(IET Energy Engineering Series, 127) Vernon Cooray (Editor), Farhad Rachidi, Marcos Rubinstein - Lightning Electromagnetics, Volume 2_ Return Electrical Processes and Effects-The Institution of Engine
(IET Energy Engineering Series, 127) Vernon Cooray (Editor), Farhad Rachidi, Marcos Rubinstein - Lightning Electromagnetics, Volume 2_ Return Electrical Processes and Effects-The Institution of Engine
Lightning
Electromagnetics
Other volumes in this series:
Volume 1 Power Circuit Breaker Theory and Design C.H. Flurscheim (Editor)
Volume 4 Industrial Microwave Heating A.C. Metaxas and R.J. Meredith
Volume 7 Insulators for High Voltages J.S.T. Looms
Volume 8 Variable Frequency AC Motor Drive Systems D. Finney
Volume 10 SF6 Switchgear H.M. Ryan and G.R. Jones
Volume 11 Conduction and Induction Heating E.J. Davies
Volume 13 Statistical Techniques for High Voltage Engineering
W. Hauschild and W. Mosch
Volume 14 Uninterruptible Power Supplies J. Platts and J.D. St Aubyn (Editors)
Volume 15 Digital Protection for Power Systems A.T. Johns and S.K. Salman
Volume 16 Electricity Economics and Planning T.W. Berrie
Volume 18 Vacuum Switchgear A. Greenwood
Volume 19 Electrical Safety: a guide to causes and prevention of hazards
J. Maxwell Adams
Volume 21 Electricity Distribution Network Design, 2nd Edition E. Lakervi and
E.J. Holmes
Volume 22 Artificial Intelligence Techniques in Power Systems K. Warwick, A.O. Ekwue
and R. Aggarwal (Editors)
Volume 24 Power System Commissioning and Maintenance Practice K. Harker
Volume 25 Engineers’ Handbook of Industrial Microwave Heating R.J. Meredith
Volume 26 Small Electric Motors H. Moczala et al.
Volume 27 AC-DC Power System Analysis J. Arrillaga and B.C. Smith
Volume 29 High Voltage Direct Current Transmission, 2nd Edition J. Arrillaga
Volume 30 Flexible AC Transmission Systems (FACTS) Y-H. Song (Editor)
Volume 31 Embedded generation N. Jenkins et al.
Volume 32 High Voltage Engineering and Testing, 2nd Edition H.M. Ryan (Editor)
Volume 33 Overvoltage Protection of Low-Voltage Systems, Revised Edition P. Hasse
Volume 36 Voltage Quality in Electrical Power Systems J. Schlabbach et al.
Volume 37 Electrical Steels for Rotating Machines P. Beckley
Volume 38 The Electric Car: Development and future of battery, hybrid and fuel-cell
cars M. Westbrook
Volume 39 Power Systems Electromagnetic Transients Simulation J. Arrillaga and
N. Watson
Volume 40 Advances in High Voltage Engineering M. Haddad and D. Warne
Volume 41 Electrical Operation of Electrostatic Precipitators K. Parker
Volume 43 Thermal Power Plant Simulation and Control D. Flynn
Volume 44 Economic Evaluation of Projects in the Electricity Supply Industry
H. Khatib
Volume 45 Propulsion Systems for Hybrid Vehicles J. Miller
Volume 46 Distribution Switchgear S. Stewart
Volume 47 Protection of Electricity Distribution Networks, 2nd Edition J. Gers and
E. Holmes
Volume 48 Wood Pole Overhead Lines B. Wareing
Volume 49 Electric Fuses, 3rd Edition A. Wright and G. Newbery
Volume 50 Wind Power Integration: Connection and system operational aspects
B. Fox et al.
Volume 51 Short Circuit Currents J. Schlabbach
Volume 52 Nuclear Power J. Wood
Volume 53 Condition Assessment of High Voltage Insulation in Power System
Equipment R.E. James and Q. Su
Volume 55 Local Energy: Distributed generation of heat and power J. Wood
Volume 56 Condition Monitoring of Rotating Electrical Machines P. Tavner, L. Ran,
J. Penman and H. Sedding
Volume 57 The Control Techniques Drives and Controls Handbook, 2nd Edition
B. Drury
Volume 58 Lightning Protection V. Cooray (Editor)
Volume 59 Ultracapacitor Applications J.M. Miller
Volume 62 Lightning Electromagnetics V. Cooray
Volume 63 Energy Storage for Power Systems, 2nd Edition A. Ter-Gazarian
Volume 65 Protection of Electricity Distribution Networks, 3rd Edition J. Gers
Volume 66 High Voltage Engineering Testing, 3rd Edition H. Ryan (Editor)
Volume 67 Multicore Simulation of Power System Transients F.M. Uriate
Volume 68 Distribution System Analysis and Automation J. Gers
Volume 69 The Lightening Flash, 2nd Edition V. Cooray (Editor)
Volume 70 Economic Evaluation of Projects in the Electricity Supply Industry,
3rd Edition H. Khatib
Volume 72 Control Circuits in Power Electronics: Practical issues in design and
implementation M. Castilla (Editor)
Volume 73 Wide Area Monitoring, Protection and Control Systems: The enabler for
Smarter Grids A. Vaccaro and A. Zobaa (Editors)
Volume 74 Power Electronic Converters and Systems: Frontiers and applications
A.M. Trzynadlowski (Editor)
Volume 75 Power Distribution Automation B. Das (Editor)
Volume 76 Power System Stability: Modelling, analysis and control A.A. Sallam and
B. Om P. Malik
Volume 78 Numerical Analysis of Power System Transients and Dynamics A. Ametani
(Editor)
Volume 79 Vehicle-to-Grid: Linking electric vehicles to the smart grid J. Lu and
J. Hossain (Editors)
Volume 81 Cyber-Physical-Social Systems and Constructs in Electric Power
Engineering S. Suryanarayanan, R. Roche and T.M. Hansen (Editors)
Volume 82 Periodic Control of Power Electronic Converters F. Blaabjerg, K. Zhou,
D. Wang and Y. Yang
Volume 86 Advances in Power System Modelling, Control and Stability Analysis
F. Milano (Editor)
Volume 87 Cogeneration: Technologies, Optimisation and Implentation
C.A. Frangopoulos (Editor)
Volume 88 Smarter Energy: from Smart Metering to the Smart Grid H. Sun,
N. Hatziargyriou, H.V. Poor, L. Carpanini and M.A. Sánchez Fornié (Editors)
Volume 89 Hydrogen Production, Separation and Purification for Energy A. Basile,
F. Dalena, J. Tong and T.N. Veziroğlu (Editors)
Volume 90 Clean Energy Microgrids S. Obara and J. Morel (Editors)
Volume 91 Fuzzy Logic Control in Energy Systems with Design Applications in
Matlab/Simulink‡ İ.H. Altaş
Volume 92 Power Quality in Future Electrical Power Systems A.F. Zobaa and
S.H.E.A. Aleem (Editors)
Volume 93 Cogeneration and District Energy Systems: Modelling, Analysis and
Optimization M.A. Rosen and S. Koohi-Fayegh
Volume 94 Introduction to the Smart Grid: Concepts, technologies and evolution
S.K. Salman
Volume 95 Communication, Control and Security Challenges for the Smart Grid
S.M. Muyeen and S. Rahman (Editors)
Volume 96 Industrial Power Systems with Distributed and Embedded Generation
R. Belu
Volume 97 Synchronized Phasor Measurements for Smart Grids M.J.B. Reddy and
D.K. Mohanta (Editors)
Volume 98 Large Scale Grid Integration of Renewable Energy Sources
A. Moreno-Munoz (Editor)
Volume 100 Modeling and Dynamic Behaviour of Hydropower Plants N. Kishor and
J. Fraile-Ardanuy (Editors)
Volume 101 Methane and Hydrogen for Energy Storage R. Carriveau and D.S-K. Ting
Volume 104 Power Transformer Condition Monitoring and Diagnosis A. Abu-Siada
(Editor)
Volume 106 Surface Passivation of Industrial Crystalline Silicon Solar Cells J. John
(Editor)
Volume 107 Bifacial Photovoltaics: Technology, applications and economics J. Libal
and R. Kopecek (Editors)
Volume 108 Fault Diagnosis of Induction Motors J. Faiz, V. Ghorbanian and G. Joksimović
Volume 109 Cooling of Rotating Electrical Machines: Fundamentals, modelling, testing
and design D. Staton, E. Chong, S. Pickering and A. Boglietti
Volume 110 High Voltage Power Network Construction K. Harker
Volume 111 Energy Storage at Different Voltage Levels: Technology, integration, and
market aspects A.F. Zobaa, P.F. Ribeiro, S.H.A. Aleem and S.N. Afifi (Editors)
Volume 112 Wireless Power Transfer: Theory, Technology and Application
N. Shinohara
Volume 114 Lightning-Induced Effects in Electrical and Telecommunication Systems
Y. Baba and V.A. Rakov
Volume 115 DC Distribution Systems and Microgrids T. Dragičević, F. Blaabjerg and
P. Wheeler
Volume 116 Modelling and Simulation of HVDC Transmission M. Han (Editor)
Volume 117 Structural Control and Fault Detection of Wind Turbine Systems
H.R. Karimi
Volume 118 Modelling and Simulation of Complex Power Systems A. Monti and
A. Benigni
Volume 119 Thermal Power Plant Control and Instrumentation: The control of boilers
and HRSGs, 2nd Edition D. Lindsley, J. Grist and D. Parker
Volume 120 Fault Diagnosis for Robust Inverter Power Drives A. Ginart (Editor)
Volume 121 Monitoring and Control using Synchrophasors in Power Systems with
Renewables I. Kamwa and C. Lu (Editors)
Volume 123 Power Systems Electromagnetic Transients Simulation, 2nd Edition
N. Watson and J. Arrillaga
Volume 124 Power Market Transformation B. Murray
Volume 125 Wind Energy Modeling and Simulation Volume 1: Atmosphere and plant
P. Veers (Editor)
Volume 126 Diagnosis and Fault Tolerance of Electrical Machines, Power Electronics
and Drives A.J.M. Cardoso
Volume 128 Characterization of Wide Bandgap Power Semiconductor Devices F. Wang,
Z. Zhang and E.A. Jones
Volume 129 Renewable Energy from the Oceans: From wave, tidal and gradient
systems to offshore wind and solar D. Coiro and T. Sant (Editors)
Volume 130 Wind and Solar Based Energy Systems for Communities R. Carriveau and
D.S-K. Ting (Editors)
Volume 131 Metaheuristic Optimization in Power Engineering J. Radosavljević
Volume 132 Power Line Communication Systems for Smart Grids I.R.S. Casella and
A. Anpalagan
Volume 134 Hydrogen Passivation and Laser Doping for Silicon Solar Cells B. Hallam
and C. Chan (Editors)
Volume 139 Variability, Scalability and Stability of Microgrids S.M. Muyeen, S.M. Islam
and F. Blaabjerg (Editors)
Volume 142 Wind Turbine System Design: Volume 1: Nacelles, drive trains and
verification J. Wenske (Editor)
Volume 143 Medium Voltage DC System Architectures B. Grainger and R.D. Doncker
(Editors)
Volume 145 Condition Monitoring of Rotating Electrical Machines P. Tavner, L. Ran and
C. Crabtree
Volume 146 Energy Storage for Power Systems, 3rd Edition A.G. Ter-Gazarian
Volume 147 Distribution Systems Analysis and Automation 2nd Edition J. Gers
Volume 151 SiC Power Module Design: Performance, robustness and reliability
A. Castellazzi and A. Irace (Editors)
Volume 152 Power Electronic Devices: Applications, failure mechanisms and
reliability F. Iannuzzo (Editor)
Volume 153 Signal Processing for Fault Detection and Diagnosis in Electric Machines
and Systems M. Benbouzid (Editor)
Volume 155 Energy Generation and Efficiency Technologies for Green Residential
Buildings D. Ting and R. Carriveau (Editors)
Volume 156 Lithium-ion Batteries Enabled by Silicon Anodes C. Ban and K. Xu (Editors)
Volume 157 Electrical Steels, 2 Volumes A. Moses, K. Jenkins, P. Anderson and
H. Stanbury
Volume 158 Advanced Dielectric Materials for Electrostatic Capacitors Q. Li (Editor)
Volume 159 Transforming the Grid Towards Fully Renewable Energy O. Probst,
S. Castellanos and R. Palacios (Editors)
Volume 160 Microgrids for Rural Areas: Research and case studies R.K. Chauhan,
K. Chauhan and S.N. Singh (Editors)
Volume 161 Artificial Intelligence for Smarter Power Systems: Fuzzy Logic and Neural
Networks M.G. Simoes
Volume 165 Digital Protection for Power Systems 2nd Edition Salman K. Salman
Volume 166 Advanced Characterization of Thin Film Solar Cells N. Haegel and
M. Al-Jassim (Editors)
Volume 167 Power Grids with Renewable Energy Storage, integration and
digitalization A.A. Sallam and B. OM P. Malik
Volume 169 Small Wind and Hydrokinetic Turbines P. Clausen, J. Whale and D. Wood
(Editors)
Volume 170 Reliability of Power Electronics Converters for Solar Photovoltaic
Applications F. Blaabjerg, A.l Haque, H. Wang, Z. Abdin Jaffery and Y. Yang
(Editors)
Volume 171 Utility-scale Wind Turbines and Wind Farms A. Vasel-Be-Hagh and
D.S.-K. Ting
Volume 172 Lighting interaction with Power Systems, 2 volumes A. Piantini (Editor)
Volume 174 Silicon Solar Cell Metallization and Module Technology T. Dullweber
(Editor)
Volume 175 n-Type Crystalline Silicon Photovoltaics: Technology, applications and
economics D. Munoz and R. Kopecek (Editors)
Volume 178 Integrated Motor Drives Xu Deng and B. Mecrow (Editors)
Volume 180 Protection of Electricity Distribution Networks, 4th Edition J. Gers and
E. Holmes
Volume 182 Surge Protection for Low Voltage Systems A. Rousseau (Editor)
Volume 184 Compressed Air Energy Storage: Types, systems and applications D. Ting
and J. Stagner
Volume 186 Synchronous Reluctance Machines: Analysis, optimization and
applications N. Bianchi, C. Babetto and G. Bacco
Volume 191 Electric Fuses: Fundamentals and new applications 4th Edition N. Nurse,
A. Wright and P.G. Newbery
Volume 193 Overhead Electric Power Lines: Theory and practice S. Chattopadhyay and
A. Das
Volume 194 Offshore Wind Power Reliability, availability and maintenance, 2nd
edition
P. Tavner
Volume 196 Cyber Security for Microgrids S. Sahoo, F. Blaajberg and T. Dragicevic
Volume 198 Battery Management Systems and Inductive Balancing A. Van den Bossche
and A. Farzan Moghaddam
Volume 199 Model Predictive Control for Microgrids: From power electronic
converters to energy management J. Hu, J.M. Guerrero and S. Islam
Volume 204 Electromagnetic Transients in Large HV Cable Networks: Modeling and
calculations Ametani, Xue, Ohno and Khalilnezhad
Volume 208 Nanogrids and Picogrids and their Integration with Electric Vehicles
S. Chattopadhyay
Volume 210 Superconducting Magnetic Energy Storage in Power Grids M.H. Ali
Volume 211 Blockchain Technology for Smart Grids: Implementation, management
and security H.L. Gururaj, K.V. Ravi, F. Flammini, H. Lin, B. Goutham, K.B.R. Sunil
and C. Sivapragash
Volume 212 Battery State Estimation: Methods and Models S. Wang
Volume 215 Industrial Demand Response: Methods, best practices, case studies, and
applications H.H. Alhelou, A. Moreno-Muñoz and P. Siano (Editors)
Volume 213 Wide Area Monitoring of Interconnected Power Systems 2nd Edition
A.R. Messina
Volume 217 Advances in Power System Modelling, Control and Stability Analysis 2nd
Edition F. Milano (Editor)
Volume 225 Fusion-Fission Hybrid Nuclear Reactors: For enhanced nuclear fuel
utilization and radioactive waste reduction W.M. Stacey
Volume 228 Digital Technologies for Solar Photovoltaic Systems: From general to
rural and remote installations S. Motahhir (Editor)
Volume 238 AI for Status Monitoring of Utility Scale Batteries S. Wang, K. Liu, Y. Wang,
D. Stroe, C. Fernandez and J.M. Guerrero
Volume 905 Power system protection, 4 volumes
Lightning
Electromagnetics
Volume 2: Electrical processes and effects
2nd Edition
Edited by
Vernon Cooray, Farhad Rachidi and Marcos Rubinstein
While the authors and publisher believe that the information and guidance given in this
work are correct, all parties must rely upon their own skill and judgement when making
use of them. Neither the author nor publisher assumes any liability to anyone for any
loss or damage caused by any error or omission in the work, whether such an error or
omission is the result of negligence or any other cause. Any and all such liability is
disclaimed.
The moral rights of the author to be identified as author of this work have been
asserted by him in accordance with the Copyright, Designs and Patents Act 1988.
Index 627
This page intentionally left blank
About the editors
We wish to thank all our colleagues who have spent a good deal of their free time
writing the chapters of this book.
We wish to express our sincere thanks to Ms. Olivia Wilkins and Ms. Nikki
Tarplett from the IET publishers and Mr. N. Srinivasan from MPS Limited for their
outstanding support throughout the publishing project. Despite unexpected delays
in our submissions, they remained patient and accommodating, always willing to
listen to our suggestions and provide valuable feedback. We are truly grateful for
their professionalism and dedication.
This page intentionally left blank
Chapter 1
Basic discharge processes in the atmosphere
Vernon Cooray1 and Liliana Arevalo2
1.1 Introduction
The main constituents of air in the Earth’s atmosphere are nitrogen (78%), oxygen
(20%), noble gases (1%), water vapour (0.03%), carbon dioxide (0.97%), and other
trace gas species. In general, air is a good insulator and it can maintain its insulating
properties until the applied electric field exceeds about 2.8 104 V/cm at standard
atmospheric conditions (i.e. T = 293 K and P = 1 atm). When the background
electric field exceeds this critical value, the free electrons in air, generated mainly
by the high energetic radiation of cosmic rays and radio active gases generated
from the Earth, start accelerating in this electric field and gain enough energy
between collisions with atoms and molecules to ionize other atoms. This cumula-
tive ionization leads to an increase in the number of electrons initiating the elec-
trical breakdown of air. The threshold electric field necessary for electrical
breakdown of air is a function of atmospheric density. For example, the critical
electric field, E, necessary for electrical breakdown of air of density d is given by
d
E ¼ E0 (1.1)
d0
where d0 is the density of air at sea level at standard atmospheric conditions and E0 is
the corresponding critical electric field necessary for electrical breakdown under the
same conditions. Since the density of air in the Earth’s atmosphere decreases with
height z (in m) as d ¼ d0 ez=lp with lp 7:64 103 m, the critical electric field
necessary to cause electrical breakdown in the atmosphere decreases with height as
E ¼ E0 ez=le (1.2)
When the electric field in the atmosphere increases beyond this critical value,
the appearance of the resulting electrical discharge depends on the pressure and the
spatial variation in the electric field. Irrespective of its apparent features the basic
constituents of an electrical discharge in air can be separated into four parts. These
1
Department of Electrical Engineering, Uppsala University, Sweden
2
Hitachi Energy Hr Dc Insulation System, Ludvika, Sweden
2 Lightning electromagnetics: Volume 2
are electron avalanches, streamer discharges, corona discharges, and leaders. When
the leaders reach an electrode of opposite polarity or a region of opposite charge
density, a rapid neutralization of the charge on the leader takes place. This neu-
tralization process is called a return stroke. The exact mechanism of the return
stroke is not yet known, but different types of models have been developed to
describe them. These models are described in several chapters of this book. Here,
we will concentrate on the four discharge processes mentioned above. Some parts
of this chapter are adopted and summarized from Ref. [1] where an extensive
description of basic physics of discharges is given.
nx ¼ eðahÞx (1.4)
This equation shows that the number of electrons increases exponentially with
distance. This exponential growth of electrons with distance is called an electron
avalanche. The equation also shows that cumulative ionization is possible only if
ða hÞ > 0. The magnitude of both a and h depends on the background electric
field and the air density. The quantity ða hÞ is known as the effective ionization
coefficient and denoted by a . For electric field values less than the critical value
necessary for electrical breakdown, ða hÞ < 0 and for higher electric fields
ða hÞ > 0. This explains the reason for the existence of a critical electric field
beyond which the air breaks down electrically.
As one can see from the above equation whether the avalanche will continue to
grow (i.e. nx continue to increase with distance) or whether it will start to decay
after an initial growth (i.e. nx will decrease with distance) depends on the spatial
distribution of the electric field. As long as the background electric field is such that
Basic discharge processes in the atmosphere 3
300 8 × 107
6 × 107
Electric field (kV/cm)
Number of electrons
200
4 × 107
100
2 × 107
0 0 × 100
0 100 200 300 400 0 100 200 300 400
Distance (µm) Distance (µm)
(a) (b)
Figure 1.1 (a) The electric field used in demonstrating the spatial variation in the
number of electrons in the avalanche head as the avalanche advances
in an electric field. (b) The variation in the number of electrons in the
avalanche head as a function of distance as the avalanche propagates
in the electric field shown in (a). Note that the origin of the avalanche
is at the point corresponding to zero distance
2 3
15
a 5:593 10 5cm2 for E 1:5 1015 V=cm2
¼ 6:619 1017 exp4 E (1.6)
N N
N
4 Lightning electromagnetics: Volume 2
h E E
¼ 8:889 105 þ 2:567 1019 cm2 for > 1:05 1015 V=cm2 (1.7)
N N N
h E E
¼ 6:089 104 2:893 1019 cm2 for 1:05 1015 V=cm2 (1.8)
N N N
As the avalanche grows the number of electrons at the head increases. These
electrons will spread out due to random diffusion causing the avalanche head to
expand. The p average
ffiffiffiffiffiffiffiffi radius of the avalanche head can be calculated from the
equation r ¼ 4Dt where t ¼ x=vd is the time of advance of the avalanche, D is the
coefficient of diffusion and vd is the drift velocity of the electrons in the electric
field. The drift velocity of the electrons can be obtained from the equation:
0:6064
16 E
vd ¼ 2:157 10 cm=s (1.9)
N
and the coefficient of diffusion D can be obtained from the following expressions [2]:
0:3441
D E E
¼ 5:645 104 V for < 2:0 1017 V=cm2 (1.10)
me N N
and
0:46113
D 7 E E
¼ 2:173 10 V 2:0 1017 < < 1:23 1016 V=cm2
me N N
(1.11)
In the above equations me is the electron mobility that can be extracted from (1.9).
generates the background electric field. First consider a source at positive polarity.
The source could be a charged graupel particle (in a thundercloud) in a background
electric field, a Franklin rod exposed to the background electric field of a thun-
dercloud or a high-voltage electrode. For clarity we will refer to it as the anode.
If the electric field in front of the anode is high enough, a photoelectron gen-
erated at a point located in front of the anode will initiate an avalanche that pro-
pagates towards the anode. The process is depicted in Figure 1.2. As the electron
avalanche propagates towards the anode, mobile positive space charge accumulates
at the avalanche head. When the avalanche reaches the anode, the electrons will be
absorbed into it leaving behind the net positive space charge. Because of the
Bazelyan and Raizer [6] suggest 109 as a reasonable value for this transfor-
mation. Thus, the condition for the transformation of an avalanche to a streamer
can be written as
Ð xc
½aðxÞhðxÞdx
e0 ¼ 108 109 (1.12)
Note that in writing down the above equation, it is assumed that the electric
field is not uniform and therefore both a and h are a function of distance. Moreover,
in the above equation the distance x is measured from the origin of the avalanche
and xc is the distance from the origin of the avalanche where the background
electric field goes below the critical value necessary for electrical breakdown.
The advancement of the streamer in a given background electric field is
facilitated by the large increase in the local electric field in the vicinity of the
streamer head and by the enhanced production of the photons from the streamer
head. The photons create secondary electrons in front of the streamer head and
these secondary electrons give rise to secondary avalanches that will move, in the
case of positive streamers, towards the streamer head. Once initiated, the streamers
have been observed to travel in background electric fields that itself cannot support
avalanche formation. Thus, the secondary avalanche formation in the streamer is
confined to a very small region around the streamer head where the electric field
exceeds 2.8 104 V/cm, the minimum electric field required for the cumulative
ionization in air at atmospheric pressure. This region is called the active region.
The dimension of the active region is about 200 mm and the streamer radius was
found to be on the order of 10–50 mm [7,8]. This value, however, may correspond
to short streamers. Since the electron multiplication in the active region is sup-
ported by the space charge electric field of the streamer head, the streamer can
propagate in electric fields that are much smaller than the critical electric field
necessary for cumulative electron ionization. In air, the background electric field
necessary for positive streamer propagation lies in the range of 4.5–5 103 V/cm
[9–11]. For negative streamers it lies in the range of 1–2 104 V/cm. Any variation
in the electron loss processes can change this electric field. For example, when air
is saturated with water vapour, the critical electric field for positive streamer pro-
pagation grows from 4.7 103 V/cm at humidity of 3 g/m3 to 5.6 103 V/cm at
18 g/cm3 [12,13]. The critical electric field necessary for streamer propagation
decreases approximately linearly with decreasing air density [6].
In background electric fields close to the critical value necessary for streamer
propagation, the speed of streamers is about 107 cm/s. However, the streamer speed
increases with increasing background electric field. No direct measurements are
available today on the potential gradient of the streamer channels. Experiments
conducted with long sparks show that the average potential gradient of the elec-
trode gap when the positive streamers bridges the gap between the two electrodes is
about 5 103 V/cm [14]. This indicates that the potential gradient of the positive
streamer channels in air at atmospheric pressure is close to this value. Note that this
value is approximately the same as the critical electric field necessary for the
propagation of positive streamers.
Basic discharge processes in the atmosphere 9
positive ions with the cathode gives out electrons that support the formation of
avalanches. The electron avalanches will move out a certain distance from the
cathode, but as they move away from the cathode the electric field decreases and
the electrons will be attached to oxygen molecules in air. This gives rise to a
negative space charge region. This negative space charge will reduce the electric
field and choke off the discharge activity. However, as the negative space charge
moves away due to the action of the electric field, the electric field at the cathode
recovers and a new discharge activity starts from the cathode. This mode of dis-
charge, relaxation and continuation, continues for a considerable potential range.
This oscillating corona discharge is called Trichel pulses because the current con-
sists of pulses separated in time. The repetition frequency of the pulses depends on
the applied voltage but can reach values about 106 s1.
As the electric field continues to increase, the negative ions are created too far
from the cathode to choke off the discharge activity and a pulseless discharge
activity, i.e. a glow discharge, starts from the cathode. This will continue for some
range of potentials and with increasing electric fields negative streamers start to
form from the cathode. Initially, they do not propagate far into the gap but with
increasing voltage they move further and further into the gap.
decreases, the gas temperature increases and very quickly all the components of the
discharge namely, electrons, ions and neutrals, will achieve the same temperature and
the discharge will reach local thermodynamic equilibrium.
produce more heating and accelerate the ionization process. With increasing ioni-
zation, the process of thermalization sets in transforming the stem into a hot and
conducting channel called the leader.
Owing to its high conductivity, most of the voltage of the anode will be
transferred to the head of the leader channel resulting in a high electric field there.
This high electric field leads to the production of streamer discharges now from a
common stem located at the head of the leader channel. With the aid of cumulative
streamer currents, the new stem gradually transforms itself to a newly created
leader section with the streamer process now repeating at the new leader head. The
streamer system located in front of the leader is the source of current that heats the
air and makes possible the elongation of the leader. The main sequences of the
propagation of a positive leader are shown in Figure 1.4.
T1 T2 T3 T4 T5
Streamer bursts
Figure 1.4 Mechanism of positive leaders. When the electric field at the surface
and in the vicinity of the anode increases to a value large enough to
convert avalanches to streamers, a bust of streamers is generated from
the anode (T1). Many of these streamers have their origin in a
common channel called the streamer stem. The combined current of
all streamers flowing through the stem causes this common region to
heat up and, as a result, the stem will be transformed into a hot and
conducting channel called the leader (T2). Owing to its high
conductivity, most of the voltage of the anode will be transferred to the
head of the leader channel resulting in a high electric field there. This
high electric field leads to the production of streamer discharges now
from a common stem located at the head of the leader channel (T2).
With the aid of cumulative streamer currents, the new stem gradually
transforms itself to a newly created leader section with the streamer
process now repeating at the new leader head (T3, T4, T5)
Basic discharge processes in the atmosphere 13
T1 T2 T3 T4 T5 T6 T7 T8
Negative
Streamer leader channel
bursts
New
leader step
Space stem
Space leader
connecting leader (a positive leader discharge) from the Franklin rod. A brief
description of how this could be achieved is given below. The description is based
on the work published previously by Becerra and Cooray [19–21].
Assume that the electric field at ground level as a function of time generated by
the down coming stepped leader is known. This can be calculated, e.g. by using the
leader charge distribution as extracted by Cooray et al. [22]. The simulation con-
sists of several main steps and let us take them one by one.
a) The first step is to extract the time or the height of the stepped leader when
streamers are incepted from the grounded rod. Since the background electric field
is known, the electric field at the tip of the grounded rod can be calculated, e.g. by
using charge simulation method. This field is used together with the avalanche to
streamer conversion criterion given in (1.12) to investigate whether the electric
field at the conductor tip is large enough to convert avalanches to streamers. The
simulation continues using the time-varying electric field of the stepped leader
until the streamer inception criterion is satisfied.
b) The moment the streamer inception criterion is satisfied a burst of streamers
will be generated by the tip of the rod. The next task is to calculate the charge
in this streamer burst. The charge associated with the streamer burst is calcu-
lated using a distance–voltage diagram with the origin at the tip of the groun-
ded conductor as follows (see Figure 1.6). The streamer zone is assumed to
maintain a constant potential gradient Estr. In the distance–voltage diagram,
this is represented by a straight line (note that the point of zero distance cor-
responds to the tip of the conductor). On the same diagram the background
potential produced by the thundercloud and the down-coming stepped leader
(taking also into account the presence of the lightning conductor) at the current
time is depicted. If the area between the two curves up to the point where they
cross is A, the charge in the streamer zone is given by
Q0 KQ A (1.13)
where KQ is a geometrical factor. Becerra and Cooray [20] estimated its value
to be about 3.5 10-11 C/V m.
c) The next task is to investigate whether this streamer burst is capable of gen-
erating a leader. This decision is based on the fact that in order to generate a
leader a minimum of 1 mC is required in the charge generated by the streamers.
If the charge in the streamer zone is less than this value, then the procedure is
repeated after a small time interval. Note that with increasing time the electric
field generated by the stepped leader increases and, consequently, the charge in
the streamer bursts increases. Thus at a certain time the condition necessary for
the leader inception will be fulfilled.
d) Assume that at time t, the condition necessary for leader inception is satisfied. The
next task is to estimate the length and the radius of this initial leader section. In
doing this it is assumed that the amount of charge you need to create a unit length
of positive leader is ql. The value of ql is about 40–65 mC/m. With this the initial
length of the leader section L1 is given by Q0/ql. It is important to point out here
16 Lightning electromagnetics: Volume 2
that Becerra and Cooray [21] utilized a more rigorous condition in which the
charge necessary to thermalize a unit leader section depends on the speed of
the leader. The initial radius of the leader, aL (t), is assumed to be 103 m and the
initial potential gradient of the leader section, EL1 ðtÞ, is assumed to be equal to
the potential gradient of the streamer region, i.e. 5.0 103 V/cm. Now we proceed
to the next time step, i.e. t ¼ t þ Dt.
e) During the time interval Dt the background potential is changed and we also have
a small leader section of length L1. Now the new charge in the streamer zone
generated from the head of the new leader section is calculated as before but now
including both the leader and its streamer zone in the distance–voltage diagram.
The leader is represented by a line with a potential gradient EL1 ðtÞ (see
Figure 1.6). The total charge is calculated from the area between this new curve
and the background potential. The charge generated in the current time step is
obtained by subtracting from this the charge obtained in the previous time
step. Let the charge obtained thus be Q1. This charge is used to evaluate the
length of the new leader section L2. Moreover, the flow of this charge through the
Background
potential
Potential
Streamer potential
gradient
Distance
(a)
Background
potential
Potential
Streamer potential
gradient
Distance
(b)
leader channel changes the potential gradient and the radius of the older leader
section L1. The new potential gradient and the radius of L1 are given by
EL1 ðt þ DtÞ and aL1 ðt þ DtÞ.
Now let us consider the nth time step. There are n leader sections and they have
there respective potential gradients and radii. The radius and the potential
gradient of ith leader section are obtained from
g1
p a2Li ðt þ DtÞ ¼ p a2Li ðtÞ þ EL ðtÞ ILi ðtÞ Dt (1.14)
g p0 i
a2Li ðtÞ
ELi ðt þ DtÞ ¼ E L ðt Þ (1.15)
aLi ðt þ DtÞ i
2
In the above equation ELi ðtÞ is the internal electric field and ILi ðtÞ is the current of
the leader section Li at time t. With these, it is possible to calculate the time evo-
lution of the internal electric field for each segment and the potential drop along the
Q3
Q2
Streamer L4
zone Q1
L3 L3
Q0
L2 L2 L2
L1 L1 L1 L1
Streamer
stem
Grounded
structure
leader channel (at a given time, with k denoting the number of leader sections) as
follows:
X
k
DUL ¼ ELi ðtÞ Li (1.16)
i¼1
The steps described above can be used to simulate the inception and propa-
gation of positive leaders. Figure 1.7 describes the basics of the process schema-
tically. The calculation can be simplified if, instead of calculating the time
evolution of leader potential gradient in each segment as above, one uses the
expression derived by Rizk [23] for the potential of the tip of the leader channel
that is given by
the inception condition of the leader is attained if the gas temperature reaches a
critical temperature of 1,500 K in the streamer stem which was considered as a
cylindrical plasma channel. Based on this study, they concluded that
(i) Photoionization is essential for the development and propagation of positive
streamers. It enables the seed electrons in the high electric field region at the
head of the streamer.
(ii) The most important mechanisms to increase the electron density, and con-
sequently incept a leader, were the fast electron detachment from negative
ions caused by oxygen atoms and the acceleration of the electron impact
ionization due to NO molecules.
(iii) The rise of temperature on the leader depends directly on the energy avail-
able in the streamer channels, the vibrational energy relaxation, and the
recombination of particles. Consequently, it is incorrect to assume that all
streamer energy is directly used for heating and a unique amount of charge is
required to heat the channel and incept a leader.
(iv) Calculations indicate that sharp tips allow more charge to flow into the
streamer channels before leader inception takes place than blunt tips.
Therefore, the amount of electrical charge required to achieve leader incep-
tion depends on the electric field distribution of the electrode arrangement
i.e., geometry, applied voltage, the space charge, and environmental condi-
tions, among others.
(v) The study shows that the criterion of a constant minimum electrical charge of
1 mC to incept a leader channel, used in lightning attachment and long gap
discharge models, is not well-founded. Even though the critical charge
necessary for leader inception for rods of 1 cm radius is about 1 mC the value
of this critical charge decreases as the radius of the conductor increases.
References
[1] Cooray, V., 2003. Mechanism of electrical discharges, in Cooray, V., (Ed.),
The Lightning Flash, The Institution of Electrical Engineers, London, UK.
[2] Morrow, R., 1985. Theory of negative corona in oxygen, Phys. Rev. A, 32,
1799–1809.
[3] Morrow, R. and Lowke, J. J., 1997. Streamer propagation in air, J. Phys. D:
Appl. Phys., 30, 614–627.
[4] Raether, H., 1940. Zur Entwicklung von Kanalentladungen, Arch.
Eleektrotech., 34, 49–56.
[5] Meek, J.M., 1940. A theory of spark discharge, Phys. Rev., 57, 722–728.
[6] Bazelyan, E.M. and Raizer, Yu.P., 1998. Spark Discharge, CRC Press, New
York.
[7] Marode, E., 1983. The glow to arc transition, in Kunhardt, E. and Larssen, L.
(Eds.), Electrical Breakdown and Discharges in Gases, Plenum Press, New
York.
20 Lightning electromagnetics: Volume 2
[8] Marode, E., 1975. The mechanism of spark breakdown in air at atmospheric
pressure between a positive point and a plane, I. Experimental: Nature of the
streamer track, II. Theoretical: Computer simulation of the streamer track, J.
Appl. Phys., 46, 2005–2020.
[9] Les Renardie´res Group, 1977. Positive discharges in long air gaps – 1975
results and conclusions, Electra, 53, 31–152.
[10] Les Renardie´res Group, 1981. Negative discharges in long air gaps, Electra,
74, 67–216.
[11] Gao, L., Larsson, A., Cooray, V. and Scuka, V., 1999. Simulation of strea-
mer discharges as finitely conducting channels, IEEE Trans. Dielectr.
Electr. Insul., 6, 1, 35–42.
[12] Griffiths, R.F., and Phelps, C.T., 1976. The effects of air pressure and water
vapour content on the propagation of positive corona streamers, Quart. J.R.
Mat. Soc., 102, 419–426.
[13] Griffiths, R.F., and Phelps, C.T., 1976. Dependence of positive corona
streamer propagation on air pressure and water vapour content, J. Appl.
Phys., 47, 2929–2934.
[14] Paris, L. and Cortina, R., 1968. Switching and lightning impulse discharge
characteristics of large air gaps and long insulator strings, IEEE Trans., PAS-
98, pp. 947–957.
[15] Gallimberti, I., 1979. The mechanism of long spark formation, J. de
Physique., 40, 7, C7–193–250.
[16] Biagi, C.J., Uman, M.A., Hill, J.D., Jordan, D.M., Rakov, V.A. and Dwyer, J.,
2010. Observations of stepping mechanisms in a rocket-and-wire triggered
lightning flash, J. Geophys. Res., 115, D23215, doi:10.1029/2010JD014616.
[17] Mazur, V., Ruhnke, L., Bondiou-Clergerie, A. and Lalande, P., 2000.
Computer simulation of a downward negative stepped leader and its inter-
action with a grounded structure, J. Geophys. Res., 105, D17, 22361–22369.
[18] Arevalo, L. and Cooray, V., 2011. Preliminary study on the modeling of
negative leader discharges, J. Phys. D: Appl. Phys., 44, 31, doi:10.1088/
0022-3727/44/31/315204.
[19] Becerra, M. and Cooray, V., 2006. A self-consistent upward leader propa-
gation model, J. Phys. D: Appl. Phys., 39, 3708–3715.
[20] Becerra, M. and Cooray, V., 2006. A simplified physical model to determine
the lightning upward connecting leader inception, IEEE Trans. Power
Delivery, 21, 2, 897–908.
[21] Becerra, M. and Cooray, V., 2006. Time dependent evaluation of the lightning
upward connecting leader inception, J. Phys. D: Appl. Phys., 39, 4695–4702.
[22] Cooray, V., Rakov, V. and Theethayi, N., 2007. The lightning striking dis-
tance—Revisited, J. Electrostat., 65, 296–306.
[23] Rizk, F., 1989. A model for switching impulse leader inception and break-
down of long air-gaps, IEEE Trans. Power Delivery, 4, 1, 596–603.
[24] Gallimberti, I., Bacchiega, G., Bondiou-Clergerie, A., and Lalande P., 2002.
Fundamental processes in long air gap discharges, C.R. Physique appliquée/
Appl. Phys., 3, 1335–1359.
Basic discharge processes in the atmosphere 21
[25] Bondiou, A. and Gallimberti, I., 1994. Theoretical modeling of the devel-
opment of the positive spark in long air gaps, Phys. D Appl. Phys., 27(6),
1252–1266.
[26] Lalande, P., 1996. Etude des conditions de foudroiement d’une structure au
soil, Ph.D. thesis, Universite de Paris-Sud U.F.R. Scientifique d’Orsay.
[27] Goelian, N., Lalande, P., Bondiou-Clergerie, A., Bacchiega, G.L., A.
Gazzani, I. Gallimberti, 1997. A simplified model for the simulation of
positive-spark development in long air gaps, J. Phys. D: Appl. Phys., 30,
2441–2452.
[28] Becerra, M. and Cooray, V., 2006, Time dependent evaluation of the light-
ning upward connecting leader inception, J. Phys. D: Appl. Phys., 39,
pp. 4695–4702.
[29] Popov, N., 2009, Study of the formation and propagation of a leader channel
in air, Plasma Phys. Rep., 35, pp. 785–793.
[30] Fofana, I. and Beroual, A., 1997, A predictive model of the positive dis-
charge in long air gaps under pure and oscillating impulse shapes, J. Phys.
D: Appl. Phys., 30, 1653–1667.
[31] Arevalo, L. and Cooray, V., 2017, Unstable leader inception criteria of
atmospheric discharges, Atmosphere, 8(9), 156.
This page intentionally left blank
Chapter 2
Modelling of charging processes in clouds
Edward R. Mansell1 and Donald R. MacGorman1
2.1 Introduction
Our goal in this chapter is to introduce the treatments of electrical processes used by
numerical cloud models that integrate dynamic, microphysical, thermodynamic and
electric processes to track what happens to several classes of water particles as they
move through the cloud and interact with each other and with the environment as the
cloud evolves. We consider primarily models that treat ice particles, as well as liquid
water, because ice is now commonly recognized as a necessary ingredient for strong
electrification. We ignore models whose winds or particle spectra are unchanging and
models that treat electrification as interactions of electric circuit elements driven by a
current, charge or voltage source unrelated to microphysics and dynamics.
When using cloud models, it is prudent to keep their limitations in mind, as no
numerical model can reasonably be expected to replicate a storm exactly. Our
computer resources and our knowledge of many relevant processes and of the
environmental state are almost certainly never adequate to do that. What is done,
therefore, is to use simplified mathematical descriptions called parameterizations to
deal with the spectrum of particle types, to estimate the effects of poorly understood
physics and to incorporate the effects of processes that occur on temporal or spatial
scales too small to be included directly. The goal is for the parameterizations to
incorporate enough physics, to treat a broad enough range of time and distance scales
and to retain enough detail in the microphysical, electric, dynamic and thermo-
dynamic fields that the model can simulate accurately the aspect of the phenomenon
we are studying. To some extent, success is judged by examining how well the model
simulates observations of related storm properties, such as the distribution and evo-
lution of precipitation, cloud particles and electric field magnitudes. However, it
often is difficult to assess how well a particular model has succeeded in simulating a
given phenomenon. Relevant properties often are not observed well enough to judge
simulations, and even if they have been observed well, it is often difficult to deter-
mine how good a match with observations is needed to establish that the model is
adequately simulating the targeted behaviour.
1
National Severe Storms Laboratory, USA
24 Lightning electromagnetics: Volume 2
Although any model involves uncertainties, models also have strengths that
help circumvent the serious difficulties one faces trying to improve understanding
of storm electrification by analysing measurements alone. A sensor may itself
distort what is being measured, and ambiguities can arise when interpreting mea-
surements by a particular sensor in terms of desired storm parameters. Furthermore,
no combination of technologies is likely ever to be able to observe simultaneously
all of the thermodynamic, kinematic, electric and hydrometeor fields in evolving
clouds with enough temporal and spatial resolution to describe all their significant
behaviours and interactions. Modelling addresses these shortcomings by attempting
to calculate all of the relevant fields in a physically consistent way from as close to
first principles as is feasible. By providing a complete set of simulated observations
of the very complex system that is a thunderstorm, modelling provides a useful
means for testing the plausibility and implications of ideas gained from theory and
observations.
Thus, observations and models often complement each other in our efforts to
advance understanding: (1) laboratory, theory and field observations provide the
knowledge needed to build a model. (2) Modelling experiments provide insight into
storm processes and lead to predictions and model sensitivities that can be tested
with observations. (3) New laboratory and field observations are acquired under
model guidance to examine the predictions and sensitivities and refine the model.
(4) The refined model is then used either to probe previously defined issues more
deeply or to begin investigating new issues not possible to address with the pre-
vious version of the model. Note that progress in model capabilities is also often a
result of increased computer memory and processing speed.
In this chapter, we limit ourselves to explaining the basics of the techniques
involved in modelling electrification, as well as giving an overview of some of the
electrification research pursued through modelling studies. We attempt to cover
most electrical processes but present only selected examples of the treatments of
each one. Beyond defining some terms, we do not review the microphysical and
dynamical frameworks of cloud models, because doing so would make this chapter
far too lengthy. An overview of microphysical and dynamical treatments is given
by MacGorman and Rust (1998). More in-depth information about cloud models is
available in several books, including Pruppacher and Klett (1997), Cotton and
Anthes (1989), Houze (1993), Stensrud (2007), Straka (2009) and in many of the
publications referenced in this chapter.
must have an ambient environment defined by soundings but has no equation for
dynamics to govern the development of the wind field. Instead, the model ingests a
cloud wind field obtained from Doppler radars. (The wind field either can be a
single wind field, assumed to be steady state, or can vary in time from an early
stage of the storm.) It then uses the thermodynamic equation of state and continuity
equations for water substance to retrieve temperature, water vapour mixing ratio
and hydrometeor mixing ratios throughout the model domain.
Models also are classified by the number of spatial dimensions of the model
grid. To study storms in which three-dimensional structure and circulations are
important, it obviously is necessary to use a three-dimensional model. However, for
simpler situations or more limited investigations, modellers can reduce the number
of spatial dimensions that they use. Because adding a spatial dimension to a model
typically increases computer memory and storage requirements by more than an
order of magnitude, reducing the number of dimensions greatly reduces the com-
puter resources required to run a model.
Some electrification models are essentially zero-dimensional cloud models
because most parameters do not vary with either height or horizontal distance.
These models normally assume that the upper and lower boundaries are infinite
horizontal planes, so that the ambient electric field is constant with the distance
between them. Microphysics and vertical winds also are kept uniform between the
horizontal planes: on each boundary there is a source of hydrometeors, usually a
source of small particles at the bottom and a source of large particles at the
top. Particles and charge can vary with time, and charges are collected at the
boundaries as particles reach them, thereby changing the ambient electric field.
The model domain normally is considered to represent only part of a cloud, with
the upper plate corresponding to the centre of the upper positive charge, and the
lower plate the centre of the lower negative charge in a thunderstorm charge dis-
tribution. Illingworth and Latham (1977) pointed out that a model with infinite
planes will overestimate thunderstorm electric field magnitudes in most situations.
In a one-dimensional model, height is the only spatial coordinate that is
retained. Cloud properties can vary with height but not with horizontal position.
The model domain normally is defined as a cylinder whose radius R(z) is specified.
The use of a finite horizontal extent makes it possible to incorporate para-
meterizations of entrainment and turbulent eddy fluxes and makes electric field
magnitudes more realistic.
In a two-dimensional model, height and one horizontal coordinate are retained,
and cloud properties over the remaining horizontal coordinate are constant. Two-
dimensional models can be either slab-symmetric or axisymmetric. A slab-
symmetric model uses Cartesian coordinates and keeps cloud properties constant
along x or y. This symmetry is applied sometimes to squall lines. An axisymmetric
model uses cylindrical coordinates and sets azimuthal variations to zero. This
symmetry is useful for small thunderstorms and some aspects of hurricanes. Two-
dimensional models have a particular limitation, however, in that turbulence
incorrectly feeds upscale growth. Proper treatment of turbulence requires a three-
dimensional model.
26 Lightning electromagnetics: Volume 2
dqN
¼ TransportðqN Þ þ SourceðqN Þ þ SinkðqN Þ (2.1)
dt
where transport is the net transport of qN into the volume by advection, turbulence
and diffusion; source is the sum of all sources of qN in the volume and sink is the
sum of all loss mechanisms for qN. Much of the challenge in parameterization is to
develop physically realistic expressions for sources and sinks, once a partition of
water into various categories is chosen.
The water substance in a cloud normally is partitioned into several categories.
For example, the following categories are described by Houze (1993):
1. Water vapour (qv) is water in the gaseous phase.
2. Cloud liquid water (qc) consists of liquid droplets that are too small to have
appreciable terminal fall speed (droplet radius less than roughly 100 mm).
3. Precipitation liquid water consists of liquid drops that are large enough to have
an appreciable terminal fall speed. This category sometimes is divided by
terminal fall speed into drizzle (qDr) (radius roughly 0.1–0.25 mm) and rain
(qr) (radius > 0.25 mm).
4. Cloud ice (qi) indicates ice particles that are too small to have an appreciable
terminal fall speed.
5. Precipitation ice consists of ice particles that have a terminal fall speed of
0.3 m/s. This category often is subdivided by density and fall speed. For
example, snow (qs) has lower density and fall speeds of 0.3–1.5 m/s, graupel
(qg) is denser and falls at 1–10 m/s and hail (qh) is larger and still denser,
with fall speeds up to 50 m/s.
Modelling of charging processes in clouds 27
Models that omit all forms of ice are referred to as warm cloud models. Models
that include equations for the ice phase are referred to as cold cloud models or
mixed-phase models.
Each category of water substance interacts with the other categories to create
sources and sinks. For example, cloud water droplets coalesce to form drizzle, a
process that is a source for drizzle and a sink for cloud liquid water. There are
several types of basic interactions:
1. Condensation or deposition of water vapour onto cloud nuclei (called nuclea-
tion) to form cloud liquid water droplets and cloud ice particles, respectively,
having size spectra characteristic of the model cloud’s environmental
conditions
2. Hydrometeor growth through vapour condensation or deposition
3. Collection of particles to form larger particles (Collection of the various types of
particles consists of collisions followed by sticking together. The ways in which two
particles stick together are labelled by specific terms depending on the types of
particles involved: coalescence involves two or more liquid water particles; aggre-
gation involves two or more ice particles; and riming is the process of cloud water
droplets freezing into ice particles on contact. Accretion is used broadly to indicate
collection of liquid particles by ice particles but often has a connotation of larger
liquid particles sticking to larger ice particles.)
4. Breakup of drops or splintering of ice (These processes can increase the
number of particles beyond what would be expected from nucleation. The
increase from splintering is referred to as ice multiplication.)
5. Freezing of liquid water (Liquid water exists as supercooled water at heights
above the 0 C isotherm, but all liquid water usually is assumed to have frozen
by the time a parcel reaches the 40 C isotherm.)
6. Evaporation or sublimation of water vapour from hydrometeors
7. Melting of ice
8. Precipitation reaching the ground
It is possible to expand the number of categories and subcategories of
hydrometeors considerably. Typical categories are cloud droplets (D < 50 mm),
rain (D > 50 mm), small ice crystals, snow (or aggregated crystals), graupel and
hail. To expand categories, e.g. a modeller might want to track different shapes
(called habits) of ice particles separately (such as columns, plates and dendrites).
However, as more categories of water are used, the number of interactions that
must be considered increases rapidly. For that reason, modellers usually use only
the categories and subcategories that are essential to simulating the particular
phenomenon being studied.
To parameterize microphysics, models handle the size distributions of the
various categories in one of two ways, referred to as bulk microphysics and bin (or
spectral) microphysics. In bulk microphysics, the size distribution of particles is
described by some simple function. The amount in a given category can be tracked
at each grid point by a single parameter, such as the mixing ratio of water substance
in that category, or by two or more parameters, such as the mixing ratio and number
28 Lightning electromagnetics: Volume 2
where n(D) is the number of particles per unit volume between D and D + dD and a
is the shape parameter, which controls the width of the spectrum. The intercept
parameter n0 is a function of the total number concentration nT and the slope
parameter l:
nT
n0 ¼ laþ1 (2.3)
Gða þ 1Þ
where l is defined as
1=d
Gða þ 1 þ dÞcnT
l¼ (2.4)
Gða þ 1Þrair q
where the constants c and d come from the mass–diameter relationship m(D) = cDd.
The moments M(j) of the distribution are given by
ð1
MðjÞ ¼ AðjÞ Dj nðDÞdD (2.5)
0
where A(j) is some constant factor. For the simplified gamma distribution, the
moments are
AðjÞnT Gða þ 1 þ jÞ
MðjÞ ¼ (2.6)
lj Gða þ 1Þ
For j = 0 and A(0) = 1, M(0) = nT, so the zeroth moment is the number con-
centration. If the mass of a particle is m(D) = cD3, then the third moment gives the
total mass of the distribution.
Single-moment schemes predict q and usually diagnose nT by setting assuming
value for n0 (the intercept parameter). Models with two-moment bulk microphysics
generally predict the mixing ratio q (third moment) and number concentration nT
(zeroth moment). Three-moment schemes can also diagnose the shape parameter a.
Setting the shape parameter a equal to zero results in the inverse exponential (IE)
distribution. This form of the size distribution is often referred to as a Marshall–
Palmer distribution (Marshall and Palmer, 1948).
In bulk microphysics, each property of the category has a single representative
value at a given grid point in the model domain that somehow averages the values
for all of the category’s particles within the corresponding volume. For example,
consider the sedimentation of liquid precipitation particles that have a terminal fall
Modelling of charging processes in clouds 29
speed relationship v(D) = aDb. The moment-weighted fall speed is used for sedi-
mentation of each bulk moment. For example, the mass-weighted fall speed VT,m is
given by
Ð1
vðDÞmðDÞnðDÞdD
VT;m ¼ 0 Ð 1 (2.7)
0 mðDÞnðDÞdD
which is used for sedimentation of the total particle mass in a grid volume. Note
that the denominator in (2.7) is the mass content (mass of condensate per volume).
Then C, the net rate of change in a grid point’s mass mixing ratio q due to sedi-
mentation, is given by
d
C¼ ðVT ;m qÞ (2.8)
dz
Moments of the distribution in a spectral bin model are calculated by an
explicit summation over the range of sizes for the water substance category being
considered. For example, the mixing ratio qx of particle type x, modelled by J size
bins and having mass increments of mx(j), is given by
1 X J
qx ¼ mx ðjÞnx ðjÞ (2.9)
rair j¼1
where nx(j) is the number density of particles in mass bin j. To represent the size
distribution of particles accurately, at least ten size bins must be included for each
hydrometeor category being parameterized.
With bin microphysics, it is unnecessary to assume that the size distribution
has a particular form. The overall size distribution is allowed to evolve naturally as
particles of different sizes gain and lose mass through their interactions. This is a
more direct approach than bulk microphysics, but even with many size bins, a bin
parameterization is not an exact treatment. There are gaps in our knowledge of
particle interactions, particularly for some size ranges and particle types, and these
gaps introduce uncertainties in bin parameterizations. Furthermore, although par-
ticles within a given size bin are more uniform than in a category that spans all
sizes in a bulk treatment, variations in properties, such as size, shape or density may
occur within an individual bin being treated explicitly and can affect interactions
involving this bin. When such variations are significant, spectral parameterizations
need to avoid treating all particles in a size bin identically, because it would
incorrectly cause all particles to transfer into a new size bin at the same time. To
avoid this, a spectral parameterization can treat particles in a given size bin as
having a distribution of sizes across the range of the bin and can use statistical
techniques to govern gradual changes in the population.
The main disadvantage of bin microphysical treatments compared with bulk
microphysics is that bin treatments require considerably more computational
resources: Each bin must satisfy its own continuity equation, and interactions must
be included between size bins, as well as between water substance categories.
30 Lightning electromagnetics: Volume 2
Thus, bin microphysics usually is reserved for addressing questions that cannot be
addressed by bulk microphysics. When specific microphysical treatments are used
in the rest of this chapter, bulk microphysics is used instead of bin microphysics.
+ θ + +
+++ +++
E
–
+
effects, contrasting surface potentials and electrical double layers arising from
physical properties of ice surfaces.
form of convergence in some region at low levels on which are superimposed thermal
bubbles, whose size, location and temperature excess above ambient conditions must
be chosen (either given specific values or chosen randomly from some range). Storms
that result from a given model experiment can be influenced by the form and mag-
nitude of forcing that is applied, as well as by assumed environmental conditions; this
is especially true of small thunderstorms driven primarily by local heating.
Besides full simulation models, which simulate the dynamics, thermodynamics
and microphysics of storms simultaneously beginning with initial forcing in a pre-
storm atmospheric environment, another type of model, called a kinematic model,
use observed wind fields of storms to estimate the accompanying temperature and
water vapour perturbations and microphysics in a particular atmospheric environ-
ment. In a kinematic model (e.g. Ziegler, 1985, 1988), the thermodynamic equation
of state and continuity equations for water substance are the same as in a full
simulation model, but there is no equation for dynamics to govern the development
of the wind field. Instead, the model ingests an observed cloud wind field from
Doppler radars and ambient environmental conditions from an atmospheric
sounding. Then, the equation of state and continuity equations for water substance
are solved to retrieve temperature, water vapour mixing ratio and hydrometeor
mixing ratios throughout the model domain. In a time-dependent, kinematic model,
the various fields are allowed to evolve from their state at an early stage of the
cloud. Changes in the wind field are calculated by interpolating between observed
wind fields at each time step. Changes in the retrieved fields then are calculated
from changes in the wind fields.
Kinematic models have the disadvantage that, because the modelled cloud
does not begin with initial cloud formation, there may be significant errors in fields
that are sensitive to the cloud’s history of evolution. Furthermore, the wind field of
the model is not influenced by microphysics and thermodynamics, and kinematic
models can be used only for periods when Doppler wind fields are available.
However, a kinematic model has the advantage that it produces a model storm
consistent with both the observed storm wind field and the complete set of ther-
modynamic and continuity equations. It can be difficult to use full simulation
models to investigate some storms, especially small storms, because of model
sensitivities to the form of initial forcing, to boundary conditions or to minor
changes in the storm environment.
Most cloud models that have been used in electrification studies have tracked
the properties of water substance categories at grid points fixed with respect to the
Earth (these are sometimes called Eulerian models). Particle tracing or Lagrangian
models (e.g. Kuettner et al., 1981), however, compute changes to the properties of
individual particles or groups of identical particles along trajectories that follow the
particles through the storm, instead of at fixed grid points. Model fields related to
the particles are derived after each model time step by interpolating particle para-
meters from particle locations to the model grid. This type of treatment typically is
used with simplifications to other aspects of the model to keep computations
tractable. Sometimes an Eulerian model uses statistical techniques to mimic a
Lagrangian treatment for a particular type of particle.
Modelling of charging processes in clouds 33
radar-derived wind fields that were available every three minutes, from early in the
storm’s lifetime throughout much of its life. They also added parameterizations of
the inductive charging mechanism and of screening layer charge on the cloud
boundary. Ziegler and MacGorman (1994) used the three-dimensional model to
study a supercell storm (features that are not slab- or axisymmetric are critical to an
adequate treatment of supercell storms).
Norville et al. (1991) developed a kinematic model that used bin microphysics.
The only electrification mechanism that they included was a non-inductive char-
ging mechanism based on the laboratory work of Jayaratne et al. (1983), Baker
et al. (1987) and Keith and Saunders (1989). The geometry of their model cloud
consisted of two concentric cylinders, in order to be able to model coexisting
updrafts and downdrafts. Conditions were horizontally uniform inside the inner
cylinder and different, but again uniform, between the inner and outer cylinders.
This configuration has been called a one-and-one-half dimensional model. Norville
et al. simulated the same storm studied by Helsdon and Farley (1987b). Their
model was able to produce electric field magnitudes comparable to observed values
by using only the non-inductive mechanism.
Most modelling studies prior to the 1990s, whether they used simulation or
kinematic models, have examined electrification only in the absence of lightning.
Without a lightning parameterization, models could simulate only the initial elec-
trification of thunderstorms, because lightning modifies the charge distribution and
limits the maximum magnitude of the electric field. Rawlins (1982) used a
threshold electric field of 500 kV m1 to initiate a lightning flash and a simple
charge neutralization scheme. He found that the electric field regenerated quickly
after a lightning flash. Helsdon et al. (1992) developed a detailed two-dimensional
unbranched lightning flash parameterization that calculated the neutralized charge
from the ambient electric field. It was used successfully to simulate a storm that
produced a single lightning flash, and later was expanded to a three-dimensional
unbranched channel (Helsdon et al., 2002). Takahashi (1987) incorporated a simple
parameterization of a lightning flash in his model to examine factors influencing
the height and location of lightning. Ziegler and MacGorman (1994) developed a
simple three-dimensional parameterization of the net effect of several flashes per
time step to simulate a supercell storm. Baker et al. (1995) added a simple para-
meterization of lightning to their one-and-one-half dimensional, kinematic model
with bin microphysics and Solomon and Baker (1996) developed an analytical
treatment of a vertical, one-dimensional lightning channel for the same model.
MacGorman et al. (2001) developed a three-dimenstional treatment to simulate
discreet flashes using the crude parameterization of neutralized charge by Ziegler
and MacGorman (1994). Mansell et al. (2002) developed a detailed three-
dimensional fractal lightning scheme that used Gauss’s law to estimate the charge
neutralized by branched equipotential lightning channels. These last two lightning
parameterizations are described in detail in Section 2.5.
The two-dimensional simulation study of Helsdon et al. (2001) broke ground
in comparing results from multiple non-inductive graupel–ice schemes. Helsdon
et al. (2001) did not include lightning, however, so results were obtained only up to
36 Lightning electromagnetics: Volume 2
the time of the first lightning flash. Helsdon et al. (2002), using a three-dimensional
version of their model and a lightning parameterization, tested the so-called ‘con-
vective’ charging hypothesis, which relies on screening layers formed by ion
attachment to be advected into the storm updraft to generate electrification via
inductive charging. Helsdon et al. (2002) found only weak, disorganized elec-
trification that could not generate electric fields sufficient for lightning without
inclusion of a non-inductive graupel–ice mechanism. A study by Mansell et al.
(2005) investigated differences in electrification from multiple graupel–ice para-
meterizations as well as inductive graupel–droplet charge separation in a three-
dimensional model with lightning.
Sun et al. (2002) employed a three-dimensional model with two-moment
microphysics (mass and number concentration) to examine the effects of elec-
trification on the microphysics and dynamics of a small storm. Feedback to
dynamics was achieved through an electrical drag term in the momentum equations
using the net charge density and electric field and by electric force adjustments to
hydrometeor terminal speeds. The results with electrification and feedback differed
appreciably from the non-electrical simulation. An increase in latent heating was
found when electrical forces were enabled. The increased latent heating was
attributed to reductions in graupel fall speeds at mid-levels of the storm, which
increased its residence time and total riming growth. The model almost certainly
overestimated electric force effects, however, in the assumption that hydrometeors
of a given class are uniformly charged, which disagrees with in situ particle charge
measurements. The maximum electric field also was allowed to increase to 250 kV
m1 before activating the lightning scheme (a three-dimensional adaptation of the
Helsdon et al., 1992, scheme). Sun et al. (2002) incorporated somewhat out-of-date
parameterizations of electrification, including drop–droplet inductive charge
separation (Chiu, 1978), which is widely considered to be ineffectual due to
enhanced drop coalescence even in relatively weak electric fields.
Mansell et al. (2002) simulated lightning behaviour in two different storms
using the same charge separation schemes. A low-shear thunderstorm had a normal
tripole charge structure (main negative charge with main positive charge above and
lower positive charge below). A high-shear supercell simulation, however, exhib-
ited an inverted tripole structure. The low-shear storm produced negative cloud-to-
ground (CG) lightning that was initiated between the main negative and lower
positive charge regions. The supercell storm simulation had positive CG lightning
that was initiated between a main positive region and lower negative charge.
A new electrified model was described by Barthe et al. (2005), who presented
initial results from a version of the Meso-NH model that included electrification
and lightning parameterizations. They also included a sensitivity study of two
graupel–ice schemes. Barthe and Pinty (2007) presented more details of the para-
meterizations used in Barthe et al. (2005), especially details concerning the light-
ning scheme.
The model used in Mansell et al. (2002) has been used in a number of later
studies. Kuhlman et al. (2006) presented simulations of a severe supercell storm
and compared results to observations of that storm. Fierro et al. (2006) investigated
Modelling of charging processes in clouds 37
φ
zk
r dz
dr r dφ
z0
Because the electric field from an infinite layer is constant with height above and
below the layer, this geometry is likely to give electric field profiles considerably
different from profiles appropriate to most physically realistic thunderstorm geo-
metries (although it might be suitable for extensive stratiform clouds).
More realistic electric field profiles usually can be obtained in a one-
dimensional model in which the size of the cloud is limited to some radius. On
the axis of a cloud with cylindrical symmetry, the only non-zero component of the
electric field is again the vertical component. The contribution of an infinitesimal
charge (including its image charge) to the vertical electric field on the axis at height
z0 (Figure 2.3) is given by
" #
rk zk z0 zk þ z0
dEz ðz0 Þ ¼ þ rdrdjdz
4pe ½r2 þ ðzk z0 Þ2 3=2 ½r2 þ ðzk þ z0 Þ2 3=2
(2.11)
The electric field on the axis due to a thin disk of charge of radius R(z) (i.e. R
can vary with height) at the kth grid level then is obtained by integrating over r and
j. Some care is needed in determining the constants of integration to get
" #
rk Dz zk z0 zk z0
Ez ðZ0 Þ ¼ þ C (2.12)
2e ½R2 þ ðzk z0 Þ2 1=2 ½R2 þ ðzk z0 Þ2 1=2
r
r2 f ¼ (2.13)
e
where r is the total space charge density at the point being evaluated. Then, the
electric field is computed from the potential by using the relationship E = rf.
Note that, in a two-dimensional model, the symmetry of the model requires the
component of the electric field perpendicular to the plane of the model to be zero.
Each point in a two-dimensional (slab-symmetric) model physically represents
an infinite line charge, which distorts the electric field relative to a three-
dimensional model.
To use numerical Poisson solvers (e.g. multi-grid iteration) to determine f, it
is necessary to specify boundary conditions for f on all sides of the grid. As an
elliptic equation, boundary conditions specify the value of f, the normal deri-
vative @f/@n or a linear combination of the two. Since the ground is a good con-
ductor, hence an equipotential, it is obvious that the potential of the bottom layer of
the grid at the ground should be set to a constant. The constant usually is chosen to be
zero, although any constant could be chosen without affecting the resulting electric
fields. It is generally inappropriate to set the potential to a constant at the lateral grid
boundaries. The upper boundary can be set to a fair-weather value if it is high enough
to approximate the electrosphere, as a constant potential at the top seems to result in
more stable solutions than setting the normal derivative. Choosing boundary condi-
tions for the other sides is not as straightforward as for the ground, and choices can
depend on the cloud size and geometry being modelled. At r = 0 in an axisymmetric,
two-dimensional model, for example, Ez is the only non-zero component of E, so
the boundary condition there would be @f/@r = 0. In three-dimensional or slab-
symmetric, two-dimensional models, however, Ex might well be large if the storm is
near the boundary at x = 0 (or x = X), so @f/@x = 0 would not be an accurate
boundary condition on that side. It often is impossible to find simple boundary con-
ditions that are exactly correct. However, if the cloud is completely contained by the
grid, suitable a priori boundary conditions can be used to produce reasonably accurate
calculations of f.
At least two strategies have been used to improve the solution of f when
charges get close to the domain boundaries. One method is to use brute force to
calculate f at the boundaries (e.g. Riousset et al., 2007); this increases solution
accuracy in the interior but is computationally expensive. An alternative method is
to solve for f in an enlarged domain that has fair-weather charge densities between
the model boundaries and the enlarged domain (e.g. Mansell et al., 2005). Pushing
40 Lightning electromagnetics: Volume 2
out the lateral and upper potential boundaries has the effect of pushing the mirror
charges farther away, thereby reducing the error they cause in the solution.
The charge density from ions is included as a category if a model treats ion
processes.
To derive expressions for the continuity equations, first consider charged parti-
cles in a parcel of air. Here, a parcel is a fixed mass of air but with a volume that can
change (e.g. can expand when rising and compress when sinking). If no charge enters
or leaves the parcel, then the total charge Qt in the parcel is invariant to parcel motion
and volume changes, but the total charge density rt = Qt/V is not invariant (where V is
the parcel volume at a given time). Thus, we can temporarily define a ‘charge mixing
ratio’ er ¼ r=rair as the charge per mass of air. Then the parcel-following
(Lagrangian) advection equation for charge on hydrometeor type n is
rn
de
¼ Ser n (2.15)
dt
where Ser n represents processes that can change the charge on the nth hydrometeor
type, such as collisional charge separation and ion attachment. Ser n also includes
transfer of charge in and out of the parcel by sedimentation (for precipitation par-
ticles) and turbulent mixing. The Lagrangian reference is then transformed to the
fixed Eulerian grid volume reference frame by the coordinate transformation
d=dt @=@t þ V r as
@e
rn
¼ V re
r n þ Ser (2.16)
@t n
For consistency with typical cloud model equations, we multiply (2.16) by air
density rair.
@e
rn
rair ¼ ¼ rair V re
r n þ rair Ser (2.17)
@t n
Modelling of charging processes in clouds 41
Expanding the divergence term using a vector identity and dividing again by rair
gives
@e
rn 1
¼ r n rair VÞ e
½r ðe r n r ðrair VÞ þ Ser (2.18)
@t rair n
Note that in the incompressible case we have rair = constant and r (rair V) = 0.
Substituting e
r n ¼ rn =rair and cancelling the factors of rair, one recovers the more
familiar expression of local charge continuity, @rn =@t ¼ r J n , with the current
density Jn replacing the charge motion rnV.
Equations (2.17) and (2.18) are mathematically equivalent and known as the
advective and flux forms, respectively. The flux form is generally preferred in finite
difference models because it has immediate conservation properties that are not
guaranteed in an advective form.
To (2.18), we now add the turbulent mixing and sedimentation terms (second
and third terms, respectively, on the right-hand side of the following):
@e
rn 1 1
¼ ½r ðe r n ra VÞ e r n r ðra VÞ þ r ðra Kh re
rnÞ
@t ra ra
r n ra V er Þ
1 @ðe
þ n
þ Ser (2.19)
ra @z n
@rn @ðrn V n Þ
¼ r ðrn VÞ þ r ½ra Kh rðrn= ra Þ þ þ Sn (2.20)
@t @z
Note that the advection part has been reduced to one term and that the second
term effectively retains e r n . In many models, however, the scalar advection for-
mulation is designed for parcel invariants (such as mass mixing ratio), so the for-
mulation of (2.19) can use the same model code without modification. The state
variable rn has more physical meaning, however, than e r n , so a compromise is to
transform rn to e r n for advection and turbulent mixing, and then back to rn for
electrification processes.
As discussed by Pruppacher and Klett (1997), V n can be generalized to include
the effect of electrostatic force in addition to gravity. If this is done, the third term
on the right of (2.19) should be replaced by ð1=rair Þr ðrair e r n Vn Þ, where Vn is
now a vector, not a scalar.
Many microphysical processes are capable of placing charge on particles. The
main processes that have been included in cloud models are the non-inductive
graupel–ice mechanism, the inductive mechanism and ion capture. The rest of this
section considers treatments of each of these. Electrification mechanisms that have
not yet been used in models are not considered here, although other mechanisms
may be significant in some situations (e.g. melting processes may play a role in
electrifying the stratiform precipitation region of storm systems).
30
20
10
Cloud water content (g/m3)
–20
1
–10
40
30 0
0.10 20
10
0.01
0 –5 –10 –15 –20 –25 –30
Temperature (°C)
*Sign (polarity) is for charge transferred to graupel. Charge on cloud-ice has opposite polarity. †PLEZ
and NLEZ refer to positive/negative low EW zones (see Figure 2.5). ‡Tr is the temperature at which the
polarity of charge gained by graupel reverses for a given value of liquid water content: Tr ¼ 15.06EW
7.38, for EW in g/m3 and Tr in oC. §For regions where q(EW,T) is negative (at EW near 1.1 for T 24oC),
use q(EW,T) ¼ 0. kMust also be outside previous ranges in T-EW space. For regions where q(EW,T) is positive
(at 16oC < T < 20oC and EW 0.1 g/m3), use q(EW,T) ¼ 0.
46 Lightning electromagnetics: Volume 2
to account for aerodynamic effects that sweep small particles around graupel, and
Estick is the fraction of colliding particles that stick to the graupel surface (i.e. do not
rebound). Saunders et al. (1991) noted that in their experiments, EW typically was
roughly equal to LWC/2.
The polarity of charge that graupel gains for each combination of temperature
and effective liquid water content under the parameterization described in Table 2.2
is shown graphically in Figure 2.5. The line dividing positive and negative charging
for 0.22 g/m3 < EW < 1.1 g/m3 can be given in terms of critical values of either EW
or T. Saunders et al. (1991) gave the following expression for EWcrit (in g/m3),
valid for 10.7 C > T > 23.9 C:
1.4 S91
1.2
+
EW (g/m3)
1.0
0.8
0.6
–
0.4
0.2
NLEZ PLEZ
0
0 –10 –20 –30
Temperature °C
Figure 2.5 Plot of the charging zones of the S91 (Saunders et al., 1991) non-
inductive ice–ice parameterization. The positive and negative low-EW
zones are indicated by PLEZ and NLEZ, respectively. Heavy dashed
line indicates extrapolation to higher temperature (no charge
separation occurs for conditions below the dashed grey line)
Modelling of charging processes in clouds 47
(determined from Formvar slides) and an assumed ice crystal collection efficiency of
approximately 0.5 (i.e. only about half of the colliding crystals will rebound). Models
typically assume much lower collection efficiencies, with correspondingly higher
rebound efficiencies. However, simulations generally have not accounted for this
difference, which could cause overestimation of charging rates by up to a factor of 2.
Brooks et al. (1997) suggested that the rime accretion rate (RAR) may affect
the sign and magnitude of charge per collision. They hoped to reconcile the results
of Takahashi (1978) and Saunders et al. (1991), which had riming rod speeds of
9 m/s and 3 m/s, respectively. They converted the Saunders et al. (1991) results
using RAR = EWDVgi/3 (units of gm2s1) in place of EW in the Saunders et al.
(1991) parameterization, so that (2.25) became
RARcrit ¼ 1:47 0:02T (2.26)
for the range 23.8 C < T < 7.4 C. (For T > 7.4 C or T < 23.8 C, one can
use RARcrit (T = 7.4 C) or RARcrit (T = 23.8 C), respectively.) Brooks et al.
(1997) also reported limited laboratory tests at 15 C that showed that charging
was similar for different combinations of V and EW that had the same RAR. Their
reformulation left out the low effective liquid water zones, citing difficulty in
reproducing results at low EW.
A follow-on study by Saunders and Peck (1998) examined charge reversal as
function of RAR. They fit a sixth-order polynomial function RARcrit to represent the
critical RAR (Figure 2.6), above which graupel gained positive charge:
4
SP98
Rime accretion rate (gm–2s–1)
3
+
2
1 –
Figure 2.6 Plot of the critical rime accretion rate RARcrit curve of Saunders and
Peck (1998). Graupel charges positively at rime accretion rates above
the curve and negatively below. Charge separation is assumed to cut
off at about 33 C
48 Lightning electromagnetics: Volume 2
which could then be used with the revised Brooks et al. (1997) charging para-
meterization. The fit is only valid to about 33 C, below which Saunders and Peck
(1998) expected charge separation to become negligible. Mansell et al. (2005) used
a piece-wise continuous function to extend charge separation smoothly to zero over
the low-temperature end of the curve from 33 C to 40 C.
For Brooks et al. (1997) and Saunders and Peck (1998), the mean separated
charge per rebounding collision can be adapted for bulk distributions as
m ðV g V 1 Þn q ðRARÞ
dq ¼ kq D (2.28)
n;1
where V g and V I are the mass-weighted mean terminal speeds for graupel and
cloud ice (or snow), respectively, and kq, m, and n are constants that depend on
crystal size as shown in Table 2.2. Mansell et al. (2005) altered the charge
separation equations, q(RAR, T) from Brooks et al. (1997) so that they smoothly
approach zero at RAR = RARcrit. For positive charging of graupel (RAR > RARcrit),
For negative charging (0.1 gm2 s1 < RAR < RARcrit),
Note that there is an implicit temperature dependence since RARcrit varies with
temperature. The negative charging equation (2.30) shifts the parabolic function
given in Brooks et al. (1997) to fit between the limits of 0.1 gm2 s1 and RARcrit,
removing the discontinuity at RARcrit in the original formulation. Charging is
assumed to be zero for RAR < 0.1 gm2 s1.
Few quantitative results are available on charge separation at temperatures less
than 30 C. (One example is Saunders and Peck, 1998, which examined the sign
of charging, but not the quantity, at lower temperature.) Therefore, lacking
experimental guidance, Mansell et al. (2005) suggested limiting charging at low
temperature by an arbitrary factor b, such as
8
<1 : T > 30 C
b ¼ 1 ½ðT þ 30=13Þ : Thom > T < 30 C
2
(2.31)
:
0 : T < Thom
The low-temperature cut-off is made at Thom –38 C because all cloud dro-
plets are homogeneously frozen by that temperature, so no further riming occurs.
Deep convective updrafts, however, may maintain ice supersaturation for T < Thom,
which might support appreciable charge separation but no laboratory data exist.
Mitzeva et al. (2006) used a one-dimensional cloud model to explore possible
Modelling of charging processes in clouds 49
If ng is the number density of graupel particles of diameter Dg, then the rate at
which ng graupel particles collide and separate from ni cloud-ice particles is given
by Kgi ng ni, and the rate at which this process charges graupel of diameter Dg and
cloud ice of diameter Di is
@rg
¼ Kgi ng ni dq
@t (2.33)
@r
¼ i
@t
The expression for non-inductive charging between graupel and snow is the
same as (2.32) and (2.33), except that the parameters for cloud ice are replaced by
those for snow (ns, Ds, Egs and DVgs). Similarly for hail and snow, graupel para
meters are changed to the values for hail (nh, Dh, Ehs and DVhs), although dq has not
yet been determined by laboratory experiments specifically for hail.
To determine the rate of change in charge density for graupel in a particular
size category of a model with bin microphysics, it is necessary to add together the
contributions given by (2.33) for every size category of snow and cloud ice that
interacts with the graupel. Likewise, the rate of change in charge density on a
particular size category of cloud ice or snow must include the contribution from all
size categories of graupel and hail.
To determine the rate of change in charge density on a particular type of
hydrometeor in a model with bulk microphysics, two approaches have been used.
50 Lightning electromagnetics: Volume 2
The first is to evaluate (2.33) at a grid point by determining mean values of dq and
Kgi and the total concentrations across all sizes for Ng and Ni at that grid point. The
second approach is to integrate the right-hand side of (2.33) across all sizes both of
that hydrometeor category and of every other category with which that category
interacts. For example, the change in charge density of graupel due to collisions
with snow or cloud ice is
ðð
@rg p
¼ ðDg þ D1 Þ2 DVgi Egi ng ðDg Þni ðDi Þdq dDi dDg (2.34)
@t 4
where the subscript i refers to either snow or cloud ice. The size distributions n(D)
often are assumed to have the form of the inverse exponential distribution (2.2 with
a = 0) with parameters n0 and L. For cloud ice, because Di Dg and ViT VgT,
the magnitude of sums and differences of these quantities can be approximated as
Dg and VgT in (2.34). To evaluate this integral for cloud ice, Ziegler et al. (1986)
assumed further that since cloud ice typically has a narrow size distribution, it can
be approximated as a population with a single average diameter D i . If Ni is the total
concentration of cloud-ice particles, this implies that
ð
@rg p
¼ Ni Egi ðDg þ D i Þ2 DVgi ng ðDg Þdq dDg (2.35)
@t 4
The general formulation of (2.34) for the non-inductive charge (density)
separation rate @rxy/@t between ice hydrometeor classes x and y is
ð1 ð1
@rxy p 0
¼ dqxy ð1 Exy ÞDVxy j
@t 0 0 4
where Dx and Dy are the diameters of the colliding particles, Exy is the collection
efficiency, DExy is the relative fall speed, nx and ny are number concentrations and
dq0 xy is the charge separated per rebounding collision. In general, dq0 xy may be a
function of ice-crystal diameter, impact speed, cloud water content and tempera-
ture. The collection efficiency Exy is the product of the collision efficiency (Ecolli,
assumed to be unity) and the probability of sticking given a collision (Estick). In
wet-growth mode, one may assume Ex,y = 1, and no charge separation occurs.
As it stands, (2.36) is not a tractable integrand. The equation can be approxi-
mated and simplified by assuming a form for dq0 xy that can be pulled out of the
integral. Also, the fall speed difference is approximated by the difference of mass-
weighted mean fall speeds. The collection efficiency is assumed to be constant.
Multiplying and dividing by Exy then isolates the number concentration collection
rate integral (nxacy):
@rxy 1
¼ bdqxy ð1 Exy ÞExy ðnxacy Þ (2.37)
@t
Modelling of charging processes in clouds 51
Where
nxacy ¼ Exy DV xy
ð1 ð1 (2.38)
p
ðDx þ Dy Þ2 nx ny dDx dDy
0 0 4
Collision rates for other size distribution functions (e.g. a general Gamma
function) are beyond the scope of this chapter. For other collision rates, readers are
referred to studies of microphysics parameterizations such as Milbrandt and Yau
(2005) and Seifert and Beheng (2006).
The expression for the induced charge gained by a spherical graupel or hail particle
in a rebounding collision with a cloud droplet in an electric field is
where rcld is the radius of the cloud droplet; qE,r is the angle between the impact
point and the electric field vector through the centre of the graupel/hail particle
(shown in Figure 2.1); Qg and Qcld are the charge already on the graupel/hail par-
ticle and cloud droplet, respectively; g1, A, and B are dimensionless functions of
rcld/rg the ratio of the radii of the two particles:
g2 ðrcld =rg Þ2
A¼ (2.42)
1 þ g2 ðrcld =rg Þ2
1
B¼ (2.43)
1 þ g2 ðrcld =rg Þ2
where the weighting factor under the first integral is an infinitesimal area in the
horizontal cross section of the graupel particle. Then the mean charge transferred to
a graupel particle by colliding with and separating from cloud droplets is
where G (1.5) = 0.886 and G (3.5) = 3.323. They chose Egc = 0.84, Er = 0.1, a = 0.022
and hcos ji ¼ 0:1, which gave a probability of rebounding collisions prbnd = EgcEra
near the lower end of the range found by Aufdermaur and Johnson (1972).
Cosmic rays are the source of most ions in the fair-weather troposphere, except
near the ground; radioactive decay from the surface contributes up to half of the ions
found near ground. In fair weather, the number of positive ions is roughly equal to the
number of negative ions, so the net charge density results from small differences of
large numbers of ions. Ions move under the influence of the electric field E and their
average drift velocity in a given E is the mobility times the electric field m E, where
subscript gives the polarity of the ions. The mobility m increases with decreasing
pressure and so increases with height (i.e. the mean free path increases).
Each polarity of ion must obey its own continuity equation, which is similar to
(2.20) for charge attached to hydrometeors. For ions, however, a term must be
added to account for average charge motion under the influence of the electric field,
since the resulting ion drift velocity can easily be much different than wind velo-
cities. Also, the source/sink term often is split into sources and sinks occurring in
fair weather and those requiring clouds or thunderstorms (e.g. Chiu, 1978; Helsdon
and Farley, 1987b). The continuity equation for free ions can be written as
@n 1 n
¼ r ½n V n m E þ r Km r þG
@t rair rair (2.50)
anþ n Satt þ Spd þ Sevap
where n+ is the number density of positive ions and n–, the number density of
negative ions. Advection (the first term in the brackets) and turbulent mixing (the
second full term) are treated the same as for the other scalar variables. The second
term in the brackets is the ion drift motion, which can be treated similarly to the
advection term. G is the background cosmic ray ion generation rate; an+n– is the ion
recombination rate and the last three terms, respectively, are ion attachment to
hydrometeors (sink), point discharge current from the surface (source) and release
of any charge as ions from hydrometeors that evaporate completely (source). If the
ion drift speeds exceed the maximum for stable transport at the time step of the
model, then the ion processes (except advection and turbulent mixing) can be per-
formed on a subdivided time step, leaving the dynamical time step unchanged.
The fair-weather state can be defined as in Gish (1944), using the modified
coefficients of Helsdon and Farley (1987b):
Ez;FW ¼ E0 ðb1 ea1 z þ b2 ea2 z þ b3 ea3 z Þ (2.51)
where G(z) is the ion generation rate by cosmic rays (held constant as a function of
altitude) and a = 1.6 1012 m3 s is the ionic recombination coefficient (Chiu,
1978).
The ion mobilities m from Shreve (1970) are given by
4
m ¼ b e1:410 z
(2.54)
where b+= 1.4 10 m V s and b– = 1.9 10 m V s
–4 2 –1 –1 –4 2 –1 –1
and z is in metres.
Diffusivity is derived from mobility by the Einstein relation
kT
D ¼ m (2.55)
e
where k is Boltzmann’s constant, T is the temperature (in Kelvin), and e is the
electron charge magnitude (ions are assumed to be singly charged).
Under steady-state (i.e. fair-weather) conditions, one may assume that the ion
currents and charge densities vary negligibly from constant values, so that
rj = 0 Therefore, from (2.52) we get
GðzÞ ¼ anþ;FW ðzÞn;FW ðzÞ ðsteady stateÞ (2.56)
(as in Takahashi, 1979) and the cosmic ray generation rate can be held constant in
time throughout a simulation.
Ion attachment to hydrometeors is a combination of diffusion, Sdiff, and con-
duction, Scond. As in Chiu Chiu (1978), the two terms are calculated separately and
added (Satt = Sdiff + Scond). The equations from Chiu (1978) (based on Whipple
and Chalmers, 1944)) for attachment by conduction can be found in Table 2.3.
*nj ¼ nj(rj). †Hydrometeor completely positive. ‡Hydrometeor completely negative; from Chiu (1978)
based on Whipple and Chalmers (1944). The minimum charge on a particle needed to prevent polar-
ization is QM 12pejEjrj2 .
Modelling of charging processes in clouds 57
In Chiu (1978), the vertical component of the electric field was used to calculate
Scond (i.e. |E| is replace by |Ez|). The equation for diffusion attachment used by
Mansell et al. (2005) is similar to Chiu (1978) and Helsdon (1980):
" #
@n X rj VTj 1=2
Sdiff ¼ ¼ j 4prj D n nj f ðXj Þ 1 þ (2.57)
@t diff 2pD
where rj, VTj, and nj are the mean radius, mean terminal fall speed, and number
concentration of the jth hydrometeor category, respectively. The factor f (Xj) is from
Helsdon (1980):
Xj
f ðXj Þ ¼ (2.58)
eXj1
where Xj = Qj/QD; Qj is the charge per particle in hydrometeor category j; and
QD = 4perjkBT/e is the hydrometeor charge at which the electric potential and thermal
energies are balanced at the surface. In Helsdon (1980), Xj was allowed to become
negative, which enhances diffusion for ions of opposite charge [i.e. f(Xj) > 1].
Alternatively, one can allow only non-negative values of Xj [i.e. f(Xj) 1] to avoid the
possibility of double-counting the ion attachment by conduction (Chiu, 1978).
A point ion discharge current jpd (corona current) from features at the sur-
face can be allowed when the vertical electric field component at the ground
exceeds a given threshold (|Ez| > E0). The formulation used here (as in
Takahashi, 1984) follows Jhawar and Chalmers (1967) with values of Standler
and Winn (1979):
Table 2.4 Variables and suggested values of the MacGorman et al. (2001)
lightning parameterization affecting lightning structure or charge
discharges in the atmosphere but had been limited to two-dimensional slab sym-
metric or axisymmetric model domains and did not include cloud models.
The lightning model simulates the step-by-step propagation of a bidirectional
discharge from an initiation point by successively creating new breakdown exten-
sions, or bonds, from the discharge structure. Figure 2.7 shows a sample two-
dimensional grid with part of a discharge. Each new extension is chosen randomly
from all possible new extensions. An extension from a grid point on the channel to
an adjacent grid point is possible (i.e. has a non-zero probability) if the electric field
vector favours propagation towards the adjacent grid point and the electric field
magnitude between the two points exceeds a critical value, Ecrit. Once all possible
extensions have been identified, the probability p of choosing a particular extension
is given by
8
<1
for Ei > Ecrit
pi ðEÞ ¼ ðE Ecrit Þh (2.64)
:F i for Ei Ecrit
0
Figure 2.7 Sample grid showing a portion of a stochastic lightning model discharge
in two dimensions. Open circles represent grid points not connected to
the discharge. Filled circles and heavy lines indicate the discharge path.
Dashed lines indicate possible new bonds. In three dimensions,
connections are allowed along all of the unit cube directions
62 Lightning electromagnetics: Volume 2
NPW had a critical field of zero (Ecrit = 0) in their model and found that the
density of branching decreases with increasing values of h. NPW found that ‘bush’
type (densely packed) discharges result for h 1, but ‘branched’ structures result
for h 3. Wiesmann (1988) found that introducing Ecrit > 0 (with h = 1) has a
similar effect as setting h > 1, since both tend to limit side branching. Dissado and
Sweeney (1993) pointed out the difficulty in physically justifying any exponent
other than h = 1, however, so Mansell et al. (2002) assumed a linear relationship
between pi and Ei (i.e. h = 1), as well as Ecrit > 0.
In the WZ model, electrical resistance of the channels is represented by an
internal electric field Eint. A perfect conductor would have Eint = 0, and its surface
would have a constant potential, whereas a conductor with resistance has Eint > 0,
and the potential varies along its length. A non-zero internal field acts to reduce the
the electric field between the channel and non-channel points and thus reduces the
overall extent of the discharge structure (all other values being held equal). When a
new grid point is added to the discharge, the electric potential fnew at the new
channel point is calculated from the potential fb of the point from which the
channel segment extends:
fnew ¼ fb sEint d (2.65)
where d is the length of the new segment and s is the sign of charge carried by the
channel. Thus, the potential at a point that is n segments away from the initiation
point can be written in terms of the reference potential fref of the starting point, so
that (2.65) becomes
X
n
fðnÞ ¼ fref s Eint ðjÞdj (2.66)
j¼1
where the sum is along the path from the initiation point to segment n. Each
channel point has a unique path from the initiation point because no branch is
allowed to rejoin to another branch. For bidirectional discharges, the initial refer-
ence potential is taken as the average of the ambient potentials of the two initial
grid points.
The electric potential f at grid points not on the channel must be updated
after each extension of the conducting channel by solving Poisson’s equation
(2.13). The electric potential is already defined by (2.66) along the discharge
structure, which is treated as a boundary with Dirichlet conditions when solving
(2.13). Because of the complex geometry of the discharge boundary, Poisson’s
equation is most easily solved with an iterative technique such as successive
overrelaxation, as described in Appendix A of Mansell et al. (2002). The charge
density induced on the channel at a grid point can be found simply from Poisson’s
equation (r = –er2 f). The charge on the channel is not needed for solving (2.67)
because the effect of this charge on f is included by satisfying the boundary
condition imposed by the channel potential (2.66). The channel charge must be
calculated periodically during development of bidirectional discharges to check
for overall flash neutrality.
Modelling of charging processes in clouds 63
Lc c1
N ðdÞ d (2.68)
Lmoy
where d is the linear distance from the starting point and c is the assumed fractal
dimension. Lc is a tunable length-scale parameter, and Lmoy was described as the
‘average length of a grid mesh’ and could be interpreted as the cube root of the
volume defined by the average horizontal and vertical grid spacings:
1=3
Lmoy ¼ DxDyDz . The Lc/Lmoy factor was designed to help maintain consistency
between simulations with different grid resolution. Barthe and Pinty (2007) found
acceptable results for 2.5 c 2.8 and of 500 Lc 1,000 in tests with a grid that
had constant Dx=Dy=1,000 m and constant Dx 500 m.
In practice, the number of allowed points are found within concentric shells
with inner radius d and thickness Dl=(DxDyDz)1/3 of the smallest grid volume in the
domain (generally at the lowest level if the vertical spacing varies with altitude).
The scheme iterates from the smallest to largest shells, randomly adding new
channel points if another site is allowed by N(d) and a point adjacent to a current
channel point meets the criteria of potential and charge density. All candidate
points are given equal probability of being chosen. Once a shell is filled or has no
more candidate points, the next larger shell is considered. Since the electric field at
the channel tips is neither calculated nor taken into account, the growth of the
structure is more akin to diffusion-limited aggregation models (e.g. Witten and
Sander, 1981; Garik et al., 1987) than dielectric breakdown models (e.g. Niemeyer
et al., 1984; Wiesmann and Zeller, 1986; Petrov and Petrova, 1993; Pasko et al.,
2000; Mansell et al., 2002).
64 Lightning electromagnetics: Volume 2
that the convective mechanism alone could produce only very weak electrification
and that the charge structure tended to be opposite of what was expected by the
hypothesis. Simulations that included ice–ice non-inductive charge separation,
however, did produce substantial electric field magnitudes. A general result of
these and other simulations (e.g. Mansell et al., 2005) is that the process of ion
attachment plays an important role in the formation of electrical screening layers,
but by itself cannot result in strong electrification.
14
Reflectivity 20 m s–1
12 Max vector
20 10
10 –40
Altitude (km)
30
8 –30
50 40
–20
6
–10
4 0
2
Temp. (°C)
0
(a) 10 15 20 25 30 35
14
Graupel/cloud mass contents
12 Cloud edge
10 30 d –40
Altitude (km)
BZ
8 –30
4
–20
6 3
2 –10
1
4 0
0.5 1 1.5
2 0.1
0
10 15 20 25 30 35
(b) Horizontal distance (km)
14
S91
12
30 dBZ
10 –40
Altitude (km)
8 –30
–20
6
–10
4 0
2
Cloud edge
Temp. (°C)
0
(a) 10 15 20 25
14
Takahashi
12
30 dBZ
10 –40
Altitude (km)
8 –30
–20
6
–10
4 0
2
Cloud edge
Temp. (°C)
0
(b) 10 15 20 25
14
SP98
12
30 dBZ
10 –40
Altitude (km)
8 –30
–20
6
–10
4 0
2
Cloud edge
0
10 15 20 25
(c) Horizontal distance (km)
Temp. (°C)
Horizontal distance (km) Horizontal distance (km) Horizontal distance (km)
(a) (b) (c)
10 15 15 25 5 –15
25 –25 –20
0 2515
8 0 35 5 –10
0 –5
6 0 0
0 –35
4 45 –35 –25 –5
25 5
15 –25 –15 +CG
2 –CG
+CG –15 –5 –5 0
15
5 –CG 5 0
0
10 15 20 25 30 35 10 15 20 25 30 35 10 15 20 25 30 35
Horizontal distance (km) Horizontal distance (km) Horizontal distance (km)
(d) (e) (f )
< –0.1 nC m–3 > 0.1 nC m–3 < –0.6 nC m–3 > 0.6 nC m–3
Figure 2.10 Simulated charge structure (top) and electric potential (bottom) using the (a,d) S91, (b,e) Takahashi and (c,f)
SP98 schemes. Contours of positive (black) and negative (grey) potential have intervals of 5 MV, 15 MV and
increasing thereafter by intervals of 10 MV. Local reversals of charge sign are caused by simulated lightning charge
deposition. The stars (d,e,f) indicate approximate initiation points of cloud-to-ground (CG) near the time of the plots
(52 min). The jagged lines connect the initiation points (stars) to approximate relative ground strike locations
70 Lightning electromagnetics: Volume 2
References
Aufdermaur, A. N. and D. A. Johnson, 1972: Charge separation due to riming in an
electric field. Quart. J. Roy. Meteor. Soc., 98, 369–382.
Baker, B., M. B. Baker, E. R. Jayaratne, J. Latham and C. P. R. Saunders, 1987:
The influence of diffusional growth rates on the charge transfer accompanying
rebounding collisions between ice crystals and soft hailstones. Quart. J. Roy.
Meteor. Soc., 113, 1193–1215.
Baker, M. B., H. J. Christian and J. Latham, 1995: A computational study of the
relationships linking lightning frequency and other thundercloud parameters.
Quart. J. Roy. Meteor. Soc., 121, 1525–1548.
Barth, M. C., S.-W. Kim, C. Wang, K. E. Pickering, L. E. Ott, G. Stenchikov, et al.,
2007: Cloud-scale model intercomparison of chemical constituent transport in
deep convection. Atmos. Chem. Phys., 7, 4709–4731.
Barthe, C., G. Molinié and J.-P. Pinty, 2005: Description and first results of an
explicit electrical scheme in a 3D cloud resolving model. Atmos. Res., 76, 95–
113.
Barthe, C. and J.-P. Pinty, 2007: Simulation of a supercellular storm using a three-
dimensional mesoscale model with an explicit lightning flash scheme.
J. Geophys. Res., 112, doi:10.1029/2006JD007484
Barthe, C., J.-P. Pinty and C. Mari, 2007: Lightning-produced NOx in an explicit elec-
trical scheme tested in a Stratosphere-Troposphere Experiment: Radiation, Aerosols
and Ozone case study. J. Geophys. Res., 112, doi:10.1029/2006JD007402
Brooks, I. M. and C. P. R. Saunders, 1994: An experimental investigation of the
inductive mechanism of thunderstorm electrification. J. Geophys. Res., 99,
10627–10632.
Brooks, I. M., C. P. R. Saunders, R. P. Mitzeva and S. L. Peck, 1997: The effect on
thunderstorm charging of the rate of rime accretion by graupel. Atmos. Res., 43,
277–295.
Canosa, E. F., R. List and R. E. Stewart, 1993: Modeling of inductive charge
separation in rainshafts with variable vertical electric fields. J. Geophys. Res.,
98, 2627–2633.
Chiu, C.-S., 1978: Numerical study of cloud electrification in an axisymmetric,
time-dependent cloud model. J. Geophys. Res., 83, 5025–5049.
Cotton, W. R. and R. A. Anthes, 1989: Storm and Cloud Dynamics. Academic
Press, San Diego, 883 pp.
Dissado, L. A. and P. J. J. Sweeney, 1993: Physical model for breakdown structures
in solid dielectrics. Phys. Rev. B, 48, 16261–16268.
Dwyer, J. R., 2003: A fundamental limit on electric fields in air. Geophys. Res.
Lett., 30, doi:10.1029/2003GL017781.
Emersic, C., 2006: Investigations into Thunderstorm Electrification Processes.
Ph.D. thesis, University of Manchester, Manchester, England, UK, 327 pp.
Fierro, A. O., M. S. Gilmore, E. R. Mansell, L. J. Wicker and J. M. Straka, 2006:
Electrification and lightning in an idealized boundary-crossing supercell
simulation of 2 June 1995. Mon. Wea. Rev., 134, 3149–3172.
72 Lightning electromagnetics: Volume 2
Keith, W. D. and C. P. R. Saunders, 1989: Charge transfer during multiple large ice
crystal interactions with a riming target. J. Geophys. Res., 94, 13103–13106.
Keith, W. D. and C. P. R. Saunders, 1990: Further laboratory studies of the char-
ging of groupel during ice crystal interactions. Atmos. Res., 25, 445–464.
Kuettner, J. P., Z. Levin and J. D. Sartor, 1981: Thunderstorm electrification-
inductive or non-inductive? J. Atmos. Sci., 38, 2470–2484.
Kuhlman, K. M., C. L. Ziegler, E. R. Mansell, D. R. MacGorman and J. M. Straka,
2006: Numerically simulated electrification and lightning of the 29 June
2000 STEPS supercell storm. Mon. Wea. Rev., 134, 2734–2757.
Latham, J. and B. J. Mason, 1962: Electrical charging of hail pellets in a polarizing
electric field. Proc. R. Soc. London, Ser. A, 266, 387–401.
Levin, Z., 1976: A refined charge distribution in a stochastic electrical model of an
infinite cloud. J. Atmos. Sci., 33, 1756–1762.
Lhermitte, R. and P. R. Krehbiel, 1979: Doppler radar and radio observations of
thunderstorms. IEEE Trans. Geosci. Electron., GE-17, 162–171.
MacGorman, D. R., A. A. Few and T. L. Teer, 1981: Layered lightning activity.
J. Geophys. Res., 86, 9900–9910
MacGorman, D. R. and W. D. Rust, 1998: The Electrical Nature of Storms. Oxford
University Press, New York, 422 pp.
MacGorman, D. R., J. M. Straka and C. L. Ziegler, 2001: A lightning para-
meterization for numerical cloud models. J. Appl. Meteor., 40, 459–478.
Mansell, E. R., D. R. MacGorman, C. L. Ziegler and J. M. Straka, 2002: Simulated
three-dimensional branched lightning in a numerical thunderstorm model.
J. Geophys. Res., 107, doi:10.1029/2000JD000244
Mansell, E. R., D. R. MacGorman, C. L. Ziegler and J. M. Straka, 2005: Charge
structure and lightning sensitivity in a simulated multicell thunderstorm.
J. Geophys. Res., 110, doi:10.1029/2004JD005287.
Mansell, E. R., C. L. Ziegler and E. C. Bruning, 2010: Simulated electrification of a
small thunderstorm with two-moment bulk microphysics. J. Atmos. Sci., 67,
171–194, doi:10.1175/2009JAS2965.1
Marshall, B. J. P., J. Latham and C. P. R. Saunders, 1978: A laboratory study of
charge transfer accompanying collision of ice crystals with a simulated hail-
stone. Quart. J. Roy. Meteor. Soc., 104, 163–178.
Marshall, J. S. and W. M. Palmer, 1948: The distribution of raindrops with size.
J. Meteor., 5, 165–166.
Marshall, T. C., M. P. McCarthy and W. D. Rust, 1995: Electric field magnitudes
and lightning initiation in thunderstorms. J. Geophys. Res., 100, 7097–7103.
Mason, B. J., 1988: The generation of electric charges and fields in thunderstorms.
Proc. R. Soc. London, Ser. A, 415, 303–315.
Mathpal, K. C. and N. C. Varshneya, 1982: Riming electrification mechanism for
charge generation within a thundercloud of finite dimensions. Ann. de
Géophys., 38, 167–175.
Milbrandt, J. A. and M. K. Yau, 2005: A multimoment bulk microphysics para-
meterization. Part II: A proposed three-moment closure and scheme descrip-
tion. J. Atmos. Sci., 62, 3065–3081.
74 Lightning electromagnetics: Volume 2
3.1 Introduction
Electrical gas discharges belong to the class of low-temperature plasmas.
Depending on particular conditions, they can be equilibrium (thermal) or non-
equilibrium (non-thermal) meaning that temperatures/energies of constituting par-
ticles (electrons, ions, neutral atoms/molecules) are either nearly equal or different,
respectively. Examples of the former are electric arcs, sparks, microwave dis-
charges, where such effects like Joule heating of the gas and its thermal ionization
are significant. In contrast, non-thermal discharge plasmas are characterized by
presence of highly energetic electrons in “cold” neutral gas and ionic component
with temperature close (or slightly higher) to that of gas. Typical examples of these
are electron avalanches, glow discharges, streamers, electrical coronas, where the
degree of ionization of the gas is much less than unity, the electron density rarely
exceeds 1014 cm3 and their mean energy is normally below 10 eV.
This chapter deals with basic principles of numerical simulations of non-
thermal electrical discharges in air, which are predecessors and indispensable
attributes of a leader discharge (see, e.g., [1]). First, processes in such discharges
and a theoretical background of so-called fluid model are considered. Further,
numerical approaches utilized for computer implementation of the model are pre-
sented. Finally, examples are given including computer simulations of coronas as
well as positive and negative streamers.
1
Department of Electrical Engineering, Chalmers University of Technology, Sweden
78 Lightning electromagnetics: Volume 2
produce an electron-positive ion pair in collisions with neutral particles. This process
provides conditions for multiplication of charge carriers and exponential growth of
their densities ne,p = n0 exp(a x), which is called an electron avalanche (in the
expression, x is the length which electrons pass in the field). Primary and newly
generated electrons drift further in the field towards the anode being concentrated in
the avalanche head while its tail is formed by the produced positive ions, which are
much slower and move in the opposite direction to the cathode. Strong gradients of
the electron density at the head lead to a diffusion that provides a typical conical
shape of an electron avalanche, which was registered in early years of studying of gas
discharges [13]. When the avalanche reaches the surface of a metallic anode, the
electronic cloud at the avalanche head is absorbed/neutralized, and only the posi-
tively charged tail is left in the gap. There are two scenarios for further development
depending upon the produced space charge density.
In case when the number of electrons in the avalanche head (and, corre-
spondingly, ions in the tail) is relatively small (less than 108), the density of the
accumulated space charge is not high enough to modify the field distribution and
so-called Townsend’s or dark discharge can occur. In such a discharge, secondary
electrons can be emitted from the cathode surface due to impacts of positive ions,
metastable exited molecules and photons left in the gap, and they give rise to the
next generation of avalanches. If the generation rate becomes sufficiently high to
compensate losses of electrons, i.e. each lost electron can be replaced by a new one
emitted from the cathode, the discharge becomes self-sustained and an electrical
breakdown occurs due to unlimited multiplication of avalanches. A conductive
channel is created after breakdown, which can be further transformed into a spark
or an electric arc depending upon the properties of the external circuit. The
breakdown criterion for this so-called multi-avalanche mechanism is given by the
well-known Townsend’s condition ad = log(1+1/g), where g is the secondary
ionization coefficient specifying electronic yield due to secondary processes on the
cathode surface and a seed electron is assumed to appear in its vicinity and travel
the whole distance d between the electrodes.
If the primary avalanche is strong and the density of the space charge produced
is high enough to generate the field comparable with the electrostatic (external)
field, conditions for a streamer discharge formation and propagation can be even-
tually created. The streamer theory is based on the concept of growth of a thin
ionized channel between the electrodes, which follows the positively charged trail
left by the primary intensive avalanche and is supported by secondary avalanches
created due to photo-ionization in the gas volume at the streamer head. Streamer
discharges in air are known to propagate with extremely high velocity up to
108 cm/s and once created, they are able to develop even in regions with low
background fields. Streamers also exhibit different properties depending on the
direction of propagation (so-called cathode- and anode-directed streamers).
Depending on the arrangement of the discharge gap and ambient conditions,
streamers developing between electrodes can either lead to a complete breakdown
forming spark or arc discharges or serve as an initial stage for a leader process
taking place in long gaps with non-uniform electrostatic fields.
82 Lightning electromagnetics: Volume 2
The concepts of both electron avalanches and streamers are widely presented
in the literature, e.g. [2,6], and are not intended to be considered in details here.
Instead, physical processes taking place in these kinds of discharges are analyzed
below in conjunction with the examples of computer simulations.
The processes in gas discharges outlined in the previous section involve effects
related to collective motion of charged species such as drift in an electric field, their
diffusion and collisions. Therefore, a self-consistent model of discharge plasma can
be developed utilizing averaged swarm parameters of the processes characterizing
particles kinetics and interactions.
In general, there are two approaches to describe dynamic behavior of electron
and ions ensembles at microscopic level. The first one is based on individual
description while another one is based on some kind of statistical treatment, where
particles distribution function f(v, r, t) is considered, which obeys Boltzmann’s
equation
@f qE @f
þ v rr f þ rv f ¼ (3.1)
@t m @t c
Here, r is the position vector in space, v is the velocity vector, m is the particle
mass, and the right-hand side represents so-called collision integral, i.e. the rate of
change of the distribution function due to collisions, e.g., between electrons and
neutrals/ions. The gradient operators in (3.1) are defined in configuration space
(symbol r) and in velocity space (symbol v).
The individual description investigates not only the most probable distribution
but also the probability of deviations from this distribution. Within this approach
the movement of each particle is traced in configuration space at all stages of
discharge development and, thus, it provides deep insight into evolution of a par-
ticle assemble. At the same time, this is the major practical drawback from the
point of view of discharge modeling because such simulations require considerable
amount of computational time and memory even when very few particles are used.
Consequently, the later stages of discharges, such as avalanche-streamer transition
and streamer propagation, cannot be modeled without introducing significant
simplifications.
The kinetic approach deals with the averaged quantities and ignores individual
deviations of parameters of motion of charge carriers. Even though such deviations
are usually very small, (3.1) is still too difficult to be solved directly for most
problems due to strong changes of the electric field in space and time. In practice, it
is usually preferable to describe behavior of charged particles in plasma on a
macroscopic level by averaged quantities like particle fluxes, densities, velocities,
etc., which are defined hierarchically by the corresponding moments of
Boltzmann’s equation (3.1). This approach leads to an infinite set of moment
equations equivalent to the distribution function in the interested time scale of
Numerical simulations of non-thermal electrical discharges in air 83
moments, and which are easier to solve. Following this way, a fluid description of
the particles assembles is obtained, which treats the macroscopic properties in
terms of the moments of particles distribution functions [14]. Thus, the evolution of
electron density, electron average momentum and electron mean energy are defined
by the zero-, first-, and second-order moments. Higher order moments may be used
in some expansions, but they usually have no physical significance for gas dis-
charge simulation problems. The lowest moment is obtained by integrating (3.1)
over dv; the conservation of the momentum equation (first-order) is obtained by
multiplying (3.1) by mv and integrating over dv; and the conservation of the energy
equation (second-order) is obtained by multiplying (3.1) by mv2/2 and integrating
over dv. In this way, the non-equilibrium fluid model for electrons in the inter-
electrode air volume can be written as
@ ne
¼ rr ðne ve Þ þ Re
@t
@ ðm ne ve Þ
¼ rr ðm ne v2e Þ q E ne rr Pe m ne ve =tm
@t
@ ðne yÞ
¼ rr ðne e ve Þ q E ne ve rr ðPe ve Þ ne ðy y0 Þ=ty
@t
(3.2)
Here, Re represents the total rate of generation and loss; y stands for electron
mean energy; y0 is the background gas thermal energy. The electron pressure is
defined by the ideal gas equation of state Pe = ne kb Te, where kb is the Boltzmann
constant and Te is the electron temperature. Note that since the relationship between
the electron mean energy and the temperature is given by 3 kb Te / 2 = y mve2/2, the
electron pressure in (3.2) can be represented as Pe = 2ne (ymve2/2)/3. In addition,
the average velocity of electrons can be replaced with their drift velocity in the field.
In general, the electron average momentum and mean energy are in non-
equilibrium states in the time scales corresponding to the characteristic times of the
moment equations. The reason is that electrons gain momentum and energy from
the local electric field and ionize the neutral gas in electron-neutral ionization
collisions. Therefore, they lose a certain amount of momentum and energy through
collisions. These losses are accounted for in the last terms of the right hand sides in
the corresponding equations (3.2) which express the fact that the average momen-
tum and mean energy will reach quasi-equilibrium states in the time scales of their
relaxation times tm and ty , respectively. For typical air discharge plasma condi-
tions, the relaxation times are of order 1013–1012 s [2]. Thus if the time scale of
interest is larger than that, one may assume that the electron average momentum
and mean energy always reach a quasi-equilibrium state and use only zero-moment
of the distribution function, namely the first equation in (3.2), to describe the
kinetics of the electron ensemble. This greatly simplifies the formulation of the
problem, but imposes some limitations, e.g. on the rate of change of the applied
field or on its frequency. Note, that this assumption is true for other kinds of par-
ticles in the discharge plasma since the characteristic time scales for processes,
84 Lightning electromagnetics: Volume 2
where ions or atoms/molecules are involved, are much larger than those for elec-
trons due to differences in their mass and velocities.
The terms forming the collision integral in (3.1) may include as many reactions
as are of physical significance and are of interest in a particular situation, and each
particle type should be described by its own equation and each reaction by the
corresponding rate. Very often such a detailed approach is not required and aver-
aged properties are used. For example, behavior of positive ions of a certain kind
can be expressed by a single continuity equation similar to the first one in (3.2) with
the kinetic parameters characteristic for this particular type of ions. Further sim-
plifications can be made by considering generic kinds of particles, e.g., positive and
negative ions in case of a discharge in air. This allows reducing the set of equation
describing dynamics of charge carriers in discharge plasma to the three partial
differential equations (PDE) expressing conservation of their densities:
@ ne
þ r ðne we þ De r ne Þ ¼ Re
@t
@ np
þ r ðnp wp þ Dp r np Þ ¼ Rp (3.3)
@t
@ nn
þ r ðnn wn Dn r nn Þ ¼ Rn
@t
In some cases the set (3.3) can be made even simpler by omitting some terms
based on physical considerations and assumptions, e.g., the ionic diffusion can be
neglected when analyzing fast processes like streamer propagation, which typically
lasts some hundreds of nanoseconds if the gap length is of order of units of cen-
timeters while the diffusion of ions is much slower. The PDEs are to be com-
plemented with boundary conditions describing interactions of discharge plasma
with surrounding objects (metallic electrodes, dielectric walls, etc.) that is usually
implemented by setting up fluxes of charge carriers or their concentrations on the
interfaces. One should stress here that a correct choice of boundary conditions is
crucial especially when surface reactions leading to generation/loss of charged
species are to be considered. In addition, initial conditions should be provided for
the equations (3.3) defining distributions of the densities of electrons and ions in
space at the instant of discharge initiation.
The so-called source terms on the right-hand side of (3.3) include rates of
processes taking place in the discharge volume discussed above. Thus, background
ionization, electron impact ionization, electron attachment, electron–ion, and ion–
ion recombination are the main phenomena to be accounted for when (3.3) is used
to describe development of electron avalanches in air. In case of streamers, addi-
tional production of secondary electrons in the gas volume is defined by the rate of
photo-ionization Rph and the source terms can be written as:
Re ¼ a ne jwe j h ne jwe j bep ne np þ R0 þ Rph
Rp ¼ a ne jwe j bep ne np bpn np nn þ R0 þ Rph (3.4)
Rn ¼ h ne jwe j bpn np nn
Numerical simulations of non-thermal electrical discharges in air 85
burst or streamer coronas [33]. In general, the corona regimes are defined by
dynamics and configuration of space charges generated in the gas volume. Thus,
the glow mode appears due to permanent presence of space charges in the close
vicinity of the corona electrode while the other two modes are determined by the
motion of space charge cloud in the gas between electrodes. The simulations below
focus on glow corona mode although problem formulation and set-up for modeling
of burst discharge is similar.
After accepting the model settings, a CAD-tool window will appear where one
may define a computational domain and draw geometry for the problem. A com-
putational domain represents a part of space for which the problem is to be defined
in terms of coefficients of (3.3)–(3.5), boundary and initial conditions. The coef-
ficients in (3.3)–(3.5) are essentially material characteristics, which can be imple-
mented using constants, expressions and functions available via “Options” in the
main menu. The geometry can be drawn using provided geometry primitives and
utilizing Boolean operations (union, difference, intersection, etc.) or, alternatively,
it can be imported from external CAD software or even from an image or sketch
(see User Guide for corresponding procedures).
Constant quantities used in the model can be split into several groups: global
(Boltzmann constant, elementary charge), conditional (pressure, temperature,
initial charge density n0, rate of background ionization R0, etc.), those related to the
applied stress (e.g., voltage magnitude and its rise/decay rates in case of step-wise
dc or voltage amplitude and time constants for double-exponential representation of
a voltage impulse), microscopic parameters of air (mobilities of ions, masses of
particles, etc.), and auxiliary (e.g., artificial diffusion factors, see below). The
constants can be implemented using corresponding “Options” as illustrated in
Figure 3.2, where typical values of air parameters are borrowed from [9,34,35].
The constants are further used either directly in the problem settings or in the
expressions for calculating field/time/pressure/etc.-dependent characteristics. In the
latter case, the expressions to be computed are specified as shown in Figure 3.3.
Here, the reduced field E/N, Td, is defined in the first line using the in-built name
for the field variable normE_es in the Electrostatics equation. This quantity is
further utilized as an argument for functions describing field dependencies of
Figure 3.2 Definitions of constants in the corona model
electron drift velocities we(E/N) and their characteristic energies De /me = f(E/N) as
well as ionization a/N (E/N) and attachment h/N (E/N) coefficients. In contrast to
electrons, most of the parameters of ions in non-thermal air discharge plasmas are
dependent not only on electric field strength but also on gas temperature [34].
Therefore, kinetic temperatures of ions Tp and Tn (positive and negative, respec-
tively) are calculated using masses of corresponding ions from “Constants” and
they are further employed for obtaining diffusion coefficients Dp,n, ion-ion
recombination coefficient bii and detachment frequency ndet in (3.3)–(3.4).
Knowing all the parameters, rates of ionization, attachment, recombination and
detachment are computed and corresponding source terms (3.4) are formed, see
Figure 3.3. The space charge density (variable rho) is calculated as it is given on the
right-hand side of (3.5). Variables cur_den_cond and cur_den_cap represent con-
ductive and capacitive current densities, respectively, which are under integral
signs in (3.6). These quantities are used for computing integration coupling vari-
ables cur_cond and cur_cap (conductive and capacitive current, respectively) by
means of corresponding option available via Options ! Integration coupling
variables ! Subdomain variables (see Figure 3.4). Finally, the total current (vari-
able current in the expressions) is obtained as a sum of the two components.
Field-dependent characteristics of electrons in air (drift velocities, mean
energies, ionization and attachment coefficients) have been studied extensively and
have been presented in numerous scientific publications. However, detailed ana-
lysis of data available in the literature showed that there are large discrepancies in
the experimental as well as calculated results obtained by different authors which
can be attributed to specific conditions of experiments and equipment used,
methods of solving Botlzmann’s equation which yielded all the microscopic
properties of interest, etc. The dependencies presented in [36] (reproduced in
Figure 3.5) seems to be most reliable and are adopted in the present analysis. In the
10–21
10–28
α/N
η/N
10–24
100 101 102
(a) Reduced field E/N, Td
5
4.5 ×10
3.5
Drift velocity, m /s
2.5
1.5
0.5
0
0 100 200 300 400 500 600
(b) Reduced field, Td
10
9
Averaged energy of electrons De/u e , eV
0
0 100 200 300 400 500 600
(c) Reduced field, Td
Figure 3.5 Field dependencies of the ionization and attachment coefficients (a),
electron drift velocity (b) and mean electron energy (c) in air [36]
92 Lightning electromagnetics: Volume 2
software, the dependencies shown in Figure 3.5 are implemented as functions of the
reduced field E/N using their tabular representations as demonstrated in Figure 3.6
(functions can be accessed via “Options” in the main window of the software). The
values between points given in the table may be requires in the solution process
and, therefore, interpolation of the data is needed. This can be implemented using,
e.g., “Piecewise cubic” interpolation method, which preserves the shapes of a
curves in Figure 3.5. To compute the functions outside the provided ranges,
extrapolation is used (in Figure 3.6, the method is set to “Interpolation function”).
Note that the specified functions can be used on any stage of the simulations
(equation settings, post-processing, etc.) by providing a name and argument as it is
shown, e.g., in lines 2–4 in Figure 3.3. Alternatively, functional dependencies can
also be read from external data files. For this, one should prepare files according to
the format used in Matlab and specify a path to the file for each function. The files
are read during model initialization stage and data are kept in computer memory
while solving the model.
Next step in building the model is to assign material characteristics for the
computational domain. For the convection–diffusion equations, these include dif-
fusion coefficients, components of the drift velocities of the charge carriers and the
source terms (3.4). An example of the settings for electrons is shown in Figure 3.7a
(note “–” sign for the velocities of negatively charged particles which should be
changed to “+” for positively charged ones). Settings for Poisson’s equation
(Figure 3.7b) include dielectric constant of the medium er = 1 and space charge
density rho as defined in the Expressions (Figure 3.3). Initial conditions are given
via “Subdomain settings” (tab “Init” in Figure 3.7). It is usually assumed that
(a)
(b)
Figure 3.7 Settings for the convection-diffusion (a) and Poisson’s (b) equations
initially (at zero field) the gas is electrically neutral and the concentrations of
electrons and positive ions are set to n0 (see Figure 3.3 for the magnitude). This is
not exactly true (see [34] for discussion) but such assumption introduces errors in
the solution only at very short instants after voltage application provided that
equilibrium concentrations are set to the values corresponding to low space charge
densities, which are not able to distort electrostatic field in the system. The con-
centrations of electrons and positive ions change to non-equilibrium distributions
94 Lightning electromagnetics: Volume 2
during several initial time steps and, also, negative ions density rises due to strong
attachment in low field regions (Figure 3.5a). Actually, the provided initial values
are used in the model just for initializing the simulation process and to compute
quantities for the first time step.
Further, the solution obtained on a current time step is utilized as initial value
when calculating densities and potential distributions for the next step and, therefore,
the specified initial state of the system is “forgotten” after several successive steps.
Boundary conditions are essential for correct representation of a real physical
picture in the model. When the computational domain is limited by physical
boundaries (like in case of coaxial cylinders), one should focus on and implement
actual processes taking place on gas–metal interfaces that may include injection of
charges, their absorption/neutralization, etc. In most of the cases, a computational
domain is limited also by artificially introduced boundaries and choice of condi-
tions there becomes non-trivial. An example is shown in Figure 3.8 for rod-plane
geometry represented in axially symmetric domain. When setting boundary con-
ditions for this case, one should keep in mind that (3.3)–(3.5) are coupled. Hence, if
a positive potential is applied to the rod (Figure 3.8b), a simplest set of boundary
conditions includes “Convective flux” for electrons (Figure 3.8a) and negative ions
and “Concentration” equal to zero for positive ions on the boundaries representing
surface of the rod. By specifying convective fluxes, one insures continuity of the
current flow between electrodes. Opposite is valid for the grounded plane (hor-
izontal boundary on the bottom): zero concentrations of electrons and negative ions
and convective flux of positive ions are specified here. Note that injection of
negative charges from the plane and positive ions from the needle is prevented
when using such conditions. The left vertical boundary between the plane and the
needle represents physical symmetry of the system and “Axial symmetry” is spe-
cified here for all equations. Finally, the choice of conditions for the open boundary
(e.g., the rightmost vertical one) should be made taking into account field behavior
which controls in-coming and out-coming fluxes of charged particles. Thus, if one
sets “Zero charge/Symmetry” for the potential here (which is natural), the field
component normal to the boundaries will be set to zero and there will be no fluxes
of charge carriers crossing the boundary that doesn’t reflect the real situation. In
reality, charges existing in the gas are attracted/repelled to/from high field region
and continuously move across the boundary. This motion is a result of the forces
associated with the field directions others then tangential (parallel) to the boundary
and can be modeled by, e.g., specifying a grounded surface there. However, this
may alter field distribution in the region of interest located around the rod tip. Thus,
a compromise between the two approaches is to be found. One way is to increase
the size of the domain allowing artificial boundaries to be located far from the
region of interest (surface of the rod) and to assign “Zero charge/Symmetry” for the
potential and “Normal fluxes” for charge carriers. The effect of the external
boundaries on the field distribution and fluxes of charges is minimized in this case,
but the size of the model can rise significantly. Another option is to apply
“Distributed impedance” conditions (available in AC/DC module) for the electro-
static equation (3.5) and “Convective flux” for (3.3).
Numerical simulations of non-thermal electrical discharges in air 95
open boundary
open
rod
rod open
plane
(a)
open boundary
open
rod
open
rod
plane
(b)
(a)
(b)
(c)
In general, one can run the model at this stage. However, most likely such an
attempt would not be successful due to specific features of the system (3.3)–(3.5)
already mentioned above. From mathematical point of view, it is very stiff (in a
sense that it is characterized by different time-scales and source terms for different
kinds of charge carriers) and strongly non-linear (most of the terms are field-
dependent). In addition, the velocities of the carriers in strong fields become
extremely high that leads to their strongly drift-dominated flows with very high
local Peclet numbers, which is a dimensionless number describing the relationship
between the convective and diffusive terms in the convection–diffusion equations.
These peculiarities make the discretized problem to be inherently very unstable that
results in non-physical oscillations in the solution and even in negative con-
centrations magnitudes, primarily in regions where steep gradients of the charge
densities are present. The oscillations can even be large enough to affect con-
vergence of the solution. To resolve these problems, special algorithms or stabiliza-
tion methods should be used. The software provides several stabilization techniques
which can be selected for each (3.3) via Physics ! Subdomain settings ! Artificial
diffusion (Figure 3.10). The experience showed that “Isotropic diffusion” with the
Tuning parameters ad defined in Constants (Figure 3.2) is one of the most appropriate
choices for the considered problem. Note that tuning parameters should not be too
high to avoid introducing large amount of artificial diffusion which may distort the
solution. This is especially important in presence of interfaces between different
materials where transport coefficients change abruptly.
After completing all the steps above, one may run solver and obtain solutions
for output times specified in Solver Parameters (Figure 3.9a). The solution progress
is indicated in log-window showing current time and some auxiliary information
from the solver. In addition, the convergence of the solution can be traced when
using iterative solvers. After successfully completing the whole specified time
interval, the software automatically switches to a post-processing mode where a
wide range of graphical tools for visualization and analysis of the results is avail-
able to create 1D-, 2D-, and 3D-plots, time dependencies, performing integration
and differentiation of different quantities, etc. These facilities are available via Post
Processing ! Plot parameters in the main menu. The post-processing tools allow
also combining different kinds of plots to facilitate the analysis of the data.
×1015
1.035
1
0.95
0.9
0.85
0.8
0.75
0.7
0.65
0.6
0.55
0.5
0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
vicinity of the corona electrode due to rising field strength and, consequently,
ionization intensity. However, the levels of the electronic concentrations are much
lower than the densities of the positive ions, Figure 3.12b, at the corresponding
voltages (time instants in the legend). The reason is that the electrons are lost on the
surface of the corona electrode where convective flux boundary condition is spe-
cified and due to their attachment to electronegative components of air. The latter
gives rise to increased densities of negative ions, Figure 3.23c, which however,
remain even lower than those for electrons. As a result, the positive space charge is
dominating in the entire discharge space, Figure 3.12d, except a very thin layer
around the internal cylinder on the surface where the concentration of positive ions
is set to zero. This picture is natural for DC coronas where the external drift region
is always filled with ions whose charge sign is similar to the sign of the potential
applied to the corona electrode.
The space charge density in the inter-electrode space becomes significant at
certain voltage level (which slightly exceeds corona inception voltage) and the
electrostatic field is distorted by the field generated by the space charge. This can
be seen in Figure 3.13a where distributions for several time instants (applied vol-
tage magnitudes) are shown. Thus, the curves for t = 100, 200, 300, and 400 s
(corresponding voltages are 20 kV, 40 kV, 60 kV, and 80 kV) are simply shifted
upwards towards higher field levels due to the increasing voltages while their shape
defined by the geometry of the system is preserved. However, at longer times
Numerical simulations of non-thermal electrical discharges in air 101
12 14
×10 ×10
10 10 0
100
9 9 200
300
400
–3
8 8 500
600
Concentration electrons, m
700
7 7 800
900
6 6 1000
5 5
4 4
3 3
2 2
1 1
0 0
0.005 0.006 0.007 0.008 0.009 0.01 0.011 0.012 0 0.05 0.1 0.15 0.2 0.25
(a) distance from wire, m (b) distance from wire, m
12
×10 ×10–4
0
2 100
1.4 200
300
1.8 400
–3
1.2 500
Concentration negative ions, m
1.6 600
3
Space charge density, C/m
700
800
1.4 1 900
1000
1.2
0.8
1
0.8 0.6
0.6 0.4
0.4
0.2
0.2
0 0
0.005 0.006 0.007 0.008 0.009 0.01 0.011 0.012 0 0.05 0.1 0.15 0.2 0.25
distance from wire, m distance from wire, m
(c) (d)
Figure 3.12 Time variations of the distributions of the densities of electrons (a),
positive ions (b), negative ions (c), and resulting space charge (d).
Arrows indicates increase of the concentrations with time (voltage).
Note that variations of the quantities at instants 0–400 s are not
visible with the linear scale used
(= higher voltages), the main changes in the distributions appear in the external
corona region (drift region) due to the accumulated space charge. The field strength
increases here in accordance with the growth of the space charge density,
Figure 3.12d. At the same time, the field remains constant on the surface of the
corona electrode, Figure 3.13b. This fact is in agreement with numerous experi-
mental findings (see, for example, [33] and the books mentioned at the beginning of
the section). From Figure 3.13b, one can identify the voltage level at which space
charge dominated corona mode appears and which is normally identified as corona
inception voltage. For the considered case of d0 = 1 cm, it is 83 kV that is in
agreement with the experimental data given in [38]. The constant field on the surface
of the corona electrode reflects the fact that the discharge is self-controlled. The
surface field in Figure 3.13b is 42 kV/cm and it only slightly exceeds the ionization
102 Lightning electromagnetics: Volume 2
×106
4.5 0
100
200
4 300
400
500
Electric field strength, V/m
3.5 600
700
3 800
900
1,000
2.5
1.5
0.5
0
0 0.05 0.1 0.15 0.2 0.25
(a) distance from wire, m
×107
1 SC controlled
electrostatic
0.9
0.8
Electric field strength, V/m
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 100 200 300 400 500 600 700 800 900 1,000
(b) Time, s
Figure 3.13 Time variations of electric field distributions (a) and the field
strength on the corona electrode surface (b)
threshold of air (31 kV/cm). This relatively low field (as compared with the elec-
trostatic one) provides rate of generation of ions needed just for compensation of
ionic losses and results in a steady-state corona at the given voltage.
Any computational model should be validated against experimental data. For
the considered case, voltage–current characteristics of the corona discharge
dependencies in [38] are used here as reference data for verification of the
Numerical simulations of non-thermal electrical discharges in air 103
Here, d is the relative air density a defined in the Expressions, Figure 3.2.
Formulas (3.7) and (3.8) are empirical and have been validated in numerous studies. As
one may see from (3.7), the corona current is proportional to V2 and it is meaningful
only at voltages exceeding Vi. In addition, the magnitude of the current is affected by
the value assigned to the mobility of positive ions mp. Actually, the mobility is the only
parameter to be chosen in (3.7) for given dimensions of the electrodes. The results of
the calculations with (3.6) and (3.7) are shown in Figure 3.14 together with the
4,500
4,000
3,500
3,000
Current, mcA/m
d0=0.64cm
2,500
2,000
1.585cm
1,500
2.54cm
1,000
500
0
50 100 150 200
Voltage, kV
experimental data. One can notice an agreement between the calculated and measured
corona characteristics. However, the best fit between the measured results and calcu-
lations with (3.6) and (3.7) is found with slightly higher value of mp = 2.8cm2/Vs than
that corresponding to typical mobility of positive ions in air. This difference can be
attributed to the special conditions of experiments [38] where the inter-electrode gap
was irradiated with a radioactive b-source containing 90Sr of strength 0.4 mCi to
achieve glow (streamer-less) mode of the corona discharge. The simulations performed
with the developed model are also agree well with the experiments, as it is seen in
Figure 3.14, and the best fit in this case is obtained with a more realistic value for the
mobility of positive ions mp = 2.3 m2/Vs. The fact that both analytical calculations and
numerical simulations provide results which are slightly different from experimental
ones demonstrates the importance of selecting reliable material characteristics for
discharge modeling from available data.
dr /2
1m
Co1
1.2 m
limited by the symmetry axis and surface of the rod on the left, horizontal line on
the bottom representing surface of the plane, and by two artificially introduced
boundaries on the top and on the right-hand side. A positive potential is applied to
the rod and the bottom plane is grounded. It is assumed that the potential changes
linearly with time from 0 to 300 kV during 1,000 s (similar to the previous case, but
with increased ramping rate 300 V/s).
The boundary conditions for the charge carriers on the electrodes are similar to
the ones used for the coaxial system above: if the signs of the charged particles
correspond to the sign of the potential on the surface, zero concentrations are
specified; otherwise, “Convective flux” conditions are used. On the open bound-
aries, “Zero charge/Symmetry” is set for the potential and inward fluxes of charge
carries are defined as normal convective fluxes. The latter condition is not strict
because the fluxes are controlled here by the direction and magnitude of the electric
field as it was discussed above. The model settings for the internal part of the
domain are similar to the ones used in the simulations of the concentric electrodes.
The computational mesh for the system in Figure 3.15 should be refined at the tip
×1014
2.5
2.375
2.25
2.125
1.875
1.75
1.625
1.5
1.375
1.25
1.125
0.875
0.75
0.625
0.5
0.375
0.25
0.125
Figure 3.16 Profile of the concentration of positive ions at 300 kV (the region
55 cm2 at the rod tip is shown)
106 Lightning electromagnetics: Volume 2
of the rod and along the symmetry axis, where most intensive discharge activity is
expected. This is done using meshing facilities provided in the software. An
example of the results obtained from the simulations is presented below for the case
of dr = 38.1 mm.
The distribution of the concentration of positive ions in the corona discharge at
the applied voltage of 300 kV is shown in Figure 3.16. One can observe that the
ionic cloud is narrow and extends into the bulk of the gap in contrast to the case of
corona around cylindrical conductor in Figure 3.11. Such shape was observed in
many experiments and it is the consequence of the distribution of the electric field
which is strongest along the symmetry axis. The maximum of the density of the
ions is located close to the rod tip and under the conditions of [40] reaches
2.51014 m3.
The time variations of the concentration profiles of charge carriers along the
symmetry axis are shown in Figure 3.17. One may see that the densities of negative
12
×10 ×1014
2.5 2.5 0
100
200
300
400
Concentration positive ions, m–3
2 500
2
Concentration electrons, m–3
600
700
800
900
1.5 1.5 1000
1 1
0.5 0.5
0 0
0 0.5 1 1.5 2 2.5 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(a) distance from rod, m (b) distance from rod, m
×1011 ×10–5
6 4 –5
0
×10
4 100
200
3.5 300
5 3.5 400
Concentration, negative ions, m–3
3
500
Space charge density, C m–3
2.5 600
3 700
2
4 800
900
2.5 1.5
1000
1
3 2 0.5
0
0 0.005 0.01 0.015 0.02 0.025 0.03
1.5
2
1
1
0.5
0 0
0 0.5 1 1.5 2 2.5 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(c) distance from rod, m (d) distance from rod, m
Figure 3.17 Time variations of densities of charge carriers along symmetry axis:
(a) electrons, (b) negative ions, (c) positive ions, and (d) resulting
space charge. The arrows indicate changes at longer instants (higher
voltages). The inset in (d) shows the distributions of the space charge
density in close vicinity of the rod tip
Numerical simulations of non-thermal electrical discharges in air 107
×106
4 0
100
200
3.5 300
400
500
3 600
Electric field strength, V/m
700
800
2.5 900
1,000
1.5
0.5
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(a) distance from rod, m
7
1.1 ×10 electrostatic
SC controlled
1
0.9
0.8
Electric field strength, V/m
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 100 200 300 400 500 600 700 800 900 1,000
(b) Time, s
Figure 3.18 Time variations of electric field distributions (a) and the field
strength on the corona electrode surface (b)
108 Lightning electromagnetics: Volume 2
carriers are much lower than the positive ones and they are localized in a thin
(<2 mm) layer at the rod surface forming the corona region (Figure 3.17a and c).
The drift region of the discharge covers practically the entire gap (Figure 3.17b). As
a result, the whole space in front of the rod is positively charged except of a small
volume around the tip of the rod where negative charges are dominating
(Figure 3.17d). The space charge density increases with time (i.e., applied voltage),
however, the shape of the profiles remains identical in the main part of the drift
region at the distance from the rod larger than 1 cm (the curves are just shifted
upwards and are parallel). Such behavior of the space charges leads to the situations
when the external field is almost completely screened out and the resulting dis-
tributions (Figure 3.18a) are characterized by extremely low field strength in the
middle of the gap (at distances from the rod tip 0.2–0.3 m) and its enhancement at
the surface of the plane. The field level at the tip of the corona electrode is the
highest and it remains constant after corona inception (Figure 3.18b). This is
similar to the case considered above for cylindrical system, however, the field
magnitude is slightly lower (compare Figure 3.13b and b). Summarizing, one may
state that corona discharge is a phenomenon completely controlled by space char-
ges produced in the corona region which define its internal as well as external
properties.
The corona current calculated using expression (3.6) is presented in
Figure 3.19 as a function of the applied voltage for three sizes of the diameter of the
corona rod together with the experimental results [40]. One may note that the best
agreement between the computed and measured data is achieved for dr = 50.8 mm
that is the largest size of the rod for which the simulations were performed. For the
250
dr= 38.1mm
200
dr= 25.4mm
Current, mcA
150
dr= 50.8mm
100
50
0
50 100 150 200 250 350
Voltage, kV
where cmax = 2 (cmTorr) 1 and cmin = 3.5102 (cmTorr) 1 are the maximum
and minimum absorption cross-sections of O2 molecule in the interval of wave-
lengths mentioned above, and pO2 is the partial pressure of oxygen in air.
The rate of the process (3.9) is to be incorporated in the expressions (3.4) for
the total rates of generation and losses of charged species. However, direct use of
(3.9) in numerical simulations is extremely costly. By its nature, photoionization is
a non-local source of charge carriers in discharge plasma meaning that productions
of ionizing agents (photons) and ionic pairs take place at different locations in
space. Thus, the intensity of photoionization in each node of the computational
mesh is to be calculated by summing up contributions from all other nodes. This
leads to a necessity to create a full matrix containing Rph values, which should be
stored in computer memory (RAM) during solution process and should be updated
on each time step. This situation can be avoided by considering the fact that the rate
of photoionization in (3.9) is dependent on the intensity of electron impact ioni-
zation (physically, it is related to a large number of excited molecules in the regions
with strong fields). Hence, the integration (3.9) can be performed only within
regions where the ionization rate Rimp is high, e.g., at the streamer head. Moreover,
the core of the integral can be approximated in some way and efficient techniques
for numerical calculations can be successfully applied as it is proposed in, e.g.,
[42]. Recently, significant progress in computing photoionization in gas discharge
related problems was achieved by applying radiation transfer theory [43–46].
Simulations of propagating streamer requires an extremely fine computational
mesh to accurately resolve charged layers in the vicinity of electrodes and in
regions with steep gradients of the densities of charge carriers. This can be
achieved by implementing static or dynamic meshes. The former may lead to
unnecessary increase of the size of the problem and even make it unsolvable if
large inter-electrode distances and higher spatial dimensions (2D or 3D) are to be
Numerical simulations of non-thermal electrical discharges in air 111
considered. A dynamic mesh is the most suitable option and it should be imple-
mented in a way allowing for an adaptive mesh refinement (AMR) in the regions of
strong variations of the carriers densities. However, implementation of AMR for
the streamer problem is not trivial and requires special consideration, particularly in
case of positive (cathode-directed) discharges where the velocity of the ionization
wave (streamer front) and the drift velocity of secondary electrons have opposite
directions. Examples of 3D simulations performed with finite-differences (FD) and
finite-elements (FE) methods using AMR can be found in [47,48].
Another computational challenge is related to the numerical algorithm for
solving drift-diffusion equations (3.3), which should be very accurate and provide
monotonic positive solutions for concentrations profiles free of unphysical features
like oscillations, ripples and numerical diffusion. Conventional high-order algo-
rithms (second order and above) being applied to (3.3) introduce dispersive ripples
in the solution, particularly at locations of steep gradients. At the same time, low
order schemes, such as donor cell, Lax-Friedrichs, etc. or high order schemes with
zero order diffusion added, produce no ripples, but suffer from excessive numerical
diffusion. Since a discharge plasma channel propagating in a neutral gas can be
considered as a kind of shock wave (ionization wave is the term commonly used in
the literature), it is natural that the best alternative would be to use one of the high-
order methods especially designed for shock capturing, such as Godunov’s method,
explicit two-step MacCormack’s method, Van Leer’s approach, etc. Discussions on
numerical schemes and techniques used in gas discharge problems can be found in
[49–52]. Among these methods, a finite-difference version of the so-called flux-
corrected transport (FCT) [53,54] has been extensively studied, tested and applied
for simulations of streamers, see e.g. [55,56]. A finite-element implementation of
FCT was developed as well [12,57,58], but has not become popular and it is still a
subject of research.
The FCT belongs to the class of non-linear flux limiter methods. It is essential
to note that FCT is not a numerical scheme, but it is rather a technique consisting of
several stages. FCT constructs the net transportive flux G (the product of particle
concentrations and their drift velocity G = nw) point by point on a computational
mesh as a weighted average of a flux computed by a low order scheme and a flux
computed by a high order scheme. The weighting is dependent upon the local
density profile (non-linearity) and it is done in a manner, which insures that the
high order flux is used to the greatest extent possible without introducing over-
shoots and undershoots in the solution. This weighting procedure is referred to as a
“flux-correction” or “flux-limiting.” The result is a family of transport algorithms
capable of resolving moving contact discontinuities over 3–4 grid points, and shock
fronts over 2 grid points, without overshoots or undershoots [54]. Formally,
application of the FD-FCT procedure to a one-dimensional continuity equation of
type (3.3) includes the following actions [53] (fully multidimensional FCT
approach has been developed in [59]):
1. Compute low order fluxes G Li1=2 and G Liþ1=2 given by some low order scheme
to guarantee monotonic (ripple-free) results.
112 Lightning electromagnetics: Volume 2
G i1=2 G i1=2 .
H L
Here, index i stands for the current position on the computational grid, and
i1/2 indicates the positions of the interfaces of the neighbouring grid cells;
indexes m and m+1 are used for the current and the following time instants,
respectively; .. is the current spatial step on the grid; D t is the time step; C is the
weighting function. One can observe that if ACiþ1=2 ¼ Aiþ1=2 , i.e. in the absence of
the flux limiter, the solution nmþ1 will simply be the time-advanced high order
solution. Flux correction (step 5) is the critical stage in the FCT procedure. In the
original algorithm [53], it is proposed to be implemented as:
(3.11)
(3.12)
whichever is larger. The “raw” antidiffusive flux Aiþ1=2 always tends to decrease
nm m
i and to increase niþ1 . Thus, the flux-limiting formula ensures that the corrected
flux cannot push ni below nm
m
i1 , which would produce a new minimum, or push
nmiþ1 above nm
iþ2 , which would produce a new maximum. The equation above is
constructed to take care of all cases for signs and slopes.
The FD-FCT technique for integrating continuity equations outlined above has
been implemented in a library of FORTRAN routines [60], which can treat one-
dimensional Cartesian, cylindrical, spherical, or generalized coordinates on uni-
form or non-uniform Eulerian or Lagrangian grids. The procedure can be extended
to higher spatial dimensions using time step splitting method.
As it was demonstrated in the examples of simulations of corona discharges,
the continuity equations (3.3) are to be solved in conjunction with Poisson’s
equation (3.5) to obtain electric field distribution in the domain. In case of
streamers, the concentrations of charge carriers are high, especially at the strea-
mer head, and the field in the inter-electrode space is fully controlled by the
generated space charge. The distribution of the field reflects structure and
Numerical simulations of non-thermal electrical discharges in air 113
Elapsed time
@ 2 j @ 2 j 1 @j
þ þ ¼ R ðz; rÞ (3.13)
@ z2 @ r2 r @ r
Here, for simplicity e = 1, and R stands for the space charge density divided with
e0. Following the steps described above, the physical domain containing a curved
boundary (Figure 3.21) can be mapped into a rectangle 0 q 1, 0 x 1
using relations q = r / Sr and x = z / f(r), where q and x are the axial and radial
coordinates in the computational domain, and f(r)= zmax = Sz+F(r) is the geometry
factor. The function F(r) determines the shape of the boundary and it is given in
Numerical simulations of non-thermal electrical discharges in air 115
Z ξ
F(r)
1
Sz
0 Sr r 0 1 θ
Table 3.1 for a variety of cases. After transformation, (3.13) in the new rectangular
domain (q,x) becomes
@j @j 1 @j
Er ¼ ¼xG (3.16)
@r @ x Sr @q
and
In general, obtaining the potential distribution from (3.14) seems to be more
complicated than by solving the original equation (3.13), mostly due to the second
term containing cross-derivatives. However, the problems are, in general, of similar
complexity and (3.14) can be solved effectively with an appropriate software
package, e.g., MUDPACK [68]. This software is a collection of FORTRAN codes
for automatic discretization and computing second- and fourth-order finite differ-
ence approximations to the elliptic PDEs using multigrid method (MG), which is a
fast iterative method based on the multilevel or multi-scale paradigm [69]. It can be
applied in combination with any of the common discretization techniques and it
does not depend on the separability or other special properties of the equation.
Extensive information about MG and available software can be found in [70].
Typical flowchart for solving the coupled system of (3.3)–(3.5) with the pre-
sented FD techniques is shown in Figure 3.22. The input parameters for initializing
calculations are: the dimensions of the physical domain, gas pressure, potentials of
the electrodes, number of nodes in the computational grid, desired output times, the
initial distribution of space charges (if any). The routine DRIVER calculates data
INPUT
reads input parameters
OUTPUT
DRIVER
writes ne(z, r), np(z, r),
E(z, r), I(t), etc.
POISSON
solves field problem
END
TRANSPORT
solves contimity equations
POISSON
Yes
t < tout
No
Figure 3.22 Flowchart of the algorithm for solving the coupled equations (3.3)–(3.5)
Numerical simulations of non-thermal electrical discharges in air 117
needed for the main solvers and calls subroutine POISSON, which solves the
Poisson’s equation (3.5). The output from it is the distribution of the electric field in
the discharge gap (after the first call it gives the electrostatic field distribution if no
space charge was set initially). Then the routine TRANSPORT is called for solving
system (3.3) and (3.4). This routine calculates the time step limited by Courant–
Friedrichs–Lewy condition, advances the profiles of the densities of charged par-
ticles taking into account the source terms and computes the space charge density
profiles using the electric field distribution obtained earlier. Then, the current time
is updated and the routine POISSON is called again to calculate the electric field
corresponding to the new space charge distribution. If the current time is not greater
than the desired output time, the loop TRANSPORT – POISSON is repeated until
the condition is fulfilled. The output includes microscopic (the particle densities,
the electric field and potential distributions, etc.) and macroscopic (the electric
current, electrostatic energy, Joule dissipation, etc.) discharge characteristics. Note
that the algorithm is different from the one used for the simulations of corona
discharges above, where the discretized PDEs are solved simultaneously. In gen-
eral, both approaches are valid as far as the correct sequence of calculations is
utilized in the latter one.
according to electrostatic calculations the field strength reached 200 kV/cm, and
it propagated further into the gap in a constant field Eb which was attained on the
distance of 0.4 cm from the anode. Thus, the streamer passed 3 cm in the
uniform weak field Eb the magnitude of which was varied in the simulations in
order to observe its effect on streamer dynamics. The problem domain was similar
to that shown in Figure 3.21 (left) which was transformed into a rectangular domain
(Figure 3.21, right) utilizing the technique discussed above. The model used was a
quasi two-dimensional one (1.5D) with the fixed radius of the discharge channel
Rc = 200 mm and, hence, uniform distributions of charge carriers’ densities within
its cross-section were assumed. The value of Rc used can be found as an overage
magnitude obtained by fully two-dimensional simulations for weak fields, e.g. [74],
and it is in agreement with experimental observations (a discussion regarding
streamer radius can be found in [61]). The field dependencies of the rate coeffi-
cients in (3.3) and (3.4) were taken from [16] and the rate of photoionization was
computed by applying (3.9) and (3.10) on the symmetry axis of the gap along
which the discharge propagation takes place. The ionic diffusion was neglected due
to its insignificance in the time scale of streamer development. The boundary
conditions for the set of PDEs (3.3) included zero densities of positive ions on the
anode and negative ions on the cathode np(Sz,t) = nn(0,t) = 0; the densities of
electrons were set to zero on surfaces of both electrodes ne(0,t) = ne(Sz,t) = 0.
Symmetry (Neumann) conditions were applied for all charge carriers on the other
boundaries. For (3.5), the potentials fc and fa were specified on the cathode and
anode, respectively, and symmetry condition was used on the boundary represent-
ing physical symmetry axis. On the outer boundary of the computational domain,
the linear potential distribution was set. The magnitudes of fc and fa and the
distribution on the outer boundary were adjusted in a way to provide the distribu-
tion of the electrostatic field in the domain close to that appeared in the experiments
[73]. The latter was justified by comparing the fields obtained by the presented
method with the results of calculations performed using commercial finite-element
software.
Solution of the set of (3.3) requires initial conditions for the densities of charge
carriers. One of the approaches utilized in streamer simulations is based on the
assumption that a quasi-neutral layer or “plasma spot” with relatively high charge
density exists in the high field region. This method allows avoiding modeling the
avalanche stage and streamer inception can be achieved during several time steps at
the beginning of the simulations. However, it appears to be artificial because the
location of initial charges and their densities are set arbitrary. A more physical
approach is presented in [75] and it employs the fact that collisional detachment of
electrons from negative ions is effective in air at atmospheric conditions. The rate
of the process is R0 = ndet nn0, where ndet is the detachment frequency and
nn0 103 cm3 is the equilibrium background concentration of negative ions. The
magnitude of ndet is determined as ndet = k nel exp(Wa/Wch), where nel is the
frequency of elastic collisions, Wa is the electron affinity, Wch is the chaotic energy
of negative ions, and k = 0.1 is the similarity coefficient. Assuming negative ions of
type O 3 [75], the following magnitudes of the parameters can be used for
Numerical simulations of non-thermal electrical discharges in air 119
evaluations: Wch = M(mn0E)2/2, where the mobility mn0= 2.55 cm2V1s1 and the
mass of the ion M = 5.31023; Wa = 2.05 eV; nel = 0.791010 s1. The rate R0
appeared to be field dependent and, e.g., for the field strength at the tip of the
needle E = 200 kV/cm, it is R0 51011 cm3s1 that is considerably lower than
the magnitudes of other terms in Re (3.4) and its influence during streamer devel-
opment is negligible. However, the term R0 is important during streamer initiation
stage for providing increased concentrations of seed electrons in the regions with
enhanced fields and facilitating streamer inception.
The problem was solved on a uniform structured computational mesh
16 mm 20 mm using the flux-corrected transport technique for (3.3) and (3.4)
implemented in [60] and multigrid solver [68] for (3.14). The time step Dt was
limited by the Courant criterion c = Dt|we|/Dz = 0.3 (here, Dz = 16 mm is the space
resolution).
The results of the simulations can be seen in Figure 3.23, where time variations
of the electron density profiles along the axis of the discharge at different magni-
tudes of the background field are presented. As one can observe, streamer inception
takes less than one nanosecond after voltage application. The density of electrons
increases during the formation stage up to 1014 cm3 (see curves for t = 1 ns) due
1015 1015
Electron density, cm–3
Electron density, cm–3
1013 64 1013
40 24 16 8 4 1 30 20 12 8 4 1
1012 1012
0 1 2 3 0 1 2 3
z, cm z, cm
1015 1015
4.3 kV/ cm 5 kV/ cm
Electron density, cm–3
1014 1014
20
1013 29 1013
24 16 12 8 4 1 16 12 8 4 1
1012 1012
0 1 2 3 0 1 2 3
z, cm z, cm
to strong field at the needle tip. During propagation stage, the streamer enters the
region with the weak uniform field in the bulk of the gap, where conditions for its
development are dependent upon the magnitude of Eb. Thus if the background field
is low, the losses of electrons exceed their production and the discharge can be
terminated as it is seen in the plots for 3 and 3.7 kV/cm. The electron densities at
the discharge front remain practically constant if propagation takes place in stron-
ger background fields of 4.3 and 5 kV/cm, indicating that the losses are compen-
sated by the production of electrons at the streamer head. In the streamer channel
(especially in the anode region), the concentrations of electrons decrease with time
continuously due to attachment and recombination. This, however, leads to just
minor variations of the conductivity s of the discharge plasma (Figure 3.24), which
is defined as a sum of contributions of all kinds of charge carriers s = q (neme + nnmn
+ npmp). One may notice that the conductivity of the channel remains on the level of
0.2 S/cm and 0.3 S/cm during propagation in the fields Eb = 4.3 kV/cm and
5 kV/cm, respectively, while it decreases rapidly in weaker background fields. The
values of the discharge plasma conductivity obtained from the simulations agree
well with known data [6,61].
As it was mentioned above, streamers are essentially ionization waves propa-
gating in an insulating medium. Conversion of a neutral gas into conductive plasma
102 102
Conductivity, S/cm
101 101
100 100
10–1 10–1
0 1 2 3 0 1 2 3
z, cm z, cm
102 102
4.3 kV/ cm 5 kV/ cm
Conductivity, S/cm
Conductivity, S/cm
101 101
100 100
10–1 10–1
0 1 2 3 0 1 2 3
z, cm z, cm
takes place at the streamer head, where extremely strong electric fields appear due
to produced space charges as it is seen in Figure 3.25. One may observe that the
field strength is low behind the discharge front due to high conductivity of plasma
in the channel and also between the front and the cathode (located at z = 0) where it
is equal to Eb. Depending on the magnitude of the background field, the peaks
associated with the streamer head may decrease during its propagation, as it hap-
pens at Eb = 3, 3.7, and 4.3 kV/cm, or remain constant if the field is strong enough
to support discharge development, see graph for Eb = 5 kV/cm. The latter indicates
a stable mode of the discharge. One should note, however, that under considered
circumstances, the streamer is also able to cross the gap between electrodes even at
lower magnitude of Eb, often referred as crossing field (see plot for 4.3 kV/cm),
although this depends upon conditions on early stages of its formation in the vici-
nity of the needle (field strength, accumulated charge, acceleration, etc.). When the
streamer front approaches the cathode region, a cathode-streamer head interaction
can be observed, which is associated with an increase of the maximum of the
electric field strength seen in the plots for Eb = 4.3 and 5 kV/cm. This field
enhancement leads to intensive ionization and to corresponding rise of the electron
density and conductivity shown in Figures 3.23 and 3.24, respectively. It has been
argued in [76] that this phenomenon is conditioned by a release of electrostatic
energy accumulated in a streamer channel – cathode system.
250 250
3 kV/cm 3.7 kV/cm
Field strength, kV/cm
Field strength, kV/cm
200 200
150 150
100 100
50 50
0 0
0 1 2 3 0 1 2 3
z, cm z, cm
250 250
4.3 kV/cm 5 kV/cm
Field strength, kV/cm
200 200
150 150
100 100
50 50
0 0
0 1 2 3 0 1 2 3
z, cm z, cm
4.3
5.0 3.7 3.0
3
Streamer length, cm
0
0 20 40 60 80
Time, ns
(a)
200
5.0
Streamer velocity, cm / μs
150
4.3
100
3.7
50
3.0
0
0 1 2 3
Streamer length, cm
(b)
Figure 3.26 Positive streamer length as function of propagation time (a) and
streamer velocity as a function of its length (b).The magnitudes of the
background field strength in kV/cm are indicated at the curves (b)
Numerical simulations of non-thermal electrical discharges in air 123
170 kV/cm to 140 kV/cm that leads to the decrease of velocity from 1.6108
cm/s to 1.1108 cm/s. In weaker background fields, the velocity decreases with
streamer length in the same linear manner as the field at its tip (se Figure 3.25). The
discharge front acceleration in the cathode region (streamer length > 3 cm) also
corresponds to the enhancement of the maximum field. It is necessary to note here
that the magnitudes of the streamer velocities obtained in present simulations are
slightly overestimated as compared with the results of fully two-dimensional
simulations [74] due to neglected radial expansion of the discharge channel.
Finally, one should mention that propagation of positive streamers in weak
fields is governed by the electrostatic energy Q accumulated in the system (Q is the
volume integral of the energy density e0E2/2). An analysis of the computed dis-
tributions of the energy along the discharge channel showed that the stable streamer
propagation is associated with a slight increase of the accumulated energy. In
contrast to this, the magnitudes of Q decrease if the propagation takes place in the
background fields weaker that the stability threshold.
Figure 3.27. As it can be seen, streamer inception takes place in the high field
region at the cathode (z > 3 cm), where the electron densities rise quickly and reach
1014 cm3 within 1 ns, similarly to the case of positive discharges (Figure 3.23).
Different behavior can be observed when plasma front enters the balk of the gap
where the background field is weak. In case of negative streamers, the minimal
crossing field is found to be 7 kV/cm and discharge propagation in this field is
associated with the decrease of the electrons density. In the magnitudes of Eb are
lower than that, a streamer is terminated rapidly in the middle of the gap as it is
observed for the strength of 5 and 6 kV/cm in which the discharge propagates for
1.9 and 2.9 cm, respectively. The concentration of electrons is kept constant if the
background field strength is raised to 8 kV/cm. In this case, loses of electrons in
discharge plasma are compensated by the higher rate of their production, which is
dependent on the local electric field strength at the streamer front. One can see in
Figure 3.28 that the field strength at the head decreases rapidly when streamer
termination takes place (graphs for 5 and 6 kV/cm) because the space charge pro-
duced is not strong enough. The situation is different when the discharge propa-
gates in the background field Eb = 8 kV/cm. In this case, the electric field
associated with the streamer head reaches 200 kV/cm and remains almost
1015 1015
Electron density, cm–3
Electron density, cm–3
5 kV/ cm 6 kV/ cm
1014 1014
1013 1013
8 4 1 8 4 1
28 16 40 28 16
1012 1012
0 1 2 3 0 1 2 3
z, cm z, cm
1015 1015
Electron density, cm–3
Electron density, cm–3
7 kV/ cm 8 kV/ cm
1014 1014
1013 15 12 8 4 1 1013 10 8 6 4 2 1
1012 1012
0 1 2 3 0 1 2 3
z, cm z, cm
250 250
5 kV/cm 6 kV/cm
150 150
100 100
50 50
0 0
0 1 2 3 0 1 2 3
z, cm z, cm
250 250
7 kV/cm 8 kV/cm
Field strength, kV/cm
150 150
100 100
50 50
0 0
0 1 2 3 0 1 2 3
z, cm z, cm
constant during propagation. Similar behavior was obtained for positive streamers,
where the peak field strength was found to be 170 kV/cm (see Figure 3.25). The
difference between these values is related to the different conditions needed to
maintain intensive production of secondary avalanches at the head of the negative
streamer as mentioned above. Note that the ionization coefficient a depends
exponentially on the field strength and, thus, even small increase of E results in a
significant amount of generated secondary electrons.
The length of the channel of the negative discharge increases linearly with time
during its stable propagation (Figure 3.29a, curve for 8 kV/cm), similar to the case
of the positive streamer (Figure 3.26a, curve for 5 kV/cm). However, in contrast to
the latter, the gap between the stability field and the minimal crossing field for
negative streamers is smaller and negative discharges seem to be more sensitive to
changes in the magnitudes of the background field strength than the positive ones.
The data presented in Figures 3.27–3.29 indicate that just a minor reduction of Eb
below 7 kV/cm may lead to rapid streamer termination. Another distinction of
negative streamers obtained from the simulations is that the velocity corresponding
to stable propagation 3.2108 cm/s (Figure 3.29b, 8 kV/cm) is two times higher
than that for the positive ones (compare with Figure 3.26b, 5 kV/cm) due to the
stronger field at the streamer front. Other features observed in Figure 3.29b, like
126 Lightning electromagnetics: Volume 2
3.5
2.5 8 kV/cm
Streamer length, cm
2 7 kV/cm
1.5 6 kV/cm
1 5 kV/cm
0.5
0
0 5 10 15
Time, ns
(a)
× 108
4.5
3.5
8 kV/cm
Streamer velocity, cm/s
2.5 7 kV/cm
2
6 kV/cm
1.5
1 5 kV/cm
0.5
0 0.5 1 1.5 2 2.5 3 3.5
Streamer length, cm
(b)
Figure 3.29 Negative streamer length as function of propagation time (a) and
streamer velocity as a function of its length (b). The magnitudes of the
background field strength in kV/cm are indicated at the curves (b)
acceleration at the very late stage when the streamer head approaches the anode
surface, and the linear decrease of the velocity with increasing length of the channel
in the background fields lower than the stability level, are similar to those discussed
above for positive discharges.
Numerical simulations of non-thermal electrical discharges in air 127
References
[1] I. Gallimberti, G. Bacchiega, A. Bondiou-Clergerie and P. Lalande,
Fundamental processes in long air gap discharges, C. R. Physique, Vol. 3,
pp. 1335–1359, 2000.
[2] E. Kuffel, W. S. Zaengl, and J. Kuffel, High Voltage Engineering:
Fundamentals, Newnes, London, 2nd ed., 2000.
[3] S. Pancheshnyi, Role of electronegative gas admixtures in streamer start,
propagation and branching phenomena, Plasma Sources Sci. Technol.,
Vol. 14, pp. 645–653, 2005.
[4] E. Nasser, Fundamentals of Gaseous Ionization and Plasma Electronics,
John Wiley & Sons Inc., New York, NY, 1971.
[5] K. L. Aplin, Instrumentation for Atmospheric Ion Measurements, Ph.D.
thesis, Univ. of Reading, 2000.
[6] Y. P. Raizer, Gas Discharge Physics, Springer Verlag, New York, NY, 1997.
[7] J. Dutton, A survey of electron swarm data, J. Phys. Chem. Ref. Data, Vol. 4,
No. 3, pp. 577–856, 1975
[8] J. W. Gallagher, E. C. Beaty, J. Dutton, and L. C. Pitchford, An annotated
compilation and appraisal of electron swarm data in electronegative gases, J.
Phys. Chem. Ref. Data, Vol. 12, No. 1, pp. 109–152, 1983.
[9] J. J. Raju, Gaseous Electronics: Theory and Practice, CRC Press, London,
2006.
[10] I. A. Kossyi, A. Yu Kostinsky, A. A. Matveyev, and V. P. Silakov, Kinetic
scheme of the non-equilibrium discharge in nitrogen-oxygen mixtures,
Plasma Sources Sci. Technol., Vol. 1, pp. 207–220, 1992.
[11] K. Becker, K. H. Becker, U. Kogelschatz, and K. H. Schoenbach (eds.), Air
plasma chemistry, in K. H. Becker, U. Kogelschatz, K. H. Schoenbach and
R. J. Barker (eds.) (eds.) , Non-equilibrium air Plasmas at Atmospheric
Pressure, IOP Publishing, Bristol, 2005.
[12] G. E. Georghiou, A. P. Papadakis, R. Morrow, and A. C. Metaxas,
Numerical modelling of atmospheric pressure gas discharges leading to
plasma production, J. Phys. D: Appl. Phys., Vol. 38 pp. R303–R328, 2005.
[13] H. Raether, The electron avalanche and its development, Appl. Sci. Res., Sec.
B, Vol. 5, No. 1, pp. 23–33, 1956.
[14] E. E. Kunhardt, Electron macrokinetics in partially ionized gases: the
hydrodynamic regime, Phys. Rev. A, Vol. 42, No. 2, pp. 803–814, 1990.
[15] R. Morrow and N. Sato, The discharge current induced by the motion of
charged particles in time-dependent electric fields; Sato’s equation extended,
J. Phys. D: Appl. Phys., Vol. 32, pp. L20–L22, 1999.
[16] R. Morrow and J. J. Lowke, Streamer propagation in air, J. Phys. D: Appl.
Phys., Vol. 30, pp. 614–627, 1997.
[17] M. C. Wang and E. E. Kunhardt, Streamer dynamics, Phys. Rev. A, Vol. 42,
No. 4, pp. 2366–2373, 1990.
[18] N. Y. Babaeva and G. V. Naidis, On streamer dynamics in dense media, J.
Electrostat., Vol. 53, No. 2, pp. 123–133, 2001.
128 Lightning electromagnetics: Volume 2
[66] H. Singe, H. Steinbigler, and P. Weiss, A charge simulation method for the
calculation of high voltage fields, IEEE Trans. Power App. Sys., Vol. PAS-
93, No. 5, pp. 1660–1968, 1974.
[67] J. C. Tannehill, D. A. Anderson and R. H. Pletcher, Computational Fluid
Mechanics and Heat Transfer, Taylor & Francis, New York, NY, 1997.
[68] J. C. Adams, MUDPACK: Multigrid software for elliptic partial differential
equations, National Center for Atmospheric Research, USA, 1998, http://
scd.ucar.edu/css/software/mudpack
[69] P. Wesseling, An Introduction to Multigrid Methods, John Wiley & Sons,
Chichester, UK, 1992.
[70] MGnet Repository, http://mgnet.org.
[71] N. L. Allen and A. Ghaffar, The conditions required for the propagation of a
cathode-directed positive streamer in air, J. Phys. D: Appl. Phys., Vol.28,
pp.331–337, 1995.
[72] L. Gao, A. Larsson, V. Cooray and V. Scuka, Simulation of streamer dis-
charges as finitely conducting channels, IEEE Trans. Diel. Elec. Insul.,
Vol.6, No.1, pp.35–42, 1999.
[73] M. Akyuz, L. Gao, A. Larsson, and V. Cooray, Streamer current in a three-
electrode system, IEEE Trans. Diel. Elec. Insul., Vol. 8, No. 4, pp. 665–672,
2001.
[74] N. Yu Babaeva and G. V. Naidis, Simulation of positive streamers in air in
weak uniform electric field, Phys. Lett. A, Vol. 215, pp. 187–190, 1996.
[75] A. F. Djakov, Y. K. Bobrov, and Y. V. Yurgelenas, Modeling of positive
streamer in air in non-uniform electric field, in A. F. Djakov (ed.) (eds.) ,
Physics and Technology of Electric Power Transmission, MPEI Publishers,
Moscow, pp. 161–200, 1998 (in Russian).
[76] I. Odrobina and M. Cernak, Numerical simulation of streamer-cathode
interaction, J. Appl. Phys., Vol. 78, pp. 3635–3642, 1995.
[77] N. Y. Babaeva and G. V. Naidis, Dynamics of positive and negative strea-
mers in air in weak uniform electric fields, IEEE Trans. Plasma Sci.,
Vol. 25, No. 2, pp. 375–379, 1997.
This page intentionally left blank
Chapter 4
Attachment of lightning flashes to
grounded structures
Vernon Cooray1
4.1 Introduction
A grounded structure can interact with a lightning flash in two different ways. It
can interact with either a downward or an upward lightning flash. The initiation of a
downward lightning flash takes place in the cloud, whereas in the case of upward
lightning flash, the point of initiation is usually at the tip of a tall structure. In other
words, upward lightning flashes are created by the grounded structure itself. In this
chapter, a brief description of various models used to study the lightning attach-
ment is given together with some of their predictions. A portion of the material
presented here is published previously in Refs. [1] and [2].
First, let us consider the events associated with the attachment of a downward
negative lightning flash (i.e. a lightning flash that transport negative charges to
ground) with a grounded structure. Experimental investigations show that a
downward lightning flash is initiated by a column of charge called the stepped
leader that travels from cloud to ground in a stepped manner. As the stepped leader
approaches the ground, the electric field at ground level increases steadily. The
electric field at the pointed tips of a grounded structure, which is immersed in this
background electric field, may reach values that are several times to several tens of
times the magnitude of the background electric field produced by the stepped lea-
der due to field enhancement. When the electric field at the tip of a structure
reaches a critical value of about 3.0 106 V/m, electron avalanches will be gen-
erated from the tip. As the background electric field and hence the local electric
field at the tip intensifies, the ionization taking place at the tip becomes more
vigorous leading to an increase in the number of charged particles in the head of the
electron avalanches. When this number reaches a value around 108 to 109, electron
avalanches will be transformed to a streamer discharge [3,4] (see also Chapter 1 of
this volume). This conversion of electron avalanches to a streamer, or the streamer
inception, is called avalanche to streamer transition. Once the conditions necessary
for streamer inception are satisfied, several streamer bursts will be issued from the
1
Department of Electrical Engineering, Uppsala University, Sweden
134 Lightning electromagnetics: Volume 2
point under consideration. These streamer bursts are generated from a common
stem and if the charge in the streamer burst is larger than about 1 mC, the streamer
stem will be thermalized leading to the creation of a leader [3]. This transition is
called streamer to leader transition. This leader, created by the action of the electric
field generated by the stepped leader, is called a connecting leader. Once incepted,
a connecting leader starts to grow towards the down-coming stepped leader. This
growth of the connecting leader is mediated by streamer bursts generated at its
tip. The charge associated with each streamer burst depends on the background
electric field and the potential gradient of the connecting leader channel. The
potential gradient of the connecting leader channel (which is positively charged in
this case) can be obtained by appealing to the thermodynamic model of the positive
leaders as described by Gallimberti [3]. Each streamer burst extends the leader by a
small amount. For example, if the charge in a streamer burst is Q, then the amount
of elongation of the positive leader is given by Q/ql, where ql is the amount of
charge necessary to thermalize a unit length of the leader channel. For positive
leaders, this is equal to about 40–60 mC/m. Indeed, both the down-coming stepped
leader and the upward-moving connecting leader moves with the aid of streamer
bursts that generate enough charge to thermalize a section of the leader. As the
positive leader approaches the negative one, the average potential gradient between
the two leader tips continues to increase, and when it reaches a value equal to 500
kV/m, all conditions necessary for the final connection are satisfied making the
final connection imminent. This condition is called the final jump condition. Once
the connection is made, the resulting rapid neutralization of the stepped leader
charge leads to the generation of a return stroke. The point of attachment of the
downward flash on the structure is the point of initiation of the connecting leader
that made the final connection with the stepped leader.
Now, let us consider the upward lightning flashes initiated by tall grounded
structures. Upward lightning flashes are initiated by the tall structures themselves due
to the enhancement, caused by the geometry of the structure, of the background electric
field generated by the thundercloud. As the electric field generated by the thundercloud
increases and once all the stages that have been described above, namely, initiation of
avalanches, initiation of streamers and initiation of a leader, had been completed, an
upward-moving leader will be initiated from a field- enhanced tip of a tall grounded
structure. Once initiated, the conditions necessary for its propagation are identical to
that of the connecting leader described in the previous section except for the fact that
here the background electric field remains more or less constant whereas in the pre-
vious case it was increasing with time as the stepped leader approaches the structure.
Once the leader initiated from the structure reaches the charge centre in the cloud, dart
leaders will follow this channel to ground initiating subsequent return strokes.
For a lightning attachment model to be self-consistent, it should take into
account all the processes mentioned above. However, due to the difficulties asso-
ciated with including all these processes into a lightning striking model, engineers
have constructed empirical models that can be applied easily in practice. In the next
section, some of the models utilized to analyse the problem of lightning attachment
are described. However, before proceeding further, let us describe the meaning of
the striking distance as applied in lightning protection studies.
Attachment of lightning flashes to grounded structures 135
Stepped leader
Striking
distance
Final jump
Connecting
leader
Grounded structure
Figure 4.1 The striking distance is defined in the figure as the separation between
the tip of the structure, where a connecting leader is generated, and
the tip of the stepped leader when the final jump condition is
established between the connecting leader and the stepped leader
136 Lightning electromagnetics: Volume 2
7:55 105
Eave ¼ (4.1)
r0:166
In the above equation, r is the gap length (in m). As one can observe from the
above equation, this average potential gradient depends on the gap length and it
decreases with increasing gap length.
In the case of encounter between the connecting leader and the stepped leader,
the final jump condition is reached when the potential gradient between the two
leader tips is equal to the specified critical potential gradient. With the definition
Attachment of lightning flashes to grounded structures 137
given earlier, the striking distance in the absence of a connecting leader from a
grounded structure is equal to the separation between the tip of the structure and the
tip of the stepped leader when the average potential gradient between them reaches
the specified critical value. With increasing length of the connecting leader, the
striking distance increases.
In order to apply this concept of striking distance, it is necessary to know when
and where from the structure a connecting leader is incepted. There are several
theories that can be utilized to find this information and some of the important ones
are summarized below.
One advantage of the critical streamer length criterion over the critical radius cri-
terion is that it can be easily implemented in any complicated structure that one
may encounter in practice.
It is important to point out, however, that the critical radius and critical streamer
length concepts are derived from breakdown characteristics of long rod– plane gaps
under the application of switching impulses of critical time to crest. In the case of
lightning attachment, the temporal variation of the electric field generated at the
grounded structure by a down-coming stepped leader is very different to that of an
electric field generated by a switching impulse. For this reason, the validity of such
concepts in the case of lightning flashes is still a topic of discussion.
240
E0 þ 12½kV=m (4.4)
1 þ h=10
Attachment of lightning flashes to grounded structures 139
where h is the height of the structure in metres. In a later study, however, Lalande
et al. [18] proposed the following equation, which is different from the above, for
the electric field necessary to initiate a stable leader:
306:7 21:6
E0 þ ½kV=m (4.5)
1 þ h=6:1 1 þ h=132:7
Unfortunately, details as to the modifications necessary both in physics and in
mathematics to change the results from (4.4) to (4.5) were not given in Ref. [18].
progression models attempt to simulate the dynamics associated with this process.
Five models that, in contrast to EGM, take into account the formation of an upward
leader exist today and they were introduced by Eriksson [21], Dellera and
Garbagnati [22], Rizk [23], Becerra and Cooray [24], and Vargas and Torres
[25–27]. Only for some of them, however, the dynamic progress of downward and
upward leader is explicitly represented, as it will be illustrated in what follows. For
ease of reference, refer to these five models as A, B, C, D, and E, respectively.
Basic features of these models are schematically depicted in Figure 4.2. Note that
the description given here for the model C is based on Ref. [23]. In a recent paper,
the model was updated and improved by Rizk [15].
● In models A, C, and D, the downward stepped leader is assumed to take a
straight path to ground without branches, while in B the path is determined step
by step by the solution of subsequent electrostatic problems in which the
boundary conditions are represented essentially by the downward and upward
leaders. In E, the downward leader channel may also be tortuous and branched,
and the channel geometry is based on the statistical characterizations of natural
lightning channels as reported by Hill [28–29] and Idone and Orville [30].
● The linear charge density on the downward stepped leader channel is
assumed to decrease upwards in models A, C, D, and E. In B, the charge per
Downward
leader Downward
leader
Striking
Striking
distance
distance
Critical
Leader inception model Streamer streamer
Streamer of Becerra and Cooray zone Upward length
zone Upward
leader leader concept
(d) (e)
unit length has two different values: one in the vicinity (last tens of metres)
of the leader tip, which is uncorrelated to the amplitude of the lightning
current and assumed equal to 100 mC/m, and the other one along the rest of
the leader channel, the magnitude of which varies with the prospective return
stroke current, along the rest of the leader channel. In D, the variation of the
charge per unit length of the stepped leader is approximated by an analytical
expression extracted by Cooray et al. [31] by analysing the charge brought to
ground by first return strokes within the first 100 ms. In E, the charge dis-
tribution along the downward stepped leader, including main channel and
branches, is estimated by an electrostatic model of the lightning leader
channel and the thundercloud [26].
● As the downward leader propagates towards the ground, the conditions at the
surface of the grounded structure are evaluated continuously to find the time of
leader inception. The criterion for upward connecting leader inception used is
either critical radius concept (A and B), Rizk’s generalized equation (C),
Becerra and Cooray procedure (D) or Vargas and Torres criteria (E). The latter
considers a streamer inception electric field assumption and the critical strea-
mer length concept [26]. It is worth noticing that models A, B and C are only
applicable to horizontal wires and vertical earthed structures [see the comment
made on model C at the beginning of this section], whereas models D and E
can be applied to any grounded structure, including complex buildings.
Recently, the procedure to apply models B and C to any grounded structure is
illustrated in references 32 and 15, respectively.
● In models A, C, D and E, the downward stepped leader path is considered unaf-
fected by the presence of upward connecting leaders. In models A, C and D, the
connecting leader travels in space in such a way that it will find the closest path for
the connection with the stepped leader. In model B, both leaders propagate in the
direction of the maximum electric field that exists along an equipotential line that
is located at the outer boundary of the streamer region. The same criterion is
applied in model E to simulate the propagation of the connecting leader.
● The ratio between the speed of propagation of the downward stepped leader and
the upward-moving stepped leader is assumed to be either 1 or 4. However, in
model D, the velocity of the upward leader is evaluated from first principles.
● In model A, the final attachment of the two leaders takes place when two tips
of the leader channel meet each other. In model B, the final jump condition is
reached when the streamer systems from the two leaders meet each other. In C,
D and E, the final jump condition is reached when the average electric field
between the two leader tips is 500 kV/m.
Recently, Mazur et al. [33] introduced a model to describe the propagation of
negative stepped leaders. They utilized the model to estimate the striking distance
of lightning flashes. In the model, the direction of a new leader step is taken to be
the direction in which the length of the negative streamers issued by the negative
leader head is longest. The direction of the positive leader is given by the direction
of the maximum electric field at the tip of the positive leader immediately outside
142 Lightning electromagnetics: Volume 2
the boundary region with an electric field of 3 MV/m. They also assumed that the
striking distance is equal to the length of the final step of the stepped leader.
The length of the final leader step was estimated by dividing the leader potential by
the electric field of the negative streamer zone. Based on their simulations, it was
claimed that the connecting leader does not play a significant role in determining
the striking distance.
When considering the complexity of the lightning phenomena, it will always
be necessary to make a large number of assumptions and simplifications in order to
formulate a usable lightning strike model. However, the creation of leader pro-
gression models has been a major step forward and they are capable of predicting
several phenomena observed in the field. Moreover, the models seem to be well
suited for sensitivity analysis where the effects of various parameters on the effi-
ciency of lightning protection procedures are being studied by changing one para-
meter at a time. It is important to stress here that self-consistent leader progression
models can also be utilized to compliment the simple engineering models that are
being used in practice.
representative of the time for the return stroke front to reach the charge centre in the
cloud and therefore the charge brought to ground during this time is a result of the
neutralization of the section of the leader channel located below the charge centre.
They found that there is a strong correlation between the two parameters. Figure 4.3
shows the results obtained by Cooray et al. [31]. The relationship between the two
parameters can be represented by the equation
Q ¼ 0:061 Ip (4.7)
where Ip is the first return stroke peak current in kA and Q is the charge, in
Coulombs, transported to ground by the return stroke during the first 100 ms.
Cooray et al. [31] extended their analysis further to obtain the distribution of the
linear charge density along the leader channel as a function of return stroke peak
current. The procedure adopted by Cooray et al. [31] to extract the leader charge
distribution corresponding to a given prospective return stroke peak current is the
following. First, they assumed that the stepped leader channel can be represented
by a vertical, finitely conducting channel with a given potential gradient. The
length of the channel was selected by Cooray et al. [31] to be 4 km, a representative
value for the height of the charge centre from the measuring station used by Berger
and Vogelsanger [36,37]. Second, they assumed that, since the negative charge
region extends more in a horizontal direction than vertical, as far as the distribution
of the charge along the leader channel is concerned, the effects of the charges in the
cloud can be represented by a conducting plane charged to a given potential.
100
80
Peak current (kA)
60
40
20
0
0 1 2 3 4 5
Charge Q
Figure 4.3 The charge dissipated by first return strokes in the first 100 ms into the
stroke (circular points). The first return stroke current waveforms used
in the study are from Berger and Vogelsanger [36] and Berger [37].
The solid line shows the linear fit to the data and the dashed line the
power fit (from Ref. [31])
144 Lightning electromagnetics: Volume 2
This assumption led to a uniform electric field below the cloud. Third, they
assumed that the charge brought to ground during the return stroke is the sum of
positive charge necessary to neutralize the negative charge on the leader channel
and the additional positive charge induced on the leader channel due to the pre-
sence of the background electric field caused by the remaining negative charge in
the cloud. Once these assumptions are made, the analysis is carried out as follows.
First, from the observed relationship between the return stroke peak current and the
charge, as depicted in Figure 4.3, the charge corresponding to a given peak return
stroke current is obtained. Second, the background electric field was adjusted so
that the estimated total charge deposited in the leader channel by the return stroke is
equal to this charge. The resulting distribution of the charge along the leader
channel is the one corresponding to the prospective return stroke current selected.
Since the background electric field corresponding to a given charge is known (or
estimated), the leader tip potential as a function of this charge can be estimated.
This potential can be expressed as a function of return stroke peak current with the
aid of (4.7). The resulting relationship between the potential of the tip of the
stepped leader and the peak return stroke current is given by
condition is reached. Thus, the EGM striking distance according to the study of
Armstrong and Whitehead [6] becomes (combining (4.1) and (4.6))
On the other hand, if one uses the leader potential given by (4.8) and assumes
that the average potential gradient at final jump is equal to 500 kV/m, the EGM
striking distance as a function of return stroke peak current becomes
Note that the similarity between the two equations, i.e. (4.9) and (4.10), is a
mere coincidence. In the above estimation it is assumed that the potential gradient
in the final jump zone is equal to 5 105 V/m. On the other hand, if the ground is
assumed to be completely flat then the final jump condition is reached when the
negative streamers of the stepped leader reach the ground. This happens when the
average potential gradient between the leader tip and the ground is equal to
(1–2) 106 V/m, the potential gradient of negative streamers. If an average value
of 1.5 106 V/m is used, the striking distance of a negative stepped leader to
completely flat ground becomes
S ¼ 2Ip0:183 (4.11)
This expression is approximately equal to the expression for the striking distance to flat
ground obtained by Cooray et al. [31]. The slight difference between this equation and
the one derived by Cooray et al. [31] is due to the fact that the estimation given here is
based on the exact charge distribution of the stepped leader channel as estimated in the
analysis whereas in Cooray et al. [31] an approximate analytical expression for
the same charge distribution is used to estimate the average potential gradient between
the leader tip and the ground. Note also that in the section where the striking distance is
calculated in Cooray et al. [31], the potential gradient of the negative streamer zone is
given as 5 105 V/m by mistake. Observe that the striking distance depends on the
assumed potential gradient of the negative streamer channels. In some studies this
potential gradient is assumed to be 750 kV/m. If we make this assumption the striking
distance would be larger than the one given by equation (4.11).
expression for the leader charge derived by Cooray et al. [31]. This analytical
expression is derived by Cooray et al. [31] from a linear fit to the measured data.
Thus it is appropriate to use the potential given by (4.8) in calculating the final
jump condition in SLIM. Moreover, in the study conducted by Cooray and Becerra
[2], what was calculated is the attractive radius of vertical and horizontal con-
ductors. The parameter under study here is the striking distance.
The definition of the striking distance that will be used is given in Section 4.2. The
structure studied here is a vertical structure of cylindrical shape. The tip of the structure
is a hemisphere with a radius equal to the radius of the cylindrical section. In the
calculation, the radius of the structure is assumed to be 0.1 m. Calculations are per-
formed for vertical conductor heights varying from 5 m up to 100 m. For a given
conductor, the striking distance radius is evaluated for prospective return stroke peak
currents of 5–90 kA. In the calculation it is assumed that the stepped leader approaches
vertical structure directly overhead. It is important to mention here that the calculated
striking distance depends also on the final jump condition used in the analysis. Here, it
is assumed that the final jump condition is reached when the average potential gradient
between the tips of the connecting leader and the stepped leader is equal to 500 kV/m.
Before proceeding further, let us consider the EGM. As mentioned previously,
according to the EGM a stepped leader will terminate at a point on a structure if the
potential gradient between the tip of the stepped leader and the point on the structure
reaches a critical value. Here it is assumed that this critical potential gradient is 500 kV/
m. Recall that according to SLIM the final attachment of the stepped leader to the
structure takes place when the potential gradient between the tip of the connecting leader
and the tip of the stepped leader reaches 500 kV/m. Since EGM neglects the presence of
a connecting leader, for a given leader potential EGM provides a lower limit to the
striking distance. As the length of the connecting leader increases, the striking distance
increases. Thus, the striking distance obtained by SLIM will be longer than the striking
distance predicted by EGM. Of course this is true as long as the same critical criterion
(i.e. same average potential gradient across the gap) is used for the final jump condition.
For example, in attractive radii calculations presented by Cooray and Becerra [38], it
was assumed that the potential gradient at every point in the gap should exceed 500 kV/
m before the final jump condition is reached. This condition is more restrictive than the
final jump condition based on the average potential gradient of 500 kV/m across the
gap. Figure 4.4 illustrates how the striking distance varies as the height of the structure is
increased. For comparison purposes, the EGM striking distance (i.e. (4.10)) is also
depicted in each diagram. Analytical expressions for the variation of striking distance as
a function of peak current for different heights are tabulated in Table 4.1.
Note, from the data given in Figure 4.4, that the difference between the striking
distances calculated using SLIM and EGM increases with increasing structure
height. For a given height, the difference is less for smaller currents than for the
larger ones. However, for structure heights less than about 30 m, the differences in
the striking distances are negligible for practical purposes even for peak return
stroke currents as large as 90 kA. For peak return stroke currents less than about 16
kA, the difference between SLIM and EGM is less than about 30% for structure
heights less than about 50 m.
Attachment of lightning flashes to grounded structures 147
250 300
250
200
200
150
150
100
100
50 50
0 0
0 20 40 60 80 100 0 20 40 60 80 100
(a) Return stroke peak current (kA) (b) Return stroke peak current (kA)
400 400
300 300
Striking distance (m)
200 200
100 100
0 0
0 20 40 60 80 100 0 20 40 60 80 100
(c) Return stroke peak current (kA) (d) Return stroke peak current (kA)
500 600
400
Striking distance (m)
Striking distance (m)
400
300
200
200
100
0 0
0 20 40 60 80 100 0 20 40 60 80 100
(e) Return stroke peak current (kA) (f ) Return stroke peak current (kA)
Table 4.1 Analytical expression for the striking distance as a function of peak
return stroke current, Ip, for different structure heights
Coefficients for both power and quadratic fits are tabulated. In the case of EGM, the striking distance is
given by S ¼ 6Ip0:813 or S ¼ 10:38 þ 2:98Ip þ :006Ip2 . The striking distance is given in metres and the
peak return stroke current in kA. The quadratic function fits the data better than the power function.
experienced by the stepped leader. In the models of Rizk [23] and Becerra and Cooray
[19,20,24], it is assumed that the path of propagation of the stepped leader is not
influenced both by the grounded structures and by the connecting leader until the final
jump condition is reached. Of course the first assumption makes some physical sense
but it is not that clear how this assumption should be applied in practice. It is difficult, if
not impossible, to determine the direction of the maximum electric field ahead of the
leader channel due to following reasons: the space in front of the leader channel is
occupied by streamer discharges that supply the current necessary for the propagation
of the leader. The electric field configuration in front of the leader channel is deter-
mined by the spatial distribution of the space charge of the streamer system, which, due
to the random nature of the electrical discharges, is not uniform. Thus, the exact dis-
tribution of the electric field in space in front of the leader cannot be determined with
certainty. Moreover, the direction of the stem of the streamer system, which becomes
the new leader segment, may to some extent be controlled by the space charge lying
ahead of it. This space charge reduces the electric field at the stem and therefore the
direction of the next section of the leader channel may lie in a direction away from the
main concentration of the space charge [40]. The situation is even more complicated in
the case of negative stepped leaders. In this case, the high electric field at the outer edge
of the negative streamer system leads to the creation of a space stem. Subsequently, this
space stem is converted to a space leader. The space leader extends in two direction,
one end moving towards the tip of the stepped leader and the other away from it. When
the two leaders meet (i.e. space leader and the stepped leader) the space leader
becomes the next step of the stepped leader. Thus, the location of the space stem with
respect to the tip of the leader channel will decide the direction of the next leader
step. The theory available at present cannot be used to predict the exact location of the
space stem with respect to the electric field configuration. Thus, the exact direction of
the new step of the leader (or its current direction of propagation) is not a parameter that
can be predicted easily. Indeed it is this random nature of the discharge that causes the
leader channel to take a tortuous path. It is important to note, however, that Mazur et al.
[33] assumed that the direction of a new leader step (and hence the direction of pro-
pagation of the stepped leader) coincides with the direction in which the length of the
negative streamers issued by the negative stepped leader is longest.
The next problem is our lack of knowledge concerning the external electric field
necessary to divert the direction of propagation of a stepped leader channel. The
engine that drives the leaders is the streamer bursts created in front of the leader
channel, and the direction of propagation of the streamers is controlled mainly by the
electric field produced by the already thermalized leader channel. The background
electric field necessary for the propagation of negative streamers is about 1 to 2 MV/
m and it is reasonable to assume that in order to divert the direction of the negative
streamer bursts the background electric field may reach values comparable to this
value. If this is correct, then the background electric field produced at the head of the
negative stepped leader channel by the grounded structures or the connecting leader
has to reach values comparable to 1–2 MV/m to divert the path of the stepped
leader channel. This can happen only when the grounded structure or the connecting
leader has approached the down-coming stepped leader almost to the point of final
jump distance. This, to some extent, justifies the assumption that the stepped leader
150 Lightning electromagnetics: Volume 2
path is not influenced by the grounded structure or the connecting leader until the
final jump condition is reached. Of course, in reality, the situation could lie some-
where in between these two extreme scenarios.
One clear example of a time-resolved stepped leader that propagated without
being much influenced by the connecting leader is shown in Figure 15 of Berger [41].
In this example, the connecting leader issued from the tower propagated first upwards
for a distance of about 10–20 m and then turned towards the down-coming stepped
leader and met it at point A (marked in Figure 15 of Berger [41]), which is at a height
of about 40–50 m from the tip of the tower. More time-resolved photographs similar
to this are needed to investigate how the electric fields created by structures and
connecting leaders can influence the path of negative stepped leaders.
the other engineering parameters such as the striking distance. The existence of
such a relationship is in agreement with physical considerations because the charge
that will be stored on the down-coming leader channel is related to its potential and
the resulting return stroke current, which, in turn, is governed by this charge. The
final jump condition between the connecting leader (or the grounded structure) and
the stepped leader is reached when the average potential gradient between the tips
of the two leaders (or between the grounded structure and the tip of the stepped
leader) is equal to the critical electric field necessary for positive streamer propa-
gation. Thus, the main parameter that influences the final jump condition is the
potential of the stepped leader channel.
Here and in the study conducted by Cooray and Rakov [35], the potential of
the stepped leader channel is evaluated first by theoretically analysing the charge
on the stepped leader channel as a function of its potential and then connecting
this charge to the charge brought to ground by the return strokes over the first
100 ms of the discharge in the current waveforms measured by Berger and
Vogelsanger [36] and Berger [37]. The detailed procedure of the analysis and the
assumptions involved were discussed by Cooray et al. [31]. The validity of
the assumptions made in the analysis was tested by Cooray et al. [31] at least for
the subsequent return strokes by checking whether the derived relationship
between the charge and the current could provide a fit to the experimentally
observed relationship between the magnitude of the close electric fields and the
return stroke peak current. Recently, Mazur and Ruhnke [42] attempted to obtain
the leader potential from the remote measurements of the electric field change
produced by the stepped leader and connecting this derived potential to the return
stroke current estimated from the lightning location systems. Unfortunately, they
could not find a strong relationship between the derived leader potential and the
return stroke current. The various assumptions made in simulating the leader
channel together with the fact that the currents used in the analysis were not
measured but were estimates from the direction finding systems may have
influenced to some extent the obtained results.
More studies should be conducted to get more accurate values of leader
potential because, as shown in this chapter, it is the most important parameter in
lightning attachment. The experimental data show that the mechanism of the
negative leader is similar to that of laboratory sparks. If this is the case, then
the step length is approximately equal to the extension of the streamer region. The
extension of the streamer region is a measure of the potential of the leader channel.
Thus, by measuring the average step length as a function of return stroke current
one may be able to derive an expression for the variation of leader potential with
return stroke current. For example, Figure 4.5 depicts the length of the leader step
as a function of prospective return stroke current calculated using the leader charge
distribution extracted by Cooray et al. [31]. In the calculation it is assumed that the
space stem is created at the edge of the streamer region. The estimated leader step
lengths correspond to leaders at a height of 300 m from ground level. Note that the
estimated leader step lengths for typical return stroke currents are considerably
smaller than the typical value of 50 m estimated in classical studies.
152 Lightning electromagnetics: Volume 2
60
40
20
0
0 20 40 60 80
Prospective return stroke peak current (kA)
Figure 4.5 The length of the leader step as a function of prospective return stroke
current calculated using the leader charge distribution extracted by
Cooray et al. [31]. In the calculation it is assumed that the space stem
is created at the edge of the streamer region
References
[1] Cooray, V., A review of simulation procedures utilized to study the attach-
ment of lightning flashes to grounded structures, Prepared on behalf of the
CIGRE working groupC4.405, Electra, 257, pp. 48–55, 2011.
[2] Cooray, V. and M. Becerra, Attractive radii of vertical and horizontal con-
ductors evaluated using a self consistent leader inception and propagation
model – SLIM, Atmos. Res., 2011, Accepted for publication in Atmos. Res.,
Available online in Science Direct since August, 2011.
[3] Gallimberti, I., The mechanism of long spark formation, J. Physique Coll.,
40, C7, suppl. 7, pp. 193–250, 1972.
[4] Bazelyan, E. M. and Y. P. Raizer, Spark Discharge, CRC Press, New York,
1977.
[5] Golde, R. H., Lightning Protection, Edward Arnold, London, 1973.
[6] Armstrong, H. R. and E. R. Whitehead, Field and analytical studies of
transmission line shielding, IEEE Trans., PAS-87, (1), pp. 270–279, 1968.
[7] Les Renardiéres Group, Research on long air gap discharges—1973 results,
Electra 35, pp. 47–155, 1974.
[8] Carrara, G. and L. Thione, Switching surge strength of large air gaps: a phy-
sical approach, IEEE Trans., PAS-95, (2), pp. 512–524, March/April, 1976.
[9] Bernardi, M., L. Dellera, E. Garbagnati, G. Sartorio, Leader progression model
of lightning: Updating of the model on the basis of recent tests results, Proc.
23rd Int. Conf. Lightning Protection (ICLP), Florence, Italy, September, 1996.
154 Lightning electromagnetics: Volume 2
5.1 Introduction
In this chapter, we present a review of recent progress in the modeling of lightning
strikes to tall structures. Since some tall structures are struck by lightning several
tens of times per year, they can be used as ground-truth to measure and calibrate the
location accuracy of lightning location systems. In addition, knowledge of the
transient processes in tall objects when they a subjected to a lightning strike allows
us to use them to calibrate the lightning return stroke currents reported by lightning
detection and location systems. Tall objects constitute also a primary source of data
from which channel-base lightning current statistics are obtained. These statistics
are in turn used to improve the design of lightning protection devices and systems.
This chapter is organized as follows: Section 5.2 presents a review of the
extension of lightning return stroke models to include the presence of an elevated
strike object. Section 5.3 deals with the computational methods for the evaluation
of the electromagnetic fields generated by lightning strikes to tall structures. A
review of available data on lightning currents from lightning to tall structures is
presented in Section 5.4. Finally, a summary is given in Section 5.5.
The presence of an elevated strike object has been included in two classes of return-
stroke models, namely the engineering models and the electromagnetic or antenna-
theory (AT) models, as defined by Rakov and Uman [1]. In the engineering return-stroke
models, the spatial and temporal distribution of the channel current is specified based on
observed characteristics such as channel-base current, return-stroke speed, and remote
electromagnetic fields. The presence of an elevated strike object in such models has been
considered by assuming the object as a uniform, lossless transmission line (e.g., [2]). In
AT models (e.g., [3–6]), the strike object and the lightning channel are represented using
thin wires. Maxwell’s equations are numerically solved using usually the method of
1
Ecole Polytechnique Fédérale de Lausanne (EPFL), Switzerland
2
School of Engineering and Management Vaud, HES-SO University of Applied Sciences and Arts
Western Switzerland
158 Lightning electromagnetics: Volume 2
moments (MoM) [7] to find the current distribution along the lightning channel, from
which the radiated electromagnetic fields can be computed. Besides the electromagnetic
and engineering model classes, the so-called Hybrid Electromagnetic/circuit theory or
HEM/circuit theory model could be considered as a third class based on a combination
of the electromagnetic approach and circuit theory (e.g., [8,9]). Each of the model classes
will be presented in turn in the next few sections.
1
X
h þ z0 2nh h þ z0 2nh
iðz0 ; tÞ ¼ ð1 rt Þ rnt rng io h; t þ rnt rnþ1
g io h; t
n¼0
c c c c
(5.2)
0
for 0 z h
where h is the height of the tower, rt and rg are the top and bottom current
reflection coefficients for upward and downward propagating waves, respectively,
given by
Zt Zch
rt ¼ (5.3)
Zt þ Zch
Zt Zg
rg ¼ (5.4)
Zt þ Zg
H0 is the total height of the extending return-stroke channel, c is the speed of light,
P(z0 ) is a model-dependent attenuation function, u(t) the Heaviside unit-step func-
tion, v is the return-stroke front speed, and v* is the current–wave speed.
Expressions for P(z0 ) and v* for some of the most commonly used return-stroke
models are summarized in Table 5.1, in which l is the attenuation height for the
MTLE model and Htot is the total height of the lightning channel.
Equations (5.1) and (5.2) are based on the concept of ‘undisturbed current’
io(t), which represents the ‘ideal’ current that would be measured at the tower top if
the current reflection coefficients at both of its extremities were equal to zero.
It is assumed that the current reflection coefficients rt and rg are constant. In
addition, any upward connecting leader and any reflections at the return-stroke
wavefront [19] are disregarded.
An extension of the engineering models was proposed by Mosaddeghi et al.
[27] that takes into account the presence of possible reflections at the return stroke
wavefront and a return stroke initiation above the structure due to an upward
connecting leader. Based on the approach proposed by Shostak et al. [19],
Mosaddeghi et al. derived closed-form iterative solutions for the current distribu-
tion along the channel and the strike object. Mosaddeghi et al. [27] used their
enhanced model to obtain simulations for the magnetic fields and compared them
Model P(z0 ) v*
BG 1 ?
TCS 1 c
TL 1 v
MTLL 1z0 /Htot v
MTLE exp(z0 /l) v
Io1 ðt=t1 Þ2
ðt=t2 Þ t=t3 t=t4
io ðh; tÞ ¼ e þ I o2 e e (5.6)
h 1 þ ðt=t1 Þ2
This undisturbed current is shown in Figure 5.1, where the values of the
parameters chosen are Io1 = 9.9 kA, h = 0.845, t1 = 0.072 ms, t2 = 5.0 ms, Io2 = 7.5
kA, t3 = 100.0 ms, t4 = 6.0 ms. These values correspond to the channel-base current
adopted in [36] to compare ground-initiated lightning return-stroke models.
Starting from the same undisturbed current, the spatialtemporal distributions of
the current along the channel and along the strike object were calculated for each
model (see Figure 5.2).
12
10
8
io(h,t)
Current (kA)
0
0 10 20 30 40 50
Time (μs)
z', m
1,000 1,000
800 800
600 600
400 t =1 μs 400 t =1 μs
200 200
0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
(a) I(z', t), kA (b) I (z', t), kA
2,000 2,000
MTLL TCS
1,800 1,800 t =10 μs
t =10 μs
1,600 1,600
1,400 1,400
1,200 1,200
z', m
z', m
1,000 1,000
800 800
600 600
400 t =1 μs 400 t =1 μs
200 200
0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
(c) I (z', t), kA (d) I (z', t), kA
2,000
MTLE
1,800
t =10 μs
1,600
1,400
1,200
z', m
1,000
800
600
400 t =1 μs
200
0
0 5 10 15 20 25 30
(e) I (z', t), kA
In the calculation, the elevated strike object was assumed to have a height
h = 168 m, corresponding to the Peissenberg tower in Germany, and the reflection
coefficients were set, respectively, to rt = 0.53 and rg = 0.7 [37].
Figure 5.2 shows the current distribution along the tower and along the chan-
nel, at different time instants (t = 1, 2, .., 10 ms), predicted by each model.
Modeling lightning strikes to tall towers 163
25 30
Top of the tower Bottom of the tower
20 25
20
Current (kA)
Current (kA)
15
15
10
10
5
5
0 0
0 5 10 15 20 25 30 35 40 45 0 5 10 15 20 25 30 35 40 45
(a) Time (Ps) (b) Time (μs)
Figure 5.3 Current at the top (a) and at the bottom (b) of a 168-m tower (adapted
from [35])
involves an infinite summation in the time domain, assuming that the reflection coef-
ficients are constant and known. Gavric [42] proposed an iterative method based on the
electromagnetic transient program (EMTP) to remove superimposed reflections caused
by a strike tower from digitally recorded lightning flash currents. Janischewskyj et al.
[43] derived reflection coefficients at the CN Tower in Toronto and stated that the
values depend on the initial rise time of the measured current, although the limited
number of points in their plots render the drawing of conclusions difficult. A depen-
dence on the risetime would suggest that at least one of the reflection coefficients is a
function of the frequency. They also proposed a method to extract the reflection
coefficients from the measured current waveform. However, their method is applicable
only assuming a simplified current waveform (double ramp) and neglecting any fre-
quency dependence for the reflection coefficients. The last consideration was relaxed in
a first approximation by Bermudez et al. [22]. They derived a frequency-domain
counterpart of expressions (5.1) and (5.2) which include the frequency-dependence of
reflection coefficients. They also derived an expression to calculate the reflection
coefficient as a function of frequency at the bottom of the lightning strike object from
two currents measured at different heights along the strike object.
Interestingly, Bermudez et al. [22] showed that, if the current and its time
derivative overlap with reflections at the top or bottom of the strike object, it is
impossible to derive the reflection coefficient at the top of the strike object exactly
regardless of the number of simultaneous current measurements and they proposed
an extrapolation method to estimate this reflection coefficient. They applied their
proposed methodology to experimental data obtained on the Peissenberg Tower
(Germany) consisting of lightning currents measured at two heights and obtained
results that suggest that the reflection coefficient at ground level can be considered
as practically constant in the frequency range 100 kHz to 800 kHz [22].
z'
H tot
v
H0
z' i(z', t)
R Observation point
ρt
h r Er
E z
ρg Ez
Perfect ground
R' az
P
ay
image az
φ
ax
ar
–H0
have little effect on the total field. However, the horizontal (radial) component of
the electric field radiated by lightning is appreciably affected by the finite ground
conductivity. Indeed, for this field component, the effects of the two contributions
subtract, and small changes in the image field may lead to appreciable changes in
the total horizontal field. Although the intensity of the horizontal field component
is generally much smaller than that of the vertical field, within the context of
certain field-to-transmission line coupling models (e.g., [50]), this component plays
an important role and, thus, its calculation requires the use of rigorous expressions
or at least reasonable approximations.
in which
– r and z are the cylindrical coordinates of the observation point,
– R is the distance between the dipole and the observation point,
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
R ¼ r2 þ ðz0 zÞ2 ,
– i(z0 , t) is the dipole current,
– c is the speed of light, and
– eo is the permittivity of free space.
The total electromagnetic fields are calculated by integrating the above equa-
tions along the tower and the channel and their image, assuming a perfectly con-
ducting ground.
In the presence of a current discontinuity, the radiation term, namely the last
term in each equation, which is proportional to the current time-derivative, intro-
duces a singularity that needs to be treated separately [52–57].
where f(z0 , z, r) can be r2/c2R3, r(zz0 )/c2R3, or r/cR2, depending on which com-
ponent of the field is being calculated [38].
The reason why an additional turn-on term must be introduced in the field
equations is that the presence of the discontinuity at the return-stroke wavefront in
(5.1) cannot be disregarded when the time-derivative of the current is calculated. Its
derivative, namely, a delta function, multiplied by the amplitude of the current at
the wavefront, needs to be added to the radiation term. In the case in which the
current distribution presents no discontinuity at the return-stroke wavefront, this
turn-on term contribution vanishes. The discontinuity can be treated considering a
continuous current wavefront of length Dz00 which reaches the level Ifront linearly in
a time Dt, and expressing the radiation integral across H taking the limit when the
front duration tends to zero [52].
168 Lightning electromagnetics: Volume 2
The final expressions for the turn-on term fields, in which the apparent front
speed appears as the reciprocal of the term between brackets, are given by [38]:
Ifront ðHÞ r 1 Ifront ðH 0 Þ r 1
HF= turnon ¼ h i þ 02
h i (5.11)
4pcR2 ðzHÞ 4pcR ðzH 0 Þ
1
v cR
1
v 0 cR
(5.12)
0
Ifront ðHÞ r 2
1 I ðH Þ r 1 2
Er= turnon ¼ h i þ front 2 0 h 0Þ
i (5.13)
4pe0 c2 R3 ðzHÞ 4pe c R 3 ðzH
1
v cR
0 1
0 v cR
In (5.11)(5.13), the two terms on the right-hand side represent the turn-on
term due to the discontinuity at the wavefront and at its image, respectively.
The general expression for the current at the wavefront is simply obtained from
(5.1), in which the time variable t appears implicitly through H [38]:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
H h 1 H h
Ifront ðHÞ ¼ PðH hÞi0 h; þ r2 þ ðH zÞ2 þ
v c v
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
H h 1 H h
rt i0 h; þ r2 þ ðH zÞ2 þ
v c c
X1 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
H h 1 H þ h 2nh
þ ð1 rt Þð1 þ rt Þ rgnþ1 rnt i0 h; þ r2 þ ðH zÞ2
n¼0
v c c c
(5.14)
It is worth observing that the first term on the right-hand side of (5.14) is
nonzero only for the BG and TCS models, and it corresponds to the inherent dis-
continuity predicted by these two models. As a consequence of this, the turn-on
term has the same expression for the TL, MTLL, and MTLE models [38].
The contribution of the turn-on term to the total field depends on many factors,
such as the height of the tower, the reflection coefficients at its extremities, the
return-stroke speed, and the position of the observation point (distance and eleva-
tion). Pavanello et al. [38] found that the contribution of the turn on term to the
total electric and magnetic fields is negligible at close distances (below 100 m) and
increases rapidly to reach an asymptotic value of about 12% at a distance of 5 km
and beyond. At these distances, the field peak is essentially due to the
radiation term.
The elevated strike object was assumed to have a height h = 168 m, corresponding
to the Peissenberg tower in Germany.
The reflection coefficients were set, respectively, to rt = 0.53 and rg = 0.7 [37].
Figure 5.5 presents electric and magnetic fields calculated at a distance of 50 m
from the tower base [35]. At this distance, the electric field is dominated (for late
times, after the initial fast transition) by its electrostatic term. The model-predicted
electric fields are very similar for the first 5 ms, beyond which the BG, TCS and
MTLL models predict a flattening of the field, typically observed at close dis-
tances, while the TL model predicts a field decay. The late-time E-field predicted
by the MTLE model exhibits a ramp, as in the case of a ground-initiated return-
stroke [35]. Note, however, that a judicious choice of the attenuation factor would
result in the flattening of the late-time E-field at close range [58].
Figure 5.5b shows that the predicted magnetic field is nearly model-
independent. At this distance, the magnetic field is dominated by its induction
term, and its waveshape is similar to the current at the base of the tower shown in
Figure 5.3b.
Figure 5.6 presents calculated electric and magnetic fields at a distance of 5 km
[35]. The electric and magnetic field waveshapes for the first 5 ms are dominated by
the radiation term and, hence, they are very similar. No significant differences are
found between the various models in this early-time region. The differences
between the model predictions become more pronounced at later times, t > 5 ms or
so, although they are unremarkable. Note that all the models predict a flattening of
the electric field at later times at a value that is significantly smaller than the initial
peak, in contrast with calculated electric fields for ground-initiated return-strokes
(see, e.g., [36]).
The electric and magnetic fields at a distance of 100 km are plotted in
Figure 5.7 [35]. At this distance, the fields are essentially radiation fields, and
electric and magnetic fields have the same waveshape. The fields associated with
ground-initiated return-strokes at such distances exhibit a zero-crossing which is
only reproduced by the MTLE and MTLL models [1,36]. As seen in Figure 5.7, for
16 100
14 90
80
12
Magnetic field, A/m
Electric field, kV/m
MTLE 70
10 60
8 TCS 50 BG, TCS, TL, MTLL, MTLE
BG 40
6
MTLL 30
4 TL
20
2 10
0 0
0 5 10 15 20 25 30 35 40 45 50 0 5 10 15 20 25 30 35 40 45 50
(a) t, μs (b) t, μs
Figure 5.5 Electric (a) and magnetic (b) fields calculated at a distance of 50 m
from a lightning return-stroke to a 168-m tower (adapted from [35])
170 Lightning electromagnetics: Volume 2
0.35 0.9
0.8
0.3
0.7
0.25
0.6
BG
0.2 0.5
TCS BG
MTLL
0.15 TL 0.4 TCS
TL MTLL
0.3
0.1 MTLE
0.2
MTLE
0.05
0.1
0 0
0 5 10 15 20 25 30 35 40 45 50 0 5 10 15 20 25 30 35 40 45 50
t, μs t, μs
(a) (b)
Figure 5.6 Electric (a) and magnetic (b) fields calculated at a distance of 5 km
from a lightning return-stroke to a 168-m tower (adapted from [35])
16 45
14 40
12 35
Magnetic field, mA/m
Electric field, V/m
30
10
25
8
20
6 BG BG
15 TL TCS
TL TCS MTLL
4 MTLL 10
2 5 MTLE
MTLE
0 0
0 5 10 15 20 25 30 35 40 45 50 0 5 10 15 20 25 30 35 40 45 50
(a) t, μs (b) t, μs
Figure 5.7 Electric (a) and magnetic (b) fields calculated at a distance of 100 km
from a lightning return-stroke to a 168-m tower (adapted from [35])
60
Measurment
Extended MTLE
50 MTLE
40
HI (mA/m), event#3
30
20
10
–10
0 5 10 15 20 25 30 35 40
Time (Ps)
recently made to find an algorithm to carry out the required numerical integrations
efficiently. A dedicated algorithm [64–67] has been developed by Delfino and co-
workers that could be applied to obtain the fields generated by lightning in the air
and in the ground.
where Erp is the radial electric field calculated at a height h, and Hjp is the azi-
muthal magnetic field computed at ground level, both computed assuming the
ground to be a perfect conductor.
Delfino et al. [65] showed that only for very low conductivities, does the Cooray–
Rubinstein formula exhibit some deviations from the reference one, but it still gives a
conservative estimate of the radial field component, since it behaves as an upper bound
for the exact curve. General limits of validity of the Cooray–Rubinstein approximation
were theoretically examined by Wait [70]. Shoory et al. [71] presented a general
equation for the horizontal electric field, from which the Cooray–Rubinstein formula
can be derived as a special case. Cooray [26] further proposed a simple modification of
(5.15) that provides a better early time response. Barbosa and Paulino [72] proposed an
approximate time-domain formula for the horizontal electric field whose range of
validity was stated to be equivalent to that of the Cooray–Rubinstein formula (which is
in the frequency domain). Caligaris et al. [73] mathematically derived the time-domain
counterpart of the Cooray–Rubinstein formula.
Petrache et al. reported a good agreement between the exact and predicted hor-
izontal electric field penetrating the ground at distances as close as 100 m [75]. The
predictions of Cooray’s formula were found to be in good agreement with exact
solutions for large values of the ground conductivity (about 0.01 S/m) [64]. For
poor ground conductivities (0.001 S/m or so), Cooray’s expression yields less
satisfactory results, especially for the late time response [64].
50 200
0 0
–200
–50
–400
Er (V/m)
Er (V/m)
–100
–600
–150
–800
–200 Simulation result
Simulation result –1000 Sommerfeld approach [17]
Sommerfeld approach [17]
–250 Cooray formula
Cooray formula –1200
–300 –1400
0 2 4 6 8 10
0 2 4 6 8 10
(a) t (Ps) (b) t (Ps)
Figure 5.9 Underground radial electric field (r =50 m, depth d = 5 m) for (a)
s = 0.01 s/m, and (b) s = 0.001 s/m lossy ground, compared with the
Sommerfeld approach and the Cooray formula (adapted from [81])
174 Lightning electromagnetics: Volume 2
50 200
0
0
–200
–50
Er (V/m)
–400
Er (V/m)
–100 –600
–200
0 2 4 6 8 10 –1200
t (Ps) 0 2 4 6 8 10
(a) (b) t (Ps)
It can be seen that the FDTD results are in excellent agreement with the exact
evaluation of Delfino et al. [67]. The comparisons also show that the results
obtained using the Cooray formula are in general in good agreement with more
exact solutions, although some discrepancies can be observed for the late-time
response of the field and for poor ground conductivities.
Baba and Rakov [33,83] also used the FDTD method to study the mechanisms
of current wave propagation along vertical conductors [83], to reproduce small-
scale experiments [33], and to study the enhancement of electromagnetic fields
measured on the top of buildings [84].
Method of Moments
The Method of Moments (MoMs) has also been extensively applied to compute
electromagnetic fields radiated by a lightning discharge, within the so-called AT
models, which belong to the class known as electromagnetic models and in which
the return-stroke channel is represented using thin wires (e.g., [4–6,44–46,71,85]).
Most of the MoM solutions are implemented in the frequency domain, which
allows taking into account the presence of a lossy ground in a straightforward way.
and Gardner’s magnetic link sensor measures only the highest current in a flash.
Wagner and McCann [89] made an improved instrument, called a fulchronograph,
which uses a number of strips attached to the edge of a wheel. As the wheel spins,
the magnetic strips pass successively close to coils through which the current to be
measured flows and, thus, different levels are recorded on different strips, provid-
ing a crude indication of the change of the current with time.
These relatively inexpensive sensors, especially the magnetic links, have been
used extensively in Germany, Poland, Czechoslovakia, and Russia and other
countries [90–94].
Although knowledge of the peak current is important, appropriate protection of
current electric and electronic systems requires more detailed information about the
waveshapes of the currents.
Measurements of the lightning current as a function of time were made by
McEachron [95] starting in 1937 on the 443-m tall Empire State Building using
oscilloscopes to measure lightning currents at a height of 389 m, on top of the
building but below its final spire. Since then, a number of towers have been instru-
mented to capture lightning currents. Without being exhaustive, but including both
active and inactive measurement stations, towers are located in Austria, Brazil,
Canada, Colombia, Germany, Italy, Japan, Russia, Spain, South Africa, and
Switzerland, and, with at least two more tower instrumentation projects in the works
as of the writing of this text, one in Serbia and Montenegro, and one in China.
In the following section, we will present a representative selection of experi-
mental lightning current data and electromagnetic fields from them.
kA
0.0
ICC
–0.5
ICC-pulses
–1.0
Return
–1.5 Strokes
–2.0
0 100 200 300 400
ms
The ICC has a typical amplitude of a few hundred Amps and a duration of a few
hundred ms. It may exhibit superimposed pulses currently called ICC pulses (in
the past called a-pulses), exhibiting amplitudes lower than 2 kA. Once the ICC
has ceased to flow, one or more return stroke pulses may appear. Return stroke
pulses are generally larger than ICC pulses, have shorter risetimes, and may be
followed by a continuing current similar to the ICC. Currents in downward
lightning striking grounded objects do not exhibit the initial ICC part. Instead,
they are characterized by the presence of a first return stroke current which fol-
lows a downward stepped leader. This first return stroke may be followed by
subsequent leader-return stroke sequences.
It is important to note that the 100 m height limit for the appearance of upward
lightning does not refer to the actual height of the tall object but to its effective
height, which takes into account the enhancement of the electric field on the object
due, for instance, to a mountain on which the object may be built. Since there is, at
present, no accurate way of estimating theoretically the effective height of a tower,
empirical methods are employed (by the ratio of upward to downward flashes or the
increase in the number of flashes to a tower with respect to the surrounding terrain).
In the following sections, we use the actual physical heights of the towers.
The towers were instrumented using resistive shunt sensors at their tops and
the currents were recorded using cathode-ray oscilloscopes installed in 1958. On
each tower, two different shunts were used in series, one with a resistance of 0.05
Ohms for currents in the 1–200 kA range and the other with a 0.8 Ohm resistance
used for currents from 50 A to 24 kA [97]. Four channels were used with two time
deflections with a resolution of 0.5 ms [98].
About 15% of the measurements reported by Berger and co-workers were due
to downward-moving stepped leaders. Most discharges to the towers were initiated
by upward-moving stepped leaders of both polarities.
0 16 32 43 64 80 µs–B
0.0
A
–0.2
–0.4
B
–0.6
–0.8
–1.0
0 80 160 240 320 400 µs–A
(a)
0 8 16 24 32 40 µs–B
0.0
–0.2
A
–0.4
B
–0.6
–0.8
–1.0
0 20 40 60 80 100 µs–A
(b)
Figure 5.12 Typical, normalized negative return-stroke current wave shapes: (a) First
return stroke and (b) subsequent return stroke (adapted from [99])
178 Lightning electromagnetics: Volume 2
99
95 1
80
50
2
20 3
They are presented in two time scales, with the label “A” corresponding to a large
scale going up to 100 ms and “B” to a shorter time scale of 40 ms (dashed lines). In
Figure 5.12, it is possible to observe that the average subsequent stroke current
exhibits a faster risetime than that of the average first stroke.
In Figure 5.13, the peak current distribution is presented for negative first
return strokes, negative subsequent return strokes, and positive return strokes. The
dashed, slanted lines represent a log normal distribution fit to the experimental data
for all three cases [100]. The value of the peak current distribution at 50% is around
30 kA for first return strokes, both negative and positive. The corresponding value
for negative subsequent strokes is close to 12 kA. Most positive flashes are single-
stroke.
Table 5.2 contains parameters of negative as well as positive lightning. There
is a controversy concerning the front duration and the maximum rate of rise, di/dt,
in Berger’s data. Indeed, the instrumentation used by Berger and co-workers had a
limited frequency bandwidth, which may have introduced inaccuracies in their
experimental observations.
95% 50% 5%
Time interval
Between negative strokes msec 133 7 33 150
Flash duration
Negative (including single stroke flashes) msec 94 0.15 13 1,100
Negative (excluding single stroke flashes) msec 39 31 180 900
Positive (only single flashes) msec 24 14 85 500
Parameter Downward
First Subsequent
strokes strokes
Sample size 42 33
Peak value (kA) 33 18
Maximum rate of rise (kA/ms) 14 33
Time to crest (msec) – (3 kA to peak) 9 1.1
Time to half value (msec) 56 28
Impulse charge (C) – (to end of impulse of 2.8 1.4
500 msec)
95% 50% 5%
Sample size 114
Peak value (kA) 4.9 12 29
10%–90% average of current steepness (kA/ms) 3.3 15 72
10%–90% time duration (msec) 0.1 0.6 2.8
Other data have been obtained using short towers in Japan [103], in Austria
[96,104,105], and in Colombia [106–108].
The Gaisberg Tower in Austria is located near the city of Salzburg. The tower
is 100 m tall and it sits on top of the Gaisberg Mountain, whose summit is at
1,287 m above sea level, and 800 m above the surrounding terrain.
The current is measured at the base of the air terminal installed on the top of
the tower using a shunt resistor of 0.25 mOhm. Two separate channels with dif-
ferent gains are used to transmit and record the data from the Gaisberg tower: One
of them, whose saturation level is 2 kA (positive and negative) is used to record the
small, slow initial continuous current ICC. The other is set for a saturation level of
plus or minus 40 kA to record pulses and return strokes.
Diendorfer and co-workers [96] analyzed currents from strikes to the Gaisberg
tower recorded from 2000 to 2007. The authors reported that the incidence of
lightning to the tower is essentially independent of the season, even though an
active thunderstorm season exists in Austria during the summer. The average
number of flashes per year recorded during the 8-year study period was somewhat
higher than 60. Overall, 93% of the flashes were negative (all strokes lowered
negative charge to ground), 4% were positive (all the strokes lowered positive
charge to ground), and 3% of the records showed bipolar current waveforms.
Based on the presence or absence of an initial continuous current, Diendorfer
et al. concluded that at least 99% of all flashes to the Gaisberg tower were upward-
initiated for the studied period. Of those, 30% of the negative flashes exhibited at
least one return stroke, the average number of strokes being 4.4. The interstroke
interval for the analyzed flashes had a geometric mean of 17.3 ms. Of the
remaining 70% of upward-initiated negative flashes, 22% consisted of an initial
continuous current with superimposed pulses greater than 2 kA and 48% had no
superimposed current pulses with peak currents greater than 2 kA.
540 m t (μs)
533 m
0 10 20 30 40
0
I, kA
–5
533 m
–10
272 m 0 10 20 30 40
0
I, kA
–10
–20 272 m
–30
0 10 20 30 40
0
I, kA
47 m –10
–20 47 m
–30
reflection coefficient can be associated with the top of the tower. This coefficient
represents the discontinuity between the tower and the “equivalent” impedance of
the lightning channel.
Rakov [109] reports a median peak value for currents measured at 47 and
533 m of 18 and 9 kA, respectively. He suggests that the effective grounding
impedance of the tower is much smaller than its characteristic impedance and that
this is appreciably lower than the equivalent impedance of the lightning channel.
Studies on lightning striking the CN Tower (553-m high) in Toronto, Canada,
have been performed and reported by the “CN Tower Lightning Studies Group
(CNTLSG)” since 1978 (e.g., [43,110,111]). The lightning return-stroke current
derivatives striking the CN tower are measured by two inductive Rogowski coils
located at 509 and 474 m height. A photograph of the CN-Tower in Toronto is
shown in Figure 5.15.
A lightning return-stroke current measured on the CN Tower in 1999 is pre-
sented in Figure 5.16. Lightning return-stroke currents and current derivatives
observed at the CN Tower have been found to exhibit multiple reflections produced
at the tower discontinuities. The observed currents and current derivatives are
therefore “contaminated” by these reflections.
The waveshapes of the currents in Figure 5.16 exhibit a positive reflection
arriving around 3.6 msec after the first current maximum. This propagation time
corresponds to a round-trip time from the tower top to ground, confirming that this
reflection was produced at the lower discontinuity level between the tower-bottom
and the grounding impedance. The positive value of the reflection implies a posi-
tive ground reflection coefficient. The observed positive reflection is less pro-
nounced for the sensor located closer to the top of the tower. This is similar to the
observations at the Ostankino tower, suggesting a negative top reflection
coefficient.
Modeling lightning strikes to tall towers 183
Figure 5.16 A lightning return-stroke current observed at (a) 509 and (b) 474 m
height along the CN Tower in Toronto (adapted from [112])
(a)
0
tower base
(5 m)
–2
tower top
(kA) (160 m)
–4
i1
i imax/top
–6
imax/bottom
–8
0 2 4 6 8
t (Ps)
(b)
Figure 5.17 (a) Peissenberg tower and (b) comparison of a lightning return-
stroke current recorded at the Peissenberg tower top and bottom
(adapted from [37])
was located at the summit of a hill 500 m above sea level. The two current deri-
vative systems and the current probe were used over a period of 5 years to record
lightning return stroke current waveshapes impacting the tower [115,116].
Figure 5.18 shows the location of the measurement systems on the tower.
186 Lightning electromagnetics: Volume 2
loop antenna
248/10nH
+248.00
current monitor
248m /25V/kA
plastic cylinder to
shelter HF-high
power antennas
+155.00
oscilloscope 3. 47 fl
GOULD 1602/47 fl oscilloscope 1.
LECROY 9450/44 fi
measures the over- 44 fl
voltage caused by oscilloscope 2.
lightning in specific LECROY 9450/44 fi
installations
39 fl
+ 103.00
30 fl
20 fl
10 fl
+ 23.74
493 m above
sea level
+ 0.00
Lightning channel
Coaxial shunt resistor
2 mΩ E/O
10 mΩ
Frequency Sensitivity E/O
Magnetic flux
5 kHz~1 MHz 0.41 μT/V
density
Optical cable
200 m
Electric field 100 Hz~100 MHz 0.87 (kV/m) / V
The 200-m high Fukui tower in Japan was also used to measure lightning
return-stroke currents and their associated electromagnetic fields at the
Fukui thermal power plant on the coast of the Sea of Japan. Two coaxial
shunt resistors (2 mW, 10 mW) were installed at the top of the tower [21]. It was
found that the measured current was affected by reflected waves at the ground
and at the top of the tower. Figure 5.19 presents a schematic representation of
the installation of the Fukui tower and the electromagnetic field recording
system.
A study conducted using data from EUCLID (European Cooperation
for Lightning Detection) over a 3 and half year time period starting in
January 1998 showed that a telecommunication tower near Saint Gallen, in
Northeastern Switzerland, is struck by lightning more than any other tower
in that country, with an average number of direct hits of more than 100 per
year.
The Säntis tower is 124 m tall and it sits on top of the 2,502 m mount Säntis
(see Figure 5.20).
188 Lightning electromagnetics: Volume 2
The tower was instrumented and the measurement system was put in operation
in June 2010, with Rogowski coils and wideband B-dot sensors installed to measure
the current at 24 m and 82 m AGL (e.g., [117–119]).
Table 5.5 summarizes statistical results for the current peak and risetime
obtained using various instrumented towers around the world.
Table 5.5 Comparison of return-stroke current peaks in downward and upward flashes measured in instrumented towers effectively
transporting negative charge to ground
Location Height (m) Location Sample size Ipeak (50%) (kA) Risetime (ms) Sensor location
CN Tower – Toronto (1992–2001) [112,120]
Upward flashes 553 Canada 387 5.06a 0.64f Top (474 m)
7.19b
Ostankino Tower – Moscow (1984) [109]
Negative upward flashes at 533 m 540 Russia 58 9 – 533 m
Negative upward flashes at 47 m 76 18 – 47 m
Empire State Building – USA (1952) [109]
Upward flashes 410 USA 84 10 – Top
Fukui thermal power plant (1989–1994) [121]
Upward flashes type A (strong luminosity) 200 Japan 22 33 (23.5c) 1–2c Top
Upward flashes type B (low luminosity) 33 3.4
Peissenberg Tower–Germany (1992/98) [37]
Negative upward flashes, ICC pulsesj 168 Peissenberg at 89 3.63b – Top
167 m
Negative upward flashes, return stroke pulsesj 68 7.97b –
Japan Transmission Towers (1994/97) [103]
Negative downward flashes, 1st stroke at top 40–140 Japan 36 39 4.5g Top
Gaisberg Tower (Austria) [122]
Downward negative, first strokes 100 Austria 14i – Top
Downward negative, subsequent strokes 2.91i –
Upward negative, ICC pulsesj 2.94i
Upward negative, return stroke pulsesj 8.57i
Mount Saint Salvatore Tower- [99]
Negative downward flashes, 1st stroke 70–90 Switzerland 101 30 5.5d Top
Negative downward flashes, 2nd stroke 135 12 1.1d
Upward flashes 70 10 –
(Continues)
Table 5.5 (Continued)
Location Height (m) Location Sample size Ipeak (50%) (kA) Risetime (ms) Sensor location
South Africa Tower (1980) [102,123]
Negative downward flashes 60 South Africa 114 12b 0.6h Bottom
Morro do Cachimbo Tower- [124]
Negative downward strokes, 1st stroke 60 Brazil 31 40.4a 5.6h Bottom
45.3b
Negative downward strokes, subs. stroke 59 16.3b 0.7h
Italy Towers [100,101]
Negative downward flashes, 1st stroke 40 Italy 42 33 9e Top
Negative downward flashes, 2nd stroke 33 18 1.1e
Negative upward flashes, 1st stroke 61 7 4e
Negative upward flashes, 2nd stroke 142 8 1.3e
a
First peak of the current.
b
Absolute peak of the current.
c
Values reported by [21].
d
Front duration defined as the time interval between the 2 kA point on the front and the first peak, the time resolution of the system was 0.5 ms.
e
Time to crest defined between 3 kA to peak.
f
Risetime to wavefront peak.
g
The time resolution of the system was 100 ns [103].
h
Time interval between instants corresponding to 10% and 90% of first current peak.
i
1 kA threshold.
j
a-Components: ICC pulses (sometimes known as a-components or a-pulses) are superimposed on the initial continuous current. Return stroke pulses (also known as
b-components or b-pulses) are pulses following a period of no current in the channel.
Modeling lightning strikes to tall towers 191
5.5 Summary
In this chapter, the recent progress in our understanding of the transient processes
that take place when lightning strikes tall structures was presented. The extension
of lightning return stroke models to include the presence of these strike structures
was given, including the engineering, electromagnetic and hybrid model classes.
The computational methods commonly used to calculate the electromagnetic fields
generated by lightning strikes to tall structures were presented, both for the idea-
lized case of a perfectly conducting ground, and for lossy ground. For the latter, in
addition to standard FDTD and MoM numerical methods and dedicated algorithms
for the evaluation of Sommefeld’s integrals, simplified, approximate approaches
were presented.
A review of available data on the lightning current and its associated electro-
magnetic fields was presented in Section 5.4, where the dataset of Berger and
co-workers was presented in detail and an overview was given of some of the
instrumented towers around the world, including the most recent lightning current
and electromagnetic field results.
It is worth noting that in this chapter, we have only considered the return stroke
phase of the lightning discharge. The effect of tall structures on the electromagnetic
fields radiated by the M-component† mode of charge transfer has also recently been
considered and analyzed [125].
References
[1] Rakov, V.A. and M.A. Uman, Review and evaluation of lightning return
stroke models including some aspects of their application. IEEE
Transactions on Electromagnetic Compatibility, 1998;40(4):403–26.
[2] Rachidi, F., V.A. Rakov, C.A. Nucci, and J.L. Bermudez, The effect of
vertically-extended strike object on the distribution of current along the
lightning channel. Journal of Geophysical Research, 2002;107(D23):4699.
[3] Podgorski, A.S. and J.A. Landt, Numerical analysis of the lightning-CN
tower interaction, in 6th Symposium and Technical Exhibition on
Electromagnetic Compatibility, 1985, Zurich, Switzerland.
[4] Heidler, F. and T. Zundl, Influence of tall towers on the return stroke cur-
rent, in Aerospace and Ground Conference on Lightning and Static
Electricity, 1995, Williamsburg, USA.
[5] Baba, Y. and M. Ishii, Numerical electromagnetic field analysis of light-
ning current in tall structures. IEEE Transactions on Power Delivery,
2001;16(2):324–8.
†
The M-component mode of charge transfer occurs during the continuing current stage in downward
lightning and during the initial stage of upward lightning initiated by tall structures or in rocket-triggered
lightning.
192 Lightning electromagnetics: Volume 2
[44] Podgorski, A.S. and J.A. Landt, Three dimensional time domain modelling
of lightning. IEEE Transactions on Power Delivery, 1987;PWRD-2(3):
931–8.
[45] Moini, R., V.A. Rakov, M.A. Uman, and B. Kordi, An antenna theory
model for the lightning return stroke, in 12th International Zurich
Symposium on Electromagnetic Compatibility, 1997, Zurich, Switzerland.
[46] Petrache, E., F. Rachidi, D. Pavanello, et al., Lightning strikes to elevated
structures: influence of grounding conditions on currents and electro-
magnetic fields, in IEEE International Symposium on Electromagnetic
Compatibility, 2005, Chicago, IL.
[47] Petrache, E., F. Rachidi, D. Pavanello, et al., Influence of the finite ground
conductivity on the transient response to lightning of a tower and its
grounding, in 28th General Assembly of International Union of Radio
Science (URSI), 2005, New Delhi, India.
[48] Rubinstein, M, An approximate formula for the calculation of the hor-
izontal electric field from lightning at close, intermediate, and long range.
IEEE Transactions on Electromagnetic Compatibility, 1996;38(3):531–5.
[49] Rachidi, F., C.A. Nucci, I. M., and C. Mazzetti, Influence of a lossy ground
on lightning-induced voltages on overhead lines. IEEE Transactions on
Electromagnetic Compatibility, 1996;38(3):250–63.
[50] Agrawal, A.K., H.J. Price, and S.H. Gurbaxani, Transient response of
multiconductor transmission lines excited by a nonuniform electromagnetic
field, in IEEE International Symposium Digest. Antennas and Propagation,
1980, New York, USA.
[51] Pavanello, D., F. Rachidi, M. Rubinstein, J.L. Bermudez, and C.A. Nucci,
On the calculation of electromagnetic fields radiated by lightning to tall
structures, in International Conference on Lightning Protection, ICLP
2004, 2004, Avignon, France.
[52] Rubinstein, M. and M.A. Uman, Transient electric and magnetic fields
associated with establishing a finite electrostatic dipole, revisited. IEEE
Transactions on Electromagnetic Compatibility, 1991;33(4):312–20.
[53] Rubinstein, M. and M.A. Uman, On the radiation field turn-on term asso-
ciated with traveling current discontinuities in lightning. Journal of
Geophysical Research, 1990;95(D4):3711–13.
[54] Le Vine, D.M. and J.C. Willett, Comment on the transmission line model
for computing radiation from lightning. Journal of Geophysical Research,
1992;97:2601–10.
[55] Thottappillil, R. and V.A. Rakov, On different approaches to calculating light-
ning electric fields. Journal of Geophysical Research, 2001;106:14191–205.
[56] Thottappillil, R., V.A. Rakov, and M.A. Uman, Distribution of charge
along the lightning channel: relation to remote electric and magnetic fields
and to return-stroke models. Journal of Geophysical Research, 1997;102
(D6):6987–7006.
196 Lightning electromagnetics: Volume 2
[57] Thottappillil, R., M.A. Uman, and V.A. Rakov, Treatment of retardation
effects in calculating the radiated electromagnetic fields from the lightning
discharge. Journal of Geophysical Research, 1998;103(D8):9003–13.
[58] Cooray, V., V.A. Rakov, C.A. Nucci, F. Rachidi, and R. Montano, On the
constraints imposed by the close electric field signature on the equivalent
corona current in lightning return stroke models, in International
Conference on Lightning Protection (ICLP 2004), 2004, Avignon, France.
[59] Pavanello, D., F. Rachidi, M. Rubinstein, N. Theethayi, and R.
Thottappillil, Electromagnetic environment in the immediate vicinity of a
tower struck by lightning, in EUROEM’2004, 2004, Magdeburg, Germany.
[60] Miyazaki, S. and M. Ishii, Influence of elevated stricken object on lightning
return-stroke current and associated fields, in International Conference on
Lightning Protection ICLP, 2004, Avignon, France.
[61] Mosaddeghi, A., A. Shoory, F. Rachidi, et al., Lightning electromagnetic
fields at very close distances associated with lightning strikes to the
Gaisberg Tower. Journal of the Geophysical Research, 2010;115:(D17101.
[62] Rakov, V.A. and F. Rachidi, Overview of recent progress in lightning
research and lightning protection. IEEE Transactions on Electromagnetic
Compatibility, 2009;51(3):428–42.
[63] Baños, A, Dipole Radiation in the Presence of a Conducting Half-Space,
1966, Oxford: P. Press.
[64] Delfino, F., R. Procopio, F. Rachidi, and C.A. Nucci, An algorithm for the
exact evaluation of the underground lightning electromagnetic fields. IEEE
Transactions on Electromagnetic Compatibility, 2007;49(2):401–11.
[65] Delfino, F., R. Procopio, and M. Rossi, Lightning return stroke current
radiation in presence of a conducting ground: 1. Theory and numerical
evaluation of the electromagnetic fields. Journal of Geophysical Research,
2008;113;D05110.
[66] Delfino, F., R. Procopio, M. Rossi, F. Rachidi, and C.A. Nucci, Lightning
return stroke current radiation in presence of a conducting ground: 2.
Validity assessment of simplified approaches. Journal of Geophysical
Research, 2008;113:D05111.
[67] Delfino, F., R. Procopio, M. Rossi, F. Rachidi, and C.A. Nucci, Evaluation
of underground lightning electromagnetic fields, in International
Symposium on Electromagnetic Compatibility EMC EUROPE 2006, 2006,
Barcelona, Spain.
[68] Rubinstein, M., An approximate formula for the caluclation of the hor-
izontal electric field from lightning at close, intermediate, and long range.
IEEE Transactions on Electromagnetic Compatibility, 1996;38:531–5.
[69] Cooray, V., Horizontal fields generated by return strokes. Radio Science,
1992;27(4):529–37.
[70] Wait, J.R, Concerning the horizontal electric field of lightning. IEEE
Transactions on Electromagnetic Compatibility, 1997;39(2):186.
Modeling lightning strikes to tall towers 197
[71] Shoory, A., R. Moini, S.H.H. Sadeghi, and V.A. Rakov, Analysis of
lightning-radiated electromagnetic fields in the vicinity of lossy ground.
IEEE Transactions on Electromagnetic Compatibility, 2005;47(1):131–45.
[72] Barbosa, C.F. and J.O.S. Paulino, An approximate time-domain formula for
the calculation of the horizontal electric field from lightning. IEEE
Transactions on Electromagnetic Compatibility, 2007;49(3):593–601.
[73] Caligaris, C., F. Delfino, and R. Procopio, Cooray-Rubinstein formula for
the evaluation of lightning radial electric fields: derivation and imple-
mentation in the time domain. IEEE Transactions on Electromagnetic
Compatibility, 2008;50(1):194–7.
[74] Cooray, V, Underground electromagnetic fields generated by the return
strokes of lightning flashes. IEEE Transactions on Electromagnetic
Compatibility, 2001;43(1):75–84.
[75] Petrache, E., F. Rachidi, M. Paolone, C. Nucci, V.A. Rakov, and M.A.
Uman, Lightning-induced voltages on Buried Cables”, Part I: theory. IEEE
Transactions on Electromagnetic Compatibility, 2005;47(3).
[76] Degauque, P. and A. Zeddam, Remarks on the transmission-line approach
to determining the current induced on above-ground cables. IEEE
Transactions on Electromagnetic Compatibility, 1988;30(1):77–80.
[77] Tirkas, P.A., C.A. Balanis, M.P. Purchine, and G.C. Barber, Finite-
difference time-domain method for electromagnetic radiation, inter-
ference, and interaction with complex structures. IEEE Transactions on
Electromagnetic Compatibility, 1993;35(2):192–203.
[78] Paolone, M., C.A. Nucci, and F. Rachidi, A new finite difference time
domain scheme for the evaluation of lightning induced overvoltages on
multiconductor overhead lines, in 5th International Conference on Power
System Transients, 2001, Rio de Janeiro.
[79] Sartori, C.A.F. and J.R. Cardoso, An analytical-FDTD method for near
LEMP calculation. IEEE Transactions on Magnetics, 2000;36(4):1631–4.
[80] Yang, C. and B. Zhou, Calculation methods of electromagnetic fields very
close to lightning. IEEE Transactions on Electromagnetic Compatibility,
2004;46(1):133–41.
[81] Mimouni, A., F. Rachidi, and Z. Azzouz, Electromagnetic environment in
the immediate vicinity of a lightning return stroke. Journal of Lightning
Research, 2007;2:64–75.
[82] Mimouni, A., F. Rachidi, and Z. Azzouz, A finite-difference time-domain
approach for the evaluation of electromagnetic fields radiated by lightning
to tall structures. Journal of Electrostatics, 2008;866:504–13.
[83] Baba, Y. and V.A. Rakov, On the mechanism of attenuation of current
waves propagating along a vertical perfectly conducting wire above
ground: application to lightning. IEEE Transactions on Electromagnetic
Compatibility, 2005;47(3):521–32.
[84] Baba, Y. and V.A. Rakov, Electromagnetic fields at the top of a tall
building associated with nearby lightning return strokes. IEEE
Transactions on Electromagnetic Compatibility, 2007;49(3):632–43.
198 Lightning electromagnetics: Volume 2
[85] Moini, R., B. Kordi, G.Z. Rafi, and V.A. Rakov, A new lightning return
stroke model based on antenna theory. Journal of Geophysical Research,
2000;105(D24):29693–702.
[86] Pockels, F., Über die Blitzentladungen erreicht Strömstärke, in Phys. Z.
1900. p. 2.
[87] Foust, C.M. and H.P. Kuehni, The Surge Crest Ammeter. General Electric
Review, 1932;35:5.
[88] Foust, C.M. and G.F. Gardner, General Electric Review, 1934;35:324327.
[89] Wagner, C.F. and G.D. McCann, Transactions on American Institute of
Electrical Engineers, 1940;59:442–449.
[90] Grünewald, H., Die Messung von Blitzstromstärken an Blitzableitern und
Freileitungsmasten. Elektrotechnische Zeitschrift, 1934;22:4.
[91] Zaduk, H., Neuere Ergebnisse der Blitzstromstärkemessungen an
Hochspannungsleitungen. Elektrotechnische Zeitschrift, 1935;17:5.
[92] Popolansky, F., Lightning current measurement on high objects in
Chechoslovakia, in 20th International Conference on Lightning Protection
ICLP, 1990, Interlaken.
[93] Kulikow, D., Blitzströme, in 14th International Conference on Lightning
Protection ICLP, 1978, Danzig.
[94] Miladowska, K., Vergleichung der Blitzresultate von hohen Schornsteinen
und den Zählern, in 14th International Conference on Lightning Protection
ICLP, 1978, Danzig.
[95] McEachron, K.B., Lightning to the empire state building. Journal of
Franklin Institute, 1939;227:149–217.
[96] Diendorfer, G., H. Pichler, and M. Mair, Some parameters of negative
upward initiated lightning to the Gaisberg Tower (2000–2007). IEEE
Transactions on Electromagnetic Compatibility, 2009. Submitted for
publication.
[97] Uman, M.A., Lightning. 1984, New York, NY: McGraw-Hill.
[98] Golde, R.H., Lightning. Vol. 1. 1977, London: Academic Press. p. 496.
[99] Berger, K, R.B. Anderson, and H. Kroninger, Parameters of lightning fla-
shes. Electra. 1975;41:23–37.
[100] Uman, M.A., The lightning discharge. Vol. 39. 1987, Florida: Academic
Press, Inc. 377.
[101] Cortina, R., E. Garbagnati, W. Serravalli, L. Dellera, A. Pigini, and L.
Thione, Quelques aspects de l’évaluation des performances des réseaux
électriques vis-à-vis de la foudre, CIGRE, 1980, Paris.
[102] Fisher, R.J., G.H. Schnetzer, R. Thottappillil, V.A. Rakov, M.A. Uman, and
J.D. Goldberg, Parameters of triggered-lightning flashes in Florida and
Alabama. Journal of Geophysical Research, 1993;98(D12):22887–902.
[103] Narita, T., T. Yamada, A. Mochizuki, E. Zaima, and M. Ishii, Observation
of current waveshapes of lightning strokes on transmission towers. IEEE
Transactions on Power Delivery, 2000;15(1):429–35.
[104] Diendorfer, G., M. Mair, and W. Schulz, Detailed brightness versus light-
ning current amplitude correlation of flashes to the Gaisberg tower, in 26th
Modeling lightning strikes to tall towers 199
6.1 Introduction
During the past decades, much attention has been paid to the problem of the
interaction between lightning electromagnetic fields and overhead and buried
conductors. This has led to the formulation of different reliable field-to-
transmission line coupling models [1,2]. All these models require an accurate
evaluation of the lightning electromagnetic fields along the line, taking into
account the effect of the ground finite conductivity, since the approximation of
perfectly conducting ground becomes unacceptable especially in the evaluation of
the horizontal electric field [3].
Different models can be adopted to represent a lossy soil; the simplest one
assumes that both the conductivity s and the relative permittivity er are constant.
Enhanced representations could then consider the influence of the working fre-
quency on s and er or their dependence on the depth, in the case of a horizontally
stratified ground. In the following, such configurations will be examined with the
final aim to derive lightning field expressions (radial and vertical electric field,
azimuthal magnetic one) and to propose numerical procedures for their efficient
and accurate calculation. It is important to highlight from the beginning that the
presence of a ground with finite conductivity will be responsible of the appearance
in field expressions of the so-called Sommerfeld’s integrals. The numerical eva-
luation of such integrals represent a hard task due to the singular, oscillating and
divergent behaviour of their integrands and, as a consequence, suitable strategies
have to be identified to guarantee at the same time precision in execution and
acceptable computational costs.
1
Department of Electrical, Electronics and Telecommunication Engineering and Naval Architecture
Department (DITEN), University of Genoa, Italy
202 Lightning electromagnetics: Volume 2
The chapter is organized as follows. In Section 6.2, the ground parameters will
be considered constant; under such assumption both the derivation and the calcu-
lation of the lightning electromagnetic field components will be presented.
Next, in Section 6.3, the influence of the frequency dependent behaviour of the
ground electrical parameters will be studied. Finally, in Section 6.4, the problem of
the derivation of the lightning electromagnetic fields over a stratified conducting
ground will be faced.
(namely the component that mostly affects line overvoltages). Among them, the
most remarkable ones are those proposed by Cooray and Rubinstein [32–34]. Such
formulas have been intensively employed for the evaluation of lightning induced
overvoltages on overhead power line [3,35]. Approximate approaches were also
developed for the numerical evaluation of Sommerfeld’s integrals appearing in the
expressions of the lightning fields [36,37] or for maximum overvoltage estimation
[38] (a detailed analysis of the available approximate expressions is the focus of the
next chapter).
In this section, our aim is to present in a systematic way the derivation of the
electromagnetic fields generated by a lightning return stroke current, respectively
in the half-space “air” and in the half-space “lossy ground,” and to propose an
efficient method for the numerical computation of the field expressions.
Lu ¼ r ðruÞ þ k2 u ¼ F (6.1)
be a partial differential equation defined in the domain A <3 , associated with the
well-known Sommerfeld’s radiation condition at the infinity [5]
pffiffi @u
lim r jku ¼ 0 (6.2)
r!1 @r
where d() is the Dirac function and P and P0 are respectively the observation and
the source point.
It can be shown that [40]
ð
u ¼ wFdP0 (6.4)
A
Equation (6.4) states that the solution of problem (6.1) for a generic right-hand-
side F can be easily obtained by solving only once problem (6.3) and calculating
the particular solution w, which is called Green’s function [39,40].
204 Lightning electromagnetics: Volume 2
az
Z P
P (r, I, Z' )
Z
ay
+ R
P' (0, 0, Z') I r
– a)
ax
ar
HPV
Air Hr,P0,V= 0
Z=0
Ground Hr,P0,V
Figure 6.1 Model geometry: the dipole radiation over a conducting ground
The numerical treatment of Sommerfeld’s integrals 205
having indicated with A and AE the !vertical components of the vector potential in
air and earth respectively and with e 3 the vertical axis unit vector. So, in the fre-
quency domain, the set of equations solving this problem is the following:
8 ! ! !
>
> DA þ k2 A ¼ m0 dðP P0 Þe 3 z > 0;
>
>
>
>
! !
DA þ kE 2 A ¼ 0 z < 0;
>
>
>
>
>
< A ¼ AE z ¼ 0;
(6.6)
>
> 2 @A @AE
>
> n ¼ z ¼ 0;
>
> @z @z
>
>
>
> lim pffiffir @A jkA
>
: _ ¼ 0;
r!1 @r
where: k2 ¼ w2 m0 e0 is the wave number in air (being w 2the angular frequency),
k
k2E ¼ w2 em0 þ jwm0 s the wave number in the earth n2 ¼ kE2 the complex refractive
index and d the Dirac function. The first and second Helmholtz equations hold
respectively in air and in earth, while the third and fourth ones are the interface
conditions, which must be satisfied in order to ensure the continuity of the tan-
gential magnetic and electric fields at ground level. Sommerfeld showed [5] that,
with the addition of the fifth equation, the well-known radiation condition, the
problem has a unique solution.
It should be observed that the set of (6.6) has been derived assuming for all
variables a time-harmonic dependence of the kind ejwt .
Once (6.6) has been solved and the vector potential spatial distribution has
been obtained, the expression of the three non-zero components of the field can be
easily derived according to the following relationships:
8
>
> jw @ 2 A
>
> E z ¼ þ k 2
A ;
>
> k2 @z2
>
>
<
jw @ 2 A
> Er ¼ 2 ; (6.7)
>
> k @z@r
>
>
>
> 1 @A
>
: Hj ¼ ;
m0 @r
A being the vertical component of the vector potential.
In order to solve problem (6.6), it is first necessary to face the simplified case
of the vertical dipole radiation in free space (Figure 6.2). Such problem is described
only by the first and the fifth equation of system (6.6), respectively the governing
equation and the condition at the infinity. Of course, due to the different symmetry
of the problem, a system of spherical coordinates is now used.
As well-known [5], the solution in terms of the vector potential spatial dis-
tribution is the following:
m0 ejkR
AðRÞ ¼ (6.8)
4p R
206 Lightning electromagnetics: Volume 2
H) Er
ϑ r Eϑ
I
di
y
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
R ¼ r2 þ ðz z0 Þ2 being the distance between the source point and the obser-
vation one. Developing (6.8) as a superposition of the Helmholtz equation eigen-
functions in a cylindrical domain [5], after some mathematical manipulations [41],
one gets:
ð
m0 ejkR m0 þ1 l mjzz0 j
¼ e J0 ðlrÞdl (6.9)
4p R 4p 0 m
being m2 ¼ l2 k2 and J0 the Bessel function of the first kind and zeroth order.
Coming back now to the primary problem of Figure 6.1, one has to modify
(6.9) by adding a further integral term, in order to satisfy the interface conditions
(at z = 0) of system (6.6). Such term can have the following form:
ð
m0 þ1 0
GðlÞ J0 ðlrÞ emðzþz Þ dl (6.10)
4p 0
This is again a superposition of eigenfunctions and the function G appearing in
it, which can be seen as the spectral distribution in the l-continuum of the eigen-
functions [5], will be determined imposing the interface conditions.
Similarly, the expression for AE can be searched in the following form:
ð
m0 þ1 mE zmz0
AE ¼ GE ðlÞ J0 ðlrÞ e dl (6.11)
4p 0
m2E ¼ l2 k2E being and GE an unknown function, whose meaning is the same as G.
The numerical treatment of Sommerfeld’s integrals 207
and finally:
l 2mE
GðlÞ ¼ 1 2 ;
m n m þ mE
(6.13)
2n2
GE ðlÞ ¼ l 2 ;
n m þ mE
which can be put into (6.10) and (6.11) respectively to obtain the expression for A
and AE.
As far as A is concerned, one therefore has that:
ð þ1
m ejkR l 2m 0
Aðr; z; z0 Þ ¼ 0 þ 1 2 E J0 ðlrÞ emðzþz Þ dl
4p R 0 m n m þ mE
(6.14)
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Now, recalling (6.9) and defining R0 ¼ r2 þ ðz þ z0 Þ2 , it readily follows that:
0 ð
m0 ejkR m0 þ1 l mðzþz0 Þ
¼ e J0 ðlrÞdl (6.15)
4p R0 4p 0 m
which means that the part of the vector potential A due to the first term of G can be
interpreted as the image effect.
Equation (6.14) can be then rewritten as:
0 ð þ1
m ejkR ejkR mE l mðzþz0 Þ
A¼ 0 þ 0 2 J0 ð lrÞ e dl (6.16)
4p R R 0 n 2 m þ mE m
Here, it should be noticed that only the first two terms would be present if the
ground conductivity was zero and so they represent the expression for the “ideal”
vector potential (i.e. the one corresponding to perfectly conducting ground), while
the last one takes into account the ground conductivity (i.e. the Sommerfeld inte-
gral). Inserting (6.16) into system (6.7), it is now possible to obtain the expression
for the fields:
8 ð þ1
>
> j mE mðzþz0 Þ l
3
>
> E z ¼ E zi J 0 ð lr Þ e dl;
>
> 2pwe0 0 n2 m þ mE m
>
> ð þ1
<
j 0 m
Er ¼ Eri l2 J1 ðlrÞemðzþz Þ 2 E dl; (6.17)
>
> 2pwe n m þ mE
>
>
0 0
>
> ð
>
> 1 þ1 mE mðzþz0 Þ l
2
: Hj ¼ Hji J1 ð lr Þ e dl;
2p 0 n2 m þ mE m
208 Lightning electromagnetics: Volume 2
where the terms Eri, Ezi, and Hji are the “ideal” fields, whose expressions can be
found in [42].
Air H0,P0,V= 0
H
i(z',t)
dz’
R
z’ P (r, I, Z)
Ground Hr,P0,V
Figure 6.3 Model geometry: the lightning radiation over a conducting ground
The numerical treatment of Sommerfeld’s integrals 209
The expressions for the “ideal field” terms, namely EziL, EriL, and HjiL, are [3]:
ð " #
Ið0; wÞ H 2ðz z0 Þ2 r2 2ðz z0 Þ2 r2 r2
EziL ¼ þ þ jw 2 3
4pe0 H cR4 jwR5 c R
0
Rþjz j
ejw Pðz0 Þdz0
c v (6.21)
ð jz0 j
Ið0; wÞ H 3rðz z0 Þ 3rðz z0 Þ r ðz z 0 Þ jw Rcþ v
EriL ¼ þ jw e Pðz0 Þdz0
4pe0 H cR4 jwR5 c2 R 3
(6.22)
and
ðH 0
Ið0; wÞ r r R jz j
HjiL ¼ 3
jw 2 ejw c þ v Pðz0 Þdz0 (6.23)
4p H R cR
Im [u]
kE/k
Re [u]
In order to get the expressions for m and mE, one can make the following
considerations.
Let us start with mE. As stated before, one must choose:
ReðmE Þ > 0 (6.25)
Recalling the definition of m2E ,
it readily follows that Im < 0. So, indi- m2E
cating with f the phase angle one has that p < f < 2p s a consequence, if q is
of mE2,
the phase angle of mE, one has that either q ¼ 2f or q ¼ 2f þ p can be chosen. In the
p
second case 3p 2 < q < 2p, while in the first one 2 < q < p. So, to meet the
f
requirement (6.25), q ¼ 2 þ p must be chosen and, as a consequence, it results:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
f f 1 þ cos f
cos q ¼ cos þ p ¼ cos ¼ þ (6.26)
2 2 2
and
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
f f 1 cos f
sin q ¼ sin þp ¼ sin ¼ (6.27)
2 2 2
Now, expressing mE in polar form, it follows:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 þ cos f 1 cos f
mE ¼ jmE jð j Þ (6.28)
2 2
being
l2 ee0 k2
cos f ¼ rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi (6.29)
2
l k
2 e
e0
2
þ s2 k2 me00
The numerical treatment of Sommerfeld’s integrals 211
1.2
1.0
0.8
0.6
real part
0.4
0.2
0.0
–0.2
–0.95 2.00 2.05 2.10 2.15 2.20
O(1/m)
Figure 6.5 Real part of the s-dependent term for k = 2.09 105
and
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 ffi
4 e m
jmE j ¼ l2 k2 þ s2 k2 0 (6.30)
e0 e0
1.5
1.0
0.5
imaginary part
0.0
–0.5
–1.0
Figure 6.6 Imaginary part of the s-dependent term for k = 2.09 105
0.05
0.04
0.03
real part of the spectrum
0.02
0.01
–0.01
–0.02
–0.03
0 1 2 3 4 5 6 7 8 9 10
frequency [Hz] ×105
Figure 6.7 Failure of the classical routine for the Hankel transform
The numerical treatment of Sommerfeld’s integrals 213
(6.32)
having indicated with lower case letters the time domain functions.
Here, the range of integration is (–vt, vt) instead of (–H, H), since the current i
(z0 ,t) is identically zero for z0 >vt. This is the reason why, if one performs the fields
calculations in the time domain and is interested in the first microseconds of the
transient, the channel height is of no use.
As far as the Sommerfeld term is concerned, the inverse Fourier transform
must be carried out numerically. This requires to evaluate numerically the corre-
sponding Sommerfeld integral appearing in the first of system (6.19) for many
different values of frequency in an assigned range (i.e. [0 Hz, 107 Hz]). Moreover,
if one observes that, for each frequency:
● the integrand contains both the Bessel function (which is highly oscillating)
and the s-dependent term (whose behaviour is shown in Figures 6.5 and 6.6);
● the integrand is singular for l = k;
● the integral must be carried out over a semi-infinite domain,
one is easily convinced that such calculation requires a huge computational effort.
To overcome the first problem, a Romberg method [47–49] has been used,
dividing the interval of integration into sub-intervals, with particular attention to
the neighborhood of k, namely:
ð1 ð 0:99 ð1 ð 1:01 ð1
k f ðkuÞdu ¼ k f ðkuÞdu þ k f ðkuÞdu þ k f ðkuÞdu þ k f ðkuÞdu
0 0 0:99 1 1:01
(6.33)
having posed u ¼ lnkand having indicated with f the integrand function of the
first integral of system (6.19).
As far as the singularity is concerned, let us consider the integral I1 between
0.99 and 1:
ð1
u
I1 ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi h1 ðuÞdu (6.34)
0:99 1 u2
214 Lightning electromagnetics: Volume 2
with
pffiffiffiffiffiffiffiffi
2
h1 ðuÞ ¼ J0 ðkruÞk3 u2 gs ðkuÞejk 1u z QðkuÞ (6.35)
pffiffiffiffiffiffiffiffiffiffiffiffiffi
Setting s ¼ 1 u2 , one has that:
pffiffiffiffiffiffiffiffiffiffiffiffi
ð 10:992 pffiffiffiffiffiffiffiffiffiffiffiffiffi
I1 ¼ h1 1 s2 ds (6.36)
0
being
pffiffiffiffiffiffiffiffi
h2 ðuÞ ¼ J0 ðkruÞk3 u2 gs ðkuÞek u 1z QðkuÞ
2
(6.38)
pffiffiffiffiffiffiffiffiffiffiffiffiffi
Setting now s ¼ u2 1, one has that:
pffiffiffiffiffiffiffiffiffiffiffiffi
ð 1:012 1 pffiffiffiffiffiffiffiffiffiffiffiffiffi
I2 ¼ h2 1 þ s2 ds (6.39)
0
×10–11
3.5
2.5
Error/Integral
1.5
0.5
0
0 1 2 3 4 5 6 7 8 9 10
Frequency [Hz] ×10–6
Figure 6.8 Ratio between the upper bound of the tail and the absolute value of the
integral between 0 and M as a function of the frequency
one has
pffiffiffi ð
2 2k3 1 kuz 3
ErrðMÞ jgs ðkMÞQðkMÞj pffiffiffiffiffiffiffi e 2 u2 du (6.42)
pkr M
Finally, since the integral in relation (6.42) is known analytically [41], it follows:
pffiffiffi 5
2 2k3 kz 2 5 kMz
ErrðMÞ jgs ðkMÞQðkMÞj pffiffiffiffiffiffiffi G ; (6.43)
pkr 2 2 2
being G the incomplete Gamma function [41].
In Figure 6.8, the ratio between the upper bound of the tail and the absolute
value of the integral between 0 and M has been plotted as a function of the fre-
quency, having set M = 4p/5k, showing that our goal has been achieved.
The position u ¼ lnk is not possible for f = 0 Hz. This implies that another
method must be used to perform the so-called static term.
Let us reconsider the Sommerfeld integral appearing in the first of (6.17);
observing that, if k approaches 0:
8
>
> n2 !
js
;
>
>
>
> we0
>
>
< m 2 ! l2 ;
E
(6.44)
>
> m 2
! l2 ;
>
>
>
>
> mE
> we0
: 2 ! ;
n m þ mE js
216 Lightning electromagnetics: Volume 2
(6.47)
having indicated with lower case letters the time domain functions.
As far as the Sommerfeld term is concerned, again the inverse Fourier trans-
form must be carried out numerically. This requires to numerically evaluate the
corresponding Sommerfeld integral appearing in the second of (6.19) for many
different values of frequency in an assigned range (i.e. [0 Hz, 107 Hz]). From a
numerical point of view, here things go better, since the integral function is not
singular for l = k.
Therefore, again a Romberg method is used, splitting the range of integration
into many sub-intervals, as done for the z-component of the electric field, but the
neighborhood of k does not require a ‘special treatment’ as before.
The only problems to solve are the ones relevant to the integral tail and the
static term. Indicating again with Err the upper-bound for the integral tail and with
The numerical treatment of Sommerfeld’s integrals 217
M the last point of the interval on which the integral is taken, one has:
ð1 pffiffiffiffiffiffiffiffi
J1 ðkruÞk3 u2 gs ðkuÞek u 1z QðkuÞdu
2
ErrðMÞ ¼ (6.48)
M
With considerations similar to the ones done in the previous section, it follows
that:
pffiffiffi 3 ð 1
2k
ek2z u2 du
u 3
ErrðMÞ jgs ðkMÞQðkMÞj pffiffiffiffiffiffiffi (6.49)
pkr M
and again [41]:
pffiffiffi 3 5
2k kz 2 5 kMz
ErrðMÞ jgs ðkMÞQðkMÞj pffiffiffiffiffiffiffi G ; (6.50)
pkr 2 2 2
ð
1 þ1 0 3r 1 r2
! J1 ðlrÞ elðzþz Þ l2 dl ¼ F 2; ; 2; (6.51)
2ps 0 2psR04 2 R02
As a consequence, the static term for the Sommerfeld integral appearing in the
second of (6.19) is given by:
ð þ1
jIð0; wÞ mE 0
limk!0 J1 ðlrÞ emz l2 QðlÞdl
2pwe0 0 n2 mþ mE
ðH
3rI ð0; 0Þ 1 1 r2
¼ F 2; ; 2; 02 Pðz0 Þdz0 (6.52)
2ps 0 R04 2 R
(6.53)
having indicated with lower case letters the time domain functions.
218 Lightning electromagnetics: Volume 2
az
z P
z
ay
+
P' (0,0,z') ϕ r
– aΦ
ax
R ar
P (r, ϕ, z)
Ground ε, μ0, σ
Assuming now the same expression for the current distribution in the lightning
channel used in the previous case (see (6.18)), with the same attenuation function P
(z0 ), the lightning fields are given by:
8 ð
>
> jIð0; wÞ þ1 l3
>
> EzL ¼ J0 ðlrÞ QðlÞemE z dl;
>
> 2pwe n 2m þ m
>
>
0 0 E
>
< ð
jIð0; wÞ þ1 l2 m
ErL ¼ J1 ðlrÞQðlÞemE z 2 E dl; (6.63)
>
> 2pwe0 0 n m þ mE
>
>
>
> ð
>
> n2 Ið0; wÞ þ1 l2
>
: HjL ¼ J1 ðlrÞ QðlÞemE z dl:
2p 0 n2 m þ mE
From a numerical point of view, the same considerations can be done as in the
previous subsection, concerning the troubles in the evaluation of the integrals, due
to the characteristics of the integrand functions.
Here, for the sake of brevity, we simply give the expressions for the upper
bound of the error that is made truncating the integral tail and for the static term for
the three nonzero components of the fields.
Vertical electric field:
– upper bound:
pffiffiffi 4 7
2k kz 2 7 kMz
EðMÞ jgsz ðkMÞQðkMÞj pffiffiffiffiffiffiffi G ;
pkr 2 2 2
– static term:
ð
jIð0; wÞ þ1 mE
limk!0 J1 ðlrÞ emE z l2 QðlÞdl
2pwe0 0 n2 m þ mE
ð
3rIð0; 0Þ H 1 1 r2
¼ 4
F 2; ; 2; 2
Pðz0 Þdz0
2ps 0 R 2 R
Azimuthal magnetic field:
– upper bound:
pffiffiffi 3
2 2k3 kz 2 5 kMz
EðMÞ jgsz ðkMÞQðkMÞj pffiffiffiffiffiffiffi G ;
pkr 2 2 2
– static term:
ð
n2 Ið0; wÞ þ1 l2
limk!0 J1 ðlrÞ emE z QðlÞdl
2p 0 n2 m þ mE
ð
rIð0; 0Þ H 1 3 r2
¼ 3
F ; 0; 2; 2
Pðz0 Þdz0
2p 0 R 2 R
R1 Rn–1 Rn
R0 C00 C1 Cn–1 Cn
1 XN
1
Y¼ þ jwC1 þ (6.65)
R0 R þ jwC
n¼1 n
1
n
and
1 X
N
Cn bn ðw=bn Þ2
s¼ þ (6.67)
R0 n¼1 1 þ ðw=bn Þ2
with bn ¼ Rn1Cn .
Now, defining
8 C1
>
> e1 ¼ ;
>
> e0
>
>
>
>
>
> 1
< s0 ¼ ;
R0
(6.68)
>
> C
> an ¼ n ;
>
>
> e0
>
>
>
>
: f ¼ bn ;
n
2p
The relative permittivity er and the conductivity can be written, respectively, in
the following form
X
N
an
e r ðf Þ ¼ e 1 þ (6.69)
n¼1 1 þ ðf =f n Þ2
X
N
an f n ðf =f n Þ2
sðf Þ ¼ s0 þ 2pe0 (6.70)
n¼1 1 þ ðf =f n Þ2
224 Lightning electromagnetics: Volume 2
Finally, in order to have good fitting between (6.69) and (6.70) and the
experimental curves obtained by Scott, Longmire and Smith [55] have shown that it
is sufficient to set:
e1 ¼ 5 (6.71)
p 1:54 S
s1 ¼ 8 103 (6.72)
10 m
p 1:28
fn ¼ 10n1 ½Hz (6.73)
10
where p is the percent water volume, which typically ranges between 2 and 30.
Lower values imply lots of “rock-like” content. The values for the coefficients an
are reported in Table 6.1. Our analysis assumes radially homogenous soil type and
moisture content.
Otherwise, it would be necessary to take into account the fact that, for any
given location, the moisture content in the top meter varies seasonally by at least
a factor of two, leading to a (roughly) factor-of-four gradient in electrical
conductivity in the top two meters of soil. This creates an additional frequency-
dependent behavior at higher frequencies due to dependence of skin depth on
frequency. The hypothesis of radially homogenous soil justifies (6.72) and (6.73)
and allows us not to consider such an effect.
As an example, Figures 6.11 and 6.12 present the variation of conductivity and
relative permittivity as a function of the frequency in the range [0–5106] Hz and
for water content p equal to 0.2%, 1% (to simulate rock-like ground), 2%, 10%, and
30% respectively.
Others formulations for the frequency-dependence of the soil conductivity and
permittivity have been proposed by Portela in [57] and Visacro et al. in [58].
Portela carried out a series of measurements which comprises experimental data
obtained in several geological areas in Brazil and considers soil samples measured
from 100 Hz up to 2 MHz. The value of the effective conductivity s, as well as
relative permittivity, is expressed as a function of the low frequency conductivity
s0 obtained from the measured 100 Hz soil resistivity according to (6.74)
a
p w
s þ jwe s0 þ Di cot ang a þ j (6.74)
2 2p 106
n 1 2 3 4 5 6 7 8 9 10 11 12 13
6 5 4 3 2 2 1 1
an 3.410 2.7410 2.5810 3.3810 5.2610 1.3310 2.7210 1.2510 4.8 2.17 0.98 0.392 0.173
The numerical treatment of Sommerfeld’s integrals 225
10–1
10–2
Sigma [S/m]
10–3
10–4
p
10–5
102 103 104 105 106 107
Frequency [Hz]
Figure 6.11 Ground conductivity as a function of the frequency for water content
p equal to 0.2%, 1%, 2%, 10%, and 30%
105
104
relative permittivity
103
102
101
P
100
102 103 104 105 106 107
Frequency [Hz]
Figure 6.12 Ground relative permittivity as a function of the frequency for water
content p equal to 0.2%, 1%, 2%, 10%, and 30%
226 Lightning electromagnetics: Volume 2
where s0 is the low-frequency conductivity (100 Hz) in mS/m, er1 is the relative
permittivity at higher frequencies and f is measured in Hz. The parameters hðs0 Þ, g
and er1 are:
8 ð
> jIð0; wÞ þ1 l3
>
> E ¼ J0 ðlrÞ QðlÞemE z dl;
>
>
zL
2pwe0 0 n2 m þ mE
>
>
>
> ð
<
jIð0; wÞ þ1 l2 m
ErL ¼ J1 ðlrÞQðlÞemE z 2 E dl; (6.78)
>
> 2pwe0 0 n m þ mE
>
>
>
> ð
>
> n2 Ið0; wÞ þ1 l2
>
: HjL ¼ J1 ðlrÞ QðlÞemE z dl:
2p 0 n2 m þ m E
for the underground one. Equations (6.77) and (6.78) are formally identical to
(6.19) and (6.63), since the dependence of the soil parameters on the frequency is
hidden in the definition of the quantities n2 and mE.
In [59], a detailed analysis on the effects of taking into account the frequency
dependent behavior of the soil characteristics in the evaluation of the electro-
magnetic fields is presented. Here, we simply limit our discussion to summarize in
Table 6.2 the main results.
Table 6.2 Summary of the results and recommended models for the calculation of
electromagnetic fields generated by lightning return-strokes
In the previous sections, the ground parameters have been considered either con-
stant or frequency dependent; in both cases, they have been treated as
homogeneous.
The aim of the present section is to investigate the effect of the soil stratifi-
cation on the evaluation of the lightning electromagnetic fields.
One of the first studies on the propagation of electromagnetic waves along a
stratified medium is due to Wait, who showed that the concepts of ground surface
impedance and attenuation function can be used to represent the effect of a multi-
layered soil [60]. For the case of the lightning radiation over a stratified conducting
ground, a recent literature review can be found in [61]. Here, the formulation
proposed by Wait in [60] for a dipole is extended to account for the presence of the
lightning channel and the derivation of the final formulas is presented in details
starting from the field problem, consisting of one Helmholtz equation for each layer
and the suitable boundary conditions [5,42].
Air ε0,μ0,σ=0
H
i(z',t)
dz'
R
z' P(r,ϕ,z)
Layer 2 ε2,μ0,σ 2
Figure 6.13 Model geometry: the lightning radiation over a multilayered ground
The numerical treatment of Sommerfeld’s integrals 229
! !
DA E1 þ kE1 2 A E1 ¼ 0 h1 < z < 0 (6.80)
! !
DA E2 þ kE2 2 A E2 ¼ 0 z < h1 (6.81)
pffiffi @A
lim r jkA ¼ 0 (6.82)
r!1 @r
1 @A 1 @AE
¼ z¼0 (6.84)
k 2 @z kE1 2 @z
1 @AE1 1 @AE2
2 @z
¼ z ¼ h1 (6.86)
kE1 kE2 2 @z
! ! !
where A is the vector potential in the upper half-space (air), A E1 A E2 is the vector
potential in the first (second) ground layer, k 2 ¼ w2 m0 e0 is the wave number in air
(w being the angular frequency), kE1ð2Þ 2
¼ w2 e0 er1ð2Þ m0 þ jwm0 s1ð2Þ is the wave
number in the first (second) ground layer and d is the Dirac distribution. Equations
(6.79)–( 6.81) are Helmholtz equations that hold respectively in air and in the two
earth layers, while (6.83)–(6.86) are the interface conditions, which must be satis-
fied in order to ensure the continuity of the tangential magnetic and electric fields at
ground level and at the transition between the first and the second layers. Finally, as
shown by Sommerfeld [5], with the addition of the well-known radiation condition
(6.82), the problem has a unique solution.
It should be observed that the set of (6.79)–(6.86) has been derived assuming,
as in the previous sections, for all variables a time-harmonic dependence of the
kind ejwt .
In order to solve these equations [5], one can first solve (82) and then modify
the obtained solution in order to meet the interface conditions. Sommerfeld himself
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
m0 ejkR
showed that Ap ¼ 4p R with R ¼ r2 þ ðz z0 Þ2 is the solution of (6.79) and
(6.82). Moreover, he proved that
ð
m0 þ1 l mjzz0 jJ0 ðlrÞdl
Ap ¼ e (6.87)
4p 0 m
where m2 ¼ l2 k2 and J0 is the Bessel function of the first kind and zeroth order.
Thus, it is sufficient to find a solution of the homogeneous Helmholtz equation,
which, added to Ap, meets the interface conditions to find the final solution of the
problem. It can be easily proven that the function u defined as:
where the functions b0, a1, b1 and b2 can be determined imposing the four interface
conditions and m2Ei ¼ l2 kEi 2
; i ¼ 1; 2. It should be noticed that the coefficient of
the term emE2 z is set to zero in order to ensure the convergence of the integral when
z ! 1. Imposing the interface conditions, one obtains that:
l
b0 ðlÞ ¼ RðlÞ (6.92)
m
R being the reflection coefficient [60], defined as:
w0 ðlÞ z1 ðlÞ
R ðl Þ ¼ (6.93)
w0 ðlÞ þ z1 ðlÞ
where
mE1
w0 ðlÞ ¼ (6.94)
kE1 2
and z1 is the surface impedance given by
w2 ðlÞ emE1 h1 þ emE1 h1 þ w1 ðlÞ emE1 h1 emE1 h1
z1 ðlÞ ¼ w1 ðlÞ (6.95)
w1 ðlÞðemE1 h1 þ emE1 h1 Þ þ w2 ðlÞðemE1 h1 emE1 h1 Þ
where the quantities w1 and w2 are defined as
mE1
w1 ðlÞ ¼ (6.96)
kE1 2
The numerical treatment of Sommerfeld’s integrals 231
and
mE2
w2 ðlÞ ¼ (6.97)
kE2 2
A brief discussion on the physical interpretation of the reflection coefficient is
in order: (6.87) states that, for each frequency, the vector potential Ap due to the
vertical dipole in the free space is a superposition of waves (belonging to a con-
tinuous spectrum) of the kind defined in (88), each characterized by a different
value of l and whose amplitude is l/m. The presence of the soil generates, for each
value of l, a reflected wave, whose amplitude is related to the corresponding
incident wave by the coefficient R.
Once the vector potential spatial distribution has been obtained, the expression
of the three non-zero components of the over-ground field can be easily derived
according to the following relationships:
jw @ 2 A
Ez ¼ 2 þ k2
A ;
k @z2
jw @ 2 A
Er ¼ 2 ; (6.98)
k @z@r
1 @A
Hj ¼ :
m0 @r
Substituting (6.89) and (6.92) into (6.98), one obtains the expression of the
fields which read:
2 3
2 ðz z 0 Þ2 r 2
6 cR4 7
6 7 R
6 7 ð þ1
1 6 6 2ð z z 0 2
Þ r 2 7 jw
7 c j 1 RðlÞ mðzþz Þ l
0 3
Ez ¼ þ e J ð lr Þe dl;
4pwe0 6 7 0
6 jwR5 7 2pwe0 0 2 m
6 7
4 r2 5
þjw 2 3
c R
2 3rðz z0 Þ 3
6 cR4 7
6 7 R ð þ1
6
1 6 3rðz z0 Þ 7 7
jw
c j
0 1 R ðlÞ
Er ¼ 6þ 7 e l2 J1 ðlrÞemðzþz Þ dl;
4pe0 6 jwR 7 5 2pwe0 0 2
6 7
4 5
rðz z0 Þ
jw 2 3
c R
0
jw R þ jz j ð
1 r r c v 1 þ1 1 RðlÞ mðzþz Þ l
0 2
Hj ¼ jw e J1 ð lr Þe dl:
4p R3 cR2 2p 0 2 m
(6.99)
232 Lightning electromagnetics: Volume 2
The first terms in the above expressions of the fields are as in the previous
sections called “ideal fields,” since they would be the only nonzero terms if the
ground was a perfect conductor.
In a similar way, one can easily determine the underground fields first by
obtaining the following expressions of the coefficients a1, b1, and b2 [which can be
done imposing again the interface conditions (6.83)–(6.86)],
8
> l l
>
> w1 þ b0 ðlÞ w0 b0 ðlÞ
>
> m m
>
> a ð l Þ ¼ ;
>
> 1
>
< 2w2
l l
w1 þ b0 ðlÞ þ w0 b0 ðlÞ (6.100)
>
> m m
> b1 ðlÞ ¼
> ;
>
> 2w2
>
>
>
> b ð lÞw
: b2 ðlÞ ¼
1 1
eh1 ðmE2 mE1 Þ ;
w1 þ w2
then inserting them in (6.90) and (6.91) and finally recalling the relationships
between the vector potential and the fields that state:
8 !
>
> jw @ 2
A
>
> Ez1ð2Þ ¼
E1ð2Þ
þ kE1ð2Þ 2 AE1ð2Þ ;
>
> kE1ð2Þ 2 @z2
>
> !
<
jw @ 2 AE1ð2Þ (6.101)
> Er1ð2Þ ¼ ;
>
> kE1ð2Þ 2 @z@r
>
>
>
> 1 @AE1ð2Þ
>
: Hj1ð2Þ ¼ :
m0 @r
where the expressions of the ideal fields EziL, EriL, and HfiL are those defined in
(6.21)–(6.23).
1 R ðl Þ m
lim ¼ 2 E ; (6.103)
h1 !1 2 n m þ mE
which confirms the fact that (6.102) reduces to the expression of the field in the
case of a one-layer soil. Then, in order to evaluate the static limit of the
Sommerfeld integrals (i.e., the DC fields), it is necessary to analyze the behavior of
the function (1R)/2 when w approaches zero. Since it is apparent that, for small
values of w
jsi l
ni 2 ! ; w0 ! 2
we0 k
l we0 (6.104)
mEi 2 ! l2 ; wi ! 2
k jsi
m2 ! l2 ;
which states that the limit is zero and the function (1R)/2 is linear in the neigh-
borhood of w = 0.
After some algebraic manipulations, (6.105) can be re-written as
1 R ðl Þ we0
lim ¼ f ðlÞ; (6.106)
k!0 2 js1
in which
0.6
0.5
h1=1 [m]
0.4 h1=10 [m]
h1=100 [m]
ratio
0.3
0.2
0.1
0 2 4 6 8 10
Iamda/k
Figure 6.14 Ratio between the function (1R)/2 and the function gs for f = 1 kHz
1.8
h1=1 [m]
1.6 h1=10 [m]
h1=100 [m]
1.4
1.2
ratio
0.8
0.6
0.4
0.2
0 2 4 6 8 10
Iamda/k
Figure 6.15 Ratio between the function (1R)/2 and the function gs for f = 1 MHz
The numerical treatment of Sommerfeld’s integrals 235
the ratios between the function (1R)/2 and the equivalent function for the case of
homogenous ground gs [62] are plotted for two different values of the frequency,
namely 1 kHz and 1 MHz. As can be seen from the figures, if the frequency
increases from 1 kHz to 1 MHz, it is sufficient to have a smaller depth of the first
layer to make the ratio become closer to one and therefore to make the corre-
sponding fields become closer one to each other.
In other words, the more high frequencies are “present” in the spectrum of the
lightning channel current, the closer are the corresponding fields to the ones which
would be present in the case of a homogeneous ground.
Figure 6.16 shows the absolute value of the reflection coefficient for four
different values of the frequency and in the case of a first layer depth of 1 m. As can
be seen, at higher frequencies the reflected wave is larger. The same observation
applies to the case of larger first layer depth, namely 10 m (Figure 6.17) and 100 m
102
100
10–2
R
10–4
Frequency
10–6
0 1 2 3 4 5 6 7 8 9 10
Iamda/k
102
100
10–2
R
10–4
Frequency
10–6
0 1 2 3 4 5 6 7 8 9 10
Iamda/k
102
100
10–2
R
10–4
Frequency
10–6
0 1 2 3 4 5 6 7 8 9 10
Iamda/k
(Figure 6.18), with the only difference that, for each frequency, the corresponding
values are higher. Extrapolating this result, we can conclude that the maximum
magnitude of the reflected wave is obtained for a homogeneous single-layer
ground.
6.5 Conclusions
The electromagnetic field due to a lightning event in presence of a lossy ground has
been studied considering three main situations:
1. homogeneous lossy ground with constant conductivity and permittivity;
2. homogeneous lossy ground with frequency-dependent soil electrical
parameters;
3. stratified lossy ground.
In all the cases, the exact field expressions have been derived starting from the
Maxwell equations both in air and in the ground. Particular attention has been
devoted to the numerical treatment of the Sommerfeld integrals involved in all the
field expressions and a suitable algorithm has been proposed in order to evaluate
them fastly and efficiently.
References
[1] C. R. Paul, Analysis of Multiconductor Transmission Lines. New York, NY:
John Wiley & Sons, 2007.
[2] F. M. Tesche, M. Ianoz, and T. Karlsson, EMC Analysis Methods and
Computational Models. New York, NY: John Wiley & Sons, 1996.
The numerical treatment of Sommerfeld’s integrals 237
[17] T. J. Cui and W. C. Chew, “Accurate model of arbitrary wire antennas in free
space, above or inside ground,” IEEE Transactions on Antennas and
Propagation, vol. 48, no. 4, pp. 482–493, 2000.
[18] T. Sarkar, “Analysis of radiation by arrays of parallel vertical wire antennas
over imperfect ground,” IEEE Transactions on Antennas and Propagation,
vol. 23, no. 5, pp. 749–749, 1975, doi: 10.1109/TAP.1975.1141153.
[19] P. R. Bannister, New Formulas that Extend Norton’s Farfield Elementary
Dipole Equations to the Quasi-Nearfield Range, Naval Underwater Systems
Center New London CT, 1984.
[20] K. A. Michalski and J. R. Mosig, “Multilayered media Green’s functions in
integral equation formulations,” IEEE Transactions on Antennas and
Propagation, vol. 45, no. 3, pp. 508–519, 1997, doi: 10.1109/8.558666.
[21] K. A. Michalski, “Extrapolation methods for Sommerfeld integral tails,”
IEEE Transactions on Antennas and Propagation, vol. 46, no. 10, pp. 1405–
1418, 1998, doi: 10.1109/8.725271.
[22] C. E. Baumann and E. E. Sampaio, “Electric field of a horizontal antenna
above a homogeneous half-space: implications for GPR,” Geophysics,
vol. 65, no. 3, pp. 823–835, 2000.
[23] J. R. Wait, “Influence of finite ground conductivity on the fields of a vertical
traveling wave of current,” IEEE Transactions on Electromagnetic
Compatibility, vol. 41, no. 1, p. 78, 1999, doi: 10.1109/15.748141.
[24] G. J. Burke, A. J. Poggio, J. C. Logan, and J. W. Rockway, “Numerical
electromagnetic code (NEC),” in 1979 IEEE International Symposium on
Electromagnetic Compatibility, 1979, pp. 1–3.
[25] R. L. Gardner, “Effect of the propagation path on lightning-induced transient
fields,” Radio Science, vol. 16, no. 03, pp. 377–384, 1981.
[26] J. R. Wait and D. A. Hill, “Ground wave of an idealized lightning return
stroke,” IEEE Transactions on Antennas and Propagation, vol. 48, no. 9,
pp. 1349–1353, 2000, doi: 10.1109/8.898767.
[27] R. K. Pokharel, M. Ishii, and Y. Baba, “Numerical electromagnetic analysis
of lightning-induced voltage over ground of finite conductivity,” IEEE
Transactions on Electromagnetic Compatibility, vol. 45, no. 4, pp. 651–656,
2003, doi: 10.1109/TEMC.2003.819065.
[28] H. K. Hoidalen, J. Sletbak, and T. Henriksen, “Ground effects on induced
voltages from nearby lightning,” IEEE Transactions on Electromagnetic
Compatibility, vol. 39, no. 4, pp. 269–278, 1997, doi: 10.1109/15.649810.
[29] C. Yang and Bihua Zhou, “Calculation methods of electromagnetic fields
very close to lightning,” IEEE Transactions on Electromagnetic
Compatibility, vol. 46, no. 1, pp. 133–141, 2004, doi: 10.1109/
TEMC.2004.823626.
[30] A. Shoory, R. Moini, S. H. H. Sadeghi, and V. A. Rakov, “Analysis of
lightning-radiated electromagnetic fields in the vicinity of lossy ground,”
IEEE Transactions on Electromagnetic Compatibility, vol. 47, no. 1,
pp. 131–145, 2005, doi: 10.1109/TEMC.2004.842104.
The numerical treatment of Sommerfeld’s integrals 239
The propagation of the lightning electromagnetic fields over and under a lossy
homogeneous ground has been intensively studied in the last decades. Researchers
have focused their attention mainly on the horizontal component of the electric
field for two reasons:
1. The finite ground conductivity mostly affects such component while the vertical
component of the electric field and the azimuthal component of the magnetic
one can be evaluated supposing perfectly conducting ground without compro-
mising the accuracy of the final result [1]. This assumption is generally valid but
some particular cases should be taken into account as presented in Section 7.1.2.
1
Department of Electrical, Electronics and Telecommunication Engineering and Naval Architecture
Department (DITEN), University of Genoa, Italy
244 Lightning electromagnetics: Volume 2
and where Er ðw; z; rÞ is the horizontal component of the electric field at angular
frequency w, height z and horizontal distance r from the lighting channel, while
Hji ðw; 0; rÞ is the azimuthal component of the magnetic field at the same frequency
and horizontal distance and at the ground level and Eri ðw; z; rÞ is the horizontal
component of the electric field both assuming perfectly conducting ground.
Moreover, m0, e, and s are the magnetic permittivity the dielectric constant and the
conductivity of the soil respectively. It should be noted that in (7.1) neglecting
propagation effects on the azimuthal magnetic field is reasonable for medium-high
values of the soil conductivity (greater than 0.001 S/m); it remains a valid
assumption only for distances lower than 200 m for conductivities of about 0.001 S/m
and lower than 100 m for conductivities of about 0.0001 S/m [6].
However, when the CR formula has to be inserted in time domain methods, an
IFFT procedure has to be adopted to evaluate the lightning horizontal field.
As well known, all the FFT-IFFT-based methods have two inherent drawbacks
to calculate the quantities of interest in time domain. One is the requirement of
huge computational resources, and the other is the error due to the truncation of the
Lightning electromagnetic field propagation 245
being s
s th s s i
K ðtÞ ¼ e 2e I0 t I1 t ¼ aeat ½I0 ðatÞ I1 ðatÞ
2e 2e 2e (7.9)
s
a¼
2e
246 Lightning electromagnetics: Volume 2
and where eri is the horizontal electric field at height z, distance r and time t,
calculated as if the ground was a perfect conductor and hfi the azimuthal magnetic
field at height z, distance r and time t, calculated as before in the case of perfectly
conducting ground. The terms eri and hfi are, respectively, the inverse Laplace
transform of Eri and Hfi appearing in (7.3) and can be evaluated directly in the time
domain, as presented in the previous chapter.
Finally, recalling the properties of the Dirac function, one has:
ðt
er ðt; z; rÞ ¼ h hji ðt; 0; rÞK ðt tÞdt þ eri ðt; z; rÞ hhji ðt; 0; rÞ (7.10)
0
With this assumption, one can compute the integral in (7.10) at each time
sample tn between t1 = 0 and tnp as:
ð tn n1 ð tmþ1
X
CIðtn Þ ¼ CIn ¼ Kðt tÞhji ðt; 0; rÞdt Kðtn tÞhm dt
t1 m¼1 tm
X
n1 ð tmþ1
¼ hm Kðtn tÞdt (7.12)
m¼1 tm
So (7.12) becomes
8
> CI ¼ 0
< 1
X
n1
(7.14)
>
: n
CI ¼ hm Knm ; n ¼ 2::np ;
m¼1
with f a suitable function such that limx!þ1 f ðxÞ ¼ 0 and where G is the incom-
plete Gamma function. It is apparent that the term ex is the responsible for the
overflow. The problem can be circumvented observing that in (7.5) the Bessel
functions In(x) are multiplied by the term ex. So, if one sets up a routine which
directly computes In(x)ex, he can avoid to calculate the exponential ex.
At time sample tn, one has to perform one more integral (7.13) and n1 pro-
ducts (and n1 sums); in other words, the number of computations to be performed
at each iteration grows linearly with the number of iterations, i.e. the algorithm has
a time complexity order n2.
For these reasons, the authors of [8] proposed the following improvement.
The surface impedance can be expressed as:
rffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffirffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sm0 m0 s stG
ZðsÞ ¼ ¼ ¼h (7.16)
se þ s e s þ s=e stG þ 1
with tG ¼ e=s.
If we can find a rational approximation (RA) for the function:
rffiffiffiffiffiffiffiffiffiffiffiffi
s0 X
NRA
rk
zðs0 Þ ¼ 0
1þ 0a
(7.17)
s þ1 k¼1
s k
where NRA is the number of poles used in the expansion, (7.3) can be approximated as:
X
NRA
rk Hji ðs; 0; rÞ
ECR ðs; z; rÞ ¼ Hji ð0; r; sÞZ ðsÞ ¼ hHji ðs; 0; rÞ h
k¼1
stG ak
(7.18)
The RA in (7.17) can be effectively computed by means of the vector fitting
(VF) technique [9]. It should be noticed that (7.17) and thus the poles ak and resi-
dues rk do not depend on tG , i.e., do not depend on the values of ground permit-
tivity and conductivity. Therefore, they have to be computed only once. The
computations illustrated as numerical examples in the last section were carried out
using an expansion with NRA=12, obtained by fitting the values of z sampled on 105
“frequencies” w0 [i.e., replacing s0 with jw0 in (7.17)], ranging from 10-4 to 100. The
fitting accuracy can be appreciated in Figures 7.1 and 7.2. All poles are real.
Table 7.1 shows the obtained values for poles and residues.
In order to re-state the problem in time domain, we rewrite (7.18) in the
classical state-space equation form. The method adopted here presents the advan-
tage of directly making use of the poles ak and residues rk from the vector fitting.
Furthermore, it allows a straightforward transformation from continuous time to
discrete time, and it is very simple to code (Table 7.1).
248 Lightning electromagnetics: Volume 2
1
[z]
0.5
0
0 10 20 30 40 50 60 70 80 90 100
N =12
ω' exact
0.4
0.3
[z]
0.2
0.1
0
0 10 20 30 40 50 60 70 80 90 100
ω'
0.8
0.6
[z]
0.4
0.2
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
N =12
ω' exact
0.4
0.3
[z]
0.2
0.1
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
ω'
ak rk
9.5767718047658079E01 2.4807255056699237E01
6.9826012409778959E01 1.4853965557307094E01
4.1509910805114025E01 6.4924454704342230E02
2.2286140573279206E01 2.4809658479110788E02
1.1343452528471004E01 8.9808401793780475E03
5.5577444158951196E02 3.1389835104789613E03
2.6246991903765934E02 1.0559668692280658E03
1.1888088239587109E02 3.3870559700515702E04
5.1013756832941235E03 1.0243347132959180E04
2.0025909738681470E03 2.8760866085666423E05
6.2900517731833501E04 6.9477373464759209E06
1.0993593611408876E04 7.8176216694433083E07
or:
8 ak 1
>
> sX ðsÞ ¼ Xk ðsÞ þ Hji ðs; 0; rÞ; k ¼ 1::NRA
>
< k t t
G G
X (7.20)
>
>
NRA
>
: CR
E ðs; z; rÞ ¼ hHji ð s; 0; r Þ h rk Xk ðsÞ
k¼1
we obtain:
Note that the coefficients eðak =tG ÞDt and ðeðak =tG ÞDt 1Þ=ak can be evaluated
only once and hold for all the iterations.
Differently from the previous approach, the algorithm corresponding to system
(7.25) requires about 3NRA products (and 2NRA sums), with NRA of the order of
ten; the number of computations at each iteration is constant, so the algorithm has a
time complexity order n. As observed in [10], the possible issue of this approach is
that the accuracy of the horizontal electric field is dependent on the performance of
the vector-fitting technique. Meanwhile, a frequency truncation is necessary when
using the vector-fitting technique, which could be an uncertain factor to influence
the final result. So, from 2011 on, researchers have mainly focused their attention
on 2 main issues: improve the computational efficiency of the convolution integral
(7.10) and analyze properties of the integral kernel (7.5) in order to find
suitable approximations that do not involve special functions.
Concerning the first research line, the first contribution was given by Zou et al.
in 2012 [10]. As summarized in [11], the method proposed in [10] proposed a
piecewise quadratic convolution and quadrature routine of a truncated integral
based on K to evaluate the convolution term in (7.10). More in details, given a
function f : D ! <, if one divides D in a finite number N of intervals (defining a
time sequence tn) and approximates f in each of them with a linear function, it
readily follows that:
X
N
f ðtÞ ðun þ vn tÞ½1ðt tn Þ 1ðt tnþ1 Þ (7.26)
n¼1
f ðtnþ1 Þ f ðtn Þ
un ¼ f ðtn Þ tn
tnþ1 tn
(7.27)
f ðtnþ1 Þ f ðtn Þ
vn ¼
tnþ1 tn
Lightning electromagnetic field propagation 251
So if one approximates both the kernel and the magnetic field as in (7.26), it
follows that:
X
N
KðtÞ ðan þ bn tÞ½1ðt tn Þ 1ðt tnþ1 Þ (7.28)
n¼1
and
X
M
hji ðt; 0; rÞ ðdm þ em tÞ½1ðt tm Þ 1ðt tmþ1 Þ (7.29)
m¼1
where the involved coefficients can be calculated as in (7.27), while M and N are
the number of intervals in which the magnetic field and the kernel are discretized,
respectively.
This way the convolution integral can be approximated as:
ðt
D
hji ðt; 0; rÞ KðtÞ ¼ hji ðt; 0; rÞKðt tÞdt
0
X X
N þ1 M þ1 ttm
b l2 g l3
’ amn l þ mn þ mn 1ðt tn tm Þ
n ¼1 m ¼1
2 3
l¼tn
(7.30)
where amn ¼ an dm , bmn ¼ an em bn dm and gmn ¼ bn em
The advantage of this method is that the numerical overflow problem and the
cancellation error due to the subtraction of two close big numbers (I0(x)I1(x)) can
be avoided. However, many time intervals are essential to approximate the kernel if
high accuracy is required. It is obvious that the computational cost would be quite
large, especially for the coupling simulation, because the horizontal electric fields
at different observation points should be calculated.
In 2017, Liu et al. proposed the following approach [12], based on the integral
form of the Bessel function:
ð
1 p x cosq
I n ðx Þ ¼ e cos ðnqÞdq (7.31)
p 0
So, the kernel K can be rewritten as:
ð
a p ð1 cosqÞat
K ðtÞ ¼ aeat ½I0 ðatÞ I1 ðatÞ ¼ e ð1 cos qÞdq (7.32)
p 0
With typical numbers of ground dielectric constant and conductivity, the
integrand in (7.32) vanishes after an angle qc which is significantly smaller than p
(the authors proposed 0.637p). This suggests the possibility of efficiently evaluat-
ing the integral in (7.32) with the standard Gauss method, that reads:
ð qc
a aX N
K ðt Þ eð1 cosqÞat ð1 cos qÞdq Wk eð1 cosqk Þat ð1 cos qk Þ (7.33)
p 0 p k¼1
252 Lightning electromagnetics: Volume 2
where Wk is the kth weight of Gauss integral according to the kth Gauss integral
point qk .
If one poses
pk ¼ ð1 cos qk Þa
a (7.34)
Ck ¼ Wk ð1 cos qk Þ
p
then
X
N
K ðt Þ ’ Ck epk t (7.35)
k¼1
which, from a formal point of view, has the same meaning as (7.17); indeed the kernel
has been developed in a sum of exponential functions in the time domain. The advan-
tage of the proposed method is that, in this case, such development does not come from
any VF technique, thus avoiding the problem of the frequency spectrum truncation.
Moreover, it is apparent from (7.34) that the poles and residues of such development are
independent on the ground conductivity and dielectric constant, which makes the
method much more appealing as such quantities have to be computed just once.
Moreover, as observed in [12], it is obvious that the resulting residues and poles are all
real numbers, while some complex residues and poles may be obtained when VF
method is used. At this point the same recursive convolution technique can be applied.
The main drawback of the proposed approach is some deviations appear in the
very early time period because of the approximation in the domain of integration
(upper integration limit equal to qc ) and the errors committed by the Gauss
numerical method to perform the integration itself. The solution proposed by the
authors is to perform a linear convolution using the exact expression of the kernel
for the first time instants and then adopt the abovementioned technique for the
remaining part of the time domain of interest.
In 2018, the same authors proposed an improvement of the procedure which is
mainly based on the definition of an upper integration limit depending on the time,
which allows to perform a recursive convolution over the entire time domain of
interest [13].
More in details, if we reconsider the integrand of (7.32), we get that:
where p is used when (7.37) does not have solution. Of course, as the procedure is
carried out in the discrete time domain, in each time interval a different value of the
angle qc is used.
The properties of the integral kernel have been mainly investigated by the
research group of the University of Naples (A. Andreotti et al.) since 2015 and the
obtained results have been published in [14–17].
In particular, in [14], a new formulation for the CR expression in time domain
is proposed, in which the kernel of the convolution integral exhibits better prop-
erties than (7.5) in terms of smoothness. As reported in [14], the kernel K is char-
s
acterized by a class of functions which describe a Dirac pulse. Indeed, when a ¼ 2e
increases, the function K becomes more and more peaked around t = 0 taking very
large values and, for increasing values of t, tends to zero more and more rapidly.
It is then apparent that any numerical integration involving such a function
becomes a hard task. However, if one observes that:
d at
K ðtÞ ¼ aeat ½I0 ðatÞ I1 ðatÞ ¼ ½e I0 ðatÞ (7.39)
dt
it follows that
ðt ðt
d
hji ðt; 0; rÞK ðt tÞdt ¼ hji ðt; 0; rÞ ½I0 ðaðt tÞÞeaðttÞ dt
0 0 dt
¼ hji ðt; 0; rÞI0 ðaðt tÞÞeaðttÞ jt0
ðt
@hji ð0; r; tÞ h i
I0 ðaðt tÞÞeaðttÞ dt
0 @t
ðt
@hji ð0; r; tÞ h i
¼ hji ðt; 0; rÞ I0 ðaðt tÞÞeaðttÞ dt
0 @t
(7.40)
The new kernel has been often indicated with Ke and has the following expression:
Given
1
f ðxÞ h1 ðxÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; (7.44)
1 þ 2px
or that
" #
ðp 1Þx
f ðxÞ h2 ðxÞ ¼ h1 ðxÞ 1 þ ðp1Þ 2
; (7.45)
1 þ 8ppþ2 x
ps
in which a ¼ :
4e0
Finally, one should note that in 2019 Liu et al. [11] further refined their
approach based on the integral form of the Bessel function and its evaluation in a
reduced integration domain (0, qc ) working on the kernel Ke.
Table 7.2 Parameters of the two Heidler’s functions used to reproduce the
channel-base current waveshape
I01 (kA) t11 (ms) t12 (ms) n1 I02 (kA) t21 (ms) t22 (ms) n2
First stroke 28 1.8 95 2
Subsequent stroke 10.7 0.25 2.5 2 6.5 2 230 2
256 Lightning electromagnetics: Volume 2
300
200
100
0
Er [V/m]
–100
–200
Ideal
–300 With the CR Formula
With Sommerfeld Integral
–400
0 0.2 0.4 0.6 0.8 1 1.2
Time [s]
× 10–5
Figure 7.3 Radial electric field calculated with the CR formula, the Sommerfeld’s
integral and under the assumption of perfectly conducting ground
(r = 250 m, z = 10 m and s = 0.0001 S/m). Channel-base current
typical of subsequent strokes. Adapted from [1]
Lightning electromagnetic field propagation 257
50
40
30
20
Er [V/m]
10
–10
Ideal
–20 With the CR Formula
With Sommerfeld Integral
–30
0 0.2 0.4 0.6 0.8 1 1.2
Time [s]
× 10–5
Figure 7.4 Radial electric field calculated with the CR formula, the Sommerfeld’s
integral and under the assumption of perfectly conducting ground
(r = 500 m, z = 10 m and s = 0.01 S/m). Channel-base current typical
of subsequent strokes. Adapted from [1]
× 104
4
3.5
2.5
Er [V/m]
1.5 Ideal
With the CR Formula
1 With Sommerfeld Integral
0.5
–0.5
0 0.2 0.4 0.6 0.8 1 1.2
Time [s]
× 10–5
Figure 7.5 Radial electric field calculated with the CR formula, the Sommerfeld’s
integral and under the assumption of perfectly conducting ground
(r = 20 m, z = 10 m and s = 0.001 S/m). Channel-base current typical
of subsequent strokes. Adapted from [1]
258 Lightning electromagnetics: Volume 2
200
Ideal
With the CR Formula
150 With Sommerfeld Integral
100
Er [V/m]
50
–50
–100
0 0.2 0.4 0.6 0.8 1 1.2
Time [s]
× 10–5
Figure 7.6 Radial electric field calculated with the CR formula, the Sommerfeld’s
integral and under the assumption of perfectly conducting ground
(r = 500 m, z = 10 m and s = 0.001 S/m). Channel-base current
typical of first strokes. Adapted from [1]
50
0
Er [V/m]
–50
Ideal
With the CR Formula
With Sommerfeld Integral
–100
0 0.2 0.4 0.6 0.8 1 1.2
Time [s]
× 10–5
Figure 7.7 Radial electric field calculated with the CR formula, the Sommerfeld’s
integral and under the assumption of perfectly conducting ground
(r = 500 m, z = 10 m and s = 0.001 S/m). Channel-base current
typical of subsequent strokes. Adapted from [1]
Lightning electromagnetic field propagation 259
field, it is shown by [26] that, for an observation point on the surface of the
ground with a conductivity of 0.001 S/m, the vertical electric field is not sig-
nificantly affected by the ground finite conductivity at distance ranges from 50 m
to 1 km. However, the peak value of the time derivative of the vertical electric
field can be attenuated by about 70% in propagating 1,000 m along the ground.
Cooray also showed that, for ground conductivities not less than 0.001 S/m and
observations points not farther than 1 km, the approximations proposed by [28]
predict adequately the vertical electric field components. On the other hand, for
observation points at about 10 m above the ground surface and as close as 100 m
to the channel-base, the authors of [1] have shown that the PEC ground is a
reasonable assumption for obtaining the vertical electric field and the azimuthal
magnetic field above a homogeneous lossy ground. However, for larger distances
and for ground conductivities smaller than 0.001 S/m, some disagreements can be
found between fields above a perfect ground and those above a homogenous lossy
ground.
having indicated, as before, with low-case letters time domain functions. The
function y is defined as:
qffiffiffiffiffiffiffiffiffiffiffiffiffi
eat=2 atz
yðtÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi I0 a t2 tz2 1ðt tz Þ þ eatz =2 dðt tz Þ (7.55)
2 t2 tz2
pffiffiffiffiffiffiffi
being a ¼ se ; tz ¼ z m0 e, d the Dirac function, 1 the Heaviside function and I0
the modified Bessel function of first type and order zero. To compute the con-
volution integral (7.55) one must know the horizontal field at ground level, which
can be calculated using the CR formula.
50
0
–50
–100 Sommerfeld approach
Er [V/m]
Cooray formula
–150
–200
–250
–300
–350
0 0.2 0.4 0.6 0.8 1 1.2
t [s] × 10–5
Figure 7.8 Underground horizontal electric field generated by the first stroke
current (r = 50 m, z = 10 m and s = 0.01 S/m). Adapted from [35]
50
–50
Er [V/m]
–100
Sommerfeld approach
–150 Cooray formula
–200
0 0.2 0.4 0.6 0.8 1 1.2
t [s] × 10–5
500
0
Sommerfeld approach
Cooray formula
Er [V/m]
–500
–1,000
–1,500
–2,000
0 0.2 0.4 0.6 0.8 1 1.2
t [s] × 10–5
Figure 7.10 Underground horizontal electric field generated by the first stroke
current (r = 50 m, z = 10 m and s = 0.001 S/m). Adapted from [35]
262 Lightning electromagnetics: Volume 2
200
–200
Er [V/m]
–400
–600
–1,200
0 0.2 0.4 0.6 0.8 1 1.2
t [s] × 10–5
solutions can be adopted: (a) the magnetic field computed with the assumption of
PEC ground and (b) the magnetic field including the propagation effects due to
lossy ground.
In order to better clarify this point, a wider comparison has been provided by
[25] through the FDTD technique. The authors studied the efficacy of the Cooray
formula at the same depth as in previous case but for different horizontal distances
from the lightning channel (from 100 m to 2 km) and for different soil con-
ductivities (from 0.0001 to 0.01 S/m), leading to the following results and key-
points:
● The assumption of perfectly conducting ground for the evaluation of the
magnetic field at ground level might lead to inaccuracies in the Cooray for-
mula. This aspect is more relevant when the soil conductivity is low
(s = 0.0001 S/m) and the observation point is distant from the stroke location.
● The use of Cooray formula taking into account propagation effects in the
computation of the magnetic field results in acceptable predictions of the
underground horizontal electric fields.
● The assumption of perfectly conducting ground for the evaluation of the
magnetic field at ground level is a valid assumption if the soil conductivity is
greater than 0.001 S/m and the observation point is not more than 2 km far
from the channel.
and theoretical bottlenecks e.g., resulting from the presence of surface waves on
“wrong” Riemann sheets is given in the review papers by Wait [36,37].
More recently studies on stratified ground have been carried out. In 2011 the
authors of [38] proposed a generalization of the C–R and Wait formulae, for cal-
culating the horizontal electric field at the ground surface showing that the C–R
formula overestimates the field value till 10% for low relative permittivity.
Irregularities in the soil have been considered in [39], where a good accordance
between model and measurements has been shown. Here FDTD procedure has been
adopted for the numerical evaluation of the Sommerfeld integrals.
In 2018 [40], an analysis of stratified frequency dependent soils has been
presented, showing that the lightning electric fields are more affected by the stra-
tification than by the frequency. In 2019, the authors of [41] showed that a two-
layered soil model is more reliable than a homogeneous one comparing their results
with measurements of the electric field in the range of distances from the lightning
strike point 69–126 km. Comparison with experimental data is carried out in [42],
confirming that the vertical electric field and the azimuthal magnetic field at close
range can be evaluated assuming the ground as a perfectly conducting plane, while
the soil stratification and frequency-dependent parameters effect becomes sig-
nificant at intermediate and distant ranges; on the contrary, the horizontal electric
field is found to be very sensitive to the ground stratification at all distances.
The formulation for obtaining lightning radiated electromagnetic fields above
a two-layer horizontally stratified ground is presented in what follows.
and
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
gi ðwÞ ¼ jwm0 ðsi þ jwe0 eri Þ (7.62)
represents the wave number in the ith layer, while g0 is the free space wave number
pffiffiffiffiffiffiffiffiffi
g0 ðwÞ ¼ jw m0 e0 (7.63)
The expression of the attenuation function F needed in (7.56)–(7.58) was
derived by Wait [48] and it involves the complementary error function (erfc ) of
complex value [49]:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Fðw; rÞ ¼ 1 j ppðw; rÞepðw;rÞ erfc j pðw; rÞ (7.64)
being
Observation
Point at
P(ρ,ϕ,z)
Air ε0, μ0
And D being defined in terms of the normalized surface impedance of the two-
layer ground
ZðwÞ
DðwÞ ¼ pffiffiffiffiffiffiffiffiffiffiffi (7.66)
m0 =e0
All the encountered square roots of complex quantities in the above expres-
sions have to be evaluated according to the roots on the upper Riemann sheet [50].
× 104 × 104
2 2
Vertical electric field (V/m)
–2 –2
–4 –4
–6 –6
–8 –8
0 10 20 30 40 50 0 10 20 30 40 50
(a) Time (μs) (d) Time (μs)
800 1,000
Horizontal electric field ( V/m)
200 0
–400 –1,000
0 10 20 30 40 50 0 10 20 30 40 50
(b) Time (μs) (e) Time (μs)
Azimuthal magnetic field (A/m)
40 40
Stratified ground (dedicated solution) Stratified ground (dedicated solution)
Stratified ground (Wait formula) Stratified ground (Wait formula)
30 Perfect ground
30 Perfect ground
20 20
10 10
0 0
–10 –10
0 10 20 30 40 50 0 10 20 30 40 50
(c) Time (μs) Time (μs)
(f )
Figure 7.13 Vertical electric (a,d), horizontal electric (b,e), and azimuthal
magnetic (c,f) field components for an observation point at the height
of z = 0.5 m above a two-layer horizontally stratified ground (see
Figure 7.3) at a distance of r = 50 m from the channel base. Height
of upper ground layer is assumed to be h1 = 2 m for (a,b,c) and
h1 = 5 m for (d,e,f) (adapted from Shoory et al. [52])
positive excursion at the later times. Detailed discussion on this effect can be
found in [52]. The authors of [42] showed that the soil stratification plays an
important role in the horizontal electric field only for observations points located
close to the ground (z = 0.5 m), while for higher vertical coordinates the ground
can be assumed homogeneous.
These conclusions have been confirmed by [53], where the electromagnetic
fields above the ground have been computed with the FDTD technique considering
different cases in terms of ground conductivity and permittivity as well as different
depth of the upper layer (from 5 to 20 m). The EM fields were computed at a height
of 10 m above the ground surface and 50 m far.
Lightning electromagnetic field propagation 267
500 500
Stratified ground (dedicated solution)
Vertical electric field (V/m)
–500 –500
–1,000 –1,000
–1,500 –1,500
0 10 20 30 40 50 0 10 20 30 40 50
(a) Time (μs) (d) Time (μs)
5
Horizontal electric field ( V/m)
0 0
–5 –10
Stratified ground (dedicated solution) Stratified ground (dedicated solution)
–10 Stratified ground (Shoory formulation) –20 Stratified ground (Shoory formulation)
Perfect ground Perfect ground
–15 –30
0 10 20 30 40 50 0 10 20 30 40 50
(b) Time (μs) (e) Time (μs)
1.5 1.5
Azimuthal magnetic field (A/m)
0.5 0.5
0 0
–0.5 –0.5
0 10 20 30 40 50 0 10 20 30 40 50
(c) Time (μs) (f ) Time (μs)
Figure 7.14 Vertical electric (a,d), horizontal electric (b,e), and azimuthal
magnetic (c,f) field components for an observation point at the height
of z = 0.5 m above a two-layer horizontally stratified ground (see
Figure 7.3) at a distance of r = 1,000 m from the channel base.
Height of upper ground layer is assumed to be h1 = 2 m for (a,b,c)
and h1 = 5 m for (d,e,f), adapted from Shoory et al. [52]
30 7
a1) a1)
25 6
20 5
Er (kV/m)
Er (kV/m)
15
2.6% WC (Homogeneous) 3 2.6% WC (Homogeneous)
10 11.6% WC (Homogeneous)
11.6% WC (Homogeneous)
2.6% WC (top) 11.6% WC (lower) 2 2.6% WC (top) 11.6% WC (lower)
11.6% WC (top) 2.6% WC (lower)
5 11.6% WC (top) 2.6% WC (lower)
2.6% WC (Homogeneous frequency-dependent) 1 2.6% WC (Homogeneous frequency-dependent)
11.6% WC (Homogeneous frequency-dependent) 11.6% WC (Homogeneous frequency-dependent)
0 2.6% WC (top) 11.6% WC (lower) frequency-dependent 0 2.6% WC (top) 11.6% WC (lower) frequency-dependent
11.6% WC (top) 2.6% WC (lower) frequency-dependent 11.6% WC (top) 2.6% WC (lower) frequency-dependent
–5 –1
0 5 10 15 0 2 4 6 8 10
Time (μs) Time ( μs)
30 7
a2) a2)
25 6
5
20
4
Er (kV/m)
Er (kV/m)
15
2.6% WC (Homogeneous)
3 2.6% WC (Homogeneous)
10 11.6% WC (Homogeneous) 11.6% WC (Homogeneous)
2.6% WC (top) 11.6% WC (lower) 2 2.6% WC (top) 11.6% WC (lower)
11.6% WC (top) 2.6% WC (lower) 11.6% WC (top) 2.6% WC (lower)
5 1
2.6% WC (Homogeneous frequency-dependent) 2.6% WC (Homogeneous frequency-dependent)
11.6% WC (Homogeneous frequency-dependent) 11.6% WC (Homogeneous frequency-dependent)
0 2.6% WC (top) 11.6% WC (lower) frequency-dependent 0 2.6% WC (top) 11.6% WC (lower) frequency-dependent
11.6% WC (top) 2.6% WC (lower) frequency-dependent 11.6% WC (top) 2.6% WC (lower) frequency-dependent
–5 –1
0 5 10 15 0 2 4 6 8 10
Time (μs) Time (μs)
Figure 7.15 Horizontal electric field (Er) at 50 m. First column: first stroke;
second column: subsequent stroke. The depth of the top layer is 2 m
in the first row and 10 m in the second row. The observation point is
set to 10 m. Adapted from [42]
water content (WC) of 2.6% (11.6%) and the second one being the viceversa of the
first (see previous chapter for the link between WC and the ground parameters). As
can be noticed from the figures, the frequency-dependent soil plays a meaningful
role only in case of observation point located at an height of 0.5 m.
1 0.5
2.6% WC (Homogeneous)
11.6% WC (Homogeneous) b1)
0.5 2.6% WC (top) 11.6% WC (lower)
11.6% WC (top) 2.6% WC (lower) 0
a1)
Er (kV/m)
Er (kV/m)
0
2.6% WC (Homogeneous)
–0.5 11.6% WC (Homogeneous)
–0.5 2.6% WC (top) 11.6% WC (lower)
11.6% WC (top) 2.6% WC (lower)
Er (kV/m)
0
2.6% WC (Homogeneous)
–0.5 11.6% WC (Homogeneous)
–0.5 2.6% WC (top) 11.6% WC (lower)
11.6% WC (top) 2.6% WC (lower)
2.6% WC (Homogeneous frequency-dependent) –1 2.6% WC (Homogeneous frequency-dependent)
–1 11.6% WC (Homogeneous frequency-dependent) 11.6% WC (Homogeneous frequency-dependent)
2.6% WC (top) 11.6% WC (lower) frequency-dependent 2.6% WC (top) 11.6% WC (lower) frequency-dependent
11.6% WC (top) 2.6% WC (lower) frequency-dependent 11.6% WC (top) 2.6% WC (lower) frequency-dependent
–1.5 –1.5
0 2 4 6 8 10 0 2 4 6 8 10
Time (μs) Time (μs)
Figure 7.16 Horizontal electric field (Er) at 50 m. First column: first stroke;
second column: subsequent stroke. The depth of the top layer is 2 m
in the first row and 10 m in the second row. The observation point is
set to 0.5 m. Adapted from [42]
The depth of the upper soil layer for r = 10 km was set to h1 = 2 m. For
r = 100 km, because of the limitations in the computational resources, the smallest
possible value for h1 was 20 m.* In the FDTD simulations, a value for the time step
of 2 ns was considered for r = 10 km and a value of 20 ns was considered for
r = 100 km. The overall time interval was set to Tmax = 50 ms for r = 10 km and
Tmax = 380 ms for r = 100 km. This corresponds to 25,000 time steps for r = 10 km
and 19,000 time steps for r = 100 km. The spatial discretization interval was
chosen to be 2 m for r = 10 km and 20 m for r = 100 km.
According to Figure 7.17, the simulation domain was truncated using the first
order Mur absorbing boundary conditions at rmax = 11 km, zmax = 7.5 km, and
zmin = 100 m for r = 10 km and at rmax = 101 km, zmax = 52.5 km, and zmin = 100 m
for r = 100 km. Making use of the axial symmtery of the problem, this corresponds
to 5,500 3,800 spatial cells for r = 10 km and 5,050 2,630 spatial cells for
r = 100 km.
The simulation results for vertical electric field are shown in Figure 7.18. For
more clarity, the curves are also given in smaller time scales. It can be seen that the
*This is due to the fact that we are simply using square cells and not adaptive meshing scheme. This
would result in a prohibitive number of cells if we decrease the depth of the upper ground layer.
270 Lightning electromagnetics: Volume 2
z
First-order Mur absorbing boundary condition
zmax Air
Lightning
return stroke
channel
Axial symmetry
ρmax
ρ
h1
σ1, ε0εr1, μ0
Figure 7.17 Side view of the simulation domain of the FDTD technique used for
the validation of the simplified approaches. For r = 10 km we used
rmax = 11 km, zmax = 7.5 km, and zmin = 100 m and and for r = 100 km
we used rmax = 101 km, zmax = 52.5 km, and zmin = 100 m (adapted
from Shoory et al. [58])
60 60
h1 = 2 m Perfect ground Perfect ground
Vertical electric field (V/m)
h1 = 2 m
50 Stratified ground (Wait formulation) 50 Stratified ground (Wait formulation)
Stratified ground (FDTD) Stratified ground (FDTD)
40 40
30 30
20 20
10 10
0 0
0 5 10 15 20 0 1 2 3 4 5
(a) Time (μs) (b) Time (μs)
5 Perfect ground 5
Perfect ground
Vertical electric field (V/m)
3 3
2 2
h1 = 20 m
1 1
0 0
0 10 20 30 40 50 0 5 10 15
(c) Time (μs) (d) Time (μs)
Table 7.3 Peak and risetimes of the far vertical electric field predicted using
Wait’s formulation and FDTD Simulations m
1
FDTD
0.5
Er (V/m) 0
–0.5
Shoory formulation
–1
–1.5
0 2 4 6 8 10
(a) Time (μs)
0.2
Shoory formulation
0.1 FDTD
Er (V/m)
–0.1
–0.2
0 5 10 15 20 25 30
(b) Time (μs)
Lightning return stroke I 01 (kA) t11 (ms) t12 (ms) n1 I 02 (kA) t21 (ms) t22 (ms) n2
RS1 16.0 0.501 9.6 5 5.6 10.0 77 3
RS4 18.0 0.359 8.9 5 5.9 9.5 60 3
RS5 15.0 0.417 6.1 6 5.0 8.0 55 3
The need of considering the soil stratification can be clear from Figure 7.20,
where a poor agreement is observed between the measured vertical electric field
and the one computed supposing homogenous ground and different values of the
ground conductivity.
The simulation results assuming a two-layer model for the ground are shown in
Figure 7.21 for all the stations and all the three return strokes. Each column presents the
results for one return stroke, for which electric field waveforms (calculated and mea-
sured) are presented from the shortest to the farthest stations (from the top to the bottom).
Lightning electromagnetic field propagation 273
6
Homogeneous ground σ = 0.001 S/m
5 Homogeneous ground σ = 0.003 S/m
Homogeneous ground σ = 0.01 S/m
4
Perfect ground
Ez (V/m)
3 Measured data at for RS1
–1
0 10 20 30 40 50
Time (μs)
Figure 7.20 Vertical electric field at 85 km: measurement vs. homogeneous soil
model computation. Adapted from [41]
In accordance with the results of [58], the two-layer soil model proposed by
Wait allows obtaining very good agreement between computed and measured
waveforms for all the considered distances and events. In particular, the late-time
response follows to a much better extent the experimental waveforms with respect
to the homogenous case. In the stations located at 87 and 100 km some dis-
crepancies can be observed, probably ascribed to the variation of the soil electrical
parameters along the considered paths, as well as due to their frequency
dependence.
7.2.2.2.1 Frequency-dependence of soil parameters
Considering the frequency-dependence of the ground parameters lead to different
results in terms of variation of the EM fields behavior. According to [42], the
vertical electric field and the azimuthal magnetic field are almost unaffected by
the frequency-dependence of the ground parameters (Figures 7.22 and 7.23). On the
other hand, the frequency dependence in case of soil stratification affects the hor-
izontal electric field in a consistent way causing a decrease of the field peak
(Figure 7.24).
The results were obtained supposing the same two different scenarios for the
two-layer soil as the previous section. The same return stroke model was adopted as
in the previous sub-section, while two cases for the channel-base current (Heidler’s
function) are considered: a first return stroke (30 kA, maximum steepness = 12 kA/
ms) and a subsequent stroke (12 kA, maximum steepness = 12 kA/ms). The
frequency-dependent soil is taken into account with the model proposed in [42] and
two different depths of the top layer are considered: 2 m and 10 m, respectively.
Figures 7.22–7.24 show the vertical electric field, magnetic field and horizontal
electric field computed 10 m above the ground at an horizontal distance of 100 km.
For each scenario the comparison between the homogenous case and the stratified
is provided considering or not the frequency-dependent soil.
274 Lightning electromagnetics: Volume 2
8 8
6
6 Stratified ground 6 Stratified ground Stratified ground
Measured data Measured data Measured data
Ez (V/m)
Ez (V/m)
Ez (V/m)
4
4 LSZ(69km) for RS1 4 LSZ(69km) for RS4 LSZ(69km) for RS5
2 2
2
0 0 0
0 10 20 30 40 50 0 10 20 30 40 50 0 10 20 30 40 50
(a1) Time (μs) (a2) Time (μs) (a3) Time (μs)
8 8 6
Ez (V/m)
Ez (V/m)
4 LPZ(72km) for RS1 4 LPZ(72km) for RS4 LPZ(72km) for RS5
2
2 2
0
0 0
0 10 20 30 40 50 0 10 20 30 40 50 0 10 20 30 40 50
(b1) Time (μs) (b2) Time (μs) (b3) Time (μs)
6 6
6
Stratified ground Stratified ground Stratified ground
4 4
Measured data Measured data Measured data
Ez (V/m)
Ez (V/m)
Ez (V/m)
4
CCZ(85km) for RS1 CCZ(85km) for RS4 CCZ(85km) for RS5
2
2
2
0
0 0
–2
0 10 20 30 40 50 0 10 20 30 40 50 0 10 20 30 40 50
(c1) Time (μs) (c2) Time (μs) (c3) Time (μs)
6
6 4
Stratified ground Stratified ground Stratified ground
4 Measured data Measured data Measured data
Ez (V/m)
Ez (V/m)
Ez (V/m)
4
CCJ(87km) for RS1 CCJ(87km) for RS4 2 CCJ(87km) for RS5
2
2
0 0 0
0 10 20 30 40 50 0 10 20 30 40 50 0 10 20 30 40 50
(d1) Time (μs) (d2) Time (μs) (d3) Time (μs)
6 6
Stratified ground Stratified ground 4 Stratified ground
4 Measured data 4 Measured data Measured data
Ez (V/m)
Ez (V/m)
Ez (V/m)
0 0 0
0 10 20 30 40 50 0 10 20 30 40 50 0 10 20 30 40 50
(e1) Time (μs) (e2) Time (μs) (e3) Time (μs)
4
4 Stratified ground 4 Stratified ground 3 Stratified ground
Measured data Measured data Measured data
Ez (V/m)
Ez (V/m)
Ez (V/m)
0 0 0
–1
0 10 20 30 40 50 0 10 20 30 40 50 0 10 20 30 40 50
(f1) Time (μs) (f 2) Time (μs) (f 3) Time (μs)
4
4 Stratified ground 4 Stratified ground 3 Stratified ground
Measured data Measured data Measured data
Ez (V/m)
Ez (V/m)
Ez (V/m)
2
2 MCZ(126km) for RS1 2 MCZ(126km) for RS4 MCZ(126km) for RS5
1
0 0 0
–1
0 10 20 30 40 50 0 10 20 30 40 50 0 10 20 30 40 50
(g1) Time (μs) (g2) Time (μs) (g3) Time (μs)
Figure 7.21 Vertical electric field for different distances (from 69 to 126 km) and
different return stroke currents (from left to right). Adapted from [41]
5
2.6% WC (Homogeneous frequency-dependent) 2.6% WC (Homogeneous frequency-dependent)
10 11.6% WC (Homogeneous frequency-dependent) 11.6% WC (Homogeneous frequency-dependent)
2.6% WC (top) 11.6% WC (lower) frequency-dependent 2.6% WC (top) 11.6% WC (lower) frequency-dependent
11.6% WC (top) 2.6% WC (lower) frequency-dependent 4 11.6% WC (top) 2.6% WC (lower) frequency-dependent
8
c1) b1)
3
Ez (V/m)
Ez (V/m)
6
2
4
2.6% WC (Homogeneous) 2.6% WC (Homogeneous)
2 11.6% WC (Homogeneous) 1 11.6% WC (Homogeneous)
2.6% WC (top) 11.6% WC (lower) 2.6% WC (top) 11.6% WC (lower)
11.6% WC (top) 2.6% WC (lower) 11.6% WC (top) 2.6% WC (lower)
0 0
0 5 10 15 0 2 4 6 8 10
Time (μs) Time (μs)
5
2.6% WC (Homogeneous frequency-dependent) 2.6% WC (Homogeneous frequency-dependent)
10 11.6% WC (Homogeneous frequency-dependent) 11.6% WC (Homogeneous frequency-dependent)
2.6% WC (top) 11.6% WC (lower) frequency-dependent 2.6% WC (top) 11.6% WC (lower) frequency-dependent
11.6% WC (top) 2.6% WC (lower) frequency-dependent 4 11.6% WC (top) 2.6% WC (lower) frequency-dependent
8
c2) c2)
3
Ez (V/m)
Ez (V/m)
6
2
4
2.6% WC (Homogeneous)
2.6% WC (Homogeneous)
2 11.6% WC (Homogeneous) 1 11.6% WC (Homogeneous)
2.6% WC (top) 11.6% WC (lower)
2.6% WC (top) 11.6% WC (lower)
11.6% WC (top) 2.6% WC (lower)
11.6% WC (top) 2.6% WC (lower)
0 0
0 5 10 15 0 2 4 6 8 10
Time (μs) Time (μs)
Figure 7.22 Vertical electric field (Ez) at 100 km. First column: first stroke;
second column: subsequent stroke. The depth of the top layer is 2 m
in the first row and 10 m in the second row. Adapted from [42]
30 14
2.6% WC (Homogeneous frequency-dependent) 2.6% WC (Homogeneous frequency-dependent)
11.6% WC (Homogeneous frequency-dependent) 11.6% WC (Homogeneous frequency-dependent)
25 2.6% WC (top) 11.6% WC (lower) frequency-dependent
12 2.6% WC (top) 11.6% WC (lower) frequency-dependent
11.6% WC (top) 2.6% WC (lower) frequency-dependent 11.6% WC (top) 2.6% WC (lower) frequency-dependent
10
20 c1)
Hphi (mA /m)
Hphi (mA /m)
c1)
8
15
6
10
2.6% WC (Homogeneous) 4 2.6% WC (Homogeneous)
11.6% WC (Homogeneous) 11.6% WC (Homogeneous)
5 2.6% WC (top) 11.6% WC (lower) 2.6% WC (top) 11.6% WC (lower)
2
11.6% WC (top) 2.6% WC (lower) 11.6% WC (top) 2.6% WC (lower)
0 0
0 5 10 15 0 2 4 6 8 10
Time (μs) Time (μs)
30 14
2.6% WC (Homogeneous frequency-dependent) 2.6% WC (Homogeneous frequency-dependent)
11.6% WC (Homogeneous frequency-dependent) 11.6% WC (Homogeneous frequency-dependent)
25 2.6% WC (top) 11.6% WC (lower) frequency-dependent
12 2.6% WC (top) 11.6% WC (lower) frequency-dependent
11.6% WC (top) 2.6% WC (lower) frequency-dependent 11.6% WC (top) 2.6% WC (lower) frequency-dependent
10
20
c2) c2)
Hphi (mA /m)
8
15
6
10
4
2.6% WC (Homogeneous) 2.6% WC (Homogeneous)
5 11.6% WC (Homogeneous) 11.6% WC (Homogeneous)
2.6% WC (top) 11.6% WC (lower) 2 2.6% WC (top) 11.6% WC (lower)
11.6% WC (top) 2.6% WC (lower) 11.6% WC (top) 2.6% WC (lower)
0 0
0 5 10 15 0 2 4 6 8 10
Time (μs) Time (μs)
Figure 7.23 Azimuthal magnetic field (Hphi) at 100 km. First column: first stroke;
second column: subsequent stroke. The depth of the top layer is 2 m
in the first row and 10 m in the second row. Adapted from [42]
276 Lightning electromagnetics: Volume 2
0.1 2.6% WC (Homogeneous frequency-dependent)
2.6% WC (Homogeneous frequency-dependent)
0.1 11.6% WC (Homogeneous frequency-dependent) 11.6% WC (Homogeneous frequency-dependent)
2.6% WC (top) 11.6% WC (lower) frequency-dependent 2.6% WC (top) 11.6% WC (lower) frequency-dependent
11.6% WC (top) 2.6% WC (lower) frequency-dependent 0.05 11.6% WC (top) 2.6% WC (lower) frequency-dependent
0
c1)
0
Er (V/m)
Er (V/m)
–0.1
c1) –0.05
–0.2
2.6% WC (Homogeneous) 2.6% WC (Homogeneous)
–0.1 11.6% WC (Homogeneous)
11.6% WC (Homogeneous)
–0.3 2.6% WC (top) 11.6% WC (lower) 2.6% WC (top) 11.6% WC (lower)
11.6% WC (top) 2.6% WC (lower) 11.6% WC (top) 2.6% WC (lower)
–0.15
0 5 10 15 0 2 4 6 8 10
Time (μs) Time (μs)
0.1
2.6% WC (Homogeneous frequency-dependent) 2.6% WC (Homogeneous frequency-dependent)
0.1 11.6% WC (Homogeneous frequency-dependent) 11.6% WC (Homogeneous frequency-dependent)
2.6% WC (top) 11.6% WC (lower) frequency-dependent 2.6% WC (top) 11.6% WC (lower) frequency-dependent
11.6% WC (top) 2.6% WC (lower) frequency-dependent 0.05 11.6% WC (top) 2.6% WC (lower) frequency-dependent
0
c2)
0
Er (V/m)
Er (V/m)
–0.1
c2) –0.05
–0.2
2.6% WC (Homogeneous) 2.6% WC (Homogeneous)
11.6% WC (Homogeneous) –0.1 11.6% WC (Homogeneous)
2.6% WC (top) 11.6% WC (lower) 2.6% WC (top) 11.6% WC (lower)
–0.3
11.6% WC (top) 2.6% WC (lower) 11.6% WC (top) 2.6% WC (lower)
–0.15
0 5 10 15 0 2 4 6 8 10
Time (μs) Time (μs)
Figure 7.24 Horizontal electric field (Er) at 100 km. First column: first stroke;
second column: subsequent stroke. The depth of the top layer is 2 m
in the first row and 10 m in the second row. Adapted from [42]
9
Homogeneous (σ = 0.003 S/m)
8
Stratified (0.001 S/m(top) and 4 S/m (lower))
7 Homogeneous frequency-dependent (σ = 0.003 S/m)
Stratified frequency-dependent (0.001 S/m(top) and 4 S/m (lower))
6 Measured data at 85 km
Ez(V/m)
0 5 10 15 20 25
Time (μs)
Figure 7.25 Comparison of the measured vertical electric field with different
models for the soil. Adapted from [42]
Lightning electromagnetic field propagation 277
(7.67)
Observation
point
ρ dl
σ1, ε0εr1, μ0 σ2, ε0εr2, μ0
in which d is a small distance over which one can assume the attenuation functions
to be constant, while Di (i = 1, 2) is the normalized surface impedance of each
ground section defined as ( e.g., by Hill and Wait [66]).
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
jwe0 ½si þ jwe0 ðeri 1Þ
Di ðwÞ ¼ (7.68)
si þ jwe0 eri
and F1 and F2 are the attenuation functions of each section of the ground repre-
sented as a function of the horizontal distance from the source to the observation
point and defined as follows:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Fi ðw; rÞ ¼ 1 j ppi ðw; rÞepi ðw;rÞ erfc j pi ðw; rÞ (7.69)
with
Case 1 Case 2
s1 0.001 S/m 4 S/m
er1 10 30
s2 4 S/m 0.001 S/m
er2 30 10
Lightning electromagnetic field propagation 279
Zmax Air
Lightning
return stroke
channel
Axial symmetry
ρmax
ρ
ρ
dl
Figure 7.27 Side view of the simulation domain of the FDTD technique used for
the validation of the simplified approaches. Adapted from [68]
respectively; curves 3 and 4 correspond to Wait’s formula and FDTD (dl = 50 m),
respectively.
The results confirm the accuracy of the simplified formula at near distances.
This, in principle, could not be obvious as the formula has been originally devel-
oped for the radiation component of the field at far distance. However, since within
tens or hundreds of meters from the lightning channel, the electrostatic and
induction field components become more and more predominant when the distance
of the observation point gets closer and they are less affected by the ground finite
conductivity, it seems that the vertical electric field at near ranges are not sig-
nificantly affected by mixed propagation path, thus Wait’s formula appears to
present a reasonable accuracy at close distances too. Therefore, in such a mixed
path, one can approximately use the vertical field waveforms of the PEC ground to
expedite the computation process. However, it is important to notice that some
minimal differences can be seen in the early-time response within a few micro-
seconds. This aspect will become more relevant in case of intermediate distances
and will be discussed in the following subsection.
6,000
Electric Field (V/m)
4,000
2,4
2,000
1,3
0
0 2 4 6
(a) Time(μs)
6,000
Electric Field (V/m)
4,000
2,4
2,000
1,3
0
0 2 4 6
(b) Time(μs)
Figure 7.28 Vertical electric field for (a) case 1 and (b) case 2 at the surface of a
mixed-path ground at a distance of d = 200 m. Curves 1 and 2
correspond to the Wait’s formula and FDTD (dl = 100 m),
respectively; curves 3 and 4 correspond to the Wait’s formula and
FDTD (dl = 50 m), respectively. The subsequent return stroke is
used. Adapted from [67]
Lightning electromagnetic field propagation 281
600
400
Electric Field (V/m)
3 4
200 12
0
0 1 2 3 4
(a) Time (μs)
600
Electric Field (V/m)
400
1 2
200
3 4
0
0 2 4 6 8
(b) Time (μs)
Figure 7.29 Vertical electric field for (a) case 1 and (b) case 2 at the surface of a
mixed-path ground at a distance of d = 1,000 m. Curves 1 and 2
correspond to the Wait’s formula and FDTD (dl = 500 m),
respectively; curves 3 and 4 correspond to the Wait’s formula and
FDTD (dl = 50 m), respectively. The subsequent return stroke is
used. Adapted from [67]
282 Lightning electromagnetics: Volume 2
As can be seen, the agreement is still excellent; however, some minimal dif-
ferences can be observed for the early time response within a few microseconds
between them. Therefore, Wait’s formula may cause errors for predicting the
lightning electromagnetic field derivates in close distances. This can be ascribed to
the high-frequency components of the subsequent strokes which are more affected
by the finitely conducting ground.
2,000
Electric Field (V/m)
1,000
12
34
0
0 2 4 6 8 10
(a) Time (μs)
2,000
Electric Field (V/m)
1,000
12
34
0
0 2 4 6 8 10
( b) Time (μs)
Figure 7.30 Vertical electric field for (a) case 1 and (b) case 2 at the surface of a
mixed-path ground at a distance of d = 1,000 m. Curves 1 and 2
correspond to the Wait’s formula and FDTD (dl = 500 m),
respectively; curves 3 and 4 correspond to the Wait’s formula and
FDTD (dl = 50 m), respectively. The first return stroke is used.
Adapted from [67]
Lightning electromagnetic field propagation 283
As stated in [67], the Wait’s formula can reproduce accurately the vertical
electric field (or azimuthal magnetic field, here not shown for sake of brevity) at
close distances within 1,000 m because the static and induction components are
dominant at close distances and they are usually less affected by finite ground
conductivity. However, if the channel-base current is characterized by an high-
frequency spectrum, the static and induction terms becomes more affected by the
finite ground conductivity and consequently the Wait’s formula partially loses its
validity.
This aspect is confirmed in Figure 7.30, where the same scenario of
Figure 7.29 is proposed for a typical first stroke, characterized by a low frequency
spectrum with respect to subsequent one.
40 40 FDTD
FDTD
30 30
20 20
Wait formulation Wait formulation
10 10
0 0
0 5 10 15 0 5 10 15
(a) Time (μs) (b) Time (μs)
Vertical electric field (V/m)
40 FDTD 40 FDTD
30 30
20 20
Wait formulation Wait formulation
10 10
0 0
0 5 10 15 0 5 10 15
(c) Time (μs) (d) Time (μs)
20 20
Wait formulation Wait formulation
10 10
0 0
0 5 10 15 0 5 10 15
Time (μs) Time (μs)
(a) (b)
Vertical electric field (V/m)
20 20
Wait formulation Wait formulation
10 10
0 0
0 5 10 15 0 5 10 15
Time (μs) Time (μs)
(c) (d)
Table 7.6 Peak and risetimes of the vertical electric field predicted using Wait’s
formulation and FDTD simulation
Table 7.6 presents the values of the field peaks and zero-to-peak risetimes
predicted by Wait’s formulation and by the FDTD method. It can be seen that
Wait’s formulation can predict field peaks with an error of less than 4.8% and field
rise times with an error of less than 18.0%.
7.4 Summary
In this chapter, we provided a review of simplified formulations for obtaining
propagation effects on lightning radiated electromagnetic fields for the case of
Lightning electromagnetic field propagation 285
(i) homogenous lossy ground, (ii) two-layer horizontally stratified ground and (iii)
two-layer vertically stratified ground. The validation has been performed using as
reference three types of approaches: full-wave approach, numerical evaluation of
Sommerfeld integrals and experimental results of rocket-triggered lightning.
For the case of a homogeneous ground, the Cooray–Rubinstein and Cooray
formula have shown to be, in general, accurate for typical values of ground con-
ductivity in the prediction of the overground (underground) horizontal electric field.
For the case of a two-layer horizontally stratified ground, it was shown that the
results obtained using the simplified approaches are in excellent agreement with
exact results in near, intermediate, and far distance ranges, with some minimal
differences of the late time response of the vertical electric field. It was seen that at
near and intermediate distance ranges and for elevated observation points, vertical
electric and azimuthal magnetic field components appear not to be appreciably
affected by the ground finite conductivity and can be evaluated assuming the
ground as a perfectly conducting ground. On the other hand, the horizontal electric
field above a horizontally stratified ground is very much affected by the ground
electrical parameters. On the other hand, in case of far distances, the vertical
electric field and the azimuthal magnetic field should be computed taking into
account the soil stratification. It is noticed that the Wait’s formula become more
precise as the distance from the lightning channel grows. For all the considered
distances, the possible influence of the frequency-dependent soil parameters has
been proposed, leading to the conclusion that it is less influent than considering the
soil stratification with respect to the PEC case.
For the case of a vertically stratified ground, the accuracy of the Wait for-
mulations was examined taking as reference full-wave simulations obtained using the
FDTD technique. It was shown that Wait’s simplified formulas are able to reproduce
the near, intermediate and distant field peak and waveshape with a good accuracy.
Some minimal differences can be noticed for the early-time response if the spectrum
of the channel-base current is characterized by a high-frequency content.
References
[1] F. Delfino, R. Procopio, M. Rossi, F. Rachidi, and C.A. Nucci, “Lightning
return stroke current radiation in presence of a conducting ground: 2.
Validity assessment of simplified approaches,” J. Geophys. Res. Atmos.,
vol. 113, no. D5, 2008, doi: 10.1029/2007JD008567.
[2] F. Delfino, R. Procopio, and M. Rossi, “Lightning return stroke current
radiation in presence of a conducting ground: 1. Theory and numerical
evaluation of the electromagnetic fields,” J. Geophys. Res., vol. 113, 2008,
doi: 10.1029/2007JD008553.
[3] M. Brignone, F. Delfino, R. Procopio, M. Rossi, and F. Rachidi, “Evaluation
of power system lightning performance—Part II: application to an overhead
distribution network,” IEEE Trans. Electromagn. Compat., vol. PP, pp. 1–8,
2016, doi: 10.1109/TEMC.2016.2601657.
286 Lightning electromagnetics: Volume 2
8.1 Introduction
The problem of lightning protection of overhead and buried power lines has been
reconsidered in recent years due to the proliferation of sensitive loads and the
increasing demand by customers for good quality in the power supply [1].
Overvoltages originated by lightning are a major cause of flashovers and
disturbances. Additionally, lightning-originated surges can also damage, depending
on their amplitude and energy content, the power components connected to these
networks as well as the relevant electronic devices.
The evaluation of lightning-induced voltages requires the knowledge of the
electromagnetic field change along the considered line. This electromagnetic field
is generally determined assuming that the lightning return stroke channel is a
straight vertical antenna above a conducting plane. Some studies have attempted to
take into account the channel tortuosity and inclination in the computation of
electromagnetic fields (e.g., [2–6]). The spatial and temporal distribution of the
current along the channel is specified using a return stroke model.
The problem of return stroke modeling and electromagnetic field computation
is beyond the scope of this chapter. It is however worth mentioning that:
● Four classes of lightning return stroke models have been defined by Rakov and
Uman [7]: (1) the gas dynamic models, (2) electromagnetic models, (3)
distributed-circuit models, and (4) engineering models. Outputs of the elec-
tromagnetic, distributed-circuit, and engineering models can be directly used
1
Department of Electrical, Electronic and Information Engineering, School of Engineering, University of
Bologna, Italy
2
Ecole Polytechnique Fédérale de Lausanne (EPFL), Switzerland
3
School of Engineering and Management Vaud, HES-SO University of Applied Sciences and Arts
Western Switzerland
292 Lightning electromagnetics: Volume 2
*Different methods based on this approach generally assume that the wire’s cross-section is smaller than
the minimum significant wavelength (thin-wire approximation).
Interaction of lightning-generated electromagnetic fields 293
→
z Ee
→
Be
y
2a
h
ZA ZB
C
x
0 x x + dx L
Figure 8.1 Geometry of the problem. C is the integration path used in the
derivation of the first Telegrapher’s equation
Interaction of lightning-generated electromagnetic fields 295
For a line of finite length, such as the one represented in Figure 8.1, the boundary
conditions for the load currents and voltages must be enforced. They are simply given by
V ð0Þ ¼ ZA Ið0Þ (8.3)
V ðLÞ ¼ ZB IðLÞ (8.4)
Note that, unlike the classical telegrapher’s equations in which no external
excitation is considered, the presence of an external field results in forcing func-
tions expressed in terms of the exciting magnetic flux and electric field in both
Telegrapher’s equations.
Equations (8.1) and (8.2) are commonly referred to as the Taylor et al. model.
They can be represented using an equivalent circuit, as shown in Figure 8.2. The
forcing functions (source terms) in (8.1) and (8.2) are included as a set of dis-
tributed series voltage and parallel current sources along the line.
8.3.3 Agrawal, Price, and Gurbaxani model
An equivalent formulation of the field-to-transmission line coupling equations was
proposed in 1980 by Agrawal, Price, and Gurbaxani [15]. This model is commonly
referred to as the Agrawal et al. model.
The basis for the derivation of the Agrawal et al. model can be described as
follows: The excitation fields produce a line response that is TEM. This response is
expressed in terms of a scattered voltage Vs(x), which is defined in terms of the line
integral of the scattered electric field from the ground to the line. The total voltage
can be obtained from the scattered voltage through
ðh
V ðxÞ ¼ V s ðxÞ þ V e ðxÞ ¼ V s ðxÞ Eze ðx; zÞdz (8.5)
0
h C'dx
ZA V(0) V(x) –jɷC′ ʃ E ez (x, 0, z)dzdx V(x + dx) V(L) ZB
0
0 x x + dx L
Note that, in this model, only one source term is present (in the first equation)
and it is simply expressed in terms of the exciting electric field tangential to the line
conductor Exe ðx; hÞ.
The boundary conditions in terms of the scattered voltage and the total current
as used in (8.6) and (8.7), are given by
ðh
V s ð0Þ ¼ ZA Ið0Þ þ Eze ð0; zÞdz (8.8)
0
ðh
V s ðLÞ ¼ ZB IðLÞ þ Eze ðL; zÞdz (8.9)
0
e
E x(x, 0, h)dx
L'dx
I(x) + I(x + dx)
+ +
h h
e
ʃ E z (0, 0, z)dz ʃ E ez (L, 0, z)dz
0 0
s(x
V s(0) V s(x) C 'dx V + dx) V s(L)
ZA ZB
0 x x + dx L
in which I s ðxÞ is the so-called scattered current related to the total current by
L'dx
s
I (x) I s(x + dx)
h
h 1 Be(L, 0, z)dz
1 Be(0, 0, z)dz
L' 0 y
L' 0 y e C'dx
h
1 дBx
ZA V(0) V(x) L' дy (x, 0, z)dzdx V(x + dx) V(L) ZB
0
0 x x + dx L
which the power line is located is perfectly conducting. The model proposed by
Cooray et al. [19] accounts for the presence of a finitely conducting ground. It is
worth noting that an extension of the Rusck model to take into account the ground
losses was recently proposed by Piantini [20]. The extension of Piantini uses the
Agrawal et al. model representation of the coupling equations in terms of the
scattered voltage.
dV s ðxÞ
þ Z 0 IðxÞ ¼ Exe ðx; 0; hÞ (8.16)
dx
dIðxÞ
þ Y 0 V s ðxÞ ¼ 0 (8.17)
dx
where Z’ and Y’ are the longitudinal and transverse per-unit-length impedance and
admittance respectively, given by [9,21]
Z 0 ¼ jwL0 þ Z 0w þ Z 0g (8.18)
ðG0 þ jwC 0 ÞY 0g
Y0 ¼ (8.19)
G0 þ jwC 0 þ Y 0g
in which
- L0 , C0 and G0 are the per-unit-length longitudinal inductance, transverse
capacitance, and transverse conductance, respectively, calculated for a lossless wire
above a perfectly conducting ground:
0 mo 1 h mo 2h
L ¼ cosh ffi ln for h >> a (8.20)
2p a 2p a
2peo 2peo
C0 ¼ 1
ffi for h >> a (8.21)
cosh ðh=aÞ lnð2h=aÞ
sair 0
G0 ¼ C (8.22)
eo
Interaction of lightning-generated electromagnetic fields 299
It has been shown [21] that the above logarithmic expression represents an
excellent approximation to the general expression (8.25) over the frequency range
of interest.
Finally, Y 0g is the so-called ground admittance, given by [9]
g2g
Y 0g ffi (8.27)
Z 0g
For typical overhead power lines, the effects of the ground admittance and the
wire impedance are negligible compared to the effect of the ground impedance and
the line inductance, and can be disregarded in the computation [1,23]†.
†
Note that for buried cables, the effect of the ground admittance is no longer negligible [78].
300 Lightning electromagnetics: Volume 2
→
Ee
→e
B
z j
2aj
i
2ai
hj
hi
rij
y
yi σg , εrg yj
The calculation in the time domain of the transient ground resistance matrix of
an overhead transmission line is addressed in [27, 28] by solving analytically the
inverse transform of the ground impedance expression in the frequency domain.
Note that in (8.28) and (8.29), the terms corresponding to the wire impedance
and the so-called ground admittance have been neglected. This approximation is
valid for typical overhead power lines [21].
The boundary conditions for the two line terminations are given by
ð hi
s
Vi ð0Þ ¼ ½ZA ½Ii ð0Þ þ Eze ð0; yi ; zÞdz (8.36)
0
ð hi
Vis ðLÞ ¼ ½ZB ½Ii ðLÞ þ Eze ðL; yi ; zÞdz (8.37)
0
in which [ZA] and [ZB] are the impedance matrices at the two line terminations.
The line coupling equations can be solved using Green’s functions to obtain
closed-form solutions in the frequency domain, as we will see in Section 8.3.8, or in
the time domain using the FDTD technique.
302 Lightning electromagnetics: Volume 2
degL
GV ðx; xs Þ ¼ 2gL
egðx> LÞ þ dr2 egðx> LÞ ðeg x< dr1 eg x< Þ (8.39)
2ð1 r1 r2 e Þ
where
● x< represents the smaller of x or xs, and x> represents the larger of x or xs,
● d = 1pfor
ffiffiffiffiffiffiffiffiffix > xs and d = 1 for x < xs,
● g¼ p Z 0ffiffiffiffiffiffiffiffiffiffiffi
Y 0 is the complex propagation constant along the transmission line,
● Zc ¼ Z 0 =Y 0 is the line’s characteristic impedance, and
● r1 and r2 are the voltage reflection coefficients at the loads of the transmission
line given by
ZA Zc ZBZc
r1 ¼ r2 ¼ (8.40)
ZA þZc ZB þZc
The solutions for the total line current I(x) and scattered voltage V s ðxÞ can be
written as the following integrals of the Green’s functions [9]
ðL ðh ðh
IðxÞ ¼ GI ðx; xs ÞVs0 dxs þ GI ðx; 0Þ Eze ð0; 0; zÞdz GI ðx; LÞ Eze ðL; yi ; zÞdz (8.41)
0 0 0
ðL ðh ðh
V s ðxÞ ¼ GV ðx; xs ÞVs0 dxs þ GV ðx; 0Þ Eze ð0; 0; zÞdz GV ðx; LÞ Eze ðL; 0; zÞdz (8.42)
0 0 0
Interaction of lightning-generated electromagnetic fields 303
Note that the second and the third terms on the right-hand side of (8.41) and (8.42)
are due to the contribution of equivalent lumped sources at the line ends (see Figure 8.3).
The total voltage can be determined from the scattered voltage by adding the
contribution from the exciting field as
ðh
V ðxÞ ¼ V s ðxÞ Eze ðx; 0; zÞdz (8.43)
0
If we are interested in the transmission line response at its terminal loads, the
solutions can be expressed in a compact way by using the so-called BLT (Baum,
Liu, Tesche) equations [9]
1
Ið0Þ 1 r1 0 r1 eg L S1
¼ 1=Zc gL
(8.44)
IðLÞ 0 1 r2 e r2 S2
1
V ð0Þ 1 þ r1 0 r1 eg L S1
¼ (8.45)
V ðLÞ 0 1 þ r2 eg L r2 S2
where the source vector is given by
0 ðL ðh ðh 1
g xs 1 eg L
B
1
2 e Exe ðxs ; 0; hÞdxs þ Eze ð0; 0; zÞdz C Eze ðL; 0; zÞdz
S1 B 0 0 2 0 2 C
¼B ð C
S2 @ L g L ðh ðh A
gðLxs Þ e e 1
1
2 e Ex ðxs ; 0; hÞdxs Ez ð0; 0; zÞdz þ
e
Ez ðL; 0; zÞdz
e
0 2 0 2 0
(8.46)
Note that in the BLT equations, the solutions are directly given for the total
voltage and not for the scattered voltage.
For an arbitrary excitation field, the integrals in (8.46) cannot be carried out
analytically. However, for the special case of a plane wave excitation field, the
integrations can be performed analytically, and closed-form expressions can be
obtained for the load responses. General solutions for vertical and horizontal field
polarizations are given in [9].
The inverse Fourier transforms of the boundary conditions written, for sim-
plicity, for resistive terminal loads read
ð hi
vsi ðx; tÞ ¼ ½RA ½ii ð0; tÞ þ Eze ðx ¼ 0; y ¼ yi ; z; tÞ (8.50)
0
ð hi
vsi ðL; tÞ ¼ ½RB ½ii ðL; tÞ þ Eze ðx ¼ L; y ¼ yi ; z; tÞ (8.51)
0
where ½RA and ½RB are the matrices of the resistive loads at the two line terminals.
The general expression for the ground impedance matrix terms in the
frequency domain does not have an analytical inverse Fourier transform. Thus, the
elements of the transient ground resistance matrix in the time domain have to be, in
general, determined using a numerical inverse Fourier transform algorithm.
However, analytical expressions have been derived by Rachidi et al. [39] and
Araneo and Celozzi [40] which have been shown to be reasonable approximations
to the numerical values obtained using an inverse FFT. More discussion on the
validity of the approximate analytical expressions can be found in [41].
One of the most popular approaches to solve the coupling equations in the
time domain is the FDTD technique (e.g., [42]). Such a technique was already
used by Agrawal et al. in [15], where partial time and space derivatives were
approximated using a first-order FDTD scheme. In [43], instead, the use of a
second-order FDTD scheme based on the Lax–Wendroff algorithm [44],[45] was
proposed. The second-order FDTD scheme shows much better stability compared
to its first-order counterpart, especially when analyzing complex systems invol-
ving nonlinearities [43, 46].
The second-order discretized solutions for the line current and scattered vol-
tage are given by Paolone et al. [13]
0 1 Dt2 0 0 1
½ii nkþ1 ½ii nk1
½vi knþ1 ¼ ½vi nk Dt Cij þ
Lij Cij
2Dx 2
½Exi nkþ1 ½Exi nk1 ½vi nkþ1 þ ½vi nk1 2½vi nk
2Dx Dx2
0h i n h in 1
0
Dt 2 v v0gi
1 B gi
k1 C
þ Lij 0 Cij 0 @ kþ1
A
2 2Dx
(8.52)
Interaction of lightning-generated electromagnetic fields 305
h i n
1 ½vi nkþ1 ½vi nk1
½ii knþ1 ¼ ½ii nk Dt Lij 0 ½Exi nk þ v0gi
2Dx k
Dt2 0 0 1 ½ii kþ1 þ ½ii k1 2½ii k
n n n
þ Cij Lij
2 Dx2
!
Dt2 0 0 1 0 ½Exi knþ1 ½Exi kn1 (8.53)
þ Cij Lij Cij
2 2Dt
0 h i n h i n1 1
Dt 2 1 B v0gi v0gi
k C
Cij 0 Lij 0 @ Cij 0 k
A
2 Dt
where
● Dx is the spatial integration step;
● Dt is the time integration step;
● k = 0,1,2, . . . , kmax is the spatial discretization index (kmax = (L/Dx)+1, where L
is the line length);
● n = 0,1,2, . . . , nmax is the time discretization index;
● ½vi knþ1 is the vector of the scattered voltages corresponding to the spatial and
time discretization indexes k and n+1, respectively;
● ½ii knþ1 is the vector of the conductors’ currents corresponding to the spatial and
time discretization indexes k and n+1, respectively;
● ½Exi knþ1 is the vector of the exciting horizontal electric field along the wires
corresponding to the spatial and time discretization indexes k and n+1,
respectively;
h in P n h i nh k k
½ii n ½ii n1
● v0gi ¼ x0gij Dt .
k h¼0 k
7 km 7 km
2 km 2 km 2 km
U0 U1 U2 U3
R R
20 km
1,000
800
Voltage (kV)
600
400
U0
2 km
200
4 km
6 km
0
0 5 10 15 20
Time (μs)
Figure 8.7 Effect of ground losses on traveling voltages along the line, due to a
direct lightning strike (adapted from [81]). The curves at 2 km, 4 km,
and 6 km correspond, respectively, to the observation points U1, U2,
and U3 in Figure 8.6
450
U0
400 U1
350 U2 U3
300
Voltage [kV]
250
200
150
100
50
0
0 5 10 15 20 25 30 35 40
Time [us]
Figure 8.8 Influence of corona on traveling voltages along the line, due to a
direct lightning strike (adapted from [63]). The curves labeled U0, U1,
U2, and U3 correspond, respectively, to the curves labeled U0, 2 km,
4 km, and 6 km in Figure 8.7 and to the observation points labeled U0,
U1, U2, and U3 in Figure 8.6
It can be seen that the traveling waves exhibit the typical distortion and
attenuation associated with corona effect. A comparison between the results of
Figures 8.7 and 8.8 shows, additionally, that the overvoltages are more significantly
affected by the corona effect, rather than by the ground losses.
308 Lightning electromagnetics: Volume 2
In the simulations presented in Figures 8.7 and 8.8, we assumed that the
magnitudes of the overvoltages are below the lightning impulse withstand voltage
of the line. In most of the cases, however, the surge propagating from the point of
strike along the line is altered by flashovers occurring between the strike location
and the point of interest. Practically all flashovers to ground occur at the poles, as
on overhead distribution lines the weakest insulation is generally at a pole structure
rather than between conductors through air [65]. Figure 8.9 shows a typical over-
voltage (evaluated by calculations) due to a direct lightning strike of 30 kA current
amplitude. The calculations have been performed using the Electromagnetic
Transient Program (EMTP) [66] for a single-wire line with no ground wires. The
line is composed of eight spans (9 poles) of 200 m length each with a characteristic
impedance of 440 W. Each pole, 8-m high, is modeled as a transmission line with a
characteristic impedance of 300 W. The footing DC resistance was assumed to be
nonlinear (current-dependent), with 30 W at zero current. The insulator flashover
voltage was fixed at 150 kV. The voltage was calculated 600 m from the strike
location. The example shows the general characteristics of a direct lightning
overvoltage which presents a few very short spikes, followed by an impulse voltage
with a smoother shape [65].
Effect of ground losses on induced overvoltages
We will now consider a single-wire overhead line matched at both ends, and
illuminated by the electromagnetic field radiated by a nearby lightning return
stroke. The computations are carried out by means of the LIOV code (see [1] for a
detailed description of models used in the code).
The wire is at a height of 10 m above a ground plane characterized by a ground
conductivity sg = 0.001 S/m and a relative permittivity er = 10. The return stroke
current has a peak value of 12 kA and a maximum time derivative of 40 kA/ms
(typical of subsequent return strokes).
160
[kV]
120
80
40
–40
–80
0 10 20 30 40 [us] 50
Figure 8.10 presents the voltage induced along a 2.8-km line for a return stroke
located in the vicinity of the left-end terminal, at both terminations and at two
intermediate distances. This configuration has been chosen because it is similar to
an event recorded by De La Rosa et al. on an experimental line in Mexico [67]. The
voltages reported in Figure 8.10 are calculated (i) considering a finitely conducting
ground as described in the previous paragraph, and (ii) considering the ground as
perfectly conducting.
The computed results show that the voltages calculated taking into account
ground losses exhibit significantly larger amplitudes than those calculated assum-
ing a perfectly conducting ground. Also, the voltages exhibit an inversion of
polarity as the observation point moves towards the far end of the line.
The reason why, for the examined case, ground losses can result in an amplitude
enhancement has been discussed thoroughly in a few papers [1,59,68]. In summary, it
can be said that the finite conductivity of the ground does not significantly affect the
vertical electric field amplitude and waveshape, but acts in modifying the horizontal
electric field waveshape and, in particular, it results in this this field component’s
polarity reversal [21,69, 70]. This inversion of polarity is more pronounced for larger
distances and for larger values of ground resistivity, as shown in Figure 8.11.
It is worth mentioning that in [71], the results obtained by the Cooray–
Rubinstein formula are compared with the calculations performed by solving the
Sommerfeld integral for the radial component of the electric field.
Figure 8.10 shows also that the voltage induced at farther distances along the line
from the strike location exhibits larger amplitudes than that close to the lightning.
Experimental data obtained by De La Rosa and co-workers [67] support this theoretical
finding, since for an event similar to the one presented in Figure 8.10, they observed a
line flashover occurring at the far end of the line and not at the close one.
4 0m
Induced Overvoltage (kV)
0
1 km
2 km 2.8 km
–4
–8
–12
430 m
–20
0 5 10 15 20
Time (μs)
Figure 8.10 Lightning-induced voltages along the line. Solid lines: perfect
ground; dashed lines: lossy ground. Adapted from [81]
310 Lightning electromagnetics: Volume 2
20
0
Ex.r 2 ( V/m * km2)
–20
–40
–60
100 m
–80 500 m
1 km
2 km
–100
0 2 4 6 8 10
Time (μs)
Figure 8.11 Radial electric field at four distances from the stroke location
calculated with the Cooray–Rubinstein formula [69,70]. Ground
conductivity 0.001 S/m, ground relative permittivity 10. Observation
point: 10 m above ground. For illustrative purposes, the values are
multiplied by the square of the distance from the stroke location.
Adapted from [59]
7.5 m 25 m
1.5 m
x V
Measured waveform
Calculated waveform (σ = 0.06 S/m)
2.25
Calculated waveform (σ = ∞)
Induced Voltage [V]
0
100 200 300
Time [ns]
–0.75
dIðxÞ
þ Y 0 V ðxÞ ¼ 0 (8.55)
dx
160
with corona
140 without corona
Obs.
Overvoltage [kV] 120 point #3
100
Obs.
80 point #2
60 Obs.
point #1
40
20
0
0 1 2 3 4 5 6
(a) Time [μs]
180
with corona
160
without corona
140 Obs.
120 point #1
Overvoltage [kV]
100
Obs.
80 point #2
60 Obs.
point #3
40
20
0
0 1 2 3 4 5 6
(b) Time [μs]
50 m 1000 m
50 m
50 m
Stroke
Location B 500 m Stroke
Location A
Top view
Side view
1.06 cm
Zc (stroke location A) Zc
Open (stroke location B)
(c) 1000 m
g2g
Yg0 ffi (8.62)
Zg0
Air
Ground σg, ε rg
2b
d
2a
Conductor
σw, ε rw
Insulating jacket
σi, ε ri
Figure 8.14 Geometry of the buried cable (cylindrical conductor of radius a with
an insulating jacket)
314 Lightning electromagnetics: Volume 2
Petrache et al. [78] have also shown that, within the frequency range of interest, the
wire impedance can be neglected due to its small contribution to the overall longitudinal
impedance of the line. The ground admittance, however, can play an important role at
high frequencies (1 MHz or so) especially in the case of poor ground conductivity. The
ground admittance needs to be taken into account in the calculation of lightning-induced
currents and voltages on buried cables. This is in contrast with the case of overhead lines
in which its contribution is generally negligible even in the MHz range.
ðL
V ðxÞ ¼ GV ðx; xs ÞVs0 dxs (8.65)
0
where GI and GV represent the Green’s function for the cable current and voltage,
respectively, which are given by [9]:
8 h i
> egL
>
> egðxs LÞ r2 egðxs LÞ ðegx r1 egx Þfor x < xs
<
2Zc ð1 r1 r2 e2gL Þ
GI ðx; xs Þ ¼ (8.66)
>
> egL h i
>
: e gðxLÞ
r e gðxLÞ
ðegxs r1 egxs Þfor x < xs
2Zc ð1 r1 r2 e2gL Þ 2
Interaction of lightning-generated electromagnetic fields 315
8 h i
>
> egL gðxs LÞ gðxs LÞ
>
< e r e ðegx þ r1 egx Þfor x < xs
2ð1 r1 r2 e2gL Þ 2
GV ðx; xs Þ ¼ h i (8.67)
>
> egL
>
: egðxLÞ þ r2 egðxLÞ ðegxs r1 egxs Þfor x < xs
2ð1 r1 r2 e 2gL Þ
pffiffiffiffiffiffiffiffiffi
where g ¼ Z 0 Y 0 is the line complex propagation constant along the cable and
pffiffiffiffiffiffiffiffiffiffiffi
Zc ¼ Z 0 =Y 0 is the line characteristic impedance.
A frequency-domain solution is particularly useful when one is interested in
calculating the inner response of a shielded cable [79], which involves the cable
transfer function, a highly frequency-dependent quantity.
0 ð jwC 0 Þ2
Yadd ¼ (8.70)
jwC 0 þ Yg0
130
loca tion #1
S troke location
474
m m
9
32 43 m
Strike
m 424 m
256
6m
Z 45 m Z
133 m
Figure 8.15 Positions of the triggered lightning strokes (top) and buried cable
experimental set-up (bottom). The cable shield is connected to the
ground rods at IS 1 and IS 2. Adapted from [80]
Figure 8.15 illustrates the positions (strike locations) for which experimental
data were recorded.
A comparison between the measured currents in the cable shield at the IS2
termination, and those predicted by simulations is presented in Figure 8.16 [80].
Figure 8.17 presents the measured and simulated currents in the inner conductor of
the buried shielded cable. The coupling to the inner conductor was evaluated using
the concept of cable transfer impedance. It can be seen that the simulation results
are in good agreement with experimental data.
8.5 Conclusions
In this chapter, we discussed the TL theory and its application to the problem of
lightning electromagnetic field coupling to transmission lines.
After a short discussion on the underlying assumptions of the TL theory, we
described seemingly different but completely equivalent approaches that have been
proposed to describe the coupling of electromagnetic fields to transmission lines.
The field-to-transmission line coupling equations were then extended to deal
with the presence of losses and multiple conductors and expressions for the line
parameters, including the ground impedance and admittance were presented. The
time-domain representation of the field-to-transmission line coupling equations,
Interaction of lightning-generated electromagnetic fields 317
160
Simulated shield current IS2
140 Measured shield current IS2
120
Shield current [A]
100
80
60
40
20
0
0 1 2 3 4 5 6 7 8
Time [microseconds]
Figure 8.16 Comparison between experimental and simulation results for the
lightning-induced current in the shield of the experimental cable for
the first return stroke of a single stroke flash recorded on August 18,
2002; strike location #1. Adapted from [80]
5
Simulated shield current IS2
4 Measured shield current IS2
3
Inner current [A]
–1
–2
–5 0 5 10 15 20 25
Time [microseconds]
Figure 8.17 Experimental data and simulation results for the lightning-induced
currents in the inner conductor of the experimental cable for the flash
recorded on August 18, 2002; strike location #1. Adapted from [79]
318 Lightning electromagnetics: Volume 2
Acknowledgments
The material presented in this chapter is mainly the results of a joint Italo-Swiss
research cooperation carried out during the last two decades involving Alberto
Borghetti, Silvia Guerreri, Michel Ianoz, Carlo Mazzetti, Mario Paolone, and
Emanuel Petrache. Their contribution and support are gratefully acknowledged.
The authors wish to express their gratitude to V.A. Rakov, F.M. Tesche, and M.A.
Uman for their precious cooperation throughout these years.
References
[1] Nucci, C. and Rachidi, F., 2014. Interaction of electromagnetic fields
generated by lightning with overhead electrical networks, in: Cooray, V.
(Ed.), The Lightning Flash. IEE, London, pp. 559–610.
[2] Le Vine, D.M. and Meneghini, R., 1978. Simulation of radiation from
lightning return strokes: the effects of tortuosity. Radio Sci. 13, 801–809.
[3] Le Vine, D.M. and Meneghini, R., 1978. Electromagnetic fields radiated
from a lightning return stroke: application of an exact solution to Maxwell’s
equations. J. Geophys. Res. 83, 2377–2384.
[4] Moini, R., Sadeghi, S.H.H., Kordi, B., and Rachidi, F., 2006. An antenna-
theory approach for modeling inclined lightning return stroke channels.
Electr. Power Syst. Res. 76, 945–952.
[5] Sakakibara, A., 1989. Calculation of induced voltages on overhead lines
caused by inclined lightning studies. IEEE Trans. Power Deliv. 4, 683–693.
[6] Wu, S.C. and Hsiao, W.T., 1994. Characterization of induced voltages on
overhead power lines caused by lightning strokes with arbitrary configurations.
In 1994 Int. Conf. on Systems, Man and Cybernetics, vol. 3, pp. 2706–2710.
Interaction of lightning-generated electromagnetic fields 319
[7] Rakov, V.A. and Uman, M.A., 1998. Review and evaluation of lightning
return stroke models including some aspects of their application. IEEE
Trans. Electromagn. Compat. 40, 403–426.
[8] Rakov, V.A. and Rachidi, F., 2009. Overview of recent progress in lightning
research and lightning protection. IEEE Trans. Electromagn. Compat. 51,
428–442.
[9] Tesche, F.M., Ianoz, M., and Karlsson, T., 1997. EMC Analysis Methods and
Computational Models. Wiley Interscience, New York, NY.
[10] Cooray, V., 2003. The Lightning Flash, IEE Power and Energy Series. IEE.
[11] Nucci, C.A., Rachidi, F., and Rubinstein, A., 2008. Derivation of telegrapher’s
equations and field-to-transmission line interaction. In Rachidi, F. and Tkachenko,
S. (eds) (eds.) , Electromagnetic Field Interaction with Transmission Lines: From
Classical Theory to HF Radiation Effects. WIT Press, Southampton.
[12] Paul, C.R., 1994. Analysis of Multiconductor Transmission Lines. John
Wiley and Sons, New York, NY.
[13] Paolone, M., Rachidi, F., Borghetti, A., et al., 2009. Lightning electro-
magnetic field coupling to overhead lines: theory, numerical simulations
and experimental validation. IEEE Trans. Electromagn. Compat. Press. 51,
532–547.
[14] Taylor, C.D., Satterwhite, R.S., and Harrison, C.W., 1965. The response of a
terminated two-wire transmission line excited by a nonuniform electro-
magnetic field. IEEE Trans. Antennas Propag. AP. 13, 987–989.
[15] Agrawal, A.K., Price, H.J., and Gurbaxani, S.H., 1980. Transient response of
multiconductor transmission lines excited by a nonuniform electromagnetic
field. IEEE Trans. Electromagn. Compat. 22, 119–129.
[16] Rachidi, F., 1993. Formulation of the field-to-transmission line coupling
equations in terms of magnetic excitation fields. IEEE Trans. Electromagn.
Compat. 35, 404–407.
[17] Rusck, S., 1958. Induced lightning overvoltages on power transmission lines
with special reference to the overvoltage protection of low voltage networks.
Trans. R. Inst. Technol. Stockh. 120.
[18] Cooray, V., 1994a. Calculating lightning-induced overvoltages in power
lines: a comparison of two coupling models. IEEE Trans. Electromagn.
Compat. 36, 179–182.
[19] Cooray, V., Rachidi, F., and Rubinstein, M., 2017. Formulation of the field-
to-transmission line coupling equations in terms of scalar and vector poten-
tials. IEEE Trans. Electromagn. Compat. 59, 1586–1591. https://doi.org/
10.1109/TEMC.2017.2657891
[20] Piantini, A., 2017. Extension of the rusck model for calculating lightning-
induced voltages on overhead lines considering the soil electrical para-
meters. IEEE Trans. Electromagn. Compat. 59, 154–162. https://doi.org/
10.1109/TEMC.2016.2601011
[21] Rachidi, F., Nucci, C.A., M., I., and Mazzetti, C., 1996b. Influence of a lossy
ground on lightning-induced voltages on overhead lines. IEEE Trans.
Electromagn. Compat. 38, 250–263.
320 Lightning electromagnetics: Volume 2
[22] Ramo, S., Whinnery, J.R., and van Duzer, T., 1994. Fields and Waves in
Communication Electronics, 3rd ed. Wiley, New York, NY.
[23] Rachidi, F., Nucci, C.A., Ianoz, M., and Mazzetti, C., 1996a. Importance of
losses in the determination of lightning-induced voltages on overhead lines.
In EMC 96 ROMA Int. Symp. Electromagn. Compat. Univ Rome Sapienza
Rome Italy 2.
[24] Tesche, F.M., 1992. Comparison of the transmission line and scattering
models for computing the HEMP response of overhead cables. IEEE Trans.
Electromagn. Compat. 34, 93–99.
[25] Sunde, E.D., 1968. Earth Conduction Effects in Transmission Systems.
Dover Publication, New York, NY.
[26] Rachidi, F., Nucci, C.A., and Ianoz, M., 1999. Transient analysis of
multiconductor lines above a lossy ground. IEEE Trans PWDR. 14,
294–302.
[27] Tossani, F., Napolitano, F., and Borghetti, A., 2018. Inverse laplace trans-
form of the ground impedance matrix of overhead lines. IEEE Trans.
Electromagn. Compat. 60, 2033–2036. https://doi.org/10.1109/TEMC.2017.
2765207
[28] Tossani, F., Napolitano, F., and Borghetti, A., 2017. New integral formulas
for the elements of the transient ground resistance matrix of multiconductor
lines. IEEE Trans. Electromagn. Compat. 59, 193–198. https://doi.org/
10.1109/TEMC.2016.2591590
[29] Hoidalen, H.K., 2003. Analytical formulation of lightning-induced voltages
on multiconductor overhead lines above lossy ground. IEEE Trans.
Electromagn. Compat. 45, 92–100.
[30] Hoidalen, H.K., 1999. Calculation of lightning-induced overvoltages using
MODELS. In Int. Conf. Power Syst. Transients Tech Univ Bp. Bp. Hung.
[31] Hoidalen, H.K., 1997. Lightning-induced overvoltages in low-voltage sys-
tems (Ph.D. thesis). Norwegian University of Science and Technology.
[32] Montano, R., Theethayi, N., and Cooray, V., 2008. An efficient imple-
mentation of the Agrawal et al. model for lightning-induced voltage calcu-
lations using circuit simulation software. IEEE Trans. Circuits Syst. – Regul.
Pap. 55, 2959–2965.
[33] Napolitano, F., Borghetti, A., Nucci, C.A., Paolone, M., Rachidi, F., and
Mahserejian, J., 2008. An advanced interface between the liov code and the
EMTP-RV. In Proc. 29th Int. Conf. Lightning Protection, Uppsala, Sweden,
pp. 1–12.
[34] Nucci, C.A., Bardazzi, V., Iorio, R., Mansoldo, A., and Porrino, A., 1994.
A code for the calculation of lightning-induced overvoltages and its interface
with the Electromagnetic Transient program. In Proc. of the 22nd
International Conference on Lightning Protection, Budapest, Hungary,
19–23 September 1994.
[35] Orzan, D., 1998. Couplage externe et interne entre un champ électro-
magnétique et un réseau de lignes multifilaires (Ph.D. thesis). Ecole
Polytechnique Federale de Lausanne, Lausanne, Switzerland.
Interaction of lightning-generated electromagnetic fields 321
[36] Orzan, D., Baraton, P., Ianoz, M., and Rachidi, F., 1996. Comparaison entre
deux approches pour traiter le couplage entre un champ EM et des réseaux
de lignes.
[37] Perez, E., Delgadillo, A., Urrutia, D., and Torres, H., 2007. Optimizing the
surge arresters location for improving lightning induced voltage perfor-
mance of distribution network, presented at the IEEE PES General Meeting,
Tampa, FL, USA, https://doi.org/10.1109/PES.2007.385697.
[38] Borghetti, A., Gutierrez, A., Nucci, C.A., Paolone, M., Petrache, E., and
Rachidi, F., 2004. Lightning-induced voltages on complex distribution sys-
tems: models, advanced software tools and experimental validation.
J. Electrost. 60, 163–174.
[39] Rachidi, F., Loyka, S., Nucci, C.A., and Ianoz, M., 2003. A new expression
for the ground transient resistance matrix elements of multiconductor over-
head transmission lines. Electr. Power Syst. Res. J. 65, 41–46.
[40] Araneo, R. and Cellozi, S., 2001. Direct time domain analysis of trans-
mission lines above a lossy ground. IEE Proc. Sci. Meas. Technol. 148,
73–79.
[41] Theethayi, N. and Thottappillil, R., 2008. Surge propagation and cross-
talk in multiconductor transmission lines above ground. In Rachidi, F.
and Tkachenko, S. (eds), Electromagnetic Field Interaction with
Transmission Lines. From Classical Theory to HF Radiation Effects.
England: WIT Press.
[42] Tafflove, A., 1995. Computational Electrodynamics: The Finite Difference
Time Domain Method. Norwood, MA:Artech House.
[43] Paolone, M., Nucci, C.A., and Rachidi, F., 2001. A new finite difference
time domain scheme for the evaluation of lightning induced overvoltages on
multiconductor overhead lines. In International Conference on Power
System Transients (IPST), pp. 596–602.
[44] Lax, P.D. and Wendroff, B., 1960. System of conservations laws. Commun.
Pure Appl. Math. 13, 217–237.
[45] Omick, S.R. and Castillo, S.P., 1993. A new finite difference time-domain
algorithm for the accurate modeling of wide-band electromagnetic phe-
nomena. IEEE Trans. Electromagn. Compat. 35, 215–222.
[46] Paolone, M., 2001. Modeling of lightning-induced voltages on distribution
networks for the solution of power quality problems, and relevant imple-
mentation in a transient program (Ph.D. thesis). University of Bologna,
Bologna, Italy.
[47] Brignone, M., Procopio, R., Nicora, M., Mestriner, D., Rachidi, F., and
Rubinstein, M., 2022. A prony-based approach for accelerating the lightning
electromagnetic fields computation: effect of the soil finite conductivity.
Electr. Power Syst. Res. 209, 108013. https://doi.org/10.1016/j.epsr.2022.
108013
[48] Darveniza, M., 2007. A practical extension of Rusck’s formula for max-
imum lightning induced voltage that accounts for ground resistivity. IEEE
Trans. Power Deliv. 22, 605–612.
322 Lightning electromagnetics: Volume 2
[49] Andreotti, A., Pierno, A., and Rakov, V.A., 2013a. An analytical approach
to calculation of lightning induced voltages on overhead lines in case of
lossy ground—Part I: model development. IEEE Trans. Power Deliv. 28,
1213–1223. https://doi.org/10.1109/TPWRD.2013.2241084
[50] Andreotti, A., Pierno, A., Rakov, V.A., and Verolino, L., 2013b. Analytical for-
mulations for lightning-induced voltage calculations. IEEE Trans. Electromagn.
Compat. 55, 109–123. https://doi.org/10.1109/TEMC.2012.2205001
[51] Mestriner, D., Brignone, M., Procopio, R., et al., 2021a. An efficient meth-
odology for the evaluation of the lightning performance of overhead lines.
IEEE Trans. Electromagn. Compat. 63, 1137–1145. https://doi.org/10.1109/
TEMC.2021.3054427
[52] Mestriner, D., Brignone, M., Procopio, R., Piantini, A., and Rachidi, F.,
2021b. A new channel-base lightning current formula with analytically
adjustable parameters. IEEE Trans. Electromagn. Compat. 63, 542–549.
https://doi.org/10.1109/TEMC.2020.3009273
[53] Mestriner, D., Brignone, M., Procopio, R., et al., 2021c. Analytical expressions
for lightning electromagnetic fields with arbitrary channel-base current. Part II:
validation and computational performance. IEEE Trans. Electromagn.
Compat. 63, 534–541. https://doi.org/10.1109/TEMC.2020.3018108
[54] Andreotti, A., Rakov, V.A., and Verolino, L., 2018. Exact and approximate
analytical solutions for lightning-induced voltage calculations. IEEE Trans.
Electromagn. Compat. 60, 1850–1856. https://doi.org/10.1109/TEMC.2018.
2812422
[55] Brignone, M., Procopio, R., Mestriner, D., et al., 2021. Analytical expres-
sions for lightning electromagnetic fields with arbitrary channel-base
current—Part I: theory. IEEE Trans. Electromagn. Compat. 63, 525–533.
https://doi.org/10.1109/TEMC.2020.3018199
[56] Napolitano, F., 2011. An analytical formulation of the electromagnetic field
generated by lightning return strokes. IEEE Trans. Electromagn. Compat.
53, 108–113. https://doi.org/10.1109/TEMC.2010.2065810
[57] Nicora, M., Mestriner, D., Brignone, M., et al., 2022. Assessment of the
lightning performance of overhead distribution lines based on Lightning
Location Systems data. Int. J. Electr. Power Energy Syst. 142, 108230.
https://doi.org/10.1016/j.ijepes.2022.108230
[58] Nicora, M., Mestriner, D., Brignone, M., et al., 2021. Estimation of the
lightning performance of overhead lines accounting for different types of
strokes and multiple strike points. IEEE Trans. Electromagn. Compat. 63,
2015–2023. https://doi.org/10.1109/TEMC.2021.3060139
[59] Guerrieri, S., Nucci, C.A., and Rachidi, F., 1997. Influence of the ground
resistivity on the polarity and intensity of lightning induced voltages, in 10th
International Symposium on High Voltage Engineering, Montreal, Canada.
[60] Hermosillo, V.F. and Cooray, V., 1995. Calculation of fault rates of over-
head power distribution lines due to lightning induced voltages including
the effect of ground conductivity. IEEE Trans Electromagn. Compat. 37,
392–399.
Interaction of lightning-generated electromagnetic fields 323
[61] Ishii, M., Michishita, K., Hongo, Y., and Ogume, S., 1994. Lightning-
induced voltage on an overhead wire dependent on ground conductivity.
IEEE Trans. Power Deliv. 9, 109–118.
[62] Rachidi, F., 2012. A review of field-to-transmission line coupling models
with special emphasis to lightning-induced voltages on overhead lines. IEEE
Trans. Electromagn. Compat. 54, 898–911.
[63] Nucci, C.A., Guerrieri, S., Correia de Barros, M.T., and Rachidi, F., 2000.
Influence of corona on the voltages induced by nearby lightning on overhead
distribution lines. IEEE Trans. Power Deliv. 15, 1265–1273.
[64] Tossani, F., Borghetti, A., Napolitano, F., Piantini, A., and Nucci, C.A.,
2018. Lightning performance of overhead power distribution lines in urban
areas. IEEE Trans. Power Deliv. 33, 581–588. https://doi.org/10.1109/
TPWRD.2017.2658183
[65] CIGRE-CIRED JWG C4.4.02: Protection of MV and LV Networks against
Lightning. Part I: Common Topics, 2006. CIGRE Tech. Broch. No 287.
[66] Dommel, H.W., 1986. Electromagnetic Transient Program Reference Manual
(EMTP Theory Book). Bonneville Power Administration, Portland, OR.
[67] De la Rosa, F., Valdivia, R., Pérez, H., and Loza, J., 1988. Discussion about
the inducing effects of lightning in an experimental power distribution line in
Mexico. IEEE Trans. Power Deliv. 3, 1080.
[68] Chowdhuri, P., Cooray, V., Correia de Barros, M.T., et al., 2005. Lightning
induced voltages on overhead power lines. Part III: sensitivity analysis.
Electra.222, 27–30.
[69] Cooray, V., 1994b. Lightning-induced overvoltages in power lines: validity
of various approximations made in overvoltage calculations. IEEE Trans.
Electromag. Compat., 40, 355–363.
[70] Rubinstein, M., 1996. An approximate formula for the calculation of the
horizontal electric field from lightning at close, intermediate, and long range.
IEEE Trans. Electromagn. Compat. 38, 531–535.
[71] Tossani, F., Napolitano, F., Borghetti, A., and Nucci, C.A., 2019. Influence
of the radial electric field appraisal on lightning-induced overvoltages sta-
tistical assessment. IEEE Trans. Electromagn. Compat. 61, 637–643. https://
doi.org/10.1109/TEMC.2019.2899656
[72] Borghetti, A., Nucci, C.A., and Paolone, M., 2009. Indirect-lightning per-
formance of overhead distribution networks with complex topology. Power
Deliv. IEEE Trans. On 24, 2206–2213.
[73] Borghetti, A., Nucci, C.A., and Paolone, M., 2007. An improved procedure
for the assessment of overhead line indirect lightning performance and its
comparison with the IEEE Std. 1410 method. IEEE Trans. Power Deliv. 22,
684–692.
[74] Dragan, G., Florea, G., Nucci, C.A., and Paolone, M., 2010. On the influence
of corona on lightning-induced overvoltages. In 30th International
Conference on Lightning Protection – ICLP 2010.
[75] Cooray, V., 1992. Horizontal fields generated by return strokes. Radio Sci.
27, 529–537.
324 Lightning electromagnetics: Volume 2
[76] Galvan, A. and Cooray, V., 1999. Lightning induced voltages on bare and
insulated buried cables. In Proceedings of the 13th International Symposium
on Electromagnetic Compatibility, Zurich, February, 1999.
[77] Vance, E.F., 1978. Coupling with Shielded Cables. New York, NY:John
Wiley and Sons.
[78] Petrache, E., Rachidi, F., Paolone, M., Nucci, C., Rakov, V.A., and Uman,
M.A., 2005. Lightning-induced voltages on buried cables”, Part I: theory.
IEEE Trans. Electromagn. Compat. 47, 498–508.
[79] Petrache, E., Paolone, M., Rachidi, F., et al., 2007. Lightning-induced
currents in buried coaxial cables: a frequency-domain approach and its
validation using rocket-triggered lightning. J. Electrost. 65, 322–328. https://
doi.org/10.1016/j.elstat.2006.09.015
[80] Paolone, M., Petrache, E., Rachidi, F., et al., 2005. Lightning-induced vol-
tages on buried cables. Part II: experiment and model validation. IEEE
Trans. Electromagn. Compat. 47, 509–520.
[81] Rachidi, F., Nucci, C.A., Guerrieri, S., and Correia de Barros, M.T., 2001.
On the amplitude enhancement of voltages induced by external EM fields on
transmission lines due to ground losses and corona phenomenon. In 2001
IEEE EMC International Symposium. Symposium Record. International
Symposium on Electromagnetic Compatibility.
[82] Ishii, M., Michishita, K., and Hongo, Y., 1997. Verification of coupling
model for calculation of induced lightning voltages, in CIGRE International
Colloquium on Insulation Coordination, Toronto, Canada.
Chapter 9
Application of scale models to the study of
lightning transients in power transmission and
distribution systems
Alexandre Piantini1 and Jorge M. Janiszewski2
9.1 Introduction
A power system’s reliability bears on its ability to supply continuous and unin-
terrupted energy without significant momentary disturbances. However, power
lines are often located in areas of high ground flash densities, being therefore prone
to lightning-caused faults.
For most overhead power transmission lines, lightning is the primary cause of
unscheduled interruptions, which may result in either shielding failure or back-
flashover. The first situation occurs when the lightning stroke bypasses the overhead
ground wire and terminates directly on a phase conductor. If the stroke current is
larger than a certain value, the resultant voltages across the insulator strings will
exceed the line critical impulse flashover voltage (CFO) and lead to flashover.
Depending on parameters such as the stroke current magnitude and front time, line
CFO, tower impedance, and footing resistance, even when lightning strikes the tower
or an overhead ground wire the voltages across the insulators may be high enough to
cause a flashover. In such a situation, it is called backflashover, as it results from a
lightning strike to part of the network which is usually at ground potential.
Concerning distribution networks, failures of power equipment, especially
transformers and pin insulators, are frequently observed, particularly when lines
cross rural areas and are consequently more exposed to direct strikes. Lightning
overvoltages on distribution lines may be originated either from direct strikes or by
induction due to nearby flashes. Although the former type is much more severe, the
latter is usually responsible for a more significant number of line flashovers and
supply interruptions on systems with rated voltage 15 kV or less due to their higher
frequency of occurrence and to the low line insulation withstand capability. The
1
Institute of Energy and Environment, Lightning and High Voltage Research Center, University of São
Paulo, São Paulo, Brazil
2
Polytechnic School, Telecommunications and Control Department, University of São Paulo, São
Paulo, Brazil
326 Lightning electromagnetics: Volume 2
namely overhead lines, transformers, and surge arresters, are described, and details
are given on the reduced system implemented at the University of São Paulo, Brazil.
The last part of the chapter is dedicated to the application of the technique to the
evaluation of lightning transients, with emphasis on the analysis of lightning-induced
voltages on overhead power distribution lines. The versatility of the scale model
technique is demonstrated and examples illustrate its usefulness in the study of
complex phenomena, either for enabling the evaluation of situations that are not
worthwhile to be treated theoretically or for giving adequate support for the valida-
tion of theoretical models and relevant computer codes.
@E m
rm H m ¼ sm E m þ em (9.4)
@tm
@H m
rm E m ¼ mm ; (9.5)
@tm
with
H m ¼ H m ðxm ; ym ; zm ; tm Þ; E m ¼ E m ðxm ; ym ; zm ; tm Þ
and sm ; em ; mm representing the characteristics of the scale model medium.
328 Lightning electromagnetics: Volume 2
In (9.4) and (9.5), the symbol rm stands for differentiation with respect to the
model coordinates. It can be shown that the scale factor for the curl operator is 1/p.
Thus,
1 b
rm H m ¼ r H m ¼ r H : (9.6)
p p
Analogously:
a
rm E m ¼ r E: (9.7)
p
and
@H m b @H
¼ : (9.9)
@tm g @t
and
a b @H
r E ¼ mm : (9.11)
p g @t
In order to correctly represent the real system by the model, (9.10), (9.11) and
(9.1), (9.2) must be equivalent. Thus:
8
> gb
>
> e ¼ e
> m pa
>
>
>
< ag
mm ¼ m: (9.12)
>
> pb
>
>
>
> b
>
: sm ¼ s
pa
The scale factor of any electromagnetic quantity can be determined from the
above definitions. If air is the medium in both systems, the following conditions
must prevail:
8
> gb
>
< em ¼ e ! pa ¼ 1
: (9.13)
>
> ag
: mm ¼ m ! ¼1
pb
Scale models to the study of lightning transients 329
Then,
a ¼ b; g ¼ p; sm ¼ s=p; (9.14)
i.e., the scale factor for conductivity must be the inverse of that of length. Clearly,
the last condition is not satisfied, as air is the medium for both systems. However,
considering that air is a good insulator, the error resulting in not taking its con-
ductivity into account can be neglected.
The scale factors for the quantities of interest, considering the general case and
the special one, when the medium (air) is the same for both systems, are reported in
Table 9.1. The results show that if p and either a or b are known, the scale factors
for all quantities can be determined. On the other hand, if only p and the ratio
between a and b are known, only some of the quantities can be directly related.
distribution lines. The scale factors relevant to the electrical quantities of this sys-
tem are reported in Table 9.2, whereas a general view of the experimental facility,
showing the ground plane and a typical configuration of an urban distribution
system with the main feeder and some laterals is presented in Figure 9.1.
Figure 9.1 General view of the scale model. Generation and measuring systems
located below the ground plane [51]
Scale models to the study of lightning transients 331
The propagation velocity along the lightning channel model and the surge
impedance of the simulated channel (Z0) were experimentally determined from the
analysis of reflected voltages resulting from the application of a step signal at one
end, leaving the other one open. The value obtained for Z0 was 2.2 kW. In order to
eliminate reflections, a non-inductive resistor R = 2.2 kW was connected between
the top of the stroke channel model and the mesh of conductors providing the return
path for the injected current, as shown in Figure 9.3. The return conductors, some
of which can be seen in Figure 9.1, were separated by a distance of approximately
1 m. Therefore, for the relevant frequencies of the current spectrum, the return
conductors can be reasonably assumed to represent a conducting surface involving
the system, so that the electromagnetic field in the region under analysis is estab-
lished by the current along the channel model. As pointed out in [35], although the
channel model was not intended to provide an accurate simulation of the lightning
channel, for points relatively far from it, the current can be assumed to propagate
upwards with the determined velocity, so that the radiated electromagnetic field can
be calculated with good accuracy.
As air is the same medium for both systems and therefore the scale factors for
time and length are equal, currents with very short rise times are needed to simulate
typical lightning currents. For a scale model with p = 1:50, the front time of the
simulated stroke current must be about some tens of nanoseconds (e.g., 40 ns for a
R = Z0
To the To the
ground plane ground plane
Figure 9.3 Schematic diagram of the experimental setup (adapted from [35])
Scale models to the study of lightning transients 333
typical first stroke front time of 2 ms). Moreover, the current amplitude must be of
the order of at least a few amperes to enable the simulation of surge arresters.
As the channel surge impedance Z0 is relatively high, it would not be possible to
obtain currents with the desired characteristics using a capacitor. Therefore, a high
voltage cable, 280 m long, charged by a dc voltage source, was used to generate
the required steep front currents [35,41,45]. Once the desired voltage was reached,
the cable was connected to the channel model through a high-speed switch. As the
circuit impedance was predominantly resistive, this technique allowed the genera-
tion of currents with the desired, relatively short, front times.
The current injected into the channel model was measured at ground level by a
current probe associated with an amplifier. The measuring system bandwidth was
from dc to 50 MHz, and its rise time was shorter than 7 ns. For the measurements
reported in [35], digital storage oscilloscopes with sampling rates of either 500 MS/s
or 1 GS/s (single-shot bandwidths of 200 MHz and 250 MHz, respectively)
were used.
9.3.2 Ground
The ground resistivity depends on the type of soil and various factors such as soil
granulometry, degree of compactness, quantity, and kind of dissolved salts, strati-
fication, water content, and temperature. It varies widely and, depending on the
condition, it may assume values lower than 100 W.m or greater than 10,000 W.m.
With a few exceptions [48], in scale model experiments, the ground is usually
assumed as a perfectly conducting plane and simulated by means of copper or
aluminum sheets [31–47]. In [59], Rachidi et al. show that the approximation of a
perfectly conducting ground is, in general, reasonable for the calculation of both the
azimuthal magnetic field and the vertical component of the electric field for dis-
tances between the lightning stroke location and the observation point shorter than
about one kilometer. However, both fields’ time derivatives may be significantly
affected by the propagation effects, as pointed out by Cooray [60,61]. On the other
hand, the earth resistivity has a remarkable impact on the horizontal electric field
[59,62–67] and, by extension, on the lightning-induced voltages. Several investi-
gations have been carried out about the influence of a lossy ground on lightning-
induced voltages on overhead lines [9,48,59,62,68–71] and, according to Nucci
[62], the assumption of a perfectly conducting ground is reasonable for distribution
systems located above a soil with resistivity lower than about 100 W m.
In [35,41], the whole scale model was placed above a conducting plane rea-
lized by means of interconnected aluminum plates covering an area of 28 9 m2,
which is equivalent to 1,400 450 m2 on a full-scale basis.
9.3.4 Transformers
A transformer can be viewed as a complex network of capacitances, resistances,
self-inductances, and mutual inductances, and the knowledge of its response
when subject to lightning and switching surges is of great importance for
designing and development. In [25], an electromagnetic model was developed to
evaluate transient voltages in the transformer windings. The model, which
represented an important advance in transient analysis in transformers, consisted
of two parts of different scale factors: an equivalent circuit of capacitances and a
geometrical model for the self and mutual inductances, even if non-linear. The
voltages measured at different points of real transformers and their respective
models were in very good agreement and the accuracy of the model was con-
sidered entirely adequate for design purposes. The electromagnetic model is
flexible, all points of interest are readily available, and measurements can be
made easily and rapidly.
Scale models to the study of lightning transients 335
100
80
Magnitude (kΩ)
60
40
20
0
1 10 100 1,000
(a) Frequency ( kHz)
–90
Phase (degrees)
–180
–270
–360
1 10 100 1,000
( b) Frequency ( kHz)
eliminate the circuit inductance effects and, therefore, the obtained peak values are
larger than the real ones. This discrepancy increases for faster front time currents,
and in [77] it is recommended to take the value of the voltage at the current peak
point, especially for front times shorter than about 4 ms.
As in the case of nearby strokes the currents through the arresters hardly ever
reach values higher than 1 kA [41,78], the similarity between the behaviors of the
model and of the actual arresters was also verified for relatively low currents. An
appropriate circuit with components selected to produce the standard current
waveform, taking into account the time scale factor, was used to test the model and
the actual arresters. Based on the analysis of the test results, the scale factor
1:18,000 was adopted for voltage and current and, as a consequence, the scale
factor for electric and magnetic fields (a) was determined as 1:360. Therefore, the
quantities in the model could be related to those of the full-scale system by
applying the relationships shown in Table 9. 1, as presented in Table 9.2.
Figure 9.6 shows the dynamic VI-characteristics, i.e., the curves relating
instantaneous values of voltages and currents, for the model and the distribution
arresters. The VI-characteristics obtained by combining all the test results are pre-
sented in Figure 9.7. As the scatter among the results relative to surge arresters of
the same type (ZnO or SiC) was small, the results corresponding to each type were
grouped. It can be readily seen that the surge arrester model represents relatively
well the behavior of the actual arresters within the range of the current peak values
considered.
The equivalent circuit of the surge arrester model is composed of a resistance
Rpr in series with an inductance Lpr, both in parallel with a capacitance Cpr, as
shown in Figure 9.8a. The behavior of the non-linear resistance Rpr is presented in
Figure 9.8b. The inductance Lpr and the capacitance Cpr were introduced to repre-
sent the model’s transient behavior more accurately. The inductance is related to
the increase in the residual voltage for currents with short front times, whereas the
capacitance is associated with the voltage wavefront.
50
SiC
40
Voltage (kV)
Model
30
ZnO
20
10
0 1 2 3 4 5
Current (kA)
Figure 9.6 Dynamic VI-characteristics for the model (scale factor 1:18,000 for
voltage and current) and actual SiC and ZnO surge arresters (adapted
from [35,41])
338 Lightning electromagnetics: Volume 2
40
30
20
10
0.1 1 10
Current (kA)
Figure 9.7 VI-characteristics for the model (scale factor 1:18,000 for voltage and
current) and actual SiC and ZnO surge arresters (adapted from
[35,41])
30
Voltage (kV)
20
Rpr
Cpr
Lpr 10
0
0 0.4 0.8 1.2 1.6
(a) (b) Current (kA)
Figure 9.8 Surge arrester model equivalent circuit (adapted from [35,41]):
(a) electric circuit; (b) characteristic of the non-linear resistor Rpr
Several preliminary simulations indicated that the circuit shown in Figure 9.8
represents the arrester model relatively well when the Lpr and Cpr values are,
respectively, 0.02 mH and 270 pF (corresponding to 1.0 mH and 13.5 nF on a
full-scale basis). Figure 9.9 presents some examples of comparisons between
calculations performed using the equivalent circuit and relevant measurements for
current amplitudes of 20 mA and 105 mA (corresponding to 0.36 kA and 1.89 kA).
The associated residual voltages were, respectively, 1.67 V and 1.83 V (30 kV and
33 kV on a full-scale basis). A very close agreement was reached in all compar-
isons, thus validating the arrester model equivalent circuit.
9.3.6 Buildings
Urban overhead distribution networks are characterized not only by a high density of
transformers, surge arresters, and laterals, but also by the presence of buildings in
their vicinity. The presence of nearby tall structures limits the line exposure to direct
Scale models to the study of lightning transients 339
0.4 40
Measured
30
Current (kA)
Voltage (kV)
0.2 20
Calculated
10
0 0
0 10 20 30 40 50 0 10 20 30 40 50
(a) Time (μs) Time (μs)
2 40
Calculated
30
Current (kA)
Voltage (kV)
Measured
1 20
10
0 0
0 10 20 30 40 50 0 10 20 30 40 50
(b) Time (μs) Time (μs)
Figure 9.9 Currents injected into the surge arrester model and corresponding
measured and calculated residual voltages (all scales referred to the
full-scale system). Adapted from [35,41]: (a) applied current: 0.36 kA;
(b) applied current: 1.89 kA
lightning strokes, but on the other hand lightning flashes may occur very close to the
line. Therefore, overvoltages of high magnitudes may be induced even in the case of
stroke currents of moderate intensity [39]. The induced overvoltages are affected by
the electromagnetic field distortion caused by neighboring buildings. The computa-
tion of lightning surges under such conditions is too complicated and cannot be done
with the existing versions of advanced codes such as the so-called LIOV-EMTP code
[79,80]. A method to take into account the effect of the attenuation of the lightning
electromagnetic pulse (LEMP) radiated by indirect lightning strokes due to buildings
has been presented by Borghetti et al. [81] and Tossani et al. [82]. Such attenuation is
represented by means of specific weighting functions applied to the LEMP analytical
expressions valid for open terrain. The weighting functions’ parameters are identified
through the least-square minimization of the differences with the results provided by
a finite-element method model that is assumed as reference for the configurations
analyzed. As shown in [82], the weighting functions can be used, with reasonable
accuracy, for lightning return stroke current waveforms and distances between the
line and the stroke location different from those used for their identification.
340 Lightning electromagnetics: Volume 2
According to Table 9.1, if the medium is the same for the model and the full-scale
system, the conductivity scale factor is the reciprocal of that of length. This means that,
for the system described in [35,39], buildings should be made of materials 50 times
more conductive than those of full-scale constructions. In [35,39], grounded aluminum
structures (cuboids) were used to simulate the buildings and analyze their effect on
lightning-induced voltages on overhead lines. As aluminum is characterized by an
electrical conductivity greater than 50 times those of the materials used in real struc-
tures, in principle the effect of the structures should be, in actual situations, less sig-
nificant than that observed in the tests [35,39,41]. Besides, as the shielding effect is due
mainly to the steel rebars embedded inside reinforced concrete, the buildings could be
better simulated by meshes of cubes with edges being constituted by metallic wires.
However, as shown in [81], the differences between the results obtained by repre-
senting the buildings as cuboids with perfectly conducting surfaces and those obtained
by simulating them by meshed structures of the steel rebars of the reinforced concrete
are small. This outcome suggests that the adopted building models are suitable and that
the results obtained from the scale model experiments can be used to assess the
shielding effect of nearby buildings on lightning-induced voltages on overhead lines.
One of the test configurations considered in [41], corresponding to structures
15 m high, is presented in Figure 9.10, where one of the return conductors shown in
Figure 9.3 can also be seen at the left side of the picture. In this particular case, the
channel model, which can be seen in the center of the figure, was only 20 m from
the closest lateral. Such a short distance is unlikely unless lightning hits an elevated
object like a tree, mast, or another structure protruding above the line, which is the
situation considered in the simulation. The distances between the buildings and the
main feeder, referred to the full-scale system, are indicated in Figure 9.11.
Figure 9.10 Scale model for studying the effect of nearby buildings on lightning-
induced voltages on overhead distribution lines (adapted from [41])
Scale models to the study of lightning transients 341
15 m
10 m
8m
6m 10 m 6m
Figure 9.11 Distances between buildings and main feeder corresponding to the
test configuration shown in Figure 9.10 (dimensions referred to the
full-scale system) [51]
role in estimating the outage rate. Although the tower surge impedance value itself
does not significantly influence the number of line outages, the tower travel time
values greatly affect the overvoltages’ characteristics and, consequently, the line
performance.
The estimation of the surge response of transmission line towers is not trivial,
and various theoretical approaches have been made for such calculations.
Approximate formulas have been proposed, sometimes with the towers represented
by either cylindrical or conical conductors. In some cases, measurements have been
made in actual towers. In this context, the use of reduced-scale models allows
verifying the validity of theoretical formulas or obtaining additional knowledge on
the towers’ behavior when submitted to direct lightning strikes.
In [83], Chisholm, Chow, and Srivastava proposed equations for tower surge
impedance and tower travel time to deal with midspan strokes, emphasizing that
previous impedance models derived for vertical strokes to tower top were not valid
for the more common midspan stroke. A new model for “horizontal” currents was
presented and verified with experiments conducted using reduced tower models.
The impedances of cylindrical, conical, and inverted-cone towers with heights of
40 cm (conical and inverted-cone) or 50 cm (cylindrical) were determined con-
sidering horizontal and vertical currents. A travelling-wave stub model for cross-
arms was also proposed, and tests with a tower model were performed to verify the
crossarm impedance values and the propagation times. The experiments clearly
showed the influence of the crossarms on the current propagation and supported the
stub models. It was shown that the crossarms divert and delay some stroke current,
and this reduces the tower impedance and increases the tower travel time.
In [84], Yamada et al. report measurements performed in both actual and
reduced models of transmission towers aiming at confirming theoretical results or
obtaining further information on their surge behavior. A step current generator was
connected to the top of a 140.5 m high tower of a UHV transmission line and
measurements were made of the voltages across the insulator strings, crossarms,
342 Lightning electromagnetics: Volume 2
and conductors. The voltage of the tower footing was also recorded. From the
measurements, a tower model was developed to fit the results of EMTP simula-
tions. In order to investigate the effect of the relative position of the current lead
and the auxiliary potential wires on the measurements, a scaled tower, 1.8 m high,
was built. A coaxial cable kept parallel to the ground connected the top of the tower
to a pulse generator on the ground. The voltage at the tower top was measured for
different positions of the auxiliary potential wire with respect to the current lead
wire, and it was shown that the surge impedance of the straight arrangement was
lower than that of the perpendicular arrangement.
In [85], a new method was proposed by Motoyama and Matsubara to calculate
the response of transmission towers to lightning surges. The method, based on the
electromagnetic field theory, showed that the tower surge response depends on the
propagation velocity of the lightning current and on the way it is injected into
the tower. Tests performed using scale models of 500 kV and UHV towers without
conductors and overhead ground wires showed the adequacy of the proposed
method and confirmed that the direction of the current injection line significantly
affects the tower surge response.
Gutiérrez et al. [86] proposed a new model to represent the transient behavior
of a tower struck by lightning. The various parts of the tower were represented
by equivalent vertical and/or horizontal transmission line segments, and complex
tower structures could be simulated, with straightforward inclusion of non-
uniformities, distributed losses, and wave speed variations. A reduced-scale model
was built and very fast pulses were applied to the top of a cylindrical copper con-
ductor with a height of 1 m and a radius of 1 mm. Due to mechanical requirements,
this was the smallest practical conductor’s radius, which was chosen to satisfy the
thin wire assumption. The voltage responses of the vertical line corresponding to
the injected current pulses were measured and compared with numerical simula-
tions carried out using the proposed model, and an excellent agreement was found.
Additional comparisons using data obtained from field experiments confirmed the
validity of the model.
Experimental results obtained using a reduced scale model of a vertical con-
ductor were used by Goni et al. [87] in comparisons with simulation results using
the FDTD method and a numerical electromagnetic code based on the method of
moments (NEC-2). The analysis involved the simulation of the surge phenomena of
a vertical conductor, including the effects of both horizontal and vertical wave
incidence. A tower was represented by a vertical cylinder with height and radius of
0.6 m and 2.5 mm, respectively, which was placed over a copper plate that simu-
lated the ground. In the case of horizontal incidence, the conductor was excited by a
pulse generator via a 2 m long horizontal wire with the remote end connected to the
ground. Bidirectional probes were used to measure the injected current and the
current flowing through the conductor. The voltage measuring wire had a hor-
izontal length of 1 m and was connected to the ground at the remote end. In the case
of the horizontal incidence, it was perpendicular to the current lead wire. For the
vertical wave incidence, the pulse generator was connected to the top of the con-
ductor, associated with a vertical current lead wire 1.5 m long. A good agreement
Scale models to the study of lightning transients 343
was found between the results of the scale model experiments and numerical
simulations.
time, on a full-scale basis, less than 0.1 ms, the measured ratios of the transfer
impedances corresponding to one or two ground wires with respect to the case of no
ground wire were, respectively, 0.6 and 0.43.
The effects of the channel impedance and of the stroke current propagation
velocity (v) were examined in [32] by measuring the voltage across the top insu-
lator for three different channel models. Two of them were helical lines; one with a
characteristic velocity of about 18% c and an impedance of 1,150 W, and the other
with a characteristic velocity of 32% c and an impedance of 750 W. The third
channel model was simply a length of wire suspended above the tower. Its char-
acteristic velocity was the same as the speed of light, and its impedance was about
500 W. The results showed that the effects of different stroke velocities and channel
impedances on the shapes of the transfer impedance curves are minor, as illustrated
in Figure 9.12. The same conclusion was obtained in [31].
Evaluations of the influence of the tower footing resistance on the transfer
impedance, considering the presence of one overhead ground wire, were carried out
in [31,32]. Three situations were considered in [31] regarding the value of the
footing resistance: 0 W (tower grounded solidly), 75 W, and 220 W. The stroke
current front time, on a full-scale basis, was less than 0.1 ms. In [32], the response of
the system was measured considering the following situations: tower grounded
solidly, grounded through 82 W (330 W at each leg), and isolated. The stroke
current front time, on a full-scale basis, was 0.05 ms; the voltage was measured at
the top insulator. As expected, both studies showed that, due to the time delay
for the arrival of the reflections originated at the tower bottom, the initial portion of
the waveforms was not affected. After a current wave had been reflected back
from the tower base, the transfer impedance and insulator voltage assumed a value
depending on the parallel resistance of the footing resistance and ground wire surge
impedances. The transfer impedances corresponding to the three arrangements
considered in [31] are presented in Figure 9.13.
The study conducted in [31] also included an evaluation of the influence of the
tower representation. Therefore, tests were performed considering three different
tower models: an exact scale reproduction, a combination of a cylinder and a cone,
and a cylinder. Very similar transfer impedances were obtained in the tests corre-
sponding to the three tower representations.
Transfer Impedance (Ω)
80 80 80
v = 0.18 c v = 0.32 c v = 1.0 c
40 40 40
0 0 0
0 0.5 1.0 1.5 0 0.5 1.0 1.5 0 0.5 1.0 1.5
Time (μs) Time (μs) Time (μs)
Figure 9.12 Transfer Impedance for top insulator as a function of stroke current
propagation velocity. Time scale referred to the full-scale system.
Adapted from [32]
346 Lightning electromagnetics: Volume 2
R = 220 Ω
50
R=0Ω
0 0.5 1
Time (μs)
Figure 9.13 Transfer impedance for different values of the tower footing
resistance (R). (Time scale referred to the full-scale system.)
Adapted from [31]
An analysis of the effects of the surge response of a tower and its conductors
and of the severity of the stroke current waveform required to produce flashover
voltages across the suspension insulators was carried out in [33] through the use of
a 1:28 scale model. A series of tests was made using a miniature representation of
an actual 51 m high transmission line tower and its associated conductors.
Lightning strikes to the tower and the overhead ground wire were simulated using a
repetitive surge generator. The applied surges had exponentially rising fronts
and essentially flat tails. The slope of the wavefront was varied and the resulting
voltage across the insulator of the model tower was measured. The results were
presented in terms of either the initial and the effective (determined from the 10%
and 90% points on the current waveform) stroke current rates of rise required to
cause insulator flashovers for the cases of the system without ground wire, with one
ground wire, and with two ground wires.
An investigation of the characteristics of lightning surges at a substation,
resulting from backflashover across an insulator string of the closest transmission
tower, considering non-horizontal or non-uniform lines incoming to the substation,
was conducted by Takami et al. [88]. The experiments were conducted in a
reduced-scale model composed of a horizontal transmission line, two towers, an
inclined incoming line, and a substation gantry. The scale factor was approximately
1:40, and the 500 kV tower models were 2 m high. The distance between the towers
was 4 m, and the gantry model stayed 3 m distant from the first tower, whose top
was connected to a pulse generator. Tests were performed simulating or not the
occurrence of backflashover, and the characteristics of the overvoltages at the
tower and the gantry were examined under both conditions. When backflashover
Scale models to the study of lightning transients 347
was considered, it was simulated by short-circuiting the crossarm and the upper
phase conductor of the first tower with a copper wire.
The experiments showed that the voltage at the tower crossarm rises gradually
in comparison with the current injected into the tower. The power line potential at
the gantry was found to be smaller than at the tower where the backflashover
occurs, in contrast with the trend observed from a circuit theory-based simulation in
which the inclined incoming line is represented by a horizontal line. Such
approximation has been usually employed in lightning surge simulations using the
Electromagnetic Transients Program (EMTP) or its alternatives. The measured
power line voltage at the gantry was about 50% of the voltage computed using
EMTP, and this would contribute to a more economical and rational design of a
substation. In [89], Takami et al. analyze the same system by FDTD simulation,
using the scale model results previously obtained to validate the computational
procedure.
20
Measured
15
Rusck
10
5
Voltage (kV)
0
Liew-Mar
–5
–10
–15
Chowdhuri
–20
–25
0 3 6 9 12 15
Time (μs)
that, owing to the characteristics of the optical-electrical converter used for the
voltage measurements, the recorded voltage waveforms presented a faster decay than
the original ones [8]. As the stroke current propagation velocity was not measured, a
constant value of 30% of that of light in free space was adopted for the calculations
presented in Figure 9.14. Other assumptions concern the soil, which was considered
as a perfectly conducting plane, and the stroke channel, which was considered ver-
tical, without branches, and 3 km long. The current distribution along the channel
was calculated according to the transmission line (TL) model [112].
The results presented in Figure 9.14 indicate that it is vital to utilize, for the
estimation of the lightning performance of a given line, a method – and the
respective computer code – whose validity has been demonstrated preferably from
comparisons with experimental results obtained under controlled conditions.
The reduced scale system developed in [41] has been used to validate two meth-
ods, namely the Extended Rusck Model (ERM) [41,97,99,100,113] and the LIOV-
EMTP [79,80,114]. The former is an extension to the Rusck model in the sense that,
unlike the original model, its features allow for taking into account situations of
practical interest such as, e.g., the case of a line with various sections of different
directions, the presences of a multi-grounded neutral or shield wire, and equipment
such as transformers and surge arresters. It allows also to take into account the elec-
trical characteristics of the soil [104]. The incidence of lightning flashes to nearby
elevated objects and the occurrence of upward leaders can also be considered [96].
Scale models to the study of lightning transients 349
12
4 calculated
1.4 m
Figure 9.15 Measured and calculated (using the ERM) induced voltages (adapted
from [115]). Stroke current with peak value of 50 kA, front time of
2 ms, and time to half-value of 85 ms. M: measuring point. (a) line
topology (top view); (b) induced voltages
350 Lightning electromagnetics: Volume 2
40 ns (2 ms), and time to half-value of 1.7 ms (85 ms). A good agreement is found
between the results derived from numerical simulation and experiment in terms of
both the voltage magnitudes and waveforms.
Figures 9.16–9.18 present comparisons between measured and calculated
induced voltages corresponding to three different currents injected into the channel
model. The single-phase straight line (hereafter referred to as “Reference Line”) was
28 m (1.4 km) long, matched at both ends, 1.4 m (70 m) from the channel model, and
its terminations were equidistant from the stroke location. The copper conductor was
supported by PVC poles spaced every 60 cm (30 m), and its diameter was 0.4 mm
(2 cm). The influence of the wire cross-section on the induced voltages is not sig-
nificant for the typical range commonly used [113]. The conductor height was 20 cm
(10 m), and its measured surge impedance was 455 W. The calculations were per-
formed with the recorded current waveforms approximated by straight lines, which in
the figures are shown superimposed to the measured currents.
Calculated
4
Voltage (V)
1.5
Current (A)
1.0
2
Measured 0.5
0 0
0 100 200 300 0 100 200 300
(a) Time (ns) (b) Time (ns)
Figure 9.16 Measured and calculated (using the ERM) induced voltages – Case
MOD-1 (adapted from [41]): (a) induced voltages; (b) measured
current and straight-line approximation superimposed
6
Calculated 1.6
4
Voltage (V)
1.2
Current (A)
0.8
2
Measured 0.4
0 0
0 100 200 300 0 100 200 300
(a) Time (ns) (b) Time (ns)
Figure 9.17 Measured and calculated (using the ERM) induced voltages – Case
MOD-2 (adapted from [41]): (a) induced voltages; (b) measured
current and straight-line approximation superimposed
Scale models to the study of lightning transients 351
1.5
6
Calculated 1.0
Current (A)
Voltage (V)
2 Measured 0.5
0 0
0 100 200 300 0 100 200 300
(a) Time (ns) (b) Time (ns)
Figure 9.18 Measured and calculated (using the ERM) induced voltages – Case
MOD-3 (adapted from [41]): (a) induced voltages; (b) measured
current and straight-line approximation superimposed
8 8
T T
6 6
Voltage (V)
Voltage (V)
4 4
R R
2 2
0 0
0 100 200 300 400 0 100 200 300 400
(a) Time (ns) (b) Time (ns)
Figure 9.19 Measured and calculated (using the ERM) induced voltages (adapted
from [41,113]). Reference Line (R): matched at both ends; Test Line (T):
one end matched and the other open. (a) Measured; (b) calculated
6 6
4 4
Voltage (V)
Voltage (V)
R R
2 2
T T
0 0
100 200 300 400 100 200 300 400
–2 –2
(a) Time (ns) (b) Time (ns)
Figure 9.20 Measured and calculated (using the ERM) induced voltages (adapted
from [41,113]). Reference Line (R): matched at both ends; Test Line
(T): one end matched and the other short-circuited. (a) measured
voltages; (b) calculated voltages
5 5
R R
0 0
Voltage (V)
Voltage (V)
–5 –5
T T
Figure 9.21 Measured and calculated (using the ERM) induced voltages (adapted
from [41,113]). Reference Line (R): matched at both ends; Test Line
(T): one end open and the other short-circuited. (a) measured
voltages; (b) calculated voltages
adjacent grounding points, the value of the ground resistance, the stroke current
steepness, and the position of the stroke location with respect to the line and the
observation point.
The examples presented in this subsection concern realistic line configurations,
representative of actual rural distribution lines. Thus, for convenience, the values of
all the parameters are referred to the full-scale system. The conversion to the values
actually recorded in the scale model experiments can be made by applying the scale
factors indicated in Table 9.2.
A comparison between measured and calculated voltages on a line with a
shield wire is presented in Figure 9.22. In this case, the Test Line had two con-
ductors, phase and shield wires, at the heights h = 10 m and hg = 9 m, respectively.
The horizontal distance between the conductors was 0.75 m. The voltages were
obtained at the point of the line closest to the stroke location, for the following
conditions: current magnitude I = 36 kA, current front time tf = 3.1 ms, distance
between the line and the lightning channel d = 70 m, distance between adjacent
grounding points xg = 450 m, and ground resistance Rg = 0 W. The stroke location
was in front of a grounding point and equidistant from the line terminations. This
test configuration is depicted in Figure 9.23a, whereas Figure 9.23b shows one of
the configurations used to evaluate the effect of surge arresters.
Figure 9.24 presents a comparison between measured and calculated voltages
for the case of surge arresters added on all the three phases of the Test Line,
which in this configuration had no shield wire. The distance between adjacent
conductors was 0.75 m. The observation point was in front of the channel model,
and the values of the parameters, referred to the full-scale system, were: I = 54 kA,
tf = 3.2 ms, xg = 450 m, and Rg = 200 W. The stroke location was equidistant from
two sets of surge arresters.
Figure 9.25 presents a configuration more representative of urban distribution
networks, in which the main feeder has various laterals. The line was three-phase
120
90
Voltage (kV)
40
Current (kA)
60 Calculated
20
30
Measured
0 0
0 5 10 15 0 5 10 15
Time (μs) Time (μs)
(a) (b)
Figure 9.22 Measured and calculated (using the ERM) phase-to-ground induced
voltages at the point closest to the stroke location for the line
configuration shown in Figure 9.23a (adapted from [41,113]).
I = 36 kA; tf = 3.1 ms, d = 70 m; h = 10 m, hg = 9 m, xg = 450 m,
Rg = 0 W, lightning channel in front of a grounding point.
(a) induced voltages; (b) measured current
354 Lightning electromagnetics: Volume 2
0.75 m 1.5 m
1m
10 m
9m
Rg Rg
(a) (b)
Figure 9.23 Configurations of the Test Line for the comparisons shown in
Figures 9.22 and 9.24 (adapted from [41,113]): (a) with a shield
wire (hg = 9 m); (b) with surge arresters
200
Measured
Voltage (kV)
150 60
Current (kA)
100
Calculated 30
50
0 0
0 5 10 15 0 5 10 15
Time (μs) Time (μs)
(a) (b)
Figure 9.24 Measured and calculated (using the ERM) phase-to-ground induced
voltages at the point closest to the stroke location for the line
configuration shown in Figure 9.23b (adapted from [41,113]). I = 54
kA; tf = 3.2 ms, d = 70 m; h = 10 m, xg = 450 m, Rg = 200 W,
lightning channel equidistant from two sets of surge arresters. (a)
induced voltages; (b) measured current
and the heights of the phase and neutral conductors were 10 m and 8 m, respec-
tively. The distance between adjacent phases was 0.75 m, and the main feeder was
matched at both ends. Each transformer was represented by capacitors connected
between each phase and the neutral. However, as the value of the simulated
transformer input capacitance (0.5 nF) was similar to the value of the voltage probe
capacitance (0.65 nF), at the measuring point the capacitor was replaced with the
voltage probe.
The neutral was grounded at every transformer and also at the middle of the
laterals through ground resistances of 50 W. At the measuring point, however, Rg
was zero and, therefore, the neutral-to-ground voltage was related only to the
inductive voltage drop across the down conductor. Owing to the low value of the
Scale models to the study of lightning transients 355
150 m
70 m
150 m
75 m
M1
20 m
Figure 9.25 Network configuration for the comparison shown in Figure 9.26
(adapted from [35,41])
400
Measured
200
Voltage (kV)
0
2 4 6 8
–200
Calculated
–400
Time (μs)
Figure 9.26 Measured and calculated (using the LIOV-EMTP) induced voltages
at point M1 of Figure 9.25 (Case 1). I = 34 kA; tf = 2 ms (adapted
from [35])
down conductor inductance (estimated of the order of ten mH), such voltage was
very small compared to the voltage across the capacitances. Hence, the differences
between phase-to-ground and phase-to-neutral voltages were negligible, as con-
firmed by simulations.
Figure 9.26, which corresponds to Case 1 [35], depicts a comparison between
the measured and calculated induced voltages at node M1 indicated in Figure 9.25
for a stroke current with peak value of 34 kA, front time equal to 2 ms and time to
half-value equal to 85 ms. The measurement and simulation regard the phase con-
ductor closest to the stroke location.
Comparisons were also made for the system configuration presented in
Figure 9.27. The voltages shown in Figure 9.28 (Case 2.1 [35]) refer to a stroke
356 Lightning electromagnetics: Volume 2
150 m
70 m
150 m
75 m
M2
Figure 9.27 Network configuration for the comparison shown in Figure 9.28
(adapted from [35,41]). The meanings of the symbols are the same as
in Figure 9.25
200
Measured
100
Voltage (V)
Calculated
0
2 4 6 8
Figure 9.28 Measured and calculated (using the LIOV-EMTP) induced voltages
at point M2 of Figure 9.27 (Case 2.1 [35]). I = 70 kA; tf = 2 ms.
Adapted from [35]
current with the same previous waveform but with a peak value equal to 70 kA.
The calculation was performed considering a simplified configuration with two
conductors (middle phase and neutral). Despite the higher stroke current mag-
nitude in relation to Case 1 (70 kA against 34 kA), the corresponding induced
voltage peak values are lower (roughly 150 kV against 300 kV), mostly because
the distance between the stroke location and the closest line section (70 m) is
larger than that corresponding to Case 1 (20 m). Moreover, in Case 1 the mea-
suring point is located at the terminal of the lateral closest to the stroke location,
while in Case 2.1, the stroke location is nearest the main feeder. The combi-
nation of the above-mentioned conditions has a significant effect on the induced
voltages.
The results obtained with the ERM and the LIOV-EMTP show a good agree-
ment between measured and calculated induced voltages. Possible reasons for
Scale models to the study of lightning transients 357
Figure 9.29 presents one of the test configurations used in [38], which con-
sisted of two lines placed at a distance of 70 m from the stroke location. The shorter
line was single-phase, straight, and with both ends matched, while the other was
four-conductor and branched. The main feeder of the branched line was matched at
both ends; the laterals were either matched or open-ended. In this configuration, the
lines had neither arresters nor transformers. The value of the ground resistances was
50 W except at the measuring point, where it was zero. The voltages shown in
Figure 9.30 illustrate the influence of the line laterals on the induced voltages for
Additional
laterals
150 m
M1
70 m
150 m
70 m
M2
128 m 39 m 104 m
Figure 9.29 Test configuration for the analysis of the influence of the line laterals
on lightning-induced voltages (adapted from [38]). The meanings of
the symbols are the same as in Figure 9.25
200 200
1 5
160 160 4
2
120 120
Voltage (kV)
Voltage (kV)
80
3 80
40
40
0
5 10 15 20
0
–40 0 5 10 15 20 25
Time (μs) Time (μs)
(a) (b)
the case of a stroke current with amplitude of 46 kA, front time of 2 ms, and time to
half-value of 85 ms.
The induced surges are affected not only by the presence of the laterals but also
by their termination conditions. The voltage magnitudes tend to decrease either in
the case of long (or matched) laterals or when they are terminated with surge
arresters. The shorter the distance between the measuring point and the nearest
lateral, the more significant the voltage reduction, as can be readily seen in
Figure 9.30a. On the other hand, if the laterals are open-ended, the induced voltage
may increase as the distance between the measuring point and the nearest lateral
becomes shorter, as shown in Figure 9.30b.
The voltages corresponding to the case of open-ended laterals tend to be
greater than those relative to long (or matched) laterals; the difference tends to
increase as the observation point approaches a line branch. For instance, although
notable differences are observed between the wavetails of voltages 2 (ends of all
laterals matched) and 4 (ends of all laterals open), measured at a distance of 342 m
from the nearest lateral, their peak values are about the same. On the other hand, a
difference of about 30% is observed between the magnitudes of curves 3 (ends of
all laterals matched) and 5 (ends of all laterals open), which were measured at a
shorter distance (132 m) from the nearest lateral.
● Influence of buildings
The presence of nearby structures causes a reduction of the lightning electro-
magnetic field around the overhead distribution line and, as a result, the induced
voltages are affected. Figure 9.31 depicts two test configurations adopted in an
investigation of the impact, on the lightning-induced voltages, of the presence of
buildings of different heights (hb) in the vicinity of the distribution network.
Aluminum structures connected to the ground were used to simulate the buildings,
as illustrated in Figure 9.10 for the case of hb = 15 m. The parameters se and sd
indicated in Figure 9.31 represent the distances between the measuring point (M)
and the closest set of surge arresters located on its left and right sides, respectively.
The same distance between the line and the stroke location, equal to 20 m, was
adopted in all the tests, although, especially for the case of hb = 0 m, a lightning
stroke so close to the line is an unlikely event, unless it hits an elevated object like a
tree, a mast or another structure protruding above the line itself.
Figure 9.32 shows the measured induced voltages at the transformer of the
main feeder located in front of the lightning strike point for the three buildings’
heights considered, namely 0 m, 5 m, and 15 m. The stroke current magnitude was
34 kA and, in order to illustrate the effects of the presence of surge arresters, two
situations were considered regarding the distances se and sd. It can be clearly noted
that the buildings provoke a reduction in the electromagnetic field around the line
and, consequently, the induced voltages may be substantially affected. This
reduction is more significant in the case of higher structures, which provide more
effective shielding of the line against the inducing field.
The influence of the distance between the observation point and the arresters on
the induced voltages tends to decrease as the building height increases. For hb = 0 m
360 Lightning electromagnetics: Volume 2
22
M 150
75
150
Se Sd
(a)
90 210 210 150 150 210 210 170
22
M 150
75
150
Se Sd
(b) 80 80 80
Stroke location 80 40 30
Figure 9.31 Test configurations (top view) [39]. Distance of 20 m between line and
stroke location. Squares and rectangles denote blocks with structures
of different heights (hb). All dimensions in meters. The meanings of the
symbols are the same as in Figure 9.25. (a) hb = 5 m; (b) hb = 15 m
(no buildings around the line), the ratio between the crest values of the induced
voltages for the two cases considered (se = sd = 75 m and se = 148 m, sd = 174 m) is
approximately 0.54. For hb = 5 m, the ratio is 0.57, while for hb = 15 m the ratio is
0.96. This can be explained by the fact that, for hb = 0 m, the difference between
the induced voltages is due only to the different distances between the transformer
and the nearest sets of arresters. This influence is very significant, especially
for discharges close to the line. On the other hand, the amplitude of the total elec-
tromagnetic field near the line diminishes as the buildings’ heights increase as a
consequence of the shielding, which becomes more effective. Therefore, in the test
conditions, for hb = 15 m the induced voltages are more influenced by the presence of
buildings than by the distance between the measuring point and the arresters.
The results shown in Figure 9.32, even that corresponding to the case of
hb = 0 m, differ from those illustrated in Figure 9.26, which was obtained for the
same lightning current. The induced voltage magnitude of Figure 9.26 (around
300 kV) is greater because the measuring point is located at the end of a lateral
where the transformer is not protected with surge arresters.
Scale models to the study of lightning transients 361
1) hb = 0 m 300
120
2) hb = 5 m
1 3) hb = 15 m 1
90 200
2
2
Voltage (kV)
Voltage (kV)
60 100
3 3
30 0
2 4 6 8 10
0 –100
2 4 6 8 10
Figure 9.32 Measured induced voltages for different buildings´ heights (test
configurations indicated in Figure 9.31) and different distances
between the observation point and the closest set of arresters
(adapted from [39]). I = 34 kA; tf = 2 ms. (a) se = 75 m; sd = 75 m;
(b) se = 148 m; sd = 174 m
22
150
75
150
80 80
M: measuring point 80
80 40 30
Figure 9.33 Test configuration (top view) for hb = 15 m and sr = 150 m (adapted
from [39]). Distances of the stroke location to the main feeder and
the closest lateral are 70 m and 20 m, respectively. Squares and
rectangles denote blocks with structures 15 m high. All dimensions in
meters. The meanings of the symbols are the same as in Figure 9.25
362 Lightning electromagnetics: Volume 2
arresters (sr), but tests were also performed for the cases of hb = 0 m, hb = 5 m, and
sr = 75 m. Figure 9.34 presents the induced voltages for a stroke current magnitude
of 50 kA and the various test conditions.
The voltage magnitudes diminish as the buildings’ height increases, and this
effect becomes more evident as the distance sr increases. For sr = 150 m the ratio
between the crest values of the voltages relative to hb = 15 m and hb = 0 m (U15/U0)
is approximately 0.48, whereas in the case corresponding to sr = 75 m the ratio is
about 0.70. If sr = 0 m, the phase-to-ground voltages at the transformer terminals
will be given by the sum of the arrester residual voltages and the voltage drop on
the grounding conductor. As the dependence of the arrester residual voltage upon
the buildings’ height is very little, the ratio U15/U0 will be close to unity. Thus, the
ratio diminishes (i.e., the influence of the buildings increase) as sr increases.
As shown in Figures 9.32 and 9.34, the presence of buildings close to dis-
tribution networks reduces the lightning electromagnetic field and, consequently,
the induced voltages. On the other hand, high structures may attract lightning fla-
shes close to the line and, therefore, voltages of large amplitudes can be induced.
The induced voltages strongly depend on the stroke current magnitude, and
this is illustrated in Figure 9.35, which presents the measured voltages for stroke
current magnitudes of 34 kA and 70 kA and the test configuration shown in
Figure 9.33, but without the presence of buildings around the line. The variation of
the peak voltages was almost linear in the test conditions, a behavior that can be
explained by the relatively high value of the ground resistance (50 W), the short
front time of the stroke current (2 ms), and the distances sr (75 m and 150 m)
between the surge arresters and the measuring point. A decrease in the value of the
ground resistance or in the distance sr would tend to increase the system
1) hb = 0 m
600 160
1 2) hb = 5 m 1
3) hb = 15 m
300 2
3 2
80
Voltage (kV)
Voltage (kV)
3
0 4 8 12
0 4 8 12
–300
Figure 9.34 Measured induced voltages for different buildings’ heights and
different distances sr between the observation point and the closest
set of arresters (Figure 9.33 corresponds to the case of hb = 15 m and
sr = 150 m). Adapted from [39]. I = 50 kA; tf = 2 ms. (a) sr = 150 m;
(b) sr = 75 m
Scale models to the study of lightning transients 363
800 200
1) I = 70 kA
1 1
2) I = 34 kA
400
2
Voltage (kV)
Voltage (kV)
100
2
0
4 8 12 16
0
4 8 12 16
–400
Figure 9.35 Measured induced voltages for different stroke current peak values
and different distances sr between the observation point and the closest
set of arresters (test configuration indicated in Figure 9.33, but with
hb = 0 m). Adapted from [39]. tf = 2 ms. (a) sr = 150 m; (b) sr = 75 m
9.5 Conclusions
The scale model technique is a very powerful and versatile tool for the analysis of
the interaction of lightning with electric power lines, and well complements other
methods such as rocket-triggered lightning and experiments with full-scale sys-
tems. It enables the simulation of a wide variety of situations and, moreover, tests
can be carried out under controlled conditions. After the system implementation, a
substantial amount of data can be obtained in a relatively short time.
An important application of scale models concerns the validation of theoretical
models of complex phenomena and their relevant codes. They can also be very
364 Lightning electromagnetics: Volume 2
useful in the evaluation of the influence of the line configuration and of various
lightning parameters on the overvoltages’ magnitudes and waveforms, which can
be assessed with satisfactory accuracy. In this chapter, the usefulness of the method
was illustrated by its application for the validation of the ERM and LIOV-EMTP
predictions, as well as for the investigation of the behavior of lightning transients
on overhead power transmission and distribution lines subjected to direct and
indirect strokes. The technique is particularly suitable for the analysis of situations
that are either too complex or not worthwhile to be treated theoretically, as e.g. the
case of lightning-induced voltages on urban power distribution networks sur-
rounded by nearby buildings.
Acknowledgments
Special thanks are due to Prof. C.A. Nucci, Prof. A. Borghetti, and Dr. M. Paolone
for the valuable discussions and for providing the simulation results obtained with
the LIOV-EMTP code.
References
[1] Piantini, A., De Carvalho, T. O., Silva Neto, A., Janiszewski, J. M.,
Altafim, R. A. C., and Nogueira, A. L. T., ‘A system for simultaneous
measurements of lightning-induced voltages on lines with and without
arresters’, In Proceedings of the 27th International Conference on
Lightning Protection (ICLP), Avignon, 1, September 2004, pp. 297–302.
[2] Michishita, K., Ishii, M., Asakawa, A., Yokoyama, S., and Kami, K.,
‘Voltage induced on a test distribution line by negative winter lightning
strokes to a tall structure’, IEEE Transactions on Electromagnetic
Compatibility, 2003, 45, (1), pp. 135–140.
[3] Fernandez, M. I., Rakov, V. A., and Uman, M. A., ‘Transient currents and
voltages in a power distribution system due to natural lightning’, In
Proceedings of the 24th International Conference on Lightning Protection
(ICLP), Birmingham, September 1998, pp. 622–629.
[4] De La Rosa, F., Pérez, H., and Galván, A., ‘Lightning-induced voltage
measurements in an experimental power distribution line in Mexico’, In
Proceedings of the 22nd International Conference on Lightning Protection
(ICLP), Budapest, Sep. 1994, pp. R 6b-09/1-6.
[5] Georgiadis, N., Rubinstein, M., Uman, M. A., Medelius, P. J., and
Thomson, E. M., ‘Lightning-induced voltages at both ends of a 448 m
power-distribution line’, IEEE Transactions on Electromagnetic
Compatibility, 1992, 34, (4), pp. 451–460.
[6] Yokoyama, S., Miyake, K., and Fukui, S., ‘Advanced observations of
lightning induced voltage on power distribution lines (II)’, IEEE
Transactions on Power Delivery, 1989, 4, (4), pp. 2196–2203.
Scale models to the study of lightning transients 365
[7] Rubinstein, M., Tzeng, A. Y., Uman, M. A., Medelius, P. J., and Thomson,
E. M., ‘An experimental test of a theory of lightning-induced voltages on an
overhead wire’, IEEE Transactions on Electromagnetic Compatibility,
1989, 31, (4), pp. 376–383.
[8] Yokoyama, S., Miyake, K., Mitani, H., and Yamazaki, N., ‘Advanced
observations of lightning induced voltage on power distribution lines’,
IEEE Transactions on Power Delivery, 1986, 1, (2), pp. 129–139.
[9] Cooray, V. and De La Rosa, F., ‘Shapes and amplitudes of the initial peaks
of lightning-induced voltage in power lines over finitely conducting earth:
theory and comparison with experiment’, IEEE Transactions on Antennas
and Propagation, 1986, 34, (1), pp. 88–92.
[10] Master, M. J., Uman, M. A., Beasley, W., and Darveniza, M., ‘Lightning
induced voltages on power lines: experiment’, IEEE Transactions on
Power Apparatus and Systems, 1984, 103, (9), pp. 2519–2529.
[11] Eriksson, A. J., Penman, C. L., and Meal, D. V., ‘A review of five years’
lightning research on an 11 kV test – line’, In Proceedings of the
International Conference on Lightning and Power Systems, IEE, London,
1984, pp. 62–66.
[12] Yokoyama, S., Miyake, K., Mitani, H., and Takanishi, A., ‘Simultaneous mea-
surement of lightning induced voltages with associated stroke currents’, IEEE
Transactions on Power Apparatus and Systems, 1983, 102, (8), pp. 2420–2427.
[13] Eriksson, A. J., Stringfellow, M. F., and Meal, D.V., ‘Lightning-induced
overvoltages on overhead distribution lines’, IEEE Transactions on Power
Apparatus and Systems, 1982, 101, (4), pp. 960–968.
[14] Perry, F. R., Webster, G. H., and Baguley, P. W., ‘The measurement of
lightning voltages and currents in Nigeria: part 2, 1938–1939’, Journal of I.
E.E. – Part II: Power Engineering, 1942, 89, (9), pp. 185–203.
[15] Perry, F. R., ‘The measurement of lightning voltages and currents in South
Africa and Nigeria: 1935 to 1937’, Journal of I.E.E. – Part II: Power
Engineering, 1941, 88, (2), pp. 69–87.
[16] DeCarlo, B. A., Rakov, V. A., Jerauld, J., et al., ‘Triggered-lightning test-
ing of the protective system of a residential building: 2004 and 2005
results’, In Proceedings of the 28th International Conference on Lightning
Protection (ICLP), Kanazawa, September 2006, pp. 628–633.
[17] Bejleri, M., Rakov, V. A., Uman, M. A., Rambo, K. J., Mata, C. T., and
Fernandez, M. I., ‘Triggered lightning testing of an airport runway lighting
system’, IEEE Transactions on Electromagnetic Compatibility, 2004,
46, (1), pp. 96–101.
[18] Rakov, V. A. and Uman, M. A., ‘Artificial initiation (triggering) of light-
ning by ground-based activity’, In Lightning: Physics and Effects
(Cambridge University Press, Cambridge, 2003), Chapter 7, pp. 265–307.
[19] Rakov, V. A., Uman, M. A., Fernandez, M. I., et al., ‘Direct lightning
strikes to the lightning protective system of a residential building:
triggered-lightning experiments’, IEEE Transactions on Power Delivery,
2002, 17, (2), pp. 575–586.
366 Lightning electromagnetics: Volume 2
[59] Rachidi, F., Nucci, C. A., Ianoz, M., and Mazzetti, C., ‘Influence of a
lossy ground on lightning-induced voltages on overhead lines’,
IEEE Transactions on Electromagnetic Compatibility, 1996, 38, (3),
pp. 250–264.
[60] Cooray, V., ‘On the validity of several approximate theories used in
quantifying the propagation effects on lightning generated electromagnetic
fields’, In Proceedings of the 8th International Symposium on Lightning
Protection (SIPDA), São Paulo, November 2005, pp. 112–119.
[61] Cooray, V., ‘Propagation effects due to finitely conducting ground on
lightning generated magnetic fields evaluated using Sommerfeld’s inte-
grals’, In Proceedings of the 9th International Symposium on Lightning
Protection (SIPDA), Foz do Iguaçu, November 2007, pp. 151–150.
[62] Nucci, C. A., ‘Lightning-induced voltages on distribution systems: influ-
ence of ground resistivity and system topology’, In Proceedings of the
8th International Symposium on Lightning Protection (SIPDA), São Paulo,
November 2005, pp. 761–773.
[63] Nucci, C. A., Rachidi, F., Ianoz, M., and Mazzetti, C., ‘Lightning-induced
overvoltages on overhead lines’, IEEE Transactions on Electromagnetic
Compatibility, 1993, 35, (1), pp. 75–86.
[64] Romero, F. and Piantini, A., ‘Evaluation of lightning horizontal electric
fields over a finitely conducting ground’, In Proceedings of the 9th
International Symposium on Lightning Protection (SIPDA), Foz do Iguaçu,
November 2007, pp. 145–150.
[65] Barbosa, C. F. and Paulino, J. O. S., ‘An approximate time-domain formula
for the calculation of the horizontal electric field from lightning’,
IEEE Transactions on Electromagnetic Compatibility, 2007, 49, (3),
pp. 593–601.
[66] Shoory, A., Moini, R., Sadeghi, S. H. H., and Rakov, V. A., ‘Analysis of
lightning-radiated electromagnetic fields in the vicinity of lossy ground’,
IEEE Transactions on Electromagnetic Compatibility, 2005, 47, (1),
pp. 131–145.
[67] Nucci, C. A. and Rachidi, F., ‘Interaction of electromagnetic fields with
electrical networks generated by lightning’, In V. Cooray (ed.), The
Lightning Flash (IEE Power Engineering Series, London, 34, 2003),
Chapter 8, pp. 425–478.
[68] Ishii, M., Michishita, K., Hongo, Y., and Ogume, S., ‘Lightning-induced
voltage on an overhead wire dependent on ground conductivity’, IEEE
Transactions on Power Delivery, 1994, 9, (1), pp. 109–118.
[69] Rachidi, F., Nucci, C. A., and Ianoz, M., ‘Transient analysis of multi-
conductor lines above a lossy ground’, IEEE Transactions on Power
Delivery, 1999, 14, (1), pp. 294–302.
[70] Borghetti, A. and Nucci, C. A., ‘Frequency distribution of lightning-
induced voltages on an overhead line above a lossy ground’, In
Proceedings of the 5th International Symposium on Lightning Protection
(SIPDA), São Paulo, May 1999, pp. 229–233.
370 Lightning electromagnetics: Volume 2
[71] Hoidalen, H. K., Sletbak, J., and Henriksen, T., ‘Ground effects on induced
voltages from nearby lightning’, IEEE Transactions on Electromagnetic
Compatibility, 1997, pp. 269–278.
[72] Nucci, C. A. and Rachidi, F., ‘Lightning protection of medium voltage
lines’, In V. Cooray (ed.), Lightning Protection (IET Power and Energy
Series, London, 58, 2010), Chapter 13 pp. 635–680.
[73] Paolone, M., Nucci, C. A., Petrache, E., and Rachidi, F., ‘Mitigation of
lightning-induced overvoltages in medium voltage distribution lines by
means of periodical grounding of shielding wires and of surge arresters:
modeling and experimental validation’, IEEE Transactions on Power
Delivery, 2004, 19, (1), pp. 423–431.
[74] Yokoyama, S., ‘Lightning protection of MV overhead distribution lines’, in
Proceedings of the 7th International Symposium on Lightning Protection
(SIPDA), São Paulo, Novemebr 2003, pp. 485–507.
[75] Nucci, C. A., (CIGRE WG C4.01). ‘Lightning-induced voltages on overhead
power lines. Part III: sensitivity analysis’, Elektra, October 2005, pp. 27–30.
[76] Piantini, A., ‘Lightning protection of low-voltage networks’, In V. Cooray
(ed.), Lightning Protection (IET Power and Energy Series, London, 58,
2010), Chapter 12, pp. 553–634.
[77] Lat, M. V. and Carr, J.: Application Guide for Surge Arresters on
Distribution Systems, (Canadian Electrical Association, Toronto, 1988).
[78] Yokoyama, S., ‘Distribution surge arrester behavior due to lightning
induced voltages’, IEEE Transactions on Power Delivery, 1986, 1, (1),
pp. 171–178.
[79] Borghetti, A., Gutierrez, A., Nucci, C. A., Paolone, M., Petrache, E., and
Rachidi F., ‘Lightning-induced voltages on complex distribution systems:
models, advanced software tools and experimental validation’, Journal of
Electrostatics, 2004, 60, pp. 163–174.
[80] Nucci, C. A., Bardazzi, V., Iorio, R., Mansoldo, A., and Porrino, A., ‘A
code for the calculation of lightning-induced overvoltages and its interface
with the Electromagnetic Transient program’, In Proceedings of the 22nd
International Conference on Lightning Protection (ICLP), Budapest,
September 1994, pp. 19–23.
[81] Borghetti, A., Napolitano, F., Nucci, C. A, and Paolone, M., ‘Effects of
nearby buildings on lightning induced voltages on overhead power dis-
tribution lines’, Electric Power Systems Research, 2013, 94, pp. 38–45.
[82] Tossani, F., Borghetti, A., Napolitano, F., Piantini, A., and Nucci, C. A.,
‘Lightning performance of overhead power distribution lines in urban
areas’, IEEE Transactions on Power Delivery, 2018, 33, (2), pp. 581–588.
[83] Chisholm, W. A., Chow, Y. L., and Srivastava, K. D., ‘Lightning surge
response of transmission towers’, IEEE Transactions on Power Apparatus
and Systems, 1983, PAS-102, (9), pp. 3232–3242.
[84] Yamada, T., Mochizuki, A., Sawada, J., et al., ‘Experimental evaluation of
a UHV tower model for lightning surge analysis’, IEEE Transactions on
Power Delivery, 1995, 10, (1), pp. 393–402.
Scale models to the study of lightning transients 371
[110] Nucci, C. A., Rachidi, F., Ianoz, M., and Mazzetti, C., ‘Comparison of two
coupling models for lightning-induced overvoltage calculations’, IEEE
Transactions on Power Delivery, 1995, 10, (1), pp. 330–339.
[111] Cooray, V., ‘Calculating lightning-induced overvoltages in power lines: a
comparison of two coupling models’, IEEE Transactions on
Electromagnetic Compatibility, 1994, 36, (3), pp. 179–182.
[112] Uman, M. A. and Mclain, D. K., ‘Magnetic field of the lightning return
stroke’, Journal of Geophysical Research, 1969, 74, pp. 6899–6910.
[113] Piantini A. and Janiszewski, J. M., ‘Lightning-induced voltages on over-
head lines – application of the extended Rusck Model’, IEEE Transactions
on Electromagnetic Compatibility, 2009, 51, (3), pp. 548–558.
[114] Paolone, M., Rachidi, F., Borghetti, A., et al., ‘Lightning electromagnetic
field coupling to overhead lines: theory, numerical simulations and
experimental validation’, IEEE Transactions on Electromagnetic
Compatibility, 2009, 51, (3), pp. 532–547.
[115] Piantini, A., Janiszewski, J. M., and Braz, C. P., ‘Utilization of reduced
models for the analysis of lightning induced overvoltages on overhead
lines’, In Proceedings of the International Conference on High Voltage
Engineering and Application, New Orleans, October 2010.
[116] Thang, T. H., Baba, Y., Piantini, A., and Rakov, V. A., ‘Lightning-induced
voltages in the presence of nearby buildings: FDTD simulation versus
small-scale experiment’, IEEE Transactions on Electromagnetic
Compatibility, 2015, 57, (6), pp. 1601–1607.
This page intentionally left blank
Chapter 10
Lightning interaction with the ionosphere
Dongshuai Li1, Alejandro Luque1, Marcos Rubinstein2
and Farhad Rachidi3
10.1 Introduction
1
Instituto de Astrofı́sica de Andalucı́a (IAA), Consejo Superior de Investigaciones Cientı́ficas (CSIC),
Spain
2
School of Engineering and Management Vaud, HES-SO University of Applied Sciences and Arts
Western Switzerland
3
Ecole Polytechnique Fédérale de Lausanne (EPFL), Switzerland
376 Lightning electromagnetics: Volume 2
thousands of kilometers [31]. The main parameters shown in (10.3) that char-
acterize the propagation of the lightning EM fields in the lower D-region iono-
sphere are the electron number density
!
Ne , the electron-neutral collision frequency
nen , and the Earth’s magnetic field B 0 , which strongly depends on the solar activity,
and that varies with the time and the geographic latitude [32].
There are several different approaches in the literature to obtain the electric
number density Ne and the electron-neutral collision frequency nen of the iono-
sphere. The simplest way is to use the following empirical formulation proposed in
[33]. According to [33], the profile of the electron number density in the undis-
turbed lower ionosphere can be approximately described by an exponential func-
tion of height based on a number of independent experimental approaches.
The exponential profile of electron number density Ne is:
Ne ¼ n0 ebðzHref Þ ; (10.4)
where b is the steepness of the profile and n0 ¼ 3 107 m3 is the electron density
where most VLF energy is reflected at the effective reflected height Href .
Figure 10.1(a) gives the electron number density Ne based on the exponential
profile (Href ¼ 83 km; b ¼ 0:35 km1 ) given by [33] with a comparison of this
profile with others in the literature. It can be seen that the exponential profile of
electron number density proposed by [33] gives a reasonable approximation for the
90 90
Wait and Spies [1964] profile (used here)
85 85
80 80
75 75
Altitude (km)
Altitude (km)
70 70
65 65
60 60
IRI lonosphere
Wait and Spies [1964] profile (used here)
55 Pasko et al. [1997] profile 1 55
Pasko et al. [1997] profile 2
Pasko et al. [1997] profile 3
Hu et al. [2007]
50 50
104 105 106 107 108 109 1010 105 106 107 108
Electron number density (m–3) Electron-neutral collision frequency (s–1)
(a) ( b)
Figure 10.1 (a) Electron number density based on the exponential profile given
by [33] with a comparison of this profile with others in the literature:
International Reference Ionosphere (IRI) 2007 model [34]; Pasko
et al. [35], Hu et al. [36], and (b) The exponential profile of the
electron-neutral collision frequency proposed by [33]
378 Lightning electromagnetics: Volume 2
θ
Er
Eφ
Eφ
Hθ
Er
RE Hr Eθ
r Eθ Jr
Earth Hφ Jφ
Jθ
Eθ Eθ Er
φ Eφ Eφ
Er
Figure 10.2 The grid of the 3D spherical FDTD model (left panel) and a unit cell
(right panel)
Lightning interaction with the ionosphere 379
nþ1=2 nþ1=2 #
Hq jiþ1=2;j;kþ1=2 Hq jiþ1=2;j;k1=2 nþ1=2 nþ1=2
ðJr jiþ1;j;k þ Jr ji;j;k Þ
Cb ;
riþ1=2 sin qj Df 2
(10.9)
nþ1
Eq ji;jþ1=2;k ¼ Ca Eq jni;jþ1=2;k
" nþ1=2 nþ1=2
Hr ji;jþ1=2;kþ1=2 Hr ji;jþ1=2;k1=2
þCb
ri sin qjþ1=2 Df
nþ1=2 nþ1=2 #
riþ1=2 Hf jiþ1=2;jþ1=2;k ri1=2 Hf ji1=2;jþ1=2;k
ri Dr
nþ1=2 nþ1=2
ðJq ji;jþ1;k þ Jq ji;j;k Þ
Cb ; (10.10)
2
nþ1
Ef ji;j;kþ1=2 ¼ Ca Ef jni;j;kþ1=2
" nþ1=2 nþ1=2
riþ1=2 Hq jiþ1=2;j;kþ1=2 ri1=2 Hq ji1=2;j;kþ1=2
þCb
ri Dr
nþ1=2 nþ1=2 #
Hr ji;jþ1=2;kþ1=2 Hr ji;j1=2;kþ1=2
ri Dq
nþ1=2 nþ1=2
ðJf ji;j;kþ1 þ Jf ji;j;k Þ
Cb ; (10.11)
2
380 Lightning electromagnetics: Volume 2
where the coefficients at the general E-field component location ði; j; kÞ are:
2ei;j;k si;j;k Dt
Ca ji;j;k ¼ ; (10.12)
2ei;j;k þ si;j;k Dt
2Dt
Cb ji;j;k ¼ ; (10.13)
2ei;j;k þ si;j;k Dt
s and e ¼ e0 er are the electric conductivity and the permittivity of the medium at
the location ði; j; kÞ, where e0 is the free-space permittivity and er is the relative
permittivity.
The field updating equations for the three magnetic field components (Hr , Hq ,
Hf ) are given as follows:
nþ1=2 n1=2
Hr ji;jþ1=2;kþ1=2 ¼ Da Hr ji;jþ1=2;kþ1=2
"
Eq jni;jþ1=2;kþ1 Eq jni;jþ1=2;k
þDb
ri sin qjþ1=2 Df
2Dt
Db ji;j;k ¼ ; (10.18)
2mi;j;k þ si;j;k Dt
Figure 10.3 further shows the grid of the 2D symmetric polar FDTD model,
where the axis of symmetry is located at the edge of i ¼ 1. The field components in
2D polar coordinates are given by:
! ! !
E ¼ Er r þ Eq q; (10.22)
! !
H ¼ Hf f; (10.23)
! ! !
J ¼ Jr r þ Jq q: (10.24)
As shown in Figure 10.3, Dr and Dq are the spatial step in the direction r and q
of 2D polar coordinates, respectively. RE is the radius of the Earth. The update
382 Lightning electromagnetics: Volume 2
i=1
r Axis of symmetry
(i, j+1)
∆r
Er
Eθ
J
r
Hφ
J
θ
θ (i, j)
Earth (i+1, j+1)
Eθ
Er
RE
(i+1, j)
∆θ
Figure 10.3 The grid of the 2D symmetric polar FDTD model (left panel) and a
unit cell (right panel). The axis of symmetry is located at the edge of
i¼1
i=1
Axis of symmetry
S = R02
Hφ R0
Er
Hφ θ
C = 2R0
Figure 10.4 Update Er on the z-axis by applying Ampère’s law. The marked blue
surface enclosed by the curve C with a radius R0 and area S
H dl ¼ e0 @@tE ds (10.28)
C S
Applying (10.28) to the field Hf j1þ1=2;jþ1=2 by assuming that Hf does not vary
along a small circle around the z-axis with radius R0 ¼ ðrjþ1=2 Dq Þ=2, we obtain:
!
!
H f j1þ1=2;jþ1=2 ð2pR0 Þ ¼ e0 @@tE r j1;jþ1=2 ðpR20 Þ
nþ1=2 nþ1=2
(10.29)
Previous studies indicate that the amplitude and phase perturbation for VLF/LF
signals of lightning EM fields in the EIWG have a complicated relationship with
many parameters. In the following sections, with the help of the FDTD models
proposed in Section 10.2, we will analyze how different factors affect the lightning
EM fields propagation through the Earth-ionosphere waveguide, including the
Earth’s curvature, the presence of the ground conductivity, the different iono-
spheric profiles in D-region ionosphere and the presence of the Earth’s
magnetic field.
where the peak current is I0 , the rise time is determined by t1 and the fall time is
determined by t2 , and h ¼ ðt21 =t22 1Þðt21 =t22 Þt1 =ðt1 t2 Þ is the amplitude correction
2 2 2
factor. Normalized lightning currents and their frequency spectrum are presented in
Figure 10.5 with the parameters I0 ¼ 1 (normalized with respect to the maximum
amplitude values), t1 ¼ 10 ms 50 ms, t2 ¼ 60 ms 100 ms chosen in the range
given by [13,46]. It can be seen that the currents with shorter rise time t1 contains
more high-frequency components. In contrast, longer fall time t2 includes more
low-frequency components.
Lightning interaction with the ionosphere 385
1
W μs, τ2= 100 μs
different W, τ2= 100 μs
0.9
W μs, different τ2
Normalized current with respective to the maximum
0.8 50μs
40μs
0.7 30μs
20μs
0.6
0.5
90μs
80μs
0.4
70μs
0.3
60μs
0.2
0.1
0
0 20 40 60 80 100 120 140 160 180 200
(a) Time (μs)
100
W μs, τ2= 100 μs
different W, τ2= 100 μs
80
W μs, different τ2
60
Normalized current Spectrum (dB)
40
20
–20
–40
–60
–80
–100
10–3 10–2 10–1 100
(b) Frequency (MHz)
Figure 10.5 The lightning current waveforms and their frequency spectrums
386 Lightning electromagnetics: Volume 2
For the ionosphere medium, we consider the electron number density Ne and the
electron-neutral collision frequency nen by following the typical exponential profiles
proposed in [33] (see (10.4) and (10.5) in Section 10.2.1 for details). The channel-
base current is based on (10.31) with I0 ¼ 30 kA, t1 ¼ 10 ms and t2 ¼ 100 ms (see
the green line in Figure 10.5). Note that the lightning channel-base current mentioned
here will be used throughout this chapter unless specifically stated.
The electric fields over a perfectly conducting ground calculated at distances
ranging from 100 km to 800 km are shown in Figure 10.6. It can be seen in the
figure that the lightning electric field waveforms in the EIWG contain a first peak
that originates from the lightning source current, known as the groudwave (marked
as “G” in Figure 10.6), followed by subsequent peaks due to the reflections from
the ionosphere, known as the reflected sky waves (marked as “S” in Figure 10.6).
As expected, when the distances increases, the reflected skywaves occur earlier in
time with respect to the groundwaves.
It can be seen in Figure 10.6 that both the amplitudes and time delays of the
groundwaves and skywaves can be affected by the presence of the Earth curvature.
Errors due to neglecting the Earth’s curvature at distances of 300 km or less can be
neglected. However, the effect of the Earth curvature becomes more significant as
the observation distances increases.
d = 100 km
2 d = 200 km
G
E field (V/m)
E field (V/m)
G
2
S
S 0
0
–2 –2
0 200 400 600 0 200 400 600
(a) Time (μs) (b) Time (μs)
1 G d = 300 km
1 d = 400 km
G
E field (V/m)
E field (V/m)
S S
0 0
–1 –1
0 200 400 600 0 200 400 600
(c) Time (μs) (d) Time (μs)
G G
E field (V/m)
E field (V/m)
0 0
–1 –1
0 200 400 600 0 200 400 600
(e) Time (μs) (f ) Time (μs)
1.0 1.0
without earth curvature without earth curvature
with earth curvature with earth curvature
d = 800 km
0.5 G d = 700 km 0.5 G
E field (V/m)
E field (V/m)
0.0 0.0
S S
0.5 0.5
–1.0 –1.0
0 200 400 600 0 200 400 600
(g) Time (μs) (h) Time (μs)
Figure 10.6 The effect of the Earth curvature on the lightning electric fields over
a perfectly conducting ground at distances ranging from 100 km to
800 km
388 Lightning electromagnetics: Volume 2
2
d = 200 km (σg = v
d = 400 km (σg = v
d = 600 km (σg = v
d = 800 km (σg = v
d = 200 km (σg = S/m
d = 400 km (σg = S/m
d = 600 km (σg = S/m
d = 800 km (σg = S/m
1
E field ( V/m)
–1
0 50 100 150 200 250 300
(a) Time (μs)
2
d = 200 km (σg = v
d = 400 km (σg = v
d = 600 km (σg = v
d = 800 km (σg = v
d = 200 km (σg = S/m
d = 400 km (σg = S/m
d = 600 km (σg = S/m
d = 800 km (σg = S/m
1
E field ( V/m)
–1
0 50 100 150 200 250 300
(b) Time (μs)
Figure 10.7 The effect of the ground conductivity on the lightning electric fields at
distances ranging from 200 km to 800 km with ground conductivity
(a) 0:01 S=m and (b) 0:001 S=m. The relative permittivity was set to
erg ¼ 10 for both cases
90
Daytime ionospheric profile
Nighttime ionospheric profile
85
80
75
Altitude (km)
70
65
60
55
50
104 105 106 107 108 109 1010
Electron number density (m–3)
Figure 10.8 The daytime and nighttime electron number density profiles in the
D-region ionosphere for the simulation
2
Free space
Daytime profile
Perfectly conducting ground Nighttime profile
G
E field ( V/m) d = 200 km
1
1S
2S
0
2S
0.0
1S
–0.5
0.6
Free space
Perfectly conducting ground Daytime profile
0.4 G Nighttime profile
0.2
E field ( V/m)
d = 600 km
0.0
–0.2
1S
2S
–0.4
–0.6
0 100 200 300 400 500 600 700 800
(c) Time (μs)
0.0
–0.2
1S
2S
–0.4
–0.6
Figure 10.9 Comparison of the lightning electric fields for free space, daytime,
and nighttime ionospheric profiles at distances ranging from 200 km
to 800 km
392 Lightning electromagnetics: Volume 2
further noticed that the waveform of the reflected skywaves in the nighttime is
sharper than in the daytime, because the ionosphere in the nighttime has a relatively
higher reflection rate at high frequencies than in the daytime.
E field ( V/m)
0.5 0.5
0.0 0.0
–0.5 –0.5
–1.0 –1.0
0 100 200 300 400 500 0 100 200 300 400 500
(a) Time (μs) ( b) Time (μs)
1.5 1.5
β= 0.3 km–1 h' = 60 km
β= 0.6 km–1 h' = 70 km
β= 0.9 km–1 h' = 80 km
1.0 1.0 –1
d = 300 km, β = 0.6 km
d = 300 km, h' = 70 km
E field ( V/m)
E field ( V/m)
0.5 0.5
0.0
0.0
–0.5
–0.5
–1.0
0 100 200 300 400 500 0 100 200 300 400 500
(c) Time (μs) (d) Time (μs)
1.5 1.5
β= 0.3 km–1 h' = 60 km
β= 0.6 km–1 h' = 70 km
β= 0.9 km–1 h' = 80 km
1.0 1.0 –1
d = 300 km, β = 0.9 km
d = 300 km, h' = 80 km
E field ( V/m)
E field ( V/m)
0.5
0.5
0.0
0.0
–0.5
–0.5
–1.0
0 100 200 300 400 500 0 100 200 300 400 500
(e) Time (μs) (f ) Time (μs)
nighttime ionospheric profile of (10.32) with b ¼ 0:8 km1 and h0 ¼ 84 km. The
other adopted parameters in this section are the same as in Section 10.3.1. The
channel-base current is based on (10.31) with I0 ¼ 30 kA, t1 ¼ 50 ms and
t2 ¼ 100 ms. To illustrate the azimuth effects of the Earth’s magnetic field, we
consider that the magnetic azimuth angle (the angle between the direction of wave !
propagation of lightning-radiated waves and the Earth’s magnetic field vector B 0 )
varies from the geographic north by following the clockwise direction: North (0 ),
Northeast (45 ), East (90 ), Southeast (135 ), South (180 ), Southwest (225 ), West
394 Lightning electromagnetics: Volume 2
Zenith
West
South North
D X
I
East Y F
B0
Nadir
Figure 10.11 The definition of Earth’s magnetic field B0 with the total intensity
F, inclination angle I and declination angle D. The vertical direction
is perpendicular to the WGS 84 ellipsoid model of the Earth
Table 10.1 The ranges of the total intensity F, the inclination angle I and
declination angle D at the Earth’s surface based on the WMM model
(270 ), and Northwest (315 ). The dip angle of the Earth’s magnetic field is taken
as 20 , 40 , and 60 with intensity 45,000 nT, which are good approximations for
the most parts of the northern hemisphere. Figure 10.12 shows the lightning electric
fields over a perfectly conducting ground calculated at distances ranging from
100 km to 300 km considering different magnetic azimuth angles and dip angles. It
is noted that the lightning electric fields can be significantly affected by the pre-
sence of the Earth’s magnetic field. Since the model is symmetric in the f-axis
direction at East (90 ) and West (270 ), the results corresponding to North (0 ) and
South (180 ), Northeast (45 ), and Southeast (135 ), as well as Southwest (225 )
and Northwest (315 ), are coincident. It can be further seen that the effect of the
Lightning interaction with the ionosphere 395
3
90˚ 270˚ 90˚ 270˚
d = 100 km, dip angle = 20˚ d = 300 km, dip angle = 20˚ 0˚ 180˚
0˚ 180˚
45˚ 135˚
1.0 45˚ 135˚
225˚ 315˚ 225˚ 315˚
2
E field (V/m)
E field (V/m)
0.5
1
0.0
0
–1 –0.5
3
90˚ 270˚ d = 300 km, dip angle = 40˚ 90˚ 270˚
d = 100 km, dip angle = 40˚
0˚ 180˚ 0˚ 180˚
45˚ 135˚ 1.0 45˚ 135˚
225˚ 315˚ 225˚ 315˚
2
E field ( V/m)
0 0.0
–1 –0.5
3
d = 100 km, dip angle = 60˚ 90˚ 270˚ d = 300 km, dip angle = 60˚ 90˚ 270˚
0˚ 180˚ 0˚ 180˚
45˚ 135˚ 1.0 45˚ 135˚
225˚ 315˚ 225˚ 315˚
2
E field ( V/m)
E field ( V/m)
0.5
1
0.0
0
–1 –0.5
Earth’s magnetic field becomes larger as the dip angle increases. The effect of the
Earth’s magnetic field is greatest for propagation toward the East (90 ) and lowest
for propagation toward the West (270 ). Moreover, the effect of the Earth’s mag-
netic field increases as the distance increases.
According to (10.3), the effect! of the Earth’s magnetic field is determined
by the ratio of the gyrofrequency w c;e and the electron-neutral collision rate nen .
When wc;e =nen 1, the ionospheric plasma is too collisional to get magnetized,
396 Lightning electromagnetics: Volume 2
electrons cannot make even a fraction of a gyration before colliding with a neutral
particle. When wc;e =nen 1, the ionospheric plasma is completely magnetized and
the EM !
waves will be significantly affected by the presence of Earth’s magnetic
field B 0 [15,26]. Further measured and frequency-dependent modeling results
about the effect of the Earth’s magnetic field on lightning EM fields can be found
in [27].
Narrow bipolar events (NBEs) (also called narrow bipolar pulses (NBPs) or com-
pact intracloud discharges (CIDs)) are energetic intracloud discharges character-
ized by narrow bipolar electromagnetic waveforms identified from ground-based
VLF/LF observations. Several properties of NBEs stand out when compared to
other lightning processes: they exhibit short time durations (10–20 ms), fast pro-
pagation speeds (107 –108 m/s), strong very high frequency (VHF) radiation and
they typically occur isolated from other electrical activity within a time frame of a
few milliseconds, although in some cases NBEs are precursors of a leader process
inside the thunderstorm [60–64]. These unique properties have turned NBEs into an
active research topic that promises to shed light into the initiation mechanism of
lightning in a thunderstorm [65]. In this section, we will investigate the propagation
of NBEs by using the full-wave FDTD approach. The obtained results are com-
pared with ground-based VLF/LF measurements from different stations in the
Jianghuai Area Sferic Array (JASA) in Southern China. JASA consists of more
than ten low-frequency electric field sensors that continuously record VLF/LF
broadband (between 300 Hz and 300 kHz) lightning signals at a sampling rate of
5MHz with a detection range of up to several thousand kilometers [66].
In the following, we consider a typical positive NBE event captured by JASA
in China at local time 08:57:29 on July 7, 2012. In the calculation domain, as
shown in Figure 10.13, we model the NBE as a vertical and straight channel located
at a height h above the curved earth surface and we consider the current distribution
along the NBE channel according to the modified transmission line model with a
linear decay of current with height (MTLL) [67]. The observation point p is located
at ground level at a distance d from the NBE source. The ground is assumed to be
homogeneous with conductivity sg and relative permittivity er . The ionosphere is
considered to be vertically stratified and horizontally homogeneous in the r direc-
tion shown in Figure 10.13. For the ionosphere medium, we consider the electron
number density Ne and the electron-neutron collision frequency nen by following
the typical exponential profiles proposed in [33] (see (10.4) and (10.5) in
Section 10.2.1 for details). To avoid artificial reflections at the upper, lower and
outer boundaries of the computational domain, we used absorbing convolutional
perfectly matched layers (CPML) [68]. The NBE current inside the discharge
channel starts at an altitude H2 and propagates downwards at a speed u with the
channel length L to reach the final, lower altitude H1 .
Lightning interaction with the ionosphere 397
Axis of symmetry
z
CPM
H2 L
1S
L NBE
H1
2S
h G Ion
osp
d he
re
Gro
und
CPM p
L
L
M
(σ
CP
g,ε
rg )
r
Re
Earth’s Center
Figure 10.13 Geometry of the 2D symmetric polar FDTD model with a ground wave
and two ionospheric reflected sky waves marked as G, 1S, and 2S
In the FDTD simulation, the computational domain is 500 km 110 km, with
spatial steps of 50 m in both the r and z directions. Considering 10 steps per
wavelength to avoid numerical dispersion, the maximum frequency that can be
simulated using our full-wave FDTD model is about 600 kHz, which is enough to
cover the VLF/LF frequency band. The expression of the NBE current is the
double-exponential waveform proposed in [65]:
eat
IðtÞ ¼ I0 ; (10.33)
1 þ eðaþbÞt
where a ¼ 1=t1 and b ¼ 1=t2 are the rise and fall time constants ofathe current, and
the peak current is normalized to one by setting I0 ¼ ð1 þ abÞðab ÞðaþbÞ . The detailed
parameters of the NBE source are based on the VHF observation results from [65]
as shown in Table 10.2.
Figure 10.14 shows a comparison between the simulated and the measured
VLF/LF waveforms at different distances ranging from about 100 km to 400 km.
The VLF/LF waveforms of NBE shown in Figure 10.14 involve a ground wave and
398 Lightning electromagnetics: Volume 2
Table 10.2 Parameters of the NBE source used for the calculation
Parameters t1 t2 L h u sg erg
Values 1 ms 6 ms 300 m 10 km 3.510 m/s7
0.01 S/m 10
Measurement
1.0 FDTD modeling
1.0 Measurement
FDTD modeling
Normalized E field
0.5 d = 169 km 0.5 G d = 245 km
1S 2S 1S 2S
0.0 0.0
–0.5 –0.5
–1.0 –1.0
300 400 500 600 700 300 400 500 600 700
(a) Time (μs) (b) Time (μs)
1.0 Measurement
FDTD modeling
1.0 Measurement
FDTD modeling
G
Normalized E field
0.0 0.0
1S 2S 2S
–0.5 –0.5 1S
–1.0 –1.0
300 400 500 600 700 300 400 500 600 700
(c) Time (μs) (d) Time (μs)
Figure 10.14 Comparison of the normalized electric fields between the FDTD
modeling results and the VLF/LF waveforms measured at different
sensors from JASA: (a) d = 169 km, (b) d = 245 km, (c) d = 340 km,
and (d) d = 413 km. The effective reflected height Href and the
steepness b of the profile are 66 km and 0:5 km1 , respectively. The
ground conductivity is sg ¼ 0:01 S=m and the relative permittivity is
er ¼ 10
two ionospheric reflected sky waves marked as G, 1S, and 2S. Notice that, by using
the appropriate physical parameters for the ground and ionosphere, all the simu-
lated results obtained from the full-wave FDTD model agree perfectly with the
VLF/LF measurements at different distances. Since our measured electric field
signals are uncalibrated and the currents of the NBE source are not available, all the
electric fields in Figure 10.14 are normalized to the maximum amplitude measured
at different sensors from JASA.
Lightning interaction with the ionosphere 399
1.0 σg = f 1.0 σg = f
Normalized E field to the case σ = ∞
0.0 0.0
–0.5 –0.5
420 490 560 630 400 450 500 550 600
(a) Time (μs) (b) Time (μs)
1.0 1.0 σg = f
σg = f
Normalized E field to the case σ = ∞
Normalized E field to the case σ = ∞
0.0 0.0
–0.5 –0.5
400 450 500 550 400 450 500 550
(c) Time (μs) (d) Time (μs)
The propagation effect of lightning EM fields over a rough surface has been inves-
tigated in many studies [42,43,49,50,52,70–74]. It was noted that the waveshape,
peak value, and time delay of the lightning-radiated fields could be strongly affected
when the propagation occurs along a mountainous terrain or a non-smooth ground.
The assumption of a finitely conducting smooth ground might result in a significant
underestimation of the peak of the electric fields in mountainous areas [43,71]. In this
section, we will investigate the propagation effects over a mountainous terrain of
lightning in the Earth-ionosphere waveguide by using a full-wave FDTD model that
includes the effect of the mountainous terrain proposed in [69]. The obtained results
are validated against simultaneous experimental data consisting of lightning currents
measured at the Säntis Tower in Switzerland and electric fields measured in Neudorf,
Austria, located at a 380 km distance from the tower.
Figure 10.16 shows the region around the Säntis Tower in the Swiss Alps. The
124-m Säntis Tower has been instrumented since 2010 and it serves as an experi-
mental station for the direct measurement of lightning currents [75]. We consider
the topography between the Säntis Tower and the 380-km distant electric field
sensor at Neudorf, Austria (see Figure 10.16) based on the global digital elevation
model version 2 (GDEM V2) data [76]. The 2D cross-section of the topographic
(m)
51˚ 4,500
50˚ 4,000
3,500
49˚
3,000
48˚
latitude
2,500
47˚
2,000
46˚
1,500
45˚
1,000
44˚ 500
43˚
4˚ 5˚ 6˚ 7˚ 8˚ 9˚ 10˚ 11˚ 12˚ 13˚ 14˚ 15˚ 16˚
longitude
Figure 10.16 The topographic map of the selected region around the Säntis Tower
(red triangle) and the 380-km distant Neudorf station (black star)
Lightning interaction with the ionosphere 401
(m)
4,000
3,000
3,500
2,500
3,000
Santis
¨
2,000 2,500
Height(m)
2,000
1,500
1,500
1,000
Neudorf 1,000
500 500
0
50 100 150 200 250 300 350
Distance(km)
Figure 10.17 2D cross-section of the topographic profiles along the direct path
between the Säntis Tower and the 380-km distant Neudorf station
(black dash line in Figure 10.16)
profiles along the direct path between the Säntis Tower and the Neudorf E-field
measurement station is given in Figure 10.17.
We adopted a 2D symmetric polar coordinate model for the FDTD simulation
(the geometry of the FDTD model can be found in Figure 10.4 of [24]). The
working space in the r and z directions is 500 km 90 km, which is divided into
square cells Dr Dz ¼ 100 m 100 m. To avoid artificial reflections at the outer
boundaries of the computational domain, we used CPML absorbing boundary [68].
The lightning channel is assumed to be straight and vertical above the ground with
a height H ¼ 8 km. The current distribution along the lightning channel was spe-
cified according to the modified transmission-line model with exponential current
decay (MTLE) [44,45]. In the MTLE model, the temporal and spatial variation of
the return stroke current Iðz; tÞ is given by Iðz; tÞ ¼ Ið0;t z=vÞexpðz=lÞ,
t z=v, where Ið0; tÞ is the current at ground level, v ¼ 1:5 108 m=s is the return
stroke speed, and l ¼ 2 km is a constant describing the current decay with height
[44,45]. The measured return stroke currents obtained from the Säntis Tower are
used directly as the input channel-base current Ið0; tÞ in the MTLE model (see
Figure 10.18(a,c)). The ground is assumed to be homogeneous and lossy with a
ground conductivity sg ¼ 0:01 S=m and relative permittivity erg ¼ 10.
Figure 10.18 shows examples of two measured lightning return stroke currents
(a,c) of an upward flash obtained at the Säntis Tower and the simultaneously
measured electric fields (b,d) at 380-km distance in Neudorf that occurred at UTC
time 17:26:00 on September 25, 2019. It can be seen in Figure 10.18 that the
waveform of the lightning-radiated electric field at 380 km distance includes both a
ground wave (marked as G) and a reflected sky wave (marked as S). For compar-
ison purposes, we consider three different FDTD simulation cases by assuming:
402 Lightning electromagnetics: Volume 2
3
Measurement
FDTD with terrian and Ne profile
0.2 FDTD with flat ground and Ne profile
RS7 FDTD with flat ground
2 G RS7
Current (kA)
Ez (V/m)
1
0.0
S
0 –0.1
0 100 200 300 400 –50 0 50 100 150 200 250 300 350 400
(a) Time (μs) (b) Time (μs)
9 1.0
Measurement
FDTD with terrian and Ne profile
FDTD with flat ground and Ne profile
RS12
FDTD with flat ground
0.5 G RS12
6
Current (kA)
3 0.0
–0.5
0
Figure 10.18 Measured lightning return stroke currents at the Säntis Tower (a, c)
and the comparison between the simultaneous measurement and the
FDTD modeling results at the 380-km distant Neudorf measurement
station (b, d). Black line: measurement data; three cases for FDTD
modeling: (i) red line: irregular terrain with an electron density
profile in the ionosphere, (ii) green line: flat ground with an
electron density profile in the ionosphere, and (iii) blue line: flat
ground with free space
(i) the irregular terrain based on the topographic map shown in Figure 10.17 with
the electron density profile in the ionosphere given by (10.4), (ii) a flat ground with
the electron density profile in the ionosphere given by (10.4), and (iii) a flat ground
with free space. It can be seen that, after taking into account the effect of the
irregular terrain between the Säntis Tower and the field measurement station, and
the electron density profile in the ionosphere, the vertical electric fields calculated
by using the FDTD model are in good agreement with the measurements obtained
from the 380-km sensor in Neudorf for both considered cases. It is further shown
that the waveforms of the lightning-radiated electric fields can be strongly affected
by the presence of both ionospheric cold plasma characteristics and the mountai-
nous terrain. Indeed, the use of either the electron density profile or the terrain
profile is not enough to obtain a complete match between the simulated and the
measured waveforms. The presence of mountainous terrain mainly affects the time
delays and amplitudes of the ground wave parts. However, the skywave parts
Lightning interaction with the ionosphere 403
mostly depend on the effect of the ionospheric cold plasma characteristics. A more
detailed analysis including the effect of the ionospheric cold plasma characteristics,
the effect of the Earth curvature, and the propagation effects over a mountainous
terrain can be found in [10,24]. Note that the measured data at the 380-km distant
Neudorf station in [10,24] was affected by a ringing which was due to a problem in
the analog integrator. The problem was solved recently and the presented data in
this section do not exhibit such ringing any more.
*BOLSIG+ uses a two-term expansion of the electron distribution function in spherical harmonics to
solve the steady-state Boltzmann equation for electrons in weakly ionized gases under uniform electric
fields in the bulk of collisional low-temperature plasmas. More details for BOLSIG+ can be found in the
website https://nl.lxcat.net/solvers/BOLSIG+/.
404 Lightning electromagnetics: Volume 2
101
Obtained from BOLSIG+ (used here)
Pasko et al. [1997]
100
μe N/N0 (m2/V/s)
10–1
10–2
10–4 10–2 100 102
Reduced E field (Td)
attachment and detachment rates to update the number density of electrons and
O ions in the ionosphere [13,82]:
@Ne
¼ ðni na ÞNe þ nd NO (10.34)
@t
@NO
¼ na Ne nd NO (10.35)
@t
where ni , na , and nd are the ionization, attachment, and detachment rates, respectively.
As shown in Figure 10.21, the rates of the ionization, attachment, and detachment are
also calculated based on the previously mentioned Boltzmann solver BOLSIG+ [78]
using the cross section database from [79,80]. All the rates are updated by following the
values of the reduced electric fields during the FDTD calculation. Similar to the electron
mobility me , the rates of ionization, attachment and the detachment are updated self-
consistently with the background electric fields. The procedure for the self-consistent
updating of the EM fields in the FDTD model is the following:
!
● Add the !current! density of the source J s .
● Update E and H at time step tn by following the standard FDTD algorithm.
Lightning interaction with the ionosphere 405
90
Total
N2
O2
85
80
75
Altitude (km)
70
65
60
55
50
1019 1020 1021 1022 1023
Gas number density (m–3)
Figure 10.20 The profile of the gas number density taken from the MSIS-E-90 model
[81] by assuming a mixed air with 78% N2 and 22% O2
1012
Ionization
Attachment
Detachment
1011
1010
νk N0/N (1/s)
109
108
107
106
100 200 300 400 500 600 700 800 900 1,000
Reduced E field (Td)
Figure 10.21 The ionization, attachment and detachment rates as a function of the
reduced electric field, where N0 ¼ 2:688 1025 m3 and N is the
background gas number density
second positive (2PN2 ) band systems of N2 , the LBH band system of N2 and the
first negative band system of N2þ (1NN2þ ). The number of molecules nk in the
excited state k corresponding to each emission band system [83] can be calculated
by:
@nk nk X
¼ nk Ne þ nm Am ; (10.36)
@t tk m
where tk ¼ ðAk þ a1 NN2 þ a2 NO2 Þ1 is the total lifetime of excited state k. a1 and
a2 are the quenching rates due to the collisions between N2 and O2 molecules,
respectively. NPN2 and NO2 are the gas number density of N2 and O2 molecules.
The sum term m nm Am represents the increase of nk due to cascading from higher
states m. Table 10.3 shows the values of Ak , a1 , and a2 used in the four considered
emission band systems [17]. The optical excitation rate nk as a function of the
reduced electric field corresponding to each emission band system is shown in
Figure 10.22, which can be obtained by using BOLSIG+ based on the cross section
database from [79,80].
Lightning interaction with the ionosphere 407
Table 10.3 The values of Ak , a1 , and a2 used in the optical emission band systems
1012
1011
1010
νk N0/N (1/s)
109
108
N2 1P
N2 2P
107
N2 LBH
N+2 1N
106
100 200 300 400 500 600 700 800 900 1,000
Reduced E field (Td)
Figure 10.22 The optical excitation rates as a function of the reduced electric
field for the main four emission band systems of elves
Finally, the photon emission intensity Ik , which is the number of photons per
volume per second (ph=cm3 =s) calculated by using:
! !
Ik ðr; tÞ ¼ Ak nk ðr; tÞ; (10.37)
where the value of Ak is shown in Table 10.3. The emission lines of the nth
vibrational state of the kth excited state into the n0 th vibrational state of the k 0 th
excited state can be calculated by:
Iðk;nÞ!ðk 0 ;n0 Þ ¼ nk qAðk;nÞ!ðk 0 ;n0 Þ ; (10.38)
408 Lightning electromagnetics: Volume 2
where q is the Frank-Condon factor and Aðk;nÞ!ðk 0 ;n0 Þ is the Einstein coefficient from
the nth vibrational state of the kth excited state into the n0 th vibrational state of the
k 0 th excited state [84]. Note that, besides the chemical reactions between the
electron and N2 /O2 molecules, other plasma chemical reactions can also play a role
in the optical emission of TLEs. A more detailed analysis of the full kinetic
chemistry model involved in 136 species interacting through 1,076 chemical
reactions can be found in [85].
Most elves are generated by intense cloud-to-ground (CG) lightnings that radiate
large EMPs directly toward the ionosphere. Both the 2D polar and the 3D spherical
model can be used to investigate the nonlinear effects of the ionospheric cold plasma
characteristics. However, due to the symmetry in 2D model, the effect of the Earth’s
magnetic field needs to be neglected if that model is selected. In order to include the
effect of the Earth’s magnetic field, in the following, we use the 3D spherical coor-
dinate FDTD model. In the calculation, the computation space in the q f r
directions is 800 km 200 km 100 km. The lightning channel is assumed to be
straight and vertical above the ground with a height H ¼ 8 km. The CG lightning
current is assumed to be a Bi-Gaussian function IðtÞ ¼ I0 ðet =t1 et =t2 Þ, where
2 2 2 2
I0 ¼ 100 kA, the rise time is t1 ¼ 30 ms and the fall time is t2 ¼ 100 ms. The current
distribution along the channel is specified according to the MTLE model [44,45].
The initial profile of electron number density Ne is taken from the International
Reference Ionosphere (IRI) 2007 model [see Figure 10.1(a)]. The ground is assumed
to be homogeneous and lossy ground with a ground conductivity sg ¼ 0:001 S=m
and a permittivity erg ¼ 10. CPML are used to avoid artificial reflections at the outer
boundaries of the computational domain [68].
We first focus on the case without the effect of the Earth’s magnetic field.
Figures 10.23 and 10.24 give the side and top views, respectively, of the vertical
electric fields at different time steps. It is noted that the different frequency com-
ponents of the EMPs from CG lightning are reflected at different heights due to
dispersion and attenuation in the ionosphere. The higher frequency components
penetrate deeper than the lower ones and are then rapidly attenuated by the higher
conductivity in the ionosphere [5,24].
The reduced electric fields at different snapshots in time are further shown in
Figures 10.25 and 10.26. The values of the reduced electric fields produced from
CG lightning are strong enough at altitudes between 80 and 90 km to accelerate the
available electrons sufficiently to affect the neutral molecules including heating,
ionization, attachment and optical emissions. In order to obtain the optical emis-
sions produced by the elves, the LBHN2 photon emission intensity of the elve
triggered by CG lightning is calculated and shown in Figures 10.27 and 10.28.
It is seen that the optical emission of the simulated elve expands around
200 km in the radial direction at an altitude of about 85 km, which is consistent
with the observation in the typical donut shape with a hole at the center as a result
of the radiation pattern of the vertical CG lightning channel. According to the
previous studies, the brightness of optical emissions of elves strongly depends on
the ambient electron number density and the peak current of the parent lightning
[13,17,26]. Furthermore, the dip angle of the Earth’s magnetic field B0 , a tilted
Lightning interaction with the ionosphere 409
60 1
0
30 –1
0 –2
Time at 0.64 ms
90 2
Altitude (km)
1
60
0
30 –1
0 –2
Time at 0.76 ms
90 2
Altitude (km)
1
60
0
30 –1
0 –2
–250 –200 –150 –100 –50 0 50 100 150 200 250
Distance (km)
Figure 10.23 Side views of the electric fields at different time steps
30 2
15 1
0 0
–15 –1
–30 –2
Time at 0.51 ms
30 2
15 1
0 0
–15 –1
–30 –2
Time at 0.64 ms
30 2
15 1
0 0
–15 –1
–30 –2
Time at 0.76 ms
30 2
15 1
0 0
–15 –1
–30 –2
–250 –200 –150 –100 –50 0 50 100 150 200 250
Distance (km)
Figure 10.24 Top views of the electric fields at different time steps at a height of
85 km
lightning discharge channel and the local chemical impact may also create sig-
nificant density perturbations in the ionosphere and further affect the shape of the
optical emission of elves [26,86]. Figures 10.29 and 10.30 further indicate the
Time at 0.38 ms Td
Altitude (km) Altitude (km)
90 25
20
70 15
10
5
50 0
Time at 0.51 ms
90 25
20
15
70 10
5
50 0
Time at 0.64 ms
Altitude (km)
90 25
20
15
70 10
5
50 0
Time at 0.76 ms
Altitude (km)
90 25
20
15
70 10
5
50 0
–250 –200 –150 –100 –50 0 50 100 150 200 250
Distance (km)
Figure 10.25 Similar to Figure 10.23, but for the reduced electric fields
Time at 0.38 ms Td
30 25
Distance (km)
20
15
15
0 10
–15 5
–30 0
Time at 0.51 ms
30 25
Distance (km)
15 20
15
0
10
–15 5
–30 0
Time at 0.64 ms
30 25
Distance (km)
15 20
15
0 10
–15 5
–30 0
Time at 0.76 ms
30 25
Distance (km)
20
15
15
0 10
–15 5
–30 0
–250 –200 –150 –100 –50 0 50 100 150 200 250
Distance (km)
Figure 10.26 Similar to Figure 10.24, but for the reduced electric fields
Time at 0.38 ms Ph/cm3/s
Altitude (km) Altitude (km) Altitude (km) Altitude (km)
90 1013
109
70
105
50 101
Time at 0.51 ms
90 1013
109
70 105
50 101
Time at 0.64 ms
90 1013
70 109
105
50 101
Time at 0.76 ms
90 1013
109
70 105
50 101
–250 –200 –150 –100 –50 0 50 100 150 200 250
Distance (km)
Figure 10.27 Similar to Figure 10.23, but for the photon emission intensity of the LBHN2 band
Time at 0.38 ms Ph/cm3/s
Distance (km) 30 1013
15 109
0
105
–15
–30 101
Time at 0.51 ms
30 1013
Distance (km)
15 109
0
105
–15
–30 101
Time at 0.64 ms
30 1013
Distance (km)
15 109
0
105
–15
101
–30
Time at 0.76 ms
30 1013
Distance (km)
15 109
0
105
–15
101
–30
–250 –200 –150 –100 –50 0 50 100 150 200 250
Distance (km)
Figure 10.28 Similar to Figure 10.24, but for the photon emission intensity of the LBHN2 band
Time at 0.38 ms Ph/cm3/s
Altitude (km) Altitude (km) Altitude (km) Altitude (km)
90 1013
109
70
105
50 101
Time at 0.51 ms
90 1013
109
70
105
50 101
Time at 0.64 ms
90 1013
70 109
105
50 101
Time at 0.76 ms
90 1013
109
70
105
50 101
–250 –200 –150 –100 –50 0 50 100 150 200 250
Distance (km)
Figure 10.29 Side view of the reduced electric field for the FDTD model at different time steps with the effect of the Earth’s
magnetic field
Time at 0.38 ms Ph/cm3/s
Distance (km) 30 1013
15 109
0
105
–15
–30 101
Time at 0.51 ms
30 1013
Distance (km)
15 109
0
105
–15
–30 101
Time at 0.64 ms
30 1013
Distance (km)
15 109
0
105
–15
–30 101
Time at 0.76 ms
30 1013
Distance (km)
15 109
0
105
–15
–30 101
–250 –200 –150 –100 –50 0 50 100 150 200 250
Distance (km)
Figure 10.30 Top view of the reduced electric fields at a height of 85km for the FDTD model at different time steps with the effect of
the Earth’s magnetic field
416 Lightning electromagnetics: Volume 2
LBHN2 photon emission intensity of the elve triggered by CG lightning with the
effect of the Earth’s magnetic field. The dip angle of the Earth’s magnetic field is
taken as 40 with an intensity of 45,000 nT, and a magnetic azimuth angle 90 . As
shown in Figures 10.29 and 10.30, the Earth’s magnetic field mainly affects the
higher altitudes above 80km. The presence of the Earth’s magnetic field constrains
electron motion and introduces anisotropy in the conductivity of the ionosphere,
which results in the typical donut shape of elves not being perfectly symmetrical
compared with the case without the effect of the Earth’s magnetic field (see
Figures 10.27 and 10.28). More details about the effect of the Earth’s magnetic
field on elves can be found in [13,26].
Moreover, recent studies indicated that narrow bipolar pulses (NBEs) whose
amplitudes are similar to CG lightning and 10 times larger than the typical
intracloud (IC) lightning [62], are also powerful enough to create elves. However,
since NBEs do not have a connection to the ground, the EMPs that they generate
reflect from both the ionosphere and the ground to generate the so-called “elve
doublets”: the first elve in the doublet is created by the EMP direct path to the
ionosphere and the second elve in the doublet is created by the ground reflection.
The time difference between the two elve peaks depends on the altitude of the
discharge, the altitude of the ionospheric reflection, and the distance to the
observer. More details about elve doublets can be found in [13,26,87]. More
recently, another special intracloud discharge, termed energetic in-cloud pulses
(EIPs), was found to be associated with elves as well [40,88]. EIPs are impulsive
in-cloud discharges with a current moment up to hundreds of kA km and with a
time duration about 10 ms. EIPs can also produce elve doublets similar to pre-
viously mentioned NBEs. According to [89,90], EIPs might have a direct rela-
tionship with terrestrial gamma-ray flashes (TGFs), which make them more
mysterious. The observation of TGF and an associated elve using the
Atmosphere-Space Interactions Monitor (ASIM, https://asdc.space.dtu.dk/) on
the International Space Station can be found in [88]. More optical elves modeling
associated with EIPs can be found in [40]. Atmospheric discharges not only exist
on Earth. They can also be found in other planets of the solar system, such as
Saturn, Jupiter, Uranus, Neptune, and Venus [91,92]. In addition, elve-like
mesospheric optical emissions also can occur in other gaseous giant planets. More
details on the lightning-induced EM pulses propagation on Venus, Jupiter, and
Saturn can be found in [15,86].
10.7 Summary
In this chapter, we first introduced the propagation theory of lightning EMPs
interacting with the ionosphere on the basis of the full-wave FDTD method and
then investigated several factors identified in the literature which could affect the
propagation of lightning EMPs in the EIWG, such as the propagation distance,
the Earth curvature, the ground conductivity, the different ionospheric profiles, and
the presence of the Earth’s magnetic field. Finally, we presented applications
Lightning interaction with the ionosphere 417
References
[1] Rakov VA, Uman MA. Lightning: physics and effects. Cambridge university
press; 2003.
[2] Cummins KL, Murphy MJ, Tuel JV. Lightning detection methods and
meteorological applications. In: IV International Symposium on Military
Meteorology, Malbork, Poland; 2000. p. 26–28.
[3] Jacobson AR, Shao X, Holzworth R. Full-wave reflection of lightning long-
wave radio pulses from the ionospheric D region: Numerical model. Journal
of Geophysical Research: Space Physics. 2009;114(A3).
[4] Jacobson AR, Holzworth RH, Pfaff R, et al. Coordinated satellite observa-
tions of the very low frequency transmission through the ionospheric D layer
at low latitudes, using broadband radio emissions from lightning. Journal of
Geophysical Research: Space Physics. 2018;123(4):2926–2952.
418 Lightning electromagnetics: Volume 2
[5] Qin Z, Chen M, Zhu B, et al. An improved ray theory and transfer matrix
method-based model for lightning electromagnetic pulses propagating in
Earth-ionosphere waveguide and its applications. Journal of Geophysical
Research: Atmospheres. 2017;122(2):712–727.
[6] Shao XM, Jacobson AR. Model simulation of very low-frequency and low-
frequency lightning signal propagation over intermediate ranges. IEEE
Transactions on Electromagnetic Compatibility. 2009;51(3):519–525.
[7] Budden KG. The wave-guide mode theory of wave propagation. Logos
Press; 1961.
[8] Cheng Z, Cummer SA. Broadband VLF measurements of lightning-induced
ionospheric perturbations. Geophysical Research Letters. 2005;32(8).
[9] Cummer SA, Inan US, Bell TF. Ionospheric D region remote sensing using
VLF radio atmospherics. Radio Science. 1998;33(6):1781–1792.
[10] Azadifar M, Li D, Rachidi F, et al. Analysis of lightning-ionosphere inter-
action using simultaneous records of source current and 380 km distant
electric field. Journal of Atmospheric and Solar-Terrestrial Physics.
2017;159:48–56.
[11] Han F, Cummer SA, Li J, et al. Daytime ionospheric D region sharpness
derived from VLF radio atmospherics. Journal of Geophysical Research:
Space Physics. 2011;116(A5).
[12] Hu W, Cummer SA. An FDTD model for low and high altitude lightning-
generated EM fields. IEEE Transactions on Antennas and Propagation.
2006;54(5):1513–1522.
[13] Marshall RA. An improved model of the lightning electromagnetic field
interaction with the D-region ionosphere. Journal of Geophysical Research:
Space Physics. 2012;117(A3).
[14] Tran TH, Baba Y, Somu VB, et al. FDTD modeling of LEMP propagation in
the earth-ionosphere waveguide with emphasis on realistic representation of
lightning source. Journal of Geophysical Research: Atmospheres. 2017;122
(23):12,918–12,937.
[15] Luque A, Dubrovin D, Gordillo-Vázquez FJ, et al. Coupling between
atmospheric layers in gaseous giant planets due to lightning-generated
electromagnetic pulses. Journal of Geophysical Research: Space Physics.
2014;119(10):8705–8720.
[16] Inan US, Marshall RA. Numerical Electromagnetics: The FDTD Method.
Cambridge University Press; 2011.
[17] Kuo CL, Chen A, Lee Y, et al. Modeling elves observed by FORMOSAT-2
satellite. Journal of Geophysical Research: Space Physics. 2007;112(A11).
[18] Lehtinen NG, Inan US. Radiation of ELF/VLF waves by harmonically
varying currents into a stratified ionosphere with application to radiation by
a modulated electrojet. Journal of Geophysical Research: Space Physics.
2008;113(A6).
[19] Lehtinen NG, Inan US. Full-wave modeling of transionospheric propagation
of VLF waves. Geophysical Research Letters. 2009;36(3).
Lightning interaction with the ionosphere 419
[20] Aoki M, Baba Y, Rakov VA. FDTD simulation of LEMP propagation over
lossy ground: Influence of distance, ground conductivity, and source para-
meters. Journal of Geophysical Research: Atmospheres. 2015;120
(16):8043–8051.
[21] Lay EH, Shao X. High temporal and spatial-resolution detection of D-layer
fluctuations by using time-domain lightning waveforms. Journal of
Geophysical Research: Space Physics. 2011;116(A1).
[22] Lay EH, Shao X, Jacobson AR. D region electron profiles observed with
substantial spatial and temporal change near thunderstorms. Journal of
Geophysical Research: Space Physics. 2014;119(6):4916–4928.
[23] Tran TH, Baba Y, Rakov VA. A Study of Earth’s Curvature Effect on LEMP
Propagation Over Distances Ranging from 100 to 1000 km. In: XVI
International Conference on Atmospheric Electricity, Nara, Japan; 2018.
[24] Li D, Luque A, Rachidi F, et al. The Propagation Effects of Lightning
Electromagnetic Fields Over Mountainous Terrain in the Earth-Ionosphere
Waveguide. Journal of Geophysical Research: Atmospheres. 2019;124
(24):14198–14219.
[25] Shao XM, Lay EH, Jacobson AR. Reduction of electron density in the night-
time lower ionosphere in response to a thunderstorm. Nature Geoscience.
2013;6(1):29.
[26] Marshall RA, Inan US, Glukhov VS. Elves and associated electron density
changes due to cloud-to-ground and in-cloud lightning discharges. Journal
of Geophysical Research: Space Physics. 2010;115(A4):A00E17.
[27] Jacobson AR, Shao XM, Lay E. Time domain waveform and azimuth var-
iation of ionospherically reflected VLF/LF radio emissions from lightning.
Radio Science. 2012;47(4):1–17.
[28] Inan US, Bell TF, Rodriguez JV. Heating and ionization of the lower iono-
sphere by lightning. Geophysical Research Letters. 1991;18(4):705–708.
[29] Taranenko YN, Inan US, Bell TF. Optical signatures of lightning-Induced
heating of the D region. Geophysical Research Letters. 1992;19(18):1815–
1818.
[30] Taranenko YN, Inan US, Bell TF. The interaction with the lower ionosphere
of electromagnetic pulses from lightning: excitation of optical emissions.
Geophysical Research Letters. 1993;20(23):2675–2678.
[31] Inan US, Cummer SA, Marshall RA. A survey of ELF and VLF research on
lightning-ionosphere interactions and causative discharges. Journal of
Geophysical Research: Space Physics. 2010;115(A6).
[32] Gurevich A. Nonlinear Phenomena in the Ionosphere. vol. 10. Springer
Science & Business Media; 2012.
[33] Wait JR, Spies KP. Characteristics of the Earth-ionosphere waveguide for
VLF radio waves. US Dept. of Commerce, National Bureau of Standards: for
sale by the Supt. of Doc., US Govt. Print. Off.; 1964.
[34] Bilitza D, Reinisch BW. International reference ionosphere 2007: improve-
ments and new parameters. Advances in Space Research. 2008;42(4):599–609.
420 Lightning electromagnetics: Volume 2
[35] Pasko VP, Inan US, Bell TF, et al. Sprites produced by quasi-electrostatic
heating and ionization in the lower ionosphere. Journal of Geophysical
Research: Space Physics. 1997;102(A3):4529–4561.
[36] Hu W, Cummer SA, Lyons WA. Testing sprite initiation theory using
lightning measurements and modeled electromagnetic fields. Journal of
Geophysical Research: Atmospheres. 2007;112(D13).
[37] Lee JH, Kalluri DK. Three-dimensional FDTD simulation of electro-
magnetic wave transformation in a dynamic inhomogeneous magnetized
plasma. IEEE Transactions on Antennas and Propagation. 1999;47
(7):1146–1151.
[38] Holland R. THREDS: A finite-difference time-domain EMP code in 3D spherical
coordinates. IEEE Transactions on Nuclear Science. 1983;30(6):4592–4595.
[39] Otsuyama T, Sakuma D, Hayakawa M. FDTD analysis of ELF wave pro-
pagation and Schumann resonances for a subionospheric waveguide model.
Radio Science. 2003;38(6):11–1.
[40] Liu N, Dwyer JR, Cummer SA. Elves Accompanying Terrestrial Gamma
Ray Flashes. Journal of Geophysical Research: Space Physics. 2017;122
(10):10,563–10,576.
[41] Thèvenot M, Bérenger JP, MonediÈre T, et al. A FDTD scheme for the
computation of VLF-LF propagation in the anisotropic earth-ionosphere
waveguide. In: Annales des télécommunications. vol. 54; 1999. p. 297–310.
[42] Li D, Azadifar M, Rachidi F, et al. Analysis of lightning electromagnetic
field propagation in mountainous terrain and its effects on ToA-based
lightning location systems. Journal of Geophysical Research: Atmospheres.
2016;121(2):895–911.
[43] Li D, Azadifar M, Rachidi F, et al. On lightning electromagnetic field pro-
pagation along an irregular terrain. IEEE Transactions on Electromagnetic
Compatibility. 2016;58(1):161–171.
[44] Nucci CA. On lightning return stroke models for LEMP calculations. In:
Proc. 19th Int. Conf. Lightning Protection, Graz, Austria; 1988.
[45] Rachidi F, Nucci C. On the Master, Uman, Lin, Standler and the modified
transmission line lightning return stroke current models. Journal of
Geophysical Research: Atmospheres. 1990;95(D12):20389–20393.
[46] Lin YT, Uman M, Standler R. Lightning return stroke models. Journal of
Geophysical Research: Oceans. 1980;85(C3):1571–1583.
[47] Delfino F, Procopio R, Rossi M, et al. Influence of frequency-dependent soil
electrical parameters on the evaluation of lightning electromagnetic fields in
air and underground. Journal of Geophysical Research: Atmospheres.
2009;114(D11).
[48] Shoory A, Mimouni A, Rachidi F, et al. Lightning horizontal electric fields
above a two-layer ground. In: 2010 30th International Conference on
Lightning Protection (ICLP). IEEE; 2010. p. 1–5.
[49] Zhang Q, Li D, Fan Y, et al. Examination of the Cooray-Rubinstein (C-R)
formula for a mixed propagation path by using FDTD. Journal of
Geophysical Research: Atmospheres. 2012;117(D15).
Lightning interaction with the ionosphere 421
11.1 Introduction
Spectacular large-scale optical flashes in the stratosphere and mesosphere above
large thunderstorm systems were first discovered by Franz et al. [1] serendipitously
during a test of a low-light television camera. Interestingly, this phenomenon was
predicted by the Nobel prize winner C.T.R. Wilson back in the 1920s [2]. And since
then, there has been a lot of evidence from airplane pilots and other eyewitnesses
about short-term light flashes of various shapes and colors over storm clouds [3].
One of the main reasons why this phenomenon was discovered so late is the
difficulty of observing this phenomenon from the earth’s surface since the optical
flashes in the stratosphere and mesosphere are closed to the observer by thunder-
clouds. In ground-based measurements, the giant optical flashes above thunder-
clouds can be seen at a small angle to the horizon at distances of hundreds of
kilometers from the flash site. In addition, they have a very short duration and occur
much less common than ordinary cloud-to-ground lightning (e.g., [4], and refer-
ences therein).
Recent researches have revealed a surprising variety of such large-scale optical
events occurring above thunderstorms at stratosphere, mesosphere, and lower
ionosphere altitudes, which have been termed Transient luminous events (TLEs)
(e.g., [5]). The most commonly observed phenomenon, the so-called red sprite or
sprite, manifests itself as a luminous red glow occurring at a 50–90 km altitude
range gradually changing to blue color below 50 km [5]. To date, a large amount of
information is available concerning these beautiful phenomena. If you type “sprite
discharge” in the Internet search bar, you will see a large number of sprite photos
taken in various conditions. To illustrate, Figures 11.1 and 11.2 show photos of
sprites captured by P. M. Smith during the 2019 Kansas storm season.
1
Pushkov Institute of Terrestrial Magnetism, Ionosphere and Radio Wave Propagation (IZMIRAN),
Russian Academy of Sciences, Russia
2
Institute of Physics of the Earth, Russian Academy of Science, Russia
3
Advanced Wireless and Communications Research Center, The University of Electro-Communications
(UEC), Japan
4
Hayakawa Institute of Seismo Electromagnetics (Hi-SEM) Co. Ltd., UEC Alliance Center 521, Japan
426 Lightning electromagnetics: Volume 2
Figure 11.1 A sprite image by P. M. Smith, which was taken from central
Oklahoma, looking over Kansas (Wichita area) in September 2019.
The photo is provided with P. M. Smith’s permission
Figure 11.2 The collection of sprites by P. M. Smith was taken from Northern
Oklahoma during a 2-hour period of storms over Central Kansas in
April 2019. In this image, all the brightest sprites are combined during
that time frame. The photo is provided with P. M. Smith’s permission
Another type of TLEs primary blue color was first discovered during the
Sprites94 aircraft campaign [6,7]. Figure 11.3 illustrates one of the first photos of
blue jets (BJ) detected over a large evening thunderstorm in the Indian Ocean on
Lightning effects in the mesosphere 427
Figure 11.3 A picture of the blue jet by P. Huet as observed in Réunion Island,
Indian Ocean (20 51:840 S latitude, 55 27:60 E longitude), on March
1997. Adapted from Wescott et al. [8]
March 1997 [8]. This kind of TLEs termed BJs propagates upward from the top of
the cloud towards the stratosphere up to altitudes of about 40 km. The BJ’s basic
properties have been well-documented despite their rare occurrence as compared to
sprites. The so-called blue starters (BSs) and gigantic jets (GJs) are propagating
upward gigantic electric discharges accompanied by luminous phenomena, which
are considered as some varieties of BJs. One of the main features of BSs is that they
propagate only a few kilometers over the cloud top to altitudes below 25 km [9].
The GJ propagates upward from the thundercloud top through the stratosphere and
mesosphere to a terminal altitude of about 8590 km. The GJ was first observed at
the Arecibo Observatory in Puerto Rico over an intense oceanic thunderstorm
200 km away from the Observatory [10]. Figure 11.4 shows an example of GJ
observed at the Mauna Kea Observatory in Hawaii and captured by Frankie
Lucena. The most intensive optical emission primary blue color is seen at the base
of GJ while the reddish color prevails in the upper portion of the GJ.
In recent decades, the study of high-altitude discharges in the atmosphere has
become an extensive, rapidly developing area of geophysical electrodynamics.
Since the sprite discovery [1], several ground-based and aircraft campaigns (e.g.,
428 Lightning electromagnetics: Volume 2
Figure 11.4 The gigantic jet image was captured by Frankie Lucena at the Mauna
Kea Observatory in Hawaii on the night of July 24, 2017. Taken from
https://spaceweathergallery.com/
[4], and references therein) as well as measurements on satellites and aboard the
International Space Station (ISS) have been conducted to study the sprites on dif-
ferent continents and oceans (e.g., [11–13]). Many other amazing varieties of TLEs
were discovered during these studies. These include “Sprite halos” which are a
brief, diffuse glow regions of light occurring at 70–85 km altitude [14] that may be
followed by a sprite although a significant fraction of haloes occurs without being
accompanied by sprites; “Elves,” which are reddish rapidly expanding rings that
sometimes appear on the lower surface of the ionosphere after an intense lightning
discharge [15–17]; so-called “Trolls” which are red spots arising after the flash of
an extremely strong sprite [18]; “Gnomes” and “Pixies” which are very small, brief
spikes and spots of light appearing above the tops of a large thundercloud and etc.
On the basis of satellite FORMOSAT-2 observations, the rate of TLE occurrence
Lightning effects in the mesosphere 429
over the globe is estimated to be several million events per year. Recent satellite
measurements have shown that a portion of UV radiation flashes, although asso-
ciated with the electrical activity in underlying thunderstorms, are not directly
related to lightning discharges and thus can be differed from traditional TLEs [19].
Thus, the list of amazing large-scale optical and electrical phenomena in the upper
atmosphere continues to expand.
This chapter presents an overview of recent high-speed video and satellite-
based observations of TLEs. The main emphasis is placed on the interpretation of
observed features of TLEs and on the recently advanced theories explaining these
features.
11.2 Sprites
11.2.1 Basic properties and morphology of sprites
As is seen from Figures 11.1 and 11.2, there can be a lot of variation among sprites
including their shape and size, vertical and horizontal dimensions, number of sprite
groups, brightness, and details of the fine structure of sprites. The majority of
sprites can be divided into two basic morphological classes [20–22]. The
columniform-shaped sprites are luminous vertical columns about 10 km long and
less than 1 km in diameter with a slightly diffuse top and abrupt bottom, which
sometimes ends in dim downward-directed streamer filaments. An example of such
columnar sprites can be found in the upper-right corner of Figure 11.2. Groups of
columnar sprites are often observed, occasionally reaching 20–30 elements.
Typically, the individual sprite flash lasted from a few to several tens of ms.
The typical “carrot” sprite shown in Figure 11.1 consists of the luminous
column with the upper diffuse region in red color at a 50–90 km altitude range and
lower tendril-like filamentary structure in bluish color. The transverse size of the
luminous area varies from 20–30 km to 50–100 km (e.g., [23]). Sometimes the
sprite occurrence is preceded by the appearance of a halo at an altitude of
approximately 75 km. The typical sprite halo is visible as a brief descending diffuse
glow in a pancake shape about 10 km in thickness and 50–80 km in diameter.
In a group of so-called “dancing sprites,” its individual elements can “dance”
over the thunderstorm during a long period of about 0.1–1 s [24–27]. The dancing
sprites are assumed to be associated with the continuing current (CC) of causative
lightning and correspond to surges in the CC moment waveform. The long-time-
delayed elements occur lower than the short-time-delayed elements and they are
horizontally shifted several tens of kilometers from the causative stroke.
It is known that most cloud-to-ground (CG) lightning discharges are negative;
that is, those discharges lower the negative charges from thundercloud to ground.
However, 5%–10% of global CG lightning activity is composed of positive cloud-
to-ground (+CG) lightning, which transfers the positive charges to ground (e.g., see
[28,29], and references therein). Interestingly, the first sprite observations showed
that 99.9% of the sprite events occurred several milliseconds after a strong +CG
[30,31]. The mean time delay between causative +CGs and the sprites is about
430 Lightning electromagnetics: Volume 2
qðz hÞ
Ez ¼ ; h r < z < h þ r;
4pe0 r3
qðz hÞ (11.3)
Ez ¼ ; 0 < z < h r or z > h þ r;
4pe0 jz hj3
where e0 is the electric constant/dielectric constant of free space. Since the ground
is a conductor, the induced charges of the opposite sign will appear on the ground
surface. The electric field of these induced charges in the atmosphere coincides
with the field of the mirror electric image of the ball located at depth z ¼ h
(Figure 11.5). The left side of Figure 11.5 shows a model of electric charges in a
cloud and their electrical images in the ground. The net electric field taken along
432 Lightning electromagnetics: Volume 2
90
z
80
70
2
60
3
50
z, km
4
40
1
30
20
E
z=h 10
E
0
–3 –2 –1 0 1 2 3 4
z = –h log Ez , kV/m
the z-axis is given by Ez ¼ Ez þ Ezþ , where Ez is defined by (11.3), and the field
of induced charges has the form:
q
Ezþ ¼ ; ðz > 0Þ: (11.4)
4pe0 ðz þ hÞ2
where q < 0 is the charge of thundercloud. Note that at high altitudes under
requirement z h the (11.3) and (11.4) are simplified. In the first approximation,
for the small parameter h=z we obtain:
qh
Ez ¼ Ez þ Ezþ : (11.5)
pe0 z3
This dependence on height is not surprising, since the cloud charges and their
electric image in the ground form an electric dipole with a dipole moment 2qh.
The direction of the vector of the total field Ez under and above the charged
ball is indicated in Figure 11.5 by vertical arrows. The dependence of the absolute
value of the vertical electric field on the height z, calculated by using (11.3) and
(11.4), is shown in Figure 11.5 with line 1. In making this plot the following
numerical values of the parameters have been used q ¼ 150 C, h ¼ 10 km, and
r ¼ 2 km. Such a thundercloud charge could arise immediately after the +CG
discharge, which results in the CMC of the order of 1,500 Ckm.
Lightning effects in the mesosphere 433
80 km
70 km
60 km
50 km
Figure 11.6 High-speed images of the sprite development from small dim
inhomogeneities as observed at Yucca Ridge Field Station on August 13,
2005, at 03:43:09.4 UT. The nucleation point of the left sprite element is
marked with a white arrow. Adapted from Cummer et al. [42]
These spatial irregularities descend rapidly along with the sprite halo and then
slow down producing spots of stationary glow. Finally, the streamer suddenly
starts from this formation. Figure 11.7 shows this scenario for the development
of the sprite. This event illustrates the sprite development from brightening
inhomogeneities at the bottom of a halo [42]. The bright inhomogeneities arise
at the lower edge of the originally homogeneous halo 2.2 ms after the return
stroke. The downward propagating sprite streamers begin to develop from these
spots and then branch in the same way as the previous example. The upward
streamers propagate at velocity ð0:52Þ 107 m/s and terminate in diffusive
emissions. The expanding bright columns in the upper portion of the sprites are
wider and more diffuse than that visible in the sprites shown in Figure 11.6. The
last images in Figure 11.7 illustrate the gradual fading of the glow and the
disintegration of the sprite.
Early observations have shown that the initial streamer’s velocities are greater
than 107 m/s [44]. Using a multi-anode photometer, McHarg et al. [45] have
observed the downward and upward propagating streamer at the velocities on the
order of 107 108 m/s. On the basis of high-speed imagining with 0:1 ms resolution,
McHarg et al. [46] has shown that the highest part of the upward sprite streamers
can accelerate on the order of 1010 m/s2 at the initial stage of streamer develop-
ment. Plots of the streamer head velocity versus time and velocity versus altitude
are shown in Figure 11.8. As is seen from this Figure, the lowest region of down-
ward streamers accelerates initially to a maximum velocity ð13Þ 107 m/s and
then immediately decelerates at approximately constant value 1010 m/s2 which is
saved until the sprite is stopped [47].
80 km
70 km
60 km
50 km
80 km
70 km
60 km
50 km
Figure 11.7 High-speed images of the sprite development from the sprite halo as
observed at Yucca Ridge Field Station on August 13, 2005, at
03:12:32.0 UT. Adapted from Cummer et al. [42]
Lightning effects in the mesosphere 435
× 107 × 107
2 3
Center Center
1× Left Left
2 1×
V, m/s
10 10
V, m/s
10 10
1 m/ 2 m/s 2
s
1
0 0
0 0.5 1.0 1.5 2.0 2.5 0 0.4 0.8 1.2 1.6 2.0
× 107 Time, ms × 107 Time, ms
2 3
Center Center
Left Left
2
V, m/s
V, m/s
1
1
0 0
75 70 65 60 55 50 75 70 65 60 55 50 45 40
A׀t, km A׀t, km
Figure 11.8 The downward streamer head velocity versus time (upper row), and
the streamer velocity versus altitude (lower row). In making the first
and second columns the high-speed images shown in Figures 11.6
and 11.7 were used. Adapted from [47]
e þ O2 ! 2e þ Oþ
2; e þ N2 ! 2e þ Nþ
2: (11.6)
In addition, free electrons can appear due to photo-ionization near the charged
streamer head, where the charge density and electric field are more significant.
436 Lightning electromagnetics: Volume 2
Under an appreciable electric field, a major role is played by the competing pro-
cesses of dissociative two-body or three-body attachment of electrons to either O2
or N2 molecules
e þ O2 ! O þ O ; e þ O2 þ A ! O
2 þ A; (11.7)
where A is another neutral molecule. It is generally accepted that the dissociative
two-body attachment is more important than the three-body attachment at meso-
spheric and lower ionospheric altitudes [14].
For simplicity, we consider a three-component plasma consisting of electrons
and single charged positive and negative ions. Let ne be the electron number density
while nþ and n be the number densities of positive and negative ions, respectively.
To treat the electromagnetic phenomena associated with sprite development includ-
ing its streamer structure and relationship with intra-cloud (IC) process, a set of
transport kinetic and electrodynamic equations are required. The equations describing
the variations in the number densities of electrons and ions must take into account not
only the ionization and electron attachment to neutrals but also the electron-ion
recombination, electron drift and diffusion, etc. Let ni and na be the ionization and
attachment rates, while me and De be the electron mobility and diffusion coefficient,
respectively. Then the basic kinetic equations are as follows:
vi – va,
2
103 s–1
0.2 1.2 E N,
0.4 0.6 0.8 1.0
0
E1 E0 E* Emin E2 10–19 V m2
–1 τ –1
–2
–3
Figure 11.9 The difference between the ionization rate and electron attachment
¼ E=N as calculated
rate as a function of reduced electric field E
from (11.12) at an altitude of 70 km. The attachment instability can
min where the graph has a minimum.
develop to left from the point E
Adapted from Surkov and Hayakawa [52]
438 Lightning electromagnetics: Volume 2
As the first term on the right-hand side of (11.15) is neglected, the solution of
this equation under zero initial condition has a form Ra ¼ ð2De tÞ1=2 , which is
typical for diffusion processes. In the model by Qin et al. [53], it is assumed that
ionization and dissociative electron attachment in the thundercloud QE field play a
major role in the excitation of the sprite streamer. In this case, only the ionization
and attachment rates can be retained in (11.8), neglecting the other terms on the
right-hand side of the equation
dne
¼ ðni na Þne : (11.16)
dt
Just after + CG lightning, the uncompensated negative electric charges appear
in a thundercloud for a short time. Equations (11.15) and (11.16) were used to
approximate the inception of sprite streamers from sprite halo in the electric field of
these charges. In this model, the condition Eс0 Ec =3 was applied as a criterion for
the avalanche-streamer transition. The numerical modeling has shown that the
sprite streamer initiation depends strongly on the CMC of causative + CG lightning
and the ambient electron density profile.
In semi-analytical models by Surkov and Hayakawa [52], a sprite is approxi-
mated through an expanding plasma ball that develops from a plasma inhomo-
geneity situated at altitudes of 70–80 km above the thundercloud. To simplify the
problem, the streamer structure of the sprite leaves out of the account, and the
plasma ball is assumed to be homogeneous and conductive. At the mesospheric
altitudes, the mobility of the electrons is much greater than that of the ions, so only
the electron conductivity is taken into account inside the ball, that is, s ¼ ene me .
Combining this relationship and (11.16) and taking into account the continuity
equation for electric current, yields
ds
r j ¼ 0; ¼ ðni na Þs; (11.17)
dt
where j ¼ sE þ e0 @ t E is the total electric current density.
Just after causative lightning, the uncompensated charges arise in the thun-
dercloud. The plasma ball is polarized in the electric field in the thundercloud
electric field E0 ðtÞ and polarization charges appear on its surface thereby producing
the time-dependent dipole-type field outside the ball. In this model, the total elec-
tric field is given by
pðtÞ r
Eðr; tÞ ¼ E0 ðtÞ r ; (11.18)
4pe0 r3
where pðtÞ is the dipole moment of the polarization charges while r denotes the
position vector drawn from the ball center which is located at the height h. The total
electric field (11.18) at the lowest point of the ball is assumed to be equal to kEc ,
where k 1 is the dimensionless factor. The polarization charges weaken the
electric field inside the ball. But, if this field exceeds the air breakdown threshold,
then the ionization of the air and the production of free electrons begin to prevail
440 Lightning electromagnetics: Volume 2
over their attachment to the neutral molecules followed by the increase in the ball
plasma conductivity in (11.17). In this model, the rates of ionization and electron
attachment to molecules are governed by the electric field (11.18) in accordance
with (11.12). That is, given some electric field E0 ðtÞ, the problem is reduced to a set
of ordinary nonlinear equations for the functions pðtÞ, sðtÞ, and RðtÞ. These equa-
tions should be supplemented by the proper boundary conditions on the ball sur-
face, that is, the continuity of electric potential and normal component of the total
current density.
The ball radius begins to increase as soon as the electric field of the thunder-
cloud exceeds the conventional breakdown threshold of air Ec ðhÞ. The analysis of
the approximate analytical solution of this problem showed that the initial accel-
eration a0 ¼ d 2 R=dt2 is practically independent of the rate of change of the thun-
dercloud electric field and is given by [54]
me0 kEc ðhÞsa ðhÞHa Ei0 Ea0
a0 ai0 exp aa0 exp : (11.19)
3e0 kEc0 kEc0
For night conditions, the conductivity of the ambient air sa ðhÞ at altitude
h ¼ 70 km is about 109 S/m while Ec ðhÞ 0:51 kV/m. Substituting the above
numerical values of the parameters as well as Ha ¼ 8 km and k ¼ 1:2 in (11.19),
we obtain that a0 1010 m/s2.
The numerical solution of this model problem is illustrated in Figure 11.10 as
the thundercloud QE field is approximated via E0 ðtÞ / ðt=tr Þexpðt=tr Þ, where tr is
× 107
4.5
4 2 ms
5 ms
3.5 9 ms
2.5
V, m/s
1.5
0.5
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45
t–t* , ms
Figure 11.10 Model calculation of the expansion velocity of the plasma ball
immersed in thundercloud QE field. Taken from Surkov and
Hayakawa [54]
Lightning effects in the mesosphere 441
the relaxation time. In this figure, the maximum expansion velocity of the plasma
ball varies within ð3:24:4Þ 107 m/s depending on tr .
Despite this model being far from perfect, the above estimates of the ball
expansion velocity and acceleration are close in magnitude to the observed values
of the sprite streamer velocity and acceleration shown in Figure 11.8. To proceed
analytically, it is necessary at this point to construct a more complete model of the
sprite which takes into account the structure of a developed sprite consisting of
individual moving streamers. Even in this simplified form, the set of (11.8)–(11.12)
describing the plasma phenomena associated with sprite evolution is rather com-
plicated. Thus, the most portion of the theoretical studies of sprites are based on
numerical simulations of the (11.8)–(11.12) or similar equations (e.g., [55–57]).
The reader is referred to that works for details about sprite numerical simulations.
Figure 11.11 A sprite telescopic image at wide (a) and narrow (b) field of view
(FOV) as observed over northwestern Mexico on July 13, 1998, at
06:00:00 UT. The small rectangle on panel (a) is shown on panel
(b) at a large scale. The images have been false-colored and
saturated at 685 kiloRayleigh (kR) in order to better represent
intensity (see color bar). Taken from Gerken et al. [59]
442 Lightning electromagnetics: Volume 2
Figure 11.11 shows sprite telescopic images taken over northwestern Mexico
on July 13, 1998, at low (a) and high (b) resolutions [59]. The small white rectangle
on panel (a) is shown on panel (b) at a large scale in order to highlight the fine
structure of the sprite. It is evident from this image that the tendril-like region of the
sprite consists of densely packed branching streamers. These high-resolution
measurements have shown that the mobile bright compact balls, shown in this
image with red color, are the highly ionized streamer heads. In the altitude range of
6064 km, the transverse scale of filamentary structures ranges from 60 to 145 m
whereas the size of streamer heads varies from 10 to 100 m and the streamer speed
lies in the range of 106 107 m/s [46,59].
The diameter of the CG lightning streamer is smaller than 1 mm and its length
amounts to tens of meters, whereas the size of sprite streamers is much greater. This
fact can be explained based on the similarity theory for streamers that establishes
similarity relations for different streamer parameters and the air number density
([40,41,60]). First of all, we consider the scaling in size of a glow discharge.
According to the similarity theory the typical streamer length Ls , the streamer head
radius Rs , and the charge qs in the head vary inversely proportional to the neutral
gas density N . For example, this means that Rs ðzÞ ¼ Rs0 N0 =N ðzÞ, where Rs0
denotes the streamer head diameter at sea level and z is the streamer altitude.
Substituting (11.1) for N ðzÞ into the above relationship and using the parameters
z ¼ 62 km, Rs0 ¼ 1 mm, and Ha ¼ 7 km, we obtain that Rs 7 m. Kanmae et al.
[61] have shown that the sprite streamer diameters have to be larger than that
predicted by this simple similarity law possibly due to the effects of the photo-
ionization and an expansion of the streamer head along its propagation over a long
distance. In this notation, the above estimate is compatible with the streamer head
size observed by Gerken et al. [59].
The dynamics of the individual sprite streamer and its internal structure are
rather complicated. After the passage of the streamer heads, the ionized channels
arise in their wake. These channels exhibit intricate luminous patterns of alternating
bright and dark spots, conventionally called beads and glows. The sprite beads
appear as persistent, localized spots of light emission that often punctuate a strea-
mer channel [62]. Their lifetime is much longer than the duration of the streamer
head propagation.
One conceivable reason for the existence of beads and glows is assumed to be
the attachment instability of electric discharges developing in the streamer channel
[63]. The attachment instability builds up as a result of the fluctuation of the
electron number density ne at certain electric field in the streamer channel (e.g.,
[39]). To explain this kind of instability, consider again (11.16) describing the
variation of electron number density due to the ionization and the dissociative
attachment of electron to neutral molecules. The difference between the ionization
rate and electron attachment rate versus the reduced electric field is shown in
Figure 11.9.
Let I ¼ ene me SE be the time-dependent electric current in a streamer channel,
where S is the cross-section of the streamer channel. For simplicity, this current is
assumed to be uniform and constant along a streamer channel. This implies that all
Lightning effects in the mesosphere 443
the parameters in the expression for I are taken in an average sense. Let E min
denotes the reduced electric field which corresponds to minimum of the graph
shown in Figure 11.9. Suppose that the electric field E min in some
0 is less than E
cross-section of the channel. Since the electric current is constant along the chan-
nel, the small negative fluctuation of the electron number per unit of the channel
length, ne S, in this section will result in a small increase of the local electric field to
the value E >E 0 . This change in the electric field, which is schematically shown
in Figure 11.9 with a blue arrow, leads to a decrease in the difference ni na . From
(11.16), it is clear that this change in ni na will cause an additional decrease of ne ,
which in turn results in the increase in E and so on thereby exciting the attachment
instability.
Following Luque et al. [63], we now examine the mechanism of beads generation
in the presence of exponentially decreasing electric current I ¼ I0 expðt=tÞ in the
streamer channel. Here t 1 ms is the current relaxation time. Assuming for the
moment that S is constant, then the equation for the current in the streamer channel can
be rewritten to the view I 1 ðdI=dtÞ ¼ n1 1
e ðdne =dt Þ þ E ðdE=dt Þ. Substituting the
expression for electric current and (11.16) for dne =dt into the above relationship yields
dE=dt ¼ E na ni t1 : (11.20)
There are three stationary points in (11.20); that is, E ¼ 0 and two points E 1
1
and E 2 which are the roots of the equation na ni ¼ t . These roots are defined
by the intersection points of the graph na ni with a red horizontal line t1 shown
in Figure 11.9. It follows from (11.20) that the electric field can enhance in the
streamer channel areas if the electric field lies within the interval from E 2
1 to E
shown in Figure 11.9. Additionally, the reduction of the free electrons due to their
attachment to neutral molecules gives rise to the resistance of these areas. This
results in the Joule dissipation and enhancement of the heat release, which can
cause the transformation of these areas into quasi-stationary bright beads.
The model based on attachment instability claims only a qualitative explana-
tion of the observed beads and glows in the streamer channel. Actually, the beads
appear in the streamer wake at different times after the streamer head passage. In
this notation, it seems likely that the streamer’s current variations and the streamer
interaction can play a significant role in bad and glow excitation.
data contained not only the signal radiated by the lightning current but also the peak
which coincides in time with the appearance of sprite luminosity [65]. This fact was
established by using the ISUAL instruments on board the FORMOSAT-2 satellite.
Figure 11.12 displays an example of ELF/VLF variations detected by the ground-
based magnetometer located approximately 2780 km from lightning discharge [66].
The radiation caused by the sprite current manifests itself as 12 ms pulses that
follow a lightning return stroke by a few milliseconds to a few hundred milli-
seconds. It should be emphasized that these pulses are caused by the currents
generated inside the sprite itself and they are not associated with the M-component
of continuing current of causative +CG discharge. In contrast to return stroke, the
sprite spectrum almost entirely belongs to the ELF region.
In theory, one of the main characteristics of the ELF field at large distances
from a lightning discharge is the so-called lightning current moment, which is
approximately equal to the product of the discharge current I ðtÞ and the length LðtÞ
of the lightning channel. The current moment waveform of the observed signals can
be extracted from the ELF data by using the model approximation of the lightning
and sprite currents. To extract this information, the solution to the problem of the
ELF electromagnetic field propagation in the Earth-ionosphere waveguide is used
[66]. The time dependence of the lightning and sprite current moments is selected
in such a way that the calculated ELF field pulse would be close to the observed
ELF magnetic field variations. On the basis of several case studies, the sprite cur-
rent moments have been estimated as much as 200300 kAkm [66].
A network of ground-based stations equipped with different sensors has been
operative during Japanese sprite campaigns in winters 2004/2005 and 2006/2007
[33]. This network included optical instruments at two-spaced observatories, VLF
and ELF electromagnetic field antennas, and magnetometers. High-sensitive cam-
eras have been used to capture optical emissions of causative lightning and sprite
images. The coordinates and onset time of lightning flashes were measured by
using three VHF antennas of the SAFIR interferometric lightning system. The
azimuthal magnetic field (nT)
0
sprite
–0.4 return current
stroke pulse
pulse
–0.8
0
HNS(+), mA/m
–0.5
–1
120
100
80
60
40
20
5 10 15 20 25 30 35
Frequency, Hz
Figure 11.14 Power spectral density of the ELF magnetic field variations which
were shown in Figure 11.13. Adapted from Surkov et al. [67], with
permission, Elsevier
446 Lightning electromagnetics: Volume 2
To explain this property one should take into account that this power spectrum
contains the contribution of the magnetic fields of both causative lightning dis-
charge, Bc , and the delayed sprite, Bs . Let t be the lag of time between the cau-
sative lightning and sprite occurrences, while rc and rs be their position vectors
with respect to the observation site. The total spectrum is then given by [67]:
52.6
51.0
50.2
49.4
48.6
47.8
–4 0 4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
Time (s)
duration smaller than 20 ms and coincides with the sprite onset time. This data
counts in favor of the suggestion that these sprite-related early/fast events are
caused by the interaction of VLF waves with the conductivity perturbation in the
upper D region. The typical recovery time of the early/fast effects varies from
10 to 300 s and is comparable to the LEP recovery time [4,77]).
It is generally believed that this recovery time is basically determined by
recombination processes in the ionospheric plasma since the dissociative recom-
bination of electrons and single positive ions prevails over electron attachment at
the altitudes of the lower ionosphere [78]. In consideration of this approximation,
(11.8) describing the perturbations of electron and ion number densities is simpli-
fied to
dne
¼ bd ne nþ : (11.21)
dt
Let ne0 be the electron number density arising just after the short-term stage of
interaction of the sprite-radiated electromagnetic wave with the lower ionosphere.
Taking into account that ne nþ and performing integration of (11.21) under the
requirement ne ð0Þ ¼ ne0 , we come to the usual law for binary plasmas
ne0
ne ¼ : (11.22)
1 þ bd ne0 t
Substituting ne ¼ ne0 =3 into (11.22) we obtain the rough estimate of the
recovery time: tr 2=ðbd ne0 Þ. Using the typical values of the parameters
bd ¼ ð13Þ 107 cm3/s [78] and ne0 ¼ 6 104 cm–3 [77] gives the value of
tr ð13Þ 102 s, which is compatible with the observed recovery time.
448 Lightning electromagnetics: Volume 2
HWV-CRE TE, dB
1
0.5
0
–0.5
–1
–1.5
Another type of the sprite effect on phase and amplitude of VLF waves pro-
pagating in the Earth–ionosphere waveguide is referred to as “early/slow” event
[4]. Figure 11.16 shows an example of such event observed in Crete at a distance of
about 2,000 km from a convective storm in central France. The sprite appears at the
time marked by the vertical dashed line in Figure 11.16. The “early/slow” events
have a long onset duration of up to 2.5 s, which is much greater than the “early/fast”
onset duration shown in Figure 11.15.
The ground-based VLF measurements have shown that the growth phase of the
early/slow events is accompanied by bursts of spherics which are commonly related to
intra-cloud (IC) lightning discharges whereas spherics were not detected during onsets
of early/fast events [74]. This implies that the long duration of the “early/slow” onset
time can be associated with the enhancement of IC electric activity before and after the
sprite appearance. Horizontal IC discharges can have the greatest impact on the
ionosphere since the horizontal currents radiate most of the electromagnetic energy in
the directions perpendicular to the current line. A series of electromagnetic pulses
radiated upward from horizontal IC discharges can accelerate sprite-produced elec-
trons followed by secondary ionization buildup in the upper D region of the iono-
sphere. This leads to a gradual buildup of conductivity changes below the nighttime
VLF wave reflection heights. One may assume such an effect could be responsible for
the long onset durations of the observed early/slow events. The VLF measurements
thus provide us with important information about the role played by IC processes in
the perturbation of the lower ionosphere above the sprite occurrence.
observed over a large evening thunderstorm in the Indian Ocean [8] is shown in
Figure 11.3. Figure 11.17 demonstrates the inverted black and white image of the
BJ shown in Figure 11.3 [8]. At the base of BJ, its diameter is about 400 m. At
30 km altitude, the jet diameter broadens to about 2 km.
It is evident that the main cause of the optical emission of BJs is the upward-
directed large-scale electric discharge in the stratosphere that cause the ionization
of air and the excitation of air molecules including N2 by electron impact. The red
N2 emissions are strongly quenched at these altitudes. So, the blue color in the
optical emission of BJs is caused by the predominance of the emission of the
excited energy levels of the second positive band of N2 (N22P) (e.g., [79]). As
compared to the sprites, BJ luminosity is brighter than the sprite ones while the BJ
tip moves slower than the sprite streamer head. Unlike sprites, the BJs do not seem
to require the occurrence of lightning before the BJ discharge.
Blue starters (BSs) are a kind of BJs, which propagate upward from cloud tops
at 1718 km altitudes and terminate abruptly at altitudes below 25 km. Basically,
they differ from BJs by a lower terminal latitude [9]. Analysis of data gathered by
40
35
Altitude (km)
30
25
20
18
0 2 4
Scale (km)
Figure 11.17 The inverted black-and-white image of this blue jet which was
shown in Figure 11.3. Adapted from Wescott et al. [8]
450 Lightning electromagnetics: Volume 2
the ISUAL payload on board the FORMOSAT-2 satellite has shown that BSs have
a length about 8 3 km with a width of ð24Þ km [80].
As compared to BJs and BSs, the GJs are more intensive discharges propa-
gating upwards from the thundercloud top through the stratosphere and mesosphere
to lower ionosphere. When the GJ terminates at altitude of about 8590 km, it
produces the electrical connection between the thundercloud top and the conduct-
ing D- and E-layers of the ionosphere [10,81,82]. The first images of GJ over an
oceanic thunderstorm were captured at the Arecibo Observatory in Puerto Rico
[10]. Figure 11.18 demonstrates an example of GJ observed over the Pilbara region
in the north of Western Australia on March 28, 2017. The bright blue/purple stem
below the streamer corona is thought to be a leader channel. The top edge of the
bright stem is usually located at 3050 km altitude [83,84]. The upper streamer
zone at mesospheric altitudes exhibits red color, in analogy to the sprites, mainly
due to the excitation of the first positive band system of N2 emission.
A majority of GJs are of negative polarity and can transfer more than 100 C of
negative charge from the cloud top to the ionosphere (e.g., [85]). The ISUAL
experiment observed global rates of 0.50, and 0.01 events per minute for sprites and
gigantic jets, respectively [86]. Although the instruments with a larger sensitivity and
coverage will detect many more such events, it is evident that the GJ occurrence is a
much rarer event compared to a sprite occurrence. The GJs are more frequent over
intense tall tropical thunderstorms of 1418 km altitude with convective surges or
even overshooting [87–89] although Yang et al. [90] have reported observations of
GJs over a mesoscale convective system in the middle latitude region in eastern
China. The GJ over a maritime thunderstorm not only with 6.5 km tall but also with
overshooting has been observed by van der Velde et al. [84].
Figure 11.18 A gigantic jet over Pilbara, Australia, on March 28, 2017, captured
by Jeff Miles. Taken from https://watchers.news/2017/03/31/gigan
tic-jets-over-pilbara-australia-on-march-28-2017
Lightning effects in the mesosphere 451
90 90
80
80 Pasko et al. [2002] 70 Soula et al. [2011]
70 60
Altitude (km)
Altitude (km)
60 50
≥ 2.3×106 m/s
≥ 1.2×106 m/s
50 40
6.3×104 m/s
40 5.7×104 m/s 30
30
20 20
50 100 150 200 250 80 100 120 140 160 180
(a) Time (ms) (b) Time (ms)
80
70
60
50
40
30
20
1 image = 1ms
90
80
Altitude (km)
70
60
50
40
30
20
altitude range of 3550 km. From this point, the GJ develops in a bidirectional
fashion in such a way that upper portion of the GJ propagates toward the iono-
sphere without interruption forming the diverging branched structure of the nega-
tive streamers whereas the power portion of the GJ transforms into downward
extending filaments/positive streamers, which gradually fade over time. Van der
Velde et al. [91] have reported that the upward negative streamers propagate at
mean speeds of ð12:5Þ 107 m/s, accelerating above 70 km up to the speed
7:5 107 m/s; that is, one order magnitude greater than that reported by Pasko et al.
[10] and Su et al. [81]. Notice that no downward-propagating return strokes were
detected upon connection to the ionosphere in contrast to CG lightning.
Simultaneous optical and ground-based ELF observations have shown that GJs
result in the generation of low-frequency magnetic fields [83,85,92]. For example,
the ascending streamers during the final jump and decay of the FDJ are accom-
panied by the appearance of a sharp peak in the magnetic field perturbation with a
duration of about 5 ms. This peak is assumed to be due to the increase in the FDJ
current up to 350 A during the final jump [91].
The TJ stage (numbered by 5–7) begins after the GJ reaches the ionosphere and
then lasts for about 200 ms. The upper part of the TJ luminescence attenuates
gradually during this stage although the middle part of the jet starts to form a
brighter section after a few milliseconds [89,91]. This bright section/trailing jet
reaches a maximum luminosity 6080 ms after the time when the FDJ stage is
completed. Beads and new patches appearing above the TJ indicate that an electric
field may exist in this region. The upward velocity of the TJ top slows down from
9 105 m/s to 2 104 m/s after reaching maximum luminosity and then the TJ ter-
minates at 5562 km altitudes. The highest current moment during the TJ stage is
estimated to be 37.5 kA km, which corresponds to the current of about 850 A while
the total CMC by the end of the TJ stage is 2,300 C km.
11.3.3 Models of gigantic jet
It is generally believed that the conditions under which a BJ or GJ can occur largely
depend on the distribution of electric charges in the cloud. Classical, normally
electrified thunderstorms have a stratiform electrical structure close to a typical
dipolar structure. Most part of negative charges predominantly accumulates in the
middle region of the thundercloud, whereas the majority of positive charges tend to
pile up at the thundercloud top [29,93,94]. These main charges are supplemented
by a small positive charge located at the bottom and negative screening charge at
the upper cloud boundary.
An example of the electric field measured by a balloon in the typical thun-
dercloud is shown in Figure 11.21, which demonstrates strong variations of the
electric field with altitude [95]. The arrow L indicates the height at which the
electric field is close to the conventional breakdown threshold. In order to explain a
normal and nonstandard electrical structure of the cloud, several models of the
charge distribution in the cloud have been proposed. The models consisting of four
charges areas located at different altitudes are illustrated in Figure 11.22 [94].
These space charges are assumed to have a Gaussian spatial distribution. The
454 Lightning electromagnetics: Volume 2
12 12
L
10 L 10
8 L
8 L
L
6 6
L
4 4
2 2
0 0
–200 –100 0 100 200 –200 –100 0 100 200
E, kV m–1 E, kV m–1
negative and positive charge areas are shown with blue and red contours while the
total charges of these areas are indicated to the right from contours (in C).
Figure 11.22a, c displays the model charge distribution at which the IC lightning (a)
can occur between the two most highly charged regions, and the model that cor-
responds to the low-altitude IC lightning (c). Negative CG discharge can arise if the
storm accumulates lower positive charge as shown in Figure 11.22b. Observations
of the so-called “bolt-from-the-blue” (BFB) lightning discharges can be explained
in terms of the model shown in Figure 11.22e. The negative BFB discharge begins
as regular upward-directed IC discharge which does not terminate in the upper
positive charge. Instead, it continues horizontally out the upper side of the storm
and turns downward to ground [94].
It is generally believed that the positive BJ discharge may occur under the
requirement that large amount of positive charge piles up near the top of cloud.
This charge can be covered from above by a negatively-charged screening layer as
shown in Figure 11.22d. It seems likely that the BJ discharges once triggered would
propagate upward through the negative-charged layer and escape the cloud top into
the stratosphere.
Since the GJ usually bears a large amount of negative charge, in contrast to the
BJ discharges, the GJ events can be associated with the accumulation of great deal
of negative charges at middle level of the thundercloud and opposite charges at
upper level. This charge distribution provides an alternative way of neutralization
the mid-level negative charge, by discharging it to the upper atmosphere due to GJ
rather than to ground due to CG lightning. Figure 11.22f shows the model dis-
tribution of space charges in the cloud, which may precede the occurrence of the
Lightning effects in the mesosphere 455
(km)
21
18
Negative 15
Intracloud cloud-to-ground
lightning lightning 12
(–15) (–20) 9
+50 +40 6
–50 –60
3
+10 +12.5
0
0 4 8 12 0 4 8 12 (km)
(a) (b) (km)
21
Positive 18
Low-altitude
intracloud lightning blue jet 15
12
(–20) (–20) 9
+40 +57.5 6
–45 –40
+15 +5 3
0
0 4 8 12 0 4 8 12 (km)
(c) (d)
Negative gigantic jet (km)
21
18
Negative (–3) 15
bolt-from-the-blue +82.5
12
(–30) 9
–120
+20 6
+30 +25
–90 3
0
0 4 8 12 0 6 12 18 (km)
(e) (f)
CG discharge [94]. This model consists of four space charges: 25, 120, 82.5, and
3 C distributed around the altitudes 4.3, 8.0, 13.2, and 15.4 km, respectively.
To study the QE field associated with the GJ events in a little more detail, the
simplest model was used, which assumes that these charges form four spherically
symmetric regions arranged one above the other along the vertical z-axis. The
spatial charges are uniformly distributed in these regions. The results of numerical
456 Lightning electromagnetics: Volume 2
20
3
18
16
14
12
, km
10
Ɀ
2
8
1
6
4
2
0
–300 –200 –100 0 100 200 300 400 500
EⱿ , kV/m
modeling of the vertical electrical field taken along z-axis versus altitude z are
shown in Figure 11.23 with solid line [60]. The dashed lines 13 indicate the
altitude dependence of the runaway breakdown threshold and minimum fields
required for propagation of positive and negative streamers in the air. The occur-
rence of GJ discharge is possible at those altitudes where the thundercloud QE field
exceeds one of these three threshold fields. Initiation of the GJ discharge due to the
runaway breakdown mechanism at low altitudes is very questionable because of the
lack of seed relativistic electrons and great length required for a relativistic ava-
lanche multiplication process (e.g., [79]). As is seen from Figure 11.23 the negative
GJ discharge can be initiated within the altitude range of 10–12 km where the
thunderstorm QE field is close to electric fields required for propagation of nega-
tive streamers in the air. The arrow indicates the most probable altitude for the GJ
initiation. The negative GJ once triggered can penetrate through the above narrow
layer with positive electric field (1015 km), and then escape the thundercloud
toward the ionosphere. So, the GJ once triggered starts to propagate as a normal IC
discharge between the main mid-level negative charge and a screening positive
upper-level charge, that continues to propagate upward out of the top of the thun-
dercloud [94].
The upward-propagating BJ and GJ discharges and normal CG lightning have
a few properties in common despite the BJ and GJ have the inverted structure and
other spatiotemporal scales. The BJ can be considered as an upward-propagating
Lightning effects in the mesosphere 457
positive leader with a streamer corona on the top in close analogy to usual CG
lightning [96] whereas the most of GJ manifest themselves during the LJ stage as
upward-propagating negative leaders (e.g., [91]). A great number of the short-lived
cold streamers produce a streamer corona of BJ and GJ discharges. In contrast to
CG discharge, the length and radius of individual streamers emitted from the hot
leader head and a typical size of the streamer corona increase with altitude because
of exponential decreasing air density.
In the model by Raizer et al. [97–99], the BJ/GJ originated from a bidirectional
leader starting at the height where the electric field is maximal. It is assumed that the
ascending leaders prevail over the descending ones due to the exponential decrease in
atmospheric pressure with altitude. The minimum field required for the streamer
propagation is given by equation similar to (11.2); that is, Es ¼ Es0 expðz=Ha Þ,
where Es0 is the critical field at sea level. Let hl be the altitude of the leader tip, which
plays a role in the streamer source. The individual streamer born at this altitude can
grow up to “infinity” if the leader tip has the potential
ð1
U¼ Es dz ¼ Es0 Ha expðhl =Ha Þ: (11.23)
hl
From here one may obtain an order-of-magnitude estimate of the escape alti-
tude hl . Such an upward-propagating streamer can reach to the lower ionosphere if
it starts from the altitude [98]
Es0 Ha
hl ¼ Ha ln : (11.24)
U
Using empirical data by van der Velde et al. [84] and by Neubert et al. [100],
Milikh et al. [101] have estimated the parameters in (11.24) and found that hl 42
km. This rough estimate is compatible with the GJ observations since the streamer
corona of the GJ start to grow up at the altitude range of 40–50 km (e.g., see
Figure 11.20). The streamer corona of the BJ discharge arises at lower altitudes and
therefore it terminates in the stratosphere before reaching the altitudes of the
ionosphere.
One of the challenges of the BJ and GJ research is to know enough about the
streamer-to-leader transition and initiation of lightning leader. Laboratory tests with
long sparks and numerical simulations have shown that the initial lightning leader
can result from the contraction of the streamer’s currents into a small radius channel
[39,102]. This contraction has been assumed to be due to the development of an
ionization-thermal instability that is well-known in the theory of a glow discharge
[40,103].
The numerical simulations of the initiation of lightning leader and streamer-to-
leader transition in the air have shown that the transition from a uniform discharge
to the contracted state occurs as the electric current in the discharge channel
exceeds the critical value which increases with the pressure decreasing [101]. This
means that the critical current required for the BJ/GJ discharge contradiction
increases with altitude. The BJ/GJ leader thus terminates at such an altitude where
the critical current becomes as high as the leader current. Taking into account that
458 Lightning electromagnetics: Volume 2
the observed altitude of GJ leader termination is about 48 km, and comparing this
altitude with that derived from their numerical calculation, Milikh et al. [101] have
estimated the critical current as 3.3 kA.
11.4 Elves
Divergent rings of brief optical emissions at the bottom of the ionosphere some-
times occurring immediately after intense return strokes of lightning discharges
were first observed from the Space Shuttle in 1991 [15]. These phenomena are
referred to as Elves that is an abbreviation for emission of light and VLF pertur-
bations due to EMP sources [16]. Generally, the Elves first arise at 90 km alti-
tude and then expand over 300700 km laterally and 1020 km in thickness for an
extremely short time of less than 0.1 ms [15,16,104,105]. Figure 11.24 illustrates an
example of Elves above a powerful thunderstorm in the Czech Republic taken from
the ground by a low-light video camera by Martin Popek.
It is generally accepted that the Elves are the visible manifestation of the
ionospheric response to a strong electromagnetic pulse (EMP) radiated by the CG
discharge current of either polarity (e.g., [104–106]). Elves are produced as a result
of heating and secondary ionization of the lower ionosphere by the EMP from a
lightning discharge [105,107]. The Elves predominantly radiate in red color basi-
cally due to the nitrogen fluorescence resulting from the molecules excitation by
the EMP. The short duration of this effect is basically due to a short lifetime of
excited states of molecules in the ionosphere.
The electromagnetic radiation pattern of a vertical lightning discharge is partly
similar to that of a vertical dipole antenna over a conducting ground. Due to that,
the lightning radiation intensity increases with increasing zenith angle. This prop-
erty of the radiation pattern can explain the fact that most observed elves have a
“doughnut” shape.
Elves are much more common than other types of TLEs. Analysis of data
gathered by the ISUAL instrument aboard the FORMOSAT-2 satellite has shown that
the global occurrence rate of the Elves can be estimated as 3.23 events per minute
while about 90% of elves happened over sea [86,108]. The data of the JEM-GLIMS
mission (Global LIghtning and sprite MeaSurements at Japanese Experimental
Module of ISS) gathered for the three-year observation period have shown that about
6.1% of lightning events are accompanied by Elves appearance [109].
The unusual Elve, which exhibits a distinct striped structure unlike the vast
majority of symmetric, torus-shaped “classic” Elves were first reported by Yue and
Lyons [110]. Figure 11.25 shows the event called a “tiger Elve”, which illustrates the
striations in the Elves luminosity observed using a high-speed camera system near Fort
Collins, Colorado on June 12, 2013 [110]. Simultaneous and independent observations
by a co-located color near infrared camera revealed a pattern of internal gravity wave
(IGW) in the airglow at the OH layer located at height of 85 km. Since the banded
structure in the Elve and the IGW roughly coincided in space, this suggested that the
Figure 11.25 A distinctly striated Elve was observed using an intensified high-speed
camera system at the Yucca Ridge Field Station near Fort Collins,
Colorado on June 12, 2013. Taken from Yue and Lyons [110]
460 Lightning electromagnetics: Volume 2
“tiger Elve” structure is due to the Elve interaction with a strong IGW propagating at
lower ionosphere altitudes. This effect can be explained by the fact that the ionization
rate in the D-region ionosphere is inversely proportional to the air density, which is
modulated by the IGWs. Numerical simulation of the lightning EMP interaction with
the ionospheric plasma has shown that the observable Elves striations at altitudes near
85 km can be generated by the IGW with a plasma density perturbation of as low as
5% [111]. So, a strong lightning discharge followed by Elves can make it visible the
IGW in the ionosphere.
About 4% of Elves events produced by strong lightning EMP are accompanied
by the so-called Elves doublets; that is, a pair of Elves occurred with a very small
time delay on the order of 100 ms [112,113]. The first Elve in the doublet is due to
EMP reflection from the ionosphere, whereas the second one is caused by triple
reflection; that is twice from the ionosphere and once from the Earth. This con-
clusion is supported by the facts that the time lag in the appearance of the second
Elves depends on the altitude of the discharge, the altitude of the ionospheric
reflection, and the distance to the camera.
cloudy areas although the most powerful flashes occurred more frequently over the
oceans at higher latitudes.
The total optical energy emitted by TAEs is estimated to be about 2550 kJ [115],
which is much less than the average optical emission energy of the sprites (300 kJ,
[116]) and lighting (1 MJ, [29]). Typical TAE may contain one burst of light or a series
of multiple bursts. The individual burst of light has a duration of about 0.1 ms or even
shorter while the multiple bursts can last for tens milliseconds. Conventionally, this
kind of faint flashes can be split into to two types: “luminous” and “dim” transients
depending on the number, Nph , of emitted photons [117]. The “dim” transients is a
portion of the short-term TAE with Nph < 1021 photons while the “luminous” tran-
sients are more intense event with Nph > 1023 photons and longer duration.
It is generally believed that the optical emission spectra of TAEs are due to the
fluorescence of the atmospheric nitrogen molecules, analogously to the emission
spectra of sprites and other mesospheric flashes, which are mainly composed of
the emission spectral lines of nitrogen molecules. Although the mechanism for the
fluorescence excitation seems to be distinct from that of TLEs. Analysis of the
“luminous” transients has shown that the photons number emitted in the UV range
was approximately three times greater than that of the red-IR range [114]. This
property is typical for the emission spectra of nitrogen molecules located at high
altitudes greater than 50 km [118]. Thus, it can be expected that the sources of the
“luminous” TAEs are located in the mesosphere.
It was hypothesized that the molecule fluorescence can be excited by electron
flows with energies on the order of tens of eV [19]. However, the reason for the
appearance of such electron flows in the atmosphere is a puzzle. Another model
explaining the TAE’s origin is based on assumption that large-scale areas with a low
space charge density and a low electric breakdown threshold can occur in the
mesosphere thereby producing electric discharges between the charged areas [115].
One possible cause for the existence of such charged areas at 5070 km altitudes is
the abnormally low air conductivity due to the presence of suspended dust particles
[119]. The decrease in air conductivity is caused by free electron and ion attachment
to both the dust particles and positive ion clusters, which build up from nitrogen ions
N þ due to their hydration with water molecules. In the model, the TAEs parameters
are best suited to the onboard observations as the charged areas are situated at alti-
tudes of 6070 km. The best-fit parameters are so that the charged areas of
1015 km in size have to contain a total charge of about 1 C. In such a case the
theory predicts that the optical energy radiated during a single electric discharge
between the charged areas can be on the order of tens kJ while the discharge duration
is about several milliseconds [115]. Despite these theoretical estimates being com-
patible with the observations much remains to be done in this area.
11.5.3 Terrestrial gamma-ray flashes
Another interesting phenomenon is the brief bursts of upward-directed X-ray and
gamma-ray radiation detected on board low-orbit satellites near active thunder-
storm. This phenomenon termed terrestrial gamma-ray flashes (TGFs) was first
observed in 1994 by the Burst and Transient Source Experiment (BATSE) on board
462 Lightning electromagnetics: Volume 2
leader just before a large current pulse heats up the leader channel and emits a main
optical pulse. In some instances, the TGF-produced lightning current had such a
high amplitude, a prerequisite for Elves, that both TGF and Elves were observed
simultaneously [150].
Recently Belz et al. [151] have reported the first high-resolution ground-based
observations of downward-directed TGFs, which were detected by the large-area
Telescope Array cosmic ray observatory in a proximity (34 km) to the lightning
flash. The TGFs consisting 510 ms duration bursts of gamma rays were observed
during strong initial breakdown pulses in the first few milliseconds of negative
cloud-to-ground and low-altitude IC flashes. The number of gamma photons pro-
duced by such TGF was estimated as high as 1012–1014; that is, several orders of
magnitude less than satellite-detected photon number [152].
We cannot come into detail about this interesting phenomenon since the
researches are still continuing. The interested reader is referred to a review by
Kumar and Pooja [153] and references herein for details.
References
[1] Franz, R. C., R. J. Nemzak, and J. R. Winkler (1990), Television image of a
large upward electrical discharge above a thunderstorm system, Science 249,
48–51.
[2] Wilson, C. T. R. (1925), The electric field of a thunderstorm and some of its
effects, Proc. R. Soc. Lond. 37, 32D.
[3] van der Velde, O. A. (2008), Morphology of sprites and conditions of sprite
and jet production in mesoscale thunderstorm systems, PhD thesis,
Toulouse, France, 202 pp.
[4] Neubert, T., M. Rycroft, T. Farges, et al. (2008), Recent results from studies
of electric discharges in the mesosphere, Surv. Geophys. 29(2), 71–137.
[5] Sentman, D. D., E. M. Wescott, D. L. Osborne, D. L. Hampton, and M. J.
Heavner (1995), Preliminary results from the Sprites94 campaign: Red
sprites, Geophys. Res. Lett. 22, 1205–1208.
[6] Boeck, W. L., O. H. Jr. Vaughan, R. J. Blakeslee, B. Vonnegut, M. Brook,
and J. McKune (1995), Observations of lighting in the stratosphere,
J. Geophys. Res. 100, 1465. doi:10.1029/94JD02432.
[7] Wescott, E. M., D. Sentman, D. Osborne, D. Hampton, and M. Heavner
(1995), Preliminary results from the Sprites94 aircraft campaign: 2. blue jets,
Geophys. Res. Lett. 22, 1209. doi:10.1029/95GL00582.
[8] Wescott, E. M., D. D. Sentman, H. Stenbaek-Nielsen, P. Huet, M. J.
Heavner, and D. R. Moudry (2001), New evidence for the brightness and
ionization of blue starters and blue jets, J. Geophys. Res. 106, 21549–21554.
[9] Wescott, E. M., D. D. Sentman, M. J. Heavner, D. L. Hampton, D. L. D.
Osborne, and Jr. O. H. Vaugham (1996), Blue starters: Brief upward discharges
from an intense Arkansas thunderstorm, Geophys. Res. Lett. 23, 2153–2156.
464 Lightning electromagnetics: Volume 2
[24] Bór, J., Z. Zelkó, T, Hegedüs, et al. (2018), On the series of +CG lightning
strokes in dancing sprite events, J. Geophys. Res. Atmos. 123, 11030–11047.
https://doi.org/10.1029/2017JD028251.
[25] Lyons, W. A (1996), Sprite observations above the U.S. High Plains in
relation to their parent thunderstorm systems, J. Geophys. Res. 101(D23),
29641–29652.
[26] Soula, S., J. Mlynarczyk, M. Füllekrug, et al. (2017), Dancing sprites:
Detailed analysis of two case studies, J. Geophys. Res. Atmos. 122(6),
3173–3192.
[27] Yang, J., G. Lu, L.-J. Lee, and G. Feng (2015), Long-delayed bright dancing
sprite with large horizontal displacement from its parent flash. J. Atmos.
Solar-Terr. Phys. 129, 1–5. https://doi.org/10.1016/j.jastp.2015.04.00.
[28] Nickolaenko, A. P. and M. Hayakawa (2002), Resonances in the Earth-
Ionosphere Cavity. Kluwer Acad., Dordrecht, Netherlands.
[29] Rakov, V. A. and M. A. Uman (2003), Lightning: Physics and Effects.
Cambridge University Press, Cambridge, UK.
[30] Boccippio, D. J., E. R. Williams, S. J. Heckman, W. A. Lyons, I. T. Baker,
and R. Boldi (1995), Sprites, ELF transients, and positive ground strokes,
Science 269, 1088–1091.
[31] Williams, E., E. Downes, R. Boldi, W. Lyons, and S. Heckman (2007),
Polarity asymmetry of sprite-producing lightning: a paradox? Radio Sci. 42,
RS2S17. https://doi.org/10.1029/2006RS003488.
[32] Lu, G., S. A. Cummer, J. Li, et al. (2013), Coordinated observations of
sprites and in-cloud lightning flash structure, J. Geophys. Res. Atmos. 118,
6607–6632. https://doi.org/10.1002/jgrd.50459.
[33] Matsudo, Y., T. Suzuki, K. Michimoto, K. Myokei, and M. Hayakawa
(2009), Comparison of time delay of sprites induced by winter lightning
flashes in the Japan Sea with those in the Pacific Ocean, J. Atmos. Solar-
Terr. Phys. 71, 101–111.
[34] Cummer, S. A., W. A. Lyons, and M. A. Stanley (2013), Three years of
lightning impulse charge moment change measurements in the United
States, J. Geophys. Res. Atmos. 118, 5176–5189. https://doi.org/10.1002/
jgrd.50442.
[35] Li. J., S. Cummer, G. Lu, and L. Zigoneanu (2012), Charge moment change
and lightning-driven electric fields associated with negative sprites and halos,
J. Geophys. Res. 117, A09310. https://doi.org/10.1029/2012JA017731.
[36] Chen, B.-C., H. Chen, C. W. Chuang, et al. (2019), On negative sprites and
the polarity paradox, Geophys. Res. Lett. 46, 9370–9378. https://doi.org/
10.1029/2019GL0838 04.
[37] Cummer, S. A. and M. Füllekrug (2001), Unusually intense CC in lightning
produces delayed mesospheric breakdown, Geophys. Res. Lett. 28, 495–498.
[38] Stanley, M., M. Brook, P. Krehbiel, and S. Cummer (2000), Detection of
daytime sprites via a unique sprite ELF signature, Geophys. Res. Lett. 27,
871–874.
[39] Raizer, Y. P. (1991), Gas Discharge Physics. Springer, Berlin.
466 Lightning electromagnetics: Volume 2
[40] Bazelyan, E. M. and Y. P. Raizer (1998), Spark Discharge. CRC Press, Boca
Raton, FL.
[41] Pasko, V. P. (2006), Theoretical modeling of sprites and jets, In: M.
Füllekrug, E. A. Mareev, and M. J. Rycroft (eds), Sprites, Eves and Intense
Lightning Discharges, NATO Science Series II: Mathematics, Physics and
Chemistry, vol 225. Springer, Heidleberg, pp 253–311.
[42] Cummer, S. A., N. Jaugey, J. Li, W. A. Lyons, T. E, Nelson, and E. A. Gerken
(2006b), Submillisecond imaging of sprite development and structure,
Geophys. Res. Lett. 33, L04104. https://doi.org/10.1029/2005G L0249 69.
[43] Qin, J., V. P. Pasko, M. G. McHarg, and H. C. Stenbaek-Nielsen (2014), Plasma
irregularities in the D-region ionosphere in association with sprite streamer
initiation, Nat. Commun. 5, 1–6. https://doi.org/10.1038/ncomms4740.
[44] Stanley, M., P. Krehbiel, M. Brook, C. Moore, W. Rison, and B. Abrahams
(1999), High speed video of initial sprite development, Geophys. Res. Lett.
26, 3201–3204.
[45] McHarg, M. G., R. K. Haaland, D. Moudry, and H. C. Stenbaek-Nielsen
(2002), Altitude-time development of sprites, J. Geophys. Res. 107(A11),
1364. https://doi.org/10.1029/2001JA000283.
[46] McHarg, M. G., H. C. Stenbaek-Nielsen, and T. Kammae (2007),
Observations of streamer formation in sprites, Geophys. Res. Lett. 34,
L06804. https://doi.org/10.1029/2006GL027854.
[47] Li, J. and S. A. Cummer (2009), Measurement of sprite streamer acceleration
and deceleration, Geophys. Res. Lett. 36, L10812. https://doi.org/10.1029/
2009GL037581.
[48] Lyons, W. A., M. Stanley, T. E. Nelson, and M. Taylor (2000), Sprites,
elves, halos, trolls, and blue starters above the STEPS domain, EOS Trans.
AGU, 81(48), F131.
[49] da Silva, C. L. and V. P. Pasko (2013), Dynamics of streamer-to-leader
transition at reduced air densities and its implications for propagation of
lightning leaders and gigantic jets, J. Geophys. Res. Atmos. 118, 13561–
13590. https ://doi.org/10.1002/2013J D0206 18.
[50] Kelley, M. C. (1989), The Earth’s Ionosphere. Academic, New York, NY.
[51] Luque, A. and U. Ebert (2009), Emergence of sprite streamers from
screening-ionization waves in the lower ionosphere, Nat. Geosci. 2(11),
757–760. doi:10.1038/ngeo662.
[52] Surkov, V. V. and M. Hayakawa (2020), Progress in the study of transient
luminous and atmospheric events: a review, Surveys Geophys. 41(5), 1101–
1142. doi:10.1007/s10712-020-09597-2.
[53] Qin, J., S. Celestin, and V. P. Pasko (2011), Onthe inception of streamers
from sprite halo events produced by lightning discharges with positive and
negative polarity, J. Geophys. Res. 116, A06305. https://doi.org/10.1029/
2010JA016366.
[54] Surkov, V. V. and M. Hayakawa (2016), Semi-analytical models of the
sprite generation from plasma heterogeneities, Geomag. Aeron. 56(6),
724–732. https://doi.org/10.1134/S0016793216050145.
Lightning effects in the mesosphere 467
[55] Luque, A. and U. Ebert (2010), Sprites in varying air density: Charge con-
servation, glowing negative trails and changing velocity, Geophys. Res. Lett.
37, L06806. https://doi.org/10.1029/2009G L0419 82.
[56] Pasko, V. P., J. Qin, and S. Celestin (2013), Toward better understanding of
sprite streamers: Initiation, morphology, and polarity asymmetry, Surv.
Geophys. 34(6), 797–830. https://doi.org/10.1007/s10712-013-9246-y.
[57] Qin, J. and V. P. Pasko (2015), Dynamics of sprite streamers in varying air
density, Geophys. Res. Lett. 42, 2031–2036. https://doi.org/10.1002/
2015GL063269.
[58] Pasko, V. P. (2007), Red sprite discharges in the atmosphere at high altitude: The
molecular physics and the similarity with laboratory discharges, Plasma Sources
Sci. Technol. 16, S13–S29. https://doi.org/10.1088/0963-0252/16/1/S02.
[59] Gerken, E. A., U. S. Inan, and C. P. Barrington-Leigh (2000), Telescopic
imaging of sprites, Geophys. Res. Lett. 27(17), 2637–2640.
[60] Surkov, V. V. and M. Hayakawa, Underlying mechanisms of transient
luminous events: A review, Ann. Geophys., 30, 1185–2012, 2012.
doi:10.5194/angeo-30-1185-2012.
[61] Kanmae, T., H. C. Stenbaek-Nielsen, M. G. McHarg, and R. K. Haaland
(2012), Diameter-speed relation of sprite streamers, J. Phys. D, Appl. Phys.
45, 275203. https://doi.org/10.1088/0022-3727/45/27/275203.
[62] Sentman, D. D., E. M. Wescott, M. Heavner, and D. Moudry, Observations
of sprite beads and balls, Eos. Trans. AGU, 77(46), F61, 1996.
[63] Luque, A., H. C. Stenbaek-Nielsen, M. G. McHarg, and R. K. Haaland (2016),
Sprite beads and glows arising from the attachment instability in streamer chan-
nels, J. Geophys. Res. Space Phys. 121, 2431–2449. doi:10.1002/2015JA022234.
[64] van der Velde, O. A., Á. Mika, S. Soula, C. Haldoupis, T. Neubert, and U. S.
Inan (2006), Observations of the relationship between sprite morphology and
in-cloud lightning processes, J. Geophys. Res. 111, D15203. https://doi.org/
10.1029/2005J D0068 79.
[65] Cummer, S. A., U. S. Inan, T. F. Bell, and C. P. Barrington-Leigh (1998),
ELF radiation produced by electrical currents in sprites, Geophys. Res. Lett.
25, 1281–1284.
[66] Cummer, S. A., H. U. Frey, S. B. Mende, et al. (2006a), Simultaneous radio and
satellite optical measurements of high-altitude sprite current and lightning
continuing current, J. Geophys. Res. 111, A10315. doi:10.1029/2006JA011809.
[67] Surkov, V. V., Y. Matsudo, M. Hayakawa, and S. V. Goncharov (2010),
Estimation of lightning and sprite parameters based on observation of sprite-
producing lightning power spectra, J. Atmos. Solar-Terr. Phys. 72, 448–456.
https://doi.org/10.1016/j.jastp.2010.01.001.
[68] Surkov, V. and M. Hayakawa (2014), Ultra and extremely low frequency
electromagnetic fields, in Springer Geophysics Series, XVI, Springer, New
York, NY, 486 p.
[69] Hobara, Y., N. Iwasaki, T. Hayashida, M. Hayakawa, K. Ohta, and H.
Fukunishi (2001), Interrelation between ELF transients and ionospheric
468 Lightning electromagnetics: Volume 2
disturbances in association with sprites and elves, Geophys. Res. Lett. 28(5),
935–938. https ://doi.org/10.1029/2000GL003795.
[70] Inan, U. S., S. A. Cummer, and R. A. Marshall (2010), A survey of ELF and
VLF research on lightning-ionosphere interactions and causative discharges,
J. Geophys. Res. 115, A00E36. doi:10.1029/2009JA014775.
[71] Otsuyama, T., J. Manaba, M. Hayakawa, and M. Nishimura (2004),
Characteristics of subionospheric VLF perturbation associated with winter
lightning around Japan, Geophys. Res. Lett. 31, L04117. https://doi.org/
10.1029/2003GL019064.
[72] Helliwell, R., J. P. Katsufrakis, and M. Trimpi (1973), Whistler-induced
amplitude perturbation in VLF propagation, J. Geophys. Res. 78, 4679–4688.
[73] Inan, U. S., T. F. Bell, V. P. Pasko, D. D. Sentman, E. M. Wescott, and
W. A. Lyons (1995), VLF signatures of ionospheric disturbances associated
with sprites, Geophys. Res. Lett. 22, 3461–3464. https://doi.org/10.1029/
95GL0 3507.
[74] Haldoupis, C., R. J. Steiner, Á. Mika, S. et al. (2006), ‘‘Early/slow’’ events:
A new category of VLF perturbations observed in relation with sprites,
J. Geophys. Res. 111, A11321. doi:10.1029/2006JA011960.
[75] Haldoupis, C., N. Amvrosiadi, B. R. T. Cotts, O. A. van der Velde, O.
Chanrion, and T. Neubert (2010), More evidence for a one-to-one correlation
between Sprites and Early VLF perturbations, J. Geophys. Res., 115,
A07304, https://doi.org/10.1029/2009JA015165.
[76] Neubert, T., T. H. Allin, E. Blanc, et al. (2005), Co-ordinated observations
of transient luminous events during the Eurosprite 2003 campaign, J. Atmos.
Solar-Terr. Phys. 67, 807–820. https://doi.org/10.1016/j.jastp.2005.02.004.
[77] Haldoupis, C., Á. Mika, and S. Shalimov (2009), Modeling the relaxation of
early VLF perturbations associated with transient luminous events, J.
Geophys. Res., 114, A00E04, doi:10.1029/2009JA014313.
[78] Glukhov, V., V. Pasko, and U. Inan, Relaxation of transient lower iono-
spheric disturbances caused by lightning-whistler-induced electron pre-
cipitation bursts, J. Geophys. Res., 97, 16971–16979, 1992.
[79] Pasko, V. P. (2010), Recent advances in theory of transient luminous events,
J. Geophys. Res. 50(6), A00E35. https://doi.org/10.1029/2009JA014860.
[80] Kuo, C. L., H.-T. Su, and R.-R. Hsu (2015), The blue luminous events
observed by ISUAL payload on board FORMOSAT-2 satellite. J. Geophys.
Res. Space Phys., 120, 9795–9804. doi:10.1002/2015JA021386.
[81] Su, H. T., R. R. Hsu, A. B. Chen, et al. (2003), Gigantic jets between a
thundercloud and the ionosphere, Nature 423, 974–976. https://doi.org/
10.1038/nature01759.
[82] van der Velde, O. A., W. A. Lyons, T. E. Nelson, S. A. Cummer, J. Li, and
J. Bunnell (2007), Analysis of the first gigantic jet recorded over continental
North America, J. Geophys. Res. 112, D20104. https://doi.org/10.1029/
2007JD008575.
Lightning effects in the mesosphere 469
This chapter will present possible effects of atmospheric lightning on the upper
atmosphere such as the ionosphere and magnetosphere composed of ionized plas-
mas. Though there exist several phenomena on the effect of lightning discharges
onto the ionosphere/magnetosphere, we introduce only two major attractive topics:
(1) lightning-induced whistlers in the ionosphere/magnetosphere, and (2) iono-
spheric Alfvén resonator (IAR) in an altitude region between the lowest ionosphere
and lower magnetosphere, where one likely candidate of its source is lightning
discharges. The former is quite a well-known phenomenon, and whistlers are bursts
of ELF/VLF waves produced by lightning discharges. Some part of VLF/ELF
lightning energy penetrates through the ionosphere, propagates along the magnetic
field line in the magnetosphere, and penetrates again through the ionosphere in the
opposite hemisphere, followed by reception on the ground as a whistler. We show
initially the phenomena of ground-based whistlers, their brief theoretical explana-
tion, and their use in the diagnostics of ionospheric/magnetospheric electron den-
sity. Also, earlier satellite observations of nonducted whistlers are presented.
Further, we will describe recent satellite observations of short-fractional hop
whistlers and VLF/ELF electromagnetic waves, with special reference to their use
in the study of global lightning activity. On the other hand, the latter phenomenon,
IAR in the ULF/ELF band is a rather new subject as compared with whistler stu-
dies, and so we pay more emphasis on IAR in this chapter. IARs exhibit an inter-
esting feature of fingerprint resonance structures on the dynamic spectra. This IAR
is apparently considered to be a kind of resonance of Alfvén waves in a region
between the lowest ionosphere and lower magnetosphere, whose resonance fre-
quencies (f = 1-10 Hz) are lower than the well-known Schumann resonances. We
1
Advanced Wireless and Communications Research Center, The University of Electro-Communications
(UEC), Japan
2
Hayakawa Institute of Seismo Electromagnetics, Co., Ltd. (Hi-SEM), Japan
3
Graduate School of Informatics and Network Engineering, The University of Electro-Communications
(UEC), Japan
4
UEC, Center for Space Science and Radio Engineering, Japan
476 Lightning electromagnetics: Volume 2
12.1 Introduction
The physics and effects of lightning discharges taking place in the atmosphere are
extensively discussed in most chapters of this monograph, and hence this chapter will
deal with the consequence of atmospheric lightning onto the upper atmosphere such
as the ionospheric/magnetospheric plasma. There are various possible effects of
lightning on the ionosphere/magnetosphere, but we restrict our attention to the fol-
lowing two interesting phenomena: (1) lightning-induced whistlers in Section 12.2
and (2) Section 12.3 concerned with ionospheric Alfvén resonator (IAR) where one
possible source is either distant or nearby lightning discharges.
In Section 12.2, we will discuss the first part of whistlers. Section 12.2.1 pro-
vides you with the history and general description of whistlers. In early times
whistlers are generally observed on ground-based stations, so they are defined by
bursts of ELF (extremely low frequency, f < 3 kHz)/VLF (very low frequency,
3 kHz < f < 30 kHz) waves, originated in lightning discharges in the opposite
hemisphere of the receiver. While the bulk of the energy from the causative light-
ning discharges propagates in the Earth-ionosphere waveguide as atmospherics (or
sferics) [1–5], some part of its energy is known to penetrate through the upper
ionosphere, propagate in the magnetosphere, probably along the magnetic field line
(ducted propagation), and penetrate again through the ionosphere in the opposite
hemisphere, to be received as a whistler. On the other hand, most whistlers
observed on board rockets and satellites (either deep in the magnetosphere or in the
ionosphere) are considered to be propagating in the nonducted mode, so these
nonducted whistlers have never been detected on the ground. There have already
been published several excellent books and reviews on the topic of whistlers [6–11].
Section 12.2.2 deals with the general theoretical background of plasma waves
including whistler mode and also Alfvén mode (this will be the topic of IAR in
Section 12.3). Particularly in Section 12.2.2.4, we will explain the whistler disper-
sion (D) as the most useful parameter of ground-based whistlers. Section 12.2.3 is
aimed at the use of whistlers as a diagnostic tool. Section 12.2.3.1 deals with the
diagnosis of magnetospheric electron density profile with the use of ground-based
whistler dispersions at different latitudes, and Section 12.2.3.2 will be concerned
with early satellite observations of various kinds of nonducted whistlers.
Section 12.2.3.3 will present recent LEO (low-Earth-orbit) satellite observations of
short-fractional hop whistlers and VLF/ELF electromagnetic waves (both wave-
form and power spectra data) to be utilized to study the global lightning activity and
lightning characteristics.
Whistlers and ionospheric Alfvén resonator 477
In Section 12.3, we will consider IARs. As compared with whistler studies, this
phenomenon of IAR in the ULF (ultra low frequency, f <10 Hz)/ELF range is
rather newly recognized as being important even in space physics [11,12]. Though
there is little doubt that IAR is a near-Earth resonance of Alfvén waves in the
altitude region between the lowest ionosphere and the lower magnetosphere, its
generation mechanism, or to be more exact, energy source or origin of IARs is
extremely poorly understood, and hence many scientists have recently paid a lot of
attention to the elucidation of this IAR phenomenon. Section 12.3.1 provides you
with the history and general description of IARs. Then, as a reference or a basis of
IAR signatures for the readers to understand what IAR looks like, we will present in
Section 12.3.2 our own statistical results of spectral resonance structures (SRSs)
and morphological features (local time dependence, seasonal variation, and
dependence on geomagnetic activity) of IARs based on a long-term (2.5 years)
observation at middle latitudes. Also, we make a brief comparison with the
observational results at high and low latitudes. In Section 12.3.3, we will discuss
the proposed generation mechanisms of IARs with special reference to geomag-
netic latitude. Generally speaking, it seems likely that magnetospheric influence
might be involved in IAR excitation at high and auroral latitudes, but the more
plausible source at middle and low latitudes might be lightning discharges. Based
on those discussions in Section 12.3.4, we will introduce our own idea of the major
role of nearby lightning discharges in exciting IARs in middle and low latitudes,
but it is needless to say that we need further appreciable efforts in the future before
a definite mechanism has been established.
the first evidence that ionized plasma was present far beyond the ionospheric
F layer. Hence, the use of whistlers opened a new era for space physics. There is a
sizable literature on whistles at different latitudes, including Helliwell [9], Walker
[7], Hayakawa and Tanaka [18], Park [10], Al’pert [19], Hayakawa and Ohta [20],
and Hayakawa [8].
Whistlers are burst electromagnetic signals with a frequency spectrum in the
range from about 100 Hz to over 10 kHz, which are originated in lightning dis-
charges in the atmosphere, penetrate into the ionosphere, and then propagate in
whistler mode in the magnetosphere approximately along the Earth’s magnetic
field lines to the opposite hemisphere. There the waves again traverse the iono-
sphere, this time in the downward direction and are observable in the conjugate
point on the Earth’s surface. The main frequency range of whistlers is ELF and
VLF, which are less than the local electron plasma frequency and the local gyro-
frequency (the definition of these characteristic frequencies will be given later), and
this frequency range is called “whistler-mode” branch. Propagation along the
Earth’s magnetic field is made possible by trapping waves in field-aligned density
irregularities (so-called ducts) (see Helliwell [9] and Hayakawa [8] for better
understanding of the physical mechanism of ducting), so that these whistlers are
named ducted whistlers as in Figure 12.1(a) and (c). Most of the ground-based
whistlers are considered to be ducted. The physics of whistler-mode propagation is
such that higher frequencies in the signal travel faster than lower ones, and hence
the higher frequencies arrive at an observing point on the Earth’s surface in the
conjugate point, earlier than the lower frequencies (this dispersion principle will be
given later). This is the reason why we observe a descending tone (as a short
(1 hop) whistler) as shown in Figure 12.1(b). Sometimes echo train whistlers are
observed (3-hop and 5-hop whistlers of a short whistler). If the whistler’s electro-
magnetic signal is converted to an audio signal, the audio signal will sound like a
whistle, and this is the reason why we call it a whistler. A less commonly observed
type of ducted whistler only at higher latitudes is the so-called “nose” whistler [21]
as shown in Figure 12.2, for which there is a minimum propagation time, corre-
sponding to a maximum propagation speed, at one “nose” frequency, the fre-
quencies below and above that value traveling slower and arriving increasingly
later. The more commonly observed whistler as in Figure 12.1(b) represents simply
the lower-frequency fraction of the nose whistler as shown in Figure 12.2(a).
Apparently, only a small fraction of lightning flashes produce ducted whistlers, but
probably a majority of flashes produce electromagnetic signals that propagate
through the ionosphere and then propagate in the magnetosphere both across and
along the magnetic field lines as “nonducted” whistlers, traveling in the magneto-
sphere until their energy is dissipated as in Figure 12.1(d). It follows that non-
ducted whistlers can only be observed on board satellites in the magnetosphere and
not on the Earth’s surface. Diagrams showing idealized paths taken by ducted and
nonducted whistlers are presented in Figure 12.1(c) and (d). An idealized time-
domain whistler spectrum is shown in Figure 12.3(a), and idealized plots of
frequency versus time in Figure 12.2(b). The square root of inverse frequency
versus time is plotted in Figure 12.3(c).
Whistlers and ionospheric Alfvén resonator 479
Northern hemisphere
1 3 5
Duct f
Dipole equator t
(c) (d)
Figure 12.1 (a) Typical path of a ducted whistler within the plasmasphere, shown
together with the waveform of the ducted whistler signal before, during,
and after its interaction with cyclotron-resonant electrons in the
magnetospheric equatorial region. (b) Frequency versus time curves
(dynamic spectra) for a whistler traversing a path that echoes from
hemisphere to hemisphere along a geomagnetic field as observed at
conjugate points in the northern and southern hemispheres. The causative
sferic is also indicated: it was produced by a lightning discharge in the
southern hemisphere. Potential ray paths (solid lines) for ducted (c) and
nonducted (d) whistler-mode propagation in the magnetosphere.
Adapted from Rakov and Uman [6]
1.0
Low-frequency θ = 0˚
0.8 approximation
High-frequency
0.6 approximation
f/fbe
0.4
Nose frequency
0.2
0
0 1 2 3 4 5 6 7
Time delay, s
(a)
Frequency, kHz
0
2000:03 UT 0.04 0.05 0.06 0.07
Time
(b)
Figure 12.2 (a) Normalized frequency versus time delay for longitudinal
propagation (q = 0 ) over a path of length c/B0. (b) An example of
observed spectrograms of nose whistlers. The left-hand arrow
indicates the causative sferic for the two whistler groups indicated by
the right-hand two arrows. Adapted from Hayakawa [8]
Above the E region is the F region, which extends to a height of 1,000 km or so.
The peak (maximum) F-region electron density has an average value of about
2 1012 m3 (2 106 cm3) during the day and 2 1011 m3 (2 105 cm3) at
night. As shown in Figure 12.4, the ionospheric electron density varies over four orders
of magnitude.
Above the F region of the ionosphere, at heights over about 1,000 km, is called
the “magnetosphere,” where the motions of plasmas (electrons and ions) are
strongly influenced by the Earth’s magnetic field [19,24]. The inner part of the
magnetosphere is called the plasmasphere. It extends outward by four or five
Earth’s radii to the plasmapause, where the plasma density decreases abruptly [25].
Atmospheric Whistler
(a)
6,000
4,000
f , Hz
2,000
0
(b)
0.05
0.04
f –1/2, Hz –1/2
0.03
0.02
0.01
0
0 0.5 1.0 1.5 2.0 2.5
Time, s
(c)
1,000
500
F
Height, km
Night Day
200
E
150
50
108 109 1010 1011 1012 1013
Electron density, m–3
Ginzburg [28], Budden [29], Al’pert [19], Parks [30], Nicholson [31], Walker [32], and
Kivelson and Russell [24], but we will make its brief review below.
We offer a general theoretical overview of the interaction of electromagnetic
waves with the charged particles of the ionosphere and magnetosphere. We con-
sider an electron having electric charge –e (e > 0) and mass m which moves with
velocity v. Only the motion of electrons can be considered because the mass of
electrons is significantly lighter than that of ions (e.g. mass of protons (the lightest
ion) is about 2,000 times heavier than that of electrons). The electrons are influ-
enced by the wave electric and magnetic fields as well as the DC Earth’s magnetic
field (B0) during the propagation of electromagnetic waves and the currents gen-
erated by the electrons re-generate electromagnetic waves. For electrons, the
equation of motion in the Cartesian coordinate is given by
dv
m ¼ eðE þ v ðB0 þ BÞÞ (12.1)
dt
The electromagnetic electric and magnetic field vectors are E and B, and the
ambient Earth’s magnetic field is B0 = B0az (az is the unit vector directed towards
positive z direction) as in Figure 12.5. Here collisions between electrons and heavy
neutral particles are ignored. We are also justified in neglecting the force on the
electron from the time-varying magnetic field (v B). The force due to the time-
varying magnetic field is generally negligible in comparison with the force due to
the electric field since the ratio of these forces is roughly v/c and v/c « 1 where c is
the speed of light. By assuming the time dependence of e jwt ( j, w, and t are ima-
ginary unit, angular frequency (w = 2p f, f frequency), and time respectively). The
equation of motion of electrons will be written as follows.
2 3 2 3 2 3
uc Ec uy
e eB0 4
jw4 uy 5 ¼ 4 Ey 5 uc 5 (12.2)
m m
uz Ez 0
z
B0
k, vp
θ
Figure 12.5 Coordinate system for the wave propagation. The Earth’s magnetic
field B0 is in the z direction, and the wave propagation (k) is in the xz
plane and makes an angle q with B0.
Whistlers and ionospheric Alfvén resonator 483
Then we obtain the three components of electron velocities driven by the wave
electric field and ambient Earth’s magnetic field are,
e jw Ec wH Ey
nc ¼
m wH 2 w2
e wH Ec þ jw Ey
ny ¼
m wH 2 w2
e Ez
nz ¼ (12.3)
m jw
where wH is the electron angular gyrofrequency given by wH (= 2pfH) = eB0/m,
which is easily calculated by the simplified form fH(Hz) = 28B0(nT). This gyrofre-
quency means the angular frequency at which the electron makes circular orbits in a
plane perpendicular to the Earth’s magnetic field. The current density is given by,
J ¼ Nev (12.4)
where N is the electron density. Each component of this current is shown by using
(12.4).
wp 2 wH
exy ¼ eyx ¼ je0 (12.7)
wðwH 2 w2 Þ
wp 2
ezz ¼ e0 1
w2
exz ¼ ezx ¼ eyz ¼ ezy ¼ 0
484 Lightning electromagnetics: Volume 2
where hIi is a unit tensor and the electron plasma frequency is defined by
w2p ¼ ð2pf p Þ2 Þ ¼ N e2 =me0 (e0, dielectric constant of free space) ðf p ðkHzÞ ¼
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
9 N ðcm3 ÞÞ:
We consider now Maxwell’s equations for the case of time-harmonic electro-
magnetic fields that vary in space as,
exp½jðwt k rÞ (12.9)
Here we have used the time dependence assumed previously and now include a
spatial variation specified by the term of –j k r, where k is the propagation vector
(or wave normal direction) (its direction is the direction of wave propagation and its
magnitude k equals 2p divided by the wavelength) and r is the position vector.
Figure 12.5 is the coordinate system of our propagation study, in which the static
Earth’s magnetic field B0 is already assumed to be in the z-axis and the wave
propagation (k) is assumed to be in the x-z plane and to make an angle (q) with
B0. For an electromagnetic wave with temporal and spatial variation given by (12.9),
we can write Maxwell’s equations as (m0, magnetic permeability of free space)
jk E ¼ jwmH (12.10)
jk H ¼ jwD (12.11)
Combining the above two equations, yields to,
The phase velocity vp is defined as the direction normal to the wave front, so
that it is easily understood as being parallel to k. On the other hand, the group
velocity vg is defined as the direction of electromagnetic wave energy (ray direc-
tion) and the most interesting peculiarity of the anisotropic plasma is that vg is not
parallel to vp for the general oblique propagation (consult Helliwell [9], Park [10],
and Hayakawa [8] for further details). Only when q = 0, the direction of vg is
parallel to vp and the value of vg is easily given by vg ¼ @ðnwÞ=@w (n: refractive
index).
This wave mode is so-called “whistler” mode when the wave frequency
is below wH.
(ii) Left-handed polarized wave
Similarly taking the second term on the left-hand side of (12.17) being zero,
the dispersion relation for this wave is,
k 2 c2 wp 2
n2 ¼ ¼1þ ¼L (12.20)
w 2 wðwH þ wÞ
and this wave is found to be left-handed circularly polarized as is evidenced by
the following relation.
Ex
j ¼1 (12.21)
Ey
wp 2
n2 ¼ 1 ¼P (12.24)
w2
This wave is linearly polarized and (12.24) is identical to the dispersion relation of
the plasma oscillation without any effect of the Earth’s magnetic field (12.16).
This is the reason why we use the terminology of “ordinary” wave.
(ii) Extraordinary wave (X wave) (k ?B0 and E ?B0 )
In order to have non-trivial solutions (Ex 6¼ 0, Ey 6¼ 0) in (12.22), the
determinant of the matrix on the left-hand side of (12.22) is set to zero, which
leads to the following dispersion relation.
Ex exy wp 2 wH
¼ ¼ j (12.26)
Ey exx wðwp 2 þ wH 2 w2 Þ
so that the wave electric field is elliptically polarized in the xy plane.
(c) Cutoffs and resonances
Cutoff is defined as a frequency at which the value k2 goes to zero (or
n2 ! 0), so these cutoff frequencies can be obtained by setting the refractive
index n2 ! 0 (or k2 ! 0). Hence we expect that the wavelength and phase
velocity becomes infinity. When the wave propagates through the plasma with
a spatial gradient of refractive index including n2 = 0, the wave is reflected at
the point of cutoff.
The cutoff frequencies can be easily estimated by setting n2 ! 0 in the dis-
persion relation for each mode. The cutoff frequencies for the longitudinal
propagation can be obtained from (12.18) and (12.20):
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
wH wH 2
wR ¼ þ wp 2 þ (12.27)
2 4
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
wH wH 2
wL ¼ þ wp 2 þ (12.28)
2 4
The cutoff frequencies for transverse propagation are obtained from (12.24)
and (12.25). For the extraordinary wave, the cutoff frequencies occur when
n2! 0 (or R = 0 or L = 0), so that they are identical to those for the longitudinal
propagation, wR and wL. For the ordinary wave, the cutoff frequency is
expected from n2 = P = 0; that is, wp.
Next, we move on to the explanation of resonances. Resonance is defined at a
frequency at which n2 = ? (k2 ! ?). When the wave approaches a resonance
condition, k2 goes to infinity so that the phase velocity becomes zero, then the
wave is mainly dissipated due to the significant interaction with plasma.
The resonance frequencies for the longitudinal propagation are obtained
when the denominator of (12.18) and (12.20) vanishes. That is, the right-
handed polarized (R) mode exhibits a resonance at w = wH, and at this fre-
quency, the wave electric field continuously accelerates electrons because the
wave field is in phase with the gyro-motion of electrons. This resonance is
called electron cyclotron resonance. Also, the left-hand polarized (L) mode is
found to have a resonance at w = wHi (ion gyrofrequency). Though in our
former equations we have only considered the motion of electrons, the
motion of ions becomes important when the frequency becomes smaller than
wHi and it is very easy for us to take into account the effect of ions in the
dispersion relation.
The resonance for the transverse propagation can be obtained exactly in the same
way as above. The ordinary wave has no resonance because the denominator of
488 Lightning electromagnetics: Volume 2
(12.24) may not vanish except w = 0. On the other hand, the extraordinary wave
(12.25) exhibits a resonance at a particular frequency of wUHR (upper hybrid
resonance frequency), so that the wave electromagnetic energy is converted to the
upper hybrid oscillation. A similar hybrid resonance is observed as well at wLHR
(lower hybrid resonance frequency).
(d) Oblique propagation
We have discussed two representative cases of propagation, that is, longitudinal (q = 0)
and transverse (q = p/2), which gives us a lot of useful ideas on the propagation
characteristics of electromagnetic waves in the anisotropic plasma. Here, the dispersion
relation of the waves propagating obliquely to B0 is obtained by solving the following
equation by setting the determinant of the coefficients of E in (12.13) to zero.
ω
n2 = R
║ n 2 = 2RL/ (R + L)
n2 = L
ωR ┴ n 2 = 2RL/ (R + L)
ωUHR
n2 = P
┴ ║ ┴
ωPe
║ ┴
ωL
whistler–mode
ωHe
n2 = R
║ ωLHR
┴ n 2 = 2RL/ (R + L)
║ n2 = L
ωHi
Figure 12.6 Dispersion curve in the form of w-k diagram for an electron-ion
plasma for the case of wp = 2wH. k and ? mean the longitudinal
(q = 0) and transverse propagation (q = p/2), and the dispersion
curve for oblique propagation must lay in the shaded area. wR and
wL are the cutoff frequencies of right-and left-handed polarized mode
waves. While wUHR and wLHR are the lower and upper hybrid
resonance frequencies. n2 = R and n2 = L for q = 0 indicate the
right- and left-handed circularly polarized waves, respectively.
waves from lightning propagate into the ionosphere, we first examine the appli-
cation of Snell’s law to the boundary between the free space and ionosphere,
sin qi ¼ ni sin qt (12.34)
where ni is the refractive index of the ionosphere and that of free space is taken as
unity. And qi is the incidence angle measured from the vertical downward, and qt is
the transmitted angle in the ionosphere measured from the vertical upward. We will
show that ni »1 and hence qt in the ionosphere must be nearly zero; that is, essen-
tially vertical for any incidence angle (qi) of the wave in free space impinging on
the boundary. Using (12.18), we can have the whistler-mode refractive index to
good approximation as,
wp 2
ni 2 ¼ (12.35)
wðwH wÞ
This refractive index exceeds much that of free space. For example, at the
ionospheric F region where the electron density is near 106 cm3, ni is close to 100
at frequencies near 5 kHz. After the whistler enters the magnetosphere, propagates
along a field-aligned duct, and reaches the ionosphere in the conjugate hemisphere,
there happens the reverse interface situation. Snell’s law can also be involved to
show that if the wave’s normal direction is inside a relatively narrow angle around
the vertical (this is called transmission cone) [8,9], then it can penetrate through the
ionosphere and hence be detected on the ground. When the wave is ducted and also
we consider the high to middle latitudes where the Earth’s magnetic field is close to
the vertical, this condition is easily satisfied. However, if not (e.g., at low latitudes),
then the wave will be reflected (totally reflected) back into the magnetosphere,
sometimes leading to the generation of echo train whistlers.
On the assumption that the whistler wave is propagating along a magnetic field
line, the time of propagation T(w) from the source to the receiver, can be computed
from the knowledge of group velocity of the wave.
ð
ds
T ðwÞ ¼ (12.36)
path vg
greatest contribution comes from the region around the apex of the propagation
path because of the smallest wH there. This equation is known as the Eckersley’s
dispersion law and is valid at frequencies far below the nose frequency. The fact
that the dispersion D is independent of frequency is illustrated in Figure 12.3(c).
By utilizing the more general (12.18) for a given path, the integral in (12.36)
enables us to deduce a nose frequency fn for which the propagation time is a
minimum. For a homogeneous plasma, the nose frequency fn = fH/4. As seen from
(12.38) the ionosphere does not contribute much to the observed time delays; the
most important contribution is expected at the highest part of the path (equatorial
plane) of the magnetosphere.
103
102
Gringauz et al.
101
1 2 3 4 5
Geocentric distance (earth radii)
Figure 12.7 Electron density profile in the equatorial plane of the magnetosphere
(full line) with the use of whistlers. White circles indicate the electron
density observed by satellites. After Darrouzet et al. [36]
492 Lightning electromagnetics: Volume 2
ground results. Furthermore, we can monitor the dynamic behavior (spatial and
temporal variations) of the magnetospheric electron density profile including the
plasmapause structure [25,33,35,36].
source to the receiver, and (2) duct characteristics and duct formation mechanism.
The first topic was to discuss the one-to-one correspondence between a causative
lightning discharge and its whistler and then the detailed propagation mechanism in
the magnetosphere. This one-to-one correspondence was attempted by many sci-
entists, especially at low latitudes [40–42]. In these studies, they made full use of
the direction finding of locating causative sferics [43] and also the direction finding
to pinpoint the ionospheric exit regions of short whistlers [44–47]. As for the sec-
ond topic, characteristics of field-aligned ducts (spatial scale, enhancement factor,
life time, etc.) are of essential interest in whistler propagation, and also the physical
mechanism of duct formation has yet to be established, though there have been
proposed a few hypotheses (e.g., [48,49]). The second topic is still poorly under-
stood even now, but the first topic has been recently studied extensively because of
the abundance of LEO satellite observations, which will be presented below.
We first present a few earlier LEO satellite observations related to whistler
mode waves (mainly 0+ whistlers [50]) in the ionosphere [51,52]. Hayakawa [51]
measured the wave intensities at a few ELF and VLF point frequencies on board the
Ariel 4 satellite. High values of the observed mean/minimum and peak/mean
intensity ratios in each sampling interval allowed him to infer that the measured
ELF/VLF radio noises at low latitudes are impulsive and so were considered to be
due to lightning discharges. His world distribution suggested that VLF noises are
localized around three lightning chimneys of the world, and also a clear day/night
asymmetry was compatible with the day/night difference of transionospheric
absorption with full-wave computations.
We move on to recent satellite observations. Chum et al. [53] investigated the
correspondence between 0+ whistlers on the DEMETER and Magion-5 satellites
and lightning discharges detected by the ground-based European Lightning
Detection Network (EUCLID). And they demonstrated that the maximum whistler
amplitude at the satellite depends primarily on the proximity to the source lightning
relative to the magnetic footpoint of the satellite, and the area in the ionosphere
through which the electromagnetic energy induced by a lightning discharge enters
the ionosphere, is up to several thousands of kilometers wide. Then, using the same
satellite and EUCLID, Fiser et al. [54] found pairs of causative lightning and cor-
responding whistlers, and processing data from 200 paths over Europe it is found
that the mean amplitude decreases monotonically with the horizontal distance up to
1,000 km from the lightning discharge, and nighttime amplitude is about three
times the daytime value. Using a similar detection technique to Fiser et al. [54]
adopted to the burst mode data on DEMETER, Compston [55] studied the global
distribution of lightning activity.
Then, using VLF wideband signals on board the DEMETER, Ferencz et al.
[56] have found a specific signal structure of numerous 0+ whistlers and they
termed them as “Spiky” whistlers as shown in Figure 12.8. These signals appear to
be composed of a conventional 0+ whistler combined with the compound mode
patterns of tweek sferics [57–60]. This finding is not so surprising, because they
have detected just the leakage of tweek sferics into the ionosphere, propagating in
whistler mode up to the satellite. A more general effect fundamentally based on the
494 Lightning electromagnetics: Volume 2
20
frequency (kHz) 15
10
0
0 0.1
time (s)
same principle, had already been observed on the ground-based short whistlers
[40,42,61]. Shimakura et al. [61] found, for the first time, such tweek mode patterns
superimposed on low-latitude short (1 hop) whistlers as observed simultaneously at
two stations of Sakushima and Kagoshima in Japan. The simultaneous observations
of whistlers and causative lightning discharges enabled them to conclude that the
causative sferics are located exactly just below the duct entrance in the southern
hemisphere and then additional tweek traces might be due to the subionospheric
propagation from the far ionospheric exit point (about 3,000 km east of the stations)
in the northern hemisphere to the VLF stations.
Next Parrot et al. [62] have studied the occurrence rate of 0+ whistlers with the
automatic detection with the use of neural network on board the DEMETER [63],
and Figure 12.9 illustrates an example of their world distribution of the occurrence
rate of 0+ whistlers during nighttime. They have shown that the whistler rate cal-
culated as a function of longitude varies between 1 and 6 s1 during nighttime and
0.5–0.7 s1 at the day. Also, whistler rate is found to be anti-correlated with the
F10.7 cm solar flux, being consistent with earlier ground-based results at low
latitudes [64,65], though there have been some publications on the positive corre-
lation between the lightning activity and solar activity [66]. The decreased light-
ning activity at the solar minimum is largely counterbalanced by the increase in
whistler rates in the ionosphere due to the decrease in ionospheric absorption.
Finally, the measurements of electromagnetic wave intensity unlike the use of
wideband spectra before, have recently been extensively utilized to study lightning
activity. Using DEMETER data, Nemec et al. [67] studied the relationship between
median intensities of electromagnetic emissions in the VLF range and lightning.
The analysis of 3.5-year data demonstrated that electromagnetic emissions may be
due to lightning activity changes. The effect of lightning activity is most
Whistlers and ionospheric Alfvén resonator 495
63191
–90 1
–180 0 180
Longitude
this IAR is not yet agreed. As you imagine, there are two probable possibilities
depending on latitude, that is (1) magnetospheric origin; any noise in the upper
magnetosphere or (2) lightning origin; that is, the effect of lightning discharges in the
atmosphere. No matter whether the origin of IAR is either magnetospheric or
atmospheric lightning effect, the mode responsible for this IAR is left-handed
polarized Alfvén mode as discussed in Section 12.2.2.3.
The findings of IARs have been obtained from ground-based observations, and we
will cite the papers mainly based on “long-term” observations. IAR was initially dis-
covered at a middle latitude station (Nizhni Novgorod, Russia; L = 2.65 (McIlwain L is
defined by the geocentric distance where the magnetic field crosses the magnetic
equator in the unit of Earth’s radius), geomagnetic latitude = 52.12) and at a high
latitude station (Kilpisjarvi, Finland; L 6, geomagnetic latitude = 66) by Belyaev
et al. [96–98]. Further experimental evidence for the existence of IAR at high latitudes
was later confirmed by Demekhov et al. [91] with the data at Kilpisjarvi observatory.
Yahnin et al. [99] studied diurnal and seasonal variations of SRS occurrence rate based
on continuous observations for more than 4 years at a high latitude station, Sodankylä
(L = 5.2, geomagnetic latitude of 64 ). They found a clear tendency of decrease in both
the resonant frequencies and difference in resonance frequencies DF from the mini-
mum to maximum solar activity. The high-resolution measurements of IAR signatures
were also made at low-latitude stations such as Crete (L = 1.3, geomagnetic latitude of
28 ) by Bösinger et al. [100] during half a year and Muroto, Japan (L = 1.206, geo-
magnetic latitude of 24.4 ) with the 2.5-year data by Nose et al. [101]. Very recently
Beggan and Musur [102] have reported on the characteristics of IAR in middle latitude
from their long-term (5 years) observation at Eskdalemuir, UK (L = 3.46, geomagnetic
latitude = 55.3), which are consistent with early works by Molchanov et al. (2014) to
be presented later. Potapov et al. [95] have attempted the simultaneous IAR observa-
tion at two mid-latitude stations, Ulaanbaatar (L = 1.9) and Mondy (L = 2.2) and one
high-latitude station, Istok (L = 6.1), but unfortunately not continuous. Their results
during 4 days are found to be consistent with early results.
While there are so many ground-based measurements of IARs, there have been
recently some attempts to make in-situ observations of IARs from data by LEO
satellites. Simoes et al. [103] have found few cases, with the data obtained by the
C/NOFS satellite, of detecting a distinct picture of IAR (and Schumann resonances)
during the minimum solar activity during the cycle 23/24 in the low-latitude
ionosphere (most evident during nighttime). Further observational evidence of IAR
signatures was obtained in low latitudes by the Ukrainian satellite, Chibis-M [104]
and by the Freja satellite in auroral latitudes [105]. Dudkin et al. [104] have shown
that their satellite observations have not revealed any long-term IAR signatures
unlike the ground-based observations (as will be shown later). Also, in contrast to
the dominant view (as will be shown later), IAR has been found to be effectively
excited on the dayside as well.
In Section 12.3.2, we initially present our statistical properties of IARs
observed at middle latitude on the basis of ground-based observations for more than
2 years at a mid-latitude observatory at Karimshino (Kamchatka, Russia, L = 2.1,
geomagnetic latitude of 46 ) [106,107] in order for the reader to understand what
498 Lightning electromagnetics: Volume 2
IAR looks like. Next in Section 12.3.4, we will review the generation mechanisms
of IAR signatures. It is definitely certain that this phenomenon is a resonance effect
of Alfvén mode waves between the lowest ionosphere and lower magnetosphere,
but the energy source (or initial agent) of those IAR signatures is very con-
troversial, seemingly strongly dependent on latitude of our observation. And, in
Section 12.3.5, we have suggested one of the most promising candidate hypothesis
or excitation of IAR at middle and low latitudes by nearby thunderstorms, which
might be very useful for the lightning community, that is, readers of this book.
12.3.2 Ground-based observations of IARs at middle latitude
12.3.2.1 Observations and data processing
The continuous registration of the ULF/ELF magnetic field variations at the geo-
physical observatory, Karimshino started in June 2000 and is still going on. The
observatory is located in Karimshino (geographic coordinates: 52.94 N, 158.25 E,
L = 2.1) at a distance of about 50 km from Petropavlovsk-Kamchatsky (see the
details in Uyeda et al. [108] and Gladychev et al. [109]). Three axial induction
magnetometers are used to measure geomagnetic field variations in the p frequency
ffiffiffiffiffiffi
range of 0.003–40 Hz. The sensitivity threshold is better pffiffiffiffiffiffithan 20 pT/ Hz at a
frequency of 0.01 Hz, and it corresponds to 0.02 pT/ Hz at frequencies above
10 Hz. The sampling rate per channel is 150 Hz, and the sampling resolution is
24 bits. The accuracy of absolute and relative (between channels) timing of digital
data is 5 ms and better than 10 ms, respectively. See Schekotov and Hayakawa
[110] for further details of the equipment and measurement.
Because of relatively high crust conductivity in the region of observations the
vertical magnetic signal amplitude is low in comparison with the horizontal ones.
The amplitude of the vertical component in the frequency range of 1–3 Hz is
comparable with the sensor sensitivity, and thus the resonant structures cannot be
detected for the vertical component with the equipment used. The technique and
results of the signal analysis for two horizontal components are given below.
The signal spectra were evaluated with Welch’s method in the frequency range
0.1–5 Hz with resolution of 0.05 Hz in the time window of 30 min and many para-
meters of resonance structures were estimated. An example of the evolution of
nighttime dynamic frequency spectra is shown in Figure 12.10 for about one week
from September 12 to 18, 2000. The bottom three panels illustrate the dynamic
spectra for the three magnetic field components (H, D, and Z), respectively. The
second panel indicates the coherency between the two magnetic components H and D
and the third panel indicates the wave polarization by using the two horizontal
magnetic field components. The top panel indicates the temporal evolution of the Kp
index (geomagnetic activity) for comparison. Figure 12.11 is another type of pre-
sentation of Figure 12.10 in the form of spectral power density versus wave fre-
quency for the two horizontal components D (in thick line) and H (in thin line). The
local time is LT = 19 h through LT = 06 h on a particular day, September 13 and 14,
2000. We can recognize from this figure the IAR spectra with typical resonance
(fingerprint) structures, especially in the local time interval from LT = 21 h to 01h,
but weak (not so conspicuous) resonance structures are also seen at other LTs.
Whistlers and ionospheric Alfvén resonator 499
4
Freq. Hz
3
2 0.25
1
0
5 0.5 Right
4
Freq. Hz
3 0 Linear
2
1
–0.5 Left
3
2 H 1
1
0.5
Dynamic Spectra
5 1.5
4
Freq. Hz
3
D 1
2
1
0.5
5 1.5
4
Freq. Hz
3
2 Z 1
1
0.5
12 13 14 15 16 17 18 Date (LT)
We can notice from Figures 12.10 and 12.11 very clear resonance structures on
the horizontal magnetic field components (H and D). However, when we use the
“polarization spectrum,” it is much easier for us to identify such resonance struc-
tures (SRS), which is clearly recognized in the second top panel of Figure 12.10,
when we compare it with the corresponding amplitude dynamic spectra (e.g., the
2nd panel from the bottom in Figure 12.10). The observed wave is found to be left-
handed polarized (or Alfvén mode wave).
The algorithm of automatic SRS detection and calculation of its main para-
meters have been described below. Various parameters obtained from the analysis
of the spectra are as follows:
● averaged separation frequencies of resonance DF,
● intensity of the resonance signal,
● “quality” parameter Q of the resonance structure.
500 Lightning electromagnetics: Volume 2
1 1 1
0 0 0
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
22 h 23 h 00 h
3 3 3
D D D
H H H
2 2 2
Field Spectral Density, Hz2*pT2/Hz
1 1 1
0 0 0
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
01 h 02 h 03 h
3 3 3
D D D
H H H
2 2 2
1 1 1
0 0 0
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
04 h 05 h 06 h
3 3 3
D D D
H H H
2 2 2
1 1 1
0 0 0
0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
Frequency, Hz
∆F
Spec. Dens., Hz 2*pT2 / Hz
0
0.5 1 1.5 2 2.5 3 3.5
(a) Frequency, Hz
Spec. Dens. Var., Hz 2*pT2 / Hz
–1
–2
0.5 1 1.5 2 2.5 3 3.5
(b) Frequency, Hz
1.5
Var. Spec. Dens., a. u.
δf
1.0
0.5
0
0 0.5 1 1.5
SRS ∆F
(c) Frequency, Hz
Figure 12.12 The method of SRS identification and determination of the SRS
parameters. (a) SRS (in solid line) and its trend (in dashed line), (b)
SRS difference, and (c) Spectrum of the SRS difference. Its
maximum location is the averaged DF in the SRS. SRS exists if the
distribution is narrow banded, that is, quality Q1. After
Molchanov et al. [107]
502 Lightning electromagnetics: Volume 2
density variation) is given in the middle panel (b). The total power can be estimated
as a sum of the resonant power and the background power approximately corre-
sponding to the curve going through the minima of spectra in Figure 12.12(a). The
intensity of the resonant signal is numerically estimated as the mean of the absolute
value of spectral density variation, but the relative accuracy of this approximation is
low. Both spectra of the total signal and its resonant part have several approximately
equidistant maxima. The average distance between the maxima is estimated with a
help of the Fourier transform of the spectrum of the resonant signal, and this spec-
trum is depicted in Figure 12.12(c) as a function of frequency. Its maximum corre-
sponds to the averaged frequency difference between the SRS maxima. We define
the “quality” Q of the resonant structure as the ratio of this maximal frequency to the
half-width of the maximum df using the formal similarity with the parameters of
damping oscillations. A resonance structure by definition exists if at least two max-
ima are found in the spectrum, that is, Q > 1. This allows us to exclude a possibility
of false IAR detection caused by some other effects like Pc1 geomagnetic pulsations.
0.4
OCCURRENCE RATE
0.3
0.2
NO DATA
0.1
0
Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
0.7
SRS FREQUENCY ∆F, Hz
0.6
0.5
NO DATA
0.4
0.3
0.2
Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
2000 2001 2002
our statistical analysis. Correlation coefficients of SRS occurrence rate with the Kp
index of global magnetic activity for different 1-year intervals are shown in
Figure 12.16 (days with data gaps and the last three months, April–June 2002 (see
seasonal dependence) were excluded from the analysis). The stable and reliable
negative correlation of IAR with Kp is noticed from the figure and it may be
possible that SRSs of IAR are masked under disturbed geomagnetic conditions by
high background activity.
0.4
OCCUR. RATE
0.2
0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
0.6
FREQ. ∆F. Hz
0.4
0.2
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
0.6
SPEC.DENS., pT/√Hz
0.4
0.2
0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
the IAR occurrence rate is estimated based on the above detection criterion.
The seasonal variation is strong: In spring–early summer only few IAR events with
SRS are registered, but the occurrence rate is found to strongly grow in the autumn-
winter period.
Whistlers and ionospheric Alfvén resonator 505
OCCURRENCE RATE
0.6
WINTER
0.4
0.2
0.6
SPRING
0.4
0.2
0.6
SUMMER
0.4
0.2
0.6
AUTUMN
0.4
0.2
0
12 14 16 18 20 22 00 02 04 06 08 10 12
Local Time, hour
Figure 12.15 Diurnal variation (or local time dependence) of the IAR occurrence
rate for different seasons (winter, spring, summer, and autumn,
from top to the bottom). After Molchanov et al. [107]
506 Lightning electromagnetics: Volume 2
0.0
–0.2
–0.4
–18 –12 –6 0 6 12 18
Time delay IAR to Kp, days
Figure 12.16 Correlation coefficients of the IAR occurrence rate with the Kp
index for different 1-year periods (shown above). All the correlation
coefficients are negative for time delay of 0–1 day. After Molchanov
et al. [107]
4 years, Bösinger et al. [100] used the data only for half a year, and Nosé et al.
[101] used the data for about 2.5 years. When we think about the generation
mechanism of IAR, the latitudinal dependence of different morphologies of IAR
would be of essential importance (this will be described in the next subsection).
Several results have been reported on the basis of a shorter database [95–97] even
at middle latitudes, but this section has provided the first results of middle latitude
IAR characteristics by using sufficiently long-term observation. Point (1) in the
above summary suggests that IAR structures occur for approximately one-quarter
of the observation periods (250 nights). This means that the IAR phenomena are not
so rare, and are rather a regular phenomenon, which seems to be consistent with the
conclusion by Bösinger et al. [100,111] and Nosé et al. [101] at low latitudes. Point
(2) on LT dependence is likely to be consistent with the nighttime maximum in
previous or recent publications at high and low latitudes ([99,101]; Bösinger et al.
[100]. Next, the seasonal variation in Point (1) seems to be consistent with Yahnin
et al. [99] at high latitudes, but Nosé et al. [101] at low latitudes have found a
slightly higher occurrence in May–September, which is considerably different from
our seasonal dependence at middle latitude. Finally, as far as Kp dependence is
concerned, we have found a clear anti-correlation at middle latitudes (Point (6))
being consistent with high-latitude results [99], but Nosé et al. [101] have noticed
no significant correlation with Kp index at low latitudes. So these existing con-
sistencies and discrepancies in morphological characteristics of IARs at different
latitudes will be of great potential in studying the energy source of IARs at different
latitudes.
These findings with ground-based measurements have been confirmed by the
LEO satellites [103,104] but few examples. Further, Dudkin et al. [104] have found
the daytime occurrence as well, which suggests that the absence of IAR during the
day on the ground may be due to the stronger daytime absorption during the
ionospheric transmission.
wave field, which means that in the ELF frequency range the discharge spectrum
from nearby lightning is more intense than that of more distant discharges [121].
This is the basic idea of trying to correlate IARs to nearby lightning.
As it is seen from the examples in Figure 12.10, there are many impulses that
can be associated with thunderstorm activity, and only some of these impulses are
accompanied by sharp impulses in the frequency range of 0.25–4 Hz. The former is
thus assumed to be a result of nearby lightning discharges. In order to estimate the
number of lightning discharges per unit of time, we choose the signal discrimina-
tion level as 5 pT. Total number of the impulses, DN, with amplitude which is
greater than this level, increases with time as shown in Figure 12.17. Lines 1 and 2
correspond to the frequency filters 6–20 Hz and 0.25–4 Hz, respectively. Averaging
over an interval of 1 hour results in a mean occurrence rate of the impulses of about
n1 = 0.144 s1 and n2 = 0.023 s1, which is typical for nighttime conditions at
Karimshino station. Note that about 2,000 thunderstorms operate simultaneously in
the whole world producing a total current of about 1,800–2,000 A [79]. Taking into
account that a lightning discharge usually brings the charge 20–30 C, one can find
that the mean occurrence rate of the lightning discharges v = 60–100 s1. Hence a
thunderstorm typically produces a rate number of about 0.03– 0.05 s1, which is
600
500
1
400
∆N
300 3
4
200
2
100
0
0 10 20 30 40 50 60
t, min
Figure 12.17 Temporal variations of the sum of magnetic impulses. The threshold
level for the impulse amplitude is 5 pT. Lines 1 and 2 correspond to
the frequency channels 6–20 and 0.25–4 Hz, accordingly. The
impulses of solely H component (6–20 Hz) are shown with dash
lines 3, while the D component is shown with dash line 4. After
Surkov et al. [121]
510 Lightning electromagnetics: Volume 2
close to n2. From here we may assume that one nearby thunderstorm and 3–6
remote ones make a major contribution to the rate number shown in Figure 12.17.
As for the signals of remote thunderstorms, it should be noted that the intensity of H
component is larger than that of D component. The impulses of H component
(6–20 Hz) that are displayed in Figure 12.17 with dashed line 3 occur more fre-
quently than in the D component shown with dashed line 4. On the other hand, in
the frequency range of 0.25–4 Hz the occurrence rate numbers of both components
are very close to each other.
Most of the intense signals which can be associated with a nearby thunder-
storm have a bipolar structure. It appears that the first impulse in the signals is due
to the primary wave radiated by the return stroke. The interval between positive and
negative impulses is typically 2 s. One may assume that such a shape of the signal
results from Alfvén wave reflection from the gradient in Alfvén velocity at the
upper boundary of the resonance cavity. If the typical size of the resonance cavity is
500–1,000 km, the arrival time of the reflected Alfvén wave is estimated as 2 s,
which is close to the signal duration. The signal occasionally contains three distinct
impulses at least. This implies a possibility for multiple wave reflections from the
IAR upper boundary.
Let N be the number of nearby thunderstorm centers simultaneously operating
around the ground-based recording station. A local coordinate system has the x-axis
eastward, the y-axis to the north, and z-axis vertically upward. Since we are inter-
ested in solely nearby thunderstorms, a plane-stratified model of the medium is
used. Let rm and jm be the polar coordinates of the thunderstorm epicenter, where
m = 1, 2, . . . N as shown in Figure 12.18. A typical size of the thunderstorm is
assumed to be smaller than the distance from the recording station. We ignore the
lightning discharge distribution inside a thunderstorm area, and this implies that all
y y'
N-S
rµ
z φµ x
Br
BI
Figure 12.18 Schematic picture of reference and local coordinate systems. There,
z and z0 axes are “out of paper.” Br and Bf are the component of
magnetic variations due to the thunderstorm located at r = rm. After
Surkov et al. [121]
Whistlers and ionospheric Alfvén resonator 511
the lightning discharges related to the same thunderstorm must have the same
coordinates; that is, the coordinates of a given thunderstorm.
Now we also introduce a reference frame x0 , y0 , and z0 fixed to the thunderstorm
with the number m. Let B (rm, t–tnm) be the magnetic field due to the lightning dis-
charge that happened at the accidental moment tnm, where nm= 1, 2, . . . is the number
of the lightning discharge. It is usually the case that the lightning discharges are
vertical and transfer a negative electric charge to the ground [6,127,128]. If we use a
cylindrical coordinate system in which the lightning discharge is in the direction of the
polar z0 -axis, the magnetic field B is independent of azimuthal angle j. According to
Belyaev et al. [97], only radial component Br among the two horizontal ones contains
the resonance factor, which dominates the IAR resonance properties. In the Cartesian
reference frame fixed to the ground-recording station, the horizontal magnetic field
can be expressed through the radial and azimuthal components as follows:
Bx ¼ Bf sin jm Br cos jm
By ¼ Bf cos jm Br sin jm (12.39)
where Bm is the horizontal magnetic field due to the thunderstorm with number m,
the matrix Âm and the vector Gnm are given by
b m ¼ cosjm
A
sinjm
; (12.42)
sinjm cosjm
512 Lightning electromagnetics: Volume 2
and
Gr ðrm ; t tnm Þ
Gnm ¼ ; (12.43)
Gj ðrm ; t tnm Þ
where wm are inverse time constants and the amplitudes Im of individual currents
P
4
term must satisfy the condition Im ¼ 0. The vertical current channel grows
m¼1
upward with the velocity dl/dt = V0 exp (Wt), where V0 = 8 107 m/s is the
current wave velocity at the ground level and W is the relaxation time parameter
whence it follows that the final channel length is l1 = V0/W. The current moment of
the single return stroke m1(t) = I(t)l(t) can be written as (see, e.g., [79,131,132])
X
4
m1 ðtÞ ¼ l1 ½1 expðWtÞ Im expðwm tÞ (12.45)
m¼1
It is usually the case that the lightning discharge contains n = 2–6 return
strokes. More frequently, there are three return strokes with a characteristic dura-
tion of about 100 ms. The mean interval between them is of the order of t0 = 40 ms.
Following Jones [129], we assume that the final length of the return stroke
increases with its number n as ln = l1 + (n–1) Dl, where l1 = 4 km and Dl = 1 km.
The current relaxation time parameter Wn is related to the channel length ln as
follows:
1 ln 1
Dl
Wn ¼ ¼ W1 þ ðn 1Þ; (12.46)
V0 V0
where W1 = V0/l1 = 2 104 s–1. The net magnetic moment of multiple discharges
can be written as m(t) = MF(t), where M = l1|I1| is the “magnitude” of magnetic
moment. while the dimensionless function F(t) describes the shape of multiple
discharge
X
n0
ln X4
Im
FðtÞ ¼ 1 exp Wn t0n h t0n exp wm t0n (12.47)
l
n¼1 1
j
m¼1 1
I j
electric current moment. To simplify the problem, we assume now that the
lightning discharge contains n0 = 3 return strokes. Taking the typical parameters
for the models of the return strokes (see, e.g., [79,131–134]) I1–4 = 28.45, 23.0,
5.0, 0.45 (in kA) and w1–4 = 6.0 105, 3.0 104, 2.0 103, 147.0 (in s1), one can
estimate the typical magnitude of magnetic moment as M = 170 kA km. The
spectrum of the function F(t) is
W1 Xn0 X4
Im exp½jwt0 ðn 1Þ
F ðw Þ ¼ (12.48)
2p n¼1 m¼1 jI1 j ðwm jwÞðWn þ wm jwÞ
z
z
B0
z=d+1+L Magnetosphere z
L F layer
z=d+l VA
σP σH
VAI VAM
l E layer
z=d x
y
Atmosphere
d
σg
z=0
σ
Earth
Figure 12.19 Schematic illustration of a stratified medium model. The plots of the
Alfvén velocity and the ionosphere/ground conductivities are shown
in the right panel. After Surkov et al. [121]
514 Lightning electromagnetics: Volume 2
is immersed in the constant Earth’s magnetic field B0. The actual profile of the
Hall and Pedersen conductivities strongly varies with altitude. The plasma con-
ductivity peak lies inside the ionospheric E layer, which is considered in the thin-
slab approximation. This condition is not so burdensome on account of smallness
of the typical wavenumber k allowing for the condition kl « 1, where l denotes the
characteristic width of the E layer. In the model, the region above the E layer
(i.e., F layer and magnetosphere) is supposed to be the area consisting solely of
cold collisionless plasma. The IAR formation in this region is due to the plasma
density fall off with altitude, which makes the Alfvén velocity, vA, rapidly
increase upward. The ionospheric resonance cavity is bounded from below by the
conductive E layer and from above by the region where the gradient of the Alfvén
velocity reaches a peak, which makes it possible to have reflection of the Alfvén
wave from the upper space. The typical vertical scale of the resonance cavity is
L 103 km.
It should be noted that the same resonance cavity can serve as a waveguide for
the magnetosonic/compressional mode (modified Alfvén wave in the previous
section) (e.g., [135]). According to Pokhotelov et al. [87], we use a suitably idea-
lized model of the resonance cavity which describes the Alfvén velocity in terms of
a piece-wise function so that vA = vAI within the resonance cavity (d < z < d + l +L)
and vA = vAM in the outer magnetosphere (z > d + l +L), where vAI and vAM (vAM » vAI)
are constant quantities referring to the ionosphere (I) and magnetosphere (M),
respectively. The geomagnetic field B0 is supposed to be vertically upward. A more
accurate model which takes into account the dip angle of the local geomagnetic field
gives rise to very complicated equations [136–138]. For simplicity, we adopt the
model of the vertical geomagnetic field in order to avoid the complexities connected
with magnetic field inclination.
On account of the axial symmetry of the field produced by the vertical current
moment, the cylindrical coordinates z, r, and f are used. In this case, all the
quantities are free of f so that @=@f ¼ 0. In the atmosphere, the primary field of
the current moment contains only three components, Er, Ez, and, Bf termed as TM
mode. When the electromagnetic wave of TM mode penetrates into the ionosphere,
it produces excitation of the TE mode, that is, Ef , Br, and Bz due to the mode
coupling via Hall conductivity in the ionosphere. The shear and compressional
Alfvén waves can get trapped in the F layer of the ionosphere thereby exciting the
IAR. Low-frequency resonant oscillations are leaking back into the atmosphere so
that all the components of the electromagnetic field, which are TM and TE modes,
can be detected on the ground.
In the framework of the model, we note that the spectra of the electromagnetic
perturbations due to a solitary lightning discharge have been obtained by [139]. The
reader is referred to those works for details about the derivation of the formulae. Let
SW ¼ 1=ðm0 vAI Þ be the Alfvén parallel conductance, ap ¼ Sp =Sw and ap ¼ SH =Sw
stand for the dimensionless height-integrated Pedersen and Hall conductivities,
respectively, and x0 = wL/vAI be dimensionless frequency. The Fourier transforms of
the magnetic perturbation dBr and dBf on the ground level z = 0 can be written as
follows,
Whistlers and ionospheric Alfvén resonator 515
0.05
1
2
0.04
2'
| δ Br | . pT . Hz –1
0.03
3
0.02
0.01
4
0
0.5 1 1.5 2 2.5 3 3.5 4
f, Hz
Figure 12.20 Model calculation of the nighttime IAR spectra excited by a solitary
cloud-to-ground lightning discharge. The radial/resonant
component Br on the ground is shown with lines 1–4, which
correspond to the distances r = 100, 300, 1,000, and 10,000 km,
respectively. The approximate analytical solution at distance
r = 300 km is shown with dotted line 2’. After Surkov et al. [121]
516 Lightning electromagnetics: Volume 2
symmetry of the problem the radial component of the magnetic perturbation must
tend to zero when r ! 0. The calculations have shown that the spectrum magnitude
reaches a peak at a distance of about 300 km.
Surkov et al. [121] have further considered the random magnetic variations
produced by the CG lightning discharges treated as a stochastic process. The
obtained frequency spectra are found to be close to the observations. The theore-
tical consideration by Surkov et al. [121] can be summarized as follows:
1. The model computations of the power spectra are in favor of nearby thunder-
storms as a possible cause for the IAR excitation at middle latitudes.
2. The solitary CG lightning discharges in the neighborhood of the ground station
may result in the impulse IAR excitation, which is capable of producing an
observable SRS signature on the ground.
3. The random lightning discharges in the range of 1–2 thousands km make a
main contribution to the mid-latitude IAR power spectrum since the predicted
resonant frequencies and peak intensities are practically consistent with those
observed.
Finally, we have to conclude that even though there have been proposed sev-
eral hypotheses as the origin of IARs at different latitudes, any mechanism is far
from being accepted. So appreciable work will be highly required observationally
in order to validate even the most promising hypothesis of nearby lightning dis-
charges in the excitation at middle/low latitudes in Section 12.3.4.
References
[1] Wait, J. R., Electromagnetic Waves in Stratified Media, Pergamon Press,
Oxford, 1970.
[2] Budden, K. G., The Waveguide Mode Theory of Wave Propagation, Logos
Press, London, 1961.
[3] Davies, K., Ionospheric Radio, Peter Peregrinus, London, 580p., 1990.
[4] Al’pert, Ya. L., Propagation of Radio Waves in the Ionosphere, Plenum
Press, New York, NY, 1974.
[5] Galejs, J., Terrestrial Propagation of Long Electromagnetic Waves,
Pergamon Press, Oxford, New York, 376p, 1972.
[6] Rakov, V. A., and M. A. Uman, Lightning: Physics and Effects, Cambridge
University Press, Cambridge, 2003.
[7] Walker, A. D. M., The theory of whistler propagation, Rev. Geophys. Space
Phys., 30, 411–421, 1976.
[8] Hayakawa, M., Whistlers, in H. Volland (Ed.), Handbook of Atmospheric
Electrodynamics, Vol. 2, CRC Press, Boca Raton, FL, pp. 155–193, 1995.
[9] Helliwell, R. A., Whistlers and Related Ionospheric Phenomena, Stanford
University Press, Stanford, CA, 1965.
[10] Park, C. G., Whistlers, In Handbook of Atmospherics, Vol. 2, Ed. by H.
Volland (eds.) , CRC Press, Boca Raton Florida, pp. 21–79, 1982.
[11] Pilipenko, V.A., Impulsive coupling between the atmosphere and ionosphere/
magnetosphere, Space Sci. Rev. 168, 1–4, 2011, doi: 10.1007/s11214-011-
9859-8.
[12] Surkov, V. and M. Hayakawa, Ultra and Extremely Low Frequency
Electromagnetic Fields, Springer, Tokyo, 486p, 2014.
[13] Barkhausen, H., Zwei mit Hilfe der neuen Verstärker entdeckte
Erscheinungen, Physik Z., 20, 401–403, 1919.
518 Lightning electromagnetics: Volume 2
[68] Colman, J. J. and M. J. Starks, VLF wave intensity in the plasmasphere due
to tropospheric lightning, J. Geophys. Res., Space Phys., 118(7), 4471–4482,
2013. doi:10.1002/jgra.50217.
[69] Zahlava, J., F. Nemec, O. Santolik, et al., Longitudinal dependence of
whistler mode electromagnetic wave in the Earth’s inner magnetosphere,
J. Geophys. Res., 123, 6562–6575, 2018. doi:10.1029/2018JA025384.
[70] Christian, H. J., R. J. Blakeslee, D. J. Boccippio, et al., Global frequency and
distribution of lightnings as observed from space by the Optical Transient
Detector, J. Geophys. Res., 108(D1), 4005, 2003. doi:10.1029/
2002JD002347.
[71] Ripoll, J. F., T. Farge, D. M. Malaspina, et al., Analysis of electric
and magnetic lightning-generated wave amplitudes measured by the Van
Allen probes, Geophys. Res. Lett., 47, e2020GL087503, 2020. doi:10.1029/
2020GL087503.
[72] Jacobson, A. R., R. H. Holzworth, R. Pfaff, and R. Heelis, Low-latitude
whistler wave spectral and polarization from VEFI and CINDI payloads on
C/NOFS satellite, J. Geophys. Res., Space Phys., 125(1), 2020. doi:10.1029/
2019JA02704.
[73] Abel, B. and R. M. Thorne, Electron scattering in Earth’s inner magneto-
sphere 1. Dominant physical processes, J. Geophys. Res., 103(A2), 2385–
2396, 1998. doi:199810.1029/97JA02919.
[74] Bortnik, J., U. S. Inan, and T. F. Bell, Energy distribution and lifetime of
magnetospherically reflecting whistlers in the plasmasphere, J. Geophys.
Res., 108(A5), 1199, 2003. doi:10.1029/2002JA009316.
[75] Meredith, N. P., R. B. Horne, S. A. Glauer, and R. R. Anderson, Slot region
electron loss timescales due to plasmaspheric hiss and lightning-generated
whistlers, J. Geophys. Res., 112(A8), A08,214, 2007. doi:10.1029/2007
JA012413.
[76] Hayakawa, M. and S. S. Sazhin, Mid-latitude and plasmaspheric hiss: A
review, Planet Space Sci., 40, 1325–1338, 1992.
[77] Sazhin, S. S. and M. Hayakawa, Magnetospheric chorus emissions: A
review, Planet Space Sci., 40, 681–697, 1992.
[78] Inan, U. S., S. A. Cummer, and R. A. Marshall, A survey of ELF and VLF
research on lightning-ionosphere interaction and causative discharges,
J. Geophys. Res., 115, A00E36, 2010. doi:10.1029/2009JA014775.
[79] Nickolaenko, A. P. and M. Hayakawa, Resonances in the Earth-Ionosphere
Cavity, Kluwer Academic Publication, Dordrecht, 2002.
[80] Schumann, W. O., Uber strahlunglosen Eigen schwingungen einer leitenden
Kugel, die von einer Luftschicht and einer Inospharenhulle umgeben ist,
Z. Naturforschung, A7, 149–154, 1952.
[81] Nickolaenko, A. P. and M Hayakawa, Schumann Resonances for Tyros:
Essentials of Global Electromagnetic Resonance in the Earth-Ionosphere
Cavity, Springer, Tokyo, 2014.
[82] Williams, E. R., The Schumann resonance: a global tropical thermometer,
Science, 256, 1184–1187, 1992.
522 Lightning electromagnetics: Volume 2
13.1 Introduction
An assessment of the global distribution of nitrogen oxides (NOx) is required for a
satisfactory description of tropospheric chemistry and in the evaluation of the
global impact of increasing anthropogenic emissions of NOx [1]. In the mathema-
tical models utilized for this purpose, it is necessary to have the natural as well as
man-made sources of NOx in the atmosphere as inputs. Thunderstorms are a main
natural source of NOx in the atmosphere and it may be the dominant source of NOx
in the troposphere in equatorial and tropical South Pacific [2].
In quantifying the production of NOx by thunderstorms, scientists have until
recently concentrated on the lightning return strokes neglecting all other processes
associated with thunderstorms [3]. This view is gradually changing as the theories
and experimental data show that not only the return strokes in ground flashes but
other discharge events in ground and cloud flashes, such as continuing currents, are
also contributing significantly to the NOx emissions [4]. For example, the study
conducted by Rahman et al. [4] shows that the contribution by continuing currents
to the NOx production in lightning flashes is comparable, if not overwhelm, to the
contribution by return strokes. However, the physics behind the process that makes
continuing currents as efficient as return strokes in producing NOx is still unknown.
Based on the ionization process that leads to the production of free electrons,
electrical discharges taking place in the atmosphere can be divided into two types,
namely, ‘cold’ and ‘hot’ electrical discharges. In cold electrical discharges, free
electrons are produced solely by the collisions between energetic electrons and
atoms. In these discharges, the electron temperature may reach several tens of
thousands of degrees whereas the gas and ion temperature remains close to ambient
temperature. Corona discharges, streamer discharges and Townsend-type electrical
*This material was published previously by the same authors in Open Atmospheric Science Journal,
vol. 2, pp. 176–180, 2008. It is reproduced here by permission from the journal.
1
Department of Electrical Engineering, Uppsala University, Sweden
2
School of Engineering, Royal Institute of Technology, KTH, Sweden
528 Lightning electromagnetics: Volume 2
discharges taking place at low pressure are several examples of cold discharges
taking place in the atmosphere. In hot electrical discharges, the gas and ion tem-
perature can also reach several tens of thousands of degrees and the main
mechanism in the discharge that generates free electrons is the thermal ionization.
In these discharges, ions and neutral atoms are heated to such high temperatures
through a process called thermalization that facilitates the transfer of energy from
free electrons to neutrals via ions [5]. This transfer of energy from electrons to
neutrals causes the temperature of the neutrals to go up. This increase in the tem-
perature of the neutrals leads to production of copious amount of electrons by the
energetic collisions between neutral particles (i.e. thermal ionization). Several
examples of hot discharge processes taking place in the atmosphere are return
strokes, leaders, M components and continuing currents.
In addition to hot discharges mentioned above, an active thundercloud also
produces cold discharges in the form of corona and streamer discharges. However,
in quantification of NOx produced by thunderstorms, scientists neglect the con-
tribution of the cold discharge processes. Moreover, the recent discovery of
thunderstorm-created ionization processes in the stratosphere and mesosphere,
known as sprites, blue jets and elves, also poses a question as to the effects of these
ionization processes on the chemistry of the upper atmosphere. These ionization
processes can also be categorized under ‘cold’ discharges because the pressure at
which these discharges take place does not support thermal ionization, making
electron impacts the main source of ionization. Indeed, there is a need today to
develop procedures to quantify the NOx production in cold discharges.
The effects of solar proton events in the NOx production in the upper atmo-
sphere and the effect of NOx on the chemical balance of the stratosphere were a
major concern of the atmospheric scientists since the discovery of the importance
of NOx in ozone production and destruction [6]. In quantifying the NOx production
from these events, scientists utilized the connection between the ionizing events in
the atmosphere and the number of resulting NOx molecules. To facilitate further
discussion, let us denote the number of NOx molecules produced per ionizing event
in the atmosphere by the parameter k. The theoretical work of Nicolet [7] set the
value of k close to unity, whereas the investigations of Jackman et al. [8] predicted
that at altitudes larger than about 80 km, k is about 1.5 and for low altitude it is
about 1.2–1.3. These results have been used extensively to study the NOx produc-
tion by cosmic rays and solar radiation impinging on the Earth’s atmosphere.
Recently, Rahman et al. [9] investigated the validity of this theoretical calculation
by studying the NOx production in air by alpha particles emitted by a radioactive
source. The results of this study confirmed these theoretical predictions fixing the
value of k to about 1.0. Since the main source of ionization during proton impacts
are high energetic secondary electrons, it is reasonable to assume that the number
of NOx molecules produced in discharge processes in which electrons are the main
source of ionization is approximately equal to the number of ion pairs produced in
the discharge. This hypothesis is tested in this chapter using the data obtained
from corona discharges and then it is utilized to quantify the NOx production in
lowpressure discharges and streamer discharges.
On the NOx generation in corona, streamer 529
8 × 1016
6 × 1016
NOx molecules/J
4 × 1016
2 × 1016
0 × 100
0 × 100 1 × 104 2 × 104 3 × 104
E/p (V/m/Torr)
Figure 13.1 NOx production per unit energy Nue as a function of E/p in electron
avalanches
the total number of ionizing events taking place in the vicinity of the streamer head
is about w/2Rs and the total number of NOx molecules created by the streamer in
moving a unit length, Ns, is
kw
Ns ¼ (13.3)
2Rs
On the other hand, the amount of energy dissipated by the streamer channel in
moving the unit distance is Eew. Thus, the production efficiency of NOx in strea-
mer discharges, PNOx , in molecules/J is
k
PNOx ¼ (13.4)
2eERS
Substituting k = 1, Rs = 10-4 m and E = 5 105 V/m (for positive streamers), we
find that the streamer will make about 6 1016 NOx molecules/J.
In a recent study, Cooray and Rahman [13] conducted an experiment in coaxial
geometry to measure the NOx production efficiency of streamer discharges. Let us
consider the cylindrical coaxial geometry. Assume that the peak voltage applied to
the central conductor of the coaxial system is V. This voltage will create an electric
field in the cylinder, which has its highest value at the surface of the conductor and
then decreases inversely with increasing radial distance. Consider a streamer dis-
charge initiated close to the inner electrode and moving in this electric field.
Assume that the streamer discharge will propagate to a distance where the back-
ground electric field reaches Es, the critical electric field necessary for streamer
propagation. If the applied voltage is V, then the charge that will be induced on a
unit length of the inner conductor of the coaxial arrangement is given by
2pe0 V
Q¼ (13.5)
lnðb=aÞ
where a and b are the radii of the inner and outer conductors. The electric field at a
radial distance r from the inner conductor is then given by
V
Er ¼ (13.6)
r lnðb=aÞ
Since the streamers propagate to a distance where the background electric field
is Es, the length of the streamers, ls, in the coaxial arrangement is given by
V
ls ¼ a (13.7)
lnðb=aÞEs
Then the number of ionizing events, Ns, generated during the propagation of
the streamer is given by
wls
Ns ¼ (13.8)
2Rs
532 Lightning electromagnetics: Volume 2
On the other hand, the energy released by the movement of the streamer head
in the gap, Us, is given by
V
Us ¼ ew lnðls =aÞ (13.9)
lnðb=aÞ
Thus, the number of NOx molecules produced per unit energy, Nue, is given by
k
Nue ¼ (13.10)
2Es Rs lnðls =aÞ
For the experimental conditions reported in [13], i.e. a = 0.001 m, b = 0.15 m
and V = 83 103 V, the calculated production efficiency is about 1.6 1016
molecules/J. The measurements produced 1–2 1016 molecules/J.
Studies on the NOx production in the atmosphere by proton impacts show that the
number of NOx molecules created is almost equal to the number of ion pairs created
during the impact. By comparing theory with available experimental data it is
shown in this chapter that this result is also valid for electrical discharges in which
electron impacts are the main source of ionization. Based on this observation, the
NOx production in electrical discharges where the electron production depends
solely on the impact of electrons with neutral atoms is evaluated. The types of
electrical discharges considered in this chapter are the corona discharges, electrical
discharges at low pressure and streamer discharges.
In a corona discharge associated with current amplitude I, the rate of occur-
rence of ionizing events (neglecting attachment and recombination) is I/e, where e
is the electronic charge. This is equal to 6.25 1018I. Thus, the number of NOx
molecules produced per second in the discharge is given by 6.25 1018kI in which
the value of k depends on polarity (i.e. k 1 for positive polarity and k 0:6 for
negative). It is of interest to note that this equation can be utilized to estimate the
global production of NOx by ground corona associated with thunderstorms. Various
studies estimate that the global corona current associated with ground corona is
about 1–2 kA [14]. Substituting this value in the above equation and using k = 1.0
for positive corona, the number of NOx molecules produced globally by ground
corona per second is estimated to be about (6–12) 1021. This is equivalent to an
annual production of 0.01 Tg (N). During thunderstorms, it is not only at ground
level that corona discharges are initiated. The thunderstorm itself is a large source
of corona discharges. For example, in a thundercloud, charges may disperse from
regions of high concentration to low concentration through corona discharges and
the same could be the vehicle that transports the charges induced in conducting
channels during neutralization events in to the bulk of the cloud. The theory as
presented in this chapter could be applied to obtain the NOx production from these
processes once the magnitude of the currents associated with these processes
are known.
On the NOx generation in corona, streamer 533
Let us consider the results presented in Figure 13.1. These results are obtained
using several assumptions. First, it is assumed that the electron generating
mechanism in the discharge is the impact ionization due to energetic electrons.
Second, it is assumed that the background electric field is uniform in the region in
which the discharge is taking place. However, the calculation procedure can be
modified rather easily to take into account the effect of non-uniform electric fields.
Third, it is assumed that the space charge accumulated in space during the dis-
charge will not distort the electric field locally. Due to these assumptions, the
results as given in Figure 13.1 are valid for ‘Townsend’s like’ low-pressure elec-
trical discharges where the space charge effects can be neglected. On the other
hand, the data given in Figure 13.1 can be utilized to study the NOx production in
any cold discharge with space charge distortions provided that the quantity E is
replaced by the effective electric field in which the electron avalanches are
generated.
An example of a cold discharge in which the space charge effects cannot be
neglected is a streamer discharge. In streamers the electron avalanches are initiated in
the electric field of the streamer head that is much stronger than the background
electric field in which streamers are propagating. In evaluating the NOx production of
streamer discharges in this chapter, instead of evaluating the NOx production in each
individual avalanche taking place at the streamer head, a simpler but an equivalent
procedure based on the known mechanism of the streamer is utilized. In a streamer
discharge, electron avalanches are generated in a specially varying electric field. The
maximum value of the field, which is about 2 107 to 4 107 V/m, occurs at the
head of the streamer (assuming 108 ions located within a radius of about 50–100 mm)
and it decreases to about 3 106 V/m at a distance of about 200 mm from the head.
Recall that our analysis gave about 2 1016 NOx molecules/J for streamer discharges
moving in a background electric field of about 5 105 V/m. Comparison of this NOx
production efficiency with the data given in Figure 13.1 indicates that in streamer
discharges the electron avalanches are generated in an effective electric field of about
5106 V/m.
Streamer discharges occurring in the atmosphere are associated mainly with
the lightning leaders in ground flashes and cloud flashes. Actually, the propagation
of the leader is mediated by streamer bursts emanating from the tip of the leader.
They may also originate during the neutralization of these leaders by return strokes.
Cooray et al. [15] utilized the theory as developed in this chapter to study the NOx
production by streamers in lightning leaders.
The experimental observations indicate that the space charge effects cannot be
neglected in upper atmospheric discharges known as sprites even though these
discharges are taking place at considerably low atmospheric pressures. For exam-
ple, observations indicate that sprites give rise to streamer-like structures.
Therefore, the electric field in which the avalanches are growing during the
development of sprites may be considerably higher than the background electric
field generated by the thunderclouds. However, if the dimension and the ion con-
centration in these streamer heads are known, an analysis similar to that utilized in
this chapter to study the NOx production in streamer discharges at atmospheric
534 Lightning electromagnetics: Volume 2
References
[1] Crutzen, P. J., The influence of nitrogen oxides on the atmospheric ozone
content, Quart. J. R. Met. Soc., vol. 96, 320–325, 1970.
[2] Gallardo, L. and H. Rodhe, Oxidized nitrogen in the remote pacific: the role
of electrical discharge over oceans, J. Atmos. Chem., vol. 26, 147–168, 1997.
[3] Chameides, W. L., The role of lightning in the chemistry of the atmosphere, in
The Earth’s Electrical Environment, National Academy Press, Washington D.C,
1986.
[4] Rahman, M., V. Cooray, V. A. Rakov, et al., Measurements of NOx pro-
duced by rocket-triggered lightning, Geophys. Res. Lett., vol.34, L03816,
2007, doi:10.1029/2006GL027956.
[5] Orville, R. E., Lightning spectroscopy, in R. H. Golde (ed.), Lightning,
Volume 1, Physics of Lightning, Academic Press, London, 1977.
[6] Crutzen, P. J., Atmospheric interactions-homogeneous gas reactions of C, N
and S containing compounds, in B. Bolin and R. B. Cook, (eds.), The Major
Bio-geochemical Cycles and Their Interactions, SCOPE, Paris, 1983.
[7] Nicolet, M., On the production of nitric oxide by cosmic rays in the meso-
sphere and stratosphere, Planet. Space Sci., vol. 23, 637–649, 1975.
[8] Jackman, C. H., H. S. Porter and J. E. Frederick, Upper limits on production
rate of NO per ion pair, Nature, vol. 280, 170, 1979.
On the NOx generation in corona, streamer 535
14.1 Introduction
An assessment of the global distribution of nitrogen oxides is required for an ade-
quate description of tropospheric chemistry and in the evaluation of the global
impact of increasing anthropogenic emissions of NOx [1]. In the mathematical
models utilized for this purpose, one needs to specify as inputs the natural as well
as man-made sources of nitrogen oxides in the atmosphere. Lightning is one of the
main natural sources of nitrogen oxides in the atmosphere, and it may be the
dominant source of nitrogen oxides in the troposphere in equatorial and tropical
South Pacific regions [2]. Thus, an accurate quantification of nitrogen oxide pro-
duction by thunderstorms is necessary for further development of the chemical
models of the troposphere and in the evaluation of the effects of the man-made
nitrogen emissions in the terrestrial atmosphere.
Because of the difficulty of making direct measurements of NOx produced by
natural lightning flashes, researchers have employed indirect methods to quantify
the global production of NOx [3–12]. Because of a large number of uncertainties
involved in these methods, the estimates of global NOx production by lightning
flashes available in the literature vary by two orders of magnitude, from 1 to 100 Tg
(N) per year. In estimating lightning produced NOx by indirect methods scientists
have usually utilized the following two procedures: (1) a laboratory measurement
of the number of NOx molecules per unit energy for a laboratory spark is made and
the result is extrapolated to lightning by multiplying this measured value by esti-
mated energy of lightning event. (2) A ground-based NOx measurement is made in
the vicinity of a natural lightning flash and from it the source strength is estimated
by making suitable assumptions concerning the fluid dynamics of the NOx flow
from the source to the measurement point.
*This material was published previously by the same authors in Journal of Atmospheric and Solar-
Terrestrial Physics, vol. 71, pp. 1877–1889, 2009. It is reproduced here by permission from the journal.
1
Department of Electrical Engineering, Uppsala University, Sweden
2
Department of Electrical and Computer Engineering, University of Florida, USA
538 Lightning electromagnetics: Volume 2
channel radius of sparks. Cooray and Rahman [15] have made a comparison of the
results predicted by (14.1) with the constant suggested by Braginskii [14] with the
experimental data published by Flowers [16] and Higham and Meek [17] and found
that it overestimates the radius of spark channels. Table 14.1 gives channel radii
observed in the experiment and constants k in (14.1) that give the best fit to the
experimental data. Based on this comparison, Cooray and Rahman [15] estimated that
k = (0.328 0.05) 103. Recently, Perera et al. [18] analysed the diameter of spark
channels of length 30 cm using a photographic technique and the maximum channel
diameter is obtained as a function of spark peak current for both positive and negative
polarities. Perera et al. [18] compared the measured maximum channel diameter with
the one obtained from (14.1) using the current waveform measured in the discharge as
input. The results of that study confirm that the Braginskii’s original constant over-
estimates the channel diameter, whereas the constant suggested by Cooray and
Rahman [15] (i.e. 0.328 103) provides a reasonable fit to the data. Based on these
experimental validations, the value of the constant suggested by Cooray and Rahman
[15] is used in the analysis presented here. Lightning currents were directly measured
on tall towers and at the triggered lightning channel base (e.g. Berger [19]; Fisher
et al. [20]). Mathematical expressions to describe the waveform of typical first and
subsequent return-stroke currents recorded by Berger [19] are found in CIGRE Study
Committee 33 Report [21] and Nucci et al. [22]. First let us utilize the typical current
waveforms found in the above references to study how the radii of the first and
subsequent return strokes vary as a function of peak current. Figure 14.1 shows how
the maximum radius of the channel given by (14.1) varies as a function of peak
current for first and subsequent return strokes. Note that the radius of a typical first
return stroke (with a peak current of 30 kA) is about 2 cm and that of a typical
subsequent stroke (peak current 12 kA) is about 1 cm. These values are in agreement
with the radii of return-stroke channels estimated in photographic studies as sum-
marized by Orville [23].
Table 14.1 Measured spark radii and values of k in (14.1) giving the best fit to data
0.035
0.03
0.025
(i)
Radius (m)
0.02
(ii)
0.015
0.01
0.005
0 20 40 60 80
Peak current (kA)
Figure 14.1 The maximum radius of the channel as given by (14.1) as a function
of peak current for first (i) and subsequent (ii) return strokes. In these
calculations, the typical current wave-shapes for first and subsequent
return strokes found in the literature were used
the cooling of the channel takes place mainly due to the entrainment of cold air
across the channel boundaries into the hot core of the channel effectively reducing
the diameter of the hot core [31]. Similarly, (14.1) predicts that the spark channel
expands initially with time and its maximum radius is attained in a few micro-
seconds to a few tens of microseconds, depending on the peak and the duration of
current. After this, the radius of the hot channel decreases, while the channel
resistance starts to increase with time. Based on the experimental and theoretical
data [24–26,28–30], we assume in our analysis that the pressure in the channel at
the time of maximum radius is close to atmospheric pressure and that the average
temperature of the hot air in the channel at this time is close to 15,000 K. In reality,
the channel temperature is not uniform across the cross-section of the channel. But
the theoretical calculations of Paxton et al. [28] and Hill [32] indicate that after a
few tens of microseconds the radial temperature distribution is relatively flat with a
sharp decrease to ambient temperature at the channel boundaries. This justifies the
use of an average value to describe the channel temperature. Further, the tem-
perature of 15,000 K estimated by Orville [27] when the channel was close to
pressure equilibrium is the average temperature across the channel justifying our
selection of 15,000 K as the average temperature. Thus, we assume that in a
cylindrical spark channel the volume of air, V, heated to a temperature of about
15,000 K is given by
V ¼ lprmax
2
(14.2)
where l is the length of the spark and rmax is the maximum radius of the spark
channel given by (14.1) with k = 0.328 103.
NO at Tf in the mixture is frozen out. Thus, if N is the number of air molecules heated
above Tf, the number of NO molecules generated by the process, MNO, is given by
MNO ¼ Nf ðTf Þ (14.3)
where f (Tf) is the fraction of NO molecules in the gas at temperature Tf [33,34].
The value of Tf depends on the cooling rate of the gas in the discharge. For
lightning-like discharges, Tf and corresponding f (Tf) were estimated to be around 2,660
K and 0.029 respectively [33,34]. These values have been used in the calculations
presented here. In order to evaluate MNO it is necessary to evaluate N, the number of
molecules heated above Tf, which can be found from the following equation:
Eh To
N ¼V No (14.4)
Ef Tf
where V is the volume of hot air in the discharge channel given by (14.2), Eh and Ef
are the internal energy of air per unit volume at the temperatures Th and Tf, respec
tively, To is the standard temperature and No is the number of molecules per unit
volume at standard temperature and pressure. In the calculations, we have assumed
that Th = 15,000 K, Tf = 2660 K, To = 273 K and No = 2.69 1025 m3. In deriving
(14.4) we have assumed that, as the channel cools after reaching the pressure
equilibrium due to the entrainment of cold ambient air into the discharge channel, not
much of the energy escapes the channel as radiation. Moreover, in the derivation we
have neglected the production of NO, if any, by the shock wave generated by the
discharge. Both these assumptions are supported by the calculations of Hill et al. [31].
The number of NO molecules produced by the discharge is then given by
Eh T0
MNO ¼ f ðTf Þ V N0 (14.5)
Ed Tf
The values of Eh and Ef are calculated using the set of equations given by
Plooster [30,35] describing the variation in internal energy of air as a function of
temperature and pressure. Since the radii of spark channels and hence the amount
of air heated to a given temperature depend not only on the peak current but also on
the temporal variation of the current, the amount of NOx produced by a spark
depends both on the peak current and the current waveshape. Thus, if the current
waveform in the discharge channel is known, then (14.1) to (14.5) can be used to
evaluate the number of NO molecules produced by the discharge.
sparks analysed by Wang et al. [40] has a rise time of 30 ms and a decay time (time
taken by the current to decay to 1/2 of its peak value) of 400 ms. The current
waveform associated with the sparks in the experiment conducted by Rehbein and
Cooray ([39]; the current waveforms are given in Ref. [42]) was oscillatory with a
frequency of 2.8 MHz and decay time of 2 ms. The current waveform in the spark
experiments conducted by Rahman et al. [41] had a rise time of about 0.3 ms and a
decay time of about 25 ms. The NOx production per unit length by these discharges
is evaluated using the equations presented in the previous section and the results,
together with the experimental data, are shown in Figure 14.2. First note that there
is a reasonable agreement between the experimental data and the theory. Second,
note how the NOx production depends on the wave-shape of current. For a given
peak current, a current with a longer duration gives rise to more NOx than a current
with a shorter duration. The reason for this is that the volume of the discharge
channel increases with increasing the duration of current waveform. Since the
volume of the discharge channel is a measure of the internal energy retained in the
1E+022
(a)
(b)
1E+021
(c)
(d)
NOx molecules/m
1E+020
(e)
1E+019
1E+018
1E+017
100 1000 10000 100000
Peak current (A)
Figure 14.2 NOx production efficiency of laboratory sparks and lightning return
strokes as a function of peak current. (a) Theoretical prediction based
on the current waveform of the study conducted by Wang et al. [40],
(b) prediction based on the typical first return-stroke current
waveform, (c) prediction based on the typical subsequent return-stroke
current waveform, (d) prediction based on the current waveform in the
sparks studied by Rahman et al. [41] and (e) prediction based on the
current waveform in the sparks studied by Rehbein and Cooray [39]).
The experimental data corresponding to different studies (Rehbein and
Cooray: hollow triangles; Wang et al.: hollow circles; Rahman et al.:
solid circles) are also shown in the figure
544 Lightning electromagnetics: Volume 2
discharge channel, in long duration currents more energy goes into the internal energy
of the discharge than in short-duration currents. As the internal energy of the dis-
charge increases the mass of air that is being heated beyond the NOx, freeze-out
temperature also increases leading to a higher NOx production. To the best of our
knowledge, this is the first time that the dependence of NOx production on the shape
of the current waveform flowing in the discharge channel is explicitly recognized.
where s is the effective conductivity of the spark channel, r(t) is the radius of the
channel at time t (given by (14.1)) and l is the length of the channel. Braginskii [14]
recommended the use of 104 S/m as the effective conductivity of the channel.
In order to test the validity of this equation, the energy in the sparks studied by
Rahman et al. [41] was evaluated by integrating the product of voltage and current
waveforms. For 35 current waveforms, the total energy calculated from the above
equation agrees within 15%, when the value of s is assumed to be 0.65 104 S/m.
Paxton et al. [28] studied the development of lightning channel taking into account
the detailed physics of the complex electrohydrodynamic and thermodynamic
processes. The current waveform used by Paxton et al. had a linear rise to peak
followed by an exponential decay. The peak value, rise time and decay time of the
current waveform used by Paxton et al. were 20 kA, 5 ms, and 50 ms respectively.
The calculated total energy dissipation in the discharge up to 50 ms was about
5 kJ/m. Equation (14.6) for the same current predicts the same energy dissipation
when s = 104S/m. These comparisons suggest that (14.6) can give a reasonable
value for the total energy dissipated in the discharge for values of s ranging from
0.65 104 to 104 S/m. In the calculations to follow, we will use s = 0.8 104 S/m.
Figure 14.3 depicts the energy dissipation per unit length in electrical dis-
charges having current signatures similar to those of typical first and subsequent
strokes, as a function of peak current. According to Figure 14.3, typical first
(30 kA) and subsequent (12 kA) return strokes will dissipate about 20 and 2.5 kJ/m,
respectively, in channel sections close to ground.
In Figure 14.4, the calculated yield of NOx as a function of the energy dissipated
in the discharge is depicted for current waveforms corresponding to the experiments
conducted by Wang et al. [40], Rehbein and Cooray [39] and Rahman et al. [41]. For
On the NOx production by laboratory electrical discharges and lightning 545
80,000
60,000
Energy (J/m)
(i)
40,000
(ii)
20,000
0
0 20 40 60 80
Peak current (kA)
Figure 14.3 The energy dissipation per unit length in (i) first return strokes and
(ii) subsequent return strokes as a function of peak current
1E+022
(a)
1E+021
(b)
NOx molecules/m
1E+020
(c)
1E+019
1E+018
1E+017
1 10 100 1,000 10,000 1,00,000
Energy (J/m)
Figure 14.4 NOx production efficiency of laboratory sparks and lightning return
strokes as a function of the energy dissipated in the discharge. (a)
Theoretical prediction based on the current waveform of the study
conducted by Wang et al. [40], (b) prediction based on the current
waveform in the sparks studied by Rahman et al. [41] and (c) prediction
based on the current waveform in the sparks studied by Rehbein and
Cooray [39]). The experimental data corresponding to different studies
(Rehbein and Cooray: hollow triangles; Wang et al.: hollow circles;
Rahman et al.: solid circles) are also shown in the figure
546 Lightning electromagnetics: Volume 2
comparison purposes the experimental data are also shown in the same diagram. It is
important to note that the experimental evaluation of the energy dissipated in spark
channels is not trivial. Errors may result from the measurement of voltage across the
spark channel either due to inductances in the circuit or due to the response char-
acteristics of high-voltage dividers. Moreover, for large currents the discharge chan-
nel transverse diameters may reach several centimetres and in the case of short gaps
(about 3 cm), as in the case of Wang et al. [40] study, the electrode effects may
influence the total energy measured. Nevertheless, as one can see in Figure 14.4,
there is a reasonable agreement between the theory and the experiment. One impor-
tant conclusion that can be made from the results of this analysis is that the energy
dissipated in the discharge cannot be used as a scaling factor in extrapolating
laboratory data to lightning. The reason for this is that for a given energy the NOx
production efficiency of a spark depends on the waveform of the discharge current.
channel core, which is at extremely high potential, into the corona sheath. The air
temperature in a streamer is close to the ambient temperature, whereas the electron
temperature can be several tens of thousands of degrees Kelvin. The collision
between energetic electrons and neutral molecules leads to the dissociation of N2
and O2 in the streamer discharges and the resulting chemistry gives rise to both
NOx and O3. However, the theory developed for the NOx generation in hot sparks
cannot be utilized here, because the NOx production process is not controlled by
temperature variation. Recently, Cooray et al. [43, see chapter 13] demonstrated
that a theory developed for studying NOx production by solar proton events [1,44]
could be utilized to calculate the NOx production from corona and streamer dis-
charges. According to this theory, the NOx production rate is approximately equal
to the rate of production of ion pairs during the proton impact. Since the bulk of
ionization in such events is produced by secondary electron impacts, Cooray et al.
[43] applied the same concept to study the NOx production in low pressure gas
discharges, corona discharges and streamer discharges in which the source of
ionization is the electron impacts. Let us assume that the radius of the streamer
channel is Rs and the number of charge particles at the streamer head is Nhead. Since
the number of ionizing events created by a streamer in moving a unit length is equal
to Nhead/2Rs, according to Cooray et al. [43], the number of NOx molecules pro-
duced by a streamer in propagating a unit distance is given by kNhead/2Rs, where k
is the number of NOx molecules generated per ionizing event. Using experimental
data for corona Cooray et al. [43] demonstrated that k 1 for positive polarity and
k 0.6 for negative polarity.
K þ r2
hstrþ ¼ (14.9)
8pe0 eEs þ Rs
where k+ is the number of NOx molecules generated per ionizing event in positive
discharges. Similarly, the number of NOx molecules generated by negative strea-
mers per unit length of the negative leader channel is given by
k r2
hsrt ¼ (14.10)
8pe0 eEs Rs
550 Lightning electromagnetics: Volume 2
where the parameters have the same definition as in previous equation but corre-
sponding to negative streamers. Thus, the total number of NOx molecules generated
by corona sheath of the leaders in the whole ground flash is given by
The evaluation of the contribution to the NOx production from non-vertical in-
cloud channels is more complicated. The channel system inside the cloud may
consist of many branches and at a given time only a few of these branches may be
developing [51–53], and hence the core current is active only along those branches
at that time. Thus, one problem in evaluating the NOx production in the core of the
non-vertical in-cloud channel is the difficulty of knowing the length of the channel
sections in which the current is flowing at a given time. Consider that the non-
vertical in-cloud channels comprise n identical branches connected to the top of the
vertical channel. We presume that these channels in the cloud are also created by
processes similar to that of lightning leaders observed in ground flashes, and
therefore, their current and speed of development are also identical to those of these
leaders. As mentioned in Section 14.5.1, we assume that the core current will flow
in each branch only when that particular branch is being formed. In reality, core
current may pass from an active branch to a previously formed branch thus
increasing the total length of the channel sections supporting a core current at a
given time. Consider the development of a horizontal channel section inside the
cloud. Let us direct the coordinate x along the channel section. Consider an element
dx on this channel located at a distance x from the origin of the section. In this
element, the current flows for a duration of (l – x)/vs where vs is the speed of
development of the channel and l is the length of the channel. Thus, the number of
NOx molecules produced in this channel element is hleaIls–H/lp(l–x)dx/vs. The total
NOx production in the channel section can be obtained by integrating the above
expression from 0 to l. The result of this integration is hleaIlsI2e–H/lp/2vs. Since we
have assumed that there are n identical branches and the total length of the hor-
izontal channels is L the number of NOx molecules produced by the current flowing
through the core of the channels inside the cloud is given by hleaIlsL2e–H/lp/2vsn. A
similar procedure can be used to evaluate the NOx pro duction along the vertical
section of the leader channel, but the mathematics is slightly more complicated due
to the fact that the pressure varies along the channel. After applying the mathe-
matics, one can show that the total number of NOx molecules produced by the core
current in the stepped leader channel of the ground flash (including the branches in
the cloud) is
h Ils 2
NOxslcore ¼ lea lp lp eH=lp ½H þ lp
vs (14.12a)
h Ils
þ lea L2 eH=lp
2vs n
Using the same procedure the number of NOx molecules produced in the dart
leader channel core can be written as
hlea Ild 2
NOxdlcore ¼ ðlp lp eH=lp ½H þ lp Þ (14.12b)
Vd
where Idl is the current in the dart leader and vd the speed of dart leaders. To
evaluate this equation, it is necessary to have values for hlea, n, vs, Ils, vd and Ild. As
552 Lightning electromagnetics: Volume 2
Figure 14.2, NOx production also depends on the peak current of the subsequent
return stroke. From this figure, we estimate that hsr = 1.4 1020 molecules/m.
The use of return-stroke current instead of the energy dissipated in return
strokes as an input parameter in quantifying NOx production in return strokes, as
done above, has at least one important advantage. The energy dissipated in light-
ning flashes cannot be measured directly but has to be inferred by indirect methods
leading to large inaccuracies in the estimated NOx production. On the other hand,
the current at the channel base of ground flashes can be measured and a large
amount of data on this parameter is available in the literature.
In deriving (14.13) and (14.14), it has been assumed that the return-stroke peak
current decreases with height. The experimental observations show that the
luminosity of both first and subsequent return strokes decreases with height indi-
cating that the return-stroke current peak also decreases with height [54,61]. Of
course, since the exact nature of how the return-stroke peak current decreases with
height is not known, one has a freedom to select any other form of decay for the
peak current than the linear decay assumed in the calculation. However, the
expression describing a linear decay with current amplitude decreasing to zero at
cloud level involves only one parameter, i.e. height of the vertical channel, and it
also specifies the boundary conditions for the current at the cloud end of the
channel. Moreover, the linear decay has been shown to produce electric fields at
both far and close distances that are similar to those measured when used in return-
stroke models [62,63].
Now, let us evaluate the magnitude of hm. A typical M component current has
a more or less symmetric bell-shaped waveform with a rise time of about 400 ms
and a peak of about 160 A [64]. Calculations done with such a current waveform
show that hm = 2 1020. Thus, hk is also equal to 2 1020. M components travel
along the vertical channel carrying continuing currents. In a typical ground flash
having a continuing current, M component may outnumber the number of return
strokes by 1 to 4. Therefore, a typical ground flash containing continuing currents
may support about 16 M components along the vertical channel. In making the
above statement, we have assumed that a typical ground flash contains four return
strokes. On the other hand, the percentage of ground flashes containing long
(longer than 40 ms) continuing currents is about 30–50% [65,66]. This fixes the
value of nm to about 5.
where hcon is the number of NOx molecules produced per Coulomb by a unit length
of the discharge channel carrying continuing current, Icon is the magnitude of
continuing current, tc is the typical duration of continuing current and kc is the
fraction of ground flashes that support continuing currents. In the calculations we
assumed that hcon = 2 1020 molecules/m/C (see Section 14.3). According to
experimental data, about 30–50% of the lightning flashes contain continuing cur-
rents and the amplitude and the duration of a typical continuing current are about
100 A and 100 ms respectively [65,66]. Thus, kc = 0.3 and tc = 0.1 s.
The exact nature of the source that drives continuing current along the vertical
channel is not known. Most likely the source is the development and the charge
transfer along the upper channel sections. Since we have taken this activity into
account in Section 14.5.2.1 under leaders, it is reasonable to consider only the
vertical channel here.
On the NOx production by laboratory electrical discharges and lightning 555
0 20 40 60 80 100
1027
Horizontal channel length (km)
(d)
(b)
1025
(e)
(f)
1024
0 10 20 30 40 50
Charge neutralized (C)
not vary with increasing length of the horizontal channels because these currents
are assumed to propagate only along the vertical channel. Second observe that the
return strokes produce the least contribution to the NOx production, whereas the
largest contribution is made by leaders with the current flowing along the core
being the main contributor. More than 90% of the contribution to the NOx pro-
duction is coming from leaders, M components and K processes. Leaders alone
contribute about 50% to the NOx production. Note that these observations are true
also for horizontal channel lengths as short as 10 km. On the other hand, in the
literature, the return stroke is often assumed to be the NOx source in ground flashes.
Our study shows that this assumption is incorrect. Using VHF lightning channel
mapping technique, Laroche et al. [69] observed that the mean total channel length
in over 20,000 cloud and ground flashes is 45 km. Taking this length as a typical
value the results presented in Figure 14.5 show that an average ground flash with
four return strokes will generate about 4 x 1025 NOx molecules per flash.
Shao et al. [51] and Shao and Krehbiel [52] made important discoveries utilizing
VHF radio imaging techniques. Based on this information, Cooray [45] summar-
ized the activities during a cloud flash as follows (see also [65]).
The cloud flash commences with a movement of negative leader discharge
from the negative charge region towards the positive one in a more or less vertical
direction. The vertical channel develops within the first 10 to 20 ms from the
beginning of the flash. This channel is a few kilometres in length and it developed
with a speed of about 2.0 105 m/s. Even after the vertical channel was formed,
one could detect an increase in the electrostatic field indicative of negative charge
transfer to the upper levels along the vertical channel.
The main activity after the development of the vertical channel is the hor-
izontal extension of the channels in the upper level (i.e. the channels in the positive
charge region). These horizontal extensions of the upper level channels are corre-
lated to the brief breakdowns at the lower levels, followed by discharges propa-
gating from the lower level to the upper level along the vertical channel. Thus, the
upper level breakdown events are probably initiated by the electric field changes
caused by the transfer of charge from the lower levels. For about 20 to 140 ms of
the cloud flash, repeated breakdowns occur between the lower and upper levels
along the vertical channel. These discharges transported negative charge to the
upper levels. Breakdown events of this type can be categorized as K changes. In
general, the vertical channels through which these discharges propagate do not
generate any radiation in the VHF range, which indicates that they are conducting.
This is so because, in general, conducting channels do not generate VHF radiation
as discharges propagate along them. Occasionally, however, a discharge makes the
vertical channel visible at VHF and then the speed of propagation can be observed
to be about (5–70) 106 m/s, typical of K changes. This active stage of the dis-
charge may continue to about 200 ms.
In the latter part of this active stage (140–200 ms), significant extensions of the
lower level channels (i.e. the channel in the negative charge region) take place, but
they occur retrogressively. That is, successive discharges, or K changes, often start
just beyond the outer extremities of the existing channels and then move into and
along these channels, thereby extending them further. These K changes transport
negative charge from successively longer distances to the origin of the flash, and
sometimes even to the upper level of the cloud flash as inferred from VHF emis-
sions from the vertical channel. Sometimes, these K changes give rise to discharges
that start at the origin of the flash and move away from it towards the origin of the
K changes. Such discharges can be interpreted as positive recoil events that trans-
port positive charge away from the flash origin and towards the point of initiation
of the K change. At the final part of the discharge, the vertical channel and the
upper level channels were cut off from the lower level channels. This is probably
caused by the decrease in the conductivity of the vertical channel. The above
description shows that a cloud flash can be described as an electrical activity that
collects the charge from the main negative charge centre and redistribute it in the
positive charge centre after transporting it along a more or less vertical channel.
The recent observations based on three-dimensional interferometry also confirm
558 Lightning electromagnetics: Volume 2
0 20 40 60 80 100
1027
Horizontal channel length (km)
1026 (c)
(d)
(b)
1025
(f)
(e)
1024
0 10 20 30 40 50
Charge neutralized (C)
NOx production from a typical positive flash is similar to that of a typical negative
flash. In the literature, the flash rate is assumed to lie in the range of 40–300
flashes/s [80,81]. In the results to be presented, we assumed a global lightning flash
rate of 100/s. There is no reason to separate the flash rate into ground and cloud
flashes because both types of flashes produced more or less the same amounts of
NOx. In Figure 14.7, we have depicted the annual NOx production by lightning
flashes as a function of the horizontal channel length. If one assumes an average
total channel length of 45 km for a lightning flash, the global NOx production by
lightning flashes will be about 4 Tg(N)/year. One has to understand that the global
lightning flash frequency is not a constant and it may vary from one year to another.
Moreover the number 100 flashes/s is based on thunderstorm observations and
satellite data suggest values in the range of 40–60 flashes/s. But, of course, this
estimation too depends on the detection threshold level of the satellite and the
possible screening of light by cloud cover. The important point however is that
the NOx production rate is proportional to the global lightning flash frequency. The
number 4 Tg(N)/year is based on 100 flashes/s and if it varies between 40 and 300
flashes/s, the global NOx production rate will also vary between 2 and 12 Tg(N)/
year. Lee et al. [82] studied the various sources and sinks of NOx in the atmosphere
and concluded that the contribution from lightning should be in the range of 4–8 Tg
(N)/year. Our results agree with this prediction. The present global estimates
of NOx based on theoretical and laboratory studies vary between 1 and about
100 Tg(N)/year. The two orders of magnitude variation in this estimate are due to
560 Lightning electromagnetics: Volume 2
0 20 40 60 80 100
12
Horizontal channel length (km)
10
0
0 10 20 30 40 50
Charge neutralized (C)
the different results obtained for the NOx production efficiency of laboratory dis-
charges and the different values assumed for the energy dissipation in lightning
flashes. As pointed out previously, the variation in the efficiency of NOx production
in laboratory discharges is probably due to the differences in the current waveforms
associated with these discharges. The energy dissipation in lightning flashes is a
parameter that cannot be measured directly and therefore is not a good scaling
quantity in NOx studies. Our estimate is free from both these drawbacks. Another
interesting point of our study is the observation that most of the NOx production in
lightning flashes is due to cloud flashes or the cloud portion of ground flashes.
Thus, the injection of NOx by thunderstorms into the atmosphere takes place pri-
marily at a height of 5–10 km. The theoretical studies conducted by Gallardo and
Rodhe [2] show that in order to account for the nitrate deposition in the remote
marine regions, the strength of the NOx source due to lightning should be about
5 Tg(N)/year and the source should be located at the cloud height. Our study
confirms this inference.
14.8 Conclusions
The results presented in this chapter show that the NOx production efficiency of
electrical discharges depends not only on the energy dissipated in the discharge but
also on the shape of the current waveform. This provides an explanation for the
different values of NOx molecules/J obtained by different researchers in different
experiments. Thus, energy dissipated in a discharge is not suitable as the scaling
On the NOx production by laboratory electrical discharges and lightning 561
quantity for extrapolating the laboratory data to lightning flashes. In this chapter,
we present a theory that can be used to evaluate the NOx production in electrical
discharges, if the discharge current is known. The results obtained are compared
with the available experimental data and a good agreement is found between theory
and experiment. The study shows that the primary contribution to NOx from
thunderstorms is coming from the electrical activity inside the cloud, with only a
small fraction being contributed by return strokes. Using the proposed theory, we
estimated the global NOx production by lightning flashes taking into account dif-
ferent lightning processes such as leaders, return strokes, M components, K changes
and continuing currents. The results show that the efficiency of NOx production in
ground flashes and cloud flashes are similar and for an average total channel length
of 45 km the global production of NOx by lightning flashes, based on lightning
flash frequency of 100 flashes/s, is about 4 Tg(N)/year.
Appendix 1
hstr – The number of NOx molecules generated in the corona sheath during the creation of
a leader channel. The current estimate is 2 1020 molecules/m.
hcon – The number of NOx molecules generated per unit length per unit charge by a
continuing current. The best estimate is 2.0 1020 molecules/m/C.
hlea – The number of NOx molecules generated per unit length per unit charge by the leader
current flowing through the channel core. The best estimate is 2.0 1020 molecules/
m/C.
lP – The decay height constant for the atmospheric pressure. This is equal to 8500 m.
H – The height of the negative charge centre. The value used in the calculations is
5000 m.
Hp – The height of the positive charge centre. The value used in the calculation is
10,000 m.
L – The total length of the horizontal sections in the cloud. Current estimates place it
somewhere between 30 and 50 km.
hfr – The number of NOx molecules generated per unit length in a discharge channel
carrying a current waveform similar to that of a typical first return stroke. The best
estimate is 5.9 1020 molecules/m.
hsr – The number of NOx molecules generated per unit length in a discharge channel
carrying a current waveform similar to that of a typical subsequent return stroke.
The best estimate is 1.4 1020 molecules/m.
ns – The number of subsequent return strokes in a typical ground flash. The best estimate is 3.
hm – The number of NOx molecules generated per unit length in a discharge channel
carrying a current waveform similar to that of M component. The best estimate is
2.0 1020 molecules/m.
hk – The number of NOx molecules generated per unit length by a K change. It is assumd
that hk = hm.
nk – The average number of K changes taking place during the development of a given
channel branch in the cloud. It is assumed to be 3.
nm – The average number of M components in a typical ground flash. The best estimate is
5. This number is based on the fact that a ground flash with a continuing current can
support about 16 M components and about 30% of the ground flashes contain
continuing currents.
(Continues)
562 Lightning electromagnetics: Volume 2
(Continued)
kc – The fraction of ground flashes containing continuing currents. The best estimate is
0.3.
tc – The average duration of continuing current. The best estimate is 100 ms. This
figure is actually valid for long continuing currents.
Icon – Magnitude of typical continuing current. The best estimate is 100 A.
Ils – Magnitude of typical stepped leader current. The best estimate is 100 A.
Ild – Magnitude of typical dart leader current. The best estimate is 1 kA.
vs – The average speed of development of lightning leader channels in virgin air inside
the cloud. The best estimate is 2105 m/s.
vd – The average speed of dart leaders. The best estimate is 107 m/s.
n – The number of major branches inside the channel. The available VHF interfero-
metric images show that it may vary from about 3 to 10. It is assumed to be 5.
References
[1] Crutzen, P. J., The influence of nitrogen oxides on the atmospheric ozone
content, Quart. J. R. Met. Soc., vol. 96, pp. 320–325, 1970.
[2] Gallardo, L. and H. Rodhe, Oxidized nitrogen in the remote pacific: the role of
electrical discharges over the oceans, J. Atmos. Chem., vol. 26, pp. 147– 168,
1997.
[3] Tuck, A. F., Production of nitrogen oxides by lightning discharges, Quart. J.
R. Met. Soc., vol. 102, pp. 749–755, 1976.
[4] Noxon, J. F., Atmospheric nitrogen fixation by lightning, Geophys. Res.
Lett., vol. 3, no. 8, pp. 463–465, 1976.
[5] Chameides, W. L., D. H. Stedman, R. R. Dickerson, D. W. Rusch and R. J.
Cicerone, NOx production in lightning, J. Atmos. Sci., vol. 34, pp. 143–149, 1977.
[6] Drapcho, D. L., D. Sisterson and R. Kumar, Nitrogen fixation by lightning
activity in a thunderstorm, Atmos. Environ., vol. 17, no. 4, pp. 729–734, 1983.
[7] Franzblau, E. and C. J. Popp, Nitrogen oxides produced from lightning,
J. Geophys. Res., vol. 94(D8), pp. l1089–11104, 1989.
[8] Stith, J., J. Dye, B. Ridley, P. Laroche, E. Defer, K. Baumann, G. Hu·bler, R.
Zerr and M. Venticinque, NO signatures from lightning flashes, J. Geophys.
Res., vol. 104, pp. 16081–16089, 1999.
[9] DeCaria, A. J., K. E. Pickering, G. L. Stenchikov and L. E. Ott, Lightning
generated NOX and its impact on tropospheric ozone production: A three
dimensional modeling study of a stratosphere-Troposphere Experiment:
Radiation, Aerosols and Ozone (STERAO-A) thunderstorm, J. Geophys.
Res., vol. 110, no. D14303, doi:10.1029/2004JD005556, 2005.
[10] Cook, D. R., Y. P. Liaw, D. L. Sisterson and N. L. Miller, Production of
nitrogen oxides by a large spark generator, J. Geophys. Res., vol. 105, no.
D6, pp. 7103–7110, 2000.
[11] Huntrieser, H., C. Feigl, H. Schlager, et al., Airborne measurements of NOX,
tracer species, and small particles during the European Lightning Nitrogen
On the NOx production by laboratory electrical discharges and lightning 563
[44] Nicolet, M. On the production of nitric oxide by cosmic rays in the meso-
sphere and stratosphere, Planet. Space Sci., vol. 23, pp. 637–649, 1975.
[45] Cooray, V., The mechanism of the lightning flash, in The Lightning Flash,
ed. V. Cooray, pp. 127–239, Institute of Electrical Engineers, UK, 2003.
[46] Gallimberti, I., The mechanism of long spark formation, J. Phys., vol. 40,
suppl. 7, pp. 193–250, 1979.
[47] Les Renardiéres Group, Research on long air gaps discharges at Les
Renardiéres, Electra, vol. 35, pp. 49–156, 1974.
[48] Jackman, C. H., H. S. Porter and J. E. Frederick, Upper limits on production
rate of NO per ion pair, Nature, vol. 280, p. 170, 1979.
[49] Zipf, E. C. and S. S. Prasad, Evidence for new sources of NOx in the lower
atmosphere, Science, vol. 279, pp. 211–213, 1998.
[50] Rahman, M. and V. Cooray, A study of NOx production in air heated by
laser discharges: Effects of energy, wavelength, multiple discharges and
pressure, Optic. Laser Tech., vol. 40, pp. 208–214, 2008.
[51] Shao, X. M., P. R. Krehbiel, R. J. Thomas and W. Rison, Radio interfero-
metric observation of cloud to ground lightning phenomena in Florida, J.
Geophys. Res., vol. 100, pp. 2749–2783, 1995.
[52] Shao, X. M. and P. R. Krehbiel, The spatial and temporal development of
intracloud lightning, J. Geophys. Res., vol. 101, pp. 26641–26668, 1996.
[53] Thomas, R. J., P. R. Krehbiel, W. Rison, S. J. Hunyady, W. P. Winn, T.
Hamlin and J. Harlin, Accuracy of the Lightning Mapping Array, J. Geo-
phys. Res., vol. 109, no. D14207, doi:10.1029/2004JD004549, 2004.
[54] Schonland, B. F. J., The lightning discharge, in Handbuch der Physik,
vol. 23, pp. 576–628, Springer, New York, 1956.
[55] Schonland, B.F.J., D. J. Malan and H. Collens, Progressive lightning, 2,
Proc. R. Soc. London, Ser. A, vol. 152, pp. 595–625, 1935.
[56] Orville, R. E. and V. P. Idone, Lightning leader characteristics in the
Thunderstorm Research International Program (TRIP), J Geophys. Res.,
vol. 87, no. C13, pp. 11,177–11,192, December 20, 1982.
[57] Wang D., N. Takagi and T. Watanabe, V. A. Rakov, and M. A. Uman,
Observed leader and return-stroke propagation characteristics in the bottom
400 in of a rocket-triggered lightning channel, J Geophys. Res., vol. 104, no.
D12, pp. 14,369–14,376, June 27, 1999.
[58] Idone, V. P. and R. E. Orville, Correlated peak relative light intensity and
peak current in triggered lightning subsequent strokes, vol. 90, pp. 6159–
6164, 1985.
[59] Cooray, V., P. Idone. and R.E. Orville, Velocity of a self-propagating dis-
charge as a function of current parameters with special attention to return
strokes and dart leaders, papers presented at the 1989 International Con-
ference on Lightning and Static Electricity, pp. 1A.3.1–1A.3.9, University of
Bath, England, September 26–28, 1989.
[60] Kodali, V., V. A. Rakov, M. A. Uman, et al., Triggered-lightning properties
inferred from measured currents and very close electric fields, Atmos. Res.,
vol. 76, no. 1–4, pp. 355–376, 2005.
566 Lightning electromagnetics: Volume 2
broadband digital interferometry, Atmos. Res., vol. 76, no. 1–4, pp. 445–454,
2005.
[77] Gallardo, L. and V. Cooray, Could cloud-to-cloud discharges be as effective
as cloud-to-ground discharges in producing NOx?, Tellus, Ser B, vol. 48,
pp. 641–651, 1996.
[78] Dye, J. E., B. A. Ridley, W. Skamarock, et al., An overview of the
Stratospheric – Tropospheric Experiment: Radiation, Aerosols and Ozone
(STERAO) – Deep convection experiment with results for the July 10,
1996 storm, J. Geophys. Res., vol. 105, pp. 10023–10045, 2000.
[79] DeCaria, A. J., E. Pickering, G. L. Stenchikov, et al., A cloud scale model
study of lightning- generated NOx in an individual thunderstorm during
STERAO-A, J. Geo phys. Res., vol. 105, pp. 11601–11616, 2000.
[80] Turman, B. N. and B. C. Edgar, Global lightning distribution at dawn and
dusk, J. Geophys. Res., vol. 87, pp. 1191–1206, 1982.
[81] Christian, H. J, R. J. Blakeslee, D. J. Boccippio, et al., Global frequency and
distribution of lightning as observed from space by the Optical Transient
Detector, J. Geophys. Res., vol. 108, no. D1, p. 4005, doi:10.1029/
2002JD002347, 2003.
[82] Lee, D. S., I. Köhler, E. Grobler, et al., Estimation of global NOx emissions
and their uncertainities, Atmos. Environ., vol. 31, pp. 1735–1749, 1997.
This page intentionally left blank
Chapter 15
Lightning and climate change
Earle R. Williams1, Joan Montanya2, Joydeb Saha3 and
Anirban Guha3
15.1 Introduction
1
Parsons Laboratory, Department of Civil and Environmental Engineering, Massachusetts Institute of
Technology (MIT), USA
2
Department of Electrical Engineering, Universitat Politècnica de Catalunya, Spain
3
Department of Physics, Tripura University, India
570 Lightning electromagnetics: Volume 2
Precipitation (mm)
Figure 15.1 Global climatologies of (a) lightning flash density (lightning imaging
sensor), (b) rainfall (NASA TRMM), and (c) aerosol concentration
(as measured with satellite aerosol optical depth). Adapted from [5]
572 Lightning electromagnetics: Volume 2
0° C
Figure 15.2 The mechanism of the thunderstorm: storm cloud with colliding ice
particles and positive electric dipole
Lightning and climate change 573
collisions of two kinds of particles: small ice crystals and larger graupel particles. Both
ice particles are the product of mixed-phase conditions involving water substance in all
three thermodynamic phases: vapor, liquid, and solid (ice). The liquid phase at “cold”
temperature (T < 0 C) is referred to as supercooled water. The ice crystals form by
diffusion of water vapor in the so-called Bergeron process based on the asymmetry in
equilibrium vapor pressure between liquid and ice. The mass of ice crystals increases at
the expense of the supercooled cloud water. The graupel particles grow by the accretion
of supercooled cloud droplets, which freeze in contact with the graupel surface. The
collisions between graupel particles and ice crystals result in the transfer of negative
charge to graupel and positive charge to the crystals, by a mechanism at the molecular
scale that has long eluded scientists [25,26], but very likely involving mobile protons as
main agents of charge transfer. The descent of negative graupel with respect to positive
ice crystals under gravity sets up the macroscopic positive dipole of the thunderstorm
(Figure 15.2). This hydrometeor-based mechanism of differential charge separation
(based on different fallspeeds of ice crystals and graupel) is immune to the effects of
turbulence, which often shows a strong presence in thunderstorms.
15.3.1 Temperature
The most commonly used thermodynamic parameter in global climate change is the
temperature of surface air, typically measured at “screen level.” The formal meteor-
ological quantity is “dry bulb temperature” to distinguish it from “wet bulb temperature”
and “dew point temperature” both of which involve the water vapor content of the air.
Traditional estimates of the global mean temperature [27,28] and global warming
involve averages of 4,000–6,000 thermometer readings of dry bulb temperature over the
Earth’s surface. This temperature parameter is also a key factor in other thermodynamic
quantities of interest here: saturation water vapor concentration, convective available
potential energy (CAPE) and cloud base height (CBH), described in greater detail below.
condensation Lv is 2.5 106 J/kg of water, sufficient energy to raise the con-
densate 250 km against gravity if this transformation took place with perfect
efficiency.
The water vapor concentration in the atmosphere in a condition of thermo-
dynamic equilibrium is controlled by temperature in an exponential dependence
known as the Clausius–Clapeyron relation. In differential form:
100
10
1
Saturation Mixing Ratio (g/kg)
0.1
Supercooled
Water
Ice
0.01
0.001
0.0001
–80 –70 –60 –50 –40 –30 –20 –10 0 10 20 30 40 50
Temperature (ºC)
Figure 15.3 Equilibrium water vapor mixing ratio (g/kg) versus temperature: the
Clausius–Clapeyron relationship
Lightning and climate change 575
consideration of the global lightning climatology shows that two of every three
lightning flashes lie within 23 of the equator [30].
The slope de*/dT in the Clausius–Clapeyron relationship in (15.1) is a valuable
benchmark for judging results on the response of lightning to temperature on var-
ious time scales (Section 15.4). This slope depends on temperature, but at the mean
Earth surface temperature (14 C), the slope is 7% per 1 C.
Cumulonimbus clouds are the primary agents for transporting water vapor
from the planetary boundary layer in the lower troposphere to the upper tropo-
sphere. Consistent with this general picture, Price [31] has shown variations in
upper tropospheric water vapor correlated with lightning variations over the
African continent.
100
θw = 24 25.8 28 –70 –60 –50 –40
EL
150
–30
Moist
Adiabat
200
–20
Temperature
250 Profile
–10
300
0
400
T
10
500
Td 20
w = 16 g kg –1 30
LFC
700
CCL θ = 39 40
850
1000
Figure 15.4 Temperature sounding, wet bulb adiabatic approximation for the
updraft temperature, and the area representing convective available
potential energy (CAPE)
but it is now recognized that for maritime convection drawing from clean boundary
layer air, the pseudoadiabatic irreversible process was a more appropriate choice.
Given the negative buoyancy contribution of condensate loading, reversible CAPE
is expected to be systematically less than irreversible CAPE. This expectation is
consistent with numerous published results [29,36]. In early considerations of the
reversible and irreversible processes [37], it was concluded that the “differences
between the products of condensation as falling out or being retained are so small
as to be negligible in practice,” but today this difference is acknowledged as being
all important in deep moist convection.
Williams and Renno [36] also showed that CAPE in the current climate was
well predicted by the wet bulb potential temperature of surface air, though different
relationships were apparent for land and ocean. Global maps of wet bulb potential
temperature show that maximum CAPE over land is greater than over ocean. Lucas
et al. [38,39] claimed that CAPE over land was similar to values over ocean but
they did not consider the diurnal variation of CAPE over land. This plays an
important role in the transition of cumulus congestus clouds to thunderstorms over
land, with important contributions from thermodynamics. A global climatology of
CAPE has been prepared [40]. These results also show that CAPE is larger over
continents than over oceans, though no consideration was given to aerosol-related
effects in condensate loading. The land–ocean CAPE contrast is qualitatively
consistent with the land–ocean lightning contrast, but on closer examination
[11,38,41,42], the contrast is not sufficient to account for the order-of-magnitude
contrast in lightning. In the latter work however, precipitation is considered at the
surface rather than in the cold region aloft where electrical energy is generated by
the relative descent of graupel particles with respect to ice crystals. This aspect will
be revisited below in the context of CBH. More recently [39] have shown evidence
for giant sea salt nuclei in suppressing lightning in oceanic convection.
Given the primary role of CAPE in the charge separation and lightning activity of
thunderstorms, the variation of CAPE with temperature on the long-time scale of
global warming is of considerable interest. This problem is nontrivial because the
entire temperature profile is involved, as well as the condensate-related ambiguities of
the wet bulb adiabat. In early work [43] CAPE was postulated to be a climate invariant.
However, many GCM results show CAPE to increase with global warming
[44,45,46,47], and Del Genio et al. [48] have found increases in cumulonimbus velo-
city in climate models in a warmer world. Furthermore, still more recent theoretical
works [42,49,50] support a scaling of CAPE with the Clausius–Clapeyron exponential
temperature dependence. On this basis alone, one expects to have more lightning in a
warmer climate. However, one recent model results [51] shows the opposite result for
the tropics. This contrast in predictions is not well understood at present.
15.3.5 Cloud base height and its influence on cloud
microphysics
The contrast in physical characteristics between land and ocean surfaces exerts an
important influence on the behavior of thermodynamic parameters of surface air.
The contrast in heat capacity and mobility between land and ocean affect the
578 Lightning electromagnetics: Volume 2
surface air temperature, with ocean water and overlying air resisting temperature
increase in response to solar heating, in comparison with dry land surfaces.
The diurnal variation of ocean surface temperature is typically a fraction of 1 C.
In contrast, the diurnal variation of land surface temperature invariably exceeds
1 C but the surface temperatures over deserts can vary by tens of C. The contrast
in available surface water between land and ocean affects the dew point tempera-
ture and relative humidity of surface air. Land surfaces are generally both hotter
(larger T) and drier (smaller Td) than oceans, and as a consequence of both of these
contributions, the dew point depression of surface air (T Td), a purely thermo-
dynamic quantity, is invariably larger over land than ocean. [Global maps of day-
time dew point depression (and equivalently CBH) would show marked land–ocean
contrasts as in Figure 15.1(a) and (c).] See [52].
The convenience of cloud physics is that the lifted condensation level (LCL)
and CBH are both proportional to T Td. If T = Td, the air is saturated (RH =
100%) and the cloud extends downward to the surface. Over oceans, typical CBHs
WT
0° C
Balance
W
Level
WT
(a)
Balance
W WT
Level
0° C
WT
(b)
Figure 15.5 Illustration of typical continental and maritime convection at (a) the
time of radar first echo and (b) at the cumulonimbus stage. The
balance level heights in all cases are also indicated
Lightning and climate change 579
are 500 m (corresponding to typical relative humidity of 80%), but over land can
vary from 1,000 to 5,000 m (with relative humidity in the range of 70%–20%).
The contrast in CBH between continental and maritime environments is illu-
strated in Figure 15.5(a), for afternoon clouds at the first-radar-echo stage. The
oceanic cloud achieves the first echo while still a warm cloud. In contrast, the con-
tinental cloud with systematically higher CBH is usually extended into the cold
region (< 0 C) of the atmosphere at first-echo stage. The heights of the 0 C iso-
therm are similar for land and ocean, near 4,500 m MSL, but the CBHs differ
markedly for thermodynamic reasons. Figure 15.5(b) shows deeper clouds for both
land and ocean that include the mixed-phase region bounded by the 0 C and 40 C
isotherms where active charge separation can occur under appropriate conditions of
cloud vertical development. The CBHs remain the same and often coincide with the
top of the planetary boundary layer. The cloud widths are different based on obser-
vations showing that continental clouds are broader than maritime ones [11,53,54].
The updrafts in clouds are fundamentally important in regulating cloud
microphysics and electrification. Scaling analysis indicates a sensitive fourth power
relationship of lightning flash rate on updraft speed [55]. Accordingly, modest
changes in CAPE can have substantial effects on lightning flash rate. Figure 15.5
also includes vertical arrows to contrast the updraft speeds in different regions
(including CBH in Figure 15.5(a)) of both shallow and deep convection.
Regarding the subcloud region, Zheng and Rosenfeld [56] have found larger
ascent speeds (by 50%–100%) in the continental boundary layer than the oceanic
one, and larger speeds at CBH, consistent with predictions based on thermo-
dynamics and the contrast in surface properties in Williams and Stanfill [11].
Puzzlingly, model results on deep clouds with greater CBH do not show evidence
for larger updraft speeds [57]. It should be remembered however that the land/
ocean lightning contrast is more strongly controlled by numbers of thunderstorms
than by lightning activity per storm [6].
Earlier studies [53,54] have shown systematically larger updraft speeds in deep
moist continental convection over land than over ocean. The contrast in ascent speeds
is unmistakably linked with a contrast in the ice phase microphysics and lightning
activity between land and ocean, but the explanation for the contrast in ascent speeds
remains a controversial issue [11,41,57]. When thunderstorms over land alone are
examined, lightning flash rate and CBH are positively correlated in global compar-
isons with the lightning imaging sensor in space and with surface thermodynamic
observations of dew point depression [58]. These measurements need to be con-
trolled for CAPE and aerosol variations to narrow down the physical causality.
The formation of precipitation within the updraft of convective clouds is important
because the precipitation can load the updraft, and ultimately reduce the updraft speed.
This issue was raised initially in the context of CAPE, but here one can be more
quantitative by estimating the precipitation content that will offset the effect of thermal
buoyancy. Simple considerations of Archimedean buoyancy show that the force per
unit mass associated with a temperature perturbation DT is simply DT/Tg, where T is
the ambient temperature. The negative buoyancy contribution (again force per unit
total mass) from additional mass loading m in a parcel with mass of air M is simply
580 Lightning electromagnetics: Volume 2
m/(m + M)g. For a parcel at the 0 C isotherm with nominal air density of 0.6 kg/m3,
the condensate loading needed to offset a thermal buoyancy of 1 C is 2.0 g/m3.
Extensive documentation on the detection of first radar echo and the formation
of precipitation in deep convection has come from radar studies (good summary in
Ludlam [32], with additional examples in [59]). The D6 dependence of the radar
cross-section of precipitation particles gives radar considerable sensitivity in detect-
ing the formation of precipitation, given the runaway nature of coalescence of cloud
droplets in the diameter range 25 mm [60] where the rate of droplet coalescence
varies steeply with droplet size as D5. The radar studies have shown that cloud depths
(cloud top height minus CBH) in the 4,000–5,000 m range over continents and as
small as 2,000 m over oceans [32,59] are needed for first echo development. These
radar-based estimates are broadly consistent with aircraft in situ measurements of
cloud depths needed to achieve critical cloud droplet size [60]. The relevance of these
results in the context of Figure 15.5 is that the warm cloud depth in the maritime case
(4,000 m), a thermodynamic effect, is large in comparison with that needed for the
formation of precipitation. In the continental case, this condition is not fulfilled.
These comparisons are consistent with observations of radar first echoes that appear
consistently in the “warm” part of the cloud over oceans, but more typically in the
“cold” part of the cloud over land [32,59,61]. In a more global context, the results are
also consistent with observations that warm precipitating clouds are prevalent over
oceans and scarce over land [11,62]. Still, the conundrum remains between a ther-
modynamic effect and an aerosol effect (Section 15.5). The lower CBHs over ocean
overlie cleaner air with more dilute CCN concentrations [63]. Accordingly, following
the discussion in Section 15.5, the cloud droplets above CBH will be larger and more
prone to form precipitation and radar first echoes at lower heights. Braga et al. [63]
give emphasis to the aerosol effect and neglect the thermodynamic effect.
The formation of precipitation in moist convection is important because it can
then descend with respect to the air parcel in which it forms, and thereby load the
updraft column beneath. (In barotropic conditions typical of tropical convection,
the updraft is vertical.) This is sometimes called “super-adiabatic” loading because
the precipitation condensate is added to the adiabatic condensate at lower levels. (In
baroclinic conditions (Section 15.3.7) more typical of convection at higher lati-
tudes, the updraft can be tilted and then the updraft can unload its precipitation, as
is assumed in the irreversible calculation of CAPE.) Based on observations with a
precipitation radar in space, warm rain clouds over oceans (where the warm cloud
depths are greatest) are capable of achieving precipitation concentrations up to
2–3 g/m3 [62]. These mass loadings are commensurate with cloud buoyancy at the
1 C level as was shown earlier. These superadiabatic loadings represent reductions
in the condensate that is delivered to the mixed-phase region by the updraft, and
where the conversion of supercooled water to ice can invigorate the updraft by the
latent heat of freezing. This process can strongly influence the nature of the vertical
profiles of latent heat release and larger ice-phase hydrometeors and help explain
the marked land–ocean contrasts in differences in lightning (Figure 15.1(a)) and
rainfall (Figure 15.1(b)). In short, the warm rain cells over ocean may be sub-
stantially prevented from becoming thunderstorms by virtue of the raindrop loading
Lightning and climate change 581
they achieve. Further evidence for this suggestion is found in the next
Section 15.3.6.
can fall out of the updraft. In a special class of supercells (called LP for “low
precipitation” but storms invariably productive of hail) with high CBHs (occa-
sionally with sub-0 C cloud base temperatures), no condensate is lost in a warm
rain process and the adiabatic cloud water is available for the growth of the hail.
15.3.7 Baroclinicity
The pole-to-equator temperature difference is an important consideration for both
weather and climate [71]. The local latitudinal temperature gradient exerts a deci-
ded influence on the organization of thunderstorm activity, and may also affect the
flash rate behavior of thunderstorms. This temperature contrast is set up by the
latitudinal imbalance between the incoming shortwave radiation from the Sun and
the outgoing longwave radiation to space. Instabilities in the prevailing westerly
winds at mid-latitudes draw on the potential energy of the pole-to-equator tem-
perature difference to produce synoptic scale weather disturbances there [72]. The
condition of a latitudinal gradient in temperature is “baroclinicity” and the thermal
wind equation in atmospheric dynamics [71,73] links a baroclinic atmosphere with
a vertical shear in the horizontal wind. This vertical wind shear can tilt the updrafts
of thunderstorms (embedded in these large-scale disturbances) from the vertical,
and thereby strongly influence the unloading of condensates from the updraft.
In the near-equatorial zone of the tropics, the latitudinal temperature gradient
nearly vanishes and the atmosphere is characterized as barotropic rather than bar-
oclinic [73]. Without the organizing effects of vertical wind shear, air mass thun-
derstorms prevail, with an expectation for vertical rather than tilted updrafts. Yoshida
et al. [74] have compared the flash rates of thunderstorm cells in continental and
oceanic zones at different latitudes. In general, the mean flash rates are increasing
with latitude away from the tropics. Bang and Zipser [75,76] have also found that
oceanic convection is more likely to produce lightning in the presence of vertical
wind shear than without it. Some of the most dramatic outbreaks of lightning over the
open ocean occur in the presence of strong baroclinicity [77,78]. These findings may
have explanation for baroclinicity in unloading of the updraft in the warm rain region
of the storms so as to provide greater invigoration of the mixed phase. On this basis,
and with all other factors the same, one might expect to have more lightning activity
globally with a larger pole-to-equator temperature contrast.
Supercells are rotating thunderstorms characteristic of strongly baroclinic
environments in the springtime, and to a lesser extent in the fall, in both northern and
southern hemispheres [79–81]. These storms are parent to a large portion of severe
weather episodes, including strong winds, large hail, and tornadoes [80,81]. Supercell
thunderstorms also produce exceptional total lightning flash rates [58,82–86].
11-year time scale of the solar cycle. In understanding global lightning’s response
to climate change on the long-time scale, it is valuable to look for consistent pat-
terns of response on other time scales whose physical origin is better understood. It
is also possible that systematic changes in aerosol may accompany these natural
variations in temperature. Here again, we encounter the problem of disentangling
aerosol and thermodynamic effects [87].
Africa
and Europe North and
110 All Oceans 80 South America
(Carnegie)
60 Asia
and Australia
100
40
Carnegie New
20 Zealand
90
Maud
0
0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
Time (GMT) Time (GMT)
Figure 15.6 The Carnegie curve of atmospheric electricity over 24 h of universal time and a global map showing three continental
lightning chimneys. Global lightning observations are taken from references [95] and [102].
Lightning and climate change 585
Clear evidence for a semiannual variation in global lightning with the same
phase as the temperature variation has been provided by optical observations of
lightning from space [102]. This signal is most conspicuous when analysis is con-
centrated in the near-equatorial zone. Here the sensitivity of lightning to tempera-
ture is 20%–30% per 1 C. Rainfall observations are also consistent with double
rainy seasons in northern South America and in near-equatorial Africa. The dis-
charge record from the extensive Congo River basin, straddling the equator in
central Africa, also shows a distinct semiannual variation.
Several independent observations of the global electrical circuit (both DC and
AC) show evidence for semiannual signals [92,101,103–105] that are plausibly
linked with the semiannual variation of temperature.
The aerosol contribution to the semiannual variation in global lightning
activity has not been studied. We are presently unaware of long-term investigations
of CCN concentration in the near-equatorial zone over continents that would shed
further light here.
methods now under development for the observation of CCN at CBH from satellite
[14] will be particularly valuable in this context.
Ambiguity about the annual phase of the DC global electrical circuit began
when Lord Kelvin observed [108] in the United Kingdom that the Earth’s electric
field was greatest in winter, contrary to the general picture based on global light-
ning activity and temperature. Whipple [7] identified this apparent contradiction
but did not resolve it. Adlerman and Williams [109] pointed out the local effect of
wintertime aerosol (in both hemispheres) in decreasing the electrical conductivity
of the atmosphere and thereby increasing the electric field. A more reliable ground-
based measurement of the DC global circuit is the air-earth current which repre-
sents the product of the electric field and conductivity. For example, long-term
measurements of the air-earth current in Athens [110] have been shown to peak in
August. Seasonal averages of the ionospheric potential [98], the preferred measure
of the DC global circuit, also show a maximum in northern hemisphere summer.
15.4.4 ENSO
Early interest in a strong relationship between the phase of the El Nino-Southern
Oscillation and lightning developed out of analysis of a single magnetic coil recording
of the first resonance mode of Schumann resonances [30], for a single ENSO event (in
1973–1974). A remarkably sensitive response of lightning to temperature was inferred
from this analysis. Though the sign of this response has been corroborated in more
recent work (e.g., [111,143]), the magnitude of the response in the data obtained by C.
Polk is unprecedented, and may not be valid. Based on the global temperature analysis
of Hansen and Lebedeff [27] and the classical analysis of rainfall variations [112]
showing less rainfall over land in the warm phase, the general picture that developed
was one in which all tropical land regions warmed in the El Nino warm phase, with the
major upwelling in the central/eastern Pacific region causing large scale subsidence
over tropical continental regions. This picture is in keeping with reductions in total river
discharge in large drainage basins (Amazon and Congo), serving as continental-scale
rain gauges, straddling the equatorial region [113] in the warm phase. This investigation
was later extended to the Nile and the Ganges [114] River basins, with similar ENSO
phasing. This picture of regional rainfall variability is also consistent with the majority
of multiple ENSO events pictorialized in Allan et al. [115].
The regional behavior of lightning over the ENSO cycle shows somewhat less
consistency overall, but a definite tendency in behavior is evident. This is most
apparent in the Maritime Continent (including southeast Asia, Indonesia, and tro-
pical Australia), where the ENSO studies are most numerous and where lightning is
more prevalent in the warm El Nino phase [116–122]. Generally speaking, in this
part of the world, the warm ENSO phase is also the drier (lower relative humidity
and higher CBH) phase and the more polluted phase. As noted above, the wet cool
phase (La Nina) is more abundant in rainfall. So here again we have the entan-
glement issue of thermodynamics and aerosol. The opposite tendencies for rainfall
and lightning on the ENSO time scale are at first peculiar, but highly variable
lightning/rainfall relationships, temporarily and regionally, are now widely recog-
nized [123,124] and are linked with differences in the vertical development of the
Lightning and climate change 587
precipitation in the cold part of the cloud. This situation of opposite ENSO phase
relationships for rainfall and lightning has also led to speculation that the two
global electrical circuits (DC and AC) may have opposite tendencies over ENSO
cycles [125,126], though coordinated synchronous measurements are lacking to
check on this prediction. In this context, it seems likely that the contribution of
electrified shower clouds [127–130] will be more potent over land during the cold
La Nina phase, when continental rainfall is greater [115].
In South America, both Chronis et al. [131] and [111] using lightning observa-
tions from space found greater amounts of lightning in northeast Brazil during the
cold La Nina phase. In contrast, Pinto [132] using thunder day observations over
many ENSO cycles found a conspicuous tendency for greater numbers in the warm
El Nino phase. An extreme El Nino event in 1926 has been documented by Richey
et al. [133] and by Williams et al. [134] in the Amazon basin, with exceptional hot
and dry characteristics but no information on lightning activity is available. Chronis
et al. [131] and Sátori et al. [111] agree in finding more lightning in the warm phase
in southern Brazil and eastern Argentina, where the discharge of the Parana River is
also maximum [113]. This region appears to be the southern component of the north–
south rainfall dipole anomaly identified by Grimm and Natori [135]. This distinctly
extra-tropical tendency for greater lightning in the warm phase was found earlier in
Argentina [136] and in the opposite hemisphere by Goodman et al. [137] in the Gulf
of Mexico region of the United States, in a similar range of latitude.
Among three tropical lightning “chimneys” [9], Africa appears to show the
weakest lightning variation on the ENSO time scale. Chronis et al. [131] found some
enhancement in the La Nina phase, whereas Sátori et al. [111] reported a modest
lightning increase in the warm phase. Given Africa’s status as the most distant lightning
chimney from the Pacific Ocean source of convective upwelling, it seems likely that
global scale subsidence would have the least effect on Africa, while leaving con-
spicuous effects on adjacent chimneys Maritime Continent and America. Dowdy [138]
has considered the effect of season on the lightning response to temperature on the
ENSO time scale and this may impact the generality of a simple positive response in
the warm phase. This seasonal aspect has also been discussed by LaVigne et al. [125].
Extreme El Nino events, such as the drought of the century in South America,
can lead to so much warming and drying as to prevent both moist convection and
lightning. Evidence for this situation may be found in the 1926 drought on the
Amazon basin [133,134,139].
Despite the great sparsity of oceanic lightning (recall Figure 15.1(a)), easily dis-
cernible regional variations are discernible on the ENSO time scale. The general ten-
dency for oceans is opposite to that for land: greater lightning over ocean in the cold La
Nina phase [111,140,141]. This has been interpreted on a basis consistent with that for
the warm phase: greater regional subsidence signifies less overall cloudiness, and
hence greater surface heating and greater instability to drive moist convection. This
situation contributes to the heat uptake of the tropical ocean during the La Nina phase.
On a worldwide basis, Satori et al. [111] documented greater lightning in the
ENSO warm phase than in the cold phase. Harrison et al. [142,143] have found
evidence for inferred increases in the global electrical circuit in the warm phase. As
588 Lightning electromagnetics: Volume 2
this book chapter was going to press, evidence for greater global lightning activity
during the 2019 El Nino was evident, followed by reductions in global lightning in
the La Nina years of 2020-2023 (See also [52]).
Recent improvements in understanding lightning variations on the ENSO time
scale Williams et al. [144] have been achieved by considering superlative events in
the Super El Nino category, of the same kind considered earlier by Williams [30]. In
contrast with earlier studies that compare lightning totals in the warm (El Nino) and
cold (La Nina) phase, the full evolution of lightning activity in individual ENSO
events has been considered, including the important transition phase from cold to
warm periods. Although more lightning is present in the warm phase than the cold
phase, as, in earlier work, the peak lightning activity is found in the transition phase,
when the vertical temperature profile is most out of equilibrium with the surface. In
another ENSO study Guha et al. [145], peak lightning activity was also found in the
transition phase, but in that case in the transition from warm to cold periods.
data (and model calculations: [155]). Thunder day observations have been under-
way at meteorological stations and airports worldwide for more than a century. One
interesting recent application here has focused on an upward lightning trend in the
Sea of Japan [156]. A rough doubling in thunder day counts since 1930 has been
linked with observed increases in sea surface temperature of 1.2 C–2.2 C.
The tendency for the current global warming to predominate at high northern
latitudes is widely recognized [28]. At Fairbanks, Alaska (latitude 64.8 N) both the
temperature and the thunder day counts have been increasing conspicuously [44].
Figure 15.7 includes plots of both quantities with least squares fits for trend. The
number of thunder days has more than doubled in 50 years. Anecdotal reports
indicate that Canadian meteorological stations at the highest latitudes have noted
thunder days for the first time on record.
The long-term record of mean global temperature [27,28,157] based on aver-
aging of surface thermometers shows an increase of the order of 1 C on the 100-
year time scale but is interrupted by shorter intervals when the temperature is flat or
declining with time. The two most notable intervals are the so-called “Big Hiatus”
from 1940 to 1975, and the more recent (and more controversial) hiatus in global
warming from 1998 to 2013. Both these intervals have been addressed recently by
Williams et al. [158]. Appeal was made to previously published thunder day
observations to address the Big Hiatus and separate analyses for both North
America and Siberia show flat or declining counts of thunder days. A 15%
decline in mean annual thunder days is evident from 1940 to 1970.
For the more recent hiatus in global warming, satellite optical observations are
available for nearly the entire interval from the lightning imaging sensor in space.
Several global temperature data sets were examined and it was shown that both the
60.0 20
°F (June–September)
57.5 16
Number of Thunder days per Year
55.0 12
Average Temperature
52.5 8
50.0 4
47.5 0
1940 1950 1960 1970 1980 1990 2000 2010 1950 1960 1970 1980 1990 2000 2010
Time (Years) Time (Years)
(a) (b)
Figure 15.7 Upward trends in (a) temperature and (b) thunder days for
Fairbanks, Alaska. Linear least squares fits are also included to
illustrate these trends
590 Lightning electromagnetics: Volume 2
temperature trend and the trend in global lightning flash rate were statistically flat
during the hiatus period [158]. This period ended with a pronounced El Nino event
in 2015 and a strong resumption in the increase in global temperature [157] that had
been underway between the time of the Big Hiatus and the recent hiatus.
Unfortunately, the Lightning Imaging Sensor is no longer operational in space to
check the lightning behavior in this later period. Overall, the available results are
not inconsistent with the hypothesis that global lightning is responsive to global
temperature.
Maritime Regime
0° C
Continental Regime
0° C
Figure 15.8 Illustration of the aerosol effect on convection and lightning in clean
and polluted situations
(2) more polluted conditions (a few hundred to 1,000 per cc) typical for continents,
and (3) ultra-polluted conditions with concentration exceeding several 1,000 per cc.
These regimes have been treated in cloud models [162]. Condition (1) may favor
the rapid formation of precipitation and the production of rain, which may ultimately
contribute to the superadiabatic loading of the air parcel (Section 15.3.5). Condition
(2) may enable the retention of condensate in the updraft (consistent with the
assumptions of reversible ascent ([13] and Section 15.3.5)) until the mixed-phase
region is attained. The most polluted condition (3) may lead to cloud droplets so
small that the formation of graupel particles in the mixed-phase region is prevented,
thereby enabling the cloud water to rise as high as the 40 C isotherm, where
homogenous nucleation of the cloud water may occur.
mediate with aerosol. Without the depletion by warm rain, the cloud water contents
in the mixed-phase region are expected to be large in such continental storms,
promoting the growth of hail and also conditions conducive to thunderstorms with
inverted electrical polarity [58,84,170,171]. Lyons et al. [172] had earlier found
that storms ingesting smoke from fire were exhibiting greater numbers of positive
ground flashes.
Additional evidence that aerosol has a pervasive influence not only on light-
ning but other aspects of meteorology are studies that show reduced activity on
weekends when the anthropogenic contribution to aerosol in industrialized regions
has been shown to be reduced. Weekend effects on rainfall [173], lightning [174],
hail [175], and even tornadoes [175] have been documented in recent years, with
statistically verified results.
Despite the general understanding that lightning activity is greatly diminished at high
latitude relative to the tropics, based on considerations of surface air temperature and
water vapor (Section 15.3.3), lightning can and does occur in the colder polar regions.
For example, consultation of the NOAA GSOD (Global Summary of Day) dataset on
thunder day observations [181] for recent decades shows that 390 surface stations
north of the Arctic circle show at least one thunder day in their records. Following the
Franklin Lecture on “Lightning and Climate” at the American Geophysical Union
meeting in 2012, one Canadian meteorologist reported that thunder days were being
recorded for the first time at the highest latitude stations in recent years.
Some improvement in the characterization of lightning trends in high latitude
regions has been achieved quite recently with regional lightning networks, for both
the Arctic as a whole using a subset of the World Wide Lightning Location
Network [182] and for the State of Alaska [183] using the Alaska Lightning
Detection Network (ALDN). In both cases, the assessment of decadal trends is
compromised by the changes in the networks over time. The quantitative results
from these studies are here discussed in turn below.
Interest in lightning at high latitudes also derives from the role of lightning in
initiating fire in relatively dry forests there [183–186]), and the possibly positive
feedback linked with aerosol and lightning [186].
stroke counts at face value to obtain a lightning sensitivity to temperature over the
past 10 years, over which Ballinger et al. (2021) [187] report a 1.5 C mean increase
in surface air temperature, that amounts to a 300% increase in lightning for 1.5 C,
or 200% per C. Clearly, this value is out of line with other estimates (Section 15.4),
and also points to the problem with time-dependent detection efficiency.
15.7.2 Alaska
Evidence for upward trends in thunder days, surface air temperature, and lightning
activity in recent decades in Alaska can be found in Section 15.4.6 (Figure 15.7),
Bieniek et al. [189] and in Bieniek et al. [183], respectively. The magnitudes of the
temperature trends show the largest values above the Arctic circle, consistent with the
evidence for the pronounced warming of the Arctic noted above. As noted by Bieniek
et al. [189], “the low frequency variability confounds the identification of trends in the
time series of temperature.” Given the network changes over time, the lightning data
are also caveated [189] by the following: “These changes in detection efficiency and
accuracy through the record make the data challenging to use for assessing the varia-
bility and trends of lightning over the historical period.” For best comparison with the
30-year lightning record in Bieniek et al. [183], our selection of representative tem-
perature increase (0.8 C) is based on the 30-year period of summertime (June, July, and
August) temperature in Bieniek et al. [189] coinciding with the months of greatest
lightning activity. Given their estimate of a 17% increase in lightning over the same
period [183, p. 1148] we derive a rough sensitivity of 17%/0.8 C = 21% per C. This
sensitivity is broadly consistent with the estimates obtained in Section 15.4.
Despite the evidence for unparalleled increases in lightning activity at high
latitude in recent years, the total contribution to global lightning from this remote
part of the world is inconsequential. To show this, one can compare the annual
lightning totals recorded by Holzworth et al. [182] and by Bieniek et al. [183] with
the totals in the tropical continental “chimneys” discussed in Section 15.1. The total
flash rates in the tropical chimneys are of the order of 10 flashes per second. In
contrast, the largest annual-averaged WWLLN stroke rate amounts to 8 103
strokes/sec. For Alaska, the peak annual stroke rate is 1 105 strokes/yr, or
3 103 strokes/s. This comparative situation gives the global VLF networks some
superiority over ELF methods in monitoring the changes in high latitude lightning
into the future, so long as the networks are stable with time.
Lightning occurrence in winter thunderstorms has been a concern due to their high
damage potential (e.g., [190]). In the winter cold period (e.g., winter in middle
latitudes), diurnal heating is weak and moisture content is much reduced over land
compared to the warm season. Low-pressure systems are more vigorous and bring
cold air masses from polar and arctic regions southward to mid-latitudes, creating
strong vertical temperature gradients as cold continental air flows over relatively
warm seas and ocean currents. Air parcels near the surface then experience no
596 Lightning electromagnetics: Volume 2
inhibiting warm layers on their way to the equilibrium level, often the tropopause,
and occur over large regions over the sea behind cold fronts. The tropopause is
found between 10 and 15 km at mid-latitudes in summer but can descend to
5–10 km in winter, limiting the vertical extent of convection. Just as in summer,
low-level convergent winds organize the triggering of storms, but over the sea,
these are often found near and upwind of coastlines, where enhanced friction and
sloping terrain create a relative stagnation and ascending flow.
The importance of the moisture source from seas and oceans in winter thunder-
storms deserves a look in the context of the ocean gyres. Every major ocean basin
(North Atlantic, North Pacific, South Atlantic, South Pacific, etc.) contains a basin-
scale closed rotating current: a gyre with clockwise (counterclockwise) rotation in the
northern (southern) hemisphere. The primary drive for these gyres is the zonal wind
stress from prevailing easterly winds in the tropics otherwise known as the trade winds,
and from prevailing westerly winds at mid-latitude. In this near-equatorial portion of
the gyre, the ocean surface is warmed substantially by sunlight and at the western limit
of this equatorial transit, this warm oceanic flow is diverted northward and southward,
depending on hemisphere. Since the surface air over continents is increasingly colder
away from the equator and tends to be moving eastward off the continents at mid-
latitude, one has a consistent situation in all gyres that warm ocean water in this
poleward current is found beneath colder air away from the equator. This configuration
is inherently unstable and can produce vigorous atmospheric convection and thun-
derstorm activity. The Gulf Stream along the North American coast and the Kuroshio
Current along the eastern coast of Asia (China, Japan, Korea, and Russia) are prime
examples in which lightning activity over warm ocean water is prevalent during win-
ter. In contrast, the return current on the eastern boundaries of oceanic gyres, moving
equatorward, is colder than the air overlying it. This situation is stable against con-
vection and accordingly lightning is absent. A prime example is the Eastern Pacific
Ocean. Similar behavior is present in the oceanic gyres of the southern hemisphere.
INTENSE
EYEWALL EYE EYEWALL THUNDERSTORM
Preliminary Data heavy
Rain Intensity
light
Figure 15.9 A vertical slice through the center of Hurricane Emily (July 17, 2005,
near Cuba) shows the rain structure across 210 miles (340 km) of the
storm. Note the lack of precipitation inside the eye of the storm
compared to the intense eyewall convection just outside of it. The
areas of heaviest rainfall are shown in red, and the lightest are in
blue. The prominent horizontal feature in red is the radar bright band
near 0C and often characterizes the outer rainbands. The rainfall
structure shown here was measured by NASA’s ER-2 Doppler Radar
much like the updraft of an exceptional thunderstorm. The vigor of the eyewall
convection and attendant lightning is often maximum during cyclone intensifica-
tion and falling central pressure (i.e., “deepening”) [203–206].
Further out in radius from the eye, the ascent speed of the air is much reduced,
as evidenced by the red horizontal feature near 5 km altitude: the radar “bright
band”, close to the 0 C isotherm. Ice phase condensate is in evidence above the
bright band to the top of the storm, an important source for charge separation and
lightning in this region. The total vertical extent of the eyewall convection, using
the bright band height as a “yardstick” is 15 km. Here the local lightning flash
densities are much reduced in comparison to the eyewall lightning, but because of
the substantially greater convective area in these outer hurricane rainbands, the
total lightning production generally exceeds that in the eyewall [205,207].
Theoretical considerations of tropical cyclones that treat these storms as giant
heat engines (transporting heat from the warm ocean surface to the cooler upper
troposphere) predict increases in cyclone intensity with global warming [208].
Numerical models also show globally-averaged increases in maximum wind speeds
by 2–11% by the year 2100 [209–211]. Using numerical global climate modeling,
Lynn et al. [212] were among the first to project increases in hurricane intensity by
the end of the century based on simulations of Category 5 Hurricane Katrina under
climate change scenarios. However, difficulties with establishing an upward trend
Lightning and climate change 599
The measured peak current in negative polarity cloud-to-ground lightning has been
shown to be greater over ocean than over continents [172,218]. The physical
explanation for this contrast, and why a similar contrast is less apparent for light-
ning with positive polarity, is presently in debate. Predictions for increases in peak
lightning current (also known as lightning “intensity”) to the ocean surface in
response to global warming have appeared [219]. These predictions are based on
expected changes in the chemical and electrical properties of seawater in a warmer
climate [219,220]. For example, increased ocean acidification may result in
increases in hydrogen ions in seawater, thereby modifying the seawater con-
ductivity. As these predictions for increases in peak current are based on
600 Lightning electromagnetics: Volume 2
31.5˚N 31.5˚N
30˚N 30˚N
28.5˚N 28.5˚N
27˚N 27˚N
25.5˚N 25.5˚N
Figure 15.10 Satellite image of record extent of a lightning flash observed with the
Geostationary Lightning Mapper over the southern United States on
April 29, 2020 covered a horizontal distance of 768 8 km
(477.2 5 mi). The horizontal structure (white line segments) and
maximum extent (gold symbols) of this megaflash are overlaid. As
of this writing, this is the record-setting megaflash for scale
602 Lightning electromagnetics: Volume 2
four separate States (Texas, Louisiana, Alabama, and Florida) and the Gulf of
Mexico. Since this distance is more than twice the threshold diameter [198] of a
Mesoscale Convective Complex, it is likely that many of the largest megaflashes
will require these superlative storms for their existence. The association of the
largest lightning flashes with the largest storms suggests that the monitoring of
megaflashes in a warmer climate may be a useful diagnostic for climate change.
Brooks [240], and Wallace [251] has systematized the local diurnal cycle in the
Great Plains of North America. Prodigious nocturnal lightning activity in both
areas may ultimately lead to some important reconsideration of the behavior of the
global electrical circuit in universal time [7,8]. In both areas, air in the prevailing
westerly flow is warmed in the afternoon in traversing the respective heated
mountain ranges, and then progresses further east over low-level flow laden with
moisture. The warm air aloft creates a temperature inversion in the local tem-
perature profiles, otherwise known as a capping inversion, which serves to suppress
the afternoon convection while allowing the buildup of temperature and moisture
below the cap throughout the daytime that contributes to a large CAPE [29]. When
the capping inversion weakens and special triggering processes occur at night, the
moist convection develops with exceptional vigor [252,253]. The prominence of
thunderstorms with large hail in the nighttime activity [252] is evidence for
exceptional updrafts in nocturnal convection.
In all of the scenarios described above, the nocturnal thunderstorms described
occur only when the usual afternoon thunderstorms are absent at the same location.
This situation has special implications for the thermodynamic boundary conditions
for the nighttime storms. One implication is that the nighttime air temperatures are
substantially larger on the evenings with earlier afternoon storms [249,253]. There
are two important reasons for this observation. One is that the customary cold
downdraft air was not present. Second, the blanket of boundary layer water vapor
that accumulates continuously in the daytime, and which is not reduced by its
transformation to rainfall in afternoon thunderstorms, serves to reduce the outgoing
longwave radiation that would otherwise serve to cool the Earth’s surface.
It is important to note that these conditions contrast starkly with the usual
nighttime conditions (without thunderstorms) in the tropics when the relative
humidity reaches 100%, the dew point depression vanishes, and dew appears at the
surface (e.g., [53,254]). In such circumstances, the contrast in CBH between land
and ocean would not serve as a viable thermodynamic contributions to the land–
ocean contrast in lightning activity, as discussed in Lucas et al. [38], Williams and
Stanfill [11], and Zipser [41,70]. But an appeal to the literature shows that noc-
turnal thunderstorms do not occur in these circumstances. As one example study in
the Great Plains [253], the development of nocturnal convection is characterized by
the presence of unsaturated air at the surface and an elevated CBH (3.4–3.5 km
MSL) linked with most unstable air that is elevated from the surface. CAPE values
are linked with the most unstable values of wet bulb potential temperature and are
also large (>3,000 J/kg) in the case of nighttime storms. The evidence that noc-
turnal thunderstorms are a subclass of “elevated thunderstorms” that feed on a
source of most unstable air removed from the surface was established earlier by
Colman [255].
Regarding the contribution of aerosol to lightning activity in the context of
nocturnal thunderstorms, generally speaking, one would not expect major changes
in CCN from daytime to nighttime. Accordingly, the reduced land/ocean lightning
at nighttime [242] is likely of thermodynamic origin. Further studies on this issue
are warranted.
604 Lightning electromagnetics: Volume 2
Vi
(a)
Schumann Resonances
Electric
Height
Magnetic
(b) Height
Figure 15.11 Illustration of the two global electrical circuits: (a) the classical DC
global circuit, and (b) the AC global circuit, otherwise known as
Schumann resonances
606 Lightning electromagnetics: Volume 2
though this change is regionally dependent. It must also be recognized that the nature
of the precipitation increase is important and that an increase in warm rain alone is
unlikely to be accompanied by an increase in lightning.
Given the evidence for CAPE as a driver for lightning in the present climate,
more lightning is expected in a warmer climate if CAPE increases. Early speculation
showed CAPE to be a climate invariant [43]. Romps et al. [291] have predicted
increases over the CONUS in a warmer climate on the basis of both increases in
CAPE and increases in precipitation. More recent global climate models show larger
CAPE in warmer climates [44]. Theoretical calculations in equilibrium atmospheres
[292] show CAPE scaling with the Clausius–Clapeyron relationship. Estimations of
CAPE changes in the western United States based on the advection of dry desert area
over the moist boundary layer with origin in the Gulf of Mexico also lead to pre-
dictions that CAPE should scale with Clausius–Clapeyron [49]. For all of these rea-
sons, one expects greater lightning in a warmer world.
A frequent assumption in the climate community is that surface relative
humidity is a climate invariant [259,264]. This assumption is based in part on the
empirical evidence in the present climate for a quasi-fixed relative humidity near
80% over large areas of the tropical oceans, with an associated CBH for moist
convection near 500 m. If the ocean surface temperature were to increase globally,
the mean dew point temperature would also increase so as to keep the dew point
depression (T – Td) and the associated CBH (above local terrain) both constant. But
the height of the 0 C isotherm would increase following a presumed lifting along a
dry adiabat. Accordingly, the warm cloud depth (distance between the 0 C iso-
therm and the CBH) would increase. Based on findings in the current climate that
the lightning flash rate is decreasing with increased warm cloud depth, in this
scenario, one might expect less lightning in the warmer world. However, the CAPE
change in this scenario also deserves consideration. The expected increases in both
T and Td both contribute to increases in the wet bulb potential temperature of the
boundary layer, a result that certainly favors greater CAPE in the current climate.
But the ultimate CAPE change depends also on the change in the overall tem-
perature profile in a warmer climate, and this aspect is indeterminate in the context
of the foregoing assumptions.
One recent study [51] predicts less tropical lightning in a warmer world. The
prediction is model-based and with the finding that less ice phase is reaching the upper
portion of the troposphere where active charge separation is known to occur. More
information about this model, and in the context of expected changes in warm cloud
depth, is needed to understand this result. Other recent observations on extremes in
tropical rainfall in a warmer climate showing enhancements exceeding the simple
predictions based on Clausius–Clapeyron [293] would seem to contradict the model
predictions of Finney et al. [51]. Further discussion on this issue and predictions for the
behavior of lightning in a warmer climate can be found in Yair [294].
Given the recent evidence for an important role of CCN in increasing lightning
activity in the present climate [23], one can surely speculate about changes in
lightning in a more polluted world on the basis of this effect. China for example
continues to undergo major industrialization and with a major reliance on coal
608 Lightning electromagnetics: Volume 2
[295] and so one can expect an increasing aerosol production from that region
alone. From what is known at present, the lightning enhancements are expected
over a finite range of CCN concentrations, from typical oceanic concentrations of a
few per cc to up to about 1,000 per cc [21,23]. Above that level (typical of con-
tinental conditions described in Section 15.5), the lightning activity is expected to
flatten and then decrease, for reasons that are evident in model calculations [162].
An assessment of where the world stands relative to this important threshold will be
much aided by a new satellite method for the estimation of CCN concentration at
CBH [14].
References
[1] Cooray, V. (ed.), Lightning Protection (Series: IET Power and Energy,
Institution of Engineering and Technology – IET, London, 2010.
[2] Herring, S.C., N. Christidis, A. Hoell, J.P. Kossin, C.J. Schreck III, and P.A.
Scott (eds.), Explaining extreme events of 2016 from a climate perspective,
Special Supplement to the Bull. Am. Met. Soc., 99(1), S1–S157 2018,
doi:10.1175/BAMS-ExplainingExtremeEvents2016.1.
[3] Aich, V., R. Holzworth, S.J. Goodman, Y. Kuleshov, C. Price, and E.
Williams, Lightning: a new essential climate variable, EOS, 99, 2018,
https://doi.org/10.1029/2018EO104583.
[4] Virts, K.S., J.M. Wallace, M.L. Hutchins, and R.H. Holzworth, Highlights of a
new ground-based, hourly global lightning climatology. Bull. Amer. Meteor.
Soc., 94, 1381–1391, 2013. doi: http://dx.doi.org/10.1175/BAMS-D-12-00082.1
[5] Kinne, S., Remote sensing data combinations-superior global maps for
aerosol optical depth, in A. Kokhanovsky and G. de Leeuw (eds.), Satellite
Aerosol Remote Sensing Over Land, Springer, New York, NY, pp. 361–380,
2009 (Chapter 12).
[6] Boccippio, D.J., S.J. Goodman, and S. Heckman, Regional differences in
tropical lightning distributions, J. Appl. Met., 39, 2231–2248, 1999.
[7] Whipple, F.J.W., On the association of the diurnal variation of electric
potential gradient in fine weather with the distribution of thunderstorms over
the globe, Quart. J. Roy. Met. Soc., 55, 1–17, 1929.
[8] Whipple, F.J.W. and F.J. Scrase, Point discharge in the electric field of the
Earth, Geophys. Mem., Lond., 68, 1–20, 1936.
[9] Williams, E.R., Lightning and climate: a review, Atmos. Res., 76, 272–287,
2005.
[10] Humphreys, W.J., The Physics of the Air, Dover Publications, New York,
NY, 1964.
[11] Williams, E. and S. Stanfill, The physical origin of the land-ocean contrast in
lightning activity, C. R.—Phys., 3, 1277–1292, 2002.
[12] Rosenfeld, D. and W.L. Woodley, Closing the 5-year circle: From cloud
seeding to space and back to climate change through precipitation physics, in
W.-K. Tao and R. Adler (eds.) , Cloud Systems, Hurricanes, and the Tropical
Lightning and climate change 609
[28] Hansen, J., R. Ruedy, M. Sato, and K. Lo, Global surface temperature
change, Rev. Geophys., 48, RG4004, 2010, doi:10.1029/2010RG000345.
[29] Emanuel, K. A., Atmospheric Convection, Oxford University Press, Oxford,
580 pp., 1994.
[30] Williams, E.R., The Schumann resonance: a global tropical thermometer,
Science, 256, 1184–1187, 1992.
[31] Price, C., Evidence for a link between global lightning activity and upper
tropospheric water vapor, Nature, 406, 290–293, 2000.
[32] Ludlam, F.H., Clouds and Storms: The Behavior and Effect of Water in the
Atmosphere, Pennsylvania State University Press, University Park, PA,
1980, 405 pp.
[33] Iribarne, J.V. and W.L. Godson, Atmospheric Thermodynamics, D. Reidel,
Dordrecht, 1981, 259 pp.
[34] Betts, A.K., Saturation point analysis of moist convective overturning, J.
Atmos. Sci., 39, 1484–1505, 1982.
[35] Xu, K.-M. and K.A. Emanuel, Is the tropical atmosphere conditionally
unstable?, Monnthly Weather Rev., 117, 1471–1479, 1989.
[36] Williams, E.R. and N.O. Renno, An analysis of the conditional instability of
the tropical atmosphere, Monthly Weather Rev., 121, 21–36, 1993.
[37] Brunt, D., Physical and Dynamical Meteorology, Cambridge, Cambridge
University Press, 1952.
[38] Lucas, C., E.J. Zipser, and M.A. Lemone, Convective available potential
energy in the environment of oceanic continental clouds: correction and
comments, notes and correspondence, J. Atmos. Sci., 51, 3829–3830, 1994.
[39] Pan, Z., F. Mao, D. Rosenfeld, et al., Coarse sea spray inhibits lightning.
Nat. Commun. 13, 4289. https://doi.org/10.1038/s41467-022-31714-5, 2022.
[40] Riemann-Campe, K., K. Fraedrich, and F. Lunkeit, Global climatology of
convective available potential energy (CAPE) in ERA-40 reanalysis, Atmos.
Res., 93, 534–545, 2009.
[41] Zipser, E.J., Some views on “hot towers” after 50 year of tropical field
programs and two years of TRMM data, in Meteorological Monographs,
American Meteorological Society, Washington, DC, pp. 49–58, 2003
(Chapter 5).
[42] Romps, D. M., A.B. Charn, R.H. Holzworth, W.E. Lawrence, J. Molinari, &
D. Vollaro. CAPE times P explains lightning over land but not the land-
ocean contrast. Geophysical Research Letters, 45, 12,623–12,630. https://
doi.org/10.1029/2018GL080267, 2018.
[43] Emanuel, K.A., J.D. Neelen, and C.S. Bretherton, On large-scale circulations
in convecting atmospheres, Quart. J. Roy. Met. Soc., 120, 1111–1143, 1994.
[44] Williams, E.R., Lightning and Climate, Franklin Lecture (invited) Fall
Meeting, American Geophysical Union, San Francisco, December, 2012,
http://fallmeeting.agu.org/2012/events/franklin-lecture-ae31a-lightning-and-
climate-video-on-demand/.
[45] Ye, B., A.D. Del Genio and K.-W. Lo, CAPE variations in the current cli-
mate and in a climate change, J. Climate, 11, 1997–2015, 1998.
Lightning and climate change 611
[46] Singha, M.S., Z. Kuang, E.D. Maloney, W.M. Hannah , and B.O. Wolding,
Increasing potential for intense tropical and subtropical thunderstorms under
global warming, PNAS, 114, 11657–11662, 2017.
[47] Chen, J., A. Dai, Y. Zhang, K.L. Rasmussen, Changes in Convective
Available Potential Energy and Convective Inhibition under Global
Warming, J. Climate, 33, 2025–2050, 2020.
[48] Del Genio, A.D., Y. Mao-Sung and J. Jonas, Will moist convection be
stronger in a warmer climate? Geophys. Res. Lett., 34, L16703, 2007,
doi:10.1029/2007GL030525.
[49] Agard, V. and K. Emanuel, Clausius-Clapeyron scaling of peak CAPE in
continental convective storm environments, J. Atmos. Sci., 74, 3043, 2017,
doi:10.1175/JAS-D-16-0352.1.
[50] Seeley, J.T. and D.M. Romps, Why does tropical convective available
potential energy (CAPE) increase with warming?, Geophys. Res. Lett., 42,
10,429–10,437, 2015, doi:10.1002/2015GL066199.
[51] Finney, D.L., R.M. Doherty. O. Wilde, D.S. Stevenson, I.A. MacKenzie, and
A.M. Blyth, A projected decrease in lightning under climate change, Nature
Clim. Change, 8, 210–213, 2018, https://doi.org/10.1038/S41558-018-0072-6.
[52] Williams, E.R., Lightning’s Response to Thermodynamics and Aerosols on
Different Time Scales, Invited presentation for Annual Meeting of the
American Meteorological Society, Denver, CO, January 11, 2023.
[53] Byers, H.R. and R.R. Braham, The Thunderstorm Project, U.S. Weather
Bureau, U.S. Department of Commerce, Washington, DC, 1949.
[54] Lemone, M.A. and E.J. Zipser, Cumulonimbus vertical velocity events in GATE.
Part I: diameter, intensity and mass flux, J. Atmos. Sci., 37, 2444–2457, 1980.
[55] Boccippio, D.J., Lightning scaling relations revisited, J. Atmos. Sci., 59,
1086–1104, 2001.
[56] Zheng, Y. and D. Rosenfeld, Linear relation between convective cloud base
height and updrafts and application to satellite retrievals, Geophys. Res.
Lett., 42,6485–6491, 2015, doi:10.1002/2015GL064809.
[57] Hansen, Z.R. and L.E. Back, Higher surface Bowen ratios ineffective at
increasing updraft intensity, Geophys. Res. Lett., 42(23), 10,503–10,511,
2015, doi:10.1002/2015GL066878.
[58] Williams, E.R., V.C. Mushtak, D. Rosenfeld, S.J. Goodman, and D.J.
Boccippio, Thermodynamic conditions favorable to superlative thunder-
storm updraft, mixed phase microphysics and lightning flash rate, Atmos.
Res., 76, 288–306, 2005b.
[59] Williams, E., D. Rosenfeld, N. Madden, C. Labrada, J. Gerlach, and L.
Atkinson, The role of boundary layer aerosol in the vertical development of
precipitation and electrification: Another look at the contrast between
lightning over land and over ocean, 11th Int’l Conf. on Atmospheric
Electricity, Guntersville, AL, June 7–11, 1999.
[60] Freud, E. and D. Rosenfeld, Linear relation between convective cloud drop
number concentration and depth for rain initiation, J. Geophys. Res., 117, D-
2207, 2012, doi:10.1029/2011JD016457.
612 Lightning electromagnetics: Volume 2
[61] Mattos, E.V., L.A.T. Machado, E. Williams, S.J. Goodman, R.J. Blakeslee, and
J. C. Bailey, Electrification life cycle of incipient thunderstorms, J. Geophys.
Res., 122(8), 4670–4697, 2017, https://doi.org/10.1002/2016JD025772.
[62] Liu, C. and E.J. Zipser, “Warm rain” in the tropics: deasonal and regional
distributions based on 9 yr of TRMM data, J. Climate, 22, 767–779, 2009,
doi:10.1175/2008CLI2641.1.
[63] Braga, R.C., D. Rosenfeld, R. Weigel, et al., Further evidence for CCN aerosol
concentrations determining the height of warm rain and ice initiation in con-
vective clouds over the Amazon basin, Atmos. Chem. Phys., 17, 14433–14456,
2017.
[64] Atlas, D., The balance level in convective storms, J. Atmos. Sci., 23,
635–651, 1966.
[65] Lhermitte, R. and E. Williams, Thunderstorm electrification: a case study,
J. Geophys. Res., 90(D4), 6071–6078, 1985, doi:10.1029/JD090iD04p06071.
[66] Williams, E. and R. Lhermitte, Radar tests of the precipitation hypothesis for
thunderstorm electrification, J. Geophys. Res., 88, 10984–10991, 1983.
[67] Lambrechts, J.d.V., The value of water drainage in upcast mine shafts and fan-
drifts, J. Chem., Metal. Mining Soc. South Africa, 56, 307–324 and 383–387, 1956.
[68] Carte, A.E., Mine shafts as a cloud physics facility, in Proc. Int’l Conference
on Cloud Physics, Toronto, Canada, August 26–30, 1968.
[69] Carte, A.E. and P.E. Lourens, Water year 1970: Convention on water for the
future; South Africa (Republic) 1970, Studies of Cloud and Rain in Mine
Shafts, Department of Water Affairs, 1970.
[70] Zipser, E.J., Deep cumulonimbus clouds in the tropics with and without
lightning, Monthly Weather Rev., 122, 1837–1851, 1994.
[71] Lindzen, R.S., Dynamics in Atmospheric Physics, Cambridge University
Press, Cambridge, 1990.
[72] Charney, J.G., The dynamics of long waves in a baroclinic westerly current,
J. Meteorol., 4, 135–162, 1947.
[73] Wallace, J.M. and P.V. Hobbs, Atmospheric Science: An Introductory
Survey, Academic Press, London, 1977.
[74] Yoshida, S., T. Morimoto, T. Ushio and Z. Kawasaki, A fifth-power relationship
for lightning activity from Tropical Rainfall Mission satellite observations, J.
Geophys. Res., 114, D09104, 2009, doi:10.1029/2008JD010370.
[75] Bang, S.D. and E.J. Zipser, Differences in size spectra of electrified storms
over land and ocean, Geophys. Res. Lett., 42(16), 6844–6851, 2015,
doi:10.1002/2015GL065264.
[76] Bang, S.D. and E.J. Zipser, Seeking reasons for the differences in size
spectra of electrified storms over land and ocean, J. Geophys. Res., 121,
9048–9068, 2016.
[77] Orville, R.E., Cloud-to-ground lightning in the Blizzard of ’93, Geophys.
Res. Lett., 20, 1367–1370, 1993.
[78] Pessi, A.T. and S. Businger, The impact of lightning data assimilation on a
winter storm simulation over the North Pacific, Ocean Monthly Weather
Rev., 137, 3177–3195, 2009.
Lightning and climate change 613
[93] Israel, H., Atmospheric Electricity, Vol. II, Israel Program for Scientific
Translations, Ltd, ISBN 0 7065 1129 8, 1973.
[94] Bailey, J.C., R.J. Blakeslee, D.E. Buechler, and H.J. Christian, Diurnal
lightning distributions as observed by the Optical Transient Detector
(OTD) and the Lightning Imaging Sensor (LIS). Proceedings of the 13th
International Conf. on Atmos. Elec., Vol. II, pp. 657–660. Organized by the
International Commission on Atmospheric Electricity (ICAE/IAMAS/
IUGG): August 13–17, 2007, Beijing Friendship Hotel, China, 2007.
[95] Cecil, D.J., D.E. Buechler, and R.J. Blakeslee, Gridded lightning climato-
logy from TRMM-LIS and OTD: dataset description, Atmos. Res.,
135–136, 404–414, 2014.
[96] Williams, E.R., Global circuit response to temperature on distinct time scales: a
status report, in M. Hayakawa (ed.), Atmospheric and Ionospheric Phenomena
Associated with Earthquakes, Terra Scientific Publishing, Tokyo, 1999.
[97] Markson, R., Ionospheric potential variation from temperature change over
continents, Proceedings of the 12th International Conf. on Atmos. Elec.,
Vol I, 283–286, Versailles, France, June 9–13, 2003.
[98] Markson, R., The global circuit intensity: its measurement and variation
over the last 50 years, Bull. Am. Met. Soc., 223–241, 2007, doi:10.1175/
BAMS-88-2-223.
[99] Phillipin, S. and E.A. Betterton, Cloud condensation nuclei concentrations
in Southern Arizona: instrumentation and early observations, Atmos. Res.
43, 263–275, 1997.
[100] Kendrew, W.G., The Climates of the Continents, Oxford University Press,
London, 3rd ed., New York, 473 pp., 1942.
[101] Williams, E.R., Global circuit response to seasonal variations in global
surface air temperature, Monthly Weather Rev., 122, 1917–1929, 1994.
[102] Christian, H.J., R.J. Blakeslee, D.J. Boccippio, et al., Global frequency and
distribution of lightning as observed from space by the Optical Transient
Detector, J. Geophys. Res., 108, 4005, 2003, doi:10.1029/2002JD002347.
[103] Füllekrug, M. and A. Fraser-Smith, Global lightning and climate variability
inferred from ELF field variations, Geophys. Res. Lett., 24, 2411–2414,
1997.
[104] Hogg, A.R., Air-earth current observations in various localities, Arch.
Meteor., 3, 40–55, 1950.
[105] Sátori, G. and B. Zieger, Spectral characteristics of Schumann resonances
observed in Central Europe, J Geophys Res., 101, 29663–29669, 1996.
[106] Nickolaenko, A.P., G. Sátori, B. Zieger, L.M. Rabinowicz, and I.G.
Kudintseva, Parameters of global thunderstorm activity deduced from the
long-term Schumann resonance records, J. Atmos. Sol.-Terr. Phys., 60,
387–399, 1998.
[107] Sátori, G., V. Mushtak, and E. Williams, Schumann resonance signatures of
global lightning activity, in H.-D. Betz, U. Schumann and P. Laroche,
(eds.) Lightning: Principles, Instruments and Applications, Springer,
New York, NY, pp. 347–386, 2009, ISBN: 978-1-4020-9078-3.
Lightning and climate change 615
[125] LaVigne, T., C. Liu, W. Deierling and D. Mach, Relationship between the
global electric circuit and electrified cloud parameters at diurnal, seasonal
and interannual time scales, J. Geophys. Res., 122, 8525–8542, 2017,
doi:10.1002/2016JD026442.
[126] Peterson, M., W. Deierling, C. Liu, D. Mach, and C. Kalb, A TRMM/GPM
retrieval of the total mean source current for the global electrical circuit, J.
Geophys. Res., 122, 10025–10049, 2017, doi:10.1002/2016JD026336.
[127] Liu, C., E.R. Williams, E.J. Zipser, and G. Burns, Diurnal variation of
global thunderstorms and electrified shower clouds and their contribution
to the global electrical circuit, J. Atmos. Sci., 67, 309–323, 2010.
[128] Rycroft, M.J., A. Odzimek, N.F. Arnold, M. Fűllekrug, A. Kulak, and T.
Neubert, New model simulations of the global atmospheric electric circuit
driven by thunderstorms and electrified shower clouds: the roles of light-
ning and sprites, J. Atmos. Solar-Terr. Phys., 69, 2485–2509, 2007.
[129] Williams, E.R. and G. Sátori, Lightning, thermodynamic and hydrological
comparison of the two tropical continental chimneys, J. Atmos. Sol. Terr.
Phys., 66, 1213–1231, 2004.
[130] Wilson, C.T.R., Investigations on lightning discharges and on the electric
field of thunderstorms, Phil. Trans. A., 221, 73–115, 1920.
[131] Chronis, T.G., S.J. Goodman, D. Cecil, D. Buechler, F.J. Robertson, and J.
Pittman, Global lightning activity from the ENSO perspective, Geophys.
Res. Lett., 35, L19804, 2008, doi:10.1029/2008GL034321.
[132] Pinto, O., III, Thunderstorm climatology of Brazil: ENSO and Tropical
Atlantic connections, Int’l J. Climatology, 35(6), 871–878, 2015,
doi:10.1002/joc.4022.
[133] Richey, J.E., C. Nobre, and C. Deser, Amazon discharge and climate
variability: 1903–1985, Science, 246, 101–103, 1989.
[134] Williams, E., A. Dall’Antonia, V. Dall’Antonia, et al., The drought of the
century in the Amazon Basin: an analysis of the regional variation of
rainfall in South America in 1926. Acta Amazonica, 35(2), 231–238, 2005.
[135] Grimm, A. M. and A. Natori, Climate change and interannual variability of
precipitation in South America, Geophys. Res. Lett., 33, L19706, 2006,
doi:10.1029/2006GL026821.
[136] Bruick, Z.S., K.L. Rasmussen, A.K. Rowe and L.A. McMurdie,
Characteristics of Intense Convection in Subtropical South America as
Influenced by El Niño–Southern Oscillation, Mon. Wea. Rev., 147, 1947–
1966, 2019.
[137] Goodman, S.J., D.E. Buechler, K. Knupp, D. Driscoll, and E.E. McCaul,
Jr.,The 1997–98 El Nino event and related wintertime lightning variations
in the southeastern United States, Geophys. Res. Lett., 27, 541–544, 2000.
[138] Dowdy, A.J., Seasonal forecasting of lightning and thunderstorm activity in
tropical and temperate regions of the world, in Nature, Scientific Reports,
11 Feb, 2016, doi:10.1038/srep20874.
[139] Marengo, J.A., E.R. Williams, L.M. Alves, W.R. Soares, and D.A.
Rodriguez, Extreme seasonal climate variations in the Amazon basin:
Lightning and climate change 617
[154] Boccippio, D.J., K. Cummins, H.J. Christian, and S.J. Goodman, Combined
satellite- and surface-based estimation of the intracloud–cloud-to-ground
lightning ratio over the continental United States, Mon. Wea. Rev., 129,
108–122, 2001.
[155] Price, C. and D. Rind, A simple lightning parameterization for calculating
global lightning distributions, J. Geophys. Res., 97, 9919–9933, 1992.
[156] Yamamoto, K., T. Nakashima, S. Sumi, and A. Ametani, About 100 years sur-
vey of the surface temperatures of Japan Sea and lightning days along the coast,
in Int’l Conf. on Lightning Protection, Estoril, Portugal, 25–30 September, 2016.
[157] Hansen, J., M. Sato, R. Ruedy, G.A. Schmidt, K. Lo, and A. Persin, Global
temperature in 2017, NASA GISS website, 2018.
[158] Williams, E., A. Guha, R. Boldi, H. Christian, and D. Buechler, Global
lightning activity and the hiatus in global warming, J. Atmos. Sol. Terr.
Phys., 189, 27–34, 2019.
[159] Alter, R.E., H.C. Douglas, J.M. Winter and E.A.B. Eltahir, Twentieth century
regional climate change during the summer in the central United States attrib-
uted to agricultural intensification, Geophys. Res. Lett., 45 1586–1594, 2018.
[160] Houghton, H.G., Physical Meteorology, MIT Press, New York, NY, 1985.
[161] Altaratz, O., I. Koren, Y. Yair, and C. Price, Lightning response to smoke
from Amazon fires, Geophys. Res. Lett., 37, L07801, 2010, doi:10.1029/
2010GL042679.
[162] Mansell, E.R. and C.L. Ziegler, Aerosol effects on simulated storm elec-
trification and precipitation in a two-moment bulk microphysics model, J.
Atmos. Sci., 70, 2032–2050, 2013.
[163] Rosenfeld, D. and I. Lensky, Satellite-based insights into precipitation
formation processes in continental and maritime convective clouds, Bull.
Am. Met. Soc., 79, 2457–2476, 1998.
[164] Orville, R.E., G. Huffines, J. Nielson-Gammon, et al., Enhancement of
cloud-to-ground lightning over Houston, Texas, Geophys. Res. Lett., 28,
2597–2600, 2001.
[165] Stolz, D.C., S.R. Rutledge, and J.R. Pierce, Simultaneous influences of
thermodynamics and aerosols on deep convection and lightning in the tro-
pics, J. Geophys. Res.: Atmos., 120(12), 6207, 2015.
[166] Altaratz, O., B.A. Kucienska, G.B. Kostinski, Raga and I. Koren, Global
association of aerosol with flash density of intense lightning, Env. Res.
Lett., 114037, 2017.
[167] Yuan, T., L.A. Remer, K.E., Pickering and H. Yu, Observational evidence
of aerosol enhancement of lightning activity and convective invigoration,
Geophys. Res. Lett., 38, L04701, 2011, doi:10.1029/2010GL046052.
[168] Keskinen, J. and T. Ronkko, Can real-world diesel exhaust particle size
distribution be reproduced in the laboratory? A critical review, Air Waste
Manag. Assoc., 60, 1245–1255, 2010, doi:10.3155/1047-3289.60.10.1234.
[169] Li, Z., F. Niu, J. Fan, Y. Liu, D. Rosenfeld, and Y. Ding, Long-term
impacts of aerosols on the vertical development of clouds and precipitation,
Nature Geosci., 4, 888–894, 2011.
Lightning and climate change 619
[185] York, A., U. Bhatt, R. Thoman, and R. Ziel, Wildland fire in high latitudes,
Arctic report card, NOAA, 2017, https://arctic.noaa.gov/Report-Card/Report-
Card-2017/ArtMID/7798/ArticleID/692/Wildland-Fire-in-High-Latitudes.
[186] Liu, Y., E. Williams, Z. Li, et al., Lightning enhancement in moist con-
vection with smoke-laden air advected from Australian wildfires, Geophys.
Res. Lett., 48(11), e2020GL092355, 2021.
[187] Ballinger, T.J., J. E. Overland, M. Wang, et al., Surface Air Temperature, Arctic
Essays, 2021, doi: 10.25923/53xd-9k68. https://arctic.noaa.gov/Report-Card/
Report-Card-2021/ArtMID/8022/ArticleID/948/Surface-Air-Temperature.
[188] Pithan, F and T. Mauritsen, Arctic amplification dominated by temperature
feedbacks in contemporary climate models, Nature Geosci., 7, 181–184,
2014, https://doi.org/10.1038/ngeo2071.
[189] Bieniek, P.A., J.E. Walsh, R.L. Thoman, and U.S. Bhatt, Using climate
divisions to analyze variations and trends in Alaska temperature and pre-
cipitation, J. Clim., 27, 2800–2818, 2014.
[190] Montanyà, J., F. Fabró, O. van der Velde, et al., Global distribution of
winter lightning: a threat to wind turbines and aircraft, Nat. Hazards Earth
Syst. Sci., 16, 1465–1472, 2016, doi:10.5194/nhess-16-1465-2016.
[191] Chen, G., J. Lu, and D.M. Frierson, Phase speed spectra and the latitude of
surface westerlies: Interannual variability and global warming trend, J.
Clim., 21(22), 5942–5959, 2008.
[192] Archer, C.L., and Caldeira, K., Historical trends in the jet streams. Geophys.
Res. Lett., 35, L08803, 2008, https://doi.org/10.1029/2008GL033614.
[193] Yin, J., A consistent poleward shift of the storm tracks in simulations of
21st century climate, Geophys. Res. Lett., 32, L18701, 2005, https://doi.org/
10.1029/2005GL023684.
[194] Norris, J. R., R.J. Allen, A.T. Evan, M.D. Zelinka, C.W. O’Dell, and S.A.
Klein, Evidence for climate change in the satellite cloud record, Nature,
536(7614), 72–75, 2016.
[195] Yang, H., G. Lohmann, U. Krebs-Kanzow, et al., Poleward shift of the
major ocean gyres detected in a warming climate. Geophys. Res. Lett., 47,
e2019GL085868. https://doi.org/10.1029/2019GL085868, 2020.
[196] Matsumura, S., K. Yamazaki, and T. Horinouchi, Robust asymmetry of the
future Arctic polar vortex is driven by tropical Pacific warming. Geophys. Res.
Lett., 48, e2021GL093440, 2021, https://doi.org/10.1029/2021GL093440.
[197] Hansen, J., Storms of My Grandchildren, Bloomsbury, New York, NY, 2009.
[198] Maddox, R.A., Mesoscale convective complexes, Bull. Amer. Meteor. Soc.,
61, 1374–1387, 1980.
[199] Jirak, I., W. Cotton, and R. McAnelly, Satellite and radar survey of
mesoscale convective system development, Monthly Weather Rev., 131.
2428, 2003, 10.1175/1520-0493(2003)131<2428:SARSOM>2.0.CO;2.
[200] Walsh, K.J.E., M. Fiorino, C.W. Landsea, and K.L. McInnes, Objectively
determined resolution-dependent threshold criteria for the detection of
tropical cyclones in climate models and reanalyses, J. Clim., 20(10), 2307–
2314, 2007, doi:10.1175/jcli4074.1.
Lightning and climate change 621
[201] Camelo J, T.L. Mayo and E.D. Gutmann, Projected climate change impacts
on Hurricane storm surge inundation in the Coastal United States, Front.
Built Environ., 6, 588049, 2020, doi: 10.3389/fbuil.2020.588049.
[202] Atlas, D., K.R. Hardy, R. Wexler, and R.J. Boucher, On the origin of hur-
ricane spiral bands: Technical Conference on Hurricanes and Tropical
Meteorology, Mexico. Geofis. Int., 3, 123–132, 1963.
[203] Williams, E.R., Meteorological Aspects of Thunderstorms, in H. Volland
(Ed.), CRC Handbook on Atmospheric Electrodynamics, Vol. I, CRC Press,
London, 1995.
[204] Squires, K. and S. Businger, The morphology of eyewall lightning out-
breaks in two Category 5 hurricanes, Monthly Weather Rev., 136(5), 1706–
1726, 2008, doi:10.1175/2007mwr2150.1.
[205] Pan, L.X., X. S. Qie, D.X. Liu, et al., The lightning activities in super typhoons
over the Northwest Pacific, Sci. China (Ser. D), 53(8), 1241–1248, 2010.
[206] Plotnik, T., C. Price, J. Saha, and A. Guha, Transport of water vapor from
tropical cyclones to the upper troposphere, Atmosphere, 12, 1506, 2021,
https://doi.org/10.3390/ atmos12111506.
[207] Pan, L.X., X. Qie and D. Wang, Lightning activity and its relation to the
intensity of typhoons over the Northwest Pacific Ocean, Adv. Atmos. Sci.,
31(3), 581–592, 2014, doi: 10.1007/soo376-013-3115-y.
[208] Emanuel, K.A., The dependence of hurricane intensity on climate, Nature,
326(6112), 483–485, 1987, doi:10.1038/326483a0.
[209] Emanuel, K., Increasing destructiveness of tropical cyclones over the past
30 years, Nature, 436, 686–688, 2005, https://doi.org/10.1038/nature03906.
[210] Knutson, T., J. McBride, J. Chan, et al., Tropical cyclones and climate change,
Nature Geosci., 3, 157–163, 2010, https://doi.org/10.1038/ngeo779.
[211] Knutson, T., S.J. Camargo, J.C.L. Chan, et al., Tropical cyclones and cli-
mate change assessment: Part I. Detection and attribution, Bull. Am.
Meteorol. Soc, 100(10), 1987–2007, 2019 doi:10.1175/bams-d-18-0189.1
[212] Lynn, B.H., R. Healy, and L.M. Druyan, Investigation of Hurricane Katrina
characteristics for future, warmer climates, Clim. Res., 39, 75–86, 2009,
doi:10.3354/cr00801.
[213] Vecchi, G.A. and T.R. Knutson, On estimates of historical North Atlantic
tropical cyclone activity, J. Clim., 2008(14), 3580–3600, 2008, doi:10.1175/
2008jcli2178.1
[214] USGCRP, Fourth National Climate Assessment, 1–470, 2018.
[215] Holland, G. and C.L. Bruyère, Recent intense hurricane response to global
climate change. Clim. Dyn., 42, 617–627, 2014, https://doi.org/10.1007/
s00382-013-1713-0.
[216] Shao, X.-M., J. Harlin, M. Stock, et al., Katrina and Rita were lit up with
lightning, EOS, 86(42), 398–399, 2005.
[217] Fierro, A.O., X.-M. Shao, T. Hamlin, J.M. Reisner and J. Harlin, Evolution
of eyewall convective events as indicated by intracloud and cloud-to-
ground lightning activity during the rapid intensification of Hurricanes Rita
and Katrina, Mon. Wea. Rev., 139, 1492–1504, 2010.
622 Lightning electromagnetics: Volume 2
[218] Cooray, V., R. Jayaratne and K.L. Cummins, On the peak amplitude of light-
ning return stroke currents striking the sea, Atmos. Res., 149, 372–376, 2014.
[219] Asfur, M., C. Price, J. Silverman, and A. Wishkerman, Why is lightning
more intense over the oceans?, J. Atmos. Sol. Terr. Phys., 252, 105259, 2020.
[220] Asfur, M., J. Silverman, and C. Price, Ocean acidification may be
increasing the intensity of lightning over the oceans, Sci. Rep., 10, 21847,
2020, https://doi.org/10.1038/s41598-020-79066-8
[221] Montanya, J., F. Molina, Pol, A. Marcelo, et al., Small scale simulation of
lightning using vertical wires deployed by drones: on the land and sea peak
current asymmetry, in Int’l Conf. on Lightning Protection, Cape Town,
South Africa, 2022.
[222] Cooray, V., M. Rubinstein and F.A. Rachidi, Self-consistent return stroke
model that includes the effect of the ground conductivity at the strike point.
Atmosphere, 13, 593, 2022.https://doi.org/10.3390/atmos13040593.
[223] Chronis, T., K. Cummins, R. Said, Climatological diurnal variation of
negative CG lightning peak current over the continental United States, J.
Geophys. Res.,Atmos., 120, 582–589, 2015, doi: 10.1002/2014JD022547.
[224] Holzworth, R.H., M.P. McCarthy, J.B. Brundell, A.R. Jacobson, & C.J. Rodger,
Global distribution of superbolts, Journal of Geophysical Research:
Atmospheres, 124, 9996–10,005. https://doi.org/10.1029/ 2019JD030975, 2019.
[225] Lyons, W.A., E.C. Bruning, T.A. Warner, et al., Megaflashes: just how
long can a lightning discharge get?, Bull. Am. Meteorol. Soc., 101, E73–
E86, 2019, BAMS-D-19-0033.1, doi:10.1175/BAMS-D-19-0033.1.
[226] Turman, B.N., Detection of lightning superbolts, J. Geophys. Res., 82,
2566–2568, 1977, https://doi.org/10.1029/JC082i018p02566.
[227] Vonnegut, B., O.H. Vaughan Jr., M. Brook, and P. Krehbiel, Mesoscale
observations of lightning from the space shuttle, Bull. Amer. Meteor. Soc.,
66, 20–29, 1985, https://doi.org/10.1175/1520-0477(1985)066<0020:
MOOLFS>2.0.CO;2.
[228] Franz, R.C., R.J. Nemzak, and J.R. Winkler, Television image of a large
upward electrical discharge above a thunderstorm system, Science, 249,
48–51, 1990.
[229] Marshall, T.C. and W.D. Rust, Two types of vertical electrical structures in
stratiform precipitation regions of mesoscale convective systems, Bull.
Amer. Meteor. Soc., 74, 2159–2170, 1993, https://doi.org/10.1175/1520-
0477(1993)074 2.0.CO;2.
[230] Stolzenburg, M., W.D. Rust, B.F. Smull, and T.C. Marshall, Electrical structure
in thunderstorm convective regions: 1. Mesoscale convective systems. J.
Geophys. Res., 103, 14 059–14 078, 1998, https://doi.org/10.1029/97JD03546.
[231] Williams, E.R., The positive charge reservoir for sprite-producing light-
ning, J. Atmos. Sol. Terr. Phys., 60, 689–692, 1998.
[232] Carey, L.D., M.J. Murphy, T.L. McCormick, and N.W.S. Demetriades,
Lightning location relative to storm structure in a leading-line, trailing-
stratiform mesoscale convective system, J. Geophys. Res., 110, D03105,
2005, doi:10.1029/2003JD004371.
Lightning and climate change 623
[266] Mühleisen, R., The global circuit and its parameters, 467–476, in H.
Dolezalek and R. Reiter (eds.), Electrical Processes in Atmospheres,
Steinkopff, Darmstadt, 1977.
[267] Bering, E.A., A.A. Few and J.R. Benbrook, The global electric circuit,
Physics Today, 51, 24–30, 1998.
[268] Chalmers, J.A., Atmospheric Electricity, 2nd ed., Pergamon Press, London,
1967.
[269] Williams, E.R., The global electrical circuit: a review, Atmos. Res., 91,
140–152, 2009.
[270] Heckman, S., E. Williams and R. Boldi, Total global lightning inferred
from Schumann resonance measurements, J. Geophys. Res., 103, 31775–
31779, 1998.
[271] Madden, T.R. and W.B. Thompson, Low-frequency electromagnetic
oscillations of the Earth-ionosphere cavity, Rev. Geophys., 3, 211, 1965.
[272] Nickolaenko, A.P. and M. Hayakawa, Resonances in the Earth-ionosphere
Cavity, Kluwer Academic Publishers, London, 2002.
[273] Polk C., Schumann resonances, in H. Volland (ed.), CRC Handbook of
Atmospherics, Vol. I, CRC Press, Boca Raton, FL., pp. 55–82, 1982.
[274] Sentman, D.D., Schumann resonances, in H. Volland (ed.), Handbook of
Atmospheric Electrodynamics, Vol. I, CRC Press, 408 pp., 1995
(Chapter 11).
[275] Williams, E.R. and E.A. Mareev, Recent progress on the global electrical
circuit, Atmos. Res., 135–136, 208–227, 2014.
[276] Williams, E., R. Boldi, R. Markson and M. Peterson, Comparative behavior
of the DC and AC global circuits, in XVI International Conference on
Atmospheric Electricity, Nara city, Nara, Japan, 17–22 June, 2018.
[277] Mühleisen, R.P., New determination of the air-earth current over the ocean
and measurement of ionospheric potential, PAGEOPH, 84, 112–115, 1971.
[278] Dyrda, M.A., M. Kulak, M. Mlynarczyk, et al., Application of the
Schumann resonance decomposition in characterizing the main African
thunderstorm center, J. Geophys. Res., Atmospheres, 119, 13338–13349,
2014.
[279] Williams, E., V. Mushtak, A. Guha, et al., Inversion of multi-station
Schumann resonance background records for global lightning activity in
absolute units, Fall Meeting, American Geophysical Union, San Francisco,
CA, December, 2014.
[280] Prácser, E., T. Bozóki, G. Sátori, E. Williams, A. Guha and H. Yu,
Reconstruction of global lightning activity based on Schumann Resonance
measurements: Model description and synthetic tests, Radio Science, 54,
254–267, 2019.
[281] Guha, A., E. Williams, R. Boldi, et al., Aliasing of the Schumann resonance
background signal by sprite-associated Q-bursts, J. Atmos. Sol. Terr. Phys.,
165–166, 25–37, 2017.
[282] Hobara, Y., E. Williams, M. Hayakawa, R. Boldi and E. Downes, Location
and electrical properties of sprite-producing lightning from a single ELF
626 Lightning electromagnetics: Volume 2
site, in M. Fullekrug, E.A. Mareev and M.J. Rycroft (eds.), Sprites, Elves
and Intense Lightning Discharges, NATO Science Series, II, Mathematics,
Physics and Chemistry, Vol. 225, Springer, New York, NY, 2006.
[283] Huang, E., E. Williams, R. Boldi, et al., Criteria for sprites and elves based
on Schumann resonance observations, J. Geophys. Res., 104, 16943–16964,
1999.
[284] Kemp, D.T., The global location of large lightning discharges from single
station observations of ELF disturbances in the Earth-ionospheric cavity, J.
Atmos. Terr. Phys., 33, 919–928, 1971.
[285] Kemp, D.T., and D.L. Jones, A new technique for the analysis of transient
ELF electromagnetic disturbances within the Earth-ionosphere cavity. J.
Atmos. Terr. Phys., 33, 567–572, 1971.
[286] Ogawa, T., Y. Tanaka, M. Yasuhara, A.C. Fraser-Smith and R. Gendrin,
Worldwide simultaneity of occurrence of a Q-type ELF burst in the
Schumann resonance frequency range, J. Geomag. Geoelectr., 19, 377–
384, 1967.
[287] Williams, E.R., W.A. Lyons, Y. Hobara, et al., Ground-based detection of
sprites and their parent lightning flashes over Africa during the
2006 AMMA campaign, Quart. J. Roy. Met. Soc., 136, 257–271, 2010
(Special Issue: Advances in understanding atmospheric processes over
West Africa through the AMMA field campaign).
[288] Yamashita, K., Y. Takahashi, M. Sato and H. Kase, Improvement in
lightning geolocation by time-of-arrival method using global ELF network
data, J. Geophys. Res., 116, A00E61, 12 PP, 2011, doi:10.1029/
2009JA014792.
[289] Markson, R. and C. Price, Ionospheric potential as a proxy index for global
temperature, Atmos. Res., 51, 309–314, 1999.
[290] Allen, M.R. and W.J. Ingram, Constraints on future changes in climate and
the hydrologic cycle, Nature, 419, 224–232, 2002.
[291] Romps, D.M., J.T. Seeley, D. Vollaro and J. Molinari, Projected increase in
lightning strikes in the United States due to global warming, Science, 346,
851–854, 2014.
[292] Romps, D.M., Clausius-Clapeyron scaling of CAPE for analytic solutions
to RCE, J. Atmos. Sci., 73, 3719–3737, 2016.
[293] Guerreiro, S.B., H.J. Fowler, R. Barbero, et al., Detection of continental-
scale intensification of hourly rainfall extremes, Nature Climate Change, 8,
803–808, 2018, https://doi.org/10.1038/s41558-018-0245-3.
[294] Yair, Y., Lightning hazards to human societies in a changing climate, Env.
Res. Lett., 13, 123002, 2018.
[295] Huaichuan, R., Globalisation, Transition and Development in China: The
Case of the Coal Industry, Routledge, London, 2004.
Index