a value proportional to current, while the characteristic
time of the process, A t l p , decreased to 1.8 ps. Calculations similar to those
described above give T~ x 1.35 x 10-4C/m, U, =20-30MV, and
VL = 2.9 x 107m/s, thereby supporting the hypothesis of a direct, though
not very strong, dependence of the leader velocity on the tip potential. For
the calculated values of linear charge and velocity at the earth, the leader cur-
rent is found to be iL = T~ VL = 3.9 kA, only an order of magnitude lower
than the current amplitude in the return stroke.
<” 0.00-
w
U
z
WO
e
-
-0.02
N
I I
-0.04-
-0.06-
.
Figure 3.14. Electric field variation at the earth, evaluated for a large distance from
the vertical channel of a descending leader.
AE(L)= -
AQH
+
27rq,(H2 R2)312 2%
+‘“I (H-x)dx
o [ ( H- x)’ + R2I3/’’ (3.7)
This expression takes into account the doubled field associated with the
earth-induced charge. The evaluation of the integral gives the known
expression
AE(L)= -
[ ( H- L)’
1 1
+ R2I1/’ - (H’ + R’)”’
-
(H‘
LH
+ R2)3/2
}. (3.8)
t This component may be ignored in field measurements at a small distance from the leader, as
described in section 3.5.
0.054
L/H
Figure 3.15. Electric field variation at the earth for a leader deviating from its vertical
path; R I H = 0.7.
the charge density here (along a length of the rod radius) rises rapidly. The
charge distribution at the rod base can be approximated as T ( X ) = ax,
where the coordinate origin x is at the rod centre. The reader will soon see
that the contribution of the end charges fq is small at larger distances, if
the rod length 2d is much greater than its radius r.
For simplicity, let us assume that the charge is concentrated on the outer
rod surface, as is the case when it has a conductivity, and that the potential
will be calculated along the longitudinal axis. The rod centre will be taken as
the zero point of the external field Eo potential. The potential p at the point x
is a sum of potentials created by the external field, -Eox, the end charges, pq,
and the charges distributed along the rod, pr:
- + + r2]1/2 + x In
[(d x12
(d - x )
-(d
+ [(a- x12 + r2]'12
+ x ) + [(d + + r 2 p 2
X)* 1
(3.10)
Here, the last approximate expression refers to the rod sites lying far from its
ends, Id f X I >> r. Here, the term pqcan be neglected, and we shall approxi-
mately have pT - Eox = 0 . With the actual charge distribution ~ ( xand ) the
end charges providing pq, the rigorous equality p ( x ) = 0 should be valid
along the whole rod length. By relating the approximate equality to the
centres of the semi-axes x = f d / 2 , we find
(3.11)
The potential at the rod ends must be calculated with the account of
their higher charges. Assuming this charge to be concentrated along the
end circumference, the potential at the centre of the end plane (at the
points x = f d on the axis) can be described as
(3.12)
smaller than the charge distributed over each half (it is an order of magnitude
smaller for d / r z 100). Therefore, the account of localized tip charge may be
necessary only for the calculation of electric field in the region close to the tip,
at a distance less than 10r from it, In a remote region, where measurements
are usually made, such a subtlety is unnecessary - it is sufficient to consider
only the charge distribution along the leader channel. Clearly, this is not a
uniform distribution, taken for granted by some researchers.
It is time to look at the shape of a field strength pulse at the earth, deter-
mined by the charge of a linearly polarized vertical axis with charge
=ra
~ ( x=) f a x per unit length. It is defined by the algebraic sum of terms
from the positively and negatively charged semi-axes and is equal to
AE(L) x ( H - X) dx
-L 2 r E o [ ( H- x)’ + R2I3/’
where L are the lengths of leader sections which have moved away from the
starting point to the earth and upwards. Integration with (3.1 1) gives
L L
AE(L) =
In (fiL/er)
Eo [ [ ( H - L)’ + +
R2]‘/’- [ ( H L)’ + R2I1/’
- In
+
[H (H’ R2)”’]2 +
{H -L + +
[ ( H- L)* R2I1/’}{H L [ ( H L)’+ + + + R2]’/’J
(3.14)
Here, H is the height of the leader start, r is its radius, and R is the
distance between the leader axis and the observation point. It can be
shown that the function A E ( L ) of (3.14) rises smoothly with L and has
no extrema.
The linear charge distribution assumed in the above illustration is, of
course, another speculation (section 4.3). Moreover, a leader goes up and
down non-uniformly, and the field in the earth-cloud gap is far from
being uniform: its strength decreases towards the earth. This limits the
linear charge growth from the start downward. The finite channel conductiv-
ity exhibits similar behaviour, reducing the tip potential. So it is impossible to
find the actual charge distribution exactly without knowing these param-
eters. Thus, a processing of field oscillograms can give nothing more than
what they actually show. The field at a point is an integral effect of the
whole combination of charges created or transported by a given moment
of time. It is probably worth speculating about registrations but one
should assess the results soberly, considering all possible variants and
insuring oneself whenever possible. The best insurance is, of course, to
increase the number of registration points and parameters determined by
independent methods.
All lightning hazards are associated with the return stroke, and this accounts
for the great effort of investigators to learn as much as possible about this
discharge stage. It has been established that the contact of a descending
lightning leader with the earth or a grounding electrode produces a return
wave of current and voltage. It travels up along the leader channel, partially
neutralizing and redistributing the charge accumulated during the leader
development (figure 3.18). The travel is accompanied by an increased light
intensity of the channel, especially at the wave front. At the earth, the
wave front intensity acquires its maximum over 3-4ps [31]. As the wave
goes up to the cloud, the wave intensity steepness and amplitude decrease
many-fold, indicating a considerable decay. Judging by streak pictures, the
region of a high light intensity at the wave front extends to 25-1 10m. The
whole wave travel takes 30-5Ops. This time is especially convenient for
electron-optical methods of streak photography. However, available
attempts to use such methods can hardly be considered successful. A serious
obstacle is the exact synchronization of a streak camera and lightning contact
with the earth. Although there are many synchronization methods, they have
no simple technical solutions and are seldom used in lightning experiments.
Continuous (e.g. sinusoidal) electron streak photography has not justified
hopes. Basic results on return stroke velocities have been obtained using
cameras with a mechanical image processing, which do not need synchroni-
zation (Boyce camera).
Figure 3.18. Scheme of the return stroke propagation after the contact of a descend-
ing leader with the earth (at moment t = 0). A leader brings potential U < 0; ZM is
return stroke current.
1.0 -
0.8 -
-3
.-O 0.6 -
cd
D
& 0.4 -
0.2 -
0.01 - . * . I 1
Initial currents ilo and i20 at the stage when current I ( t ) is stabilized to Z
are generated over a very short time equal to the I risetime. The branch
currents, therefore, rise from zero very quickly. The reactive components of
voltage drop -di/dt produced by them are much larger than the ohmic
ones -i that can be neglected for the time being. Hence, we have
i1o/i20 = (L2- M ) / ( L 1- M ) , and the initial current, say, in the first branch
is ilo = Z(L2- M ) / ( L 1+ L2 - 2 M ) . When the transitional process, whose
duration is defined by the time constant At = (Li + L2 - 2 M ) / ( R I+ R 2 ) ,
is over, currents ilx = IR2/(R1+ R2) and i2x = ilxR1/R2 are established
in the circuits. The durations of lightning currents are usually comparable
with the time constant At. Therefore, a magnetic detector placed in one of
the branches will register a current intermediate between the initial and estab-
lished values having a maximum amplitude, since the residual magnetization
of the rod contains information only about the maximum magnetic field of
current. For this reason, one can calibrate a magnetodetector for deriving a
full current amplitude only if the impulse shape is known. This cannot be
done in a real experiment, so one has to resort to a rough estimation of current
distribution over metallic constructions and use it in data processing.
Second, the operating range of the rod magnetization curve is not large,
and the transition from a linear to saturation region may produce additional
errors in data processing. To avoid saturation, the magnetodetector is placed
far from the conductor, which creates difficulties in data processing of light-
nings with low current and low magnetic field. Besides, when the distance
between a conductor and a detector is large, the magnetic field effects of
other metallic elements with current are hard to take into account. So a
100% error does not seem too high for magnetodetectors, even when several
detectors are placed at different distances from a current conductor. Their
records provide sufficient material for engineering estimations or for a
qualitative comparison of storm intensity in different regions, but they are
insufficient for theory. Organization of direct registrations takes much time
and effort. There are no more than a hundred successful registrations
made over a decade. Let us see what information can be derived from them.
Current impulse amplitudes vary widely, from 2-3 to 200-250 kA.
Some magnetodetector measurements give even 300-400 kA, but these
amplitudes seem doubtful. According to [42,46], the integral amplitude dis-
tributions for the first and subsequent lightning components obey the so-
called lognormal law, in which it is current logarithms, rather than currents
themselves, that meet the normal distribution criterion. The probability of
lightning with a current larger than ZM, is defined as
(3.16)
where (lg I)av is an average decimal logarithm of the currents measured and
olg is the mean square deviation of their logarithms. This approximation
cannot be considered accurate. The relative deviation of the value of (3.16)
from the real one may be several tens percent; it may be even more for prac-
tically important current ranges. Nevertheless, lognormal distributions allow
measurement comparison and serve as a guide to engineering estimations.
For example, about 200 current oscillograms for lightnings that struck the
70m tower on the San Salvatore Mount in Switzerland [42] satisfactorily
obey the lognormal law with (lgZ)av= 1.475 and olg= 0.265 for the first
component currents of a negative lightning discharge. This means that the
50% current value is estimated to be 30 kA; 95% of lightnings must have
currents exceeding 4kA and 5% of lightnings 80kA. The probability of
higher currents rapidly decreases: 100 kA is expected in 2% of cases and
200kA in less than 0.1% of cases (figure 3.20). It should be emphasized
again that the distribution boundaries must be treated with caution. The
curve shape in the low current range strongly depends on the sensitivity of
the measuring instruments used (its left-hand limit is usually taken to be
1-3 kA in distribution plots). There are few measurements in the high current
0.01
I ' I . I . I . 1
-I Eo -
Figure 3.22. The distribution of current impulse parameters in the return stroke,
based on oscillograms.
charged region of a cloud. But we can only speculate about the nature of this
process producing final current of 100 kA and ask why it is manifested only in
positive lightnings.
double time of the wave run along the channel. After passing the zero point,
the pulse part opposite in sign rises and then decreases with nearly the same
rate; its amplitude is 2-3 times smaller that the first ‘half period’.
The inverse proportionality of radiation components to the distance
from the radiation source was the reason why measurements are presented
in the above form: they are normalized to the basic distance Ybas = 100 km
as E:,, = 10-5Em,,r with r in metres. For the first lightning component,
the average values of the initial pulse peak of the vertical component,
EA,,, lie within 5-10 V/cm [49-541 (compare: radio receivers detect well
signals of lmV/m in the medium bandrange). The electric component of
subsequent lightning components is 1.5-2 times smaller. The spread of
measurements is as large as that of lightning currents. The standard deviation
oE is in the range 35-70% for the first lightning component and 30-80% for
the channel. These are appreciable changes, since current of lOOA extracts,
from a cloud, charge A Q x 1OC over the time 0.1 s. The field on the earth
right under a cloud changes by the value A E = AQ/(27qH2) M 200V/cm if
the height of the charged cell centre is H = 3 km; at distance r = 10 km from
the lightning axis, A E = A Q H / [ 2 q , ( H 2+ Y~)~'~] zz 5V/cm. Similar values
were registered during observations.
Continuous current flow is accompanied by slowly rising and as slowly
decreasing current impulses with an amplitude up to 1 kA. These are M-
components of lightning. The risetime of a typical M-component is about
0.5ms, an average impulse duration (on the level 0.5) is twice as much, an
average amplitude is 100-200 A, although M-components with current up
to 750 A have also been registered [56,57]. Pulsed current rise is always
accompanied by an increase in light emission intensity of the whole channel,
from the cloud down to the earth. Streak photographs (even taken slowly) do
not show the propagation of a well defined emission wave front similar, say,
to the tip of a dart leader. It seems as if most of the channel flares up
simultaneously, although excitation, no doubt, propagates down from a
cloud with a high velocity, (2.7-4)x lo7mjs (from measurements of [58]).
Two M-components were identified in [58] as ascending ones. In later
measurements, the existence of ascending processes were questioned, because
there were no clear physical reasons for the appearance of an inducing
perturbation at the earth's surface.
Variations in current and electric field of M-components were registered
in triggered lightning flashes at a short distance from the channel (r = 30 m)
[57]. The field variation of a vertical component at the earth is shown in figure
3.24. The pulse A E rises to its maximum 70ps earlier than the current
impulse. The field rises and decreases at nearly the same rate. The pulse
component of field perturbation is nearly completely damped while the
current still has a high amplitude.
I- --
2?0 I 400 600
b
m t, PJ-
,e-
0.5
loO1
1.5
E, kV/m
Figure 3.24. Superimposed schematic oscillograms of M-component electric field and
current at the earth [ 5 7 ] .
probabilities, it exceeds, respectively, 20, 80 and 350C. One cannot say that
the charge transported by a flash is very large. For comparison, even a very
large lightning charge of 350C flows through the arc of a conventional
welding unit for 3-5 s.
Charge transport is accompanied by energy release. An average negative
flash with a charge Q = 1OC and gap voltage 50 MV dissipates about QU =
5 x 10sJ, which is equal to the energy released by a 100 kg trinitrotoluene
explosion. While most energy is released within the lightning trace, the
problem of energy release and heating of metal constructions is of much inter-
est. Normally, the resistance of metallic conductors and that of a grounding
electrode are much less than the equivalent resistance of a lightning channel
RI = U / I M (IM is the impulse amplitude of a return stroke); RI = 1 kR if
U x 50 MV and ZM = 50 kA. Therefore, lightning can be regarded as a current
source, assuming that current IM is independent of the object’s resistance. Any
conductor with lightning current flow releases the energy
K =R 1; i2 dt (K/R)R. KIR = 1; i2 dt
other impurities. No doubly charged ions have been detected, indicating that
the temperature does not exceed 30 000 K. Measurements of time resolved
NI1 (N') line intensities show that the return stroke temperature reaches
30 000 K for the first 10 ps [59,62] and drops to 20 000 K in 20 ps. Average
temperatures are estimated to be about 25 000 K. These results are obtained
assuming that a plasma channel is optically transparent and that the
excitation of atoms in the plasma is equilibrium (of the Boltzmann type).
The estimations justify this assumption.
Electron densities found from the Stark broadening of the Ha lines
are 1 0 ' * ~ m -for
~ the first 5ps of the stroke life. Under thermodynamic
equilibrium conditions at T = 30 000 K, this value of ne corresponds to
the pressure of 8atm [63]. About lops later, ne decreases to l O ' ' ~ m - ~ ,
corresponding to the pressure drop down to the atmospheric pressure.
Then the value of ne remains unchanged over the time of the N I I line
registration. This does not seem strange. Equilibrium electron density in
air at p = const = 1 atm changes only slightly in a wide temperature
range 15000-30 000 K, remaining about 1017~ m - As ~ . the channel cools
down, the ionization degree x = n e / N certainly decreases, but when the
pressure reaches the atmospheric value, the gas density N rises simul-
taneously. For this reason, ne = x N does not change much. High intensity
radiation is observed for about loops (from 40 to lOOOps). The first
peak is often followed by another one several hundreds of microseconds
later.
Spectroscopic measurements were mostly made during a return stroke,
but some authors [64] managed to register the spectrum of a 2-m portion of a
stepwise leader. The leader tip temperature calculated from the N 11 lines lies
within 20 000-35 000 K. The diameter of the radiation region is less than
35 cm. More accurate evaluations are unavailable. It seems unlikely that
this temperature is characteristic of the whole leader channel. Rather, the
experiment registered a short temperature rise during a powerful step
which was akin to a miniature return stroke (section 2.7). The step-induced
perturbation involving part of the channel region is most likely to be
damped rapidly along the leader length.
It is not only the plasma dynamics but the channel radius, too, which
still remains enigmatic. In making evaluations of the radius, one usually
relies on photographs. But in this case, it is very important to agree on the
kind of radius being evaluated. This may be the radius of the channel,
through which current flows during the leader and stroke stages. Clearly,
such a radius will include the best conducting and, hence, the hottest core
of the plasma channel. Or, one can follow another approach. When solving
the problem of electric field variation during the lightning development, one
has to deal with the radius of the leader cover where most of the space charge
is concentrated. This is the charge radius of lightning. Therefore, each time
we speak of radius, we must define exactly what we mean.
Here, we shall use the concept of channel radius as applied to the region
where the lightning current is accumulated and the concept of cover radius to
the region where most of the space charge is concentrated. The former can.be
determined, to some extent, by using optical methods, although this is a com-
plicated task. With reference to the optical measurements [65], one usually
deals with radii of several centimetres. This resolution is accessible to
modern cameras at a distance of about a kilometre, but the cameras must
have the highest resolution possible. Anyway, we have never heard about
the application of such perfect optical equipment in lightning research.
In addition to using special-purpose optics, the experimentalist must
match perfectly the sensitivity of photographic materials and exposures.
A longer exposure produces a halo, increasing the actual radius. Unless
special measures are taken, the error may be very large, especially for
flashes with a high light intensity. For some reasons, the optical radius of
a lightning channel may exceed manifold the thermal radius. Such an
effect was observed in studies of spark leaders in laboratory conditions
[66]. Registration of the thermal radius appears problematic even for trig-
gered lightning, with a fixed point of contact with the earth. For natural
lightning, this task is much more complicated. As for the cover radius,
there is no reliable technique for its registration at all. So lightning radius
measurements cannot provide unquestionable data, and the researcher is
to rely on theoretical evaluations only.
It was not our task to review all experimental studies on lightning: this has
been well done in [l, 591. We believe that the latest experimental data will
be presented in a new Uman book now in preparation. But the basic facts
have been discussed here, and we can now ask ourselves whether the
available data are sufficient to build lightning theory and to check it by
experiment.
The situation with lightning is somewhat similar to that for a long
laboratory spark, i.e., experiments give mainly external parameters of a dis-
charge. In the laboratory, these are velocities of the major structural elements
(streamers and leaders), their initiating voltages, currents, transported
charges, and, possibly, some other characteristics Sometimes, we have
some information on channel radii, or on the time variation of radii, or
scarce data on plasma parameters. But that is all.
The arsenal of lightning researchers is much smaller. First, they have no
information about the voltage in the cloud-earth gap at the lightning start,
and there are no data on the initial distribution of electric field. Both literally
and figuratively, the bulk of a storm cloud, where a descending leader
originates, is obscure. Measurements made at the earth’s surface cannot
References
[l] Uman M 1987 The Lightning Discharge (New York: Academic Press) p 377
[2] Golde R H (ed) 1977 Lightning (London, New York: Academic Press) vols 1, 2
[3] Imyanitov I M 1970 Aircraft Electrization in Clouds and Precipitation (Lenin-
grad: Gigrometeoizdat) p 210
[4] Gunn R 1948 J . Appl. Phys. 19 481
[5] Gunn R 1965 J . Atmos. Sci 22 498
[6] Evans W H 1969 J . Geophys. Res. 74 939
[7] Winn W P, Schwede G W and Moore C B 1974 J . Geophys. Res. 79 1761
[8] Winn W P, Moore C B and Holmes C R 1981 J. Geophys. Res. 86 1187
[9] Kazemir H W and Perkins F 1978 Final Report, Kennedy Space Center Contract
CC 69694A
[lo] Newman M M, Stahmann J R, Robb J D et all967 J . Geophys. Res. 72 4761
[ l l ] Kito Y, Horii K, Higashiyama Y and Nakamura K 1985 J . Geophys. Res. 90
6147
[12] Hubert P and Mouget G 1981 J . Geophys. Res. 86 5253
[13] Hubert P, Laroche P, Eybert-Berard A and Barret L 1984 J . Geophys. Res. 89
251 1
[14] Idone V P and Orville R E 1984 J . Geophys. Res. 89 7311
[15] Fisher R G, Schnetzer G H, Thottappillil R et a1 1993 J . Geophys. Res. 98 22887
[16] Wang D, Rakov V A, Uman M A et a1 1999 J . Geophys. Res. 104 4213
[17] Malan D J and Schonland F G 1951 Proc. R. Soc. London Ser. A 209 158
[18] Malan D J 1963 Physics of Lightning (London: English Univ. Press) p 176
[19] Malan D J 1963 J . Franklin Inst. 283 526
[20] Bazelyan E M and Raizer Yu P 1997 Spark Discharge (Boca Raton: CRC Press)
p 294
[21] Chalmers J A 1967 Atmospheric Electricity (2nd edn) (Oxford: Pergamon) p 418
[22] Berger K and Fogelsanger E 1966 Bull. S E V 57 13 1
[23] Schonland B 1956 The Lightning Discharge. Handbuch der Physik 22 (Berlin:
Springer) p 576
[24] Schonland B, Malan D and Collens H 1938 Proc. Roy. Soc. London Ser. A 168
455
[25] Schonland B, Malan D and Collens H 1935 Proc. Roy. Soc. London Ser. A 152
595
[26] Jordan D M, Rakov V A, Beasley W H and Uman M A 1997 J . Geophys. Res.
102 22.025
[27] Fisher R G, Schnetzer G H, Thottappillil R et a1 1992 Proc. 9th Intern. Conf. on
Atmosph. Electricity 3 (St Peterburg: A I Voeikov Main Geophys. Observ.) p 873
[28] McCann G 1944 Trans. A I E E 63 1157
[29] Berger K and Vogrlsanger E 1965 Bull S E V 56 No 1 2
[30] Bazelyan E M, Gorin B N and Levitov V I 1978 Physical and Engineering Funda-
mentals of Lightning Protection (Leningrad: Gidrometeoizdat) p 223 (in Russian)
[31] Saint Privat d’Allier Research Group 1982 Extrait de la Revue Generale de
I’Electricite, Paris, September
[32] Gorin B N and Shkilev A V 1974 Elektrichestvo 2 29
[33] Idone V P, Orville R E 1985 J . Geophys. Res. 90 6159
[34] Rubinstein M, Uman M A and Thomson P 1992 Proc. 9th Intern. Conf. on
Atmosph. Electricity 1 (St Peterburg: A I Voeikov Main Geophys. Observ.) p 276
[35] Rubinstein M, Rachidi F, Uman M A et a1 1995 J. Geophys. Res. 100 8863
[36] Thomson E M 1985 J . Geophys. Res. 90 8125
[37] Kolechizky E C 1983 Electric Field Calculation for High-Voltage Equipment
(Moscow: Energoatomizdat) p 167 (in Russian)
[38] Jordan D M and Uman M A 1983 J . Geophys. Res. 88 6555
[39] Schonland B and Collens H 1934 Proc. Roy. Soc. London Ser. A 143 654
[40] Idone V P and Orville R E 1982 J . Geophys. Res. 87 9703
[41] Idone V.P, Orville R E, Hubert P et a1 1984 J . Geophys. Res. 89 1385
[42] Berger K, Anderson R B and Kroninger H 1975 Electra 41 23
[43] Berger K 1972 Bull. Schweiz. Elekrtotech. Ver. 63 1403
[44] Gorin B N and Shkilev A V 1979 in Lightning Physics andLightning Protection
(Moscow: Krzhizhanovsky Power Engineering Inst.) p 9
[45] Gorin B N and Shkilev A V 1974 Elektrichestvo 2 29
[46] Eriksson A J 1978 Trans. South Afr. ZEE 69 (Pt 8) 238
[47] Anderson R B and Eriksson A J 1980 Electra 69 65
[48] Alizade A A, Muslimov M M et a1 1974 in Lightning Physics and Lightning
Protection (Moscow: Krzhizhanovsky Power Engineering Inst.) p 10
[49] Master M J, Uman M A, Beasley W H and Darveniza M 1984 ZEEE Trans. P A S
Pas-103 2519
[50] Krider E P and Guo C 1983 J . Geophys. Res. 88 8471
[51] Cooray V and Lundquist S 1982 J . Geophys. Res. 87 11203
[52] Cooray V and Lundquist S 1985 J . Geophys. Res. 90 6099
[53] McDonald T B, Uman M A, Tiller J A and Beasley W H 1979 J . Geophys. Res.
84 1727
[54] Lin Y T, Uman M A et a1 1979 J . Geophys. Res. 84 6307
[55] Krehbiel P R, Brook M and McCrogy R A 1979 J. Geophys. Res. 84 2432
[56] Thottappillil R, Goldberg J D, Rakov V A, Uman M A et a1 1995 J . Geophys.
Res. 100 25711
[57] Rakov V A, Thottappillil R, Uman M A and Barker P P 1995 J . Geophys. Res.
100 25701
[58] Malan D J and Collens H 1937 Proc. R. Soc. London A 162 175
[59] Uman M A 1969 Lightning (New York: McGraw-Hill)
[60] Orvill R E 1968 J . Atmos. Sci. 25 827
[61] Orvill R E 1968 J. Atmos. Sci. 25 839
[62] Orvill R E 1968 J . Atmos. Sci. 25 852
[63] Kuznetsov N M 1965 Thermodynamic Functions and Shock Adiabata for High
Temperature Air (Moscow: Mashinostroenie) (in Russian)
[64] Orvill R E 1968 J. Geophys. Res. 73 6999
[65] Orvill R E 1977 in Lightning, vol 1, R Golde (ed) (New York: Academic Press)
p 281
[66] Positive Discharges in Air Gaps at Les Renardieres - 197s 1977 Electra 53 31
Physical processes in a
Iight ning discharge
138
external field Eo, a field El x Eoh/r >> Eo is created by the induced charge at
the conductor top (see section 2.2.7). This field rapidly decreases in air (for a
distance of several r values), creating a potential difference between the
conductor end and the adjacent space, AU x Eoh. When the cloud bottom
is charged negatively and the vector Eo is directed from the earth up to
the cloud, the grounded conductor becomes positively charged, since the
field makes some of the negative charges leave the metal to go down to
the earth.
No stringent conditions are necessary for the field E , to initiate the air
ionization (at sea level El x E, x 30 kV/cm) or for a corona discharge to
arise at the pointed parts of a high structure (it is necessary to have
El x40-31kV/cm for r = 1-10cm). The conditions for a leader to be
initiated in the streamer corona stem are much more rigorous. The energy
estimations made in section 2.6 show that there is no chance for a leader to
arise if the leader tip potential U,, or, more exactly, its excess over the external
potential at the tip, A U = U, - U,, is less than AVrm,,x 300-400 kV. This
estimate is supported by experiments with leaders, whose streamer zones
have no contact with the electrode of opposite sign at the initial moment of
time. Therefore, for the desired potential difference A Ut,, to be produced,
the structure must have, at least, h x AUrm,,/EO x 20-30m if the average
field of the storm cloud at the site of the grounded object is -150 V/cm.
On the other hand, even if a leader is produced at such a low potential,
AVfm,,,it can hardly travel for a large distance. The leader current will be too
low to heat the channel to a sufficiently high temperature. As a result, the
channel resistance will be too high so that a very strong field will be required
to support the current in the channel. The channel field E, is, however,
limited by the external field Eo. Indeed, a grounded body of height h, from
which a positive leader has started, possesses zero potential. Having covered
the distance L, the leader tip acquires the potential U , = -E,L. Here, the
+
potential of the unperturbed external fields is U. = -Eo(L h), and we have
AUt = Ut - U0 = AU, + (Eo - E,)L. AU, = Eoh. (4.1)
For a leader to develop from the initial threshold conditions, the potential
difference A U , should not decrease relative to the initial value of AU,. For
this, the average channel field E, must be lower than the external field Eo.
However, a mature channel possesses a falling current-voltage characteristic
E, (i). A decrease in E, to 100 V/cm requires a channel current higher than
N
1A. We discussed this issue in sections 2.5.2 and 2.6. With the approximation
accepted there (E, x b / i and b =300 VAjcm), the leader current is to exceed
,,i = b/Eo x 2 A at Eo x 150 V/cm.
Let us see how large the potential difference A u t should be to make the
current exceed i,. In chapter 2, we derived formula (2.35) relating the
channel current behind the leader tip to the tip potential U, and the leader
velocity vL. That formula was applicable to the laboratory conditions
t The streamer theory has been advanced further. Note, for comparison, that the streamer
velocity is V, RZ AU, from formulae (2.6) and (2.8) with AU, instead of U, and E, = const.
the leader becomes longer, especially when the leader reaches the region of a
very high cloud field. The rising current heats the channel more, so that its
linear resistance and field drop. With time, the channel becomes a nearly
perfect conductor. Grounded at its base, the channel possesses the same
potential U ( t ,x) everywhere along its length, including the tip, which is
low relative to the absolute external potential IUo(x)I. In this case, the
value of A U = U ( t , x )- Vo(x)z -Uo(x)varies only slightly with time at
every point x along the channel.
The linear leader capacitance C1, given by formula (2.8) with length L
and cover radius R instead of 1 and Y, also varies very little. Indeed, the
cover radius behind the tip is about the same as the streamer zone radius
which, according to (2.39), is R = AUt/2E,, where E, M 5 kV/cm is the
streamer zone field under normal conditions. The height of the charged
region centre in the cloud, H PZ 3 km, is much greater than that of the
leader starting point, h PZ 200m. Suppose the leader length is greater than
h but smaller than H by such a value that the cloud field non-uniformity
along the channel can be neglected. We then have AU, M IUo(L)I M IEoLI,
and the value of LIR M 2E,/Eo under the logarithm in C1 of (4.2) and
(4.4) is independent of time, If this relation does change, which happens
when a leader rises so high that it enters the region of a rapidly increasing
external field, the logarithm changes much more slowly. Thus, the linear
charge ~ ( x M) C , A U remains nearly constant in time at every leader point.
But if there is no charge redistribution along the channel, the current in it,
i(t, x), does not change along its length but changes only in time. Entering
the channel through its grounded base, the current supplies charge only to
the front leader portion. The current in the base is the same as in the channel
right behind the tip. It is defined by formula (4.2) with Aut M I Uo(L)I,close
to the unperturbed potential of the cloud charge at the tip site. Similarly, the
leader velocity can be found from (4.3).
Therefore, the velocity and current of a fairly long leader (long relative
to the start height), which develops in the average field Eo, are described as
1‘ = 2TEoa(AUt)3’2
WL = a(ALJ,)1/2, AU, = EoL, e = 2.72.. . (4.5)
ln(2&E,/eEo) ’
The boundary conditions for the set of integral differential equations (4.6)
and (4.7) are described by the third equality in (4.6) and U ( 0 ,t) = 0, since
the leader base is grounded. Practically, it is convenient to subdivide the
channel into N fragments and consider the charge density in each fragment
to be dependent only on time, thus replacing the integral equation of (4.7) by
a set of linear algebraic equations. Each of them will relate the potential
U(xk) at the middle point xk of the kth fragment to the intrinsic and all
other linear charges. After integrating (4.7), one can easily see that radius
R enters logarithmically the factors of the set of equations (compare with
(4.2)), thereby justifying the use of linear leader charge r instead of its
cover space charge. The set of algebraic equations for U(xk) and '(Xk) is
solved in time at each step, and the progress is made by using equations
(4.6). We are presenting the result of this solution.
As the leader tip approaches the cloud, the external field at the tip site
becomes stronger and the ever increasing portion of the channel finds itself
in a strongly non-uniform field. Since the velocity and current are largely
defined by the potential Uo(L) at the tip site, formulae (4.5), in which Eo is
an average field, remain valid. In a simple model of a cloud with a spherical
unipolar charged region, the potential distribution in the space free from
charges is the same as for a point charge. If H is the height of the spherical
charge centre, Q,, the potential at height x at the point displaced from the
vertical charge axis for distance r (with the account of the mirror reflection
by the earth's plane) is
"{ 1
- }.
(4.8)
= 4T&o [ ( H - x ) 2 + r2I1l2 [ ( H+ + y2I1l2
Figure 4.1 presents the parameter calculations for an ascending leader
propagating in such a non-uniform field. The calculations were made from
300
- 1.4
3
- 1.2 e
E . E
- 1.0 *z
Y
~ 200
s
8 -8 3 .
-0.8 'g6
5
U 3
d
?
100 2 - 0.6 L
1 3
- 0.4 2
- 0.2
the set of equations (4.6) and (4.7), as described above. The current at the
channel base is defined by the total charge Q and the velocity by expression
(4.5):
For comparison, the current was also calculated from (4.5). The results show
the good accuracy of this simple formula, so the use of average linear
capacitance C1 can be considered justifiable in the calculation of
T ( L )= CIAUo(L)and in the case of a sharply non-uniform field.
from (4.8) for H = 3 km, the field near the earth under the charge centre,
EO= -Qc/27qH2 = 150V/cm, is created by charge Q, = -7.4C and
~Uo,,,I = 340MV for R, = 300m. From (43,a leader that has reached
the charged region centre acquires the velocity wL = 2.8 x 105m/s and the
current i,, = 5 kA (the field at the sphere boundary is EOmax = 7.5 kV/cm,
decreasing to zero towards the centre; in the current estimation, the
logarithm was taken to be 1). The lifetime of this high current is short,
about Rc/w~,,, lO-'s, with the total duration of the leader ascent of
N
-
+
1
+
+
[(H D - x ) ~ r2I1l2 [(H+ D
1
+ x ) +~ r2I1l2
}. (4.10)
Without allowance for the voltage drop across the channel (U, 0) the
leader tip will reach the point x,, where the absolute potential U. drops to
AU, E 400 kV, remaining negative as before. Since AUlmnis small relative
to huge potentials of charged regions (I Uolm,, N 100 MV), a positive ascend-
ing leader halts at a slightly lower height than the zero equipotential surface
of the external field. Because of the effect of charges reflected by the earth, the
zero potential line lies somewhat lower than the dipole centre. For example,
at D = H , which corresponds, more or less, to reality, we have x, = 1.4868
exactly on the vertical axis (Y = 0) instead of 1.5H, as would be the case with
a solitary dipole. With greater radial displacement Y, the zero equipotential
line comes closer to the earth, slowly at first but then more rapidly at
r > H (figure 4.2). This is the reason why ascending leaders taking different
vertical paths halt at different heights.
It has just been mentioned that the equipotential line U o ( x ,Y) = 0
corresponds to the maximum height attainable by a single ascending
leader. With allowance for the voltage drop across the channel, which
may appear appreciable in some situations, AU, drops to the threshold
value AUc, below the maximum height. This is supported by numerical
0.00
0.0 0.5 1.o 1.5 2.0
r/H
Figure 4.2. The zero potential line of a cloud dipole with the allowance for charges
reflected by the earth for D = H.
calculations made from the set of equations (4.6) and (4.7) and illustrated in
figure 4.3. They also indicate the leader retardation rate. As the leader
velocity decreases, the channel current becomes lower, causing the field E,
to rise. The tip potential decreases respectively, together with the potential
difference AU,, which limits the current still more, and so on.
800
Q 1.5
E
150 s
-, +
0
d 1.0 $ 600
a
.s
-
0
0
loo z 400 $
el
8
U
8 0.5 50
el 200
0.0 0 0
Leader length, km
Figure 4.3. Numerical simulation of an ascending leader propagating in a cloud
dipole field (Qc = 12 C, H = D = 3 km, r = 0.5 km), with allowance for the voltage
drop across the channel.
When a leader goes beyond the lower cloud charge region, the external
field changes its direction along the channel: below the negative charge
centre, x = H , its vertical component is directed upwards but above the
centre it is directed downwards. Correspondingly, the external potential U,
is non-monotonic and has an extremum at height H (absolute maximum).
The leader continues to develop beyond the maximum point, as long as
the relation
Aut ut(L) - Uo(L)= IUo(L)I- Iut(L)/> AUt,,,,,
is valid. But now, the leader velocity decreases continuously, because I Uo(L)I
drops with the leader elongation and I U,(L)I rises due to the rising channel
field. It is clear that a leader can develop successfully in any other direction,
since it is capable of propagating in the direction strictly opposite to the
external field. The calculations show the leader path along the equipotential
line in a zero external field. Here, AU,, i and wL decrease slowly, only due to
the greater voltage drop across the channel; otherwise, the leader would
travel for an infinitely long time.
We have focused on this circumstance because it is here that the princi-
pal features of a leader process manifest themselves clearly. The external field
at the tip site is usually low and cannot affect the instantaneous leader velo-
city, current and direction of motion. The direction may vary randomly, a
fact well known to those making lightning observations. What is important
is the voltage U, created by this field along the leader path, rather than the
field strength. The propagation of a positive leader is provided by the trans-
port of a fairly high positive charge to its streamer zone. The current of many
streamers taking the charge out accumulates in the channel, heating it and
providing its viability. But for many streamers to be excited off from the
leader tip, the latter must possess a high potential relative to the unperturbed
potential AU, = U,(L) - U o ( L )x IUo(L)I.This is indicated by the absence
of appreciable discrepancies between the current calculations made straight-
forwardly from the linear density of induced charge T(X) in a strongly non-
uniform external field and from formula (4.5)containing only AU,.
Therefore, it is not surprising that the lightning paths exhibit the diversity
illustrated in figure 4.4.No random change of the leader path can disturb its
viability. A leader can follow any direction: it can move along the external
field or in the opposite direction, along the equipotential line, etc. - all ways
are open as long as the condition AU, > AUtmnis valid. But the leader
acceleration does depend, of course, on its direction of motion. Moving
along the field, the leader is accelerated, because the voltage drop is compen-
sated excessively by the increase in I U, 1, When the leader moves in the opposite
direction, it is decelerated. The maximum acceleration is achieved in the
direction of the maximum gradient AU,, and this seems to be the reason for
the fact that the main leader branch darts in the direction of the rising field,
i.e., towards a charged cloud, a high object, etc. (for details see section 5.6).
running up far from the charged cloud region, has reached the maximum
height x , ( r ) and stopped. The channel plasma cannot decay immediately
but persists for some time. During this time, another, luckier branch has
approached the negative cloud bottom and even partly penetrated it. In
this region, the cloud has potential U,,,,, so that the branch portion that
has entered it acquires the positive charge
which may be as large as S x 10% of the negative cloud charge. Due to the
partial compensation of the lower cloud charge, with the upper charge
being constant, the zero equipotential surface will become lower by the
length Ax, which is about the same percentage of x, - H , so that
Ax x S(x, - H ) . As a result, the upper portion of the first halted branch
(from the tip down to the new zero equipotential surface) will be in a field
directed downwards. The new external potential at the tip site,
U; x IdUo/dxlxsAx, will become positive and the potential difference
AU, = U,(L)- U ; ( L )will be negative. For the example given in the previous
section with x, - H x 1.5 km at S x 0.1, we have Ax x 150 m; from formula
(4.10) with r << H , we have IdUo/dx~xs x 600V/cm, so that eventually
U;(L) x 9MV. Even if the branch penetrating the cloud charge misses its
centre to enter a region with a potential several times lower than U,,, (as
a result, UA(L) will be reduced as much), this will still be sufficient to
revive the first leader branch.
Therefore, the halted leader has a chance to revive and move on up to
the upper positive cloud charge but as a negative leader this time. The
leader position at the point of the first stop is unstable. Even a slight
perturbation, such as a decrease in the lower cloud charge (in the example
presented, due to the penetration of another branch) may stimulate its
further growth with the opposite sign. As the leader develops, it will
penetrate into an increasingly higher field of the upper charge and
become accelerated. Having passed the upper charge centre, H D, it +
will be retarded and stop, for good this time, at a height H,,, > H D, +
where the potential of (4.10) will drop to a relatively low value of AUtmin.
The height H,,, may be 10-20km or higher if one accounts for the air
density decrease. The currents flow in different directions in different
portions of this leader. Above the equipotential surface, the current flows
downwards, as in a negative leader. In the lower leader portion which
serves as a stem for many positive branches, the current remains directed
upwards. The observer, who registers the current at the earth, may not
suspect the sign reversal occurring up in the clouds. The channel field is
established in accordance with the current. It reverses in the upper channel
portion, thereby reducing the total voltage drop across the branch that
went far up and stimulating its further development.
Like a high grounded body, a large object isolated from the earth can become
a source of lightning in a high electric field of a storm cloud. Discharges can
be induced by fields extending not only between the earth and a charged
cloud but also between oppositely charged clouds. A lightning discharge
can be excited by a large aircraft, rocket or spacecraft when it travels through
the troposphere, and this is a serious hazard to its flight. Therefore, this
phenomenon is of primary practical importance.
Figure 4.5. Streak photographs of leaders from the ends of a metallic rod placed in a
uniform electric field: (a) general view; (b) fast streak photograph demonstrating the
relationship between the positive and the negative leaders; (1) rod, (2), (3) tip and
streamer zone of positive leader, (4),(5) tip and streamer zone of negative leader,
(6), (7) negative and positive leader flashes.
zero in a system isolated from the voltage source. The discharges appear to be
interrelated. Any fluctuation - say, a flash - of one leader appreciably acti-
vates the other: the space charge (e.g. positive) incorporated in front of the
rod stimulates the accumulation of negative charge across the conductor,
thereby enhancing the field at its negative end. The high-speed streak pictures
in figure 4.5(b),resolving individual streamer flashes, show an activation of
the positive leader channel following a negative leader flash.
The conditions for the start of leaders from a long isolated conducting
body are the same as from a grounded conductor, and they are also defined
by expression (4.4). But now, when estimating the threshold field Eo from the
value of AVimin = Eod, one should keep in mind that d is a half length of an
isolated leader. For a field capable of exciting a discharge from a conductor
of length 2d, we find
E o = - [1 bln(L/R) ] 2/5 .
(4.11)
d3/5 27qa
As in the illustration in section 4.1, we take the ratio of the channel length of
a young leader to the equivalent charge cover radius to be L/R M 10. Then
we have Eo = 440 V/cm for an aircraft of length 2d = 70 m. This estimate
describes the external field component along the aircraft axis. But an aircraft
often flies at an angle to the field vector, so that the threshold external field
may be several times higher. Fortunately, the lightning excitation threshold is
not very low, otherwise airline companies would suffer tremendous losses
from lightning damage. On the other hand, fields of this scale are not very
rare: much higher fields were registered during airborne cloud surveys. For
this reason, the problem of lightning protection in aviation is regarded as
being very serious.
Having started from an isolated body, each leader develops as long as
the external field permits. This process is basically the same as that discussed
in section 4.1 for an ascending leader. Below, we shall consider the specific
behaviour of two differently charged leaders developing simultaneously.
This specificity becomes especially clear in a non-uniform field typical of a
storm cloud.
Figure 4.6. A schematic diagram of a binary leader channel in a cloud dipole field.
x1: descending leader tip coordinate; x2: ascending leader tip coordinate; xo:position
of zero charge point.
assumed to be the same along the conductor length, the general condition for
an uncharged conductor is
1 x*
C 1 [ U - Uo(x)]dx=O; U=-/ U ~ ( Xdx.
) (4.12)
x2 - X I XI
= -a[U - Uo(x1)]lI2
dx2
-=a[Uo(x2) - U ]112 . (4.13)
dt ’ dt
Together with (4.12), expressions (4.13) describe the evolution of the two
leaders starting from the body ends, whose coordinates xl0 and x20 are
given as the initial conditions for equations (4.12) and (4.13). The sign
reversal point of the conductor, xo(t), defined by the equation
U(t) = UO(x0) is displaced, during the leader propagation, in accordance
with the nature and degree of field non-uniformity along the channels.
Having solved the equations, one can find the currents at the leader tips
from (4.3) with L = x2 - x l . Generally, they differ quantitatively from one
another and from the current in other channel cross sections, including the
sign reversal at point xo,through which the total charge flows during the
This current is used for changing the charge of the old leader portions,
increasing or decreasing them as A U ( x , t), and for supplying charge to its
new portions. This leads to the current variation along the channels, which
can be found by solving the problem.
For some simple distributions of Uo(x),the division of equations (4.13)
by one another
(4.15)
-
too. This becomes possible if the cloud field is approximated by the point
charge field Uo(x) IxI-', and if a new variable z = x2/x1 is introduced.
The resultant formulas allow an analytical treatment of some characteristic
relationships. To avoid cumbersome derivations, we invite the reader to do
this independently, while we, instead, shall present some numerical calcula-
tions for several variants.
The calculations prove to be quite simple in integrating the set of
equations (4.13) and (4.14) as well as in the case of a more rigorous approach
to the problem, when the charge distribution along the conductor length is
found from an equation similar to (4.7). Figure 4.7 demonstrates the propa-
gation of vertical leaders in the field of a cloud dipole (with the allowance
for the earth's effect). The calculation was made using an equation similar
to (4.7). The initiating vertical body is located between the lower negative
charge of the dipole and the earth, being displaced horizontally by
Y = 500m from the charge line. As the ascending leader moves up, its tip
approaches the bottom charge centre and enters a region of an ever increas-
ing field. The descending leader moves more slowly towards a weaker field.
The external field potential approaches zero at the earth but increases rapidly
near the charged cloud. As a result, the negative potential of the conductor
made up of the leader channels, U , rises with time, with the sign reversal
point xo going up closer to the cloud. At the initial moment of time, the
potential is U = -27 MV and the point is at an altitude xo = 1603m.
When the ascending leader reaches the charged centre 17 ms later, we have
the altitude xo = 2040 m and U = -64 MV. The absolute potential rise
stimulates the descending leader, increasing its velocity by a factor of three
during this time in spite of its propagation through an ever decreasing
external field. The calculations made with (4.13) and (4.14) have yielded
similar results.
"."I . I . , . I . I ,
0 5 10 15 20
Time, mc
Figure 4.7. The propagation of leaders from a metallic body located between the
cloud and the earth (Q, = -1OC, H = D = 3 km, Y = 0.5 km).
1.4-
1.2-
2 1.0-.
0 .
60.8-
.e
S
8
0.4-
0.2 -
.
. \
0 . 0 - I
10 15 20 25
Time, ms
4.0 1.6
1.4
3.5
E.
l3.0
J
.s7
U
2.5
7
upward leader
/rd downward
leader
1.2
1.0 2
-
0.8 3
5
0
*
0.6
$
.I
0.4
2.0
0.2
1.5 0.0
I 1'5 ' 20 25
Time, ms
Figure 4.8. The development of a leader pair from a metallic body at 1.5 km above the
earth in a cloud dipole field ( H = 3 km, D = 3 km, Q, = - 10 C ) .At the moment of time
N 10 ms and 1 km altitude, the descending leader turned to follow an equipotential path:
(top) leader velocities; (bottom) position of the zero charge point (T = 0), the altitude of
the ascending leader tip and the length of the portion along the equipotential path.
must apparently be ‘attracted’ by the negative cloud. (In the two situations
above, it was the negative leader that travelled to the earth, ‘naturally’
extracting a negative charge from the cloud). This exotic situation arises
when two leaders are excited from a body located somewhat above the
negative charge centre. The negative leader then goes up to a positive
cloud. The strong field created by the cloud dipole induces a large negative
charge in the ascending leader, displacing the positive charge down. The
average external potential between the two leader tips, U < 0, appears to
be ‘more positive’ than the external potential of the lower tip, Uo(xl) < 0,
so that AU,, = U - Uo(xl) > 0. This is the reason why the descending
leader is positive. With time, when the ascending tip comes closer to the
positive charge, the potential U of the binary system does become slightly
positive (3-3.5 MV), making up several percent of 1 Uol,,,. Then it persists
as such, sustaining the descending leader travel to the earth. By the
moment of contact with the earth, the positive charge is distributed along
the channel in about the same way as in a grounded conductor in a negative
cloud field, being mainly concentrated at the height of this charge. For this
reason, the return stroke current, which only slightly contributes to the
charge after the contact, is weak. The return stroke can be said to make no
contribution to the charge redistribution, since the channel potential
should be corrected only slightly (as compared with I Uolm,, x 100 MV), by
reducing it from 3 MV to zero. This reduction enhances, though only slightly,
the ascending leader, which travels on until it stops high above the positive
charge of the storm cloud.
When an isolated body initiates two oppositely directed leaders, it does
not always happen that the descending positive leader reaches the earth. For
the contact with the earth to take place, the potential U of the conductor
made up of the two leaders must become positive at a certain moment. Other-
wise, the descending leader will stop at the point xls, where the negative
leader tip potential U,, will be by a small value of AVtm,, higher than the
negative potential of the external field (assuming AVtm,,= 0, when U is
equal to U o ( x l s ) )The
. condition for the average conductor potential to be
positive at the moment of contact with the earth is described by the inequality
JZ dx > o
u0(x) (4.16)
which follows from (4.12). Here, Uo(x) is the cloud dipole potential given by
(4.10) with the allowance for its reflection from the earth, and x2 is the
altitude the ascending leader tip has reached by that moment. In principle,
there are no reasons for this inequality to be violated, since the integral of
(4.16) in the limit x2 = CO is necessarily positive (and equal to
-21n(l + D/H)Q,/4mo at r << H ) for the negative lower cloud charge
(Q, < 0). This means that the descending positive leader has a chance to
reach the earth - this only requires that the ascending leader should reach
threshold (Ei M 20-25 kV/cm at the height of the cloud charge) or the cloud
is to contain inclusions enhancing the field locally via the polarization charge.
It seems that neither mechanism should be discarded entirely, although cloud
probing rarely registered fields exceeding several kilovolts per centimetre.
These results do not testify to the absence of higher fields, because most of
the measurements concerned fields averaged over lengths of several dozens
of metres. No measurements were made at the moment of lightning initia-
tion, because the probability of a detector registering a field at the right
place at the right moment is extremely low. On the other hand, the conditions
necessary for the excitation of a leader process in a cloud are quite rare;
otherwise, the number of lightning strikes per square kilometre of the earth’s
surface would greatly exceed 2-5 per storm season.
Let us estimate the volume to be occupied by a cloud charge capable of
creating an ionization field. It was mentioned above that the field Eo at the
earth was often found to be lOOV/cm during thunderstorms. This value
should not be considered to be the cloud dipole field, since the near-earth
charge provided by microcoronas from various pointed objects attenuates
the cloud field at the earth. A similar value is obtained from a small positive
charge supposed to lie under the principal negative charge [5]. Taking, for
estimations, the intrinsic dipole fields Eo to be 200V/cm and the heights of
the lower (negative) and the upper (positive) charges to be x = H = 3 km
and H + D = 6 km, respectively, we find, from (4.10), the dipole charges
Q, = 13.3C. These values will serve as guidelines in further numerical calcu-
lations. The charge Q, can create field Ei M 25 kV/cm at its boundary if it is
distributed throughout a sphere of radius R, = 220 m. Measurements show
that the charged region is, in reality, 2-3 times larger, but one should not
discard the possibility of a short accidental charge concentration in a smaller
volume due to the action of some flows in the clouds.
More probable is the situation when a macroscopically averaged maxi-
mum field of cloud charge is several times lower than Ei and local fields,
enhanced to Ei x 25 kV/cm, arise near polarized macroparticles. Note that
the maximum field near a metallic ball polarized in an external field E is
E,,, = 3E. Similarly enhanced is the external field of a spherical water
droplet, since water possesses a very high dielectric permittivity E = 80 and
E,, = 3 E ~ / ( + 2 E ) . Therefore, if charge Q, is concentrated in a sphere of
& times larger radius, R, = 380 m, the field three-fold enhanced by polari-
zation can achieve the ionization threshold. Following the ionization onset,
streamers may be produced around large droplets, giving rise to a possible
leader, because streamers may be branched and extended in an average
field of -10 kV/cm.
Leaving aside the mechanisms of ionizing fields and leader origin,
because they are still poorly understood, we shall take for granted only the
mere fact that a leader does occur. At its start, a descending leader is
devoid of the possibility of taking the charge it needs away from the cloud.
Observations show that this charge is quite large: an average negative leader
transports to the earth a charge QL RZ -5 C and, sometimes, it is as large as
-20 C [l], a value close to the evaluations of Q, for the storm cloud. But if the
cloud charge remains ‘intact’, the only thing that can provide the charge
balance is the ascending leader of opposite sign, which is to develop simulta-
neously with the descending leader. This idea was suggested in [6],which
presented a qualitative distribution of the charge induced along a vertical
conductor made up of two leaders prior to and following its contact with
the earth. What happens is principally the same as in the excitation of two
leaders by a conducting body isolated from the earth and is affected by an
external field (section 4.2). This process is independent of the descending
leader sign; therefore, one should not think that a negative cloud can produce
only a negative leader while a positive cloud always produces a positive one.
In any case, two oppositely charged leaders are produced simultaneously,
and which of them will travel to the earth depends on the charge position
in the cloud and on the leader starting point.
A binary leader is most likely to be initiated near the external boundary
of the charged region, because the field there is highest. The field at the
centre of an isolated charged sphere is zero. In the case of a uniform
charge distribution throughout its volume, the field rises along the radius
as E r but decreases from the outside as e rP2 with the maximum
N N
earth is quite low. This actually cancels the return stroke current. Positive
lightnings with very low currents of about 1 kA present no danger and are
quite frequent. The number of their registrations by the observer increases
with increasing sensitivity of the detectors used.
The potentials from the upper dipole charge and from charges reflected by
the earth can be found as from point charges. They do not contribute
much to U,$. For example, for the centre of a negative sphere with
Q, = -13.3C and R, = 500m, we have U,, = -360MV and at the
5
boundary U,, = U,, = -240 MV. With all other charges taken into
account, we get U,, = - 196 MV for the bottom edge of the lower sphere
at H = D = 3km.
Figure 4.9 presents the results of this calculation including those for the
charge distribution along the conductor length from an equation similar to
(4.7). We have evaluated the development of both leaders along the dipole
axis, following the start from the bottom edge of the lower negative
Figure 4.9. (Opposite) The model of a descending leader from the lower boundary
of the negative dipole charge (Qc = -12.5C, H = 3 km, D = 3 km, R, = 0.5 km).
Vertical channels have no branches: (top) tip positions of the negative descending
leader, x l , and its positive ascending partner, x2,with the points of zero potential
differences, xo; (centre) charge distribution along the leader channel; (bottom)
potential and velocity of the descending leader.
2-
1-
--.
E
U 0
E 4
I-" -1 -
-2 -
-3
Time, ms
sphere. The dipole potential from (4.10) was used as Uo(x), except for the
length in the charged region, where expression (4.17) was employed with
r = jx - HI. The descending negative leader of a binary system is accelerated
quickly after the start. Having covered about 500 m, it travels farther to the
earth with a slightly decreasing velocity wL x (1.6-1.7) x 105m,’s, a value
close to observations. The leader strikes the earth in 16ms. By that time,
the ascending leader has reached the height x2 = 3.6 km. This is far even
from the zero potential point located at x x 4.5 km, let alone from the
upper positive charge located at an altitude of 6km. The descending
leader, which started from a site with the local potential - 185 MV, trans-
ports to the earth nearly half of this value, U1 % - 105 MV, in spite of the
initial assumption of the zero voltage drop across the channel assumed to
be a perfect conductor.
The reader should not feel discouraged by the large calculated value of
U1, which is more appropriate to record strong lightnings rather than to a
common lightning discharge, especially considering that the cloud param-
eters taken for the calculation were quite moderate. It will be demonstrated
in section 4.3.3 that leader branching, which is a rule rather than an excep-
tion, reduces considerably the potential transported down to the earth. It
is quite likely, however, that lightnings of record intensities are produced
in ordinary clouds rather than in those having a record high charge, but
only if the descending leader does not branch (or does so slightly).
The above calculation for an ideal situation with unbranched leaders is
interesting and useful for two reasons. First, one should understand the
physics of a simple observable phenomenon before one turns to its complex
modifications. The other reason is, probably, more important. Practical
lightning protection requires the knowledge of both typical average lightning
parameters and their record high values. It is the latter that become more
important in designing prospective measures for especially valuable con-
structions and objects. As was pointed out above, the case of an unbranched
leader just discussed is likely to be one of the rare but most hazardous
phenomena.
The potential U1 transported by a lightning leader to the earth is an
important parameter for practical lightning protection. The return stroke
current (section 4.4), the most destructive force of lightning, is proportional
to U1. The nature of U1 becomes clear from the above conception of
descending leader development in a binary leader process. Ideally, potential
U1 is that of a perfect conductor, made up of two leader channels, at the
moment of its contact with the earth. But the ascending and descending
leaders develop differently, because their paths cross regions possessing
different distributions of cloud potentials Uo(x). The descending leader
travels nearly without retardation because the potential difference at its
tip, AU,, = U - Uo(xl), remains almost constant (a decrease in 1 U1 is largely
compensated by 1 Uo(x)Idecreasing towards the earth). The ascending
I”
Figure 4.10. Estimation of the potential transported by a negative leader to the earth.
positive leader moves ‘against’ the field and soon enters a region of a rapidly
rising external potential; as a result, AU,, = U - U0(x2) becomes relatively
low soon after the start. This leads to a lower velocity of the ascending
leader, which now goes up ‘unwillingly’, being affected by its more active
twin which moves faster, pumping its charge into it. For this reason, just
before the descending leader contacts the earth, the total potential of the
system nearly coincides with the external field potential U0(x2) at the site
of the ascending leader tip (AU,, = U - Uo(x2) << 1 Ul).
This circumstance makes it possible to determine the transported poten-
tial U1 = U just from the condition U = U0(x2) at x 1 = 0. The condition has
a clear geometrical interpretation (figure 4.10). The shaded regions between
the external potential curve Uo(x) and the horizontal line intercepting it must
be identical on both sides of the left-hand interception point (the point of the
conductor sign reversal, xo).This results from the net polarization charges of
both signs being identical; they are proportional to the shaded regions (see
formula (4.12)). This approach can be used to find U1 in different charge
distribution models and for different horizontal deviations of the vertical
leader path from the dipole axis. In a simple case when both leaders propa-
gate along the dipole axis, formulae (4.12) and (4.10) with D = H and r = 0,
together with expression (4.17), yield a dimensionless equality for finding
point x2 and then U 1 :
(4.18)
For the variant shown in figure 4.9 with n = i, expressions (4.18) give
e2 = 1.27 and U / U O R= 0.63, in a fairly good agreement with the
calculations of the leader evolution (note that U,, RZ U,, is the external
field potential at the bottom edge of the lower cloud charge that has
triggered both leaders).
A negative descending leader can start from any point on the lower
hemisphere of the bottom negative charge of the cloud. The location of
this point is quite likely to be a matter of chance, since the field across the
surface is nearly uniform. Depending on the location of the starting point,
the ascending twin crosses the charged region along chords of different
lengths, and this, along with the other factors, affects the potential trans-
ported to the earth. The maximum potential U1 at the moment of contact
with the earth is characteristic of a leader that has started from the lowermost
point of the charged sphere, when the paths of both twins cross the regions of
maximum potentials and the ascending leader path in it is longest. The value
of U1max is about 60% of the external potential U,, at the start and 40% of
the maximum U,, value at the centre of the negative cloud charge. But even
if the descending leader is initiated near a lateral point of the hemisphere
located at a maximum distance from the dipole axis, it transports a
considerable potential found from calculations to be 0.65U1max x O.4UoR.
Therefore, an unbranched negative leader transports to the earth a high
potential, (0.6-0.4)UOR, no matter where it has started from the lower
hemisphere.
cc+ CL Cc
where UoR would be the boundary potential of a charged sphere, were it
isolated (the small capacitance gain due to charges induced in the earth are
ignored). For L/R x 100 and H/Rc = 6, as in the previous numerical
illustration, the potential U1 x O.6UoRis nearly the same as that transported
to the earth in the absence of ascending leader branching. The potential of
the cloud-leader system drops because of the outflow of some of the cloud
charge to the new capacitance of the descending leader just produced. A
similar effect has been observed in long laboratory sparks. The capacitance
of an extremely long spark is often only one order of magnitude smaller
than the output capacitance of a impulse voltage generator, connected
directly to a gap without a large damping resistor. The charge inflow into
the leader is quite appreciable and reveals itself as a voltage drop across
the gap.
A branched descending leader possesses a larger capacitance than an
unbranched one; it takes away a higher charge from the cloud and decreases
the potential more. To estimate this effect, let us represent a branched leader
as a bunch of n identical conductors of radius R and length L, spaced at
distance d (L > d >> R). Supplied by the same power source, they possess
the same potential U and linear charge T . The potential at the centre of
any of these conductors is found by summing the potentials of all charges
of all conductors, including the intrinsic potential. Integration with the
neglect of the small effect of the earth yields
The total capacitance of the n conductors, Ctn = nrL/U, is larger than that
of a single isolated conductor, but this gain is less than n-fold:
Ct n - nln(L/R)
ct1 ln[(L/R)(L/d)n-']
The reduction in the potential transported to the earth roughly follows
the distribution of the cloud charge Q, between the capacitances of the
charged cloud cell, C,, and of the leader, Ctn,described by the first equality
of (4.19). For n = 10 branches separated at distances d = L/3 and
L/R x 100 derived from photographs, the capacitance is Ctl0x 3.2Ct1. This
well-branched leader will transport to the earth potential U1 x 0.3Uo~.In
view of the real length of a leader (especially, a well-branched one) which is
about 1.5 times longer than the charge height H , i.e. L M 1.5H, the
where H1and H2are the heights of the bottom and top charge centres and x2
is the ascending leader height at the moment the descending leader contacts
the earth. We mentioned at the end of section 4.2.2 that an ascending leader
must go up at least to x2 = 2.47H1 at H2 = 2H1; then we have U = 0.
At x 2 M 4H1, the function U ( x 2 ) crosses the smooth maximum,
U,,, KZ -Qc/207r~OH1M 8MV, if Q, = -13.3C and H1 = 3 km, as in the
previous examples. Even the maximum potential transported to the earth
is small. This means that the return stroke current of a descending positive
leader travelling along the dipole axis will be low. The potential and the
current will be still lower, with the real voltage drop U, across a channel of
total length 4H x lOkm taken into account. Even for the channel field
E x 10 V/cm, the value of U, 10 MV is comparable with U,,. This light-
ning is so weak that it has little chance of being registered and included in the
statistics.
Vertical channels demonstrate maximum positive potentials transported
to the earth. They go through more or less identical regions of negative (at
the bottom) and positive (at the top) external potentials, and the respective
contributions to the integral of (4.12) are mutually compensated. Positive
lightnings, however, can possess very high currents. With the foregoing
taken into account, one can suggest at least two reasons for this. One is a
favourable random deviation of the channel path from the vertical line.
Suppose the ascending leader of a binary system, starting from the upper
positive charge point closest to the earth, xo = H2 - R,, moves up vertically,
while the other leader, having descended to the zero potential point between
the charges, turns aside and goes along the zero equipotential line. After it
has deviated for a large distance r from the dipole axis, it turns down
vertically to contact the earth this time. In this case, the descending leader
misses the region of high negative potential, and positive contribution to
the integral of (4.12) remains uncompensated. Calculations with formulae
(4.12), (4.10) and (4.17) made at H I = 3 km, H2 = 6km, R, = 0.5 km, and
r = 1 km show that the descending leader will transport to the earth a poten-
tial 4.3 times greater than that to be transported along the dipole axis.
Another principal possibility is the deviation of the dipole axis itself
from the vertical line, with the vertical leader path preserved. The centres
of the top and bottom charges can be shifted from the same vertical line
because of the difference in the wind forces at different heights. Then the
leader that has started up vertically from the top charged region passes
through the region of high positive potential, while its twin, descending ver-
tically, will appear to be shifted aside relative to the bottom charge and go
through the region of low negative potential. The effect will be the same as
in the first case. Quantitatively, it may even appear to be stronger, since
the length and capacitance of the descending leader are smaller due to the
lack of an extended path along the zero potential line.
4.3.5 A counterleader
The descending lightning leader does not reach the earth or a grounded body,
because it is captured by the ascending leader developing in the electric field
of cloud and earth-reflected charges. This field is enhanced by the charge of
the descending leader approaching the earth. This can also happen in labora-
tory conditions, especially if the descending leader is negative. Then the
counterleader is positive and requires a lower field for its development.
Streak pictures of laboratory sparks clearly show the counterleader start
Figure 4.11. A streak photograph of a long spark with a counterleader coming from a
grounded electrode.
and motion towards the descending leader (figure 4.11). The altitude at which
their encounter occurs depends on the descending leader sign and charge.
The length of the counterleader at the moment of their contact is important
for lightning protection practice, because it defines the number of strikes at
bodies of different heights and, to some extent, the current rise parameters
of the return stroke from the affected body.
Let us estimate the altitude z, which the descending leader tip is to
reach to be able to create a field at the earth high enough to produce a
viable counterleader. The latter does not differ from any other ascending
leader, and its development from a body of height d requires that the
near-terrestrial field should exceed the value of Eo from formula (4.1 1).
For the height d = 30m characteristic of industrial premises, the field
must be Eo x 480V/cm. If the cloud field is -lOOV/cm, the field
AE = 380V/cm must be created by the descending leader with charge.
The main contribution to the near-terrestrial field is made by the charge
concentrated at the leader channel bottom. Therefore, the calculation of
the field AE under a very long vertical conductor should utilize the constant
value of r averaged over this bottom of length -z, rather than the linear
density of the non-uniform charge r(x).With the charge reflected by the
earth, we have
r "dx - r u-uo -U
(4.21)
*E(z) =GIZ T - G - z l n ( L / R ) -zln(H/R)
N
where U is the channel potential and L x H is its length, which is about the
cloud height at the tip height z << H. Here, we have used the conventional
expression T = C1( U - U o ) with average linear capacitance and accounted
for the near-terrestrial potential of the cloud, 1 Uoi << 1 UI. For an unbranched
descending leader carrying high potential U x 50 MV, we obtain z x 260 m
at ln(H/R) x 5.
The counterleader arises at the last stage of the descending leader
development, i.e., near the earth. Its velocity is not high and is equal to
wL, x 2 x 104m/s from the first formula of (4.5), because the potential
difference on the leader tip is quite low, A U x Eod x 1.5 MV. The descend-
ing leader has an order of magnitude higher velocity. For this reason, the
counterleader acquires the length L1 x (wL,/vL)z x 25 m by the moment of
encounter. This is a large value, since the length L1 is summed with the
body’s height d, so that the total height of the grounded conductor becomes
nearly doubled. This affects the frequency of the body’s damage by lightning
strikes.
It follows from formulae (4.21) and (4.11) that the height z,to which the
leader descends before it can initiate a counterleader, is greater for higher
premises, from which the counterleader starts, z d3I5,although this depen-
N
dence is not very stringent. It is important that as the altitude of a body and z
become greater, the counterleader has more time for its acceleration and can
acquire a longer length. It is important for applications that it is not only the
length L1 which increases but also the L1 / d ratio.
The simple estimation obtained from (4.21) and (4.11) can be refined by
accounting for the T ( X ) non-uniformity in the integral of (4.21) arising from
the proportionality T U - U 0 ( x ) and by rejecting the approximation of
N
Figure 4.12. Schematic recharging of a lightning channel after the contact of the
descending leader with the earth. Shaded regions, charge; ( a ) moment of the leader
contact with the earth; (b) the return stroke reaching the upper channel end; (c)
charge change.
A T = T ~- ~ o = [ ( C l- C i ) ( U i - U o ) + C ’ 1 ( 0 - U o ) ] - [ C 1 ( U i - U o ) ]
= -c:ui (4.22)
remains the same in order of magnitude. Consequently, when considering
fundamental stroke mechanisms, one can take C: % C1 and assume the
equivalent line to be charged uniformly.
Measurements made at the earth show that a descending leader is
discharged with a very high current. For negative lightnings, the current
impulse of a return stroke with an amplitude ZM -10-100 kA lasts for 50-
100 ps on the 0.5 level. A short bright tip of the return channel well seen in
streak photographs runs up for approximately the same time. Its velocity
v, M (0.1-0.5)~is only a few times less than light velocity c. It would be
natural to interpret this fact as the propagation of a discharge wave along
the channel; this wave is characterized by a decreasing potential and rising
current. Due to an intensive energy release, the channel portion close to
the wave front, where the potential drops from U, and a high current is
produced, is heated to a high temperature (from 30000 to 35000K, as
shown by measurements). This is why the wave front is so bright. The
channel behind it is cooled due to expansion and radiation losses, becoming
less bright. A return stroke has much in common with the discharge of a
common metallic conductor in the form of a long line. The line discharge
also has a wave nature, and this process was taken to be a model discharge
in shaping the ideas concerning the return lightning stroke.
A lightning channel is discharged much faster than it was charged
during its development with the leader velocity vL % (10-3-10-2)~,. But
the variations in potential and linear charge during the charging and the
discharge are expressed as values of the same order of magnitude: ro AT.
In agreement with the velocity, the channel is discharged with current
-
1, M ATV, by a factor of v,/uL %102-103 higher than the leader current
iL M r0uL ~ ~ 1 0 0 The
A . linear channel resistance R Idecreases approximately
as much during the leader-stroke transition. This decrease is due to the
average electric field in the leader channel and behind the discharge wave
in the return stroke, E, M R1IM x RILiL,have the same order of magnitude.
This is consistent with the conclusion to be made from a straightforward
analysis of the established states in both channels. The situation there is
similar to that in a steady state arc. But the channel field E, in a high current
arc does vary but slightly with the current [ 111.
It follows from the foregoing that if a leader has iL M 100A, E, z 10Vi
cm and ROLM 0.1 R/cm, the return stroke must have Ro x 10-3-10-4 R/cm
in the steady state behind the wave front; the total resistance of a channel of
several kilometres in length appears to be lo2 R. This value is comparable
with the wave resistance of a long perfectly conducting line in air, 2,whereas
the total ohmic resistance of a leader of the same length is two orders of
magnitude larger than Z . The ratio of the ohmic resistance of the line portion
behind the wave to the wave resistance indicates the degree of the wave
attenuation during its travel along the line (section 4.4.2). If the channel
resistance were constant and remained on the leader level, the lightning chan-
nel discharge wave would attenuate, being unable to cover a considerable
channel length. The current through the point of the channel closing on
the earth would also attenuate too quickly. Experiments, however, point to
the contrary: the visible bright tip has a well-defined front, and a high current
is registered at the earth during the whole period of the tip elevation. The
transformation of the leader channel during the wave travel decreases its
linear resistance considerably, determining the whole return stroke process.
While the wave propagates along the line, a detector mounted at its
beginning will register direct current. If voltage Ui is low and there is no
charge cover around the conductor, the capacitance and inductance are
characterized by the same radius Y, in the logarithms of (2.8) and (4.26). In
this case, we have
ti
Figure 4.13. Distributions of potential and current during the discharge of a perfectly
conducting line.
i ( x ,t ) = -
Z
(4.29)
(4.31)
(4.32)
This current decreases much more slowly than at the wave front, because in
spite of the negligible front current, the line far behind the wave front is
discharged all the same, and all the charge that flows down from it goes
through the base.
The wave front propagates at a rate of electromagnetic perturbation. It
is independent of the line ohmic resistance but is determined exclusively by
its reactive parameters and is close to light velocity. This is a 'precursor'
which exists under any conditions, no matter whether the line has a
resistance or whether it changes behind the wave front. The precursor
carries information about the changes in the line, in our case about the
line grounding. If the resistance is zero or, more exactly, has no effect yet
because it is much less than the wave resistance (Rlxf << Z), the line is
discharged in a resistance-free way, and its initial potential and charge
practically vanish right behind the front of the primary electromagnetic
signal, the precursor. When the resistance becomes much higher (practically
several times higher) than the wave resistance, the charge and potential dis-
appear gradually, and the rate of their reduction decreases as the linear
resistance RI increases. At R1 = 10 n / m corresponding to the leader chan-
nel resistance, the time constant is a = 2ps-l and the precursor current
decreases, in accordance with (4.3 l), by an order of magnitude as compared
with the initial value of i(0, 0) over the period of time t 1 ps, for which the
N
vi-
s/(4xr)’:2
U ( x ,t ) exp( -t2)d[ = U, erf
f ?i [0
Expressions (4.33) and (4.34) have a clear physical sense, demonstrating the
nature of a non-ideal line discharge.
When the perturbation front (the precursor due to the action of induc-
tance) goes far away, the current decreases slowly from the line base to the
front. It also varies slowly in time at every point, except for the region
close to the front. This is the reason why the inductance effects in the main
discharge region are very weak. With the neglect of the inductance term,
equations (4.24) transform to equations similar to those for heat conduction
or diffusion:
(4.35)
Figure 4.14. The potential wave ( a )in linear diffusion with x = const and (b) in non-
linear diffusion with rising x.
propagates in such a way that its characteristic point x l , say, where the
potential is reduced by half relative to the initial value of Vi, x 1 / ( 4 ~ t ) ~=/ *
0.477, obeys the diffusion law x1 x ( x t ) l l 2 with a decreasing velocity
v1 x ( ~ / t ) M” ~x / 2 x 1 .From expressions (4.33), the current at point x1 is
20% lower than that at the channel base x = 0. Substituting the leader
resistance R1 x 10R/cm and C1 = 10pF/m into the formulae, we get
x = 1010 m2/ S . Over the time t = 10 ps (at at = 20), during which a weak
precursor will cover a distance of 2000 m, the half-potential point character-
izing the propagation of the line discharge wave will diffuse for 3 15 m only
and will be moving at velocity v1 x 0 . 0 5 ~ By . that time, the base current
i(0, t ) will have dropped by a factor of 11 relative to the initial current
i(0,O) (formula (4.32)).
The calculated values of x l , wl, and i(0,t ) can be brought closer to
measurements at a certain stage of the lightning discharge. Instead of the
leader resistivity, one should then deal with a lower resistivity averaged
over the perturbed region. This makes sense in some evaluations. But the
illusion of a satisfactory numerical agreement with measurements in a
short stage of the process is destroyed, as soon as we recall one of the
important qualitative observations. At the return stroke stage, a bright
and well-defined wave front - the channel tip, which becomes smeared
only slightly with time - is moving up to the cloud. This indicates that the
energy release and, hence, the current rise occur faster than in the solution
to (4.33). Clearly we deal with a wave possessing a steep front, at least for
powerful lightnings, rather than with diffuse current profiles. This contra-
diction can be resolved by rejecting the approximation R1 = const and by
including, in the theoretical treatment, the time evolution of the leader
channel and its transformation to a return stroke channel.
Note that the simple and attractive model of an immediate transforma-
tion of the leader channel at the wave front to an ideal conductor cannot
rectify the situation. This model would take us back to equalities (4.26)
and (4.27) describing the wave of immediate voltage removal and sustained
current, which propagates with the velocity of an electromagnetic signal
close to light velocity. But this possibility was already refused above. It
was mentioned in section 4.4.1 that the key to the phenomenon of return
stroke should be the analysis of the channel transformation dynamics.
The effect of the gradual resistance reduction during the Joule heat
release can be understood from equations (4.35) and (4.24) without the
inductance term. It would be justifiable to replace U by the potential varia-
tion AU = U - Ui, since Ui = const in our approximation, so we get
- - aX - d A U
dAU dAU 1
-- i = - C 1 X X ’ x = - (4.36)
at ax ax RlCl .
The resistance decreases while x increases, as the amount of charge flowing
through the particular channel site becomes larger, or with the increase in
AU. Consequently, the rear sites of the diffusion wave, where A U and
diffusion coefficient x are already higher, propagate faster than the front
sites, where A U and x are still low. To supply current to the region close
to the discharge wave front (a weak precursor is out of the question now),
the potential gradient there must be large because of the small diffusion
coefficient. Both circumstances indicate that the wave acquires a sharp
front, its profile becomes steeper and convex. In contrast to the gradual
asymptotic approximation at x = const, the curves U ( x ) and i ( x ) for a
given moment of time look as if they stick into the abscissa (figure 4.14;
the same will be seen from the numerical simulation illustrated in figure
4.17). The effect described here is well known [12]; this is a non-linear heat
wave driven, for example, by radiative heat conduction, whose coefficient
drops with decreasing temperature T approximately as x = T3.t
The variant with R I = const, for which the solution to (4.33) and (4.34)
is valid, probably corresponds to low current lightnings, when the energy
release is too small to provide an essential reduction in the former channel
resistance. In any case, there are streak pictures of return strokes with unclear
wave fronts or those becoming smeared after the propagation for a few
hundreds of metres [13,14]. To obtain conclusive evidence, stroke streak
pictures should be analysed at different currents. Regretfully, no simulta-
neous recordings of currents and stroke waves are available.
One can draw another conclusion from the solution to the set of
equations (4.24) at R I = const # 0, which is important for the analysis of
observations and for the formulation of boundary conditions necessary for
finding a numerical solution. According to (4.29), when the line closes on
the earth instantaneously the discharge current through the closed end also
reaches its maximum instantaneously. As mentioned above, the maximum
is independent of R I , being determined exclusively by the wave resistance.
Clearly, the same will also be true for any time-variable resistivity, and the
only question is how fast the current will decrease after the maximum.
However, the current in a real return stroke rises for several microseconds,
sometimes for several dozens of microseconds, and this time may become
even comparable with the total impulse time. Such a slow current rise may
t Equations (4.36) with A U ( x . 0 ) = 0, AU(0. t ) = -L'i and no inductance terms allow self-
similar solutions. The simplest of them are (4.33) and (4.34) for x = const. The process is self-
similar in a more complex approximation for x = b( 1 U,l)ntv,which corresponds qualitatively
-
to the R I x-' evolution during the channel transformation. Constants 6 , n, and U can be
chosen from the analysis of R I behaviour (section 4.4.3):n x 1-2; U 0.5-1. The wave front
follows the relations
x/ = E[b(lC:l)"]'!?t'"-'~:?
Vf = i ( V + l)E[b(lC:l)"]':2t-('-V'".
be only due to the properties of the commutator, whose role is played by the
streamer zones of the descending leader and the counterleader. Their contact
actually gives rise to the return stroke. The streamer zone field rises, as the
streamer zones are reduced and the leader tips come close to each other or
as the descending tip approaches the earth with no counterleader formed.
The streamers are accelerated to a velocity lo7m/s, transporting kiloampere
currents even in laboratory conditions [ 15,161. Thus, the rate of current rise
and the impulse front duration at the earth, 9, are determined by processes
occurring in the vanishing streamer zone rather than in the former leader
channel. Measurements provide indirect evidence for this, showing that the
impulse rise time tf in positive lightnings possessing a longer streamer zone
than negative ones, at the same voltage, is several times longer.
10 15 20 25 30 35
T, kK
Figure 4.15. The conductivity of thermodynamically equilibrium air at atmospheric
pressure.
tThe problem of a short laboratory spark is much simpler. The set should include a simple
discharge equation for a capacitor bank as a high-voltage source for a spark gap with the desired
resistance and allowance for the circuit inductance. Note that this kind of LRC circuit usually
registers damped oscillations unobservable in lightning.
25- t=5p
20 -
a
$15
20
10 -
50
5-
4
sa
2
0
0
r, cm r, cm
Table 4.1 presents, among other parameters, the linear resistance values
obtained from T(r) data borrowed from [23,24]. One can see that the
resistivity drops at first for 1 ps but then falls rather slowly. This decrease
ceases closer to the pulse tail, and the resistance begins to rise gradually.
The dramatic initial drop in R I is due to the primary heating of a very thin
initial channel by high density current.? As T increases to about 20000K,
the conductivity a rises but remains nearly constant with further temperature
t The gas is assumed to be in thermodynamic equilibrium at every moment of time. This assump-
tion is justified by a fast energy exchange (for lO-*-lO-'s) between electrons and ions, resulting
in a small difference between the gas and electron temperatures. The ionization is of thermal
nature: a Maxwellian distribution is established in the electron gas, and the amount of ionizing
electrons is defined directly by the electron temperature, rather than by the field. The electron
temperature, in turn, is determined by the Joule heat release and energy balance of the gas.
Equilibrium ionization is also established rapidly (for details, see [ll]).
Note. T,,, is the temperature along the channel axis, Teffis the average temperature in the
conductive channel, oeffis an average channel conductivity, p is channel pressure, reffis the effec-
tive radius of the conductive channel, W is the total energy released (no data for the second
variant; the given values was estimated as W x i ~ a x R 1 t p )and
r Q is the charge transported
during the current impulse.
-
rise. In a strongly ionized plasma, with ions of constant charge o T3I2,but
doubly charged ions appear with increasing T . Since U ZC2,where Zi is the
-
ion charge multiplicity, the two effects compensate each other. The resistance
of a highly heated channel decreases with time due to its expansion only.
Some time later, however, the pressure at the channel centre drops to
atmospheric pressure, and the expansion ceases. The conductive channel
cross section is reduced gradually because of the gas cooling caused by
thermal radiation. The channel resistance begins to rise slowly because of
decreasing reff and Teff.
The expansion time of the channel becomes longer and its minimal
resistance decreases for the stronger current impulses. Physically, the linear
resistance is affected by the energy released per unit channel length, W ,
rather than by the current. This value is not described unambiguously by
- - -
the current amplitude; what is more important is the amount of the
transported charge Q: W1 i 2 R l t QiRl QE and the field does not
vary much. The calculations, however, deal with the current impulse but
not with W1. Semi-quantitatively, the time dependence of resistance can be
understood using the relations for the shock wave of a powerful cylindrical
explosion. The explosion can be considered to be strong as long as energy
is released in a thin channel and the pressure of the explosion wave does
not fall close to the atmospheric pressure. In this case, the flow is self-similar.
The shock front radius rs and pressure p in the affected region depend, within
the accuracy of numerical factors, on W 1 and t as r, ( W 1 / p o ) 1 / 4 t and
N 1/2
p N Wl/rz ( Wlpo)'/2t-',where po is cold air density. The channel
N
this occurs for the time t W'/'. These relationships are qualitatively
N
consistent with the calculations for the two variants described in table 4.1.
whose current is not low on the arc scale; then the conductivity is
G,, = i / E L . In a mature gas-dynamic process when the shock wave is still
strong, the resistance will drop with time. As the shock wave becomes
weaker, the decrease in R1 and the increase in G become slower. These
tendencies are described by the relaxation-type formula
dG
- - i/EL - G(t) -
- Gst(i) - G(t) (4.38)
dt Tg Tg
where Tgis the characteristic time of linear conductivity variation (relaxation
time). In a simple case with i = const, Tg = const, and G(0) = 0, we have
G = GSt[1- exp(-t/T,)].
Equations (4.24) are solved with the initial conditions U ( x ,0) = U, and
i ( x ,0) = 0, RI(x30) = RI,-;the reactive parameters are taken to be constant:
C1 = 10 pF/m and L1 = 2 pH,”. The channel does not close on the earth
in an instant but does so through the time-decreasing resistance of the
commutator (similarly to the real lightning length decreasing through the
streamer zone). The accepted values of R,,, = R(0)exp(-cut), R(0) = 10 R
and cu = 1 ps-’ provide a typical duration of the negative current impulse
front tf FZ 5 p . The boundary condition at the grounded end of the line
raises no doubt: U ( 0 .t ) = i(0,t)R,,,. The problem of the far end up in the
clouds, x = H , is much more complex. Conventionally, it is considered as
being open, assuming i ( H ,t) = 0. In reality, the situation is far from being
self-evident. When the line gets discharged and its end in the clouds takes
zero potential, a high electric field must arise near it due to the voltage
difference A U = - U o ( H ) . This gives impetus to very intensive ionization
processes, probably involving high current. This situation will be partly
discussed below. Now, we shall assume the upper end to have no current.
The results to be presented were obtained for a vertical unbranched channel
with the total length H = 4 km. This is the height the ascending leader tip
reaches when the descending leader, which has started from the point closest
to the earth in the bottom negative sphere with the centre 3 km high, contacts
the earth. It is normal practice to use the following averaged parameters
of the leader prior to the contact: EL = 10V/cm, iL = lOOA, and
RIL = EL/iL= 10 O/m. For a realistic description of the resistivity dynamics
(section 4.4.3), the relaxation time should be taken to be Tg = 40 ps, when the
t The motion of a high current wave with attenuation but without noticeable distortions
facilitates the electromagnetic field calculation necessary in many applied problems of lightning
protection and in substantiation of remote current registration methods.
0 1 3 4
X,2km
t=8p
0 1 2 3 4
x,km
I,,,= = 67 kA
U
3
\
.3 0.4 -
0.2 -
0.0 I . , . , . , . ,
0 1 2 3 4
x, lan
1- = 8,15 kA
0.8
1
0.6
.I
0.4
0.2
0.0
0 1 2 3 4
Figure 4.18. Deformation of the current wave front at leader potential (top)
Ui = -50MV and (bottom) -10 MV; for the other parameters, see figure 4.17.
propagation along the channel. There is no damping at a very high current and
the impulse front becomes steeper, as was discussed in section 4.4.2 (figure
4.18). Non-linearity is also observed in the current amplitude dependence
on the initial potential U, at the earth. If the commutator were perfect
(R,,, = 0), the current at the earth at the moment of contact would instantly
200 1 r 0.6
Voltage, MV
Figure 4.19. Calculated dependencies of the current amplitude and average wave
velocity in the return stroke on the leader potential Vi.
0 1 2 3 4
x, km
-8a& 15--
10-
5-
0-
-5 -
Figure 4.20. Current and potential distributions during the propagation of waves
reflected by the cloud end of the channel.
When the reflected wave reaches the earth, delivering a positive potential
to it, a new discharge cycle begins. The newly acquired positive charge flow-
ing into the earth is equivalent to the extracted negative charge. The current
sign at the grounded end is reversed (figure 4.21). In the absence of
dissipation in a distributed system such as a long line, undamped oscillations
with a period T = 4H/w, would arise similar to those in an LC circuit.
Nothing of the kind is observed in lightning registrations, nor is there a
single change in the current direction. This means that the discharge wave
is either not reflected by the upper end of the line or the reflected wave
becomes so damped on the way back to the earth that it is unable to manifest
.-
-0.4 j U
Figure 4.21. Calculated current impulse through the grounded channel end.
end of the line, which generates a reflected current wave of the same sign and
amplitude as the incident one. As the capacitance becomes discharged, the
reflected wave amplitude decreases and then the sign is reversed. The com-
pletely discharged capacitance, incapable of supporting current, eventually
becomes equivalent to an open line end. It is clear even without a numerical
calculation how much the current changes at the earth after the arrival of a
reflected wave of such complexity. It should be emphasized again that
nothing of the kind has ever been observed in real lightning.
One can also try to rectify the situation by complicating the boundary
condition with the allowance for the final resistivity of the ‘metallized’
cloud region. The streamer and leader branches filling the cloud possess a
resistance at the moment of their generation. A leader branch can hardly
be heated as much as a single descending leader. The resistance of the
‘metallized’ cloud, R,,,is quite likely to be high during the whole return
stroke stage. If this is so, the boundary condition should be formally repre-
sented as i ( H ,t) = C, dU/dt - R,,dildt. Strictly, it is not only the boundary
condition that changes in this case, like in the case of ideal metallization, but
also the set of equations. The cloud potential U. can no more be considered
as being constant in time. The second equation of (4.24) should be re-written
as
having taken into account the change in U, due to the change in the cloud
charge Q,. Then the function Uo(Q,) must allow for the delay because of
the finite rate of the electromagnetic field propagation. The problem becomes
extremely complicated. Although radar registrations do indicate the develop-
ment of a wide network of branches in clouds, there has been no investigation
of cloud ‘metallization’. The reason for this, no doubt, is the lack of initial
data.
One should not discard two other factors unaccounted for by the numer-
ical model. First, a leader channel can practically never be single. Owing to
the numerous branches of different lengths developing at different heights,
numerous reflected waves will arrive at the earth at different moments of
time, creating a sort of ‘white noise’ with a nearly zero total signal. This
will deprive the current of its characteristic bending which is usually created
by a single reflected wave at the earth. Second, constant linear capacitance
only approximately describes the real re-charging of a lightning leader. We
have mentioned above that the cover charge around a leader channel is
changed by numerous streamers starting from it. Their velocity decreases
rapidly when the streamer tips go away from the channel surface with its
high radial field. So, when the voltage at the wave front changes, the
charge near the channel changes almost immediately, while its change at
the external cover boundary occurs with a delay. In other words, a lightning
discharge can proceed for a fairly long time. The quasi-stationary current
from the discharge of the cover periphery, having the same direction as the
current in the forward wave, can compensate for the reverse current induced
by the wave reflected by the earth.
To conclude, the model of a single long line with varying linear resis-
tance allows elucidation of many aspects of the return stroke but cannot
claim to give reliable quantitative description of all characteristics of this
phenomenon. The much more simplified models of return stroke are usually
used calculating electromagnetic field for technical application. A review of
this model is given [30].
However, when the ascending leader penetrates the charged cloud region
(positive, in this case), a large potential difference arises between the front
end of its grounded channel and the space around it, so the leader current
has been found from many registrations to rise to several kiloamperes.
This event seems to be triggered by the same mechanism, but its effect is
greatly enhanced by the leader hitting the very centre of a large cloud
charge of, say, Q, x 30C and radius R, M 500m, where the potential is as
high as U, M 500-800MV. At such voltages, the streamer zone and cover
appear much extended. Negative streamers develop until the average field
in their streamers drops below E, M 10 kV/cm under normal conditions (or
1.5 times less at a 5-6km height [16]). Streamers elongate very quickly
when the field is higher. Therefore, a very powerful streamer corona con-
sisting of numerous branched streamers (they are likely to originate not
only from the stem but from its numerous branches, too) will fill up a
space of size R M Uo/EsM R,. The negative charge of the streamer zone
will partly neutralize the positive charge of the cloud cell. If the streamers
have velocity U, x 106m/s, they will fill the charged cloud region for
t x Rc/vU,M s. Since the capacitance of the leader portion inside the
cloud, CL,is comparable with that of the charged cloud region, Ccl,
When discussing the negative leader in section 4.3.2, we put off the considera-
tion of its stepwise behaviour until the reader has become familiar with the
return stroke, since a similar phenomenon is the principal event occurring
in each step. It would be reasonable, at this point, to turn to the nature
and effects of the stepwise leader behaviour. But we should like to warn
the reader that there is no clear answer to the question why a negative
leader has a stepwise structure while a positive one has not. Nonetheless,
some observations of stepwise positive lightning leaders were presented in
[32]. This phenomenon has never been observed in laboratory conditions.
t The fact that intercloud discharges do neutralize charged regions is supported by electric field
measurements, and high currents that flow through lightning channels are indicated by peals of
thunder.
Figure 4.22. The potential distribution for various stages of a step formation. The main
leader potential U ( x )is counted from the external potential U,:(top) secondary leaders
1 and 2 are formed at point 3 at the streamer zone end; (centre) the tip of a positive sec-
ondary leader has reached tip 4 of the main leader, and a discharge wave has started its
travel along the secondary leader channel (dashed line); (bottom) the main leader tip
after the step formation has taken a new position, and the process is repeated.
potential U 1 and the external (for the tip) potential U ( x l )at the tip site x1
increases (figure 4.22 (top)). Second, the streamer zone field of the main
leader, E, % 10 kV/cm, which must support the generated negative streamers,
is higher than the field E, x 5 kV/cm required for the development of positive
leaders. For this reason, the streamers generated by the secondary positive
leader tip develop in a fairly strong field, become accelerated and all reach
the main leader tip. Since the long channel of the main leader has a
capacitance greatly exceeding that of the short secondary leader, it absorbs
completely all charges carried by the positive streamers. In other words,
the secondary positive leader develops in the final jump mode. We know
from section 2.4.3 that this leads to its acceleration. The secondary negative
leader, on the contrary, develops in a decreasing field beyond the streamer
zone of the main leader, whose streamers stop in space, so it moves much
more slowly, similarly to the main leader.
When the tip of the secondary positive leader comes in contact with the
main channel, the positive leader experiences the transition to the return
stroke. Charge variation waves run along both channels, as described in
section 4.4, and their potentials tend to become equalized (figure 4.22
(centre)). But the capacitance of the kilometre length channel is much
higher than that of the shorter secondary channel, so their fusion results in
establishing a potential only slightly differing from the initial potential of the
main leader tip, U,.
The moment at which the potential U1 is taken off the secondary leader
channel and the latter joins the main leader, manifests the end of the step.
The main leader tip ‘jumps’ over to a new place, the one occupied previously
by the tip of the secondary negative leader, delivering to it its potential U ,
(figure 4.22 (bottom)). The tremendous potential difference that arises in the
vicinity of the newly formed tip at this moment produces a flash of a powerful
negative streamer corona, which transforms to the novel streamer zone of the
main leader. Then the sequence of events is repeated. The combination of the
charge utilized for the short recharging of the secondary positive leader and
for charging the secondary negative one, plus the charge incorporated into
the new streamer zone, create the step current impulse. (Recall that there is
always a local current peak at the streamer tip or in the leader streamer
zone, related to the displacement of the charge in this region; see sections
2.2.3 and 2.3.2.) Part of the step impulse creates a current impulse in the
main channel and the other part is spent for the formation of a new cover
portion. The charge Q, pumped into the main channel can be evaluated in
terms of the mean velocity of the stepwise leader, wL M 3 x 10m/s, the
length of a step Ax, M 30m, and the current iL x lOOA averaged over the
whole duration of the process. Since the time between two steps is
At, x Ax,/wL M lop4s, the charge is Q, M iLAt, x lop2C.
pumps some of the charge into the leader cover. This process, the cover
ionization in particular, requires much energy. But even if the energy released
in the channel is assumed to be W M lo4 J, t h s is still a very large energy.
The power required to support an average current of lOOA in remote
channel portions, where the effects of current impulses are averaged and
smeared, should be P1 M lo5Wjm. It is removed from the channel by heat
conduction and, partly, by radiation. These parameters correspond to the
maximum channel temperature T M 10 000 K, field E x 10 Vjcm, and resis-
tance R1 M 10R/m taken for the above estimations (section 2.5.2). At this
power, the energy released between two flashes per unit channel length will
be Wl,, M 10J/m for the time At, M lOP4s. Therefore, the single pulse
energy W would be sufficient to support a channel 1 km long. In reality, a
step pulse is damped at a much shorter length. At a distance of about
1 km, the step effects are smeared and the energy released in the channel
becomes totally averaged. But at a short distance from the tip, the energy
effect of the step is very strong, as is indicated by the intensive flash. The
temperature registered in some measurements was as high as 30 000 K [35],
i.e., the same as at the wave front in a return stroke.
Let us evaluate the distance at which the energy effect of an individual
step is still essential. When a short step joins a long channel, charge Q, is
pumped into the channel for a short time. Since we are interested in distance
and time much larger than the real length and duration of a charge source, let
us assume the source to be instantaneous and point-like, as is usually done in
physics: a point charge Q, is introduced at the initial point of the line, x = 0,
at the initial moment of time t = 0. The resistance of not very short channel
fragments, Rlx, is higher than the wave resistance, so the inductance effect
will be neglected. At an average resistance of 10R/m, this distance is just
about the step length Ax,. At shorter distances, the instantaneous point
source model is invalid, since it implies an infinite initial voltage and
energy. They drop to realistic values only if the charge affects a length
exceeding Ax,, at which the source was actually placed. Therefore, with
the neglect of inductance (and the precursor), the line charging to potential
U,(x, t ) above the background potential is described by equations (4.36).
On the assumption of R1 = const ,t they have an exact solution correspond-
ing to heat flow from an instantaneous lumped source:
I.The value of resistance R Ito be taken for evaluations may be smaller than that in the leader,
having in mind a transformation of the channel due to the step current.
The power released by the current step per unit channel length is
described as
(4.42)
(4.43)
For the time of the pulse action, the energy released per unit length at point x
is
W, M 1; Q2
iiR, dt = 2x - -
W Ax,
d 1 x 2 Ax, (
x
, x > Ax, (4.44)
where W is the total energy injected into the channel by the pulse, with its
upper limit given by formula (4.39). The effective duration of energy release
from a single step at point x is expressed as
X2
At, Wl
- 2.2t - -. (4:45)
- 2.7%
N
PImax
Consequently, the contribution of charge injection to the energy release at
a given channel site decreases in the direction of perturbation propagation as
Wl x x - ~and is independent of R 1 . The latter fact justifies the use of
R1 = const without reservations concerning the resistance variation during
the current impulse passage. The energy pulses released at point x owing to
the two subsequent steps superimpose at x > (2.7xAts)ll2; t h s critical
distance follows from the condition At, M At,. For example, at the average
resistance R I = 10 R/m, with x = 1010 m2 /s and step frequency At, = lop4s,
this happens at a distance x x 1.6 km in t , M AtJ2.2 M 45 ps, after the pulse
arrival here. Thus, the effects of energy release from individual steps are
detectable even along an extended lightning path, and this is the cause of
observable flashes of almost the whole channel. For a flash to arise, there is
no need for a strong energy effect. A short temperature rise of, say, above
l000K over l0000K would be sufficient for a flash to be detected by
modern photographic equipment.
The channel energy is affected by the temperature rise above the average
background, rather than by the time separation of the energy pulses between
two subsequent steps. In this respect, the impulse effect on the channel during
the wave propagation is damped at distances close to the tip. The plasma
temperature modulation determining the flash intensity at large distances is
due to the imbalance between the energy release and heat removal from
the channel during the pauses between the steps. There is no imbalance
at a large distance from the tip after the channel development has been
established. At T x 10 000 K, the losses for air plasma radiation are not par-
ticularly great but become appreciable at T M 12- 14 000 K. The Joule heat of
current is eliminated from the channel primarily by heat conduction. This
process occurs at constant (atmospheric) pressure when the energy release
is moderate, as is the case for distances of hundreds of metres from the tip.
At T M l0000K, the air heat conductivity is X x 1 . 5 ~ WjcmK and
the thermal conductivity at pressure p = 1 atm is XT = X/pcp x 180cm2/s,
where p is air density and cp is heat capacity. The average conductivity in
the channel corresponds to a temperature lower than the maximum tem-
perature. To illustrate, at c x 10 (Cl cm)-’ corresponding to T = 8000 K,
the effective radius of a conductive channel is r M (mrR1)-”2M 0.6cm
for R1 = 10R/m. The heat is removed from the channel for time
t r 2 /2xT
N N s, an order of magnitude longer than the pause between
the steps. The repeated energy pulses dissipate rather slowly, and the
temperature modulation relative to its average value T M 10000K is not
large at long distances x.Indeed, the energy released in the remote channel
portions during a pause is Wla, x PlavAt,x 10 Jim at an average power
PlavM lo5 Wjm. Even if we assume that all energy of a step is released in
the channel and W/Ax, x 2500 Jim in (4.45), the excess of the pulse release
over the average heat removal, which is equal to the average energy release
Wla,, will be small at x > Ax,( Wl/Wlav)1’2x lOAx, zz 500m. With allow-
ance for other energy expenditures, this reduction in the pulse effect will be
evident even at shorter distances. This circumstance makes the use of average
parameters reasonable in the consideration of the evolution of long stepwise
lightning leaders, ignoring the stepwise behaviour effects. In any case, labora-
tory experiments show that there is no appreciable difference between a
positive continuous and a negative stepwise spark discharge as for the
velocity, average leader current or breakdown voltage in superlong gaps.
However, even a small excess of the average temperature over its aver-
age value may be sufficient for a flash to be registered optically. As for chan-
nel portions located at a distance of one or two steps from the tip, the energy
pulses and outbursts of temperature and brightness are found to be very
strong there. A gas-dynamic expansion of the channel is also possible, as
happens in the return stroke (section 4.4), although it occurs on a smaller
scale. No doubt, a flash is also produced by a powerful impulse corona
giving rise to a new streamer zone of the elongated leader. Photographs
show that the transverse dimension of a step flash is about 10m [38].
The processes in the lightning channel following the first component are
known as subsequent components. Of interest among these are so-called
M-components and dart leaders. In the first case, the current impulse
registered at the earth has a very smooth front (0.1-1 ms), a similar duration
and an amplitude of several hundreds of amperes, sometimes of 1-2 kA. The
channel radiation intensity increases abruptly during the impulse, but one
can hardly identify in the photographs a structure similar to the impulse
front. The current impulse of an M-component is always registered against
the background of about 100A continuous current of the interpulse
pause. For a dart leader to arise, this current must necessarily be cut off
[39,40]. A few microseconds after the cut-off, a short high-intensity region
- a dart leader tip - runs down to the earth along the previous channel
with a velocity of -107m/s. The contact of the dart leader with the earth
produces a return stroke with its typical characteristics but having a much
shorter impulse front than in the first component (less than 1 ps or even
0.1 ps in some impulses). It is hard to say anything definite about the lower
limit of the front duration: it is likely to lie beyond the time resolution of
the measuring equipment.
The papers published almost simultaneously [41, 421 interpret the sub-
sequent component qualitatively as representing the discharge, into the
earth, of an intercloud leader after its contact with the upper end of the preced-
ing grounded but still conductive channel. Here we describe the evolution of an
M-component in terms of a numerical simulation.
The model underlying the simulation is as follows (figure 4.23). Initially,
there is a grounded plasma channel of length HI with zero potential, which
was left behind by the preceding lightning component. At time t = 0, a leader
channel of length H 2 and potential Ui joins it in the clouds (the voltage drop
from the leader current and from the intercomponent current is neglected).
The short process of the channel commutation through the streamer zone
Figure 4.23. The formation of a subsequent component: (a) the grounded channel of
the previous component and an intercloud leader; (b) channel charging-discharging
waves.
25001
2000 f
d
g 1 500
C
-
E
61000-
500-
0-
is damped, the channel resistance increases gradually due to the gas cooling.
But if the intercomponent current is comparable with the leader current, as
is usually the case by the moment the M-component arises, the increased
resistance of the grounded channel may be suggested to be limited by the
value of R I , M r l L . The reactive parameters of both lines, which are not
very sensitive to the channel plasma state, can be taken to be identical to
those of the leader: C1 M 10pF/m and L1 M 2.7 pH/m.
During the intercloud leader discharge into the earth via the preceding
channel path, the channel resistances change, as in the return stroke
14
12
.
E
3 10
28
g6
b
7-
& 4
0
0 i 2 3 4
x, km
800 -
<
600-
m
a,
a,
c,
4-3
400-
E
t,
200-
(section 4.4.4).Suppose that these changes follow the relaxation law expressed
by formula (4.38). This formula describes adequately the qualitative tenden-
cies; there are no quantitative results to compare them with.
This process is described by the long line equations of (4.24) with the
following initial and boundary conditions:
U ( x , O ) = 0 at 0 < x < H I ,
t The current impulse of a return stroke is of a different form. The current amplitude is registered
right after the short-term commutation process when the leader contacts the earth via its
reducing streamer zone, which takes a few microseconds.
w, M 10’ m/s and M 0.1 ps, the length of the region with an abrupt poten-
tial drop in the dart leader front is Sx M 10m. It is quite possible that this
value is actually smaller because the return wave cannot at first gain the
full return stroke velocity U, M lo’,/,. On the other hand, the potential
drop region should not be shorter than Ax M vdL?f x 1 m, since the cross
section of a channel with total potential U approaches the earth at velocity
‘udL. Such a steep front of 1-10m is unattainable not only by a diffusion
wave with its potential varying along many hundreds of metres (figure
4.24) but even by an ‘ordinary’ leader of the first lightning component, in
which Ax is determined by the streamer zone length. At the moment of
contact with the earth, the latter is measured in dozens of metres at the tip
potential of 20-30MV. This is the reason why the time necessary for the
return wave current to grow to its maximum value is dozens of times
longer than that for the dart leader.
It follows from the foregoing and the fact that a dart leader travels as
fast as a very fast streamer that the former has no streamer zone which
would serve as the primary prerequisite for a leader mechanism. It appears
that the dart leader, contrary to its name, is essentially not a leader, although
it has a charge cover, which it has to acquire under somewhat different
circumstances (see below). Nor does it look like a diffusion wave of the M-
type. The latter would have an order of magnitude higher velocity and a
very diffuse front.
A dart leader looks more like the oldest of the known types of propa-
gating plasma channel - a streamer, whose head represents an ionization
wave. The velocity of a dart leader is close to that of a high voltage streamer.
The principal reason for a streamer channel being non-viable in air - a rapid
loss of conductivity by the cold plasma - is very weak in this case. A dart
leader follows the track heated by the preceding component, so that the
still-hot track serves as a kind of waveguide to the leader. The high gas
temperature greatly retards electron losses. Therefore, the possibility for the
region behind the tip to be heated to arc temperatures increases considerably.
This provides a stable highly conductive state inherent in a ‘classical’ leader.
The preheated air pipe serves another, probably more important, func-
tion. Its hot and rarefied air is surrounded laterally by cold dense air, Since
the rate of ionization due to the field is described by the E / N ratio, the
radial expansion of the channel region behind the streamer tip is abruptly
retarded as compared with the forward motion of the ionization wave. So
the air mass to be heated by the current is reduced, permitting the channel
gas to be heated to a higher temperature. The cold air restricts the channel
expansion because it acts as a charge cover produced by the streamer zone
of the leader.
One should not think that the channel does not expand through the
ionization mechanism at all. This process is just much slower than the
forward motion of an ionization wave towards the earth, so most of the
Joule heat is released into the yet unexpanded channel having a smaller
radius. The radial field leads to the channel expansion only at the beginning,
as is the case with common streamers (section 2.2.2). When the radial field is
somewhat reduced, the channel becomes the source of a radial streamer
corona which does not require a high field. Radial streamers rapidly lose
their conductivity in cold air, and their immobile charges form a cover of
the type that surrounds a common leader channel. Now, though with some
delay, the mechanism of radial field attenuation and hot channel stabilization
comes into action. Thus, a dart leader, being essentially a streamer (i.e., an
ionization wave having no streamer zone in front of the tip), must possess
a charge cover, as a leader. Unlike the case with a common leader, the
cover is not inherited from a streamer zone but is formed entirely behind
the tip, which is the seat of the principal processes driving the dart leader.
(In a classical leader, the cover formation partly continues behind the tip,
(4.47)
(4.48)
(4.49)
and is expressed directly through the local potential. Indeed, reducing the
rank of the set of equations (4.48) and (4.49) by dividing them by one
another, we find
w, - w,o = c1u2/2 (4.50)
where Wlo is the initial energy in the channel far out the wave front. Thus the
statement repeatedly used in evaluations that the energy dissipated in the
channel is of the same order as that stored in its capacitance is valid exactly
in the stationary case.t
We shall consider moderate waves, when the gas is heated at constant
pressure, and the Joule heat is released at constant mass m = r r 2 p = mopo2
per unit channel length (yo and p are the initial radius and gas density in
front of the wave). Then we have W1 = mh, where h is the specific gas
enthalpy. Assume for simplicity that thermodynamically equilibrium ioniza-
tion is established at every point of the wave, so that conductivity 0 and
x = ma(pC1)-’ are the functions of temperature T or h ( T ) . Then x is
unambiguously related to U through formula (4.50).$ With (4.48)-(4.50),
finding the distributions along the wave reduces to the quadrature
1- = -2vx. U=[
2m(h - h,)
cl ] 1’2
. ho=-.w
ml o (4.51)
t This is quite natural because under the problem conditions the channel is not created anew but
exists from the very beginning with its linear capacitance C , . Then every channel portion is
charged as lumped capacitance (cf. the comment o n formula (2.17) in section 2.2.4).
$In sections 4.7 and 4.4, the quantity x was related to electrical parameters through relation
(4.38) which refers to strong waves with a high energy release. If desired, one can use this relation
after substituting aG/ar by -U dC/dx and doing the above operations.
start, in front of which (x > 0) the channel transforms very slightly, so that x
increases slightly, followed by (x < 0) where it changes noticeably. The
parameters of the initial front point will be marked by the subindex 1, assum-
ing for definiteness x1 = 2x0, where xo corresponds to the initial channel
conductivity. Then we have h l - ho = Sho, where S = 2l/" - 1. An exponen-
tially damping tail of the electric field and current extends forward along the
wave where the diffusion is 'linear':
This point is closer to the a priori position of the front start x = 0 than Ax,
which justifies the approximations.
Let us illustrate this situation numerically with reference to the conditions
typical of the M-component (figure 4.25). Suppose the diffusion wave has
velocity v = 10' mjs running along a channel with the initial radius ro = 1 cm,
temperature To = 5900 K (h, = 14.8 kJ/g) and po = 5 x lop5g/cm3 which
is by a factor of 25 less than the normal; m = 1.54 x 10-4g/cm;
the initial linear resistance R1 = 10R/m, xo = 10 m /s
10 2
ne 1 . 8 10'4cm-3,
~
and C1 = 10pF/m. For the temperature range T x 6-10000K in air at
1 atm, we have a l p = 17h3 (where .[(a cm)-'], p[g/cm3], and h[kJ/g]).
a
x, 0 A x
Figure 4.25. Schematic diagram of the nonlinear diffusion wave front.
the effective field length before the wave front Ax = 100m. The point behind
the wave with U = lOMV (h M 50 kJ/g, T M 100OOK) lies at a distance
x = 500m from the front. There, x M 3x10" m2/s, the resistance is by a
factor of 30 lower than before the front, and the field drops to 33 V/cm. The
field maximum lies near the initial front point. The qualitative picture presented
in figure 4.25 agrees with the numerical results of figure 4.24(a).
(4.55)
E = iR1 =
(..xot) lI2
exp (- A).
t4x0
At every point x, the field first rises with time but then falls after the
maximum E,, = 2(7re)-li2U,/x at moment t = x2/2x0. The point E,,
moves at velocity wg = xo/x, and the potential at this point is U , = 0.33U1.
An ionization wave can be formed if the maximum field is sufficiently
high and exceeds a certain critical value E,. The ionization wave is assumed
to propagate at velocity w, supposed to be equal to that of a dart leader. Since
E,, Nx-l tC1I2, the ionization wave could principally arise at earlier
N
times when E,, > E,, but if its velocity is U, < wg,is immediately overcome
by a diffusion wave. This will not happen if wg drops below U, while E,, is
still higher than E,, i.e., if the conditions vs 3 wg, E,, 2 E, are fulfilled
together. For this to happen, the diffusion coefficient must be smaller and
the linear channel resistance larger:
(4.56)
For this, the gas temperature in the initial channel should not be high. On
the other hand, for the 'waveguide' properties to manifest themselves, the
temperature must be as high as possible to make the air rarefied. Because
of the very sharp temperature dependence of conductivity (when it is
low), these conditions are met only in a very short temperature range,
T M 3000-4000K, where the air density is by a factor of 10-15 lower than
-U-
dX = uix, vj =Nf(E/N). (4.57)
dx
This equation describes a new law for the channel transformation. Owing to
(4.48), the ionization frequency 4 turns to the potential function. Then, by
dividing (4.48) and (4.57), the problem is reduced to the equation for x( U )
and the quadrature, as in section 4.8.2.
(4.58)
where Axi is the extension of the ionization region from the initial to the final
point of the wave. For a wave to survive, its parameters must meet the last
inequality of (4.59). Otherwise, the field within the wave will be unable to
exceed E*, so no ionization will occur.
The capabilities of an ionization wave are limited, and this limit increases
with increasing initial conductivity of the medium. For example, if the initial
parameters ne0 M lOl2cmP3 and x0 x 107m2/s were even lower than the
critical values found in section 4.8.3 and if the threshold field was 3 kV/cm,
it would be necessary to have the ionization frequency 4 = 2.1 x lo6 s-'
and potential U2 = 300MV in order to increase ne and x by three orders of
magnitude (Ax = 1 m, U 1 = 0.3 MV) and U2 = 30 MV by two orders. The
wave width in t h s case is Axi E 22m, i.e., it is very extended. Only when
t Sometimes, it seems better to describe vi the function of E and, on the contrary, to remove U
from (4.48)and (4.57). Instead of (4.48), we then get
the initial conductivity is still an order of magnitude lower (nCo= 10" cmP3,
xo = lo6m2/s, and R = lo5fl/m), the wave width begins to approach what
would be desired for a dart leader. In the same medium at the same U
and E*, the parameters necessary for the ratio x 2 / x 0= lo3 would be
~ sP1, U2 = 30 MV, Ax = 10 cm, Axi = 5 m, and U , = 30 kV.
vi = 1 . 4 lo7
A still narrower region of the potential rise would be obtained at a still
lower initial conductivity. But then we approach the applicability limits of
the basic concepts of the theory of perturbation propagation in a conductive
medium and of the long line theory, and we are probably coming closer to the
understanding of criteria for the dart leader production.
where Er0is the radial field on the surface of a conductor of length I >> yo:
(4.62)
If the longitudinal field varies along the channel so slowly that the axial diver-
gence can be neglected (the characteristic length for the variation of E, is
Ax >> ro), we arrive at one of the basic conceptions of the long line theory,
~ ( x=) C1U(x), whose implication is the potential diffusion mechanism. It
is suggested implicitly that the resistance varies very slowly along the
channel, so this variation cannot be an obstacle to a charge flux, making
the flux velocity decrease abruptly and create a space charge due to its
local accumulation (a long line has no 'jams').
However, space charge does accumulate at a sharp boundary between a
poorly- and a well-conducting channel portion. A charged tip is formed at
the end of an ideal (or non-ideal) conductor, the potential in front of it
drops abruptly, at distances about equal to ro, inducing there a strong field
capable of sustaining an ionization wave. This is what happens in a
common streamer in a non-conductive medium. It is then clear what is
necessary to support a sharp potential drop at the ionization wave front for
a long time. The conductivity along the perspective trajectory must drop to
a value low enough for the diffusion field tongue to be unable to smear the
sharp potential drop. Therefore, the tongue length must become comparable
with the channel radius Ax x xo/v N yo. Because of the strong temperature
dependence of the degree of equilibrium ionization in air at low temperatures,
a drop to T x 3000K would be sufficient. The equilibrium electron density
established for the long zero-current pause will be neo 10'o-lO1l cmP3;
- -
hence, Rlo 105-1060jm, and xo 106-105m2/s. But the air density in
-
the cooled channel of the previous component at T x 3000K is by an order
of magnitude lower than that of cold air, so that the conductivity drop will
not interfere with the 'waveguide' properties of the track.
The velocity of a dart leader as an ionization wave is defined, in order
-
of magnitude, by the same formula (2.2) as the streamer velocity. But the
'pre-ionization' in this case (nee 10'o-lO'l cmP3) is considerable, and a
much smaller number of electron generations (1n(ne2/neo)x 5 ) is to be
produced in the wave. With the account of the similarity law for vi at
vi - --
an order of magnitude lower gas density, the ionization frequency is
10" sC1 and ro 1 cm; then we obtain a correct order of the velocity
U = qro/ In(ne2/ne0) io7 mjs.
One cannot say that all the details of the dart leader behaviour have been
clarified by the above considerations. For the dart leader channel to be well
conductive, the electron density in it must be at least 5-6 orders of magnitude
higher than the initial value for the track. But the capabilities of the ioniza-
tion wave to produce more electrons are limited. The maximum conductivity
of an ionization wave propagating through a non-conductive medium is
defined, in order of magnitude, by the relation ~ F ~ ~vi (section~ / -
E 2.2.2),
because the space charge of the streamer tip, providing a strong ionization
~
field is dissipated with the Maxwellian time TM = Eo/cF.t After the wave
t It also determines the rate at which the linear charge T = C1U is established, if it is, in the
channel. Let us integrate the relation for charge conservation in the conductor cross section.
Neglecting, for simplicity, the variation in 0 along the channel length, we obtain
Using (4.61) and (4.62), we arrive at a refined equation for the relation between T and U :
The postulate of the long line theory, T = C , U, is valid if the changes in the system, which also
define ~ ( t )occur
, slower than with T , =~ co/u. When applied to the wave front moving at
velocity U in a line with conductivity uo,this happens at 00 >> u i o / r o and xo >> vr0.
has passed, the channel still needs to be heated and ionized, but both pro-
cesses are to occur in a moderate electric field, as in a classical leader channel.
Besides, this must take place before a strong radial field makes the channel
expand beyond the hot gas tube, or if it has already become enveloped by
a stabilizing charge cover (section 4.8.1).
There are still many questions about the processes in a dart leader that
remain to be answered; the development of its quantitative theory is also a
task of further research.
To conclude, it is worth noting some specific features of a current
impulse in the return stroke of subsequent components. Generally, the
impulse duration is related to the time it takes the return stroke to run
along the whole channel. For the subsequent components, this time must
be longer than for the first component due to the attached intercloud
leader. But the impulse duration in the subsequent components is about
twice as short, although the return wave velocities are generally the same.
The reason for this difference is likely to be the absence of branches in a
dart leader. It is quite possible that the relatively slow process of their re-
charging elongates the current impulse tail of the first component. The
impulses of the subsequent components do not reverse the sign, similarly
to those of the first one. In the absence of branches, the action of the reflected
wave can no longer be screened by the randomly reflected waves of the
numerous branches (section 4.4.5). The hypothesis of ‘white noise’ should,
probably, be discarded as being inadequate. This problem, like the others
above, awaits its solution.
AE(r)= -
aq [ +
lnH ( H 2 +r 2 y 2
- H-h ] (4.63)
27rEO +
h (h2 + r2)1'2 +
( H 2 r2)1/2
where H is the height of the grounded channel. If H is, at least, several times
larger than r, the dependence of field A E on the distance between the
registration point and the channel line will be only logarithmic. The same
is true of the front duration of a field pulse.
The situation must be quite different for a dart leader with the abrupt
potential drop at the wave front, since the first approximation in the field
calculation may assume a uniform potential along the channel and
~ ( x= ) const at x > h. This gives formulae (3.6) and (3.7), which yield the
maximum value AE,,,(r) N r - ' . Such a large difference in the field variation
is easily detectable experimentally, especially if we remember that it concerns
not only the field pulse amplitude but also its front rise time. To see that this
is so, it is sufficient to introduce into (4.63) and (3.6) the h-coordinate for the
wave front, expressed through the respective velocities: h = H - ut.
Triggered lightning is a perfect source for such measurements. A triggered
lightning is initiated by launching a small rocket raising a very thin wire which
evaporates during the development of the first component. The point of
contact of the lightning with the earth is strictly defined, so it is easy to
position current detectors at the necessary distances. Besides, the channel at
the earth follows the wire track and is strictly vertical, as is implied in
the numerical formulae. Such measurements have been partly made [44-451.
In section 3.5, we discussed the measurements of field A E at distances
r1 = 30m and r2 = 500m from the channel during the dart leader develop-
ment. These measurements were not synchronized. However, the ratio
AE(30)/AE(500) = 17.4 for approximately equal currents is nearly the
same as r2/r1 = 16.7.
The field measurements for M-components have been reported only for
r = 30m [42]. The oscillogram of A E ( t ) is accompanied by a simultaneous
registration of a current impulse with the amplitude of 800A and the front
rise time -100 ps. The duration of the impulse front A E is approximately
the same, but the field reaches its maximum of 1350 V/m earlier, when the
current has reached half of its maximum amplitude (until the potential
wave arrives, the current at the earth is zero, whereas the field begins to
rise since its start). Figure 4.26 shows the calculated functions i(t) and
A E ( t ) at the observation points with r = 30m and 500m. The long line
model described in section 4.7.1 was used with the same C1 = 10pF/m,
L1 = 2.7 pH/m, and R I (0) = 10 njm. The length of the grounded channel
was 4000m and that of the intercloud leader contacting it was 2000m. The
experimentally observed current of 800 A was reproduced in the calculation
at the leader potential U, = 9.7MV. Under these conditions, the field
Figure 4.26. Calculated variations of the electric field at the earth’s surface due to the
M-component under the conditions of figure 4.24. The dashed curve shows the
current impulse I .
References
[41] Bazelyan E M 1995 Fiz. Plazmy 21 497 (Engl. transl.: 1995 Plasma Phys. Rep. 21
470)
[42] Rakov V A, Thottappillil R and Uman M A 1995 J . Geophys. Res. 100 25701
[43] Sinkevich 0 A and Gerasimov D N 1999 Fiz. Plazmy 25 376 (Engl. transl.: 1999
Plasma Phys. Rep. 25 339)
[44] Rubinstein M, Rachidi F, Uman M A et a1 1995 J . Geophys. Res. 100 8863
[45] Rakov V A, Uman M A, Rambo K J et a1 1998 J. Geophys. Res. 103 14117
In this chapter, we shall describe the way a lightning channel chooses a point
to strike (a terrestrial or a flying body). This is the principal issue for lightning
protection technology. In any case, a direct stroke is more hazardous than a
remote lightning effect via the electromagnetic field or shock wave in the air.
Historically, direct lightning strokes were observed earlier than indirect ones,
and the first research into lightning protection problems was associated with
direct strokes.
Everyday experience and scientific observations, including those made
as far back as the 18th century, indicate that lightning most often strikes
individual structures elevated above the earth. These may be towers,
churches, houses on high open hills, and just high trees. Today, this list is
much longer and includes power transmission lines, transmitting and
receiving antennas, and the like. The experience in maintaining such
structures indicates that the frequency of strokes increases with the
object’s height. This observation was used as a basis for the most common
lightning protection techniques. A grounded rod higher than the object to
be protected - a lightning rod - put up in the vicinity of the object is
supposed to attract most strokes, thus protecting the object. The underlying
principle of this approach has not changed since the first lightning rod was
constructed two and a half centuries ago. What has changed is the require-
ment for the protection reliability, which have become extremely stringent.
For this reason, the specialists have to deal with exceptions rather than the
rules, focusing on the rare cases of lightning breakthroughs to the object
being protected, because they lead to emergencies and sometimes to
catastrophes.
The study of lightning attraction mechanisms is extremely time-
consuming and expensive. Even a simple measurement of the number of
lightning strokes at objects of various heights is very hard to arrange.
Most apartment houses and industrial premises in Europe are less than
222
50m high. On the average, a lightning strokes a 50m building once in five
years. Every kilometre of a power transmission line 30m high attracts
approximately one lightning discharge per year. Long-term observations of
a large number of buildings and multi-kilometre transmission lines are
necessary to accumulate a representative statistics. The difficulties increase
many-fold when one needs to extract information on the protection reliabil-
ity from the observational statistics. To illustrate, 10-20 years of continuous
observations of a 50 m building would be required to obtain information on
the lightning discharge frequency, and at least 1000 years would be necessary
to check whether its lightning rod can really provide a ‘99% protection’
promised by the rod producers.
In a situation like this, one has to resort to theoretical evaluations,
and this is one reason why lightning attraction theory has been the focal
point of research for many lightning specialists. Here, as in many other
lightning problems, there is an acute lack of factual data. The available
evidence obtained from laboratory investigations on long sparks does not
always provide an unambiguous interpretation, and this makes one treat
with caution many, even generally accepted, concepts. We shall focus on
the most advanced approaches, discussing, where necessary, alternative
hypotheses.
The distance Re, is known as the equivalent attraction radius for an object of
height h. It indicates the surface area, from which lightning discharges that
have descended to the altitude Ho are attracted by the object. For a compact
object of small cross section, this is a circle of area SeqM T R ; ~for; an
extended object of length L >> h and width b << h (e.g., a power transmission
line), this is a stripe of area Seq x 2ReqL.The average number of strokes per
storm season is evaluated from Seqas
NI = nlSeq (5.4
where nl is the year density of lightning discharges into the earth at the
object’s site. Global and regional maps of storm intensity are made from
meteorological survey data [l, 21. The n1 data are usually given per 1 km2
per year. Quite often, the maps indicate the number of storm days or
hours, together with empirical formulae to relate this parameter to nl.
The equidistance principle, simple and clear as it may seem, is of little
use, because one can employ formulas (5.1) and (5.2) to advantage only if
one knows the altitude Ho (the attraction altitude), at which a descending
lightning leader begins to show its preference and selects the point to
strike. The condition of the earth’s surface and the objects located on it
cannot influence the lightning behaviour high up in the clouds. A lightning
develops by changing its path randomly. As it approaches the earth, the
field perturbation by charges induced by terrestrial objects become
increasingly comparable with random field fluctuations. Eventually, the
perturbation begins to play the dominant role, determining the channel
path more or less rigorously. The average altitude Ho at which this happens
is known as the attraction altitude.
It is unlikely that the altitude Ho should be determined only by the
terrestrial object’s height h. It must also depend on the leader field varying
statistically with the lightning due to the variations in the storm cloud
charge, the starting point of the descending leader, its path, number of
branches, etc. This diversity of lightning conditions is uncontrollable. The
only parameter that can, to some extent, depend on observations is the
attraction altitude averaged over all descending discharges. It deserves
attention because the averaging will require only the statistics of descending
lightnings which have struck objects of various heights. These statistics
cannot be said to be reliable but it provides some factual information
important for lightning protection practice.
Before we use the stroke statistics in a theoretical treatment, we think it
worthwhile defining the range of object heights. Unfortunately, one has to
discard the stroke data concerning high constructions. Ascending discharges
become dominant at heights h > 150m. Data on such strokes cannot be
included in the statistics without reservations, even though they were
obtained from well-arranged observations, in which every discharge was
identified unambiguously. The point is that ascending lightnings partly
discharge the clouds, reducing the number of descending discharges. This
interference into the storm cloud activity is so appreciable that a further
increase of h above 200 m does not practically change the stroke frequency
of an object by descending lightnings. Of little use are the data on low
structures (10-15m). The number of strokes in this case is greatly affected
by the nearest neighbours and the local topography. Account should be
taken of the statistics for low buildings, but such observations are scarce.
The overall data have a too large spread.
The authors of [3] selected the most reliable data and, by averaging
many observations, derived the relationship between the number of descend-
ing strokes and the terrestrial object height. Figure 5.2 shows individual
representative values to demonstrate the data spread. All of the results are
normalized to the intensity of the storm cloud activity, which is 25 storm
days per year. In the range h 9 150 m considered, we can admit, with some
reservations, the existence of a quadratic height dependence of the number
of lightning strokes for concentrated objects and a linear dependence for
extended ones. Both dependencies mean Re,/h const.
Figure 5.2 shows that the expression Re,= 3h, sometimes used for
rough estimations of the expected number of strokes, agrees fairly well
with the averaged values of Re, derived from observations. The substitution
of Re, = 3h into (5.1) yields for the average attraction altitude for descending
leaders
Ho = 5h. (5.3)
This does not seem to be a large height. A lightning is insensitive to the
earth’s surface along most of its path and it is only its last 50-500m which
are predetermined. Below, we shall discuss the mechanism of the more or
less rigid determination of the leader behaviour by a particular site on the
earth (section 5.6).
0.
Figure 5.2. The average number of strokes per year for compact (top) and extended
(bottom) objects of height h. The dashed curves bound spread zones in observation
data. Solid curves are plotted using the equivalent attraction radius.
Figure 5.4. The dependence of striking distance on lightning currents. The lower
curve is plotted using [4] data, the upper using [5] data. The spread region is hatched.
where a2is the integral distribution defining the probability of the gap break-
down at a voltage less than U . If the distributions cpl and p2 are described by
the normal law with the standard deviations u1 and u2 and by the average
values of Uavl and Uav2, the breakdown voltage difference AU = U1 - U,
also obeys this law, with AU,, = Uavl - Uav2 and a = (a: +
This
allows us to rewrite (5.6) using the tabulated probability integral:
P1 = A2 [I - g / : e x p ( - x 2 / 2 ) d x 1 , A= Uavl - u a v 2
(0: + a;,li2.
The expressions of (5.7) are valid for Uavl 2 Uav2.Otherwise, one should find
(5.7)
the breakdown probability P2 for the second gap, writing for the first one
P1 = 1 - P2.
The formal relations of the probability theory (5.6) and (5.7) are valid if
the discharge processes in the gaps do not affect one another and if every
individual breakdown can be considered as an independent event. Multi-
electrode systems of this kind can be termed uncoupled. A classical example
of an uncoupled multi-electrode system is an insulator string of a power
transmission line. The distance between the adjacent towers is so large that
there is no electrical or electromagnetic effect of discharges occurring in
one string on those of its neighbours. The earth’s surface and an object
located on it can also be regarded as an uncoupled system, with a descending
lightning leader acting as a common high voltage electrode. Such systems
have been studied in laboratory conditions [9],in which the distribution of
breakdown voltage was used as the indicator of an uncoupled nature of
the system. If the individual gaps comprising a system are tested individually
and have the integral distributions Q1( U ) and a2(U ) with the probability
densities cpl(U) and cp2(U),the system will have the following distribution
of the breakdown voltages:
(5.8)
Figure 5.5. The breakdown voltage probability for the uncoupled multielectrode
system involving the high-voltage and two grounded electrodes. x : measured
QsYs(U),0 : measured @ ( U )for the single gap. The dashed curve is evaluated for
the system using @ ( U ) .
object, d o ( x , y ) :
(5.10)
The expected total number of lightning strokes at the object, NI, is found by
integrating Pa over the attraction plane. If the earth's surface is flat and the
lightning discharge density nl is constant, then we have for a compact object
of height h and for an extended object of average height h and length L ,
respectively:
The relations obtained from the equidistance principle are identica, to (5.11)
at cr, = 0.
In virtue of the approximate symmetry of the function Pa(r)relative to
the point with Pa = 0.5 ( r / h E 3; see figure 5.6), the calculations of NI slightly
depend on the standard attraction deviation 0,. When cra varies from zero
to 10% (there are practically no greater deviations in pure air), the value
of NI increases only by 15% for a compact object and by less than 5% for
an extended one. It would be unreasonable to discard the simple and clear
equidistance principle for the sake of this small correction, but for the greatly
inclined (almost horizontal) paths of lightnings attracted by objects.
rh
Figure 5.6. Evaluated attraction probability for the attraction altitude Ho = 5h ( h is
the object height, Y is the object-lightning stroke distance).
lightning
rod f
hr
I / , , a ,,,//,/,,/;
h0
I
Ir I-
Figure 5.7. Why a lightning rod is less effective when an inclined lightning approaches
from a side of the protected object.
It is clear from the foregoing that the larger the distance between the
lightning and the object (compared with that from the lightning to the rod)
the greater the protective effectiveness of a lightning rod. For a lightning
travelling in the attraction plane strictly above the lightning rod, the
difference between the two paths (figure 5.7) is largest:
Ad = [(Ho- ho)’ + - Ho + h,
A d z h, - ho at a << [2(h,- ho)(Ho- h0)]1/2.
With increasing side shift of the lightning in the attraction plane r , the value
of A d decreases, and this decrease is especially noticeable when the lightning
approaches from the side the object being protected, as is shown in figure 5.7:
Ad = [(Ho- h0)’ + ( r - a)2]1/2- [(Ho- A,)’ + ~’1”’.
In the limit r + x,the distance to the object is smaller than to the lightning
rod ( A d x -a), and it is quite ineffective for lateral strongly inclined lightnings
coming from the object side. In the equidistance approach, there should be no
events like ths, since lightnings are not to strike an object at a distance r > Req.
In reality, the proportion of lateral strokes is found to be fairly large. The fact
that the probability method considers this circumstance correctly (figure 5.6) is
very important for the evaluation of the lightning rod effectiveness.
rods and objects to be protected on the grounded floor [lo, 111. At that
time, experimental researchers expected to derive information necessary for
a numerical evaluation of lightning rod effectiveness. The naive optimism
has long vanished. The measurements showed that the attraction process
did not obey similarity laws. Essentially different results were obtained
from gaps of different lengths and different time characteristics of the voltage
pulses applied [12-141. But the interest in laboratory investigations of
lightning has survived, and they are currently performed in an attempt to
understand the attraction mechanism of long leaders.
The primary question is when the attraction begins. Clearly, the
condition of the earth’s surface does not affect the leader propagation
while its tip is far from the earth. Here, the spark paths become distributed
randomly. If one projects a multiplicity of paths on a sheet of paper and
finds the mean deviation Ax from the normal passing through a high-voltage
electrode with the account of the sign (e.g., plus on the right and minus on the
left), one obtains Ax = 0 for altitudes z > Ho. The mean path in a gap
perfectly symmetrical relative to the normal proves strictly vertical down
to the altitude Ho. The attraction onset is indicated by the mean path
deviation towards the electrode simulating a terrestrial object (figure 5.8).
Data treatment for determining the attraction altitude was made in [14] for
spark discharges of up to 12 m in length. The path statistics involved different
time characteristics of the voltage pulse. In the case of a steep pulse front
( 6 p ) , a leader was attracted from the moment of its origin; for a smooth
front (250p), it had enough time to cover an appreciable gap length
before the deviation towards a grounded electrode became noticeable
(figure 5.9).
I I
I- r-I
Figure 5.8. Determination of the attraction altitude Ho by the bend point of a mean
path deviation onset.
1.0
0.8
0.6
0.4
0.2
Figure 5.10. The deviation Ar at various gap length d for tf = 250ps under the
conditions of figure 5.9.
(p, a, [ ( H l - h ) In 2H1 - Ho - h - ( H l
=-
4T&o Ho - h
+ h ) In 2H1Ho- Ho
+ h-t '1. (5.13)
The field average in the rod height is E,, = p q / h + Eo (with the account of
the cloud field Eo). By equating Eay to the threshold field necessary for the
excitation of a viable counterleader (formula (4.1 l)), we find the attraction
altitude Ho from (5.13), assuming the attraction to begin at the moment of
the counterleader start.
We shall not be interested in the quantity Ho linearly dependent on the
poorly known parameter aq to be averaged over all descending lightnings.
Rather, we shall focus on the tendency in the variation of the Ho/h ratio
with varying h in the range 10-150m. Buildings lower than 10m are rarely
affected by lightning, while the picture for high structures is greatly distorted
by ascending lightnings, as pointed out above. Suppose that the attraction
altitude for an object of average height, say, h = 50 m, is found from (4.11)
and (5.13) to be really close to the experimental value Ho = 5h. This yields
the estimate for a, (which is aq/4neo E 1.5 kV/m at Eo = lOOV/cm and
There is no doubt that lightning attraction is due to the electric field which is
related to the object. It is difficult to imagine another remote way to affect a
leader. As for the field source, the evaluations made in section 5.5 show that
the field created by the charge induced in the object itself proves very weak.
When the distance between the descending leader tip and the object is
sufficient for an attraction effect to reveal itself, the object charge field at
the leader tip is by a factor of 102-103 lower than the cloud charge field.
There are no reasons why such a slight perturbation should make the
leader change its path, which is subject to various random bendings even
without the influence of any terrestrial objects. No doubt, a counterleader
excited by the object serves as a mediator between the object and the
descending lightning. It looks as if it elongates the object, thereby increasing
the charge acting on the descending leader. The counterleader travelling
towards its tip attracts it to itself, and this eventually results in the lightning
stroke at the object. The mutual attraction of the two leaders becomes
especially pronounced when the fields they excite at the tip are comparable
with or, better, exceed the differently directed cloud field. It is only then
that the descending leader changes its path to go to the object, and the
counterleader is attracted by the descending leader rather than by the
cloud charge centre, as is usually the case. It is the excess of the perturbation
field over the cloud charge field which imparts a quasi-threshold character to
the attraction process.
This unquestionable and fairly trivial reasoning is certainly useful for
lightning protection practice. Physically, however, it remains quite meaning-
less until the mechanism of the external field effect on the leader is known.
This is equally true of the cloud field which is also involved in the attraction
of the leader, generally directing it to the earth. It is not clear at first sight
what exactly is affected by the external field, which may be very weak. The
fact is that the leader moves along the field even at Eo M lOOV/cm. Fields
of this scale cannot affect directly the leader development - we have
emphasized this several times above. The leader propagation, which occurs
via turning the air into the streamer and leader channel plasmas, requires
much stronger fields. These are present in the leader tip, in the tips of numer-
ous streamers, as well as in the streamer zone where the strength (the lowest
of the three) exceeds 10 kV,km in a negative leader and 5 kV/cm in a positive
one. High driving fields are created by the charges of the tips, streamer zones
and, partly, by the nearest portions of the channel and leader cover. They
dVL
dt :i (
=&- dUt f -duo
--
dt
- d r ) =*y-
dr dt
(: dUt
dt
E0VL) (5.14)
where plus refers to a negative leader and minus to a positive one. The first
term in the sum of (5.14) does not depend on the direction of the external
field. One of the reasons for the variation of U, with time was discussed in
section 4.3.2. Another reason is the increasing voltage drop across the
channel with its elongation. Normally, a variation in U, has a retarding
effect on descending leaders of both signs.
The second term in (5.14) leads to acceleration if the negative leader
moves in the direction opposite to the field vector, with the positive leader
moving along the field. The accelerating effect of the external field increases
as the field becomes higher and the angle between the field and velocity
vectors becomes smaller. Both terms have been estimated to have the same
order of magnitude (10’ m/s2); the second term may sometimes be even
larger. For this reason, the attractive action of the external field proves
essential.
We can now make clear the attraction mechanism. The actual mechan-
ism, by which a leader chooses its propagation direction, has a statistical
nature. This is indicated by numerous random path bendings and branching.
Clearly, there is a high probability that the leader moves towards a site where
it can acquire the greatest acceleration or the least retardation. It will be able
to develop a maximum velocity in this direction, bypassing other competitors
on its way. Large-scale leader photographs taken with a very short exposition
nearly always show several leader tips on short, variously oriented branches
Figure 5.12. A still photograph of the leader channel front with exposure of 0.3 ps.
(figure 5.12). Among these, only one tip has a real chance of survival - for a
positive leader, it is the one which belongs to the branch oriented along the
external field; for a negative leader, the respective branch must be oriented
against the field vector. The other tips usually die.
The mutual attraction of the descending leader and the counterleader,
mediated by the electric fields created by their charges, is a self-accelerating
process. This is due to a positive feedback arising between them. An
enhanced field of one leader accelerates the other leader towards the first
one. Because the distance between the leaders becomes shorter, the field of
each leader rises at the site of the other leader tip, and the mutual acceleration
proceeds at an increasing rate. This goes on until the streamer zones of the
leaders come in contact and their channels unite. As a result, the common
channel appears to be tied up to the object, from which the counterleader
started.
lightning usually bypasses them to strike one of the rods. This can be predicted
from the equidistance principle. But when designing lightning protection
devices, one usually focuses on exceptions rather than the rules. So the question
arises of how large is the probability that the leader will miss the rod and strike
the object, having taken a longer path. It seems justifiable to apply the concepts
of a multi-electrode system to this problem. The lightning which has become
oriented towards a group of grounded ‘electrodes’ has to choose among
them. Let us make an estimation from formulas (5.10) and (5.1l), substituting
the distance from the leader tip to the earth, de,by the distance to the rod top, d,
(do is, as before, the distance to the object’s top). Lightning protection
experience shows that there is no need to make a lightning rod much higher
than typical terrestrial constructions (ha < 50m). Arranging them close to
each other, one can provide a reliable protection of the 0.99 level (of 100 light-
nings, 99 are attracted by a protection rod) if the rod height h, is only 15-20%
larger than ho. For an ‘average’ lightning, displaced at a distance equal to the
attraction radius Re, x 3ho relative to the grounded system, we have
Ad = do - d, RZ (0.12-0.15)hr at Ho = 5h, (figure 5.13(a)). The substitution
of these values into (5.10) with oa RZ 10% gives A , x 0.2. After taking the
integral of (5.7,) one gets the probability of the lightning stroke at the object
Po x 0.4 instead of the experimental value 0.01.
The complete failure of the theory was predictable. A system with a close
arrangement of grounded electrodes cannot be considered to be discon-
nected, Its counterleaders affect one another. The first leader that has started
from one of the electrodes decreases the electric field behind it, via its cover
space charge, preventing the upward development of counterleaders from the
other electrodes. Appearing with a delay, if they do, these counterleaders
cannot retard their faster competitor, because the field is enhanced in the
direction of the first leader propagation (figure 5.13(b)). This makes all of
Figure 5.14. The oscillogram shows how the counterleader started from the 'active'
grounded electrode of 1.1 m height screens an electric field on similar 'passive'
electrode located at a distance of 10cm. The gap length is 3 m. E the field at the
passive electrode tip. Q - the counterleader charge. I . the gap voltage.
x
.&
3
c,
’g
0.8
0.6-
s
E!
o.21.r/
a 0.4-
0.0 , , , . , ,
0.90 0.95 1.00 1.05 1.10
U”%
Figure 5.15. The breakdown voltage distributionfor the system of figure 5.5, but with a
small distance 10 cm between the grounded electrodes,which makes the system coupled.
A, =
do - dr (5.15)
+
a,(di &)‘I2
defines, as in (5.7), the probability of choosing the stroke point on grounded
electrodes:
(5.16)
? A t A , >> 1 and, hence, P, << 1, one can use the approximate expression P, Y
(27r-’/*A;’ exp(-Af/2), which is valid and more convenient for estimations.
object is ho and the distance between their top projections on to the earth’s
surface is A r , we have
A, =
+
[(Ho- h ~ )( r -~ A r ) 2 ] 1 ’ 2- [ ( H ,- +r2]1/2 (5.17)
+
o,[(Ho- h0)2 ( r - Ar)’ + ( H o- h,)’ + r2I1l2 ’
1
described as
x
Nb = 27rnl
sol Pa(r)Pc(r)rdr, Nb = 27rnl
0
dy.
Pa(y)PC(y) (5.18)
5.8 Why are several lightning rods more effective than one?
The answer to this question can be found geometrically. Let us consider two
lightnings which travel in the same vertical plane going through an object
and its lightning rod in opposite directions. Suppose both leader tips are at
an attraction altitude Ho at the same distance from the rod. They have, there-
fore, an equal chance to be attracted by the object-rod system. The only
difference is that one leader will approach it on the lightning rod side
(version 1) and the other on the side of the object to be protected (version 2).
Assume, for definiteness, that the displacement of the lightnings relative to the
rod axis is equal to the attraction radius Re, = 3h, (i.e., an average displace-
ment), Ho = 5h,, and the horizontal distance between the rod and the object
is AY= h, - ho << h,. From (5.17), the upper limit of the probability integral
for version 2 is nearly seven times less than for version 1:
7Ar Ar
4 1 = u,25fih, > 4 2 = u,25&hr ’
cost. No rod palisades are known from the protection practice; nevertheless,
it is tempting to surround the object of interest with a protecting wire,
especially if it is not very high but occupies a large area.
As an illustration, let us consider a case simple for the calculations. This
will allow us to get numerical results and demonstrate the calculation
procedures. Suppose a circle of radius Ro = lOOm is densely filled by
constructions of height ho = 10m. All of them must be protected with a
0.99% reliability, i.e., the probability of a lightning stroke should not exceed
@bmu = Let us now place a circular grounded wire at a distance of
10m from the external perimeter of the premises. This distance is necessary
for technical considerations. For example, we must prevent a sparkover
between the grounded wire and the communications systems and other struc-
tures, whether it occurs across the earth’s surface or through the air due to high
current pulses of the lightning discharge. Therefore, the circular grounded wire
will have a radius R, = 110 m. Let us find the wire height h,, whch will provide
the necessary value of Bb,,,. For the radial symmetry, the probability of the
lightning breakthrough is found from formulae (5.11) and (5.18) as
(5.20)
The probabilities of attraction P,(r) and point choice P,(r) for a lightning,
whose tip (in the horizontal plane at the attraction altitude Ho) is at the
instantaneous distance r from the area being protected, are defined by similar
expressions (5.7) and (5.16). These differ only in the values of the upper limit
of the probability integral. For the attraction probability, the limit A , ,
according to (5.10), is described by the difference between the minimal
distances from the leader tip at the attraction altitude Ho to the system of
grounded electrodes and to the earth, Ad, = d, - de. In the case being
considered, A , is defined by the smaller of the values (at Y < Ro):
Ad,, = [(R,- r ) 2 + (Ho - hr)2]1’2 - Ho, Ad,, = ho.
At r > Ro,we have Ad, = Ad,, . In the calculation of the choice probability,
the upper integral limit is given by formula (5.15). When calculating the
difference between the minimal distances to the object and the protector,
A d = do - d,, one has to keep in mind that we have domi,= H - ho at
+
r < R and do,,, = [ ( r - R0)2 ( H o- hO)2]1’2at r > Ro.
The calculation procedure reduces to finding, for every value of r , the
upper limits A , and Act in the integrals of (5.7) and (5.16) to calculate
(extract from tables) these integrals, which give P,(Y)and P , ( r ) , and to
calculate the integrals of (5.20). Practically, it is sufficient to make the
f The value of the choice standard oc necessary for the calculation of A , is found from formula
(5.19). taking into account the distance D between the protector top and the point on the object’s
surface nearest to the lightning with the instantaneous coordinate r; ua 0.1.
I .1
Figure 5.17. The object of 10 m height and of 100 m radius is protected by a bounding
circular wire. In the graph is presented the evaluated wire height h necessary to
decrease the probability of a lightning breakthrough to the object up to the value
of CJshown on the abscissa axis.
calculations with the step Ar M (0.1 - 0.2)hr and finish them when P,(r)
drops to 10-6-10-7 with growing r . If the probability integral is given reason-
ably (by an empirical formula or by borrowing it from a table, e.g., using a
spline), the volume of calculations proves so small that they can be made with
a programmed calculator. With a modern computer, the time necessary for
numerical computations is only that for the data input.
The calculations made for the above example are shown in figure 5.17.
The probability of a lightning breakthrough to the object decreases to the
given value of lo-* when the protective wire is suspended at a reasonable
height h, 34m. Note, for comparison, that a single lightning rod placed
at the centre of a similar area provides the same protection reliability only
if its height is h, > 150m. Even if one builds such a rod, the result may
prove disappointing. Quite often, it is impossible to provide a safe delivery
to the earth of a high lightning current impulse, when conductors with
current pass close to structures being protected. Electromagnetic induction,
sparking capable of setting a fire, etc. may also be dangerous.
Figure 5.18. Positive and negative protection angles. A : grounding wire; B power
wire; C: insulator string.
This determines the distance between the grounded wire and the power wire
at fixed a, which defines, through the choice standard a,, the degree of the
system connectivity. Of lines with an identical protection angle, the best pro-
tected line is the one with the largest value of Ah and the lowest value of h,.
Empirical formulas, which relate the lightning breakthrough probability
to wires with a and h,, have found wide practical application. Their accuracy,
however, is not very high because they do not include Ah. For example, there
are expressions identical in composition [21,22]:
ah;'=
lg@b =--- 4, lg@b =-- 3.95, h, [m], a [degree]. (5.21)
90 75
They give the probability of a lightning stroke with a 300% error related to a
value supported by practical observations. These formulae should be treated
with caution when the line supports are higher than 50m at small positive
and, especially, at negative protection angles. This is because most main-
tenance data refer to lines of up to 40m high with positive protection
angles of 20-30". Besides, very few of the data used for deriving empirical
formulas represent direct measurements. Usually, the data are derived
from registrations of storm cut-offs minus the calculated return sparkovers
(section 1.6.1). The latter calculations often give a large error. Still, expres-
sions (5.21) demonstrate that negative protection angles are quite attractive.
The action of protection wires placed farther from the tower axis than line
wires (cy < 0) is similar to that of a closed grounded wire surrounding a
region being protected (section 5.8). This type of protector could provide
an exceptionally low probability of a lightning stoke at line wires, but the
implementation of negative angle protection requires larger towers and,
hence, a higher cost. This approach is, for this reason, unpopular.
their values are lower. For objects of regular height (-30m), we have
ua lo-' and uc x This enables us to focus on the choice only. The
effects of the high potential action of the object, U&, will noticeable at
U& u,U,, where U , is the tip potential at the moment the leader has
descended to the attraction altitude. An 'average' lightning has
U , = 50MV. Therefore, in order to get an effect on the process of choice,
one must apply Uob M 500 kV to the object (or to the protecting wire). In
order to affect the attraction process, the applied voltage must be -5 MV.
The latter value is, certainly, not feasible for the present power industry,
but available operation voltage of an ultrahigh voltage (UHV) line is high
enough to affect the lightning preference to a protecting wire or to a line
wire [22,23].
Most UHV lines operate at alternative voltage of frequency f = 50 Hz
(60Hz in the USA). Over the time H o / V Lx s along the flight path
Ho, during which the lightning chooses a point to strike, the wire potential
changes but little, and its values U & ( [ )= Ufmax sinwt (U = 2759 can be
taken to be equally probable. By the initial moment of attraction, u o b ( t ) ,
may have the same or opposite sign relative to the lightning. If the sign is
the same, the development of a counterleader from the wire will be delayed,
so the probability of the lightning striking the wire will be reduced. In the
other situation, the effect will be opposite. To get a total result over a
long-term observation of the line operation (or a short-term observation of
a very long line), one should average the operating voltage effects over an
oscillation period. For this, expression (5.15) for the parameter A , must be
extended to the case in question. Expression (5.15) was based on the differ-
ence in the average fields along the lengths from the leader tip at the attrac-
tion altitude to the protector and to the object. Now, this difference can be
calculated with the potential U& to get, instead of (5.15),
Figure 5.19. Effect of the AC transmission line operation voltage on the lightning
breakthrough probability. The lower and upper curves correspond to probabilities
of and lo-', respectively, without voltage.
only half of the on-voltage time. The unfavourable effect of positive half
periods may be much stronger. In principle, the potential difference U,, > 0
and U, < 0 at a high alternative voltage amplitude may produce such a
strong 'attracting' field that all lightnings going to the power line will strike
its wires. The probability of a strike at the line wire during the positive half
periods may rise by 2-3 orders of magnitude (even as much as unity) against
its two-fold reduction during the negative half periods. As a result, the
stroke probability averaged over a long time for the line wire grows. The
operating voltage effect on power lines reduces the reliability of lightning
protection.
The numerical calculations of this effect are illustrated in figure 5.19.
The probability of lightning breakthrough to AC lines increases by an
order relative to the probability x lop2 for the off-voltage mode at
y = Uobmax/ccUtx 3.75; at ab x lop3, this effect is produced at 1.5 times
lower voltage. For the typical size of modern power line towers with
oc x 0.008, the stroke probability for the power wire at U , x 30MV rises
from lop3 to lop2 at phase voltage amplitude Uobmaxx 625kV. Such are
the line voltages (750 kV) in some countries. Only the next generation of
power lines with 1150 kV can be expected to produce as strong effect on light-
ning at U , x 50MV. An experimental line of this kind has been in use in
Russia for a short time.
Direct current line has a more pronounced effect on lightning. Lightning
separation is possible in D C lines: a positive line wire more strongly attracts
negative lightnings and a negative wire more strongly attracts positive ones.
Since the frequency of positive descending lightnings is an order of magni-
tude smaller, a positive UHV DC line wire will attract a larger fraction of
strokes. This effect may become well pronounced at the wire potential of
f 5 0 0 kV and higher.
The treatment of a hot air flow from an object to be protected is
generally similar to the above analysis. The density and electrical strength
of hot air are lower, and the strength is proportional to the density in the
first approximation [25-261. Formally, this is equivalent to the reduction
of distance do from the lightning to the object in expression (5.15), as if the
object height were increased. As a result, the lightning protector has a
lower efficiency. Consider, as an illustration, a chimney lOOm high with a
10m lightning rod fixed on its top. With the practically zero horizontal
distance between the rod and the object and in the absence of hot smoke
gases, the rod will intercept about 90% of all lightnings attracted by the
chimney (figure 5.16). But if the chimney ejects a hot gas flow with the
temperature of 100°C along the length of 30m, the probability that
the lightning will miss the rod to strike the chimney will rise from 10 to
50%. Actually, the lightning rod becomes ineffective.The question is whether
it is worth constructing this purely decorative device on the chimney top.
Protection of aircraft and spacecraft has always been a complex and demand-
ing problem - poor protection may have serious repercussions. It has been
mentioned that an aircraft can be damaged by an ascending lightning starting
from its surface or by an attracted descending discharge in the atmosphere, as
happens with a terrestrial construction. Naturally, the concept of attraction
refers only to descending lightnings. There are no observational data on the
interaction between aircraft and descending lightnings, and one has to resort
again to laboratory experiments. Figure 5.20 gives a set of static photographs
taken from the screen of an electron optical converter. Of many pictures, we
have selected the most typical ones. The electronic shutter was shut at differ-
ent moments of time, so the result is not exactly a movie film but something
close to it. One can see that a vertical rod insulated from the earth has
attracted one of the leader branches together with its streamer zone,
having first excited a streamer flash and then a counterleader. Its contact
with the descending leader has produced a short luminosity enhancement
of their, now common, channel, like a step of the negative leader with its
miniature return stroke (sections 2.7 and 4.6). As a result, the channel and
the rod have become the extension of a high voltage electrode. The leader
has started off towards the earth from the lower end of the rod which now
seems to be part of the leader channel.
Figure 5.20. Attraction of the spark leader by the isolated metal rod suspended in the
gap middle.
size I is not larger than that for a grounded object of the height h = 1. This
limit for the number of strokes does not follow only from the experimental
fact of a certain decrease in Re, with H . Of greater importance are the
possible variations in the aircraft position relative to the external field
vector, Eo, during the flight. The field enhancement at the ends of its fuselage
of length I is defined by field projection on to the aircraft axis, rather than by
the value of EO.High terrestrial constructions are always aligned with the
field since it is vertical at the earth.
Let us now estimate the possible number of descending lightning strokes
at an aircraft of length I = 70m, using the concept of attraction radius
Re, M 31. We shall have Nd x n17rR&kh, where n1 is an average annual
frequency of lightning strokes at the earth and kh is the ratio of the total
flight hours per year to the total number of hours in a year. For kh = f
and nl x 3 kmP2 per year, we get Nd M 0.1 per year. This is at least an
order of magnitude less than what follows from official statistics. One
should not think that the discrepancy is due to the neglect of intercloud
discharges, whose number is 2-3 times larger than that of lightnings striking
the earth. In order to be attacked by intercloud lightnings, aircraft must
penetrate through the storm front, but this is absolutely forbidden and
may happen only as an accident. Rather, the result was overestimated
because any pilot tries to stay as far away from a storm as possible.
Therefore, descending lightnings are responsible for fewer than 10% of
strokes at aircraft. The other 90% or more are due to ascending lightnings
excited by aircraft and spacecraft themselves (section 4.2). However, the
interest in descending lightnings remains active because of the poor predict-
ability of the stroke points on the aircraft surface. A similar situation but for
high terrestrial constructions was discussed in section 5.7. The probability of
a lightning striking much below the top is rather high. This situation can be
readily simulated in the laboratory for a long positive spark excited by a
voltage pulse with a smooth front, tf M loops and higher. The photograph
in figure 5.21 illustrates a spark stroke almost at the rod centre, together
with the integral distribution of the stroke points along its length. The
wide, if not random, spread of stroke points over the aircraft surface creates
additional problems. The aircraft has many vulnerable areas. In addition to
the cockpit and fuel tanks, these are hundreds of antennas and external
detectors providing a safe flight. It would be desirable to hide them from
descending lightnings but the chances for this are quite limited. One con-
solation is that most lightnings affecting aircraft are of the ascending type
starting mostly from the ends of the fuselage and wings, where the external
electric field is greatly enhanced.
The excitation of ascending lightnings by aircraft was considered in
section 4.2. Formula (4.11) allows estimation of the hazardous field Eo for
an aircraft of length I = 2d. The field Eo decreases with growing d, some
slower than d-315.Note that the parameter 2d is not necessarily the fuselage
Figure 5.21. The stroke probability at various points of an isolated rod for two
voltage front durations. The photograph shows how the spark has struck at the
rod centre.
length; this may be the wing length if it is larger. In general, the experience
indicates a direct relationship between the aircraft size and the frequency
of lightning strokes. There are exceptions, of course. The statistics of flight
accidents shows that aircraft of identical size may differ considerably in the
capacity to excite lightnings. In one design, the engines are mounted on
the wing pylons, and the ejected hot gas jet passes near the metallic fuselage,
where the low fields cannot excite a leader. In another design characteristic of
rockets also, the engine nozzle is placed in the tail, so that the hot jet serves as
the fuselage extension. This is a perfect site for a counterleader to be excited
since the leader development needs a lower field in a low density gas.
In the estimation, we shall assume the jet length to be half the fuselage
length, lj = d , and its average temperature to be twice as high as the ambient
air temperature. Suppose that the jet radius is large enough for the streamer
zone to be entirely within it and that the leader develops in a gas of relative
density S = 0.5. When the gas density becomes lower due to the heating, the
field providing the streamer propagation decreases at a rate 6 [25,26].The
rate of the electric strength decrease in long gaps is approximately the same
for mountainous regions, although the density variation range in these
experiments was narrower, 6 M 0.7 [25]. We shall assume from these data
that a leader developing within a hot jet requires a potential drop 5-' times
smaller than that given by formula (2.49), i.e., AU = 36A3/5(3hd/2)2/5
(here, the leader length L has been replaced by the jet length d ) . The total
length of a conductor consisting of a fuselage of 2d long and a leader of
top and the lightning-rod, since the mutual effects of the components in a
multi-electrode system become weaker. Formally, this weaker effect mani-
fests itself in increasing standard cc.It appears that the lightning control is
easier for objects of low height and area, when conventional protectors are
sufficiently effective. It is much more difficult to deviate a lightning from
an object without mounting a metallic rod on top of it. The application of
destructive technologies to storm clouds and their charge neutralization
are not discussed in this book, because this is a special problem having no
direct relation to lightning processes.
The physics of the effect of a voltage pulse rise on a descending lightning
leader is clearer than that of other effects. The effect can be expected to be
favourable when the potential applied to the lightning rod is of opposite
sign to that of the lightning, or the potential applied to the object is of the
same sign. In the former case, the conditions for a counterleader to start
from the rod are quite favourable. To initiate a preventive start of a counter-
leader from a lightning rod is to deviate the stroke point from the object. But
in order to produce a noticeable effect, the counterleader must have a channel
length comparable with the length difference between the object and the rod,
or between their tops (the latter quantities are comparable). Only then does
the effective rod height really grow and the charge space of the counterleader
considerably limits the field at the object top. Therefore, one deals with
channels of metre lengths, sometimes of tens or even hundreds of metres,
especially if one takes into account the multi-fold increase in the radius of
the area to be protected. This is a fairly complicated task.
A short-term ‘elongation’ of the rod by exciting a plasma channel from its
top is very similar to the counterleader behaviour. A laser spark or a short-
term long plasma jet would be sufficient for ths. Laboratory studies have
shown that a man-made plasma conductor affects a long spark path as a
metallic conductor. The problem is the technological complexity and consider-
able cost of the project rather than the principal possibility of control.
Imagine an ideal pulse generator, whose effectiveness is so high that it
blocks a lightning breakthrough to the object with 100% probability. The
protection reliability will then be determined by the reliability of the genera-
tor itself, primarily by its synchronizing unit. It is a difficult task to design a
reliable synchronizing unit capable of responding to a nearby descending
lightning leader. A leader always chooses a complicated, poorly predictable
path and has many branches. It is necessary either to distinguish a branch
from the main channel or to trigger the control unit repeatedly. The latter
is undesirable not only because this is resource-consuming. A control pulse
can stimulate a branch to become the main channel, which is the first to
reach the grounded electrode, producing a powerful return stroke pulse.
The close vicinity of a strong current may be as hazardous to the object
being protected as a direct stroke. Finally, we should not discard multi-
component lightnings - 50% of subsequent components do not follow the
(5.23)
where z is the altitude of the descending leader tip and VL = -dz/dt is its
velocity.
Let us find the maximum rate of the field rise at which the ultracorona
can still survive. Assume, for simplicity, that a corona (positive, for definite-
ness) arises at a sphere of radius yo, attached to the electrode top. TO
prevent the corona transformation to an ionization wave capable of
initiating a streamer flash and then a leader, the field on the sphere
should not rise in time with AEo. The surface with maximum field should
not detach from the sphere to move into the gap interior. In an ultracorona,
the field on the sphere is stabilized by space charge on the level of E,
depending on radius yo. The sphere concentrates a constant charge
Q, = 4neor&. A short time A t after the corona ignition, the voltage
increases by the value A U = A,At, which is supposed to increase the posi-
tive sphere charge by AQ1 = CAU = 4mOrOAuAt. To avoid this, the sphere
charge AQ1 must be compensated. The compensation occurs owing to the
gas ionization in the thin surface layer. Positive ions transport the charge
AQ for the distance Ar = plE,At (where p, is the ion mobility), so that
+
the negative charge induced in the sphere AQ, = - A Q r o / ( r o Ar) is able
to neutralize AQ1. The charge actually induced in the sphere is transported
into it by electrons produced in the near-surface layer, whose number is
excessively large since lAQll = AQ, < AQ. ‘Excessive’ electrons leave for
the external circuit and then to the ‘opposite’ electrode - the earth. The
+
field on the radius r = yo Ar now becomes equal to E ( r ) = (Q, AQ)/ +
+
[4mO(r0 AY)^] and should not exceed E,. To the small value of about
A r / r o , this requirement is met at A, d 2p,E; M 3.6 k V / p (E, M 30 kV/cm,
p1 x 2 cm2/VSI.
We have analysed the other extrema1 situation when the corona exists so
long that charge Q >> Q, is incorporated into space and the ion cloud radius
becomes r1 >> YO. A well-developed corona can exist at the sphere for a long
time if the voltage U. does not grow in time faster than U, = Aut. The
maximum admissible growth rate A, coincides, in order of magnitude,
with the above estimate but is slightly lower. At a fast voltage growth,
say, U M t” with n > 1, there necessarily comes the moment when the
ion cloud field becomes higher than E,, stimulating the transition to a
streamer flash. For the typical values of h = 50m, rLx 5 x Cjm, and
VL x 3 x lo5m/s, the voltage growth rate reaches the estimated critical
value when the leader descends to the altitude z x 200m, at which the
attraction process begins. A little later, A , N z-* becomes even more critical,
and the ultracorona dies giving way to a counterleader.
To conclude, lightning can be controlled but this task is costly and very
complicated technologically. So it would be unreasonable to discard
conventional protection technologies where they can solve the problem suc-
cessfully. One should not expect miracles in lightning protection. If particular
circumstances make one turn to unconventional measures, one must be ready
to create complex devices, whose protection reliability will be determined by
their operation, rather than by the interaction with a lightning.
This is likely to happen more often than direct strokes at an object. Some-
times, the object attracts a lightning branch which could hit the object if it
had enough time before the return stroke develops from the main channel.
Such a situation is illustrated in figure 5.22. The counterleader, which has
started from the television tower top, has no time to transform to an ascend-
ing lightning or intercept the descending leader, because the latter has struck
a metallic tower below the tower top. As a result, the counterleader remains
uncompleted. The counterleader channel has, however, become several
dozens of metres longer. This is now a mature channel, whose temperature
is at least 5000-6000K. If it had touched a hot gas jet, it would inevitably
ignite the gas. Practically a leader of any length is suitable for ignition of
inflammable exhausts into the atmosphere. To excite and develop a leader
in air under normal conditions, a voltage of 300-400 kV would be sufficient.
Such a potential difference AU = Eoh can be produced in objects of height
h > 30m even in the absence of lightning because this would require a
storm cloud field of Eo M lOOV/cm. If the object is lower, uncompleted
counter-leaders can be excited even by remote lightnings. From formula
(3.7), a descending leader that has started at an altitude of H = 3 km and
has touched the earth creates a field Eo = lOOV/cm at a distance R = 1 km
from the stroke point if it carries the linear charge T~ M 8 x lop4Cjm. This
charge is characteristic of a descending leader with average parameters.
This is one of the long-range mechanisms of lightning, which should be
Figure 5.22. The long incomplete counterleader (2) started from the top of the
Ostankino Tower while the descending lightning struck lower than the top (1).
taken into account when treating possible emergencies for objects containing
large amounts of inflammable fuels.
References
[l] Uman M A 1987 The Lightning Discharge (New York: Academic Press) p 377
[2] Operating Instruction for Lightning Protection of Buildings and Works RD
31.21.122-87 1989 (Moscow: Energoatomizdat) p 56 (in Russian)
[3] Bazelyan E M, Gorin B N and Levitov V I 1978 Physical and Engineering Funda-
mentals of Lightning Protection (Leningrad: Gidrometeoizdat) p 223 (in Russian)
141 Golde R H 1967 J . Franklin Inst. 286 6 451
[5] Linck H and Sargent M 1974 CIGRE, Sec. N 33/09 (Paris) 11
[6] Wagner C F 1963 AZZZ Trans. 83 (Pt 3) 606
[7] Wagner C F 1967 J . Franklin Inst. 283 (Pt 3) 558
[8] Darveniza M. Popolansky F and Whitehead E R 1975 Electra 41 39
[9] Bazelyan E M, Levitov V I and Pulavskya I G 1974 Elektrichestvo 5 44
[lo] Stekolnikov I S 1943 Lightning Physics and Lightning Protection (Moscow,
Leningrad: Izdatelstvo Akademii Nauk SSSR) p 229 (in Russian).
[ l l ] Akopyan A A 1940 Res. All-Union. Electr. Inst (Moscow) 36 94
[12] Bazelyan E M, Sadychova E A and Filippova E B 1968 Elektrichesrvo 1 30
[13] Bazelyan E M and Sadichova E A 1970 Elektrichesrvo 10 63
[14] Aleksandrov G N, Bazelyan E M, Ivanov V L et a1 1973 Elektrichesrvo 3 63
[I51 Bazelyan E M. Burmistrov M V, Volkova 0 V and Levitov V I 1973 Elektri-
chesrvo 7 72
[16] Cann G 1944 Trans. AIEE 63 1157
[17] Gorin B N and Berlina N S 1972 Elektrichesrvo 6 36
[18] Gorin B N, Levitob V I and Shkilev A V 1977 Elektrichesrvo 8 19
[19] Bazelyan E M 1967 Elektrichesrvo 7 64
[20] International Standard Protection Structures against Lightning 1990 IEC 1021
P 48
[21] Burgsdorf V V 1969 Elektrichesrvo 8 31
[22] Kostenko M V, Polovoy I F and Rosenfeld A N 1961 Elektrichesrvo 4 20
[23] Bazelyan E M 1981 Elektrichesrvo 5 24
[24] Larionov V P, Kolechitsky E S and Shulgin V N 1981 Elektrichesrvo 5 19
[25] Bazelyan E N, Valamat-Zade T G and Shkilev A V 1975 Zzvestiya. Akad. Nauk
S S S R , Energetika i transport 6 149
[26] Aleksandrov N L and Bazelyan E M 1996 J. Phys. D: Appl. Phys. 29 2873
[27] Rakov V A, Uman M A and Thottappillil R 1994 J. Franklin Inst. 99 10745
[28] Bazelyan E M and Raizer Yu P 1997 Spark Discharge (Boca Raton: CRC Press)
p 294
[29] Uhlig C A 1956 Proc. High Voltage Symp. Nut. Res. Council of Canada
265
due to an abrupt current change in the construction and the mutual induction
produced by the current wave running through the lightning channel). But
lightning overvoltages may result not only from a direct stroke but from
remote lightning discharges as well. Their effect is associated with electro-
static and electromagnetic inductions. In the former case, an overvoltage
results from the time variation of the electric field strength at the object,
created by the lightning channel charges during the leader and return
stroke stages (sometimes, by the slowly changing charge of the storm
cloud). Another reason for a remote excitation of overvoltage is the varying
magnetic field of the rapidly changing lightning current. Overvoltages
became a very serious hazard at the beginning of the twentieth century
when the first power transmission lines were built, and the engineer still
associates an overvoltage with a powerful effect of tens and hundreds of kilo-
volts. This is true of transmission lines of high and ultrahigh voltages (UHV
lines). However, an overvoltage as small as several hundreds or dozens of
volts may become hazardous to electric circuits with a low operating voltage.
Especially vulnerable in this respect are the circuits of microelectronic
devices.
Historically, the theory of overvoltages has developed with reference to
power transmission lines. Naturally, the mechanisms of ultrahigh voltage
excitation were the first to attract the researchers’ attention. So this theory
is now very detailed [l-41 and the numerical procedures suggested are
capable of solving engineering problems with a desired accuracy. We shall
not describe these approaches here but rather focus on the physical aspects
of the overvoltage problem, because in many practical applications they
are not as self-evident as in a lightning stroke at a power line.
The calculation of overvoltage includes the solution of two equally
important problems. One is to find the electromagnetic field of a lightning
discharge at the site where the object to be protected is located. These calcu-
lations may prove very cumbersome and time-consuming, especially when
one tries to take into consideration such parameters as the real path and
length of a leader channel, the non-uniform charge distribution along the
channel length, and the lightning current spread over the metallic parts of
a particular object and underground service lines. The physical aspects of
this problem, however, are quite clear and the numerical methods are well
known. The other problem is to determine the response of an object and
its electrical circuits to the electromagnetic field of lightning. The physical
aspects of this problem are much more diverse, and the basic mechanisms
of overvoltage excitation are not always obvious. So the latter are the subject
of special interest in this chapter.
An induced overvoltage is normally smaller than an overvoltage pro-
duced by a direct stroke, especially by remote strokes, but it affects the
object more frequently. When one calculates the frequency of emergencies
for a high-voltage circuit with an insulation designed for hundreds of
kilovolts, one usually deals with direct strokes, because induced overvoltages
cannot damage the insulation. Objects with metallic shells which can screen
well the internal electric circuits (including low-voltage ones) are designed in
a similar way. However, unscreened low-voltage circuits suffer equally from
overvoltages due to direct strokes and from induced overvoltages. Since the
latter are more numerous, they should not be discarded when choosing the
protective measures.
the field, Eo(t),at the object's site will not differ from that calculated neglect-
ing the delay. The phase delay which acts for the time A t = r / c does not
affect the overvoltage.
Let us make a direct evaluation of the 'electrostatic' component of
overvoltage during the return stroke, assuming that a rectangular charge
neutralization wave (section 4.4) is moving along a vertical, perfectly
conducting channel towards a cloud. At any point of the channel behind
the wave front z = wrt, the charge changes by the same value r.The electric
field follows the charge variation. Without the account of the delay, its
change AE, at the distance r from the channel is described by an expression
similar to (3.5) (with h = 0, H = z and R = z):
The time constant for real electric circuits, R,C < 0.1 ps, is several orders of
magnitude smaller than the time of the return stroke flight from the earth to
the cloud. Then, according to (6.2), the electric component of the overvoltage
(relative to the earth) for a compact object is defined as
dAE, r R Ch vr2 t
U, R,Ch- -
dt 2 m (vft2 + r2)3'2
where h is an average object height. The short-term action of this overvoltage
load must be endured by all the insulation gaps separating the object from
the adjacent constructions and service lines, whose potentials were not
changed by the lightning or, if they were, to a different extent.
At the moment of time tmax, the pulse Ue(t) reaches its maximum
--In-.
r1
The emf induced in the circuit, U , = -dQ/dt, is
At the maximum rate of the current change, Ai z 10" A/s, characteristic of the
return stroke of subsequent lightning components, the emf'induced in a circuit
I Pr
with the sides h = d = 10 m at the distance rl = 100 m from the conductor with
current is U , = 19 kV. The emf for a smoother current impulse of the first
component of a moderate lightning with Ai = 5 x lo9A/s is U , = 1 kV.
Overvoltages excited electrostatically and electromagnetically are gener-
ally comparable. The former can be coped with using an effective grounding
of the object, but overvoltages due to the electromagnetic mechanism do not
respond to the grounding efficiency. Imagine metallic columns buried deep in
the ground, which support rails for a mobile overhead-track crane mounted
high up at the ceiling of industrial premises. The whole construction has a
perfect grounding owing to the metallic columns which provide a complete
absence of electrostatic overvoltages from close lightning strokes. However,
a pair of columns with a rail and the conducting earth forms a closed circuit
with an area of several hundreds of square metres, in which the time-variable
lightning current excites an emf. The same thing occurs in a circuit formed by
columns, fixed at the opposite sides of the premises, and an overhead crane.
A possible disconnection at any site of the metallic construction cannot be
ignored either. A disconnection may arise due to metal erosion, poor welding
or inadequate contact between the crane wheel and the rail. In that case,
practically all emf of the circuit will appear to be applied to the site of
defect, provoking a spark discharge through the air or a creeping discharge
across the surface to bypass the defective site. A spark-induced emergency is
inevitable if there is an explosive gas mixture in the premise.
The fact that any construction may serve as a circuit capable of inducing
an emf increases the hazard - this may be a metallic ladder on a conductive
floor, a metallic pipe leaning against a wall, etc. Such casual circuits present
an even more serious hazard, because their parts may have only a slight
contact between them, so that the probability of a spark gap is extremely
high. An explosion would, no doubt, destroy the casual circuit, creating a
mystery to the fire brigade in the spirit of Agatha Christie’s stories.
The sequence of procedures for the calculation of overvoltages due to
lightning current is similar to that for lightning charge calculation. One
should first find the magnetic flux through the circuit in question and
calculate the induced emf. The magnetic flux is often replaced by the
vector potential A(t) to simplify the calculations. For current i in a thin
conductor such as a lightning channel, the vector-potential is
where the integral is taken in the conductor length and r is the distance from
the current element id1 to the point, at which A is determined. The emf
induced in the circuit of interest is defined as
(6.10)
Figure 6.2. Lightning current flows along a pipe with a conductor inside.
Suppose lightning current runs along a closed metallic shell of an object, inside
which there is a conductor connected to the shell at one of its ends, say, at the
lightning current input. The potential at this contact will be taken to be zero. If
R I is the linear resistance of a shell of length 1 and L1 is its linear inductance,
the voltage applied to the shell will be Uf = -(L1 di/dt + Rli)l.The lightning
current does not branch into the inner conductor disconnected at the other end
(the capacitance is neglected). The conductor potential changes only due to
the mutual inductance, U, = Mlldi/dt. Since the magnetic flux of the shell
current is entirely attributed to the inner conductor, the linear mutual induc-
tance M I is equal to the linear inductance of the frame, L1. Then the potential
difference between the shell and the inner conductor at the far end of the latter
is described as
U, = U, - Uf = iR1l. (6.13)
The remarkable property of a cylindrical system with an inner wire to
compensate completely the induction emf is well known to impulse measure-
ment technology. This property is the basis for making shunts for measuring
current impulses with very short fronts (to a few fractions of a nanosecond).
The respective theory, useful for the understanding of the overvoltage
mechanism, is discussed in detail in [ 5 ] . We shall turn to it when evaluating
the skin-effect in a shell. Here, it should be noted that the shape of an
overvoltage pulse, U,(t), in the absence of a skin-effect is similar to that of
a current impulse, i(t). This is valid as long as the time of the electromagnetic
wave propagation along the frame is much shorter than the impulse
duration.
No principal changes will occur when the conductor ends are connected
to the shell via resistances Rkl and Rk2.The voltage U, will then appear to be
operative in the inner closed circuit consisting of the shell, conductor, and
resistors. When the resistances of the conductor and the shell are small, the
current i = U e / ( R k l+ Rk2) arising in the circuit will distribute the over-
voltage U, between the turned-on resistors in reverse proportion to their
values. The same will happen when the conductor is connected to the shell
via spurious capacitances. Of course, if a massive aluminium shell has a
cross section of 100 cm2 and the linear resistance is R1 E 3 x lop6n/m, the
(6.15)
Figure 6.3. Streak photographs of a leader creeping along the soil (top) and an air
leader (bottom). 1: channel, 2: tip, 3: streamer zone.
the plasma column with the conducting soil. So, the leader eventually stops.
Let us evaluate the maximum length I of the leader channel. Suppose current
,Z is delivered to the channel through its base at the stroke point. The current
value is determined by the recharging of the lightning leader channel at the
return stroke stage and is independent of the creeping spark length. For
simplicity, we take the delivered current I.w and the longitudinal field E,,
supporting the creeping leader current, to be constant. At high currents
(i > 1 A), the dependence E,(i) is, indeed, not particularly strong. By the
moment the leader has stopped, the tip potential U, and current it are low
relative to U(x) and i(x) at distances .x from the tip, comparable with the
channel length. We then have
di GIEcx2
U ( x )M E,x, _
dx
-- Zl= G,U(.x), 2 .
i(x) = (6.16)
~
With i(Z) = 1
, at the channel base, the maximum channel length is defined,
1
with the account of (6.15), as
1%
( 2zM - [
-
)'/2
N
21,p In ( I / ~ J 'I2
(6.17)
GI E, ..E,
sheath and the other the propagation of waves, excited by this current, inside
the cable.
Let us first follow the fate of lightning current i ( x ,t ) in the cable sheath.
Its variation along the length due to the displacement current associated with
the charging of the sheath linear capacitance C1, to the voltage U,(x,t ) can be
assumed to be negligible, as compared with the large current leakage into the
soil through the linear conduction G1 of the sheath grounding. One can also
neglect the mutual induction emf in the sheath, produced by the core current
i,, because it is small compared to the self-induction emf. Since the total
magnetic flux of the sheath current i involves both the sheath and the core,
the mutual inductance M 1 per unit length of the sheath-core system is
equal to the linear sheath inductance L1. However, the current in the core
is i, << i. Indeed, the current in the core screened from the earth by the
sheath is only due to the charging of the cable capacitance C1, to the voltage
U, acting between the core and the sheath. The value of U, does not exceed
the electrical strength of the cable insulation, U, M 2 kV. Even if the current
wave velocity in the core were close to light velocity, the core current would
be of the order i, M C1,Uec M 10-30A (for a cable of a small cross section,
C1, M 20-50 pF/m), which is much lower than the lightning current
i M 10kA. So, one can ignore the current deviation into the core even
when its insulation is damaged at the lightning current input into the cable
so that the core appears to be connected to the sheath.
Therefore, on the above assumptions, the current flow along the sheath
is defined by the equations
aU, ai ai
(6.18)
ax
-L1-+Rli,
at ax
-- = G1 us
where L 1 is given by formula (4.25) and R 1 is its linear resistance. If the cable
were on the earth's surface, with the lower half of the sheath touching the
earth, formula (6.15) would be valid for G1. When a cable is buried at a
large depth, the current spreads radially from it in all directions uniformly,
so the value of GI is doubled. In intermediate situations, one can use the
empirical formula
2n
G1 = Y Q h K 114 (6.19)
p In ( 12/2hr)'
where h is the cable depth. The boundary condition for (6.18) is expressed by
the equality i(0,t ) = Io, where Z,,is one half of the current delivered by the
lightning to the cable at the input x = 0 (the current flows in both directions
from this point).
The cable sheath possesses a low active resistance and a fairly high
inductance because it is a solitary conductor. The self-induction emf has
a greater effect on the distribution of the rapidly varying lightning current
in the sheath than the active voltage drop. If Rli is neglected in the first
The latter expression for xo corresponds to a surface cable; at a large depth, this
would be xo = p / p o : xo x 160-600m2/ps at p x 102-103Rjm ( p x 500 s2jm
for a common sandy soil). The point with a fixed value of illois shfted with
a decreasing velocity ‘U x xo/x x ( ~ , / t ) ” ~in, agreement with the diffusion
law x (4xor)’/*. The current covers a 1 km cable length for t x 2000-
200 ps, decreasing rapidly at the wave front (figure 6.4). The sheath potential
from (6.18) and (6.20) is
The potential at the current input drops with time, from an ‘infinite’value at
t = 0, which results from the neglect of Cls.tThe equivalent sheath resistance
t If C1, is taken into account, there is a weak precursor which propagates with the velocity of an
electromagnetic signal (L, overtaking the diffusion wave described by (6.20) and (6.21)
(cf. section 4.4.2).
also decreases with time. It is defined by the resistance of the soil around the
elongating cylindrical surface, through which the current leaks.
Now turn to the wave process inside the cable. The type of overvoltage
under consideration is dangerous only to communications lines, whose linear
inductance L1, is small because of the narrow gap between the core and the
sheath. The self-induction term is usually small relative to the voltage drop
on the active core resistance RI,. The current leakage through a high quality
insulation can also be neglected, since it is small relative to the displacement
current charging the linear core capacitance C1, (relative to the sheath). With
these assumptions, the core potential U, relative to infinity and the core
current i, are described by the equations
au, - Rlcic+ L1di ai, = Clca(uc - Us)
-- (6.23)
dX at!
ax at
which account for M 1 = L1. The electrical signal induced in the core by the
lightning stroke has a much higher propagation rate than the process of filling
the sheath with current. Indeed, we have U, = 0 and a i l a t = 0 far ahead of
the filled part of the sheath. Equations (6.23)transform to the diffusion equation
for U, and i, with the coefficient xc = (RlcClc)-lx 2.5 x 106-2 x lo5m2/ps
( R I ,x 0.01-0.1 fl em) exceeding xo by several orders of magnitude. This
means that the charging of the cable capacitance occurs very quickly, and a
quasi-stationary mode is established in the cable, in which U, and i, follow a
relatively slow variation of the sheath current.
By subtracting the first equalities of (6.18) and (6.23) from one another
and keeping in mind R I , R I and i, << i, we obtain the equations for over-
N
U 2 ( 0 .t ) = -
sox i ( x . t)R1 dx KZ -ZoRlxl,
where x1 is the equivalent sheath length with the lightning current at the
= ( 4 ~ ~ t ) ” ~(6.25)
current is still high at the moment t.t For example, at Io = ZM/2 = 10kA and
R1 = 3.5 x lop4sljm (the aluminium sheath is 1 mm thick and 30“ in
diameter), we have U, M ZoRlxl x 2 kV at a distance x1 M 600m from the
current input. This happens at the moment of time t M x:/4x0 x 6 0 p (at
xo x 1000m2/ps, if the lightning current is still high relative to its amplitude.
This is the duration of comparatively short current impulses of negative
lightnings. For anomalously long impulses (- 1000ps) of positive lightnings,
the length of the ‘active’ cable portion where the overvoltage arises can
increase to 1-10 km, with the overvoltage amplitude becoming appreciably
larger. It is clear now why the repair of the damaged insulation at the
lightning input is insufficient and other damaged sites must be found and
removed along several kilometres of the cable length. In regions with
poorly conducting soils (rocks, permafrost), a damaged line may extend to
dozens of kilometres.
So far, we have evaluated the overvoltage for a rectangular current
impulse. To calculate it for a real lightning pulse, we should first find a
more rigorous solution for the current input into the cable with an intact
insulation. This will provide the maximum value of U,. We shall apply the
operator approach to equation (6.18), omitting the term R1i, as before. As
a result, we get the expression
A = (P/xo)1/2= (PPo/2P)1/2
in which the last term corresponds to a cable on the earth’s surface. If unit
current i(0,t ) = Io = 1 flows into the sheath, the integration constant is
A = 1. The operator form of the overvoltage is
(6.28)
which coincides, within the accuracy of the numerical coefficient of the order
of unity, with (6.25) at Io = 1. Expression (6.28) for unit current Z o ( t ) = 1
represents the unit step function of y ( t ) providing the solution for arbitrary
lightning current i( t ) by taking the Duhamel-Carson integral. In particular,
we get the following expression for a bi-exponential current impulse
0.01 . 1
0.01 0.1 1 10 100
Z
I .
5b
" I
drops by half in 230 ps. At a distance of 1000 m from the lightning current
input, the overvoltage pulse is somewhat higher, smoother and longer. For
current IM = 30 kA, its amplitude rises to 1.1 kV in an aluminium sheath
with R1 = 3.5 x lop4O/m and to 7.5 kV for a cable with a lead sheath of
the same cross section. All of the calculated parameters are quite comparable
with those estimated from (6.29).
this fact, the effective resistance of the conductor is higher than in the case of
direct current. The formal use of this fact in (6.13) would result in an increase
in the overvoltage which is proportional to R I . But an opposite effect is
observed in reality. Owing to the skin effect, the overvoltage pulse front
becomes smoother than the current pulse front, reducing the overvoltage
at a finite pulse duration.
The reason for this paradox is that the last equality of (6.13), which is
strictly valid only for an infinitely thin sheath or for direct current, should
not be used in any situation. If the current varies in time and the sheath
has a finite thickness, its voltage can also be represented as a sum of the
resistance U R ( t )= f ' j l ( j is the current density) and the induction
Ui(t) M d@/dt components (@ is the magnetic flux). But with the same sum
+
U , = UR Vi, the value of each component varies with the point r of the
sheath cross section, for which the calculation is being made, since the
proportion between the current density j ( r ) and the magnetic flux @ ( r )
varies when the total current over the cross section changes. Calculations
of overvoltages between the conductor and the sheath, U, = U, - U,, are
generally indifferent to which r the value of U, is being found, because the
potential does not vary with the thickness. For simplicity, however, it is
reasonable to make calculations for the inner sheath surface: this surface
and the internal conductor are the only elements of the system affected by
equal magnetic fluxes, mutually excluding the induction components of
overvoltage on the conductor and the sheath. Consequently, formula
(6.13) can be replaced, without any restrictions, by the expression
U,(?) =j&)cr-ll = Ein(?)l (6.30)
where j,, and Einare the current density and longitudinal electric field, respec-
tively, on the inner surface of the object's sheath.
The current penetration into a thin sheath is described by the equation
for one-dimensional plane diffusion. It has been mentioned that the diffusion
coefficient is expressed by the quantity xs = (pea)-'. For a rectangular
current impulse of infinite duration, the longitudinal field strength on the
inner surface of a sheath of thickness d can be written as
The exponential series at t > y-' converges very rapidly, so one can restrict
oneself to the first term only. Therefore, the field Ei, rises with the time
constant T : = y-l = p0ad2/7r2; its value is 6 ps at cr FZ 5 x lo7 (Cl. m)-'
and d M 1 mm. This permits the neglect of the skin-effect action on over-
voltages in long underground cables, in which the current diffusion along
the sheath and, hence, the time of the overvoltage rise to the maximum
take 1 or 2 orders of magnitude longer than T : . However, the skin-effect
in objects located on the earth's surface and having relatively short sheaths,
0.0 ' .
50 100
I
150
I
200
Time, p
Figure 6.7. Overvoltage pulse deformation in a cable sheath due to skin-effect with
the time constant T, = 10 p. An exponential current impulse is duration of 100 ps
(dashed curve).
in which lightning current propagates over the time t << T i , decreases the
overvoltage with a greater efficiency in the case of a shorter current impulse.
For an exponential current impulse i(t) = ZMexp(-at), we have from for-
mula (6.31) with the first series term only and the Duhamel integral
exp(-at) -
27 exp(-yt)
-
7-a 1 t > y-'.
The results of the calculations presented in figure 6.7 show that the skin effect
(6.32)
the sheath into a set of N parallel conductors of a short length Ark along the
cross section perimeter, such that the current J k per perimeter unit length in
the kth conductor could be regarded as varying only with time (the total
current in the kth conductor is ik = JkArk). In a steady-state mode when
the current becomes direct, all J k values are the same, since they are
determined by equal ohmic voltage drops in all of the conductors. A mere
summing of the magnetic fields of the conductors will indicate that a
magnetic field may be present inside a sheath of an arbitrary cross section
geometry.
When lightning current is introduced into the sheath very quickly, the
magnetic induction emf in the conductors greatly exceeds the ohmic voltage
drop. But in this approximation, all of the conductors will indeed form an
integral ‘perfectly conducting’ sheath, namely, they will be connected in
parallel. This means that all of them will have equal potentials. Hence, the
magnetic Aux coupling @ for each conductor is the same. This provides a
set of equations for finding the currents ik at the initial stage of the process:
N N
Lkik(0) + Mkmim(0) m # k, im(0)= I M (6.33)
m= 1 k= 1
The current in the central conductor is lower than in the end conductors
because of M12> M23.
It is easy to solve a set of equations of the type (6.33) even for a large
number of conductors simulating a sheath. Only the calculation of inter-
conductor distances is somewhat cumbersome, requiring knowledge of the
cross section profile coordinates. We shall leave this problem to the reader
and illustrate, instead, the analytical solution for the current distribution
in a long cylindrical sheath with an elliptical cross section [9]. This solution
is useful for the evaluation of many real profiles and for testing computation
programmes:
J(x) = (6.34)
2 4 a 2 - x2(1 - b2/a2)]’I2‘
Here, a is the large and b the small semiaxis of the ellipse and x is the distance
between the ellipse centre and the calculation point projection on the large
axis. The ratio of the minimum linear current density (on the plane part of
the ellipse) to the maximum one (on its tip) is Jmax/Jmin = a / b . The current
non-uniformity may be great in real structures, such as the aircraft wing,
a / b > 100.
There is no magnetic field in the sheath at the moment of time t = 0. This
is the result of the initial current distribution among the conductors owing to
the magnetic induction emf. With the redistribution of the currents under the
action of ohmic resistance, a magnetic field will gradually arise in a non-
circular sheath. The field becomes the source of overvoltages in the inner
circuits of the object. By integrating numerically the set of equations
where U (t) is also the unknown voltage drop along the length of the sheath
‘made up’ of conductors, one can find the variation in the current distribu-
tion along the sheath perimeter. The initial condition for the integration is
the solution to (6.33). The calculation accuracy increases with the number
N of simulating conductors. But the limiting case of N = 1 is also suitable
for the evaluation of the time constant of a transient process: TI, = L 1 / R 1 ,
where L1 and RI are the linear sheath inductance and resistance. The current
is redistributed slowly, T,, 0.1 s, in the sheaths of large objects with radius
N
current impulse with t, 100 ps, the current distribution along the sheath
N
perimeter differs but little from the initial distribution profile. The results
of a computer simulation support this conclusion. The computation for
a sheath of complex geometry (figure 6.8) with L 1 = 0.57 pH/m and
RI = 1.05 x lop5Rjm (Ttz= 54ms) has shown that the linear current
density at all characteristic points of the sheath takes the steady-state value
for a time about 20 ms. During the first 200 ps typical of lightning current,
the density cannot change appreciably.
Let us consider overvoltages across the insulation between an inner
conductor and the sheath. Suppose the conductor is placed very close to
the inner sheath surface. The contour area between the conductor and the
wall will be very small, and the internal magnetic flux will be unable to
create an appreciable induction emf. The voltage between the conductor
and the sheath will be equal to the integral of the ohmic component of the
O’ fo 20 30 4-0 t,ms
Figure 6.8. Evaluated evolution of a linear current density at indicated points of the
wing-like sheath shown. J , = J ( t + XI).
longitudinal electric field &(x) at the site of the conductor location x. But
now, the evaluation of Einshould not be based on the average current density
in the sheath, using the total current and linear resistance R I .For a sheath of
thickness d , we have
=J(x)~/d. (6.36)
The nearer the current line with the maximum linear density, the higher the
overvoltage across the conductor insulation relative to the object’s shell. One
practical conclusion is quite evident. To reduce manifold the overvoltage in
the electrical line inside an aircraft wing and along its thinnest back end,
where the current density is maximal, it is sufficient to shift the wire closer
to the upper wing plane or, better, to the lower one, where the current density
is minimal due to the wing curvature (figure 6.8). Laboratory measurements
have confirmed this suggestion [lo].
Note the seemingly ambiguous character of the evaluations. The sheath
cross section is practically equipotential, so the inner conductor must be
under the same voltage with respect to any point of the sheath in a particular
cross section. However, the ohmic overvoltage component for a conductor
inside an elliptical cylinder (figure 6.9) with respect to points 1 and 2 of the
large and small semiaxes differ by a factor of Jmax/Jmin = a/b, in agreement
with (6.34). This contradiction is superficial. In the presence of a magnetic
field, there is the magnetic component, in addition to the electrical one,
U = U, + U,. The distance between the conductor and current line 1 is
practically zero, and the magnetic flux induces nothing in such a narrow
circuit, U , = 0. On the contrary, a wide circuit, made up of a conductor
and remote current line 2, is affected by most of the internal magnetic flux.
The emf induced by the flux adds the ohmic voltage to the necessary value U .
The evaluation of the magnetic flux direction will show that the signs of U,
and U , coincide if the circuit, in which U , is induced, is composed by the
current line with a linear density less than the average value; otherwise, U,
and U, have opposite directions. Therefore, the values of U, and U, vary
with the design circuit chosen, but the sum remains the same.
We shall make use of this circumstance to find the time variation of the
magnetic field inside the sheath. It has been pointed out above that lightning
current i ( t ) acts for such a short time that it cannot be redistributed radically
along the sheath perimeter; therefore, we have J ( t ) w i ( t ) at any point. Hence,
we get Ein(t) i ( t ) for a t h n sheath where the slun effect is inessential. Choos-
N
i
-OV
0.2 0.'4 0.'6 ' t, ms
Figure 6.10. Oscillograms of the test current and magnetic field inside the wing-like
sheath.
sheath simulating an aircraft wing [l 11. The response time of the magnetic
field detector did not exceed 0.5ps, so that the H ( t ) pulse front close to
300ps and an order of magnitude higher than the current impulse front
causes no doubt.
and, hence, in the linear current density. A compact paclung of cable assem-
blies at sites of minimum surface curvature is a good and nearly free means
of limiting overvoltages in internal circuits of objects with metallic shells.
Overvoltages rise considerably if the shell is made from a plastic and if
its electric circuits are located in a special outer metallic jacket extending
from the head to the tail. The linear resistance of the jacket may be 1-2
orders higher than that of the totally metallic fuselage. The ohmic component
of overvoltage will increase respectively. To eliminate the magnetic compo-
nent, associated with the penetration of the magnetic field into the jacket,
it is very desirable to make it as a pipe with a circular cross section.
Point number
01
1 2 3 4 1
Figure 6.11. Measured angular distributions of the linear current density along the
circular pipe perimeter at various locations of the reverse current conductor.
Marked points (on the pipe scheme) are presented on the abscissa axis. The curve
A I corresponds to a single reverse wire A for a = 2r, curve A2 is that for a = 4r,
curve B depicts three reverse wires B placed as shown in the scheme. Uniform
distribution C corresponds to the coaxial reverse current cylinder of radius 2r.
Figure 6.12. Multilayer cable (a) and solid sheath model (b).
For this reason, a set of circular sheaths can be regarded as a solid conductor,
and the current penetration along its radius (from the second layer to the first
one) can be considered as a skin effect. Such a system can also be treated as a
set of discrete circular layers.
In the latter case, the current distribution among the layers at the initial
moment of time t = 0 can be found from the condition of magnetic flux
coupling equality (6.33). Equations (6.35) are valid at t > 0 and have the
following solution for two layers at constant current IM = il i2 = const:+
R2rM IM
[l - exp(-Xt)],
+ R2 [Rl + R2 exp(-Xt)l
i l ( t ) = ____ i2(t) =
R1+ R2 RI
~
(6.37)
where X = (R, + R2)/(L1 - L 2 ) . Equations (6.35) allowed for the mutual
inductance of the layers, M12= L2, as in the treatment of the screen-wire
system in section 6.2. In accordance with the skin-effect law, the lightning
current first loads the outer sheath and then gradually penetrates into the
inner sheath. The current is distributed uniformly between the individual
screens in each circular layer, iSl= il/nl and is2 = i2/n2.The overvoltage
across the insulation between a wire and its own screen (providing that the
skin-effect in an individual screen is neglected) is similar to the current in the
layer, U , ( t ) = Rlil(t) and U 2 ( t )= R2i2(t),but not to the lightning current.
If a double wire circuit uses the cores of one layer, there is no over-
voltage in the instruments connected to it, because the potentials of the
layer cores are identical. If the instruments are connected to the cores of
different layers, the voltage between them is
U12 = U2 - U1 = I M R eXp(-Xt).
~ (6.38)
At I M = 1, expression (6.38) is a unit step function for the set of equations
providing the solution for the lightning current impulse of an arbitrary
shape. In particular, at i(t) = IM[exp(-at) - exp(-Pt)], we have
U12= IMR2[Bexp(-Pt) - A exp(-at) - ( B - A ) exp(-At)]
(6.39)
A = ./(A - a), B = p/(X - p).
Owing to the relatively small value of L1 - L2 x (p0/27r) In (r2/r1) at close
layer radii r2 and r l , the layer current ratio is redistributed rapidly, for
T = A-' FZ l o p . This is the reason for a fast damping of the overvoltage
pulse U12(figure 6.13), which may be remarkably shorter than the current
impulse. It follows from (6.39) that the pulse U12 reverses the sign; its
opposite tail is damped approximately at the rate of lightning current reduc-
tion. The overvoltage amplitude in a double wire cable is close to that in a
wire-shell system, exactly as in a sheath with a sharply non-uniform current
distribution. If the screens are thin and have a high resistance, the hazard of
damaging the connected measuring instruments is fairly great.
-0.2 J
Figure 6.13. Overvoltage pulse on a two-layer cable for the bi-exponential current
impulse with cy = 0.007 ps, ,3 = 0.6 ps and the redistribution time constant T = 50 ps.
The problem for a multilayer cable can be solved in a similar way. The
overvoltages between the cable cores grow with distance between the respec-
tive layers. Other conditions being equal, the overvoltages drop with the layer
depth in the cable. The use of cores of one cable layer reduces considerably
the overvoltage in a double wire system but does not eliminate it entirely.
There are no perfectly circular cables - the cable is pressed under its own
weight and becomes deformed during its winding on a drum. The result is
that the current distribution along the sheath cross section perimeter
becomes non-uniform, producing additional overvoltages between the
cores of the same layer. To minimize these overvoltages, it is desirable to
connect the equipment to the adjacent cores of the same layer. High precision
equipment should be connected to the cores of deeper layers. Overvoltages
arising in a multilayer cable can be evaluated from the same set of
equations (6.35).
Figure 6.14. Underground pipe as the pathway for a lightning current and the design
circuit for a simple evaluation of the object potential.
At the beginning, while the effect of self-induction emf is still noticeable, the
current largely flows through the equivalent resistance Rel at the front end of
the line. After time T = A-', the current gradually penetrates to the far end
of pipe. Some of it, i82 = i 2 R / ( R+ Re]!,finds its way to the grounding elec-
trode of the object of interest, raising its potential to the value U2 = ig2Rg1
relative to a remote point on the earth. For a longer line, the values of ig2
and U2 decrease for two reasons. An increase in L = L1l and G = G 1 l
raises the time constant T , and by the time the current has reached the far
end of the pipe, the initial lightning current is considerably damped. Besides,
a smaller portion of the current i2 that has reached the far end enters the
object's grounding electrode because of the greater pipe leakage. The depen-
dence of ig2 and U2 on 1 proves to be rather strong, especially when the
effective duration of the lightning current, t, x a - ' , is comparable with
T = A-'. Suppose we take t, = 100 ps on the 0.5 level ( a = 0.007 ps-'), the
grounding resistances Rgl = Rg2= 10 R, and L1 = 2.5 pH/m. A metallic
pipe with a lOcm diameter and lOOm in length, lying at the surface of the
soil with p = 200 R/m (G1 = 2.1 x (a/m)-', R = 9.7 R), will deliver
the current igz 0.171ZAvto the ground of the object located at its far end.
The object's potential will be raised to U , x 50kV at ZM = 30kA. At
I = 200m, we have ig2x 0.0861Zjw and, at the same lightning current,
U2 z 25 kV. But even this voltage is quite sufficient for a spark to be ignited
between closely located elements of two metallic structures, provided that
one of them is connected to the grounding electrode and the other is not.
Such a spark can induce an explosion or fire in explosible premises.
In low conductivity soils, current can be transported through metallic
pipes for many kilometres. This refers, to a still greater extent, to external
pipes and rails mounted on a trestle which are grounded only locally, through
the supports separated by dozens of metres. Here, evaluations can also be
made with expression (6.42), putting R = 2Ri/n, where RL is an average
resistance of the support grounding and n is the number of supports.
A comparison of the estimates and computations is shown in figure 6.15
for the above example with I = 200m. The estimates for the current
amplitude at the far end of the pipe and for the moment of maximum
current show a satisfactory agreement with the numerical computations.
The computations will be unnecessary if one finds it possible to ignore the
initial portion of the pulse front and can put up with a 20-25% error.
Let us calculate the potential at the far end of the pipe unconnected to
the grounding electrode at either end. This may happen due to careless design
or poor maintenance of communications lines. The soil will be considered to
have a low conductivity, p = lOOOQ/m; L1 = 2.5 pH,". The curves in
figure 6.16 show the variation in the voltage and current amplitude ratio
UmaX/ZAv for impulses of negative lightnings with tp = 100 ps and for those
of 'anomalous' positive lightnings, which are an order of magnitude
longer. The pipe is capable of delivering a potential of dozens of kilovolts
Time, ~s
Figure 6.15. Portion of a lightning current passed to the object through the communi-
cation pipe of 200 m length. Curve 1: numerical computation, 2: simple evaluation.
20 -
f
E
0 15-
+-
J
.
10-
5-
the potential at the current input into the protector). We have discussed, at sev-
eral points in the book, the two physical mechanisms affecting differently the
ability of a metallic conductor to tap off the lightning current to the earth: the
self-inductance and ionization expansion of the surface contacting the soil.
The voltage drop across the inductance prevents current flow into the
conductor. A long conductor has to be treated as a line with distributed
parameters. The input resistance of the line, Ri, = U ( 0 ,t)/i(O:t ) varies in
time, since the current diffuses along the line, and it takes some time for
the whole conductor to be loaded by current more or less uniformly. As
the limiting case, consider an infinite conductor in a soil with resistivity p .
From formulae (6.21) and (6.22), the voltage at the conductor input is
U e ( t )z U ( 0 ,t ) t - 1 / 2 for the current i(0,t ) = const = lo and t > 0. At
N
U ( 0 ,t ) = 210 (g y 2 1 1 ( 3 t ) (6.43)
where h ( @ ) is a function given by the last integral in (6.29) and figure 6.5. Its
maximum h,, at pt, M 0.9 permits the calculation of the maximum voltage
drop across the grounding electrode:
(6.44)
The effective input resistance of an extended horizontal grounding electrode,
corresponding to U,,,, is expressed as
(6.45)
Extending the grounding bus beyond the limit leff,we are still unable to
reduce appreciably the maximum voltage drop across the bus. For this
reason, it is better to introduce current at the centre of a long bus rather
than at its end, such that two current waves would run in opposite directions
along the half-length conductors. Still more effective are three conductors
arranged at an angle of 120”, and so on. When a grounding mat with the
lowest possible value of R,, is desired, it is preferable to load, more or less
uniformly, the whole of the adjacent soil volume. For this aim, a set of
horizontal conductors or a conductor network is combined with vertical
rod electrodes. To avoid the interaction effect of the grounding elements
and to achieve the maximum loading of them by current, the distance
between the elements should be made comparable with their length (or
with the height, for vertical rods). But even in that case, only part of the
grounding mat, within the radius of leEfrom the current input, will operate
effectively at the impulse front.
Thus, the resistance of a grounding electrode for rapidly varying cur-
rents is much higher than for direct current. A grounding mat network
with numerous horizontal buses and vertical rods is able to reduce the effec-
tive resistance to the value of R,,, x 1 0. But when a large number of objects
is being constructed, for example, the towers of a power transmission line,
one has to deal with resistances as high as R,, x 10 R and more.
Laboratory experiments show that the grounding resistance of an
electrode delivering to the earth very high currents is lower than for low
currents. The grounding resistance decreases with the current rise. The
grounding resistance ratio of a high impulsed current and low direct current,
cui = R,/Rpo, is often called the impulse coefficient of a grounding. The
coefficients ai used in the literature are sometimes as small as aix 0.1. To
illustrate, we shall cite the generalized function cui =f(plM) which has
been suggested for a vertical rod of 2.5m in length from the results of
small-scale laboratory experiments [7] (figure 6.17). The grounding resistance
-
is reduced by a factor of four at p = 1000 R m and IM = 30 kA.
In principle, this reduction in resistance might be due to a larger effective
radius of the grounding electrode because of the soil air ionization. In section
6.2.2, we gave formula (6.15) for the linear conductivity of a long rod lying on
the earth with one half of its surface contacting the soil. If the rod is fixed in
the vertical position, the whole of its surface contacts the soil but its leakage
conductivity is lower by a little less than a factor of 2 at the same length (due
to the poorer operation of the upper end of the rod located at the earth’s
surface, because current cannot flow upward into the air). The linear conduc-
tivity GI and the grounding resistance R, of a rod of radius ro, fixed vertically
into the earth for a length 1, are
(6.46)
ai I
1.0’
0.8- 2.5 m
0.6-
0.4.
0.2 -
PI, M V m
0, 1 I
Figure 6.17. Impulse coefficient for the grounding rod of 2.5 m length.
develops in the air adjacent to the surface of a conductive soil and is shown by
laboratory experiments to be devoid of a streamer zone and a charge sheath
-
(section 6.2.2). In t h s respect, it is similar to a dart leader whch develops a
velocity of about lo7 mls at current i 1 kA. Assuming that the development
of a fast leader along the soil surface requires this current in the leader tip, it, let
us estimate the tip radius, at which this appears possible.
If the grounding resistance is R, and the lightning current is I,, the leader
is supported by the voltage U zz R,IM applied to its base. The tip possesses
approximately the same potential, because the leader channel is a good
conductor and not a large part of the voltage drops across it. The resistance
of the leakage current from a hemispherical tip is equal, from formula
(6.14), to R, % p/27rrt. The tip current is it U / R , . Only part of the lightning
N
current enters the channel, Io. This current mostly leaks into the soil through
the lateral channel surface possessing, according to (6.15), a leakage resistance
RI, = (G,L)-' = pln(L/ro)/rL. Keeping in mind U Rl,Zo. we obtain the
N
(6.47)
-
yo zz rt, appears to be quite reasonable. The field at the channel lateral surface
behind the tip, E pZo/*irr,L 80 kV:cm, is high enough for the ionization
expansion of the leader channel to occur there. Radii much larger than
those of a conventional leader in air under similar conditions have been
registered for laboratory leaders creeping along the soil. The photographs
in figure 6.18 illustrate this quite clearly.
To conclude, the spread of high lightning currents reduces the ground-
ing resistance, probably due to the excitation of one or several leaders
creeping along the soil surface, thereby increasing the length of the ground-
ing electrodes. But for a fast leader growth (otherwise, the leakage surface
has no time to become larger for the short lightning rise time), a high
current of about 1 kA must be delivered to the leader tip. This restricts
the process of grounding resistance reduction by the condition under
which the electrodes are arranged in compact groups. A fast reduction is
hardly possible for a modern substation having an extended grounding
network. The reduction is, however, quite possible in the case of a con-
centrated protector consisting of 2-3 horizontal conductors or several
vertical rods.
It is worth saying a few words about the testing of lightning grounding.
The great complexity of a large-scale simulation of lightning current makes
one turn to model tests. in which the surface current density of small
electrodes is preserved while the total current is reduced manifold. The
laboratory studies indicates that the similarity laws are invalid for the
leader process. The questions of how to interpret the small-scale simulation
Figure 6.18. Still photographs of leader creeping along the soil during its develop-
ment ( a ) and at the moment of gap bridging ( h ) .
results and how well they reproduce the real process of lightning current
spread are open to speculation.
vortex electric field EM = -aA/dt for the calculations. Suppose the current
rises linearly with a distance from the current wave front, i ( x ) = b(xf - x),
where xf = v,t and b = const. The current at the point x of the channel
rises linearly with time, Ai = ai/& = bur. Neglecting the delay time, as was
done in (6.9) and (6.1 l), we have
where ro is the average radius of the object affected by lightning. One can see
that the formula for L1 in (6.48) coincides with the one above, provided 1 is
understood as the length of the channel loaded by current by the moment of
time t .
If the finite velocity of an electromagnetic signal is taken into account
and the object is located directly under the lightning channel, which happens
in the case of a direct stroke, the evaluations made in section 6.1.1 give
Once again, we should like to emphasize the small contribution of the delay:
the logarithms in formulae (6.48) and (6.49) differ less than by 3 % at 3 = 0.3,
tf = 5 ps, and yo M 1 m.
Expressions (6.48) and (6.49) define rigorously the vortex electric field
EM at the earth's surface. When the object's height is h << urtf, which is
valid for many practical situations, the variation in EM along the object
can be ignored and the induction emf is U , = EMh.The emf rises linearly
with increasing h. In particular, if we have h = 30m, w, = 0 . 3 ~and the
lightning current rises to the amplitude IM = 100 kA for the time tf = 5 ps,
the maximum value of the induction component of the voltage at yo = 1 m
is U.Mm,,= 780 kV, a value comparable with the electrical component
Uem,, = R,ZM M 1000 kV at R, M 10 R. Of course, the effects of the electrical
component may be more serious because of its longer action. Indeed, in the
first approximation, the pulse Ue(t) is similar in shape to the lightning current
impulse and U,w(t) to its time derivative.
Formula (6.49) can also be used to evaluate the magnetic component
after the current impulse maximum. For this, the real current entering the
lightning channel should be represented as a sum of the two components:
i, = Ait, i2 = -A,(t - tf), and i2 = 0 at t < t f . It is not surprising that U ,
is non-zero behind the impulse front, since the magnetic field continues to
grow, as the lightning channel is filled by current. The total overvoltage
pulse of a direct stroke, U d ( t )= U,(t) + U M ( t ) ,is very different from the
2.0 -
2.01
1.5-
$1.0-
8 .
0.5 -
0 .40 t .
0.0 1 . , q. , . , , 1
I
0 2 4 6 8 10
Time, ks
Figure 6.19. Computed overvoltage of the direct stroke at the transmission line tower
with the grounding resistance R, = 10 s2 for the lightning current with tf = 5 ps and
amplitude of 100kA; U, = 0 . 3 ~ .
lightning current impulse because of an abrupt rise and an equally abrupt fall
of U , with time (figure 6.19).
(6.50)
Expression (6.50) allows for the value v,t >> d at the moment this maximum
occurs. The magnetic component of the overvoltage across the insulation
Figure 6.20. Estimation the voltage between a lightning rod struck and an object.
object in formula (6.6) can be taken to be r = 25m. Assuming that the height
of a typical object is h = 30m, its linear capacitance C1 z 10pF/m,
C = Clh = 300pF, and R , = l o a 2 ,we shall have UeDb, 30kV at the light-
ning current Z,v = 30 kA. Although Uerodand Ueobihave different signs and
iAU,I > IUerodI,the additional value is not essential because it is an order
of magnitude less than R,ZM in the above example. The situation when a
lightning rod is put up at a distance sufficient for the separation of its own
grounding grid and that of the object is quite realistic. This is done for the
protection of especially important constructions to avoid pulse noises or
sparking due to the induction emf, when some of the current finds its way
to the object’s grounding through the soil.
Another extreme case, in which the electrical component of the object
overvoltage is dominant, is a lightning stroke at a metallic grounded tower
of a power transmission line. The direct stroke overvoltage affects an
insulator string, to which a power wire is suspended. Consider first a
simple and frequent variant (in lines with an operation voltage below
1lOkV) when the line has no protecting wire. In that case, we do not have
to solve the difficult problem of lightning current distribution between the
affected tower and the wire, repeatedly grounded by the adjacent towers.
Nor should we bother about the electromagnetic effect of the protecting
wire on the power wire (section 6.4.4). As in the previous situation, the
insulator string is affected by the overvoltage A U equal to the potential
difference of the tower at the point of the string suspension and the power
wire. The calculation of the tower overvoltage is similar to that for a light-
ning rod, just described. A specific feature of this problem is the existence
of the wire. Being suspended horizontally, it does not respond to the mag-
netic field of the current in the lightning vertical channel. The power wire
is well insulated from the tower grounding by the insulator string. Owing
to its far end being grounded, it would be able to maintain zero potential,
but for the current created by the redistribution of the charge induced on
the wire. The induced charge is very high because some of the wire length
is located close to the lightning channel, and the total capacitance of a
long wire is very large. Naturally, the small distance between the wire and
the lightning channel does not mean the existence of a direct contact between
them, so we can speak only of the effect of electrical induction on the wire.
Even though the wire is connected to the earth at zero resistance, the
induced charge cannot respond immediately to the lightning charge variation
and the wire potential cannot remain at zero. The grounding point is located
far away, at the end of the wire, so the charge liberated by the induction
cannot be delivered to it faster than with light velocity c. For the induced
charge q,n to appear at the point x, a current wave must be excited at this
point, which will eventually transport the charge -q,n out of the wire to
the earth. This wave will propagate at light velocity. During its motion,
the potential at the wave front will rise due the voltage drop on the wave
resistance of the line with distributed parameters, i.e., along a long wire.
Elementary current and potential waves arise at any point on the wire,
where the induced charge is changed by the lightning field. Propagating
with light velocity to the left and to the right of the origin, the currents of
elementary waves are summed, raising the voltage between the wire and
the earth. After the waves are damped, this voltage, naturally, drops to
zero, because the wire is grounded. The response of a long line to the external
field Eo,(x, t ) acting along a horizontally suspended wire is described by the
equations
di di
aue = R l i + L l - - E o , ( x , t ) ,
-- --= C aue
1 T (6.51)
dx at dx
where the potential U e ( x ,t ) is due exclusively to the line response to the
field Eo,(x, t ) . The total potential of the wire relative to the earth, U p e ( xt,) =
U o ( x ,t ) + U e ( x ,t ) , contains another component, Uo(x,t ) , defined by the
charges of the lightning return stroke. Neglecting the ohmic voltage drop
relative to the induction term and taking b’Uo/b’x = -Eox into account, we
arrive at the wave equation with a distributed driving force and containing
no damping term:
zIo
to the two identical waves propagating in opposite directions along the line:
= 1 ‘ d 8O U o ( X l , O ) d O +
Uge(x,t)
(6.53)
XI = x - C ( t - 0 ) . x,= x + c(t - 0 ) .
The integrals give the sum of the above elementary waves moving at light
velocity. The waves are excited by the time variation of the external field
potential U,. For the elementary wave to arrive at the point x at the
moment of time t , the causative variation in U0 must occur at the points
x f Ax earlier, by the time 0 = A x / c . If the time is counted from the
moment of the lightning contact with the line tower, the lower, zero limit
of the integrals of (6.53) should be replaced by the time-of-flight of light
for a minimum distance between the lightning channel and the wire,
In the general case, the difficulties that arise in the calculation of the
integrals depend on how one approximates the lightning current related to
the linear charge in the return stroke wave inducing the field Eo, as well as
on the lightning channel position relative to the wire. Of significance are
the following factors: what object the lightning strikes (the earth or an
element of a power transmission line raised above the ground), the channel
deviation from the normal, as well as its bendings and branching. It is impos-
sible to solve this problem without numerical computations. The question
then arises as to the stage in the study, at which a computer simulation is
most helpful. One should not ignore a numerical integration of initial
equations (6.51), allowing the control of the effect of active resistance R I
which sometimes has a large value. The effective value of R Imay be much
higher than the resistance of the line wire to direct current because of the
skin-effect, the soil resistance, used by the wave as a ‘return wire’, and due
to the energy consumed by the impulse corona. The corona is excited in
the wire by overvoltages and absorbs some of the propagating wave
energy, contributing to its damping. The impulse corona also increases the
effective linear capacitance of the wire, since the electrical charge is localized
not only on the wire surface but in the adjacent air. The charge is delivered
there by streamers of metre lengths. The capacitance CleRdepending on the
local wire overvoltage varies together with the velocity of perturbations in
the wire, U = (C1,,L1)-1/2.This greatly distorts the wave front, since different
sections of the wave front have different velocities. The problem becomes
greatly non-linear and definitely requires a numerical solution.
The calculation formulas given below describe simple situations
neglecting the wave damping in the wire. They have been derived by direct
integration of (6.53) and borrowed from [3]. The lightning channel is
considered to be vertical; the return stroke wave moves towards the cloud
at constant velocity U,. For a rectangular charge wave in the channel of
lightning that has struck the earth (but not a line tower) at a horizontal
distance r from the wire, we obtain for the wire point nearest to the lightning
channel
where p = ur/c, U, is the return stroke velocity, and h is the wire height above
the earth. The time in (6.54) is counted from the moment of the lightning
channel contact with the earth. This formula can be used at t > r / c , i.e.,
after the electromagnetic signal has covered the distance between the channel
and the wire. The overvoltage is still active at a large distance from the stroke
point (x + CQ), where the lightning field effect is negligible, Eox x 0. The
wave reaches that point through the wire, as in the case of a communications
line (this occurs without damping at R I= 0). Such overvoltage waves are
known as wandering waves. For these waves, we have
(6.55)
where the time is counted from the moment of the wave front arrival at the
‘infinitely’ remote point of interest. At t , = r/wr, the function - U g e ( m , t ) has
Figure 6.21. Evolution of an overvoltage at the wire point nearest to the lightning
stroke point (solid curves) and a wandering wave voltage (dashed curve). For a
return stroke, a rectangular current wave model is used.
a maximum:
(6.56)
A wandering wave also arises when a lightning strikes a power wire. The
lightning current spreads along the wire from the stroke point in both
directions, producing very strong overvoltage waves, U ( x ,t ) = Z i ( x , t ) / 2 .
Since the wave resistance is 2 x 250-400R (the smaller value is typical of
ultrahigh voltage lines with split wires of a large equivalent radius), the
current ZM = 30 kA would produce an overvoltage with an amplitude of
3.5-6MV. In reality, the overvoltage is limited to the value of breakdown
voltage for the tower insulator string closest to the lightning stroke, where
the flashover does occur, A wave with an amplitude equal to the string break-
down voltage is a wandering wave in this case. Of course, the overvoltage
may rise again, after the string flashover, due to the self-induction emf of
the tower and to the voltage drop across its grounding, to which the lightning
current runs after the string flashover. Wandering waves are damped by the
same processes that determine the resistance RI in (6.51).
It is important for lightning protection practice to compare the over-
voltages due to direct strokes at a line tower and a wire. In the former
case, the voltage drop across the insulator string is the sum of three
components. The voltage drop across the tower grounding and the induced
voltage of the wire are approximately the same quantitatively but have
opposite signs. This totally gives about -2R,ZM over the string. The mag-
netic component L,Ai has a real effect only on the current impulse front,
and its average value is equal to L,Z,w/tf (L, is the tower self-inductance).
The magnetic component for the first leader of a moderate negative lightning
(ZM = 30 kA, t f = 5 ps) and for a tower of standard size (h, M 30 m) does not
exceed 200 kV but 2R,Iy > 600 kV because of R, 2 10 Cl. It appears that
overvoltages due to a direct moderate stroke at a tower can flashover the
insulation only in lines with voltages less than llOkV, which have strings
less than 1 m in length. For a 220 kV transmission line, a hazard may arise
when the currents are twice as high as the average value, but such lightnings
occur only with a 10% frequency. The hazard of a lightning stroke at a tower
is not high for 500-750 kV transmission lines, since they have long strings. A
reverse flashover may arise from a lightning with 100 kA currents and more,
but their number is less than 1YOof the total. If the lightning current strikes
the wire, the current spreads in both directions along it. With the wave resis-
tance Z > 200 52, we get ZZM/2> 3 MV even for a moderate lightning. This
is sufficient to flashover the insulation of any of the currently operating lines.
A lightning stroke at a wire should always be considered to be hazardous.
(6.59)
Figure 6.22. The design circuit for a current a tower of the line with the protective
cable.
The considerable simplification of the real process can be justified only at low
grounding resistances of the towers, when the current in the wire circuit is
limited mostly by its inductance, and one can neglect the current branching
off to the grounding resistances of all other towers except the one nearest to
the affected tower.
The value of R, in a real transmission line in areas with low conductivity
soils may be several times higher than the normal value, reaching 100 0. Then
the current distribution problem must also take into account the removal of
some of the lightning current to 2-5 towers away from the stroke point. The
equivalent circuit becomes more complicated (chain-like), representing a
series of link circuits identical to the first one. For a more rigorous solution,
the ground-wire is to be considered as a long line with a wave resistance Z,
and many local non-uniformities produced where the ground-wire contacts a
tower. Each tower is then represented as a chain of L, and R, connected in
series. Figure 6.23 illustrates the variation of the current impulse in the
tower with the design circuit. For a circuit with lumped parameters, the neglect
of the tower grounding resistance R, x 10 R does not affect the result much,
while at R, x 1000, the tower current impulse shape changes radically. A
circuit with distributed parameters permits one to follow the effect of con-
secutive wave reflections at the contacts between the ground-wire and the
towers. The current impulse distortion by the reflected waves is especially
Time, ps
Figure 6.23. Current impulse on the struck tower with (solid curves) and without
(dashed curves) allowance for a grounding resistance of the nearest tower. The
model with a linearly raising current front is used for return stroke.
(6.60)
We finish this chapter and the book by describing the lightning effect on
power transmission lines. Scientists are still unable to offer a clear
mathematical description of its complicated mechanism. Modern computer
simulations can infinitely specify and refine a mathematical model of the
lightning effect, with respect to both the electromagnetic field and the object’s
response to a stroke. This is, to some extent, interesting, useful and makes
sense. The process of refining computations has no limit. To illustrate, a
detailed analytical treatment of long line parameters with the account of
the earth’s effect has taken several hundreds of pages in the work by
Sunde [6]. Suppose a superprogramme has been created for the solution of
the lightning protection problem; its application will immediately show
that the great efforts it has required can change but little the existing low
predictability of lightning-induced cut-offs. The key problem today is not a
rigorous mathematical solution of the available equations but an adequate
physical description of the principal physical processes producing a lightning
discharge, its electromagnetic field, and the object’s response to it. For this
reason, we have tried to present simple qualitative models rather than
stringent solutions to the equations. On the other hand, many aspects of
this problem have been omitted, partly because they are not directly related
to lightning as a physical phenomenon and partly due to the lack of space or
to the limited knowledge about the key physical phenomena.
Let us look back at the material presented in this book in order to
emphasize the points of primary importance. After the numerical value of
an overvoltage has been calculated, it is necessary to compare the result
obtained with the flashover voltage of the insulation in order to identify its
possible flashover. Most of the voltage-time characteristics of insulation
strings have been found from tests by standard 1.2/5Ops impulses (here,
the first value is the front duration and the second is the impulse duration
on the 0.5 level). Such a refined impulse has little to do with lightning over-
voltages, and this is clear from figure 6.24. A lightning-induced overvoltage
has necessarily a short-term overshoot arising not only from the current wave
reflection by the grounding of the neighbouring towers but also from the
magnetic induction emf. It is not quite clear how this rapidly damping
04-
0 5 10 15 20
Time, ps
Figure 6.24. Current impulses in a struck tower ( i s ) and in the few neighbouring
towers (il-i3).The wave problem was solved allowing for wave reflection from the
places where grounded wire is connected with the towers.
Refe rences