00293729

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

785

IEEE TRANSACTIONS ON INDUSTRY APPLICATIONS. VOL. 30, NO. 3, MAY/JUNE 1994

Natural Lightning
Martin A. Uman, Fellow, IEEE
Abstruct- The present understanding of natural lightning is
reviewed. Recent research on lightning has been motivated, in
part, by the need to protect advanced ground-based and airborne
systems that utilize low voltage, solid-state electronics; by the
desire to prevent spectacular accidents, such as occurred in 1%9
during the launch of Apollo 12 and in 1987 during the launch of
Atlas-Centaur 67; and by the desire to elucidate the physics of
one of natures most impressive phenomena.

?d
A--

I. INTRODUCTION

ENJAMIN Franklin, more than 200 years ago, proved


that lightning was an electrical discharge and measured
the sign of the cloud charge that produced it [l]. Modem
research on the physics of lightning began in the early 20th
century with the work of C.T.R.Wilson [2], [3], the same
scientist who received the Nobel Prize for his invention of
the cloud chamber. Wilson, by making and analyzing remote
measurements of thunderstorm electric fields, was the first to
infer the charge structure of the thundercloud and the amount
of charge involved in lightning. In the 1930s, lightning
research was motivated primarily by the need to reduce the
effects of lightning on electric power systems and by the
desire to understand an important meteorological process.
The pace of that research was fairly steady until the 1960s
when there was renewed interest because of the generally
unexpected vulnerability of solid state electronics to damage
from lightning-induced voltages and currents with the resultant
hazard to both modem ground-based and airborne systems. A
good deal of the recent lightning research has been motivated
by three spectacular accidents involving aircraft or spacecraft:
(i) In 1963, a Boeing 707 jetliner flying near Elkton, MD,
at an altitude of about 5000 feet was struck and, via a fuel
explosion, was destroyed by lightning, killing all occupants
[4]. (ii) In 1969, the Apollo 12/Satum V space vehicle, during
launch through a weak cold front that was not producing
natural lightning, artificially initiated (or triggered) two
lightning flashes, one to ground and one intracloud discharge
[5], [6]. The vehicle and its crew survived various major
system upsets and minor permanent damage and were able
to complete successfully their mission to the moon. (iii) In
1987, an unmanned Atlas-Centaur space vehicle (AC167)was
launched into weather conditions similar to those present at
the launch of Apollo 12 and triggered a lightning discharge
that compromised the vehicles control system, resulting in
the destruction of the Atlas-Centaur [7].
Paper ICPSD 93-1, approved by the Power Ssytems Engineering Department of the IEEE Industry Applications Society, for presentation at the
1993 IEEE PROCEEDINGSand the 1993 I&CPS Dept. Technical Conference.
Manuscript released for publication October 7,1993. This work was supported
by the National Science Foundation under Grant Am-9014084.
M. A. Uman is with the Department of Electrical Engineering, University
of Florida, Gainesville, FL 3261 1 USA.
IEEE Log Number 9400166.

-)
2

Fig. 1. Thundercloudcharge distribution and categorization of the four types


of lightning between cloud and ground. (Reprinted from [26] with permission.)

In this article, we will survey our present knowledge of


natural lightning. A companion paper in this issue will address
artificially initiated lightning.
11. SOURCES OF LIGHTNING
Most research on the electrical structure of clouds has
focused on the cumulonimbus, the familiar thundercloud or
thunderstorm, because this cloud type produces most of the
lightning. There have been limited studies of the electrical
properties of other types of clouds such as stratus, stratocumulus, cumulus, nimbostratus, altocumulus, altostratus, and
cirrus clouds that might potentially produce lightning or, at
least, long (1 to 100 m) sparks [8], [9].
The classic model for the charge structure of a thundercloud
was developed in the 1920s and 1930s from ground-based
measurements of both thundercloud electric fields and the
electric field changes that are caused when lightning occurs
[2], [3], [lo], [ l l ] . In this model, the thundercloud forms a
positive electric dipole as shown in Fig. 1 and Fig. 2; that
is, a primary positive charge region is found above a primary
negative charge region. By the end of the 1930s, Simpson and
co-workers [ 121, [ 131 had verified this overall structure from

0093-9994/94$04.00 0 1994 IEEE

IEEE TRANSACTIONS ON INDUSTRY APPLICATIONS, VOL 30, NO 3. MAY/JLI\IE 1994

786

CLOUDCHARGE
DETRIBUTION

PRELIMINARY
BREAKDOWN

1.00 ms

t=O

STEPPED
LEADER

1.10 ms

1.15 rns

20.00 ms

19.00 ms

;. .

RETURN

20.10 ms

,-

*.
?
.

1.20 ms

20.15

ms

20.20 ms

-{-

KANDJ
PROCESSES

40.00 ms
Fig. 2.

60.00 ms

61.00 ms

62.00 ms

62.05 ms

Various processes that make up a negative CG lightning discharge. (Reprinted from (261 with permission.)

measurements made with sounding balloons inside clouds and


had also identified a small localized region of positive charge
at the base of the cloud. Subsequent measurements of electric
fields both inside and outside the cloud have confirmed the
general validity of this double-dipole structure [ 141-[24].
111. NATURALLIGHTNING

Lightning is a transient, high-current discharge whose path


length is measured in kilometers. Well over half of all flashes
occur wholly within the cloud and are called intracloud (IC)
discharges. Cloud-to-ground (CG) lightning has been studied
more extensively than other forms of lightning because of its
practical importance (for instance, as the cause of injuries and
death, disturbances in power and communication systems, and
the ignition of forest fires) and because lightning in the clear
air below the cloud base is more easily studied with optical
techniques. Cloud-to-cloud and cloud-to-air discharges occur
less frequently than either IC or CG lightning. All discharges
other than CG are often combined under the general term cloud
discharges.
Four different types of lightning between cloud and Earth
have been identified, the ways by which these are initiated

being shown in Fig. 1. [25]. CG flashes initiated by downwardmoving negatively-charged leaders probably account for about
90% of the CG discharges worldwide (Fig. 1, category I ) ,
while less than 10% of lightning discharges are initiated by
a downward-moving positive leaders (category 3). Ground-tocloud discharges are also initiated by leaders of either polarity
that move upward from the Earth (categories 2 and 4). These
upward-initiated flashes are relatively rare and usually occur
from mountain peaks and tall man-made structures.
The number of cloud-to-ground flashes per square kilometer
per year has a maximum of 30 to 50, and a typical over-land
value of 2-5. Between 10 and 20 million CG flashes strike
the continental United States annually. Worldwide there are
about 100 total (cloud and ground) flashes per second for a
worldwide average flash density of about 6 km-2yr-' [26].

IV. NEGATIVE
CG LIGHTNING
A negative CG discharge (Fig. I , category I ) begins in
the cloud and effectively lowers some tens of coulombs of
negative charge to Earth. The total discharge is termed a
flash (as is the total discharge for other types of lightning).
Flash durations are typically about half a second. A flash

UMAN NATURAL LIGHTNING

has several components, the most significant being three or


four high-current pulses called strokes. Each stroke lasts
about a millisecond, and the separation between strokes is
typically several tens of milliseconds. Lightning often appears
to flicker because the human eye can just resolve the
individual pulses of luminosity that are produced by each
stroke.
The sequence of luminous processes that are involved
in a typical negative CG flash is shown in Fig. 2. The
stepped leader initiates the first return stroke after it propagates
downward in a series of discrete steps. The stepped leader is
itself initiated by a preliminary breakdown within the cloud,
although there is no agreement about the exact form and
location of this process. High-speed photographs show that
leader steps are typically 1 ps in duration, tens of meters in
length, and that the pause time between steps is 20 to 50 ps. A
fully developed stepped leader can effectively lower 10 C or
more of negative charge toward the ground in tens of ms. The
average downward speed of propagation is about 2 x lo5 m/s.
The average leader current is between 100 and lo00 A. The
leader steps have peak pulse currents of at least 1 kA. During
its progression toward ground, the stepped leader produces a
downward-branched geometrical structure.
The potential difference between the lower portion of the
negatively charged leader and the Earth has a magnitude in
excess of lo7 V. As the tip of the leader nears ground, the electric field at sharp objects on the ground or at irregularities on
the surface increases until it exceeds the breakdown strength
of air. At that time, one or more upward-moving discharges
are initiated from those points, and the attachment process
begins. When one of the upward-moving discharges contacts
the downward-moving leader, some tens of meters above the
ground, the leader is effectively connected to ground potential.
The leader channel is then discharged by an ionizing wave of
ground potential that propagates up the previously charged
leader channel. This process is the first return stroke. The
electric field across the difference in potential between the
return stroke, which is at ground potential, and the channel
above, which is near cloud potential, is what produces the
additional ionization. The upward speed of a return stroke
is typically one-third to one-half the speed of light near the
ground, and the speed decreases with height. The total transit
time between ground and cloud is on the order of 100 ps.
The first return stroke produces a peak current of typically 30
kA at the ground, with a time from zero to peak of a few
microseconds. Currents measured at the ground decrease to
half the peak value in about 50 ps, and currents of the order
of hundreds of amperes may flow for times of a few to several
hundred milliseconds. We will discuss these long-duration,
low-amplitude currents later in this section.
The rapid release of return-stroke-energy heats the leader
channel to a peak temperature above 30 000 K and produces
a high-pressure channel that expands and creates the shock
waves that eventually become thunder. The return stroke
effectively lowers to ground the negative charge originally
deposited on the stepped leader channel including all the
branches, as well as other negative cloud charge that may
become available at the top of the channel.

I81

When the return-stroke current ceases, the flash, including


various discharge processes within the cloud, may end. In that
case, the lightning is called a single-stroke flash. On the other
hand, if additional cloud charge is available, a continuous dart
leader can propagate down the residual first-stroke channel
and initiate another return stroke. During the time between
the end of the first return stroke and the initiation of a dart
leader, so-called J- and K-processes occur in the cloud. The
J-process involves charge motion in the cloud on a tens of
milliseconds time scale, while the K-process moves charge
on a time scale ten times shorter. The dart leader has a peak
current of 1 kA or more and lowers a total charge on the order
of 1 C at a speed of about 3 x lo6 m / s . Some dart leaders
become stepped leaders toward the end of their progression
toward ground and do not follow the previous return stroke
channel. Dart leaders and return strokes subsequent to the first
are usually not branched.
The time between successive strokes in a flash is usually
several tens of milliseconds, but can be tenths of a second if a
continuing current persists in the channel after a return stroke.
Continuing currents are of the order of 100 A and represent
a direct transfer of charge from cloud to ground [27]-[31].
Between 25 and 50% of all CG flashes contain a continuing
current component [27]-[3 13.
A comprehensive list of lightning current parameters has
been derived from tower measurements in Switzerland during
strikes initiated by downward leaders [32], [33]. In these tower
studies the measured maximum rate-of-rise of current and the
duration of the current front are limited by the bandwidth of
measuring apparatus. The best data for these parameters are
probably derived from triggered lightning studies, where the
highest maximum rate-of-rise of current were found to be 4
x 10l1 A / s with a typical value of 1011 A / s and typical rise
times in the 100 ns range [34], [35]. Additional discussion of
these parameters is found in the last section of this paper.

V. POSITIVECG LIGHTNING
Positive flashes to ground (category 3 in Fig. 1) are of
considerable practical interest because their peak currents and
total positive charge transfer to ground can be much larger
than the more common negative flashes [36]. Positive flashes
to ground are initiated by leaders that do not exhibit as distinct
steps as their negative counterparts. Rather, they exhibit a more
or less continuous luminosity that is modulated in intensity.
Positive flashes usually contain only a single return stroke
followed by a period of continuing current [37], [38]. Positive
flashes are probably initiated by the upper positive charge in
thunderclouds (Fig. 1) where this charge has been separated
horizontally from the lower negative charge by wind shear
[36], [39], but this may not always be a necessary condition
[40]-[42]. Positive flashes are the majority of flashes to ground
in winter thunderstorms (and snowstorms) even though these
storms produce few flashes overall; and they are relatively rare
in summer thunderstorms, only 1 to 15% of the flashes [43],
[44], although storms with predominantly negative lightning
often end with positive discharges. The fraction of positive
discharges in summer thunderstorms apparently increases with

Iht-1; TRAKSACTIONS O N INDUSTRY APPLICNIONS, VOL.. 30. NO 3. MAY/JUNE 1Y93

788

increasing geographic latitude and with increasing height of


the local terrain; that is, the closer the cloud charge is to the
ground, the more probable is positive lightning, but again, not
enough is known about positive lightning to be able to say
that this is always a necessary condition.

VI. UPWARD LIGHTNING


The leaders in upward-initiated lightning are usually positive (Fig. 1 , category 2). Positive upward leaders show a
continuous luminosity that is modulated in a fashion similar to
positive downward-stepped leaders. Negati1.e upward leaders
(category 4) exhibit a stepped behavior that is similar to
negative downward-stepped leaders.
Positive upward leaders often enter the cloud and produce
only a more or less continuous flow of current, of the order
of 100 to 1000 A, at ground. In about half of the upwardinitiated events, however, the continuous current is followed
by a sequence of dart leaders and retum strokes that are similar
to those following first strokes in natural CG discharges that
are initiated by negative downward-moving leaders. A detailed
review of the measurements that have been made prior to 1987
on upward-initiated lightning is found in [26].
Lightning that is artificially initiated, as discussed in a
companion paper in this issue, begins with an upward leader
from the triggering rocket.

VII.

CLOUD

DISCHARGES

Cloud discharges can be subdivided into IC, intercloud,


and cloud-to-air flashes, but there are no experimental data at
present to distinguish between these three types. Indeed. on the
basis of electric field records, there is considerable similarity
between these discharges [4SJ.The term cloud discharge could
also be applied to those portions of a flash to ground that take
place within the cloud. In some cases, flashes that are primarily
within the cloud, and are best characterized as cloud flashes.
produce a channel to ground, seemingly as an unimportant
by-product.
Intracloud flashes typically occur between positive and
negative charge regions or represent discharges away from
concentrated regions of positive or negative charge and have
total durations that are nearly the same as ground flashes, about
half a second. A typical cloud discharge effectively moves tens
of coulombs of charge over a distance of 5 to I O km. The
discharge process is thought to consist of a breakdown phase
followed by a continuously propagating leader that generates
weak return strokes called recoil streamers when the leader
contacts pockets of space charge opposite to its own. The
electric field changes that are associated with recoil streamers
are termed K-changes. K-changes are thought to be similar,
but usually of opposite polarity, to the Kchanges that occur
in the intervals between retum strokes in CG discharges. A
detailed review of the literature on cloud discharges is found
in [26j.
VIII. TOP-OF-THEc 1 2 0 U D AND CLEAR AIR LIGHTNING
There are numerous reports of lightning propagating upward from the tops of clouds and perhaps to the ionosphere

[46]-[52]. Lightning has also been reported when there is a


clear blue sky [531-[58), commonly referred to as a bolt from
the blue. Most of these reports, however, refer to a situation
where there is blue sky overhead and the thunderstorm is 10 or
more km away, out of viewing range, from where the lightning
originates. However, there are photographs and supporting
charge
discharge can
- locations that show that a triggered
__
occur entirely in clear air near a thunderstorm 1581. In the
case cited, there was a thunderstorm about 10 km away, and
the lightning was artificially initiated by firing a small rocket
upward that trailed a grounded wire. There were high electric
fields, but the aky overhead where the charge appears to have
been located was mostly clear with broken altocumulus and
altostratus clouds at higher altitudes.

1x. LIGHTNING AVOIDANCE


AND PROTECTION
When one is faced with the possibility of a lightning hazard,
there are two methods of protection: (i) identify and avoid the
hazard and (ii) harden tile system of interest to withstand the
effects of nearby and direct strikes.
In the last 10 years, exceptional progress has been made
in the technology for identifying and locating natural CG
lightning, and commercial nationwide networks of lightning
sensors with a wideband magnetic direction-finding technology that produce lightning locations in real time are now
operational in the United States and in about IS other countries
[ 59]-[63]. Many networks are being used both for research
and for providing warnings of natural lightning in a variety
of applications. Two addition techniques for lightning location
have recently become available commercially, time-of-arrival
and interferometry, with these technologies now being used in
the U.S. and in several other countries 1641-[691.
In order to design better lightning protection systems to
improve the capability of ground-based facilities and of aircraft
and space vehicles to withstand the effects of nearby or direct
strikes, it is essential to understand the nature of the lightning
environment near and at the point o f a direct strike. Among the
most important parameters responsible for coupling lightning
signals to electronic systems is the maximum current and the
maximum rate of change of current with respect to time.
Return stroke current measurements made on towers with
wideband transient recorders I 701-1 72 1, currents measured
during lightning strikes to a F-l06B research aircraft (731.
and retum stroke current measurements in rocket-initiated
discharges [ 341, 13.51 show that the peak current rate-of-rise
is considerably larger than was believed to be the case a
decade or two ago. A maximum value of 3.8 x 101A/s has
been measured on the F-106 B aircraft in flight, and values
above 2 x 10 A/s occurred frequently 1731. Only a few
measurements of current rate of rise have been made with
sufficient bandwidth on natural retum strokes. but from this
small sample a peak value of 1.8 x lO1]A/s has been observed
1701-[72]. Finally, the rocket-triggered discharges to saltwater
at the NASA Kennedy Space Center have produced a peak
current rate-of-rise for return strokes of 4.1 x lo1 A/s [ 341,
[ 3 S ] .Thus, from the limited data that are available, we expect

U M A N NATURAL LIGHTNING

that a peak current rate-of-rise near 5 x 10I1 A / s is possible,


even at flight altitudes.
Although considerable progress has been made recently in
understanding lightning vis-a-vis lightning protection, we still
need to understand more thoroughly the following important
aspects of lightning: the total wave shapes of the currents that
are associated with all of the salient lightning processes in both
natural and triggered events, the physical processes by which
natural lightning attaches to ground-based structures and to
aircraft, and the role of the structures and aircraft in initiating
lightning.

REFERENCES
B. Franklin, Experiments and Observations on Electricity Made at
Philadelphia. London, Great Britain: E. Cave, 1774; see also I. B.
Cohen, Benjamin Franklins Experiments. Cambridge, MA: Harvard
Univ. Press, 1941.
C. T. R. Wilson, On some determinations of the sign and magnitude
of electric discharges in lightning flashes, Proc. Roy. Soc., vol. A92,
pp. 555-574, 1916.
C. T. R. Wilson, Investigations on lightning discharges and on the
electric field of thunderstorms, Phil. Trans. R. Soc., vol. A221, p. 73,
1920.
Aircraft Accident Report, Boeing 707-121, N709PA Pan American
World Airways, Inc., near Elkton, MD, Dec. 8, 1963, Civil Aeronautic
Board File No. 1-0015, Feb. 25, 1965.
R. Godfrey, E. R. Mathews, and J. A. McDivitt, Analysis of Apollo
12 Lightning Incident, NASA MSC-01540, Jan. 1970.
E. P. Krider, R. C. Noggle, M. A. Uman, and R. E. Orville, Lightning
and the Apollo 17/Satum V exhaust plume, J. Spacecraf? Rockets, vol.
11, pp. 72-75, 1974.
J. Bussey, Report of AtlaslCentaur-67FLTSATCOMF-6 Investigation
Board, vol. 11, NASA, 1987.
G. C. Simpson, Atmospheric electricity during disturbed weather,
Geophys. Mem. vol. 84, p. 1, 1949.
I. M. Imyanitov, Ye V. Chubarina, and Ya M. Shvarts, Electricity in
Clouds, Rep. IT F-718, NASA, Washington, DC, 1972, (translation
from Russia).
B. F. J. Schonland and J. Craib, The electric fields of South African
thunderstorms, Pmc. R. Soc., vol. A114, p. 229, 1927.
E. J. Workman, R. E. Holm, and G. T. Pelsor, The Electrical Structure
of Thunderstorms, tech. notes of the National Advisory Committee on
Aeronautics, no. 864, US. Government Printing Office, Washington,
DC, 1942.
G. S. Simpson and F. J. Scrase, The distribution of electricity in
thunderclouds, Proc. R. Soc., vol. A161, p. 309, 1937.
G. C. Simpson and G. D. Robinson, The distribution of electricity in
the thunderclouds, Part 11, Pmc. R. Soc., vol. A177, p. 281, 1941.
J. Kuettner, The electrical and meteorological conditions inside thunderclouds, J. Meteorol., vol. 7, p. 322, 1950.
D. J. Malan, Les descharges dans lair et la charge inferieure positive
dun nuage orageux, Ann. Geophys., vol. 8, p. 385, 1952.
S. E. Reynolds and H. W. Neil, The distribution and discharge of
thunderstorm chargecenters, J. Meteorol., vol. 12, p. 1, 1955.
A. Huzita and T. Ogawa, Charge distribution in the average thunderstorm cloud, J. Meteorol. Soc. Japan, vol. 54, p. 285, 1976.
W. P. Winn, C. B. Moore, and C. R. Holmes, Electric field structure
in an active part of a small, isolated thundercloud, J. Geophys. Res.,
vol. 86, p. 1187, 1981.
T. C. Marshall and W. P. Winn, Measurements of charged precipitation
in a New Mexico thunderstorm: Lower positive charge centers, J.
Geophys. Res., vol. 87, p. 7141, 1982.
T. Taniguchi, C. Magnono, and T. Endoh, Charge distribution in active
winter clouds, Res. Lett. Amos. Electr., vol. 2, p. 35, 1982.
M. E. Weber, H. J. Christian, A. A. Few, and M. F. Stewart, A
thundercloud electric field sounding: Charge distribution and lightning,
J. Geophys. Res., vol. 87, p. 7158, 1982.
M. E. Weber, M. F. Stewart, and A. A. Few, Corona point measurements in a thundercloud at Langmuir Laboratory, J. Geophys. Res.,
vol. 88, p. 3907, 1983.
G. J. Byne, A. A. Few, and M. E. Weber, Altitude, thickness and
charge concentration of charged regions of four thunderstorms during
TRIP 1981 based upon in situ balloon electric field measurements,
Geophys. Res. Lett., vol. 10, p. 39, 1983.

789

1241 W. J. Koshak and E. P. Krider, Analysis of lightning field changes


during active Florida thunderstorms, J. Geophys. Res., vol. 94, p. 1165,
1989.
[25] K. Berger, Blitzstrom-Parameter von AutwWsblitzen, Bull. Schweiz.
Elektrotech. Ver., vol. 69, pp. 353-360, 1978.
[26] M. A. Uman, The Lightning Discharge. San Diego, CA: Academic
Press, 1987.
1271 P. R. Krehbiel, M. Brook, and R. McCrory, Analysis of the charge
structure of lightning discharges to the ground, J. Geophys. Res., vol.
84, pp. 2432-2456, 1979.
1281 M. Brook, N. Kitagawa, and E. J. Workman, Quantitive study of strokes
and continuing currents in lightning discharges to ground, J. Geophys.
Res., vol. 67, pp. 649-659, 1962.
[29] N. Kitagawa, M. Brook and E.J. Workman, Continuing currents in
cloud-to-ground lightning discharges, J. Geophys. Res., vol. 67, pp.
637-647, 1962.
[30] T. Shindo and M. A. Uman, Continuing current in negative cloudto-ground lightning, J. Geophys. Res., vol. 94, p. 5189, 1989.
[31] V. A. Rakov and M. A. Uman, Long continuing current in negative
ground flashes, J. Geophys. Res., vol. 95, pp. 5455-5470, 1990.
[32] K. Berger, R. B. Andersonand H. Kroninger, Parameters of lightning
flashes, Electra, vol. 80, pp. 23-37, 1975.
[33] K. Berger, Methoden und Resultate der Blitzforschung auf dem Monte
San Salvatore bei Lugano in den jahren 1963-1971, Bull. Schweiz.
Elektrotech. Ver., vol. 63, pp. 1403-1422, 1972.
[34] C. Leteinturier, C. Weidman, and J. Hamelin, Current and electric field
derivatives in triggered lightning return strokes, J. Geophys. Res., vol.
95, pp. 811-828, 1990.
[35] Leteinturier, C., J. H. Hamelin, and A. Eybert-Berard, Submicrosecond
characteristics of lightning return-stroke currents, ZEEE Trans. EMC,
vol. 33, 351-357, 1991.
1361 M. Brook, M. Nakano, P. Krehbiel, and T. Takeuti, The electrical
structure of the Hokuriku winter thunderstorms, J. Geophys. Res., vol.
87, pp. 1207-1215, 1982.
1371 W. D. Rust, W. L. Taylor, D. R. MacGorman, and R. T. Arnold,
Research on electrical properties of severe thunderstorms in the Great
Plains, Bull Am. Met. Soc., vol. 62, pp. 1286-1293, 1981.
[38] D. M. Fuquay, Positive cloud-to-ground lightning in summer thunderstorms, J. Geophys. Res., vol. 87, pp. 7131-7140, 1982.
1391 N. Takagi, T. Takeuti, and T. Nakai, On the Occurrence of positive
ground flashes, J. Geophys. Res., vol. 91, pp. 9905-9909, 1986.
[40] S. A. Rutledge and D. R. MacGorman, Cloud-to-ground lightning
activity in the 10-1 1 June 1985 mesoscale convective system observed
during Oklahoma-Kansas pre-storm project, Mon. Weather Rev., vol.
116, p. 1393, 1988.
[41] D. R. MacGorman and W. L. Taylor, Positive cloud-to-ground lightning
detection by a direction-finder network, J. Geophys. Res., vol. 94, pp.
13313-13318, 1989.
1421 S. A. Rutledge, C. Lu, and D. R. MacGorman, Positive cloud-toground lightning in mesoscale convective systems, J. Amos. Sci., vol.
47, pp. 2086-2100, 1990.
1431 R. E. Orville, R. A. Weisman, R. B. Pyle, R. W. Henderson, and R. E.
Orville, Jr., Cloud to ground lightning flash characteristics from June
1984 through May 1985, J. Geophys. Res., vol. 92, pp. 56465644,
1987.
1441 J. Huse and K. Olsen, Some characteristics of lightning ground flashes
observed in Norway, in Lighfning and Power Systems, ZEEE Conference Publication Number 236. London and New York Institution of
Electrical Engineers, 1984, pp. 72-76.
1451 T. Ogawa and M. Brook, The mechanism of the intracloud lightning
discharge, J. Geophys. Res., vol. 69, pp. 5141-5150, 1964.
[46] 0. H. Vaughan and B. Vonnegut, Lightning to the ionosphere, Weatherwise, vol. 35, pp. 7-72, 1982.
[47] B. Vonnegut, Vertical lightning, Weatherwise, vol. 37, p. 61, 1984.
[48] M. Brook, C. Rhodes, 0. H. Vaughan, Jr., R. E. Orville, and B.
Vonnegut, Nighttime observations of thunderstorm electrical activity
from a high altitude airplane, J. Geophys. Res., vol. 90, pp. 61 11-6120,
1985.
[49] 0. H. Vaughan, Jr. and B. Vonnegut, Recent observations of lightning
discharges from the top of a thundercloud into the clear air above, J.
Geophys. Res., vol. 94, pp. 13179-13182, 1989.
[50] P. A. M OLightning
,
above a large thunderstorm, Weatherwise, vol.
43, no. 6, p. 363, June 1990.
[51] R. C. Franz, R. J. Nemzek and J. R. Winckler, Television image
of a large upward electrical discharge above a thunderstorm system,
Science, vol. 249, pp. 48-51, 1990.
[52] 0. H. Vaughan, Jr., R. Blakeslee, W. L. Boeck, B. Vonnegut, M. Brook,
and J. McKune, Jr., A cloud-to-space lightning as recorded by the

790

1531
1541

1551
[56)
1571

1581
[59]
[60]
[61]

[62]

[631
1641
[65]

[66]

[67]

IEEE TRANSACTIONS ON INDUSTRY APPLICATIONS, VOL. 30. NO. 3. MAYIIUNE 1994

Space Shuttle payload-bay TV cameras, BUN. Am. Meteord. Soc., vol.


120, pp. 1459-1461, 1992.
2. A. McCaughan, A lightning stroke far from the thunderstorm cloud,
Mon. Weather Rev., vol. 54, p. 344, 1926.
H. F. Gishorne, Lightning from clear sky, Mon. Weather Rev., vol.
56, p. 108, 1928.
F. Myers. Lightning from a clear sky, Mon. WenlhPr Rev., Vol. 59,
pp. 3 9 4 0 , 1931.
T. Gifford, Aircraft struck by lightning. Mereorol. Mag., Vol. 79. PP.
121-122, 1950.
D. Baskin, Lightning without clouds. Bull. Am. M e t e o r d . Soc., Vol.
33, p. 348, 1952.
P. Waldteufel, P. Metzger, J. L. Boulay, P. Laroche, and P. Hubert,
Triggered lightning strokes originating in clear air. J . Geophys. Res.,
vol. 85, pp. 2861-2868, 1980.
E. P. Krider, R. C. Noggle, and M. A. Uman, A gated wideband mag.
netic direction finder for lightning return strokes, J. A .p.i ~ / Meteoro:.,
vol. 15, p. 301, 1976.
E. P. Krider, R. C. Noggle, A. E. Pifer, and D. L. Vance, Lightning
direction-finding systems for forest fire detection, Bull. Am. Mereorol.
Soc., vol. 61, p. 980, 1980.
S. Israelsson, T. Schutte, E. Pisler, and S. Lundquist, Increased occurrence of lightning flashes in Sweden during 1986, J. Geophys. Res.,
vol. 92, pp. 10996-10998, 1987.
J. Hojo, M. Ishii, T. Kawamura, F. Suzuki, H. Komuro and M.
Shiogama, Seasonal variations of cloud-to-ground lightning flash characteristics in the coastal area of the Sea of Japan, J. Grophys. Rer.,
vol. 94, pp. 13207-13212, 1989.
R. E. Orville, Lightning ground flash density in the contiguous United
States-1989, Mon. Weather Review,, vol. 119, pp. 1557-1574, 1991.
A. C. L. Lee, An operational system for the remote location of lightning
flashes using a VLF arrival time difference technique, J. A m . und
Oceanic Tech., vol. 3, pp. 63C642, 1986.
R. B. Bent and W. A. Lyons, Theoretical evaluation5 and initial
operational experiences of LPATS (lightning position and tracking
system) to monitor lightning ground strikes using a time-of-arrival
(TOA) technique, in 7rh Int. Conj Atmospheric Electricitv, June 3-8,
1984, Albany, NY, pp. 317-324.
E. M. Thomson and P. J. Medelius, A system for mapping sources of
VHF and electric field pulses from in-cloud lightning at KSC, in Pror.
JY9J Int. Aerospace nnd Ground Conj Lightning nnd Static Elrctricih,
Cocoa Beach. FL, Apr. 16, 1991, pp. 70-10.
C. 0. Hayenga, Characteristics of lightning VHF radiation near the
time of return strokes,J. Geophys. Res., vol. 89, pp. 1403-1410, 1984.

[68) C. 0. Hayenga and J. W.Warwick, Two-dimensional interferometric


positions of VHF lightning sources, J. Ceophys. Res., vol. 86, pp.
7451-7464, 1981.
(691 P. Richard. A. Delannoy, G. Labaune, and P. Laroche, Results of spatial
and temporal characterization of the VHF-UHF radiation of lightning,
J . Geophys. Res.. vol. 91, pp. 1248-1260, 1986.
[70] A. J . Eriksson, Lightning and tall structures, Trans. S. Afr. inst. N e c .
Eng.. vol. 69, pp. 2-16, Aug. 1978.
1711 0. Beierl, Lightning cumnt measurements at the Peissenberg tower,
in 7th ISH, Dresden, Germany, ref. 81.02, 1991.
of negative subsequent strokes mea.
~ 7 2 10 ~ ~ Front
i ~ ~
shape
l ,
sured at the Pcissenherg lower, in 21st Inf. Conf: Lightning Protecrion,
Berlin, Germany, Sept. 21-25, 1992, pp. 19-24.
1731 F. L. Pitts, L. D. Lee, R. A. Perala, and T. H. Rudolf, New Methods
and Results for Quantification of Lightning-Aircraft Electrodynamics,
NASA Tech. Paper 27/37, June 1987.

Martin A. Uman, (SM73, F88) received the


bachelors, masters and doctoral degrees from
Princeton University, Princeton, NJ, USA, the latter
in 1961.
He is currently a Professor of Electrical
Engineering and Chairman of the Department
of Electrical Engineering at the University of
Florida, Gainesville, FL. He has also headed that
Departments Lightning Research Laboratory for
more than 20 years. He has written three books on
lightning (two in second editions) as well as a book
on plasma physics, and has published more than 125 reviewed journal papers
on lightning and long sparks. He was an associate professor of Electrical
Engineering at the University of Arizona, Tucson, AZ, from 1961 to 1964.
He joined the University of Florida faculty in 1971 after working as a fellow
physicist at Westinghouse Research Labs in Pittsburgh. He co-founded and
served as president of Lightning Location and Protection, Inc. from 1975
to 1983.
Dr. Uman was named the 1990 Florida Scientist of the Year by the Florida
Acadetny of Sciences. Honored as 1988-89 University of Florida TeacherScholar of the Year (the highest faculty award at UF), he is a Fellow of three
professional societies: the American Geophysical Union (AGU). American
Meteorological Society (AMS), and the IEEE.

You might also like