Вы находитесь на странице: 1из 462

Basics of

Atmospheric Science

A. Chandrasekar
Senior Professor and Head
Department of Earth and Space Sciences
Indian Institute of Space Science and Technology
Thiruvananthapuram

PHI Learning Private Limited


New Delhi-110001
2010
BASICS OF ATMOSPHERIC SCIENCE
A. Chandrasekar

© 2010 by PHI Learning Private Limited, New Delhi. All rights reserved. No part of this book may
be reproduced in any form, by mimeograph or any other means, without permission in writing from
the publisher.

ISBN-978-81-203-4022-0

The export rights of this book are vested solely with the publisher.

Published by Asoke K. Ghosh, PHI Learning Private Limited, M-97, Connaught Circus,
New Delhi-110001 and Printed by Mohan Makhijani at Rekha Printers Private Limited,
New Delhi-110020.
In memory of
My respected teacher
Padmashri (Late) Professor R. Ananthakrishnan
Former Director, Institute of Tropical Meteorology, Pune
Contents

Foreword ................................................................................................................................. xv
Preface .............................................................................................................................. xvii

1. Introductory Survey of the Atmosphere .................................................. 1–15


1.1 Introduction 1
1.2 Origin and Composition of the Atmosphere 2
1.3 Distribution of Pressure and Density 4
1.4 Ionosphere, Atmospheric Electric Field and Magnetosphere 6
1.5 Distribution of Temperature and Winds 8
1.5.1 Distribution of Temperature 8
1.5.2 Distribution of Winds 11
1.6 Atmosphere as a Fluid and Fluid Continuum 13
1.7 Physical Laws 13
1.8 Determinism and Chaos 14
Review Questions 15

2. Atmospheric Observations...................................................................... 16–34


2.1 Overview and Importance of Meteorological Observation 17
2.2 Measurement of Temperature and Humidity 18
2.2.1 Temperature Measurement 18
2.2.2 Humidity Measurement 21
2.3 Measurement of Wind and Pressure 22
2.3.1 Wind Measurement 22
2.3.2 Atmospheric Pressure Measurement 24
2.4 Measurement of Precipitation 25
2.5 Modern Meteorological Instruments 26
v
vi u CONTENTS

2.6 Surface and Upper Air Observational Network 27


2.6.1 Surface Observational Network 27
2.6.2 Radar Network 28
2.6.3 Upper Air Observational Network 28
2.7 Satellite Observation 31
Review Questions 33

3. Atmospheric Thermodynamics .............................................................. 35–75


3.1 Gas Laws and Equation of State for a Mixture of Ideal Gases 36
3.1.1 Mixture of Gases 37
3.2 Work, Heat and First Law of Thermodynamics 38
3.2.1 Work 38
3.2.2 Work Done by a System Expanding against External Forces 39
3.2.3 Heat 41
3.2.4 First Law of Thermodynamics 41
3.2.5 Internal Energy and Enthalpy 42
3.2.6 Specific Heat Capacity 44
3.3 Adiabatic Processes 46
3.4 Moist Thermodynamics and Latent Heats 47
3.4.1 Measures of Water Vapour in Air 47
3.4.2 Equation of State of Moist Air 48
3.4.3 Latent Heat 49
3.5 Hydrostatic Equilibrium 49
3.5.1 Geopotential 50
3.5.2 Scale Height and Height Computations Using the
Hypsometric Equation 51
3.5.3 Reduction of Pressure to Sea Level 54
3.6 Thermodynamic Diagram 55
3.6.1 Emagram 55
3.6.2 Tephigram 55
3.6.3 Skew T–log p Diagram 56
3.6.4 Stuve Diagram 56
3.7 Hydrodynamic Stability—Parcel and Slice Methods 56
3.7.1 Saturated Adiabatic and Pseudoadiabatic Processes 56
3.7.2 Saturated Adiabatic Lapse Rate 57
3.7.3 Equivalent Potential Temperature 57
3.7.4 Stability Criteria Using Parcel Method 59
3.7.5 Stability Criteria Using Slice Method 63
3.7.6 Entrainment Effects 65
3.8 Entropy and Second Law of Thermodynamics 67
3.8.1 Entropy 67
3.8.2 Second Law of Thermodynamics 68
3.8.3 Heat Engines and Refrigeration Cycles 69
CONTENTS u vii

3.9 Carnot Cycle and Clausius Clapeyron Equation 70


3.9.1 Carnot Cycle 70
3.9.2 Clausius Clapeyron Equation 71
Solved Examples 72
Review Questions 74

4. Radiation ................................................................................................ 76–104


4.1 Spectrum of Radiation 76
4.1.1 Overview 77
4.1.2 Electromagnetic Spectrum of Radiation 78
4.2 Black Body Radiation 78
4.2.1 Planck’s Law 79
4.2.2 Local Thermodynamical Equilibrium 81
4.2.3 Radiometric Quantities 82
4.3 Atmospheric Absorption of Solar Radiation 85
4.3.1 Absorption and Emission of Radiation by Molecules 85
4.3.2 Absorptivity and Emissivity 85
4.3.3 Kirchhoff’s Law 87
4.3.4 Reflectivity and Transmittivity 87
4.3.5 Absorption of Solar Radiation by Atmosphere 87
4.3.6 Indirect Estimate of Solar Irradiance at the Top of the
Atmosphere 89
4.3.7 Vertical Profile of Absorption 90
4.4 Scattering of Solar Radiation 91
4.5 Atmospheric Absorption and Emission of Infrared Radiation 92
4.6 Remote Temperature Sounding from Space 94
4.6.1 Calculation of the Surface Temperature 95
4.6.2 Two-layer Atmospheric Temperature Profile 96
4.6.3 Multi-layer Atmospheric Temperature Profile 97
Solved Examples 100
Review Questions 103

5. Clouds and Precipitation ..................................................................... 105–148


5.1 Atmospheric Aerosols 105
5.1.1 Aerosol Size and Concentration 106
5.1.2 Sources and Sinks of Atmospheric Aerosol 108
5.2 Nucleation of Water Vapour Condensation 109
5.2.1 Thermodynamic Potentials 109
5.2.2 Nucleation Theory of Water Vapour Condensation 111
5.2.3 Cloud Condensation Nuclei 117
5.3 Droplet Growth in Warm Clouds 118
5.3.1 Overview 118
5.3.2 Growth of Cloud Droplets in Warm Clouds
by Condensation 119
viii u CONTENTS

5.3.3 Growth of Cloud Droplets by Collision and Coalescence 122


5.3.4 The Continuous Collision Model 124
5.3.5 The Stochastic Collision Model 126
5.4 Formation and Growth of Ice Crystals in Cold Clouds 127
5.4.1 Homogeneous Nucleation of Ice Particles 127
5.4.2 Ice Nuclei 128
5.4.3 Bergeron Process 128
5.4.4 Growth Rate of Ice Crystals by Deposition 129
5.4.5 Hail Formation 130
5.4.6 Radiative Effects of Clouds 131
5.5 Mechanisms of Cloud Formation and Cloud Seeding 132
5.5.1 Mechanisms of Cloud Formation 132
5.5.2 Types of Clouds 133
5.5.3 Convective Clouds 135
5.5.4 Cloud Seeding 137
5.6 Role of Clouds and Precipitation Products in Charge Separation 138
5.6.1 Distribution of Charges in a Thunderstorm 138
5.6.2 Mechanisms for Charge Separation 139
5.6.3 Lightning Discharge 141
Solved Examples 142
Review Questions 147

6. Governing Laws of Atmospheric Motion .......................................... 149–186


6.1 Equation in a Rotating Coordinate System—Centripetal and
Coriolis Acceleration 150
6.1.1 Introduction 150
6.1.2 Rotating Frame of Reference 151
6.1.3 Equation of Motion in an Inertial Frame of Reference 153
6.2 Gravity and Pressure Gradient Forces 154
6.2.1 Pressure Gradient Force 154
6.2.2 Gravitational Force 155
6.2.3 Equation of Motion in a Rotating Coordinate System 155
6.2.4 Effects of Coriolis Force 156
6.2.5 Flow of Rivers on the Surface of the Earth 156
6.2.6 Effects of Coriolis Force Due to Relative Motion
along a Latitude Circle 157
6.2.7 Effects of Coriolis Force Due to Relative Motion
along a Meridian 158
6.2.8 Effects of Coriolis Force Due to Vertical Motion 159
6.2.9 Rossby Number 159
6.2.10 Gravity 160
6.3 Total, Local and Convective Derivatives 161
CONTENTS u ix

6.4 Continuity Equation 162


6.4.1 Eulerian Approach 162
6.4.2 Lagrangian Approach 164
6.5 Equations of Motion and Equations for Horizontal Flow 165
6.5.1 Equations of Motion in Spherical Coordinates 165
6.5.2 Scale Analysis of the Equations of Motion 167
6.5.3 f-plane and b-plane Approximations 168
6.5.4 Geostrophic Wind 169
6.5.5 Isallobaric Wind 170
6.5.6 Natural Coordinate System 170
6.5.7 Inertial Flow 172
6.5.8 Cyclostrophic Flow 173
6.5.9 Gradient Flow 174
6.6 Thermal Wind 176
6.7 Thermodynamic Energy Equation 178
Solved Examples 180
Review Questions 185

7. Atmospheric Motion ............................................................................ 187–213


7.1 Circulation and Vorticity 188
7.1.1 Vorticity 188
7.1.2 Decomposition of a Linear Velocity Field 188
7.1.3 Circulation 190
7.1.4 Kelvin’s Circulation Theorem 192
7.1.5 Bjerknes’ Circulation Theorem 193
7.1.6 Applications of Circulation Theorem 194
7.1.7 Vorticity in the Natural Coordinate System 195
7.2 Isobaric Coordinate System 196
7.2.1 Horizontal and Time Derivates in Isobaric
Coordinate System 197
7.2.2 Continuity Equation in Isobaric Coordinate System 199
7.2.3 Horizontal Equation of Motion in Isobaric
Coordinate System 200
7.2.4 Geostrophic and Thermal Wind Equations in
Isobaric Coordinates 200
7.3 Vorticity and Divergence Equations 201
7.3.1 Vorticity Equation 201
7.3.2 Divergence Equation 203
7.4 Absolute and Potential Vorticity 203
7.4.1 Absolute Vorticity 204
7.4.2 Potential Vorticity 204
Solved Examples 207
Review Questions 212
x u CONTENTS

8. Atmospheric Boundary Layer ............................................................ 214–238


8.1 Brief Consideration 214
8.2 Definition of Viscosity 214
8.3 Expression for Viscosity from Kinetic Theory 215
8.4 Viscous Forces in the Equation of Motion 217
8.5 Turbulence 219
8.6 Turbulence and Diffusion 220
8.7 Equations of Mean Motion in Turbulent Flow 221
8.8 Mixing Length 223
8.9 Surface and Ekman Layers 225
8.9.1 Surface Layer 225
8.9.2 Ekman Layer 227
8.10 Secondary Circulations and Spin-down in the Atmosphere 229
8.11 Secondary Circulations and Spin-down in a Teacup 231
Solved Examples 233
Review Questions 238

9. Waves in the Atmosphere .................................................................... 239–267


9.1 Rossby Waves 239
9.1.1 Perturbation Method 240
9.1.2 Some Basic Properties of Waves 240
9.1.3 Rossby Waves in a Barotropic Atmosphere 241
9.2 Gravity Waves in Shallow Water 244
9.3 Orographic and Sound Waves 246
9.3.1 Orographic Waves 246
9.3.2 Sound Waves 248
9.4 Internal Gravity Waves 250
9.5 Equatorial Waves 252
9.5.1 Shallow Water Equations 252
9.5.2 Equatorial Rossby and Rossby Gravity Waves 256
9.5.3 Mixed Rossby Gravity Waves 260
9.5.4 Equatorial Kelvin Wave 261
Solved Examples 263
Review Questions 266

10. Large-scale Meteorological Systems in Mid-Latitudes .................... 268–279


10.1 General Considerations 268
10.1.1 Air Masses 269
10.2 Fronts 270
10.2.1 Warm Front 270
10.2.2 Cold Front 272
10.2.3 Stationary Front 272
10.2.4 Occluded Front 273
CONTENTS u xi

10.3 Extratropical Cyclone 273


10.4 Jet Streams 275
Review Questions 278

11. Meteorological Systems in Low Latitudes ......................................... 280–326


11.1 General Considerations 280
11.2 Monsoons 281
11.2.1 Differential Heating of Land and Sea 287
11.2.2 Compressibility, Rotation and Moisture Effects 288
11.2.3 Tropical and Oceanic Convergent Zones 289
11.2.4 Intraseasonal and Interannual Variability of the
Indian Monsoon 290
11.3 Monsoon Disturbances and Semipermanent Monsoon Systems
Over India 295
11.3.1 Monsoon Disturbances 295
11.3.2 Semipermanent Monsoon Systems Over India 298
11.4 Tropical Cyclones 300
11.4.1 Factors Responsible for the Formation of
Tropical Cyclone 301
11.4.2 Climatology of Tropical Cyclones 302
11.4.3 Movement of Tropical Cyclones 303
11.4.4 Life Cycle of a Tropical Cyclone 306
11.4.5 Tropical Cyclone Structure 308
11.4.6 Eye and the Eyewall 311
11.5 Thunderstorms and Tornadoes 312
11.5.1 Thunderstorms 312
11.5.2 Life Cycle of Thunderstorms 313
11.5.3 Severe Thunderstorms and Squall Lines 315
11.5.4 Tornadoes 316
11.6 El Nino-Southern Oscillation 316
11.6.1 Overview of ENSO 317
11.6.2 Indian Ocean Dipole 323
11.6.3 ENSO and Indian Monsoon 323
Review Questions 324

12. Global Energy Balance ........................................................................ 327–344


12.1 Globally-averaged Atmospheric Energy Balance 327
12.1.1 Global Energy Balance Requirement for the
Earth’s Atmosphere 328
12.1.2 Global Energy Balance at the Earth Surface 330
12.1.3 Estimates of the Global Energy Balance for the
Earth–Atmospheric System 332
12.1.4 Energy Processes in the Upper Atmosphere 333
xii u CONTENTS

12.2 Internal, Potential and Kinetic Energy 334


12.2.1 Internal and Potential Energy 334
12.2.2 Kinetic Energy 335
12.3 Conversion of Potential and Internal Energies to Kinetic Energy 336
12.3.1 Available Potential Energy 336
12.4 Generation and Frictional Dissipation of Kinetic Energy 338
12.4.1 Generation of Kinetic Energy 338
12.4.2 Frictional Dissipation of Kinetic Energy 340
12.5 Atmosphere as a Heat Engine 343
Review Questions 343

13. General Circulation ............................................................................. 345–359


13.1 General Consideration 346
13.1.1 Overview 346
13.1.2 Observed Meridional Cross-section of Longitudinally-
averaged Zonal Wind and Temperature 347
13.1.3 Longitudinally-dependent Flow 349
13.1.4 Requirement on Theories of General Circulation 350
13.2 Meridional Circulation Model—Hadley Circulation 351
13.3 Angular Momentum Balance 353
13.4 Dishpan Experiments 358
Review Questions 359

14. Numerical Modelling of the Atmosphere ........................................... 360–389


14.1 General Considerations 360
14.1.1 Overview 361
14.1.2 The Finite Difference Method 362
14.1.3 Partial Differential Equations 363
14.2 Modern Numerical Weather Prediction 370
14.2.1 Overview 370
14.2.2 Observations 371
14.3 Data Assimilation 371
14.3.1 Overview 371
14.3.2 Objective Analysis 372
14.3.3 Initialization 373
14.3.4 Data Assimilation Cycle 374
14.4 Spectral and Finite Element Methods 375
14.4.1 Galerkin Method 375
14.4.2 Spectral Method 377
14.4.3 Finite Element Method 378
14.5 Challenges in Weather and Climate Forecasts 380
14.5.1 Weather Forecasting—A Historical Perspective 380
14.5.2 Ensemble Forecasting 383
14.5.3 Climate Forecasting 385
Review Questions 388
CONTENTS u xiii

15. Chaos in the Atmosphere .................................................................... 390–409


15.1 Illustrative Example of Chaos—Forced Pendulum 391
15.2 Poincare Section and Lyapunov Exponents 393
15.3 Period Doubling and Route to Chaos 397
15.4 Bouncing Ball Problem 400
15.5 Lorenz Attractor 403
15.6 Limits of Deterministic Predictability 407
Review Questions 408

Appendix 1 Useful Universal Physical Constants ................................................. 411–412


Appendix 2 Vector Identities ..................................................................................413–415
Appendix 3 Atmospheric Ozone .............................................................................416–418
Appendix 4 Equations of Motion in Spherical Coordinates .................................419–420
Appendix 5 Relaxation Methods ............................................................................421–422
Appendix 6 Von Neumann Stability Analysis ........................................................423–425
Appendix 7 Fortran Computer Program for Numerical
Solution of the Barotropic Vorticity Equation ...................................426–429
Appendix 8 Fortran Computer Program for Numerical
Solution of the Shallow Water Equation ...........................................430–434
Appendix 9 Fortran Computer Program for Numerical
Solution of the Forced Damped Pendulum .......................................435–436
Appendix 10 Fortran Computer Program for Numerical
Solution of the Lorenz System ...........................................................437–438

Bibliography................................................................................................................439–440
Index ...........................................................................................................................441–446
Foreword

I consider it a great privilege to write this Foreword to this extremely timely and useful book
on “Basics of Atmospheric Science” by Prof. A. Chandrasekar. There are many introductory
books in the market on Atmospheric Science that are very attractive with coloured
illustrations, but lack in depth. On the other hand, there are books that deal in depth only on
certain aspects of Atmospheric Science. During my long teaching career, I felt the need of a
book on Atmospheric Science that will not only introduce the different aspects of the field
with some depth, but also highlight the exciting challenges. This book by Prof. Chandrasekar
is going to fill the much needed gap for such an introductory textbook for undergraduate and
postgraduate students in Atmospheric Science.
The book does a great job of laying the foundation of all aspects of atmospheric science
related to weather and climate. While the first few chapters (Chapters 1–5) discuss the
fundamental processes such as the origin of the atmosphere, atmospheric thermodynamics,
atmospheric radiation and cloud and precipitation, the next few chapters (Chapters 6–8) lay
the foundation for theoretical understanding of weather and climate. The following few
chapters (Chapters 9–11) deal with large scale systems such as waves and synoptic
disturbances in both tropics and extra-tropics, while the driving factors for the observed
climate and the general circulation are introduced in Chapters 12 and 13. Finally, the book
ends with discussing advanced numerical modelling of the atmosphere and the challenging
problem of deterministic limit on weather predictability.
Today, the atmospheric science has emerged as a highly quantitative science. I am very
happy to see that Prof. Chandrasekar’s book attempts to make the learning quantitative by
introducing questions at the end of every chapter together with some model solutions. Also
for any student of meteorology, it is fundamental to learn and understand the differences
between the tropical and extra-tropical systems. I am happy to see that the book introduces
the students to both tropical and extra-tropical systems.

xv
xvi u FOREWORD

I was responsible for introducing atmospheric science to Prof. Chandrasekar. After


completion of his Ph.D. in Applied Mathematics, we worked together for a couple of years at
the Centre for Atmospheric and Oceanic Sciences, IISc Bangalore. He then joined IIT
Kharagpur and taught Atmospheric Science for nearly two decades. I am very happy to see
that he took time off to write this book that will be extremely valuable for the students.

B.N. Goswami
Indian Institute of Tropical Meteorology
Dr. Homi Bhabha Road
Pashan, Pune-411008
Preface

Atmospheric Science over the years has evolved into an exciting field of study with far-
reaching scientific, economic and societal implications. In recent times too, there has been a
greater appreciation of the importance of this branch of science. This is understandable
considering the grave issues which confront mankind, ranging from global warming arising
from man-induced activities affecting climate change to depleting food resources for the ever-
growing world population. Although many of us in the field of atmospheric science, had for
long, felt the need for a book on atmospheric science, dealing with not only the different
aspects of the field in depth, but also providing the readers with a comprehensive treatment of
the underlying physical principles. I did not imagine that I would indeed venture into writing
such a book.
I am indebted to numerous persons including my teachers, colleagues, students and
others who have played an important role in my career and development. I have dedicated this
book in memory of my teacher Prof. (Late) R. Ananthakrishnan, an inspiring teacher and a
person who epitomized both the desirable qualities of “high thinking” and “simple living”.
I have also benefited immensely from my association with Prof. G. Nath, my Ph.D. supervisor
at the Indian Institute of Science, Bangalore. I acknowledge with gratitude the help and
encouragement I received from Prof. B.N. Goswami, Director, Indian Institute of Tropical
Meteorology, Pune, over the years. After my Ph.D., I worked with Prof. B.N. Goswami at the
Indian Institute of Science, Bangalore during 1987–1988 and got introduced to some very
interesting research problems in the area of atmospheric science. Prof. B.N. Goswami, despite
his very busy schedule, has been kind enough to write the Foreword for this book and I thank
him for the same. Prof. B.N. Goswami provided several helpful suggestions on the contents as
well as the subject matter to be covered in this book. His timely comments and
encouragement finally led to the fruitful completion of the book.
I would be failing in my duty if I did not acknowledge the help and assistance I received
from Professor J. Srinivasan, Centre for Atmospheric and Oceanic Sciences, Indian Institute
of Science, Bangalore, Prof. T.S. Murty, University of Ottawa, Canada and Dr. Suresh,
xvii
xviii u PREFACE

India Meteorological Department, Chennai, in the preparation of this book. Both


Prof. Srinivasan and Prof. Murty were kind enough to go through the contents of this book
and provide pertinent and useful suggestions which helped in the overall improvement of the
book. Dr. Suresh took time off in going through the entire book and besides providing useful
comments, he also patiently corrected the errors which appeared in the equations as well as in
the text. I wish to thank my Ph.D. student Dr. S. Sandeep, Mr. Liju and Mr. Sai of
the Indian Institute of Technology Kharagpur who all helped in the preparation of the figures
in this book. But for their help, this book project would not have been completed in time. I
wish to acknowledge the help and assistance which I received from DGM, TERLS, VSSC,
Thiruvananthapuram and Dr. K.V.S. Namboodiri and his group at Meteorological facility at
VSSC, Head TDAD, VSSC and the members of the photography unit of VSSC for their help
in the preparation of the photos for the cover page. I also acknowledge the help I received
from Ms. Mary Maxine Browne, Purdue University who read the first few chapters and
carefully edited the contents.
I take this opportunity to thank Prof. B.K. Mathur and Prof. R.N.P. Chowdhury (the
earlier and the present) Head of the Department of Physics and Meteorology, Indian Institute
of Technology Kharagpur and Prof. S.K. Dube and Prof. D. Acharya (the earlier and the
present) Director, Indian Institute of Technology Kharagpur for their unstinted support and
encouragement in this book project. I took leave from the Indian Institute of Technology
Kharagpur and joined the Indian Institute of Space Science and Technology (IIST),
Thiruvananthapuram in July 2009 during the final stages of this book project. I acknowledge
with gratitude the help and encouragement I since received from Dr. B.N. Suresh, Director,
IIST towards this project.
I wish to place on record the excellent help and encouragement I received from the editors
of PHI Learning, the publishers of this book. In this connection I would like to specially
thank Mr. Surajit Sarkar, Ms. Babita Mishra, Mr. Darshan Kumar and Mr. K.K. Chaturvedi, all
of PHI Learning, who have been very supportive during the period of this project.
I have also received support and encouragement from well wishers, collaborators,
friends and colleagues, a list too long to mention here. However, I wish to acknowledge the
encouragement I received from Mr. G. Srinivas, Prof. P.C. Pandey, Dr. A. Gambheer,
Dr. Kiran Alapaty, Dr. D. Niyogi, Dr. K. Srinivasan, Dr. D. Lohar, Dr. S.P. Namboodiri,
Dr. Akio Kitoh and Dr. Panos G. Georgopoulos. I also wish to acknowledge the quiet support
I received from my family members and thank them for their forbearance. And finally, I seek
the blessings from ALMIGHTY and hope that the book will be well received.

A. Chandrasekar
1 Introductory Survey
of the Atmosphere

In recent times there has been a pronounced increase and appreciation of the importance of
the science of the atmosphere. The reasons for such increased interests in the earth’s
atmosphere are due to the increasing concern of man’s role in the emerging global warming
scenario as well as issues related to world food resources in the light of the increasing human
population. This chapter presents an introductory survey and overview of the study of the
earth’s atmosphere in eight sections. Section 1.1 introduces the various disciplines of
atmospheric science, and Section 1.2 summarizes the origin and composition of the earth’s
atmosphere. While Section 1.3 explains the basic elements of (vertical) distribution of
pressure and density, Section 1.4 outlines the components of the ionosphere, the earth’s
electric field, and the magnetosphere. Section 1.5 reviews the (vertical) distribution of
temperature and winds, while Section 1.6 outlines the behaviour of the atmosphere as a fluid.
Section 1.7 introduces the fundamental physical laws on which the science of the atmosphere
is based. Chapter 1 concludes with Section 1.8, which introduces the concepts of determinism
and chaos as these are understood in atmospheric science.

1.1 INTRODUCTION
The study of the atmospheric sciences is primarily devoted to the description and
understanding of phenomena in the earth’s atmosphere and to a lesser extent on that of the
other planets in the solar system. Atmospheric Sciences refer to the study of the physical,
chemical and dynamical aspects of the earth’s atmosphere, which extends upwards several
hundred kilometres from the earth’s surface. The term “atmospheric sciences” is usually used
in a broad sense and it includes atmospheric chemistry, aeronomy, magnetospheric physics,
and solar influences on the entire atmospheric system of the earth. The underlying postulate in
the study of the atmospheric sciences is that the atmospheric phenomena can be understood in
terms of the basic laws of physics. The physical laws of fluid dynamics, radiation and
thermodynamics are the most readily applicable to the study of atmospheric phenomena.
1
2 u BASICS OF ATMOSPHERIC SCIENCE

Meteorology can be defined as the scientific study of the earth’s atmosphere that
focuses on weather processes and forecasting. Meteorological phenomena are primarily
observable weather events, which are explained by the science of meteorology. Atmospheric
Sciences and Meteorology have been traditionally divided into the following three broad
disciplines: physical, synoptic and dynamic meteorology. Physical meteorology is devoted to
the study of atmospheric structure and composition, atmospheric optics, atmospheric
electricity as well as the physical processes involving radiation, cloud and rain formation.
Both, synoptic and dynamic meteorology deal with atmospheric motion and their evolution in
time. However, while the former employs empirical approaches to forecast large-scale
atmospheric motion, the latter utilizes approaches based on the physical laws of fluid
dynamics. In this book an effort has been made to provide an overview of the various facets
of the behaviour of the atmosphere in combination with an emphasis on the fundamental laws
of physics that aid in the understanding of the atmospheric behaviour.

1.2 ORIGIN AND COMPOSITION OF THE ATMOSPHERE


The earth’s atmosphere, in contrast to the sun’s atmosphere, is very much deficient in noble
gases such as helium, neon, argon, xenon and krypton. Atmospheric scientists generally agree
that during the early history of the earth the gaseous material (most probably hydrogen and
helium) which formed part of the original atmosphere of the earth was lost into space due to
(i) the earth’s weak gravity that was not strong enough to retain these lighter gases, and
(ii) the lack of differentiation between the earth’s solid inner and liquid outer core. The earth’s
first atmosphere apparently did not exist for a long period of time, and was followed by a
second atmosphere, which had formed due to volcanic out gassing. Volcanic eruptions and
their violent expulsion of volcanic substances from the earth’s interior gave rise to the second
atmosphere. The constituents of this atmosphere were quite similar to the gases emitted by
modern volcanoes, and were composed of 85% water vapour, 10% carbon dioxide, and a few
per cent nitrogen and sulphur together with sulphur compounds such as sulphur dioxide and
hydrogen sulphide, as well as very small amounts of carbon monoxide, hydrogen, chlorine,
ammonia and methane. It is to be noted that no evidence exists of free oxygen being present
in these volcanic emissions, and consequently free oxygen was absent in this second
atmosphere. Since the atmosphere is capable of holding only a very small amount of the mass
of water vapour released through volcanic eruptions, it can be easily surmised that a large part
of the water vapour present in the second atmosphere must have condensed as clouds. This
would have led to torrential rains and the formation of large water bodies on the earth’s
surface. The formation of the hydrosphere, however, does not explain the presence of oxygen
in our present atmosphere.
Two possible sources of atmospheric oxygen are associated with the absorption of solar
radiation: (i) the photodissociation of water due to the absorption of ultraviolet radiation, and
(ii) the action of plants through photosynthetic reaction due to the absorption of visible
radiation. Oxygen formation using the photodissociation process is less likely to be the chief
cause of oxygen formation as studies indicate that not more than 1–2% of the oxygen
presently seen in our atmosphere could have been produced using the photodissociation
process. However, it is generally agreed that the photosynthesis reaction (CO2 + H2O +
INTRODUCTORY SURVEY OF THE ATMOSPHERE u 3

visible sunlight = organic compounds + O2) by cyanobacteria and eventually higher plants is
the main source of production of significant amounts of the present oxygen. While there were
two sources of oxygen production in the atmosphere, three oxygen sinks have been identified
which occurred at different stages of the earth’s evolution. The first sink of oxygen, which
happened in the early stages of the earth’s atmospheric evolution, is attributed to chemical
weathering through the oxidation of surface materials. Both animal respirations as well as the
burning of fossil fuels, the latter happening in very recent times, are the other additional sinks
of oxygen. If the atmospheric oxygen did form due to photosynthesis reaction, then the
possibility exists for plant life to form in an oxygen-free atmosphere. Scientists believe that in
the initial stages, single-celled organisms existed that did not require oxygen. Subsequently
primitive forms of plant life formed, which released oxygen through photosynthesis. Since
nitrogen is chemically inert and has low solubility in water, most of the nitrogen released
during the early volcanic eruptions has managed to remain in the atmosphere. Due to the
nearly complete removal of water vapour and carbon dioxide associated with the process of
condensation and photosynthesis, nitrogen has become the dominant constituent of the earth’s
atmosphere. Quite striking is the fact that the atmospheres of the earths nearest neighbours,
Venus and Mars, are entirely different from that of the earth. In contrast to the earth’s
nitrogen–oxygen dominated atmosphere, the atmospheres of both Venus and Mars are
composed primarily of carbon dioxide. Also, the atmosphere of Venus is one hundred times
more massive than the earth, and the atmosphere of Mars is one hundred times less massive
than the Earth’s atmosphere. The differences between the atmospheric histories of these
planets are indeed intriguing because, despite sharing a common birth, the atmospheres of all
these planets have evolved along very different paths.
Table 1.1 presents the composition of the earth’s atmosphere in the well-mixed region
up to a height of 100 km. Nitrogen contributes up to 78% by volume, while oxygen and
argon contribute 21% and 0.93% by volume. The earth’s atmosphere also contains a variable
amount of water vapour (accounting for a maximum of 4%) as well as very small amounts of
carbon dioxide and ozone. Due to the effective mixing associated with the turbulent fluid

TABLE 1.1 Composition of earth’s atmosphere in the homosphere


(ppmv is parts per million by volume)

Gas Volume mixing ratio Molecular weight


Nitrogen (N2) 0.78 28.02
Oxygen (O2) 0.21 32.0
Argon (Ar) 0.0093 39.95
Water vapour (H2O) < 0.04 18.02
Carbon dioxide (CO2) 360 ppmv 44.01
Neon (Ne) 18 ppmv 20.18
Ozone (O3) 12 ppmv 48.0
Helium (He) 5 ppmv 4.0
Krypton (Kr) 1 ppmv 83.7
Hydrogen (H2) 0.5 ppmv 2.02
4 u BASICS OF ATMOSPHERIC SCIENCE

motion of the atmosphere up to a height of 100 km, the atmosphere has a uniform
composition; this atmospheric layer is known as the homosphere. Above 100 km, the mean
free path is no longer small, and the process of molecular diffusion by random molecular
motion dominates more than the macroscopic turbulent mixing of air parcels, therefore,
random molecular motion determines the composition of the earth’s atmosphere. The region
of the atmosphere above 100 km is known as the heterosphere and is characterized by a
gradual decrease with the height of the mean molecular weight of the mixture of the gases.
The earth’s atmosphere above 120 km is predominantly atomic oxygen; at heights above
100 km, helium and hydrogen dominate. Unlike the other constituents, both water vapour and
ozone are known to vary widely both in space and time even within the homosphere.

1.3 DISTRIBUTION OF PRESSURE AND DENSITY


Atmospheric pressure at a given point is defined as the height of the overlying atmosphere of
a column of unit cross-sectional area around that point extending to the top of the atmosphere.
The above definition of pressure implies a decrease of pressure with increase in height, and
hence it is not surprising that pressure is observed to decrease with height. Also, the vertical
variation of pressure with height is large in comparison with its horizontal and temporal
variations. Hence, for the sake of convenience, a standard atmosphere is defined; representing
the horizontal and time-averaged structure of the atmosphere as a function of height only.
Table 1.2 presents the details of pressure, temperature, and density variation based on the US
Standard Atmosphere.
TABLE 1.2 US Standard Atmosphere

Height (km) Temperature (°C) Pressure (hPa) Density (kg m–3)


0 15.0 1013.15 1.225
1 8.5 900.0 1.1
2 2.0 800.0 1.0
3 –4.5 700.0 0.91
4 –11.0 620.0 0.82
5 –17.5 540.0 0.74
6 –24.0 470.0 0.66
7 –30.5 410.0 0.59
8 –37.0 360.0 0.53
9 –43.5 310.0 0.47
10 –50.0 260.0 0.41
11 –56.5 230.0 0.36
12 –56.5 190.0 0.31
13 –56.5 170.0 0.27
14 –56.5 140.0 0.23
15 –56.5 120.0 0.19
(Contd.)
INTRODUCTORY SURVEY OF THE ATMOSPHERE u 5

TABLE 1.2 U.S. Standard Atmosphere (Contd.)

Height (km) Temperature (°C) Pressure (hPa) Density (kg m–3)


16 –56.5 100.0 0.17
17 –56.5 90.0 0.14
18 –56.5 75.0 0.12
19 –56.5 65.0 0.10
20 –56.5 55.0 0.088
21 –55.5 47.0 0.075
22 –54.5 40.0 0.064
23 –53.5 34.0 0.054
24 –52.5 29.0 0.046
25 –51.5 25.0 0.039
26 –50.5 22.0 0.034
27 –49.5 18.0 0.029
28 –48.5 16.0 0.025
29 –47.5 14.0 0.021
30 –46.5 12.0 0.018
31 –45.5 10.0 0.015
32 –44.5 8.7 0.013
33 –41.7 7.5 0.011
34 –38.9 6.5 0.0096
35 –36.1 5.6 0.0082
40 –22.75 2.87 0.0039
46 –6.25 1.31 0.0017
50 –2.55 0.79 0.0010
56 –15.15 0.37 0.0005
60 –26.15 0.22 0.0003
66 –42.65 0.09 0.0001
70 –53.55 0.05 0.00008
76 –66.75 0.02 0.00003
80 –74.55 0.01 0.000018
86 –86.25 0.003 0.000007

In an isothermal atmosphere, both the pressure and density decrease exponentially with
height. In the real atmosphere, up to a height of 100 km, the logarithm of pressure is nearly
linear with height. The atmospheric pressure averaged over the surface of the earth at mean
sea level has a value of 1.0132 × 105 Pa or 1013.2 hPa. The averaged value of density of the
air at mean sea level has a value of 1.225 kg m–3. The vertical distribution of air density in
the earth’s atmosphere closely follows the vertical distribution of pressure. As is commonly
known, density depends on both pressure and temperature. Consequently, if density variations
with height closely follow pressure variations with height, the variations of air temperature in
6 u BASICS OF ATMOSPHERIC SCIENCE

the vertical are logically much less smaller than pressure variation in the vertical. Also, the
horizontal and temporal variation of the air density is much smaller compared to the vertical
variation.

1.4 IONOSPHERE, ATMOSPHERIC ELECTRIC FIELD AND


MAGNETOSPHERE
On December 12, 1901, Marconi successfully demonstrated trans-Atlantic communication by
receiving a radio signal in St. John’s Newfoundland that had been sent from Cornwall,
England. A year later, Oliver Heaviside and Arthur Kennelly independently proposed the
existence of a conducting layer—the ionosphere—in the upper atmosphere that would allow
an electromagnetic signal to be reflected back to the earth. At heights of 80 km or above, the
atmospheric density is so low that free electrons can exist for short periods of time before
they are captured by a nearby positive ion. The existence of charged particles at this altitude
(80 km and above) indicates the beginning of the ionosphere, a region having both the
properties of a gas and of plasma. Solar radiations at ultraviolet (UV) and shorter wavelengths
(X-rays) are considered to be ionizing because photons of energy at these frequencies are
capable of dislodging an electron from a neutral gas atom or molecule during a collision. In
the above process, known as photoionization, the interaction of electromagnetic radiation such
as solar UV and X-ray radiation with matter results in the dissociation of that matter into
electrically charged particles made up of a free electron and a positively charged ion. Cosmic
rays and solar wind particles are also known to play a role in the photoionization process, but
their effect is minor compared with the effect of the sun’s electromagnetic radiation on photo-
ionization. At the highest levels of the earth’s outer atmosphere, (greater than 300–400 km),
solar radiation is very strong, but only a few atoms exist with which to interact, so the amount
of ionization is small. As the altitude decreases, more gas atoms are present, so the
photoionization process increases. However, at the same moment an opposing process known
as recombination begins, wherein a free electron is “captured” by a positive ion if it moves
close enough to it. Since the density of the atmosphere increases at lower altitudes, the
recombination process becomes dominant, since the ions are at this lower altitude relatively
closer to each other. Ultimately, the balance between these two conflicting processes
determines the degree of ionization present at any given time. At still lower altitudes (80 km
or lower), the number of gas atoms (and molecules) increases further, and more opportunity
exists for absorption of energy from a photon of UV or X-ray solar radiation. However, the
intensity of the above-mentioned solar radiations is less at these lower altitudes because some
of UV and X-ray solar radiation have been absorbed at the higher altitude levels. Finally a
situation is reached where the lower radiation intensity, greater gas density, and greater
recombination rates balance out, and the ionization rate begins to decrease with decreasing
altitude. This leads to the formation of several distinct ionization peaks or layers, such as the
D, E, F1, and F2 layers centred approximately at 80 km, 105 km, 175 km and 250 km,
respectively. The ionosphere is still used by international broadcasters to reflect radio signals
back to the earth, so that programs can be heard around the entire world. The ionosphere
provides long-range capabilities for commercial ship-to-shore communications, trans-oceanic
aircraft links, and military communication and surveillance systems. Also, all signals
INTRODUCTORY SURVEY OF THE ATMOSPHERE u 7

transmitted to and from satellites for communication and navigation purposes must pass
through the ionosphere; hence ionospheric irregularities (disturbances) can have a major
impact on system performance and the reliability of satellite equipment. Since the ionosphere
is almost entirely made up of plasma, also known as the fourth state of matter, research on
the ionosphere can increase our understanding of plasma.
The region of the atmosphere—extending upwards from a few tens of kilometres to the
top of the ionosphere—is characterized by very large electrical conductivity at a constant
electric potential and is known as the electrosphere. Under conditions of fair weather,
a downwardly directed atmospheric electric field exists with an averaged magnitude of
120 V m–1 near the earth’s surface. The downwardly directed electric field, under fair weather
conditions, implies that the electrosphere carries a net positive charge and an average
potential of 300 kV with respect to the earth’s surface. To maintain the above voltage, the
earth has a negative charge of about a million coulombs on its surface and an equal net
positive charge is distributed throughout the atmosphere. Careful measurements have indicated
that the earth’s negative charge remains roughly constant over time. Considering the
magnitude of the leakage current flowing in the air, which amounts to 2 to 4 × 10–12 A m–2,
the electrosphere together with the earth both constitute a spherical capacitor, and should get
discharged in a matter of minutes. However, the fact the fair weather electric field is by and
large constant suggests the existence of electrical generators, which maintain the fair weather
electric field. Atmospheric scientists generally agree that thunderstorms serve as electrical
generators and maintain the fair weather field by separating the electric charges with positive
and negative charges concentrated at the top and the base of the thunderstorm cloud. While
the positive charges found in the upper regions of the thunderstorm cloud get leaked to the
electrosphere, lightning flashes ensure that negative charges are transported to the ground
from the base of the thunderstorm cloud. The point discharge current and the precipitation
current account for the remaining two components of the global electrical circuit. Point
discharge currents transport positive charges from pointed obstacles on the earth’s surface
upward through the air beneath and above a thunderstorm. Precipitation currents bring
positive charges to the earth’s surface with thunderstorm precipitation. The approximate values
of the various components of the electrical budget for the earth in units of C km–2 yr–1 are:
90 units of positive charge gained during the fair weather conditions, 100 units of positive
charge lost due to point discharge current, 30 units of positive charge gained due to
precipitation current and 20 units of negative charge gained due to the lightning discharge.
At very high levels of 500 km and more, the motion of the charged particles is very
much influenced by the presence of the earth’s magnetic field; hence this region of the
atmosphere at and above 500 km is known as the magnetosphere. In the magnetosphere, very
little interaction occurs between the charged particles and the neutral atoms/molecules due to
the infrequent collisions between them. The sun’s extremely hot atmosphere consists of
nothing but plasma—a gas consisting of charged particles—mostly electrons and protons.
Solar plasma streams radially into space at high speed and pulls the sun’s magnetic field
along with it. The electrified particles, and the solar magnetic field that they pull along, is
called the solar wind. These solar wind particles come streaming towards the earth at very
high velocities at 450 km s–1 or more and take about 2–3 days to reach the earth. The solar
wind particles flowing directly from the sun towards the earth come across the
8 u BASICS OF ATMOSPHERIC SCIENCE

magnetosphere, which acts as an obstacle, which the solar wind particles must go around.
Due to their high speed, however, they have no time for an orderly detour. Instead, their
direction is changed abruptly in the so-called bow shock region just outside the sunward
magnetic field. The abrupt passage of the solar wind particles through the bow shock region
reduces the speed and changes the motion of the particles. Most of these solar wind particles
are deflected around the magnetosphere through an area known as the magnetosheath. Acting
as a shield, the magnetosphere effectively screens the earth from most of the direct solar wind
particles. These charged solar wind particles do not travel readily across the magnetosphere,
but are deflected at angles to the magnetic field. Some of the solar wind particles, however,
can travel along the earth’s magnetic field lines, and leak through the earth’s magnetic screen.
These particles cause disturbances within the earth’s magnetosphere and are associated with
the structure of the interplanetary magnetic field, which rotates with the sun.

1.5 DISTRIBUTION OF TEMPERATURE AND WINDS


The earth’s atmosphere is mostly transparent to the incoming solar radiation, and hence the
above radiation from the sun is absorbed by the earth surface. The earth’s atmosphere in turn
absorbs the terrestrial long wave radiation from the earth surface resulting in a situation,
where the atmosphere is primarily heated, from below. This results in a vertical distribution of
air temperature, wherein the air temperature decreases with height. In addition to the vertical
distribution of air temperature, there also exists horizontal distribution of air temperature, due
to the differential heating of the incoming solar radiation between the low and high latitudes.
Typically, the air temperature decreases poleward in the troposphere. In addition to giving rise
to the horizontal distribution of air temperature, the differential heating between the low and
high latitudes also results in atmospheric motions of various scales. The thermally direct
circulations associated with the differential heating also give rise to vertical distribution of winds.

1.5.1 Distribution of Temperature


Vertical Distribution of Temperature
Figure 1.1 depicts the vertical distribution of air temperature for the standard atmosphere, for
typical mid-latitude conditions. Four distinct layers can be identified in the figure and these
are known as troposphere, stratosphere, mesosphere and thermosphere, respectively. The
boundaries (transitions) between the different layers are known as tropopause, stratopause,
and mesopause, respectively. The heights of these boundaries vary in both time and space,
primarily with latitude and season. The observed, globally averaged surface temperature for
the earth is about 15°C. Incoming solar radiation entering the earth’s atmosphere has a short
wavelength with a maximum energy centred around the wavelength range of 0.2 mm to 2 mm.
The radiation emitted by the earth and its atmosphere have longer wavelengths, known as
infrared radiation, with its maximum energy centred on > 4 mm. The earth’s atmosphere is
mostly transparent to the solar radiation in the wavelength range of 0.2 mm to 2 mm, but absorbs
infrared radiation due to the properties of water vapour, carbon dioxide, methane, and other
trace gases. While the incoming solar radiation and outgoing infrared radiation tend to remain
nearly in balance, the net effect of the earth’s opacity to infrared radiation and its near
INTRODUCTORY SURVEY OF THE ATMOSPHERE u 9

100

Thermosphere

90

Mesopause

80

70

60
Height (km)

50 Stratopause

40

Stratosphere
30

20

Tropopause

10 Troposphere

180 190 200 210 220 230 240 250 260 270 280 290 300 310
Temperature (K)

FIGURE 1.1 Vertical temperature profile in the atmosphere.


transparency to the solar radiation results in warming the surface of the earth. This effect is
known as the greenhouse effect and is responsible for warming the surface of the earth by
about 35°C (i.e. from –20°C to 15°C). The differing absorption responses of the earth’s
10 u BASICS OF ATMOSPHERIC SCIENCE

atmosphere with respect to the incoming solar and the outgoing infrared radiation result in a
general decrease of temperature with increase in height.
The atmosphere is a fluid, and in a situation, where this fluid is heated from below at
the earth’s surface, one expects convection to manifest. The first 10–15 km layer above sea
level, known as the troposphere (literally, the turning sphere), is characterized by convection
and is associated with strong vertical flows of air. The troposphere has nearly 90% of the
entire atmospheric mass and has almost all of the water vapour, clouds, and rainfall found in
the earth’s atmosphere. In short, the troposphere is that layer of the atmosphere where the
weather occurs. In the troposphere, air temperature decreases sharply and continuously with
height, with an average drop of temperature being 6–6.5°C km–1. The height of the
tropopause is about 9 km in high latitudes (polar regions), 12 km in midlatitudes (extra-
tropical regions) and 16 km in low latitudes (tropical regions). The temperature at the
tropopause is nearly –55°C in mid to high latitudes, and –75°C in the tropics.
The stratosphere (literally, the layered sphere), unlike the troposphere, is characterized
by very little vertical motion and mixing. The sharp decrease in water vapour amounts and
sharp increase in the ozone content across the transition between the troposphere and the
stratosphere (tropopause) make this clearly evident. In the lower stratosphere, temperature
remains nearly constant (isothermal) with height, in the atmospheric region extending from
the tropopause up to approximately 20 km above sea level. Above this altitude, the
temperature increases with height by about 2°C km–1 to the upper stratosphere. The
temperature of the stratopause at a height of 50 km is about 0°C. The increase of temperature
with height in the stratosphere is not conducive for convection and its associated vertical
motion and is contrary to the general overall picture presented earlier. This unusual behaviour,
i.e. temperature increase with height in the stratosphere, results from the presence of
atmospheric ozone, which absorbs ultraviolet radiation from the sun in the wavelength range
of 0.2 mm to 0.31 mm. The ozone’s absorption of UV radiation is responsible for
the temperature maxima observed at the stratopause level. The atmospheric pressure at the
stratopause level is about 1 hPa and consequently, the troposphere in combination with
the stratosphere, account for about 99.9% of the entire atmospheric mass.
Above the stratopause, in the mesosphere (literally, the middle sphere), the general
overall picture presented earlier is restored as far as the temperature change with height is
concerned. The temperature decreases with height in the mesosphere until the mesopause
is reached. The mesopause is at a height of 80 km and has a temperature minimum of about
–90°C. Within the mesosphere, the gain of energy due to absorption of short wave solar
radiation is less than the energy lost to space in the infrared part of the electromagnetic
spectrum. This deficit in the radiational energy balance is compensated by the convection of
heat upward from the stratopause level. As in the troposphere, vertical air motions are not
inhibited in the mesosphere, and during twilight hours over the polar regions, mesospheric
clouds known as noctilucent clouds are visible from the ground. The mesosphere together
with stratosphere is also referred to as the middle atmosphere.
The thermosphere (literally, the heating sphere) extends upwards from 80 km to altitudes
of several hundred kilometres with temperatures ranging from 500 K to as much as 2000 K.
The thermosphere is among the first to get exposed to the solar radiation among the four
layers previously described. The few molecules that are present in the thermosphere receive
INTRODUCTORY SURVEY OF THE ATMOSPHERE u 11

very large amounts of energy from the sun, and this causes the layer to warm to very high
temperatures. The temperature in the thermosphere is critically dependent on solar activity.
Due to the very low density of the air, thermospheric air temperature cannot be measured
directly. Instead, the density of the air in the thermosphere is measured by calculating the drag
the air exerts on satellites, and this calculation is used to determine thermosphere temperature.
The thermosphere is also referred to as the upper atmosphere.

Horizontal Distribution of Temperature


Apart from the vertical variation of temperature with height, pronounced variability of
temperature with latitude and seasons also exists. Furthermore, major changes in the
horizontal distribution of air temperature manifest due to the presence of fronts, a narrow
boundary zone which separates warm air and cold air. Such fronts are common only over the
midlatitudes and higher latitudes. The temperature within the troposphere decreases with
increase of latitude, with the largest latitudinal temperature gradients associated with the
winter hemisphere. Poleward of 30°N and 30°S, the temperature gradients are quite steep as
compared to temperature gradients at the lower latitudes in the troposphere. As mentioned
earlier, the tropical tropopause is very much higher (16 km above sea level) and, therefore, is
considerably colder, than the tropopause at high latitudes (8 km–9 km). A pole-to-pole
gradient; the summer pole having the maximum temperatures, and the winter pole having
the minimum temperatures, characterize the latitudinal temperature distribution in the strato-
sphere. The nearly six months of continuous sunlight over the summer pole heats up the
stratosphere at the summer pole more than anywhere else, and the lack of sunlight during
winter over the winter pole cools the stratosphere at the winter pole more than anywhere else.
The above atmospheric behaviour results in a monotonic temperature gradient between
the warm summer pole to the cold winter pole in the stratosphere and stratopause. Since the
summer and winter hemispheres alternate every six months, the temperature gradients in
the stratosphere get reversed twice in a year. The latitudinal temperature distribution in the
mesosphere, like the stratosphere is again characterized by a pole-to-pole gradient, but
opposite to that observed in the stratosphere. Hence, in the mesosphere, maximum
temperatures are observed at the winter pole and minimum temperatures over the summer
pole with the summer poles at the mesopause region having the coldest temperature. Similar
to the stratosphere, temperature gradients in the mesosphere also reverse direction twice a
year. In addition to the above-mentioned latitudinal distribution temperature changes do occur
along a latitude circle, especially in the troposphere due to the presence of land-sea contrasts
and the existence of large-scale orography.

1.5.2 Distribution of Winds


Winds can be conveniently defined as the horizontal movement of air and occur at various
scales ranging from the local breeze arising out of local heating to global winds due to
differential global scale heating. The trade winds are observed at low levels within the low-
latitude region (within 30° latitude N/S) and are northeasterlies in the northern hemisphere
(winds from the north-easterly direction) and southeasterlies in the southern hemisphere
(winds from the south-easterly direction). Surface easterlies are seen over the polar regions,
12 u BASICS OF ATMOSPHERIC SCIENCE

and in the lower troposphere, while westerlies dominate in the mid-latitude region of both the
hemispheres. The above distribution of surface winds (easterlies in the low latitudes and polar
regions and westerlies in the mid latitudes) in both the hemispheres is consistent with a tri-
cell meridional structure of the atmospheric circulation. The three cells which comprise the
above tricell structure are known as Hadley cell, Ferrel cell and the Polar cell. The Hadley
cell is manifested as the ascending motion over equator and descending motion over the
subtropics (30°N/30°S), while the Ferrel cell has ascending motion over 60°N/60°S and
descending motion over the subtropics. The Polar cell has ascending motion over 60°N/60°S
and descending motion over the poles. The ascending branches of the Hadley cell are
associated with the convergence of the northeasterlies from the northern hemisphere and
southeasterlies from the southern hemisphere in a broad convergence zone called the
Intertropical Convergent Zone (ITCZ). The increase of westerlies (or weakening of the
easterlies) with height are related to the north-south temperature gradient. Whenever,
temperatures decrease, towards the poles from the equator (as is the usual case in the
troposphere), surface westerlies strengthen with height, and surface easterlies weaken with
height. Since surface westerlies are seen in mid-latitudes, and the north-south temperature
gradient contributes to the strengthening of the westerlies with height, as might be anticipated,
very strong westerly winds (known as subtropical jet streams) develop in the upper
troposphere around 45°N/S.
The above jet stream coincides with the location of the maximum north-south
temperature gradients. Since the north-south temperature gradient is stronger in the winter
hemisphere, the subtropical jet stream over the winter hemisphere is stronger than that seen in
the summer hemisphere. The winter polar regions are very cold in the stratosphere, and hence
the north-south temperature gradients near the winter polar regions are very high. This
situation contributes to a westerly jet stream in the lower and middle stratosphere over the
winter polar regions known as the polar night jet. The zonal (east-west) winds over the
summer stratosphere are characterized by very strong easterly winds (the temperature gradient
here is opposite to that in the troposphere), while the zonal winds over the winter stratosphere
are characterized by very strong westerlies. Again, the stratospheric westerlies seen in the
winter hemisphere are stronger than the stratospheric easterlies seen in the summer
hemisphere.
In addition to stronger north-south temperature gradients in the winter hemisphere,
tropospheric westerlies in the winter hemisphere strengthen with increase of height rising right
up to the stratosphere, while, the tropospheric westerlies in the summer hemisphere weaken
with increase of height into the stratosphere. Since the temperature gradients in the
stratosphere reverse direction twice a year, the associated zonal winds also alternate between
easterlies and westerlies twice a year. The zonal winds over the mesosphere can also be
understood from considerations of the north-south temperature gradient. Since the temperature
gradients in the mesosphere are exactly opposite to those seen in the stratosphere, the
easterlies and westerlies seen in the summer and winter hemispheres in the stratosphere
weaken with height, giving rise to strong westerlies in the mesosphere during summer, and
easterlies in the mesosphere during winter.
INTRODUCTORY SURVEY OF THE ATMOSPHERE u 13

1.6 ATMOSPHERE AS A FLUID AND FLUID CONTINUUM


The earth’s atmosphere consists of a mixture of ideal gases mainly comprised of nitrogen and
oxygen, and small amounts of carbon dioxide, ozone and water vapour. The earth’s
atmosphere, together with the oceans, provides one of the grandest manifestations of fluids in
nature. Fluids are defined as, material continuum that is unable to withstand static shear
stress. In other words, fluids yield to shear stress. The effect of applied force is referred to as
stress and the resulting deformation is termed as strain. Shear stress is simply the force it
takes for one layer of the material (fluid) to slip over another. The above behaviour is at
variance with a solid, which responds to a shear stress with a recoverable deformation. Within
the elastic limit, while stress is proportional to strain for a solid; shear stress is proportional to
rate of shear for a (Newtonian) fluid. A Newtonian fluid is a fluid in which the shear stress is
linearly proportional to the velocity gradient in the direction perpendicular to the plane of
shear, with the coefficient of viscosity being the constant of proportionality.
The dynamics of fluids comprised of gases and liquids have been the subject of study
for over two centuries. In order to study fluid behaviour, atmospheric scientists use a model
known as the continuum model, in which the existence of molecules is ignored, and the fluid
is assumed to be a continuous medium. Essentially, the fluid is assumed to be made up of
particles, called fluid particles, each of which contain a large number of molecules.
This continuum model enables atmospheric scientists to ascribe macroscopic properties
such as temperature and density, etc. to “fluid particles” and to ignore the motion of the
individual molecules. The applicability of the continuum model to fluid problems is restricted
to situations where the characteristic length scales of the fluid problem are very much larger
than the mean free path, i.e. larger than the average distance traversed by a molecule between
two successive collisions in the fluid medium. In this book, the atmosphere will be assumed
to satisfy the requirement of a fluid continuum

1.7 PHYSICAL LAWS


As mentioned earlier, the physical laws of thermodynamics, radiation and fluid dynamics find
ready applications in the study of the atmosphere. The principles of thermodynamics are
essential to the description of various atmospheric processes. The equation of state of the
atmosphere is deduced using the principle of thermodynamics. The first law of ther-
modynamics, a statement of conservation of energy becomes important when considering
situations involving the heating and cooling of the atmosphere. The second law of
thermodynamics specifies the direction of heat flow and introduces “entropy”, a concept that
is important in dealing with the efficiency of the atmospheric heat engine. Also, the
principles of thermodynamics play an important role in understanding processes leading to a
change of phase.
Atmospheric forcing and its associated motion are primarily due to solar radiation. A
small part of the solar radiation coming into the earth’s atmosphere is scattered back to space
by atmospheric gases and is also reflected back-to-space by clouds and the earth’s surface; the
remaining amount of radiation is absorbed by water vapour and ozone in the atmosphere and
by the earth’s surface. Beer’s law provides the relationship of the fractional absorption of
14 u BASICS OF ATMOSPHERIC SCIENCE

radiation in terms of the density of the medium as well as the distance traversed. Atmospheric
gases, as well as clouds and the earth’s surface, emit and absorb long-wave infrared radiation;
the functional dependence of the rate of emission of black body radiation on the temperature
and the wavelength is given by Planck’s law. The above-mentioned radiative processes of
absorption and re-emission between the different layers of the atmosphere can also be
understood by invoking the equation of radiative transfer.
The physical laws of fluid dynamics as applied to atmospheric sciences include the
Newton’s laws of motion as applied to rotating coordinates and the statement of conservation
of mass. Newton’s second law of motion is a statement of the conservation of linear
momentum, and the equation of continuity is the statement of mass conservation. For large-
scale atmospheric motions, the concept of hydrostatic equilibrium, a balance between the
vertical pressure gradient force and gravity, finds wide applicability. (The pressure gradient
force is the three-dimensional force vector that is usually responsible for accelerating air
parcels away from regions of high atmospheric pressure towards regions of low atmospheric
pressure.) For an isothermal atmosphere, the combination of an ideal gas law together with
the equation characterizing hydrostatic equilibrium gives rise to conditions in which the air
density and pressure decrease exponentially with height. Even for the real atmosphere, the air
density decreases with height, suggesting the existence of a continuously stratified
atmosphere. Newton’s second law of motion, which characterizes the conservation of linear
momentum when deduced in rotating coordinates, gives rise to additional apparent forces such
as the Coriolis force. (Coriolis force is a force exerted on any moving air parcel due to the
rotation of the earth.) The horizontal frictionless flow, parallel to the isolines of pressure
known as geostrophic flow, results from a balance between the Coriolis force and the pressure
gradient force. The gradient flow resulting from the balance between Coriolis, centrifugal and
pressure gradient forces manifests as a counterclockwise circulation around lows in the
northern hemisphere; in the southern hemisphere this circulation around low-pressure systems
is clockwise.

1.8 DETERMINISM AND CHAOS


Since the late eighteenth century, eminent philosophers and mathematicians have claimed that
an accurate knowledge of the present state of any system together with a complete
understanding of all the laws governing the system must necessarily lead to perfect and
complete knowledge of the system for all future times. The above hypothesis (principle) of
determinism had complete and total support from all scientists and researchers for a century
and half until the beginning of the twentieth century. In the 1920s with the advent of quantum
mechanics and the Heisenberg uncertainty principle, the principle of determinism was given a
body blow as far as atomic scale behaviour was concerned. Even such a celebrated theoretical
physicist as Albert Einstein was not enamoured with the apparent failure of the above
principle of determinism and expressed his dismay with the famous comment, God does not
play dice.
Despite scientists’ reluctant acceptance of the failure of the principle of determinism as
manifested in atomic scale phenomena, the scientific community widely believed that large-
scale systems were forecastable—in other words, the future behaviour of a system was
INTRODUCTORY SURVEY OF THE ATMOSPHERE u 15

completely predictable. With the development of digital computers in the middle of the
twentieth century, however, many researchers started utilizing numerical methods to solve
physical problems which did not possess an analytically closed-form solution. Lorenz, a
meteorologist working at MIT in the USA in the 1960s, accidentally found that the crude
computer model which he was working on to simulate the atmosphere showed considerable
sensitivity to very small differences in the initial conditions.
Lorenz felt that this property of extreme sensitivity to initial conditions as manifested in
his computer model was indeed a property of the real atmosphere and coined the name chaos
to describe the behaviour of systems that are sensitive to initial conditions. Lorenz suggested
that this extreme sensitivity to initial conditions was due to nonlinear interactions within parts
of the system. It is important to note that chaotic behaviour is not random behaviour. What
chaos essentially means to the real atmosphere is that the present state of the atmosphere is
still related to all the future states of the atmosphere through the governing laws of
atmospheric motion. However, due to the extreme sensitivity of the initial conditions, any
inaccuracies or uncertainties in the initial observations of the atmosphere can grow rapidly
with the evolution of the system with time until the errors reach a magnitude that can swamp
the solution. That is, any chaotic system has limited predictability.
This new conclusion has led to the idea of a limit in the deterministic predictability of
the state of the atmosphere. Studies have indicated that this deterministic limit is
approximately two weeks. While highly-specific forecasts from numerical weather prediction
models have a limited predictability of up to two weeks, current research indicates that less-
specific seasonal forecasts in the tropical regions may be achievable as the tropical climate is
primarily driven by slowly varying anomalous boundary forcing such as sea surface
temperature.

REVIEW QUESTIONS

1. If one were given the vertical temperature profile of the atmosphere at a place up to an
elevation near 20 km, how would one be able to locate the tropopause?
2. If there were no ozone layer, how would the vertical distribution of temperature look like,
say at the lower latitudes from 0 to 120 km.
3. What are homosphere and heterosphere?
4. How does the density of the air vary with height?
5. Mention the percentage of the total mass contained in the troposphere and stratosphere.
6. Despite having a smaller vertical extent, why is it that the troposphere contains more mass
than the stratosphere?
7. Is it possible to correctly define the absolute top of the atmosphere?
8. Name the variable gases within the homosphere.
9. Why has the concentration of carbon dioxide in the atmosphere been increasing since the
20th century?
10. Mention the beneficial and the harmful effects of atmospheric ozone.
11. Are the aerosols produced only due to human activities or they occur naturally?
2 Atmospheric
Observations

Observations or measurements of any system such as the atmosphere provide important and
valuable information of the system. Such information of the atmosphere through
meteorological observations is then used in the development of theories which explain the
observed atmospheric behaviour. Also, the information of the atmosphere as obtained through
meteorological observation further our understanding of the processes involved in the science
of the atmosphere. There are undoubtedly various challenges confronting “weather prediction”
as a physical and mathematical problem. One of the fundamental requirements for weather
prediction is to accurately prescribe the initial state of the atmosphere. “Observations” of the
various atmospheric and oceanic variables through conventional as well as otherwise, i.e. non-
conventional platforms provide an important component in prescribing the initial state of the
atmosphere. As the oceans account for as much as 70% of the entire surface area of our planet
earth and they are mostly out of bounds for our conventional meteorological (and oceanic)
observations, weather satellites launched from 1960s onwards have provided very important
information of the atmosphere over these data sparse regions.
The chapter presents a very brief account of the various types of meteorological
observations including conventional as well as non-conventional observations. Conventional
meteorological observations can be broadly divided into surface and upper air measurements.
The chapter begins with Section 2.1 on the overview of meteorological observations and their
importance. Observations in addition to their role in providing the initial conditions for a
numerical weather prediction model are also very useful as they help us in gaining
understanding of the various physical processes, which occur in the atmosphere. Temperature,
humidity, wind speed and direction, pressure and precipitation are some of the most important
meteorological variables that are routinely observed in the atmosphere. While Section 2.2,
provides for the measurement of atmospheric temperature and humidity, the next Section 2.3
outlines the measurement of wind and pressure, and Section 2.4 gives the measurement of
precipitation. Section 2.5 discusses briefly modern meteorological instruments. It is obvious
that the observation from a single surface or a single upper air meteorological station at a
16
ATMOSPHERIC OBSERVATIONS u 17

fixed time cannot completely capture the various meteorological systems and associated
weather. This requires simultaneous observations from several surface/upper air mete-
orological stations at the same universal time. The above measurements from several
meteorological stations at the same universal time can reveal important patterns/structures as
well as characteristics of the size and movements of the various meteorological systems. The
above brings out the need for a surface and an upper air meteorological network of stations,
which are conveniently dealt in Section 2.6. The chapter concludes with Section 2.7 on
satellite observations.

2.1 OVERVIEW AND IMPORTANCE OF METEOROLOGICAL


OBSERVATION
Meteorological observations are the main source of providing meaningful information of the
physical and chemical aspects of the atmosphere. In addition to furthering our understanding
of the atmospheric processes, the meteorological observations are also used in the
development of theories which ultimately explain the observed behaviour of the atmosphere.
Also, our understanding of the atmosphere through meteorological observations has helped in
gaining greater insights of the various aspects of the weather prediction problem. The state of
the atmosphere at any instant and at any place is completely determined by the values
assumed by the meteorological variables such as temperature, pressure, humidity, wind speed,
wind direction, etc. at that time, and at that place. Hence knowledge of the state of
atmosphere requires measurements of all the meteorological variables. This makes it
mandatory for us to understand the different methods by which the various meteorological
variables are measured. The advancements of the understanding of any branch of science are
indeed intimately related to careful measurements and subsequent detailed analysis of the
measurements. Unlike, other branches of science, the science of the atmosphere is even more
dependent on careful meteorological observations for the following reason. In most situations,
an atmospheric phenomenon or a meteorological system over a region may not be identified
by a single observation over that region. However, if several meteorological stations measure
the atmospheric variables at the same universal time, and if one were to carefully plot and
analyze the measurements, a spatial pattern of the atmospheric phenomenon or a
meteorological system can be easily discerned. In addition, if the same meteorological stations
measure the atmospheric variables at the same (selected) intervals of time, and carefully plot
and analyze the measurements at different times, a very definite time evolution of the
atmospheric phenomena or time evolution of a meteorological system can be easily obtained.
Since weather knows no political boundaries and many of the meteorological systems are
global in scale, there is a definite need for collecting, and cooperative sharing of the
meteorological information across the globe. The above requires setting-up of a global body
or organization, which ensures that data all over the world are collected, disseminated and
fruitfully utilized by the international meteorological community. Another important reason for
careful measurements of meteorological variables is intimately related to the task of providing
both short- and medium-range forecasting of atmospheric phenomena or meteorological
systems by the operational meteorological forecasting centres around the world.
18 u BASICS OF ATMOSPHERIC SCIENCE

It is certainly not an easy task to clearly pinpoint the beginning of the science of
observational meteorology. Man, during the very early stages of his development must have
scanned the skies with his eyes, sensed the warmth or otherwise of the environment, and
obtained direct exposure to the strength of winds and changes in the wind direction. If one
were to take into account all of the above-mentioned sensations as the beginning of
observational meteorology, then the latter is, simply put as old as the existence of man on
earth. The onset of revolution in the agriculture, which happened about ten thousand years
ago, must have led to the beginning of the awareness of the science of observational
meteorology. Furthermore, the development and successful completion of long distance ocean
voyaging undertaken by mariners in the last millennium as well as the rapid advances in the
aviation industry in the last century have provided an important fillip to the development of
observational meteorology.
It was becoming clear in the early days of observational meteorology that to obtain
definite and unambiguous meteorological observations, one needs to extend beyond the sphere
of human sensations and actually go in for mechanical aids and/or instruments. The
information of the atmosphere derived from the instrumented observations are more accurate
and are less likely to suffer from the subjective inadequacies associated with the observations
not employing instruments. Furthermore, the meteorological observations from instruments
provide reliable quantitative information of the various meteorological variables that are not
possible through the observations not employing instruments. This resulted in extensive use of
meteorological observations derived from the instruments. Once instruments became
mandatory and an important aid in observing the atmosphere, the following question does
arise. What are the requirements for proper exposure of the instruments as well as issues
regarding the instrumental accuracy and its associated error? Furthermore, it is important to
realize that all the meteorological variables are not routinely measured. For example,
quantities such as atmospheric density, atmospheric heating rate due to diabatic sources, as
well as several microphysical cloud parameters, to name a few are not routinely observed in
the atmosphere. In addition to problems of inadequate accuracy of the individual
meteorological observations, there are also issues of representatives and consistency, which
arise naturally when meteorological data from widely scattered observing stations are
collected. The following section provides brief discussion of a few basic instruments
employed for making observations of the atmosphere near the surface.

2.2 MEASUREMENT OF TEMPERATURE AND HUMIDITY


While the air temperature measurements provide a measure of how hot or cold the weather is,
the humidity measurements give a measure of how humid or dry the atmospheric conditions
are. Together, the combined values of air temperature and air humidity determine the “human
comfort” or the “feel-good comfort” for most of us.

2.2.1 Temperature Measurement


The air temperature near the surface is one of the most important of the atmospheric
parameters, which is routinely measured. Although our human body itself is sensitive to air
ATMOSPHERIC OBSERVATIONS u 19

temperature, a less ambiguous temperature sensor is necessary to measure accurately the air
temperature. The most important sensors, which are used to measure the ambient air
temperature, are: (i) mercury in glass thermometers, (ii) resistant temperature detectors, and
(iii) thermocouples. Although Fahrenheit conceived the idea of using mercury in glass
thermometer in the early eighteenth century for the measurement of air temperature, it was not
until almost a century later that there was recognition for the need of careful exposure of the
mercury-in-glass thermometer. To isolate the effect of air temperature, it is important to shield
the thermometer from both solar and terrestrial radiations. Furthermore, it is important to
allow adequate ventilation to ensure that the air in contact with the thermometer is at the same
temperature as the air outside the shield. The requirements for adequate ventilation of air and
proper shielding from both the solar and terrestrial radiations led to the development of the
Stevenson screen. The Stevenson screen, shown in
Figure 2.1, is widely used as a perfect enclosure for
instrument shelter for surface air temperature
measurement. The Stevenson screen has thick wooden
louvred walls, which are painted white. Very little
sunlight is absorbed by the white surface ensuring that
the albedo is maximized. The thick wooden walls
insulate the interiors of the Stevenson screen from the
warming effects of both the residual solar radiation
and terrestrial radiation. Due to the above reason, the
shelter is panelled with slats rather than solid
sidewalls. The door to the shelter is mounted on the
north side of the box (in the northern hemisphere) so
that direct sunlight will not strike the temperature
instruments, whenever the door to the Stevenson
screen is opened during the coarse of the day. The
instrument shelter is of such a height that the FIGURE 2.1 Stevenson screen.
temperature instruments are kept at a height of about 5
feet above the ground.
The main disadvantage of the mercury-in-glass thermometer is that it cannot provide a
continuous record of temperature observations automatically and also cannot be employed at
very high latitudes and altitudes, since mercury freezes at –40°C. For very low temperatures, a
better option would be to replace mercury by alcohol and employ an alcohol-in-glass
thermometer. The alcohol-in-glass thermometer is presently employed in the minimum
temperature measurement. In the minimum thermometer, the alcohol meniscus moves a metal
index inside the bore of the thermometer. If the index is at the end of alcohol and the
temperature is dropping, surface tension pulls the metal index towards the bulb. However,
with the rise in temperature, the index remains at its present position as the alcohol expands
away from the bulb. The minimum thermometer is usually mounted horizontally and the
minimum thermometer is usually reset by turning it upside down, allowing the index to slide
down to the end of alcohol. The maximum thermometer also utilizes the mercury-in-glass
thermometer. The maximum thermometer has a very narrow constriction in the tube just
beyond the bulb which allows the mercury to expand outward whenever the temperature is
20 u BASICS OF ATMOSPHERIC SCIENCE

rising but prevents it from contracting back into the bulb when the temperature decreases. The
temperature shown on the maximum thermometer indicates the highest temperature
experienced by the air since the last time the maximum thermometer was reset. Shaking the
maximum thermometer can force the mercury to go down the bulb ensuring the resetting of
the maximum thermometer.
For recording temperature on a continuous basis automatically, a thermograph is usually
employed. Figure 2.2 shows a thermo-hygrograph, which provides for continuous recording of
temperature and humidity. The thermograph contains a bimetallic strip, which moves a pen
over a clockwork powered chart drum. The bimetallic strip consists of two thin strips of
different metals bonded together. Normally, either bronze and invar or steel and brass are used
as bimetallic strip. Since different metals have different rates of expansion and contraction
with temperature, one of the two metals undergoes a greater change in length as compared to
the other, causing the bimetallic strip to bend. A pointer and scale are attached to the
bimetallic strip, whose bending can be amplified by a lever. The lever is coupled with a
rotating drum and a pen arrangement providing continuous record of the air temperature.

FIGURE 2.2 Thermo-hygrograph.

In addition to the above, there exist other accurate instruments to measure air
temperature, such as the resistance thermometer. In this resistance thermometer, an electric
current is sent through a thin filament made of either a conductor or a semiconductor
material, which is exposed to the air. The temperature of the thin filament is the same as the
temperature of the air in contact with the filament and the resistance to the electric current
depends on the air temperature. The resistance thermometer as the name suggests, registers
the amount of resistance and from the above information the air temperature is determined.
The resistance thermometers show a nearly linear increase of resistance with increase of
ATMOSPHERIC OBSERVATIONS u 21

temperature. Also, resistance thermometers provide for a calibration over a wide temperature
range. A particular type of resistance thermometer that uses a ceramic semiconductor instead
of a metallic wire filament is known as the thermistor. The thermistor unlike the resistance
thermometer provides for a larger change of resistance with change of temperatures. However,
unlike the resistant thermometers the functional relationship between resistance and
temperature in a thermistor is nonlinear. Due to the above reason, a combination of two or
more thermistors and fixed resistors are utilized to provide for a near linear response over a
given temperature range. Thermistors are widely used in situations where there is a need for a
fast response temperature instrument. For example, thermistor is used to measure temperature
at different levels of the atmosphere in a radiosonde comprising of weather instruments,
carried by a balloon, to observe the upper levels of the atmosphere. Another sensor to
measure air temperature known as thermocouple, however, has found very limited application
in actual field measurements. Thermocouples function on the principle of a temperature-
dependent electrical current flow between two dissimilar metals.

2.2.2 Humidity Measurement


The simplest and the most reliable of the
instruments, which measure the amount of
water vapour in the air, is known as psychro-
meters or sling psychrometers, shown in
Figure 2.3. These consist of a pair of thermo-
meters, one of which has a cotton wick
around the bulb that is saturated with water.
The other thermometer has no such cotton
wick and hence simply measures the air FIGURE 2.3 Sling psychrometer.
temperature. The two thermometers, known
as the wet bulb and the dry bulb thermometers are mounted to a pivoting device that allows
them to be circulated through the surrounding environmental air. Both the wet bulb and the
dry bulb thermometers can also be kept mounted inside the Stevenson screen or can be part of
the sling psychrometer. In both cases, of either natural ventilation (Stevenson screen) or the
slinging circulation (sling psychrometer), water is allowed to evaporate from the saturated
cotton wick of the wet bulb thermometer. The resulting cooling of the wet bulb thermometer
with respect to the dry bulb thermometer will directly depend on the humidity of the
surrounding air. If the air is unsaturated, water evaporates from the wet bulb, resulting in
cooling as the latent heat is used up for evaporation. The amount of heat lost by evaporation
is then offset by the input of sensible heat from the surrounding warm air. A balance between
the above two processes will cease the cooling resulting in the wet bulb maintaining a
constant temperature, known as the wet bulb temperature. The difference between the dry
bulb and the wet bulb temperatures called the wet bulb depression depends on the humidity
content of the air. If the air were completely saturated, there will be no net evaporation from
the wet bulb resulting in the wet bulb temperature being equal to the dry bulb temperature. If
the moisture content of the air is very low, considerable evaporation will take place from
the wet bulb resulting in the wet bulb temperature attaining an equilibrium value much lower
22 u BASICS OF ATMOSPHERIC SCIENCE

than the dry bulb temperature. From the knowledge of the dry bulb temperature and the wet
bulb depression, it is possible to determine the humidity content of the air using the basic
thermodynamic principle which embodies the wet bulb process. A modification of the sling
hygrometer is known as the aspirated psychrometer, a psychrometer, which is equipped with
fans that circulate the air across the bulbs of the wet and dry thermometer, thereby avoiding
the need to sling the two thermometers through the air.
Tissues especially those involving animal tissues such as hair or skin respond directly to
changes in the relative humidity of air. This dependence forms the basis for several simple
hygrometers such as the hair hygrometer. In the hair hygrometer, a hair is kept under slight
tension so that the decrease in the hair length with the increase of relative humidity is
recorded. Since the length variation of the hair to relative humidity changes is regular, a
bunch of hairs can be calibrated in several known relative humidity environments. At room
temperatures, hair hygrometer exhibits a lag of about ten seconds. In the hair hygrometer, the
hair is connected to a lever mechanism from which the humidity content of the air can be
determined. When the lever is coupled to a rotating drum and a pointer arrangement,
continuous recording of the humidity can be obtained in what is known as hygrograph. Often,
the hair hygrometer is coupled with a metallic strip and a rotating drum, and pen arrangement
to give a continuous record of both temperature and humidity. Such an instrument, called
the hygro-thermograph or thermo-hygrograph is shown in Figure 2.2. Also available are
fast response humidity sensors for rapid and accurate measurements of air humidity. These
fast response humidity sensors measure the relative humidity through its effect on the
electrical resistance of a hygroscopic surface. Such electrical psychrometers are utilized in
routine measurements as well as in the study of atmospheric turbulence.

2.3 MEASUREMENT OF WIND AND PRESSURE


Typically, wind measurements account for the measurement of wind speeds and wind
direction. The direction of wind, in most situations, can provide valuable information on the
nature of the air, such as warm and moist or cold and dry, etc. which is moving into the
region. Atmospheric pressure differences in the horizontal generally determine the magnitude
of the horizontal wind speed and hence it is important to measure the atmospheric pressure at
different locations.

2.3.1 Wind Measurement


Horizontal wind movements or more notably, the horizontal wind direction and wind speed
are among the important meteorological variables to be observed, for they characterize
transport of air and are important in air quality studies. The wind direction is always defined
as the direction from which the wind blows and is expressed by its azimuth. The degree of
angle from due north and measured clockwise is referred as the azimuth. Thus, while a
northerly/southerly wind has an azimuth of 0° (or 360°)/180°, the easterly/westerly winds have
an azimuth of 90°/270°. A simple instrument, which measures the wind direction, is called the
wind vane. The wind vane, shown in Figure 2.4(a), consists of a horizontal aluminium arm
ATMOSPHERIC OBSERVATIONS u 23

carrying a fin at one end counter balanced about its axis of rotation by means of a cylindrical
mild steel balance weight having a much smaller surface. The wind vane exposes its broad
surface to the wind, while the other end is narrow and points to the direction from which the
wind is blowing. The entire assembly is mounted over a cup having a ball bearing. When the
wind changes direction, the wind vane pushes against the tail, and points the wind vane
towards the wind. Under normal winds, the typical wind vane has a lag of several seconds.
Anemometer of the cup type, called the cup anemometer, shown in Figure 2.4(b),
usually measures wind speed. The principle of wind speed measurement by cup anemometers
is based on the differential drag force exerted by the wind on the cups which cause them to
rotate, and this rotation is a measure of the wind speed. These have rotating semiconical cups
mounted at the edges of three rods placed on a moving shaft. These cups are mounted
symmetrically about a vertical axis such that the diametric planes of each cup are vertical.
The cups of the anemometer rotate continuously since the drag of the wind is greater when
blowing into the mouth than in the back of each cup. By careful design, it is possible to relate
directly the rate of the cup rotation to the wind speed over a wide-range of wind speeds. The
number of cup axle revolutions in a certain time period is noted in the cup anemometer and
this is converted into the length of the wind. Dividing the length of the wind by the time
period used, one gets the average wind speed in this period. Typically, three-minute or ten-
minute averages are used to define the wind speeds near the surface in conventional
observation. Unlike the cup anemometer, which is insensitive to the azimuth or the horizontal
wind direction, the propeller type anemometer must be kept pointed into the wind by a
steering vane. To observe the vertical component of the wind, an additional propeller on a
vertical axle needs to be included.

FIGURE 2.4(a) Wind vane. FIGURE 2.4(b) Cup anemometer.

A different type of wind measuring instrument is the pitot tube anemometer, which
measures the excess pressure developed in the air as the air rams into the mouth of a small
tube steered into the wind by a wind vane. Since the rammed air is brought to a state of rest,
the air loses all its kinetic energy. Using the Bernoulli principle, the excess pressure can be
related to the square of the wind speed and the latter can be determined. A more accurate
measurement of the horizontal wind speed is obtained from the electrical anemometer. The
electrical anemometer consists of a small electrical generator maintained in a weatherproof
24 u BASICS OF ATMOSPHERIC SCIENCE

housing and is driven by the rotation of a three-cup rotor carried on a vertical spindle. The
voltage generated increases with the wind speed and is primarily used in remotely located
observation sites.

2.3.2 Atmospheric Pressure Measurement


There are primarily two different types of instruments, which measure atmospheric pressure,
these are the mercury-in-glass barometer, shown in Figure 2.5, and the aneroid barometer.

FIGURE 2.5 Mercury-in-glass barometer.

The principle of pressure measurement by mercury-in-glass barometer is that the pressure


of the air acting on an exposed surface of mercury containing an inverted tube, raises
the mercury in the inverted tube and the weight of the raised mercury per unit area in the
inverted tube is the atmospheric pressure. The mercury-in-glass barometer consists of a glass
tube of about 84 cm height, closed at one end and open at the other. The above tube is filled
with mercury and then inverted in a trough full of mercury. The height of the mercury level in
the tube above its level in the trough is a measure of atmospheric pressure as proposed by
Torricelli in the year 1644. According to the Torricelli principle, atmospheric pressure acts on
the exposed surface of mercury in the trough to maintain a column of mercury in the glass
tube. The atmospheric pressure is simply the product of the vertical height h of the top of the
mercury column above the exposed surface of mercury, the gravitational acceleration g and
the density of mercury r and is measured in units of hPa. (1 hPa = 100 Pascal). The mercury
level in the mercury barometer is read with a special form of a graduated scale attached to it.
Mercury is preferred to other liquids in barometer in view of its high density, requiring
shorter columns and its property of not sticking to the walls of the container. The above
readings from a mercury barometer need to be corrected for variations in temperature,
gravitational acceleration and altitudes. Temperature correction is required since the mercury
in the barometer expands with increase of temperature. However, the temperature corrections
are usually very small and are of the order of 0.1 hPa or smaller. The value of acceleration of
gravity is slightly greater at higher latitudes and to standardize the readings from all latitudes,
reference latitude of 45°N or 45°S is considered and the local gravity is converted to the
reference latitude gravity value. The most important correction is the altitude correction to
compensate for the influence of the height of the station on the pressure and is done by
converting surface station pressure values to sea level values.
The second type of pressure instrument called the aneroid barometer consists of a
partially evacuated drum (aneroid) that has an elastic cover on it. The drum has a spring with
a pointer attached to it. Associated with a decrease/increase in the atmospheric pressure, the
aneroid expands/compresses and this movement of the drum is detected by a manually
operated probe. The aneroid barometer is usually calibrated against a mercury-in-glass
barometer. Aneroid barometers are relatively inexpensive and are reasonably accurate. Due to
ATMOSPHERIC OBSERVATIONS u 25

the absence of an expandable fluid in the aneroid barometer, these instruments do not require
any temperature correction. Also, the effects of height and latitude are already taken care of
when the aneroid barometer is first calibrated and hence these instruments directly give the
sea level pressure without any further corrections. A continuous record of atmospheric
pressure can be obtained from a barograph, shown in Figure 2.6. The barograph is similar to
the aneroid barometer in which the movement of the aneroid drum is communicated
mechanically to a pen writing on a graph fixed to a drum rotating by clockwork.

FIGURE 2.6 Barograph.

2.4 MEASUREMENT OF PRECIPITATION


The amount of precipitation being received at the surface over a typical observation period
is among the most important meteorological variables. In addition to meteorologists,
the precipitation measurements are of interest to hydrologists who are concerned with the
management of rivers and water reservoirs. Even though most of the precipitation, which
reaches the surface, is actually rainfall, the precipitation term truly does include all the various
forms by which water and ice fall to the surface. Hence precipitation includes rainfall, drizzle,
snowfall and hail as well. Rain gauges are instruments utilized to measure the quantity of
rainfall received at the surface in a 24-hour observation period. The rain gauge consists of a
circular collector of about 100 or 200 square cm cross-sectional area in the form of a funnel
base, a polythene bottle and a measuring glass. The collectors as well as the base are made up
of fibreglass-reinforced polyester. The depth of water in millimetres collected on a horizontal
area specifies the quantity of rainfall in a 24-hour period. During days of snowfall or hail, or
other forms of frozen precipitation, measured warm water is poured into the rain gauge to
melt the snow or frozen precipitation. The actual amount of precipitation is then obtained by
subtracting the amount of known warm water from the total measured precipitation.
The standard rain gauge can provide only the 24-hour accumulated precipitation and not
the timing and the intensity of the observed precipitation. For obtaining the timing and the
26 u BASICS OF ATMOSPHERIC SCIENCE

intensity of the precipitation, an automated


collector called the tipping bucket gauge is
used. Figure 2.7 shows a continuous
recorder rain gauge of the siphon type. The
tipping bucket gauge, like the standard rain
gauge funnels the precipitation from the top
and the accumulated water is stored in one
of the two pivoting buckets. One of the two
buckets is initially kept upright, while the
other bucket, mounted on the opposite end of
a pivoting lever, is tipped downward and
away from the collector. When the upright
bucket collects precipitation, equivalent to a
certain depth, the weight of the water-
collected causes the upright bucket to tip FIGURE 2.7 Continuous recording rain gauge
of the siphon type.
over and empty its contents and bring the
other bucket to the upright position. The tipping of the pivoting buckets triggers an electric
current to a computer that notes precisely the time of the tipping. The number of tips per unit
of time provides the intensity of the observed precipitation. Another form of automated
rainfall measurement is by a weighing bucket rain gauge. This instrument has a weighing
mechanism that communicates the weight of the accumulated water in the rain gauge to a
precipitation depth and this information is recorded automatically.

2.5 MODERN METEOROLOGICAL INSTRUMENTS


Modern devices are available which measure temperature using the property of the change of
electrical resistance with temperature or employ the principle of thermocouple. In any case,
the modern devices are designed so that the air temperature can be directly recorded in digital
form in a computer. Modern pressure sensors also called micro-barographs employ tiny
silicon diaphragms attached to a capacitance gauge. Depending on the external pressure, the
diaphragm bends by different degrees altering the capacitance of the attached capacitor. These
so-called microbarographs can detect pressure change of the order of 2 Pa or less. Modern
humidity metres use a capacitor which consists of two metal plates separated by a thin
polymer film. The film releases or absorbs water vapour with the decrease/increase of the
humidity, changing the dielectric constant of the film. The above change in the dielectric
constant causes a change in the capacitance of the unit which can be electronically recorded.
Electrical hydrometers, the dew point hygrometer, the infrared hygrometer and the dew cell
are some of the other modern humidity measuring devices available. The electric hygrometer
is based on the principle of a change in resistance across a carbon-coated plate due to
absorption/release of water vapour with the change of humidity as an electric current is
passed through the plate. While the dew point hygrometer measures the temperature at which
a change of phase (condensation) takes place over a cold plate, the dew cell measures directly
the vapour pressure of the air. The principle of infrared hygrometer is based on the absorption
of the infrared light as it passes through the air with the absorption being dependent on the
humidity content of the air. An example of the modern rainfall measuring device which
ATMOSPHERIC OBSERVATIONS u 27

allows for automation and digitization is known as the tipping bucket rain gauge. In the
above device, at regular intervals of time, the gauge tips its contents onto a measuring scale
and the contents are electrically weighed with the rainfall amounts stored digitally. Fast
response wind information can be obtained from the sonic anemometer. In this device, a
sound signal from a fixed transmitter to a fixed receiver is sent by measuring the time for the
sound to travel; from which the speed of sound can be found. Depending on whether the
wind is a tail or a head wind, wind speed can either increase or decrease. By measuring the
speed of sound in both the directions, the wind speed along the axis can be measured from
the difference of the two measurements. For measuring the horizontal or the three-dimen-
sional component of wind, a two-axis or a three-axis sonic anemometer instrument can be
employed. Sound speed is a function of air density which depends mainly on the temperature
and to a lesser extent on the humidity and hence can be envisaged as a measure of
temperature. Hence the sonic anemometer can be used for air temperature measurement also.

2.6 SURFACE AND UPPER AIR OBSERVATIONAL NETWORK


The work by European meteorologists showed that the weather in middle latitudes is
organized in patterns whose horizontal extent ranged to about 1000 km or more. By the
middle of the nineteenth century, the development of electric telegraph in the year 1844
provided a means of effective communication of meteorological data from a network of
surface observing stations. The establishment of a network of upper air stations followed
this much later. Presently, the above network of meteorological observing stations, which
covers the land areas of the globe, is organized under the aegis of the World Meteorological
Organization (WMO). A complete system of observing, communicating, analyzing and
forecasting the weather is known as World Weather Watch (WWW), which functions under
WMO. The purpose of WWW is to observe and record carefully the present state of the
atmosphere, analyze the data to ultimately enable useful weather forecasts. Since the
processes associated with the various weather systems are complex and our understanding of
these processes remains somewhat incomplete, it is important to expand our network of
meteorological stations. However, the costs of personnel and the availability of funds for
procuring the required meteorological instruments together limit the number of meteorological
observing stations in the network. Presently, the network of surface and upper air
meteorological stations are just about adequate to define the large-scale weather patterns as
manifested in the synoptic scales. Synoptic scales have typically horizontal scales of the order
of 105 m to 106 m. It is important, however, to emphasize the need for uniform standardization
of all the observational practices within the entire network of meteorological stations.

2.6.1 Surface Observational Network


Observations of wind speed and direction, temperature and humidity, atmospheric pressure
and visibility, 24-hour accumulated rainfall, types and amount of cloud, nature of the present
and past weather are typically made every hour at all surface stations over the globe with
respect to a universal time (UTC). For regions where hourly surface data are difficult to
observe and report, the surface observations are restricted to 0000, 0006, 0012 and 0018 UTC
everyday, which are called main synoptic observations. However, four more observations are
taken at 0300, 0900, 1500 and 2100 UTC by some selected main observatories and these are
28 u BASICS OF ATMOSPHERIC SCIENCE

known as auxiliary synoptic observations. All the instruments in the surface observational
network conform to uniform international standards and are all operated by trained staff.
Furthermore, the surface observational stations are designed to ensure that they provide for
uniform standard exposures consistent with international requirements.
The distribution of surface meteorological observational stations in the developed
western countries has an average separation of about 50 km. In relatively poor countries, the
average separation between the surface meteorological stations may be of the order of 75 km
to 100 km. Many of the developed countries are installing fully automatic weather stations to
increase the surface network of meteorological stations. These automatic weather stations, in
addition to reducing the costs incurred towards manpower, also require minimal maintenance
and repair costs. Surface meteorological observations over the sea are on voluntary basis by
the merchant ships. While the worldwide total number of land-based surface meteorological
network of stations may be a few thousand, the average number of ships providing surface
meteorological observation over the sea is lower than the land-based stations. In India, the
India Meteorological Department (IMD) has about 559 surface observatories over land and
has access to about 203 ships, which provide voluntary assistance of surface meteorological
observation over the sea.

2.6.2 Radar Network


Many of the western developed countries presently have also a surface network of weather
radars. Since the radars provide meteorological information for about 400 km, the average
separation between coordinated network of surface radars may be much higher than the
surface meteorological network of stations. In India, the IMD’s network of X-band radars
consists of twenty-nine radars comprising about twelve storm detection radars and seventeen
dual-purpose weather-cum-wind finding radars. The X-band radars operate in a wavelength
range of 2.5 cm to 4 cm and a frequency of 8 GHz to 12 GHz. Due to their smaller
wavelengths, the X-band radars are used to detect small water droplets and light precipitation.
Since the wavelengths associated with the X-band radars get attenuated easily, they are used
only for short-range weather observations. Furthermore, the IMD has a network of ten S-band
cyclone detection radars covering the Indian coast with six radars along the east coast and the
remaining four radars along the west coast. The S band radars operate in a wavelength range
of 8 cm to 15 cm and a frequency of 2 GHz to 4 GHz. Due to the above choice of frequency
and wavelengths, the S-band radars are not attenuated much and hence are widely used for
near- and far-range weather observations. Of the six S-band radars on the east coast of India,
S band analogue cyclone detection radars at Chennai, Kolkata, Machilapatnam and
Visakhapatnam have already been replaced by S-band Doppler Weather Radars (DWRs). In
addition, a new DWR has been installed at Sriharikota during 2004.

2.6.3 Upper Air Observational Network


The most important instrument intended for the upper air observations of wind speed and
direction, air temperature, relative humidity and pressure, is known as the radiosonde. The
radiosonde carries a small radio transmitter together with sensors to measure pressure,
temperature and humidity and the entire assembly is made to go up by a free-flying balloon
released from the surface at 0000 and 1200 UTC daily. The transmitter produces a carrier
wave, which is amplitude, modulated by signals from thermistor, hygristor sensors. The
ATMOSPHERIC OBSERVATIONS u 29

released balloon rises at a rate of 5 m s–1, and can reach heights of about 20 km to 30 km
above the mean sea level. While an aneroid capsule measures pressure, temperature and
humidity are measured by a thermistor and hygrister/humicap, respectively. A small electric
motor turns the switch, which couples the individual oscillators to the transmitter circuit,
while a small battery provides power for the assembly. The information of air temperature,
humidity and pressure at various upper levels of the atmosphere are measured and transmitted
to the ground receiving station. The radiosonde receiver consists of a radio receiver and a
computer-controlled processor, which measures and decodes the signals. The sensors in the
radiosonde are calibrated before the launch and these calibration factors are used to derive the
values of the meteorological parameters. The balloon carries a metallic reflector which acts as
a target for the ground-based radar. As the balloon freely ascends, it is carried along
horizontally by the winds. The information obtained from successive positions at frequent
intervals of time is used to derive the vertical profile of the horizontal wind. A typical upper
air observational network of meteorological station in the developed western countries has an
average separation of about 200 km, while the same for the developing countries is about 300 km.
In India, the IMD has an upper air network of thirty-nine radiosonde stations (refer Figure 2.8).
It is virtually impossible to reuse the radiosonde sensor and the transmitter after it has fallen
back to the ground after providing the upper air meteorological data. Due to the above
reasons, the costs associated with the upper air radiosonde network are indeed very high.

39N

36N

SRINAGAR
33N

30N PATIALA

NEW DELHI
27N LUCKNOW GORAKHPUR DIBRUGARH
SILIGURI
JODHPUR GUWAHATI
GWALIOR PATNA
24N
AHMEDABAD AGARTALA
BHOPAL RANCHI
KOLKATA
21N NAGPUR RAIPUR
AURANGABAD BHUBANESWAR
MUMBAI JAGADALPUR
18N HYDERABAD
VIZAG

15N MACHILIPATANAM
GOA
BANGALORE
MANGLORE CHENNAI
12N PORT BLAIR
AMINI DIVI KARAIKAL
KOCHI
9N
MINICOY THIRUVANANTHAPURAM

6N

66E 69E 72E 75E 78E 81E 84E 87E 90E 93E 96E 99E

FIGURE 2.8 Present radiosonde network of upper air stations over India.
30 u BASICS OF ATMOSPHERIC SCIENCE

Upper winds are also commonly determined by an alternate manner called the pilot
balloon. Here, a small rubber balloon is filled with either hydrogen or helium gas to enable it
to rise at a known constant rate. During the balloon ascent, observations of the elevation and
azimuth of the balloon are made at frequent intervals of time using an instrument called as
theodolite. From the knowledge of the elevation, azimuth and the constant rate of ascent of
the balloon, the wind speed and wind direction at different levels can be determined. In
the presence of light wind, the balloon can easily reach altitudes of about 9 km. In India, the
IMD has an upper air network of sixty-two pilot balloon stations, which are shown in
Figure 2.9. Although only two to four pilot balloon observations are taken over a day and the
network is relatively coarse, it turns out that the present upper air network is not totally

AMRITSAR SUNDERNAGAR

AMBALA DEHRADUN
SRI GANGANAGAR
CHURU MOHANBARI
GANGTOK
BIKANER DELHI BAREILLY

JAIPUR BAHRAICH
JAISALMER SILIGURI
JODHPUR GORAKHPUR
LUCKNOW GUWAHATI
BARMER
KOTA PATNA IMPHAL
UDAIPUR ALLAHABAD
BHAGALPUR AGARTALA
GAYA
BHUJ
DEESA RANCHI BANKURA

AHMEDABAD BHOPAL JABALPUR JAMSHEDPUR


CALCUTTA
JHARSUGUDA
VERAVAL SURAT BALASORE
AURANGABAD NAGPUR RAIPUR
BHUBANESHWAR
MUMBAI JAGDALPUR
GOPALPUR
PUNE
VISHAKHAPATNAM
RATNAGIRI
HYDERABAD

MACHILIPATNAM
GADAG

ANATAPUR

CHENNAI PORT BLAIR


MANGALORE
BANGALORE

KARAIKAL
AMINI
TIRUCHIRAPALI

MINICOY THIRUVANATHAPURAM

FIGURE 2.9 Present pilot balloon network of stations over India.


ATMOSPHERIC OBSERVATIONS u 31

inadequate. Unlike the upper air network, one requires a very large number of surface station
networks. The reason for the above behaviour is that the atmospheric structure is much
smoother and is predominantly large scale at the upper levels of the atmosphere than it is
close to the surface. Due to the presence of surface inhomogeneties, the atmospheric structure
is significantly smaller and relatively more transient, i.e. less smoother close to the surface.
This is why, despite the relatively fine mesh and more frequent surface observations, the
surface meteorological network is just about adequate to provide for atmospheric structures
close to the surface.

2.7 SATELLITE OBSERVATION


Oceans cover about seventy percent of the surface area of the earth. Furthermore, the oceans
are typically data-sparse regions for conventional surface and upper air meteorological
observations. The only means of overcoming the above difficulty is to take recourse to
observations of a non-conventional nature, such as those available from satellites to fill in the
gaps of meteorological data over the data-sparse oceanic regions. The first meteorological
satellite was successfully launched in the year 1960 and in the last half of the century the
satellite observation of meteorological and oceanic variables has provided an invaluable
source of data for both the weather analysts and the forecasters. The earlier meteorological
satellites provided cloud pictures in the visible and in the infrared part of the electromagnetic
spectrum. Furthermore, the relatively high resolution of the above satellite pictures of the
order of a few kilometres revealed the detailed structure of the various cloud systems, which
were associated with the meteorological system.
Meteorological satellites can be broadly classified into two types: (i) Polar orbiting, and
(ii) Geostationary satellites. Polar orbiting satellites orbit from pole-to-pole to scan the
earth and its atmosphere from a height of 850 km. These satellites follow orbits nearly fixed
in space and are able to scan the entire surface of earth as the earth rotates beneath these
satellites. The areas scanned by the polar orbiting satellite on each pass, known as swath is
typically about 2600 km wide. The polar orbiting satellites complete 14 orbits per day and can
provide coverage over the entire globe twice a day. Geostationary satellites orbit around the
earth once in 24 hours. They orbit over the equator at a height of 36000 km and hence remain
over the same location on the equator. The main advantage of the geostationary satellites is
the high time scale resolution (30 minutes) of their data. However, their disadvantage is that
they have limited spatial resolution and cannot provide global coverage. The sun-synchronous
orbit which most polar orbiting satellites use ensures that the satellite passes over given
latitude at the same local sun time. Meteorological satellite sensors can be classified broadly
into the following two types: (i) active, and (ii) passive sensors. The active sensors use their
own source of electromagnetic radiation to illuminate targets and from the data of reflected
radiation derive information of the target. The passive sensors do not have their own source of
electromagnetic illumination and depend on the radiation emitted or reflected from objects
about whom the information is sought.
Satellite Remote Sensing is the means of getting information about the earth surface
(land, oceans) and the earth atmosphere using sensors in space-borne platforms. Active
precipitation radar flown on a satellite can provide valuable information on the intensity and
32 u BASICS OF ATMOSPHERIC SCIENCE

distribution of rainfall. Narrow beam radar flown on a satellite transmits radiation which after
striking raindrops in the atmosphere gets echoed back to the satellite. From the knowledge of
the returned radar pulse, the size and the height of the raindrops are inferred. Scatterometer
measures the wind speed and direction over the sea surface and is based on the principle that
electromagnetic radiation transmitted towards the sea surface gets scattered back and the
intensity of backscatter is dependent on the sea surface roughness. The backscattered intensity,
for a given wind speed depends on the azimuth angle between the wind direction to the
emitting antenna. By measuring backscattered intensity from the sea surface at several
azimuth angles, the wind speed and wind direction over the sea can be obtained. Cloud
motion vectors can be derived from successive images of cloud charts from geostationary
satellites by assuming that the duration and the amounts of cloud development are due to the
wind vectors at those altitudes. In addition, various components of the earth radiation budget
can also be retrieved from satellites.
The earlier meteorological satellites provided a convenient platform for scanning
passively the electromagnetic radiation emanating from the earth atmosphere system. The
sensor utilized in such passive satellites is sensitive to one or more wavelengths bands in the
visible and the infrared part of the electromagnetic spectrum. The field of view from the
passive satellite sensor scans line-by-line either by utilizing the equivalent of a television
camera or by physically sweeping a very narrow-field radiometer across the view. The data is
then sent in a sequential manner to a receiving station on earth for reconstruction of the entire
picture. The limit of resolution of the picture is determined by the width of a scan line on the
earth surface. Typically, the limit of resolution of the image is a few km at a location
vertically beneath the satellite, while the limit changes with increasing obliqueness of the
view. When visible wavelengths are used, the visible picture is primarily governed by
differences in the reflected sunlight due to differing albedos of the reflecting surface. This
yields a visible picture having a brightness scale which corresponds to various albedo scales.
The disadvantage of the visible picture is that it is not available in the night side of the earth.
Radiometers that are sensitive to wavelength in the far-infrared (wavelengths greater
than 3 mm) are almost unaffected by the short-wavelength solar radiation, but are quite
sensitive to the terrestrial radiation emitted by the earth surface and the atmosphere. The
intensity of the terrestrial radiation emitted increases with the increase in the temperature of
the emitting surface and this results in a infrared image having a brightness scale which
corresponds to a temperature scale. The cloud-free atmosphere while being opaque to the
infrared radiation over a certain part of the infrared spectrum is also mostly transparent to
other parts of the infrared spectrum. Hence, a radiometer measuring terrestrial radiation over
wavelengths for which the atmosphere is transparent yields infrared picture highlighting
the temperature of the ground or sea surface in the absence of clouds or the temperature of
the surface of the cloud top when clouds are present. However, the infrared pictures corres-
ponding to the wavelengths in the opaque atmosphere primarily reveal the distribution of
temperature for the middle and upper troposphere.
Passive sensors are also utilized to infer the temperature profiles of the atmosphere by
making measurements at a number of different wavelengths in the infrared and microwave
regions of the spectrum. Wavelengths are chosen at which carbon dioxide or oxygen strongly
absorbs and emits radiation. Since these gases are evenly distributed throughout the atmos-
ATMOSPHERIC OBSERVATIONS u 33

phere, the amount of radiation received by the satellite sensor is dependent on the temperature
of the emitting gas. By judicious choice of suitable wavelengths it is possible to obtain the
temperature profile of the atmosphere. A similar method is used to infer the vertical profiles
of water vapour content, except that the infrared wavelengths used here are particularly
sensitive to water vapour emissions. The sea surface temperature can be obtained from the
infrared measurements corresponding to the atmospheric window region if there were no
cloud. From the knowledge of the infrared measurements of the cloud top temperature and
comparing it with the derived vertical atmospheric profile, it is possible to obtain the cloud
top height. Active satellite sensors are in principle like conventional radars, which send
electromagnetic pulses, directed towards the target. From the analysis of the received
backscattered electromagnetic radiation, one derives the physical properties of the target
surface and the intervening atmospheric medium. Active satellite sensors are utilized to
determine the variations in salinity of water as well as in radio altimetry.
Meteorological satellites are launched in one of the two types of orbits called as the
near-polar orbit and the geostationary orbit. The polar orbiting satellites are usually launched
in what is known as sun-synchronous type where the satellite orbit is about 860 km above the
earth surface. The above orbits pass near the poles but make an angle to the meridians, which
is just enough to allow the orbit to remain effectively fixed relative to the sun. This sun-
synchronous type of orbit ensures that the sun is no longer likely to introduce any variations
in the satellite measurements. The above orbit takes about 102 minutes between successive
passes near the same pole, while its radiometers scan the swath of the planet passing
continuously below it. In the geostationary orbit, the satellite orbits at about 36000 km (the
height at which they will have the same orbital speed as that of the earth) over the earth
surface above the equator and moves in the same direction as the rotation of the earth. In this
orbit, the satellite maintains a fixed position relative to the earth. A satellite in this position
covers about one third of the earth surface. A series of five such geostationary satellites
equally spaced around the equator can monitor all but not the polar regions. The various
sequence of images obtained at frequent intervals of time can be processed to provide for the
movements of cloud systems and can be used to forecast the development of cloud/weather
systems. Tracking the movement of cloud patterns from successive geostationary satellite
images can help in deriving the upper level winds. The basic assumption is that the cloud
cluster has moved with the appropriate layer wind and such satellite-derived winds are an
important source of wind information over the oceans.

REVIEW QUESTIONS

1. Write down the wind direction for the following winds: (i) westerly, (ii) easterly, (iii) south-easterly,
(iv) north-easterly.
2. Assume that the air temperature and the pressure inside a room are constant. Assume that
the water vapour content inside the room is increasing. Mention whether the following
would increase or decrease: (i) vapour density, (ii) relative humidity, (iii) wet bulb
depression, (iv) dew point depression.
3. Why is the Stevenson screen painted white?
34 u BASICS OF ATMOSPHERIC SCIENCE

4. Why does one incorporate temperature corrections while reading the pressure from a
mercury barometer?
5. While reading the pressure from a mercury barometer, how does one account for the greater
acceleration of gravity at higher latitudes?
6. Why should one incorporate altitude corrections while reading the pressure from a mercury
barometer?
7. Why is alcohol used in minimum thermometers?
8. What is the principle behind using bimetallic strip for continuous temperature
measurements?
9. Name an instrument which can provide the three components of the velocity.
10. Name the instruments which can provide for continuous measurement of precipitation?
11. Mention the nature of the errors associated with the raingauge.
3 Atmospheric
Thermodynamics

Thermodynamics is a practical science based on a small number of principles that are deduced
from common everyday experience. The science of thermodynamics is concerned primarily
with the large-scale or macroscopic properties of matter and makes no hypothesis about the
micro- or small-scale structure of matter. Thermodynamics is also mostly phenomenological in
its spirit in the sense that it describes and treats phenomena and not their causes. The present
state of knowledge of thermodynamics is due to a combination of inductive–deductive
thought. Certain postulates were put forward after observing the behaviour of matter through
observations of everyday experience. The consequences of these postulates/assumptions were
then examined and checked with actual experiments. Wherever the consequences of the
postulates were found to be at variance with the actual experiments, the postulates were either
modified to fit in with actual reality or they were given up entirely. Albert Einstein, the
celebrated physicist had this to say on thermodynamics: A theory is the more impressive the
greater the simplicity of its premises is, the more different kinds of things it relates and the
more extended its area of applicability. Therefore the deep impression, which classical
thermodynamics, made upon me. It is the only physical theory of universal content
concerning which I am convinced that within the framework of the applicability of its basic
concepts it will be never overthrown.
The science of thermodynamics had its beginnings in the early nineteenth century and
was put forward primarily to address questions on improving the efficiency of steam engines.
The scope and application of thermodynamics broadened with the growth of the subject and it
now finds application in design of internal combustion engines, conventional thermal and
nuclear power plants, refrigerator and air-cooling systems, propulsion systems for rockets,
missile, aircraft, ships and land vehicles, science of chemical physics as well as in the
production of extremely low temperature in the neighbourhood of absolute zero.
The atmosphere is a uniform mixture of gases except for water vapour. The atmosphere
and its constituent gases are assumed to be ideal and the equation of state for the above
mixture is discussed in Section 3.1. The important thermodynamical concepts of work and
35
36 u BASICS OF ATMOSPHERIC SCIENCE

heat as well as the first law of thermodynamics are outlined in Section 3.2. An important
thermodynamical process—the adiabatic process is introduced in Section 3.3. Water vapour
as one of the constituents of the atmosphere which can exist in all its three phases is
introduced and its thermodynamics and concepts of latent heat are discussed in Section 3.4.
The balance between the vertical pressure gradient force and gravity—hydrostatic equilibrium
is introduced in Section 3.5. A brief discussion of thermodynamical diagrams is outlined in
Section 3.6. The concepts of hydrodynamic stability with respect to both the parcel and slice
approaches are discussed in Section 3.7. Entropy and the second law of thermodynamics are
outlined in Section 3.8, while the Carnot cycle and the Clausius Clapeyron equation are
introduced in the final section of this chapter.

3.1 GAS LAWS AND EQUATION OF STATE FOR A MIXTURE OF


IDEAL GASES
It is better to put in place some basic definitions before starting any discussions. The term
“system” refers to that part of the universe (or physical world) to which we direct our
attention. The region outside the “system” refers to the surroundings. System can be open
(exchange matter with surroundings) or closed. The property of a system is defined as
outlining a procedure for performing an experiment, the result of the measurement being the
numerical value of that property. Extensive properties of a system are those whose
measurement requires observing the entire system, while intensive properties are those which
are not extensive. Volume, energy and mass are extensive properties, while temperature,
pressure and density are examples of intensive properties. The state of a thermodynamic
system is specified by the values of all the properties of that system. The simplest type of
system is one in which the properties of the system do not change with time; such systems are
said to be in a state of equilibrium. A closed system with no thermal interaction with the
surroundings and not subjected to any mechanical action is known as an isolated system.
Every (isolated) system when left to it will ultimately reach its state of equilibrium. A process
is known as reversible when the direction of the process can be reversed by an infinitesimal
change in some property of the system. A process is known as adiabatic if there are no heat
interaction between the system and the surroundings.
The atmosphere behaves as an ideal gas to a reasonably good approximation by obeying
the ideal gas law

Q7 N35 (3.1)
.
where p is the pressure, V is the volume, m and M are the mass of the gas and its molecular
mass respectively, R is a gas constant and T the temperature. The ideal gas law was arrived at
after careful experiments by Robert Boyle in 1660 and Jacques Charles in 1787 led to the so-
called Boyle’s law and Charles’ law, respectively. Boyle’s law states that at constant
temperature, pressure of a given mass of a gas is inversely proportional to its volume, while
Charles’ law refers to the direct proportionality between the temperature and volume of a
ATMOSPHERIC THERMODYNAMICS u 37

mass of a given gas at constant pressure. A gas, which completely satisfies both Boyle’s and
Charles’ law, is known as an ideal gas. For any ideal gas, the following relation should hold
Q7
DPOTUBOU (3.2)
5

The above equation incorporates Boyle’s law ( Q — at constant temperature) as well
7
as Charles’ law (V µ T at constant pressure) and hence has the form of the equation of state
for an ideal gas. It is clear that the size of the constant in the right-hand side of Eq. (3.2)
depends on the amount of the gas used (i.e. mass m), the nature of the gas (molecular mass M)
used and any universal constant (universal gas constant R). The universal gas constant R has
a constant value 8314.3 J kg–1 K–1. It is important to note that in practice, gases obey the
ideal gas laws only at moderate pressures and at temperatures well above the temperatures at
which the gas would liquefy. The critical temperature of a gas is the temperature at or above
which no amount of pressure, however great, will cause the gas to liquefy. The minimum
pressure required to liquefy the gas at the critical temperature is called the critical pressure.
Our atmosphere is primarily made up of nitrogen and oxygen and the critical temperature and
pressure of nitrogen is 126 K and 3.4 MPa, while the same for oxygen is 155 K and 5 MPa.
Also, the normal boiling points (temperature at which the vapour pressure is one atmosphere)
are about 90 K and 77 K for oxygen and nitrogen, respectively. It is now clear why our
atmosphere can be considered as an ideal gas.

3.1.1 Mixture of Gases


To deduce an equation of state for a mixture of ideal gases, it is necessary to invoke the so-
called Dalton’s law. This law states that the total pressure exerted by a mixture of gases is
equal to the sum of partial pressures, which would be exerted by each constituent if it alone
filled the entire volume at the same temperature as that of the mixture. Hence for a mixture of
k components
L
Q Q  Q  "  QL ÇQ O (3.3)
O 

where pn’s are the partial pressures of the constituents and p is the total pressure. Assume that
the mass and molecular mass of each constituent of the mixture are mn and Mn (n varies from
1 to k) and let V be the total volume of the mixture. Assuming that each constituent obeys the
ideal gas laws individually, one has
NO 35 (3.4)
pnV =
.O
From Dalton’s law,
L
35 L N
p= Ç QO 7 Ç .O (3.5)
O  O  O

The specific volume (a), reciprocal of density in this case is related to the volume V and
the total mass of all the constituents m. That is
38 u BASICS OF ATMOSPHERIC SCIENCE

7 7
B (3.6)
N L

Ç NO
O 

Substituting Eq. (3.6) in Eq. (3.5), results in


L
N
Ç .O
QB 35 O L O (3.7)
Ç NO
O 

Equation (3.7) shows that a mixture of ideal gases also obeys a gas law and an equation of
state, which is of the same form as the ideal gas law for a single constituent of the mixture
provided a mean molecular mass of the mixture of gases M is defined as
L
N

Ç .O
O  O
(3.8)
. L

Ç NO
O 

Hence the equation of state for a mixture of ideal gases is given by


35
QB (3.9)
.
The mean molecular mass for a mixture of ideal gases is just the mass weighted harmonic
mean. The mean molecular mass of dry air calculated in the above manner from
3
Eq. (3.8) has a value of 28.96 kg k mol–1. The ratio = Rsp is known as the specific gas
.
constant for dry air and has a value of 287 J kg–1 K . –1

3.2 WORK, HEAT AND FIRST LAW OF THERMODYNAMICS


The important concepts in thermodynamics such as work and heat are introduced in the
following subsections. Both work and heat are path functions; meaning that both depend not
only on the initial and final states but also on the nature of the thermodynamic process which
connect the initial and final states of a thermodynamic system. However, under special
situations work ceases to be a path function. The statement of the first law of thermodynamics
under such special situations is discussed in the following subsection 3.2.4.

3.2.1 Work
Work can be associated with a variety of processes and each form of work can be converted
to other forms. For example, gravitational work may be done on water while falling over
a dam. The falling water may then perform mechanical work by exerting a force through a
ATMOSPHERIC THERMODYNAMICS u 39

distance against the turbine of an electric generator, the latter being used to perform electrical
work. The convertibility of the various forms of work from one to another provides a means
to compare the various kinds of work against a standard, i.e. how each form of work raises a
standard weight against the earth’s gravitational field. It is of course convenient to relate the
definition of work in thermodynamics with concepts such as systems, properties and
processes. Work is said to be done by a system if the sole effect on the surroundings could be
the raising of a weight. It is to be noted that rising of a weight is in effect force acting
through a distance. Also, the above definition of work does not state that a weight was
actually raised or that a force actually acted through a given distance, but the sole effect
external to the system could be rising of a weight. By convention, the work done by a system
is considered positive, while the work done on the system is considered negative. The reason
for the above convention is that in the initial stages of the development of thermodynamics,
the primary interest was the work done by a system in a process in which steam in a cylinder
is expanded against a piston. Conveniently, the work done by the system in the above process
was considered positive.

3.2.2 Work Done by a System Expanding Against


External Forces
Let the continuous line represent the boundary of dFe = Pe dA
a system of volume V and arbitrary shape acted on
by a uniform external hydrostatic pressure Pe. Let
the system expand against pressure to the shape as ds dA
shown in the dotted line in Figure 3.1. The
external force acting on an element of the
V
boundary surface of area dA is equal to the
product of the external pressure Pe and the surface
area dA, i.e. Pe dA. When the element expands
outward through distance ds, work of the force
equals Pe dA ds. Integrating over all such elements
over the entire surface, the work done equals
Pe dV where dV is the increase in the volume of FIGURE 3.1 Work done by expansion
against pressure.
the boundary surface. If the process is reversible,
system is in equilibrium at all times and the external pressure equals the pressure P exerted
against the boundary by the system and so for a reversible process
dW = PdV (3.10)
It is to be noted that work can be identified only at the boundaries of the system. It is
clear that work is a path function. Consider a system that is undergoing a reversible process,
which takes the system from the state a (refer Figure 3.2) to a state b through the path acb.
Then the work done by the system during the process acb equals
C 7C
8 Ô E8 Ô QE7 (3.11)
B 7B
40 u BASICS OF ATMOSPHERIC SCIENCE

The above turns out to be the area under the curve acb, extending up to the volume axis.
Assume that the system can be subjected to an alternate reversible process which can again
take the system from the state a to the state b through the path adb. For the second reversible
process, the work done by the system would correspond to the area under the curve adb. It is
clear that the work done by the system in any reversible process depends not only on the
initial and final equilibrium states, but also on the process (path) to which the system is
subjected, confirming that the work is a path function.
Imagine that a system found at the initial
equilibrium state a is subjected to a reversible
process along acb and then is subjected to an
alternate reversible process along bda until the
system reaches the same initial state a. Now the
entire process acbda is known as a cyclic process
and the work done during this cyclic process is
simply equal to the area acbda in Figure 3.2. The
differentials of those thermodynamic variables,
which are also properties of the system, are known
as exact differentials. A function Q is a state
function or the property of a thermodynamic
system if dQ is an exact (or total) differential. In
two dimensions, for a differential dQ = A(x, y)dx +
B(x, y)dy to be an exact differential in a certain
FIGURE 3.2 Graphical depiction of
region of the xy space, the following relationship
work done in a reversible
È ˜" Ø È ˜# Ø expansion process as de-
ÉÊ ˜Z ÙÚ ÉÊ ÙÚ must hold. Their importance lies picted in a pV diagram.
Y
˜Y Z
in the fact that if one integrates an exact differential from an initial state to a final state, then
the result of the integration depends only on the initial and final states and not on the process
by which the change from the initial to the final state was brought about, i.e. exact
differentials are independent of the path followed. Also, if one integrates an exact differential
along some closed path (cyclic process), the result must be zero. The above forms the
statement behind a theorem in advanced calculus, whose proof we will not go into.
A necessary and sufficient condition that a differential be exact is that its integral
around a closed curve is zero.
It is clear that dW is not an exact differential, since the work done along a closed path
acbda in Figure 3.2 is non-zero. Hence W or work done is not a property of the system. Work
done can either be of the following two types: (i) configuration work, and (ii) dissipation
work. Any form of work that is of the form of Y dX, where Y is an intensive variable and X is
an extensive variable is known as configuration work. It is possible that the configuration of
the system can change without the performance of work, as in the example of free expansion,
which is an irreversible process. If the work in a process cannot be expressed in terms of
change of some property of the system, then such a work is known as dissipative work. Such
dissipative work is always done on the system. Dissipative work is intimately related to
friction and dissipative processes.
ATMOSPHERIC THERMODYNAMICS u 41

3.2.3 Heat
Consider a non-adiabatic process (i.e. exchange of heat with the surroundings is permitted)
between a given pair of equilibrium states. Heat flow Q into the system in any process is
defined quantitatively as the difference between the work W in the above-mentioned non-
adiabatic process and the adiabatic work Wad in any adiabatic process between the same pair
of the equilibrium states.
Q = W – Wad (3.12)
Depending on the nature of the process, the work W may be greater or less than the
adiabatic work Wad, the sign of Q may be positive or negative. In the trivial case where the
process connecting a given pair of equilibrium states happens to be adiabatic, the
corresponding heat flow will vanish identically. By convention, heat flow into a system is
considered positive, while the heat flow out of the system is considered negative. Since the
work is not a property of the system and is an inexact differential, and from Eq. (3.12) the
heat flow is related to the difference of work and the adiabatic work, it is clear that heat flow
also like work is not a property of the system and is also an inexact differential. Furthermore,
like work, heat flow can be identified only at the boundaries of the system.

3.2.4 First Law of Thermodynamics


There are different processes by which a thermodynamic system can be taken from one
equilibrium state to another and in general the work done by the system is different in
different processes. Out of all the possible processes between the two given states let us
choose only those processes which are adiabatic. Consider the system to be enclosed by
adiabatic boundaries all around.
Furthermore, assume that the boundary
be not rigid. This would allow con-
figuration work to be done by the
system or on the system. Also, assume
that dissipation work may be done on
the system and that there are no
changes in the kinetic and potential
energies of the system.
Let the system in an initial state
a be subjected to an adiabatic-free
expansion (shown as hatched region in
Figure 3.3) from a to c. Since free
expansion is an irreversible process, it
cannot be represented on a (pV) diagram
as a curve. In a free expansion, no
configuration work is done and let it be
assumed that no dissipation work is Figure 3.3 Different possible all adiabatic pro-
performed during the process ac. Let the cesses (acb, adb and adeb) as re-
system perform a reversible adiabatic presented graphically in a pV diagram.
42 u BASICS OF ATMOSPHERIC SCIENCE

expansion from state c to state b. In this process, the area under the curve cb, extending up to
the volume axis, gives the configuration work. Furthermore, since dissipation work is zero in
any reversible process, the total work equals the area under the curve cb. Envisage a second
process again starting from the same initial state a, and assume that the system is subjected to
a reversible adiabatic expansion to a state d. The state d is chosen in such a way that a
subsequent adiabatic-free expansion (again assuming no dissipative work and shown as
hatched region in Figure 3.3) from d will terminate at the same final state b. The total work
done in this second process adb will be the area under the curve ad extending up to the
volume axis. Although the two processes (paths acb and adb), are very different, it can be
proved experimentally that the work represented by these two paths or the area under the
curves of cb and ad are exactly the same. One can even imagine a third process wherein the
system starting from the same initial state a, is subjected to a reversible adiabatic expansion,
continues beyond the point d and reaches the point e having the same configuration value
(volume here) as the final state b. To reach the final state b from the point e, the dissipative
work has to be done on the system (since both states e and b have the same configuration,
configuration work between the states e and b would be zero). The total work done in the
third process adeb would equal to the configuration work during ade (area under the curve
ade extending up to the volume axis) minus the dissipative work eb done on the system. It
turns out that the total work done in all the three processes (acb, adb, adeb) are exactly the
same and hence the work done by the system during the reversible expansion from d to e is
exactly equal in magnitude to the dissipative work done on the system during the process e to b.
In fact, the entire structure of thermodynamics is consistent with the above conclusion that
irrespective of the nature of the process.
The total work done is the same in all adiabatic processes between any two
equilibrium states having the same kinetic and potential energy.
The above remarkable statement is called the First Law of Thermodynamics.

3.2.5 Internal Energy and Enthalpy


The concepts of internal energy and enthalpy are widely used in atmospheric thermodynamics
as they are both thermodynamic properties. Internal energy can be simply thought of as the
energy stored within the atoms and molecules. The internal energy of any thermodynamic
system increases/decreases as the system gains/loses heat or work is done on the system/does
work on the surroundings. When no work is done by the system on the surroundings or vice
versa and these are under constant pressure conditions, the change in enthalpy equals the
change in internal energy.

Internal Energy
It was noted in the earlier section that the differential dW is inexact and the work done W has
different values for different processes (paths). However, the first law of thermodynamics
asserts that the differential dWad is exact in the sense that the work done is the same along all
the different adiabatic paths between a given pair of equilibrium states having the same
kinetic and potential energy. Due to the above reason, it is convenient to define a property of
the system represented by U such that the difference between the values of the property U in
ATMOSPHERIC THERMODYNAMICS u 43

the two equilibrium states a and b is equal to the total work done by the system along any
adiabatic path from a to b. The property U is known as the internal energy of the system.
Since U is a property of the system, the differential dU is an exact differential. It is
convenient to define dU, as the negative of the adiabatic work done by a system or in other
words dU is equal to the adiabatic work done on the system.
dU = – dWad
Integrating between the states a and b, one gets
6C C
Ô E6 6C  6B  Ô E8BE  8BE  6B  6C 8BE (3.13)
6B B
The total work done by a system in any adiabatic process between any two equilibrium states
a and b having the same kinetic and potential energies is equal to the decrease
(Ua – Ub) in the internal energy of the system. It is to be noted that not all states of a system
can be reached from a given state by adiabatic process. The above comment will become
clearer after the introduction of the concept of entropy. For a system undergoing a process,
not necessarily adiabatic, the first law of thermodynamics is written as
dQ = dU + dW (3.14)
where dQ, dU and dW are the increment of heat, change in internal energy and the work
done. It is to be noted that only dU is an exact differential in the above equation. Dividing
Eq. (3.14) by the mass of the system enables one to write Eq. (3.14) as
dq = du + dw = du + pda (3.15)
where dq, du and dw represent the heat added per unit mass, change in internal energy
per unit mass and work done per unit mass, and a is the specific volume. The first law of
thermodynamics simply states that the heat added to a system goes towards increasing the
internal energy of the system as well as towards performing some work by the system.
An ideal gas is made up of a very large number of small units called molecules, which
are almost always in random perpetual motion. These molecules which comprise the ideal gas
have negligibly small dimensions and have very negligibly small attractive forces with each
other. Furthermore, the collisions of these molecules with other molecules as well as with the
walls of the container are completely random events. Also, the collisions of the molecules
with one another and with the walls are elastic. Due to the above, the energy (internal energy)
of an ideal gas is a measure of the mean kinetic energy of the random motion of the
molecules, since the potential energy of interaction is negligible. Also, the pressure of an
ideal gas is the average force per unit area exerted by the random bombardment of the
molecules on the walls. The kinetic theory of matter asserts that the mean kinetic energy of
the random motion of the molecules is directly proportional to the temperature of the ideal
gas, thereby indicating a direct relationship between the internal energy and the temperature
for an ideal gas. It is clear that the internal energy of the gas is independent of any bulk
motion of the gas. If this were not true, then a cylinder containing an ideal gas taken in a
train or by road will have higher internal energy as well as higher temperature. Assume that
44 u BASICS OF ATMOSPHERIC SCIENCE

the volume of an ideal gas is kept constant when it is warmed. Since the volume is kept
constant, no work is done and the heat added to the system goes to increase the internal
energy only. The change in the internal energy of an ideal gas, by the earlier arguments, is
proportional to the temperature change of the gas, no matter how the change has occurred.
Hence the internal energy of an ideal gas is independent of its volume and hence it depends
only on its temperature. Some books define an ideal gas as one that obeys both the Boyle’s
and Charles’ law and whose internal energy is independent of volume.

Enthalpy
It was noted that the internal energy of a gas is also a state variable, i.e. a variable which
depends only on the state of the gas and not on any process that produced that state. One is
free to define many more additional state variables which are combinations of existing state
variables. Such new variables often make the analysis of a system much simpler. For a gas, a
useful additional state variable is the enthalpy (H), which is defined to be the sum of the
internal energy U plus the product of the pressure p and the volume V, i.e. H = U + pV.
Dividing by the mass can make the enthalpy into an intensive, or specific variable called
the specific enthalpy (h). Consider a system (gas) with heat transfer dQ and work dW. The
change in the internal energy for such a system from the first law of thermodynamics is
given by
U2 – U1 = dQ – dW
where the subscript 2 and 1 refer to the final and the initial states of the system. It is to be
noted that the heat transfer and work done depend on the process. For the special case of a
process where the pressure is constant (isobaric process), the work done by the system (gas)
is given as
dW = p (V2 – V1)
Substituting the above relation in the statement of the first law of thermodynamics and
equating the quantities with their respective states, one gets
(U2 + pV2) – (U1 + pV1) = dQ
Using the definition of the enthalpy, H = U + pV, one gets
H2 – H1 = dQ (3.16)
Hence for systems at constant pressure, the change in enthalpy is the heat received by the
system plus any non-mechanical work that has been done. Therefore, the change in enthalpy
can be envisaged without the need for compressive or expansive mechanics; for a simple
system, with a constant number of particles, the difference in enthalpy is the maximum
amount of thermal energy derivable from a thermodynamic process in which the pressure is
held constant.

3.2.6 Specific Heat Capacity


By definition, the heat capacity (denoted by C) of a body is the amount of heat required to
raise its temperature by 1 degree and has units of Joules K–1. The specific heat capacity of a
ATMOSPHERIC THERMODYNAMICS u 45

substance is the amount of heat required to raise the temperature of 1 kilogram of the
substance through 1 degree. Essentially, it is the heat capacity per unit mass of the substance.
The amount of heat necessary to raise the temperature of a unit mass (1 kg) of an ideal gas
(or for that matter any gas) by one degree will vary for the same gas, depending on whether
it is allowed to expand or not. When the gas is not allowed to expand (i.e. at constant
volume), there is no work of expansion and all the heat added would go towards increasing
the internal energy of the gas. However, if there are no constraints regarding constancy of
volume, the amount of heat added will go towards increasing the internal energy as well as
towards the work of expansion. From the above discussion, it is obvious that the amount of
heat required to raise the temperature of a unit mass of a gas through unit temperature raise
will be lower for situations under constant volume than under constant pressure. Thus, the
specific heat capacity of an ideal gas at constant pressure (denoted by cp) is always higher
than the specific heat capacity of an ideal gas at constant volume (denoted by cv). Here, cp
and cv are defined as
È ER Ø È ER Ø
DW ÉÊ ÙÚ  D Q ÉÊ ÙÚ (3.17)
E5 W E5 Q

where the subscripts v and p above refer to volume and pressure being held constant. For an
isosteric process (constant specific volume process)
pda = 0, and
dq = du in Eq. (3.15) results in
È EV Ø
cv = É Ù BOE
Ê E5 Ú
the first law of thermodynamics for an ideal gas can be written as
dq = cvdT + pda (3.18)
Another form of the first law of thermodynamics for an ideal gas can be written using
the differentiated form of the equation of state
pda + adp = RspdT
where, Rsp is the specific gas constant. Substituting the above in Eq. (3.18), one gets
dq = (cv + Rsp)dT – adp (3.19)
For an isobaric process, dp = 0 and using the relation of cp from Eq. (3.17), one gets

È ER Ø
DQ ÉÊ ÙÚ DW  3TQ (3.20)
E5 Q
Also, for an isobaric process, cp can be defined in terms of specific enthalpy as dh = cp T.
The value of cp and cv for dry air is given by 1004 J deg–1 kg–1 and 717 J deg–1 kg–1,
respectively. By combining Eq. (3.19) and Eq. (3.20), one gets another form of first law,
dq = cpdT – adp (3.21)
46 u BASICS OF ATMOSPHERIC SCIENCE

3.3 ADIABATIC PROCESSES


Although the concept of an adiabatic boundary and adiabatic process have been introduced in
the earlier section, it is important to deal with adiabatic processes in the atmosphere in more
detail. It is important to clearly define the concept of an air parcel. Air parcel can be defined
as an imaginary body of air to which may be assigned any or all of the basic dynamic and
thermodynamic properties of atmospheric air. Furthermore, a parcel is assumed large enough
to contain a very great number of molecules, but small enough so that the properties assigned
to it are approximately uniform within it. Also, it is assumed that the motions of the air parcel
with respect to the surrounding atmosphere do not induce marked compensatory movements.
Radiation, turbulent transfer of heat, latent heat of condensation of water vapour as well as
friction can all contribute to adding an increment of heat to an air parcel. There is evidence
that for time periods up to a day or two the above-mentioned heating processes are not
sufficiently important in adding an increment of heat to an air parcel and hence one can
investigate the impact of the first law of thermodynamics for the adiabatic processes. The first
law of thermodynamics (Eq. (3.21)) for an adiabatic process, becomes
0 = cp dT – adp (3.22)

3TQ5
From the equation of state for an ideal gas, B and substituting for a in
Q
Eq. (3.22), one gets
E5 3TQ EQ
  (3.23)
5 DQ Q
Representing Rsp/cp = k, Eq. (3.23) is integrable and one obtains
T = const pk (3.24)
Equation (3.24) is known as the Poisson equation and is used to define the potential
temperature (q ). For an ideal gas undergoing adiabatic process, the const in Eq. (3.24)
depends on the initial values of pressure and temperature and if one assumes that the initial
pressure in 1000 hPa and the initial temperature has a value of q, then Eq. (3.24) becomes
L
5 È Q Ø
ÉÊ Ù (3.25)
R  Ú
The potential temperature (q) is defined as the temperature a parcel of ideal gas would have
if it were expanded (or compressed) adiabatically from a given state characterized by the
pressure p and temperature T to a standard pressure of 1000 hPa. The potential temperature
is obviously a property of the parcel of air, which is invariant during the adiabatic process. A
property, which is invariant during a process, is defined as a conservative property associated
with that process. The value of k for dry air is about 0.286. Isolines of constant potential
temperature are known as dry adiabats.
ATMOSPHERIC THERMODYNAMICS u 47

3.4 MOIST THERMODYNAMICS AND LATENT HEATS


Processes involving water in their liquid phase as well as in their other phases play a very
important role in the thermodynamics of the atmosphere. Clouds having condensation
products such as water and or ice affect the incoming and the outgoing radiation.
Furthermore, heat in the form of latent heat is either taken in or liberated out due to phase
changes associated with water. Such heating of the atmosphere due to latent heat release can
affect the circulation and impact on the atmospheric dynamics.

3.4.1 Measures of Water Vapour in Air


It would be convenient to first familiarize ourselves with the various measures of water
vapour in the atmosphere. The most important measures of moisture in air are given below.
(i) Saturation vapour pressure (es): Imagine a closed container containing some
pure water at its bottom. If the air in the container is dry initially, water will
evaporate and the pressure of the water vapour will start increasing. Some of the
water vapour molecules will also return to the liquid phase. The evaporation
process in the above closed container will continue until there are as many
molecules returning to the liquid as there are escaping. The above condition
corresponds to an equilibrium condition wherein the rates of evaporation of the
molecules from the water equal the rate of condensation on the water from the
moist air above it. Once the above equilibrium state is reached, the air is said to be
saturated with respect to a plane surface of pure water. The saturation vapour
pressure value depends only on temperature and it increases with the increase of
temperature.
(ii) Absolute humidity (rv): It is defined as the mass of water vapour in 1 m3 of air.
(iii) Mixing ratio (w): It is defined as the ratio of the mass of water vapour present to
the mass of dry air in a certain volume of air. Hence,

X NW
NE
Saturation mixing ratio (ws) is the value of the mixing ratio at saturation
conditions. It is clear that the mixing ratio is unitless. However, due to the small
amounts of water vapour present in a parcel of air, in practice the mixing ratio is
conveniently expressed as the number of grams of water vapour per kilogram of
dry air.
(iv) Specific humidity (q): It is defined as the ratio of the mass of water vapour
present to the mass of moist air (i.e. mass of dry air plus the mass of water vapour)
in a certain volume of air. Hence,

R NW NW
N NW  NE
Like the mixing ratio, the specific humidity also is unitless. However, due to
the small amount of water vapour present in a parcel of air, the specific humidity is
48 u BASICS OF ATMOSPHERIC SCIENCE

also conveniently expressed in practice as the number of grams of water vapour per
kilogram of moist air. Saturation specific humidity (qs) is the value of the specific
humidity at saturation conditions.
(v) Relative humidity (r): It is the ratio of the actual mixing ratio of a parcel of air
at a given temperature and pressure to the saturation mixing ratio at that
temperature and pressure. Hence,

S X
XT
The relative humidity is usually multiplied by 100 and is expressed as a percentage.

3.4.2 Equation of State of Moist Air


In the range of atmospheric conditions that we would be interested in, it is perfectly possible
to assume that water vapour satisfies the ideal gas law. The specific gas constant for water
vapour (Rv) would be simply the ratio of the universal gas constant to the molecular mass of
water and has a value 461 J deg–1 kg–1. For our purposes, it is sufficiently accurate to assume
that both dry air and water vapour separately satisfy the equation of state for an ideal gas.
Hence it would be important to obtain the equation of state for moist air (i.e. a mixture of dry
air and water vapour). While deriving the equation of state for dry air, we defined a mean
molecular mass for the dry air by taking the weighted harmonic mean of the molecular masses
of the constituents (Eq. 3.8) of the dry air. In a similar manner the mean molecular mass of
moist air can be defined as

  Ë NE N Û
Ì  WÜ (3.26)
. NPJTU NE  NW Í . E . W Ý
where the subscripts d and v refer to dry air and water vapour, respectively. The above

expression can be written in term of Md, mixing ratio X ÈÉ X NW ØÙ and the ratio of the
Ê NE Ú
molecular masses
. W
F as
. E

  NE Ë NW NE Û

= Ì  (3.27)
. NPJTU . E NE  NW
Í . W . E ÜÝ


    X F

= (3.28)
. NPJTU .E   X

Using the above expression for the mean molecular mass of moist air, the equation of state
for moist air can be written as

QB 3   XF
5 3   XF
5
(3.29)
.E   X

TQ
  X

ATMOSPHERIC THERMODYNAMICS u 49

Hence the equation of state for the moist air has the same form as the equation of state for
dry air except for the correction to the specific gas constant for dry air. Instead of dealing
with a variable gas constant involving the mixing ratio, it would be convenient to incorporate
the above change, in the temperature and define a new temperature, known as the virtual
temperature

5W   XF
5 (3.30)
  X

Substituting e = 0.622, one gets Tv = T(1 + 0.61w).


The virtual temperature is defined as the temperature that dry air must have in order to
have the same density as that of the moist air at the same pressure. Since the moist air is less
dense than the dry air, the virtual temperature is always greater than the temperature. The
maximum difference in Tv and T can never be greater than 7 K and is usually much less.
Using Tv, the equation of state for moist air becomes
pa = RspTv (3.31)

3.4.3 Latent Heat


Under conditions involving a change of phase, heat may be supplied to (or taken away from)
a substance without altering its temperature. The quantity of heat supplied to or taken away
per unit mass of the substance is called the latent heat of phase change. The latent heat of
vaporization (Lvap) is the amount of heat required to convert a unit mass (1 kg) of a substance
from the liquid to the vapour phase without a change in temperature. The latent heat of
melting (Lmelt) is the heat required to convert a unit mass of a substance from the solid to the
liquid phase without a change in temperature. Lvap and Lmelt for water at normal atmospheric
pressure and 0°C have values of 2.5 ´ 106 J kg–1 and 3.3 ´ 105 J kg–1, respectively. The latent
heat of condensation (Lcond) and latent heat of fusion (Lfus) have the same values as that of
Lvap and Lmelt. The latent heat associated with change of phase from solid to vapour phase is
known as the latent heat of sublimation (Lsub) and is the amount of heat required to convert a
unit mass of a substance from the solid to vapour phase without a change in temperature.
From this above discussion, it is clear that
Lsub = Lmelt + Lvap (3.32)

3.5 HYDROSTATIC EQUILIBRIUM


For most situations in the atmosphere, the upward directed force acting over a thin layer of air
due to the decrease in pressure with height is in very close balance with the downward force
due to gravity. When an exact balance exists between the above two forces, the atmosphere is
said to be in hydrostatic balance or hydrostatic equilibrium. Consider a vertical column of air
between heights z and z + dz (Figure 3.4) with unit cross-sectional area. The downward force
acting on this column due to the weight of the air is g r dz, where r is density of the air at this
height and g is the acceleration due to gravity. The upward pressure on the lower portion of
the vertical column is higher than the downward pressure at the upper portion of the vertical
50 u BASICS OF ATMOSPHERIC SCIENCE

column since pressure decreases with height. The net force on the vertical column due to
differences in pressure on the upper and lower portions is hence upward and is denoted by the
positive quantity –dp. Hydrostatic balance then requires that –dp = g r dz, or
EQ
 HS (3.33)
E[
The above equation is known as the hydrostatic equation and is applicable to all
meteorological situations with negligible or very small vertical accelerations.

FIGURE 3.4 Change of pressure with height for a vertical column of air.

3.5.1 Geopotential
Since acceleration due to gravity which is the resultant of gravitational force and centrifugal
force can change with latitude, it is convenient to define a new variable df = gdz, where f is
known as the geopotential. The geopotential is the gravitational potential per unit mass and is
defined at any point in the atmosphere as the work that must be done to raise a unit mass
(1 kg) of air from sea level to that point against the earth’s gravitational field. The unit of f
is J kg–1. One can define a unit of geopotential called the geopotential height (Z) by the
following relation
[
G [

;
H H
Ô H E[ (3.34)


where g0 is taken as 9.8 m s–2 and is the globally averaged acceleration due to gravity at the
earth surface. It is to be noted that Z is a unit of energy per unit mass and has been so
defined so that Z in these energy units is numerically quite close to the geometric height in
metres.
ATMOSPHERIC THERMODYNAMICS u 51

3.5.2 Scale Height and Height Computations Using the


Hypsometric Equation
Density is not a quantity normally measured in the atmosphere and hence it is convenient to
eliminate r from the hydrostatic equation, Eq. (3.33) using Eq. (3.1)

EQ  QH
E[ 3TQ5W
EQ  HE[
Q 3TQ5W
EQ
EG  3TQ5W (3.35)
Q
Integrating the above relation between two pressure levels p1 and p2 with appropriate
geopotentials f1 and f2 respectively, gives
Q
EQ
G  G  3TQ Ô 5W Q
Q

Dividing both sides by go, one gets


Q
3TQ EQ
;  ;  Ô 5W Q (3.36)
H Q

(i) For a homogeneous atmosphere (density constant with height), integrating the
hydrostatic equation with the depth of the atmosphere
 )
Ô EQ  SH Ô E[  Q S H) (3.37)
Q 

where H = p0 (rg)–1; and from the equation of state, H = Rsp T0 g–1, H is called the
height of the homogeneous atmosphere and has a value of about 8 km for T0 = 273 K.
Since pressure decreases with height and density is constant with height, the
equation of state requires that temperature decrease with height. The rate of
decrease of temperature with height known as temperature lapse rate g for a
homogeneous atmosphere is obtained from differentiating the equation of state with
respect to z, to get
EQ E5
S 3TQ (3.38)
E[ E[
Substituting the hydrostatic equation in the above expression,

È E5 Ø H
H IPNP É Ù  , LN  (3.39)
Ê E[ Ú 3TQ
52 u BASICS OF ATMOSPHERIC SCIENCE

The above temperature lapse rate is about six times larger than the observed lapse
rate, the difference attributed to the fact that the density actually decreases with
height.
(ii) For an isothermal and dry atmosphere
Q
3TQ5W EQ Q
Z2 – Z 1 =  Ô Q ) MO (3.40)
H Q
Q
Ë  ;  ;
Û
p2 = Q FYQ Ì ÜÝ (3.41)
Í )
3TQ5W
where ) 5W is called the scale height. For every increase of H in
H
geopotential height, the pressure decreases by a factor e (= 2.718). For an iso-
thermal and dry atmosphere, the pressure decreases exponentially with height.
Hence there is no definite upper boundary to this atmosphere. The temperature
lapse rate for isothermal atmosphere is by definition zero.
(iii) For a dry atmosphere with constant temperature lapse rate, the temperature T
varies linearly with height and is given by
T = T0 – g z (3.42)
where T0 is the temperature at sea level and g is the constant temperature lapse rate.
The hydrostatic equation with density replaced from the equation of state becomes
 QH
dp = E[ (3.43)
3TQ5W
EQ H E[
= (3.44)
Q 3TQ 5  H [

Integrating the above equation between the levels z = 0 when p = p0 and an


arbitrary height z where the pressure is p, one gets
Q [
EQ H E[
Ô Q
=
3TQ Ô 5  H [

Q 

Q H Ë 5  H [ Û H 5
MO = MO Ì Ü MO
Q 3TQH Í 5 Ý 3TQH 5
H3TQH
Ë5 Û
p= Q Ì Ü (3.45)
Í5 Ý



For normal tropospheric situations (g > 0, temperature decreases with height), from
the above relation one obtains a decrease of pressure with height. For cases of
ATMOSPHERIC THERMODYNAMICS u 53

inversion (g < 0, temperature increases with height), T > T0, the exponent is
negative and again the pressure decreases with height. An atmosphere with a
constant positive temperature lapse rate has only a finite vertical extent while an
atmosphere with negative temperature lapse rate has an infinite vertical extent with
no upper limit.
(iv) For a dry adiabatic atmosphere, the temperature lapse rate is obtained by
logarithmic differentiation of the Poisson equation
3TQ D Q
5 Ë Q Û
= Ì , gives
R Í ÜÝ
 E5 3TQ EQ
=
5 E[ D Q Q E[
Substituting from the hydrostatic equation and rearranging terms, one gets

È E5 Ø H Ë 3TQ5 S Û H
ÉÊ ÙÚ =  Ì Ü 
E[ BE DQ Í Q Ý DQ
È E5 Ø H
or gad = É Ù
Ê E[ Ú BE (3.46)
DQ
A sample of air subjected to a dry adiabatic process has a constant potential
temperature and whose temperature will decrease at a rate of gad, i.e. at a rate of
9.8 K km–1.
(v) The real atmosphere is neither dry nor homogeneous in density or temperature. In
such situations it may be possible to define a mean virtual temperature <Tv> with
respect to ln p for any two pressure levels as shown in Figure 3.5 as
MO Q Q
EQ
Ô 5W E MO Q
Ô 5W Q
Q Q
 5W !
MO

Q (3.47)
MO
ÈQ Ø
Ô

E MO Q
MO
ÉÊ Q ÙÚ

MO Q

From Eq. (3.47), one gets after substituting in Eq. (3.36)

3TQ ÈQ Ø ÈQ Ø
;  ;  5W ! MO É  Ù  ) ! MO É  Ù (3.48)
HP Ê Q Ú Ê Q Ú

3TQ  5W !
where the scale height < H > is defined as   5W !  Equation (3.48) is
HP
called the hypsometric equation. The difference in the geopotential heights Z2 – Z1
between any two-pressure levels in the atmosphere, called, as the thickness of the
layer is proportional to the mean virtual temperature of the layer.
54 u BASICS OF ATMOSPHERIC SCIENCE

A B
ln P2

C
ln P1 E
D
ln P

Virtual Temperature (K) Tv

FIGURE 3.5 Mean virtual temperature calculation using


graphical methods for an actual sounding.

3.5.3 Reduction of Pressure to Sea Level


One of the important components of weather and synoptic analysis is to construct the isolines
of pressure, since horizontal variation of pressure have important effects on the future
evolution of weather system. It is known that measurements of surface station pressure over
different meteorological stations cannot be compared directly as these stations are located at
widely varying altitudes above the sea level. Also since the variation of pressure with height
is large as compared to the horizontal variation, it is very essential to account for the different
altitudes of the meteorological stations by reducing the observed station pressure to a common
fixed reference height such as the mean sea level.
To implement the above reduction of observed station pressure at a known height to
mean sea level, one assumes the existence of a hypothetical column of air extending from the
station height to the mean sea level and calculates the pressure difference between the top and
the bottom of this hypothetical layer using the hypsometric equation, Eq. (3.48) as
Q
;H  ) ! MO
QH
where Zg is the station altitude, <H> is the scale height and p0 and pg refer to mean sea level
pressure and the observed station pressure. The above equation can be used to get the sea
level pressure
È ;H Ø È HP ; H Ø
Q QH FYQ É QH FYQ É Ù
Ê  ) ! ÙÚ Ê 3TQ  5W ! Ú
;H ;
For   the exponential term can be approximated by   H and
)! )!
ATMOSPHERIC THERMODYNAMICS u 55

;H È H ; H Ø
Q  QH  QH É Ù (3.49)
)! Ê 3TQ  5W ! Ú
With < H > » 8 km and pg » 1000 hPa, the above approximation provides for a reduction of
sea level pressure by 1 hPa for every 8 m of altitude at levels close to mean sea level.

3.6 THERMODYNAMIC DIAGRAM


Thermodynamic diagrams provide a convenient means to represent major atmospheric
processes, which take place in the atmosphere. As energy changes associated with the
atmospheric process are important, it is desirable that an area in a thermodynamic diagram
represent changes in energy or work done associated with that process. The thermodynamic
diagrams, which satisfy the above requirements, are known as energy diagrams. A second
desirable requirement of a thermodynamic diagram is that the lines that characterize the
various atmospheric processes such as isothermal, isobaric, dry adiabatic etc. are straight. A
third and further requirement of a thermodynamic diagram is that the angles between the
isotherms (lines of constant temperature) and the dry adiabats (lines of constant potential
temperature) be as large as possible.

3.6.1 Emagram
A thermodynamic diagram with temperature (T) as the abscissa and a logarithmic scale of
pressure (– Rsp ln p) along the ordinate decreasing upward as in the atmosphere is known as
the Emagram. In this diagram, both the isotherms and isobars are straight lines and are
perpendicular to one another. A typical area in an Emagram will be of the form Td (Rsp ln p)
EQ
3TQ5 B EQ satisfies the requirement of an energy diagram. The dry adiabats or
Q
constant potential temperature lines are logarithmic curves as can be seen by taking the
logarithm of Poisson equation
DQ
MO Q  MO 5 DPOTU
3TQ

The angle between the isotherms and dry adiabats in an emagram is about 45°.

3.6.2 Tephigram
The tephigram is the most widely employed among all thermodynamic diagrams. The
tephigram has temperature (T) as the abscissa and entropy (s), a logarithmic scale of potential
temperature (cP lnq) along the ordinate, increasing upward as in the atmosphere. In a
tephigram, both the isotherm and dry adiabats are straight lines and are perpendicular to one
another. This satisfies the requirement of the maximum angle between the isotherm and dry
adiabat. An area in the tephigram, like the emagram, has units of J kg–1, and is also an
energy diagram. Also, the saturation mixing ratio lines are nearly straight in the tephigram.
56 u BASICS OF ATMOSPHERIC SCIENCE

3.6.3 Skew T–log p Diagram


An improvement over the emagram so as to ensure that the angle between the isotherms and
dry adiabats is nearly 90°, results in a new thermodynamic diagram known as Skew T–log
pressure diagram. The ordinate of this diagram is Rsp ln p and the abscissa is T + k ln p,
where k is a constant. An area in this diagram also represents energy and the angle between
the isotherms and dry adiabats is nearly 90°.

3.6.4 Stuve Diagram


È 3TQ Ø
É Ù
Ê DQ Ú
This thermodynamic diagram has temperature (T) on a linear scale as the abscissa and Q
as the ordinate with pressure increasing downward as in the atmosphere. Unlike the other
thermodynamic diagrams, an area in the Stuve diagram does not represent energy and hence it
is not an energy diagram. The angle between the isotherm and the dry adiabats for the Stuve
diagram is nearly 45°.

3.7 HYDRODYNAMIC STABILITY—PARCEL AND SLICE


METHODS
The discussion, which follows in this section, will assume that the atmosphere is in
hydrostatic balance. In such an atmosphere under hydrostatic equilibrium it is important to
study the hydrostatic stability of a sample of air subjected to vertical displacements under
adiabatic conditions. The criteria for hydrodynamic stability of a sample of air undergoing
dry/saturated adiabatic vertical displacements would involve use of dry/saturated adiabatic
lapse rate and hence there is a need to derive an expression for the saturated adiabatic lapse
rate. When unsaturated air is subjected to vertical adiabatic motion, its temperature changes at
a rate of gad and its potential temperature is conserved.

3.7.1 Saturated Adiabatic and Pseudoadiabatic Processes


When a sample of moist unsaturated air rises in the atmosphere adiabatically, its temperature
decreases with height at the rate of gad (9.8 K km–1) until the sample of air becomes saturated
with water vapour. Further, lifting of this saturated air leads to a change of phase in the form
of condensation of liquid water or deposition of ice, which releases latent heat of
condensation or latent heat of deposition. Due to the above release of latent heat, the rate of
decrease in the temperature of the rising sample of air becomes less, indicating that the
saturated adiabatic lapse rate is smaller than gad. Now two different possibilities exist; one, the
condensation products remain in the sample of air and the process may be considered
adiabatic and reversible, despite the release of latent heat provided no heat passes through the
boundaries of the sample of air. In such a situation, the air is said to undergo a saturated
adiabatic process. In the second possibility, all the condensation products fall out of the
sample of air as and when they form, making the process irreversible as well as strictly not
adiabatic since the heat content associated with the condensation products are removed from
ATMOSPHERIC THERMODYNAMICS u 57

the system. The air in the second situation is said to undergo a pseudoadiabatic process.
However, since the amount of heat carried away by the condensation products is quite small,
as compared to the air sample, the saturated adiabatic lapse rate is more or less the same as
the pseudoadiabatic lapse rate.

3.7.2 Saturated Adiabatic Lapse Rate


To derive an expression for the saturated adiabatic lapse rate (gs), the first law of
thermodynamics is used along with the hydrostatic equation as
dq = cp dT + gdz (3.50)
If the saturated mixing ratio of the air is ws, the quantity of heat dq released into (or absorbed
from) a unit mass of dry air due to condensation (or evaporation) of liquid water during
vertical ascent (or descent) of the sample of saturated air is – L dws, where L is the latent heat
of condensation/evaporation. Hence Eq. (3.50) becomes
– L dws = cp dT + gdz (3.51)
By neglecting the warming (or cooling) of the small amounts of water vapour which are
present in the sample of the saturated air during the release (or absorption) of latent heat, the
cp in the above equation (Eq. (3.51)) will specifically refer only to the specific heat at
constant pressure of dry air only. Dividing Eq. (3.51) by gdz and after some rearrangements of
terms, one gets
E5 - EXT H
 
E[ D Q E[ DQ
which can be written as
E5 - EXT E5 H
 
E[ D Q E5 E[ D Q
From the above, the saturated adiabatic lapse rate (gs) is defined as
E5 H BEJ
HT  (3.52)
E[ È - Ø È EX Ø
 É Ù É T Ù
Ê D Q Ú Ê E5 Ú
EXT
Since is always positive, gs < gad, the largest difference with gs » 4 K km–1 between the two
E5
occurring in the warm humid air near the ground. Also, gs has typical values of 6–7 K km–1
in the middle tropospheric region. Lines, which depict the change in temperature with height
of a sample of air undergoing, saturated adiabatic or pseudoadiabatic vertical displacements
are known as saturated adiabats or pseudoadiabats. The pseudoadiabats are depicted as
curved lines in all the four thermodynamic diagrams discussed in the earlier section.

3.7.3 Equivalent Potential Temperature


Dividing the first law of thermodynamics by temperature and using the equation of state, yields
58 u BASICS OF ATMOSPHERIC SCIENCE

ER E5 EQ
DQ  3TQ (3.53)
5 5 Q
Taking the logarithmic derivative of the Poisson equation, one gets
ER E5 EQ
DQ DQ  3TQ (3.54)
R 5 Q
Equating Eqs. (3.53) and (3.54) and using dq = –L dws, one gets
- ER
 EX (3.55)
D Q5 T R
-
Assuming is independent of temperature, and employing an order of magnitude argument
DQ
E5 EX T
for small incremental displacements along saturated adiabats it can be verified that 
5 XT
-
The validity of the above inequality together with the independence of temperature of
DQ
leads to the following approximate relation:

- È -XT Ø
EX EÉ Ù (3.56)
D Q5 T Ê D Q5 Ú
Using Eq. (3.56) in Eq. (3.55), one gets after integration
-XT
 MO R  DPOTUBOU
D Q5
The constant of integration is determined by requiring that at low temperatures
XT   R  R
5
F  )FODF

-XT ÈRØ
 = MO
ÉÊ R ÙÚ
D Q5 F

È -XT Ø
qe = R FYQ É Ù (3.57)
Ê D Q5 Ú
The quantity qe is called the equivalent potential temperature and is simply the potential
temperature q of a sample of air when its saturation mixing ratio ws is zero. The equivalent
potential temperature (qe) can be calculated for a sample of air using any of the thermodynamic
diagrams as follows. The air from the initial state is expanded through a pseudoadiabatic
process until all the water vapour has condensed and the condensed products are removed
from the sample. The air is then subjected to dry adiabatic compression (along a dry adiabat)
ATMOSPHERIC THERMODYNAMICS u 59

to a standard pressure of 1000 hPa. The temperature of the above sample of air at 1000 hPa
is then qe. If the sample of air is initially unsaturated, ws and T are the appropriate saturation
mixing ratio and temperature at the point where the sample of air first becomes saturated after
being lifted dry adiabatically. The equivalent potential temperature is conserved during both
the dry and saturated adiabatic processes.

3.7.4 Stability Criteria Using Parcel Method


A parcel is a sample of air and the hydrostatic stability of such a parcel subjected to vertical
adiabatic displacements is being discussed in this section. The following assumptions are
assumed to hold while outlining the stability criteria using the parcel method:
(i) There are no compensating motions in the environment associated with the parcel
motion.
(ii) The parcel retains its individual identity and does not mix with its environment. It is
true that neither of the above two assumptions hold strictly in the atmosphere.
However, the above assumptions ensure that the approach to deriving the stability
criteria remains simple.
Consider an atmosphere in hydrostatic equilibrium having a certain observed virtual
temperature lapse rate. Envisage a small parcel of air (which forms part of the environment),
to have the same pressure, temperature and density as that of the surroundings. Assume that
the above parcel is subjected to a small upward displacement. If the air parcel is unsaturated,
the parcel will expand, as it encounters lower atmospheric pressure. The energy towards the
work of expansion is taken from the internal energy, resulting in a decrease of its virtual
temperature, assuming that the upward displacement is adiabatic. The parcel will cool at the
rate of 9.8 K km–1, the dry adiabatic lapse rate gad. If the surrounding (observed) lapse rate g
is less than gad, the air parcel will be at a lower virtual temperature than its new immediate
surroundings. One can easily see that the pressure in the air parcel will adjust quickly to the
surrounding pressure, for otherwise sound waves will manifest and propagate bringing out
equalization of pressure. This causes the air parcel to be less buoyant and denser than the
surroundings resulting in the parcel sinking back to its original level. The above situation
represents stable conditions. The parcel will reach its original level at the same temperature
and density with which it started. However, the small downward momentum of the parcel will
carry it further beyond the original level. The parcel, which will now be warmed at the rate of
gad will find itself warmer (since g < gad) and lighter than the surroundings. This will lead to
a net upward motion and using the same arguments outlined above, the parcel will undergo
vertical oscillations about the original level. For an air parcel subjected to downward adiabatic
displacements with g < gad, similar arguments as outlined above will indicate that the air
parcel will again undergo vertical oscillations about the original level. This suggests that
g < gad will provide for stable conditions for both upward and downward adiabatic
displacements of unsaturated air.
For the situation g > gad, a parcel displaced vertically will find itself warmer than the
new immediate surroundings and hence more buoyant. The air parcel will continue to move
upward and will never return to its original level. This situation represents the unstable
conditions. For the air parcel subjected to downward displacements with g > gad, the air parcel
60 u BASICS OF ATMOSPHERIC SCIENCE

will continue to move down signifying the unstable condition. For the situation g = gad , an air
parcel subjected to upward or downward adiabatic displacements will have the same
temperature and density as that of the new immediate surroundings. There will be no net
force in either upward or downward direction signifying neutral conditions.
The above qualitative discussion can be put in a quantitative framework as follows. The
environment is assumed to be in hydrostatic balance and hence
 ˜Q „
 H  (3.58)
S „ ˜[
The properties of the environment are denoted by primes, while the same for the air parcel
will be unprimed. Newton’s second law of motion for the air parcel is then

E[  ˜Q
H  (3.59)
EU  S ˜[
Since the pressure of the parcel adjusts quickly to the environment,

E [  ˜Q „
H  (3.60)
EU  S ˜[
˜Q „
Elimination of from Eqs. (3.58) and (3.60), one gets
˜[
E [ HS „ È S  SØ
h
H  HÉ
EU  S Ê S ÙÚ
Substituting from the equation of state for moist unsaturated air, one gets

E[ È 5  5W„ Ø
HÉ W (3.61)
EU  Ê 5W„ ÙÚ
Expanding the virtual temperature in a Taylor series expansion and retaining only up to the
linear term
E5 „
5W„ [
5W„ [
 [  [
W  "
E[
where z = z0 is the original level and z = z is the level to which the parcel has risen. Since
5W„ [
5W [
and taking z0 = 0 without any loss of generality, one gets

E5W„
5W„ [
5W 
 [ 5W 
 H [ 5W   H [ (3.62)
E[
E5W„
where  H is the environmental lapse rate.
E[
The dry adiabatic lapse rate gad is the rate at which the air parcel is cooling
5W [
5W 
 H BE [ 5W   H BE [ (3.63)
ATMOSPHERIC THERMODYNAMICS u 61

Substituting, Eqs. (3.62) and (3.63) in Eq. (3.61), one gets



E [ Î  H BE [  H [ Þ Î H  H BE
[ Þ

HÏ ß HÏ ß (3.64)
EU Ð 5W   H [ à Ð 5W   H [ à
  Î H[Þ H[
But,  Ï  ß TJODF is a small quantity. Using the above
5W   H [
5WP Ð 5W  à 5W 

E [
H Ë H  H BE

Ì H  H BE
[  H[ Ü (3.65)
5WP Í
EU  5W  Ý
Neglecting quadratic and higher order terms in z, one gets

E [ H
 H BE  H
[  (3.66)
EU 
5W

H
If the coefficient of [ H BE  H
is positive, then the solution of Eq. (3.66) is a sinusoidal
5W 
function of time. Hence, the parcel for the stable condition (g < gad) will oscillate about its
original level with a period given by
Q
U 5W (3.67)
H H BE  H

If the coefficient of z is negative (g > gad), which corresponds to unstable conditions, the
solution will be expressed in terms of exponentials of time. Further, for the case of the
coefficient of z being zero (g = gad), which corresponds to neutral conditions, there is no net
force on the air parcel.
The qualitative discussion on the various stability criteria for unsaturated air can be very
easily extended for saturated air. The saturated air subjected to vertical upward displacement
will cool according to the saturated adiabatic lapse rate gs and the environmental lapse rate
need to be compared with gs to deduce the stability criteria. Since gs is always less than gad, it
is easier to achieve instability for saturated air as compared to unsaturated air.
Consider the following situation when the environmental lapse rate lies between gs and
gad, i.e. gs < g < gad. The parcel of air if unsaturated will be stable to vertical upward
displacements since g < gad. However, if the parcel is initially saturated or if it is unsaturated
and is lifted sufficiently to saturation, then it will cool according to gs and the air parcel will
be unstable since g > gs. This situation where the parcel is stable/unstable to unsaturated/
saturated vertical lifting processes is known as conditional instability.
There are also situations when the entire atmospheric layer may be lifted up or lowered
down. Examples of this are the forced ascent/descent of air across a mountain barrier. It
would be important to discuss the associated stability/instability conditions when an entire
layer is lifted up or lowered down. It is assumed that during the entire lifting, no additional
mass of air is brought in or removed from this layer. This would ensure that the vertical
pressure difference between the top and the bottom of the layer remains unchanged during the
lifting process.
62 u BASICS OF ATMOSPHERIC SCIENCE

Consider first the unsaturated process where the entire atmospheric layer of unsaturated
air is lifted up. Taking the logarithmic derivative with respect to height of the Poisson equation
 ˜R „  ˜5 „ L ˜Q „

R „ ˜[ 5 „ ˜[ Q „ ˜[
3TQ
where, L and where primed quantities refer to the environment.
DQ
Substituting from the hydrostatic equation and the equation of state, one gets

 ˜R „  ˜5 „ 3TQ S „ H  ÎÑ ˜5 „ H ÞÑ
=   (3.68)
R „ ˜[ 5 „ ˜[ Q„ DQ 5 „ ÏÑÐ ˜[ D Q ßÑà
 ˜R „ 
= \H  H ^ (3.69)
R „ ˜[ 5 „ BE
H
where H BE . The stability consideration of unsaturated parcel depended on the sign of
DQ
˜R „
(gad – g), i.e. on the sign of  Hence, the stability characteristics of unsaturated parcel can
˜[
also be restated in terms of potential temperature of the environment.
˜R
> 0, indicating stable conditions
˜[
˜R
< 0, indicating unstable conditions (3.70)
˜[
˜R
= 0, indicating neutral conditions
˜[
It is clear from the above that the stability characteristics can be immediately discerned from
the plot of the observed sounding in the Tephigram from the nature of the slope, since the
ordinate of the Tephigram happens to be the logarithm of potential temperature. Since the
pressure differs across the top and the bottom of the layer lifted does not change during the
lifting process, it is better to rewrite Eq. (3.69) in pressure coordinates, as
˜R „ ˜ R „ ˜[
˜Q ˜[ ˜Q „
Substituting from the hydrostatic equation, one gets
 ˜R „ 
\H BE  H ^ (3.71)
R „ ˜Q „ H S „5 „
Using the equation of state in Eq. (3.71)
 ˜R „  3TQ
\H BE  H ^ (3.72)
R „ ˜Q „ HQ „
ATMOSPHERIC THERMODYNAMICS u 63

Since q ¢ is conserved in a dry adiabatic process, and since the pressure difference across
the top and bottom of the layer lifted is constant, the difference in potential temperature
 ˜R „
between the top and the bottom of the lifted layer will also be conserved. Hence is
R „ ˜Q
constant for the layer and so
(gad –ÿ g) = constant ´ p¢ (3.73)
The above equation suggests that during the lifting of the layer when pressure decreases,
(gad – g) decreases and the lapse rate becomes more nearly dry adiabatic. Further, during
downward descent of the entire layer, the pressure increases, (gad – g) increases and hence the
lapse rate moves farther from gad. Hence, an initial stable layer when lifted is destabilized,
while an initial stable layer gets further stabilized due to descending of the entire layer.
It is interesting to examine the stability situations when the result of lifting of an initial
stable unsaturated layer results in saturation throughout the layer. Three such cases are
possible for the above situation. In the first case, the layer has uniform equivalent potential
È ˜R Ø
temperature É F Ù  Thus, every point in the layer after an initial dry adiabatic expansion
Ê ˜[ Ú
reaches condensation along the same moist adiabatic line. This corresponds to conditions in
which the layer becomes neutral with respect to saturated parcel displacements. In the second
È ˜R F Ø
case, the layer is assumed to have an increase of qe with height É ! Ù  The top of the
Ê ˜[ Ú
layer reaches saturation along a moist adiabatic line, which is to the right of the one along
which the bottom of the layer reaches saturation. Due to the above, the final lapse rate of the
lifted air parcel is less than gs and the lifted layer is said to be stable with respect to saturated
parcel displacements. In the third final case, the layer is assumed to have a decrease of qe with
È ˜R F Ø
height ÉÊ  Ù  In this case, the top of the layer reaches saturation along a moist adiabatic
˜[ Ú
line, which is to the left of the one along which, the bottom of the layer reaches saturation.
Hence, the final lapse rate of the lifted air parcel is more than gs and the lifted layer is said to
be unstable with respect to saturated parcel displacements. The above discussion of the
stability characteristics of an entire layer lifted to saturation is called convective instability and
can be summarized as
˜R F
> 0, indicating convective stability for which the saturated layer will be stable
˜[
˜R F
< 0, convective instability for which the saturated layer will be unstable (3.74)
˜[
˜R F
= 0, indicating convective neutral for which the saturated layer will be neutral
˜[

3.7.5 Stability Criteria Using Slice Method


It is clear that the parcel method does not incorporate the effects of compensating vertical
motion in the environment as an air parcel or an entire layer of the atmosphere rises. We shall
64 u BASICS OF ATMOSPHERIC SCIENCE

consider an initial horizontal layer of saturated air. It is assumed that within such a horizontal
layer there are several regions of ascending as well as descending motion of saturated air. It is
assumed that the ascending air has a total horizontal area A and an upward speed w, while the
corresponding total horizontal area and the downward speed of the descending air are A¢ and
w¢, respectively. Further, it is assumed that the rate at which the saturated air descends across
a certain reference level is the same as the rate at which the saturated air ascends through the
same level within the initial horizontal layer. Within a small interval of time dt, the air masses
transported upward and downward are
dM = rAw dt = rAdz = – Adpg–1
dM ¢ = r¢A¢w ¢dt = r¢A¢dz¢ = –A¢dp¢g–1 (3.75)
where the hydrostatic equation has been used and dz and dz¢ are the vertical distances through
which the ascending and descending saturated air moves in a time, dt and dp and dp¢ are the
corresponding hydrostatic changes in pressure. The saturated layer in the initial horizontal
layer is assumed to be horizontally homogeneous and so r = r ¢. Since the rates are assumed
equal, dM = dM ¢ and hence in the initial stage of the process
"„ E[ EQ
" E[ „ EQ „
It is assumed that the horizontal advection effects are small and can be neglected.
Figure 3.6 illustrates the case for an initial conditionally-unstable atmosphere where the air
sinks from Z0 + dz to Z0 dry adiabatically in time dt and air rises from Z0 – dz¢ to Z0 along a
moist adiabat. The ascending air having a temperature T at level Z0 – dz rises along a
saturation adiabat and reaches the reference level Z0 where it will have a temperature T1 = T –
gs dz. The descending air having a temperature T ¢ at level Z0 – dz¢ reaches the reference level
Z0 where it will have a temperature 5„ 5 „  H BE E[ „ For the establishment of instability, a
situation of positive buoyancy is necessary, i.e. the temperature T1 of the rising air must be
greater than 5„ the temperature of the adjacent subsiding air at the reference level Z0.

FIGURE 3.6 Changes in temperature associated with a conditionally unstable


slice of air undergoing vertical displacements.
ATMOSPHERIC THERMODYNAMICS u 65

Hence for the unstable case


T– gs dz > T¢ + gad dz¢ (3.76)
However, T = T0 + g dz and T ¢ = T0 – g dz¢, where g is the initial lapse rate in the layer
(gs < g < gad) and T0 is the initial temperature at the reference level Z0. Substituting the above,
one gets
H  H T
E[ ! H BE o H
E[ „ (3.77)
"„ E[
However, and so
" E[ „
(g – gs) A¢ > (gad – g) A for unstable conditions
(g – gs) A¢ = (gad – g) A for neutral conditions (3.78)
(g – gs) A¢ < (gad – g) A for stable conditions
By defining gn, the slice lapse rate for the case of neutral equilibrium the stability criteria
can be written in a more compact form as
g > gn for unstable conditions
g = gn for neutral conditions (3.79)
"„H T  "H BE
g < gn for stable conditions, where H O
"„  "
It is to be noted that gn plays the same role in the slice method as played by gad and gs and
that gn is the weighted mean of gad and gs with the areas of ascent and descent serving as the
weighting factors. For the conditionally unstable case we are considering gs < gn < gad. From
the above, it is clear that once the compensation motions are taken into account it is relatively
easy to obtain instability than in the case of dry uncompensated parcel displacements. Further,
it is more difficult to obtain instability in the case when compensating motions are taken into
account than in the case of saturated uncompensated parcel displacements. The slice method
indicates that all things being equal, the chances of slice instability are greater when A¢ is
large and A is small, i.e. small areas of ascent and large areas of descent are most favourable
for the development of instability. Conditions favourable for the formation of cumulus clouds
resemble the above requirement. The compensating subsiding motion in the slice method is
likely to reduce the excessive buoyancy of the parcel method since the subsiding air will get
warmed dry adiabatically and will reduce the temperature of the warm ascending air with
respect to the environment.

3.7.6 Entrainment Effects


The parcel method had two major limitations; one being that compensating motion in the
environment was not being taken into account and the other due to the neglect of entrainment
or mixing of environmental air into an ascending saturated air mass. The slice method
incorporated the compensating motion while the present section will seek to investigate the
effects of entrainment. In the treatment outlined here it is assumed that the condensation
products remain within the system, i.e. the ascending air mass can be considered to be an
66 u BASICS OF ATMOSPHERIC SCIENCE

ascending cloud mass. Laboratory experiments have established the existence of the
entrainment phenomena; wherein a jet injected into a fluid medium draws into it some of the
surrounding medium.
Assume a sample of saturated cloud of mass m which moves vertically a distance dz and
is subjected to a pressure change dp. As the cloud mass rises, it entrains a mass dm of
unsaturated environmental air, usually colder and drier. It is assumed that the ascent and
mixing are made up of (i) a saturated adiabatic cooling of the ascending cloud mass m,
(ii) isobaric cooling of moist ascending air of mass m and warming of the environmental air
of mass dm due to entrainment, and (iii) evaporation of a part of the liquid water content of m
into dm so as to produce a final saturated mixture. Further, it is assumed that the cloud mass
has uniform horizontal distribution of temperature, water vapour and liquid water content.
Also, the cloud mass together with the entrained air is considered as a thermodynamically
isolated system. That is, except for the heat lost due to work done during expansion, the heat
gained by the entrained air must be equal to the heat lost by the cloud mass. The cloud mass
loses heat due to the following two reasons: (i) warms the entrained air which is cooler as
compared to the cloud temperature, and (ii) heat of evaporation required to saturate the
entrained air which is usually dry. The cloud mass gains heat due to the release of latent heat
of condensation due to the ascent of the cloud mass. In quantitative terms, the heat required to
warm the entrained environmental air from its original temperature T ¢ to the cloud
temperature T is
dQ1 = cp (T – T ¢)dm (3.80)
It is assumed that heat changes of water vapour and liquid water are small compared to heat
change of dry air. The amount of heat required to evaporate some of the liquid water to
saturate dm of the entrained air is
dQ2 = L(ws – w¢)dm (3.81)
where ws is the saturation mixing ratio in the cloud, w¢ is the mixing ratio of the
environmental air and L is the latent heat of vaporization. The heat gained by the cloud mass
due to the ascending motion and consequent release of the latent heat of condensation is
dQ3 = – mLdws (3.82)
The first law of thermodynamics for the cloud mass is written as
Î EQ Þ
 D Q 5  5 „
EN  - XT  X „
EN  N-EX T N ÏD Q E5  3TQ5 ß (3.83)
Ð Qà
Substituting the approximate relationship
EXT EFT EQ EQ

XT FT Q and dividing Eq. (3.83) by m and using the hydrostatic equation Q
HE[
  one gets
3TQ5 „

 EN Ë EFT HXT E[ Û
<D Q 5  5 „
 - XT  X „
>  - Ì XT  Ü D Q E5  HE[ (3.84)
N ÌÍ FT 3TQ5 ÜÝ
ATMOSPHERIC THERMODYNAMICS u 67

EFT EFT E5
Dividing by dz throughout and substituting for one gets
E[ E5 E[

 EN
D 5  5 „
 - XT  X „
>  - Ì XT EFT E5 HXT Û D E5  H
Ë
< Q  Ü (3.85)
NE[ ÍÌ FT E5 E[ 3TQ5 ÝÜ Q E[
Dividing Eq. (3.85) by cp and rearranging the terms one gets the expression for the lapse rate
in an entraining cloud as

E5 ÎÑ H È -XT Ø  EN Ë - XT  X „
Û ÞÑ ÎÑ -XT EFT ÞÑ
 Ï É  Ù Ì 5  5 „
 Ü ß Ï  ß (3.86)
E[ ÑÐ D Q Ê 3TQ5 Ú N E[ ÌÍ DQ ÜÝ Ñà ÑÐ D Q FT E5 Ñà
EN
It is to be noted that if  Eq. (3.86) reduces to the same form as saturated adiabatic
E[
EN
lapse rate. Also, when entrainment of environmental air occurs !  5 ! 5 „ XT ! X „
E[
Since the environmental air is cooler and drier than the ascending cloud mass and these
cause the lapse rate in the entraining cloud, Eq. (3.86) to be larger than gs. Hence the
entraining of air into an ascending cloud mass will reduce the buoyant lift and the cloud mass
will rise to lower heights as compared to the situation where entrainment effects are absent.

3.8 ENTROPY AND SECOND LAW OF THERMODYNAMICS


The first law of thermodynamics is simply a statement of conservation of energy. The first law
is silent on the direction of a thermodynamic process and it is only the second law of
thermodynamics through the concept of entropy that provides for the direction of a
thermodynamic process. Entropy is a property of state which increases in association with an
irreversible process in an isolated system.

3.8.1 Entropy
The first law of thermodynamics can be written in the following form for any reversible process:
ER D Q E5  B EQ
Using the equation of state and dividing throughout by temperature T, one gets

ER D Q E5 3TQ EQ
 D Q E MO 5
 3TQ E MO Q
(3.87)
5 5 Q
Both the terms on the right-hand side are differentials of the function of thermodynamic
variables or exact differentials. The property of an exact differential, say cp d(lnT ) is that the
integration of an exact differential from an initial state where the temperature is T1 to a final
È 5 Ø
state where the temperature is T2 results in D Q MO
ÉÊ 5 ÙÚ which depends only on the initial and

68 u BASICS OF ATMOSPHERIC SCIENCE

final conditions and not on the process by which the change has been brought about.
Extending the above argument, if the integration were performed of an exact differential over
a closed path, the result must be zero. The integral of Eq. (3.87) about a closed path should
yield
ER
vÔ 5 = 0 (3.88)
ER
Defining ds = (3.89)
5
as the specific entropy, it is clear that entropy itself is a state variable. Entropy is also related
to the potential temperature as can be seen by logarithmically differentiating the Poisson
equation and rearranging terms

ER E5 EQ
DQ = DQ  3TQ
R 5 Q
cp d (ln q) = cp d(ln T) – Rsp d(ln p) (3.90)
From Eqs. (3.87) and (3.90), one gets
ds = cp d(ln q) (3.91)
or s = cp ln q + constant
Since we are concerned about the changes in the entropy, the additive constant is of no
consequence. The above discussion suggests that processes in which the potential temperature
remains constant are also processes in which entropy does not change and such processes are
known as isentropic processes. It is to be noted that specific entropy is depicted as an
ordinate in the Tephigram.

3.8.2 Second Law of Thermodynamics


The first law of thermodynamics is simply a statement of the conservation of energy, i.e. the
amount of heat added to/removed from a thermodynamic system during a process goes partly
towards increasing/decreasing the internal energy of the system and the remaining towards the
work done by/work done on the system during the above said process. The first law is
completely silent on the direction of the thermodynamic process. It would be important to
answer the following question: Given any two states of an isolated system in both of which
the energy is the same which are connected by a process. Can a criteria be provided that
determines which of the two states one is a possible initial state and the other one is a
possible final state? The answer would be available if there existed some property of the
system that is a function of state of the system, which have different values at the beginning
and at the end of the possible process. Energy cannot be the property as it is same at both the
states. Clausius proposed that the above property be called as the entropy of the system. The
second law of thermodynamics through the introduction of entropy provides for the direction
of the thermodynamic process by stating the following: Processes in which the entropy of an
isolated system would decrease do not occur at all. It is to be noted that whereas work can
always be transformed completely into heat, the opposite is not quite true, i.e. heat cannot be
ATMOSPHERIC THERMODYNAMICS u 69

completely transformed into work. The above implies that during certain situations, when any
system is undergoing an irreversible process, some of the energy becomes unavailable for the
production of work. It can be shown that the amount of unavailable energy during irreversible
process is equal to T Ds, where Ds is the difference in the entropy between the final and the
initial states of the system and its surroundings. Hence entropy is a measure of the
unavailability of energy and the entropy of an isolated system undergoing an irreversible
process increases. This can be interpreted as a consequence of the second law of
thermodynamics. Since work is an orderly process of motion, one can conclude that the
entropy of any thermodynamic system is a measure of the disorder of the system. Also, since
all natural processes are more or less irreversible, the above implies that nature tends towards
a state of greater disorder.
Since entropy is a state variable, it is defined only for equilibrium systems and any
change in entropy from one state to another is associated with a reversible process connecting
the two states. For a system undergoing an irreversible process between any two given states,
the change in the entropy for the above case is exactly equal to the change in entropy for the
equivalent reversible process connecting the same two states. However, it can be shown that
ER ER
Ô 5 is not the same for both the processes with 'T … Ô 5 with the equality/inequality sign
holding for the reversible/irreversible process. The above indicates that to achieve a given
change in entropy for any system through an irreversible process, more heat energy (greater dq)
is required as compared to a reversible process. This obviously suggests that reversible
processes are more efficient than irreversible processes.

3.8.3 Heat Engines and Refrigeration Cycles


A heat engine is a device which operates in a cycle and converts heat energy to work. The
science of thermodynamics rapidly grew in the early part of the nineteenth century as
scientists and engineers were attempting to improve the efficiency of steam engines. Steam
engines are devices into which a certain input of energy in the form of heat is given and
whose output is mechanical energy. Carnot showed in the nineteenth century that even with
an ideal engine it is never possible to convert more than a certain percentage of heat into
work. One form of the second law of thermodynamics attributed to Kelvin and Planck states,
It is impossible to construct a device that will operate in a cycle and produce no effect other
than the raising of a weight and exchange of heat with a single reservoir. That is, an amount
of heat in a reservoir at higher temperature T1 cannot yield any work unless it flows through
some heat engine and unless part of the heat is deposited to another reservoir at a lower
temperature T2. The efficiency of a heat engine (expressed in percentage) is the ratio of the
work done to the amount of heat input (i.e. the ratio of the energy sought to the energy cost)
and from the Kelvin and Planck statement it is clear that the efficiency of a heat engine has to
be less than 100%. If one reverses the direction of the heat flow and work in a heat engine
cycle, one obtains the refrigeration cycle. The refrigerator is an example of a device which
operates in a cycle (refrigeration cycle) in which a receipt of a certain input of work ensures
successful transfer of heat from a cold body to a hot body. Another form of the second law of
thermodynamics attributed to Clausius states, No process is possible whose sole result is a
70 u BASICS OF ATMOSPHERIC SCIENCE

heat flow out of one system at a given temperature and a heat flow of the same magnitude
into a second system at a lower temperature. The heat extracted/delivered from a hot/cold
reservoir in a heat engine cycle becomes the heat delivered/extracted to a hot/cold reservoir in
a refrigerator cycle. Also, the amount of work output in a heat engine cycle becomes the work
input to a refrigerator cycle. The efficiency of a refrigerator cycle known as coefficient of
performance (COP) is again the ratio of the energy sought to the energy cost and can have
values greater than 100%. It is clear from Clausius statement that any refrigerator cannot have
infinite COP.

3.9 CARNOT CYCLE AND CLAUSIUS CLAPEYRON EQUATION


Heat engine is a device that makes use of heat to do work. Carnot engine is one of the most
important among the ideal heat engines. The thermodynamic processes associated with the
Carnot engine is known as Carnot cycle. It is known that the saturation vapour pressure
depends on temperature. The exact dependence of the saturation vapour pressure on
temperature is given by the Clausius Clapeyron equation. Both the Carnot cycle and the
Clausius Clapeyron equation will be discussed in the following subsections.

3.9.1 Carnot Cycle


The Carnot cycle is an ideal heat engine cycle which provides for the maximum efficiency of
any reversible heat engine operating between any two given heat reservoirs. The Carnot cycle
is usually described using a cylinder of the engine having a perfect gas (working substance)
and a heat reservoir at a temperature T and a heat sink at a lower temperature T ¢. It is
assumed that the cylinder has a non-conducting piston and the walls of the cylinder are non-
conducting with only the bottom of the cylinder being a perfect conductor. Figure 3.7
illustrates the various stages of the Carnot cycle. Let the initial temperature of the gas within
the cylinder be T (indicated by point A in Figure 3.7) and let it be placed in contact with the
heat reservoir and the gas be subjected to reversible isothermal expansion until the point B is
reached. The contact with the heat reservoir is removed and the gas is then subjected to
reversible adiabatic expansion until the point C is reached with the temperature of the gas
being T ¢. The cylinder is now brought into contact with the heat sink and the gas is subjected
to a reversible isothermal compression until the point D is reached. Finally the contact with
the heat sink is removed and gas is then subjected to reversible adiabatic compression until
the gas gets back to its initial state (point A). It can be shown that in a Carnot cycle the ratio
of heat Q1 absorbed from the heat reservoir at temperature T K to the heat Q2 rejected to the
5
heat sink at temperature T¢ K is equal to  The efficiency of the Carnot cycle is given by


2 

2 5
indicating that the work done is Q1 – Q2. The Carnot cycle can be used to derive the well-
known Clausius Clapeyron equation which provides for the variation of the equilibrium
vapour pressure with temperature.
ATMOSPHERIC THERMODYNAMICS u 71

FIGURE 3.7 Various stages of the Carnot cycle as depicted graphically in a pV diagram.

3.9.2 Clausius Clapeyron Equation


The following derivation of Clausius
Clapeyron equation follows the use of the
Carnot cycle. Let ABCD and EFGH (refer
Figure 3.8) represent two adjacent
isothermals at temperatures T and T + dT
in a pa diagram. Construct two dry
adiabats from points F and G in the
isotherm having temperature T + dT
meeting the other isotherm at points M and
N. The cycle FGNM can be envisaged as a
Carnot cycle. Consider a unit mass of a
substance be subjected to a Carnot cycle
by allowing it to expand isothermally
along FG, expand adiabatically along GN, Figure 3.8 Graphical representation of a
and an isothermal compression along NM Carnot cycle in a pa diagram.
followed by an adiabatic compression
along MF. The substance at F is in the liquid state and at G is in the form of saturated
vapour. Let the amount of heat taken in by the substance during the Carnot cycle be L + dL
at temperature T + dT, where L is the latent heat of vaporization. Hence the work done during
the Carnot cycle is the product of the heat taken multiplied by the efficiency of the Carnot
cycle and is given by
È 5 Ø -E5
-  E-
É  to the first order (3.92)
5  E5 ÙÚ

Ê 5
Also, the work done during the Carnot cycle is simply the area of the parallelogram FGNM
that can be approximated by the product of FG and the perpendicular distance between
FG and NM. FG is the change in volume due to evaporation of unit mass of liquid and is
hence equal to (a2 – a1), where a2 and a1 refer to the specific volumes of the vapour and the
72 u BASICS OF ATMOSPHERIC SCIENCE

liquid, respectively. The perpendicular distance between FG and MN is nothing but dp, the
increase in saturation vapour pressure (des) due to increase in temperature dT. Hence the
area of the parallelogram is (a2 – a1) des and this when equated with Eq. (3.92) gives the
Clausius Clapeyron equation
EFT -
E5 5 B   B
(3.93)

where L is the latent heat of vaporization.

SOLVED EXAMPLES

1. Show that the mass of a vertical column of air of unit cross-sectional area in a hydrostatic
Q
atmosphere extending from the ground to very great heights equals  where p0 is the
H
surface pressure and g is the acceleration due to gravity. From the above expression,
estimate the total mass of the atmosphere.
Solution: The mass of the atmosphere having unit cross-sectional area and a thickness
dz is dm = rdz, where r is the density of air. Integrating the above expression over the entire
atmospheric column from z = 0 to very great height H and using the hydrostatic equation
EQ
 S H one gets
E[
)  Q
EQ  Q
. Ô SE[  Ô H H Ô EQ H
 Q 
Q
The total mass of the atmosphere is obtained by multiplying with the surface area of the
H
earth, i.e. with 4pR , where R is the radius of the earth. Using R = 6400 km = 64 ´ 105 m,
2

p0 = 1013 hPa = 1013 ´ 102 Pa, g = 9.81 ms–2, one gets for the total mass of the
atmosphere a value 5.3 ´ 1018 kg.
2. Find the density of water vapour, which exerts a pressure of 10 hPa at 25oC.
Solution: The equation of state for water vapour is, eav = RvT
where e is water vapour pressure (in Pascals), av is the specific volume of water vapour
(in m3 kg–1). Rv is the gas constant for 1 kg of water vapour (in J deg–1 kg–1).
Since Rv = 461 J deg–1 kg–1, T = 298 K and e = 1000 Pa,
rv = e/Rv ´ T = 1000/461 ´ 298 = 0.00728 kg m–3 is the density of water vapour.
3. For air at 0 oC and a pressure of 500 hPa, find the saturation values of mixing ratio, specific
humidity, absolute humidity
Solution: Triple point temperature and triple point pressure correspond to 0oC and
6.11 hPa. Hence given, T = 0oC, pressure = 500 hPa, and es = 6.11 ´ 100 Pa.
FT  – 
The saturation mixing ratio ST  H H   H LH 
Q  FT   
ATMOSPHERIC THERMODYNAMICS u 73

FT  – 


Saturation specific humidity RT  H H   H LH 
Q 
FT  – 
Absolute humidity SW  LH N 
3W 5  – 
4. Find the relative humidity for air at T = 18oC and actual vapour pressure equals 10 hPa.
Given the saturation vapour pressure at T = 18oC is 20.88 hPa.
F  – 
Solution: Relative humidity =   
FT 
5. For air at 30oC,
pressure of 1000 hPa and a relative humidity of 20%, find the dew point
temperature given the saturation vapor pressure at 30oC is 43.67 hPa.
Solution: Using the Clausius Clapeyron equation, the dew point temperature can be found
from the following equation:

Ë 3W Î F ÞÛ
5E Ì   MO Ü
ÍÌ 5P - ÏÐ F ßàÝÜ
3W

where T0 = 273 K, e0 = 6.11 hPa, and  , 
-
Y
The actual vapour pressure e = 0.2 ´ 43.67 = 8.734 hPa.
Substituting the above values, one gets, Td = 278 K
6. Find the saturation adiabatic lapse rate for air at 26°C, pressure of 1013 hPa and having a
saturation-mixing ratio of 21.85 g kg–1. Use the specific heat at constant pressure for dry air
= 1004 J kg–1 K–1.

Ë Û
Ì È ST -W Ø Ü
ÉÊ  
H Ì 3E 5 ÙÚ Ü
Solution: The saturation adiabatic lapse rate * T Ì Ü
DQ Ì ÎÑ -W ST 
Þ

ÑÜ
Ì Ï   ßÜ
D Q 3E 5
ÍÌ ÐÑ àÑ ÝÜ
Using cp = 1004 J kg–1 K–1, Lv = 2.5 ´ 106 J kg–1, Rd = 287 J K–1 kg–1, T = 299 K,
p = 1013 hPa, g = 9.8 m s–2, and rs = 21.85 g kg–1, one gets the saturation adiabatic lapse
rate is 3.727 = 3.73 K km–1.
7. Calculate the ratio of the number densities of oxygen atoms and hydrogen atoms at a height
of 1500 km, given the above ratio is 105 at a height of 200 km above the earth surface.
Assume an isothermal atmosphere of T = 2100 K.
Solution: The distribution of individual gases at such high altitudes is determined by
diffusion from the following relation

Q LN
0 Q LN
0 FYQ<  LN) LN
> Ë È   ØÛ
0 
FYQ Ì  LN
Q LN
) Q LN
) FYQ<  LN) 0 LN
>

ÉÊ )  ) ÙÚ Ü
ÍÌ 0 ) ÝÜ
The scale heights at 2100 K are defined as
74 u BASICS OF ATMOSPHERIC SCIENCE

3 –  3 –   
)0 N ) ) N   –   N o
 –   –  )0 ))
Q LN
0

Substituting in the first equatuon,  FYQ  

Q LN
)

8. Given the mean virtual temperature of the layer between 1000 hPa and 600 hPa over a
certain region is 7°C, find the thickness between the above pressure levels.
Solution: From the following equation, thickness can be found

3E 5W È  Ø  – 


'; ; I1B  ; I1B MO É MO 
 N
H Ê  ÙÚ 
9. Given the pressure at sea level is 1013 hPa, find the geopotential height of the 1000 hPa
pressure surface assuming the scale height of the atmosphere as 8000 m.
Solution:
È Q Ø È Q   Ø È Q   Ø
) MO É  Ù ) MO É  
; I1B Ù )É      
 N
Ê  Ú Ê  Ú Ê  ÙÚ
10. Given the density of dry air and that of water vapour alone at 0oC, being 1.27 kg m–3 and
4.77 ´ 10–3 kg m–3, find the total pressure exerted by the mixture of dry air and water
vapour at the above temperature.
Solution: The equation of state for dry air, pd = rd RdT = 1.27 ´ 287 ´ 273 = 99505 Pa
The equation of state for water vapour, pv = rvRvT = 4.77 ´ 10–3 ´ 461 ´ 273 = 600 Pa
Hence total pressure = 995.05 + 6 = 1001.05 hPa
11. A parcel of dry air has a temperature of 222 K at 300 hPa. What is its potential
temperature? What would be temperature the parcel would acquire if it was to be
adiabatically compressed to 700 hPa.
3E D Q
È  Ø È  Ø


Solution: From Poisson equation, R 5É  É  ,


Ê Q ÙÚ Ê  ÙÚ

È  Ø
The temperature at 700 hPa equals, 5 5 ÉÊ Ù  ,
 Ú
 

12. Calculate the virtual temperature correction for moist air at 20oC having a mixing ratio of
15 g kg–1.
Solution: The virtual temperature correction equals
5W  5 5   X
    – 
   ,

REVIEW QUESTIONS

1. Consider a column of air that is above a unit horizontal area of 1 m2. What is the mass of
air in that column: (i) above the earth’s surface, (ii) above a height where the pressure is
500 hPa, and (iii) between pressure levels of 700 hPa and 500 hPa.
ATMOSPHERIC THERMODYNAMICS u 75

2. Imagine a sample of air comprising of only nitrogen, oxygen and argon. What is the mean
molecular weight of the above sample of air? Let the molecular weights of nitrogen, oxygen
and argon be 28.01, 32.00, and 39.95, respectively.
3. The mean lapse rate of the troposphere is given as 6.5°C per km. If the surface temperature
is 30°C, what is the temperature at the top of Mt. Everest which is about 9 km above the
surface? Also, if the surface temperature is 30°C and the temperature at the top of
Mt. Everest is – 30°C, what is the mean tropospheric lapse rate?
4. How does the height of the tropopause change as one moves north from Trivandrum, India
at 8oN to Srinagar at 35oN? Explain your observations.
5. Why is absolute humidity rarely used?
6. Consider a unit mass (1 kg) of air. Assume that within this 1 kg of air, there are 50 grams of
ice. (i) How much heat is required to sublimate the ice? (ii) Assuming that the heat required
to sublimate the ice comes from the unit mass (1 kg) of air, how much colder does the air
become after all the ice has sublimated?
7. Why does the surface temperature of air increase on a clear, calm night as a low cloud
moves overhead?
8. It can be shown that the saturation vapour pressure (es) depends only on temperature:

Ë - È   ØÛ
FT F FYQ Ì É  ÙÜ
ÌÍ 3W Ê 5 5 Ú ÜÝ
where e0 = 6.11 hPa and T0 = 273 K are constant parameters. Rv is the gas constant for
water vapour and is equal to 461 J K–1 kg–1. L is the latent heat of condensation if water is
present or the latent heat of deposition if ice is present. T is the temperature in Kelvin.
(i) For water, make a plot of es (y-axis) versus temperature (x-axis) for T = 250 K to 273 K.
How does es depend on temperature? (ii) Plot es versus temperature for ice, on the same
graph (iii) Is es larger for ice or for water?
9. Would one expect water in a glass to evaporate more quickly on a windy, warm, dry summer
day or on a calm, cold, dry winter day? Explain.
10. Discuss how and why each of the following will change as a parcel of air with an
unchanging amount of water vapour rises, expands, and cools: (a) absolute humidity,
(b) relative humidity, (c) actual vapour pressure, (d) saturation vapour pressure.
11. Assume that the dew point of cold air outside is the same as the dew point of warm air
inside a room. If the door of the room is opened, and cold air replaces some of the warm air
inside, what would happen to the new relative humidity inside the room? Will it be: (a) lower
than before, (b) higher than before, or (c) the same as before? Explain.
12. Where would one go if one is interested to experience the least variation in dew point from
January to July?
4 Radiation

Almost all of the energy transfer between the earth and the rest of the universe is through
radiative processes. The earth and its atmosphere are absorbing the incoming short-wave solar
radiation and emitting long-wave infrared radiation to space. The earth’s atmospheric
constituents such as water vapour, carbon dioxide and ozone do absorb the long-wave
radiation emitted by the earth, while the earth in turn absorbs a large part of the infrared
radiation emitted by the atmosphere. The average rate of absorption and emission are very
nearly equal, leading to the conclusion that the earth-atmosphere system is very nearly in
radiative equilibrium.
The transfer of energy through radiative processes is an important mechanism for the
exchange of energy between the different layers of the atmosphere and also between the
atmosphere and the underlying surface. High energy radiations with wavelengths less than 0.1 mm
and ultraviolet radiations get absorbed in the upper regions of the atmosphere and play a vital
role in the formation of ozone. The satellite sensors duly monitor the radiation lost to space
by the earth atmospheric system and such measurements form the basis of the remote sensing
of the atmospheric structure.
This chapter begins with a short discussion on the spectrum of electromagnetic
radiation. Section 4.2 introduces the concept of black body radiation, while Section 4.3
outlines the atmospheric absorption of the solar radiation. The next section introduces briefly
the physical aspects of the scattering of the solar radiation. While Section 4.4 provides for the
absorption and emission of the infrared radiation, the last section discusses briefly the aspects
of remote sensing from space.

4.1 SPECTRUM OF RADIATION


A hot black body emits radiation in a wide range or spectrum of wavelength of varying
intensities. The radiation spectrum of a hot black body has a typical peak corresponding to a

76
RADIATION u 77

particular wavelength, which depends on the temperature of the hot black body. At any given
temperature, the radiation emitted by a black body is the maximum.

4.1.1 Overview
Before initiating a detailed discussion of the various topics outlined in this chapter, it would
be worthwhile to provide a broad overview of some important physical concepts. Firstly,
radiation is a process of transfer of energy by photons or equivalently by electromagnetic
waves without the need for any medium. One way in which the effects of the atmospheric
radiation can be studied is to investigate the interaction between photons and the atmospheric
gases. The amount of energy associated with a photon equals hn, where n is the frequency of
radiation and h is the Planck’s constant (h = 6.626 ´ 10–34 J s). A photon being part of the
incoming solar radiation may interact with the gas molecules at certain discrete frequencies,
resulting in absorption of the photon leading to the excitation of the gas molecule. In such a
case, the energy levels are related to the frequency n, as given below
DE = hn (4.1)
Since radiation travels at the speed of light c, (c = 3 ´ 108 m s–1), the wavelength is related to
the frequency through the relation
l = cn—1 (4.2)
The interactions of the photons in the solar radiation with the gas molecules may lead to an
excited state for the latter. The excitation energy associated with the excited state may
however, be lost through the following processes: (i) the electron falls back to the ground
state and in that process re-emits a photon in a random direction of the same energy and
frequency as the solar photon, and (ii) the molecular collisions occur before the re-emission of
the excitation energy can happen, leading to the transfer of the excitation energy DE to other
forms of energy, resulting in the solar photon being absorbed. In the latter case, the excitation
energy is delivered to the molecules as kinetic energy. The absorption of the solar photon
usually occurs at sufficiently high pressure. Both the processes involving loss of energy as
outlined above are denoted as extinction.
The first process outlined above is an example of scattering of a photon having a given
discrete frequency by an atmospheric gas molecule. In the real atmosphere, scattering of
photons occurs continuously over broad range of frequencies by atmospheric molecules
(Rayleigh scattering), by aerosols (Mie scattering) and by cloud droplets and raindrops
(treatment is then by geometrical optics). Absorption of photons in the high frequency
incoming solar radiation may cause photodissociation or breakdown of the molecules and
photoionization, or the removal of the outermost electrons from atoms resulting in the
formation of ions.
Photons in the infrared region of the spectrum may be scattered and absorbed in a
similar manner as outlined for the solar photon. The photon in the infrared part of the
spectrum can also be emitted, wherein the energy is being drawn from the molecular kinetic
energy. However, it is the vibrational or the rotational energy changes, rather than the orbital
transition energy, which is relevant in the case of photon in the infrared part of the spectrum.
78 u BASICS OF ATMOSPHERIC SCIENCE

4.1.2 Electromagnetic Spectrum of Radiation


Electromagnetic radiation may be envisaged as the totality of all the waves which travel
through vacuum at a speed of light. The electromagnetic waves may exhibit a continuous
range of wavelengths and the totality of all possible wavelengths (or frequencies) is known as
the electromagnetic spectrum. Table 4.1 provides the various parts of the electromagnetic
spectrum. The visible part of the spectrum can be further classified by different wavelength
TABLE 4.1 Various portion of the electromagnetic spectrum
Name Typical wavelength (mm)
X-ray <0.03
Ultraviolet 0.03–0.4
Visible 0.4–0.7
Near-infrared 0.7–8
Infrared 8–100
Microwave >100

intervals which are identified with a particular colour. Table 4.2 gives the wavelength interval
and the associated colour within the visible part of the electromagnetic spectrum.
TABLE 4.2 Wavelength interval and colour in the visible region
Colour Wavelength interval (mm)
Violet 0.39–0.455
Blue 0.455–0.505
Green 0.505–0.55
Yellow–green 0.55–0.575
Yellow 0.575–0.585
Orange 0.585–0.62
Red 0.62–0.76

Diffuse radiation is the name given to radiation emanating from a source that subtends a finite
arc of solid angle. The case of radiation emanating from a concentrated source, wherein the
angle subtended by the source is zero, is referred to as parallel beam radiation. For all
practical purposes, it is reasonably accurate to treat the solar radiation reaching the earth as a
parallel beam radiation. The above assumption effectively eliminates the need to integrate
over the solid angle.

4.2 BLACK BODY RADIATION


Black body refers to an idealized body or a system which absorbs all radiation which is
incident on it. Furthermore, no radiation passes through a black body and there is no
reflection associated with the incidence of radiation on a black body. At a given temperature,
a black body emits a spectrum of radiation. The peak radiation emitted depends on the
temperature of the body. The peak radiation also changes with change of temperature of the
emitting black body.
RADIATION u 79

4.2.1 Planck’s Law


Consider a cavity whose walls are maintained at a uniform temperature T. Under the condition
of thermodynamic equilibrium, the electromagnetic radiation within such a cavity is in
equilibrium with the walls of the cavity. It can then be shown that under the thermodynamic
equilibrium condition the spectral energy density, which is defined as the energy per unit
volume per unit frequency interval, is solely dependent only on the frequency and temperature.
Furthermore, under the condition of thermodynamic equilibrium, the electromagnetic radiation
is isotropic, i.e. the radiation is independent of direction within such a cavity. If a small opening
is made in the above-mentioned cavity, the radiation emanating from the cavity will have the
same characteristics as the radiation within it and is known as the black body radiation.
A black body is a body, at best hypothetical in nature, which is capable of absorbing and
emitting the electromagnetic radiation in all the portions of the electromagnetic spectrum and
having the following characteristics: (i) all incident radiation is completely absorbed, and
(ii) in all wavelengths and in all the directions, the maximum possible emission is obtained.
The example of the cavity serves as a black body, while the electromagnetic radiation inside
the cavity corresponds to the black body radiation. Hence, the nature of the spectral energy
density of black body radiation at a temperature T can be investigated by performing careful
experiments on the radiation emanating from the small opening of the cavity. Max Planck in
1900 investigated the black body radiation and derived an expression for the spectral energy
density of black body radiation at a temperature T, known as Planck’s law and is given as

 Q IO 
VO 5
(4.3)
Ë È IO Ø Û
D Ì FYQ É Ù  Ü
Í Ê L5 Ú Ý
where k is the Boltzmann constant (k = 1.38 ´ 10– 23 J K–1), h is the Planck’s constant, n is
the frequency and c is the speed of light. Since the black body radiation is isotropic, the
photons associated with the black body radiation can move in any direction. The energy
density associated with a group of photons moving within a small solid angle DW steradian is
':
 Since the photons travel at the speed of light, the rate of energy per unit area, per unit
Q
solid angle, per unit frequency interval can be calculated by multiplying the right-hand side of

Eq. (4.3) by c (to get the rate of energy per unit area) and also by multiplying by (to get
Q
per unit solid angle) to obtain the spectral radiance Bn (T) as

IO 
#O 5
(4.4)
Ë È IO Ø Û
D  Ì FYQ É Ù  Ü
Í Ê L5 Ú Ý
The spectral radiance given in Eq. (4.4) is also known as the Planck function. The Planck
function can also be written in terms of the rate of energy per unit area, per unit solid
angle and per unit wavelength interval and is denoted as Bl (T) and is given from Eq. (4.4) as
80 u BASICS OF ATMOSPHERIC SCIENCE

ID
#M 5
(4.5)
Ë IO Ø Û
M  Ì FYQ ÈÉ 
Í Ê L M5 ÙÚ ÜÝ
Plotting Eq. (4.5) as a function of wavelength for a given temperature, the characteristic curve
for the Planck function (refer Figures 4.1(a) and 4.1(b) reveals a sharp short wavelength
cut-off, a steep rise to the maximum and a gentler drop towards the longer wavelength.
Further, the maxima of the Planck function shifts towards the lower wavelength for higher
temperatures. Figures 4.1(a) and 4.1(b) depict the Planck black body function for emission
from the sun (T = 5780 K) and from the earth-atmospheric system (T = 255 K). It is clear
from Figure 4.1(a) that the peak emissions from the sun are in the visible range of
wavelengths, while the peak emissions from the earth-atmospheric system are in the infrared
range of the electromagnetic spectrum.

FIGURES 4.1(a) Planck black body curves for emission from the sun.

FIGURES 4.1(b) Planck black body curves for emission from the earth-atmospheric system.
RADIATION u 81

Except for the large wavelength end of the electromagnetic spectrum, the exponential
term in Eq. (4.5) is much larger than unity and hence Eq. (4.5) can be approximated as

  Ë  IO Û
#M 5
ID
M FYQ (4.6)
ÌÍ L M5 ÜÝ

For the peak emission of black body radiation, the variation of the Planck function with
respect to the wavelength at a fixed temperature has to vanish. Using the expression for the
Planck function in Eq. (4.6), one obtains the wavelength corresponding to the maximum black
body emission at a temperature T to be given by

MN (4.7)
5
where lm is expressed in mm and T is in K. Equation (4.7) is called Wiens displacement law
and provides for an estimate of the temperature of a radiation source from a knowledge of its
emission spectra.
One obtains the black body irradiance by integrating Eq. (4.5) over all possible
wavelengths to get
‡
# 5
Ô #M 5
E M T5  (4.8)


where s is the Stefan Boltzmann constant and is given by s = 5.67 ´ 10–8 W m–2 deg–4.
Equation (4.8) expresses the well-known Stefan Boltzmann law.

4.2.2 Local Thermodynamical Equilibrium


The expression for the Planck function Eq. (4.5), is applicable for a cavity with a uniform
temperature containing electromagnetic radiation, but not containing matter. However,
statistical physics shows that the energy levels of a material system, say, for example a gas, in
equilibrium at temperature T will be populated according to the Boltzmann distribution when
the radiation is neglected. The Boltzmann distribution is given by
O H
FYQ  < &  & > L5 (4.9)
O H
where n1 and n2 are the number of molecules in states of energy E1 and E2, while g1 and g2
are the degeneracies of the states 1 and 2, respectively.
The following question arises immediately. What happens to the Planck’s law and the
Boltzmann distribution when both matter (say, a gas) and the electromagnetic radiation are
present in the cavity of uniform temperature? The answer to the above-mentioned question is
that if the interaction between the matter and the radiation is sufficiently weak, then under
conditions of thermodynamical equilibrium, the electromagnetic radiation will satisfy the
Planck’s law, while the matter (say, the gas) will satisfy the Boltzmann distribution. While the
interaction between the matter and the radiation is essential in getting to the thermodynamical
equilibrium, the interaction should not be very strong as it may lead to significant departures
82 u BASICS OF ATMOSPHERIC SCIENCE

of the electromagnetic radiation from the Planck’s law and of the matter from the Boltzmann
distribution. If the interaction between the matter and the radiation is strong, no theoretical
assumption is possible and radiative transfer will ever remain an unsolved problem of
theoretical physics. It is to be noted that the interaction between the matter and radiation will
be sufficiently weak if the average time between collisions for a given molecule is very much
smaller than the lifetime of the excited electron for the given energy levels. The average time
between the collisions for a given molecule is inversely proportional to pressure and hence
the state of the thermodynamical equilibrium will hold if the pressure is large enough for a
given transition between the energy states.
It is convenient to distinguish between the global thermodynamical equilibrium (GTE)
and local thermodynamical equilibrium (LTE). Unlike the GTE, which requires all the
intensive parameters to be homogeneous throughout the system, the LTE allows the intensive
parameters to vary very slowly in space and time, i.e. vary so slowly that for any point, one
can assume thermodynamic equilibrium in some neighbourhood about that point. For
example, it is well known that a particle requires a certain number of collisions to equilibrate
to its surroundings. If the average distance moved by the particle during these collisions
removes it from the neighbourhood it is equilibrating to, it will never equilibrate, and there
will be no LTE. An atmosphere is said to be in LTE when it is possible to define a
temperature T at each point in the atmosphere in such a way that the coefficient of emission is
just the product of the coefficient of absorption and the Planck function at that temperature.
Furthermore, both the absorption and the emission coefficients under the conditions of LTE
are functions of temperature and density only. The energy associated with a molecule has two
main components; the kinetic energy associated with translational motions (which determines
the thermal temperature) and the energy associated with molecular scale energy transitions.
The latter include the vibrational, rotational and the electronic energy transition components.
Imagine a molecule which maintains the Boltzmann distribution, which is a function of its
radiating Planck temperature. Let the molecule be bombarded by electromagnetic radiation at
frequencies corresponding to the molecular modes. This will disturb the Boltzmann
distribution and result in the raising of the molecular energies to a higher state. Now, either of
the following two processes can happen. In the first process, the molecules, if left to
themselves, will release energy (photons) and reestablish the Boltzmann distribution.
However, in the second process, if the pressure of the gas is large enough, molecules will
collide before the above-mentioned release of energy takes place. Due to the molecular
collisions, the molecular energy gets redistributed as kinetic energy. If equilibrium can be
maintained between the continued absorption of radiation energy and its subsequent
redistribution as kinetic energy in such a way that the thermal temperature equals the Planck
temperature, then we say that a LTE exists. For atmospheric pressure above 0.05 hPa, i.e. for
almost 99.5% of the earth’s atmosphere, the condition for LTE holds.

4.2.3 Radiometric Quantities


In the description and measurement of radiative transfer, several quantities are usually used,
and these are defined as follows.
(i) Spectral radiance (or monochromatic radiance), Ln (r, s): It is the power per
unit area, per unit solid angle, per unit frequency interval in the vicinity of the
RADIATION u 83

frequency n, at a point r in the direction of the unit vector s. Figure 4.2 illustrates
the definition of the spectral radiance at a point r in the direction of the unit vector s.
The spectral radiance is measured in units of watts per square metre per steradian
per hertz. The spectral radiance may be visualized in terms of photons, which
emerge from a small area DA with unit normal s, centred at a point r. Considering
only these photons whose momentum vectors lie within a cone of small solid angle
DW centred in the direction of s (refer Figure 4.2), and whose frequencies lie
between n and Dn, the energy transported by these photons per unit time from
“below” the area DA to “above”, is then given by Ln DA DW Dv. This simply
follows from the definition of the spectral radiance. The Planck function is a
special case of the spectral radiance corresponding to isotropic black body radiation
within a cavity of uniform temperature, the former being independent of the
position and direction due to the isotropic nature of the radiation. The spectral
radiance, like the Planck function can also be expressed in terms of per unit
wavelength interval, instead of per unit frequency interval.

DW

DA

FIGURE 4.2 Figure illustrating the definition of spectral radiance.

(ii) Radiance L(r, s): It is the integral of the spectral radiance, integrated over all the
frequency and is nothing but the power per unit area, per unit solid angle at a point r
in the direction of the unit vector s. The relation of L and Ln is then expressed as
‡
- S T
Ô -O S T
EO (4.10)


The units of radiance are watts per square metre per unit steradian. The radiance is
a property solely of the radiation field and does not depend on the orientation of
the surface.
84 u BASICS OF ATMOSPHERIC SCIENCE

(iii) Spectral irradiance (or monochromatic irradiance) Fn (r, n): It is defined as


the power per unit area, per unit frequency interval in the vicinity of the frequency
n, at a point r, through a surface of normal n. The spectral irradiance is obtained
from the spectral radiance by integrating the latter over a hemisphere on one side
of the surface and is expressed as

'O S O
ÔQ -O S T O ¹ TE: T



(4.11)

where dW(s) is the element of solid angle in the direction of s (refer Figure 4.3).
The spectral irradiance is expressed in watts per square metre, per hertz.
Essentially, the spectral irradiance provides for the energy transferred by a group
of photons per unit time that emerge onto the region above the surface. The spectral
irradiance, like the Planck function can be expressed per unit wavelength interval
instead of per unit frequency interval.

FIGURE 4.3 Figure illustrating the calculation of spectral irradiance.

(iv) Irradiance (or flux density) F(r, n): It is defined as the power per unit area at a
point r through a surface of normal n. Essentially, irradiance is obtained from the
spectral irradiance by integrating the latter over all frequencies. Alternatively, the
irradiance can also be obtained from the radiance by integrating the latter over a
hemisphere. That is, the irradiance is expressed as
‡
' S O
Ô 'O S O EO ÔQ - S T O ¹ TE : T





(4.12)

The units of irradiance are simply watts per square metre.


The relation between the irradiance and the radiance can be understood as follows.
Since the irradiance represents the combined effects of the normal component of the radiation
coming from the whole hemisphere, it can be expressed as
Q
' S O
Ô - S T
DPT G E: (4.13)

RADIATION u 85

where f is the zenith angle defined as the angle between the direction of the radiation and the
normal to the surface in question and hence the component of the radiance normal to the
surface is given by L cosf. Imagine an infinitesimal area located at the centre of a sphere of
unit radius, which is emitting or receiving radiation. Let the orientation of the infinitesimal
area be coincident with the equatorial plane so that the normal to the surface passes through
the poles of the sphere. In this case, the zenith angle f simply becomes the co-latitude and the
solid angle dW can be expressed in spherical coordinates as
dW = sinf df dq (4.14)
where q is the longitude or the azimuth. The integral of dW in Eq. (4.14) over the upper
hemisphere is then equal to
Q  Q

Ô TJO G EG Ô ER Q (4.15)
 

which is equal to half the area of the sphere of unit radius. For the above case, the
relationship between the irradiance and radiance as given in Eq. (4.13) is then given by
Q  Q
' S O
Ô Ô - S T
DPT G TJO G EG ER (4.16)
 

4.3 ATMOSPHERIC ABSORPTION OF SOLAR RADIATION


Solar radiation provides the main source of energy that is responsible for the circulation of the
atmosphere. Although the atmosphere is by and large transparent to the solar radiation, there
exists strong absorption of solar radiation by some atmospheric constituents in certain select
wavelength intervals. The following subsections introduce the concepts of absorptivity and
emissivity together with concepts of reflectivity and transmittivity. Furthermore, the following
subsections also introduce the well-known Kirchoff’s laws and the Beer’s laws.

4.3.1 Absorption and Emission of Radiation by Molecules


It was earlier mentioned that an isolated molecule could absorb and emit energy only in
discrete amounts corresponding to the allowed changes in its energy level. Hence, an isolated
molecule can interact only with radiation having certain discrete wavelengths, thereby
providing for the emission and absorption properties of an isolated molecule in terms of a line
spectrum, consisting of a finite number of narrow absorption or emission lines where
absorption and emission of radiation are possible. The absorption lines associated with the
orbital changes correspond to X-ray, ultraviolet and the visible parts of the electromagnetic
radiation. While the vibrational changes are associated with the near infrared, the rotational
changes are associated with infrared and microwave radiation.

4.3.2 Absorptivity and Emissivity


While black body radiation represents the maximum possible emission in all the wavelength
bands and in all directions, most real substances emit radiation at a given temperature which
86 u BASICS OF ATMOSPHERIC SCIENCE

is lower than the black body temperature. Due to the above reason, a measure of how strongly
a given body radiates at a given wavelength can be proposed using the concept of emissivity.
Hence, emissivity depends on the body temperature, wavelength of emitted energy and also
on the angle of emission. Emissivity is normally measured in a direction normal to the surface
and as a function of wavelength. While, considering the entire energy loss by a body, one
requires an emissivity which includes all directions and all wavelengths. For situations
involving radiation exchanges between surfaces, it is useful to consider emissivity, which
average out all wavelengths while still depending on directions. In other situations where
spectral effects are large, it is convenient to define emissivity by averaging the spectral values
over all directions. Emissivity el of a body at a given wavelength l is simply defined as the
ratio of the spectral radiance emitted by the body to the spectral radiance of the black body
radiation at the same wavelength. Strictly speaking, the definition of emissivity should be
qualified by the frequency, direction, and even the polarization state of the emitted radiation.
This is usually done by including qualifiers such as monochromatic (spectral) emissivity (at a
given frequency) as opposed to total (over a broad range of frequencies), and directional
emissivity (the ratio for a particular direction) as opposed to hemispherical (the ratio for a
hemisphere of directions, i.e. integrated over all directions). Furthermore, one refers to the
emissivity, which includes both direction and wavelength (frequency) as directional spectral
emissivity. Hence,
-M
FM (4.17)
#M
Contrary to the common opinion, the upper limit of the emissivity of any real substance
(body) is not really one. The above upper limit of one is only valid, that too approximately,
for bodies large as compared with all the relevant wavelengths. It is sometimes convenient to
delink the wavelength dependence by defining the gray body emissivity, e as the ratio of the
irradiance to the black body irradiance
'
F (4.18)
#
In a similar manner, the absorptivity al of a real (body) substance can be suitably defined as
the ratio of the fraction of the energy incident on a body that is absorbed by the body. The
absorptivity may be a function of wavelength and/or direction, and is related to the emissivity
of the region. It is to be noted that the incident radiation depends on the radiative conditions
at the source of the incident energy. However, it turns out that the spectral distribution of
incident radiation is indeed independent of the temperature and/or the physical nature of the
absorbing surface unless the radiation emitted from the surface is partially reflected back to
the surface. The absorptivity has additional complexities as compared to the emissivity, since
for the former the directional and spectral characteristics of the incident radiation must be
considered. In a similar manner, the definition of absorptivity should be qualified by the
frequency (spectral absorptivity), direction (directional absorptivity) and both frequency
(wavelength) and direction (directional spectral absorptivity). The absorptivity and the
emissivity of a black body are equal to unity for all wavelengths. Furthermore, the
absorptivity is independent of wavelength for the so-called gray body. A gray body is a
RADIATION u 87

hypothetical body, which absorbs some constant fraction, of all the electromagnetic radiation,
which is incident upon it, i.e. for which the absorptivity is a constant independent of
wavelength. The absorptivity of a gray body is termed as gray body absorptivity.

4.3.3 Kirchhoff’s Law


Kirchhoff’s law states that at a given temperature and wavelength el = al, i.e. for a given
body the emissivity equals the absorptivity. Essentially, Kirchhoff’s law states that substances
(bodies) that are strong absorbers at a particular wavelength are also strong emitters at that
wavelength. One can also conclude from Kirchhoff’s law that weak absorbers at a particular
wavelength are also weak emitters at that wavelength. It is important to realize that the
validity of Kirchhoff’s law is not dependent on the requirement for thermal or radiative
equilibrium of the body. It is also possible to apply Kirchhoff’s law to gases provided that the
average time between collisions for a given molecule is much smaller than the lifetime of the
individual absorption and the emission events. For the earth’s atmosphere, the above condition
holds up to heights of about 60 km.

4.3.4 Reflectivity and Transmittivity


The reflectivity rl, of any surface is the ratio of the flux that is reflected to the flux that is
incident, while the transmittivity tl of any layer is the fraction of the flux that is transmitted
through the layer without any absorption. For any opaque surface, when the incident
monochromatic radiation is either absorbed or reflected, one has the following relation:
Ll (absorbed) + Ll (reflected) = Ll (incident) (4.19)
Dividing each of the above terms by Ll, one gets
al + rl = 1 (4.20)
Equation (4.20) states that for an opaque surface and for all wavelengths, the sum of
absorptivity and reflectivity equals unity. Equation (4.20) can be extended for a non-opaque
layer by including the transmittivity to obtain,
al + rl + tl = 1 (4.21)

4.3.5 Absorption of Solar Radiation by Atmosphere


In this section, an expression for the transmittivity of a layer of the atmosphere above a height z
is derived using the Beer’s law. For this derivation, only the effects of absorption are
considered, while the effects of scattering are assumed unimportant. In this case, the
absorption of solar radiation (parallel beam radiation) as it passes downward through a
horizontal layer of the atmosphere of infinitesimal thickness dz is proportional to the number
of molecules per unit area, corresponding to the area absorbing the solar radiation along its
path and the same can be expressed as
 E'M
EBM  LM S TFD G E[ (4.22)
'M
88 u BASICS OF ATMOSPHERIC SCIENCE

where r is the density of the atmosphere, f is the zenith angle, and kl is the absoption
coefficient. Since dFl signifies the depleted amount of radiation and since the radiation is
passing downward, both dLl and dz are < 0, giving positive value for daë. In Eq. (4.22), r sec f dz
denotes the mass within the volume swept out by a unit cross-sectional area of the incident
solar radiation as it passes through the atmospheric layer of thickness dz as shown in Figure 4.4.
In general, the absorption coefficient is a function of the composition, temperature and
pressure of the atmosphere within the layer. Integrating Eq. (4.22) from a height to the top of
the atmosphere (z = ¥), one gets
‡
MO 'M ‡  MO 'M TFD G
Ô LM S E[ (4.23)
[
which becomes
'M 'M‡ FYQ<  T M > (4.24)

Fl

Volume sec f dz

dz

Fl – dll

FIGURE 4.4 Depletion of an incident beam of radiation of unit cross-section


while passing through an absorbing layer of thickness dz.

‡
where TM TFD G Ô LM S E[ (4.25)
[

Equation (4.24) expresses the Beer’s law which states that the monochromatic irradiance
decreases monotonically with increasing ‘optical path length’ through the layer. The
expression sl is called as optical depth or optical thickness and is a measure of the totality of
depletion that the incident beam of radiation has experienced due to the passage of the
radiation through an atmospheric layer. The transmittivity of the layer of the atmosphere, at a
height z is then given as
'M
UM FYQ<  T M > (4.26)
'M‡
RADIATION u 89

In the absence of scattering, from Eq. (4.20), it follows that


al = 1 – tl = 1 – exp[– sl] (4.27)
Equation (4.27) clearly shows that with increasing optical thickness, the absoptivity
approaches the value unity.
The spectral irradiance of the incoming solar radiation at the top of the atmosphere and
at mean sea level are compared with the Planck black body spectral irradiance of the solar
photosphere temperature corresponding to 6000 K in Figure 4.5. It is clear from Figure 4.5
that the spectral irradiance of the solar radiation at the top of the atmosphere is in good
agreement with the Planck black body irradiance at 6000 K for most wavelengths. However,
comparing the top of the atmosphere irradiance and the mean sea level irradiance of the solar
radiation clearly shows significant deviations (refer Figure 4.5) between the two due to the
absorption and scattering effects in the atmosphere. The sharp dips seen in Figure 4.5 in the
mean sea level irradiance curve at certain wavelengths correspond to absorption of gases such
as ozone in the ultraviolet and the visible parts of the spectrum while the absorption in the
infrared parts of the spectrum is due to carbon dioxide and water vapour.

2200
2000
Solar Spectral Irradiance (Wm–2 /mm)

1800
1600
1400
1200
1000
800
600
400
200
0
0 0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.2 3.6 4.0
Wavelength (mm)

0.3 0.5 0.7 11 0.94 38 87 2.7 3.2


O3 O3 O2 H2O H2O H2O H2O H2O–CO2 H2O

FIGURE 4.5 Observed solar spectral irradiance distribution at the top of the atmosphere
(upper curve) as well as the observed solar spectral irradiance distribution at the
sea level (lower curve). The shaded areas represent absorption due to various
gases in a clear atmosphere.

4.3.6 Indirect Estimate of Solar Irradiance at the Top of the


Atmosphere
It is possible to obtain estimates of the solar irradiance incident at the top of the atmosphere
using direct measurements of the solar irradiance measured at frequent intervals of time at a
ground station.
90 u BASICS OF ATMOSPHERIC SCIENCE

Equation (4.23) can then be rewritten as


‡
MO 'M MO 'M‡  TFD G Ô LM S E[ (4.28)
[

When the solar irradiance Fl at the ground station is measured at frequent intervals of time
over a given day under clear, stable atmospheric conditions, the numerical value of the
integral changes much less as compared to the large changes in the solar zenith angle. Hence
to a good approximation, the above equation can be rewritten as
ln Fl = A – BZ (4.29)
where Z = sec f and A and B are constants. Plotting the measured Fl at the ground on a log
scale at different times (different Z) shows a linear relationship between ln Fl and Z. Since
the path length is directly proportional to Z, one can deduce the solar irradiance on the top of
the atmosphere by simply extending the straight line (making the best fit to the data points) to
Z = 0.

4.3.7 Vertical Profile of Absorption


Chapman had proposed a simple mathematical model to explain the vertical distribution of
absorption of monochromatic radiation in the earth’s atmosphere. Consider the parallel beam
radiation from the sun, directly overhead and incident on a well-mixed isothermal layer in
which the absorption coefficient kl has a constant value in the vertical. For an isothermal
atmosphere, combining the equation of state and the hypsometric equation, one gets

Ë [ Û
S S FYQ Ì Ü (4.30)
Í)Ý
where r0 is the air density at the mean sea level. Substituting for the density from Eq. (4.30)
in Eq. (4.25), one gets
‡
[
T M LM S Ô FYQ ËÌ ÛÜ E[ (4.31)
[
Í) Ý
Integrating the integral in Eq. (4.31), one gets

Ë [ Û
TM )LM S FYQ Ì (4.32)
Í ) ÜÝ
The absorbed radiation within an atmospheric layer is given by combining Eqs. (4.22), (4.24)
and (4.26), to get
dFl = Fl¥ tl dal (4.33)
where tl is the transmittivity of the portion of the atmosphere which lies above the
considered atmospheric layer. Substituting for tl from Eq. (4.26), dal from Eq. (4.22) and
r from Eq. (4.30), one gets
dFl = (Fl¥ kl r0) exp (–z/H) exp (– sl) (4.34)
RADIATION u 91

Substituting from Eq. (4.32) for exp (–z/H) in Eq. (4.34), one gets an expression for the
absorption per unit thickness of the layer as a function of optical depth
E'M 'M‡
T M FYQ T M
(4.35)
E[ )
E'M
At the level of the strongest absorption the first derivative of has to identically vanish, i.e.
E[
E È E'M Ø 'M ‡ E
ÉÊ <T M FYQ  T M
> (4.36)
E[ ÙÚ

E[ ) E[
The above differentiation results in
'M‡ ET M
FYQ  T M
  T M
 (4.37)
) E[
implying sl = 1, i.e. the strongest absorption occurs at the level corresponding to unit optical
E'M
thickness. Hence, it is clear that the vertical profile of the absorption ratio has a
E[
E'M
maximum corresponding to sl = 1. It is to be noted that according to Eq. (4.22), — 'M S
E[
For levels corresponding to sl <<1, the incoming solar radiation is absorbed very little due to
the low values of density and availability of very few air molecules required to produce
adequate absorption. Again, for levels corresponding to sl >>1, the density of air is much
higher. However, there is very little of radiation available to absorb and hence the radiation is
again virtually undepleted. The larger the value of the absorption coefficient, the smaller is the
air density required to provide for significant absorption and hence the higher the level of unit
optical thickness. In a similar manner, for smaller values of absorption coefficient, the
radiation may reach the bottom of the atmosphere long before it reaches the height
corresponding to unit optical thickness. Despite, the simplifying assumption of isothermal
atmosphere and a constant absorption coefficient, it is found that the overall results of the
above model are still qualitatively valid for realistic vertical profiles of the temperature and
absorption coefficients.

4.4 SCATTERING OF SOLAR RADIATION


Analogous to the absorption of solar radiation (refer Eq. (4.22)), one can also formulate a
similar approach to studying the scattering of the solar radiation. Like Eq. (4.22), the
scattering of solar radiation (parallel beam radiation) as it passes downward through a
horizontal layer of the atmosphere of infinitesimal thickness dz is given as
E'M
ETM L" TFD G E[ (4.38)
'M
where k is a dimensionless coefficient called scattering area coefficient and A is the cross-
sectional area that the particles in a unit volume expose to the beam of incident radiation.
92 u BASICS OF ATMOSPHERIC SCIENCE

Here Asecf dz represents the fractional area occupied by the particles, if all the particles
which the incident beam encounters in its passage through the differential layer, were to be
projected onto a plane perpendicular to the incident beam. The scattering area cross-section k
effectively measures the ratio of the effective scattering cross-section of the particles to their
geometric cross-section. Assuming, that only the scattering effects are considered, while
ignoring the absorption effects, Eq. (4.38) can be integrated to obtain expressions similar to
Eqs. (4.23)–(4.27). In the real atmosphere, a variety of particle shapes and a whole system of
particle sizes are always present. However, it is convenient to consider the idealized case of
scattering by particles of uniform radius r which are spherical and for which the scattering
area coefficient k can be prescribed on the basis of theory. The scattering area coefficient k
depends on the index of refraction of the particles responsible for the scattering together with
Q S
the particle size parameters, the latter expressed as B 
M
For the case of a << 1, (Rayleigh scattering regime), the scattering area coefficient is
proportional to the fourth power of the size parameter and for this case, the scattered radiation
is evenly distributed between the forward and backward hemispheres. The scattering of the
solar radiation by air molecules and the scattering of microwave radiation by raindrops
correspond to this case. For values of the size parameter a lying between 0.1 and 50
(Mie scattering regime), the scattering phenomenon utilizes the more general Mie theory of
scattering. In the Mie scattering regime, the angular distribution of scattered radiation varies
in a complicated and rapid manner with the size parameters, leading to a result where the
forward scattering dominates over the back scattering. The scattering of sunlight by particles
of haze, smoke, smog, and dust correspond to the Mie scattering regime. For high values of
the size parameter a > 50 (geometrical optics regime), the angular distribution of the
scattered radiation requires the principle of geometrical optics for its description. The
scattering of visible radiation by cloud droplets, raindrops and ice particles correspond to the
geometric optics scattering regime. Some of the important optical phenomena such as
rainbows, halos etc. are due to the scattering where the size parameter is larger (a > 50).

4.5 ATMOSPHERIC ABSORPTION AND EMISSION OF


INFRARED RADIATION
Some of the constituents of the atmosphere such as water vapour and carbon dioxide absorb
the long wave infrared radiation emitted by the earth and in turn radiate the infrared radiation
partly to earth as well as to space. The presence of these above-mentioned constituents of the
atmosphere called the greenhouse gases ensures that the average temperature of the earth
system is higher and the above effect of the presence of greenhouse gases is known as
greenhouse effect.
Unlike the solar radiation, which can be considered as parallel beam radiation, the
transfer of radiation from the earth surface through the different layers of the atmosphere is
complicated since the terrestrial radiation is of a different type. This necessitates the need to
consider the integration over the solid angle to calculate the irradiance. However, to make the
analysis simple in this section, the geometrical issues associated with the treatment of diffuse
RADIATION u 93

radiation are disregarded. This will enable us to concentrate on the physical processes, which
govern the absorption and emission of infrared radiation in the earth’s atmosphere.
It is to be noted that all the relationship derived for parallel beam radiation carry over to
the diffuse radiation except that the radiance takes the role of irradiance in the diffuse
radiation case. Furthermore, in the diffuse radiation case, the absorptivity and emissivity are to
be interpreted as the radiance along a given path length which is absorbed or transmitted
through the given atmospheric layer. Unlike the solar radiation where only the absorption
effects was to be considered since the atmospheric emission is negligible at these
wavelengths, for the terrestrial radiation both the absorption and emission effects need to be
simultaneously considered.
The absorption of terrestrial radiation along an upward path through the atmosphere is
given by Eq. (4.22) with the radiance replacing irradiance and with the sign reversed and is
given by
 E-M -M LM S TFD G E[ (4.39)
Using Kirchoff’s laws and employing similar arguments as in Eq. (4.39) for the emission of
radiation from the atmosphere, one gets
E-M #M EF M #M EBM #M LM S TFD G E[ (4.40)
where Bl is the black body monochromatic radiance given by Planck’s law.
Obtaining, the net contribution of the layer, to the monochromatic radiance, of the
terrestrial radiation passing upward, through a layer of atmosphere, from Eqs. (4.39) and
(4.40), one gets
E-M  LM -M  #M
S TFD G E[ (4.41)
Equation (4.41) is called the Schwarzchild’s equation and is extensively used as the basis for
the computation of the transfer of infrared radiation.
For an isothermal atmosphere with constant absorption coefficient, Eq. (4.41) can be
integrated to give
-M  #M
-M   #M
FYQ  T M
(4.42)
where Ll0 is the radiance incident on the atmospheric layer from below. Equation (4.42)
suggests that Llÿ should approach Bl exponentially with the increase of the optical thickness.
Furthermore, for a layer of infinite optical thickness, the emission from the top is Bl,
irrespective of the value of Ll0; implying that such a layer behaves as a perfect black body.
Figure 4.6(a–b) shows that the absorption of terrestrial radiation by water vapour, (refer
Figure 4.6(a), and carbon dioxide (refer Figure 4.6(b)), in the atmosphere. A broad wavelength
region of weak absorption between 8 mm and 12 mm, except possibly for a band near 9.6 mm
associated with absorption of ozone is noticed. Water vapour absorbs strongly over a wide
band of wavelength near 6.3 mm (about 1200 to 2000 cm–1) and over a narrower band near
2.7 mm. The above two absorption bands of water vapour are associated with transitions
involving the vibrational modes. Water vapour has another absorption band above 16 mm
(less than 500 cm–1) associated with transitions involving the rotational modes, as shown in
Figure 4.6. Carbon dioxide absorbs strongly (refer Figure 4.6(b)) in a broad band near 15 mm
(about 600 to 800 cm–1) associated with the vibrational modes while there are narrow bands
94 u BASICS OF ATMOSPHERIC SCIENCE

of carbon dioxide absorption near 4.3 mm and 2.7 mm. This spectral region in the wavelength
interval 600 to 800 cm–1 also corresponds to the maximum intensity of the Planck function in
the wave number domain. Except for ozone, which has a strong absorption band near 9.6 mm,
the atmosphere is relatively transparent from 800 to 1200 cm–1. This region is referred to as
the atmospheric window.
1.00

0.80
Transmittance

0.60

0.40

0.20

0.00
320 328 336 344 352 360 368 376
(a) Wave number (cm–1 )

1.00

0.80
Transmittance

0.60

0.40

0.20

0.00
680 688 696 704 712 720 728 736
(b) Wave number (cm–1 )

FIGURE 4.6 Infrared absorption spectrum for two strongly absorbing gases for a vertical beam
passing through the atmosphere in the absence of clouds. The top panel (a) shows
the water vapour rotational band and the bottom panel (b) shows the 15 mm
carbon dioxide band.

4.6 REMOTE TEMPERATURE SOUNDING FROM SPACE


In this section, the physical principles underlying the basis for the sea surface temperature and
the remote temperature sounding from space (i.e. satellites) are outlined. It is to be noted that
the passive sensor in the satellite measures the radiation lost to space by the earth atmospheric
system. Let us consider a satellite sensor receiving the infrared emission from the atmosphere
along a vertical path. Assuming local thermodynamical equilibrium and using Eq. (4.41), the
spectral radiance Ln at a frequency n received by the satellite sensor is given by
‡
˜U O [ ‡

-O Ô #O <5 [
> ˜[
E[  #O 5T
U O  ‡
(4.43)

RADIATION u 95

where tn (z, ¥) is the spectral transmittance between the height z and the satellite sensor, Ts is
the surface temperature and T(z) is the temperature of the atmosphere at the height z. For the
infrared portion of the electromagnetic spectrum, it can be assumed that the surface acts as a
black body. The first term in the right-hand side of Eq. (4.43) refers to the atmospheric
extinction term, while the second term in the right-hand side of Eq. (4.43) corresponds to the
surface contribution.

4.6.1 Calculation of the Surface Temperature


The surface temperature can be calculated from Eq. (4.43) if the atmospheric extinction term
can be completely neglected. This will correspond to tn (0, ¥) = 1 every where and the
satellite only receives the surface contribution with no atmospheric interference. This would
correspond to the wavelengths in the atmospheric window region wherein the atmospheric
interference is least. From a knowledge of the radiance measurement Ln, one directly gets
Ln = Bn (Ts). All one has to do is to invert the Planck function and from the knowledge of the
radiance measurement, one can calculate the surface temperature Ts as
IO
5T (4.44)
È IO  Ø
L MO É   Ù
Ê D -O Ú
where k is the Boltzmann constant, h is the Planck’s constant and c is the speed of light.
It is possible to define a brightness temperature with respect to the measured radiance
Ln, using Eq. (4.44), even in the presence of atmospheric interference. Consider an isothermal
atmosphere at constant temperature T0 which is not the same as the surface temperature Ts.
Let the frequency n be close to a spectral line centred at a frequency n0 so that the absorption
by the atmosphere (the first term in the right-hand side of Eq. (4.43)) cannot be disregarded.
In the above case, the integral in Eq. (4.43) can be calculated and Eq. (4.43) becomes
-O #O 5
<  U O  ‡
>  #O 5T
U O  ‡
(4.45)
Writing Eq. (4.45) in terms of the total atmospheric absorption An (0) where An (0) = 1 – tn (0, ¥),
Eq. (4.45) becomes
-O #O 5
"O 
 #O 5T
<  "O 
> (4.46)
Equation (4.46) may be rearranged as
-O #O 5T
 < #O 5
 #O 5T
> "O 
(4.47)
Hence, from Eq. (4.47) the measured radiance Ln, equals the Planck function contribution
from the surface plus a term equal to [Bn (T0) – Bn (Ts)] An (0). It is to be noted that any
rapidly varying frequency dependence of Ln, must be attributed to the total absorptance An,
rather than to the Planck function Bn (Ts).
Figure 4.7(a) shows the schematic plot of the observed satellite radiance (solid line) near
a spectral line for an isothermal atmosphere that is warmer than the surface temperature Ts.
Since T0 > Ts, Bn (T0) > Bn (Ts) and the observed satellite radiance shows a “hump” around the
96 u BASICS OF ATMOSPHERIC SCIENCE

line centred at n0 as shown in Figure 4.7(a). Since in this case, T0 > Ts, the spectral line at n
= n0 corresponds to an emission line. Figure 4.7(b) is similar to Figure 4.7(a) except that the
former corresponds to the case when the isothermal atmosphere T0 is colder than the surface
temperature Ts. Since T0 < Ts, Bn (T0) < Bn (Ts) and the observed satellite radiance shows a
“dip” around the line centred at n0 as shown in Figure 4.7(b). Since T0 < Ts, the spectral line
at n = n0 corresponds to an absorption line.

FIGURE 4.7(a) Schematic diagram of satellite observed radiance (shown as solid line) near a
spectral line for an isothermal atmosphere at higher temperature than the
surface temperature. The Planck’s function at the air temperature and surface
temperatures are shown as dotted lines.

FIGURE 4.7(b) Schematic diagram of satellite observed radiance (shown as solid line) near a
spectral line for an isothermal atmosphere at lower temperature than the
surface temperature. The Planck’s function at the air temperature and surface
temperatures are shown as dotted lines.

4.6.2 Two-layer Atmospheric Temperature Profile


Consider a two-layer atmospheric temperature profile comprising the troposphere and the
stratosphere. Assume that both the troposphere and the stratosphere have uniform
RADIATION u 97

temperatures in the vertical with the tropospheric temperature T2 being less than the
stratospheric temperature T1. Let the surface temperature Ts be greater than the stratospheric
temperature T1 and let the transition between the troposphere and the stratosphere correspond
to a certain height z = z1. Applying Eq. (4.43) to this case, one gets
-O #O 5T
 < #O 5
 #O 5T
> "O 
 < #O 5
 #O 5
> "O [
(4.48)
Equation (4.48) shows that the observed satellite radiance for a two-layer atmosphere is given
by the “background” term corresponding to the Planck function at the surface temperature
Bn (Ts). As shown in Figure 4.8, superposed on the background Planck function of the surface
temperature is a large dip since T2 < Ts and Bn (T2) < Bn (Ts). Furthermore, Figure 4.8 clearly
shows that in the middle of the above-mentioned large dip is a small hump, since T1 > T2 and
Bn (T1) > Bn (T2). In the figure, it is assumed that the total atmospheric absorptance is unity at
the surface and equals 0.6 at the tropopause corresponding to the central frequency n0.

FIGURE 4.8 Schematic diagram of satellite observed radiance (shown as solid line) for a
two-layered atmosphere having temperature T2 at lower levels and
temperature T1 at upper levels. The Planck function at the two air temperatures
T2 and T1 are shown as dotted lines. It is assumed that T2 < T1, while the
surface temperature Ts > T1.

4.6.3 Multi-layer Atmospheric Temperature Profile


The physical basis for deriving the temperature profile from the observed satellite radiance
can be simply stated as follows. At wavelengths close to the centre of the absorption lines,
the absorption coefficient is large and hence a very short path length is adequate to absorb
virtually all the incident radiation. However, at wavelengths away from the centre of the
absorption lines, a very large path length may be necessary to produce any significant
absorption. From the above, it is clear that the contribution to the observed variation at the
centre of the absorption line must necessarily come from the top regions of the atmosphere,
while the contribution to the observed radiance at wavelengths away from the centre of the
absorption lines, must come from the bottom layers of the atmosphere.
Extending the two-layer atmospheric temperature profile to the multi-layer atmosphere, it
is convenient to utilize the log pressure coordinates
98 u BASICS OF ATMOSPHERIC SCIENCE

È Q Ø
; MO
ÉÊ Q ÙÚ (4.49)

where p0 is the surface pressure. Substituting the above equation in Eq. (4.43), one gets
‡
Ln = Ô #O <5 ;
>, ;
E;  #O 5
UO  ‡

T (4.50)


˜U O ; ‡

where K(Z) = (4.51)


˜;
is the weighting function. It is to be noted that the weighting function K weights the
contribution to Ln from the Planck function at different log-pressure altitudes Z.
It is possible to calculate the weighting function Z for a simple case such as the case for
which the absorption coefficient kl is constant with height and the absorber has a constant
SB [

mass mixing ratio N B where r is the total atmospheric density and ra is the density
S [

of the absorbing gas. The expression for the transmittance for the satellite sensor receiving
infrared emission from the atmosphere along a vertical path is given by

ˇ Û
U O [ ‡
FYQ Ì
Ì
Ô LO [ „
SB [ „
E[ „ÜÜ (4.52)
Í[ Ý
where kn is the absorption coefficient. When the scattering processes are also considered, kn
will refer to the extinction coefficient in Eq. (4.52). For the above simple case, the
transmittance takes a simple form

Ë LO N B Q Û
U O [ ‡
FYQ Ì  Ü FYQ<  Y Q> (4.53)
Í H Ý

LO N B
where Y (4.54)
H
From Eq. (4.53) it is clear that x p is the optical thickness from the satellite sensor down to
the pressure p. The weighting function K is given in terms of pressure from Eq. (4.49) and
Eq. (4.51) and is given as
˜U [ ‡

, Q O Y Q FYQ<  Y Q> (4.55)


˜Q
The weighting function K can also be written in terms of Z and is given as

, ;
Y Q FYQ<  \;  Y Q F  ; ^> (4.56)
The variation of the weighting function with respect to Z as given in Eq. (4.56) is shown in
Figure 4.9. From Eq. (4.55) it is clear that the weighting function takes a maximum value of
RADIATION u 99

 H Z
e–1 at a pressure of Q QN and at
Y LO N B
È LO N B Q Ø
; ;N Y Q

MO MO
ÉÊ H ÙÚ corresponding Z1

to unit optical depth. In the neighbourhood of


this level, the integral in Eq. (4.50) tends to be
dominated by the Planck function. From the
above, it is clear that pm is inversely proportional Zm
to the absorption coefficient kn while Zm is linear
in ln(kn). If the measurements of the radiance at
the satellite sensor are taken at various different Z2
frequencies n, corresponding to different absorption
coefficients kn, it is clear that different vertical Weighting function
regions of the atmosphere will therefore dominate
FIGURE 4.9 Variation of the weighting
and contribute to the integral in Eq. (4.50) function with Z.
corresponding to the different frequencies. The
above provides the physical basis of a standard method of remote sounding. That is, from the
observed radiance measurements at different frequencies, one can infer the temperature
measurements of different vertical regions of the atmosphere.
In order to estimate the thickness of the weighting function in the neighbourhood of the
maximum Zm, one needs to find the pressure at which the weighting function equals to half its

maximum value. Since the maximum value of the weighting function is we need to find
F

the pressure corresponding to a weighting value of  From Eq. (4.55), one gets
F



Y Q FYQ<  Y Q> (4.57)
F
Solving Eq. (4.57), one gets the half maximum pressure points at p1 = 0.23 pm and p2 = 2.68 pm.
This corresponds to Z1 = Zm + 1.46 and Z2 = Zm – 0.99, which provides for a width in Z of
Z1 – Z2 = 2.45 scale heights. Assuming a typical value of 8 km as scale height the width of
the weighting function has a value about 20 km. The above broad nature (~ 20 km) of the
weighting function is a major limitation as the averaged temperature corresponding to this
thick 20 km atmospheric layer may not be a very useful quantity. Furthermore, the weighting
function corresponding to different frequencies also overlap among one another. It is possible
to obtain relatively narrow weighting functions by considering cases where the absorption
coefficient is varying with height or by using the limb rather than the nadir sounding. Also,
due to poor signal-to-noise ratio, the infrared remote sounding based on satellite
measurements from isolated parts of single spectral lines are not utilized. Hence, averages
over many lines need to be considered in a manner that preserves the desirable characteristics
of the weighting function.
It is important to utilize an absorbing gas whose mixing ratio is uniform in the
atmosphere for remote soundings of the temperature profile. For the infrared region, the
100 u BASICS OF ATMOSPHERIC SCIENCE

emission from carbon dioxide is usually utilized to derive the vertical temperature structure of
the atmosphere since the mixing ratio of carbon dioxide is almost independent of height in the
lower and in the middle atmospheres. Once the temperature profile T(z) has been calculated, a
variant of the above-mentioned approach provides for the measurement of composition since
the absolute density ra can now be regarded as the unknown quantity. While downward
looking sounding, also called nadir sounding is common, many satellites also observe the limb
of the atmosphere for retrieval of atmospheric properties of temperature and composition.
Limb sounding of the atmosphere has some advantages over the nadir sounding, namely
(i) better vertical resolution of a limb sounder over nadir sounding, (ii) higher sensitivity in
the limb sounding to trace gases with low atmospheric concentration. The disadvantages of
limb sounding are to do with the quality of limb retrieval as well as the increased cloud
interference in limb sounding.

SOLVED EXAMPLES

1. Using a formal integration over solid angle, calculate the arc of solid angle subtended by the
sky when viewed by a point on a horizontal surface.
Solution: Assume a spherical coordinate system centred on a point on the surface with j
and q being the azimuth angle and the zenith angle. The required arc of solid angle is then
given as
 Q Q  Q 

Ô EX Ô Ô TJO R ER EG  QÔ TJO R ER Q

Q
 G  R  R 

2. The flux density Fs of the incoming solar radiation, assumed isotropic, incident upon a
horizontal surface at the top of the atmosphere at zero zenith angle is called the solar constant
and has a value of 1368 W m–2. Find the intensity of solar radiation assuming that the radius
of sun Rs is 7 ´ 108 m and the distance between the sun and earth d is 1.5 ´ 1011 m.
Solution: Since the zenith angle is zero, the sun is directly overhead. Also, given that the
solar radiation is isotropic, the flux density of the solar radiation at the top of the earth’s
atmosphere Fs can be written in terms of the intensity of solar radiation Is as follows:

'T Ô* T DPT R EX
EX

where dw is the arc of the solid angle subtended by the sun in the sky and q is the angle
between the incident radiation and the direction normal to the surface. Since dw is very
small, one can ignore the variations of cos q in the above integrations, getting the so-called
parallel plane approximation, given by
'T * T DPT R EX
Since zenith angle is zero, cos q is 1, resulting in

'T * T EX
RADIATION u 101

The fraction of the hemisphere of solid angle occupied by the sun is the same as the fraction
of the area of the hemisphere of radius d, centred on the earth, yielding

EX Q 3T  Ë  –  Û
Ì  Ü
 EX  –   TS
Q Q E   ÌÍ –  ÜÝ

'T 
*T  –  8 N o TS o
EX  –  
3. Find the flux density of the emitted radiation, assuming that the radiation is emitted from a
plane surface with a uniform intensity in all directions.
Solution: The following expression gives the flux density of emitted radiation
 Q Q  Q 
' Ô * DPT R EX Ô Ô * DPT R TJO R ER EG Q * Ô DPT R TJO R ER Q*
 Q G  R  R 

4. Assuming the wavelength of maximum solar emission to be 0.475 mm, find the colour
temperature of sun using the Wien’s displacement law.
Solution: From the Wien’s displacement law,

 
5  ,
MN 
5. Find the black body monochromatic irradiance of green light of wavelength 0.53 mm from
an object of temperature 2000 K.
Solution: Planck’s law giving the amount of black body radiative flux, i.e. irradiance is
given by

D Ë  –  


 Û   
&M Ì Ü  –  8 N PN
Ë Û Î  Þ
M  Ì FYQ ÈÉ
D Ø ÌÑ Ë  –  Û Ñ Ü
 Ü
Í
Ù
Ê L5 Ú Ý Ì ÏFYQ Ì  
Ü  ß Ü
ÌÍ ÑÐ Í Ý Ñà ÜÝ
where c1 = 3.74 ´ 108 W m–2 mm–4 and c2 = 1.44 ´ 104 mm K
6. What is the total irradiance emitted from a black body earth of temperature 255 K.
Solution: Using Stefan Boltzmann law, one finds
E = s T 4,
E = 5.67 ´ 10–8 ´ (255)4 = 239.7 W m–2
7. Estimate the value of the solar constant, given that the radius of sun is 7 ´ 108 m, earth orbit
radius is 1.5 ´ 1011 m and the temperature of the sun is 5780 K.
Solution: From Stefan Boltzmann law, one finds
E = s T 4,
E = 5.67 ´ 10–8 ´ (5780)4 = 6.3284 ´ 107 W m–2
102 u BASICS OF ATMOSPHERIC SCIENCE

From inverse square law, one finds the solar constant as


Ë  –  Û
&T  – 
Ì  Ü
 8 N 
ÍÌ –  ÝÜ
8. Calculate the equivalent black body temperature TE of the outer visible surface of the sun,
assuming that the flux density of the solar radiation reaching the earth surface is 1368 W m–2,
the radius of sun Rs is 7 ´ 108 m, and the average sun–earth distance d is 1.5 ´ 1011m.
Solution: The flux density at the top of the outer layer of sun is calculated using inverse
square law as follows:

3  Ë –  Û
'UPQ 'T ËÌ T ÛÜ  Ì  Ü
 –  8 N 
ÍE Ý ÍÌ  –  ÝÜ
Using the Stefan Boltzmann law, one gets
Ftop = T 5&
 
Ë' Û Ë  –  Û
 


TE = Ì  ,
UPQ

Ü Ì  Ü
Í T Ý ÍÌ  –  ÜÝ


9. Calculate the equivalent black body temperature of the earth assuming that the earth is in
radiative equilibrium, i.e. experiences no net loss or gain of energy through radiative
transport. Also assume that the planetary albedo is 0.35.
Solution: Let Fs be the flux density of solar radiation incident on earth given by 1368 W m–2,
FE the flux density of the terrestrial radiation emitted by the earth, RE the radius of earth, A
the planetary albedo, TE the equivalent black body temperature of the earth, then from
Stefan Boltmann law

Ë   "
Q 3& 'T Û Ë  
Û 
'& T 5& Ì Ü ÌÍ ÜÝ  8 N
Q 3&

ÌÍ ÜÝ 

Solving for TE, one gets


 
Ë '& Û Ë  Û
 

5& ÌT Ü Ì  Ü  ,
Í Ý ÌÍ  –  ÜÝ


10. Assuming Rayleigh scattering by air molecules, find the relative efficiencies with which red
light (l = 0.64 mm) and blue light (l = 0.47 mm) are scattered.
Solution: Using the wavelength dependence of Rayleigh scattering, one finds the ratio of
relative efficiencies as
, CMVF È  Ø


ÉÊ Ù 
, SFE  Ú
11. Parallel beam radiation is passing through a layer of thickness 50 m, containing an absorbing
gas with an average density of 0.1 kg m–3. The beam is directed at an angle of 60°
with respect to the normal to that layer. Calculate the optical thickness, transmittivity and
RADIATION u 103

absorptivity of the layer at the three wavelengths for which the mass absorption coefficient
are 10–3, 10–1 and 1 m2 kg–1.
Solution: The mass of the absorbing gas encountered by the parallel beam radiation along
its slant path length equals
;5
V TFD R Ô S SE[
;#

where ZB and ZT are the heights of the bottom and top of the layer. Substituting sec q = 2, r
= 0.1 kg m–3, r = 1 and layer of thickness = 50 m, one gets

u = 2 ´ 0.1 ´ 1 ´ 50 = 10 kg m–2
Assuming the mass absorption coefficient kl to be uniform within throughout the layer, the
transmittivity of the layer can be written as

Tl = exp (– tl) = exp [– kl u]


Similarly, the absorptivity is defined as al = 1 – Tl = 1 – exp [– kl u], where the slant path
thickness
;5
UM LM TFD R Ô S SE[ LM V
;#

Substituting for u and kl for the three wavelengths, one gets


Transmittivity Tl = 0.99, 0.368, 2.06 ´ 10–9.
Absorptivity al = 0.01, 0.632, and 1.0
Slant path optical thickness tl = 0.01, 1.0 and 10

REVIEW QUESTIONS

1. Consider the wall (4 m tall and 6 m wide) in a room that has a temperature of 25°C.  If the
wall acts as a black body, what is the total amount of radiation it is emitting? And, at what
wavelength is it emitting most of its radiation?
2. Let the surface body temperature average 37°C. How much radiant energy in W m–2 would
be emitted from your body? What is the total radiant energy emitted by your body in watts?
At what wavelength is this radiant energy emitted?
3. It is known that moon has no atmosphere. How is the heat transferred away from the surface
of the moon?
4. Which of the below-mentioned processes would have the greatest effect on the earth’s
greenhouse effect: removing all of the CO2 from the atmosphere or removing all of the water
vapour?
5. An increase in the cloud cover surrounding the earth would naturally increase the earth’s
albedo, yet not necessarily lead to a lower earth surface temperature. Explain.
6. The solar declination angle ds is defined as the latitude where the sun angle is 90°. For
example, the solar declination angle on 22 June of every year would be 23.5°N.  For any
104 u BASICS OF ATMOSPHERIC SCIENCE

given day of the year, the expression for the solar declination angle is given by:

Ë Q E  
Û
ET )S DPT Ì Ü
ÍÌ EZ ÝÜ
where Fr is the tilt angle of the earth’s axis of rotation, d is the Julian Day (relative day of
the year), and dy is the number of days in the year. Find the solar declination angle on January 1,
March 1, May 1, July 1, September 1, and November 1. Also make a plot of solar declination
angle versus time of the year.
7. Consider the following two hypothetical situations: (a) the tilt of the earth is decreased to
10°, (b) the tilt of the earth is increased to 40°. How would the above change in tilt alter
the summer and winter temperatures at 23°N, in qualitative terms?
8. At the top of the earth’s atmosphere during the early summer (Northern Hemisphere), above
what latitude would one expect to receive the most solar radiation during a day? During the
same time of year, where would one expect to receive the most solar radiation at the surface?
9. The solar constant has a value of 1368 W m–2. What would be the value of the solar constant,
if the distance between the sun and the earth were to (i) double, and (ii) become half of the
present distance, with everything else being the same.
10. Name the most important factor responsible for the existence of seasons on earth.
11. What is the significance of the Arctic and Antarctic circles?
12. The equator always has 12 hours of sunlight. Why?
13. Arrange the following in increasing order of wavelengths: (i) visible, (ii) ultraviolet, (iii) X-ray,
(iv) infrared, and (v) microwave radiation.
5 Clouds and
Precipitation

The development of cloud droplets (of typical size of 10 mm) from aerosol particles having a
size range between 0.01 mm and 0.1 mm and the subsequent growth of cloud droplets to a rain
drop having a size of 1000 mm is a fascinating study by itself. Furthermore, the fact that the
above growth through such a wide range of sizes happens in a very short time (of the order of
ten minutes for convective clouds) provides adequate evidence to the extreme complexity of
the physical processes involved in the above growth.
The earlier studies dealing with the physics of clouds were invariably limited to the
laboratory experiments. In these laboratory experiments, attempts were made to isolate
physical processes of possible importance in clouds and study their characteristics. In recent
times, however, properties of clouds and precipitation are investigated using sophisticated
instrumented platforms such as aircraft and other remote sensing devices. Furthermore, with
the advent of availability of high-speed computers, a concerted effort is being made towards
developing numerical models of clouds. The results of such numerical simulation of cloud
processes are then verified with actual observations of real clouds in the atmosphere.
The importance of the atmospheric aerosol in cloud microphysical processes is presented
in Section 5.1. The next section is concerned with the formation of cloud droplets. Sections 5.3
and 5.4 outline the structure and the growth of cloud droplets in warm and cold clouds. While
Section 5.5 outlines the different mechanism of formation of clouds and cloud seeding, the
last Section 5.6 deals with the role of the cloud and precipitation products in the separation of
the electrical charges in the atmosphere.

5.1 ATMOSPHERIC AEROSOLS


Aerosols are minute particles which are suspended in the air. Aerosols do scatter and absorb
the solar radiation. Aerosols also play an important role in the heterogeneous nucleation
leading to the development of cloud particles. The size of the cloud particles in turn

105
106 u BASICS OF ATMOSPHERIC SCIENCE

determines the reflection and scattering of solar radiation by clouds. Hence the aerosols play a
very important role both directly and indirectly in the earth’s energy budget.

5.1.1 Aerosol Size and Concentration


A suspension of solid or liquid particles in the atmosphere, other than precipitation products is
known as an aerosol. Aerosols originate from both natural as well as man-made
(anthropogenic) sources. Aerosols can be directly emitted as particles; they are then called
primary aerosols. The primary aerosols are emitted into the atmosphere by any one of the
following processes: (i) volcanic emissions, (ii) the effect of wind lifting dust particles in arid
regions, (iii) through combustion during biomass burning, and (iv) from sea spray, vegetation
etc. Aerosols can also be produced as a result of chemical reactions (gas-to-particle
conversion) and these aerosols are then called secondary aerosols. While the sources of
primary aerosols which originate from natural sources are: (i) volcanic dust, (ii) sea salt,
(iii) mineral aerosol, and (iv) organic aerosol, the sources of primary aerosols which originate
from anthropogenic sources are: (i) biomass burning, (ii) industrial dust, and (iii) soot. The
sources of secondary aerosols which originate from natural sources are: (i) sulphates from
volcanic SO2, (ii) nitrates from NOx, (iii) organic aerosols from volatile organic compounds
(VOCs), and (iv) sulphates from biogenic gases. Three of the important sources of secondary
aerosols which originate from anthropogenic sources are the same as the first three listed
above under natural sources. In the troposphere, anthropogenic sources contribute significantly
to the aerosols. The chemical composition of aerosols due to both natural and anthropogenic
sources can be extremely variable, depending on the nature of the geographical region.
However, typical chemical compositions for the different geographical regions can be found.
Mostly, the aerosols found in the rural regions are made up of ammonium sulphate as well as
insoluble mineral material.  The aerosols found in the urban regions are made up of 80% rural
aerosols, while the remaining 20% is accounted by soot. While the small maritime aerosols
are mainly ammonium sulphate due to gas-to-particle conversion of SO2, the large maritime
aerosols are mostly in the form of sodium chloride formed due to sea spray evaporation.
Atmospheric aerosol ranges in size from about 10–4 mm to tens of micrometre and are
typically found in concentrations ranging from about 107 cm–3 to 10–6 cm–3. The smallest of
the aerosols is called the Aitken nuclei. These Aitken nuclei have sizes less than 0.2 mm, and
vary widely in concentration near the earth surface. The average concentration of the Aitken
nuclei near the earth surface are about 103 cm–3 over the oceans, 104 cm–3 over rural land
areas and 105 cm–3 or higher over polluted city areas. The concentration of the Aitken nuclei
over the land areas decreases considerably with height, having average values of 102 cm–3 at
12 km and 10 cm–3 at height of 18 km or above. The above rapid decrease in the
concentration of the Aitken nuclei with height conclusively shows that the land is an
important source of aerosol with human and industrial activities being the dominant source of
Aitken nuclei.
A very convenient method to obtain the concentration of the atmospheric aerosol is by
using an instrument called the Aitken nuclei counter. In this counter, saturated air is expanded
very rapidly to become supersaturated with respect to water by several hundred percent.
Associated with these high supersaturation values, a change of phase occurs almost
CLOUDS AND PRECIPITATION u 107

immediately with water condensing onto almost all of the aerosols present to form a cloud of
small water droplets. The concentration of droplets in the cloud which are nothing but the
concentration of aerosol can be easily determined by either an optical method or by allowing
the droplets to settle out onto a substrate, where they can be counted under a microscope.
Since the aerosol sizes exist over a wide range from 10–4 mm to 0.1 mm, it is observed
that different methods need to be employed to obtain measurements of these atmospheric
aerosols. The electrical aerosol analyzer is used to classify aerosols whose sizes range from
0.003 mm to 1 mm. In the above device, the aerosols are first given a known electrical charge
and are then collected in a controlled manner by the application of electric fields. The
variation in the magnitude of the collected charge is then related to the aerosol size
distribution. Aerosols of 0.3 mm to 30 mm are usually classified by measuring the amount of
light energy they scatter. Furthermore, aerosols of size greater than 0.1 mm may also be
collected by impaction onto different surfaces and sized using either optical or electron
microscopes. The composition of the larger aerosols whose size is greater than 1 mm may be
determined from analytical techniques such as energy dispersive analysis of X-rays. Since the
smaller aerosols with size less than 0.1 mm follow the streamlines more readily around the
collecting obstacle in the direct impaction methods, they are not collected as efficiently as the
larger aerosols with sizes greater than 0.1 mm. Due to the above reason, the determination of aerosol
concentration by direct impaction methods is generally considered only for the larger aerosols.
Figure 5.1 shows the number distribution of aerosols obtained from averaging several
sets of measurements made in continental, marine and the urban polluted air. The number

FIGURE 5.1 The number distribution of aerosols obtained from averaging many sets of
observations made in continental air ( ), marine air (– – –) and the urban
polluted air (..........). Also shown is Eq. (5.1) with b = 3 seen as dashed line which
is displaced from the other curves.
108 u BASICS OF ATMOSPHERIC SCIENCE

E/
distribution is given as and is plotted on a logarithmic scale with log D. Here, D is
E MPH %

the diameter of the aerosol and N is the concentration of aerosol with diameter greater than D.
Figure 5.1 clearly reveals that the aerosol concentration falls off very sharply with increasing
size. Furthermore, the straight-line portion of the curve in Figure 5.1 can be expressed as
E/
$%  C (5.1)
E MPH %

where C is a constant dependent on the concentration of the aerosol and – b is the slope of
the curve in Figure 5.1. The values of D typically lie between 2 and 4. Aerosols larger than
the Aitken nuclei have a value of b ~ 3. Also, the giant aerosols whose size is greater than
2 mm have similar concentration over continental, marine and urban polluted air. It is to be
noted that the smallest Aitken nuclei contributes only 10% to 20% of the total mass of the
aerosols, despite their large concentrations. Also, the large aerosol with sizes between 0.2 mm
and 2 mm and the giant aerosol (size is larger than 2 mm) make similar contributions to the
total mass of aerosols in the continental air.
The maximum particle concentration for the urban aerosol is seen in the small particle
sizes (radius of 10–2 mm) and drops markedly to a value of a few particles for aerosols of
10 mm radius. In addition to natural sources, the urban aerosol is mainly due to emissions
from industries, power plants, transportation and also from gas-to-particle conversion. The
maximum particle concentration for the rural aerosol is again seen in the small particle sizes
(radius of 10–2 mm). Also, for the rural aerosol, there is a lower concentration of small size
particles as compared to the urban aerosol. However, unlike the urban aerosol, the
concentration of rural aerosols remain high for radius up to 0.1 mm. The rural aerosols are
mostly from natural sources, however, anthropogenic rural aerosols are also observed. The
maximum particle concentration for the maritime aerosol is seen for a radius of 0.1 mm with
a typical concentration of 100 per cm–3. The maritime aerosol is mainly composed of the sea
salt which results from the evaporation of sea spray.

5.1.2 Sources and Sinks of Atmospheric Aerosol


The smallest Aitken nuclei originate mainly from combustion processes. Even though natural
causes such as volcanoes and forest fires do contribute to the formation of Aitken nuclei, the
chief source of Aitken nuclei are those associated with human activities. It is no wonder that
the highest concentration of the smallest Aitken nuclei is seen in the urban polluted air. Since
the concentration of the Aitken nuclei is considerable over marine air as well, it is seen that
sources other than combustion also play an important role in the formation of Aitken nuclei.
An important source for the formation of Aitken nuclei is associated with the conversion of
trace gases in the atmosphere into aerosols through the gas-to-particle-conversion. The above
conversion can occur through the nucleation of aerosol from supersaturated gases and by
photochemical reactions associated with the absorption of incoming solar radiation by
molecules. Furthermore, the presence of large relative humidity and liquid water further
enhances the above gas-to-particle conversion. The principal sources of large and giant
aerosols on the earth surface are wind-blown dust in arid regions, emission of pollens and
CLOUDS AND PRECIPITATION u 109

spores from plants and the bursting of air bubbles over oceans. There has been some evidence
of an increase in the aerosol concentration in the northern hemisphere during the twentieth
century. The above increase in the aerosol concentration in the northern hemisphere may be
due to anthropogenic sources.
Aerosols have important effects on the various chemical processes in the atmosphere.
The aerosol in the solid form readily provides a surface upon which the trace gases can be
absorbed, while the liquid aerosol can absorb gases and react in the solution. Also, aerosols
can react with sulphur dioxide to form sulphates and sulphuric acid. The above pollutants, if
found in stable moist air under a temperature inversion can persist for a long time leading to
serious health hazards. Aerosols also participate in the scattering and absorption of radiation.
The above effects can degrade (lower) the atmospheric visibility as well as alter the transfer
of solar radiation in the atmosphere. The exact consequences of the supposed increase in the
aerosol concentration in the northern hemisphere are rather hard to conclude. While many of
the aerosols are known to absorb the incoming solar radiation leading to an increase in the air
temperature, some of the aerosols are also known to scatter solar radiation back to space
causing a decrease in the air temperature. Due to the above cooling and heating effects, it is
difficult to predict the net changes in the average global temperature due to the supposed
increase in the aerosol concentration in the northern hemisphere.

5.2 NUCLEATION OF WATER VAPOUR CONDENSATION


Changes of phase from vapour to water occur when the air becomes supersaturated with
respect to the liquid water, while change of phase from vapour to ice occurs when the air
becomes supersaturated with respect to ice. Since most clouds form due to the ascending
motion of air parcels, ascending air parcels give rise to supersaturation. The change of phase
from water vapour to liquid water associated with the supersaturated air manifests when the
water vapour condenses onto some of the atmospheric aerosol particles present in the air to
form a cloud of small water droplets.

5.2.1 Thermodynamic Potentials


Before we discuss the theory of condensation, it is important to introduce the so-called
thermodynamic potentials. The Helmholtz free energy f of a unit mass of a body is defined as
f = u – Ts (5.2)
where u is the internal energy per unit mass, s is the specific entropy and T is the temperature.
Differentiating Eq. (5.2), one gets
df = du – Tds – sdT (5.3)
Combining the first and second laws of thermodynamics, one obtains
Tds ³ du + pda (5.4)
where the inequality sign holds for irreversible spontaneous transformation, while the equality
sign refers to reversible equilibrium transformation. Using Eq. (5.4) in Eq. (5.3) for a
110 u BASICS OF ATMOSPHERIC SCIENCE

reversible equilibrium transformation, one gets


df = – sdT – pda (5.5)
For a body at equilibrium under constant temperature and constant volume, Eq. (5.5) becomes
df = 0 (5.6)
Using Eq. (5.4) in Eq. (5.5) for a body undergoing spontaneous irreversible transformation,
one gets
df < –sdT – pda (5.7)
For an irreversible transformation at constant temperature and constant volume, from Eq. (5.7),
one gets
df < 0 (5.8)
From Eqs. (5.6) and (5.8), it follows that a body at constant temperature and constant volume
is in stable equilibrium when its Helmholtz free energy has a minimum value. The Gibbs free
energy g of a unit mass of a body is defined as
g = u –Ts + pa (5.9)
Using exactly analogous arguments as above, it can be shown that for a body in equilibrium
at constant temperature and constant pressure
dg = 0 (5.10)
Also, for an irreversible transformation at constant temperature and constant pressure, one
gets
dg < 0 (5.11)
It follows from Eqs. (5.10) and (5.11) that a body at constant temperature and constant
pressure is in stable equilibrium when its Gibbs free energy has a minimum value.
If a single molecule is removed from a material in a certain phase during which the
temperature and pressure remain constant, the resulting change in the Gibbs free energy of the
material is called the chemical potential m of that phase. Hence, the chemical potential m of a
given phase is simply the Gibbs free energy per molecule. The above-mentioned
thermodynamic functions play an important role while considering the changes of phases,
which occur in the atmosphere. It can be shown easily (refer to solved example of this
chapter) that when a plane surface of liquid is in equilibrium with its vapour, the chemical
potential corresponding to both the liquid and the vapour phases are the same. Furthermore,
for a plane surface of liquid at temperature T, it can be shown (refer to solved example of this
chapter) that the difference in the chemical potentials of the vapour and the liquid phases are
related by the following expression:
F
NW  N M
L5 MO (5.12)
FT
where mv and ml are the chemical potentials of the vapour and the liquid phases, respectively,
CLOUDS AND PRECIPITATION u 111

k is the Boltzmann constant, e is the actual vapour pressure, T is the temperature and es is
the saturated vapour pressure corresponding to a plane surface of liquid at temperature T.
For a reversible transformation, the first law of thermodynamics is expressed as
Tds = du + dwtot (5.13)
where dwtot is the total work done by a unit mass of a body. From Eqs. (5.5) and (5.13), for
a reversible transformation, one gets
dwtot = –df – sdT (5.14)
which at constant temperature, becomes
dwtot = –df (5.15)
Equation (5.15) shows that the total external work done by a body in a reversible isothermal
process is equal to the decrease in the Helmholtz free energy of the body.
Let da be the external work done by a unit mass of a body over and above any work of
expansion, i.e. pda. Hence, da can be written in terms of the total work done as
da = dwtot –pda (5.16)
Differentiating Eq. (5.9), one gets
dg = du – Tds – sdT + pda + adp (5.17)
Combining Eqs. (5.13), (5.16) and (5.17), one gets
da = –dg – sdT – adp (5.18)
For an isothermal and isobaric process, Eq. (5.18) becomes
da = –dg (5.19)
Equation (5.19) shows that the external work done by a body over and above the pda work
in a reversible isothermal isobaric process is equal to the decrease in the Gibbs free energy
of the body.

5.2.2 Nucleation Theory of Water Vapour Condensation


In this section, the theory of initiation (nucleation) of condensation of water vapour to liquid
water in the atmosphere is discussed. Consider the formation of a pure water droplet by
condensation from a supersaturated vapour without the presence of atmospheric aerosols. It
can be shown experimentally that nucleation in the absence of aerosols can occur only at very
high relative humidity, say of the order of 400%. The above nucleation in the absence of
atmospheric aerosols is known as homogeneous nucleation, or spontaneous nucleation of
condensation. Homogeneous nucleation occurs due to a number of water molecules in the
vapour phase coming together through chance collisions to form small embryonic water
droplets large enough to remain intact without dissipating.
The supersaturation expressed as a percentage with respect to liquid water is
ÈF Ø
ÉÊ F  ÙÚ –  where e is the water vapour pressure in the air and es is the saturation vapour
T
112 u BASICS OF ATMOSPHERIC SCIENCE

pressure over a plane surface of liquid water. The supersaturation with respect to ice is also
defined in a similar manner except that the saturation vapour pressure is now over a plane
surface of ice.
If mv and ml are the chemical potentials in the vapour and liquid phases, respectively, and
n is the number of water molecules per unit volume of liquid, the decrease in the Gibbs
energy of the system due to condensation of water vapour is given by nV (mv – ml). It is to be
noted that other than work associated with the change in volume of the system, work is done
in creating the surface area of the water droplet. The above-mentioned work required to create
the surface area of the water droplet is given by As, where s is the surface energy of water.
The surface energy can be interpreted as the work necessary to create a unit area of the
liquid–vapour interface.
If this transformation were an equilibrium transformation, which it is not, Eq. (5.10) will
hold. The external work done over and above the pdá work, which in this case equals to As
would be equal to – dg; the latter in this case equals nV(mv – ml). Since the transformation in
not equilibrium one, the change in the Gibbs free energy will differ from the work term As
and can be written as
DE = As – nV(mv – ml ) (5.20)
where DE is the net increase in the energy of the system due to the formation of water
droplet. Combining Eqs. (5.20) and (5.12), one gets
F
'& "T  O7L5 MO (5.21)
FT
where e and T are the vapour pressure and temperature of the system and es the saturation
vapour pressure over a plane surface of water at temperature T. For a spherical droplet of
radius R, Eq. (5.21) becomes
 F
'& Q 3 T  Q 3 OL 5 MO (5.22)
 FT

F
In air, which is sub-saturated, e < es, relative humidity < 100% and hence MO   and DE is
FT
always positive and increases with increasing radius (refer Figure 5.2). That is, the larger the
embryonic droplet that forms in the sub-saturated vapour, the greater is the increase in the
energy of the system. Since a system tends to equilibrium state by reducing its energy, the
above formation of a droplet under sub-saturated condition is highly unlikely. Even if some
very small embryonic droplets were to form due to random collisions of water molecules in
sub-saturated air, they will not grow large enough to become cloud droplets.
However, if the air is supersaturated, e > es, relative humidity > 100% and hence
F
MO !  From Eq. (5.22), it is seen that DE can be either positive or negative depending on
FT
the values of R. As seen in Figure 5.2, for the case, e > es, DE initially increases with
increasing R, reaches a maximum at R = r, called the critical radius and then decreases
subsequently with increasing R. Hence, for the case of e > es, small embryonic droplets with
CLOUDS AND PRECIPITATION u 113

Increase in Gibbs free energy (DE)


RH = 90%

RH = 110%

R=r Radius R (in mm)

FIGURE 5.2 Schematic diagram showing the increase DE in the energy of the system due to the
formation of a droplet of radius R from water vapour with relative humidity (RH) for
90% and 110%. The saturation vapour pressure with respect to a plane surface of
water at the same temperature of the system is indicated by es.

radius less than the critical radius (R < r) tend to evaporate. However, in supersaturated air, if
droplets manage to grow by chance to a radius, which just exceeds the critical radius, they
will continue to grow spontaneously by condensation from the vapour phase, since this way
they tend to decrease the total energy of the system. One can then obtain the expression for
˜ '&

the critical radius r in terms of e by setting,  at R = r. From Eq. (5.22), this gives
˜3
an expression for the critical radius r as follows:
T
S (5.23)
ÈFØ
OL5 MO É Ù
Ê FT Ú
Equation (5.23) is known as the Kelvin’s formula or Kelvin’s equation. The Kelvin’s formula
can be interpreted in two different ways. Firstly, it can be used to calculate the radius r of a
droplet which will be in unstable equilibrium with air at a given water vapour pressure.
Further, the Kelvin’s formula may be used to calculate the saturation vapour pressure e over a
droplet of spherical radius r. It is to be noted that the above-mentioned equilibrium is unstable
since if the embryo droplet gains a molecule, it will continue to grow by condensation, while
if it loses a molecule, it will continue to evaporate. The relative humidity, with which a
F
droplet of radius r is in unstable equilibrium, is given by 
FT
F
where is to be obtained from Eq. (5.23). Such a variation of relative humidity with droplet
FT
radius is shown in Figure 5.3.
It can be clearly seen from Figure 5.3, that a pure water droplet of radius, say equal to
0.01 mm requires a relative humidity of 111–5% in order to be in unstable equilibrium with its
114 u BASICS OF ATMOSPHERIC SCIENCE

environment, while a droplet of radius


1 mm requires 100.12% relative humidity. 114
Further, it is observed from Figure 5.3 that
112
as the radius r increases beyond 1 mm one

Relative humidity (%)


recovers the “plane surface” result, i.e. 110
relative humidity becomes 100%.
108
Since the supersaturation, which
develops in the real atmosphere, due to 106
ascending motion of air parcels in observed
104
clouds do not exceed 1%, (i.e. relative
humidity is less than 101%), it is observed 102
that even embryonic droplets of size such
100
as 0.01 mm will be well below the critical 0.01 0.1 1.0 10.0
radius necessary for survival at 1% Radius (mm)
supersaturation. Hence, from the above FIGURE 5.3 The relative humidity with respect
discussion, it is seen that droplets do not to a plane surface of liquid water
form in the natural clouds by the with which pure water droplets are
homogeneous nucleation of pure water. in unstable equilibrium at 5°C.
This leads to the conclusion that droplets
form in natural clouds by heterogeneous nucleation on atmospheric aerosols.
Since the real atmosphere contains a large number of atmospheric aerosols with sizes
ranging from 10–4 mm to 10 mm, it is observed that droplets can form in natural clouds by
heterogeneous nucleation on atmospheric aerosols. Those atmospheric aerosols, that are
wettable, i.e. those aerosols, which allow the water to spread out on it as a horizontal film,
can serve as centres upon which the water vapour can condense using heterogeneous
nucleation. Furthermore, droplets can form and grow on these wettable aerosols at much
lower supersaturation than are required for homogeneous nucleation. From Figure 5.3, it is
seen that if sufficient water condenses onto a completely wettable aerosol of size 0.3 mm, to
form a thin film of water over its surface, this water film will be in unstable equilibrium with
air if the supersaturation is 0.4%. Since, such supersaturation values are observed in the real
atmosphere, more water will condense onto the film and the droplet will grow in size,
provided the supersaturation is greater than 0.4%.
Further, a few of the atmospheric aerosols have an affinity to water and hence are
soluble in water. When water condenses on this soluble aerosol, the aerosol dissolves giving
rise to a solution droplet. It is to be noted that the equilibrium saturation vapour pressure over
a solution droplet is less than that over a pure water droplet of the same size. The above
behaviour is attributed to the fact that the saturation vapour pressure is proportional to the
concentration of solute molecules on the surface of the droplet. It is known that in a solution
droplet, some of the surface molecular sites are occupied by the molecules of salt (or ions if
the salt dissociates), and thus the vapour pressure is reduced by the presence of the solute.
The reduction in the vapour pressure, expressed as a fraction is given by

G (5.24)
F
CLOUDS AND PRECIPITATION u 115

where e¢ is the saturation vapour pressure over a solution droplet containing a kilomole
fraction f of pure water and e is the saturation vapour pressure over a pure water droplet of
the same size and at the same temperature. The kilomole fraction of pure water is then
defined as the ratio of the number of kilomoles of pure water in the solution droplet to the
total number of kilomoles (salt plus pure water) in the solution.
Consider a solution droplet of radius r, which contains a mass m of the dissolved salt of
molecular weight M. If one assumes that each molecule of the salt dissociates in water into
JN
i ions, the effective number of kilomoles of the salt in the droplet is  If the density of the
.T

solution is r¢ and Mw is the molecular weight of water, the number of kilomoles of pure water
È Ø
in the droplet is then given by É Q S  S „  NÙ .X  Hence, the kilomole fraction of water is
Ê Ú

È  Ø Ë Û
ÉÊ Q S S „  NÙÚ .X Ì Ü
 JN. X
G Ì  Ü (5.25)
È  È 
Q S S „  NØÙ .X  JN . T Ì ØÜ
. T É Q S S „  NÙ
ÊÉ  Ú ÌÍ Ê Ú ÜÝ
Combining Eqs. (5.23) and (5.25) and replacing s and n by s ¢ and n¢ to denote the surface
energy and the number density of water molecules in the solution, one gets the following
expression for the saturation vapour pressure e¢ over a solution droplet of radius r:

Ë Û
F „ Ë T „ Û Ì JN. X
Ü
ÌÍ FYQ O „L5S ÝÜ Ì  È  ØÜ
Ü (5.26)
FT Ì .T É Q S S „  NÙ Ü
ÌÍ Ê ÚÝ
Equation (5.26) can be used to determine the vapour pressure e¢ (or relative humidity 100 e¢/es
or supersaturation [e¢/es – 1] 100) of the air adjacent to the solution droplet of the given radius r.
If the variation of the relative humidity or of the supersaturation of the air adjacent to a
solution droplet is plotted as a function of its radius, one gets what is known as Kohler curves.
Figure 5.4 shows several such Kohler curves. Below a certain droplet size, the vapour
pressure of the air adjacent to a solution droplet is less than the value corresponding to
equilibrium with a plane surface of pure water at the same temperature. As the droplets
increase in size the solution becomes weaker and the curvature effect (Kelvin’s equation)
becomes the dominant influence. This leads to the situation where the relative humidity of the
air adjacent to the droplet becoming essentially the same as that over pure water droplets.
As a means of interpreting the Kohler curves, consider a solution droplet containing
10–19 kg of sodium chloride, which corresponds to a dry radius of 0.022 mm, i.e. curve 2 in
Figure 5.4. If the solution droplets were 0.05 mm in radius, the relative humidity of the air
adjacent to the surface would be 90%. If an initially dry sodium chloride aerosol of mass
10–19 kg were placed in air with 90% relative humidity, water vapour would condense onto
the aerosol. The sodium chloride salt will then dissolve and a solution droplet with radius
0.05 mm would appear. However, if the initially dry sodium chloride particle of the same mass
116 u BASICS OF ATMOSPHERIC SCIENCE

(10–19 kg) were to be placed in air, slightly supersaturated (say relative humidity is 100.1%), a
solution droplet with radius of about 0.1 mm would form on the sodium chloride salt. In a
similar manner, as initially dry sodium chloride salt of higher mass (say, 10–18 kg), when
placed in an air of 90% relative humidity, a solution droplet with radius of about 0.1 mm
would form on the sodium chloride salt. All the above-mentioned cases correspond to solution
droplets that form in stable equilibrium with the air. The equilibrium is stable in the sense that
if the droplets grew a little less or grew a little more, their respective vapour pressure would
fall below or rise above that of the ambient air and hence they would either grow back or
evaporate back to their respective equilibrium size. All the droplets, which lie on the left-hand
side of the maxima of their respective curve, would be in a state of stable equilibrium with
the air.

0.5

0.4 1
Relative humidity (%)

supersaturation (%)

0.3
2
0.2
or

0.1 3
100
4
90

80
0.01 0.1 1.0 10.0
Droplet radius (mm)

FIGURE 5.4 The variations of the supersaturation and relative humidity of the air adjacent to
droplets of (1) pure water and of solution droplets containing 10–19 kg of sodium
chloride (NaCl) in (2), containing 10–18 kg of sodium chloride (NaCl) in (3), and
containing 10–17 kg of sodium chloride (NaCl) in (4).

The situation for a solution droplet in a state represented by the peak of its respective
curve is very different to the droplet state lying to the left of the maximum. In the former
case, corresponding to a certain value of saturation, if the droplet evaporated slightly, the
supersaturation of the air in the vicinity of the droplet would fall below that of the ambient
air. This would enable the droplet to grow by condensation back to its original equilibrium
size. However, if the droplet corresponding to the peak of its maximum curve were to grow
slightly, the supersaturation of the air in the vicinity of the droplet would again fall below that
of the ambient air, and the droplet would grow further by condensation. This will cause the
supersaturation of the air in the vicinity of the droplet to drop down even further resulting in
the droplet growing and passing through the states represented by the points to the right of
the peak. A solution droplet which has grown beyond the peak in its respective Kohler curve
is said to be activated and can grow rapidly to form cloud droplets by condensation. Figure 5.4
also reveals that the activated solution droplet asymptotically approaches the Kelvin’s formula
Eq. (5.23), which provides the variation of the relative humidity of air in the vicinity of
CLOUDS AND PRECIPITATION u 117

droplets corresponding to pure water droplets at the same temperature. This is understandable,
considering that the activated solution droplets, which grow rapidly, resemble a pure water
droplet with the decreased effect of the role of the solute as compared to the role of the
curvature effect.

5.2.3 Cloud Condensation Nuclei


Cloud condensation nuclei (CCN) is the name given to all the atmospheric aerosols which act
as nuclei, upon which water vapour condenses in the atmosphere. The nuclei, which can act
as a CCN, can be either soluble or insoluble but wettable aerosol. Wettable but insoluble
aerosols are those aerosols which are readily wetted by water. Wettable aerosols are those in
which water spreads out as a water film over the aerosol surface. It is seen from the earlier
discussion that aerosols, which are larger in size and those with affinity to water (i.e. either
soluble or if insolvable, readily wettable), can serve as CCN at much lower supersaturations
as compared to smaller aerosols and those which do not have any affinity to water.
Furthermore, CCN can also be made up of a mixture of both soluble and insoluble
components. Many of the CCN seen in the atmosphere are made up of sulphates.
It is to be noted that only a small fraction of the atmospheric aerosols act as CCN. This
is because, to act as a CCN, say at a particular supersaturation value of 1%, an insoluble but
wettable aerosol need to have a radius slightly greater than 0.1 mm, while a completely
soluble aerosol requires a size of 0.01 mm or higher. Hence, only 1% of the continental
aerosols and about 10% to 20% of the maritime aerosols can serve as CCN. A greater
percentage of maritime aerosols can serve as CCN since the sea salt particles, which are one
of the most important of the maritime aerosols, are completely soluble in water. Figure 5.5
depicts the averaged CCN spectrum based on measurements of continental and maritime air
samples. The above spectrum provides the averaged CCN concentrations active at
supersaturations, expressed in terms of supersaturations. Earlier, measurement studies have
shown that the CCN concentration
worldwide has no systematic seasonal or
Cloud condensation nuclei (cm– 3 )

latitudinal variations, except for a diurnal


r
maximum which occurs at 6 PM local l ai
10
3
n e nta
time. However, over continental regions, nti
Co
the concentration of CCN shows distinct a ir
ine
variations in the vertical with significantly M
a r
–3
larger values of about 500 cm close to the
surface and a reduction by a factor of five 10
2
–3
(100 cm ) at a height of 5 km. Unlike the
continental behaviour, the CCN over the
maritime region shows a fairly constant 10
0.1 1.0 10
value of 100 cm–3 with the vertical up to a Supersaturation (%)
height of 5 km. The large increase of the
FIGURE 5.5 The averaged CCN spectra based
concentration of CCN near the surface of
on observations of continental and
the continental region, as compared to the maritime air samples as a function
maritime region is due to the presence of of supersaturation.
118 u BASICS OF ATMOSPHERIC SCIENCE

additional sources of CCN over land. The higher concentration of CCN over the land surface
is attributed to forest fires, soil and dust particles, and existence of certain industries such as
paper mills over land. While the additional source of CCN over the land does contribute to
the increase in its concentration, the above can still not account for completely the observed
CCN concentration seen over land. Over the oceans, measurements have shown that the sea
salt particle, although an important nuclei, constitute only a few percent of the CCN. From
the above discussion, it is seen that there exists a widespread and uniform source of CCN
over both the oceans and land. The only candidate (additional source) to explain the observed
concentration of CCN over both the land and the oceans is the gas-to-particle-conversion
mechanism. Since, the above-mentioned mechanism requires the presence of solar radiation,
the rate of production of CCN would peak during the afternoon hours resulting in a local
maximum of the CCN concentration in the early evening hours. The concentration of CCN
active at different supersaturations can be measured using the thermal diffusion chamber.

5.3 DROPLET GROWTH IN WARM CLOUDS


5.3.1 Overview
Warm clouds are those clouds which lie well below the 0°C isotherm and have only the liquid
water droplets as condensation products. The description of the microstructure of the warm
clouds is usually done in terms of quantities such as the liquid water content, droplet
concentration and the droplet spectrum. The liquid water content expressed as gm m–3, is
defined as the amount of liquid water per unit volume which has manifested as condensation
product. The total number of liquid water droplets per unit volume is referred as the droplet
concentration and is expressed in units of m–3. It is to be noted that the CCN has a range of
sizes and the liquid water droplets formed due to condensation onto the CCN also exhibit a
range of sizes. A histogram is then necessary to delineate the number of droplets per unit
volume for the various droplet size intervals and such a representation is referred as the
droplet spectrum.
The direct impaction methods to measure the microstructure of the warm clouds provide
for the collection of all the liquid water droplets in a measured volume of the cloud. The
above collected droplets in the volume are then sized and counted under a microscope. The
direct impaction techniques are prone to error since they are biased against the smaller
droplets, which tend to follow the streamlines around the collector and hence avoid capture.
One way of circumventing the above difficulty is to incorporate corrections to account for the
above bias, based on theoretical calculations. In a real cloud, it is important to measure the
cloud microstructure and its changes in both space and time. This requires that the cloud be
sampled continuously and hence a method to size the droplets without having to collect them
is very much essential. Such automated techniques are presently available. A recent method to
measure the cloud liquid water content from an aircraft is based on the following technique; a
wire heated electrically to a high temperature is exposed to the air stream within the cloud.
When the cloud droplets collide with the wire, they are evaporated and contribute to the
cooling of the wire. Hence, a magnitude of the cooling, which can be measured, is a measure
of the number of liquid water droplets, i.e. to the liquid water content.
CLOUDS AND PRECIPITATION u 119

Previous measurements of the cloud microstructure in warm cumulus clouds over both
continental and maritime regions have revealed the important role of CCN in the detailed
cloud microstructure of warm clouds. While most of the warm marine cumulus clouds have
droplet concentration of less than 100 cm–3, very few of these maritime warm clouds have
droplet concentration larger than 200 cm–3. However, many of the continental cumulus clouds
have droplet concentrations of the order of a few hundreds per cm–3, with some continental
cumulus clouds even having values of droplet concentration of as high as 900 cm–3. The
above values are entirely on expected lines considering the very large number of CCN
(an increase of a factor of five close to the land surface) over the continental regions.
However, detailed measurements of the cloud microstructure in warm cumulus clouds have
revealed that the liquid water content values of maritime clouds do not vary significantly from
those of the continental clouds. With little variation in the liquid water content between the
warm clouds of continental and maritime origin, the relatively smaller droplet concentration
seen in the maritime warm cumulus clouds essentially implies that the average droplet sizes of
the maritime warm cumulus clouds are larger as compared to the droplet size of warm
continental clouds. Furthermore, the droplet size spectrum for the maritime warm cumulus
cloud is much broader as compared to the droplet size spectrum for the continental cumulus
clouds. The above difference in the warm cloud microstructure between the continental and
the maritime clouds do contribute to differences in the development of rainfall in the
continental and the maritime cumulus clouds.

5.3.2 Growth of Cloud Droplets in Warm Clouds by


Condensation
The previous discussion in Section 5.2 considered only the water vapour in the immediate
vicinity of the water droplet. However, for a droplet to grow there must be a continuous
supply of water vapour to its surface. The above transport of water vapour can be realized
through the process of diffusion, provided there exists a vapour density gradient in the region
surrounding the cloud droplet, with the vapour density increasing with distance. In this
subsection, the growth of cloud droplets in warm clouds by condensation is considered.
Consider an isolated cloud droplet with radius r = a, at time t. Let the above droplet
be present in a supersaturated environment where the Ficks law of diffusion holds. The above
diffusion law represents the transfer of water vapour onto the droplet surface and can be
written as
f = –D Ñrv (5.27)
where f is the water vapour flux vector, rv is the water vapour density and D is the diffusion
coefficient which can be assumed constant. The diffusion coefficient D of water vapour in air
can be defined as the rate of mass flow of water vapour across and normal to a unit area in
the presence of a unit gradient of water vapour density. Considering the distribution of the
water vapour density to be spherically symmetric, and assuming the radius of the droplet to be
‘a’ at some instant, the inward flux of mass of water vapour through a sphere Sr of radius r > a is
E SW
 Ô G ¹ OE4 Q S  % (5.28)
4S
ES
120 u BASICS OF ATMOSPHERIC SCIENCE

where n is the outward normal to Sr (refer f


Figure 5.6). However, since the water vapour
n
is lost only by condensation at r = a, the flux
of water vapour for r > a, must be inde- r
pendent of r and equal to the rate of increase a
Sr
of the mass of the droplet. That is, if the
above system can be assumed to be in steady
state equilibrium, the rate of increase of the
mass M of the droplet at time t is equal to
the rate of flow of the water vapour across FIGURE 5.6 Schematic diagram showing
any spherical surface of radius r centred on inward diffusion of water
vapour through a sphere Sr
the droplet. The rate of increase in the mass
onto a droplet of radius a.
of the droplet is then given by
E. E SW
Q S  % (5.29)
EU ES
where rv is the water vapour density at distance r (r > a) from the droplet. Let the water
vapour density in the immediate vicinity of the droplet be rv (a) while the water vapour
E.
density far away from the droplet be rv (¥). Since is independent of r, Eq. (5.29) can be
EU
integrated as given below.

. ‡ ES
E
= Q %
SW ‡

EU ÔB S  Ô E SW (5.30)
SW B

E.
= Q B% < SW ‡
 SW B
> (5.31)
EU

 
Substituting for

.Q S B M
(5.32)

where rl is the density of liquid water into Eq. (5.31), one gets
EB %
< SW ‡
 SW B
> (5.33)
EU BS M

Utilizing the ideal gas equation for water vapour, rv = e/Rv T in Eq. (5.33), one gets
EB %SW ‡
<F ‡
 F B
>
EU BS F ‡
(5.34)
M

where e(¥) is the water vapour pressure in the surrounding air far from the droplet and e(a)
the vapour pressure in the immediate vicinity of the droplet. The above Eq. (5.34) has been
arrived at by assuming (i) the droplet is at rest, (ii) all the molecules, which impinge on, the
droplet remains there, and (iii) the water vapour adjacent to the droplet is at the same
temperature as the ambient air.
CLOUDS AND PRECIPITATION u 121

In reality, the e(a) in Eq. (5.34) has to be replaced by e¢, where e¢ is given by Eq. (5.26).
However, it can be seen from Figure 5.4, that for cloud droplets of radius 1 mm or more, both
the solute effect and the curvature effects are not very important and hence one can
approximate the vapour pressure e(a) by the saturation vapour pressure es over a plane surface
of pure water at the same temperature. Hence, if e(¥) is not too different from es, one can
approximate the following:
F ‡
 F B
F ‡
 FT
 4 (5.35)
F ‡
FT
where S is the supersaturation, expressed as a fraction of the ambient air. Using Eq. (5.35) in
Eq. (5.34), one gets
EB
B = (4 (5.36)
EU M

%SW ‡

where Gl = (5.37)
S
M

For a given environment at a fixed temperature, Gl may be considered a constant.


Assuming Gl to be a constant, it is seen from Eq. (5.36) that for a given supersaturation S,
EB
is inversely proportional to the radius a of the droplet. Hence, the growth of the cloud
EU
droplets in warm clouds by condensation is characterized by a rapid initial increase of radius
and a small rate of growth with further increase in time. Such a growth of cloud droplets by
condensation is schematically depicted by curve a in Figure 5.7. In reality, within a cloud, one
is concerned with the growth of cloud droplets associated with a rising parcel of air. As the
air parcel is undergoing adiabatic ascent, it encounters lower pressure and hence expands. The
work of expansion requires energy, which is taken from the internal energy of the parcel, and
hence the ascending air cools and attains
saturation with respect to liquid water.
Additional uplift will cause the relative
humidity to increase beyond 100% leading
to supersaturation. The above-mentioned
supersaturation initially increases at a rate
proportional to the updraft velocity.
Associated with the increase in the
supersaturation, the CCN present in the air
parcel gets activated, beginning from the
most efficient nuclei corresponding to
larger mass and also those which have a
higher affinity to water. The super-
saturation reaches a maximum value when
the rate at which the moisture is made FIGURE 5.7 Growth of droplet with time by
available by the adiabatic cooling equals to condensation (a) and by collision-
the rate at which the water vapour is coalescence mechanism (b).
122 u BASICS OF ATMOSPHERIC SCIENCE

condensing onto the CCN and the droplets. The concentration of cloud droplets is then
determined at this stage and it equals to the concentration of CCN which got activated up to
this stage. The supersaturation then begins to decrease as the growing droplets use water at a
greater rate than is made available by the adiabatic cooling of the air. The droplets, which
were not activated, (i.e. which lie on the left side of their respective peak in the Kohler curve)
will then evaporate slowly, contributing to the growth by condensation of the activated
droplets. Due to the inverse relationship of the growth rate with the radius of the droplet, the
smaller activated droplets grow faster than the larger activated droplets. This finally results in
a size distribution of the droplets becoming more or less uniform with increase in time.
The comparison between the cloud droplet size distributions as observed above a few
hundred metres above the cloud base of a non-precipitating warm cumulus cloud and the
droplet size distributions calculated by assuming growth by condensation using Eq. (5.36)
show a good agreement if the integration of Eq. (5.36) over time is restricted to a few
minutes, i.e. if the radius of the droplet grows up to a size of 10 mm. This means that the
growth rate by condensation can provide for reasonable verification with observed cloud
droplet distribution only for the initial growth of the cloud droplets. Furthermore, for a typical
cloud droplet of radius 10 mm, to grow into a raindrop having a typical radius of 1 mm,
through condensation, will require according to Eq. (5.36) a time period of several tens of
hours. This time is typically, two orders of magnitude larger than the time required in reality.
In the real atmosphere, a warm cumulus cloud over the tropics can develop and precipitate in
a matter of thirty minutes. Hence, it is observed that the growth of cloud droplets by
condensation alone cannot account for the formation of raindrops and other mechanisms of
growth do exist. These new mechanisms must ensure that the growth from a cloud drop size
to raindrop size takes place in a matter of thirty minutes. Furthermore, in a litre of air, i.e. in
103 cm3 of air, typically, the number of typical cloud droplets is about 106, while the number
of raindrops is about 1. Assuming the droplets to be spherical, it is seen that a cloud droplet
of a typical radius of 10 mm has to grow one million times to grow into a raindrop size of
1000 mm. However, only one of the one million cloud droplets in a litre of air has to grow a
million times to form a raindrop and initiate precipitation. The mechanisms dealing with the
selective growth of cloud droplets in warm clouds will be discussed in the next subsection.

5.3.3 Growth of Cloud Droplets by Collision and Coalescence


The observed rapid growth of the cloud droplets to a size of raindrops in warm clouds can be
explained through the mechanism of the collision and coalescence of droplets. The steady
settling velocity of a cloud droplet, (known as the terminal velocity), as it falls under the
influence of gravity is proportional to the square of the radius of the droplet. The above result
can be arrived at by equating the Stokes drag force given by 6ph av to the force due to
gravity, where v is the terminal velocity, a is the radius of the droplet, and h is the coefficient
of viscosity of air. Assuming the droplet to be spherical, with radius a and equating the Stokes
drag force with the force of gravity, one gets
 
QIBW Q B SH (5.38)

CLOUDS AND PRECIPITATION u 123

Here r is the density of the water droplet. It is assumed that the density of air is much less
than r. Solving Eq. (5.38) for the terminal fall speed v, one gets

B  S H
W (5.39)
I
Hence, it is seen from Eq. (5.39), that everything else being equal, those droplets in the cloud
that have sizes somewhat larger than the average droplet size, will have a higher than average
terminal fall speed. This will result in the larger droplets undergoing collisions with smaller
droplets lying in their path.
It is possible to define a measure of the strength of collisions, in terms of collision
efficiency. The following relation can express the collision efficiency E, of a large drop of
radius r1, called the collector drop overtaking a smaller droplet of radius r2,

Z
& (5.40)
S  S

where y is a parameter which is a measure of the effective cross-section. The parameter y
represents the critical distance between the centreline of the collector drop and the centre of a
small droplet measured at a large distance from the collector drop, such that the droplet just
makes a grazing collision with the collector drop. When the centre of the small droplet of
radius r2 is more distant/closer than y to the centreline of the collector drop, the smaller
droplet of radius r2 will not collide/will collide with the collector drop.
The determination of collision efficiency of a collector drop and a droplet of smaller
radius involve very involved theoretical calculations. Previous theoretical studies have shown
that collision efficiencies for collector drops, which have a radii less than 20 mm, are indeed
quite small. Furthermore, the collision efficiency increases quite significantly with the increase
of the size of the collector. Also, when the collector drop is very much larger than the droplet,
the collision efficiencies are again very small. The reason for the above behaviour is that the
much smaller droplets tend to follow closely the streamlines around the collector drop thus
evading collision. Further, for situations where the collector drop and the droplets have similar
sizes, say 20 mm or more, the collection efficiency becomes larger since both the drop and the
droplet strongly affect each other’s motion.
The important question, however, is the following. Will the collision of a collector drop
of radius r1 with a droplet of smaller radius r2, lead to a capture of the smaller droplet, i.e.
whether the collision would lead to coalescence? The coalescence efficiency E¢ of a droplet
of radius r2 with a collector drop of radius r1 will depend on the fraction of collisions which
have resulted in coalescence. When both collisions and coalescence occur, the droplet gets
captured (collected) by the collector drop and hence collection efficiency Ec can be defined.
Ec can be simply defined as the product of the collision and coalescence efficiencies.
Laboratory experiments have indicated that collision leads to coalescence between two
droplets if the cushion of air trapped between them when they collide can be squeezed out.
This will lead to collection and growth of the collector drop. Furthermore, laboratory
experiments have indicated that the presence of electric fields has a positive effect on
coalescence.
124 u BASICS OF ATMOSPHERIC SCIENCE

5.3.4 The Continuous Collision Model


A theoretical model, called the continuous collision model can provide for the growth of a
collector drop through the collision coalescence process. Let a collector drop of radius r1
having a terminal fall speed v1 fall through still air through a cloud of droplets of uniform
radius r2 and terminal fall speeds v2. The schematic diagram of this model is depicted in
Figure 5.8. In this continuous collision model, it is assumed that the droplets are distributed
uniformly in space. Furthermore, this model also assumes that all collector drops of the same
radii falling through the same distance through the uniformly distributed droplets will collect
the droplets uniformly at the same rate. Assuming the collector drop and the droplet to be
spherical, the rate of increase of mass M of the collector drop due to collision and
coalescence as it falls through the cloud of smaller droplets is
E.
Q S W  W
X & (5.41)
EU
M D

where wl is the liquid water content of the cloud droplets of radius r2 and Ec is the collection
 
efficiency. Since . 
S
Q  S where rl is the density of liquid water, Eq. (5.41) becomes
M

ES W  W
X &D
(5.42)
M

EU S
 M

Assuming v1 >> v2 since r1 >> r2 and also assuming that the coalescence efficiency is unity,
Ec = E and Eq. (5.42) becomes
ES WXM &
(5.43)
EU S M

where E is the collision efficiency. It is seen from


Eq. (5.42) that the above equation describes the
accelerating growth of the collector drop, since v1
increases with increasing r1 and E also increases with r1
increase of r1. Hence the growth rate of the collector
drop through the collision coalescence process can
increase with time resulting in a growth to a raindrop
size in a matter of about half an hour. The nature of
this growth through the collision-coalescence process
is schematically shown in Figure 5.8.
It is possible to obtain the radius of the v1
collector drop at any desired height above the cloud
base or for that matter, the height that would
correspond to a given collector drop radius. To v2
obtain the above information, knowledge of the
steady updraft velocity, and the dependence of the FIGURE 5.8 Schematic diagram illus-
trating the collision-coal-
collision efficiency, and the terminal fall speed of the
escence mechanism.
collector drop as a function of collector drop size are
CLOUDS AND PRECIPITATION u 125

required. Let w be the steady updraft velocity in the cloud and the velocity of the collector
drop and droplet in the presence of updraft is then given by w – v1 and w – v2, respectively.
Since the updraft velocity is assumed to be steady and uniform, the relative velocity between
the collector drop and the droplet in the presence of updraft velocity is still the same as the
case when there is no updraft velocity. Hence the growth rate of the collector drop is still
given by Eq. (5.43). However, the motion of the collector drop is given by
EI
X  W (5.44)
EU
where h is the height above a fixed reference level, say a cloud base at some time t.
Eliminating dt between Eqs. (5.43) and (5.44) and assuming that v1 >> v2 and Ec = E, one gets
ES WXM &
(5.45)
EI  S X  W

Let the radius of the collector drop be r0 at the cloud base and rH at a height H, where H is
the height above the base of the cloud. Assuming the steady upward velocity to be
independent of height in the cloud and integrating Eq. (5.45) from the cloud base to some
height H above the cloud base yields

 SM Ë ) X ES Û
S S)

) ÌÔ ES  Ô  Ü (5.46)
XM Ì S W & )Ü
Í S Ý
Given the functional dependence of E and v1 as a function of r1 and the value of wl, one can
find from Eq. (5.46), H given rH, or find rH, given H. When the collector drop is still quite
small, the updraft velocity is much larger than its terminal velocity and the first integral
dominates as compared to the second integral. Increase of rH causes an increase of H and the
collector drop is carried up by the updraft. During its upward motion, the collector drop
collides, coalesces and collects the smaller droplets leading to its increase in size. Growth of
the collector drop increases its terminal fall speed and at some stage when its terminal fall
speed is greater than the updraft velocity, the collector drop begins to fall. At this stage, the
second integral dominates over the first, since an increase of rH corresponds to a decrease in
H, indicating that the collector drop is falling. Again, the collector drop while falling through
the cloud of droplets of smaller size, collides, coalesces and collects the smaller droplets. The
collector drop ultimately falls through the cloud base and reaches the earth surface as a raindrop.
Theoretical calculations of the continuous collision model indicate that the precipitation
can be initiated through the collision-coalescence process in a reasonable time period of an
hour or less. Typically, a collector drop of radius 10 mm from cloud base in a cloud having
1 g m–3 liquid water content and a steady updraft velocity of 1 m s–1, grows to a size of 0.15
mm in forty-five minutes as the drop is carried up to a height of 2.2 km. The drop then falls
back to the cloud base in another fifteen minutes by which its radius increases to 0.75 mm.
With the same updraft velocity and lower liquid water content, say half of the earlier value,
the collector drop needs to be carried up to a height of 3.2 km and will have a size of 0.65
mm as it falls back to cloud base. With the same liquid water content (i.e. 1 g m–3) and a
much lower updraft velocity (10% of the earlier value, i.e. 0.1 m s–1), the collector drop will be
126 u BASICS OF ATMOSPHERIC SCIENCE

taken up only up to a height of 0.5 km and can only attain a size of 0.1 mm at the cloud base.
Furthermore, smaller the updraft velocity, longer will it take for the drop to reach back to the
cloud base. Hence, it is seen that strong updraft velocities in warm clouds can initiate
precipitation in a much shorter time than in clouds with low updraft velocities. Furthermore,
as mentioned above, warm clouds with strong updrafts must be quite deep to attain the
raindrop size. Also, warm deeper clouds with strong updraft velocities will provide larger size
raindrops than produced by shallower warm clouds with weaker updraft velocities. The above
discussion clearly reveals that all the collector drops of the same size will grow at the same
rate as they rise and fall through the cloud of droplets. This is because it is assumed that the
smaller droplets are distributed uniformly in size and that the collector drop collides coalesces
and collects smaller droplets in a continuous and uniform manner.

5.3.5 The Stochastic Collision Model


An improvement over the continuous collision model known as the stochastic collision model,
results in a broad droplet size spectra. In this stochastic collision model, it is assumed that all
collisions are individual events distributed statistically in space and time. Figure 5.9 depicts a
schematic diagram, which illustrates the broadening of the droplet sizes due to statistical
collision model.

Line 1
100

Line 2

80 20

16 16
Line 3

64
32
4

FIGURE 5.9 Schematic diagram indicating the broadening of the droplet spectrum using the
stochastic collision model.

Consider for example, 100 small cloud droplets initially of the same size as shown in the first
top line of Figure 5.9. After a certain time interval, let us assume that 20% of the total
droplets would have collided with other smaller droplets and would have grown as depicted
second line of Figure 5.9. Due to their larger sizes, these twenty large droplets are more likely
to undergo further collision. Assuming that the second set of collisions are again statistically
distributed, these result in three different size categories of droplets. In the second set of
collisions, it is assumed that sixteen of the smaller drops and four of the larger drops undergo
collision. Unlike, the continuous collision model, the stochastic collision model leads to not
CLOUDS AND PRECIPITATION u 127

only a broad droplet size distribution, but also provides a mechanism for a small fraction of
the cloud droplets to grow very much faster as compared to the average growth of the droplets.

5.4 FORMATION AND GROWTH OF ICE CRYSTALS IN


COLD CLOUDS
Even over the tropical regions, it is important to realize that not all the precipitation is
obtained from warm clouds only. In fact, a large amount of non-convective precipitation is
obtained from middle level and upper level clouds, which clearly lie above the 0°C isotherm.
Furthermore, the most important of the clouds, the cumulonimbus cloud or the thunderstorm
cloud is a strong convective cloud with a large vertical extent reaching right up to the
tropopause. While the base of the above cloud is below the 0°C isotherm, its cloud top is
much above the 0°C isotherm. Due to the above reason, the hydrometeors or precipitation
products in this thunderstorm cloud are made up of rain, ice crystals and hail. Hence, it is
important to study in detail the formation and growth of the other hydrometeors such as ice
crystals and hail.

5.4.1 Homogeneous Nucleation of Ice Particles


Cold clouds are those which extent vertically above the 0°C isotherm. It is to be noted that
water droplets can exist in the liquid phase in clouds even at temperatures below 0oC and the
water droplets in such cases are referred as supercooled droplets. Typically, a cold cloud will
contain both the ice particles as well as supercooled droplets. However, water droplets cannot
exist in the liquid phase at temperatures below – 40°C. Hence, clouds having temperatures at
and below – 40°C will consist entirely of ice particles and such clouds are said to be glaciated.
If saturation occurs at temperatures between 0°C and –4°C, the excess water vapour
undergoes a change of phase to form supercooled water. Ice particles, however, do not form
within the above range of temperature. However, as the temperature decreases, the possibility
of formation of ice particles increases and at temperatures between – 10°C and –40°C,
saturation can lead to the formation of ice particles, supercooled droplets or both. Lower the
temperature in the above range more will be the proportion of ice particles. At temperatures
below –40°C isotherm, saturation leads to the formation of ice particles only.
In the absence of foreign particles in a water droplet, an ice embryo of critical size can
form by the chance collections of water molecules coming together within the water droplet.
Such a process is known as the homogeneous nucleation of ice particles through freezing. The
above process is entirely analogous to the formation of a water droplet from the vapour phase
as discussed in Section 5.2. Similar to the previous discussion involving the homogeneous
nucleation of condensation, if the ice embryo exceeds a certain critical size, its further growth
will contribute to a decrease in the total energy of the system. With decreasing temperature,
the number and size of the ice embryos that form by chance collection within the droplet
increase. Hence, unlike the homogeneous nucleation of condensation, at low temperatures, the
homogeneous nucleation of ice particles through freezing is highly likely. For droplets of
radius between 20 mm and 60 mm, homogeneous nucleation occurs at –36°C, while for the
droplets of a few mm, the homogeneous nucleation occurs at – 39°C. Hence, the formation of
homogeneous nucleation of ice particles through freezing is possible only for high clouds.
128 u BASICS OF ATMOSPHERIC SCIENCE

5.4.2 Ice Nuclei


Just as the heterogeneous nucleation of liquid droplets was possible at relative humidity near
100%, using the CCN, the heterogeneous nucleation of ice particles through freezing at
temperatures near 0°C is aided by the presence of freezing nuclei. Freezing nuclei, can both
be contained within the water droplet or be outside the droplet. In the latter case, the cloud
droplets may be frozen if freezing nuclei in the air come into contact with the droplet. Such a
freezing of the cloud droplet due to contact is known as contact nucleation. The freezing
nuclei responsible for the contact nucleation are then referred as contact nuclei. Laboratory
experiments have revealed that a foreign particle can cause a drop to freeze by contact
nucleation at a higher temperature than it would if it were contained within the drop.
Deposition nuclei are particles in air, which aid in the formation of ice crystals directly from
the vapour phase. Freezing nuclei, contact nuclei and deposition nuclei are together commonly
called as ice nuclei. However, unlike the CCN, which are abundant, the ice nuclei are
somewhat rare in the atmosphere. This is because of the requirement that the ice nuclei need
to have a molecular spacing and crystallographic arrangements, which are similar to the
hexagonal structure of ice. Furthermore, most ice nuclei are virtually insoluble in water. Many
organic materials as well as inorganic soil substances, mainly clay, can serve as ice nuclei at
temperatures above –15°C. Decayed plant leaves contain enough ice nuclei and these can be
active even at temperatures as high as – 4°C. Ice nuclei active at – 4°C are also found in sea
water rich in plankton. Since the ice embryo also has the same dimensions as the freezing
nuclei, to start with, and the formation of ice particles is aided by the freezing nuclei, the
heterogeneous nucleation of ice particles can occur at much higher temperatures than the
corresponding homogeneous nucleation of ice particles.
A useful method to measure the concentration of particles in air, which can serve as ice
nuclei at a given temperature, is to draw a known volume of air into a container. Cooling of
air in the container will lead to a formation of a cloud from which the number of ice crystals,
which have formed in the cloud at a particular temperature, can be measured. Cooling may be
produced either by compressing the air and suddenly expanding the air as in expansion
chambers or by refrigeration as in mixing chambers. Illuminating a certain volume of the
chamber and determining visually the number of ice crystals in the light beam may provide
the number of ice crystals, in the chamber. An alternate method is to allow the ice crystals to
fall onto a dish of supercooled soap or sugar solution, where they can be counted easily.
Measurements have indicated that the number of ice nuclei is more in the northern as
compared to the southern hemisphere. The average concentration of ice nuclei are about 103
in m–3, i.e. one nuclei per litre at –20°C. Furthermore, the concentration of ice nuclei
increases by a factor of ten for every four degree fall in temperature. Considering that the
total concentration of aerosol is about 108 litre–1, it is observed that only one particle in 108
acts as an ice nuclei at –20°C.

5.4.3 Bergeron Process


Typically, in a well-developed cumulus cloud, the lower and the upper portions of the cloud
will consist entirely of water droplets and ice crystals, respectively, while the middle portion
will consist of a combination of supercooled water droplets and ice crystals. The co-existence
CLOUDS AND PRECIPITATION u 129

of supercooled water droplets and ice particles are essential to the formation of precipitation
in cold clouds and this process is called the Bergeron process.
In a cold cloud containing the supercooled droplets and ice particles, the air that is
saturated with respect to the liquid water is, however, supersaturated with respect to ice. That
is, the saturation vapour pressure over ice, which is the amount of water vapour required to
keep the ice in equilibrium, is less than the saturation vapour pressure over the supercooled
water at the same temperature. This is due to the fact that the water molecules in an ice
crystal are bound to each other more tightly than the molecules of liquid water. For example,
air saturated with respect to liquid water at –10°C, is actually supersaturated with respect to
ice by 10%, while for air saturated with respect to liquid water at –20°C, it is actually
supersaturated with respect to ice by 21%. The above causes some of the water vapour in the
air to be deposited directly onto the ice particles. This leads to a drop in the vapour content of
the air which causes the supercooled liquid droplets to evaporate. The above evaporation
ensures that the vapour pressure of the air with respect to the liquid water remains saturated.
Again, air saturated with respect to the liquid water is supersaturated with respect to ice and
this causes the excess water vapour to be directly deposited onto the ice particles. When the
above process is repeated, there results a continuous transfer of vapour through which the ice
crystals grow at the expense of supercooled droplets. As the ice crystals grow, they attain a
size, which is large enough to be supported by air currents, and hence they fall. As they fall
through the cloud, they collide with droplets and other ice crystals.
Since ice itself can act as effective ice nuclei, ice crystals, which fall through a cloud,
are likely to collide with supercooled liquid droplets. The above collision results in the liquid
water freezing onto the ice crystals. The above process, called riming or accretion, leads to a
rapid growth of the ice crystals, which further increases their terminal fall velocities and
promotes further accretion or riming. Another process in the development of precipitation in
cold clouds is called as aggregation, in which two ice crystals join to form a single larger
crystal. Aggregation is highly likely when the ice crystals have a thin coating of liquid water.
The above thin coating ensures the liquid water to be more adhesive. Since water is likely to
be present in cold clouds having temperatures below 0°C, adhesion is highly in the lower
portion of the well-developed cumulus clouds. Both the above process of riming and
aggregation ensure that the ice crystals grow much faster than by the deposition of water
vapour to ice particles. It turns out that the growth rate from the combination of all the three
processes ensures that the formation of precipitation sized ice crystals can happen within
about half an hour from the initial formation of ice crystals.

5.4.4 Growth Rate of Ice Crystals by Deposition


The growth rate of an ice crystal by deposition from the vapour phase is quite similar to the
process that controls the droplet growth by condensation except for the fact that the ice
crystals are not spherical. However, if one considers the special case of a spherical ice particle
of radius r, one can write the mass growth of the ice crystal as
E.
Q S%< SW ‡
 SWD > (5.47)
EU
130 u BASICS OF ATMOSPHERIC SCIENCE

where rvc is the density of the vapour in the immediate neighbourhood to the surface of the
ice crystal. For ice crystals of arbitrary shapes, one can utilize the analogy between the vapour
field around a crystal and the electrostatic potential field around a charged conductor having
the same shape and size of the crystal. For a sphere, the leakage of charge from the conductor
is proportional to the electrostatic capacity C of the conductor and is given by
$ Q S (5.48)
F
where e0 is the permittivity of free space given by 8.85 ´ 10–12 C2 N–1 m–2. Using Eq. (5.48)
in Eq. (5.47), one gets
E. %$
EU F  <S ‡
 S
W WD > (5.49)

Assuming that the vapour pressure corresponding to rv(¥) is not very much greater than the
saturation vapour pressure esi over a plane surface of ice and assuming that the ice crystal is
not very small, Eq. (5.49) becomes
E . $
EU F  ( 4 J J (5.50)

where Si is the supersaturation with respect to ice and is given by


F ‡
 FTJ
4 (5.51)
F
J
TJ

while Gi = Drv(¥) (5.52)


The variation of the product of Gi Si with respect to temperature for the case of an ice crystal
growing in air saturated with respect to liquid water shows that the product attains a
maximum value at – 14°C. Hence from Eq. (5.50), it is seen that the ice crystals grow most
rapidly by vapour deposition in clouds having both supercooled droplets and ice crystals at
temperatures around – 14°C.

5.4.5 Hail Formation


Hail or hailstone is the chief form of frozen precipitation observed during the prime warm
season, and always accompanies a severe thunderstorm. Hail can be found in the middle and
upper portions of almost all thunderstorms. However, most hail either melts before hitting the
ground, or being very soft, disintegrates in the violent interior of the thunderstorm itself. Hail,
generally, begins forming on seeds of small frozen raindrops or soft ice particles known as
graupel, which are hardened conglomerates of snowflakes. It is not as if graupel or frozen
droplets are the only embryos for hail formation. Hail sometimes does contain foreign matter
such as pebbles, leaves, and twigs that have been lofted into the thunderstorm cloud by strong
updraft winds. When a hail is sliced through its centre, its cross-section reveals an onion-like
layering which is more evident in the larger hail. These distinctively different onion-like
layers indicate the type of ice formed as the hail grew in size. The above-mentioned layers
CLOUDS AND PRECIPITATION u 131

usually alternate between opaque ice and clear ice. The opaque ice layer forms when the hail
collects small, supercooled liquid water drops that freeze rapidly on impact, thereby trapping
air bubbles within the ice and giving it a “milky” texture. However, when larger supercooled
water drops impact on a hail, the freezing is slower, allowing the air bubbles to escape, and
thus forming clear ice. The size of hail is critically dependent on the intensity of the
thunderstorm cell, within which it forms. To form hailstones of the size of golf ball requires
over ten billion (1010) supercooled droplets to be collected. For this, they must remain in the
thunderstorm cloud for at least 5 to 10 minutes. It is worthwhile to compare the above figure
of ten billion droplets for the formation of large hail to the one million droplets needed to
form the typical raindrop. Hence, from the above discussion, it is observed that large
hailstones of diameter greater than 5 cm, form mostly in very intense supercell thunderstorms,
which have strong updraft winds. The largest hailstone to have ever fallen was observed in
Coffeyville, Kansas on September 3, 1970, weighing 0.75 kg, and having a diameter of
14.4 cm. The magnitude of the updraft velocity is important in determining the size of hail.
While an updraft of around 36 to 54 km hour–1 may be necessary for a small hail to be
formed, larger hailstones having diameter of about 5 cm require very strong updraft of the
order of 88 km hour–1. Well, within the environment of the thunderstorm interior, hailstones
of various sizes may collide. Due to the strength of the force of collision, the hailstone may
break into smaller sizes. Also, at times the collision results in joining of the individual
hailstones. It is these violent collisions within a thunderstorm cloud, which are responsible for
the formation of the irregular, large hailstones often observed. The hailstone may fall as a
result of an increase to a size, which is so large that it cannot be supported by the
thunderstorm cell’s updrafts. Also, the hailstone may get caught in a downdraft and be hurled
downward towards the earth. Either way, large hailstones fall at great speeds, faster than
160 km hour–1.

5.4.6 Radiative Effects of Clouds


Clouds do alter the horizontal and vertical distributions of short-wave heating due to solar
radiation together with the distribution of latent heating as well as the distribution of
terrestrial long-wave cooling. Clouds are responsible for reducing the net absorption of the
short-wave solar radiation by increasing the earth’s albedo. Furthermore, clouds also cause a
reduced loss of long-wave radiation by decreasing the effective radiation temperature of the
earth. At the top of the atmosphere both the above effects due to solar radiation and terrestrial
radiation act in an opposite manner causing the net effect to be small. Clouds also modify
transports of moisture through the precipitation process. Most of the earlier studies have
investigated the relationship between the radiation budget at the earth surface and the total
cloud cover. While the total cloud cover is an important measure of the effect of the clouds
on the earth radiation budget, there are other important characteristics such as different types
of clouds and their variations in terms of water content and cloud height which have an
important bearing on the earth radiation budget. Clouds play an important and at the same
time a complex role in the earth’s radiation budget. It is known that low clouds reflect much
of the sunlight that falls on them, while high clouds reflect much less. Furthermore, low
clouds are known to have little effect on the emitted energy, while high clouds trap most of
132 u BASICS OF ATMOSPHERIC SCIENCE

the energy emitted by the surface. Since the low clouds reflect much of the sunlight, and have
little effect on the emitted energy, it is suggested that low clouds act to cool the current
climate. Not only the type of cloud, but also its amount and the cloud microphysical
properties have a bearing on the changes in the radiative forcings and hence such cloud
radiative-forcings have important implications for the climate of earth.

5.5 MECHANISMS OF CLOUD FORMATION AND CLOUD


SEEDING
The following subsections briefly present the various mechanisms of cloud formation ranging
from the ascent of moist warm air due to heating by the underlying surface to the forced
ascent of moist air due to orography. Clouds can also form due to the forced ascent of stable
air as well the dynamical uplift of warm moist air due to convergence at lower levels. For
precipitation to occur, the particles in the cloud have to grow to the size of a raindrop, the
latter being no longer maintained by the ascending updrafts. Section 5.5.4 discusses cloud
seeding procedures; the various ways to hasten the initiation of rain by seeding selectively
the clouds.

5.5.1 Mechanisms of Cloud Formation


From the earlier discussions in this chapter, it is seen that clouds form when the air has
become supersaturated with respect to liquid water and or/ice. Furthermore, whenever moist
air ascends upwards, it experiences adiabatic expansion and the associated cooling leads to
supersaturation of air and hence to cloud formation. The following mechanisms provide for
the different ways by which moist air can be lifted to form clouds:
(i) Local ascent of warm moist air in a conditionally unstable environment can produce
clouds, which are convective in nature, and these clouds are known as cumulus
clouds. Such convective lifting occurs when moist air heated at the earth’s surface
rises in the form of thermal currents or bubbles. These convective clouds have
horizontal dimensions ranging from 0.1 km to 10 km and have vertical velocities of
the order of a few meters per second. The vertical extent of a well-developed
convective cloud can extend to heights of the order of 16 to 20 km. Typically, these
convective clouds have liquid/ice water content of the order of one gram per cubic
metre of air and these clouds have a typical lifetime lasting from a few minutes
to an hour.
(ii) Forced lifting of stable air produces clouds, which have a typical “layered” structure
and these clouds are known as stratus clouds. These layered clouds can occur at
heights ranging from the ground level up to heights of the tropopause and can extend
over areas ranging up to hundreds of thousands of square kilometres. Unlike the
convective clouds, the stratus clouds do not extend vertically and have smaller
vertical velocities of the order of a few centimetres per second. Typically, these
stratus clouds have liquid water content of the order of a few tenths of a gram per
cubic metre of air and these clouds have a typical lifetime lasting over periods of
tens of hours.
CLOUDS AND PRECIPITATION u 133

(iii) Orographic or forced lifting of moist air occurs when the moist air is forced upward
by a barrier of mountains or hills, resulting in the so-called orographic clouds. The
heights to which these orographic clouds can rise is not limited by the mountain
height; and can at times even extend into the lower stratosphere. Downwind of the
mountain barrier, on the lee side of the hill, one encounters a rain shadow region;
regions of lower precipitation caused by the descent of the air down the slope and
the associated adiabatic warming by compression. The vertical updraft velocities of
the orographic clouds depend on the height of the mountain/hills as well as on the
speed and direction of the wind and can have values of the order of several metres
per second. Typically, the orographic clouds have water content of the order of a
few tenths of a gram per unit cubic metre of air. The lifetime of a orographic cloud
depends primarily on the nature and persistence of winds and can last for long
periods of time provided steady winds exist.
(iv) Frontal lifting or widespread lifting results from the interaction of air masses along
the frontal boundaries. Such frontal lifting results in cloud formation and these are
associated with both cold fronts as well as warm fronts. When the cold air advances
towards the warm air in the cold front, the denser cold air displaces and forces the
warm air to rise ahead of it. A situation similar to the orographic lift occurs in a
warm front when the warm air flows towards the wedge of the cold air, resulting
in the forced uplift of the warm air.
(v) Dynamical upward lifting of moist air occurs at lower levels due to the horizontal
convergence of moist air into an area. Such dynamical upward lifting of air is seen
in the neighbourhood of a low-pressure area near the surface. Since winds in the
lower troposphere will tend to converge onto the centre of the low-pressure area,
these will result in the accumulation of the mass of air causing horizontal
convergence of air at low-levels. Such horizontal convergence of air at lower
troposphere lead to upward vertical motions which can cause adiabatic expansion
and cooling leading to cloud formation.
(vi) Fog forms when surface air cools to its dew point as it comes into contact with a
cold surface. While the radiation fog forms during clear windless nights when the
ground is cooled by loss of terrestrial radiation, advection fog forms when warm
air moves over a relatively colder surface. With the advent of insolation, and the
resultant mixing, the fog may lift upwards to form a stratus cloud.
(vii) Adiabatic expansion and cooling due to a rapid local reduction in pressure can result
in the localized formation of funnel clouds which are associated with tornadoes and
water spouts.

5.5.2 Types of Clouds


There are primarily ten main cloud types, which are classified according to their height,
shape, colour and associated weather. From considerations of height, clouds are classified as
low clouds (those which form from the earth’s surface to a height of 2.5 km), middle clouds
(those which form from 2.5 km to 6 km), or high clouds (those which form above 6 km).
Furthermore, clouds are given Latin names, which describe their characteristics, e.g. cirrus
134 u BASICS OF ATMOSPHERIC SCIENCE

(a hair), cumulus (a heap), stratus (a layer) and nimbus (rain-bearing). It is to be noted that all
clouds are white, however, when viewed from the ground some clouds do appear gray or dark
gray depending on their depth and shading from higher cloud. The ten main cloud types are
the following:
(i) Stratus cloud is a low-level layered cloud. The stratus cloud has a gray uniform
base and is referred as fractostratus if it has a ragged structure. Drizzle is the
common precipitation associated with the stratus cloud.
(ii) Cumulus cloud is a low-level convective cloud consisting of individual cells,
vertical rolls or towers. The chief characteristic of the cumulus cloud is a flat base
and the typical heaped appearance. Rain or snow is the common precipitation
associated with the cumulus cloud.
(iii) Stratocumulus cloud is a low-level layered cloud consisting of a series of rounded
rolls. Stratocumulus clouds appear generally white in colour. Drizzle is the common
precipitation associated with the stratocumulus cloud.
(iv) Nimbostratus clouds appear typically as thicker and darker clouds having a sheet
like appearance in their lower base. As the name “nimbus” indicates, “rain-
bearing”, these nimbostratus clouds are associated with heavy intensity precipitation
in the form of rain or snow.
(v) Altostratus clouds are typically middle-level clouds which form at heights between
2.5 km to 6 km. These altostratus clouds appear as a thin gray sheet. Due to the
thin layer the altostratus clouds allow the sun to appear as through ground glass.
Rain or snow is the common precipitation associated with the altostratus cloud.
(vi) Altocumulus clouds are typically middle-level clouds which form at heights
between 2.5 km to 6 km. Altocumulus clouds have a layered structure, and are
made up of rippled elements, which are generally white in colour with some
shading. Altocumulus clouds may produce light showers.
(vii) Cirrus clouds are typically high-level clouds which form at heights above 6 km. As
the name indicates these cirrus clouds resemble white tufts or filaments resembling
hair. Since they are high-level clouds these clouds do contain ice crystals. Usually,
these cirrus clouds do not provide any precipitation.
(viii) Cirrocumulus clouds are also high-level clouds which form at heights above 6 km.
These cirrocumulus clouds resemble small rippled elements and due to the
elevations at which they form, these clouds do contain ice crystals. Again, these
cirrocumulus clouds do not provide any precipitation.
(ix) Cirrostratus clouds are again high-level clouds which form at heights above 6 km.
These cirrostratus clouds appear as transparent sheet or veil and are associated with
the halo phenomena. Furthermore, due to the elevations at which they form, the
cirrostratus clouds do contain ice crystals. Like, the other high-level clouds such as
cirrus and cirrocumulus, the cirrostratus clouds also do not provide any
precipitation.
(x) Cumulonimbus clouds typically have their bases at low levels, while their cloud
tops can extend right up to the tropopause or even to the lower statrosphere. Hence
these clouds have the largest vertical development among all clouds. These clouds
have very large cauliflower-shaped towers extending to great heights and are often
CLOUDS AND PRECIPITATION u 135

accompanied by an anvil structure at their top. These cumulonimbus clouds are


associated with lightning, thunderstorms, and squalls. The nature of the
precipitation associated with these clouds corresponds to intense showers of rain
or snow.

5.5.3 Convective Clouds


Convective clouds are those clouds that form directly due to convection. Such convective
clouds form due to warm rising thermals which are heated by the underlying surface. These
convective clouds can develop vertically to great heights extending right up to the tropopause.
Such vertically developed clouds which arise due to deep convection over tropics are one of
the primary mechanisms by which the solar heating of the earth surface is transported to
upper troposphere from where they can be sent to higher latitudes. Furthermore, deep
convective clouds over tropics act as the source of upper tropospheric moisture.

Lifting Condensation Level


The base of a cloud layer is usually indicated by a decrease in dew-point depression, while
the top of a cloud layer is usually indicated by an increase in the dew-point depression. The
Lifting Condensation Level (LCL) is defined as the level at which an air parcel at the surface,
if lifted dry adiabatically, would reach saturation. The LCL, at a given place can be
conveniently found from the Tephigram if the observed sounding is available. Using the
Tephigram, start from the surface air temperature, and follow the dry adiabat through the
surface temperature upward to where it would cross the constant mixing ratio line that runs
through the surface dew-point temperature.  The pressure at which the above-mentioned lines
cross one another is the LCL. The LCL is approximately the height (or rather the pressure
level) at which one would expect cloud bases to form provided the mechanism for upward
motion was due to mechanical lifting (such as through orographic lifting, frontal wedging,
or convergence).

Convective Condensation Level


The Convective Condensation Level (CCL) is the level at which one would expect to find the
bases of the convective clouds. Like the LCL, the CCL at a given place can be conveniently
found from the Tephigram if the observed sounding is available. Usually, two different
methods exist to determine CCL from the Tephigram. The first method known as parcel
method is primarily utilized for finding the bases of shallow convective clouds. In the parcel
method, one uses the Tephigram to follow the mixing ratio line that passes through the
surface dew-point temperature up to where it intersects the observed temperature sounding.
The pressure at which the mixing ratio line from the surface crosses the temperature sounding
is the CCL. The above level as determined by the parcel method is also sometimes known as
the Mixing Condensation Level (MCL). The second method called the mixing method is
mainly used for finding the bases of deep convective clouds, such as thunderstorms. In the
mixing method, one uses the Tephigram to determine the average mixing ratio in the lowest
50 hPa (by inspecting the dew-point line). Once the average mixing ratio in the lowest 50 hPa
is found, one follow this mixing ratio line up to where it intersects the observed temperature
136 u BASICS OF ATMOSPHERIC SCIENCE

sounding. Both the above-mentioned methods provide more or less the same CCL although at
times the CCL obtained from both these methods may differ. The convective temperature is
determined by following the dry adiabat from the CCL to the surface.  The convective
temperature is the temperature, which the surface would have to reach in order for convective
clouds to form. From the knowledge of the convective temperature from an early morning
sounding, it is possible to forecast the likelihood of a thunderstorm during the afternoon/
evening hours at a given place.

Level of Free Convection


The Level of Free Convection (LFC) is defined as that pressure level in the atmosphere where
the observed air temperature of the environment decreases faster than the moist adiabatic
lapse rate of a saturated air parcel at the same level. The LFC can be easily determined from
the Tephigram. Starting from the LCL, using the Tephigram follow the moist adiabatic lapse
rate until the temperature of the parcel reaches the observed air temperature. The pressure
level at which the moist adiabatic lapse rate from the LCL cuts the environmental air
temperature is the LFC. One can indeed confirm that the LFC has been reached by
ascertaining that the temperature of the parcel along the moist adiabat is warmer than the
environment on further lift beyond LFC. Further, lifting along the moist adiabatic lapse rate
beyond the LFC results in a warmer air parcel as compared to the environment. At some
higher level, called the Equilibrium Level (EL), the temperature of the parcel along the moist
adiabat again reaches the observed environmental air temperature.

Convective Available Potential Energy


The Convective Available Potential Energy (CAPE) can be thought of as the maximum
available energy available to an air parcel obtainable during its vertical ascent. CAPE is found
within the conditionally unstable layer of the troposphere, and manifests whenever an
ascending air parcel is warmer than the ambient environmental air. In simple terms, the
greater the value of CAPE, more likely the atmosphere will produce convection and
potentially thunderstorms. CAPE is measured in units of J kg–1 of air. Typically, the value of
CAPE ranging from 1500 J kg–1 to 2500 J kg–1 is considered large; while a value above
2500 J kg–1 is considered an extremely large value. This is not to mean that a CAPE value
below 1500 J kg–1 cannot produce severe weather, in fact, any positive value of CAPE
indicates instability and the possibility of thunderstorms. CAPE is usually calculated by
integrating vertically the local buoyancy of a parcel from the LFC to the equilibrium level
(EL) and the expression for CAPE is given by

Ë 5WQBSDFM  5WFOW Û
[FM

$"1& Ô HÌ Ü E[
[GD Í 5WFOW Ý
where zfc and zel refer to the height of the LFC and EL, respectively, Tvparcel and Tvenv refer to
the virtual temperature of the specific parcel and the virtual temperature of the environment,
respectively and g refers to the acceleration due to gravity. The CAPE value for a given
region is most conveniently found from the Tephigram by calculating the positive area above
the LFC, the area between the air parcel’s virtual temperature line along a moist adiabat and
CLOUDS AND PRECIPITATION u 137

the environmental virtual temperature line where the ascending parcel is warmer than the
environment. CAPE is sometimes referred to as positive buoyant energy. The negative area
closest to the ground refers to the area between the environmental temperature sounding and
the air parcel’s line consisting of a dry adiabat till the LCL and along the parcel’s virtual
temperature line along a moist adiabat, beyond the LCL, where the ascending parcel is colder
than the environment. One can have additional negative areas at heights above the equilibrium
level. The negative area is referred as the Convective Inhibition (CIN). Let CAPE exist along
with a layer of CIN. In such a situation, the positive buoyant energy is unavailable to deep,
moist convection until CIN is overcome. When there is mechanical forced lift of moist air,
cloud base forms at the LCL. In the absence of forced lifting, the cloud base forms at the
CCL, where heating from below causes spontaneous buoyant lifting. When CIN is absent or
is overcome, saturated air parcels at LCL or CCL, will continue to rise to the LFC. The
saturated air parcels having positive buoyancy at LFC will rise spontaneously up to the stable
layer of the equilibrium level. The end result is the manifestation of a deep, moist convection,
or a thunderstorm cell.

5.5.4 Cloud Seeding


It is seen from the discussions of the earlier sections that the growth of a cloud drop to a rain
drop or for that matter the growth of ice crystals from supercooled droplets is indeed a very
complex process which occurs in nature. In the previous twentieth century, since 1940s, many
researchers have tried to induce precipitation from clouds, mostly to overcome perennial
drought type situations. The process, by which precipitation can be induced from clouds, is
known as cloud seeding. Cloud seeding, which can be thought of as a form of weather
modification is the attempt to change the amount or type of rainfall that falls from clouds, by
dispersing (introducing or seeding) substances into the air that serve either as cloud
condensation nuclei or ice nuclei. The material used to seed the cold clouds to initiate
precipitation uses either dry ice or silver iodide. Dry ice which is nothing but frozen carbon
dioxide is used to seed clouds as it promotes freezing. At –78°C, carbon dioxide changes
directly from a solid phase to a gaseous phase through sublimation, or from gaseous to solid
through deposition. The sublimation of dry ice cools the air to such an extent that ice crystals
can nucleate spontaneously from the vapour phase. That is, when a certain amount of dry ice
is introduced into a cloud, the dry ice lowers the temperature of the droplets so that freezing
can occur by homogeneous nucleation. The above-mentioned spontaneous nucleation does not
require any existing droplets or particles because it produces extremely high vapour
supersaturations near the seeding substance. However, the existing droplets are needed for the
ice crystals to grow into large enough particles to precipitate out.
The other agent for cloud seeding in cold clouds, which is commonly used, is silver
iodide. Silver iodide acts as ice nuclei at temperatures as high as –5°C and induces freezing
through heterogeneous nucleation. Silver iodide can serve as effective ice nuclei since it has a
six-sided crystalline structure similar to that of ice. By acting as ice nuclei in cold clouds
having supercooled droplets, silver iodide can help initiate the Bergeron process. This will
lead to the growth of ice crystals at the expense of supercooled droplets. When the ice
crystals grow into large enough size, they will precipitate out.
138 u BASICS OF ATMOSPHERIC SCIENCE

Seeding of warm clouds seeks to exploit the latent heat released by freezing. This
strategy of dynamic seeding assumes that the additional latent heat adds buoyancy, strengthens
updrafts, ensures more low-level convergence, and ultimately causes rapid growth of properly
selected clouds. Cloud seeding agents such as silver iodide may be dispersed by aircraft or by
dispersion devices located on the ground (generators). For release by aircraft, silver iodide
flares are ignited and dispersed as an aircraft flies through a cloud. However, aircraft usually
disperses dry ice into a cloud for seeding purposes.
Seeding is also utilized to clear fogs along aircraft runways. When fog temperatures are
below 0°C, then dry ice can serve as a useful fog-seeding agent, by initiating the Bergeron
process. Here, some of the water droplets freeze into ice particles which can then grow at the
expense of supercooled droplets. For dispersing warm fogs, salt particles are introduced into
the fog using an aircraft. The objective is to make some of the droplets to have a larger size;
thereby allowing for the growth by collision-coalescence process. Seeding has also been
attempted to suppress and reduce the intensity of hail. The idea to seed a hail producing cloud
is to increase the number of growing ice pellets. Since the thundercloud has a limited amount
of water that can freeze onto the growing hailstones, increasing the number of ice pellets and
hence the number of hailstones would automatically bring down the average size of the
hailstone and thereby reducing drastically the damage from large hailstones.

5.6 ROLE OF CLOUDS AND PRECIPITATION PRODUCTS IN


CHARGE SEPARATION
It is a fact of common everyday experience that lightning flashes and thunder manifest during
the development of thunderstorms. In this section, the microphysical mechanisms responsible
for the electrification of thunderstorm are briefly outlined. About 80% of all the lightning
results from discharge of electricity within the clouds, while the remaining 20% is accounted
by the lightning discharge from cloud to ground. The cloud-to-cloud lightning discharge
occurs when the voltage gradient within a thundercloud or between clouds overcome the
electrical resistance of air. A large and powerful spark that partially equalizes the charge
separation characterizes the lightning discharge. The cloud-to-ground lightning discharge
occurs when the negative charges accumulate in the lower regions of the cloud. Positive
charges are then attracted to a relatively small area in the ground directly beneath the
thundercloud. The above results, in the establishment of a large potential gradient, between
the cloud base and the ground. All lightning discharges require the initial separation of
positive and negative charges into different regions within the cloud. Before, addressing the
mechanism responsible for charge separation, it would be pertinent to outline the distribution
of charges in a thunderstorm.

5.6.1 Distribution of Charges in a Thunderstorm


The fair-weather electric field owes its existence to the presence of a conducting layer in the
upper atmosphere called the electrosphere. The electrosphere, together with the earth, act as a
huge spherical capacitor. While all clouds are to some extent electrified, the electrification of
a well-developed convective cloud is of such a magnitude that electric charges actually
CLOUDS AND PRECIPITATION u 139

separate within the thundercloud. Direct measurements from instrumented aircraft, or special
radiosondes, known as alto-electrographs, have provided a clear picture of the distribution of
the electrical charges in a thunderstorm. A schematic diagram showing the distribution of
electrical charges in a thunderstorm is shown in Figure 5.10.

FIGURE 5.10 Distribution of electrical charges in a thunderstorm cell.

The above figure clearly shows that an average thunderstorm contains positive charge
(~ +24C) in the upper regions of the cloud and a negative charge (~ – 20C) in the lower
region of the cloud, but above the 0°C isotherm. Furthermore, the average thunderstorm also
contains a small region of positive charge (~ +4C) at a height below the 0°C isotherm. The
above distribution of charges in a thunderstorm provides for leaking of the upper positive
charges to the base of the electrosphere through the highly conducting atmosphere at regions
above the thunderstorm cloud top. The electrical conductivity of the air, however, is low at
levels below the base of the thundercloud. When the electric voltage between the positive and
the negative charges is large enough, discharges (lightning discharges) take place between
clouds or between cloud and the earth surface. Lightning flashes from cloud to base primarily
transport negative charges from the base of thunderstorm clouds to the ground. Hence, the
above distribution of charges in a thunderstorm ensures that the fair-weather electric field is
maintained. Nearly 2000 thunderstorm cells are estimated to be present over the earth at any
given time. An average thunderstorm generates charge at the rate of 1 C km–3 minute–1. The
above global distribution of thunderstorms generates enough charges to maintain the fair-
weather electric field.

5.6.2 Mechanisms for Charge Separation


From common experience, it is known that lightning flashes mostly occur only in clouds that
extend above the 0°C isotherm, i.e. those clouds which contain both ice particles and
supercooled droplets. Furthermore, it is also known that the beginning of strong electrification
follows the manifestation of heavy precipitation in thunderclouds, i.e. after the development of
graupel and hailstones. From the above, it may be presumed that the processes of formation
of ice crystals or the formation of solid precipitation such as graupel and or hail, influence
charge separation. The initial theories, which were proposed to explain charge separation,
140 u BASICS OF ATMOSPHERIC SCIENCE

were based on the thermoelectric effect. The thermoelectric effect deals with the direct
conversion of temperature differences to electric voltage and vice versa and gives rise to a
potential difference DV across a rod of ice which is maintained at a steady temperature
difference between its two ends. Usually, the cold end of the ice becomes positively charged
and the warm end becomes negatively charged with the voltage difference DV in mV of the
order of twice the temperature difference DT expressed in °C (i.e. DV » 2 DT). Assume a
hailstone or a graupel particle, whose surfaces are at a relatively higher temperature, falls
through a cloud consisting of small ice crystals and supercooled droplets. The surfaces of a
hailstone or graupel are warmer due to the latent heat of freezing released by the large
number of supercooled liquid droplets colliding with the hail or graupel. Due to the
thermoelectric effect, during the collision of the small ice crystals and hail/graupel, the former
(ice crystals) gets positively charged, while the hailstone/graupel becomes negatively charged.
The ice crystal retains the above positive charge after rebounding from the collision with hail/
graupel and since it is small, and has relatively lower terminal fall speed is carried up to the
upper regions of the cloud by the updraft. The hailstone/graupel on the other hand, being
heavier and having higher terminal fall velocities carry these negative charges to the lower
regions of the cloud.
The other process, which also utilizes the thermoelectric effect to explain charge
separation, happens when a supercooled droplet collides with a hailstone and is about to
freeze. It was mentioned earlier that the cross-section of hail consists of alternate layers of
opaque ice and clear ice, and the opaque layer forms when small supercooled droplets freeze
rapidly on impact with the hail. During this freezing of small supercooled droplets, a large
number of ice splinters may be thrown into the air. The inner surface of the shell in contact
with liquid water is then at a higher temperature (0°C), while the outer surface of the shell is
at a lower temperature and cooling towards the environmental temperatures. Due to the
thermoelectric effect, positive charges accumulate on the shells at outer surface. During this
freezing of small supercooled droplets, a large number of ice splinters may be thrown into the
air and these splinters having predominantly positive charge are taken to the upper regions of
the cloud by the updraft. The remaining shell of the hailstone is negatively charged and being
heavier is taken to the lower regions of the cloud by its larger terminal fall speed.
A mechanism based on induction charging theory is also proposed to explain charge
separation. Due to the normal fair-weather field, all precipitating solid and liquid particles are
polarized such that their upper surfaces are negatively charged, while their lower surfaces are
positively charged. When cloud particles collide with downward moving precipitating
particles, negative charges will be transferred to the precipitating particles, while the cloud
particles become positively charged. The negatively charged precipitating particles are taken
to the lower regions of the cloud, while the positively charged cloud particles are taken to the
upper regions of the cloud by updrafts.
The above three theories, while explaining the major positive and negative charges at
the upper and lower regions of the cloud are in no position to explain the small region of
positive charge seen at heights below the 0°C isotherm. The small region of positive charges
is attributed to the charging of solid precipitation during melting. Laboratory experiments
have indicated that ice particles can receive positive charge during melting due to bursting of
air bubbles. Also, splashing of water droplets on melting ice particles can result in the ice
particles having large positive charges
CLOUDS AND PRECIPITATION u 141

5.6.3 Lightning Discharge


The dielectric breakdown refers to a rapid reduction in the resistance of an electrical insulator
that can lead to a spark and has a value of three million volts per metre for air. An average
bolt of lightning discharge carries a negative electric current of 40 kiloamperes (kA) although
some lightning discharges can carry an electric current of up to 120 kA. The average
lightning discharge duly transfers a charge of five coulombs and is equal to the power
necessary to light a 100-watt light bulb for about two months. Lightning discharges heat up
the nearby air to about 10,000°C nearly instantly and such intense heating creates a shock
wave that is heard as thunder. Studies with high-speed cameras have shown that most
lightning flashes are multiple events, consisting of as many as forty-two main strokes, each of
which is preceded by a leader stroke. All strokes follow an initial ionized path, which may be
branched, along with the current flows. The average interval between successive lightning
strokes is 0.02 s, while the average flash lasts for about 0.25 s. Lightning flashes travel at a
speed of one third the speed of light, i.e. at a speed of 1 ´ 108 m s–1. Since the duration of
one powerful stroke is no more than 0.0002 s, the intervals between strokes account for most
of the duration of a lightning flash.
The first stage in the formation of a lightning discharge from cloud-to-ground is the
development of a rapid and staggered advance of a shaft of negatively charged air called as
the stepped leader. The stepped leader is not a single column of ionized air; rather it branches
off from the main trunk at several places. The stepped leader is, however, not visible and
follows the path of least resistance. Each section of the stepped leader moves downward
about 50 m in about one microsecond with the diameter of the stepped leader being about
10 cm. The stepped leader pauses for about 50 ms as electrons pile up at its tip and generate
a strong electric field in the surrounding area. When the stepped leader approaches the
ground, electrons on the surface retreat from the leader creating a region of positive charge.
Corona discharges or sparks are released from tall objects on the surface and reach out to the
approaching stepped leader. When the downward moving stepped leader connects with a
surface corona discharge or a spark discharge, a continuous path between the cloud and the
ground is established and a powerful return stroke is triggered. The return stroke, which is
illuminated rapidly, moves upwards into the cloud following the ionized trail of the stepped
leader, stripping the electrons from its path. The above electrical discharge of the first return
stroke neutralizes some, but not all of the negatively charged ions near the base of the cloud.
Hence, another leader (now called a dart leader) forms within about a tenth of a second and
starts coming down from the cloud following a direct path to the surface. Subsequently, a dart
leader triggers a second return stroke. The combination of several return strokes is called a
lightning flash, whose net effect is to transfer electrons from the cloud to the ground. The
above-mentioned lightning flash where electrons are brought from the cloud to the ground is
called a negative cloud-to-ground flash. The positive cloud-to-ground flash is less common
than the negative cloud-to-ground flash. Since the positive charges are concentrated near the
top of the cloud, positive flashes are to come from the higher altitudes in the cloud. The
positive flashes, which account for about 10% of all lightning flashes are usually composed of
a single stroke, and typically lasts longer, making forest fires more likely.
142 u BASICS OF ATMOSPHERIC SCIENCE

SOLVED EXAMPLES

1. Show that for a plane surface of liquid in equilibrium with its vapour, the chemical
potentials in the liquid and vapour phases are the same.
Solution: If the liquid is in equilibrium with its vapour, molecules may condense or
evaporate at constant temperature and pressure. Under the above conditions, the external
work done by a unit mass of a body over and above any pda work equals to the decrease in
the Gibbs free energy of the body, i.e. da = – dg.
Since the only work done during evaporation is the work of expansion, pda, da = 0.
Hence dg = 0. If there is no change in the Gibbs free energy during evaporation, the
chemical potential of the molecules in the liquid phase must be the same as the chemical
potential of the molecules in the vapour phase.
2. Derive an expression for the difference in the vapour and liquid phases, in terms of the
actual vapour pressure e, temperature T and the saturated vapour pressure es corresponding
to a plane surface of liquid at temperature T.
Solution: For a pressure e and temperature T, let the chemical potentials in the liquid and
vapour phases be denoted by ml and mv. If the pressure changes irreversibly by de at a
constant temperature, the expression for a chemical potential as applied to the case of a
single vapour molecule is given as
dmv = vv de
where vv is the volume occupied by a single molecule in the vapour phase at temperature T
and pressure e. In a similar manner, the change in the chemical potential for the liquid phase
corresponding to an irreversible change in pressure de at constant temperature is given as
dml = vlde
where vl is the volume occupied by a single molecule in the vapour phase. Subtracting the
above two equations from one another, one gets
d(mv – ml) = (vv – vl)de
Since vv >> vl, the above equation becomes
d(mv – ml) = vv de
Applying the ideal gas equation for one molecule in the vapour phase, one gets
evv = kT
where k is the Boltzmann constant. Combining the last two expressions, one gets
L5
E NW  N M

EF
F
In the earlier solved problem it was shown that ml = mv, when e = es. Hence integrating
N W  NM F
L5
Ô E NW  N M

= Ô EF
 FT
F

ÈFØ
mv – ml = L5 MO
ÉÊ F ÙÚ
T
CLOUDS AND PRECIPITATION u 143

3. How many cloud condensation nuclei (CCN) are expected over the continental region
per unit m3 volume of air if the radii of the nuclei are (i) 0.3 mm, (ii) 0.7 mm, and
(iii) 1.0 mm. Let the constant c that appears in the radius-number density relationship be
equal to 5 ´ 106 mm4 m–3.
Solution: The relationship between the number densities of the CCN as a function of
radius over the continental region is given by
n = cR–4
For R = 0.3 mm, n = 5 ´ 10 6 ´ 0.3 –4 = 6.17 ´ 108 m–3.
For R = 0.7 mm, n = 5 ´ 106 ´ 0.7–4 = 2.08 ´ 107 m–3.
For R = 1.0 mm, n = 5 ´ 106 ´ 1.0–4 = 5.0 ´ 106 m–3.
4. Given the aerosol number distribution as
E/
$%  C
E MPH %

where N is the concentration of aerosols with diameter greater than D, C is a constant and
E/
value of b having a value between 2 and 4, derive the expressions for (i) (ii) the
E%
E4 E7
surface and (iii) the volume distributions 
E MPH %
E MPH %

Solution:
E/
(i) Since $%  C we have
E MPH %

E/ E% –b
E% E MPH %
= CD
E/ ÿ E MPH %
$%  C E $  $
= $%  C MO %
% C %  C 

E% E% MO E% MO % MO

(ii) Since the surface distribution is dS = pD2 dN, one gets


E4 E/
Q % Q $%   C
E MPH %
E MPH %

Q
(iii) Since the volume distribution is E7 %  E/ one gets

E7 Q  E/ Q
% $%  C
E MPH %
 E MPH %

5. Find the equilibrium relative humidity over a pure droplet of radius 0.1 mm, and at tem-
perature 20°C.
Solution: The equilibrium relative humidity over a pure water droplet is given by
F È D Ø
FYQ É
FT Ê 53 ÙÚ
144 u BASICS OF ATMOSPHERIC SCIENCE

where c1 = 0.3335 K mm, T is the temperature in K and R is the drop radius in mm.
F È  Ø
FYQ É 
FT Ê  – ÙÚ
The equilibrium relative humidity over the pure water droplet expressed as a percentage is
given by 101.145%.
6. Find the equilibrium relative humidity over a droplet of radius 0.1 mm, temperature 20°C
containing 10–15 g of ammonium sulphate. Given the molecular weight and the approximate
ion count of ammonium sulphate are 132.13 and 3, respectively.
Solution: The equilibrium relative humidity over a solution droplet is given by
È D Ø
FYQ É
FT„„ Ê 53 ÙÚ
FT DJNT
 
.T3
where FT„„ is the actual saturation vapour pressure in equilibrium over a solution with a
curved surface, while es is the saturation vapour pressure over a plane water surface at the
same temperature. Also, R is the drop radius, T is the temperature, Ms is the molecular
weight, ms is the mass, i is the ion count and c2 = 4.3 ´ 1012 mm3 g–1. The equilibrium
relative humidity is given by
ÈD Ø È  Ø
FYQ É  Ù FYQ É
FT„„ Ê 53 Ú Ê  – ÙÚ

FT
 Ë  –  

Û
D JNT
. T 3 Ì  Ü
Í  
 Ý
The equilibrium relative humidity of ammonium sulphate expressed as a percentage is 92.15%.
7. Find the critical radius (drop radius at the peak of the Kohler curve) for 10–16 g of
ammonium sulphate at 0°C. Given the molecular weight and the approximate ion count of
ammonium sulphate are 132.13 and 3, respectively.
Solution: The critical radius (drop radius at the peak of the Kohler curve) expression is
given by
DJNT5
3
.T
where Ms is the molecular weight, ms is the mass, T is the temperature, i is the ion count and
c3 = 3.8681 ´ 1013 mm2 K–1 g–1. The drop radius at the peak of Kohler curve is

 –  
 


3  PN

8. Find the supersaturation value for 10–16 g of ammonium sulphate at 0°C. Given the molecular
weight and the approximate ion count of ammonium sulphate are 132.13 and 3, respectively.
Solution: The supersaturation value expression is given by
D . T
4 
JNT5
CLOUDS AND PRECIPITATION u 145

where c4 = 1.278 ´ 10–15 K3 g. The supersaturation value for ammonium sulphate is

 –   

4 
  


The supersaturation expressed as a percentage is 0.53%.
9. Find the number density of cloud condensation nuclei activated in continental air, given a
certain supersaturation percentage of 0.4%.
Solution: The number density of cloud condensation nuclei activated in continental air is
given as
nCCN = c ´ (100 S)k,
where S is the supersaturation fraction, c = 6 ´ 108 m–3 and k = 0.5.
Substituting the values, one gets
nCCN = (6 ´ 108) (0.4)0.5 = 3.79 ´ 108 m–3.
10. Find the number density of sea salt cloud condensation nuclei activated in maritime air,
given a certain supersaturation percentage of 0.4%.
Solution: The number density of sea salt cloud condensation nuclei activated in maritime
air is given as
nCCN = c ´ (100 S)k
where S is the supersaturation fraction, c = 1 ´ 108 m–3 and k = 0.7. Substituting the values,
one gets
n CCN = (1 ´ 108) 0.40.7 = 5.27 ´ 107 m–3.
11. Given the available liquid water as 5 g kg –1 and air density as 1 kg m–3 inside a cloud, find
the final drop size for a given number density of activated CCN being (i) 108 m–3, (ii) 109 m–3
and (iii) 1010 m–3.
Solution: The final drop size can be obtained from the following relationship:

Ë  SBJS SM Û
3 Ì Ü
Í Q SXBUFS O$$/ Ý

where rl is the available liquid water expressed as kgwater kg–1air, nCCN is the number density
of CCN, rair and rwater are the densities of air and water, respectively.
 
Ë  
 –  
Û


For (i), 3 Ì Ü  –   


 PN
ÌÍ Q ÜÝ
 



 
Ë  
 –  
Û


For (ii), 3 Ì Ü  –   


 PN
ÌÍ Q ÜÝ
 



 
Ë  
 –  
Û 

For (iii), 3 Ì Ü  –   PN
ÍÌ Q ÝÜ
 


146 u BASICS OF ATMOSPHERIC SCIENCE

12. What should be the updraft velocity needed to keep a typical cloud droplet of radius 10 mm
from falling?
Solution: The expression for the updraft velocity for typical cloud droplets of radius < 40 mm
is given by the Stokes drag law and is of the form
w = kR2,

where k = –1.19 ´ 108 m–1 s–1, R is the droplet radius in m and w is the updraft velocity in
m s–1. The negative sign on k indicates that the drop is falling. Substituting, one gets,
w = (–1.19 ´ 108) ´ (10–6)2 = – 1.19 ´ 10–4 m s–1
13. Find the terminal velocity of a rain droplet of radius 1400 mm at a pressure of 700 hPa?
Solution: Unlike the cloud droplet, the terminal velocity of a falling raindrop is not
determined by Stokes drag law. For rain droplets having radius in the range of 40 mm to
2500 mm, an empirical expression for the terminal velocity is given as

Ë È 3  3Ø Û
X  D Ì X  FYQ É ÙÚ Ü
Í Ê 3 Ý
where w0 = 12 m s–1, R0 = 2500 mm, R1 = 1000 mm, and c =1.

Ë È    Ø Û
X   – Ì  FYQ É
Ê ÙÚ Ü   NT 
Í  Ý
14. If an average of 4 g kg–1 of water existed in the troposphere between 1000 hPa and 300 hPa,
find the precipitable water depth.
Solution: The expression for the precipitable water depth is given by

S5 Q#  Q5

EX
H SM

where rT is the average total water vapour mixing ratio expressed in kg kg–1, rl is the density
of water, g is the acceleration due to gravity and pB and pT are the ambient air pressures at
the bottom and top of the column. Substituting, one gets

 –  
  
– 
EX  N



15. A drop enters the base of a cloud with a radius r0 and grows by the collision-coalescence
mechanism while travelling up and down in the cloud until it reaches the cloud base with a
radius rf. Show that rf is a function only of r0 and the updraft velocity w in the cloud. It is
assumed that the updraft velocity in the cloud and the collection efficiency are assumed
constant.
Solution: The following relation gives the growth of a drop in a cloud:

ES WXM &


EI  S X  W

M
CLOUDS AND PRECIPITATION u 147

where r1, v1 are the radius and velocity of the collector drop, wl is the liquid water content,
E is the collection efficiency, rl is the density of water and h is the height above a fixed
level (say cloud base). Integrating the above equation between cloud base and a height H
above the cloud base, one gets

X  W

) S)

Ô XM EI S
 M
Ô W& ES
 S


Substituting H = 0 and rH = rf in the above expression, the above equation becomes

X  W

SG

 S
 M
Ô W& ES
SP

Since E and w are assumed constant,


SG SG
ES
X Ô W
Ô ES SG  S
S S

SG
ES
Hence, the final radius SG S X Ô W
S

REVIEW QUESTIONS

1. Cloud and raindrops are accelerated downward towards the ground by gravity. The
equilibrium velocity resulting from a balance between gravity and frictional drag is called
the terminal velocity. For cloud droplets, the terminal velocity can be expressed as:

X  L 3
where k1 = –1.19 ´ 108 m–1 s–1. The negative sign indicates that the drops are falling. Given
cloud droplets 15 mm in size, how strong would an updraft have to be to keep the droplets
suspended in the cloud?
2. An alternate method for bringing air to saturation is by mixing two air parcels together. Is
the above-mentioned method responsible for the effect that one can often see one’s breath on
a cold morning?
3. During winter, there is an important wintertime expression, “clear moon, frost soon.”. Is
there any rationale for the above expression?
4. Is it true that relative humidities never/ever reach 100% in polluted air?
5. Are advection fogs rare in tropics? Explain.
6. Let the sky be overcast and let it be raining. Can one accurately tell whether the cloud
responsible for the rain is a nimbostratus or a cumulonimbus without venturing outdoors?
7. Why can’t the Bergeron process happen in warm clouds?
148 u BASICS OF ATMOSPHERIC SCIENCE

8. Calculate the time it would take for a drizzle having droplets of diameter of 250 mm to
reach the surface if it falls at its terminal fall velocity from the base of a cloud 1.5 km above
the ground. Let the air be saturated beneath the cloud. Also, assume that the drizzle does not
evaporate, and the air is still.
9. Suppose the drizzle evaporates on its way to the ground. If the drop size is 250 mm for the
first 400 m of descent, 150 mm for the next 400 m, and 50 mm for the final 200 m, how
long will it take the drizzle to reach the ground if it falls in still air?
10. Incorporating an updraft (w = 0.1 m s–1) to Problem 9, would the drizzle safely reach the
ground?
11. At what relative humidity will pure water droplets of the following radii grow by
condensation through homogeneous nucleation: (i) 12 mm, (ii) 6 mm, and (iii) 2 mm.
12. For cloud droplets in warm clouds growing by condensation, the radius of the droplet (R)
increases as the square root of time (t) and the expression for the growth is given by
R » c × (DSt)0.5
where c is a dimensionless constant, D is a diffusion coefficient, and S is a measure of
supersaturation of the air far away from the droplet. (i) Plot R (in mm) versus time for t = 0
to 2 hours, in steps of 10 minutes, assuming c = 0.0023, D = 2 ´ 10–5 m2 s–1, and S = 1%
supersaturation, (ii) Based on the above plot, would you expect a droplet growing by
condensation only to eventually fall out of the cloud?
13. Why is a warm, tropical cumulus cloud more likely to produce precipitation than a cold,
stratus cloud, all other conditions being equal?
14. Haze particles can form when the relative humidity is less than 100%. Are these haze
particles made up of pure water droplets or solution droplets? Explain.
15. Consider a thick nimbostratus cloud containing both ice crystals and cloud droplets having
about the same size. Mention the precipitation process, which is likely to produce rain from
this cloud.
16. Why do very small cloud droplets of pure water evaporate even when the relative humidity
is 100%? Explain.
17. Maritime clouds, i.e. clouds that form over water are usually more efficient in producing
precipitation than clouds that form over land. Explain?
18. The radar reflectivity factor (Z in units of mm6 m–3) is expressed as:

6 %
;
7
where D is the diameter of the raindrops in some volume (V). The radar reflectivity is
expressed as decibels dB of Z or dBZ and is defined as dBZ = 10 log (Z). Assume that one
has (i) one drop 4 mm in diameter, (ii) two drops 2 mm in diameter, and (iii) one drop
3 mm in diameter, all in unit volume (1 m–3) of air, then calculate the value of the radar
reflectivity in dBZ.
6 Governing Laws of
Atmospheric Motion

Atmospheric dynamics is the study of those motions of the earth’s atmosphere that are
associated with the weather and climate over our planet earth. In atmospheric dynamics, the
atmosphere (fluid) is regarded as a continuous medium, and the fundamental laws of fluid
mechanics and thermodynamics are expressed in terms of conservation laws of mass,
momentum and energy along with the equation of state. These conservation laws, which
govern the motion of our earth’s atmosphere, are expressed in terms of partial differential
equations involving the fluid velocity, density, pressure, and temperature. Newton’s second
law of motion (after incorporating the acceleration terms of the rotating frame) forms the
basis for the principle of conservation of momentum, while the laws of thermodynamics
provide the basis for the conservation of energy.
The atmosphere is in general in a state of motion and this atmospheric motion can be
described using two broad approaches. The first approach known as the Eulerian approach is
to describe the fluid motion at fixed points in space. Essentially, this means that the frame of
reference, which is used to describe the atmospheric motion, is fixed with respect to the earth.
A reference frame, which is fixed to the earth, is actually rotating and one has to account for
the acceleration terms of the rotating frame before applying Newton’s second law of motion.
The Eulerian approach is a very convenient means of describing the atmospheric motions and
will be discussed in this book. The second approach known as Lagrangian approach
considers a frame of reference, which moves with the fluid, describing the various forces and
their effects on an air parcel.
In the atmosphere, the rate of change of any atmospheric variable following the fluid
motion is different from the rate of change of the same variable at a fixed point in the
associated reference frame and there is a need to relate the above two rates of changes. The
atmospheric circulations can be described using simple models based on horizontal balance of
forces. Different combination of forces give rise to varied horizontal flows such as the
geostrophic, gradient and cyclostrophic flows. The vertical variation of geostrophic flow
known as thermal winds provides the relationship between the vertical shear and the average
149
150 u BASICS OF ATMOSPHERIC SCIENCE

horizontal temperature gradient of the layer. Section 6.1 presents the governing equations in a
rotating coordinate system and introduces the two additional acceleration terms, which arise
due to the non-inertial nature of the rotating coordinate system, while in Section 6.2, both the
gravity and the pressure gradient forces are discussed. In Section 6.3, the concepts of total,
local and convective derivatives are presented, while the continuity equation is derived in
Section 6.4. Section 6.5, discusses the different examples of horizontal balanced flow such as
geostrophic, gradient and cyclostrophic flow. Lastly, Section 6.6 presents the description of
the vertical variation of geostrophic wind, viz. thermal wind.

6.1 EQUATION IN A ROTATING COORDINATE SYSTEM—


CENTRIPETAL AND CORIOLIS ACCELERATION
It becomes necessary in certain situations to recast Newton’s equation of motion with respect
to frames of reference which are rotating with constant angular velocity, i.e. with respect to a
frame of reference which is essentially an accelerating frame of reference. One way of
handling the above is to determine the acceleration of the frame of reference in such a
situation and equate the above acceleration of the frame of reference to fictitious or pseudo
forces. Such a procedure gives rise to the manifestation of two fictitious forces which are
dealt in detail in the following subsection.

6.1.1 Introduction
Newton’s laws of motion are known to be valid in a class of reference frames called the
inertial frames. Newton’s first law of motion (also known as the law of inertia) states, In an
inertial frame every particle not acted upon by a net external force, i.e. a free particle has a
constant velocity. The motion of a free particle in inertial frames will be described by straight
lines, for otherwise, the velocity of the free particle would change violating the law of inertia.
Since a free particle traverses an equal measure of space in equal measure of time, the above
implies that along the straight line path of a free particle, space is uniform and homogeneous
and so also time. Also, since the direction of the straight line path of a free particle can be
any, the same implies that space is also isotropic. A closed system is one which is not acted
by an external field of force. Hence an inertial frame ensures homogeneity of space (every
point in space is as good as any other point for the description of a closed system), isotropy of
space (every direction in space is as good as any other direction for the description of a
closed system), and homogeneity of time (every moment of time is as good as any other
moment of time for the description of a closed system). Since the free particles travel in
straight lines, an inertial frame with all its axes as straight lines would be convenient. Hence
a rectangular Cartesian coordinate system is usually used to represent an inertial frame.
To describe the atmospheric motion with respect to an observer fixed to the earth, it
would be convenient to utilize a frame of reference, which is fixed to the earth. The earth is
rotating around an axis passing through its geographical north pole with a constant angular
velocity. A reference frame attached to the earth rotates with the earth and hence is no longer
an inertial frame. However, by taking into account the additional acceleration terms of the
rotating frame of reference as fictitious forces, it is possible to apply Newton’s laws of motion
to a non-inertial frame of reference fixed to the earth.
GOVERNING LAWS OF ATMOSPHERIC MOTION u 151

6.1.2 Rotating Frame of Reference


Consider a fixed rectangular frame S with (i, j, k) as the fixed unit vectors and a frame S ¢
with unit vectors (i¢, j¢, k¢), originally coincident with S and rotating with a velocity w (= nw)
around their common origin O (Figure 6.1).

k

n

w

j
O S

i

FIGURE 6.1 Rectangular Cartesian axes of an inertial S and a rotating S¢ reference frames. The
frame S¢ rotates about their common origin O with an instantaneous angular
velocity of rotation w with respect to the inertial frame S.

The question posed is the following: How are the components of a given vector A and
its time derivative related in the two frames S and S¢. Although the vector A is the same in
both the frames, the components of the vector A will be different in both the frames since the
unit vectors in both the frames are pointing in different directions. The convention followed
for vectors and scalars are that while vectors are written boldly, scalars are not written boldly.
Hence,
A |fixed = A |rot (6.1)
A |fixed = A1 i + A2 j + A3 k = A1¢ i¢ + A2¢ j¢ + A3¢ k¢ = A|rot (6.2)
Since the time derivatives of Afixed and Arot are also physical vectors, they must also be equal
in the two frames
E E
< " GJYFE > < " SPU > (6.3)
EU EU
Ë E" Û
The left-hand side of Eq. (6.3) is usually denoted by Ì since this is the rate of
Í EU ÜÝ GJYFE
variation of A as measured by an observer in the fixed frame S and is given by

Ë E" Û
E
< " GJYFE > ÌÍ EU ÜÝ
E"
J  E" K  E" L (6.4)
EU GJYFE EU EU EU
152 u BASICS OF ATMOSPHERIC SCIENCE

The unit vectors in the fixed frame S do not change their orientation with time. However, the
same is not true for the rotating frame S¢ and so
E E
< " > = "„ J „  "„ K„  "„ L „
(6.5)
EU EU
SPU

E"„ E"„ E"„ EJ „ EK„ EL „


= J„  K„  L „  "„  "„  "„ (6.6)
EU EU EU EU EU EU
Ë E" Û
The first three terms of Eq. (6.6) can be denoted as Ì as they represent the time
Í EU ÜÝ SPU
variation of A as measured by an observer in the rotating frame S¢.
An arbitrary small rotation Dq about some direction with unit vector n in time Dt
passing through the origin brings about a change in any vector B which is given by
'# È 'R Ø
ÉÊ OÙ – # (6.7)
'U 'U Ú
Taking the limit as Dt ® 0, one gets
E#
XO – # Z–# (6.8)
EU
Applying Eq. (6.8) to i¢, j¢ and k¢, one gets
EJ „ EK„ EL „
Z – J „ Z – K„ Z – L„ (6.9)
EU EU EU
Using Eq. (6.9) in Eqs. (6.6) and (6.7)

Ë E" Û Ë E" Û
ÌÍ EU ÜÝ = Ì  Z – "„ J „  "„ K„  "„ L „
(6.10)
GJYFE Í EU ÜÝ SPU

Ë E" Û Ë EA Û
ÌÍ EU ÜÝ = Ì Z–A (6.11)
GJYFE Í EU ÜÝ SPU
Let A = r be the position vector, then one gets

Ë ES Û Ë ES Û
ÌÍ EU ÜÝ = Ì Ü Z–S (6.12)
GJYFE Í EU Ý SPU
or [v]fixed = [v]rot + w ´ r (6.13)
Denoting [v]fixed as v0 and [v]rot as v, Eq. (6.13) becomes
v0 = v + w ´ r (6.14)
Let A = v0. Substituting in Eq. (6.11)
Ë EW  Û Ë EW  Û
Ì EU Ü = Ì Ü  Z – W (6.15)
Í ÝGJYFE Í EU Ý SPU
GOVERNING LAWS OF ATMOSPHERIC MOTION u 153

E
= W  Z – S
 Z – W  Z – S
(6.16)
EU
EW E Z
=  – S   Z – W  Z – Z – S
(6.17)
EU EU
EW
Let us examine the terms in the right-hand side of Eq. (6.17). The term is the acceleration
EU
EZ
as measured in a rotating coordinate system S ¢. The term – S is known as the Eulerian
EU
term which vanishes for uniform rotation (uniform w) as in the case of the rotation of the
earth. The term 2w × v represents the Coriolis acceleration; this term manifests whenever
there is a relative motion with respect to
the rotating frame. The Coriolis w
acceleration is perpendicular to both the
velocity in the rotating frame v as well as
the angular velocity w. The term w ´ (w ´ r)
represents the centripetal acceleration; this
w´r
term manifests even when there is no
– w ´ (w ´ r)
relative motion with respect to the rotating
frame. The centripetal acceleration has a
magnitude equal to w2 R, where R is the
perpendicular distance from the point r to
the axis of rotation and is directed away
from the rotation axis (Figure 6.2). Eq.
(6.17) can be used to write the governing
equations in the rotating frame of reference Figure 6.2 Centrifugal force directed away
once the form of the Newton’s second law from the axis of earth’s rotation.
of motion is known in the inertial frame.

6.1.3 Equation of Motion in an Inertial Frame of Reference


In order to obtain the equations of motion which govern atmospheric motion in an inertial
frame of reference, one has to apply Newton’s second law of motion to a fluid element of
density r moving with velocity v0 in the presence of a pressure gradient Ñp and a gravitational
field g. The forces exerted on a fluid element can be broadly classified into two types— body
forces and surface forces. Body forces (for example gravitational forces), as applied to a fluid
element are those external forces, which act throughout the mass of the body. In other words,
the body forces do not depend on the presence of other surrounding fluid elements. Surface
forces (for example, pressure gradient forces), as applied to a fluid element, on the other
hand, depend on or are affected by the presence of the surrounding fluid elements. Newton’s
second law of motion as applied to a fluid element in an inertial frame of reference gives the
following equation:
EW 
'HSBW  'QSHSBE  'GSJDU (6.18)
EU
154 u BASICS OF ATMOSPHERIC SCIENCE

The first term in the right-hand side of Eq. (6.18) is the gravitational force per unit mass
while the other two terms represent the pressure gradient force per unit mass and the frictional
force per unit mass. The expressions for the pressure gradient force and the gravity force will
be derived in the next section.

6.2 GRAVITY AND PRESSURE GRADIENT FORCES


In addition to the two fictitious forces introduced earlier, a fluid particle is subjected to two
other forces, namely the gravity and the pressure gradient forces. While the pressure gradient
forces (an example of a surface force) is a real force, the gravity force (an example of a body
force) is the resultant of the Newtonian gravitational force and the centrifugal force. The
following subsection derives the expression for the pressure gradient force and the gravity force.

6.2.1 Pressure Gradient Force


One of the most important forces exerted on a fluid element is the pressure gradient force and
an expression for the same for per unit mass will be derived in this subsection. Consider a
small fluid element within a fluid medium in the shape of a parallelepiped whose sides are dx,
dy and dz as shown in Figure 6.3.

p dy dz Ê ∂p ˆ
-Áp + dx ˜ dy dz
dz Ë ∂x ¯
dy

dx
FIGURE 6.3 The component of the pressure gradient force in the x direction.

The force due to the surrounding fluid on the left face normal to the yz plane is
p dy dz (6.19)
since the force is acting in the positive x direction. The force due to the surrounding fluid on
the right face normal to the yz plane can be written by expanding the pressure in a Taylor
series expansion about x and retaining only the linear term, i.e.

È ˜Q Ø
Q Y  EY
Q Y
 EY É
Ê ˜Y ÙÚ
and multiplying the same with the area of the right face dy dz. Hence the pressure force on
the right face normal to the yz plane is

È ˜Q Ø
É Q  EY Ù EZ E[ (6.20)
Ê ˜Y Ú
GOVERNING LAWS OF ATMOSPHERIC MOTION u 155

The negative sign in Eq. (6.20) indicates that the force on the right face is in the negative
x direction. The net pressure force in the x direction is the sum of Eqs (6.19) and (6.20)
which is
È ˜Q Ø
 É Ù EY EZ E[ (6.21)
Ê ˜Y Ú
The net pressure force per unit mass in the x direction is obtained by dividing Eq. (6.21) with
the mass of the small fluid element and is expressed as
 ˜Q
 (6.22)
S ˜Y
where r is the density of the fluid element (parallelepiped). Using similar arguments as
above, the expressions for the pressure force per unit mass in the y and z directions are
respectively,
 ˜Q  ˜Q
 BOE  (6.23)
S ˜Z S ˜[
Equations (6.22) and (6.23) show that the pressure force per unit mass is a three-dimensional
vector and is expressed as

 ³Q (6.24)
S
Equation (6.24) shows that the pressure force is due to the gradient of pressure and is directed
in the opposite direction of the gradient (ascendant), i.e. from a region of high pressure
towards low pressure. Due to the above reason the pressure force is also known as the
pressure gradient force.

6.2.2 Gravitational Force


Consider a fluid element at rest on a rotating earth. Assume that the fluid element is not
subjected to any other external forces other than the gravitational, i.e. the attractive force
between the bodies as described in Newton’s law of universal gravitation. Let us indicate the
gravitational force per unit mass by g.

6.2.3 Equation of Motion in a Rotating Coordinate System


Utilizing Eq. (6.24) in Eq. (6.18), one gets
EW  
H ³Q  ' SJDU (6.25)
EU S G

Substituting Eq. (6.25) in Eq. (6.17) for a rotating earth with constant w, one gets
EW 
³Q H o Z – Z – S
o Z – W 'GSJDU (6.26)
EU S
156 u BASICS OF ATMOSPHERIC SCIENCE

Equation (6.26) gives the rate of change of velocity of a fluid element as observed from a
rotating frame of reference. The term –2w ´ v in Eq. (6.26) is regarded as the Coriolis force,
while the term – w ´ (w ´ r) is known as the centrifugal force. The above two forces
(Coriolis and centrifugal) arise due to the non-inertial nature of the rotating frames of
reference and are considered as pseudo or fictitious forces. The real forces are the pressure
gradient force, the friction forces and the gravitational forces.

6.2.4 Effects of Coriolis Force


The Coriolis force per unit mass –2wÿ ´ v is a velocity-dependent force and does not do any
work, as the instantaneous velocity v is always perpendicular to the Coriolis force. The earth
rotates from west to east around an axis passing through its geographical north pole. The
magnitude of the angular velocity of the earth’s rotation relative to fixed stars is

È Q Ø È  Ø
X ÉÊ  –  ÚÙ ÉÊ  ÙÚ  –   SBEJBOT T 

The first factor is the angular velocity of the earth’s rotation relative to the radius vector to
the sun, while the second factor is the correction factor to provide the angular velocity of the
earth’s rotation with respect to the fixed stars. The second correction factor is the ratio of the
number of sidereal days in a year to the number of solar days.
A fluid element with velocity v will experience a Coriolis force per unit mass given by
–2w ´ v. The magnitude of the Coriolis force can never exceed 2w v = 1.46 ´ 10–4v. Even for
an extremely high velocity of the order of 1 km s–1, the maximum magnitude of the Coriolis
acceleration is only 0.15 ms–2, which is much smaller than the acceleration due to gravity.
However, in many natural situations, the period of time over which this small acceleration
acts can be quite large resulting in marked effects.

6.2.5 Flow of Rivers on the Surface of the Earth


At the outset, it has to be mentioned that the Coriolis force does not solely govern the flow of
rivers although it does influence the river flow. The earth rotates from west towards east and
hence at any place on earth the angular velocity vector w is directed towards the north and is
parallel to the axis of rotation of the earth (Figure 6.4). The angular velocity w at any place
having a latitude j will be resolved along the vertical direction (w sin j) and the north-south
direction (w cos j). For convenience, let us choose the direction of flow of the river to be the
x-axis and the transverse horizontal axis to the left of the flow direction as the y-axis. Let the
direction of the river flow make an angle l (in the anticlockwise sense) with respect to the
geographical north direction. Then the angular velocity component directed to the north w cos j
can be resolved along and perpendicular to the direction of the river flow as w cos j cos l
and – w cos j sin l. Hence the earth’s angular velocity w with respect to the above reference
frame is
w= w [cos j cos l i – cos j sin l j + sin j k] (6.27)
GOVERNING LAWS OF ATMOSPHERIC MOTION u 157

The velocity of the river flow is v = v i and so the Coriolis force becomes
–2v w cos j sin l k – 2v w sin j j (6.28)
The vertical component of the Coriolis force Eq. (6.28) is very small in magnitude as
compared to gravity and can be neglected. However, the component of the Coriolis force
along the j axis has a value – 2v w sin j. The above quantity is independent of l and hence
does not depend on whether the river flow is
towards east or west, north or south. However, it
depends on the latitude and is negative (i.e. to
w
the right of the river flow) for northern
hemisphere (j > 0) and is positive (i.e. to the left
of the river flow) for southern hemisphere (j < 0).
Needless to mention the effects of Coriolis force
vanish at equator (j = 0). Hence the water in the
river flowing in any direction will experience a
Coriolis force causing a deviation towards the
right of the flow direction in the northern
hemisphere and to the left of the flow direction in
the southern hemisphere. Due to the above effect
of the Coriolis force, it is observed that the
corresponding banks of the river will be more
denuded. Thus the effect of Coriolis force on the
river flow is to raise the right/left banks of the FIGURE 6.4 The direction of w at any
rivers in the northern/southern hemisphere to a arbitrary point on the
higher level as compared to the left/right banks. earth surface.

6.2.6 Effects of Coriolis Force Due to Relative Motion


Along a Latitude Circle
Consider that a fluid element is set in motion in the eastward direction due to an impulsive
force. Since the fluid element is now rotating faster than the earth, the centrifugal force on the
fluid element will be increased. If w is the magnitude of the angular velocity of the earth’s
rotation and if R is the position vector from the axis of rotation to the fluid element, and u is
the eastward speed of the fluid element relative to the ground, then the total centrifugal force
may be expressed as

3 X  3   ÈÉÊ X V ØÙÚ 3  ÈÉÊ V ØÙÚ 3


 
È VØ
ÉÊ X  ÙÚ (6.29)
3 3 3
The first term in the right-hand side of Eq. (6.29) is just the centrifugal force due to the
V
earth’s rotation, while the third term is the same acceleration due to an additional rotation 
3
Normally, for large-scale atmospheric motion | u| <<ÿ w R and hence the third term is
negligible. The second term is the Coriolis force due to the relative motion parallel to a
158 u BASICS OF ATMOSPHERIC SCIENCE

latitude circle. The Coriolis force term can be resolved into components along the vertical and
the meridional direction (Figure 6.5). Hence the relative motion along the east-west
coordinate will produce acceleration in the north-south direction given by

È EW Ø È EX Ø
ÉÊ ÙÚ  X V TJO K K  É X V DPT K L
EU $PS Ê EU ÙÚ $PS (6.30)

A fluid element moving eastward in a horizontal plane in the northern hemisphere will be
deflected southward by the Coriolis force, while a westward moving fluid element in the
northern hemisphere will be deflected northward. In both the cases in the northern
hemisphere, the deflection is to the right of the direction of motion. The vertical component of
the Coriolis force is considerably smaller as compared to gravity and hence can be neglected.

FIGURE 6.5 Components of the Coriolis force arising due to relative motion along a latitude circle.

6.2.7 Effects of Coriolis Force Due to Relative Motion


Along a Meridian
Let a fluid element initially at rest on the earth be set in motion equatorward by an impulsive
force. In the absence of torques in the east-west direction, the fluid element will conserve its
angular momentum. Let the fluid element move from a radius R and latitude j 0 to a radius
R + dR and corresponding latitude j 0 + dj, equatorward in time d t. Since the fluid element
moves towards a region with a high angular momentum (larger R), its velocity will change by
d u in order to conserve the angular momentum. Writing down the conservation of angular
momentum as
È EV Ø
X 3 ÉÊ X  3  E 3 ÙÚ 3  E 3

The second factor in the right-hand side of the above equation can be written as
E3
3 ÈÉ   ØÙ and substituting in the above equation and neglecting the second-order
Ê 3Ú
GOVERNING LAWS OF ATMOSPHERIC MOTION u 159

differentials, one gets E V  X E 3 where dR = – a dj sin j 0, where a is the radius of the


earth. The above result shows that a relative westward velocity must develop for an
equatorward displacement and hence the Coriolis acceleration for the above case is obtained in
the limit d t ® 0, as

È EV Ø È EK Ø
ÉÊ ÙÚ X B É Ù TJO K  X W TJO K  (6.31)
EU $PS Ê EU Ú

EK
where W B is the velocity component in the north-south direction. Hence when a fluid
EU
element moves along a meridian, the Coriolis force acts along the latitude circle. The
horizontal accelerations are proportional to the quantity 2w sin j0 which is known as the
Coriolis parameter and is denoted by the symbol f.

6.2.8 Effects of Coriolis Force Due to Vertical Motion


Consider a fluid element being launched vertically up at a latitude j0 with a vertical velocity w.
Invoking the conservation of absolute angular momentum and following the same arguments
as in the earlier section (since both equatorward displacement and launching vertically gives
rise to increase of the distance to the axis of rotation from the fluid element), one arrives at
the result
du = –2w dR
However, dR in this case of a fluid element launched vertically is given by dR = dz cos j0
and so d u = –2w dz cos j0 gives rise to a Coriolis acceleration in the zonal direction given by
È EX Ø
ÉÊ Ù  X X DPT K  (6.32)
EU Ú $PS

E[
where X is the vertical velocity. Hence a fluid element launched vertically upwards
EU
experiences a Coriolis deflection along the latitude circle towards west at all latitudes in both
Q Q
the hemispheres since DPT K  !  GPS   K   For a fluid element launched vertically
 
downwards, the Coriolis deflection is along the latitude circle towards east at all latitudes in
both the hemispheres.

6.2.9 Rossby Number


Large-scale motion can be conveniently defined as those which are significantly affected by
the earth’s rotation. A measure of the importance of the Coriolis force for a particular
situation in the atmosphere can be obtained using the Rossby number. Let L and U be the
characteristic length scale and the characteristic horizontal velocity scale of a particular
motion in the atmosphere. The characteristic time scale or the time it takes for the fluid
-
parcel moving with a speed U to traverse a distance L is  If the above characteristic time
6
160 u BASICS OF ATMOSPHERIC SCIENCE

is much less than the period of rotation of the earth, it can be safely surmised that the fluid
parcel will hardly sense the earth’s rotation over the characteristic time scale of the motion.
Rather, for the earth’s rotation (Coriolis force) to be important, one requires that
- 6
!! X  PS F …
6 X -
where e is the Rossby number. Rossby number can also be defined as the ratio of the inertial
force to the Coriolis force. It is clear from the above that smaller the value of the Rossby
number, the more significant is the importance of the earth’s rotation (Coriolis force) on the
motion.

6.2.10 Gravity
The centrifugal force –wÿ ´ (w ´ r) can be written as a gradient of a scalar function similar
to the gravitational force. Also, both these forces do not depend on the relative motion with
respect to the rotating frame. Hence these two forces can be conveniently combined to get,
geff = g – w (w ´ r) (6.33)
where geff is known as the effective gravity and can be derived from the effective potential,

7FGG 7  N ] Z–S 
] (6.34)


where V is the actual gravitational potential energy of the fluid element due to the earth.
Equation (6.33) and (6.34) are effectively related by
m geff = –ÑVeff (6.35)
Figure 6.6 depicts the combined effect of the w
gravitational and the centrifugal force on earth.
The earth before cooling down to its present
form was in the form of a fluid state rotating
about its axis. The free surface of such a fluid
in a steady state assumes an equipotential
surface. Since the gravitational force as well as
the centrifugal force can be expressed as
gradient of scalar function (i.e. potentials), the
nature of the equipotential surface must include
the contribution of all potentials including that
due to gravitational and centrifugal forms. The
gravitational force is directed towards the
centre of the earth while the centrifugal force
is directed away from the axis of rotation and
FIGURE 6.6 The relationship between the
except for equator and poles the effective
Newtonian gravitational force
gravity is not directed towards the centre of the g*, acceleration due to gravity
earth providing an equatorward component geff and the earth’s shape.
GOVERNING LAWS OF ATMOSPHERIC MOTION u 161

(Figure 6.6). However, on such an equipotential surface, the effective gravity must be normal at
every point on the earth. For this to happen, the surface of the earth must change from a
perfect sphere to an oblate spheroid as shown in Figure 6.6. Our present earth has an
equatorial radius of 21 km larger than the polar radius confirming the oblate spheroid shape.
The angle between the gravitational force and the effective gravity is quite small reaching a
maximum value of 0.1° at 45° latitude and decreasing to zero at both equator and poles. Since
the centrifugal force vanishes at the poles and is a maximum at equator, the effective gravity
has a maximum at poles and a minimum value at equator.

6.3 TOTAL, LOCAL AND CONVECTIVE DERIVATIVES


In the earlier section, a mention was made of two different approaches, Eulerian and
Lagrangian to describe fluid motion. Consider a fluid element, which contains a very large
number of molecules and moves as an entity. Let us assume that one can ascribe properties
such as density etc. to this fluid element. If measurements are made at fixed locations in space
then the resulting description of fluid motion is Eulerian. However, if one follows the motion
of a simple fluid element in time, the resulting description of fluid motion is Lagrangian.
Consider the kinematics of the fluid element by considering the trajectory of its motion
between times 0 and t by ignoring any distortions to its shape. The centre of mass velocity at
time t is the Eulerian velocity v evaluated at the current position r(t) of the fluid element.
This velocity is also the current rate of change of time of the fluid element’s position. Hence
ES
W <S U
U > (6.36)
EU
The acceleration of the fluid element at time t, a[r(t), t] also equals the second derivative of
the position vector, i.e.
E S
B <S U
U > (6.37)
EU 
Let us try to relate the description of both the approaches. Let r = (x, y, z) and v = (u, v, w).
The x component of Eq. (6.36) is
EY
V< Y U
Z U
[ U
U > (6.38)
EU
Differentiating Eq. (6.38) with respect to time and using the chain rule, one gets

EY ˜V EY ˜V EZ ˜V E[ ˜V
 = ˜Y EU
  
EU ˜Z EU ˜[ EU ˜U

˜V ˜V ˜V ˜V
= V W X
˜U ˜Y ˜Z ˜[
˜V EV
=  W  ³V (6.39)
˜U EU
162 u BASICS OF ATMOSPHERIC SCIENCE

E S EW ˜W
In vector form, B  W  ³
W (6.40)
EU  EU ˜U
From Eqs. (6.39) and (6.40) one obtains an important operator relationship
E ˜
 W  ³
(6.41)
EU ˜U
E
in Eq. (6.41) is known as the total or material derivative and represents the rate of
EU
˜
change with respect to time following the motion (or following a fluid element). in
˜U
Eq. (6.41) is known as the local derivative and it represents the rate of change with respect
to time at a fixed point. The term (v × Ñ) in Eq. (6.41) is known as the convective derivative.
Notice that the convective derivative term (v × Ñ)v in the velocity acceleration Eq. (6.40) is
nonlinear in v and is responsible for the difficulties in the numerical forecast of the
atmosphere. The concept of total, local and convective derivatives, although applied with the
velocity in Eq. (6.40) can also be applied to scalars as well.

6.4 CONTINUITY EQUATION

6.4.1 Eulerian Approach


The continuity equation is the mathematical statement of the conservation of mass and this
section will derive the above equation. Consider an infinitesimal parallelepiped of fluid fixed
in space whose sides are dx, dy and dz as in Figure 6.7. The idea is to compute the net rate at
which mass enters the above parallelepiped and relate this to the rate of change of mass with
time contained in the parallelepiped. In an interval of time dt, the amount of mass transported
into the parallelepiped by the fluid flow in the x direction through the face ABCD will be
(r u) dt dy dz (6.42)
During the same interval of time, the gain in mass by the fluid flow through the face EFGH

C
G

H
D

Ê ∂( ru ) ˆ
ru - Á ru + dx ˜
Ë ∂x ¯
B F
dz
dy
A dx E
FIGURE 6.7 Mass inflow of fluid into a fixed Eulerian control volume, due to motion parallell to x-axis.
GOVERNING LAWS OF ATMOSPHERIC MOTION u 163

is obtained by expanding r u as a Taylor series expansion in x and retaining only up to the


linear terms and multiplying that with dt dy dz, to get

Ë ˜ SV
Û
 Ì SV  EY Ü EU EZ E[ (6.43)
Í ˜Y Ý
The negative sign of Eq.(6.43) indicates that fact that a positive value of u signifies a
decrease in mass. The net gain of mass due to the fluid motion along the x-axis will be the
sum of Eqs (6.42) and (6.43) and is
˜ S V

 EU EY EZ E[ (6.44)
˜Y
By similar arguments, the net gain of mass due to the fluid motion along the y- and z-axis are
˜ S W
˜ S X

 EU EY EZ E[ BOE  EU EY EZ E[ (6.45)
˜Z ˜[
The increase in the rate of change with time of the mass contained within the parallelepiped
during this time, is given by
È ˜S Ø
ÉÊ ÙÚ EU EY EZ E[ (6.46)
˜U
Since the parallelepiped is fixed in its location, the local derivative is being utilized in
Eq. (6.46). Equating, Eq. (6.46) to Eqs. (6.44) and (6.45), one gets

˜S Ë ˜ S V
˜ S W
˜ S X
Û
Ì   Ü (6.47)
˜U Í ˜Y ˜Z ˜[ Ý
In vector form, Eq. (6.47) is written as
˜S
 ³ S W
 (6.48)
˜U
Expanding Ñ(rv) using vector identities, Eq. (6.48) is rewritten as
˜S
 W¹³S  S³¹W  (6.49)
˜U
The first two terms in Eq. (6.49) can be combined to get the total derivative and so
ES
 S³¹ W  (6.50)
EU
Equations. (6.48) and (6.50) are both statements of conservation of mass and are known as
the mass divergence and the velocity divergence form of the continuity equation. It is possible
to derive Eq. (6.50) from the framework of the Lagrangian approach and will be discussed in
the next section.
164 u BASICS OF ATMOSPHERIC SCIENCE

6.4.2 Lagrangian Approach


In the Lagrangian approach to derive the continuity equation, there is a need to focus
attention on a moving infinitesimal parallelepiped of fluid element instead of considering a
fixed parallelepiped in space as in the earlier section. It is assumed that the fluid element has
a fixed mass dm and a volume d V = d x d y d z which can change with time. For simplicity it
is assumed that the above fluid element is constrained to move in an one-dimensional flow
along the x direction with u(x) being its velocity along the x direction (Figure 6.8). Hence the
focus of our attention is a section of the parallelepiped which can be considered as a
rectangular section of the fluid element. The flow is assumed to vary in the x direction only.
Let the rectangular section of the fluid element be represented by ABCD at time t when its
left and right sides AD and BC are at positions x and x + d x, respectively. Assume that
within a small time interval d t, the side AD moves from the position x to a position x1
(indicated by the side EH in Figure 6.8) and the right side BC moves from a position x + d x
to a position x2 (indicated by the side GH in Figure 6.8). The positions x1 and x2 are related
to x and x + d x as follows:
x1 = x + u(x) d t (6.51)
x2 = x + d x + u (x + d x)d t (6.52)

D C
x x + dx

t dy

dx B
A

u (x + dx)
u (x)

E F

t + dt

H G
FIGURE 6.8 A rectangular fluid element, ABCD at time t becomes the fluid element EFGH at
time t + d t in a fluid flow moving parallel to the x-axis and varying in the x direction.
Unlike a fixed parallelepiped in the Eulerian description, a box (or section of the
box) does move in the Lagrangian description.

The rectangular section of the fluid element has its length elongated in the x direction by an
amount x2 – x1 which equals
[u(x + d x) – u(x)] d t (6.53)
Since the velocity component is restricted to the x direction only, the new volume of the
parallelepiped is
GOVERNING LAWS OF ATMOSPHERIC MOTION u 165

dV¢ = {d x + [u(x + d x) – u(x)] d t} d y dz (6.54)


where the length scales in the y and z directions remain unchanged.
Equation (6.54) can be approximated using a Taylor series expansion as follows:

È ˜V Ø È ˜V Ø
E7 „  É E Y  E Y E U Ù E Z E [ = E7 É  EUÙ (6.55)
Ê ˜Y Ú Ê ˜Y Ú

E Ë E 7 „  E7 Û ˜V
Hence, E7
= E7
(6.56)
EU
MJN
EU   ÌÍ EU ÜÝ
˜Y
Allowing the flow to be three-dimensional and varying in all three directions, v = v (r, t),
Eq. (6.56) can be rewritten the three-dimensional case as

E È ˜V ˜W ˜X Ø
E7
E 7   E7
³ ¹ W (6.57)
EU É
Ê ˜Y ˜Z ˜[ ÙÚ

EN
Since E7 BOE E N is a constant (conserved quantity) and hence can be cancelled on
S
both sides of Eq. (6.57) giving
E È Ø 
ÉÊ S ÙÚ ³¹W (6.58)
EU S
which reduces to the velocity divergence form of the continuity equation
ES
 S³ ¹ W (6.59)
EU

6.5 EQUATIONS OF MOTION AND EQUATIONS FOR


HORIZONTAL FLOW
The equations of motion are mathematical equations which state the physical conservation
principles such as conservation of momentum, conservation of mass, conservation of energy.
These equations of motion can be integrated numerically from the knowledge of the initial
state of the atmosphere to predict the future state of the atmosphere. The earth’s atmosphere is
a thin sheet of air with the vertical extent very much smaller compared to the horizontal. The
above indicates that the atmospheric flow is quasi-horizontal and hence it is important to
derive the equations of motion for horizontal flow. The following subsections attempt to
derive the equations of motion in a convenient spherical coordinates as well as derive the
equations of motion for horizontal flow.

6.5.1 Equations of Motion in Spherical Coordinates


It is convenient to express the equation of motion as applied to the earth in spherical
coordinates (r, j, l), since the earth is spherical. Here r is the distance from the centre of the
166 u BASICS OF ATMOSPHERIC SCIENCE

earth, j is the latitude and l is the longitude.


The triad unit vectors (i, j, k) can be drawn at North

any point in the earth surface with unit vector i y z


pointing eastwards, j pointing northwards and k
pointing vertically upwards as shown in Figure 6.9. x
Let us introduce small incremental East
distances in the east-west, north-south directions
as dx = r cos j dl and dy = r dj. Also, let the
vertical distance from the earth’s surface be z so
that r = z + a, where a is the radius of the earth
and dz = dr. The velocity components of the
relative velocity then become v = ui + vj + wk,
where the components are defined as
EM EK E[ Figure 6.9 Local Cartesian coordinate
V S DPT K W S X  axes defined at any arbitrary point
EU EU EU
on the spherical earth surface.
Then Eq. (6.26) with the effective gravity using
Eq. (6.33) can be written in the component form using the above coordinate system as

EV È V Ø  ˜Q
 É X  W TJO K  X DPT K
 ' Y

EU Ê Ù
S DPT K Ú S ˜Y = GSJD
(6.60)

EW XW È V Ø  ˜Q
  X  V TJO K 
S ˜Z = 'GSJD
Z

(6.61)
EU S ÉÊ S DPT K ÙÚ

EX V  W   ˜Q
  X V DPT K  H = ' [

(6.62)
EU S S ˜[ FGG GSJD

E ˜ ˜ ˜ ˜
where V W X (6.63)
EU ˜U ˜Y ˜Z ˜[

The terms proportional to are curvature terms and are due to the curvature of the earth.
S
The above equations are complex and difficult to handle. However, simpler approximations of
the above equations are adequate to model many important atmospheric motion. Since the
radius of the earth (a) is 6400 km, and our region of interest in the atmosphere is within 100 km,
V
we can replace the distance r by the earth’s radius a. In Eq. (6.60), the term,  X
B DPT K
since 2w ~ 10–4 rad s–1, and a ~ 6.4 ´ 109 m, and the zonal wind, u << 100 m s–1, in
magnitude. Also the vertical velocity is very much less as compared to the horizontal velocity
and so | w cos j | < | v sin j | .
GOVERNING LAWS OF ATMOSPHERIC MOTION u 167

È V  W  Ø
Also, in Eq. (6.62) the terms É Ù and 2w u cos j are very much smaller as
Ê B Ú
compared to geff. The vertical frictional force term 'GSJD
[

is considered to have very small


values and hence can be neglected. Using the above approximations, one gets

EV  ˜Q
 GW  = ' Y

(6.64)
EU S ˜Y GSJD

EW  ˜Q
EU
 GV 
S ˜Z
= ' Z

GSJD
(6.65)

EX  ˜Q
  HFGG = 0 (6.66)
EU S ˜[

6.5.2 Scale Analysis of the Equations of Motion


One can further simplify the equations of motion by resorting to the method of scale analysis.
Consider a typical synoptic scale motion in the mid-latitude region. Table 6.1 provides the
characteristic scales of meteorological variables based on the observed values associated with
mid-latitude synoptic systems.
With the help of the characteristic scales of Table 6.1, it must be possible to estimate the
magnitude of each term in Eqs. (6.64) and (6.65) with the Coriolis parameter f = 2w sin j0 =
10–4 rad s–1 for j0 = 45°
From Eq. (6.64),
˜V 6 ˜V 6  ˜V 6  ˜V 86
   V    W    X    GW   
˜U 5 ˜Y - ˜Z - ˜[ %
TABLE 6.1 The characteristic scales of meteorological variables based on the observed
values associated with mid-latitude synoptic systems
Symbol Meteorological variable Characteristic scale
U Horizontal velocity scale 10 m s–1
W Vertical velocity scale 1 cm s–1
L Length scale 106 m
D Depth scale 104 m
dp/r Horizontal pressure fluctuation scale 103 m2 s–2
LU–1 Time scale 105 s

Also 'GSJD
Y

is considered negligible. Hence in Eq. (6.64) the Coriolis term fv has the largest
magnitude and so must be balanced by the remaining pressure gradient force term
 ˜Q
GW (6.67)
S ˜Y
168 u BASICS OF ATMOSPHERIC SCIENCE

A scale analysis of Eq. (6.65) using similar arguments yields


 ˜Q
 GV
(6.68)
S ˜Z
Equations (6.67) and (6.68) together provide the so-called geostrophic approximation or
geostrophic balance in which the horizontal pressure gradient force is exactly balanced by
the Coriolis force. A scale analysis of Eq. (6.66) gives

˜X 8 ˜X 68 ˜X ˜X 8 
   V    W   X   
˜U 5 ˜Y - ˜Z ˜[ %
EX
All the acceleration terms in are much less than geff and hence geff is balanced by the
EU
only remaining pressure gradient force term
˜Q
 HS (6.69)
˜[
which is the hydrostatic balance equation.

6.5.3 f-plane and b-plane Approximations


Equations (6.60)–(6.62) provide the governing equations of motion in spherical coordinates.
There is a possibility of a useful simplification for situations, which involve only small
regions in the vicinity of a point P having latitude j0 and longitude l0. For the above
situation involving small regions around point P, one can define a Cartesian coordinates (x,y,z)
on the tangent plane at the point P. Since, in the Eqs. (6.60)–(6.62) dx and dy were
considered as eastward and northward distances (similar to the x and y Cartesian coordinates)
it is quite possible to consider Eqs. (6.60)–(6.62) as being applied to the Cartesian coordinates
on the tangent plane defined at the point P. The above tangent plane coordinates can handle
complications due to the spherical nature of the earth. However, the Coriolis parameter
( f = 2 w sin j) has to be properly taken into account in the tangent plane coordinates. If the
meridional extent of the region of our interest is not large, one can replace the Coriolis
parameter f by the constant value
f0 = 2w sin j0 (6.70)
Equation (6.70) is known as the f-plane approximation. However, if our region of interest is
larger, then instead of taking a constant f value it is a better approximation to allow some
variation of f with latitude and expand f with latitude and retain only the linear term
f (j) = 2w sin j = 2w {sin j0 + (j – j0) cos j0 + …} (6.71)
Z
with the result, on the tangent plane, where K  K   where a is the radius of earth, and
B

X DPT K  È EG Ø
G Z
 G Z
 C Z G   C Z XIFSF C (6.72)
B ÊÉ EZ ÚÙ Z 
GOVERNING LAWS OF ATMOSPHERIC MOTION u 169

Equation (6.71) is known as the b-plane approximation. Unlike, the f-plane approximation
where the Coriolis parameter ( f ) is considered a constant, the b-plane approximation allows
the Coriolis parameter ( f ) to vary linearly with latitude.

6.5.4 Geostrophic Wind


The geostrophic balance, Eqs. (6.67) and (6.68), is a diagnostic expression providing the
approximate relationship between the pressure field and the horizontal velocity valid for
synoptic scale systems. The geostrophic balance cannot be used to predict the evolution of the
velocity field since the above geostrophic balance does not involve time explicitly. It is
possible to define a horizontal wind velocity Vg = iug + jvg, called the geostrophic wind and
Eqs. (6.67) and (6.68) can be written in a vector form as follows:

7H L – ³Q (6.73)
SG
Figure 6.10 indicates schematically the geostrophic balance.
Let the isobars have values p + Dp and p
and be oriented in a parallel direction as shown
in the figure. The pressure gradient force
È  Ø
ÉÊ  S ³QÙÚ will act to accelerate the fluid from Vg
Fco

the region of high pressure towards the region p + Dp


of low pressure. Before the fluid starts moving,
the Coriolis force is zero as there is no relative —p
motion with respect to the rotating frame of
reference. Once the fluid starts moving towards
the region of low pressure, the Coriolis force p
will manifest and will act to the right of the
direction of the fluid flow in the northern FIGURE 6.10 Balance of forces between
the Coriolis and the
hemisphere and will begin to deflect the fluid
pressure gradient force for
away from moving in a direction perpendicular geostrophic equilibrium in
to the isobars. With increase in the velocity, the Northern hemisphere.
Coriolis force also increases until an exact
balance between it and the pressure gradient force ensures that the fluid moves in a direction
parallel to the isobars with lower pressure to the left and higher pressure to the right in the
northern hemisphere. The geostrophic balance does not generally hold over the equator. The
geostrophic approximation is a very useful diagnostic relationship which can help one obtain
the horizontal velocity if the pressure field is known. Hence it is clear that the geostrophic
approximation alone does not contain enough information to determine dynamically the fluid
motion. For the above purposes there is a need to incorporate small deviations from
geostrophy and an example of this is known as the isallobaric wind, which will be discussed
in the next sub section.
170 u BASICS OF ATMOSPHERIC SCIENCE

6.5.5 Isallobaric Wind


In order to allow deviation from geostrophy to determine fluid motion, we need to retain the
acceleration term in the horizontal equation of motion and neglecting frictional terms again,
one gets
E7 
 G L – 7  ³Q (6.74)
EU S
where the wind velocity is made up of the geostrophic and ageostrophic parts
V = Vg + Va (6.75)
Substituting Eq. (6.75) in Eq. (6.74) and using Eq. (6.73), one gets
E7
 G L – 7B (6.76)
EU
Equation (6.76) indicates that the ageostrophic component is always perpendicular and to the
left of the acceleration in the Northern hemisphere. Solving Eq. (6.76) for Va, one gets

7B 
L – E7 (6.77)
G EU

Substituting the decomposition in Eq. (6.77), we get

L – E7B
 ˘ È  ØÛ 
7B  Ì É ³Q Ù Ü  (6.78)
G Í ˜U Ê S G ÚÝ G EU
As a first approximation, we can neglect the second term in the right-hand side of Eq. (6.78)
and can write the deviation from the geostrophic wind in component form as
 ˜ Q
u – ug =  (6.79)
S G  ˜U ˜Y
 ˜ Q
v – vg =  (6.80)
QG  ˜U ˜Z
The departure from the geostrophic components of the wind depends on the time rate of
change of the pressure gradients, also known as the tendency. The isolines of tendency are
known as isallobars. The isallobaric wind is normal to the isobars and oriented towards the
isallobaric minimum. The combined total wind can be regarded as made up of both the
geostrophic component as well as the isallobaric component.

6.5.6 Natural Coordinate System


The geostrophic flow required that the pressure to decrease to the left of the motion in the
northern hemisphere and the flow to be along the isobars. The geostrophic flow can be put in
a very simple form in the so-called natural coordinate system. Since the directions of
motion of the fluid and the norm to it are the two primary orientations, the natural coordinate
system uses the above orientation with the t-axis tangent to the flow along which t increases
GOVERNING LAWS OF ATMOSPHERIC MOTION u 171

in the direction of flow and a normal n-axis perpendicular to the flow along which n increases
to the left of the direction of motion (Figure 6.11). In terms of the natural coordinate system
the geostrophic wind speed Vg becomes

t
s
n

FIGURE 6.11 Natural coordinates with t and n along and normal to the motion.

 ˜Q
7H  (6.81)
S G ˜O
The horizontal velocity vector may be written as V = Vt where V is the horizontal wind speed.
The acceleration in this reference system following the motion is
E7 E EU E7
7U
7 U (6.82)
EU EU EU EU
From geometrical considerations, the rate of change of t following the motion (Figure 6.12) is
ET
EZ EU (6.83)
3
where R is the radius of curvature following the motion of the fluid element. R is considered
positive when the centre of curvature is in the positive n direction (i.e. for a fluid element
turning to the left, R > 0 and vice versa). Since d t is directed parallel to n, in the limit d t ® 0,
EU O
= (6.84)
ET 3
EU EU ET O
Hence, = 7 (6.85)
EU ET EU 3
E7 E7 O7 
= U  (6.86)
EU EU 3
The Coriolis force can be written as f k ´ V = f V n (6.87)
172 u BASICS OF ATMOSPHERIC SCIENCE

dt

t + dt
dy

ds
t
dy

R
n

FIGURE 6.12 Dependence of unit vector t along the path.

The horizontal momentum equations in the directions t and n are


E7  ˜Q
=  (6.88)
EU S ˜T
7  ˜Q
 G7 =  (6.89)
3 S ˜O
Equations (6.88) and (6.89) can be regarded as providing the balance between the pressure
È7  Ø
gradient force, Coriolis force and the centrifugal force. If the centrifugal force É Ù is
Ê 3Ú
disregarded in Eq. (6.89), one obtains the geostrophic wind, the same as Eq. (6.81).

6.5.7 Inertial Flow


If the pressure gradient is horizontally uniform, the pressure gradient force identically
vanishes and Eq. (6.89) provides a balance between the Coriolis and the centrifugal forces

7
 G7 = 0 (6.90)
3
7
giving R=  (6.91)
G
For uniform pressure field, Eq. (6.88) requires constant horizontal velocity field and hence the
radius of curvature R has to be constant (neglecting the variation of f with latitude) in Eq. (6.91).
That is, fluid elements follow circular path in an anticyclonic sense (the sense of rotation is
opposite to the sense of rotation of the earth). In the northern hemisphere ( f > 0 and R < 0),
flow turns to the right, hence clockwise motion, while in the southern hemisphere ( f < 0 and
GOVERNING LAWS OF ATMOSPHERIC MOTION u 173

R > 0), flow turns to the left, hence anticlockwise motion. The time period of the circular
motion is
Q 3 Q Q
5  QFOEVMVN EBZ (6.92)
7 ] G ] ]X TJO K ]
since one pendulum day which is defined as the time required for the plane of a freely
suspended (Foucault) pendulum to complete an apparent rotation about the local vertical is
Q
given by ]X TJO K ] where w is the angular speed of the rotating earth and j is the latitude.
The above motion resulting from a balance between the Coriolis and the centrifugal forces is
known as inertial flow, the circle of radius |R| is called an inertial circle and T is known as
the inertial period.

6.5.8 Cyclostrophic Flow


For motion with large Rossby number, the Coriolis force can be neglected and Eq. (6.89)
provides a balance between the pressure gradient and the centrifugal force known as the
cyclostrophic balance
7  ˜Q
 (6.93)
3 S ˜O
Solving the above Eq. (6.93) for V

È 3 ˜Q Ø 
7D É (6.94)
Ê S ˜O ÙÚ
The cyclostrophic flow may be either cyclonic (refer Figure 6.13(a) or anticyclonic (refer
Figure 6.13(b). In general, cyclostrophic equation is valid for all cases where the curvature
radius is small and for large Rossby number. Examples of cyclostrophic flows are tornados,
waterspouts and dust devils.

Ce P Ce
P
L L

∂p ∂p
(a) R > 0, < 0, (b) R < 0, > 0,
∂n ∂n
FIGURES 6.13 Various cases of cyclostrophic flow: (a) cyclonic flow in northern hemisphere,
and (b) anticyclonic flow in northern hemisphere. L at the centre indicates ‘low-
pressure’, P and Ce refer to pressure gradient force and centrifugal force,
respectively while V indicates the velocity.
174 u BASICS OF ATMOSPHERIC SCIENCE

6.5.9 Gradient Flow


For motion with a three-way balance between the pressure gradient force, Coriolis force and
the centrifugal force the resulting balance is known as gradient balance and Eq. (6.89)
describes the gradient balance
7  ˜Q
 G7  (6.95)
3 S ˜O
The gradient wind speed Vgr as obtained from Eq. (6.95) is

È  
G3 G 3 3 ˜Q Ø 
 “É  (6.96)
S ˜O ÙÚ
7HS
 Ê 
It is to be noted that not all the mathematically possible roots mentioned above (or) possible
roots of Eq. (6.95) correspond to physically realizable observed flows.
First, let us consider the positive root of the square root term in Eq. (6.96). This root
describes the observed cases of cyclonic flow around a low-pressure centre (Figure 6.14(a))
and the anticyclonic flow around a high-pressure centre (Figure 6.14(b)). For Figure 6.14(a),
˜Q
R > 0 (centre of curvature is in the positive direction of n) and since   (a low), the
˜O
G3
square root term exceeds in magnitude and Vgr is found by considering ‘+’ of the square

˜Q G3
root. For Figure (6.14(b)), R < 0 and   (a high), the square root term is less than in
˜O 
magnitude and Vgr is found by considering the ‘+’ of the square root.
Consider next the negative roots of the square root term in Eq. (6.96).
˜Q
As approaches zero in Eq. (6.96), Vgr ® – f R. This represents anticyclonic flow (refer
˜O
˜Q
Figure 6.14(c)). For Figure 6.14(c), !  3   and both the terms in Eq. (6.96) will still
˜O
be negative and the motion will be anticyclonic. This anticyclonic rotation about a low has
˜Q
not been observed in large-scale atmospheric motion. For, Figure 6.14(d),   3   In
˜O
this case, the gradient wind speed is high because the two terms of Eq. (6.96) complement
one another by having the same sign (unlike the case of Figure 6.14(b), where the two terms
of Eq. (6.96) were of opposite sign) and the speed was less. The cases of Figure 6.14(c) and
(d) are termed anomalous cases, as they are not normally observed. The anomalous cases
(negative root cases) are not observed since they correspond to high wind speeds and very
high kinetic energy values are required for these kinds of flows.
Consider the gradient wind balance Eq. (6.95) for the positive root. For the flow around
a low, the square root term can never become imaginary so that all values of pressure gradient
are allowed and hence there are no restrictions on the magnitude of the pressure gradient for
a low. However, for a flow around a high, there is a possibility for the square root term to
GOVERNING LAWS OF ATMOSPHERIC MOTION u 175

Ce
P
C0 P
C0
Ce
L H

V
(a) (b)

P
Ce C0 P

L H
C0
Ce

V V
(c) (d)

FIGURE 6.14 Various cases of gradient balance in northern hemisphere (a) and (c) flow around a
low pressure and (b) and (d), flow around a high pressure. L and H indicate ‘low-
pressure’ and ‘high-pressure’, C0, Ce and P denote the Coriolis force, centrifugal
force and pressure gradient force and V denotes the velocity.

become imaginary and such cases need to be excluded from the discussion since Vgr is real.
The condition that Vgr shall be real is
˜Q S ] 3 ] G
… (6.97)
˜O 

That is for a flow around a high, the magnitude of the pressure gradient should not exceed a
certain value determined by the latitude and the distance from the centre. Due to the above
restriction, the observed pressure gradient is very small near the centre of a high. One can
reverse and restate the above restriction for a high and conclude that for a given pressure
gradient, the radius of curvature of the flow must not fall below a minimum value given by the
inequality Eq. (6.97).
One can rewrite the gradient wind balance Eq. (6.95) in terms of the geostrophic wind
7HS
 G7HS G7H (6.98)
3
Hence the ratio of the geostrophic wind to the gradient wind is
7H 7HS
 (6.99)
7HS G3
176 u BASICS OF ATMOSPHERIC SCIENCE

For normal cyclonic flow ( f R > 0), Vg is greater than Vgr, while for anticyclonic flow ( f R < 0)
Vg is less than Vgr. Hence the geostrophic wind is an overestimate of the balanced wind over
a region of cyclonic circulation and an underestimate over a region of anticyclonic circulation.

6.6 THERMAL WIND


It can be shown from physical arguments based on the hydrostatic balance, that the
geostrophic wind will have a variation in the vertical in the presence of a horizontal
temperature gradient. Using the hydrostatic approximation and the equation of state, a
relationship between the height increment äz corresponding to a given pressure increment dp
is given by
EQ HQ E Q 3TQ5W
 HS  E[   (6.100)
E[ 3TQ5W Q H
where Tv is the averaged virtual temperature of the layer. From Eq. (6.100) it is clear that for
a given pressure increment, the thickness of a layer (dz) is directly related to the average
virtual temperature of the layer. For a given pressure increment, a greater thickness must be
associated with a warm layer.
From the geostrophic wind Eq. (6.67) and the equation of state, one gets

 ˜Q 3TQ5W ˜Q ˜ MO Q

GW 3TQ5W (6.101)
S ˜Y Q ˜Y ˜Y
From the hydrostatic equation and the equation of state
˜Q HQ H  ˜Q ˜ MO Q

 HS   (6.102)
˜[ 3TQ5W 3TQ5W Q ˜[ ˜[
Neglecting vertical variation (but not horizontal) in the virtual temperature (Tv) (vertical
variation in Tv need not be neglected if pressure is used as the vertical coordinate instead
of z), differentiate Eq. (6.101) with respect to z and differentiate Eq. (6.102) with respect to x,
one gets the following equations:
˜W ˜  MO Q

G = 3TQ5W (6.103)
˜[ ˜[ ˜Y

H È  ˜5W Ø ˜ MO Q

 É   Ù = ˜Y˜[ (6.104)
3TQ Ê 5W ˜Y Ú

Equating Eqs. (6.103) and (6.104), one gets


G ˜W H ˜5W
= PS
3TQ5W ˜[ 3TQ5W  ˜Y

˜W H ˜5W
G = (6.105)
˜[ 5W ˜Y
GOVERNING LAWS OF ATMOSPHERIC MOTION u 177

Continuing with similar analysis using Eq. (6.68) and the equation of state
 ˜Q 3TQ5W ˜Q ˜ MO Q

 GV 3TQ5W (6.106)
S ˜Z Q ˜Z ˜Z
Differentiating Eq. (6.106) with respect to z and again neglecting the vertical variation of Tv
and differentiating Eq. (6.102) with respect to y, one gets
˜V ˜  MO Q

G = 3TQ5W (6.107)
˜[ ˜[ ˜Z

H È  ˜5W Ø ˜ MO Q

 É   Ù = (6.108)
3TQ Ê 5W ˜Z Ú ˜ Z ˜[

Equations (6.107) and (6.108) can be equated to get


˜V H ˜5W
G  (6.109)
˜[ 5W ˜Z
Equations (6.105) and (6.109) are called as the thermal wind equations. They provide a useful
relationship between the horizontal temperature gradients and the vertical gradients of the
horizontal wind whenever geostrophic and hydrostatic balances are valid. As illustrations of
thermal wind, consider a situation where the temperature varies only in the north-south
direction and is constant along the east-west direction (Figure 6.15). From Eq. (6.105) the
meridional wind component has no vertical shear, while from Eq. (6.109) the zonal
component has a positive vertical wind shear. For this case the difference between the wind
vectors at any two vertical levels is parallel to the x-axis and is also parallel to the isotherms.
This is an interesting and important result. The thermal wind is the vertical shear of the
geostrophic wind and as shown above, the direction of the thermal wind is parallel to

FIGURE 6.15 The thermal wind balance. Relationship between the vertical
wind shear and the horizontal temperature gradients.
178 u BASICS OF ATMOSPHERIC SCIENCE

isotherms with the lower temperature to the left and higher temperature to the right in
northern hemisphere. From the knowledge of the behaviour of the winds with altitude, the
thermal wind relationship can be used to predict the temperature advection. It can be surmised
that if the wind changes in the anticlockwise direction with height (also called as backing),
the wind would blow across the isotherms from the cold to the warm side contributing to cold
air advection. Alternatively, if the wind changes in the clockwise direction with height (also
called as veering), the wind contributes to warm air advection.

6.7 THERMODYNAMIC ENERGY EQUATION


The thermodynamic energy equation is simply the statement of the first law of
thermodynamics after taking into account the motion of the atmosphere. The first law of
thermodynamics is written in the form
dq = cp dT – a dp (6.110)
Dividing Eq. (6.110) by dt and taking to the limit, one gets
ER E5
DQ  BX (6.111)
EU EU
EQ ER
where w is the vertical velocity, X  It is to be noted that
EU
is not a total derivative
EU
since Dq is not a perfect differential. The total derivative of temperature in the coordinate
system using the pressure as the vertical coordinate can be written as
E5 ˜5 ˜5
 7) ¹ ³ Q5  X (6.112)
EU ˜U ˜Q
where VH indicates the horizontal wind and Ñp indicates gradient on a pressure surface.
Substituting Eq. (6.112) in Eq. (6.111), one gets

˜5 È B ˜5 Ø  ER
 7) ¹ ³ Q5  X É  Ù (6.113)
˜U Ê D Q ˜Q Ú D Q EU
Equation (6.113) shows that the local temperature change is contributed due to heat advection
(first term in the right-hand side), the vertical motion (second term) and the heat input to the
XB
system (third term). For w < 0, upward motion, the term is the work due to expansion
DQ

term contributing to the decrease in temperature and X ˜5 is the vertical temperature


˜Q

advection term. If X ˜5
  and temperature decreases with height, the result is local
˜Q
warming or increase in temperature. The last term in Eq. (6.113) is known as the diabatic term
and takes into account the energy exchanges due to non-adiabatic processes such as absorption
and emission of radiation as well as latent heat due to condensation and evaporation of water
GOVERNING LAWS OF ATMOSPHERIC MOTION u 179

vapour. Taking the logarithmic derivative of the Poisson equation with respect to pressure and
rearranging terms, one gets
 ˜5  ˜R 3TQ
= 
5 ˜Q R ˜Q D Q Q
˜5 5 ˜R B
= 
˜Q R ˜Q D Q
Substituting the above, one gets
˜5  ER
 7) ¹ ³ Q5  X 4 Q  (6.114)
˜U D Q EU
˜ MO R 3TQ B ˜R
where 4 Q 5 where Sp is related to a static parameter T 4Q  
˜Q Q R ˜Q
Equation (6.114) is a form of the thermodynamic energy equation. An alternate form of the
thermodynamic energy equation can be got from the first law of thermodynamics of the form,
dq = cvdT + pda (6.115)
Dividing Eq. (6.115) by DT and taking to the limit, one gets
ER E5 EB
R DW Q (6.116)
EU EU EU
Using the equation of state, we get
EB EQ E5
Q B 3TQ (6.117)
EU EU EU
Using Eq. (6.117) in Eq. (6.116), one gets

 ER D Q E5 B EQ
 (6.118)
5 EU 5 EU 5 EU
The left-hand side of Eq. (6.118) is nothing but the entropy change
ET D Q E5 3TQ EQ
 (6.119)
EU 5 EU Q EU
Since s = cp ln q,
ET E MO R
E MO 5
E MO Q

DQ DQ  3TQ (6.120)
EU EU EU EU
Substituting Eq. (6.120) in Eq. (6.116) one gets another form of the thermodynamic energy
equation of the form
ET ER
5 R (6.121)
EU EU
ET R
or (6.122)
EU 5
180 u BASICS OF ATMOSPHERIC SCIENCE

SOLVED EXAMPLES

1. Let the surface pressure at station A be 1001.5 hPa, while the surface pressure at station B
100 km west of station A is 1000.5 hPa. Assuming air density to be 1.2 kg m–3, find the
pressure gradient force per unit mass. It is assumed that both stations have same elevation.
Solution: The pressure gradient force per unit mass
 Q"  Q#
  
– 
= 
S'Y  –  – 

= –8.33 ´ 10–4 m s–2.


2. Let the surface horizontal wind components at station A be 8 m s–1, 4 m s–1, while the same
for another station B lying 200 km south of station A has values of 5 m s–1, 5 ms–1).
Assuming that air density to be 1.2 kg m–3, find the advection force per unit mass.
Solution: The components of the advection force in the x and y directions are equal to
'6 '6
6 7
'Y 'Z
BOE

'7 '7
6 7
'Y 'Z
Dy = 200 ´ 103 m, while Dx = 0.
Average U = (8 + 5)/2 = 6.5, while average V = (4 + 5)/2 = 4.5
Components of advection force in x direction

'6 '6 '6  

6 7 7

    –   NT 
'Y 'Z 'Z  – 

Components of advection force in y direction


'7 '7 '7   

6 7 7    –   NT o
'Y 'Z 'Z  – 
3. Find the Coriolis force at a station at 30°N having a zonal wind of 10 m s–1.
Solution: Coriolis paramter at 30°N, f = 2 ´ 7.29 ´ 10–5 sin (30) = 7.29 ´ 10–5
Coriolis force per unit mass in the north-south direction = –f u
= – 7.29 ´ 10–5 ´ 10 = – 7.29 ´ 10–4 ms–2
Negative sign indicates force is from north to south.
4. If the height of the 500 hPa pressure surface decreases by 20 m northward across a distance
of 1000 km, what is the pressure gradient force?
Solution: The pressure gradient force per unit mass in the x and y directions are given by
'[
H BOE
'Y
'[
H
'Z
GOVERNING LAWS OF ATMOSPHERIC MOTION u 181

where Dz/Dx and Dz/Dy are the slopes of the isobaric surfaces and g = 9.8 m s–2.
Also, given Dz = –20 m, Dy = 1000 ´ 1000 m and Dx = 0.
The pressure gradient force per unit mass in the x and y directions are 0 and
'[  
H     –   NT 
'Z 
5. If the pressure increases by 2 hPa eastward over a distance of 400 km, at latitude of 30°N,
find the geostrophic wind assuming that air density is 1.2 kg m–3.
Solution: Using the following expressions for the geostrophic wind components, one gets
 'Q  'Q
VH  W
S G 'Z H S G 'Y
Also, given Dp = 200 Pa, Dx = 400 ´ 1000 m and Dy = 0.
Substituting, one gets
 'Q  
WH  N T 
S G 'Y < –  –  TJO 
>  

The geostrophic wind is towards north with a speed of 5.72 m s–1. Since Dy = 0, there is no
zonal component of the geostrophic wind.
6. If the height of a pressure surface at 900 hPa, at latitude 30 oN increases by 100 m eastward
across a distance of 1000 km, find the geostrophic wind at 900 hPa.
Solution: Using the following expressions for the geostrophic wind components, one gets
H '[ H '[
VH   WH
G 'Z G 'Y
Also, given Dz = 100 m, Dx = 106 m and Dy = 0. Since Dy = 0, there is no zonal component
of the geostrophic wind. The meridional component of the geostrophic wind is

H '[  


WH 
 N T 
G 'Y < –  –  TJO 
> 
The geostrophic wind is towards north with a speed of 13.44 m s–1.
7. If the geostrophic wind speed around a low-pressure region is 15 m s–1, find the gradient
wind speed given the Coriolis parameter is 10–4 s–1, and a radius of curvature of 400 km.
Solution: The gradient wind speed around a low-pressure region is given by

Ë 7H Û
Vgr =  G3 Ì      Ü
ÍÌ G3 ÝÜ
Ë  –  Û
=  –    – 
Ì        Ü
 N T 
ÍÌ   – 
ÝÜ
8. If the geostrophic wind speed around a high-pressure region is 10 m s–1, find the gradient
wind speed given the Coriolis parameter is 10–4 s–1, and a radius of curvature of 800 km.
182 u BASICS OF ATMOSPHERIC SCIENCE

Solution: The gradient wind speed around a high-pressure region is given by

Ë 7H Û
Vgr =  G3 Ì o  o Ü
ÌÍ G3 ÜÝ
Ë  –  Û
=  –    – 
Ì o  o   Ü
 N T 
ÍÌ   – 
ÝÜ
9. A 60 kg man steps onto a train moving towards west at a speed of 10 m s–1, along the
equator. Find the resulting change in his apparent weight.
Solution: The man’s weight before stepping onto the train is given by W = mg = m(g* –
w2RE), where g* is the force per unit mass of Newtonian gravitational attraction, m is the
mass, g is the force per unit mass of the effective gravity, w is the angular velocity of the
rotation of earth and RE is the radius of the earth. After the man steps onto the train, his
apparent weight becomes
\X 3&  V
 Û
Ë
8 „ NÌH  Ü
ÌÍ 3& ÜÝ
where u is the speed of the train. Since the train is west-bound, a direction that is opposite
to the direction of the earth’s rotation, a negative sign appears before u. The apparent
change in the man’s weight is then given by W¢ – W = m [2 w u – u2/RE]. Substituting for
the mass from the first equation, one gets the apparent change in the man’s weight is given by

=
8 <X V  V 3& >
H
< –  –   –   

  – 
>
=  –   LH

The man’s weight has increased by 8.83 g.
10. Repeat the above problem except that the train is bound eastward with the same speed?
Solution: The man’s weight before stepping onto the train is given by W = mg = m (g* – w2RE)
After the man steps onto the train, his apparent weight becomes

Ë \X 3&  V
 Û
8 „ N ÌH  Ü
ÍÌ 3& ÝÜ
where u is the speed of the train. Since the train is east-bound, a direction that is in the
direction of the earth’s rotation, a positive sign appears before u. The apparent change in the
man’s weight is then given by W ¢ – W = – m[2w u + u2/RE]. Substituting for the mass from
the first equation, one gets the apparent change in the man’s weight is given by

=
 8 <X V  V 3& >
H
< –  –   –   

  – 
>
=   –   LH

The man’s weight has decreased by 9.02 g.
GOVERNING LAWS OF ATMOSPHERIC MOTION u 183

11. At a certain station, the horizontal gradient of sea level pressure is 6 hPa per 100 km. Find
the slope of the 1000 hPa pressure surface at this station assuming that the air density is
1.25 kg m–3.
Solution: The pressure gradient force per unit mass is given by

 ˜Q   – 
QO   –   N T 
S ˜O   – 
The slope of the 1000 hPa pressure is related to the pressure gradient force per unit mass as

˜[ ˜[  ˜[ 
QO H JF   –    – 
˜O ˜O ˜O
That is the magnitude of the slope of the 1000 hPa surface is 49 m per 100 km.
12. A projectile of mass m initially at rest is dropped to the earth surface from a height h which
is small compared to the radius of earth. Assuming that the angular velocity of the earth
about its axis w is a constant and in the absence of friction prove that the projectile is
 
 È I Ø
deflected east of the vertical by the amount X H DPT G É Ù as it reaches ground where
 Ê HÚ
j is latitude.
Solution: The equations of motion of a particle is given by

EW
H   Z – W

EU
where g is the effective gravity. Since the Coriolis force is much smaller than g in
magnitude, one can use the method of successive approximations to solve the above
equation. In this method, one writes
v(t) = v1(t) + v2(t) such that

] W U
] !! ]W  U
] W  H W   W – Z
Integrating first of above equation, one gets v1 = g t + v0, where v0 is a constant of
integration. Substituting of this v1 in the second equation and integrating, one gets

v2 = (g ´ w) t2 + 2(v0 ´ w) t
Hence v(t) = v0 + (g + 2v0 ´ w) t + (g ´ w) t2
Integrating the above equation with respect to time, one gets

 
S U
H  W  – Z
U   H – Z
U 
S  W  U 
 
where r0 and v0 are the initial position and velocity of the projectile.
For a projectile dropped from height h with initial velocity zero,

v0 = 0, r0 = h k, g = –g k and w = (w cos j) j + (w sin j) k


Substituting these, one gets
184 u BASICS OF ATMOSPHERIC SCIENCE

 
   È I Ø
S U
IL  H U L  X H DPT K É Ù J
  Ê HÚ
 
The projective will land in the earth as soon as the condition I H U is satisfied and at


that moment, S U
X DPT K H U  J

 

Hence the amount of deflection is to the east and equals E X H DPT G ÈÉ I ØÙ
 Ê HÚ
13. Suppose the projectile with mass m is sent vertically up with velocity v0, reaches a height h
above the ground and returns to the ground. Find the amount of deviation of the projectile at
the ground.

Solution: In this case, W  W  L  HI L S  H  H L and w = (w cos j) j +


(w sin j) k. Substituting these, in the expression for r(t), one gets


S U

È
ÉÊ W  U  HU

ÙÚ L   HI U

X H DPT K J X H DPT K U

J
 

The coefficient of k in the above equation must vanish as the projectile returns to the
ground. Hence the deviation at the ground equals to
 W
S U
 X DPT K J
 H
Hence the deflection is to the west in both the hemispheres. The deflection in this case is
four times the deflection in the earlier problem.
14. Let the zonally averaged tropospheric temperature gradient over extra tropics be about 1 per
degree of latitude and let the mean longitudinally averaged zonal component of the
geostrophic wind at the ground be close to zero. Find the mean zonal wind at the jet stream
level of 200 hPa.
Solution: The longitudinally averaged zonal wind can be determined from
˜ >[@ H
<V H > 
G ˜Z
where ug is the zonal component of the geostrophic wind and the square bracket indicates
longitudinal averaging. Applying the above relation to 200 hPa and 1000 hPa, pressure
levels and subtracting, one gets
H ˜
<V H >  <VH > \< [ >  < [ > ^
G ˜Z
Using R = Rad = 287 J deg–1 kg–1, and using the hypsometric equation, one gets

3 ˜<5 >    

<VH >  MO   MO   N T 


G ˜Z    – 

15. Assume that a cold front has just passed a station A and the air temperature be 10°C and is
GOVERNING LAWS OF ATMOSPHERIC MOTION u 185

falling at a uniform rate of 2° h–1. Let the wind be a northerly having a speed of 50 km h–1.
At another station B located north of station A by 100 km, the temperature is 2°C. Find the
time rate of change of the temperature of the air parcel as it moves southward behind the
front. Assume that the meridional wind velocity and the temperature gradient are uniform in
the vicinity of stations A and B.
Solution: The equation governing the total rate of change of temperature will be of the
form
E5 ˜5 ˜5     LN   

W  – <    > ’I  ’ I  (warming)


EU ˜U ˜Z I I  LN
16. At a certain station A located at 40°N, the geostrophic wind at the 1000 hPa level is a south-
westerly (direction is 230°) with a speed of 12 m s–1, while at the 500 hPa level the wind is
a west-northwesterly (direction is 300°) with a speed of 28 m s–1. Find the geostrophic
temperature advection
Solution: The thermal wind is the vector difference between the geostrophic wind at 500 hPa
and the geostrophic wind at 1000 hPa and the thickness (or isotherms of the mean
temperature in the layer) contours are parallel to the thermal wind with lower temperatures
to the left in the northern hemisphere. From the thermal wind relationship,

3 Ë  Û @@
75 Ì MO Ü L – ³ 5

G Í  Ý
where 5 is the mean layer temperature. The absolute value of the temperature gradient is
 
 –   N T EFH
G
]³5 ] "75 where "   

3 MO

The geostrophic temperature advection equals AV500 V1000 sin q, where q is the angle
between the geostrophic wind vectors between 1000 hPa and 500 hPa. Given q = 70°, the
geostrophic temperature advection is 5.01 ´ 10–7 ´ 12 ´ 28 ´ sin (70°) = 1.58 ´ 10–4 deg s–1
= 13.65 deg yr–1.

REVIEW QUESTIONS

1. On a given day, let the wind be a westerly with a speed be 25 m s–1 at a given station at 500 hPa. 
(i) Assuming that the flow is geostrophic, find the magnitude of the pressure gradient force
responsible for producing the above-observed wind. (ii) Mention the directions of the
pressure gradient force and the Coriolis force necessary to produce the above-observed flow.
2. The flow within a tornado is due to the resultant balance of forces between the pressure
gradient force and the centrifugal force. If the wind speeds have magnitudes of about
90 m s–1 within a tornado that is 400 m in diameter, what must be the pressure change from
outside to the centre of the tornado, assuming that the air density is 1 kg m–3.
3. Can a hurricane cross the equator? Explain.
4. Assuming the hypothetical situation where the earth is not rotating, how would the winds
186 u BASICS OF ATMOSPHERIC SCIENCE

blow with respect to centres of high and low pressure? Is the direction of the winds in such
a situation same in both the hemispheres?
5. Consider wind blowing over a land surface that ultimately crosses a coastline and then blows
over a lake. How will the wind speed and direction change as the air moves from the land
surface to the lake surface?
6. Are the surface winds that blow over the ocean closer to being geostophic than those that
blow over the land? Why?
7. What is the inertial period over the North Pole?
7 Atmospheric Motion

The state of the atmosphere is characterized more often than not by its ceaseless motion. In
the earlier chapter, various forces which influence atmospheric motion have been discussed.
Due to the earth’s rotation, the overlying atmosphere also rotates. Vorticity and circulation,
measures of the rotation of the air, or more generally a fluid parcel are well known in the
study of fluid dynamics. Vorticity or rotation of the air can arise either due to the effects of
the shear or curvature. A convenient way of isolating the above effects of rotation is to
examine vorticity using the natural coordinate system.
Most meteorological upper air observations such as radiosondes/rawinsondes report
meteorological data at constant pressure levels and not at constant height levels. Also, the
equation of continuity assumes its simplest form when pressure is used as the vertical
coordinate system. Hence it is important to introduce pressure as the vertical coordinate
system, the isobaric coordinate system. Another important concept well known in the study of
fluid dynamics is the horizontal divergence. The convergence/divergence patterns of the
atmosphere at lower levels provide important information on the development/dissipation of
weather systems. Both the vorticity and the horizontal divergence can play the role of
prognostic variables in describing atmospheric motion and hence it is important to derive
equations to describe the evolution of vorticity and also for the horizontal divergence in the
atmosphere. These equations are known as vorticity and divergence equations.
In the earth’s atmosphere, a few variants of vorticity known as absolute vorticity and
potential vorticity find important applications. Section 7.1 presents the basic concepts of
circulation and vorticity as applied to the atmosphere. In Section 7.2, the isobaric coordinate
system is introduced and its advantages are mentioned. In Section 7.3, the vorticity and
divergence equations are derived, while Section 7.4 introduces the concepts of both the
absolute vorticity and the potential vorticity.

187
188 u BASICS OF ATMOSPHERIC SCIENCE

7.1 CIRCULATION AND VORTICITY


7.1.1 Vorticity
The equations, which govern the motion of the earth’s atmosphere, are complex and are not
amenable to direct closed form solutions. In order to simplify the problem of solving the
governing equations of motion, one introduces quantities which provide for more general
properties of the fluid. One such quantity is the so-called vorticity, well known in the study of
fluid dynamics. Vorticity (W), which is a measure of the rotation of the air, is simply defined
as the curl of the velocity field.
W=Ñ´V (7.1)
If the velocity field V in Eq. (7.1) is the relative velocity, then the vorticity W in Eq. (7.1) is
known as the relative vorticity. Relative velocity of an air parcel refers to the velocity of the
air parcel relative to the rotating earth. A fluid flow having zero vorticity is known as
irrotational flow. In irrotational flow, the fluid elements do not spin around, but maintain their
upright orientations as shown in Figure 7.1. The above clearly implies that for fluid flows
having non-zero vorticity, the fluid elements do spin around and do not maintain their upright
orientations.
G H

G H
G H
G H E F

E F
E F
E F

FIGURE 7.1 Fluid elements do not spin around, but maintain their upright orientations in fluid
flows with zero vorticity.

7.1.2 Decomposition of a Linear Velocity Field


It can be shown that every motion of a fluid particle at a point in space and time can be
divided into the following fundamental modes, such as rotation, translation and deformation.
To be more specific, it can be shown that a linear velocity field is made up of a linear
combination of four quantities, such as vorticity, divergence, deformation and translation. For
simplicity, and convenience we shall restrict our discussion to two-dimensional horizontal
motion as referred to a Cartesian coordinate system. Expanding the horizontal velocity
components u and v in a Taylor series expansion about the origin and restricting the
expansion up to the linear term, one gets

È ˜V Ø È ˜V Ø
u » V0  É Ù Y  É Ù Z (7.2)
Ê ˜Y Ú 0 Ê ˜Z Ú 0

È ˜W Ø È ˜W Ø
v » W0  É Ù Y  É Ù Z (7.3)
Ê ˜Y Ú 0 Ê ˜Z Ú 0
ATMOSPHERIC MOTION u 189

where the subscripts O refer to the quantities being evaluated at the origin. It is worthwhile to
examine whether any of the above derivatives or their combinations remain invariant under a
rotation of the coordinate axes. Let the initial coordinate system (x, y) be rotated through a
fixed angle q as in Figure 7.2 to a rotated coordinated system denoted by primed quantities
(x¢, y¢).

FIGURE 7.2 Two cartesian systems of coordinates differing by an angle q.

The two coordinate systems are then related by the following equations
x = x¢ cos q – y¢ sin q (7.4)
y = x¢ sin q + y¢ cos q (7.5)
Differentiating, Eqs. (7.4) and (7.5) with respect to time, one gets
u = u¢ cos q – v¢ sin q (7.6)
v = u¢ sin q + v¢ cos q (7.7)
For any function G,
˜( ˜ Y „ ˜ ( ˜ Z „ ˜( ˜( ˜(
=  DPT R  TJO R (7.8)
˜Y ˜Y ˜Y „ ˜Y ˜Z „ ˜Y „ ˜Z „
˜( ˜ Y „ ˜ ( ˜ Z „ ˜( ˜( ˜(
=  TJO R  DPT R (7.9)
˜Z ˜Z ˜Y „ ˜Z ˜Z „ ˜Y „ ˜Z „
Equations (7.8) and (7.9) follow from
x¢ = x cos q + y sin q (7.10)
y¢= – x sin q + y cos q (7.11)
The four derivatives in Eqs. (7.2) and (7.3) are got from Eqs. (7.6)–(7.7) and are given by

˜V Ë ˜V „ ˜W h Û Ë ˜V „ ˜W „ Û
= DPT RÌ DPT R TJO RÜ  TJO RÌ DPT R TJO RÜ (7.12)
˜Y Í ˜Y „ ˜Y „ Ý Í ˜Z „ ˜Z „ Ý
˜V Ë ˜V „ ˜W „ Û Ë ˜V „ ˜W „ Û
TJO R
˜Z
= ÌÍ ˜Y „ DPT R  ˜Y „ TJO R ÝÜ  DPT R Ì ˜Z „ DPT R  ˜Z „ TJO R Ü (7.13)
Í Ý
190 u BASICS OF ATMOSPHERIC SCIENCE

˜W Ë ˜V „ ˜W „ Û Ë ˜V „ ˜W „ Û
= DPT R Ì TJO R  DPT R Ü  TJO R Ì TJO R  DPT R Ü (7.14)
˜Y Í ˜Y „ ˜Y „ Ý Í ˜Z „ ˜Z „ Ý
˜W Ë ˜V „ ˜W „ Û Ë ˜V „ ˜W „ Û
TJO R
= ÌÍ ˜Y „ TJO R  ˜Y „ DPT R
ÜÝ  DPT R Ì TJO R  DPT R Ü (7.15)
˜Z Í ˜Z „ ˜Z „ Ý
Subtracting Eq. (7.13) from Eq. (7.14), one gets
˜W ˜V ˜W „ ˜V „
  (7.16)
˜Y ˜Z ˜Y „ ˜Z „
Adding Eq. (7.12) and Eq. (7.15), one gets
˜V ˜W ˜V ˜W
h h
  (7.17)
˜ Y ˜Z ˜Y ˜ Z
h h

Equations (7.16) and (7.17) provide for two combinations of the derivative of the velocity
components which remain invariant under the rotation of the coordinate axes. The quantity in
Eq. (7.16) is known as the vertical component of vorticity and many times simply referred to
˜W ˜V
as vorticity and is denoted as [ L ¹ ³ – 7
 while the quantity in Eq. (7.17) is
˜Y ˜Z
known as divergence. The following combination of the sum of the squares of the derivatives
in Eqs. (7.2) and (7.3) are also invariant under the rotation of the coordinate axes:
   
È ˜W ˜V Ø È ˜V ˜W Ø È ˜W „ ˜ V „ Ø È ˜V „ ˜W „ Ø
ÉÊ ˜Y  ˜Z ÙÚ  ÉÊ ˜Y  ˜Z ÙÚ ÉÊ ˜Y „  ˜Z „ ÙÚ  ÉÊ ˜Y „  ˜Z „ ÙÚ (7.18)

Each of the combination of derivatives whose squares appear in Eq. (7.18) is each known as
deformation. It is possible to rewrite Eqs. (7.2) and (7.3) in terms of vorticity, divergence and
deformation fields as follows:

 È ˜W ˜V Ø  È ˜V ˜W Ø  È ˜ V ˜W Ø  È ˜W ˜V Ø
u = V  É  Ù Z É  Ù Y É  Ù Y É  Ù Z (7.19)
 Ê ˜Y ˜Z Ú   Ê ˜Y ˜Z Ú   Ê ˜Y ˜ Z Ú   Ê ˜Y ˜Z Ú 

 È ˜W ˜V Ø  È ˜V ˜W Ø  È ˜V ˜W Ø  È ˜W ˜ V Ø
v = W  É  Ù Y É  Ù Z É  Ù Z É  Ù Y (7.20)
 Ê ˜Y ˜Z Ú   Ê ˜Y ˜Z Ú   Ê ˜Y ˜Z Ú   Ê ˜Y ˜ Z Ú 
The first term in the right-hand side of Eqs. (7.19) and (7.20) refer to translation, while the
remaining terms refer to vorticity, divergence and deformation fields

7.1.3 Circulation
Unlike vorticity, a vector field, which gives a measure of the rotation at any point in the fluid,
the circulation is a scalar field which provides a measure of the rotation for a finite area of
the fluid. The circulation is always defined around a closed contour in a fluid and is defined
as the line integral around the contour of the component of the velocity vector which is
ATMOSPHERIC MOTION u 191

locally tangent to the contour. Hence, for a contour in a fluid in the horizontal plane, the
circulation is given by
$ vÔ 7 ¹ EM vÔ 7 ] ] DPT B EM (7.21)
where dl represents an element of the contour in the fluid and V is the velocity field (refer
Figure 7.3). The circulation is considered positive, by convention if C > 0 for the counter-
clockwise integration around the closed loop.

FIGURE 7.3 Circulation around a closed contour.

EXAMPLE 7.1 For a small element of area in two dimensions show that the vertical
component of the vorticity of the fluid in the area is equal to the circulation around the
perimeter per unit area.
In two dimensions, the circulation around the perimeter of an element of area is given by

$ vÔ V EY  W EZ
where dx and dy are the components of the line element ds in x and y directions. Consider a
closed loop of fluid formed by an infinitesimal rectangle as seen in Figure 7.4.

∂u
u+ dy
∂y

dy
∂v
v v + dx
∂x

dx

u
FIGURE 7.4 Relationship between circulation and vorticity for an area element.
192 u BASICS OF ATMOSPHERIC SCIENCE

The circulation and the perimeter of the infinitesimal rectangle are:


È ˜W Ø È ˜V Ø
dC = V EY  É W  EY Ù EZ  É V  EZ EY  WEZ
Ê ˜Y Ú Ê ˜Z ÙÚ
È ˜W ˜ V Ø
dC = É  Ù EY EZ
Ê ˜Y ˜ Z Ú
dC = z dx dy
where z is the vertical component of vorticity and dx dy is the area of the infinitesimal area of
the rectangle.
The relationship between the circulation and vorticity in general is given by applying the
Stokes theorem to the velocity vector.

vÔ 7 ¹ EM ÔÔ " ³ – 7
¹ O E" (7.22)
where A is the area enclosed by the contour of the fluid particles and n is a unit normal to the
area element dA. Equation (7.22) states that the circulation around any closed circuit is equal to
the integral of the normal component of vorticity over the area enclosed by the closed circuit.

7.1.4 Kelvin’s Circulation Theorem


Kelvin’s circulation theorem is obtained by taking the line integral of the equation governing
the conservation of momentum in a coordinate system fixed with respect to the stars
(absolute or inertial coordinate system). Neglecting viscous forces, the above becomes
EB 7B ³Q ¹ EM
vÔ EU
¹ EM  vÔ S
 vÔ ³G ¹ EM (7.23)

where the gravitational force g is represented as a gradient of the geopotential f and the
subscript a refers to the absolute coordinate system. The integral on the left-hand side of
Eq. (7.23) can be written as
EB 7B E
EM = EU 7B ¹ EM
 7B ¹ EB EM

EU EU
EB 7B E
EM = 7B ¹ EM
 7B ¹ E 7B
EU EU
EB M
since 7B with l being the position vector. Substituting the above equation in Eq. (7.23),
EU
one gets the Kelvin’s circulation theorem
E$B E 7 ¹ EM EQ
EU vÔ B
v
Ô (7.24)
EU S
It is to be noted that the line integral of a perfect differential vanishes and this has been used
in deriving Eq. (7.24) since

vÔ ³G ¹ EM vÔ EG  BOE vÔ 7B ¹ E7B
 vÔ
E 7B ¹ 7B

ATMOSPHERIC MOTION u 193

For a barotropic fluid, the density of the fluid is a function of only the pressure and not
dependent on pressure and temperature. Hence the term on the right-hand side of Eq. (7.24),
known as the solenoidal term, also vanishes for a barotropic fluid giving rise to the Kelvin’s
circulation theorem which can be stated as: the absolute circulation is conserved following the
motion for a barotropic non-viscous fluid.

7.1.5 Bjerknes’ Circulation Theorem


It is more convenient and appropriate to utilize a coordinate system which is fixed with the
rotating earth and find the rate of change of relative circulation following the motion. It is
obvious that the relative circulation equals the absolute circulation minus the circulation due
to the rotation of the earth around its axis, since the absolute velocity equals the sum of
relative velocity and the velocity of the rotating earth. The velocity of the rotating earth is
given by V = w ´ r, where w is the angular velocity of the earth. The circulation due to
earth’s rotation around its axis can be obtained from the Stokes’ theorem as follows:

$F vÔ 7F ¹ EM ÔÔ" ³ – 7F
¹ O E" (7.25)
where A is the area enclosed by the closed circuit of fluid particles and n denotes the normal
as defined by the anticlockwise sense of the line integration using the right-hand screw rule.
If the line integral is calculated in the horizontal plane, n is directed along the local vertical
and ³ – 7F
¹ O Z TJO G is the Coriolis parameter. From Eq. (7.25), the circulation due to
the rotation of the earth is Z TJO G " XIFSF G is the average latitude over the area element A.
One can then write for the relative circulation as follows:

$ $B  $F $B  Z " TJO G $B  Z "O (7.26)


where An is the projection of the area element A in the equatorial plane (refer Figure 7.5).

FIGURE 7.5 Area An subtended on the equatorial plane by a horizontal area A at a latitude.
194 u BASICS OF ATMOSPHERIC SCIENCE

Differentiating Eq. (7.26) following the motion and using Eq. (7.24), one gets the Bjerknes’
circulation theorem
E$ EQ E"O
vÔ  Z (7.27)
EU S EU
For a barotropic fluid in the absence of viscous forces, Eq. (7.27) becomes
E$ E"O
 Z (7.28)
EU EU
If Eq. (7.28) is integrated from an initial state 1 to a final state 2 at latitudes f1 and f2 with
corresponding areas A1 and A2, one gets
$  $  Z " TJO G  " TJO G
(7.29)
Equation (7.29) relates the change in relative circulation in a barotropic non-viscous fluid for
a closed chain of fluid particles to changes in latitude and to the changes in the area enclosed
by the chain of fluid particles.

7.1.6 Applications of Circulation Theorem


The land and sea breeze circulation is a circulation commonly experienced by people who live
in the coastal areas. Most of us have an intuitive idea of the cause of the land and sea breeze
circulation. The following subsection utilizes the circulation theorem to explain the land and
sea breeze circulation.

Land and Sea Breeze Circulation


Occurrence of winds close to the surface from the sea during the day and from the land to the
sea during the night near the seacoast constitutes sea and land breeze circulation. The above
circulation from the sea and the land can be readily explained using the term of the right-hand
side of Eq. (7.24) called the solenoidal term. Figure 7.6 shows the vertical cross-section of
the atmosphere in the vicinity of a seacoast during the day. Since the land becomes warmer
r
r+1
P
4 3 r+2

P+1 r+3

P+2
r+4

P+3
r+5

P+4

1 2
Land
Sea
FIGURE 7.6 Isobars and isolines of density (isoteres) in a baroclinic fluid.
ATMOSPHERIC MOTION u 195

than the sea during the day and since the density of the air decreases with height, surfaces of
constant density slope upward from land to sea. However, the surfaces of constant pressure
are more nearly horizontal over both land and sea. The rate of change of circulation around a
vertical closed circuit consisting half over the land and half over the sea is obtained by taking
the line integral of the closed vertical circuit of Eq. (7.24). Line integration along the circuit
from 1 to 2 and 3 to 4 do not contribute since these correspond to dp = 0, while the line
integral from 2 to 3 contribute positively (since dp < 0) and the line integral from 4 to 1
contribute negatively (since dp > 0). Since the mean density over the land is less than over
the sea and since density appears in the denominator of the integrand in Eq. (7.24), the
positive contribution from 2 to 3 is less than the negative contribution of 4 to 1 and the right-
hand side of Eq. (7.24) makes a negative contribution to the rate of change of circulation.
This will result in an increase of circulation in the negative (clockwise) sense and cause the
circulation to develop such that the dense fluid will sink, while the lighter fluid will rise
giving rise to the sea breeze.

7.1.7 Vorticity in the Natural Coordinate System


The natural coordinate system introduced in Chapter 6 can be conveniently utilized to express
the vertical component of vorticity. Consider an infinitesimal loop ABCD as shown in Figure 7.7.

FIGURE 7.7 Circulation for an infinitesimal loop in the natural coordinate system.

Evaluating the circulation about the closed contour ABCD, one gets
˜7
E$ 7 <E T  E E T
>  ÈÉ7  Ø
E OÙ E T (7.30)
Ê ˜O Ú
However, from Figure 7.7, one gets
E E T
EC E O (7.31)
where db is the angular change in the wind direction for a distance d s. Substituting Eq. (7.31)
in Eq. (7.30), one gets
È ˜7 EC Ø
E$ É  7 E OE T (7.32)
Ê ˜O E T ÙÚ
196 u BASICS OF ATMOSPHERIC SCIENCE

Taking the limit of dn ® 0, ds ® 0, in the above equation, one gets


E$ ˜7 7
[   (7.33)

MJN
E O E T  E OE T
˜O 3T
where Rs is the radius of curvature of the line tangential to the velocity. Equation (7.33) can
be physically interpreted as follows:
The vertical component of the vorticity is the resultant of (i) the rate of change of wind
˜7
speed normal to the flow  called the shear vorticity (n is the unit vector in the
˜O
horizontal plane directed to the left of the local flow direction in the natural coordinate
7
system), and (ii) the turning of the wind tangential along the flow direction called the
3T
curvature vorticity.
The above shows that shear vorticity is present even in straight line motion of the fluid
speed changes in a direction normal to the flow direction. A paddle wheel kept in the vicinity
of a jet stream shows the existence of a cyclonic/anticyclonic vorticity to the north/south of
the velocity maximum (refer Figure 7.8(a)). For curved flow, consider a frictionless fluid with
zero relative vorticity upstream around the bend as shown in Figure 7.8(b). The bend in the
flow will ensure that around the bend, the flow will acquire anticyclonic vorticity due to
curvature. However, it may so happen that the above curved flow can have zero vorticity
provided that the shear vorticity is equal and opposite to the curvature vorticity. This would
require that the shear vorticity gives rise to cyclonic vorticity, i.e. the flow along the inner part
of the flow along the bend flows faster in just the right amount to cancel the vorticity due to
the curvature.

(a) (b)

FIGURES 7.8 Two different types of two-dimensional flow: (a) linear shear flow with
vorticity, (b) curved flow with no vorticity.

7.2 ISOBARIC COORDINATE SYSTEM


Atmospheric scientists find it convenient to utilize pressure as the vertical coordinate system
(isobaric coordinate system) instead of height since the meteorological upper air observations
are measured in constant pressure levels. Also, utilizing pressure as the vertical coordinate
ATMOSPHERIC MOTION u 197

system results in some simplification in some of the relations and hence it is worthwhile to
study the isobaric coordinate system.
The hydrostatic equation, which provides for the relationship between the pressure
increments and the height increments, can be utilized to obtain the relationship between the
derivatives in the height coordinate system to the derivates of the isobaric coordinate system.
We shall adopt a Cartesian coordinate system in which x and y axes are the horizontal axes
and in which the position on the vertical axis is specified in terms of pressure. A constant
pressure surface will be inclined with respect to the horizontal and not normal to the pressure
(vertical) coordinate. Along such a vertical, the z coordinate depends only upon one
independent coordinate p. Hence a derivative with respect to z may be written as follows:
˜ ˜Q ˜
(7.34)
˜[ ˜[ ˜Q
Utilizing the hydrostatic equation in Eq. (7.34), one gets
˜ ˜
 SH (7.35)
˜[ ˜Q

7.2.1 Horizontal and Time Derivates in Isobaric Coordinate


System
The isobaric coordinate system appears as in Figure 7.9(a), where d p is the pressure
increments. Let any two constant pressure surfaces be denoted by p2 and p1 (with p2 > p1).
Consider any arbitrary function f (x, y, z) which we would like to consider in the isobaric
coordinate system as f (x, y, p(x, y, z)) (refer Figure 7.9(b)).

FIGURES 7.9 Isobaric coordinate system: (a) isobars drawn at thickness of dp, (b) a quantity f is
measured at three points at the intersection of a isobaric and a constant height
surface.
198 u BASICS OF ATMOSPHERIC SCIENCE

One can notice that the following equation holds:


G  G G  G  G  G
Q  Q

 (7.36)
Y  Y Y  Y Q  Q
Y  Y

Taking the limits as x2 ® x1 and p1 ® p2, one gets

È ˜G Ø G  G È ˜G Ø G  G  È ˜G Ø G  G È ˜Q Ø Q  Q
ÉÊ ÙÚ É Ù 
É Ù  ÉÊ ÙÚ (7.37)
˜Y [ Y  Y Ê ˜Y Ú Q Y  Y Ê ˜Q Ú
Y
Q  Q ˜Y [ Y  Y

Substituting Eq. (7.37) into Eq. (7.34), one gets

È ˜G Ø È ˜G Ø È ˜G Ø È ˜Q Ø
ÉÊ ÙÚ ÉÊ ÙÚ  É Ù ÉÊ ÙÚ (7.38)
˜Y [ ˜Y Q Ê ˜Q Ú Y ˜Y [
A similar relation can be derived in the y direction of the form:

È ˜G Ø È ˜G Ø È ˜G Ø È ˜Q Ø
ÉÊ ˜Z ÙÚ ÉÊ ˜Z ÙÚ  ÉÊ ˜Q ÚÙ ÊÉ ˜Z ÚÙ (7.39)
[ Q Z [

Combining Eqs. (7.38) and (7.39), one gets


˜G
³[ G ³Q G  ³[ Q (7.40)
˜Q

È ˜G Ø È ˜G Ø
where ³Q G J ÉÊ ÙÚ  K É Ù (7.41)
˜Y Q Ê ˜Z Ú Q
It is assumed in the above derivation that neither pressure nor the function f depends on time.
However, it is relatively easy to extend the same to include time also. Equations (7.38) and
(7.39) can be rewritten using the hydrostatic equations as follows:

È ˜Ø È ˜Ø È ˜[ Ø ˜
ÉÊ ÙÚ = ÉÊ ÙÚ  S H ÉÊ ÙÚ (7.42)
˜Y [ ˜Y Q ˜ Y Q ˜Q

È ˜Ø È ˜Ø È ˜[ Ø ˜
ÉÊ ˜Z ÙÚ = ÉÊ ˜Z ÙÚ  S H ÉÊ ˜Z ÚÙ ˜Q (7.43)
[ Q Q

Partial derivatives with respect to time can be obtained in the same manner as follows:

È˜Ø È˜Ø È ˜[ Ø ˜
ÉÊ ÙÚ ÉÊ ÙÚ  S H ÉÊ ÙÚ (7.44)
˜U [ ˜U Q ˜U Q ˜Q

È ˜[ Ø
The term É Ù refers to the rate at which the height of the isobaric surfaces changes with
Ê ˜U Ú Q
time. Equations (7.37), (7.42),(7.43) and (7.44) can be utilized to transform the basic
governing equations to the isobaric coordinate system. In the height coordinate system,
ATMOSPHERIC MOTION u 199

E ˜ ˜ ˜ ˜
V  W X (7.45)
EU ˜U ˜Y ˜Z ˜[
Substitution of Eqs. (7.37), (7.42), (7.43) and (7.44) in Eq. (7.45), one gets

E È˜Ø È ˜Ø È ˜Ø ˜ Ë È ˜[ Ø È ˜[ Ø È ˜[ Ø Û ˜
ÉÊ ÙÚ  V ÉÊ ÙÚ  W É Ù  X S H  S H ÌÉ Ù  V É Ù  W É Ù Ü (7.46)
EU ˜U Q ˜Y Q Ê ˜Z Ú Q ˜Q ÌÍ Ê ˜U Ú Q Ê ˜Y Ú Q Ê ˜Z Ú Q Ü ˜Q
Ý
EQ
Applying the above operator to pressure itself and defining X one gets
EU

EQ Ë È ˜[ Ø È ˜[ Ø È ˜[ Ø Û
X  XS H  S H ÌÉ Ù  V É Ù  W É Ù Ü (7.47)
EU ÌÍ Ê ˜U Ú Q Ê ˜Y Ú Q Ê ˜Z Ú Q Ü
Ý
EQ
It is unfortunate that and the angular velocity of earth’s rotation have the same symbol.
EU
E
Substituting Eq. (7.47) into Eq. (7.46), one gets the operator in isobaric coordinate system as
EU
E È˜Ø È ˜Ø È ˜Ø ˜
ÉÊ ÙÚ  V ÉÊ ÙÚ  W É Ù  X (7.48)
EU ˜U Q ˜Y Q Ê ˜Z Ú Q ˜Q
Equation (7.48) shows that the total derivative in a isobaric system has the same form as in
the height coordinate system except for the last term in the right-hand side of Eq. (7.48).

7.2.2 Continuity Equation in Isobaric Coordinate System


The continuity equation in isobaric coordinates assumes a simple form as can be seen below.
The equation of continuity in the height coordinate system is given as
 ES ˜V ˜W ˜X
  
S EU ˜Y ˜Z ˜[

Using Eqs. (7.37), (7.42), (7.43) and (7.48) in the above equation, one gets

 Ë È ˜S Ø È ˜S Ø È ˜S Ø È ˜S Ø Û
 ÌÉ Ù  V É Ù  W É Ù  X É Ù Ü
S ÌÍÊ ˜U Ú Q Ê ˜Y Ú Q Ê ˜Z Ú Q Ê ˜Q Ú ÜÝ

È ˜V Ø È ˜W Ø ˜X Ë È ˜[ Ø ˜V È ˜[ Ø ˜W Û  X ˜S ˜
ÉÊ ÙÚ  É Ù   SH ÌÉ Ù É Ù Ü  SH
˜Y Q Ê ˜Z Ú Q ˜Q ÌÍ ˜Y Q ˜Q Ê ˜Z Ú Q ˜Q ÜÝ S ˜Q
Ê Ú ˜Q
ËÈ ˜[ Ø È ˜[ Ø È ˜[ Ø Û
ÌÉ Ù  V É Ù  W É Ù Ü (7.49)
ÌÍÊ ˜U Ú Q Ê ˜Y Ú Q Ê ˜Z Ú Q Ü
Ý
200 u BASICS OF ATMOSPHERIC SCIENCE

È ˜V Ø È ˜W Ø È ˜X Ø
ÉÊ ÙÚ  É Ù  É
˜Y Q Ê ˜Z Ú Q Ê ˜Q ÚÙ

 Ë È ˜S Ø È ˜S Ø È ˜S Ø Û Ë È ˜[ Ø ˜V È ˜[ Ø ˜W Û
 ÌÉ Ù  V É Ù  W É Ù Ü  S H ÌÉ Ù É Ù Ü
S ÌÍ Ê ˜U Ú Q Ê ˜Y Ú Q Ê ˜Z Ú Q Ü
Ý ÌÍ Ê ˜Y Ú Q ˜Q Ê ˜Z Ú Q ˜Q ÜÝ
ËÈ ˜ Ø È ˜Ø È ˜ Ø Û ˜[ Ë È ˜[ Ø È ˜V Ø È ˜[ Ø È ˜W Ø Û
 S H ÌÉ Ù  V É Ù  W É Ù Ü  S H Ì É Ù É Ù  É Ù É Ù Ü (7.50)
ÌÍ Ê ˜U Ú Q Ê ˜Y Ú Q Ê ˜Z Ú Q Ü ˜ Q
Ý ÌÍ Ê ˜Y Ú Q Ê ˜Q Ú Ê ˜Z Ú Q Ê ˜Q Ú ÜÝ
The first term in the right-hand side of Eq. (7.50) cancels the third term of Eq. (7.50) by
using Eq. (7.37) resulting in
È ˜V Ø È ˜W Ø È ˜X Ø
ÉÊ ÙÚ  É Ù  É  (7.51)
˜Y Q Ê ˜Z Ú Q Ê ˜Q ÙÚ
Equation (7.51) is the continuity equation in the isobaric coordinate system and has its
simplest form.

7.2.3 Horizontal Equation of Motion in Isobaric Coordinate


System
The horizontal equations of motion in the isobaric coordinate system can be written down as

È ˜V Ø È ˜V Ø È ˜V Ø ˜V È ˜[ Ø
 VÉ Ù  WÉ Ù  X GW  H É Ù (7.52)
ÊÉ ˜U ÚÙ Q Ê ˜Y Ú Q Ê ˜Z Ú Q ˜Q Ê ˜Y Ú Q

È ˜W Ø È ˜W Ø È ˜W Ø ˜W È ˜[ Ø
ÉÊ ÚÙ  V ÊÉ ÚÙ  W É Ù  X  GV  H É Ù (7.53)
˜U Q ˜Y Q Ê ˜Z Ú Q ˜Q Ê ˜Z Ú Q
Equations (7.52) and (7.53) are simpler than the corresponding equations in the height
coordinate system since the density r does not appear explicitly in the above equations.

7.2.4 Geostrophic and Thermal Wind Equations in Isobaric


Coordinates
The geostrophic equations in the isobaric coordinate system can be immediately written from
Eqs. (7.52) and (7.53) and they are as follows:

È ˜[ Ø È ˜[ Ø
GW HÉ GV HÉ Ù (7.54)
Ê ˜Y ÙÚ Q

Ê ˜Z Ú Q
The thermal wind equations in the height coordinate system were earlier derived in Section 6.6
and they are
˜V  H È ˜5 Ø ˜W H È ˜5 Ø
É Ù (7.55)
G5 ÉÊ ˜Z ÙÚ

˜[ ˜[ G5 Ê ˜Y Ú
ATMOSPHERIC MOTION u 201

Employing Eq. (7.37) and the equation of state on the above equation, one gets the thermal
wind equations in isobaric coordinate system.
˜V 3 È ˜5 Ø ˜W  3 È ˜5 Ø
É Ù  É Ù (7.56)
˜ MO Q
G Ê ˜Z Ú Q ˜ MO Q
G Ê ˜Y Ú Q

7.3 VORTICITY AND DIVERGENCE EQUATIONS


By taking the dot product of the gradient operator with the conservation of momentum
equation, one gets the divergence equation. By taking the curl of the gradient operator with
the conservation of momentum equation, one gets the vorticity equation. By taking the dot
product of the unit vector in the vertical direction with the vorticity equation, one obtains the
vertical component of the vorticity equation. Even though the vorticity equation and the
divergence equations are alternate expressions of the conservation of momentum equation,
these equations can provide greater physical insight into the dynamics and hence are widely
employed. The following subsections derive the vorticity and the divergence equations.

7.3.1 Vorticity Equation


The horizontal equations of motion in height coordinate system are
˜V ˜V ˜V ˜V ˜Q
V W X = GW  B  'Y (7.57)
˜U ˜Y ˜Z ˜[ ˜Y
˜W ˜W ˜W ˜W ˜Q
V W X =  GV B  'Z (7.58)
˜U ˜Y ˜Z ˜[ ˜Z
The vertical component of the curl of the horizontal equations of motion provides the rate of
˜W ˜ V
change of the relative vorticity [  by taking the partial derivatives with respect to y
˜Y ˜ Z
of Eq. (7.57) and subtracting it from the partial derivative with respect to x of Eq. (7.58) to get
˜[ ˜[ ˜[ ˜[ ˜G È ˜ V ˜ W Ø Ë ˜ X ˜ W ˜X ˜ V Û
V W X W  [  ÉÊ ˜Y  ˜Z ÚÙ  Ì ˜Y ˜[  ˜Z ˜[ Ü
G

˜U ˜Y ˜Z ˜[ ˜Z Í Ý
Ë ˜B ˜Q ˜B ˜Q Û Ë ˜'Z ˜'Y Û
Ì  ÜÌ  Ü (7.59)
Í ˜Y ˜Z ˜Z ˜Y Ý Í ˜Y ˜Z Ý
˜G EG
Since f depends only on y, one can replace W and hence, Eq. (7.59) becomes
˜Z EU

E [  G
È ˜V ˜W Ø Ë ˜X ˜W ˜X ˜V Û Ë ˜B ˜Q ˜B ˜Q Û Ë ˜'Z ˜'Y Û
 [  G
É      Ì  Ü
˜U Ê ˜Y ˜Z ÚÙ ÍÌ ˜Y ˜[ ˜Z ˜[ ÝÜ ÍÌ ˜Y ˜Z ˜Z ˜Y ÝÜ Í ˜Y ˜Z Ý
(7.60)
Equation (7.60) is called the vorticity equation and it provides for the total derivative of
z + f, also known as absolute vorticity.
202 u BASICS OF ATMOSPHERIC SCIENCE

The first term on the right hand side of Eq. (7.60) is known as the divergence term. This
È ˜V ˜W Ø
term indicates that positive divergence É  ! Ù is associated with the decrease of the
Ê ˜Y ˜Z Ú
absolute vorticity. This is understandable since with positive divergence, the area enclosed by
a circuit of fluid particles will increase with time. If the circulation were to be conserved this
would contribute to the decrease of the average absolute vorticity of the fluid mass within the
circuit of fluid parcel. The second term of Eq. (7.60) known as the tilting term represents
generation of vertical vorticity due to tilting of the horizontally oriented vorticity components
into the vertical by a vertical motion field which varies non-uniformly in the horizontal.
Figure 7.10 illustrates the tilting mechanism.
˜ X ˜W
Consider the first tilting term 
˜ Y ˜[


Consider a region where the above term is


positive (i.e. where the y component of
velocity increases with height and where the
vertical velocity decreases with increasing x).
Increase of v with height will give rise to a
component of vorticity in the negative x
direction (as shown by the broad arrow in
Figure 7.10). The fact that the vertical
velocity decreases with increasing x will tend
to tilt the vorticity vector (oriented initially in FIGURE 7.10 Vorticity production by
the negative x direction) to a position the tilting of a horizontal
velocity vector (indicated
as indicated by the dotted broad arrow in
by double arrow).
Figure 7.10.
The dotted broad arrow has a component in the vertical, indicating a generation of
˜W ˜X
positive vertical vorticity, if !  BOE   By similar arguments it can again be shown
˜[ ˜Y
˜W ˜X
that there will be generation of positive vertical vorticity, if !  BOE   The third term
˜[ ˜Z
of the right-hand side of Eq. (7.60) is known as the solenoidal term of the vorticity equation.
The solenoidal term in the circulation theorem (Eq. 7.24) may be written as

EQ È ˜Q ˜Q Ø
vÔ S Ô ÉÊ B ˜Y EY  B ˜Z EZÚÙ
v (7.61)

Applying the Stokes theorem to the above equation, one gets

EQ È ˜B ˜Q ˜B ˜Q Ø
vÔ S
 ÔÔ É 
Ê ˜Y ˜Z ˜Z ˜Y ÚÙ
EYEZ (7.62)

Equation (7.62) indicates that the solenoidal term in the vorticity equation is the limit of the
solenoidal term in the circulation theorem divided by the area in the limit when the area goes
to zero. The final term in the right-hand side of Eq. (7.60) represents the effect of viscous
ATMOSPHERIC MOTION u 203

forces on the generation of vertical vorticity. The vorticity equation (Eq. (7.60)) provides for
the various mechanisms by which the absolute vorticity of a parcel of air can change. The
vorticity equation expressed in isobaric coordinate system is of the same form as Eq. (7.59)
except for the absence of the solenoidal term and is as follows (the terms corresponding to
the viscous forces are neglected):

È ˜[ Ø È ˜[ Ø È ˜[ Ø È ˜[ Ø EG È ˜V ˜W Ø
ÉÊ ÙÚ  V ÉÊ ÙÚ  W É Ù  X É Ù W  [  G
É  
˜U Q ˜Y Q Ê ˜Z Ú Q Ê ˜Q Ú Q EZ Ê ˜Y ˜Z ÚÙ Q
È ˜X Ø ˜W È ˜X Ø ˜V
ÉÊ Ù  (7.63)
˜Y Ú Q ˜Q ÊÉ ˜Z ÚÙ Q ˜Q
The absence of the solenoidal term in the above equation is due to the fact that density does
not appear as a coefficient in the isobaric coordinate system.

7.3.2 Divergence Equation


The divergence equation is obtained by taking the divergence of the momentum equations, i.e.
by taking the partial derivatives with respect to x of Eq. (7.57) and adding it to the partial
derivative with respect to y of Eq. (7.58) to get

˜% È ˜% Ø È ˜% Ø È ˜% Ø È ˜ V ˜ W ˜ W ˜ V Ø ˜ X ˜ V ˜X ˜ W
 VÉ  WÉ  XÉ  %   É    
˜U Ê ˜Y ÙÚ Ê ˜Z ÙÚ Ê ˜[ ÙÚ Ê ˜Y ˜Z ˜Y ˜Z ÙÚ ˜Y ˜[ ˜Z ˜[

È ˜B ˜Q ˜B ˜Q Ø
ÉÊ ˜Y ˜Y  ˜Z ˜Z ÚÙ  C V  G 7 

È ˜ Q ˜ Q Ø È ˜'Y ˜'Z Ø
BÉ  Ù  (7.64)
Ê ˜Y  ˜Z  Ú ÉÊ ˜Y ˜Z ÙÚ
˜V ˜W
where the horizontal divergence D is given by %   The divergence equation in the
˜Y ˜Z
isobaric coordinate system with viscous terms neglected has a similar form as Eq. (7.64)
except for the absence of terms involving the derivatives of specific volume a and is given by

È ˜% Ø È ˜% Ø È ˜% Ø È ˜% Ø È ˜V ˜W ˜W ˜V Ø È ˜X Ø ˜V
ÉÊ ÙÚ  V ÉÊ ÙÚ  W É XÉ  %   É  É Ù 
˜U Q ˜Y Q Ê ˜Z Ú Q
Ù Ê ˜Q ÙÚ Ê ˜Y ˜Z ˜Y ˜Z Ú Q Ê ˜Y Ú Q ˜Q
Ù

È ˜Z Ø ˜ W È ˜  [ ˜ [ Ø
ÉÊ ˜Z ÙÚ ˜Q  C V  G [  H É    Ù (7.65)
Q Ê ˜Y ˜Z Ú Q

7.4 ABSOLUTE AND POTENTIAL VORTICITY


The following subsections introduce the important concepts of absolute vorticity and potential
vorticity. The most important of the meteorological waves, the Rossby wave, arises as a result
204 u BASICS OF ATMOSPHERIC SCIENCE

of the conservation of absolute vorticity in a barotropic environment. However, in a baroclinic


environment, the Rossby wave arises as a result of the conservation of potential vorticity.

7.4.1 Absolute Vorticity


The absolute vorticity is simply the sum of the relative vorticity and the planetary vorticity,
i.e. the sum of the vorticity about the vertical of the motion relative to the earth and the
vorticity about the vertical of the earth’s rotation
h=z+f (7.66)
The total derivative of the absolute vorticity was derived in Eq. (7.60). Applying the scale
analysis to the vorticity equation for mid-latitude synoptic scale motion, only the divergence
term appears to be important among the terms in the right-hand side of Eq. (7.60). However,
for appreciable periods of time even this divergence term may be considered negligible
resulting in
E
[  G
 (7.67)
EU

Equation (7.67) expresses the conservation of absolute vorticity and is valid under the
following conditions of (i) negligible divergence, (ii) small vertical motion, (iii) weak
solenoidal fields, and (iv) negligible viscous forces. Under the conditions when conservation
of absolute vorticity holds, a poleward moving air parcel (increasing f ) should have its
relative vorticity decreased. Assuming that the relative vorticity is determined primarily by the
curvature, this corresponds to a situation where a poleward moving air parcel will lose its
cyclonic curvature and ultimately curve anti-cyclonically. If one were to construct the
trajectory of an air parcel subjected to the conservation of absolute vorticity, such trajectories
called constant absolute vorticity trajectories, can form the basis of applying diagnostic
reasoning to weather forecasting.

7.4.2 Potential Vorticity


Using the Poisson equation and the equation of state, it can be shown that the density varies as

QD7 D1

3TQ D1
S 
(7.68)
R 3TQ
On an adiabatic surface (where q is constant), the density of an air parcel is a function of
pressure alone and hence the solenoidal term in the circulation theorem for an air parcel along
an adiabatic surface is of the form
EQ  D D

vÔ S
—v
Ô EQ 7 1  (7.69)

Hence for adiabatic motion, the circulation theorem as applied on a constant q surface reduces
to the same form as that for a barotropic fluid, i.e.
E
$  : " TJO G
 (7.70)
EU
ATMOSPHERIC MOTION u 205

where C is evaluated for a close circuit of fluid within an area A on an adiabatic surface.
Assuming the adiabatic surface to be horizontal and using the definition of the relative
vorticity z, the integral of Eq. (7.70) for an infinitesimal air parcel becomes
A (z + f ) = constant (7.71)
Let the above air parcel be confined between two potential temperature surfaces q0 and q0 + dq,
"E Q
duly separated by a pressure difference d p. Since the mass of the air parcel . is
H
conserved following the motion, one gets

"
.H È ER Ø È .H Ø
DPOTUBOU É
È ER Ø
(7.72)
ÉÊ E Q ÙÚ ÉÊ Ù
EQ ER Ú Ê E Q ÙÚ

since
.H is a constant. Substituting Eq. (7.72) in Eq. (7.71) and taking the limit of dp ® 0,
ER
one gets
˜R
[  G
DPOTUBOU (7.73)
˜Q
Equation (7.73) states that in an adiabatic frictionless motion the potential vorticity is
conserved. The exact expression for the potential vorticity can assume different forms
depending on the nature of the fluids. For a homogeneous incompressible fluid since the
density is constant, one has

SE
" .[ È Ø
DPOTUBOU É Ù
ÊE Ú [ (7.74)

where d z is the depth of the fluid parcel. Substituting Eq. (7.74) in Eq. (7.71), one gets
another form of potential vorticity conservation
[ G
DPOTUBOU (7.75)
E[
The conservation of potential vorticity exerts a powerful constraint on the large-scale motion
in the atmosphere as can be illustrated by considering the airflow over a mountain range. Let
us assume a straight uniform westerly current with z = 0 impinges on a north-south oriented
mountain range. If the airflow is adiabatic and frictional forces can be neglected, each column
of air confined between the potential temperature surfaces q0 and q0 + dq remains between
these potential temperature surfaces as it crosses the mountains. This results in a potential
temperature surface close to the ground following the contour of the local topography.
However, the potential temperature surface very much above the ground by several kilometres
will not follow the ground contours. We shall assume that any vorticity, which may appear in
the uniform westerly current, will arise due to curvature and not due to shear. When the air
column approaches the mountain range, its vertical extent usually increases (refer Figure
7.11(a)), and conservation of potential vorticity (Eq. (7.75)) requires the relative vorticity must
increase an indicated in Figure 7.11(a). The air parcel then acquires cyclonic vorticity initially
206 u BASICS OF ATMOSPHERIC SCIENCE

as shown in Figure 7.11(b). When the air column begins to cross the mountain range, its
vertical extent decreases (refer Figure 7.11(a)). Since the behaviour of the westerly adiabatic
frictionless flow is governed by Eq. (7.75), the relative vorticity must decrease in order to
conserve potential vorticity. The air parcel will hence acquire anticyclonic vorticity and must
move southward as shown in Figure 7.11(b).

q0 + dq
z
z=0 z>0 z<0 z>0 z<0

q0
x
(a)

x (b)

FIGURE 7.11 Example of a westerly flow over a mountain barrier: (a) the depth of a fluid column
with respect to the downwind direction, and (b) the trajectory of an air parcel in the
x, y plane.

Once the air parcel has passed over the mountain range and the air column has got back
to its original depth, the parcel will be south of its original latitude. This reduces the Coriolis
parameter and in order to conserve potential vorticity the air parcel must acquire cyclonic
curvature. Once the air parcel gets back to its original latitude, it will still have a northward
velocity component and continued northward movement will increase f and reduce z thereby
acquiring anti-cyclonic curvature. Thus, the air parcel forming part of a uniform westerly
current with zero relative vorticity upstream of the mountain range follows a wave-like
trajectory in the horizontal plane after crossing the mountain barrier. Hence, a steady westerly
flow over a mountain range results in a cyclonic flow pattern immediately on the other side of
the mountain range followed by an alternating pattern of ridges and troughs.
The above cyclonic flow pattern on the lee side of the mountain range is called the lee
side trough and is frequently observed when well-defined atmospheric currents flow over
mountain ranges.
In the case of a uniform easterly flow approaching a mountain range, it is clear that in
order to conserve potential vorticity, an air column has to acquire anti-cyclonic curvature as it
begins to move over the mountain. Hence the fluid parcel would tend to curve northwards
towards higher latitudes. This leads to an increase in f and to a further decrease in z. This
would eventually cause the fluid parcel to curve back towards the east and hence no steady
state potential vorticity conserving flow pattern would be possible. Due to the above reason, a
uniform easterly flow (unlike the westerly flow) must feel the influence of the mountain range
upstream. As shown in Figure 7.12, a uniform easterly air column becomes to curve
cyclonically before it reaches the mountain range.
ATMOSPHERIC MOTION u 207

q0 + dq
z
z=0 z>0 z<0 z>0 z=0

q0
x
(a)

x (b)

FIGURE 7.12 Example of an easterly flow over a mountain barrier: (a) the depth of a fluid
column with respect to the downwind direction, and (b) the trajectory of an air
parcel in the x, y plane.

The above cyclonic curvature is produced as the airflow follows the isolines of the
pressure field generated by the mountain range. The potential vorticity is conserved since the
cyclonic vorticity is balanced by a decrease in f. As the column of air moves to the top of the
mountain, it continues in its equatorward movement with the decrease in depth balanced by a
decrease in the Coriolis parameter. The above-mentioned process is simply reversed as the air
parcel moves down the mountain towards the west and hence the air column is moving
westward at its original latitude at the same distance downstream from the mountain range.
Hence the conservation of potential vorticity imposes a powerful constraint on a uniform
easterly current impinging on a north-south mountain range and damps out the disturbances
on the lee side of the mountain range.

SOLVED EXAMPLES

1. Show that for the case of a circular disk of fluid in solid body rotation, the circulation is the
product of the area enclosed by the circular disk and twice the angular speed of rotation.
Solution: Consider a circular disk of fluid of radius r undergoing solid body rotation at
angular velocity w about the vertical z-axis. The circulation about the circular loop of the
fluid is then given by
Q
$ vÔ 7 ¹ EM Ô X S  EG XQ S  X
Q S 


Since the tangential wind is V = w ´ r, where r is the distance to the axis of rotation and f
is the angle subtended by an element. Hence for the case of a fluid undergoing solid body
rotation, the circulation is the product of the area enclosed by the circular disk and twice the
angular speed of rotation
2. In the free atmosphere, i.e. neglecting frictional forces, what would be the relative velocity
at 30°N for an air parcel moving northward from equator?
208 u BASICS OF ATMOSPHERIC SCIENCE

Solution: Using the conservation of angular momentum, one gets m Us Rs = m Ud Rd,


where subscript d and s refer to the destination and source, respectively. While Us = w RE
cos fs, where w is the angular velocity of earth, RE is the radius of earth and fs is the
latitude of source. Again Rd = RE cos fd, and Ud = U¢ + w RE cos fd, and so
Ë DPT GT Û Ë  Û
6„X 3& Ì  DPT GE Ü  –   – 
Ì  DPT ’
Ü  N T 
ÍÌ DPT G E ÝÜ Í DPT ’
Ý
3. If wind rotates as a solid body around the centre of a low-pressure system and the tangential
velocity is 10 m s–1 at a radius of 200 km, find the relative vorticity.
Solution: For solid body rotation,
7
 – 
[   T 
S
 – 
4. A 10 km deep layer of air at 45°N latitude has no curvature, but has a shear of –10 m s–1
across a distance of 300 km. What is the potential vorticity?
Solution: The absolute vorticity is sum of relative vorticity and planetary vorticity.
Potential vorticity is then defined as the ratio of absolute vorticity and the depth of the air. Hence
[ S  [D
zp =
'[

  –  –  
TJO 

 – 
=  –   N  T 
 – 
5. Find the isentropic potential vorticity for the above example assuming the density of air as
'R
0.5 kg m–3 and  , LN 
'[
Solution: The isentropic potential vorticity is given by
'R [ S  [ D 'R
zIPV = [ 1
S S '[

 –   –  

= 
 –  LN T

LH

,


6. Show that for a barotropic flow under conservative body force the vorticity equation can be
E:
written as : ¹ ³
W where W is the vorticity vector and v is the velocity vector.
EU
Solution: Euler’s equation of motion for a conservative body force can be written as
˜W 
 W ¹ ³W  ³ Q  ³G
˜U S
where f is the geopotential. From vector identity, one gets
È W Ø
W ¹ ³W ³É Ù  W – ³ – W
Ê Ú
ATMOSPHERIC MOTION u 209

By definition, vorticity W = Ñ ´ v
Substituting, one gets

˜W È W Ø 
 ³É Ù  : – W  ³Q  ³G
˜U Ê Ú S
Taking the curl of the above equation and using the identity curl(div v) = 0, one gets

˜: È Ø
 ³ – : – W
 ³ – É ³QÙ
˜U ÊS Ú
From vector identity, ³ – " – #
"³ ¹ #  # ¹ ³
"  #³ ¹ "  " ¹ ³
# and Ñ ´ ( f v) =
f Ñ ´ v + Ñf ´ v, one gets

˜:  
 W ¹ ³:  :³ ¹ W  : ¹ ³W  ³ – ³Q  ³S – ³Q
˜U S S
For barotropic fluid r = r(p) only and so Ñr ´ Ñp = 0.
Also, from the continuity equation, ³ ¹ W  Substituting all the above, one gets

E:
: ¹ ³
W
EU
7. Show that in a flow, which is not isentropic, any moving fluid particle carries with it a
È Ø
constant value of the product É Ù ³T ¹ ³ – W
where s is the specific entropy. The above
Ê SÚ
quantity in its most general form is known as Ertel’s potential vorticity.
Solution: When the fluid flow is not isentropic, the right-hand side of Euler’s equation of

motion  ³Q cannot be written in terms of –Ñw, where w is the specific enthalpy and
S
hence the vorticity equation becomes

˜X 
³ – W – X
 ³S – ³Q
˜U S
Multiply the above equation scalarly by Ñs. Since s = s(p, r), Ñs is a linear function of Ñp
and Ñr and hence ³T ¹ ³S – ³Q
 The expression on the right-hand side of the vorticity
equation can be transformed as follows:
˜X
³T ¹ = ³T ¹ ³ – W – X
 ³ ¹ <³T – W – X
>
˜U
=  ³ ¹ <W X ¹ ³T
>  ³ ¹ <X W ¹ ³T
>
=  < X ¹ ³T
³ ¹ W
>  W ¹ ³ X ¹ ³T
 X ¹ ³ W ¹ ³T

210 u BASICS OF ATMOSPHERIC SCIENCE

˜T
Assuming adiabatic conditions for the ideal fluid, W ¹ ³T   Substituting, one gets
˜U
˜
X ¹ ³T
 W ¹ ³ X ¹ ³T
 X ¹ ³T
³ ¹ W 
˜U
E
The first two terms can be combined as X ¹ ³T

Using from the continuity equation
EU
ES
S³ ¹ W  and substituting, one gets
EU
E È X ¹ ³T Ø

EU ÉÊ S ÙÚ
8. Determine the shape of the surface of an incompressible fluid subject to a gravitational field,
which is contained in a cylindrical vessel, which rotates around its vertical axis with a
constant angular velocity w.
Solution: Assuming the axis to coincide with the z-axis, one gets vx = –y w, vy = x w and vz
= 0. The above velocity components identically satisfy the continuity equation. Substituting
in the Euler’s equations of motion, one gets

 ˜Q
xw2 =
S ˜Y
 ˜Q
yw2 =
S ˜Z
 ˜Q
H =0
S ˜[
The general integral of the above three equations is
Q 
X  Y   Z 
 H[  DPOTUBOU
S 

Since at the free surface, p = constant and hence it is clear that the surface is a parabloid
whose equation is given by

[ X  Y  Z

H
Here it is assumed that the origin is taken at the lowest point of the surface.
9. Show that for a steady motion of a rapidly rotating ideal incompressible fluid, is a
superposition of two independent motions: two-dimensional flow in the transverse plane and
an axial flow independent of the vertical axis z. The above result is known as the Taylor
Proudman theorem.
Solution: For rapidly rotating steady ideal flow, the equation of motion becomes

X – W  ³Q or in component form
S
ATMOSPHERIC MOTION u 211

 ˜Q  ˜Q ˜Q
X W Z   X W Y   
S ˜Y S ˜Z ˜ [
where x and y are Cartesian coordinates in the plane perpendicular to the axis of rotation z.
From the above, it is clear that p and hence, vx and vy are independent of the longitudinal
coordinate z. Eliminating p from the first two equations by cross-differentiating, one gets
˜W Y ˜W Z
 
˜Y ˜Z
Combining the above equation with the equation of continuity for incompressible flow, i.e.

˜W Y ˜W Z ˜W [
  = 0, one gets
˜Y ˜Z ˜[
˜W [
=0
˜[
Thus, the steady motion of a rapidly rotating ideal incompressible fluid is a superposition of
two independent motions: two-dimensional flow in the transverse plane and an axial flow
independent of the vertical axis z.
10. A two-dimensional fluid flow field is given by the streamfunction y = x y. (a) Show that the
fluid flow is irrotational. (b) Find the velocity potential f, and (c) verify that the
streamfunction and velocity potential satisfy the Laplace equation.
Solution:
(a) The velocity components are related to the streamfunction and velocity potential in
two-dimensional flows as
˜G ˜Z ˜G ˜Z
V   W
˜Y ˜Z ˜Z ˜Y
Hence u = –x, and v = y. The curl of the velocity field, the latter given by ui + vj can
be shown to be zero and hence the fluid flow is irrotational.
˜G ˜Z Y

(b) Since   Y Integrating one gets G   G Z

˜Y ˜Z 

˜G ˜Z Z

Since Z Integrating one gets G  G Y

˜Z ˜Y 
 
Z Y
Hence G Z
 DPOTUBOU G  Y
  DPOTUBOU
 
 
Y Z
Hence G    DPOTUBOU
 
The Laplace equation in terms of y and f are Ñ2y = 0 and Ñ2f = 0. Substituting the
expression for the strem function and velocity potential it is clear that both y and f
satisfy Laplace equation identically.
11. Neglecting body force, derive the equation for the velocity potential for steady barotropic
irrotational flow.
212 u BASICS OF ATMOSPHERIC SCIENCE

˜G
Solution: For steady flow, the continuity equation with W  is
˜Y
J
J

˜G ˜S ˜G
S  BOE
˜Y ˜Y J
˜Y ˜Y J J J

the equation of motion is


˜G ˜G  ˜Q  EQ ˜S

˜Y ˜ Y ˜ Y
K K J
S ˜Y J
S E S ˜Y J

EQ
Let D  the local sound speed, one gets
ES
 ˜S  Ë ˜G ˜G Û
  Ì Ü
S ˜Y J D ÍÌ ˜Y ˜Y ˜Y ÜÝ
K K J

Substituting the above equation in the continuity equation, one gets

˜G ˜G ˜G ˜G
  D 
˜ Y ˜Y ˜ Y ˜ Y
J K
˜Y ˜ Y
K J J J

E[
12. From the full vorticity transport equation, <³W >[  O³ [ derive the vorticity transport
EU
equation in two-dimensional flows.
Solution: For the two-dimensional flows, let the vertical component of velocity be v3 = 0
and the remaining two velocity components become a function of x1 and x2, only. Then the
vorticity vector has only the vertical component to be non-vanishing and is given by

Ë ˜W ˜W Û
[ Ì  Ü F [ F
Í ˜Y ˜Y Ý
For the above two-dimensional flows it turns out that the vector [Ñv] z is zero and hence the
vorticity transport equation in two-dimensional flows reduces to the scalar equation

E[
W³[
EU

REVIEW QUESTIONS
1. Is the Coriolis parameter same as the planetary vorticity. Explain.
2. Why is counterclockwise rotation in northern hemisphere said to have positive vorticity?
3. What is the difference between relative vorticity and absolute vorticity?
4. In a typical trough and ridge pattern, can one identify zones of negative, positive and zero vorticity?
5. Write down the most general relationship between vorticity and circulation.
6. Why do cities near large bodies of cold water in summer experience well-developed sea
breezes, but only poorly-developed land breezes. Why?
ATMOSPHERIC MOTION u 213

7. What are the advantages of using isobaric coordinate system over height coordinate system?
8. Let the air whirl at 90 m s-1 around a tornado having a radius of 120 m. What would be the
slope of an isobaric surface associated with the tornado?
9. What is the difference between a barotropic fluid and a baroclinic fluid?
10. What is potential vorticity?
11. What is Taylor Proudman theorem?
12. The tendency of cyclones to move poleward is known as beta drift. Suggest the basis for the
above tendency.
8 Atmospheric
Boundary Layer

The layer of the atmosphere close to the earth surface where viscous forces are important is
defined as the Atmospheric Boundary Layer (ABL). The atmospheric boundary layer plays a
very important role in the transport of momentum, mass (moisture) and energy through
processes that are inherently turbulent. The turbulent flow in the atmospheric boundary layer
can effectively transport mass, momentum and energy through turbulent eddies; a concept
which was first introduced by Ludwig Prandtl.

8.1 BRIEF CONSIDERATION


It is possible to investigate the transport phenomena based on the mechanism of the free path.
This would eventually mean that a molecule traversing a mean free path l is transporting its
momentum, energy and mass over a distance l.

8.2 DEFINITION OF VISCOSITY


Consider a fluid layer confined between two horizontal plates as shown in Figure 8.1.
t

FIGURE 8.1 Fluid shear between two parallel plates.

214
ATMOSPHERIC BOUNDARY LAYER u 215

Let the bottom plate be fixed and the upper plate be set in motion with a constant speed
V (refer Figure 8.1) in the positive x direction. It is clear that a certain force needs to be
exerted upon the upper plate to keep it moving with constant speed V. Experiments indicate
that the above force is directly proportional to the area of the upper plate and to the constant
speed V and is also inversely proportional to the distance z between the two plates, i.e. F µ A,

F µ V and ' — i.e. ' —
"7 Also, the fluid layer near the upper plate will move with

[ [
speed V along the positive x direction, while the fluid layer near the lower plate has zero
speed. The fluid layer between the upper and the lower plates will all move in the positive x
directions with speeds which will vary from V to zero as seen in Figure 8.1. A steady state
situation is defined as one in which the fluid characteristics (fluid properties) do not change
locally with time. In such a steady state situation, the variation of fluid speed with height is
found to be linear.
The applied force can be written as
N "7
' (8.1)
[
where the constant of proportionality is called the dynamic coefficient of viscosity. In the
steady state situation, the applied force is proportional to the vertical shear of the velocity and
from Eq. (8.1) one can write
˜7
U N (8.2)
˜[
where t is the stress or the applied force per unit area. Equation (8.2) was suggested by
Newton and is known as Newton’s law of viscosity. All fluids, which satisfy Newton’s law of
viscosity, are known as Newtonian fluids. Unlike the case of solids where stresses are directly
proportional to the strain, Newton’s law of viscosity suggests that the stress is directly
proportional to the rate of strain for fluids. A fluid is considered non-viscous (or inviscid) if
its dynamic coefficient of viscosity is zero. An ideal fluid is one which is both inviscid and
incompressible.

8.3 EXPRESSION FOR VISCOSITY FROM KINETIC THEORY


Maxwell, in 1860 successfully explained viscosity on the basis of kinetic theory and obtained
an expression for the same. A simple derivation for viscosity, which is correct except for a
factor, is as follows:
Consider a gas moving parallel to the horizontal plane in the positive x direction. To
bring out the effects of viscosity alone, let the gas have uniform temperature and
concentration. Assume that the mass velocity u0 in the positive x direction increases with
increasing z, where z-axis is perpendicular to the xy plane. The molecules of the gas above
any reference plane z = z0, possess an average x momentum greater than the molecules below
it. Since there is no mass motion along the z-axis, the number of molecules per unit volume
traversing upward or downward in the z direction and crossing the reference plane z = z0 is
the same. This means that there is always a greater transport of x momentum downwards
216 u BASICS OF ATMOSPHERIC SCIENCE

whenever molecules cross the reference plane from both sides. Hence any particular fluid
layer gains in x momentum and thereby experiences an accelerating viscous force parallel to
the direction of mass motion.
Let it be assumed that every molecule traverses a distance l equal to the mean free path
before it suffers a collision. Since the molecules can move in all possible directions, i.e. with
all possible inclinations to the vertical z-axis, to make our discussion simple we shall assume
M
that the average projection of the mean free path on the z-axis is half the mean free path, i.e. 

That is, any molecule crossing the reference plane z = z0, is assumed to have suffered its
M
previous collision at a distance both above and below the reference plane, and also

contributes the momentum appropriate to its original plane to the reference plane. For any
layer, one can denote V V where V is the component velocity averaged over all the
EV
molecules in that layer. If the vertical velocity gradient is  the difference in the mean
E[
M M EV
molecular velocity across two planes which are distant from one another is  It is
  E[
assumed that l is much smaller as compared to the scale of variation of u0, and hence one can
very well retain only up to the linear term in the Taylor series expansion. If m is the mass of
M EV
the molecule, the excess of momentum transported downwards is m  From kinetic
 E[
theory of gases, it is known that for an ideal gas in equilibrium, the number of molecules
M
crossing per unit area of any layer per unit time is OD where n is the molecular density and

D is the mean molecular velocity of the gas. Hence, the excess of positive momentum
transported downwards across unit area per unit time of the reference plane z = z0 is
 EV
NO D M   Arguing on similar lines it is clear that an equal number of negative momentum
 E[
is transported upwards across unit area per unit time of the reference plane. Hence the net
 EV
transfer of positive momentum downwards equals to NO D M  and this will exert an
 E[
accelerating force on the lower layer. This accelerating force from Newton’s law of viscosity
EV
equals N and hence equating, one gets
E[
 EV EV
NO D M  = N  
 E[ E[

N NO D M =  S D M (8.3)
 
where r is the density of the gas. The rigorous derivation for the dynamic coefficient of
 
viscosity has the same form as Eq. (8.3) except for the factor replacing  It is clear that
 
ATMOSPHERIC BOUNDARY LAYER u 217

 M
if the average projections of the mean free path were taken as M BOE OPU one would have
 
obtained the correct expression for the dynamic coefficient of viscosity. From kinetic theory,
one gets an expression for the mean free path,

M (8.4)
 O QT 
where s is known as the collision cross-section or the sphere of influence. Combining
Eqs. (8.3) and (8.4), one gets

N
ND
(8.5)
  QT 

Since D — 5 only, the dynamic coefficient of viscosity is independent of the pressure or the
density of the gas and depends only on the temperature. The above dependence was
theoretically predicted by Maxwell, and remains one of the first successes of the kinetic
theory.

8.4 VISCOUS FORCES IN THE EQUATION OF MOTION


Consider a parallelepiped of fluid of lengths z
tzz
dx, dy and dz as shown in Figure 8.2.
It is to be noted that typically nine ∂t zy
t zy + dz
kinds of stresses manifest with three of ∂z
them being normal, while the remaining six
are tangential stresses. For simplicity and
con-venience, the first subscript of the stress
shall indicate the nature of the face of the
x
ty

dz
parallelepiped on which the stress acts or y
rather the perpendicular axis to the face of
dx
the parallelepiped, while the second tzy
subscript denotes the direction of the fluid dy
motion producing the stress. The nine x
possible stresses constitute a stress tensor FIGURE 8.2 Stresses (both tangential and
and is usually denoted as follows: normal) on a fluid cube.

txx txy txz


tyx tyy tyz
tzx tzy tzz

The above stress tensor can be shown to be symmetrical, tyx = txy; tzx = txz; tzy = tyz, with
the result that there exist only six independent stresses for us to work on. When the fluid is in
a state of rest or under rigid body motion, the state of stress for a fluid is characterized by
absence of shearing stresses (txy = txz = tyx = tyz = tzx = tzy = 0) and the normal stresses
(txx = tyy = tzz = –p), where the scalar p is the magnitude of the compressive normal stress
218 u BASICS OF ATMOSPHERIC SCIENCE

and is called as hydrostatic pressure. In this situation, the stress tensor components are given
by ti,j = –pdij where dij is the Kronekar delta (di,j = 1, if i = j, while ti,j = 0 otherwise). For a
fluid in general body motion, the stress tensor components are given by t ¢i, j where t ¢i, j is
known as the deviatoric viscous stress tensor and is related to the rate of deformation tensor
 È ˜W ˜W Ø
Di, j. For a Newtonian fluid, the rate of deformation tensor is given by %  J K

É Ù
 Ê ˜Y ˜Y Ú
J K
K J

where vi and vj are velocity components and xi and xj are coordinates with i and j varying
from 1 to 3. For a Newtonian fluid, the deviatoric viscous stress tensor is related to the rate of
deformation by the relation
t i¢,j = l (D11 + D22 + D33) di,j + 2m Di, j

where m is the coefficient of viscosity and l is called the second coefficient of viscosity. For
an incompressible fluid, D11 + D22 + D33 is zero at all times and so the total stress tensor
assumes the form
t ¢i,j = –p di,j + 2 m Di,j

We shall utilize Figure 8.2 to calculate the net viscous force on an element of fluid by
considering the stresses on the faces perpendicular to the z-axis due to motion along the x
direction. The stress across the bottom z face by the fluid element below the bottom z face is
tzx. The tangential stress exerted across the upper z face by the fluid above the top z face can
be obtained by expanding the same in Taylor series expansion and retaining only up to the
linear term. The above approximation is valid if the parallelepiped can be assumed to be very
small. Hence the total force on the parallelepiped is the difference between the two stresses

È ˜U [Y Ø ˜U [Y
ÉÊ U [Y  E[ Ù EY EZ  U [Y EY EZ EY EZ E[
˜[ Ú ˜[
The force per unit mass due to the above stress is

 ˜U [Y  ˜ È ˜V Ø
N (8.6)
S ˜[ S ˜[ ÉÊ ˜[ ÙÚ
Using analogous arguments, it is possible to obtain the force per unit mass due to the other
stresses and using Newton’s law of viscosity, and assuming m is a constant, one gets the
viscous force per unit mass as
N È ˜  V ˜  V ˜ V Ø
x direction:   (8.7)
S ÊÉ ˜Y  ˜Z  ˜[  ÚÙ

N È ˜ W ˜  W ˜ W Ø
y direction:   (8.8)
S ÉÊ ˜Y  ˜Z  ˜[  ÙÚ

N È ˜ X ˜  X ˜  X Ø
z direction:   (8.9)
S ÉÊ ˜Y  ˜Z  ˜[  ÙÚ
ATMOSPHERIC BOUNDARY LAYER u 219

The above terms are to be added to each of the equations of motion expressing conservation
N
of momentum in all the three directions. S is called the kinematic coefficient of viscosity
and is denoted by n. The above viscous force together with the pressure gradient force can
 È ˜U Ø
JK
be directly arrived by noting that the surface fluid force is given by É Ù
S Ê ˜Y Ú
K

8.5 TURBULENCE
We encounter turbulent flow in our day-to-day experience. If one opens a kitchen tap just a
little bit, one finds the flow of the water to be smooth and regular, called the laminar flow. If
the tap is opened a little further, the flow of water becomes more irregular or sinuous and is
called the turbulent flow. Again close to the top of a burning cigarette, the smoke emanates in
a regular laminar flow, while high above the burning cigarette the flow of the smoke is
irregular and highly diffusive indicating the manifestation of turbulence. Turbulence is
composed of eddies, which are nothing but patches of zigzagging, and very often swirling
(rotating) fluid, moving randomly around and about the overall direction of motion.
Turbulence has been one of the most difficult problems in classical physics and despite the
efforts of the finest of hydrodynamicists for more than a century, it is still an unsolved
problem. Richard Feynman, the celebrated physicist called turbulence as, the most important
unsolved problem of classical physics. In fact, an anecdote attributed to the physicist Horace
Lamb, captures the formidable difficulty attributed to turbulence. During an address in 1932
to the British Association for the Advancement of Science, Horace Lamb reportedly
mentioned, I am an old man now, and when I die and go to heaven there are two matters on
which I hope for enlightenment. One is quantum electrodynamics, and the other is the
turbulent motion of fluids. And about the former, I am rather optimistic. Osborne Reynolds, a
British engineer performed very systematic and careful experiments on fluid motion to unravel
the facets of turbulence. Reynolds suggested that the regular laminar fluid motion changes to
the highly irregular turbulent flow when the Reynolds number of the flow, exceeds a critical
value. The Reynolds number is the ratio of the inertial force to the viscous force and is given by
6  - 6-

(8.10)
O6 -
3F
 O
where U is the characteristic velocity scale, L is the characteristic length scale and n is the
kinematic coefficient of viscosity of the fluid. For example, in the flow of water in a circular
pipe, the diameter of the pipe and the velocity of water at the centre of the pipe will serve as
the characteristic length and the characteristic velocity. The transition to turbulence typically
occurred in the Reynolds experiments when the Reynolds number exceeded the value of
2300, although it is possible to delay the onset of the transition to turbulence by ensuring that
the roughness associated with the surfaces of the circular pipe (in Reynolds experiment) is
kept at the minimum. Turbulent flows exhibit random velocity fluctuations and the patterns of
turbulent flow do not repeat themselves. Just observing the smoke emanating out of a stack
height can convince us of the above features of turbulent flow. Although it is not
220 u BASICS OF ATMOSPHERIC SCIENCE

straightforward to provide a correct definition of turbulent flow, it turns out that it is fairly
easy to outline the chief characteristics of the turbulent flow. The fundamental characteristic
of the turbulent flow is its irregularity. Also, turbulent flows can be very effective (much
more as compared to the molecular transport) in transporting momentum, energy and mass.
Turbulent flows are also highly diffusive and are composed of eddies or vortices in a very
broad range of sizes. These eddies are continually forming and breaking down. Large vortices
break down into smaller ones, which break down into yet smaller vortices, and so on,
ultimately the smallest eddy would dissipate into heat due to viscosity. The celebrated
meteorologist Lewis F. Richardson described the above process in poetic verse as follows:
Big whorls have little whorls,
Which feed on their velocity,
And little whorls have lesser whorls,
And so on to viscosity.
Turbulence is not something which needs to be conveniently avoided because of its
formidable difficulty. In fact, many engineers work hard to increase turbulence in real
situations. For example, the mixing of fuel and oxidizer in the cylinders of an internal-
combustion engine can provide for enhanced mixing and can produce a cleaner and more
efficient combustion if the flow were to be turbulent. The Reynolds number can also be
defined as the ratio between the diffusion time due to the turbulence and the time scale due to
molecular diffusion as can be seen from the following discussions.

8.6 TURBULENCE AND DIFFUSION


Consider a room with dimensions of the order of L having a radiator (heat source). Assume
that the air is at rest and the transfer of heat is effected only by the molecular diffusion which
is governed by the diffusion equation
˜5
% ³ 5 (8.11)
˜U
where D is the diffusion coefficient, T is the temperature and Ñ2 is the Laplacian operator.
Since the diffusion coefficient D and the kinematic coefficient of viscosity n have the same
order of magnitude, one can write Eq. (8.11) as
˜5
O³5 (8.12)
˜U
Dimensional analysis of Eq. (8.12) provides for the characteristic time scale of diffusion

-
UN (8.13)
O
where tm represents the time it would take for the heat to be transported across the room by
molecular diffusion. Assuming L ~ 5 m, and n = 2 ´ 10–5 m2 s–1, one obtains a characteristic
diffusion time scale of 14–15 days, underlying the very slow and ineffective means of
transporting heat through molecular diffusion.
ATMOSPHERIC BOUNDARY LAYER u 221

A more effective and alternate mechanism for the transport of heat is as follows:
The radiator (heat source) may heat the air in contact with it and the associated buoyant
circulation can then have a characteristic time scale given by
-
U (8.14)
U
6
where L and U are the associated characteristic length and velocity scales. The order of
magnitude of the characteristic velocity scale U can be calculated as follows:
Let us assume that the heated layer has a thickness of h = 0.1 m. For a temperature
H '5
change DT, the acceleration of an air parcel is  If h is the thickness of the heated layer,
5
HI'5
the energy gained per unit mass of the heated layer is  Assuming DT = 10 K, T = 300 K
5
and h ~ 0.1 m, one obtains an energy of the order of 0.0326 m2 s–2. If this is equated to the
kinetic energy per unit mass, then the velocity has an order of 0.255 m s–1. This gives from
Eq. (8.14) tt ~ 20 s. The average velocity U in fact is somewhat lower than the value
indicated above and is only of the order of a few cm s–1. This would correspond to a
characteristic time scale for turbulence diffusion to be of the order of a few minutes. From
Eqs. (8.13) and (8.14), it is clear that the Reynolds number Re can be defined as the ratio
between the molecular diffusion time scale to the turbulent diffusion time scale.

3F
U N 6-
(8.15)
OUU
The Reynolds number for the above example is of the order of 62500. The reason for the
generation of turbulence at large Reynolds number can be envisaged by the following
discussion. Due to the viscous nature of the fluid, there is no relative motion between the
fluid layer close to the surface at rest and at the surface—the so called no slip condition. At
high Reynolds number, the viscous forces are effectively confined to a thin layer close to the
boundary called the boundary layer. Outside the boundary layer the fluid velocity reaches its
free stream value. The velocity shear in the boundary layer is a source of vorticity and any
vorticity produced is diffused in the fluid only if the molecular diffusion time is less than the
turbulence diffusion time (i.e. tm < tt). Under large Reynolds number when the above
condition no longer holds, the vorticity remains confined in the shear layer which can later
detatch from the surface and generate turbulence.

8.7 EQUATIONS OF MEAN MOTION IN TURBULENT FLOW


Since the chief characteristic of turbulent flows is its irregularity, it is not surprising to
observe that any fluid variable (for example, the velocity components at any given point in
space) in turbulent flows shows irregular and widely fluctuating behaviour. However, careful
observations of the velocity components with time reveal that the turbulent flows have
underlying the irregular fluctuations, a component which is less rapidly varying with time.
Hence, if one were to average the velocity components over a certain period of time, it is
clear that the averaged velocity component will be a relatively more slowly varying function
222 u BASICS OF ATMOSPHERIC SCIENCE

of time than that of the velocity itself. One could define a mean velocity for the x component
of velocity as follows:
U 5 

V
5 Ô VEU (8.16)
U 5 

The averaging time T is chosen to be such that it is long enough so that the underlying widely
fluctuations are smoothed out and only the slowly varying component is observed. Hence, one
can write the instantaneous x component of velocity as the sum of the average x component
of velocity V and a widely fluctuating component u¢, which is considered as a departure from
the mean,
V V  V„ (8.17)
It is of course quite obvious for a given chosen interval T, the average of the departure will
vanish, i.e. V „  Since the averaged quantities are slowly varying in time and do not exhibit
the marked fluctuation, the averaged quantities would be more amenable to analysis. Hence,
one needs to obtain the equation of the averaged (mean) motion in turbulent flows. For
simplicity, we shall derive the equations of mean motion for a turbulent incompressible fluid.
The momentum conservation equation in the x direction, in the absence of molecular viscous
force terms is
˜V ˜V ˜V ˜V ˜Q
S  SV  SW  SX S GW  (8.18)
˜U ˜Y ˜Z ˜[ ˜Y
Multiplying the equation of continuity with r u, one gets
˜V ˜W ˜X
SV  SV  SV  (8.19)
˜Y ˜Z ˜[
Summing the above two equations, one gets the x momentum conservation equation in the
flux form
˜ SV
˜ SVV
˜ SVW
˜ SVX
˜Q
   S GW  (8.20)
˜U ˜Y ˜Z ˜[ ˜Y
The mean equation for the x momentum conservation equation follows from applying
Eq. (8.17) on Eq. (8.20). For example, the term

VV V  V „
V  V „
V V  V „V „ (8.21)

since the mean of the departures vanishes. Terms such as V „V „ do not vanish since there is a
non-zero correlation between the two departures u¢ and u¢ whose product is averaged. Hence
the mean x component momentum equation becomes
˜ ˜ ˜ ˜
SV
 SV V
 S V W
 S V X

˜U ˜Y ˜Z ˜[
˜Q ˜ ˜ @@@@@ ˜
SG W   S V „V „
 S V „W „
 S V „X „
(8.22)
˜Y ˜Y ˜Z ˜[
ATMOSPHERIC BOUNDARY LAYER u 223

Equation (8.22) can be written in its non-flux form by subtraction of Eq. (8.22) with the
equation obtained by multiplying V with the mean equation of continuity,
˜V ˜W ˜X
V V V  (8.23)
˜Y ˜Z ˜[
Subtraction of Eq. (8.23) from Eq. (8.22) yields

˜V ˜V ˜V ˜V  ˜Q  Ë ˜ ˜ ˜ Û
V W X GW   Ì S V „V „
 S V „W „
 S V „X „
Ü (8.24)
˜U ˜Y ˜Z ˜[ S ˜Y S Í ˜Y ˜Z ˜[ Ý
The y momentum conservation equation follows from similar arguments and is

˜W ˜W ˜W ˜W  ˜Q  Ë ˜ ˜ ˜ Û
V W X GV  Ì QW „V „
 QW „W „
 QW „X „
Ü (8.25)
˜U ˜Y ˜Z ˜[ S ˜Z S Í ˜Y ˜Z ˜[ Ý
Equations (8.24) and (8.25) are the horizontal equations of motion for a turbulent flow of an
incompressible fluid. Equation (8.24) is similar to Eq. (8.18) except for the fact that the
former equation refers to mean quantities and also due to the appearance of three additional
terms in the right-hand side of Eq. (8.24). These additional terms in Eq. (8.24) are similar to
the viscous stress terms due to molecular viscosity described earlier in this chapter and hence
are known as Reynolds stress or eddy stress terms. Since we had disregarded the molecular
viscous terms while deriving the mean equations of motion, they are not present in Eqs. (8.24)
and (8.25). However, it turns out the eddy stress terms are very much larger than the
molecular stress terms that the latter stress terms can be conveniently neglected. The
additional eddy stress terms, unless expressed in terms of other dependent variables pose
problems regarding the mathematical closure of the system of equations. If one sets out to
write the dynamical equations for the eddy stress terms (i.e. for the second-order
correlations), one finds that the terms with higher moments (i.e. third-order correlations such
as V „V „X „ ) appear in the equation and if the dynamical equations for the third-order
correlations were written, one finds that fourth-order correlation terms appear in them. It
appears that there is no end to this dilemma and for this reason the above difficulty is known
as the closure problem. By writing the eddy stress terms in terms of the other dependent
variables, one can successfully close the system at the first stage itself; this is known as the
first- order closure. It is pertinent to note that the eddy stress terms have appeared due to the
averaging of the nonlinear terms of the original equations.

8.8 MIXING LENGTH


Consider the situation as depicted in Figure 8.1 and assume that the fluid motion is turbulent.
It is clear that the force required to keep the upper plate in motion is much more than the case
for orderly laminar flow. Since the turbulent flow is much more effective in transporting
momentum, energy and mass as compared to the molecular transport, it is convenient to
assume that the above-mentioned properties are transferred in the turbulent flow by bodily
movement of large masses of fluid conveniently defined as eddies. Prandtl used the molecular
224 u BASICS OF ATMOSPHERIC SCIENCE

transport of momentum as analogy to study the turbulent momentum exchange by eddies. It is


known that the molecules retain the properties of the level at which they originate, move a
distance equal to the mean free path, collide with another molecule and transfer their
properties (momentum, energy and mass) to their new levels. Prandtl proposed a similar
mechanism with eddies playing the role of the molecules and a new concept mixing length
corresponding to the mean free path. Let the mean turbulent motion be in the x direction with
˜V
speed increasing with [ !  Prandtl, in analogy with the molecular transport suggested
˜[
that an eddy carrying the average x momentum of its original level moves a distance vertically
l ′ producing a turbulent fluctuation u¢ and getting absorbed at its new level.
A mass of fluid originating at the level z + Dz with mean speed V[ '[ and having moved
downward to a reference level z where the mean speed is V[ the turbulent fluctuation
produced will be V „ V[ '[  V[  If Dz is small, then one can expand by Taylor series
˜V
expansion and retain only up to the linear term, then V „ ' [  However, the sign of Dz
˜[
depends upon whether the fluid mass originates above or below the reference level z. The
vertical displacement can be expressed as l¢ = –Dz since the eddy originating above z
(Dz positive) is displaced downward with l¢ being negative. Thus,
˜V
V„ M„ and
˜[
the eddy stress is
˜V
U [Y  SV „ X „ SX „ M „ (8.26)
˜[
When an eddy moves from one level to a second level, one can expect an equivalent fluid
mass to move away horizontally with same speed to make way under the requirement of mass
continuity. Hence the horizontal and vertical eddy velocities can be considered equal and so
| w¢| » | v ¢h | where v ¢h is the horizontal eddy velocity. Since we have assumed that the mean
motion is in the x direction, one can consider | w¢| » | u¢|. Hence,

˜V
X„ M„
˜[
where the absolute magnitude is considered since w¢ and l¢ have to have the same sign (eddy
moving downward will have w¢ < 0 and l¢ < 0). Equation (8.26) becomes
˜V ˜V
U [Y S M „ (8.27)
˜[ ˜ [
The transport of momentum in Prandtl’s mixing length theory is performed by eddies in the
turbulent flow and in analogy with Newton’s law of viscosity, one can write
˜V
U [Y "[Y (8.28)
˜[
ATMOSPHERIC BOUNDARY LAYER u 225

where Azx is the eddy exchange coefficient. Defining a root mean square mixing length

MY M „  Eq. (8.27) becomes


˜V ˜V
U [Y SMY (8.29)
˜[ ˜ [
The eddy exchange coefficient Azx can be written from Eqs. (8.28) and (8.29) as

˜V
"[Y SMY (8.30)
˜[
Relations similar to Eqs. (8.27) and (8.30) corresponding to the mean flow in the y direction
can be derived using similar arguments and are as follows:

˜W ˜W ˜W
U [Z SM Z  "[Z SM Z (8.31)
˜[ ˜[ ˜[
Although Prandtl’s mixing length hypothesis is in analogy with the molecular transport, it is
clear that the eddy stresses are expected to pose greater difficulties, since unlike the molecular
case, the eddy exchange coefficients themselves depend on the mean state of the motion.
Taylor later hypothesized that vorticity rather than the momentum is the property transferred
by turbulence.

8.9 SURFACE AND EKMAN LAYERS


The following subsections provide brief discussions on the surface layer and the Ekman layer.
While the chief characteristic of the surface layer is the constancy in the wind direction with
the flux values not changing markedly within it, the Ekman layer is characterized by the
change in the wind direction and marked decrease of the flux values within it. The surface
layer due to the above reason is also known as the constant flux layer.

8.9.1 Surface Layer


Surface layer is the lowest part of the atmospheric boundary layer which is strongly
influenced by surface temperature. We shall investigate the characteristics of the surface layer
along with the vertical variation of the mean wind in this layer. For simplicity, assume that the
flow close to the ground is parallel to the x-axis. The surface stress divided by the density of
È U [Y Ø
air has units of velocity squared and is defined as V  ÉÊ S ÙÚ where u* is known as the
T
friction velocity. Actual measurements close to the ground show that the surface stress in the
atmosphere has typical magnitudes of t ~ 0.1N m–2. This gives a value of the friction velocity,
u* = 0.3 m s–1. Applying a scale analysis for the horizontal equation of motion in x direction,
one finds that the Coriolis term has an order of magnitude of 10–3 (since f ~ 10–4 s–1 and
v ~ 10 m s–1). The pressure gradient force also has the same order (since Dp ~ 10 hPa ~ 103N
m–2 and Dx ~ 103 km ~ 106 m) as the Coriolis term. If these terms are to balance the eddy
stress term, it is necessary that
226 u BASICS OF ATMOSPHERIC SCIENCE

ÈU Ø
' É [Y Ù
Ê S Ú
…   N T 
'[
ÈU Ø
For Dz = 10 m, ' [Y …   N T o  Hence the change in the vertical eddy stress in the
ÉÊ S ÙÚ
lowest 10 m of the atmosphere close to the ground is less than 10% of the surface stress.
Hence it is plausible to assume that in the lowest metres of the atmosphere, the shear stress is
constant and is more or less equal to this surface value. From Eq. (8.28),
U [Y "[Y ˜V
V  (8.32)
S S ˜[
In deriving Azx, we assumed that the horizontal and vertical scales of the eddy are
approximately equal. Near the earth’s surface, the vertical eddy scale is limited by the
distance from the earth surface and the simplest assumption is to assume that the mixing
length is proportional to the height above the ground as lx = kz, where k is a constant called
the Von Karman constant. Using this in Eq. (8.30), one gets
˜V
"[Y S L[
 (8.33)
˜[
Substituting Eq. (8.33) in Eq. (8.32), one gets
˜V V
(8.34)
˜[ L[
Integrating Eq. (8.34) with respect to z, yields the logarithmic wind profile for the vertical
variation of wind with height
V È [Ø
V MO
ÉÊ [ ÙÚ (8.35)
L 

where z0 is called the roughness length and is the height where V vanishes. The value of the
roughness length depends on the roughness of the surface and may change from 0.5 cm for
smooth snow surface to values of the order of 4.5 cm for wheat field. Based on a variety of
experiments, the Von Karman constant k is found to have a value of 0.38. The logarithmic
wind law (Eq. (8.35)) is found to be in reasonable agreement with the observed variation of
wind in the surface layer under neutral conditions. When the lapse rate in the surface layer
departs from neutral conditions, a power law profile of the form
N
È [Ø
V V É Ù (8.36)
Ê[ Ú 

where 0 £ m £ 1, is a good approximation to the observations with m depending on the


stability. The exponent m has been found to decrease with increasing lapse rates with a value

of for the neutral case.

ATMOSPHERIC BOUNDARY LAYER u 227

8.9.2 Ekman Layer


At heights of 50 m or above, experiments have indicated that the turbulent stress is no longer
independent of height and actually decreases with increase in height. The turbulent stress
ultimately becomes very small at heights known as the ‘gradient level’ which is of the order
of 1000 m. Above the gradient level, it can be safely assumed that the flow will be
determined by the pressure gradient force and the Coriolis force. This part of the atmospheric
boundary layer above the surface layer and extending up to the gradient level is known as
the Ekman layer. We shall derive the expression for the vertical variation of the mean wind
in the Ekman layer. For this, we assume that the following assumptions hold good:
(i) The mean motion is horizontal
(ii) The horizontal mean wind shears are smaller compared to the vertical mean wind
shears
(iii) A three-way balance between the Coriolis, pressure gradient and the eddy viscosity
forces exists at every level in the Ekman layer.
(iv) The eddy exchange coefficient is independent of height.
The horizontal equations of motion are then given by

 ˜Q E V
0 = GW   "[Y  (8.37)
S ˜Y E[
 ˜Q E W
0 =  GV   "[Z  (8.38)
S ˜Z E[
The bars are not explicitly written for the above equations for convenience, although all the
variables are average quantities. Multiplying Eq. (8.38) by i, where i equal   and adding it
to Eq. (8.37), and using Azx = Azy = A, one gets

È ˜Q ˜Q Ø E
  JG V  JW
 É  J Ù  "  V  JW
(8.39)
Ê ˜Y ˜Z Ú E[
We shall assume that in the Ekman layer the pressure gradient and the density do not vary
appreciably with height. If the x-axis is so chosen to be parallel to the surface isobars with a
geostrophic wind ug in the positive x direction, we have
˜Q ˜Q
  S GVH (8.40)
˜Y ˜Z
Substituting Eq. (8.40) in Eq. (8.39), one gets

E JG
V  JW  V H
 V  JW  V H
 (8.41)

E[ "

Equation (8.41) is a second-order homogeneous linear differential equation. Since the vertical
extent of the Ekman layer is much larger than the surface layer, one can ignore the surface
228 u BASICS OF ATMOSPHERIC SCIENCE

layer and impose the boundary condition that the wind vanishes exactly at the earth surface
and equals the geostrophic value for large z. Thus, the boundary conditions are as follows:
z = 0; u + iv = 0 (8.42)
z ® ¥ u + iv = ug (8.43)
The general solution of Eq. (8.41) is
È JG Ø È JG Ø
V  JW  VH # FYQ É [
Ù  $ FYQ É  [
Ù (8.44)
Ê " Ú Ê " Ú
where B and C are arbitrary constants to be determined from the boundary conditions. Since
  J

J one can write Eq. (8.44) as



u + iv – ug = B exp[(z)] + C exp[–a(1 + i)z] (8.45)
G
where B  Applying the boundary condition Eq. (8.43) in Eq. (8.45) gives B = 0,
"
while applying the boundary condition Eq. (8.42) yields C = –ug. Hence the solution is
u + iv = ug {1 – exp[– z] exp [– i a z]} (8.46)
Applying the Euler’s formula exp [–i a z] = cos (a z) – i sin (az) in Eq.(8.46) and equating
the real and the imaginary parts, one gets
u = ug[1 – exp(–az) cos (az)] (8.47)
v = ug [exp (–az) sin (az)] (8.48)
It is clear that due to the presence of cos and sin terms, the wind direction will change
Q "
continuously with height and at a certain height [ Q the v component will vanish
B G
and u will equal to the geostrophic wind ug. At z = 0, the angle the wind makes with the
eastward direction which is determined by the tangent inversion of the ratio of v to u is
indeterminate. Applying the L’ Hospital’s rule, one finds that the ratio of v to u approaches +1
as z approaches zero as can be seen below
EW
MJN
E[ MJN DPT B[  TJO B[
 
[  EV [   DPT B[  TJO B[

E[
The above indicates that the wind direction at and close to the surface (z = 0) makes an angle
of 45° with the eastward direction and hence points towards the lower pressure. Thus, the
wind vector turns clockwise with increase of height (from an angle of 45° with the eastward
direction towards low pressure close to the ground to a direction which coincides with the
eastward direction) in the northern hemisphere. For the southern hemisphere, f is negative and
since the sine function is an odd function, the expression for v in Eq. (8.48) will have a
negative sign in the southern hemisphere. Again at z = 0, the ratio of v to u is indeterminate
ATMOSPHERIC BOUNDARY LAYER u 229

and again by applying L’Hospital’s rule, one finds that the ratio v to u approaches –1 as z
approaches zero. Thus, the wind direction at and close to the surface (z = 0) makes an angle
of 45° with the eastward direction (and 135° with the northward direction measured
clockwise) and hence again points towards lower pressure. Thus, the wind vector turns
anticlockwise with increase of height in the southern hemisphere. Observations indicate that
the gradient level is at a height of about 1 km and hence using f ~ 10–4 s–1, one gets a value
of A ~ 5 m2 s–1. This is about five orders of magnitude (105 times) larger than the molecular
kinematic viscosity value. The vertical variation of mean wind in the Ekman layer is clearly
brought out by means of a polar coordinate plot of the direction of wind tangentially against
the wind speed called the hodograph and is shown in Figure 8.3.

FIGURE 8.3 Ekman spiral. The velocity components are shown at different heights.

The above hodograph reveals a spiral called the Ekman spiral in honour of Ekman who
first obtained the analogous result for the surface layer of the ocean. Actual observations
indicate that a reasonable agreement exists with the theoretical Ekman spiral in the lowest
part of the Ekman spiral. However, in the case of strong cold or warm air advection, the
pressure gradient will no longer be constant with height in the Ekman layer and hence a
departure from the theoretical Ekman spiral cannot be ruled out. Another major assumption
made in the derivation, regarding the constancy of the eddy exchange coefficient within the
Ekman layer may not hold in the real atmosphere and can cause departures from the
theoretical Ekman spiral.

8.10 SECONDARY CIRCULATIONS AND SPIN-DOWN IN


THE ATMOSPHERE
One of the important results of the earlier section is that the wind within the atmospheric
boundary layer has a component directed towards lower pressure, i.e. to the left of the
geostrophic wind in the northern hemisphere. The above result leads to mass convergence in
low-pressure zones (cyclonic circulations), and mass divergence in high-pressure zones
(anticyclonic circulations). Considerations of mass conservation then yield vertical motion, out
of and into the boundary layer. It is possible to estimate the magnitude of these induced
230 u BASICS OF ATMOSPHERIC SCIENCE

vertical motion by noting that the cross isobaric mass transport per unit area at any level in
the atmospheric boundary layer is given by r0 v, if vg = 0. The net mass transport for a
column of unit width extending vertically from the surface to the top of the Ekman layer is given as
%F %F
. Ô SWE[ Ô SVH <FYQ  Q [ %F
> TJO Q [ %F
E[ (8.49)
 

where De is the depth of the Ekman layer. The above Eq. (8.49) can be put in relation with
the vertical velocity through the equation of continuity
˜ ˜ ˜
SX
 SV
 SW
(8.50)
˜[ ˜Y ˜Z
Assuming vg = 0, ensures that ug is independent of x and hence u is independent of x in the
above relation. Substituting in Eq. (8.49) and noting that w = 0, at z = 0, one gets, assuming
constant density, the following relation:
%F
SX
%F
˜ .
˜

Z Ô SVH FYQ< Q [ %F > TJO Q [ %F
E[
˜

Z
˜
(8.51)

The above equation shows that the mass flux of the boundary layer is equal to the
˜VH
convergence of the cross-isobaric mass transport in the layer. If the vorticity  [ H is
˜Z
assumed constant in the Ekman layer, then the integral in Eq. (8.51) has a constant value

  FQ
%F and the vertical velocity at the top of the boundary layer becomes
Q

Ë "Û
X% [H Ì Ü (8.52)
Í G Ý
F

where the following approximation (1 + e–p) = 1 has been used. Equation (8.52) provides the
important result that the vertical velocity at the top of the boundary layer is proportional to
the geostrophic vorticity. It is now clear from the above discussion that the effects of the
boundary layer fluxes are communicated directly to the free atmosphere through a forced
secondary circulation. The above forced secondary circulation is called as the boundary layer
pumping and often dominates over the turbulence mixing. It is important to note that the
boundary layer pumping occurs only in rotating fluids and provides one important distinction
between rotating and non-rotating fluid flows. For a typical synoptic scale system, zg = 10–5 s–1,
f = 10–4 s–1, De = 103 m. The vertical velocity at the top of the boundary layer has a value of
the order of a few millimetres per second. The above boundary layer pumping damps the
vorticity around the low- and high-pressure systems and the characteristic time scale of the
above damping can be easily calculated by considering a barotropic atmosphere. For a
synoptic scale motion, the vorticity equation can be written as

E[ H È ˜V ˜ W Ø È ˜X Ø
G É  G É Ù (8.53)
EU Ê ˜Y ˜Z ÚÙ Ê ˜[ Ú
ATMOSPHERIC BOUNDARY LAYER u 231

In Eq. (8.53), we have neglected zg as compared to f in the divergence term and have
neglected the latitudinal variation of the Coriolis parameter f. Since the geostrophic vorticity
is independent of height in a barotropic atmosphere, one can integrate Eq. (8.53) from the
top of the Ekman layer (z = De) to the top of the troposphere (z = H), to give

E[ H Ë X )
 X %F
Û
GÌ Ü (8.54)
EU Í )  %F
Ý
Substituting for the vertical velocity at the top of the boundary layer from Eq. (8.54), and
assuming that H >> De and the vertical velocity at the top of the troposphere vanishes,
one gets

E[ H G"
 [H (8.55)
EU ) 

Equation (8.55) gives a relation for the decaying time scale for the vorticity. Equation (8.55)
can be easily integrated in time to give
Ë U Û
zg (t) = zg (0) exp Ì Ü (8.56)
ÍU Ý

where zg (0) is the geostrophic vorticity at time t = 0, and U

) ÈÉ
is the time that it
Ê G" ÙÚ
takes for the vorticity to decrease to e–1 of its original value. The above damping time scale is
also called as the spin-down time. Considering typical values of H = 104 m, f = 10–4 s–1, and
A = 10 m2 s–1, one finds that the spin-down time is about 4 days. Hence for a mid-latitude
synoptic scale system in a barotropic atmosphere the characteristic spin-down time is of the
order of a few days. The above time scale is also roughly the lifetime of a cyclone.

8.11 SECONDARY CIRCULATIONS AND SPIN-DOWN IN


A TEACUP
A somewhat similar secondary circulation, i.e. boundary layer pumping is noticed when a cup
of tea is stirred. Again the boundary layer pumping is accompanied by a decay of the
circulation in the teacup as in the atmosphere. An approximate balance exists between the
radial pressure gradient and the centrifugal force away from the bottom and the sides of the
cup. It is to be noted that since water is an incompressible fluid, the radial pressure gradient is
independent of the depth. However, close to the bottom of the cup, the pressure forces are no
longer balanced by the centrifugal forces and this ensures that radial inflow takes place near
the bottom of the cup. This is easily observed by noticing the cluster of the tea leaves near the
centre at the bottom of the cup when stirred. From considerations of mass conservation, radial
inflow at the bottom of the cup must be accompanied by upward vertical motion and a slow
compensating outward radial flow throughout the remaining depth of the tea. This slow
outward radial flow approximately conserves angular momentum, and by replacing high
angular momentum fluid by low angular momentum fluid serves to spin-down the vorticity in
the teacup far more rapidly than could be possible by mere diffusion. The chief difference
232 u BASICS OF ATMOSPHERIC SCIENCE

between the secondary circulations seen in the atmosphere and the teacup is that in synoptic
scale motions it is the Coriolis force which balances the pressure gradient forces away from
the boundary in the atmosphere.
Consider a reference system with the z-axis coinciding with the vertical rotating axis and
the x- and y-axes being normal to the z-axis. Assume at a time t = 0, the angular velocity on
the lateral wall of the cup changes by a small quantity e W either in acceleration (spin-up) or
deceleration (spin-down). The pressure forces are no longer balanced by the centrifugal forces
at the bottom of the cup resulting in a radial inflow with a velocity of the order of e Wr. The
above radial inward motion occurs within a boundary layer whose thickness is of the order of
 
È W Ø
E ÉÊ Ù  Equating the Coriolis acceleration terms to the viscous force terms in the
: Ú
horizontal momentum equations, one arrives at the above order of boundary layer thickness.

È ˜ W Ø È ˜ V Ø
:V O É  Ù   :W O É  Ù (8.57)
Ê ˜[ Ú Ê ˜[ Ú
Here n is the kinematic coefficient of viscosity and W is the angular velocity (different
notation used in Chapters 6 and 7), u and v are the horizontal velocity components. We have
assumed that u ~ v within the boundary layer to arrive at the order of the boundary layer
thickness. One can arrive at an estimate of the time it takes to establish the boundary layer by
performing a scale analysis of the equation of motion
EW È ˜ W Ø
 :V O É  Ù (8.58)
EU Ê ˜[ Ú
Since the velocity component is of the order of e Wr, one finds from Eq. (8.58) that the time
to establish the boundary layer is of the order of W–1.
Writing the continuity equation in the radial plane, one gets
˜X ˜WS
 (8.59)
˜[ ˜S
where vr is the radial velocity. Integrating Eq. (8.59) across the boundary layer, one gets
E 
˜W S È O Ø
X Ô E[  F : É (8.60)
˜S Ê : ÙÚ


Within the fluid, one can determine the velocities by integrating the continuity equation and
applying the boundary conditions. Integrating the continuity equation, and noting that the
radial velocity is independent of depth, one gets
È ˜W Ø
X  É S Ù [  DPOT (8.61)
Ê ˜S Ú
where cons is a constant of integration. Assuming that the fluid is contained between the free
surface z = H and the bottom z = 0, one gets from Eq. (8.60), that
ATMOSPHERIC BOUNDARY LAYER u 233


È O:Ø
"U [  X FÉ
Ê  ÚÙ
BOE
at z = H, w = 0 (8.62)
It is to be noted that the vertical velocity is zero at the free surface outside the bottom
boundary layer. Using Eq. (8.62), in Eq. (8.61), one gets the constant of integration and
substituting this the expression for the vertical velocity and the radial velocity gradient

ÈO :Ø Ë [ Û
w = FÉ
Ê  ÙÚ ÌÍ  ) ÜÝ

ÈO : Ø
ËSÛ
vr = F É (8.63)
Ê  ÙÚ
ÌÍ ) ÜÝ
It is pertinent to note that the vertical velocity is independent of the radius. The spin-down
time for the teacup can be estimated from the following argument:
It is clear in the spin-down case, the secondary circulation transports fluid with lower
angular momentum to regions of high angular momentum. Considering a ring of fluid of
radius r and mass d m the angular momentum is then given by J = d mWr2. Neglecting the
friction within the fluid, conservation of angular momentum requires that the following
relation hold, i.e.
': 'S F :
  (8.64)
: S :
The radial distance travelled in the spin-down time will be

FS ÈO : Ø È SØ
'S WS U FÉ Ù ÉÊ ÙÚ U (8.65)
 Ê  Ú )

where t is the spin-down time scale. From Eq. (8.65), the spin-down time scale will be

È ) Ø
U É Ù (8.66)
Ê O : Ú
Comparisons of Eqs. (8.66) and (8.56) indicate the analogous nature of the secondary
circulation between the atmosphere and the teacup.

SOLVED EXAMPLES

1. An anemometer located at 10 m above the ground measures a wind speed of 6 m s-1 within
an orchard on an overcast day. Assuming a roughness length of 0.3 m, find the wind speed
at a height of a 25 m smoke stack.
Solution: Since conditions are overcast, one can safely assume neutral stability conditions
to prevail and hence use the logarithmic wind profile
MO  N   N

V
[ N V
[ N  N T
MO  N   N

234 u BASICS OF ATMOSPHERIC SCIENCE

2. Given the following instantaneous temperature measurements T at a place, find the average
temperature and the departure of the temperature from the average, i.e. T¢.

Time (minute) Temperature (°C)


1 12
2 14
3 10
4 15
5 16
6 13
7 10
8 11
9 9
10 10

Solution: Adding the given 10 temperatures and dividing by 10 gives the average
temperature, i.e. 120 divided by 10 equals 12. Hence average temperature is 12°C.
Subtracting the above average temperature from each instantaneous temperature gives T¢
and is given as

Time (minute) Departure from the average


temperature T¢ (°C)

1 0
2 2
3 –2
4 3
5 4
6 1
7 –2
8 –1
9 –3
10 –2

3. Find the drag coefficient in statically neutral conditions with the wind speed at 10 m having
a value of 6 m s–1 over the village environment. Assume the roughness length to be 1.0 m.
Find also the friction velocity and the surface stress.
Solution: The expression for drag coefficient under statically neutral air can be related to
the aerodynamic roughness length z0 as follows:

L
$% 
MO [ 3  [

ATMOSPHERIC BOUNDARY LAYER u 235

where k is the Von Karman constant, (k = 0.4) and zR is a reference height usually taken as
10 m.



$% 

MO 

The square of the friction velocity is then given as

u* 2 = CD (u)210 m = (0.03) (6) (6) = 1.086 m2 s–2


The friction velocity is then 1.04 m s–1
The surface stress can be determined from the square of the friction velocity assuming a
value of density of air, r = 1.2 kg m–3

t = ru* 2 = (1.2) (1.086) = 1.30 Pa


4. Repeat the above problem for the case of grass prairie having a roughness length of 0.03 m.
Solution: The drag coefficient in this case becomes



$% 

MO  

The square of the friction velocity becomes

u* 2 = CD(u)210 m = (0.0047) (6) (6) = 0.17 m2 s–2


The friction velocity is then 0.41 m s–1
The surface stress becomes

t = ru* 2 = (1.2) (0.17) = 0.20 Pa


5. Find the friction velocity, the velocity standard deviation, and the turbulent kinetic energy in
a statically stable air at a height of 50 m in an atmospheric boundary layer of thickness 250 m.
Assume the drag coefficient is 0.002 and the wind speed at a height of 10 m is 6 m s–1.
Solution: The square of the friction velocity is then given as
u* 2 = CD(u)210 m = (0.002) (6) (6) = 0.072 m2 s–2; u* = 0.27 m s–1
For statically stable air, standard deviations in an atmospheric boundary layer of depth h
have been empirically found to vary with height z as
su = V <  [I
>
sv = V <  [I
> 

sw = V <  [I


>

where u* is the friction velocity.


Using the above relations, one gets the velocity standard deviations as

su = V <  [ I
>  
<  
>  N T 
236 u BASICS OF ATMOSPHERIC SCIENCE

sv = V <  [ I
>  
<  
>  N T 
sw = V <  [ I
>  
<  
>  N T 
The turbulent kinetic energy is then given by

5,& <T V  T W  T X > < 


  
  
 >  NT 
6. Within the lowest one kilometre in middle latitudes, the cross-isobaric angle (angle between
the isobar and the wind vector) is about 20°. Assuming that the frictional forces have a
linear form –a v, and that the other forces do not exist, estimate the time required for the
above friction force to reduce the horizontal velocity by a factor of e.
Solution: In the absence of other forces, the horizontal equation of motion reduces to
EW
 BW
EU
A solution of the above equation is v = v0 e–a t, where v0 is the velocity at the initial time
t = 0. The time required to reduce v by a factor e is a–1. From the balance of forces in the
boundary layer, with the Coriolis force C and frictional force F together balancing the
pressure gradient force P and the cross-isobaric angle y is the angle between the isobar and
the wind vector. From the above balance, one notes that
| F | = | P| sin y, and
| P| cos y = | C| = f |V|
Combining the above two relations, one gets
| F | = f |V| tan y
From the above, it is clear that a = f tan y. Hence the time required for the horizontal
velocity to reduce by a factor e is

 
 –   T
G UBO Z  –  –   UBO ’

7. Consider the laminar flow of a viscous Newtonian fluid with constant coefficient of viscosity
m (m = 1 mPa s). The velocity components are given as
vx = –x –y; vy = y – x; vz = 0.
For a plane whose normal is in the x-axis, find (i) excess of the total normal compressible
stress over the pressure p, and (b) the magnitude of the shearing stress.
Solution:
(i) From the given velocity fields, it is clear that the velocity divergence (vx + vy + vz = 0)
and so the normal stress along x-axis is t11 = – p + 2 m D11.
Ë ˜W Y Û
We know % Ì ˜Y Ü   in this case. The excess of the total normal compressible
Í Ý
stress whose normal is in x-axis over the pressure p is – (t11) – p = –2m D11 = –2 ´ 1
´ –1 = 2 m Pa.
ATMOSPHERIC BOUNDARY LAYER u 237

Ë ˜ W Y ˜W Z Û
(ii) t12 = N % NÌ  Ü  N   N1B
Í ˜Z ˜Y Ý

˜W Y ˜ W [ Û
t13 = N % N ËÌ  
Í ˜[ ˜Y ÜÝ
Thus the magnitude of the shearing stress is 2 mPa
8. Repeat the above-mentioned problem for a plane whose normal is in the y-axis, everything
else the same.
Solution:
(i) From the given velocity fields, it is clear that the velocity divergence (vx + vy + vz = 0)
and so the normal stress along y-axis is t22 = –p + 2 m D22.
Ë ˜W Z Û
We know % Ì Ü  in this case. The excess of the total normal compressible
Í ˜Z Ý
stress whose normal is in y-axis over the pressure p is – (t 22) – p = – 2m D22 = –2 ´ 1
´ 1 = –2 mPa.
Ë ˜W Z ˜W Y Û
(ii) t21 = N % NÌ  Ü  N   N1B
Í ˜Y ˜Z Ý

Ë ˜W Z ˜W [ Û
t23 =  N % NÌ  Ü 
Í ˜[ ˜Z Ý
Thus, the magnitude of the shearing stress is 2 mPa.
9. Repeat the above-mentioned problem for a plane whose normal is in the z-axis everything
else being the same.
Solution:
(i) From the given velocity fields, it is clear that the velocity divergence (vx + vy + vz = 0)
and so the normal stress along z-axis is t33 = –p + 2m D33.
Ë ˜W Û
We know % Ì [ Ü  in this case. The excess of the total normal compressible
Í ˜[ Ý
stress whose normal is in z-axis over the pressure p is –(t33) – p = –2m D33 = – 2 ´ 0 = 0.

˜W [ ˜W Y Û
(ii) t31 = N % N ËÌ  
Í ˜Y ˜[ ÜÝ
Ë ˜W [ ˜W Z Û
t32 = N % NÌ  Ü 
Í ˜Z ˜[ Ý
Thus, the magnitude of the shearing stress is 0.
10. A parallel or a unidirectional flow is one for which all fluid elements have their velocity
vector parallel to a particular fixed direction. For a parallel flow of an incompressible
laminar flow of a viscous fluid, show that the total normal compressive stress on any plane
parallel to and perpendicular to the direction of fluid flow is the pressure p.
238 u BASICS OF ATMOSPHERIC SCIENCE

Solution: Without loss of generality, let the velocity vector of the parallel flow point in the x
direction. Then the components of the velocity along y and z directions vanish, i.e. vy = vz = 0
˜W Y
and from the continuity equation for the incompressible fluid,  yielding, vx = vx (y, z);
˜Y
vy = vz = 0.
For this parallel flow, D11 = D22 = D33 = 0. Hence,
t11 = –p; t22 = –p; t33 = –p.

REVIEW QUESTIONS

1. Define a boundary layer.


2. Draw a schematic diagram indicating the balance of forces in the Ekman layer.
3. Why is the surface layer also known as the constant flux layer?
4. Indicate the change in wind direction with increase in height within the Ekman layer in the
northern hemisphere.
5. Indicate the change in wind direction with increase in height within the Ekman layer in the
southern hemisphere.
6. What is Reynolds number?
7. Within a boundary layer can one experience potential flow? Explain.
8. What is the mixing length?
9. What is Richardson number?
10. What do we mean by the closure problem in turbulence modelling?
11. What are Reynolds stress terms?
12. Under what conditions is the logarithmic wind profile actually observed?
13. What is Monin-Obhukov length?
14. What are secondary circulations?
15. What are the assumptions, which go into the Ekman layer theory?
16. Do we have boundary layers over the oceans as well?
17. How does one define the atmospheric boundary height under different stability conditions?
18. What is an eddy?
9 Waves in the
Atmosphere

It is important to study the waves in the atmosphere as many of them are commonly observed
in the atmosphere. Waves also allow effective communication between different parts of the
atmosphere. It is convenient to study the atmospheric waves using the “perturbation method”,
a technique, which will be introduced in Section 9.1. Just as waves can form both within and
on the surface of a fluid medium, waves in the atmosphere can form internal as well as on the
free surface separating two fluids of different density.
Since the atmosphere is a fluid, waves arise naturally in the fluid medium due to the
action of restoring forces on air parcels which are displaced from their equilibrium positions.
Acoustic waves manifest when the restoring forces are due to compressibility, while gravity
waves are produced when the restoring force is buoyancy. The meridional gradient of the
Coriolis parameter is responsible for the existence of the most important meteorological
wave—the Rossby wave in a barotropic atmosphere. Section 9.1 presents the discussion of
Rossby waves, while Section 9.2 outlines the derivation of the phase speed of shallow water
gravity waves. Both the orographic as well as the acoustic waves are presented in Section 9.3,
while the gravity waves internal to the atmospheric medium are introduced in Section 9.4.
Finally, the various equatorial waves are described in Section 9.5.

9.1 ROSSBY WAVES


Rossby waves, first identified by Carl-Gustaf Rossby in 1939, are a large, slow moving
planetary scale waves which manifest in the troposphere due to orographic forcing as well as
the temperature contrasts between the land-ocean-atmosphere. Rossby waves are affected by
the Coriolis force and arise due to the conservation of absolute vorticity. Rossby waves are
known to influence the weather in the mid-latitudes since they play an important role in the
formation of low-pressure as well as high-pressure regions. The phase of the Rossby waves
has a westward component and this can be explained using the conservation of absolute
vorticity.
239
240 u BASICS OF ATMOSPHERIC SCIENCE

9.1.1 Perturbation Method


In this subsection, we introduce the perturbation method—a technique widely used for the
qualitative analysis of the various atmospheric waves. In this technique, all the meteorological
variables such as velocity, pressure, density, etc. are divided into two parts, one a basic state
part and another a perturbation part. The basic state part is normally assumed to be
independent of time and longitude, while the perturbation part is the local deviation of any
meteorological variable from the basic state. Let V be the time and longitude averaged basic
zonal velocity, while u¢ be the deviation from the basic state. Since u¢ can in general be a
function of both time and longitude V „ V  V „ Y U
 With the above definition, the inertial
˜V
acceleration term V is written as
˜Y
˜V ˜ V  V „
˜V „ ˜V „
V V  V „
V  V„ (9.1)
˜Y ˜Y ˜Y ˜Y
The perturbation method assumes the following: (i) the basic state meteorological variables
themselves must satisfy the governing atmospheric motions when the perturbation parts
vanish, (ii) the perturbation fields are small and hence the terms involving the products of
perturbation in the governing equations can be safely neglected. Applying the above
˜V „
assumption, on Eq. (9.1), results in the nonlinear term V „ being neglected. This
˜Y
requirement ensures that the nonlinear governing atmospheric equations get reduced to linear
differential equations in the perturbation meteorological variables in which the basic state
fields are assumed to be known coefficients. These linear equations can then be solved using
standard methods to determine the perturbation fields in terms of the specified basic state.

9.1.2 Some Basic Properties of Waves


Wave motions are characterized by oscillations in the meteorological field variables that
propagate in space. The solution of the simplest wave motion is given by a linear harmonic
oscillator, as manifested in the simple pendulum and is given as
q = q0 cos (n t – a) (9.2)
where q0 is the amplitude, n is the frequency and the solution is given in terms of the
amplitude q0 and the phase n t – a. Unlike the above wave motion, progressive waves are
waves having phases depend on time and also one or more space variables. Thus, for a wave
propagating in the x direction, the phase can be expressed as (kx – n t), and the surface of
O
constant phase will propagate with a velocity c, known as the phase velocity c given by 
L
Here, k is known as the wave number of the progressive wave. While the frequency of a
linear oscillator depends only on the physical properties of the oscillator, the frequency of
progressive waves depends on the wave number as well as the physical properties of the
medium. Waves for which the phase speed varies with the wave number are known as
dispersive waves and the relation between the frequency and wave number is known as the
WAVES IN THE ATMOSPHERE u 241

dispersion relation. For dispersive waves, the various sinusoidal components of a disturbance
starting at a given point are found at various different locations at a later time. This causes a
wave group—a spatially localized disturbance made up of a number of wave components, to
change its shape as it propagates. Hence, for dispersive waves the speed of a wave group is in
general different from the average phase speed of the individual wave components. The
velocity at which the observable disturbance and hence its energy propagates is known as the
group velocity cg and can be obtained as follows.
Consider the superposition of two waves of equal amplitude with different frequencies
n1 and n2 and wave numbers k1 and k2.
cos (k1x – n1t) + cos(k2x – n2t) = 2 cos[0.5(n1 – n2)t – 0.5(k1 – k2)x] ´
cos[0.5(n1 + n2)t – 0.5(k1 + k2)x] (9.3)
If the wave numbers and frequencies differ very slightly from one another, then

È O  O Ø
n1 – n2 = 'O  O  É   O (9.4)
Ê  ÙÚ
È L  L Ø
k1 – k2 = 'L  L  É  L (9.5)
Ê  ÙÚ
Using Eqs. (9.4) and (9.5) in Eq. (9.3), one gets
cos(k1x – n1t) + cos(k2x – n2t) = 2 cos[0.5 (Dn t – Dk x)] cos [n t – kx] (9.6)
Since n1 is close to n2, we have used n1 » n2 = n and, similarly, since k1 is close to k2, we
have used k1 » k2 = k, in Eq. (9.6). Equation (9.6) may be interpreted as a wave propagating
O
with a phase speed and a modulated amplitude with maxima and minima that moves with
L
'O
the group velocity cg, where DH and in the limit Dk going to zero, cg becomes
'L
EO
DH (9.7)
EL
Hence the observed disturbance or energy associated with a wave propagates at the group
velocity.

9.1.3 Rossby Waves in a Barotropic Atmosphere


The most important meteorological wave happens to be the Rossby wave, a wave which owes
its existence in a barotropic atmosphere to the meridional gradient of the Coriolis parameter.
We had introduced the concepts of absolute and potential vorticity in Chapter 7. In a
barotropic atmosphere, the Rossby waves are absolute vorticity conserving trajectories.
Rossby waves play a very important role in the large-scale meteorological processes and are
used widely in diagnostic analysis of the large-scale meteorological motions. Rossby waves in
a baroclinic atmosphere owe their existence to the isentropic gradient of potential vorticity
and are a potential vorticity conserving motion.
242 u BASICS OF ATMOSPHERIC SCIENCE

Consider a closed ring of air parcel aligned initially along a latitude circle. Since Rossby
waves follow a constant absolute vorticity (CAV) trajectory, it is convenient to note that
the absolute vorticity equals the sum of the relative vorticity and the Coriolis parameter, i.e.
h = z + f. Let the closed ring of air parcel have z = 0 at the initial time t0. If d y is the
meridional displacement of the above air parcel from the original latitude, then at a later time
t1, the conservation of absolute vorticity requires that
[  G
= G (9.8)
 U  U

[ =
U
GU

 GU

 CE Z (9.9)

EG
where C is the meridional gradient of the Coriolis parameter calculated at the original
EZ
latitude. It is pertinent to note that b is positive in both the northern and the southern
hemispheres and hence from Eq. (9.9) it is clear that the perturbation vorticity is cyclonic
(positive vorticity) for a southward displacement (d y < 0) and anticyclonic (negative
vorticity) for a northward displacement (d y > 0). As indicated in Figure 9.1, the perturbation
vorticity field will induce in the northern hemisphere a meridional velocity field which
ensures that the ring of air parcel is advected southward, west of the cyclonic vorticity and
northward, west of the anticyclonic vorticity. Hence the air parcels oscillate back and forth
about the original latitude and the pattern of the alternating cyclonic and anticyclonic
vorticity propagates to the west, constituting a Rossby wave.
w


+

Figure 9.1 The westward displacement of the perturbation pattern associated with a
meridionally displaced ring of fluid particles. The dashed lines show the
perturbation vorticity field and the induced velocity field, while the heavy wavy lines
show the initial perturbation and the westward displacement of the pattern.

The conservation of absolute vorticity following the horizontal motion is written as


E [  G

=0 (9.10)
EU
Ș ˜ ˜Ø
ÉÊ ˜U  V ˜Y  W ˜Z ÙÚ [  C W = 0 (9.11)
WAVES IN THE ATMOSPHERE u 243

EG
where C  We shall assume that the motion consists of a basic state zonal velocity
EZ

together with a small horizontal perturbation

V V  V „ W W  W „ [ [ [„ (9.12)
Defining a perturbation streamfunction y ¢ as follows:
˜Z „ ˜Z „
u¢ =   W„ (9.13)
˜Z ˜Y
one obtains z¢ = ³Z „ (9.14)
Applying the perturbation technique to Eq. (9.11) and retaining only the linear terms, one gets

Ș ˜Ø  ˜Z „
ÉÊ  V ÙÚ ³ Z „  C  (9.15)
˜U ˜Y ˜Y

We seek a wave type solution of the form


y ¢ = Re [y exp (if)] (9.16)
where Re stands for the real part of the expression in the square bracket, J   and j is
the phase kx + ly – n t. k and l are the wave numbers of the zonal and the meridional
directions, while n is the frequency. Substituting Eq. (9.16) into Eq. (9.15), one gets after
solving for the frequency,
CL
O VL  (9.17)
L  M


Dividing by k on both sides of Eq. (9.17), one gets for the expression for the zonal phase
speed of the Rossby wave in terms of the mean wind, b and the zonal and the meridional
wave numbers as
C
D V (9.18)
L  M


Hence the zonal phase propagation of the Rossby wave is always westward (since b > 0)
relative to the mean zonal flow. The Rossby wave speed depends inversely on the square of
the horizontal wave vector, the latter being the resultant of the zonal and the meridional wave
numbers. Since the zonal phase speed of the Rossby wave varies with the horizontal wave
number, Rossby waves are dispersive in nature with zonal phase speeds increasing with
increasing wavelengths. For a typical mid-latitude synoptic scale disturbance, with zonal wave-
length being 6000 km, meridional wavelength being 3000 km and b = 1.61 ´ 10–11 m–1 s–1, the
Rossby wave zonal phase speed relative to the zonal flow is approximately – 18 m s–1. It is
possible that for longer wavelength, the westward zonal phase speed of the Rossby wave
equals the eastward mean zonal wind resulting in a disturbance, which is stationary relative to
the surface. For this condition to hold, one requires that

L   M CV (9.19)
244 u BASICS OF ATMOSPHERIC SCIENCE

Since Rossby waves are dispersive, their zonal group velocity may be either eastward or
westward relative to the mean flow depending on the ratio of the zonal and meridional wave
numbers. Also, it can be shown very easily that for stationary Rossby wave condition, the
zonal group velocity is always in the eastward direction.

9.2 GRAVITY WAVES IN SHALLOW WATER


Gravity waves owe their existence to the restoring mechanism of buoyancy. Gravity waves in
shallow water can form only if the fluid has a free surface or an internal density discontinuity.
Consider the wave produced in a shallow pond as an example of gravity wave in shallow
water (Figure 9.2). Consider two incompressible fluids of varying density r1 and r2 with
r1 > r2 and confined on a semiinfinite plane. Since the densities r1 and r2 are constant, the
horizontal pressure gradient is independent of the distance from the free surface and one
obtains
˜ È ˜Q Ø ˜S
É Ù  H  (9.20)
˜[ Ê ˜ Y Ú ˜Y

FIGURE 9.2 A two-layer fluid system.

For simplicity and convenience one can assume that there is no horizontal pressure gradient in
the upper fluid layer. Further, assume that the fluid motion is restricted to two dimensions in
the x–z plane. The x momentum equation for the lower fluid layer is then
˜V ˜V ˜V  ˜Q
V X  (9.21)
˜U ˜Y ˜[ S ˜Y
As can be seen readily from Figure 9.2, the horizontal pressure gradient in the lower layer
arises primarily due to the differing weights of fluid column. For the lowest two points at the
same height as indicated in Figure 9.2, one can write
p + dp1 = p + r1 g Dz; p + dp2 = p + r2 g Dz (9.22)
From Eq. (9.22) the horizontal pressure gradient in the lowest fluid layer is given by
Q  E Q
 Q  E Q
'[ ˜I
S  S
H
 H 'S (9.23)
'Y 'Y ˜Y
WAVES IN THE ATMOSPHERE u 245

˜I
where is the slope of the interface in the limit, when Dx ® 0. Substituting Eq. (9.23) in
˜Y
Eq. (9.21), one gets for the x momentum equation in the lowest fluid layer
˜V ˜V ˜V H 'S ˜I
V X  (9.24)
˜U ˜Y ˜[ S ˜Y
The continuity equation is of the form
˜ V ˜X
  (9.25)
˜ Y ˜[
Since the horizontal pressure gradient in Eq. (9.24) is independent of height, the x component
of velocity, u, will also be independent of height provided that u did not vary with height
initially. This enables one to integrate Eq. (9.25) vertically from the lower boundary z = 0 to
the interface between the two fluids at z = h to get

X I
X 
 I ÈÉÊ ˜V ØÙÚ (9.26)
˜Y
w(h), which is the rate at which the height of the fluid interface is changing is given by
EI ˜I ˜I
X I
V (9.27)
EU ˜U ˜Y
Also, for a flat lower boundary, w(0) = 0 (9.28)
Substituting Eqs. (9.27) and (9.28) in Eq. (9.26), one gets
˜I ˜ IV

  (9.29)
˜U ˜Y
Equations (9.24) and (9.29) are two closed set of equations in u and h. If one applies the
perturbation techniques by assuming as before
V V  V „ I )  I„ (9.30)
where H is the mean depth of the lower fluid layer, V is the basic state zonal velocity and u¢
and h¢ are the associated perturbations about the basic state. Substituting Eq. (9.30) in
Eqs. (9.24) and (9.29) and dropping the nonlinear terms, one gets
˜V „ ˜V „ H 'S ˜I „
V  =0 (9.31)
˜U ˜Y S ˜Y
˜I „ ˜I „ ˜V „
V ) =0 (9.32)
˜U ˜Y ˜Y
Eliminating u¢ from the above two equations, one gets

Ș ˜Ø H) 'S ˜  I „
ÉÊ  V ÙÚ I „   (9.33)
˜U ˜Y S ˜Y 
246 u BASICS OF ATMOSPHERIC SCIENCE

Assuming a wave-type solution of the form


h¢ = A exp [ik (x – ct)] (9.34)
one gets the expression for the phase speed as

Ë H) 'S Û
D V “Ì Ü (9.35)
Í S Ý

If the upper and the lower fluid layers refer to air and water, then Dr ~ r1 and the phase
speed expression for the shallow water gravity wave become

D V “ < H) > 
(9.36)
The above results are applicable only for waves whose wavelengths are very much greater
than the vertical depth of the fluid, since otherwise the hydrostatic approximations will not be
valid. If one assumes an average depth of 4 km for the oceans, Eq. (9.36) provides
propagation velocities of about 200 m s–1, which are really very high values. Such values can
occur only if the waves are to be triggered by tsunamis, since only in the latter case the
vertical wavelength is much larger than the average depth of the ocean. However, shallow
water gravity waves can also be produced when the density differences are due to the
'S
differences in temperature as in a thermocline where the density varies by  and the
S
propagation velocities are much smaller.

9.3 OROGRAPHIC AND SOUND WAVES


In addition to the meteorologically important Rossby waves and gravity waves, two other
waves are also considered important while dealing with waves in the atmosphere. These are
the orographic waves and sound waves and they are dealt briefly in the following sub-
sections. Sound waves are longitudinal waves and require the medium to be compressible for
their propagation. Orographic waves form when strong terrain induced winds occur on the
downwind side or lee of the mountains or when the strong terrain induced winds flow
perpendicular to the mountain ridgelines.

9.3.1 Orographic Waves


One of the most illuminating examples of atmospheric waves is produced by a wave,
manifested by the striking alignment of clouds, when a steady current is forced to flow over a
mountain range as shown in Figure 9.3(a, b). Figure 9.3(a) depicts a westerly current flowing
over a north-south oriented mountain barrier and resulting in individual air parcels being
displaced upwards and downwards from their equilibrium levels. Hence these air parcels
undergo buoyancy oscillation as they cross over the mountain barrier. It is clear that these
waves are stationary and in a coordinate system moving with the wind, the surfaces of
constant phase move away from an observer in that coordinate system.
WAVES IN THE ATMOSPHERE u 247

FIGURE 9.3 (a) Top panel shows a schematic diagram of generation of waves over orography.
The dashed line indicates the top of the boundary layer. (b) Bottom panel shows
the lines of constant phase when the corrugated profile moves with a velocity equal
to the phase velocity c.

An alternate way of studying the above is to consider the corrugated surface moving
below the atmosphere. If one considers an air parcel along one of the lines at constant phase,
this air parcel will oscillate with a frequency proportional to the Brunt Vaisala frequency N. It
is known that the restoring force along the vertical for an air parcel undergoing buoyancy
oscillation is given by
F = –N 2d z (9.37)
From Eq. (9.37), the restoring force along a direction forming an angle q with the vertical is
given by
F = – N2 d z cos q (9.38)
where d z = d s cos q. Along the direction of s (refer Figure 9.3(b)), the equation of motion
can be written as
E  E T


 /  DPT RE T (9.39)
EU
Equation (9.39) represents an oscillation with frequency (kc)2 = (N cos q)2; that determines
the angle. However, the same angle q can be obtained by equating the horizontal and vertical
wave lengths by the relation
-) M
UBO R (9.40)
-7 L
248 u BASICS OF ATMOSPHERIC SCIENCE

where k and l are the horizontal and vertical wave numbers. Using the above sample relation
it is possible to derive the dispersion relation as given below

  DPT R ÈMØ   LD /

UBO  R  ÉÊ ÙÚ  (9.41)
DPT R L LD /

From Eq. (9.41),


È / Ø 
M É   Ù L (9.42)
ÊT Ú
Where sÿ = kc. From Eq. (9.42) it is clear that if N > s, the vertical wave number l is real and
vertical propagation is possible. From Eq. (9.42),

T “
/L
(9.43)
L  M

  

From Eq. (9.43), the expression for the phase and the group velocities in the horizontal and
the vertical directions can be written as follows:
T
“
/  D Q[
T
“
/L
cpx = (9.44)
L L  M

 
M M<L  M  >


ET /M  ET /LM
cgx = “  DH[ “ (9.45)
     
EL L  M
EM <L  M >

Since the above relation for cpx and cgx are obtained relative to the air and in the presence of
a wind which is along the x direction, the expressions cpx and cgx of Eqs. (9.44) and (9.45) become

D QY V “ /  DHY V “
/M 
(9.46)
L  M
<L  M >
     

with expressions for cpz and cgz remaining the same. The positive sign in Eqs. (9.44)–(9.46)
refer to the eastward propagation waves relative to the average wind.

9.3.2 Sound Waves


Sound waves are longitudinal waves, which are propagated by the alternating adiabatic
compression and the rarefactions of the medium. Hence the sound waves require that the
medium be compressible to propagate. Sound waves are not important in the meteorological
context and can be studied using the perturbation method. For simplicity, assume that the
sound waves are one-dimensional in nature and propagate in a direction parallel to the
positive x-axis. Assume that the motion is restricted in the longitudinal direction only with the
result that v = w = 0. Also, since the sound waves are one-dimensional in this case, the
dependence on y and z will vanish. Under the above assumptions, the governing conservation
equations of momentum, mass and energy for adiabatic motions are
EV  ˜Q
 =0 (9.47)
EU S ˜Y
WAVES IN THE ATMOSPHERE u 249

ES ˜V
S =0 (9.48)
EU ˜Y
E MO R
=0 (9.49)
EU
From the Poisson equation and the equation of state, one gets

È Q Ø È  Ø 3TQ D Q 

R É ÙÉ Ù (9.50)
Ê S 3TQ Ú Ê Q Ú
Eliminating the potential temperature q from Eq. (9.49) using Eq. (9.50), one gets
DW E MO Q E MO S
  (9.51)
DQ EU EU

Eliminating the density r from Eq. (9.51) using Eq. (9.48), one gets
DW E MO Q ˜V
  (9.52)
DQ EU ˜Y
From the perturbation technique, one writes
u(x, t) = V  V „ Y U
(9.53)
p(x, t) = Q  Q „ Y U
(9.54)
r (x, t) = S  S „ Y U
(9.55)
Substituting Eqs. (9.53)–(9.55) in Eqs. (9.51) and (9.52) and neglecting the nonlinear terms

and noting that since   one can approximate the density term by binomial expansion as
S

  È S„Ø È S„Ø
   Ù (9.56)
S  S„ S ÉÊ S ÙÚ É
SÊ SÚ
Using Eq. (9.56), one gets the following perturbation equation:

Ș ˜Ø  ˜Q „
ÉÊ  V ÙÚ V „  =0 (9.57)
˜U ˜Y S ˜Y
Ș ˜Ø DQ ˜V „
ÉÊ  V ÙÚ Q „  Q =0 (9.58)
˜U ˜Y DW ˜Y
Ș ˜Ø
Eliminating u¢ from Eqs. (9.57) and (9.58) by operating on Eq. (9.58) with É  V Ù and
Ê ˜U ˜Y Ú
substituting from Eq. (9.57), one gets

Ș ˜Ø D Q Q ˜ Q „
ÉÊ  V Ù Q „   (9.59)
˜U ˜Y Ú DW S ˜Y 
250 u BASICS OF ATMOSPHERIC SCIENCE

Assume a wave-type solution of the form


p¢ = Re[A exp{ik(x – ct)}] (9.60)
Substituting Eq. (9.60) in Eq. (9.59), one obtains an expression for the phase speed c as

H Q
D V“ (9.61)
S
DQ H Q
where H  The speed of the sound wave relative to the zonal motion is “ and this
DW S
latter quantity is known as the adiabatic speed of sound.

9.4 INTERNAL GRAVITY WAVES


Section 9.2 introduced the gravity waves in shallow water and an important characteristic of
these waves is that they cannot propagate vertically. A more general class of waves whose
phase can propagate vertically is called the internal waves. Unlike the ocean which is
bounded both above and below and wherein the gravity waves can propagate mainly in the
horizontal plane, the atmosphere has no clear upper boundary, and hence can allow for the
gravity waves to be propagated both vertically as well as horizontally. Internal gravity waves
play an important role in transporting energy and momentum to higher levels and are
associated with the formation of mountain lee waves as well as clear air turbulence.
Internal gravity waves can be studied on a vertical plane (x, z plane) using the
perturbation technique by writing down the momentum conservation equation in the
x direction as
˜V ˜V ˜V  ˜Q
V X  (9.62)
˜U ˜Y ˜[ S ˜Y
Equation (9.62) can be linearized by assuming V V  V „ Q Q  Q „ S S  S „ X X „ and
where V is independent of height. Neglecting the nonlinear terms,

Ș ˜Ø  ˜Q „
ÉÊ  V ÙÚ V „  (9.63)
˜U ˜Y S ˜Y
The vertical equation of motion is arrived at by assuming that the hydrostatic equilibrium
holds for the average values, while the fluctuation in the potential temperature q ¢ gives rise to
HR „
the buoyancy acceleration of the form and hence can be written as
R
Ș ˜Ø  ˜Q „ HR „
ÉÊ  V ÙÚ X „    (9.64)
˜U ˜Y S ˜[ R
E MO R
Utilizing the adiabatic form of the thermodynamic energy equation  one can
EU
ER
express the linearized form of the above equation in the form  to get
EU
WAVES IN THE ATMOSPHERE u 251

Ș ˜Ø ˜R
ÉÊ  V ÙÚ R „  X „  (9.65)
˜U ˜Y ˜[
Eliminating p¢ between Eqs. (9.63) and (9.64) by differentiating Eq. (9.63) with respect
to z and differentiating Eq. (9.64) with respect to x, and subtracting one from the other,
one gets

Ș ˜ Ø È ˜V „ ˜X „ Ø H ˜R „
ÉÊ  V ÙÚ ÉÊ  Ù  (9.66)
˜U ˜Y ˜[ ˜Y Ú R ˜Y
Eliminating q ¢ from Eqs. (9.65) and (9.66) and using the expression for Brunt Vaisala
H ˜R
frequency, /  in the resultant expression, one gets
R ˜[
˜ Ø È ˜V „ ˜X „ Ø
/  ˜X „

Ș
ÉÊ  V ÙÚ ÉÊ  Ù (9.67)
˜U ˜Y ˜[ ˜Y Ú ˜Y
The continuity equation is of the incompressible form since we wish to isolate only the
gravity waves and is of the form
˜V „ ˜ X „
  (9.68)
˜Y ˜[
Eliminating u¢ from Eqs. (9.67) and (9.68) by cross-differentiating, one gets

˜ Ø È ˜ X „ ˜ X „ Ø  ˜ X„
 
Ș
ÉÊ  V ÙÚ É   Ù  /  (9.69)
˜U ˜ Y Ê ˜Y ˜[  Ú ˜Y 
Assume a wave-like solution of the form

X„ X J LY  M[  X U^>
3F < ~ FYQ \ (9.70)
where X~ is a complex quantity and kx + lz – wt is the phase. Substituting Eq. (9.70) in
Eq. (9.69), one gets dispersion relation

X~ X  VL “
/L
(9.71)
L   M 

Here, the horizontal wave number k is real, while for vertical propagation one requires that
the vertical wave number l be complex. Also X~ is the wave frequency with respect to the
average wind and so the phase velocities relative to the wind in the x and z directions are

given by D QY
X~ BOE D X~ and have the same form as given by Eq. (9.44). The group
Q[
L M
velocities relative to the wind in the x and z directions again have the same form as
Eq. (9.45). It is pertinent to note that the solution obtained here includes also the orographic
waves.
252 u BASICS OF ATMOSPHERIC SCIENCE

9.5 EQUATORIAL WAVES


Theoretical investigations have shown that for weather systems to grow by converting
potential energy to kinetic energy, their horizontal and vertical scales need to scale as
) G (9.72)
- /
where f is the Coriolis parameter and N is the Brunt Vaisala frequency. Typically, the aspect

ratio,
) which is about  for mid-latitude weather systems becomes smaller and smaller
- 
and ultimately tends to zero as the equator is approached, since the Coriolis parameter
vanishes at equator. Hence, weather systems with a fixed horizontal scale would become
much smaller in the vertical extent as the equator is approached. Atmospheric tropical waves
play a very important role in the adjustment of the mass and the wind fields following the
response of the tropical atmosphere subjected to diabatic heating. These tropical waves can
propagate both horizontally and vertically with the latter having important effects on the
circulation of the equatorial stratosphere. In this section, only the horizontally propagating
equatorial waves will be investigated. The nature and the properties of the tropical waves are
determined primarily by the rotation of the earth and the density stratification of the earth’s
atmosphere. At low latitudes, the equatorial beta plane approximates f = b y, where
EG
C Z finds widespread applications and where a and w are the radius and the angular
EZ

velocity of the earth. Before we take up the study of equatorial waves in detail, it is important
to derive the shallow water equations which will be utilized in the study of equatorial waves.

9.5.1 Shallow Water Equations


The shallow water model and its associated equation provides for the description of some
important aspects of both atmospheric and oceanic motions, and hence will be dealt below.
Consider a sheet of inviscid fluid with uniform density as shown in Figure 9.4. Here, the
height of the surface of the fluid above the reference level z = 0 is h(x, y, t), while the rigid
bottom is defined by the surface z = hB(x, y). Let the mean depth of the fluid layer be D and
let u, v and w be the velocity components along x, y and z directions, respectively. Let the
pressure of the fluid surface be constant and equal to p0. Let L and T be the characteristic
horizontal length scale and the time scale for the motion and w be the angular velocity of the
rotating fluid. The fundamental requirement for the application of the shallow water model is
that the aspect ratio is very small, i.e.

E
%  
(9.73)
-
Assumption of uniform density reduces the continuity equation to its simplest form
˜V ˜W ˜X
   (9.74)
˜Y ˜Z ˜[
WAVES IN THE ATMOSPHERE u 253

FIGURE 9.4 Schematic diagram representing the shallow water model.

If U and W are the characteristic scales for the horizontal and vertical velocities, then from
8
Eq. (9.74),
% cannot be larger than 0 ÈÉÊ 6 ØÙÚ and so
-
W £ O (d U) (9.75)
Writing down the momentum equations in all the three directions and applying scale analysis,
one finds for the x momentum equation
˜V È ˜V ˜V ˜V Ø  ˜Q~
 V W  X Ù  GW  (9.76)
˜U ÉÊ ˜Y ˜Z ˜[ Ú S ˜Y
6 68
Here, the local derivative term as an order the vertical advection has an order of
%

5

1
the Coriolis term has an order f U, pressure gradient term has an order of S - where P is

the scale for the variation of the pressure field, and the horizontal advection terms have an

order of magnitude of
6 The y momentum equation is of the form
-


˜W È ˜W ˜W ˜W Ø  ˜Q~
 ÉV W  X Ù  GV  (9.77)
˜U Ê ˜Y ˜Z ˜[ Ú S ˜Z
and has exactly the same order of magnitudes for the various terms as the x momentum
equation. The z momentum equation is
˜ X È ˜X ˜X ˜X Ø  ˜Q~
 ÉV W X  (9.78)
˜ U Ê ˜Y ˜Z ˜[ ÙÚ S ˜[

Here, the local derivative term has an order of


8 the vertical pressure gradient term has
5

1
and the vertical advection term has
88 while the horizontal advection term has an
S% %
order of
68 The vertical advection term is extremely small and can be neglected.
-

254 u BASICS OF ATMOSPHERIC SCIENCE

One can write the total pressure to be made up of a term, which balances the
gravitational force and a departure term as
Q Y Z [ U
 S H[  Q~ Y Z [ U
(9.79)
From Eq. (9.75), one gets
68 0 È 6  Ø
(9.80)
% É - Ù
Ê Ú
From the horizontal momentum equations, one requires that the scale for the variable pressure
field have to be the largest of the three entities of the right-hand side of the equation below

-
1 S6 ËÌ 6 G-
Û
ÜÝ (9.81)
Í5 NBY
in order that the pressure gradient force term be not negligible. The ratio of the inertial term
to the vertical pressure gradient force term in Eq. (9.78) is bounded by the larger of

S ËÌ
8 86 Û Ë EX Û
Í5 - ÜÝ 0 ÌÌ S EU ÜÜ

˜Q~
(9.82)
1 % Ì Ü
ÍÌ ˜[ ÝÜ
Substituting from Eq. (9.81) to Eq. (9.82), one gets

EX Û Ë È 6Ø Û
S ËÌ Ì ÉÊ 5 - ÙÚ Ü
Í EU ÜÝ E Ì NBY Ü
˜Q ÌÈ  6 Ø Ü
(9.83)
Ì ÉÊ 5 - G ÙÚ
~

˜[ Í
Ü
NBY Ý
6
If the Rossby number is O (1), i.e. moderate and not very large, Eq. (9.83) implies an
G-
estimate of O(d 2) for the ratio of the right-hand side of Eq. (9.83). For very small Rossby

numbers, the ratio of the right-hand side is


È
0 É E  ËÌ 6Û Ø which is much smaller
Ü Ù
Í G5 G- Ý NBY Ú

Ê
˜Q~
than O(d 2). Since for shallow water model d << 1, it follows that is negligible and at
˜[
least to O(d 2) and so the total pressure is
˜Q
 S H  0 E 
(9.84)
˜[
Integrating Eq. (9.84), one gets
p = – r gz + A(x,y,t) (9.85)
WAVES IN THE ATMOSPHERE u 255

Applying the boundary conditions at the surface p(x,y,h) = p0, where p0 has a constant value,
one gets
p = rg(h – z) + p0 (9.86)
It is to be noted that the horizontal pressure gradient is independent of height z,
˜Q ˜I
= SH (9.87)
˜Y ˜Y
˜Q ˜I
= SH (9.88)
˜Z ˜Z
Equations (9.87) and (9.88) imply that the horizontal acceleration must be independent of
height. If one assumes that the horizontal velocity fields are independent of heat initially, then
they would be independent of the height at later times as well. From Eqs. (9.87) and (9.88)
and from the fact that both u and v velocities are independent of z, the horizontal momentum
equations become
˜V ˜V ˜V ˜I
V W  GW =  H (9.89)
˜U ˜Y ˜Z ˜Y
˜W ˜W ˜W ˜I
V W  GV =  H (9.90)
˜U ˜Y ˜Z ˜Z
Integrating the continuity equation (Eq. (9.74)) in z gives

È ˜V ˜W Ø
X Y Z [ U
 [É   X~ Y Z U
(9.91)
Ê ˜Y ˜Z ÙÚ
where the last term in the right-hand side of Eq. (9.91) is the constant of integration.
Utilizing the condition at w = 0 at z = hB, one gets

E[
w(x, y, hB, t) = 
EU [ I#

˜I# ˜I
w(x, y, hB, t) = V W # (9.92)
˜Y ˜Z
Hence, from Eq. (9.91) for z = hB, one gets

˜I# ˜I È ˜ V ˜W Ø
X~ Y Z U
V  W #  I# É  (9.93)
˜Y ˜Z Ê ˜Y ˜Z ÙÚ
Substituting Eq. (9.93) in Eq. (9.91), one gets
È ˜ V ˜W Ø ˜I ˜I
X Y Z [ U
I#  [
É  Ù V # W # (9.94)
Ê ˜Y ˜Z Ú ˜Y ˜Z
256 u BASICS OF ATMOSPHERIC SCIENCE

The kinetic condition at the surface z = h(x, y, t) is


˜I ˜I ˜I
X V W (9.95)
˜U ˜Y ˜Z
Equating Eqs. (9.94) and (9.95), one gets for z = h,
˜I ˜< I  I#
V> ˜< I  I#
W >
   (9.96)
˜U ˜Y ˜Z
By defining the total depth H = h – hB, Eq. (9.96) can be written in terms of H as
˜) ˜ V)
˜ W)

   (9.97)
˜U ˜Y ˜Z
or in another form
E) È ˜ V ˜W Ø
)É   (9.98)
EU Ê ˜Y ˜Z ÙÚ
Equations (9.89), (9.90) and (9.97) are known as shallow water equations. These equations
are much simpler and more amenable to analysis since there is a reduction by one (w and z)
of the number of dependent and independent variables. Furthermore, there has been a
reduction by one in the number of dynamical equations.

9.5.2 Equatorial Rossby and Rossby Gravity Waves


Assuming the mean basic state is motionless and having a mean depth hc, the linearized
shallow water equations (Eqs. (9.89), (9.90) and (9.97)) for the perturbations in an equatorial
b plane can be written as
˜V ˜G
 C ZW =  (9.99)
˜U ˜Y
˜W ˜G
 C ZV =  (9.100)
˜U ˜Z
˜G È ˜V ˜W Ø
 HID É  =0 (9.101)
˜U Ê ˜Y ˜Z ÙÚ
where f = ghc is the geopotential perturbation field. Using the equatorial Rossby radius of

È HI Ø
deformation 3F É D Ù as the characteristic length scale, U C  HID
 as the
ÊC Ú 

characteristic time scale and ghc as the characteristic geopotential height, it is possible to non-
dimensionalize Eqs. (9.99)–(9.101) to get the non-dimensional equations (designated by the
primed quantity) as
˜V „ ˜G „
 Z „W „ =  (9.102)
˜U „ ˜Y „
WAVES IN THE ATMOSPHERE u 257

˜W „ ˜G „
 Z „V „ =  (9.103)
˜U „ ˜Z „
˜ G „ ˜V „ ˜ W „
 
˜ U „ ˜Y „ ˜ Z „ = 0 (9.104)

The solution of Eqs. (9.102)–(9.104) may be expressed in terms of zonally propagating


waves as
(u¢, v¢, f ¢) = Re[U(y), V(y), F(y)] exp [i(kx – w t)] (9.105)
Substituting, Eq. (9.105) into Eqs. (9.102)–(9.104), and after eliminating U(y) and F(y), one
gets an equation for V(y) as follows:

E 7 L
 < X   L 
  Z  > 7  (9.106)
EZ 
X
Note that in Eq. (9.106), k2 ¹ w2 was assumed. It is necessary that the meridional structure
function V(y) be bounded as y ® ± ¥ since the equatorial b plane approximation is not a
valid approximation for spherical geometry for large y. Equation (9.106) along with the
bounded requirement of V(y) for large y, i.e. as y ® ± ¥ is an eigenvalue problem similar to
the Schrödinger wave equation for a simple harmonic oscillator and has a solution only if
L
X  L  O   O     (9.107)
X
Equation (9.107) is the non-dimensional dispersion equation relating the frequency w and the
wavenumber k. Figure 9.5 depicts the above dispersion curve for the equatorial waves. The
frequency w is assumed to be always positive and so k > 0 (k < 0), implies eastward
(westward) propagation of the wave relative to the ground.
For n ³ 1, Eq. (9.107) becomes

 
LO  “ X   O  
O    ž (9.108)
X X


For real k, corresponding to propagating non-decaying neutral waves, one requires that
O  
O
either X•  •

PS X … 
 Hence, as can be seen from Figure 9.5,
   


then there are two distinct groups of waves for which X •  called high-frequency


waves and the other group for which X …  called the low-frequency waves. For

L
n = 0, the dispersion relation Eq. (9.107) reduces to X  L   which yields one
X

meaningful root L X as can be seen from Figure 9.5. The second root for n = 0, k = – w
X
is not admissible as the frequency cannot be negative. The high-frequency waves are known
258 u BASICS OF ATMOSPHERIC SCIENCE

as the inertia-gravity wave and the low-frequency waves are known as the equatorial Rossby
waves, while the case n = 0 corresponds to the mixed Rossby gravity waves. The high-
frequency inertia-gravity waves are not dealt in detail in this chapter. In dimensional form
taking ghc = 2500 m2 s–2, the high-frequency waves have a period shorter than 1.28 days,
while the low-frequency waves have a time period larger than 7.3 days. Differentiating
Eq. (9.107) with respect to k, one gets the group velocity in the x direction as
˜X L X  
$ HY (9.109)
˜L L
X  
X

FIGURE 9.5 Dispersion curves for equatorial waves (up to n = 3), in a basic state at rest as a
function of the non-dimensional zonal wavenumber k and frequency w. Positive
(negative) wavenumber k refers to eastward (westward) propagating waves. The
dotted line indicates zero group velocity.
WAVES IN THE ATMOSPHERE u 259

L
From Eq. (9.109), it is clear that Cgx vanishes at 2 kw = –1 provided X   ›  Figure 9.5
X

shows the case X  as a dashed line indicating zero group velocity in the x direction.
L
The meridional structure of the zonally propagating Rossby, inertia-gravity and the
mixed Rossby-gravity waves are given in terms of the solution of Eq. (9.106) which are
known as the Weber-Hermite functions

7 Z
%O Z
F  Z ) O Z
O
< >
   ž (9.110)
where Hn(y) are the Hermite polynomial of order n and is given by
O 


M
O  Z
 
O M

) Z

O Ç M  O  M

(9.111)
M 

The solution for the other meridional structure functions U(y) and F(y) can be obtained from
Eqs. (9.102)–(9.104) in terms of Dn(y). The equatorial Rossby waves always propagate
westward as can be seen from the negative sign of the dimensional dispersion relation
obtained from Eq. (9.108) for the low-frequency waves
X 
C
(9.112)
L L   O  
C  HID

However, for the equatorial Rossby waves the group velocity can be either eastward or
È EX Ø
westward. The slopes of the dispersion curve in Figure 9.5 ÉÊ $ H
EL ÙÚ represent the
corresponding group velocity. A positive slope indicates an eastward group velocity, while the
negative slope shows a westward group velocity. The dashed curve corresponds to zero group
velocity. Since in Figure 9.5, the Rossby waves denoted as curves to the right side of the zero
group velocity, have negative slope, their energy (group velocity) propagates westward. These
Rossby wave curves to the right side of the zero group velocity have small wavenumber and
hence longer wavelength. Hence it is clear that the energy associated with the long Rossby
waves propagates westward, while the energy associated with the short Rossby waves
propagates eastwards. The above differences in the energy propagation of the long and short
Rossby waves have important effects when the reflection of the equatorial Rossby waves at
the oceanic lateral boundaries are to be considered. For long equatorial Rossby waves, k ® 0

and hence from Eq. (9.112), one gets


X   HID
 

O   ž suggesting that the long


L O  

Rossby waves are approximately non-dispersive. Also, the dimensional westward phase speed
is (2n + 1)–1 times the long gravity wave speed (ghc)0.5. Hence, for the fastest long Rossby
wave (n = 1) phase speed is about one third of the long gravity wave speed. Rossby waves
satisfy the geostrophic balance between pressure and the meridional as well as the zonal wind
as can be seen from Figure 9.6 which depict the horizontal structure for n = 1 equatorial
Rossby waves. Strong zonal winds are seen near the equator for the n = 1 Rossby mode as is
260 u BASICS OF ATMOSPHERIC SCIENCE

to be expected since both the Coriolis and the pressure gradient force approach zero as the
equator is approached. For the n = 1 long Rossby mode as can be seen from Figure 9.6, both
the zonal wind and the geopotential height f are symmetric about the equator, while the
meridional wind is antisymmetric. However, for the n = 2 long Rossby mode (not shown here
due to brevity), both the zonal wind and the geopotential height fields are antisymmetric
about the equator while the meridional wind is symmetric. At the equator there is no
meridional motion for the n = 1 long Rossby mode, while there is no zonal motion for the
n = 2 long Rossby mode at the equator.

FIGURE 9.6 The horizontal velocity and the height perturbations associated with an n = 1
equatorial Rossby wave.

9.5.3 Mixed Rossby Gravity Waves


The solution for the n = 0 mode, known as the mixed Rossby gravity waves, with a

dispersion relation, L X  has the following meridional structure:
X
Ë ËÈ
X   ØÙÚ X ÛÜ Û
X
J Y U
 Ì ÉÊ
W„ = 3F ÌF  Z 
F Í ÝÜ
(9.113)
Ì Ü
Í Ý

Ë  <ÉX 
È Ø
Ù Y  XU
Û
3F JX F ZF X
J

V„ = G„ Ì  Z  Ê Ú
Ü (9.114)
Ì Ü
Í Ý
WAVES IN THE ATMOSPHERE u 261

The mixed Rossby gravity wave corresponds to the n = 0 curve in Figure 9.5. For large w,
one gets from the dispersion relation k = w, which is the asymptotic limit of the high
wavenumber gravity waves. However, for small w, one has from the dispersion relation

L  which corresponds to the high wavenumber limit of the Rossby waves. Due to the
X
above characteristics, the n = 0 mode is called the mixed Rossby gravity wave mode and is
unique only to the equatorial region. For ghc = 2500 m2 s–2, the crossover transition from the
positive to negative k occurs at a dimensional time period of 2.1 days and corresponds to a
stationary wave in the y direction. For waves with time periods less than 2.1 days, k > 0, and
the waves propagate eastwards, while for the waves with time periods longer than 2.1 days,
k < 0, the waves propagate westwards. Since the slope is positive corresponding to the n = 0
mixed mode (refer Figure 9.5), the energy (group velocity) associated with the mixed Rossby
gravity waves propagate only to the east. Figure 9.7 depicts the horizontal distribution of the
velocity and pressure perturbation for an equatorial mixed Rossby gravity wave. The pressure
and the zonal velocity are antisymmetric about the equator, while the meridional component
of the velocity is symmetric. Figure 9.7 also shows that the largest meridional flow occurs at
the equator giving rise to the cross equatorial flow.

FIGURE 9.7 The horizontal velocity and the height perturbations associated with an equatorial
Rossby gravity wave.

9.5.4 Equatorial Kelvin Wave


Consider a wave motion without the meridional wind. Equations (9.102)–(9.104) with v¢
vanishing, then becomes
˜V „ ˜G „
=  (9.115)
˜U „ ˜Y „
˜G „
y¢u¢ =  ˜Z „ (9.116)

˜G „ ˜ V „
 =0 (9.117)
˜ U „ ˜Y „
262 u BASICS OF ATMOSPHERIC SCIENCE

Equation (9.115) when combined with Eq. (9.117) results in a wave equation of the form

˜ V „ ˜G „
(9.118)
˜U „ ˜Y „
which has the general solution V „ ' Y „ B U „
: Z „
where F is an arbitrary function. From

Eq. (9.115), one gets f ¢ = ±u¢. Using Eq. (9.116), one gets : Z „
: 
F B Z   Only the 

minus sign in the exponent is valid since the positive sign leads to an unbounded solution for
large values of y. From the above, the solution of the Kelvin wave in dimensional form for
C0 = (ghc)0.5, is given as
G 
$
V ' Y  $ U
F  C Z 
BOE W  (9.119)
$
The chief characteristic of the equatorial Kelvin wave is the vanishing of the meridional
velocity. The dimensional form of Eq. (9.118) is of the same form as that of the ordinary
shallow water gravity waves and hence the Kelvin wave dispersion relation in dimensional
form is identical to that of the ordinary shallow water gravity wave

$ XL
 HID (9.120)
The phase speed corresponding to Eq. (9.120) can be either positive or negative. But, from
Eq. (9.119) it is clear that if the solutions are to remain bounded and need to decay away
from the equator, the phase speed C0 must be positive. Thus, Kelvin’s waves phase speed is
eastward propagating. Figure 9.5 illustrates the equatorial Kelvin wave as a straight line in the
k > 0 domain denoted as the n = –1 wave. It is clear that the phase speed of the equatorial
Kelvin wave is about three times the phase speed of the fastest Rossby wave corresponding to
n = 1 mode.
The horizontal distribution of the perturbation zonal wind and the pressure are depicted
for an equatorial Kelvin wave in Figure 9.8. In the meridional direction as can be seen in
Figure 9.8, an exact geostrophic balance exists between the zonal wind field and the
meridional pressure gradient, while the zonal direction is characterized by an eastward
propagating shallow water gravity wave. Kelvin waves are equatorially trapped waves which
exist due to the change in the sign of the Coriolis parameter across the equator.

L H Equator

FIGURE 9.8 The horizontal velocity and the height perturbations associated with an equatorial
Kelvin wave.
WAVES IN THE ATMOSPHERE u 263

SOLVED EXAMPLES

1. A background jet stream of speed 60 m s–1 meanders with 6000 km wavelength and 1500 km
amplitude, centred at 45°N. Find the phase speed relative to the ground of the barotropic
Rossby wave?
Solution: The beta parameter

X  –  –  
C DPT G DPT ’
 –   N  T 
3&  – 
The phase speed relative to the ground,

 
È M Ø Ë  –  Û
D 6  C É    –  
Ì Ü     N T 
Ê Q ÙÚ ÍÌ Q ÝÜ
2. Find the wavelength of the path of air over a mountain given the following information.
The wind speed is 25 m s–1, the air temperature is 10°C, and the actual lapse rate is
6 K km–1.
Q6
Solution: The wavelength of the path of air over a mountain is given as M where
/ #7
U is the wind speed and NBV is the Brunt Vaisala frequency in units of s–1. The Brunt
Vaisala frequency is given as

H È '5W Ø Ë  Û
/ #7 ÉÊ  *E Ù   –     –  
Ü  T 
5W '[ Ú ÍÌ  Ý

Q 

The wavelength is then M  LN



3. For a natural wavelength of 104 m, and a hill width of 10 ´ 103 m, find the Froude number.
M
Solution: The Froude number is defined as 'S where W is the hill width and l is
8
 
the natural wavelength. From given data, 'S 
 – 
4. For the Froude number < 1, the mountain wave crests tilt upward with altitude. For an air
flow with Froude number of 0.7, calculate the angle a of the tilt of the wave crests relative
to the vertical.
Solution: For Froude number < 1, the relation of the angle a of the tilt of the wave crests
relative to the vertical with the Froude number is given by
cos(a) = Fr; cos(a) = 0.7, a = 45.57°.
5. Consider the surface of a liquid whose area is unlimited. Determine the velocity of
propagation of gravity waves propagating along the x-axis and uniform in the y direction.
Assume that the wavelength is small in comparison with the depth of the fluid.
264 u BASICS OF ATMOSPHERIC SCIENCE

Solution: For the potential flow v = grad f. Substituting in the Euler’s equation of motion,
one gets
È ˜G   Ø
³É  W  XÙ  where f is the velocity potential and w is the enthalpy.
Ê ˜U  Ú
˜G  
Integrating above equation,  W  X G U
where f (t) is a function of time.
˜U 
Assuming wave amplitude to be small, v2 term can be neglected. Putting f(t) = 0 and
including r gz term due to the gravitational field, pressure becomes
˜G
Q  S H[  S
˜U
Assuming a constant pressure p0 acts on the surface and let z be the vertical displacement of
the surface in its oscillations, one gets
˜G
Q  S H[  S
˜U
È Q UØ
By redefining the potential É BEEJOH  Ù the constant p0 can be eliminated and equation
Ê S Ú
becomes
È ˜G Ø
H[  É Ù 
Ê ˜U Ú [ [

Since the displacement z is small, the vertical velocity at the surface can be taken as
˜[
vz =
˜U
˜G
But, vz = BOE
˜U
È ˜G Ø ˜[ È  ˜ G Ø
so ÉÊ ÙÚ = É Ù
˜[ [ [ ˜U Ê H ˜U  Ú [ [

Since oscillations are small, the derivatives can be taken at z = 0 instead of z = z. Hence the
governing equations are
Ñ2f = 0, and

È ˜G  ˜ G Ø
É ˜[  H  Ù =0
Ê ˜U Ú [ 
We seek a simple periodic solution of the form, f = f (z) cos [kx – wt]
Substituting in Laplace equation, we have

E G

 L G 
E[

The solution of the above equation is of form


f = Aekz cos [kx – w t]
WAVES IN THE ATMOSPHERE u 265

Substituting the above solution to the boundary condition at z = 0, one gets the dispersion
relation w2 = kg
Since the velocity of propagation of the wave is given by

˜X
6
˜L
Substituting, one gets
 H  HM
6
 L  Q
6. Determine the propagation of long waves in a channel, where the cross-sectional area may
have any shape and may also vary along its length. Assume that the wavelength is long in
comparison with the depth and the width of the channel.
Solution: Let the cross-sectional area of the liquid in the channel be denoted by S(x, t).
Since the long waves move along the channel, the velocity component vx along the channel
is large as compared to vy and vz, respectively. Denoting vx by v, the x and the z components
of the Eularian equations are

˜W  ˜Q  ˜Q
   H
˜U S ˜Y S ˜[
Since the pressure of the free surface z = z is p0, from the z equation, one gets

p = p0 + gr (z – z)
Substituting the above in x momentum equation, one gets

˜W ˜[
H
˜U ˜Y
The second equation is the continuity equation expressed as

˜4 ˜ 4W

 
˜U ˜Y
Let S0 be the equilibrium cross-sectional area of the liquid in the channel. Then S = S0 + S¢,
where S¢ is the change in the cross-sectional area caused by the wave. Since the change in
the liquid level is small, one can write S ¢ = bz, where b is the width of the channel. The
continuity equation then becomes

˜[ ˜ 4 W

C  
˜U ˜Y
˜W
Differentiating the above equation with respect to t, and substituting from the equation
˜U
˜W ˜[
H one gets
˜U ˜Y
˜[ H ˜ Ë ˜[ Û
 4 
˜U 
C ˜Y ÍÌ ˜Y ÝÜ
266 u BASICS OF ATMOSPHERIC SCIENCE

If the channel cross-section is the same at all points, S0 = constant and the above equation
becomes
˜[
H4 ˜ [
 
˜U  C ˜Y 
The above equation has the same form as the wave equation and hence the velocity of
propagation of the long waves in a channel is given by

H4
6
C
7. Considering long waves in a large tank; assume the tank is infinite in the two directions
x and y. Assuming the depth of the tank to be h, find the velocity of propagation of the long
waves in such a large tank.
Solution: Since the vertical component of the velocity is small, the Euler’s equations of
motion become
˜W Y ˜[ ˜W Z ˜[
H  H 
˜U ˜Y ˜U ˜Z
The continuity equation in terms of the depth is of the form

˜I ˜ IW Y
˜ IW Z

  
˜U ˜Y ˜Z
Writing the depth h = h0 + z, where h0 is the equilibrium depth, one gets from the above
equation
˜[ ˜ I W Y
˜ I W Z

  
˜U ˜Y ˜Z
Assuming that the tank has a horizontal bottom given by h0 = constant, and differentiating
the above equation with respect to t and substituting from the x and y component of the
Euler equations, one gets

˜[ Ë ˜ [ ˜ [ Û
 HIÌ  Ü 
˜U  Í ˜Y

˜Z Ý
The above equation corresponds to a two-dimensional wave equation and hence the velocity
of propagation of the long waves in a large tank is given by

6 HI

REVIEW QUESTIONS

1. Identify the restoring force for the following waves: (i) sound waves, (ii) gravity waves, and
(iii) Rossby waves.
2. What are internal gravity waves?
WAVES IN THE ATMOSPHERE u 267

3. Why are gravity waves considered a noise and meteorologically not very significant?
4. What is the fastest wave?
5. What is the characteristic of Kelvin wave?
6. What is the mixed-wave? What is the significane of the mixed-wave with respect to
numerical weather prediction?
7. The barotropic Rossy wave is also known as Constant Absolute Vorticity (CAV) trajectories?
8. What is the direction of the movement of the Rossby wave?
9. When stable air flows over a hill or mountain, mountain waves develop in the flow with
wavelength l. The vertical displacement of a parcel of air moving over the mountain can be
expressed as
È  YØ È Q Y Ø
[ [ FYQ É DPT É
Ê CM ÙÚ Ê M ÙÚ
where z is the height of the air above its starting equilibrium height, z1 is the initial
amplitude of the wave, x is distance downwind of the mountain crest, and b is a damping
factor. (i) At what downwind distances x does the parcel height decrease by a factor of e–n,
where n is a positive integer? (ii) Make a plot of z versus x for a wavelength of 12 km and
a damping factor of 4. Assume that the initial wave amplitude is 600 m.
Large-scale
Meteorological

10 Systems in
Mid-latitudes

It is not uncommon for regions over mid-latitudes to experience extremely cold and dry
conditions during winter and extremely hot and humid conditions during summer. At times,
these conditions can last as long as three weeks without any break and can be observed over
extensive regions. The above situations correspond to conditions where large areas are
covered by an extensive body of air having more or less uniform temperature and moisture.
Such large volumes of air are known as air masses. Further, an extensive region, as large as
a continent, may be covered by several air masses at the same time, with the cold and dry
conditions to the north and warm and moist conditions to the south. Narrow boundary
regions called fronts commonly separate the contrasting air masses from one another. The
passage of fronts at a place is normally accompanied by sudden changes in temperature,
moisture and wind. It is often observed that systems form and intensify along the boundary
region separating the polar air from the warm air down south; the above boundary region is
referred to as polar front. Within the polar front, due to strong horizontal temperature
gradient, there are associated intense pressure gradients which ultimately give rise to strong
meandering air currents known as jet streams. Section 10.1 outlines the general
considerations of the large-scale meteorological systems in mid-latitudes and Section 10.2
introduces the concepts of fronts. Section 10.3 provides the life cycle and the structure of the
extratropical systems, while Section 10.4 discusses the characteristics of the jet stream.

10.1 GENERAL CONSIDERATIONS


The middle and the high latitudes are characterized by abrupt changes in the day-to-day
weather, involving marked changes in the temperature, humidity and wind following the
passage of a weather system. Most of the winter storm systems in the middle and high
latitudes are frontal in nature and are characterized by strong gradients of temperature and
moisture across the frontal zone.

268
LARGE-SCALE METEOROLOGICAL SYSTEMS IN MID-LATITUDES u 269

10.1.1 Air Masses


The continuous exchange of energy and moisture near the earth surface determines to a
large extent the temperature and humidity characteristics of the atmosphere. When large
masses of air, called the air masses, stagnate over a particular place for a reasonably long
time, of the order of many days, they tend to acquire the characteristics of the underlying
surface. The region over earth where air masses form is known as source regions. To act as
a source region, the area must be quite large, of the order of tens of thousands of square
kilometres. Also, air masses tend to form either in high or low latitudes. This is because the
atmospheric conditions in mid-latitudes can vary quickly on a day-to-day scale. Air masses
tend to have uniform distributions of temperature and humidity in the horizontal. Air masses
are normally classified depending on the temperature and humidity characteristics of their
source regions. Air masses, based on their humidity content are considered to be either
continental (dry) or maritime (moist). Further, based on their temperature, air masses are
tropical (warm), polar (cold) and arctic (extremely cold). According to convention, a two-
letter short-hand scheme is used to classify the air masses. A small letter c (denoting
continental) or m (denoting maritime) depending on the humidity conditions followed by a
capital letter T, P or A to represent temperature is used. According to the above notation, cP
stands for continental polar air mass, while mT refers to maritime tropical air mass. With
the exception of maritime arctic (mA) air mass, which does not exist, there are five
remaining air masses of the following types: cT, cP, cA, mT and mP. Air masses are not
necessarily confined to their source regions and they do move to other regions. This
migration causes an abrupt change in the temperature and humidity conditions where the air
masses move. Further, the air masses do acquire some of the characteristics of their new
regions.
Continental Polar (cP) air masses form over extensive land masses located at high
latitudes such as Siberia and northern parts of Canada. The winter cP air masses are cold,
extremely dry and are highly stable. The summer cP air masses are similar, except that they
are both warmer and more moist than their winter counterparts. Continental Arctic (cA) air
masses are colder than cP air masses and form over the arctic land masses. The cA and cP
air masses are separated by a front, called the arctic front. The arctic front is quite shallow
and does not extend to more than a kilometre in the vertical. Maritime Polar (mP) air
masses form over the high-latitude oceans and are similar to cP air masses except that they
are somewhat moderate in both temperature and in the extent of dryness. mP air masses
form over North Pacific as cP air mass moves air from the interiors of Asia. Continental
Tropical (cT) air masses form during the summer season over hot, low-latitude desert areas.
cT air masses are extremely hot and very dry. Maritime Tropical (mT) air masses form over
the warm tropical seas. mT air masses, unlike the cT air masses, are warm, moist and
unstable near the surface. The above combination provides the ideal conditions for the
development of clouds and rainfall. Table 10.1 outlines the general characteristics of the
different air masses.
270 u BASICS OF ATMOSPHERIC SCIENCE

TABLE 10.1 General characteristics of the air masses

Air masses Source region Source characteristics


Continental Arctic (cA) Greenland, Antarctica, Highest latitudes Very cold and very dry, extremely
of Asia and North America stable, very little cloud cover.
Continental Polar (cP) High latitudes, continental interiors Cold and dry, very stable, very little
cloud cover
Maritime Polar (mP) High-latitude oceans Cold, moist and cloudy, somewhat
unstable
Continental Tropics (cT) Low-latitude deserts Hot and dry, very unstable
Maritime Tropics (mT) Subtropical oceans Warm and humid

10.2 FRONTS
Fronts are boundary regions that separate air masses of contrasting temperature and humidity
characteristics. Mostly, fronts separate the tropical and the polar air masses. When two
contrasting air masses meet, the air in the two air masses does not readily and completely
mix. Typically, warm air glides over the cold air or the cold air undercuts the warm air. A cold
front occurs when a wedge of cold air advances towards the warm air ahead of it. A warm
front occurs when a warmer air mass moves towards a colder air. A stationary front is a front
that does not move. Unlike the above-mentioned fronts, an occluded front does not usually
separate tropical and polar air masses. Instead, occluded fronts form at the surface on the
boundary region between any two air masses, with the colder polar air mass usually
advancing on a slightly warmer air mass ahead of it. Fronts can be considered within a
reasonable approximation to behave as material surfaces in the atmosphere. This means that
the air parcels which form part of the frontal surface at a given instant of time will continue
to define the frontal surface at future times as well, indicating that air does not normally move
through a frontal surface. The passage of a front over a place is normally accompanied by an
abrupt change in temperature, shift in the wind speed and direction, change in the humidity
and a change in the cloud cover. It is to be noted that the transition between the contrasting
air masses takes place within a zone of finite width, called the frontal zone. The front can be
defined as the warm air boundary of the frontal zone. The horizontal temperature gradient
across the frontal zone can be as high as 10°C, over a distance of 100–200 km. Further, there
is no discontinuity in the temperature field across the front.

10.2.1 Warm Front


Fronts are named in terms of their direction of movement. If the air on the cold side of the
frontal zone is retreating and is being replaced by warm air, such a front is called a warm
front. Warm fronts are indicated on the weather maps by semicircular symbols, which point in the
direction of the frontal movement, in this case, in the direction of the cold air. Figures 10.1(a–c)
shows the idealized vertical cross-section through the frontal zone for all the three fronts with
the left panel of Figure 10.1(a) corresponding to a warm front. The above cross-section has
been drawn with warm and cold air to the left and right respectively. For all the three fronts
LARGE-SCALE METEOROLOGICAL SYSTEMS IN MID-LATITUDES u 271

as can be seen from Figures 10.1(a–c), the frontal surface slopes in the direction of the cold
air with increasing height; resulting in the frontal zone lying below the frontal surface and
warm air overlying the frontal surface. As can be seen, the warm air can either gently glide
over the frontal surface (Figure 10.1(a–b)) or can be forcibly lifted up as in Figure 10.1(c).
However, the air within the frontal zone (below the frontal surface) is trapped in a shallow
wedge and hence cannot move relative to the front. This ensures that the direction and the
speed of the movement of the front are solely determined by the air within the frontal zone.

T T T
T + dT T + dT T + dT

(a) (b) (c)

Figure 10.1 Vertical cross-section through frontal zones showing isotherms (dotted lines) and
air motions relative to the ground (continuous arrow). (a) Warm front, (b) stationary
front with overrunning warm air, and (c) cold front. Heavy arrows indicate the
direction of frontal movements.

In the warm front (Figure 10.1(a)), the warm air flows up along the frontal surface in a
process known as overrunning, a flow similar to the flow of air ascending over a mountain
range. The slope of a warm front is about 1:200 indicating a gentle and gradual slope. As the
warm air rises along the frontal surface, adiabatic cooling of the warm air leads to the
formation of low-level stratus clouds. As the warm air continues to raise, low and medium
clouds such as nimbostratus, altostratus, cirrostratus and finally cirrus develop at the sequence
mentioned above. As the warm front moves eastward and approaches a place, the cirrus cloud
is first noticed at that place, followed by cirrostratus, altostratus, nimbostratus and then
nimbus in that order. The clouds along a warm front exist in the warm air and these clouds
have smaller droplets and lower liquid water content. Hence the precipitation associated with
the warm front is generally considered to be light and steady. If the falling droplets from the
clouds in the warm air associated with the warm front are much warmer than the cold air
through which they fall through, the droplets can evaporate rapidly and give rise to frontal
fog. Warm fronts typically move slowly with speeds of about 20 km h–1. Since the slope of
the warm front is not steep, the lifting of the warm air can extend to much greater horizontal
distances. This together with slow movement of the warm front results in light and persistent
rainfall that can even last up to several days. Fronts, though plotted as a line on the surface
maps, are actually three-dimensional in nature and usually extend upward much above the
500 hPa level. The position of fronts is not usually plotted on upper-level weather maps since
adequate upper-level data are not available to delineate their locations. Further, the surface
location of a front is more significant than its upper-level location. For identifying warm front
location on surface maps, the following features are useful:
(i) Look for an area or zone where warm air advances towards cooler air,
(ii) Dew point temperatures typically increase behind the position of the warm front,
(iii) Wind direction usually shifts from southwesterly ahead of the front to south-
easterlies behind the front,
272 u BASICS OF ATMOSPHERIC SCIENCE

(iv) Bands of clouds and precipitation which nearly coincide with the position of the
warm front, and
(v) Area ahead of the warm front has decreasing atmospheric pressure, while the air
behind the warm front has stable air pressure.

10.2.2 Cold Front


If the air on the cold side of the frontal zone is advancing into a region formally occupied by
warm air, such a front is called a cold front (refer Figure 10.1(c)). Cold fronts are indicated
in a weather map by triangular shaped teeth which point in the direction of the frontal
movement. In the cold front, the warm air is forcibly lifted up as the cold air undercuts the
warm air. The slope of a cold front is about 1:100, and is higher than the slope associated
with the warm front. The cold front results in the convergence of two opposing air masses.
The difference in the wind speed and direction provide for cold northwesterly air to converge
on the warm air ahead of the cold front and force it upwards. Also, the warm air ahead of a
cold front tends to be unstable and hence can be easily lifted, leading to the formation of
cumuliform clouds along these boundaries. Due to their large vertical extent, the cumuliform
clouds can give rise to intense precipitations. Cold fronts typically move at widely varying
speeds ranging from even zero to about 50 km h–1. Since the slope of a cold front is very
steep, and also due to the nature of the convergence of the two opposing air masses, the
forced lifting of the warm air extends to much smaller horizontal distances. This together
with relatively faster movement of the cold front at most times results in intense
precipitations which are often of short duration. The cold front is normally accompanied by
well-organized cloud patterns, which assume the form of layered clouds such as altostratus
and nimbostratus or in the form of convective clouds such as cumulonimbus depending on
the static stability as well as the abruptness of the vertical lifting which results when the cold
air undercuts the warm air. In the rear of cold fronts, the weather clears up considerably,
however, in the marine polar air mass, the convective clouds can at times produce very heavy
precipitation locally. For identifying the cold front location on surface maps, the following
features are useful:
(i)Look for an area or zone where colder air advances towards warmer air,
(ii)Lower temperatures behind cold front,
(iii)Dew point temperatures typically decrease behind the cold front,
(iv) Wind direction typically changes from northwesterly in the cold sector to
southwesterly in the warm region,
(v) Bands of clouds and precipitation which nearly coincide with the position of the
cold front, and
(vi) Decreasing atmospheric pressure ahead of the cold front and increasing atmospheric
pressure behind the cold front.

10.2.3 Stationary Front


It may so happen that at certain time, the front continues to remain at the same place for
extended time periods. Such non-moving frontal surfaces are called stationary fronts
(refer Figure 10.1(b)). Thus, stationary fronts are similar to cold or warm fronts in terms of
LARGE-SCALE METEOROLOGICAL SYSTEMS IN MID-LATITUDES u 273

the relationship between the respective air masses. Stationary fronts are indicated in a weather
map by alternating cold and warm front symbols on either sides of the line. Similar to warm
and cold fronts, the frontal surface in a stationary front also slopes in the direction of the cold
air with increasing height. The process of identifying whether, a front is stationary, or not is
somewhat subjective. In practice, a meteorologist will determine whether a front is stationary
or not by examining any two neighbouring surface maps valid at three hourly intervals.

10.2.4 Occluded Front


The term occlusion means closure or overlapping; referring to the cutting off a warm air mass
from the surface due to the convergence of two cold fronts. Occluded fronts are indicated in
a weather map by alternating cold and warm front symbols on the side which points to the
movement of the occluded front. Occluded fronts form when a faster-moving cold front
overtakes the warm front ahead and in that process the warm air rises. At the surface, cold air
mass of fast-moving cold front merges with that of the other front with the result, smaller
differences in temperature (and also in dew point temperatures) are seen across the occluded
front. The warm air mass is lifted aloft and hence is not seen in the surface temperature.
In the cold-type occlusion (refer Figure 10.2(a)), the cold front extends to the ground
and the warm front exists only aloft. In the warm-type occlusion (refer Figure 10.2(b)), the
reverse is true. It is to be noted that the weather conditions prior to the frontal passage are
similar to those ahead of a warm front and conditions after the front passage are similar to
conditions behind a cold front. While the passage of a cold-type occlusion results in increase
in the static stability of the lower troposphere the passage of a warm-type occlusion is
accompanied by a decrease in the static stability.

T
T T + dT
T + dT T + 2dT
T + 2dT

(a) (b)

FIGURES 10.2 Occluded fronts: (a) cold type, and (b) warm type. The above schematic diagram
shows the frontal surface (continuous line) and isotherms (dotted line) in a vertical
section normal to the occluded fronts, which are moving from left to right.

10.3 EXTRATROPICAL CYCLONE


Extratropical cyclones are very large weather systems which form along a front in the middle
and high latitudes. These systems travel very great distances and very often provide extensive
precipitation over a very large region. These extratropical systems typically last for a week or
more and can cover extensive regions, amounting to even large portions of continents. The
passage of extratropical cyclones over a place is accompanied by, in addition to precipitation,
abrupt changes in the wind speed and direction, temperature, moisture and cloud conditions.
274 u BASICS OF ATMOSPHERIC SCIENCE

The above extratropical cyclones were first thoroughly investigated by the Norwegian school
led by Vilhelm Bjerknes in the early part of the twentieth century. Bjerknes and his
colleagues proposed a theory called the polar front theory, for the formation, growth and
dissipation of the extratropical cyclones. The above polar front theory has successfully
withstood the test of time. Despite a very large increase in the observations which we
presently have compared to the beginning of the twentieth century, the life cycle of an
extratropical cyclone is still described in much the same way as was proposed by Bjerknes in
his polar front theory.
The life cycle of an extratropical cyclone has the following sequence of events with the
respective characteristics. Figure 10.3(a–c) illustrates the various stages (initial, mature and
occluded) of the idealized model of the development of an extratropical cyclone as proposed
by the Norwegian school, shown as (a), (b) and (c) panels of Figure 10.3. The term
“cyclogenesis” refers to the formation of extratropical cyclones. Consider the development of
an extratropical cyclone in the northern hemisphere. Initially, the polar front separates the cold
polar easterlies from the warm mid-latitude westerlies. With the beginning of cyclogenesis, a
small “kink” appears along the frontal boundary. The cold air north of the front starts pushing
southward behind the cold front. Further, the air behind the warm front advances northward.
The above situation thus leads to a cyclonic anticlockwise rotation around a weak low-
pressure system which develops on the crest of a wave-like undulation. The above stage with
more intensification leads to a situation depicted in Figure 10.3(a) when the low pressure
deepens and distinct cold and warm fronts manifest from the original polar front.

L
L

(a) (b) (c)

FIGURE 10.3 The various stages (initial stage (a)), mature stage (b) and occluded stage (c) of a
middle-latitude cyclone showing the frontal surface and isobars of sea level
pressure. The directions of the arrows indicate the direction of the geostrophic
wind.

Low-level convergence associated with the low pressure can lead to vertical ascending
motion and cloud formation in this cyclogenesis stage. Figure 10.3(b) shows the mature
cyclone stage. The distribution of the isobaric pattern within the extratropical cyclone is
interrupted only along the cold and the warm fronts. This leads to abrupt changes in the wind
direction along the front boundaries. While the isobars are more or less straight in the warm
sector lying between the two fronts, they are mostly curved in the larger colder region. For the
warm front, the winds shift from the south-westerly in the warm sector to south-easterly on
LARGE-SCALE METEOROLOGICAL SYSTEMS IN MID-LATITUDES u 275

the cold side, while across the cold front, the winds shift from north-westerly on the cold side
to south-westerly in the warm sector. Bands of cumuliform cloud run along and ahead of the
cold front caused by the cold air undercutting and forcing aloft the warm air. Precipitation
along the cold front is highly likely towards the centre of the low pressure since the large-
scale convergence associated with the low-pressure complements the forced uplift of the warm
air in the cold front. Due to the high moisture content and unstable conditions, which prevail
ahead of the cold front, precipitation in the form of rain, snow, sleet or hail can be intense.
However, since the band of cumuliform clouds is rather narrow, the intense precipitation lasts
only for a brief period. The precipitation associated with the warm front tends to be light,
steady, extensive and long-lasting as earlier mentioned in the previous section.
Typically, clear skies characterize the region over the warm sector although squall lines
can occur over the warm sector under certain conditions. The third stage (refer Figure 10.3(c)),
called the occluded stage occurs when the low-pressure centre propagates towards the cold air
as it deepens. An occluded front forms connecting the low-pressure centre to the junction of
the warm and the cold fronts. West of the occluded front boundary, air flow is northwesterly
and is excessively cold. Relatively warm air approaches the occluded front from the east,
however, this air had originated in the cold sector of the cyclone. Thus, the temperature
difference across the occluded front is less than the temperature difference associated with the
original warm or cloud fronts. The occlusion stage signifies the end of the process of
cyclogenesis and the beginning of the dissipation of the system.
At any particular place, the passage of the extratropical cyclones which generally moves
eastward with either a northward or a southward component provides a reasonable predictable
sequence of changing sky conditions. With the approach of a warm front, cloud cover deepens
and increases accompanied by light to moderate precipitation. With the passage of the warm
front, warmer, sunny and relatively cloud-free conditions prevail. Further, with the passage of
the warm front, the wind shifts from a southerly to a south-westerly direction. These clear,
warm and sunny conditions, which occur when the warm sector air overlies the place, may
persist for a day or two. With the approach of the cold front, fast moving band of clouds,
chiefly of the cumuliform type and intense short-lived precipitation in the form of rain and
snow are seen. The rear of the cold front brings very cold and absolutely cloud-free
conditions. On an average 1071 extratropical cyclones form each year out of which 579 form
in the northern hemisphere, while 492 form in the southern hemisphere. The extratropical
cyclones of the southern hemisphere, during their intense stage, on an average have much
lower minimum pressure of about 972 hPa as compared to the corresponding value of
988 hPa for northern hemisphere.

10.4 JET STREAMS


Jet streams are meandering air currents, seen typically at heights of 9 to 12 km above the sea
level. Their average wind speeds are about 180 km h–1 in winter and about half the above
value in summer. It is known that the horizontal temperature gradient within an air mass is
quite less irrespective of the nature of the air mass. However, within the frontal zone, due to
sharp horizontal changes in temperature, one would expect to see steeply sloping pressure
surfaces and consequently strong pressure gradient. For the purpose of illustration, consider a
276 u BASICS OF ATMOSPHERIC SCIENCE

cold front. The pressure will decrease at a lower rate over the regions where the warm air
resides, while the pressure will decrease at a greater rate over regions where cold air is
present. This causes pressure surfaces to slope steeply across the frontal zone giving rise to
strong winds culminating in a jet stream. One would expect to find a jet stream as a
consequence of the polar front according to the above discussion. It turns out that such a jet
stream does exist and is known as the polar jet stream. The jet streams are characterized by
the existence of very strong vertical wind shear, i.e. a very large vertical gradient of the
horizontal wind vector. Like the polar front, the polar jet stream also affects greatly the day-
to-day weather in the mid-latitudes. The polar jet stream is westerly and can be explained on
the basis of the thermal wind relationship. The polar jet stream is found centred at 300 hPa
and closely follows the position of the polar front. The polar jet stream is at its most intense
in the winter months, since the horizontal temperature gradients are stronger in the winter
months.
The formal World Meteorological Organization (WMO) definition of the jet stream is as
follows:
A jet stream is a strong narrow current concentrated along a quasihorizontal axis
characterized by strong vertical and lateral wind shears and featuring one or two velocity maxima.
The speed of the wind should be greater than 30 m s–1. Jet streams extend about 1000 km
long, about 100 km wide and about 1 km deep with a vertical wind shear of about 5–10 m s–1
per km and a lateral wind shear of about 5 m s–1 per km. For the westerly jet stream in the
northern hemisphere, the left exit (III Quadrant in Figure 10.4) and the right entrance
(I Quadrant in Figure 10.4) regions are susceptible to disturbed weather. Figure 10.4 shows
the representation of relative vorticity in a horizontal plane in the neighbourhood of a westerly
jet stream in Northern Hemisphere. The Quadrants III and II are known as left exit and right
exit regions, while Quadrants IV and I are referred to as left entrance and right entrance
regions, respectively.

FIGURE 10.4 Schematic diagram of relative vorticity in the vicinity of a westerly jet stream in
southern hemisphere
LARGE-SCALE METEOROLOGICAL SYSTEMS IN MID-LATITUDES u 277

It is clear that the relative vorticity z is positive/negative for regions north/south of the
jet core, as can be shown by envisaging a paddle wheel and observing the shear vorticity so
produced. Since the maximum wind shears are associated with the jet core, it is clear that the
maximum z (largest positive value) is along AO and minimum z (largest negative value) is
along OB. Assuming a purely zonal flow with no meridional motion in southern hemisphere
E[ E
[  G

EU EU

and retaining only the divergence term in the vorticity equation, one gets for h = zÿ + f, the
following equation:
 EI
 ³ ¹ 7
(10.1)
I EU
For the left exit region (Quadrant III of Figure 10.4), one has
˜[ ˜[
[ !    V !  G   V   O  
˜Y ˜Y
˜[
Neglecting one obtains
˜U
E[ ˜[
V  
EU ˜Y
EI  EI
This means that   and so !  That is  ³ ¹ 7
!  and so ³ ¹ 7
  i.e. the
EU I EU
upper-level divergence in the left exit region is negative. For the left entrance region
(Quadrant IV of Figure 10.4), by similar arguments one has
˜[ ˜[
[ !  !  V !  G   V !  O  
˜Y ˜Y
That is
E[ EO  EO
!  
EU EU I EU
making  ³ ¹ 7
  ZJFMEJOH ³ ¹ 7
! 
Using very similar arguments for right entrance region (Quadrant I in Figure 10.4), one has
z < 0, f < 0
Since f is more than z in magnitude, the sign of h will be the same as that of f in this case.
Since z becomes more negative with increase of x,
˜[ ˜[ E[
  BOE TJODF V !  V   "MTP I  
˜Y ˜Y EU
 EI
This means !  and so  ³ ¹ 7
!  yielding ³ ¹ 7
 
I EU
278 u BASICS OF ATMOSPHERIC SCIENCE

Lastly, for the right exit region (Quadrant IV of Figure 10.4), one has
z < 0, and
˜[
again assuming that f is more than z, increase of x makes z less negative, making ! 
˜Y
This means
˜[ E[
V !  "MTP I  
˜Y EU
This leads to
 EI
  PS  ³ ¹ 7
  ZJFMEJOH ³ ¹ 7
! 
I EU
Hence, it is clear that both the left entrance and the right exit regions of the westerly jet
stream in southern hemisphere are associated with upper-level divergence. This upper-level
divergence will correspond to vertical upward motion and consequently cloud formation and
precipitation. A similar argument on the above lines for the westerly jet stream in northern
hemisphere yields just the opposite: the left exit and the right entrance regions are associated
with upper-level divergence and hence with disturbed weather. The above difference is
primarily due to the fact that h is considered negative in the southern hemisphere (again f is
assumed to be larger than z in magnitude), unlike the northern hemisphere where h is
considered positive.
Over subtropics another jet stream called the subtropical jet stream is seen. Like the
polar jet stream, the subtropical jet stream is a westerly air current. The subtropical jet
stream is associated with the Hadley circulation (the thermally direct meridional circulation
with ascending regions close to equator and descending regions over the subtropics). As the
upper air flow of the Hadley circulation moves northward, it gets deflected by the Coriolis
force giving rise to air currents which can assume jet-like strength. The subtropical jet
stream is found centred at 200 hPa and is at its most intense during the winter months.
During the summer months, the subtropical jet stream moves poleward and weakens
considerably.

REVIEW QUESTIONS

1. Mention the contrasts, if any, in the formation of air masses in the northern and southern
hemispheres?
2. Which type of fronts is most likely to have inversions?
3. Which type of fronts is least likely to have inversions?
4. While continental polar air masses can migrate to the south-eastern parts of the
United States of America, in winter, the same is not possible over northern India. Why?
5. Name the system which provides for most of the winter-time precipitation over India.
6. Imagine that a region is occupied by the continental air mass. What would the height of any
upper tropospheric isobaric surface, say 200 hPa level be like?
7. Mention the type of clouds one is likely to associate with a warm front.
LARGE-SCALE METEOROLOGICAL SYSTEMS IN MID-LATITUDES u 279

8. Mention the type of clouds one is likely to associate with a cold front.
9. What are occluded fronts?
10. What are jet streams?
11. Name the various jet streams seen over the Indian region.
12. Mention the regions where the precipitation is most likely to be seen within an extratropical
cyclone.
13. Consider a typical trough and a ridge system. In which part of the above system, will one
find regions of increasing and decreasing vorticity?
14. Consider a westerly jet stream in the northern and southern hemispheres. Where are the
upper-level convergence and divergence most likely to occur near the westerly jet stream?
11 Meteorological Systems
in Low Latitudes

Unlike the middle-latitude and high-latitude regions, which are differentially affected by the
seasonal march of the sun, the regions in low latitudes are much less affected with the result
seasonal temperature variations are often small here. As much as half of the earth surface lies
between 30°N and 30°S. Further, the presence of warm tropical oceans and the inherent
complexity of the atmospheric processes which determine low-latitude weather and climate
make the study of low-latitude meteorological systems a very important one. One of the most
illuminating of the large-scale meteorological systems in the tropics is the tropical monsoon
circulation which has its spectacular manifestation over the Indian subcontinent. Within the
overall planetary scale monsoon circulation over the Indian region are embedded weather
systems such as monsoon depressions and low-pressure systems, which provide large amounts
of rainfall over several regions of India. One of the important weather systems found in the
tropics and known for its destructive fury to life and property is the tropical cyclone or hurricane.
Two additional weather systems, which are typically small-scale but known for their
destructive damage, are the thunderstorm and the tornado. One of the most important
manifestations of the ocean–atmosphere interaction is the El Nino-Southern Oscillation—an
event which forms in the low-latitude regions. Section 11.1 outlines the general consideration
of the meteorological systems in the low latitudes. Section 11.2 introduces the tropical
monsoon circulation and its characteristics, while Section 11.3 provides the description of
monsoon depression. Sections 11.4 and 11.5 address the characteristics of some of the most
destructive of the low-latitude weather systems such as tropical cyclones, thunderstorms and
the tornado, respectively. The last Section 11.6 provides for the description of the El Nino-
Southern Oscillation events in the low-latitude regions.

11.1 GENERAL CONSIDERATIONS


The meteorological systems associated with the low-latitude regions are very different from
those which originate in the middle and high latitudes. Due to the large amounts of insolation
280
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 281

received at low-latitude regions, convection is the most important physical mechanism


operating in the tropics. Unlike the middle and high latitudes, meridional temperature
gradients in the low-latitude regions are lower and weaker. Hence, the concepts of fronts are
unimportant for studying the low-latitude meteorological systems. Further, the geostrophic (as
well as quasi-geostrophic) balance relationships do not very well hold in the low latitudes.
Discounting diurnal changes, typically, 24-hour changes in surface temperature and pressure
in the low latitudes are much less (of the order of 1°C and 1 hPa). The distinct seasonality of
weather with dry and wet seasons is an important feature of the low latitudes. Further, within
each season in the low-latitude regions one has the manifestation of a well-marked diurnal
cycle of weather.

11.2 MONSOONS
The livelihood of over sixty per cent of the world’s global population is directly dependent on
the monsoons. The Asian monsoon, one of the largest among the global monsoons, plays a
key role in maintaining the earth’s climate system. Also, the Asian monsoon has important
teleconnections with global weather and climate. Monsoons can be thought of three-
dimensional circulations associated with the global distribution of sea and land. A few
common characteristic features of monsoons are:
(i) The winds over the lower troposphere blow from a high-pressure region to a low-
pressure region.
(ii) The low-pressure region is normally manifested as a trough in the surface pressure
field.
(iii) The monsoonal circulation, has a vertical structure.
The following four criteria were proposed in 1971 by Ramage in his monograph titled
Monsoon Meteorology to determine whether a region can be defined as monsoonal, and these are:
(i) The prevailing wind direction shifts by at least 120° between January and July.
(ii) The mean frequency of the prevailing wind direction in January and July exceed
40%, i.e. if a histogram is made of wind direction in the month of January/July of
a monsoon location, more than 40% of the time would correspond to the direction
of the prevailing wind.
(iii) The mean resultant winds in January and July exceed 3 m s–1.
(iv) Fewer than one cyclone–anticyclone alternation occurs every two years in either
January or July in a 5o latitude ´ longitude rectangle.
Nowhere in the world are the seasonal changes in the mean pressure pattern more
dramatically manifested than in its effects over the Asian continent in general and over the
Indian subcontinent in particular. The above seasonal changes in the mean pressure pattern are
associated with the mean seasonal changes in the atmospheric circulation, the so-called
monsoonal circulation. The surface wind charts for the months of January (winter) and July
(summer) over the Indian Ocean are shown in Figures 11.1 and 11.2.
282 u BASICS OF ATMOSPHERIC SCIENCE

January Wind
30°N

25°N

20°N

15°N

10°N

5°N

E0

5°S

10°S

15°S

20°S

25°S

30°S
30°E 40°E 50°E 60°E 70°E 80°E 90°E 100°E 110°E 120°E 130°E

FIGURE 11.1 Surface wind chart for the month of January over the Indian Ocean.

FIGURE 11.2 Surface wind chart for the month of July over the Indian Ocean.
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 283

The surface wind flow shown in the month of July is typical of the monsoonal flow
pattern over the Indian subcontinent. Figures 11.3 and 11.4 depict the normal onset and the
normal withdrawal dates of the summer monsoon over the Indian region.

SRN LEH

JUL 1

CNG

JUL 15
DLH JUN 15
BRL
JUN 5 DBH
JUN 10
JUL 15 JPR LKN
JDP GHT
PTN
IMP
JUL 1
AGT JUN 1
BHJ RNC CAL
BHP JBP
JUN 15 AHM
RJK
NGP
BWN

JUN 10
AGT
MUM

PNE
VSK
HYD

PNJ

JUN 5 ANT
CNN
MNG
BNG
PBL
AMN

TRP

TRV
JUN 1
MNC

FIGURE 11.3 Normal onset dates of the summer monsoon over the Indian region.
284 u BASICS OF ATMOSPHERIC SCIENCE

SRN LEH

SEP 1 SEP 15

CNG

JUL 15
DLH BRL
DBH
OCT 1
JPR LKN
JDP GHT
SEP 1
PTN
IMP

BHJ AGT
SEP 16 BHP JBP RNC CAL
AHM
RJK

NGP BWN

MUM AGT
OCT 15
PNE
VSK
HYD
OCT 1
PNJ

ANT
CNN
OCT 15 MNG
BNG
AMN PBL

TRP

TRV
MNC

FIGURE 11.4 Normal withdrawal dates of the summer monsoon over the Indian region.

The arrival of the monsoon over Bombay, in the words of the first Prime Minister of
India, convey the sense of anticipation of an average Indian to the onset of monsoon rains
over India. Jawaharlal Nehru, the first Prime Minister of India, wrote:
I looked forward to the coming of the monsoon and I became a watcher of the
skies, waiting to spot the heralds that preceded the attack. A few showers came.
Oh, that was nothing, I was told; the monsoon has yet to come. Heavier rains
followed, but I ignored them and waited for some extraordinary happening. While
I waited, I learnt from various people that the monsoon had definitely come and
established itself. Where was the pomp and circumstance and the glory of the
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 285

attack, and the combat between cloud and land, and the surging and lashing sea?
Like a thief in the night the monsoon had come to Bombay, as well it might have
done in Allahabad or elsewhere. Another illusion gone.
The normal onset date of the summer Indian monsoon over land occurs on June 1 over
Kerala and by July 15, the entire country (India) is covered and comes under the spell of
monsoon. The normal withdrawal of the summer monsoon begins on September 1 from the
northwestern part of India and by October 15 the monsoon has withdrawn from all parts of
the country except for the southern peninsula. The onset of Indian monsoon over Kerala is
primarily based on precipitation and is defined as the date of the transition from light to heavy
precipitation sustained over a few days and encompassing several rain gauge stations over
Kerala. The actual definition of the onset over Kerala is as follows:
If after May 10 of any year, 60% of the available fourteen rain gauge stations, such as
Minicoy, Amini, Thiruvananthapuram, Punalur, Kollam, Allapuzha, Kottayam, Kochin, Trissur,
Kozhikode, Talassery, Cannur, Kasargode and Mangalore report rainfall of 2.5 mm or more
for two consecutive days, the onset of monsoon over Kerala may be declared on the second
day, provided the following additional criteria are also satisfied:
(i) Depth of the westerlies must extend to 600 hPa over the region from equator to
10°N and 55°E to 80°E with the zonal wind speed at 925 hPa over the region from
5°N to 10°N and 70°E to 80°E having values 15–20 knots.
(ii) The outgoing long wave radiation having value less than 200 W m–2 over the
region from 5°N to 10°N and 70°E to 75°E.
It is clear from Figures 11.3 and 11.4 that while the progress of the monsoon across
India, is rather swift, the withdrawal of the monsoon across the country is relatively slower.
Figure 11.5 shows the mean monsoon rainfall over India extending from the month of June to September.
It is clear from Figure 11.5 that the mean monsoon rainfall shows large spatial variation
with as much as 250 cm rainfall over the west coast of India and Assam, a state in North-East
India and as low as 15 cm rainfall over Rajasthan, a state in North-West India. The normal all
India summer monsoon rainfall (AISMR) from June to September for the entire country using
data from 1901 to 2003 is about 877 mm which accounts for nearly three-fourths of the mean
annual rainfall of 1183 mm. The interannual standard deviation of the AISMR is 87 mm,
about 10% of the seasonal mean. The mean July rainfall contributes the maximum (24.2%) to
the annual rainfall, while the August, September and June rainfalls contribute, respectively,
21.2%, 14.2% and 13.8% of the total annual rainfall. The balance between the annual all
India rainfall and the AISMR is contributed nearly equally by the pre-monsoon (April and
May) and the post-monsoon (October and November) rainfall. The individual contributions of
July, August, June and September to AISMR works out to 32.6%, 28.6%, 18.6% and 19.1%).
This section, while outlining the physical processes, which determine monsoonal circulation,
will discuss only the Indian monsoon, in some detail, due to brevity. One reason for the
spectacular manifestation of the seasonal reversal of surface winds that characterizes the
monsoon to take place over the Indian subcontinent is the role played by the Himalayan
mountains to the north and the geographical location of India which is surrounded on the
three sides by waters of the Indian Ocean. The Himalayan mountain ranges act as a barrier,
blocking the northward flow of the moisture-laden air during the summer months. Further, the
Himalayas also block the cold southward outflow from regions of Siberia during the winter
286 u BASICS OF ATMOSPHERIC SCIENCE

months. The Himalayan Mountains also alters the upper-level tropospheric flow over the
Indian monsoon regions which ultimately influence the surface conditions during the monsoon
season. The name “monsoon” has come from the Arabic word mausin, which means,
“season”. The monsoon refers to the atmospheric climate pattern in which heavy and
extensive precipitation alternates with very hot and dry conditions on an annual basis. During
the northern winter months such as January, the low-level winds typically flow as
northeasterlies towards the Indian Ocean from the hills of the Himalayas. The descending air
from the Southern Himalayas gets adiabatically warmed resulting in very dry conditions over
India and South-East Asia. During January, the subtropical jet stream is found south of the
Himalayas and this jet stream maintains the offshore flow away from the land regions. Over
the India region in January the subtropical jet stream winds are convergent at upper levels
leading to subsidence and warming.

10
33°N 75 30
40
20
100 20
10 20 30
30°N
50
100 30
50 40 50
27°N 75 100 75
10 20 150
100

75 100 200
24°N 75
100
50
200
21°N
150
75
40
100
18°N 75

50 150
50
15°N 40
100
50
200
12°N
30 75
20 100
150
9°N 100
75 20
30
75
10
6°N 40
20
66°E 69°E 72°E 75°E 78°E 81°E 84°E 87°E 90°E 93°E 96°E

Figure 11.5 The mean monsoon rainfall in cm over India extending from the month of June to
September.
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 287

The above atmospheric pattern changes dramatically during the late spring/early summer
season when the seasonal reversal of surface winds occurs, due to heating of the Indian land
mass. At the lower levels, winds blow onshore bringing warm, moist and unstable air from
the Indian Ocean to the southern parts of the Indian peninsula. Though the initial land sea
thermal contrast (with land and sea over the summer and winter hemispheres), initiates the
monsoonal circulation, the above alone cannot sustain the deep planetary scale monsoonal
circulation seen over the Indian region. For, land sea contrasts due to surface heating effects
can only result in a shallow circulation confined to the low levels of the atmosphere and
cannot produce the deep vertical circulation associated with the Indian monsoon. Further,
once the Indian summer monsoon has set in over India, the surface temperatures over the
Indian land mass are lower than the temperatures over the sea. The above discussion
provides conclusive evidence for the existence of deep tropospheric heat sources over the
Indian land mass for the sustenance of the planetary scale Indian monsoon. The generation of
deep tropospheric heat sources can be explained as follows. The large amounts of
accumulated energy (in sensible and latent heat form) stored in the warm moist low-level air
due to its passage over a vast expanse of the Indian Ocean is released very selectively over a
localized Indian land mass region in the form of latent heat of condensation. The release of
the latent heat of condensation over the Indian land mass of relatively limited geographical
extent heats up the atmosphere so much that the meridional temperature gradient in the
troposphere gets reversed. The extensive heating over the Tibetian Plateau—an elevated land
mass with an average height of 4000 m, also plays an important role in the seasonal changes
of the meridional gradient of heating.
The three most fundamental driving mechanisms for the planetary scale monsoon
circulations are:
(i) The differential heating of the land and ocean in the summer and the winter
hemispheres and their associated pressure gradients that drive the monsoonal winds
from high pressure to low pressure.
(ii) The vorticity or the swirl introduced to the monsoonal winds due to the earth’s
rotation.
(iii) The moist processes which determine the intensity and the location of the monsoon
precipitation.

11.2.1 Differential Heating of Land and Sea


The term “differential heating of land and sea” refers to the non-uniform distribution of
heating and cooling over the land and sea areas of the earth. If some regions of the earth are
heated or cooled more than some other regions, this will give rise to motion of the
atmosphere. The above differential heating is due to the following:
(i) The space–time variation of the solar insolation to earth due to the nature of the
underlying earth surface and to the annual cycle of the sun.
(ii) The space–time variation of the cooling of the earth to space.
Since the earth is a sphere and is tilted 23.5° to the solar plane, there is a latitudinal
distribution of the incoming solar energy. The actual amount of solar energy received at the
earth surface, after accounting for cloud, aerosols and ground reflection, depends on the
288 u BASICS OF ATMOSPHERIC SCIENCE

intensity of the incoming solar radiation as well as the length of the day. These lead to an
uneven distribution of incoming solar energy received at the earth surface with the largest
energy received over the equatorial regions and the summer hemisphere, while the smallest
energy is received over the winter hemisphere. The loss of energy from the earth to space
would after accounting for cloud absorption depend on the temperatures themselves. That is,
regions with warmer temperatures such as equatorial regions and the summer hemisphere will
lose more radiation than the winter hemisphere. However, the loss of radiation to space varies
by less than 30% over the entire globe. The net radiation (gain minus loss) shows that the
winter hemisphere loses far more heat than it gains, while the opposite is true for the summer
hemisphere, indicating that a strong differential heating exists between the summer and the
winter hemispheres.
Further, the nature of the underlying surface (land or sea) and its response to a given
heat input varies providing additional reasons for differential heating. The reasons for the
differential response by land/sea to a given heat input is due to
(i) specific heat of water is twice that of soil,
(ii) the capacity of the seas (effective heat capacity) to store heat is very much larger
than that of land,
(iii) transfer of heat through the soil is brought out by the slow and inefficient process
of molecular conduction while the same over the sea is due to the efficient process
of turbulent mixing, and
(iv) over land, all the incident solar radiation is absorbed at the top soil surface, while
over the seas, the solar radiation can penetrate through a certain depth of water.
The continuous differential heating will successfully maintain a column of dense air
over the ocean and a column of relatively lighter air over the land together with
effective maintenance of the pressure gradient force ensuring a continuous
monsoonal circulation.

11.2.2 Compressibility, Rotation and Moisture Effects


Since the atmosphere is a gas, the density of air is inversely proportional to its temperature, at
constant pressure. Due to the above, the height of a constant pressure surface (say 500 hPa) is
lower over regions where temperatures are lower. A consequence of differential heating would
lead to a difference in temperature with latitude. This would correspond to horizontal
meridional gradients of heights of constant pressure surfaces or horizontal pressure gradients
at constant height surfaces. Hence, due to compressibility and differential heating effects,
there exists a strong meridional pressure gradient between the monsoonal regions in the
summer hemisphere and the regions in the winter hemisphere. Over the land in the summer
monsoon region, surface pressure is lower, while in the middle and upper troposphere,
pressure is higher over the summer monsoonal land regions. Thus, the pressure gradient
actually reverses with height and also increases in magnitude with height. The surface
pressure gradient directs air from the winter hemisphere towards the heated continent, while
the strong reversed pressure gradient at the upper troposphere returns the air to the winter
hemisphere. The above circulation is completed with air rising over the heated continents and
sinking air over the winter oceanic regions. The monsoon circulation is also known as the
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 289

reverse Hadley circulation as the ascent is over higher latitudes and descent over the low
latitudes—a situation exactly reversed to the Hadley circulation. The effect of rotation ensures
that the winds move from the winter hemisphere to the heated continent at low levels and out
of the heated continent at the upper levels in a less direct manner. The Coriolis force deflects
air to the right in the northern hemisphere and to the left in the southern hemisphere. Since
the effects of the Coriolis force are very small at and near equator, air close to equator will
directly flow from the region of high pressure to the region of low pressure. At the low levels,
the monsoonal flow after crossing the equator gets deflected to the right by Coriolis force and
approaches India as southwesterlies. By the same arguments, the return flow from India at
higher levels is in the north-easterly direction.
The effects of moisture constitute enormously to the intensification of the monsoonal
circulation as can be seen from the following discussion. When the monsoonal air is moist
and saturated, the decrease of temperature with height, the so-called saturated adiabatic lapse
rate is much less than that for dry air. Further, the warmer the monsoonal air, the larger the
amount of water vapour it can hold. Hence, when the warm and moist monsoonal air is forced
to ascend over the heated continent, the resulting change of phase of water vapour to water
and/or ice is accompanied by the release of the latent heat of condensation. The hotter and
more saturated a rising air parcel, more will be the release of latent heat and this will
intensify the pressure gradient force and the associated monsoonal circulation.

11.2.3 Tropical and Oceanic Convergent Zones


The low-latitude regions are characterized by the existence of low-level trade winds, north
easterlies to the north and south easterlies to the south of equator. These trade winds converge
in a broad band of convergent zone. This convergent zone which is east-west oriented is
characterized by a broad belt of precipitation maxima and is called the Tropical Convergent
Zone (TCZ). Some books refer the above as Intertropical Convergent Zone (ITCZ). The TCZ
does migrate with the motion of the sun in its journey from the Tropic of Cancer to the
Tropic of Capricon and back. During the northern winter, the mean position of the TCZ over
the Indian Ocean is 5°S and it experiences its largest migration over the Indian region where
it is seen around 20°N over the Indian continent during the northern summer season. Once the
summer monsoon has set in over India, the large-scale monsoonal flow at lower levels gives
rise to a large-scale cyclonic vorticity. The above is associated with a pressure trough, called
the monsoon trough, oriented in the north-west-south-east direction in the surface map
running from Ganga Nagar in Rajasthan, North-West India to the northern Bay of Bengal.
The low level, large-scale cyclonic vorticity aids in the organization of convection and in
maintaining the TCZ over the Indian continent.
A second TCZ is seen over the equatorial Indian Ocean and is called the Oceanic
Convergent Zone (OCZ). It appears that the sea surface temperature (SST) over the northern
Indian Ocean plays an important role in sustaining and controlling the OCZ. During the
northern summer (Indian monsoon) season, the OCZ is maintained by the meridional gradient
of the SST at the location of the SST maxima over the eastern equatorial Indian Ocean. The
SST over the Indian Ocean is essentially determined by the net heat flux at the sea surface
and the forcing due to the wind stress. The wind stress is related to the low-level monsoonal
290 u BASICS OF ATMOSPHERIC SCIENCE

flow which in turn is determined by the amount of release of the latent heat associated with
the precipitation over the continental TCZ. Observations indicate that when the continental
TCZ is well developed and strong, active Indian monsoon conditions prevail and during this
phase the OCZ is found to be weak. During the break like situation (lull in the monsoonal
rainfall) over India, the OCZ is very strong with a corresponding weakening of the continental
TCZ. It is apparent that both the continental and the oceanic convergent zones compete for
the moisture in the low-level monsoonal flow and together determine the active and break
spells of the Indian monsoon season.

11.2.4 Intraseasonal and Interannual Variability of the Indian


Monsoon
Although the Indian summer monsoon shows up unfailingly year-after-year, the fact remains
that the Indian summer monsoon is characterized by variations in the rainfall amounts both
from year-to-year as well as within the monsoon season. The variability of the Indian summer
monsoon from year-to-year is called interannual variability, while the variability of the Indian
summer monsoon within the monsoon season is called intraseasonal variability. Both the
intraseasonal and the interannual variability of the Indian summer monsoon will be discussed
in the following subsections.

Intraseasonal Variability
The so-called active (wet) and break (dry) spells of the monsoon precipitation over the Indian
region are nothing but the manifestation of the successive, northward propagation of the TCZ
from an oceanic position close to the equator to the continental position over the Indian land
mass. Spectral analyses of meteorological data over the Indian region during the summer
season have yielded two dominant time periods: 10–20 days oscillation and a 30–60 days
oscillation. The 10–20 days oscillation is characterized by a westward propagation, while the
30–60 days oscillation is characterized by a northward propagation over India. While the
10–20 days oscillation has a smaller zonal scale and is somewhat regional in nature, the
30–60 days oscillation has a very large zonal scale encompassing regions of both South and
East Asia. It is found that both the 10–20 days as well as the 30–60s day oscillation
contribute more or less equally to the total intraseasonal variability of the Indian monsoon. It
is found that there exists similarity between the large-scale spatial patterns of the dominant
intraseasonal oscillation (ISO) with the spatial structure of the seasonal mean monsoon.
Hence, it is suggested that the ISO, through their relative frequency of occurrence of the wet
and dry spells actively influence the seasonal mean and also contribute to the interannual
variability (IAV) of the Indian monsoon. Studies have indicated that the 10–20 days
oscillation is nothing but an n = 1 equatorial Rossby wave with about 6000 km wavelength
and a period of 14–16 days driven primarily by a convective feedback involving boundary
layer convergence. The mechanism to explain the 30–60 days oscillation involves eastward
propagation of convection in the equatorial Indian Ocean in the form of Kelvin wave and a
west-north-west propagation of Rossby waves originating from the western Pacific. A second
mechanism not invoking the wave dynamics but based on the “convective–thermal relaxation
feedback mechanism” has also been put forward to explain the 30–60 days oscillation.
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 291

Interannual Variability
Figure 11.6 presents the time series evolution of all India summer monsoon rainfall (AISMR)
anomalies, expressed as per cent departures from its long-term mean, for the years 1871–
2003. Some of the recent El Nino and La Nina events are also indicated in the same figure.
El Nino/La Nina events will be discussed in greater detail in Section 11.6. Typically, the
flood/drought years are defined as years with the AISMR in excess of/less than, one standard
deviation above/below the mean (i.e. anomaly exceeding 10%). During the period 1871–2007,
there were 19 major flood years and 23 major drought years. The major flood years are: 1874,
1878, 1892, 1893, 1894, 1910, 1916, 1917, 1933, 1942, 1947, 1956, 1959, 1961, 1970, 1975,
1983, 1988, and 1994, while the major drought years are: 1873, 1877, 1899, 1901, 1904,
1905, 1911, 1918, 1920, 1941, 1951, 1965, 1966, 1968, 1972, 1974, 1979, 1982, 1985, 1986,
1987, 2002 and 2004. It is important to note that there have been alternating periods
extending to 3–4 decades with less and more frequent weak monsoons over India.

Figure 11.6 Time series evolution of all India summer monsoon rainfall anomalies, expressed
as per cent departures from its long-term mean, for the years 1871–2003. Some of
the recent El Nino and La Nina events are also indicated.

Table 11.1 presents the decadal mean, expressed as a percentage departure from normal,
as well as the frequency of drought and flood years. It is interesting to note that over a
44-year period from 1921–1964, India as a whole experienced only two drought years (1941
and 1951), while for the period from 1965–1987, a period of 23 years, the country witnessed
as much as ten droughts (1965, 1966, 1968, 1972, 1974, 1979, 1982, 1985, 1986, 1987).
Since 1987, India did not experience a major drought until the year 2002, i.e. after a 15-year
292 u BASICS OF ATMOSPHERIC SCIENCE

TABLE 11.1 Decadal mean (expressed as a percentage departure from normal,


as well as the frequency of drought and flood years)

Decade Percentage departure Frequency of Frequency of


from normal deficient years excess years
1901–1910 –2.2 3 0
1911–1920 –2.5 4 3
1921–1930 –0.4 1 0
1931–1940 1.7 1 1
1941–1950 3.3 1 1
1951–1960 2.5 1 3
1961–1970 –0.1 2 1
1971–1980 –0.8 3 1
1981–1990 –0.3 2 2
1991–2000 0.6 0 1
2001–2003 –5.9 1 0

period. The interannual variability of the Indian monsoon is not very marked with the
interannual standard deviation being about 10% of the seasonal mean. Further, the amplitude
of the intraseasonal variability of the Indian monsoon is much larger than its amplitude of the
interannual variability. One of the important causes for the interannual variability of the Indian
monsoon is the El Nino-Southern Oscillation (ENSO) phenomenon. Although the ENSO will
be dealt in greater detail in a later section, it is to be noted that the warm/cold phase of the
ENSO called the El Nino/La Nina is generally associated with droughts/floods over the Indian
monsoon region. The above interaction between the ENSO and the Indian monsoon is made
possible through changes in the equatorial Walker circulation affecting the regional reverse
Hadley circulation associated with the Indian monsoon. Another important large-scale forcing
responsible for the interannual variability of the Indian monsoon is the snow cover over
Eurasia. Observations indicate a weak negative correlation between the Eurasian snow cover
and the strength of the Indian monsoon. Increased snow cover over central and southern
Eurasia would reduce the land–ocean temperature contrast and hence decrease the strength of
the Indian monsoon. The Indian monsoon also possess a biennial tendency—the above
biennial tendency is explained through a physical mechanism based on ocean—atmospheric
coupling.

Seasonal Prediction of the Indian Summer Monsoon


India, in the year 1877, experienced one of its worst famines, caused by a major drought due
to highly deficient summer monsoon rainfall for that year. The Government of India invited
H.F. Blanford, who had helped establish the India Meteorological Department (IMD), in the
year 1875, to assist the government with the preparation of monsoon forecasts. Blanford was
the first to attempt a monsoon forecast in the year 1882 and his forecasts were based on the
hypothesis that the horizontal extent and thickness of snow in the Himalayas had a profound
effect on the weather and climate over the plains of northwest India. Blanford continued his
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 293

monsoon forecasts till he demitted office in the year 1895. The people who followed him,
such as Sir John Elliot, and Sir Gilbert Walker, to name a few, continued and expanded the
scope of the monsoon forecast work initiated by Blanford.

Statistical Methods
Sir Gilbert Walker, the then Director General of IMD in the beginning of the 20th century had
initiated wide-ranging studies of the worldwide variation of weather parameters such as
pressure, temperature, rainfall, etc. Walker’s chief goal was to develop an objective method
for long-range forecasting (LRF) of the summer monsoon rainfall over India. Walker made
seminal contributions to the LRF methods of forecasting monsoon rainfall over India by first
introducing the concept of correlation. Walker also, was responsible for identifying three
large-scale pressure seesaw patterns; two in the northern hemisphere (North Atlantic
Oscillation, (NAO) and North Pacific Oscillation (NPO)) and one in the southern hemisphere
(Southern Oscillation (SO)). Unlike the NAO and NPO, which are essentially regional in
nature, the SO has emerged as a large-scale phenomenon with global-scale influences.
Most of the seasonal forecasting of the monsoon rainfall over India is based on
statistical and empirical methods. Despite their limitations in predicting very accurately the
extremes in the Indian summer monsoon, one season ahead, the statistical methods are still
very much in use in India. Diagnostic studies of historical data sets, spread over many years,
have helped identify several reliable predictors for the monsoon rainfall forecasting by
analyzing the relationships between AISMR and regional/global fields of several surface/
upper-air parameters. These predictors corresponding to the previous winter/spring season are
associated with both regional and global conditions. Among the regional predictors used are,
the pre-monsoon surface pressure and thermal fields over India, latitudinal location of the
April 500 hPa ridge position over India, and strength of the May 200 hPa winds over India.
Also, the widely used predictor over Pacific Ocean is the Southern Oscillation Index (SOI),
which represents a measure of the strength of the Walker Circulation across the Pacific, and is
defined as the normalized difference between the SLP anomalies at Tahiti in East Pacific and
Darwin from West Pacific. Additional global scale predictors are, the northern hemisphere
winter (January and February) surface air temperature anomaly, the wintertime Eurasian/
Himalayan snow cover and the quasi-biennial oscillation.
The various commonly used statistical techniques for seasonal forecasting of AISMR are
those based on, (i) linear regression using both simple as well as multiple regressions,
(ii) autoregressive integrated moving average (ARIMA) method, (iii) parametric and multiple
power regression (MPR) models, and (iv) dynamic stochastic transfer (DST) models. Attempts
have also been made to use the “feed-forward neural network methods” for forecasting AISMR.

General Circulation Models


The number of studies on the seasonal prediction of Indian summer monsoon rainfall using
general circulation models (GCMs) is quite small, when compared with the number of studies
using statistical and empirical methods. The reason for the fewer number of GCM studies are
partly due to the lack of skill in the simulation of mean monsoon rainfall over the Indian
subcontinent by most of the GCMs. Furthermore, there are also marked differences in the
mean monsoon rainfall simulated over India by the different GCMs. Also, it turns out that the
294 u BASICS OF ATMOSPHERIC SCIENCE

simulation of the summer monsoon rainfall over the Indian region is very sensitive to the
initial conditions. However, some seasonal GCM studies have successfully simulated both the
strong and weak monsoon circulations over India based on the sea surface temperature (SST)
distributions over tropical Pacific and Indian Oceans. Hence, while an accurate and realistic
simulation of the mean monsoon rainfall over India in absolute terms is yet to be achieved by
most GCMs, the sensitivity of the model monsoon to interannual variability in tropical
circulation appears to be in good agreement to the observed characteristics of the monsoon.
This is indeed a positive development for the seasonal integrations of the Indian summer
monsoon using GCMs.
It was noted in the earlier section that the Indian monsoon intraseasonal oscillations are
associated with quasi-periodic oscillations and the amplitude of the intraseasonal variability is
much larger than the amplitude of the interannual variability of the Indian monsoon. The
above characteristics provide some indication of the potential predictability of the
intraseasonal oscillation phases beyond the medium-range weather prediction of two weeks.
Studies have indicated that the break phase of the ISO is potentially more predictable with a
predictability of three weeks while that of the active phase is smaller. It is no wonder that the
break spell corresponding to the cessation of convection is more predictable than the active spell.
Several studies have indicated that the interannual variability of the tropical climate is
determined primarily by the slowly varying boundary conditions such as the SST anomalies.
Due to the above dominance of the boundary forced variability, tropical climate is assumed to
be more predictable than the extratropical climate. While the above observation is more or
less true in most parts of the tropics, some studies indicate that the Indian monsoon is an
exception to the above general rule. This indicates that the interannual variability of the
seasonal Indian mean monsoon is determined by variability due to internal dynamics in
addition to the effect of anomalous boundary conditions. Modelling and observation studies
have indicated that the contribution of the component, which corresponds to the internal
dynamics, is as large as the boundary force component as far as the Indian summer monsoon
variability is concerned. The above would naturally limit the predictability of the Indian
summer monsoon. On the seasonal time scale, observations indicate that the SST has a
negative feedback with precipitation over the eastern Indian Ocean and the North-West Pacific
Ocean. Most global atmospheric models, when forced with observed SSTs, however, reveal an
opposite response to the one shown by observations. Hence, many of the global atmospheric
models have inherent inadequacies as far as predicting the Indian summer monsoon is
concerned. A solution to the above problems would be to utilize a coupled ocean atmosphere
model for achieving better predictive skill of the Indian summer monsoon. After all, on a
seasonal time scale, the ocean is an important component of the monsoon climate system and
a dynamic ocean coupled to the atmosphere seems to be the right choice. However, most of
the current coupled ocean atmosphere models have not reached a stage where they can be
used as a viable option to achieve improved skill in predicting the Indian summer monsoon.
This is because most of the current coupled ocean atmosphere models are unable to simulate
accurately the annual cycle over the Indian monsoon region as they exhibit very large
systematic biases. Any progress in the use of the coupled ocean atmosphere models for
seasonal prediction of the Indian summer monsoon can only happen if significant
improvements are made in these coupled models to reduce the systematic biases.
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 295

11.3 MONSOON DISTURBANCES AND SEMIPERMANENT


MONSOON SYSTEMS OVER INDIA
The Indian summer monsoon, in addition to its planetary scale monsoon circulation linking
the summer and the winter hemisphere is also characterized by the existence of several
monsoonal disturbances and existence of some semipermanent meteorological monsoon
systems. The following sections briefly discuss the monsoonal disturbances and the
semipermanent meteorological monsoon systems seen over India during the Indian summer monsoon.

11.3.1 Monsoon Disturbances


The monsoon disturbances which form over India during the Indian summer monsoon include
the monsoon depression, the onset vortex, the mid-tropospheric cyclone, and the offshore
trough/vortex, respectively. All the above-mentioned monsoon disturbances over India during
the Indian summer monsoon provide copious rainfall over the regions they form and move to.
The following subsections discuss briefly the various monsoonal disturbances which form
over the Indian region during the Indian summer monsoon season.

Monsoon Depressions
Monsoon depressions are low-pressure systems that form in the vicinity of the monsoon
trough. Depressions are cyclonic systems, which are intermediate in intensity between the
weak low-pressure systems (wind speed less than 17 knots) and the tropical cyclones (wind
speed from 34 to 47 knots). The above term is most frequently used to describe the weak
cyclonic disturbances that form over the North Bay of Bengal and which generally move
west-north-westward over the Indian subcontinent. Further, the term monsoon depression is
also used to describe depressions, that form within the monsoon trough near Australia and in
the western North Pacific region. According to the India Meteorological Department,
depressions are cyclonic disturbances with an associated wind speed of 17 to 27 knots (1 knot
equals 0.514 m s–1). An intense form of depression called the Deep Depression has an
associated wind speed of 28 to 33 knots. The preferred region of formation of monsoon
depression over India is between 20° and 30°N and 80° and 90°E. Table 11.2 gives the
monthwise distribution of depressions over the Indian region for 100 years from 1891–1990.
The central parts of India receive about 80% to 90% of their annual rainfall during the
summer monsoon months of June to September. A major portion of the above rainfall over
Central India is associated with the passage of monsoon depressions which can provide
widespread and copious rainfall to these areas. The maximum numbers of monsoon
depressions which form during the monsoon months of June to September, each year are 3, 4,
4 and 3, respectively with the highest frequencies during July and August. While some of the
monsoon depressions owe their origin to weak easterly waves travelling from the east, others
have their genesis in situ over the North Bay of Bengal region. The passage of an easterly
wave intensifies the seasonal monsoon trough over India and often a closed low-pressure
system forms over the North Bay of Bengal. The above low-pressure system then intensifies
to a monsoon depression in about a day’s time. The monsoon depressions over North Bay of
Bengal have a horizontal extent of about 1000 km. Monsoon depressions are typically cold
core systems below 700 hPa level and warm core systems aloft. Hence, the strongest winds
296 u BASICS OF ATMOSPHERIC SCIENCE

TABLE 11.2 Monthwise distribution of monsoon


depressions over India (1891–1990)

Month Number of depressions


January 9
February 3
March 1
April 10
May 29
June 87
July 131
August 171
September 144
October 99
November 53
December 30
Total 767

associated with the monsoon depressions are seen in the neighbourhood of 700 hPa. The
cyclonic circulation associated with the monsoon depression becomes very weak at the upper
levels and is absent at and above 300 hPa. The location of the centre of the monsoon
depression slopes south-westwards with height. The wind fields associated with the monsoon
depression are asymmetric with stronger winds found south of the centre at lower levels. The
maximum horizontal convergence and associated spatial precipitation pattern are generally
found in the south-west sector of the monsoon depression. The primary zone of the heaviest
precipitation (occurring in the south-west sector) is about 200 km to 400 km away from the
centre, while a secondary zone of relatively less rainfall is located about 800 km west of the
monsoon depression centre. The area around the first hundred kilometres of the monsoon
depression centre is relatively free from intense precipitation.
The normal movement of monsoon depressions is in the west-to-west north-west
direction. During the months of June and September, the movements of the depressions are
somewhat spread out and they follow either the northerly direction or they recurve over the
Bay of Bengal. During the other two months of July and August, most of the depressions
move in the west-north-westerly direction and their tracks are confined to within a narrow belt
along Central India. During the month of July, the average speed of a depression is between
1.2 m s–1 and 2.4 m s–1, to the east of 85°E, while the average speed of the depression is
between 4.8 m s–1 and 9.6 m s–1, to the west of 85°E. The movement of the depression during
its initial formative stage over the Bay is rather slow and the depressions increase in speed as
they move over the land. The average life period of a monsoon depression, which has formed
over the Bay of Bengal, is about five days, while the average life of a depression, which has
formed over Arabian Sea and over land, is about three days.
The depressions and deep depressions, which form during the Indian summer monsoon
season, do not generally intensify into a tropical cyclone. This is because for the formation of
a tropical cyclone, there is a need for large-scale organization of cumulus convection in the
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 297

form of cumulonimbus clouds. For penetrative convection to happen, the cumulonimbus


clouds have to grow vertically to great heights. The vertical development of cumulonimbus
cloud is, however, inhibited by the existence of strong wind shear in the vertical direction
during the Indian summer monsoon season. The existence of low-level westerly winds
accompanied by strong upper-level easterlies during the Indian summer monsoon season is
responsible for the strong wind shear in the vertical. Over the north Bay of Bengal region, the
preferred region of formation of monsoon depression, the difference in the horizontal wind
between 850 hpa and 200 hPa during the monsoon months is greater than 15 m s–1, a value
large enough not to favour penetrative convection. Monsoon depressions generally weaken in
intensity after reaching the central parts of India due to lack of supply of moisture. These then
weaken to low-pressure systems and move in a west-north-west direction and merge with the
seasonal low over North-West India. At times, these depressions over the central parts of India
may experience a fresh lease of life due to the supply of fresh moisture from the Arabian Sea.
Some studies have investigated the role of monsoon disturbances (depressions and lows),
which form over the Bay of Bengal in initiating floods over the Godavari basin. The
Godavari, one of the largest of Indian peninsular rivers, spreads over an area of 312870 km2.
The above river originates in the Western Ghats and flows in a south-easterly direction across
five Indian states before joining the Bay of Bengal. A large part (50% to 65%) of the seasonal
rainfall in the northern and eastern sectors of the Godavari basin is attributed to the passage
of the monsoon disturbance. The passages of monsoon disturbances (depressions/lows) also
cause significant flooding related hazards. A study, which identified 22 flood situations from
the daily river discharge data during 1962–1990 over the eastern Godavari districts, found that
thirteen of them were associated with the presence of monsoon depressions, while the
remaining nine were associated with monsoon lows. Thus, the number of monsoon
depressions, which occur over India, their strength and their longevity, are the primary
contributors to the quantity of Indian summer monsoon rainfall and to the occurrence of
flooding throughout the central river basins of India.

Onset Vortex
The onset of the south-west summer monsoon is heralded over India due to the formation of
a major cyclonic disturbance or low-pressure system which forms in the Arabian Sea. Such a
cyclonic disturbance forming in the Arabian Sea is known as the onset vortex. The northward
movement of the onset vortex over the Arabian Sea is primarily responsible for the initial
progress of the monsoon circulation and associated precipitation to northern latitudes. The
onset vortex generally moves northwards or north-westwards. The onset vortex forms and
dissipates generally over the Arabian Sea and hence the above cyclonic disturbance has not
been extensively studied due to paucity of conventional upper air meteorological observation
over the sea. Further, detailed observational and modelling studies of the monsoon onset
vortex over the Arabian Sea were performed during the summer time of the Monsoon
Experiment (MONEX) in 1979.

Mid-tropospheric Cyclone
Mid-tropospheric cyclone (MTC) is the name assigned to an important cyclonic disturbance
which has its maximum intensity in the middle troposphere. Over India, the mid-tropospheric
298 u BASICS OF ATMOSPHERIC SCIENCE

cyclones occur mostly over northeast Arabian Sea and the state of Gujarat. The main cyclonic
circulation of the MTCs is to be found between 700 hPa and 500 hPa with the strongest
cyclonic circulation at about 600 hPa. The vertical temperature structure of the mid-
tropospheric cyclone reveals that the MTCs are cold core system at 700 hPa and warm cored
at 500 hPa. These MTCs were first identified and studied in 1968 by Miller and
Keshavamurty during the International Indian Ocean Expedition.

Offshore Trough/Vortex
The progress of the summer monsoon into Kerala is mostly associated with a weak trough
seen off the west coast of India, called as the off-shore trough. The offshore trough is also
seen subsequent to onset conditions, when it manifests as a trough over the Arabian Sea
running from north Kerala to south Gujarat coast. The offshore trough provides copious
rainfall over the west coast of India during the summer monsoon season. Furthermore, small
closed circulations, embedded within the off-shore trough can provide very heavy rainfall
along the western coasts of India. Such closed circulations in the offshore trough are called as
offshore vortices. At times, the onshore low-level westerlies from the Arabian Sea prefer to go
around the Western Ghats. The Western Ghats are an orographic barrier running parallel to
the west coast of India, and oriented in the north-south direction, about 1000 km long,
250 km broad and about 1 km in height. When the onshore westerlies go around the Western
Ghats, the return current causes an offshore vortex over Arabian Sea. Such offshore vortices
have a linear dimension of about 100 km long and can provide extensive rainfall over the
west coast and in the neighbourhood.

11.3.2 Semipermanent Monsoon Systems Over India


The semipermanent meteorological monsoon systems which form over India during the Indian
summer monsoon include the heat low, monsoon trough, tropical easterly jet, Tibetan
anticyclone, Mascarene high, low-level jet and southern hemispheric equatorial trough,
respectively. The following subsections discuss briefly the semipermanent meteorological
monsoon systems which form over India during the Indian summer monsoon.

Heat Low
Heat lows are observed during the summer months over the north-west parts of India and the
central parts of Pakistan. These lows are manifested in the surface pressure chart as a low
pressure. However, the above lows are indeed very shallow and can extend only up to a few
kilometres in the vertical. Above these surface lows are seen well-marked ridge extending up
to the upper troposphere. The heat low weakens during the night due to radiational cooling.
These heat lows are also regions where subsiding motion and diabatic warming takes place.
The regions where the heat lows develop are characterized by desert-type conditions and these
regions do receive a large amount of insolation. These regions do serve as a radiative sink,
due to the high albedo of the desert soil and the loss of the reflected radiation. These regions
are dominated by subsiding motions and diabatic warming; processes necessary to compensate
for the cooling due to the deficit of radiation. Such regions of heat lows are normally close to
the latitudes associated with the subtropical highs.
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 299

Monsoon Trough
A trough in the surface pressure chart is seen over India during the summer monsoon season,
oriented in the northwest-southeast direction and running from the heat low over Pakistan to
the southeast Bay of Bengal, and is called the monsoon trough. Over India and over land, the
monsoon trough runs from Ganga Nagar in Rajasthan to Kolkata through Allahabad. The
monsoon trough does have a vertical structure and normally extends up to 6 km above and is
sometimes seen up to 9 km also. The monsoon trough does slope southwards with increase
in height. The monsoon trough is fully established by the end of June. A few studies are
associating the monsoon trough with the continental tropical convergent zone which was
discussed in an earlier section of this chapter. Even though the monsoon trough is a quasi-
permanent system, its position does vary from day-to-day and these variations in the position
of the monsoon trough have important effects on the daily monsoon rainfall received over
India. When the axis of the monsoon trough is south of its normal position with its eastern
end dipping into the Bay of Bengal, active monsoon conditions are supposed to prevail over
India. During such an active phase, heavy rainfall over the plains of northern India, the
central parts of India as well as along the west coast are observed. Also, monsoon
depressions can form over North Bay of Bengal during such an active phase, especially when
the eastern end of the monsoon trough dips into the Bay. However, when the axis of the
monsoon trough moves to the north of its normal position and is observed near the foothills
of Himalayas, break monsoon conditions prevail over India. The heavy belts of monsoon
rainfall during the break conditions shifts to the foothills of Himalayas, while the rainfall
over the northern and central India is either absent or subdued. Moreover, Tamil Nadu
(a State in extreme southeast peninsular India), a rain shadow region during southwest
monsoon season may get copious rainfall during break monsoon conditions. Only for about
30 to 47% of the total time, is the monsoon trough seen in its normal position.

Tropical Easterly Jet


In an earlier chapter, the polar and the subtropical jet streams have been discussed. Unlike
the above jet streams, which are westerly, an easterly jet stream is seen over India around
13°N during the summer monsoon season at a height of 100 hPa. It has been mentioned
earlier, that an intense anticyclonic circulation dominates the Indian summer monsoon
circulation at upper levels over Tibet. A diverging part of the anticyclonic circulation over
Tibet in the upper atmosphere will get deflected by the Coriolis force to the right in the
northern hemisphere resulting in a tropical easterly jet stream. Just prior to the onset of the
Indian summer monsoon, the northern hemispheric subtropical jet steam moves further
northward. The easterly, flow south of the subtropical ridge over Asia, concentrates into a jet
stream with two branches. While the first branch is located around 15° to 18°N extending
from the northeastern parts of the South China Sea westwards to the Bay of Bengal and
southern India, the other branch is located between 5°N and 10°N over the southern parts of
South China Sea.

Tibetan Anticyclone
Tibetan plateau, having an average height of 4 km, with its complex topography and
containing several individual peaks, plays a very important role in the Asian winter as well as
300 u BASICS OF ATMOSPHERIC SCIENCE

in the Indian summer monsoon. In addition to acting as a mechanical barrier to air currents,
the Tibetan plateau also serves as a heat source (i.e. an effective recipient of heat for
atmospheric motions). The mechanical barrier effect of the Tibetan plateau is considered
important for the Asiatic winter monsoon circulation, while the role of the heat source played
by the Tibetan plateau has important consequences for the Indian summer monsoon
circulation. Since Tibetan plateau is an elevated land mass, it receives a large amount of solar
radiation, which is absorbed by the mountain surface. The air in the immediate neighbourhood
of the plateau is heated considerably, and this results in the establishment of strong horizontal
temperature gradients with the surrounding free atmosphere. Hence, the Tibetan plateau acts
as a sensible heat source at the middle tropospheric levels and is responsible for the formation
of a heat low over the mountain surface and a warm core anticyclone (high pressure) in the
upper troposphere. At 500 hPa, a ridge line is seen east of 80°E with its axis close to 28°N.
The above high pressure at 500 hPa has its centre at 28°N, 98°E on the eastern periphery of
the Tibetan plateau and covers the entire plateau. At 300 hPa, the high pressure is seen
between 70°E and 110°E with its centre at 30°N, 90°E. At 200 hPa, the high pressure extends
from 78°E to 140°E and is centred at 30°N 88°E. At 100 hPa, over the Indian region a broad
belt of high pressure is seen at 35°N extending from 30°E to 150°E. The high pressure over
Tibet from 500 hPa to the upper troposphere is usually referred as the Tibetan anticyclone.
The strength of the Tibetan anticyclone provides a measure of the strength of the large-scale
monsoonal circulation, indicating that a weak monsoon is associated with a less intense
Tibetan anticyclone.

Mascarene High, Low-level Jet and Southern Hemispheric Equatorial


Trough
The earlier mentioned semipermanent systems associated with the Indian summer monsoon
were seen over India and its immediate neighbourhood. However, the Indian summer
monsoon is also associated with other semipermanent systems which occur far away from the
Indian subcontinent. A few of the semipermanent systems occur in the southern hemisphere
and have important influence on the variability of monsoon rainfall over India. Among these
are the intense high pressure over the Madagascar area, in southern hemisphere called the
Mascarene high, from where the low-level monsoon circulation originates in the southern
hemisphere. Also, there exists a strong low-level cross equatorial jet over the Somalia coast
of Africa called the Somali jet and an east-west trough on the surface near south of equator
called the southern hemispheric equatorial trough (SHET). The Somali jet is a narrow current
of air at a height between 1 km and 1.5 km, off the coast of east Africa. The above jet flows
from the northern parts of Madagascar and reaches the northern Arabian Sea and India in the
month of June. The strongest cross-equatorial flow during the Indian summer monsoon occurs
in the immediate vicinity of the low-level jet. The SHET is also referred as the oceanic
convergent zone which is discussed in an earlier section of this chapter. The intensity of
SHET determines greatly the intraseasonal variation of the Indian monsoon rainfall.

11.4 TROPICAL CYCLONES


A tropical cyclone is usually used in a generic sense for an intense vortex or a whirl in the
atmosphere (synoptic scale low-pressure system) over tropical (or sub-tropical) regions having
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 301

a cyclonic circulation at the lower levels and possessing organized convection. The word
“Cyclone” is derived from Greek word Cyclos meaning “the coils of a snake”. A cyclonic
system with associated wind speed of 34 knots (17 m s–1) or above is denoted as a tropical
cyclone, hurricane (in North Atlantic Ocean, North-East Pacific Ocean and in South Pacific
Ocean, east of 160oE), and typhoon in North-West Pacific Ocean. The more intense tropical
cyclones having associated wind speed of 48 to 63 knots are known as severe cyclonic storms
and those with wind speeds greater than 64 knots are known as very severe cyclonic storms
and those with more than 119 knots as super cyclones in India.

11.4.1 Factors Responsible for the Formation of Tropical


Cyclone
Vast quantities of heat energy are necessary to fuel the heat engine of the tropical cyclone.
The primary source of this energy is the release of latent heat supplied by the evaporation
from the ocean surface. Since the presence of warm surface waters only ensures large
evaporation rates, tropical cyclones form only over warm ocean waters having a sea surface
temperature (SST) greater than or equal to 27°C. Due to the above requirement, tropical
cyclones cannot form pole ward of 20° latitude as the SST are usually lower over these
regions. Further, tropical cyclones usually form in late summer or in early fall when the sea
surface temperatures are warmest. Tropical cyclones also cannot form within 5° latitude north
and south of equator since the Coriolis force is too small there to maintain the low-pressure
system. Another requirement for the formation of a tropical cyclone is an atmosphere, which
is potentially unstable to moist convection. In other words, unstable conditions must exist
throughout the troposphere. Further, the mid-tropospheric levels (~ 500 hPa) must be
sufficiently moist for the formation of a tropical cyclone. Tropical cyclone also ideally forms
in an environment with low-vertical wind shear values (less than 10 m s–1) between the
surface and the upper troposphere. Presence of large vertical wind shear affects the vertical
transport of latent heat and disrupts the organization of deep convection around the cyclone
centre. This is the chief reason, why no tropical cyclone forms during the Indian summer
monsoon season (June to September), over the Indian seas since the monsoonal winds are
characterized by strong vertical wind shears with westerly winds at lower levels and strong
easterlies at upper levels. Finally, the formation of a tropical cyclone requires a pre-existing
system with sufficient vorticity and low-level inflow. The tropical cyclones cannot form in situ
without any initial incipient disturbance. Hence, for the development of a tropical cyclone,
one requires a weakly organized system with moderate vorticity, convergence at low level and
divergence at upper level. Over the Bay of Bengal region, the remnants of a typhoon from
North-West Pacific, which has weakened may provide the initial disturbance for the
regeneration of a tropical cyclone. The Intertropical convergence zone can also provide the
low-level convergence and the upper-level divergence can also provide the initial environment
for the formation of a tropical cyclone. Further, an important source for the formation of a
tropical cyclone is the easterly wave. Easterly waves, schematically shown in Figure 11.7, are
wave-like disturbances which are embedded in the trade winds and as the name suggest travel
from east to west. They have a horizontal extent of about 2000 km and have associated a
region of low-level convergence and cloudiness on the trailing side of the wave (east side of
302 u BASICS OF ATMOSPHERIC SCIENCE

Convergence

Divergence

FIGURE 11.7 Schematic diagram illustrating an easterly wave.

the ridge). The above can be easily understood from the following discussion. Air over the
trough axis of the easterly wave has negative relative vorticity, while the same over the ridge
axis corresponds to positive relative vorticity. Also, the air gains planetary vorticity by moving
from the trough axis to the ridge axis. Hence, in effect the air gains in absolute vorticity as it
moves from the trough axis to the ridge axis. From the vorticity equation and by retaining
only the divergence term, it is clear that the gain of increasing absolute vorticity would
correspond to convergence and hence the east side of the ridge has associated with its
convergence and cloudiness. The tropical cyclone can then easily form on the trailing side of
the easterly wave (i.e. on the east side of the ridge), where conditions of low-level
convergence are present.

11.4.2 Climatology of Tropical Cyclones


On an average, a total of about 84 tropical cyclones form in over the entire globe in a year.
About 26 of these cyclones in a year form over the western North Pacific, while about 17
form over the eastern North Pacific. While North Atlantic and South-West Indian Oceans have
each an average of about 10 cyclones in a year, the South-West Pacific Ocean has about 9
cyclones forming in a year. Also, the North Indian Ocean (which includes the Indian seas of
Bay of Bengal and Arabian Sea) and the South-East Indian Ocean have each an average of
about 6 cyclones in a year. Out of the 6 cyclones over the North Indian Ocean, the Bay of
Bengal accounts for about 4 cyclones, while the Arabian Sea accounts for the remaining 2
cyclones. Nearly 45 of these cyclones (about half of the total) are quite intense with an
associated wind speed of 64 knots or greater. However, no tropical cyclone forms over South
Atlantic and the eastern South Pacific.
In North Indian Oceans (Bay of Bengal and Arabian Sea), tropical cyclones form during
the pre-monsoon (April and May) and the post-monsoon (October to December) seasons.
Presence of large vertical wind shear during the Indian summer monsoon season (June to
September) inhibits systems to intensify to tropical cyclone intensity. Over Atlantic (i.e. North
Atlantic) a pronounced tropical cyclone maxima is seen in the months of August through
October, precisely early to mid-September, although the tropical cyclone season is from June
to November. There is a near total absence of cyclones during the months of December to
April. While, the Eastern North Pacific hurricane season runs from May 15 to November 30,
the same has a broader peak with activity beginning in late May or early June and going until
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 303

late October or early November with a peak in tropical cyclone number in late August/early
September. The main season for North-West Pacific tropical cyclones runs from July to
November with a peak in late August/early September. However, the North-West Pacific basin
has tropical cyclones occurring all year round regularly though there is a distinct minimum in
February and the first half of March. Table 11.3 presents the monthwise distribution of
cyclonic storms and severe cyclonic storms over the Indian seas for a 100-year period from
1891–1990.
TABLE 11.3 Monthwise distribution of cyclonic storms and severe cyclonic storms
over the Indian seas for a 100-year period, from 1891–1990.

Month Number of Number of severe Total


cyclonic storms cyclonic storms
January 5 2 7
February – 1 1
March 2 2 4
April 13 14 27
May 20 50 70
June 40 17 57
July 37 8 45
August 30 3 33
September 30 17 47
October 54 45 99
November 46 68 114
December 26 21 47
Total 303 248 551

The South-West Indian and Australian/South-East Indian basins have very similar
tropical cyclone annual cycles beginning in late October/early November, and reaching a
double peak in activity—one in mid-January and one in mid-February to early March, and
then ending in May. The Australian/South-West Pacific basin season for tropical cyclones
begins in late October/early November, reaches a single peak in late February/early March,
and then fades out in early May. Studies have indicated that there has been no increase in
tropical cyclone frequency globally over at least the past several decades.

11.4.3 Movement of Tropical Cyclones


Tropical cyclones have a tendency to move westward and poleward in both the hemispheres
(refer Figure 11.8). Figure 11.8 shows the general movement of tropical cyclones, and their
source regions, over the various oceans. However, once they are over the subtropics, tropical
cyclones re-curve towards north-east/south-east in northern/southern hemispheres. Although
the tropical cyclones form over the tropical regions, the low-level trade winds have very little
role in influencing their movement. Instead, the upper-level winds and the spatial distribution
of the sea surface temperature to some extent determine the speed and direction of movement
304 u BASICS OF ATMOSPHERIC SCIENCE

FIGURE 11.8 General direction of movement of tropical cyclones, and their source regions, over
the various oceans.

of the tropical cyclone. While, the larger and more intense tropical cyclones interact and do
influence the surrounding environment, for the less intense storms, the surrounding
environmental flow particularly in the middle and upper troposphere does determine the storm
movement through the concept of “steering”. Further, the track of the tropical cyclone varies
considerably in response to the weather patterns occurring at the time. The movement of a
tropical cyclone also affects the speed of the winds that circulate about the storm centre. On
one side of the storm, where the circulating winds and the tropical cyclone are moving in the
same direction, the forward movement of the cyclone increases the wind speed. On the
opposite side of the storm, the forward motion decreases the circulating wind speed. In the
northern hemisphere, the right side of a tropical cyclone, looking in the direction in which it
is moving, has the higher wind speeds since the circulating air is rotating in the anticlockwise
direction. Using similar argument, the left side of a tropical cyclone in the southern
hemisphere will have the higher wind speeds.
The westward movement can be explained using the advection of planetary vorticity.
Consider a cyclonic vortex in northern hemisphere with air rotating in counterclockwise
direction. Consider the linear nondivergent barotropic vorticity equation in an equatorial b plane
˜[
 WC (11.1)
˜U
where z is the vertical component of relative vorticity, b is the latitudinal variation of
Coriolis parameter, and v is the meridional component of velocity. It is to be noted that b is
positive in both the hemispheres. Since the meridional component is negative/positive to the
west/east of the centre of the vortex, in northern hemisphere from Eq. (11.1), it is clear that
z will increase/decrease to the west/east of the vortex. In eff ect, the advection of planetary
vorticity will cause an increase in z to the west of the vortex centre with time, effectively
causing the vortex to diffuse to the west. A similar argument for the cyclonic vortex in
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 305

southern hemisphere due to the advection of planetary vorticity again results in a diffusion of
the vortex westward.
The poleward movement of the cyclonic vortex can be explained based on the
nondivergent barotropic vorticity equation on an equatorial b plane by considering a cyclonic
vortex in northern hemisphere with air rotating in counterclockwise direction. The
nondivergent barotropic vorticity equation on an equatorial b plane is given by
E [  G

 (11.2)
EU

To the east of the vortex centre, the counterclockwise rotation results in an increase in the
Coriolis parameter f and hence from Eq. (11.2), the relative vorticity z decreases. This sets up
a secondary anticyclonic circulation (clockwise) over the eastern part of the vortex. To the
west of the vortex centre, the counterclockwise rotation results in a decrease in the Coriolis
parameter f and hence from Eq. (11.2), the relative vorticity z increases. This sets up a
secondary cyclonic circulation (counterclockwise) over the western part of the vortex. The
above combination of cyclonic/anticyclonic secondary circulations to the west/east of the
vortex centre will ensure that the vortex centre gets advected northward. A similar argument
for the cyclonic vortex in southern hemisphere due to the conservation of absolute vorticity
again results in the vortex centre being advected southward. The above-mentioned poleward
movement of a cyclonic vortex is attributed to the conservation of absolute vorticity and is
known as the b-effect.
The likelihood of tropical cyclone having typical tracks over North Atlantic Ocean
depends on the month in which the cyclone forms. During the month of August, the most
likely track of a tropical cyclone over North Atlantic will pass over the West Indies islands.
From here, the tropical cyclones are equally likely to move either towards the Texas coast or
along the Atlantic coasts from Florida to North Carolina. There are two distinct cyclone tracks
during the month of September in North Atlantic. While one cyclone track reaches the central
Gulf of Mexico by moving northward between the Yucatan peninsula and western Cuba, the
other track moves northward from around Haiti, and Puerto Rico into the western Atlantic.
During the month of October, the typical cyclonic tracks start from eastern Mexico and move
northward to Florida and to the rest of the south-eastern United States.
Tropical cyclones off the north-west coast of Australia do exhibit a typical track. Mostly
these cyclones move to the west-south-west direction and as they move further to the south,
they take a more southerly direction of movement. Once these tropical cyclones reach south
of 22°S, they tend to recurve to the south-south-east direction. Also, the tropical cyclones of
the South-West Pacific (off the Queensland coast) are known to exhibit widely erratic tracks,
which make it extremely difficult for forecasters involved in real-time tropical cyclone
prediction. Typically the tropical cyclones, which form in North-West Pacific, move in a west-
north-westward movement. Once they are over the subtropical regions, they recurve into the
mid-latitudes of the North Pacific well offshore of Japan.
Isolated tropical cyclones can form in March over South Bay of Bengal and generally
move west-north-westwards and hit Tamil Nadu and Sri Lanka coasts. During the pre-
monsoon months of April and May, tropical cyclones form in the South and adjoining Central
Bay of Bengal and move initially north-west, north and then recurve to the north-east striking
306 u BASICS OF ATMOSPHERIC SCIENCE

the Arakan coasts in April and the Andhra–Orissa–West Bengal–Bangladesh coasts in May.
During the post-monsoon months of October to December, tropical cyclones form mostly in
the South and the Central Bay of Bengal, recurve between 15° and 18°N affecting Tamil
Nadu–Andhra–Orissa–West Bengal–Bangladesh coasts. Over Arabian Sea, the tropical
cyclones generally form in South-East Arabian Sea and adjoining Central Arabian Sea in the
months of May, October, November and December and in East Central Arabian Sea in the
month of June. Further, some of the tropical cyclones, which originate in the Bay of Bengal
move across the peninsula, weaken and enter Arabian Sea as low-pressure areas. These weak
low-pressure systems may again intensify into tropical cyclones. Most of the tropical cyclones
in Arabian Sea move in west-north-westerly direction towards the Arabian Coast in the month
of May and in a northerly direction towards Gujarat Coast in the month of June. During other
months, the tropical cyclones generally move north-west-north and then recurve north-east
affecting Gujarat–Maharashtra coasts. However, a few of the tropical cyclones can move
west-north-westwards towards Arabian coast.

11.4.4 Life Cycle of a Tropical Cyclone


Tropical cyclones have an average longevity of about 4–7 days. While some weak tropical
cyclones can only reach severe cyclonic stage with hurricane winds, very briefly, there are
other storms, which can remain in the severe cyclonic stage for a period of week or more. For
tropical cyclones that have reached the severe cyclonic stage, (equal and greater than 48 m s–1),
the life cycle can be divided broadly into four different stages.
(i) Formative stage: This stage is characterized by the appearance of the system as
an active and disorganized area of convection in satellite images. Further, in this
stage, the circulation centre is not well defined. However, the spiraling cumulus
cloud bands as manifested in the satellite image do indicate the location of the
cyclone centre. In this stage, the strongest surface winds are confined to one
quadrant only and appear far away from the cyclone centre. Further, in this stage,
the maximum wind speed is usually less than gale force (14 m s–1 to 28 m s–1). The
tropical cyclones, which suffer landfall, during this formative stage, produce little
damage to life and property except for some heavy rain on the coastal regions.
(ii) Immature stage: During this stage, the convection becomes more organized,
while the system intensifies rapidly. The lowest sea level pressure associated with
the system drops rapidly to values below 1000 hPa, while the organized convection
manifests as long bands spiraling inwards. In this stage, winds that are gale force
(14 to 28 m s–1) are observed within the system. The maximum winds are found
concentrated in a tight band close to the cyclone centre. In this stage, the cyclone
centre is well defined and the formation of the eye is just visible. Figure 11.9 shows
a satellite image of the Orissa Supercyclone on October 29, 1999 04 UTC. The eye
of the tropical cyclone can be seen in the above satellite image. Further, in this
stage, satellite images reveal the existence of well-organized curved bands of active
convection, spiraling in towards a central dense mass of clouds covering the centre
of the cyclone. During this stage, the minimum sea level pressure and the
maximum wind speed associated with a tropical cyclone are usually observed. The
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 307

Figure 11.9 Satellite image of the Orissa Supercyclone on October 29, 1999
04 UTC. The eye of the tropical cyclone can be seen in the above
satellite image.

tropical cyclones, which suffer landfall, during this immature stage, can produce
very devastating wind and storm surge effects, although the areal extent of the
damage is usually small.
(iii) Mature stage: During this stage, the process of intensification ceases and the
tropical cyclone becomes a quasi-steady state system. Overlying the above-
mentioned quasi-steady state is manifested random fluctuations in the central
minimum pressure and maximum wind speed. However, the horizontal extent of the
tropical cyclone increases rapidly during this stage. That is, the horizontal extent of the
cyclonic circulation as well as the area of the gale type winds (14 m s–1 to 28 m s–1)
increase markedly. In this stage, the satellite images reveal the existence of a highly
organized cloud field which is more or less symmetrical. A distinct and circular eye
is seen during this stage for most severe cyclones. Since this stage corresponds to
the maximum intensity stage in terms of the extent of the cyclonic circulations, most
cyclones last for a day or so in this mature stage before they weaken.
(iv) Dissipative stage: This is the final stage in the life cyclone of a cyclone and is
characterized by the dissipation of the system. During this stage, the warm core
(seen prominently at upper levels) vanishes. Further, the central minimum pressure
starts rising and the region of maximum winds moves away from near the cyclone
centre. In this stage, the satellite images reveal the marked weakening of the
organized convection near the cyclone centre together with the absence of the
major curved convection bands. However, the narrow bands of low clouds, as seen
in satellite images, may still manifest in the existence of the low-level circulation
centre. The tropical cyclones, which suffer landfall, during this dissipative stage can
cause heavy rains at coasts and even at large distances inland. Tropical cyclones,
308 u BASICS OF ATMOSPHERIC SCIENCE

over seas, can reach the dissipative stage as they move over regions of cooler sea
surface temperature or when they encounter strong vertical wind shear regions.

11.4.5 Tropical Cyclone Structure


Tropical cyclones contain a very large number of thunderstorm cells arranged in the wall
cloud (or eyewall) region, outside the eye, consisting of thick bands of clouds and associated
heavy precipitation, spiraling cyclonically (counterclockwise) in the northern hemisphere
around the cyclone centre. Areas of weaker uplift and less-intense precipitation or even air,
which is descending, separate the bands of thick convective clouds. Approaching the cyclone
from outside, the magnitude of wind speed as well as the intensity and amount of
precipitation, increase towards the centre of the tropical cyclone with the maximum winds
and rainfall occurring at the eyewall region, a distance of 10 to 20 km away from the eye
(centre) of the cyclone. Figures 11.10 shows the schematic diagram of a typical cross-section
of rainfall intensities (refer Figure 11.10(a)), atmospheric pressure (refer Figure 11.10(b)),
and wind speed (refer Figure 11.10(c)), through a tropical cyclone. The maximum intensity
of rainfall (Figure 11.10(a)), is seen over the eyewall region, while there are no rains at the
centre of the eye of the cyclone. The intensity of rainfall decreases from the eyewall towards
the outer portions of the cyclone with the descending portions of air reporting minimum
intensity of rainfall. The surface pressure gradient (Figure 11.10(b)), and the surface wind
speed (Figure 11.10(c)), both increase gradually when approaching the centre of the cyclone
from outside. However, both the pressure gradient and the wind speed show a rapid increase
near the eyewall region.

FIGURE 11.10 Schematic diagram of a typical cross-section of rainfall intensities (a) (top panel),
(b) pressure (middle panel), and (c) wind speed (bottom panel) through a tropical
cyclone. The horizontal scale extends to about 700 km (i.e. 350 km from the centre
of cyclone).
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 309

Tropical cyclones are typically warm core systems, i.e. the temperature difference
between a cyclone and the environmental air shows positive values. Close to the surface,
however, the temperature increases only very slightly towards the cyclone centre. With
increase in height, the temperature differences become more and more positive until they
attain a value of 10 to 11oC in the upper troposphere. Figure 11.11 shows the magnitude of
the temperature differences across a tropical cyclone with respect to the environmental air. As
the low-level air flows towards the centre of the cyclone close to the surface, the warm ocean
surface supplies large amounts of both sensible and latent heat to the surface air. The above
heat input compensates for any cooling experienced by the low-level air as it moves towards

12

11

10
10

2
Height (km)

6
10
1
5 9

4 8

7
3
6
2 2 5

4
1
–2 3
2
1
68°E 69°E 70°E 71°E 72°E
Centre
FIGURE 11.11 Temperature differences across a tropical cyclone with respect to the environ-
mental air.
310 u BASICS OF ATMOSPHERIC SCIENCE

lower pressures and this ensures that the temperature differences are very small close to the
surface between the cyclone and the environment. However, at upper levels due to the release
of latent heat of condensation, some thermal energy gets added, which manifests as positive
temperature differences, giving rise to a warm core system. Due to the system being a warm
core low, the pressure inside the tropical cyclone decreases somewhat slowly with height.
Hence, the horizontal pressure gradient within the tropical cyclone decreases slowly with
height. At 400 hPa, the pressure within the cyclone is nearly the same outside it. However, at
heights above 400 hPa, i.e. in the upper troposphere, there exists upper-level divergence and
hence the air over these regions rotates anticyclonically, i.e. rotates clockwise in the northern
hemisphere.
Figure 11.12 shows a schematic diagram of the air trajectories near the eyewall at both
the lower and upper levels of the tropical cyclone. It is clear from Figure 11.12 that the
surface air parcels spiral inward towards the eyewall, rotate rapidly around the eyewall as they
rise to higher levels. At the upper troposphere, these ascending air parcels near the eyewall
spiral outward anticyclonically.

Figure 11.12 Schematic diagram of the air trajectories near the eyewall of the tropical cyclone at
the lower and upper levels.

Figure 11.13 presents the schematic diagram of the vertical cross-section of the
circulation associated with a tropical cyclone, low-level inflow, ascend over the eyewall region
and outflow at the upper troposphere and the subsiding motion far from the eyewall region.
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 311

Tropopause

Eyewall

FIGURE 11.13 Schematic diagram illustrating the vertical cross-section of the circulation
associated with a tropical cyclone.

11.4.6 Eye and the Eyewall


The eye is one of the most distinctive features of a tropical cyclone. The eye is a roughly
circular area of comparatively light winds and fair weather with relatively clear skies found at
the centre of a severe tropical cyclone. Also, the eye is a region of subsidence where the air
descends down, however, the subsidence does not extend till the surface, lest it may fill up
the system. The average diameter of an eye in a tropical cyclone is about 30 km, while the
eye diameter can vary from 20 km to 50 km. The change in the size of the eye does provide
some indication of the intensification/weakening of the eye. Generally, a shrinking eye is
considered to indicate an intensifying cyclone. The eye is the region of the lowest surface
pressure and warmest temperatures aloft.
The eye is surrounded by the eyewall, a roughly circular area of deep convection, which
is the area of highest surface winds, thickest cloud cover and the most intense precipitation in
the tropical cyclone. While the eye is composed of air that is sinking slowly, the eyewall has
a net upward flow as a result of many strong (and occasionally moderate) updrafts and
downdrafts. The eye’s warm temperatures are due to adiabatic compression of the subsiding
air. Most observations from radiosondes taken within the eye show a low-level layer, which is
relatively moist, with an inversion above; clearly indicating that the subsidence in the eye
does not reach the surface, and reaches to around 1–3 km of the surface.
The physical mechanisms responsible for the formation of the eye and eyewall are not
yet completely understood. A few of the mechanisms suggested for the subsidence in the eye
is presented. It has been assumed that supergradient winds, i.e. rotating winds that are
stronger than the gradient winds obtained from the local pressure gradient, prevail near the
eyewall of the tropical cyclone. It has been suggested that the above-mentioned supergradient
wind flow causes the air to be centrifuged out of the eye into the eyewall, giving rise to the
subsiding air in the eye. However, observations indicate that the rotating winds within several
tropical storms and hurricanes are more or less in gradient wind balance (within 1–4% of
gradient balance). However, it may be possible that the amount of supergradient flow needed
312 u BASICS OF ATMOSPHERIC SCIENCE

to cause such centrifuging of air is only on the order of a few per cent and hence may be
difficult to measure.
Another important characteristic of tropical cyclones that probably plays an important
role in the formation and maintenance of the eye is the eyewall convection. It is well known
that convection in tropical cyclones is organized into long, narrow rainbands, which are
oriented in the same direction as the horizontal wind. Because these bands seem to spiral into
the centre of a tropical cyclone at low levels, they are sometimes called spiral bands. Along
these bands, low-level convergence is a maximum and, hence, upper-level divergence is also
well pronounced above. A direct circulation can develop over these bands wherein the warm,
moist air converges at the surface ascends through these bands, diverges aloft, and descends
on both sides of the bands. Subsidence is distributed over a wide area on the outside of the
rainband but is concentrated in the small inside area. As the air subsides, adiabatic warming
takes place, and the air becomes drier. Since subsidence is concentrated on the inside of the
band, the adiabatic warming is stronger inward from the band causing a sharp contrast in
pressure falls across the band since warm air is lighter than cold air. Because of the pressure
falls on the inside, the tangential winds around the tropical cyclone increase due to increased
pressure gradient. Eventually, the band moves toward the centre of the cyclone and encircles
it leading to the formation of the eye and the eyewall. Thus, from the above two hypotheses,
the cloud-free eye may be due to a combination of both dynamically forced centrifuging of
mass out of the eye into the eyewall and to a forced descent caused by the moist convection
of the eyewall.

11.5 THUNDERSTORMS AND TORNADOES


Tornadoes are among nature’s violent manifestations of severe weather which can cause loss
of human lives as well as extensive loss of property. Tornadoes are among the most difficult
of the severe weather systems which can be accurately forecasted. Thunderstorms provide
very intense and showery type of rainfall and are usually accompanied by lightning, thunder,
hail, etc. The following subsections briefly discuss the general characteristics of thunderstorms
and tornadoes.

11.5.1 Thunderstorms
In addition to maintaining the earth’s fair-weather electric field, thunderstorms also play an
important role in transporting heat and moisture from the lower levels to the upper levels of
the atmosphere. Strong winds and intense precipitation including hail as well as lightning and
thunder usually accompany thunderstorms. Severe thunderstorms occur in large numbers and
in their most violent form over India during the pre-monsoon months of April and May as
well as during the month of June. The above months are characterized by hot and humid
conditions and hence the above characteristics are conducive for the formation of
thunderstorms. The most intense of the pre-monsoon thunderstorms as well as the highest
number of thunderstorms occur over the eastern parts of India comprising the states of
West Bengal and Orissa and the north-eastern parts of India.
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 313

For a thunderstorm to form, the following conditions need to be satisfied. Firstly, there
must be plenty of moisture available in the lower to middle levels of the troposphere together
with large amount of insolation. Furthermore, the atmospheric conditions must be conducive
for the establishment of instability. In addition, it is necessary to have a mechanism for
initiating an updraft in a moist unstable atmospheric environment. The above mechanism
for lift can either be mechanical or dynamical. The source of lift is mechanical when the
moist air is forced to flow up, say, over a mountain. In addition to the above, advancing
cold fronts, outflow boundaries, dry lines and sea breeze fronts may act as triggers to lift
the moist low level air. Dynamical effects become important for lifting, with the existence
of an upper-level divergence region, which can provide for low-level convergence and
dynamical lifting. The existence of upper-level jet streams can initiate lifting from low levels
through dynamical uplift.
The nature of the atmospheric instability necessary for the formation of thunderstorm
has been discussed in Chapter 3 and is known as the potential instability. Potential instability
occurs when a layer of dry air overlies warm moist air. The above instability is caused by
advection of the mid-level dry air over the top of a warm moist air. When a mechanism
(either dynamical or mechanical) exists for the uplift of the entire layer of air, the warm moist
air at low levels, which is nearly saturated, when lifted cools at the moist adiabatic lapse rate,
say 5°C km–1. However, the dry air at the mid-levels will be unsaturated and when lifted will
cool at the dry adiabatic lapse rate of 9.8°C km–1. That is, the layer of dry air at mid-level
will cool much more rapidly than the warm moist air at low levels, thus, destabilizing the
entire air column. Hence, a situation can arise when the air, which is initially statically stable,
can (potentially) become statically unstable when given sufficient lift.
Associated with the passage of a thunderstorm, rapid and marked changes in surface
pressure and also changes in air temperature are generally seen. As the thunderstorm
approaches a station, an abrupt fall in the surface pressure is noticed at the station. While the
thunderstorm is directly overhead the station with the station experiencing rain showers, there
is an abrupt rise in the surface pressure. However, as the thunderstorm moves over from the
station and the rainfall ceases, the surface pressure returns to its normal value. If the rainfall
associated with the thunderstorm reaches the surface, the air temperatures will start to
decrease due to evaporative cooling. If the rainfall lasts long enough, the air temperature may
cool to its dew point temperature.

11.5.2 Life Cycle of Thunderstorms


All thunderstorm cells undergo a sequence of three distinct stages of their life cycle and these
are referred as the cumulus, mature and the dissipative stages. Figure 11.14 outlines the three
stages of the life cycle of a thunderstorm cell. The first stage (refer Figure 11.14(a)), known
as the cumulus stage begins when the first updraft is initiated. The above updraft, i.e. the rise
of the unstable air is attributed to localized convection, which can arise when some surfaces
undergo more heating than others. The rising unstable air cools adiabatically and forms
cumulus clouds. The initial cumulus clouds, so formed may last for a few minutes before
quickly evaporating. Although these cumulus clouds may not directly initiate precipitation,
they play an important role in the development of thunderstorm by transporting the water
314 u BASICS OF ATMOSPHERIC SCIENCE

vapour from the levels close to the surface to middle troposphere. Hence, ultimately a
situation is reached where the atmosphere becomes humid enough, that newly formed clouds
undergo marked vertical development without evaporating. The above vertical growth
represents the cumulus stage in the development of a thunderstorm cell. The vertically
developing clouds in the cumulus stage can grow upward at the rate of 5 m s–1 to 20 m s–1.
Within the vertically growing clouds, the temperature decreases with height at a rate close to
the saturated adiabatic lapse rate. Also, the top of the vertically growing cumulus clouds is
above the freezing level. During the cumulus stage, there is growth by collision and
coalescence process leading to the formation of bigger drops. When the precipitation starts
falling down, it signals the end of the cumulus stage and the beginning of the mature stage.
The cumulus stage typically lasts for up to ten minutes and is not associated with rainfall,
severe weather, lightning or thunder.
The beginning of the mature stage (refer Figure 11.14(b)), is characterized by the falling
precipitation, which initiates a downdraft. During the mature stage, both the ice particles and
the supercooled droplets are present within the growing cumulus cloud. This allows for the
growth of ice crystals at the expense of supercooled droplets through the Bergeron process,
leading to formation of precipitation. The cooling of air as the falling precipitation evaporates
also strengthens the downdrafts. The mature stage is characterized by the co-existence of both
updrafts and downdrafts within the thunderstorm cell. The formation of intense precipitation
products such as hailstone as well as lightning and thunder occur during the mature stage.
The mature stage also determines the maximum heights to which the top of the vertically
developing convective clouds can grow. Also, the familiar anvil shape is seen during this
stage. The anvil shape is due to the effect of strong winds at the top of the clouds pushing the
ice crystals forward. During the first two stages of the development of the thunderstorm cell,
an abrupt transition is seen between the edge of the cloud and the surrounding unsaturated air.
During the mature stage, while updrafts are seen in the interior of the cloud, the downdrafts
are observed just outside the cloud. The above distribution of updrafts and downdrafts create
conditions where entrainment of unsaturated air along and into the cloud margins is possible.
The entrainment of dry unsaturated air into the cloud margin will cause the droplets to
evaporate and consequently cool the margins of the cloud. This results in the outer parts of
the cloud becoming denser and consequently less buoyant inhibiting further updraft of air.
The mature stage typically lasts from ten to twenty minutes.
The final stage of the development of the thunderstorm cell, called the dissipative stage
(refer Figure 11.14(c)), is characterized by the occurrence of the downdrafts over the entire
base of the cloud. The downdrafts increase in their horizontal extent and ultimately occupy
the entire cloud base, since more and more of the heavy precipitation associated with the
clouds begins to fall. During the dissipative stage, the amount of precipitation starts
diminishing, while the sky begins to clear up as the remaining droplets get evaporated. Even,
when a thunderstorm cell is in its dissipative stage, its downdrafts while hitting the ground
can spread out creating a wedge of cold and dense air called the gust front. However, the gust
front can also be seen in the mature stage as well. The gust front associated with the
dissipative stage of a thunderstorm cell may trigger new thunderstorm cells as it lifts warm
moist unstable air.
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 315

Mature stage
Dissipating stage

Updraft
Developing stage

Weak downdraft

(a) (b) (c)

Figure 11.14 Various stages of the life cycle of a thunderstorm cell. The left panel (a) shows the
cumulus stage, while the middle (b) and the right panels (c) illustrate the mature
and the dissipating stages.

11.5.3 Severe Thunderstorms and Squall Lines


Severe thunderstorms are by definition, very intense thunderstorms having wind speeds that
exceed 23 m s–1 and which can produce hailstones larger than 1.9 cm in diameter. The
associated updrafts and downdrafts reinforce one another in a severe thunderstorm and this
reinforcement requires suitable conducive conditions over an extensive region, known as
mesoscale of horizontal dimensions from 10 km to 1000 km. Hence, for the development of
severe thunderstorms, one requires a favourable pattern of a mesoscale atmospheric system.
Since the severe thunderstorms manifest over a large area, they typically appear in groups
with individual storms separating the groups. Such a group of thunderstorms is called the
mesoscale convective system (MCS). When the MCSs occur as line bands they are called
squall lines and when they occur as roughly oval or circular clusters, they are called
mesoscale convective complexes (MCCs). All the individual storm cells of an MCS form part
of a single system. The individual cells within an MCS either have all developed from a
common cause or have developed due to the existence of other cells within the same MCS.
The MCS can cause intense weather conditions over large areas and can last up to 12 hours.
Squall line thunderstorms consist of a large number of individual thunderstorm cells
organized in a linear band and having horizontal dimensions of about 500 km. A strong
vertical wind shear is an important component of the squall line thunderstorm. Usually, the
wind velocities in the direction of the thunderstorm movement increase with height. The
presence of strong winds at the upper levels push the updrafts ahead of the downdrafts and
allow the rising air to feed additional moisture into the storm; thus increasing the severity of
the squall line thunderstorm. Strong downdrafts associated with intense squall line
thunderstorms can create the so-called downbursts. These downbursts are very intense gusts
of winds that can reach speeds in excess of 65 m s–1. Downbursts with diameters less than
4 km are called microbursts, and these are considered very hazardous for aviation when they
occur near airports.
316 u BASICS OF ATMOSPHERIC SCIENCE

11.5.4 Tornadoes
Heavy precipitation, including hail as well as strong winds associated with a severe
thunderstorm can cause widespread damage. However, the above magnitude of damage is
considered insignificant when compared to the destructive capability of the so-called
tornadoes. Tornadoes are columns of extremely rapid and violently rotating air extending
beneath the base of a thunderstorm (cumulonimbus) cloud. While some tornadoes appear as a
very thin rope-shaped column of air, most tornadoes have the familiar funnel shape that
narrows from the cloud base to the ground. While a large number of tornadoes rotate
cyclonically (counterclockwise) in northern hemisphere, a few of the tornadoes do rotate
anticyclonically as well. Strong tornadoes are associated with very large differences (of the
order of 100 hPa) in atmospheric pressure over very short distances (of a few tenths of a
kilometre) between the tornado core and the region outside the tornado. Tornadoes exhibit a
very wide range of sizes with diameters varying from about 100 m to 1500 m. Normally, most
tornadoes last only for a few minutes, while some of them may last for several hours.
Tornadoes move across the surface at speeds of about 12 m s–1. Typically, a tornado covers a
horizontal extent of about 3 km to 4 km over land from the time it touched the ground until
it has dissipated. While typically tornadoes move from south-west to north-east, they have
been known to move in any direction. Over India, tornadoes occur very rarely and can occur
only over the north-eastern parts of India. The maximum numbers of tornadoes, which occur
over the whole globe, are seen over the United States of America and they amount to about
750 tornadoes in a year. The months of May and June account for the largest number of
tornadoes, which occur in USA, while the number is somewhat less in April and July. In
USA, while the months of May and June account for about 170 and 165 tornadoes in a year,
April and July account for 100 and 85 tornadoes every year. Also, in USA, the months of
August and March have about 60 and 50 tornadoes every year. Canada averages much less,
about a 100 tornadoes every year.
When tornadoes occur over warm water bodies they are known as waterspouts.
Waterspouts are much smaller than tornadoes and typically have diameters between 5 m and
100 m and are considered somewhat weaker as compared to the tornadoes. The waterspouts
can have wind speeds of the order of 38 m s–1 and hence has the capability to severely
damage boats and small ships. While most of the waterspouts form over the water bodies
themselves, some of them may develop as land-based tornadoes, which move offshore. The
visible water seen in the waterspout is not associated with the sucked-up water from the ocean
below, but chiefly arises from the water vapour in the air.

11.6 EL NINO-SOUTHERN OSCILLATION


El Nino-Southern Oscillation (ENSO) is a global coupled ocean–atmosphere phenomena,
which is known to affect the interannual variability in weather and climate through most of
the world. While the El Nino indicates the ocean signature of the ENSO event, the Southern
Oscillation refers to the atmospheric signature of this global coupled ocean–atmosphere
phenomena. The El Nino Southern Oscillation phenomenon is briefly discussed in the
following subsection.
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 317

11.6.1 Overview of ENSO


El Nino-Southern Oscillation (ENSO) as it is called is one of the most important
manifestations of the interaction of tropical atmosphere with warm tropical ocean. Since
ENSO embodies the complex interaction of the atmosphere and oceans over the equatorial
Pacific Ocean; it is conveniently described with respect to the two components—the oceanic
component denoted by El Nino and the atmospheric component known as Southern
Oscillation. The El Nino event is characterized by the appearance of unusually warm waters
over the eastern Pacific Ocean, especially near the coast of Peru. Since the above warm event
over East Pacific coincides in its occurrence with the Christmas season, the event is named as
El Nino, a reference to the Christ child. However, the El Nino is much more than a simple
oceanic warming over the equatorial East Pacific. It can be shown to result from a complex
interaction with the tropical atmospheric circulation over Pacific, the latter being called the
Walker circulation. Subsequent investigations have lead to the evidence that the warming over
Pacific is not confined over the Peru coast but extends from eastern Pacific right up to central
Pacific (180°).
Southern Oscillation refers to the changes in the sea surface conditions related to the
seesaw change in the atmospheric pressure distribution over the tropical Pacific Ocean. Under
normal conditions, the western Pacific is characterized by warm sea surface temperatures
(SSTs), low sea level atmospheric pressure (SLP), cloud formation and associated rainfall and
weather activity. Figure 11.15 shows a schematic diagram of the normal conditions, which
prevail over Pacific. The eastern Pacific, on the other hand is characterized under normal
conditions by cold SSTs, high SLP, relatively cloud-free conditions and very little weather.
Further, under normal conditions, the easterly trade winds, which blow from east to west over
the equatorial Pacific Ocean, move warm surface water near the equator westward causing
higher temperatures over the western Pacific Ocean as well as higher sea level elevation over

Normal Conditions

Convective
Circulation

Equator

Thermocline

120°E 80°W

FIGURE 11.15 Schematic diagram of the normal conditions


which prevail over the Pacific Ocean.
318 u BASICS OF ATMOSPHERIC SCIENCE

the western Pacific Ocean. Warmer water over western Pacific leads to higher air temperature,
lower SLP and development of convective rainfall. The Walker circulation in the equatorial
Pacific Ocean is an east-west zonal circulation with easterly trade winds over the lower levels
of the atmosphere and ascending/descending regions over the western/eastern Pacific Ocean.
Although, many texts indicate westerly (west to east) flow at upper levels to complete the
Walker circulation, recent evidence indicates that the above simple picture of the return flow
may not be completely correct. Sir Gilbert Walker in 1924 first documented the seesaw in the
atmospheric SLP over the Pacific Ocean and hence the above zonal circulation over Pacific
Ocean is named after him. Sir Gilbert Walker, then working for the India Meteorological
Department identified the seesaw in the pressure pattern over Pacific while undertaking a
study to investigate the cause of the failure of the Indian summer monsoon. Sir Walker found
that whenever the SLP over western Pacific Ocean was low/high, the same over the eastern
Pacific Ocean was exactly opposite, i.e. high/low.
Before dwelling on more details of the ENSO, it is important to introduce a few
definitions. The vertical temperature structure of the tropical oceans are characterized broadly
by three layers: the top mixed layer extending by a few tens of metres and having a uniform
temperature in the vertical, and the lowest bottom layer which extends from a few to several
hundreds of metres till the ocean bottom and is characterized by very small and gradual
decrease in temperature with depth. The above two layers are separated by a sharp layer of
transition called the thermocline which can extend up to a few hundreds of meters in depth
and is characterized by the presence of very strong decrease of temperature with depth. When
the waters over the ocean surface diverge or separate from one another, mass conservation
requires that the subsurface water is welled up and replaces the surface divergent water. This
process called upwelling, generally brings cooler and nutrient-rich subsurface water on to the
surface. Fishing industry is generally known to thrive in coastal upwelling regions, as fishes
are abundant in nutrient-rich surface waters. Further, the Coriolis force produces an apparent
deflection on all moving bodies (including fluid parcels) irrespective of their direction of
motion. The deflections are to the right/left in the northern/southern hemisphere. The effect of
the Coriolis force increases with increasing latitude and is absent over equator. Due to the
drag exerted on the surface water of oceans by winds, ocean currents are observed. The
surface water moves at an angle of 45° to the right (left) of the winds in the northern
(southern) hemisphere. The ocean current turns increasingly to the right (left) in the northern
(southern) hemisphere with depth and decreases in speed at greater depths. Around 100
meters below the surface, the wind-driven ocean current dies down and this layer is known as
the Ekman layer. Upwelling of subsurface water can occur near coasts if the direction of the
wind is parallel to the coast. Consider the example of the wind blowing from north to south
parallel to an eastern coast in northern hemisphere. The Ekman deflection will be to the right
and results in propelling surface waters offshore (from land to ocean). This will enable cooler
subsurface water to rise and replace the displaced surface water. The coasts of Peru
experience a northward ocean current, which from the Ekman displacement will lead to
conditions conducive for coastal upwelling. Further, the equatorial easterlies (trade winds)
blowing over regions north and south of equator will lead to surface divergence due to
Coriolis effect. This will lead to equatorial upwelling over Pacific. Also, due to the piling of
the water westward by the trade winds, the thermocline is shallow over the eastern Pacific and
relatively deeper over the western Pacific (refer Figure 11.15).
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 319

To sum up, under normal conditions the western Pacific has a higher elevation of sea
level, a deeper thermocline, lower SLP, warmer SST and ascending motion in the atmosphere
giving rise to cloud formation and weather. The conditions at the eastern Pacific are lower
elevation of sea level, a shallow thermocline, higher SLP, colder SST and descending motion
and absence of weather. Further, during normal conditions, easterly trade winds are strong, the
Walker circulation is strong, coastal upwelling is seen over Peru and equatorial upwelling is
observed over equatorial Pacific. Now, imagine a slight weakening of the easterly trade winds.
If the above weakening leads to a situation wherein a further weakening of the easterly trade
winds happens, one says that the above is an example of a positive feedback process. That is,
a change in any variable (parameter) leads to processes in the ocean–atmosphere system,
which tends to reinforce or amplify the change. Now, a weakening of the easterly trade winds
is likely to cause less of piling of warm water to the west. This is likely to result in lowering
of the sea level elevation in western Pacific and a corresponding increase of the same in
eastern Pacific. Also, this leads to lesser transport of heat by westerly moving water and
consequently intensification of warm SST anomaly over the Eastern and Central Pacific
Oceans. Further, this is likely to cause deepening of the thermocline in the Eastern Pacific and
a weakening of the coastal upwelling over Peru and Eastern Pacific. Also, weakening of the
easterly trade winds will also reduce equatorial upwelling. The reduced upwelling over
Eastern Pacific causes the warm surface layer to deepen and with reduced equatorial
upwelling causes the deepening of warm surface layers to extend throughout the tropical
Pacific basin up to 180° (date line). Warmer temperatures over Eastern Pacific will lead to
higher air temperatures, lower SLP and development of convective rainfall there. Higher air
temperature and lower SLP will lead to ascending air over Eastern Pacific, further weakening
the easterly trade winds. Meanwhile, conditions in the Western Pacific change over to lower
sea level elevation, lower SST, and a relatively shallow thermocline. Lower SSTs over
western Pacific lead to lower air temperature, higher SLP, and cessation of weather activity.
The continued application of the positive feedback mechanism leads to an ENSO (warm)
event. Figure 11.16 shows a schematic diagram illustrating a warm (El Nino) event in the
El Nino Conditions

Equator

Thermocline

120°E 80°W
FIGURE 11.16 Schematic diagram illustrating a warm (El Nino) event in the Pacific Ocean.
320 u BASICS OF ATMOSPHERIC SCIENCE

Pacific Ocean. It is clear from the above discussions that the existence of positive SST
anomalies over eastern and central Pacific Ocean during the warm event is attributed to
internal processes responsible for the interaction within the combined ocean–atmosphere
system. The ENSO event is not due to any change in the external forcing, say, change in
solar radiation, impact of an asteroid or say a volcanic eruption. Jacob Bjerknes in 1960s
proposed the above hypothesis to explain the occurrence of an ENSO (warm) event in the
Pacific Ocean.
Figure 11.17 shows a schematic diagram of the progression of the warming in terms of
intensity and horizontal extent of the sea surface temperature (SST) anomalies over the
Pacific Ocean for the recent El Nino event of 1997–1998. The warming shows up as a small
positive anomaly in the SST of about 1°C off the coast of northern Peru and Ecuador in the
month of March. In another couple of months, i.e. by May 1997, while the magnitude of the
SST anomaly has risen to 2°C off the coast of northern Peru and Ecuador, the warming is
seen extending over the entire equatorial Pacific Ocean with a SST anomaly of about 1°C. It
is quite possible that the SST anomaly may weaken during subsequent months leading to
normal conditions. It is still too premature to declare that a warming (El Nino) event has
occurred over the Pacific. However, during some years, say for example 1997, the warming
continues with time both in its horizontal extent as well as in its intensity, i.e. magnitude of
the SST anomaly. As can be seen in Figure 11.17, the warming (El Nino) event has its
maximum intensity (with SST anomaly of about 5°C) and its largest horizontal extent over
Pacific during January 1998. By March 1998, after reaching its maximum intensity, the SST
anomalies start decreasing in their magnitude as well as in their horizontal extent.
The ENSO (warm) event reappears again in 2 to 7 years and lasts for a duration, which
ranges from 12 to 18 months. The ENSO (warm) event starts as a slight warming over eastern
Pacific during the northern spring (March to May) season and has its maximum amplitude
during the southern summer (November to March) season. When an El Nino (warm) event
dissipates, there can either be a return to normal surface conditions or a situation where there
is further cooling in the tropical eastern Pacific. The latter resulting event is designated as a
La Nina (cold) event in which the conditions over the Pacific Ocean are reversed with respect
to the El Nino event. Figure 11.18 shows a schematic diagram illustrating a cold (La Nina)
event in the Pacific Ocean. A measure of the ENSO activity is defined as the Southern
Oscillation Index (SOI), which is defined as the monthly sea level pressure departure from
normal at Tahiti in eastern Pacific minus the departure from normal at Darwin in western
Pacific. Positive SOI indicates normal or La Nina conditions, while negative SOI indicates
El Nino conditions have occurred.
The same positive feedback mechanism can be invoked to explain the formation of a
La Nina (cold) event from normal conditions. Instead of a weakening of the easterly trade
winds, imagine a strengthening of the easterly trade winds and use of similar arguments as
advanced earlier will lead to the formation of La Nina (cold) event. The following questions,
however, remain to be answered. If the appearance of the El Nino and the La Nina can be
explained on the basis of positive feedback mechanisms, how does one explain their
termination? Also, why is there an oscillation from one state to the other? It would be
tempting to suggest that there must exist negative feedback mechanisms, which will be
responsible for the cessation of the events. For oscillation to occur naturally in the ocean–
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 321

January 1997 March 1997


30°N 30°N
20°N 20°N 1
1
10°N 10°N 1
1
1
E0 E0
1
2
10°S 10°S
20°S 1 20°S –1
1 2 –1
1
30°S 30°S
180 150°W 120°W 90°W 60°W 180 150°W 120°W 90°W 60°W
May 1997 July 1997
30°N 2 30°N
3
20°N 20°N 1
1 1
10°N 1 10°N
1
E0 1 E0 2
–1 –1 1
10°S 10°S
20°S –2 20°S
–2
30°S 30°S
180 150°W 120°W 90°W 60°W 180 150°W 120°W 90°W 60°W
September 1997 November 1997
30°N 30°N
–3
20°N –1 20°N –2

10°N 1 10°N –1
32 1 3 2
E0 5 E0 4 5
2 43
10°S 10°S 2
1
20°S 1 20°S 3
1 1 1
2 2
30°S 30°S
180 150°W 120°W 90°W 60°W 180 150°W 120°W 90°W 60°W
January 1998 March 1998
30°N 30°N
20°N 20°N
10°N 10°N
2 1
1
E0 3 E0 1
3 1
10°S 2 10°S 1
–1
1
20°S 1 20°S –2 –1
–3
30°S 30°S
180 150°W 120°W 90°W 60°W 180 150°W 120°W 90°W 60°W

1 2 3 4 5
FIGURE 11.17 Schematic diagram of the progression of the warming in terms of intensity and
horizontal extent of the SST anomalies over the Pacific Ocean for the El Nino event
of 1997–1998.
322 u BASICS OF ATMOSPHERIC SCIENCE

La Nina Conditions

Equator

Thermocline

120°E 80°W
FIGURE 11.18 Schematic diagram illustrating a cold (La Nina) event in the Pacific Ocean.

atmosphere system from one state to the other (El Nino to La Nina or vice versa), there must
exist delayed negative feedback mechanisms which are out of phase with the above-described
positive feedback processes. Hence, one requires one or more restoring processes that lag
behind the positive feedback processes, which tend to take the system to either of the two
extreme states. As the system tends to either of the two extreme states, the delayed restoring
processes also grow and can become large enough to completely dominate and swamp the
positive feedback processes, bringing the system back either to the normal conditions or
towards the opposite extreme state. A few of the negative feedback processes proposed are the
delayed oscillator mechanism or the recharge oscillator mechanism.
Teleconnections is the term utilized to relate statistically significant correlations between
weather events that occur at different parts of the earth. When one finds enhanced
precipitation or drought over some regions of the globe during an El Nino/La Nina event,
then one speaks of El Nino/La Nina Teleconnections. These enhanced precipitation or drought
may occur in the same year as the El Nino/La Nina event or may occur a year after the
El Nino/La Nina event began. These patterns of teleconnections associated with El Nino/La
Nina occur due to the shifting of the major heat source over the equatorial Pacific. Such a
shift of the major heat source in the equatorial Pacific modifies the atmospheric circulation
causing changes in the weather patterns worldwide. The effects of El Nino are stronger over
regions in South America as compared to regions in North America. While southern Brazil
and northern Argentina have wetter than normal conditions during the spring and early
summer season, the coasts of northern Peru and Ecuador, experience major flooding during a
strong El Nino event. While drier and hotter weather is seen in parts of the Amazon River
Basin, Colombia and Central America, during an El Nino event, central Chile experiences a
mild winter with large rainfall. El Nino is responsible for drier conditions to occur in parts of
South-East Asia and Northern Australia, while wetter than normal conditions are seen between
March to May over East Africa. While drier than normal conditions are seen from December
to February in South-Central Africa during an El Nino event, central and southern California,
north-west Mexico and the south-western United States experience wetter and cooler
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 323

conditions during northern winter. However, winters in the United States are warmer than
normal in the upper Mid-West states, the North-East, and Canada during an El Nino event.

11.6.2 Indian Ocean Dipole


The Indian Ocean Dipole (IOD) is a climate mode that occurs interannually in the tropical
parts of the Indian Ocean. The IOD is a coupled ocean atmospheric phenomena characterized
by anomalous cooling of SST in the south-eastern equatorial Indian Ocean and anomalous
warming of SST in the western equatorial Indian Ocean. Associated with these changes in the
SST distribution, the normal convection situated over the eastern Indian Ocean warm pool
shifts to the west causing heavy rainfall over East Africa and severe droughts over the
Indonesian region. A measure of the intensity of the IOD is expressed in terms of the
anomalous SST gradient between the western equatorial Indian Ocean (50°E–70°E and
10°S–10°N) and the south-Eastern equatorial Indian Ocean (90°E–110°E and 10°S–0°N). The
above-mentioned SST gradient is known as Dipole Mode Index (DMI). When the DMI is
positive/negative, the phenomena are referred as the positive/negative IOD. When a dipole is
forming, the wind patterns change in the equatorial Indian Ocean from months of April–May
and the dipole peaks in the month of October. When a positive dipole is evolving, the winds
in the equatorial Indian Ocean blow from east to west. These anomalous easterly winds lead
to changes in the ocean currents. The changes in the ocean currents cause the Arabian Sea off
the Somali coast in Africa to become unusually warm, enhancing cloud formation over there,
while the eastern Indian Ocean around Indonesia becomes colder than normal and more dry.
During a negative dipole year, the opposite happens, and the Arabian Sea near Africa
becomes cooler and less cloudy, while the Indian Ocean is warmer and rainier around
Indonesia. Furthermore, studies have indicated that the IOD events are largely independent of
ENSO and arise out of the coupled ocean–atmosphere dynamics in the tropical Indian Ocean
itself. Only 35% of the IOD events have co-occurred along with ENSO which clearly
indicates the independent nature of the IOD. While, in 1961/1997, a positive dipole occurred
without/with an El Nino; in 1967, a positive IOD did occur along with a La Nina. Hence, it is
clear that a positive IOD can occur under normal, El Nino and La Nina conditions over Pacific.

11.6.3 ENSO and Indian Monsoon


Studies in the early 1980s on interannual variation of the Indian monsoon showed a clear link
between the All India Summer Monsoon rainfall and the ENSO. The interannual variability of
a planetary-scale circulation like the Indian monsoon which links two hemispheres is related
to ENSO, the most dominant signal of interannual variation of the coupled atmosphere–ocean
system over the Pacific. It was shown that there is an increased tendency over India towards
floods during the cold (La Nina) event and of excess deficit of rainfall over India, i.e.
droughts during the warm (El Nino) event. The SST anomaly over the Nino-3, a region of
eastern Pacific from 150°E–90°W and 5°S–5°N, can be considered as an index of the ENSO
variability. Studies have shown that the long-term correlation of the seasonal anomalies of the
India monsoon and the Nino-3 SST is 0.63, significant at 99.9%. That is, the above
correlation represents a tendency for the Indian monsoon rainfall to be less than normal when
the eastern Pacific is warm. The physical mechanism by which ENSO is related to the Indian
324 u BASICS OF ATMOSPHERIC SCIENCE

monsoon can be explained as follows. It is to be noted that the interannual variations of the
Indian monsoon is characterized by fluctuations of a regional Hadley (meridional)-type
circulations. While a strong Indian monsoon is characterized by anomalous ascent of air
around 25°N, a weak Indian monsoon is associated with an anomalous ascent of air over
equator. It is speculated that ENSO affects the Indian monsoon through an interaction
between the equatorial Walker circulation and the regional Hadley circulation. The warm
ENSO events are characterized by a shift of the equatorial Walker circulation eastward
towards eastern Pacific. The above shift in the Walker circulation results in low-level
convergence over equatorial Indian Ocean. The above convergence results in an anomalous
regional Hadley circulation with ascent over equator and descent over the Indian continent.
Such a descent over the Indian continent leads to below normal monsoon rainfall. Figure 11.6
shows the El Nino and the La Nina years for the period of 1871–2003 as seen in the rainfall
anomaly of the Indian summer monsoon season. It is clear from Figure 11.6 that the Indian
monsoon ENSO relationship is indeed a complex one.
Consistent with the above relationship between ENSO and the Indian monsoon, the cold
(La Nina) event of 1988 was associated with large excess in the all India summer monsoon
rainfall, while the El Nino events of 1982 and 1987 were associated with droughts over India.
However, starting from 1988, India did not experience a single drought in the next fourteen
years despite the occurrences of the El Nino. In fact, the Indian monsoon was above normal
in 1997, despite the occurrence of the strongest El Nino of the 20th century, which occurred
in 1997. Again, while a weak El Nino was known to be forming in 2002, none of the
predictions of the Indian monsoon forecasted a large deficit (19%) of the all India rainfall as
was observed in 2002. It is clear that the relationship between ENSO and the Indian monsoon
is indeed a complex one and it might as well be that ENSO is just one (and possibly one of
the most important) and not the only important influence on the Indian monsoon.
Studies have indicated that the atmospheric component of the IOD, called the Equatorial
Indian Ocean Oscillation (EQUINOO), could be playing an important role in influencing the
Indian monsoon. It is proposed that the anomalous wind patterns during the Indian monsoon
associated with EQUINOO can seriously affect rainfall. Studies have indicated that the
combination of EQUINOO and ENSO account for much of the Indian monsoon years with
large excess or deficit rainfall. However, the same studies indicate that neither EQUINOO nor
ENSO separately or in combination could explain small variations in the Indian monsoon
rainfall. The EQUINOO has been invoked to explain not only the droughts that occurred in
the absence of El Nino or in the presence of a weak El Nino, but also excess rainfall seasons
in which ENSO did not predominate.

REVIEW QUESTIONS

1. Name the region which has the greatest incidence of major tropical cyclones.
2. Name the typical tropical cyclone season over the North Indian Ocean.
3. Why are there no tropical cyclones during the Indian south-west monsoon season?
4. Name the regions in tropics where tropical cyclones are absent.
5. What is monsoon?
METEOROLOGICAL SYSTEMS IN LOW LATITUDES u 325

6. Describe the differences between tropical cyclones and extratropical cyclones.


7. What are easterly waves?
8. The right side (relative to its direction of movement) of a tropical cyclone in northern
hemisphere is supposed to be the most dangerous region of a tropical cyclone? Why?
9. What are the ocean conditions necessary for the formation and maintenance of tropical
cyclones?
10. What is CISK?
11. How many tropical cyclones form on an average throughout the world in a year?
12. How many tropical cyclones form on an average over the Indian seas in a year?
13. Tropical cyclones have a tendency to move poleward in both the hemispheres. Why?
14. Tropical cyclones have a tendency to move westward in both the hemispheres. Why?
15. What is the nature of the vorticity distribution near the centre of a tropical cyclone?
16. What is the radius of maximum wind?
17. What is the eye of the tropical cyclone?
18. During which time of the year, do tropical cyclones form worldwide?
19. Mention the balance of forces associated with a tropical cyclone.
20. Why don’t tropical cyclones form over the equator?
21. What is North-East Indian monsoon?
22. How is the North-East Indian monsoon different from the South-West Indian monsoon?
23. Define the onset of monsoon over Kerala (India).
24. What is “Onset Vortex”?
25. What is the average South-West monsoon rainfall over India?
26. Which are the states of India which experience North-East Indian monsoon?
27. Which are the states of India which do not experience South-West Indian monsoon?
28. Are there any special reasons for the Indian summer monsoon onset to be spectacular?
29. Define intraseasonal monsoon variability and interannual monsoon variability.
30. How many monsoon depressions form over the Indian region during the monsoon season?
31. Unlike the tropical cyclones, monsoon depressions do not weaken as they cross land. Why?
32. Mention the average lifetime of the monsoon depression?
33. Name the typical months when thunderstorms prevail over the eastern parts of India.
34. What is a squall line?
35. What are the conditions necessary for the formation of thunderstorms over land?
36. Mention the typical lifetime of a thunderstorm cell.
37. What are Kal Baisakhis?
38. What is hail?
39. Thunderstorms form over land typically during late afternoons or early evenings. Why?
40. What are tornadoes?
41. What are water spouts?
42. What is the typical value of the Rossby number associated with a tornado?
43. What is El Nino?
326 u BASICS OF ATMOSPHERIC SCIENCE

44. What is Southern Oscillation?


45. What is El Nino Southern Oscillation?
46. What is a thermocline?
47. What is La Nina?
48. Does the El Nino affect the Indian summer monsoon?
49. What is the average timescale of an El Nino event?
50. Can we predict the El Nino event?
51. What is ENSO index?
52. What is Indian Ocean Dipole?
53. What is EQUINOO?
12 Global Energy
Balance

The earth receives almost all of its energy from the sun. The earth does absorb a good part of
the incident radiation from the sun. To maintain equilibrium, the earth must warm up and
radiate energy at the same rate at which the energy is received. Due to the presence of
greenhouse gases such as water vapour and carbon dioxide, the earth’s surface temperature is
considerably higher than would be the case in the absence of the greenhouse gases.
The first law of thermodynamics is a statement of the conservation of energy and this
energy conservation provides the basis for conversion of one form of energy to another form,
following a process. It turns out in the real atmosphere, not all the potential energy is
available for conversion to kinetic energy. The above gives rise to a new concept of “available
potential energy” (APE). The production and dissipation (through friction) of kinetic energy
of the atmosphere are important issues to be addressed. The atmosphere can be treated as a
heat engine as it extracts energy from a warm source (low-latitude regions), performs work
and delivers the unused energy at a cold sink (high-latitude regions).
It is important to identify and take into account the various sources and sinks of energy
from a standpoint of global averaged energy balance. Section 12.1 outlines in detail the
globally-averaged energy balance. The definition of the internal energy per unit cross-sectional
horizontal area of an atmospheric layer and its relationship to the potential energy per unit
area of the atmospheric layer is presented in Section 12.2. While Section 12.3 deals with the
conversion of potential and internal energy to kinetic energy and introduces the concept of
available potential energy, Section 12.4 is concerned with the production and frictional
dissipation of kinetic energy. The last Section 12.5 presents the atmosphere as a heat engine
and gives the efficiency of the atmospheric heat engine.

12.1 GLOBALLY-AVERAGED ATMOSPHERIC ENERGY BALANCE


The solar energy is the chief source of energy for maintaining the atmospheric and the
oceanic circulations. Since the earth is a sphere, the solar energy input is not uniform over the
327
328 u BASICS OF ATMOSPHERIC SCIENCE

entire earth. While the maximum solar energy is received over the equatorial regions, the
polar regions receive much lower solar energy. However, the long-wave energy radiated by
the earth and its atmosphere which is lost to space is much less latitude-dependent with the
result; the equatorial regions have a net surplus of radiation while the polar regions have a net
deficit. The atmospheric and the oceanic circulations ensure that the energy is transported
from the surplus regions to the deficit regions so that a globally-averaged energy balance is
achieved. The following subsections discuss briefly the globally-averaged atmospheric energy
balance.

12.1.1 Global Energy Balance Requirement for the Earth’s


Atmosphere
Figure 12.1 shows the annual mean distribution of the short-wave (continuous line) and
long-wave (dotted line) radiation from 90°S to 90°N. One can see that, the low latitudes
(30°S to 30°N) receive a larger amount of solar radiation, while the regions over the poles
receive much less solar radiation. The amount of energy received from the sun at the earth’s
surface depends on the intensity of the short-wave solar radiation and the length of the day. It
is well known that the earth is a sphere, tilted at 23.5° to the orbital plane and rotates in an
elliptical orbit having a small eccentricity of 0.017 around the sun. Also, the earth axis is
oriented in a fixed direction in space as it travels around the sun.

FIGURE 12.1 Annual mean distribution of the latitudinal distribution of net short-wave solar
radiation received at the earth surface and the net long-wave terrestrial radiation
lost to space.

Once a year (June 21) the North Pole points directly towards the sun, while the South
Pole does not receive any radiation from sun. Another six months later (December 21) the
North Pole does not receive any solar radiation, while the South Pole points directly towards
the sun. This is the reason for the occurrence of seasons. The changes in season or the
seasonal cycle can then explain the change in solar radiation received at the earth surface.
Figure 12.1 already incorporates the effect of change in the incidence angle of the incoming
radiation with latitude, as well as the number of hours of sunlight in a day in determining the
solar radiation received at the earth surface.
GLOBAL ENERGY BALANCE u 329

Unlike the incoming solar radiation, which depends very strongly on latitude, the
outgoing long-wave radiation loss by the earth varies by less than 30% over the entire
globe. The distribution of net radiation (the difference between incoming solar radiations
gain and the outgoing long-wave earth radiation to space) in Figure 12.1 shows a net gain
of radiation over the low-latitude regions and a net deficit of radiation over the high-latitude
regions. If the low-latitude regions were to get continuously heated and the high-latitudes
regions continuously cooled, the temperatures over the earth would be very different in the
low and high latitudes than the observed. This is because heat is transported from the
regions of excess heat (low-latitude regions) to regions of heat deficit (high-latitude
regions) by the atmospheric winds and the ocean currents. Further, despite the large
amounts of energy being added or removed from the atmosphere by all the energy transfer
processes, there is evidence to suggest that the amount of energy in the atmosphere is not
increasing or decreasing systematically. That is over a long-time scale (of the order of an
year) there is a balance between the energy sources and the energy sink for the entire
atmosphere as a whole.
A form of the first law of thermodynamics in terms of specific enthalpy h is
dq = dh – a dp (12.1)
where the terms have their usual meaning. It is assumed that the only work done is due to the
work of expansion. Since h = cP T and from hydrostatic relation involving geopotential
dj = g dz = – a dp, one gets

dq = d(h + j) = d(cP T +ÿ j)
which is written as
E D Q5  K

R (12.2)
EU
From Eq. (12.2), it is clear that for a hydrostatic atmosphere, the net diabatic heating or
cooling of an air parcel is related to the rate of change of specific enthalpy and the
geopotential of the air parcel. The quantity cP T + j is called dry static energy, and is
conserved during unsaturated vertical and horizontal motion. The diabatic heating rate is due
to the following processes, occurring in the atmosphere, such as (i) absorption of short-wave
incoming solar radiation as well as absorption and emission of long-wave radiation from the
earth–atmosphere system, resulting in net radiative heating rate SR, (ii) release of latent heat
associated with the various phase changes of water, SLH, and (iii) exchange of heat with the
surroundings through conduction (due to random molecular motion) and convection (due to
mixing by organized fluid motion), SH. Equation (12.2) can then be represented from the
above processes as
E D1 5  K

R 43  4-)  4) (12.3)
EU
EX
The rate of change of water vapour mixing ratio following the motion, can be made up
EU
of two components—the first associated with phase changes between vapour to liquid and
330 u BASICS OF ATMOSPHERIC SCIENCE

vapour to solid, while the second is related to the exchange of water vapour molecules with
the surroundings due to exchange processes and can be written as
EX È EX Ø È EX Ø
ÉÊ ÙÚ É Ù (12.4)
EU EU QIBTF DIBOHF Ê EU Ú FYDIBOHF
The heating due to latent heat release associated with the phase changes of water can be
written as

4-)  - ÈÉÊ EX ØÙÚ (12.5)


EU QIBTF DIBOHF
where L is the appropriate latent heat and the exchange term can be written as

4N - ÈÉÊ EX ØÙÚ (12.6)


EU FYDIBOHF
Combining Eqs. (12.3), (12.4), (12.5) and (12.6), the following relation is got
E D1 5  K  -X

4 3  4 )  4N (12.7)
EU
The quantity cP T + j + Lw is called moist static energy, and is not affected by the process of
condensation. Over a long-time scale and for the entire atmosphere, the moist static energy
can be assumed to be neither increasing nor decreasing and hence
E
\D1 5  K  -X ^  (12.8)
EU
where the over bar above a quantity indicates a long-term average of the quantity of the order
of an year and the double brackets refer to the summation over the entire mass of the
atmosphere. From Eqs. (12.7) and (12.8), one gets

\ 43  4)  4N ^  (12.9)
Equation (12.9) is a statement of the energy balance requirement for the entire earth’s atmosphere.

12.1.2 Global Energy Balance at the Earth Surface


The long-term averaged (of the order of an year) and the globally-averaged net upward flux
of moist static energy per unit area from the earth surface is determined by the net irradiance
En at the earth surface, the fluxes of the sensible heat ESH and the latent heat flux ELH,
respectively and the same is expressed as
< &> & & &
< O >  < 4) >  < -) > (12.10)
The square bracket in Eq. (12.10) indicates that the averaging is performed over the surface
of the whole earth. The net irradiance is made up of both the short-wave and long-wave
components and is given by
< &O > < & - >  < & - ‘ >  < & 4 > (12.11)
GLOBAL ENERGY BALANCE u 331

where ES is the incoming solar radiation that is actually absorbed at the surface, EL ­ is the
emission of long-wave infrared radiation from the earth surface and EL ¯ is the downward
emission from the atmosphere to the earth surface. ES is the incident solar radiation minus the
reflected portion due to the albedo of the earth surface.
The total energy ET is the sum of internal, potential, kinetic, chemical and nuclear
energy of the solid earth as well as the energy of the ocean caps and the biosphere.
Considering the earth to be spherical with radius RE, the total outgoing flux averaged over the
entire surface of earth equals Q 3& < & > and from the conservation of energy principle, the


local decrease in the combined total energy ET stored in the system will be related to the total
outgoing flux at the surface, i.e.
˜&5
 Q 3& < & > (12.12)
˜U
The globally-averaged and time (over a decade)-averaged net flux of energy through the earth
surface (on the left-hand side of Eq. (12.10)) is at least two orders of magnitude smaller
compared to each of the right-hand side terms.
The net flux of energy through the earth surface is due to (i) burning of fossil fuels,
nuclear energy and waste energy from human activity, (ii) leakage of geothermal energy
through the earth’s crust, (iii) fluxes associated with temperature change in the upper layers of
the ocean, and (iv) energy associated with changes of continental ice sheets. The rate of
release of globally-averaged energy and time-averaged energy over a decade for each of the
four processes mentioned above is about two orders less than the right-hand side of
Eq. (12.10). Hence, the left-hand side of Eq. (12.10) can be neglected compared to the right-
hand side terms of Eq. (12.10) which are of the order of 30–100 W m–2. This results in the
following requirement for the energy balance at the earth surface, globally-averaged and time-
averaged over a decade, of the form

< &4 >  < & - ‘> < &- >  < &4) >  < &-) > (12.13)
Since the net flux of energy through the earth surface is very small and negligible and the
energy stored in the atmosphere is not increasing or decreasing, it can be safely presumed that
taken over the entire globe, the net flux of energy through the top of the atmosphere should
also be very small and negligible. Also, since almost all the transfer of energy between the
earth system and the rest of the universe is through radiative processes, it can be safely
concluded that the earth–atmospheric system must be fairly close to radiativve equilibrium. If
A is the planetary albedo, S is the irradiance incident upon a plane normal to the incident ray
at the top of the atmosphere, RE is the radius of earth, then the outgoing infrared irradiance of
planetary radiation emitted to space under radiative equilibrium can be calculated as follows.
Equating incoming flux to the outgoing flux, one gets

  "
4 Q 3&
& Q 3& (12.14)
Averaging over the globe and time averaging over a long-time scale, one gets

< &> <   "


4 > (12.15)
332 u BASICS OF ATMOSPHERIC SCIENCE

12.1.3 Estimates of the Global Energy Balance for


the Earth–Atmospheric System
Figure 12.2 presents the average estimates of the global energy balance based on the actual
observational data for the earth–atmospheric system. At the top of the atmosphere, the amount
of solar radiation received is effectively S/4 or 345 W m–2 and the above S/4 irradiance of
energy is assumed to represent 100 units of incident solar radiation at the top of the atmosphere.

Incoming solar Reflected solar radiation Outgoing irradiation


SPACE radiation 100 6 20 4 6 38 26
Back-scattered
Atmosphere by air
Net emission
by H2O, CO2
16
Absorbed by Reflected by Emission
H2O, Dust, O3 clouds by clouds

3
Absorbed in 15 Absorption
by H2O, CO2
clouds Reflected by
surface
Net emission of
irradiation Latent
from surface heat flux
Sensible
heat flux
Ocean, Land 51 21 7 23

FIGURE 12.2 Average estimates of the global energy balance for the earth–atmospheric system.

About 30% (30 units) of the incident solar radiation is reflected back to space. Two
thirds of the reflected radiation, i.e. 20 units is reflected by clouds, 4 units are reflected by
earth surface and the remaining 6 units are reflected by the atmosphere. Out of the remaining
70 units of the incident solar radiation, 19 units are absorbed by the atmosphere and clouds,
while the remaining 51 units are absorbed at the earth surface. Out of 19 units of incident
solar radiation absorbed during the passage through the atmosphere, 16 units are absorbed in
cloud-free air by water vapour, ozone and dust particles, while 3 units are absorbed by clouds.
The earth loses the 51 units of solar radiation, which it has absorbed through 21 units of net
emission of long-wave infrared radiation from the earth surface (i.e. upward emission of long-
wave infrared radiation from the earth surface minus the downward emission of long-wave
infrared radiation from the atmosphere to the surface), 23 units of latent heat flux through
melting and evaporation of water and 7 units of sensible heat flux through conduction and
convection processes.
The atmosphere loses 64 units to space through 38 units of net upward emission of
long-wave infrared radiation by the greenhouse gases (primarily from water vapour and
carbon dioxide), and 26 units of upward emission of long-wave infrared radiation by clouds.
The atmosphere gains 15 units due to absorption of greenhouse gases from the 21 units of net
upward emission of infrared radiation from the earth surface while the remaining 6 units are
lost to space. The atmosphere gains in total 64 units—15 from long-wave infrared radiation
GLOBAL ENERGY BALANCE u 333

from the earth, 19 units of incoming solar radiation, 23 units from latent heat flux and 7 units
from sensible heat flux. At the top of the atmosphere, the amount of radiation lost to space is
100 units in which 30 units are reflected solar radiation, while the remaining 70 units are
outgoing long-wave infrared radiation. Thus, the global energy balance of the earth–
atmospheric system is achieved with the atmosphere, earth surface and outer space having
neither deficit nor surplus of energy.

12.1.4 Energy Processes in the Upper Atmosphere


The absorption of incoming short-wave solar radiation plays a very important role in the
energy balance above the tropopause. The above absorption of solar radiation are due to
photoionization and photodissociation of the various gases present in the upper atmosphere by
the ultraviolet and X-ray radiation regions of the solar spectrum. The above absorption
processes of photoionization and photodissociation, account for less than 2% of the solar
radiation incident upon the earth’s atmosphere. Almost all of the incoming solar radiation
having wavelength up to 0.31 mm are absorbed in the upper atmosphere. Solar radiation
having wavelength up to 0.1 mm are absorbed at heights above 90 km by the gaseous
constituents such as molecular nitrogen N2, molecular oxygen O2, and atomic oxygen O. The
absorption of these wavelengths less than 0.1 mm results in the photoionization of N2, O2, and
O giving rise to the E and the F layers of the atmosphere at heights of 110 km and 160 km,
respectively.
Incoming solar radiation in the wavelength region (0.1 < l < 0.2 mm) is absorbed by
molecular oxygen giving rise to the photodissociation of molecular oxygen to atomic oxygen.
O2 + hn ® 2O (12.16)
At levels above 100 km, the atomic oxygen produced by the photodissociation process is a
major constituent of the atmosphere. This atomic oxygen is highly reactive, that is, it
recombines with molecular oxygen in the presence of a third molecule N to form ozone. The
probability of the above chemical reaction depends on the square of the density of the gas,
with the result; the atomic oxygen recombines quickly to produce ozone at lower levels
(stratosphere) where the density is higher.
O2 + O + N ® O3 + N (12.17)
Ultraviolet solar radiation with wavelengths larger than 0.2 mm is not absorbed strongly in the
photodissociation process and hence penetrates successfully lower down into the atmosphere
until it meets ozone. The above ultraviolet radiation having wavelengths greater than 0.2 mm
(and less than 0.31 mm) is rapidly absorbed by ozone to become molecular and atomic oxygen.
O3 + hn ® O2 + O (12.18)
The free oxygen atom then quickly recombines with molecular oxygen to form another ozone
molecule. When the reactions characterized by Eqs. (12.17) and (12.18) occur one after the
other sequentially, there corresponds no net chemical change, but only an absorption of
ultraviolet radiation and a resultant heat input. Due to the high absorptivity of ozone for
wavelengths less than 0.31 mm and the ease with which the dissociated ozone molecule
(represented as in Eq. (12.18)) are regenerated using Eq. (12.17) , even the small amounts of
334 u BASICS OF ATMOSPHERIC SCIENCE

ozone present in the upper atmosphere are capable of absorbing all the solar radiation in the
wavelength range 0.2 < l < 0.31 mm. In the atmosphere, the maximum absorption of solar
radiation in the wavelength range (0.2 < l < 0.31 mm) takes place around 50 km and this
corresponds to the large input of energy giving rise to a temperature maximum (known
as stratopause). Due to the density dependence of the chemical reaction (Eq. (12.17)), ozone
formation is more at lower levels and hence the maximum ozone concentrations occur at
around 25 km. The upper atmosphere is by and large transparent to solar radiation of
wavelengths larger than 0.35 mm.

12.2 INTERNAL, POTENTIAL AND KINETIC ENERGY


The forms of energy that are of at most importance in the earth–atmospheric system are the
internal energy, the potential energy and the kinetic energy. While the kinetic energy is
manifested in the movement of air and ocean currents, the potential energy is chiefly in the
form of gravitational potential energy. The gravitational potential energy is the potential
energy associated with the gravitational force. The internal energy of the atmosphere is the
energy stored in atoms and molecules of what constitutes the atmosphere. The following
subsections discuss briefly the above-mentioned various forms of energy.

12.2.1 Internal and Potential Energy


In this section, the energy of the atmosphere from a general point of view will be considered.
The total energy of the atmosphere is the sum of internal energy, gravitational potential
energy and the kinetic energy. It turns out that for the case of a hydrostatic atmosphere, a
relationship exists between the internal and the gravitational potential energy and so it makes
sense to combine the above two energy and call it as the total potential energy. The
relationship between the internal and the gravitational potential energy for a hydrostatic
atmosphere can be shown by considering a column of air of unit cross-sectional area
extending from the surface to the top of the atmosphere.
The internal energy per unit horizontal cross-sectional area of an air column of thickness
dz is
dEI = r cV Tdz (12.19)
Integrating Eq. (12.19) from the surface to the top of the atmosphere and assuming cV to be
constant yields the internal energy for the entire column and is given as
‡
& *
D7 Ô S5 E[ (12.20)


EQ
Using the hydrostatic equation, S E[  in Eq. (12.20) one gets
H

D
&*  7 Ô 5 EQ (12.21)
H Q

GLOBAL ENERGY BALANCE u 335

where p0 is the surface pressure. The gravitational potential energy per unit horizontal area of
an air column of thickness dz at a height z is
dEP = rgzdz (12.22)
Integrating Eq. (12.22) over the entire atmospheric column yields the gravitational potential
energy for the entire column and is given as
‡
&1 Ô S H[ E[ (12.23)


Using the hydrostatic equation, Eq. (12.23) becomes



&1  Ô [ EQ (12.24)
Q

Integrating Eq. (12.24) by parts and using the ideal gas law, one gets
‡ ‡
&1 Ô Q E[ 3TQ Ô S5 E[ (12.25)
 

where Rsp is the specific gas constant. Comparing Eqs. (12.20) and (12.25), one gets

3TQ
&1 & (12.26)
D7 *
The physical meaning of Eq. (12.26) is very much evident. Let the temperatures of a column
of air increase. This increases the internal energy EI and the column of air must expand
vertically. The above raises the centre of mass of the air column and consequently increases
the gravitational potential energy as well.

12.2.2 Kinetic Energy


The kinetic energy per unit horizontal area of an air column of thickness dz is
dEK = 0.5 r (u2 + v2)dz (12.27)
where the variables have their usual meaning. Integrating Eq. (12.27) over the entire
atmospheric column yields the total kinetic energy of the entire atmospheric column and is
given as
‡

&, Ô S V  W
E[
 
(12.28)



Using the hydrostatic equation, Eq. (12.28) becomes




&,  Ô V

 W 
EQ (12.29)
 H Q
336 u BASICS OF ATMOSPHERIC SCIENCE

12.3 CONVERSION OF POTENTIAL AND INTERNAL ENERGIES


TO KINETIC ENERGY
The conversion of potential energy to kinetic energy in the atmosphere at low latitudes is
associated with the mean meridional Hadley circulation. However, the mean meridional
circulations are not important at higher latitudes and the conversion of potential energy to
kinetic energy in the atmosphere at higher latitudes is associated with atmospheric
disturbances. When warm air moves poleward, the above air is cooled by conduction, and
undergoes sinking motion and resultant lowering of its centre of gravity which causes a
conversion to kinetic energy. The following subsections briefly discuss the conversion of
potential energy and internal energy to kinetic energy.

12.3.1 Available Potential Energy


The following question naturally arises. Given any state of the atmosphere, whether a suitable
redistribution of atmospheric mass can lead to a state of a lower total potential energy, thereby
allowing for energy to be released, which can be converted to kinetic energy of motion? The
answer for the above question is in the affirmative. However, it turns out that only a small
fraction of the total potential energy is available for conversion to kinetic energy. Lorenz
demonstrated that an adiabatic redistribution of the atmospheric mass leads to a state of
minimum total potential energy called as the reference state. The adiabatic redistribution of
mass ensures that the mass of air above any given constant potential temperature surface
(isentropic surface) will not change. The difference in the total potential energy between the
actual and the reference states can be determined and this difference is known as available
potential energy. The available potential energy represents the maximum amount of potential
energy that may be available for release to be converted to kinetic energy. To introduce the
concept of available potential energy, consider a model atmosphere made up of two equal
masses of dry air separated by vertical partitions as seen in Figure 12.3. Let the two air
masses be at uniform potential temperature q1 and q2, with q1 < q2, and let the pressure in the
ground level on either side of the partition be equal to say, one atmospheric pressure, i.e. say,
1000 hPa. It is of interest to investigate the maximum amount of kinetic energy that can be
obtained by an adiabatic redistribution of mass within the same volume on removal of the partition.
Since the process of redistribution of mass is adiabatic, one can invoke the conservation
of total energy
EP + EI + EK = constant (12.30)
Assuming that the air masses are initially at rest and letting the final state be denoted by
primed quantities
E¢P + E¢I + E¢K = EP + EI (12.31)
From Eq. (12.26), total potential energy is
D1
&1  & &
(12.32)
*
D
7
*
GLOBAL ENERGY BALANCE u 337

FIGURE 12.3 Schematic representation of a model atmosphere made up of two equal masses of
dry air separated by vertical partitions

Using Eq. (12.32) in Eq. (12.31), one gets the kinetic energy realized by the removal of the
partition as
D1
&,„ &*  &*„
(12.33)
D7
Since the potential temperature is conserved for an adiabatic process, the two air masses
cannot mix after the partition is removed. It is clear from Eq. (12.33), that for a given initial
internal energy, the maximum kinetic energy realized by removal of partition will correspond
to the situation where the final internal energy is minimal. The final internal energy is
minimal when the air masses are rearranged such that the air with lower potential temperature
q1 lies completely beneath the air with higher potential temperature q2. Designating the
D1
minimum E¢I by E²I, it is clear that a certain amount of the total potential energy & „„ is not
D
7
*

available for conversion to kinetic energy since no adiabatic process can reduce E²I still
further. To state the above in an alternate way, the adiabatic redistribution of mass leading to
the final state lowers the centre of mass of the system allowing for a certain realization of the
kinetic energy. The centre of mass of the system in the final state cannot be at the ground
level and hence only a small part of the total potential energy is available for conversion to
kinetic energy.
The available potential energy is defined as the difference between the total potential
energy of a closed system and the minimum total potential energy that could be realized from
an adiabatic rearrangement of mass. Hence, for the illustrated example given in Figure 12.3,
the available potential energy denoted by P is given as
338 u BASICS OF ATMOSPHERIC SCIENCE

D1
1 &  & „„
(12.34)
D7
* *

Essentially, Eq. (12.34) indicates that the available potential energy is equivalent to the
maximum kinetic energy obtained by an adiabatic process. It can be shown that the available
potential energy of the atmosphere is approximately given by the volume integral of the
variance of potential temperature on an isobaric surface taken over the entire atmosphere and
is given by
1 — 7  Ô R „ R 
E7 (12.35)

where R is the average potential temperature for a given pressure surface, q ¢ is the local
deviation of potential temperature from the average, V is the total volume and 1 is the
average available potential energy per unit volume. For the entire atmosphere as a whole,
observations indicate that

1 ,
  –      (12.36)
D1 1
&
D7
*

where , is the total kinetic energy of the atmosphere. Equation (12.36) clearly shows that
only about 0.5% of the total potential energy of the atmosphere is available and only 10% of
the available portion is actually converted to kinetic energy.

12.4 GENERATION AND FRICTIONAL DISSIPATION OF


KINETIC ENERGY
The generation of kinetic energy in the atmosphere from internal energy and gravitational
potential energy, the latter is made available through differential heating and the resultant
lowering of the centre of gravity. The dissipation of kinetic energy to heat is through frictional
dissipation. Both the generation and frictional dissipation of kinetic energy are briefly
discussed in the following subsections.

12.4.1 Generation of Kinetic Energy


It is generally accepted that the kinetic energy is produced in the atmosphere from conversion
of total potential energy, the latter is made available through differential heating. In this
section a simple model is considered to investigate the generation of kinetic energy. Further,
in this model only the mechanical generation of kinetic energy is considered, while the
thermal effects are not included. Consider an air mass defined by vertical sides extending
from the ground to the top of the atmosphere. The horizontal momentum equations in
component form are then given as
˜V ˜V ˜V ˜V  ˜Q
V W X = GW   'Y (12.37)
˜U ˜Y ˜Z ˜[ S ˜Y
GLOBAL ENERGY BALANCE u 339

˜W ˜W ˜W ˜W  ˜Q
V W X =  GV   'Z (12.38)
˜U ˜Y ˜Z ˜[ S ˜Z
where Fx and Fy are the components of the frictional force per unit mass in x and y directions.
Multiplying Eq. (12.37) by u and Eq. (12.38) by v and adding both equations, one gets an
equation for the kinetic energy of horizontal motion, as given below

˜ È D Ø ˜ È D Ø ˜ È D Ø ˜ È D Ø È ˜Q ˜Q Ø
S  S V  S W  S X  ÉV  W Ù  SV'Y  S W'Z
˜U ÉÊ  ÙÚ ˜Y ÉÊ  ÙÚ ˜Z ÉÊ  ÙÚ ˜[ ÉÊ  ÙÚ Ê ˜Y ˜Z Ú
(12.39)
where c2 = u2 + v2. Using the equation of continuity in mass divergence form
˜S ˜ SV ˜S W ˜S X
V W X  (12.40)
˜U ˜Y ˜Z ˜[
Equation (12.39) may be written as

˜, ˜V, ˜W, ˜X, È ˜QV ˜QW Ø È ˜V ˜W Ø


   É  Ù  QÉ   SV'Y  S W'Z
(12.41)
˜U ˜Y ˜Z ˜[ Ê ˜Y ˜Z Ú Ê ˜Y ˜Z ÚÙ

SD 
where K is the horizontal kinetic energy per unit volume  The last term in the right-

hand side of Eq. (12.41), expressed as ‘– b’ represents the rate at which K is dissipating due
to viscosity. Equation (12.41) when integrated over the entire volume, becomes

˜ Ë ˜V, ˜W, ˜X, Û Ë ˜QV ˜QW Û


, EY EZ E[ Ü EY EZ E[  ÔÔÔ Ì Ü EY EZ E[
˜U ÔÔÔ
 ÔÔÔ Ì   
Í ˜Y ˜Z ˜[ Ý Í ˜Y ˜Z Ý

Ë ˜V ˜W Û
 ÔÔÔ Q Ì  Ü EY EZ E[  ÔÔÔ C EY EZ E[ (12.42)
Í ˜Y ˜Z Ý
Utilizing the Gauss divergence theorem, in its three-dimensional and two-dimensional forms,
the first and second terms in the right-hand side of Eq. (12.42) can be rewritten as

˜ Ë ˜ V ˜W Û
 ÔÔ ,7O ET  ÔÔ Q WEY  VEZ
EY  ÔÔÔ Q Ì  Ü EY EZ E[
˜U ÔÔÔ
, EY EZ E[
Í ˜Y ˜Z Ý
 ÔÔÔ C EY EZ E[ (12.43)
The first term in the right-hand side of Eq. (12.43) represents the energy advected across the
vertical walls, while the second term in the right-hand side represents the effect of work done
by pressure forces on the boundary. The third and fourth terms in the right-hand side of
Eq. (12.43) refer to the primary source of generation of kinetic energy within the volume and
the dissipation rate of the kinetic energy for the entire volume due to frictional forces.
Equation (12.43) can be simplified further by either considering the above simple model
to be a mechanically closed system or by considering the volume to be made up of the entire
340 u BASICS OF ATMOSPHERIC SCIENCE

atmosphere. Either way, no energy can be advected into the volume and no pressure work can
be done on the boundary, resulting in the vanishing of the first and second terms in the right-
hand side of Eq. (12.43), which then becomes
˜ È ˜V ˜W Ø
˜U ÔÔÔ , EY EZ E[ ÔÔÔ Q ÉÊ ˜Y  ˜Z ÙÚ EY EZ E[  ÔÔÔ C EY EZ E[ (12.44)

The first term in the right-hand side of Eq. (12.44) representing the production is balanced by
the dissipation (second term in the right-hand side of (Eq. 12.44)) when the whole atmosphere
is taken into account over a long-term scale.
The production term may be envisaged as the integral vertical sum of contributions from
all horizontal layers as follows:
Ë È ˜V ˜W Ø Û
Ô ÌÍ ÔÔ Q ÉÊ ˜Y  ˜Z ÙÚ EY EZ E[ ÜÝ
However, for a mechanical closed system (no flow across lateral boundaries) or for the
actual atmosphere (no lateral boundaries), the surface integral of the horizontal divergence
must vanish in each of the horizontal layers. Hence, the production term will vanish unless
the pressure p assumes large values where horizontal divergence is present and the pressure p
assumes small values where horizontal convergence is present. Thus, from the production
term it is clear that kinetic energy is generated/lost in regions of horizontal divergence/
convergence. This may appear inconsistent with our every day experience that low-pressure
regions are the centres of weather activity. The inconsistency can be explained as follows.
Since the net divergence is zero in a mechanically-closed system, a divergent region must be
balanced by another convergent region. If we confine our attention to a high-pressure
divergent region, the kinetic energy produced within a divergent high-pressure region is
rapidly transported by the work done due to pressure forces at the boundary and also advected
by the outward directed winds associated with the high pressure. The large kinetic energy
observed over low-pressure areas is due to the convergent transport of energy into the centre
of the low pressure together with the work done by the pressure forces on the winds
converging towards the centre of the low.

12.4.2 Frictional Dissipation of Kinetic Energy


It is well recognized that the kinetic energy of the atmosphere is lost (dissipated) through
frictional effects. It is generally assumed that this frictional loss or dissipation occurs as a
result of the final end product of a turbulent cascade of energy from the large scales to the
small scales, the molecular viscosity acting at the latter small scales. Essentially, this implies
that molecular viscosity converts the energy at the small scales to heat. The above turbulent
cascade is expected to occur in the atmospheric turbulent boundary layer or in patches of
turbulence in the free atmosphere.
A simple model is invoked to estimate the rate at which frictional and turbulent effects
contribute to the dissipation of the kinetic energy of the atmosphere. For simplicity, the
following assumptions are made:
(i) Only horizontal non-accelerating motions considered
(ii) Balance of forces prevail at all levels
GLOBAL ENERGY BALANCE u 341

(iii) The isobars are oriented in the east-west direction at all levels leading to
˜Q

˜Y
(iv) Above the atmospheric boundary layer, geostrophic balance holds,
 ˜Q ˜Q
 GVH  WH  with ug being assumed constant, and
S ˜Z ˜Y
(v) Within the atmospheric boundary layer, the Ekman layer balance holds.
Utilizing the above assumptions, Eq. (12.39), the energy equation becomes
0 = –f v ug – (uFx + vFy) (12.45)
Above the atmospheric boundary layer, v = 0, and also the frictional forces are small
and may be ignored ensuring that Eq. (12.45) is identically satisfied. Within the atmospheric
boundary layer, the rate of frictional dissipation of energy (the last term in brackets in the
right-hand side of Eq. (12.45), balances the other term which represents the work done by the
pressure field associated with the cross isobaric flow. Hence, the pressure field ensures that
the energy is generated at the same rate to compensate for the frictional loss of energy. The
total rate of change of the kinetic energy per unit horizontal area due to friction can be
written as
% %
E& ,
EU Ô S V'Y  W'Z
E[  G Ô SWVH E[ (12.46)
 

where D represents the Ekman layer depth corresponding to the height at which the wind first
becomes parallel to the isobars. Within the Ekman layer (atmospheric boundary layer), the
density of the air can be considered constant. Using the Ekman layer solution for v,
v = ug e–az (12.47)
where a = p D–1
Equation (12.46) becomes
%
E& ,  B[
 G SVH ÔF TJO B[ E[ (12.48)
EU 
Using D = pa–1 and integrating the right-hand side of Eq. (12.48) results in

E& , G SVH % Q
   F
(12.49)
EU Q

Using r = 1.29 kg m–3 and D = 1 km,


E&,
  GVH (12.50)
EU
The total kinetic energy of the whole atmospheric column needs to be calculated and
compared with the rate of dissipation of kinetic energy. In the atmospheric boundary layer, the
kinetic energy is given by
342 u BASICS OF ATMOSPHERIC SCIENCE

%
V  W 
& L SÔ E[ (12.51)


Using the Ekman layer solution for u,
u = ug(1 – e–az cos az), and
v as given in Eq. (12.47) in Eq. (12.51), one gets (12.52)

SVH %  B[
Ek1 = Ô   F DPT B[  F  B[
E[ (12.53)


or
S %VH Ë   F Q
 Û 
Ek1 = Ì  Ü  V H (12.54)
 Í Q Ý
Above the atmospheric boundary layer, since vg = 0, the kinetic energy is
‡ VH VH Q%
Ek2 = Ô S E[ Ô EQ (12.55)
%
 H 
or
Q% VH
Ek2 =  VH (12.56)
H
where pD = 899 hPa is the pressure at a height of 1 km in the U.S. Standard atmosphere.
The total kinetic energy is then
EK = Ek1 + Ek2 = 5120 u2g (12.57)
Hence, the fractional rate of dissipation is given as

 E& ,  G  –  


    –   T  (12.58)
&, EU  
where the Coriolis parameter is assumed to have a value of 10–4 s–1. If there were no
mechanisms for the production of the kinetic energy, the entire atmosphere would lose energy
at the rate of 36% in a day. The above simple model does not allow for any turbulent loss of
kinetic energy above the atmospheric boundary layer. Further, the observed ug is not constant,
but increases with height.
Since the solar energy is the source of all the atmospheric circulations, it would be
worthwhile to compare the above estimated dissipation rate to the rate of absorption of solar
energy by the earth–atmospheric system. The average rate of absorption of the solar energy as
4
given in Eq. (12.15) is   "
 Using the value of S = 1380 W m–2 and albedo A = 0.35,

the above average rate of absorption of the solar energy is 224 J m–2 s–1. Hence, the above
simple analysis shows that the average rate of dissipation of kinetic energy is about 1% of the
rate of absorption of solar energy.
GLOBAL ENERGY BALANCE u 343

12.5 ATMOSPHERE AS A HEAT ENGINE


Heat engines essentially convert heat energy into work by taking in heat at a higher
temperature source and delivering it to a lower temperature sink. The source regions of energy
for the atmosphere are regions where the temperatures are higher and the sink regions are of
lower temperature. In this sense, regions at higher latitudes and at higher elevations are the
energy sinks, while the regions of lower latitudes and at lower heights are the energy sources.
It is known that the motions in the atmosphere that transport energy are poleward as well as
upward. The work done by the atmospheric heat engine maintains the kinetic energy of the
overall general circulation against the frictional dissipation of energy.
In Chapter 3, the expression of thermal efficiency of a heat engine (say, a Carnot
engine) was given as
2
  

2 5
Here, Q1 is the heat absorbed at the higher temperature T, while Q2 is the heat rejected
at the lower temperature T¢. In the case of the atmosphere, since the temperature difference
between T and T¢ is not much, the efficiency of the atmosphere as a heat engine is quite low.
An appropriate quantitative estimate of the efficiency of the atmospheric heat engine can be
got by determining the globally-averaged rate of generation or the dissipation of kinetic
energy and dividing it by the rate of absorption of solar radiation per unit area averaged over
the entire globe. This estimate was made in the earlier section of this chapter, which yielded a
value of 1% for this ratio. The above clearly indicates that to compensate and balance the
turbulent (frictional) dissipation of kinetic energy, not more than 1% of the absorbed solar
radiation is required. The above estimate clearly reveals that the atmospheric heat engine
operates at a low level of efficiency.
While the entire atmosphere can indeed be regarded as a heat engine, albeit a low-
efficiency engine, there are certain small-localized regions, which behave in a different
manner. These regions actually behave in a manner exactly opposite to a heat engine, in a
thermodynamical sense, i.e. these regions accept heat at low temperatures and deliver heat at
higher temperatures. These regions which act as refrigerators are dominated by thermally
indirect circulation which converts the kinetic energy transported from the other regions to
potential energy. The equatorial tropopause and the mesopause over the summer pole are two
of the regions which function as refrigerators. The thermally indirect local circulations
prevailing over these regions are maintained by the available potential energy produced by
diabatic heating gradients in other regions of the atmosphere.

REVIEW QUESTIONS

1. Do radiative processes alone determine the global energy balance of the earth–atmospheric
system?
2. What is available potential energy?
3. What are the mechanisms through which kinetic energy is generated from available potential
energy?
344 u BASICS OF ATMOSPHERIC SCIENCE

4. Mention the reason as to why the atmosphere is considered a heat engine?


5. Is the atmosphere an efficient heat engine?
6. In a hydrostatic atmosphere, how are the internal energy and the gravitational potential
energy related?
7. Does the atmosphere, in some regions, play the role of a refrigerator?
8. Do the aerosols affect the energy balance?
9. In a heat engine, what happens to the heat, which is not converted to useful work?
10. What causes the dissipation of kinetic energy?
11. Why only a part of the potential energy is available for conversion to kinetic energy?
12. Mention the thermal efficiency of the atmospheric heat engine.
13. Mention the processes which govern the energy balance in the atmosphere.
14. In the absence of greenhouse gases, what would be the average surface temperature of the
earth?
15. What determines the albedo of the earth–atmospheric system?
16. Mention the non-radiative processes which ensure energy balance for the earth–atmospheric
system.
17. What is internal energy?
13 General Circulation

The general circulation of the atmosphere can be envisaged to denote the totality of all the
atmospheric motions that characterize global scale atmospheric flow. In other words, the
general circulation is concerned with the time-averaged structure of atmospheric fields such as
winds, temperature, humidity and precipitation. The above-mentioned flow fields are averaged
over a time scale which is sufficiently long to exclude the effects of the individual weather
events and at the same time small enough to retain monthly and seasonal variations.
One of the earliest models proposed to describe the general circulation was suggested by
Hadley, in which the longitudinal perturbations of the flow were neglected resulting in a
single meridional cell. The existence of surface westerlies over the mid-latitude region led to
a refinement of the single meridional cell model by the three-cell meridional circulation
model. It is to be noted that the three-cell meridional circulation model is also required from
considerations of the angular momentum conservation. Further studies have revealed the
existence of large-scale eddies in the mid-latitude region which play a very important role in
the maintenance of the zonal mean circulation.
In addition to investigating the budget of momentum, it is also important to study the
overall balance of the angular momentum for the combined earth–atmosphere system. Such a
budget of angular momentum provides information of the transport of angular momentum in
the atmosphere together with the transfer of angular momentum between the earth and the
atmosphere. Laboratory experiments on a rotating differentially heated fluid have provided
important insight into the overall features of the general circulation of the atmosphere.
Section 13.1 presents the overall general considerations of the general circulation of the
atmosphere. While Section 13.2 introduces the meridional circulation models of the single cell
and the three-cell variety, the next section, Section 13.3 discusses the angular momentum
balance of the earth–atmosphere system. The last section, Section 13.4 introduces the
laboratory dishpan experiments and highlights their importance in revealing insights into the
general circulation of the atmosphere.

345
346 u BASICS OF ATMOSPHERIC SCIENCE

13.1 GENERAL CONSIDERATION


The general circulation of the atmosphere comprises the global patterns of air movements,
from the meridional circulation of ascending air over the equatorial region and descent over
the subtropical regions as well as the cyclones in the extratropics which transport energy and
moisture through the mid-latitudes. One can interpret the general circulation of the
atmosphere as the average air flow around the globe. The earth is considered to be in
radiative equilibrium on a global scale with the incoming energy equaling to the outgoing
energy. However, if one considers a particular latitude, it is clear that the radiative equilibrium
does not exist and the role of the general circulation of the atmosphere is to transport heat
poleward. The following subsections deal briefly on the overview and other facets of the
general circulation of the atmosphere.

13.1.1 Overview
It is known that the earth’s atmosphere exhibits motion in various scales ranging from the
planetary scale of the order of several thousands of kilometres to the microscale turbulent
motion of the order of a few centimetres. The scales lower to the planetary scales in the
hierarchy with respect to the horizontal scales are the synoptic and the mesoscale motions.
Synoptic scales extend to a few thousands of kilometres, while the mesoscale motion is
characterized by motions, which lie between a few hundreds to a few tens of kilometres.
While the monsoon depressions and the extratropical cyclones characterize the synoptic
scale, systems such as tropical cyclones, land and sea breezes constitute some examples of
meso-beta scale motion. Meso-alpha scales typically have horizontal scale ranging from
200 km to 2000 km while meso-beta scales vary from 20 km to 200 km Scales of horizontal
extent ranging from 1 km to 20 km are known as meso-gamma scale. Thunderstorms,
tornadoes and dust devils are examples of meso-gamma scale motions. Still lower in the
hierarchy of scales are motions associated with macroscopic turbulence and finally to the
smallest scales of motion which are characterized by molecular transport of mass, heat and
momentum. Table 13.1 provides the various different scales of the atmospheric motion.
TABLE 13.1 Different scales of the atmospheric motion

Name of scale Length scale(m) Time scale(s) Examples of systems


Planetary scale > 106 106 Long waves in the westerlies, sea-
sonal circulation, planetary monsoon
Synoptic scale (also referred 105 to 106 105 Mid-latitude, highs and lows, major
as meso-alpha scale) storms
Mesoscale (also referred as 104 to 105 104 Hurricanes and tropical storms, land
meso-beta scale) and sea breeze, coastal fronts
Convective (also referred as 103 to 104 103 Thunderstorm, tornadoes, dust devils
meso-gamma scale)
Turbulent 10–2 to 103 101 Sea surface fluctuation, wind stress
and wave formation
Microscale 10–7 to 10–2 10–1 Molecular transport of mass, heat
and momentum
GENERAL CIRCULATION u 347

It is important to note that each of the above-mentioned scales do influence and affect
the other scales. Further, the largest scales are invariably the slowest, while the smallest scales
are associated with relatively rapid motion. If one were to average the whole atmosphere over
a long-time period as well as average over large-spatial scales, the averaged in space and time
structure of the atmospheric motion, i.e. the general circulation involves primarily the largest
and the slowest scales. Due to the mutual interaction and the dependence of the various
scales, the general circulation cannot be completely indifferent to the influence of the smaller
scales. However, as a first step in the study of the general circulation of the atmosphere, it is
convenient to invoke a simplified picture of the global atmospheric motion by retaining only
the largest and the slowest scales and totally neglecting the faster and the smaller scales.
Despite not retaining the faster and the smaller scales, the above-simplified picture does
provide important information of the various facets of the general circulation of the atmosphere.
Due to reasons of convenience and simplicity, the first observational and modelling
studies on the general circulation undertook to examine the zonally-averaged circulation.
However, the longitudinal dependence on the general circulation cannot be ignored since the
forcings caused by orography as well as land–sea contrasts are asymmetric in the longitudinal
direction. Furthermore, longitudinal dependence is important as far as the general circulation
of the atmosphere is concerned due to the existence of important circulations such as the east-
west oriented Walker circulation. It turns out that the longitudinal dependent part of the
general circulation may be classified into quasi-stationary motions, which do not change with
time, and monsoonal circulations, which reverse seasonally and other subseasonal and
interannual components. It is imperative to understand not only the zonally-averaged
circulation, but also the longitudinal dependent part of the general circulation for a thorough
and complete description of the general circulation of the atmosphere.

13.1.2 Observed Meridional Cross-section of Longitudinally-


averaged Zonal Wind and Temperature
It is well known that the incoming solar radiation at the top of the atmosphere is essentially
independent of longitude. It is to be noted that the solar radiation averaged over seasonal
time-periods is independent of longitude. If the above is not true, then the diurnal variation is
the strongest variation one is likely to observe. Since the latitudinal distribution of the above
radiation is responsible for driving the atmospheric motions, it would be appropriate to
investigate the time-averaged general circulation, which is longitudinally-independent.
Figures 13.1 and 13.2 provide the meridional cross-section of the observed global distribution
of the longitudinally-averaged zonal wind and temperature for the northern winter (December,
January and February) and the northern summer (June, July and August) months, respectively.
The behaviour of the global distribution of the longitudinally-averaged zonal wind and
temperatures are shown for the entire troposphere in Figures 13.1 and 13.2, respectively. The
stronger zonal wind vis-à-vis meridional winds is due to geostrophic effect as stronger
meridional gradient of temperature leads to stronger meriodional pressure gradients and hence
higher zonal winds. The zonally-averaged meridional temperature gradient in northern
hemisphere is clearly stronger in the winter season as compared to the summer season.
However, due to the large fractional cover by the oceans in the southern hemisphere and their
associated large heat capacities, the differences between the summer and the winter
348 u BASICS OF ATMOSPHERIC SCIENCE

100
10
200 10
30 20 15
300 55
35
Pressure (hPa)

400 10 45
10 40 15
30 5
500 35
25 30 10
5
600
20 25 20
10
700 5
800
10
900 15 15
10
1000
90°S 75°S 60°S 45°S 30°S 15°S EQ 15°N 30°N 45°N 60°N 75°N 90°N

100 200
210
200 220 215 220
225
225 230
300 240 235
245 220
250
230
Pressure (hPa)

400 255
260
265 230
500 270 235
600 275 240
280 245
700
285
800 250
290
900
295
1000
90°S 75°S 60°S 45°S 30°S 15°S EQ 15°N 30°N 45°N 60°N 75°N 90°N
Latitude
FIGURE 13.1 Meridional cross-section of longitudinally-averaged zonal wind (top panel in unit of
m s–1) and temperature (bottom panel in unit of K) for December, January and
February.

meridional temperature gradients are much less marked in the southern hemisphere. Also, due
to the thermal wind relationship, the maximum zonal wind in the northern hemisphere is
higher during the winter season as compared to the summer season. Furthermore, in both the
seasons and over both the hemispheres, a core of maximum westerly zonal winds, called jet
streams are found just below the tropopause at subtropical latitudes (i.e. between 30°N/S and
35°N/S) during winter and slightly higher latitudes (between 40°N/S and 45°N/S) during the
summer season.
The surface winds are primarily easterlies at low latitudes and near the polar regions
over both the hemispheres and in both the winter and the summer seasons. However, the
surface winds are westerlies over mid-latitude regions in both the hemispheres and over both
the seasons. Essentially, the nature of the zonally-averaged flow in each hemisphere during
each of the seasons can be characterized as having three cells, consisting of a shallow layer of
polar easterlies and a deep layer of mid-latitude westerlies as well as deep layers of low-
latitude easterlies.
GENERAL CIRCULATION u 349
100
15
20
200 20
25 15 15 15
300
50 10 25
25
Pressure (hPa)

400 45
40 10
35
500 35
a
30 30
600 25
25 20
700 15
800 20 15 10
10
900 10
1000
90°S 75°S 60°S 45°S 30°S 15°S EQ 15°N 30°N 45°N 60°N 75°N 90°N

100
210
200 220 215 225
225
230 230
205 235
245
300 245 240 235
250 250 240
255
Pressure (hPa)

400 240 255 260 245


260 265 250
500 270 255
275 260
600 230 265
265 280 270
700 285
275
800 280
900 290
1000
90°S 75°S 60°S 45°S 30°S 15°S EQ 15°N 30°N 45°N 60°N 75°N 90°N
Latitude
FIGURE 13.2 Meridional cross-section of longitudinally-averaged zonal wind (top panel in unit of
m s–1) and temperature (bottom panel in unit of K) for June, July and August.

13.1.3 Longitudinally-dependent Flow


Examining the time-averaged surface winds during the summer and winter months, provide us
additional evidence of the longitudinally-varying flow. Large divergent anticyclones
characterize the surface wind distribution during both the summer and the winter seasons over
the oceans associated with subtropical highs as well as prominent equatorial convergent
zones. Furthermore, during the northern hemisphere winter season, a very strong divergent
anticyclonic circulation is seen over Siberia and convergent cyclonic circulation is observed
near 60°N in both the Pacific (Aleutian low), and the Atlantic (Icelandic low) oceans. Also,
the seasonal monsoonal circulation over Asia is another example of longitudinally-varying flow.
The longitudinally-varying flow would develop due to the presence of forcings due to
large-scale orography, and to forcings due to the uneven distribution of continents and oceans
and their associated heating contrasts. The above-mentioned time-averaged, longitudinally-
asymmetric large-scale flow known as stationary waves, is very prominent in the northern
hemisphere during the winter season. If one were to examine a typical time-averaged upper
tropospheric (200 hPa) zonal wind component for northern winter (December, January and
February) in a latitude–longitude cross-section, it is immediately apparent that over some
350 u BASICS OF ATMOSPHERIC SCIENCE

regions (longitudes) there are marked departures of the time-averaged zonal flow from the
zonally-averaged circulation pattern. For example, over 30°N at 200 hPa there are regions
having a very prominent minimum in the time-averaged zonal wind speed in the eastern
Pacific and the eastern Atlantic, while over the same latitudes and heights, very strong
maxima in the time-averaged zonal wind speed are seen east of the Asian and the North
American continents, i.e. over the western Pacific and the western Atlantic oceans. The
above-mentioned strong subtropical jets east of the Asian and North American continents are
manifestation of a stationary wave pattern, apparently forced primarily by orography, i.e. by
the Himalayas and Rockies. The above subtropical jets also appear to be influenced by the
continent–ocean heating contrasts. However, there is some lack of agreement among the
researchers regarding the relative importance of heating (continental–oceanic contrasts) and
orography in forcing the observed stationary wave pattern.

13.1.4 Requirement on Theories of General Circulation


A careful examination of the long time-averaged global distribution of the meteorological
variables shows that the global atmospheric circulation is subject to certain important
balances. The most important balance is the global energy balance outlined in Chapter 12.
Without any adequate mechanism to maintain the global energy balance through latitudinal
and upward transport of heat, any general circulation model will fail miserably in its efforts to
provide a credible explanation of the observed behaviour of the general circulation.
Another important requirement that must be met by a general circulation model is to
ensure that the conservation of angular momentum holds. This is because the earth’s angular
velocity is observed to have no important variation in time. An additional requirement for a
general circulation model is that it should provide for a long-term steady distribution of
atmospheric mass. This requirement is necessary since there are no observed appreciable
secular variations in pressure. Lastly, any general circulation model must ensure that the long-
term global distribution of water vapour must be constant in time, discounting of any effects
happening in a climate change world. It is to be noted that a single transport mechanism may
not be able to completely explain the observed distribution of mass, angular momentum and
water vapour. That is, a transport mechanism meant to explain the distribution, of say, angular
momentum cannot be used to explain the distribution of another variable. In short, no single
transfer mechanism may be adequate to satisfy all the four requirements mentioned above.
Quantitative modelling of the general circulation of the atmosphere is based on solving
the most general (primitive) equations of atmospheric motion in spherical coordinates through
advanced numerical models. The chief objective of the above quantitative modelling efforts is
to simulate faithfully the features of the general circulation. Furthermore, one also requires
these models to accurately predict the consequence of any changes, in say, the concentration
of greenhouse gases on the global climate. The current generation of global models can
provide fairly accurate simulation of the present climate and also can provide for reasonable
prediction of the response of the earth’s climate system to changes in the external parameters.
However, uncertainties in the representation of the physical processes involving clouds and
precipitation restrict the ability of these global models in making accurate quantitative climate
change predictions.
GENERAL CIRCULATION u 351

13.2 MERIDIONAL CIRCULATION MODEL—HADLEY


CIRCULATION
The earliest theoretical explanation proposed on the nature of the general circulation was by
George Hadley in 1735. Hadley proposed that the general circulation is a thermally-direct,
zonally-symmetric overturning in which the heated air at equatorial latitudes rises and flows
poleward where it cools, sinks and flows towards the equator. The above picture of a simple,
thermally-direct, zonally-symmetric overturning was modified by Coriolis to explain the
presence of trade winds. Due to the Coriolis force, the poleward moving air at upper levels is
deflected to the east, while the equatorial moving air gets deflected to the west. The above-
mentioned single, zonally-symmetric, thermally-direct circulation modified by the earth’s
rotation called the Hadley circulation does explain the presence of the trade winds
(northeasterlies in the northern hemisphere and southeasterlies in the southern hemisphere) in
the surface flow.
The effect of the Coriolis force, due to the earth’s rotation can also be deduced from
consideration of the conservation of the angular momentum. In the absence of torques, a
zonal ring of air moving poleward will conserve the angular momentum, i.e.

È V Ø  
ÉÊ Z  B DPT G ÙÚ B DPT G DPOTUBOU (13.1)

where a is the radius of the earth, w is the angular velocity of the earth,ÿ f the latitude and
u is the relative zonal wind speed which manifests due to conservation of angular momentum.
From Eq. (13.1),
  
V <DPOTUBOU  Z B DPT G > (13.2)
B DPT K
It is clear from Eq. (13.2) that for a zonal
ring of air moving poleward, u will
increase, while for a zonal ring of air
moving equatorward u will decrease. This
essentially leads to a strong westerly
component in the upper levels and an
easterly zonal wind component in the
lower levels as shown in Figure 13.3. The Equator
simple picture shown in Figure 13.3 is
inconsistent with the observed zonal wind
structures at the lower levels over the
mid-latitudes where zonal westerlies
prevail throughout the year. Furthermore,
assuming that the above simple picture of
Figure 13.3 is a correct one, it can be
readily shown that it is not a plausible FIGURE 13.3 Schematic diagram of a single
description of the nature of the general zonally-symmetric, thermally-direct
circulation. Since the earth rotates from circulation modified by rotation.
352 u BASICS OF ATMOSPHERIC SCIENCE

west to east, the presence of easterlies near the surface over the whole earth would take away
the westerly angular momentum from the earth, ultimately slowing the angular velocity of the
earth’s rotation. Since observations indicate that the earth’s angular velocity is a constant and
is not subject to any marked time variations, it is clear that the surface winds cannot be
entirely easterlies (slowing the angular velocity of the earth) or entirely westerlies (enhancing
the angular velocity of the earth). This fits in with the observed surface winds distribution,
which is characterized by easterlies at low latitudes and polar regions and westerlies at mid-latitudes.
Theoretical studies suggest that a zonal symmetrical hemispherical Hadley-type
circulation would be unstable and break down outside the tropics. A simple extension of the
single cell Hadley circulation to a tricell (Figure 13.4) meridional circulation was proposed
with the objective of explaining the surface winds over the entire earth. The three cells which
make up the tricell structure are named as Hadley cell, Ferrell cell and the Polar cell
respectively. The northward moving equatorial air at upper levels cools at a rate of 3°C to
2°C per day. This causes the upper air current to cool by the time it reaches the subtropical
latitudes, say 30°, leading to sinking at these subtropical latitudes. The sinking of the air at
30° latitude leads to adiabatic warming and the sinking air spreads out horizontally at the
ground. A part of the sinking air associated with the Hadley circulation returns equatorward to
complete the thermally-direct Hadley circulation, restricting essentially the Hadley circulation
to the low-latitude tropical regions only. The equatorward directed air according to Eq. (13.2)
would cause the relative zonal wind to decrease leading to surface easterlies in the tropical
regions. This explains the surface trade winds over the low-latitude tropical regions.

FIGURE 13.4 Schematic diagram of a tricell meridional circulation model.

The poleward branch of the subsiding air current over the subtropics, according to
Eq. (13.2), would contribute to an increase in the zonal wind, leading to surface westerlies
over the mid-latitude regions. The presence of surface westerlies over the mid-latitudes and
surface easterlies over the tropics as suggested by the tricell model will ensure that the net
flux of the angular momentum between the earth–atmospheric systems remains zero.
GENERAL CIRCULATION u 353

The polar regions are characterized by a net deficit of radiant energy and hence near the
poles there is a net loss of heat by the atmosphere. This is manifested near the polar regions
as cold air subsiding and spreading towards the lower latitudes. Such spreading equatorward
winds at low levels over the high latitudes give rise to surface easterlies according to
Eq. (13.2). The polar front gets established when the cold low-level easterly air meets the
warmer westerly air of the mid-latitudes. The tricell model is fully determined by the
completion of the thermally-direct circulation of the polar cell and the completion of the
indirect circulation, called as Ferrell cell over the mid-latitude region. The equatorial and the
polar cells of the tricell model are thermally direct circulations and hence energy producing
while the mid-latitude cell (Ferrell cell) is a thermally-indirect circulation (air sinking over the
warmer temperature regions and rising over the cooler temperature regions) and hence is
energy consuming one.
The above tricell model has explained the presence of the observed surface tropical and
polar easterlies as well as the surface westerlies over the mid-latitudes. The tricell model also
accounts for the polar front and the presence of upper-level westerlies in the equatorial and
the polar regions. However, the tricell model fails to explain the presence of the observed
upper-level westerlies over the mid-latitude regions since the equatorward directed air at upper
levels over the mid-latitude region in the Ferrell cell, according to Eq. (13.2) leads to
easterlies.
It was mentioned earlier in this section that the zonally-symmetric hemispherical Hadley
circulation extending over the entire hemisphere would breakdown outside the tropical
regions. The above breakdown results in the development of baroclinic eddies which transport
heat poleward. Baroclinic eddies; in simple terms refer to the circular patterns that one gets to
see on weather maps—the cyclones (lows) and anticyclones (highs).  It is well known that
baroclinic instability is the principal mechanism through which the cyclone–anticyclone
weather systems develop in the earth’s atmosphere. Baroclinic instability also occurs in the
oceans and is responsible for the most energetic mesoscale eddies, which transport heat and
other tracers across the oceans. Such baroclinic eddies evolve slowly and are typically close
to being in hydrostatic and geostrophic balance. These baroclinic eddies will intensify until
this poleward heat transport together with the heat transported by the ocean currents and the
planetary waves are adequate to compensate for the radiation deficit in the polar regions. This
ensures that the meridional temperature gradient between the equator and the polar regions
does not increase. Furthermore, these baroclinic eddies convert potential energy to kinetic
energy and assist in maintaining the kinetic energy of the atmosphere against frictional
dissipation of energy. To sum up, the heat is transported to higher latitudes (poleward)
through the direct thermal circulation of the Hadley-type as well as through the development
of the above-mentioned baroclinic eddies.

13.3 ANGULAR MOMENTUM BALANCE


Considerations of the conservation of angular momentum suggest that in the absence of
significant external torques about the earth’s rotation axis, the angular momentum of the
earth–atmospheric system will be conserved, i.e. it will not change with time. Since the earth
is in mechanical contact with its atmosphere, it is quite likely that at the earth–atmosphere
354 u BASICS OF ATMOSPHERIC SCIENCE

interface, a transfer of angular momentum from one to the other occurs through frictional
interaction. If there were net flow of angular momentum either to the earth or to the
atmosphere over a period of time, this would result in a change in the angular velocity of the
earth. However, since the observations clearly indicate that the angular velocity of the rotating
earth is constant and is not subject to any time variations; one can infer that the total
atmospheric angular momentum remains constant.
Due to the observed surface easterly winds over the tropics and polar regions, it follows
that there is a transfer of angular momentum from the earth to the atmosphere due to
frictional interactions. In order to conserve the angular momentum, it is necessary that surface
westerlies must be present in other regions, say in mid-latitudes which will transfer angular
momentum back to the earth, ensuring that the total atmospheric angular momentum remains
a constant. Since, however, the mean surface westerlies and easterlies are maintained and do
not die away with time, it is important that any good model or theory of the general
circulation should be in a position to explain satisfactorily the transport of atmospheric
angular momentum from the easterly belt to the mid-latitudes or the westerly belts.
In the low-latitude regions, the poleward momentum transport are due to both the axially-
symmetric Hadley circulation as well as the transport by eddies. However, over mid-latitudes
it is the eddy motion that transports the momentum poleward. The maximum poleward flux of
angular momentum during the winter season occurs near 30° latitude while the maximum
horizontal convergence of angular momentum occurs at about 45° latitude. In the discussion,
which follows, the transport and balance of angular momentum is provided. The absolute
angular momentum per unit mass of air is
m = (w a cos f + u) a cos f = w a2 cos2 f + ua cos f (13.3)
where the variables have their usual meaning. The first term in the right-hand side represents
the angular momentum due to earth’s rotation, while the second term represents the angular
momentum due to zonal air motion relative to the earth.
In the absence of torques, the angular momentum of the atmosphere is constant. The
only torques, i.e. tangential forces that possess a moment about the rotation axis, are those
due to zonal pressure gradients and those due to friction. Hence, the rate of change of angular
momentum in the presence of torques due to zonal pressure gradients and frictional forces are
given by
EN È  ˜Q Ø
EU ÉÊ  S ˜Y  'Y ÙÚ B DPT G (13.4)

The rate of change of angular momentum per unit mass of the atmosphere multiplied by the
density is given by
EN ˜ ˜ ˜ ˜
S = ˜U S N
 ˜Y SVN
 ˜Z S WN
 ˜[ S XN

EU

N ËÌ ˜S  ˜ SV
 ˜ SW
 ˜ SX
ÛÜ (13.5)
Í ˜U ˜Y ˜Z ˜[ Ý
GENERAL CIRCULATION u 355

However, from the equation of continuity, the term in the square bracket in Eq. (13.5)
EN
vanishes identically. Replacing from Eq. (13.5) in Eq. (13.4), one gets an equation for
EU
the local rate of change of the absolute angular momentum per unit volume and is written as

˜ ˘ ˜ ˜ Û ˜Q
S N
 Ì SVN
 S WN
 S XN
Ü  B DPT G  S 'Y B DPT G (13.6)
˜U Í ˜ Y ˜ Z ˜ [ Ý ˜Y
If Eq. (13.6) is integrated over the entire volume of the atmosphere poleward of a certain
latitude fo, one gets after replacing the term in the square bracket in Eq. (13.6) by Gauss
divergence theorem to get
˜ ˜Q
S N E7  Ô S N7O ET  Ô B DPT G E7  Ô S 'Y B DPT G E7
˜U Ô
(13.7)
˜Y
where Vn is the outward directed normal to the surface bounding the volume. Since the entire
volume is bounded by the earth surface, the top of the atmosphere and a vertical surface of
latitude fo, the contribution to the first integral in the right-hand side of Eq. (13.7) must be
zero since the normal component of velocity vanishes at a solid boundary. Due to the
negligible value of the density, contribution of the first term at the top of the atmosphere is
also zero and hence the only contribution from this term arises from the velocity component
perpendicular to the vertical surface at latitude fo. Hence, the first term in the right-hand side
of Eq. (13.7) is responsible for the transport of angular momentum per unit volume out of the
volume considered and hence this term is called as the meridional transport term. The second
term on the right-hand side of Eq. (13.7) represents the torque due to asymmetrical pressure
distribution. Such east-west asymmetric pressure distributions are present across mountain
barriers and hence this term is referred as mountain–pressure torque term. The third term is
referred to as friction–torque term. Since the frictional force is proportional to the vertical
gradient of the zonal shear stress, the integration of the last integral in Eq. (13.7) about the
vertical will reduce it to a surface integral involving the zonal shear stress. The above-
mentioned surface integral is to be determined only at the lower boundary of the atmosphere
since only here is the zonal shear stress non-vanishing. Since the direction of the zonal shear
stress is opposite in direction to the wind direction, the frictional torque term will be negative/
positive for westerlies/easterlies.
Over the mid-latitude regions, the mountains extract angular momentum from the
atmosphere through mountain pressure torque, while the earth surface extracts angular
momentum from the atmosphere through the frictional torque term. The mountain pressure
torque acts to transform angular momentum from the atmosphere to the ground, if the surface
pressure and the slope of the ground are positively correlated. Observations indicate that this
is true in mid-latitudes since the surface pressure is higher on the western side of the
mountain as compared to the eastern side. In the mid-latitudes of northern hemisphere, the
mountain torque term accounts for 50% of the total atmosphere–surface mountain exchange,
while in the tropics and southern hemisphere the exchange is dominated by turbulent eddy
stresses. To ensure that the surface westerlies over the mid-latitudes are maintained, the above
356 u BASICS OF ATMOSPHERIC SCIENCE

loss of angular momentum of the atmosphere must be balanced by the meridional transport
term. Since the contribution from the meridional transport term arises only due to the velocity
component perpendicular to the vertical surface at latitude fo, the negative of the outward
directed normal to the surface bounding the above volume is nothing but the northward
directed meridional velocity component in northern hemisphere. Hence, the meridional
transport term can be written as

ÔÔ SNW EY E[
where v is the northward directed meridional velocity component in the northern hemisphere.
Hence, the meridional term can be expanded over x and z by
‡ Q ‡ Q
ÔÔ SNWB DPT G E M E[ B DPT G Ô Ô V  Z B DPT G
SW E M E[ (13.8)
   

where dx = a cos fÿ dl is utilized and dl is the increment in the longitude. Using the hydro-
EQ
static equation SE[  in Eq. (13.8) and extracting the average value of the variables
H
over all longitudes around a latitude circle, the meridional transport term becomes
Q
Q B DPT G
H Ô VW  Z WB DPT G
EQ (13.9)


where p0 is the surface pressure. Using VW V W  V „W „ where a prime indicates a departure,


and has values when averaged in Eq. (13.9), one gets

Q
Q B DPT G
H Ô V W  V „W „  Z BW DPT G
EQ (13.10)


The first three terms in the integral of Eq. (13.10) are called the drift term, the eddy
momentum flux term and the w-momentum flux term. In the drift term, W represents the slow
non-geostrophic meridional drift flow and this term can transport angular momentum of the
mean zonal current if a mean meridional flow layer exists. The eddy momentum flux term can
transport the zonal eddy momentum by meridional eddies, while the w-momentum flux term
can transport the angular momentum of the earth’s rotation by the mean meridional drift
velocity W . While the drift term is important over the tropical region, it is much smaller and
can be neglected over the mid-latitude region. Furthermore, the w-momentum flux term has
no contribution to the vertically-integrated flux of angular momentum.
Over the mid-latitude region in the northern hemisphere, the eddy momentum flux term
is positive and decreases with the increase in latitude. Furthermore, over the mid-latitude
region, the eddy momentum flux term is the most dominant of the meridional transport term.
The positive eddy momentum flux over the mid-latitude region requires that eddies be
asymmetric in the horizontal plane with the resulting troughs and ridges tilted. The tilting of
GENERAL CIRCULATION u 357

the trough and ridges in the upper levels has to assume a pattern such that there is poleward
transport of angular momentum. Essentially, for the poleward transport of angular momentum
through the eddy momentum flux term, V „W „ has to be positive. This requires that strong
zonal flow (u¢ > 0) must occur where strong poleward flow (v¢ > 0) is already present.
Alternatively, this requirement is that there is less than the average zonal flow (u¢ < 0) where
the meridional flow is equatorward (v¢ < 0). In either of the above two cases, the eddy
momentum flux term V „W „ is positive, ensuring poleward transport of the angular momentum.
The above two cases can occur only if the troughs and ridges are oriented in a southwest-to-
northeast phase tilt as shown in Figure 13.5.

ge
Rid
h
ug
Tro

x
FIGURE 13.5 Schematic diagram illustrating the streamlines for a positive eddy momentum flux.

If one were to obtain the flux form of the angular momentum equation, one finds that
this equation depends on the convergence of the horizontal flux of angular momentum and the
convergence of the vertical flux of the angular momentum. The latter is made up of the flux
owing to the large-scale motion, the flux due to the pressure torques as well as the flux due to
the small-scale turbulent stresses. Above the atmospheric boundary layer the w-momentum
flux is the main contribution to the vertical momentum transport in the troposphere. However,
the flux due to the pressure torque and the flux due to the small-scale turbulent stresses are
responsible for the transfer of momentum from the earth to the atmosphere in the low
latitudes and from the atmosphere to the earth in the mid-latitudes.
Several earlier studies have estimated the torques due to the turbulent transfer from the
surface, the torques due to the large-scale topography and the latitudinal variation of the
torque exerted by the earth on the atmosphere and is shown in Figure 13.6(a). The flux of
angular momentum directed northward and required to balance the above estimated total
surface torques is shown as the dashed curve in Figure 13.6(b). Despite the uncertainties in
the observed wind measurements, the observed transport of angular momentum shown as the
continuous curve in Figure 13.6(b), is in very good agreement with the required northward
transport of angular momentum, as can be seen from Figure 13.6(b). It is to be noted that
except for the equatorial region within 10° of the equator almost all of the northward flux is
contributed by the eddy flux term, underlining the importance of the northward transport by
the eddies.
358 u BASICS OF ATMOSPHERIC SCIENCE

Average eastward torque

(¥ 10 kg s )
per unit area
–2

0
5

–5
90° 60° 30° 0° 30° 60° 90°
North South

FIGURE 13.6(a) Averaged eastward torque per unit horizontal area exerted on the
atmosphere by surface friction (solid line) and by mountain (dashed line).
Transport of angular momentum

50
kg m2 s )
–2

0
18
(¥ 10

–50
90° 60° 30° 0° 30° 60° 90°
North South
FIGURE 13.6(b) Observed transport of angular momentum (solid line) and required
transport from surface observations (dashed line).

13.4 DISHPAN EXPERIMENTS


Laboratory experiments on a rotating annulus have also been utilized by experimentalists to
provide important insights into the features of the general circulation of the atmosphere. Since
the mathematical models using the b plane have successfully simulated the essential features
of the general circulation without considering explicitly the earth’s sphericity, it may be
adequate to utilize rotating differentially heated fluids to demonstrate the essential features of
the general circulation of the atmosphere.
The dishpan experiments refer to a group of experiments which utilize a cylindrical
vessel (the earliest of these experiments used a dishpan), that is rotated about its vertical axis.
The fluid in the vessel represents the atmosphere over one hemisphere with the rim of the
cylindrical (dishpan) vessel resembling the equator, while the centre corresponds to the pole.
In these experiments, the outer rim is heated and the centre cooled to represent differential
heating. The whole set-up is then placed on a rotating turntable and rotated at a constant
angular velocity w about its axis.
GENERAL CIRCULATION u 359

For a certain combination of rotation speed (slow rotation) and heating rates, the flow in
the cylindrical (dishpan) vessel manifests the axially-symmetric circulation with a steady
azimuthal flow superposed on a direct meridional circulation with rising motion near the rim
and sinking near the centre. Since the above direct meridional circulation resembles the
Hadley circulation, it is known as the Hadley regime. This Hadley regime has no flow
variation in the azimuthal direction. At faster rates of rotation, the above flow changes from
the symmetric pattern to form a number of wave-like patterns that start drifting in the
azimuthal direction. The number of such waves found in the pattern depends on the rotation
speed. In addition to the wave-like structure, a pattern of meandering zonal jets is also seen.
Tracers on the surface of the fluid in the dishpan reveal pattern quite similar to those on
upper air mid-latitude westerly charts, while the tracers on the lower boundary of the dishpan
reveal fluid structures similar to fronts seen in the extra tropics. This flow pattern is called the
Rossby regime. At certain rotation rates, the amplitude of the waves or the number of waves
may pulsate slowly and this pulsation is known as vacillation. At much higher rotation rates,
the wave-like flow suffers a breakdown and changes to a disordered turbulent state, somewhat
similar to conditions, which prevail during disturbed atmospheric flows. Hence, it is clear that
the laboratory dishpan experiments do reveal the general features of the general circulation
and can provide important insights into the dynamics of the general circulation.

REVIEW QUESTIONS

1. What is “General Circulation of the Atmosphere”?


2. What are the single cell and the three-cell models of the general circulation of the atmosphere?
3. Why are the trade winds north easterlies and south easterlies and not simply northerlies/
southerlies?
4. What is Hadley cell? Where is it seen?
5. What are Ferrell and Rossby cells and where are they found?
6. What is Walker circulation?
7. Why the Hadley circulation cannot extend right up to the poles?
8. Why cannot we have surface easterlies or surface westerlies over the entire globe?
9. What is meant by thermally-direct circulation?
10. What is a baroclinic eddy?
11. What is the role of baroclinic eddies in transporting energy from the lower to the higher
latitudes?
12. State the principle of conservation of angular momentum?
13. Can the conservation of angular momentum principle hold for a single-cell model?
14. Do the oceans play an important role in transporting energy from the lower to the higher
latitudes?
15. What are the broad requirements expected from a general theory of the general circulation
of the atmosphere?
16. What is hydrologic cycle?
17. What are stationary waves?
14 Numerical Modelling
of the Atmosphere

In recent years, due to the economic importance of having accurate weather predictions, a large
amount of effort has gone into studying the complex nature of the weather. Such an
investigation into the complex nature of weather is best accomplished by developing
atmospheric models, mostly mathematical in nature and solving those using numerical methods.
Modern numerical weather prediction (NWP) deals with the development and the ultimate
solution of these mathematical models. The techniques of NWP require the accurate knowledge
of the initial state of the atmosphere. Initialization and data assimilation procedures are
necessary to prepare an accurate initial state of the atmosphere. Numerical methods are used
to solve the governing equations of atmospheric motion. The governing atmospheric equations
are expressed as a system of partial differential equations (pde) and these pde’s can be
broadly classified into three categories: elliptic, parabolic and hyperbolic types. Further, the
various numerical methods to solve the atmospheric equations are three in number and they
are the method of finite difference, the spectral method and the finite element method. The
science of accurate weather prediction is a highly specialized research area, which has grown
extensively and is still evolving. With the increase in the computational resources,
atmospheric scientists are addressing scientific issues related to climate prediction as well.
Section 14.1 briefly presents the general considerations related to the numerical
modelling of the atmosphere and Section 14.2 outlines the various facets of NWP. Section 14.3
introduces the important procedures of initialization and data assimilation and Section 14.4
outlines the spectral and finite element methods. This chapter concludes with Section 14.5
which outlines the various challenges in weather and climate forecasts.

14.1 GENERAL CONSIDERATIONS


The dynamical and thermodynamical behaviour of the atmosphere and the oceans are best
represented by a system of partial differential equations. Due to the inherent nonlinear nature
of these partial differential equations, it is almost impossible to solve these systems of
360
NUMERICAL MODELLING OF THE ATMOSPHERE u 361

equations in closed form and one has to resort to numerical methods to solve these systems
of equations. The study of the application of numerical methods in solving the governing
equations of atmospheric motion is briefly discussed in the following subsections.

14.1.1 Overview
Atmospheric scientists, unlike experimental laboratory scientists, cannot perform controlled
laboratory experiments in the large-scale atmosphere. To underline the difference between
atmospheric sciences and other sciences, let us imagine a phenomenon (physical, chemical or
biological) is observed. To explain the above-observed phenomenon, scientists put forward
hypothesis and test the above hypothesis by performing very careful controlled experiments in
the laboratory. If the results of the controlled laboratory experiments are at variance with the
hypothesis, the latter is suitably revised so that it is consistent with the results of the
controlled laboratory experiment. Such a “standard scientific method” available to
experimental laboratory scientists is unfortunately not available to atmospheric scientists due
to the inability to perform controlled laboratory experiments in the large-scale atmosphere.
The atmospheric scientists after collecting a large amount of data, which correspond to an
observed atmospheric phenomenon, take recourse to developing models. These models take
into account the most important and significant atmospheric processes, necessary for
providing the plausible explanation of the observed atmospheric phenomenon. These
atmospheric models are usually formulated as a system of partial differential equations. The
above equations are solved numerically under various general conditions and these solutions
are interpreted in terms of the observed atmospheric behaviour. When the results of the
numerical solution of these atmospheric models are at variance with the observed atmospheric
behaviour, the models are revised by incorporating additional significant atmospheric
processes. Hence, the “models” play the same role in atmospheric sciences as played by the
laboratory-controlled experiments in the other sciences. In this chapter, we shall confine our
discussions only to the mathematical models of the atmosphere.
The mathematical models can be broadly classified into three types: simple, intermediate
and complex. Simple models are those which can be solved without the help of a computer.
Also, simple models involve a very small number of physical variables and atmospheric
processes and are described by mathematical equations, which have a simple form. The
intermediate models have a small number of physical variables and atmospheric processes and
require a computer to numerically solve the resulting mathematical equations. Both the simple
and intermediate models provide very valuable information of the underlying atmospheric
processes responsible for an observed atmospheric phenomenon. However, the simple and
intermediate models do not provide for very accurate simulation of the observed atmospheric
phenomenon. Only the complex models, known as general circulation models (GCMs), can
provide for very accurate simulation of the observed atmospheric behaviour. However, the
complex models have a large number of physical variables and incorporate a very large
number of atmospheric processes in them. All the three models are necessary and complement
one another. For example, assume that the complex model is unable to provide a reasonably
accurate simulation of the observed atmospheric behaviour. In such situations, an intermediate
model may provide valuable information on the reasons for the complex model’s poor
performance. Also, the simple models provide the necessary physical interpretation of the
results of the intermediate model.
362 u BASICS OF ATMOSPHERIC SCIENCE

14.1.2 The Finite Difference Method


As mentioned earlier, the intermediate and complex models involve a system of partial
differential equations, whose solution requires a numerical method to integrate the equations
as well as a computer. The governing equations of atmospheric motion belong to a general
class of systems known as initial value problems (IVPs). IVPs are basically propagation
problems where the future behaviour of the system is to be predicted based on certain initial
conditions. One of the well-known numerical methods of solving ordinary or partial
differential equations is the finite difference method. In the finite difference method, the
derivatives (ordinary or partial) are replaced by differences and hence the differential equation
becomes a difference equation.
For better appreciation of the finite difference method, imagine a meteorological variable
u(x). Let u(x) be a solution of some differential equation in the interval 0 £ x £ L. If the
above interval is divided into I subintervals of length h, then u(x) can be approximated by a
set of I + 1 values as wi = u (ih), where i = 0, 1, … I. Here, wi are the values of the
meteorological variable at the I + 1 grid points defined by x = ih, i = 0, 1, 2…, I and, where
h = L/I. If the chosen h is sufficiently small compared to the scale on which the
meteorological variable u varies, the I + 1 grid point values provide a reasonably good
approximation to u(x) and its derivatives.
The grid point meteorological variable wi can also be used to construct a finite
difference approximation to any differential equation. This is accomplished by using the
Taylor series expansion to represent the first- and second-order derivatives of u. For a
sufficiently small h, the Taylor series expansion about any point x0 up to the third order can
be written as
EV I E V I E V
u(x0 + h) = V Y
 I  
 
 0 I
(14.1)
EY Y Y  EY
Y Y
 EY
Y Y

EV I E  V I E  V
u(x0 – h) = V Y
 I  
 
 0 I 
(14.2)
EY Y Y  EY
Y Y
 EY
Y Y

where O(h4) indicates that further terms are of magnitude h4 or lower and they will be
neglected. Subtracting Eq. (14.2) from Eq. (14.1) gives the expression for the central finite
difference form for the first derivative

EV V Y  I
 V Y  I

 0 I
(14.3)
EY Y Y I

Adding Eq. (14.1) and Eq. (14.2) gives the expression for the central finite difference form
for the second derivative

E V V Y  I
 V Y
 V Y  I

 0 I
(14.4)
EY  Y Y
I
NUMERICAL MODELLING OF THE ATMOSPHERE u 363

If h is exceedingly small and one disregards the terms with order of magnitude of h2 and
higher order, Eq. (14.1) provides the expression for the forward difference form for the first
derivative
EV V Y  I
 V Y

 0 I
(14.5)
EY Y Y I


In a similar way, if one retains only the linear term in the Taylor series expansion in
Eq. (14.2), one obtains the expression for the backward difference form for the first derivative

EV V Y
 V Y  I

 0 I
(14.6)
EY Y Y I
Consider, for example, the x component of the momentum equation after disregarding the
friction forces
˜V ˜V ˜V ˜V  ˜Q
V W X  GW  (14.7)
˜U ˜Y ˜Z ˜[ S ˜Y
By retaining only up to the linear term in the Taylor series expansion in time, the term in the
left-hand side of Eq. (14.7) can be expressed as
˜ V Y Z [ U

V Y Z [ U  L
V Y Z [ U
 L  0 L 
(14.8)
˜U
As in Eq. (14.5), for exceedingly small k, Eq. (14.8) becomes
˜ V Y Z [ U
V Y Z [ U  L
 V Y Z [ U

 0 L
(14.9)
˜U L
Assume that all the quantities on the right-hand side of Eq. (14.7) are known at some time t.
˜V ˜V
This means that is known at time t. Substituting the values of u and at time t in
˜U ˜U
Eq. (14.8) provides for the values of u at time t + k with an error of the order of k2.
Effectively, the variable u is obtained at a later time t + k, i.e. clearly, a time marching by a
time step k is successfully performed. A repetition of the above procedure results in the
prediction of the meteorological variable u at all later times.
The right-hand side of Eq. (14.7) does involve spatial derivatives with respect to x, y
and z. The above spatial derivatives are to be replaced by differences using the finite
difference approximation. To realize the above, the atmosphere is envisaged to be represented
by a three-dimensional grid with grid spacing in the x, y and z (eastward, northward and
vertical) directions, respectively.

14.1.3 Partial Differential Equations


The governing equations which describe the motions of atmosphere form a set of coupled
nonlinear partial differential equations subject to boundary and initial conditions. Due to the
complex nature of the above governing equations, these equations are usually solved by
employing numerical methods. The following subsections provide a brief description of the
364 u BASICS OF ATMOSPHERIC SCIENCE

classification of partial differential equations into parabolic, elliptic and hyperbolic equations
and briefly outline the method of finite difference to solve simpler forms of the above
equations.

Boundary and Initial Value Problems


Before discussing partial differential equations, we need to define the initial and boundary
value problems (BVPs) mathematically. BVPs require an unknown function u, to be
determined in a domain D. Let the boundary of D be denoted by GD. Let u be a function of
x and y and let the differential equation within D be given by
L(u) = f (x, y) (14.10)
subject to the boundary condition, B(u) = g(x, y) on GD (14.11)
where L and B are differential operators and u, f and g are functions of the same independent
variables. In an IVP, it is required to predict the behaviour of an unknown function u in a
domain D for all future times based on the knowledge of u at some initial time t = 0. This
problem could also have boundary conditions on GD, the boundary of D. Let u be a function
of x, y and t and let the differential equation within D be given by
L(u) = f (x, y, t) (14.12)
The initial condition on u at time t = 0 is given as
u(t = 0, x, y) = h(x, y) on D and GD (14.13)
while the boundary condition on u is given as
B(u) = g(x, y, t) on GD (14.14)
Here again, L and B are differential operators, u, f and g are functions of x, y and t, while
h is a function of x and y only.

Classification of Partial Differential Equations


For a second-order linear partial differential equations (PDEs) with two independent variables
x and y and a dependent variable u, the form is

˜V ˜ V ˜ V ˜V ˜V
B  C  D  E  F  GV  (14.15)
˜Y  ˜Y˜Z ˜Z  ˜Y ˜Z
where a, b, c, d, e and f are functions of x and y only, the classification of the above PDE
depends on the values of the expression b2 – ac. The above PDE is hyperbolic if b2 – ac > 0,
is elliptic if b2 – ac < 0 and is parabolic if b2 – ac = 0. While elliptic equations are typically
boundary value problems, hyperbolic and parabolic equations belong to the initial value
problem category. Examples of well-known elliptic equations are the Poisson equation,
while that of the hyperbolic and parabolic equations are the wave equation and the diffusion
equation.
NUMERICAL MODELLING OF THE ATMOSPHERE u 365

Parabolic Equation (Diffusion Equation)


The one-dimensional parabolic (diffusion) equation can be written as

˜V Y U
˜  V Y U

,  … Y … M U !  , !  (14.16)
˜U ˜Y 
subject to the conditions
u(0, t) = 0, u(l, t) = 0, t > 0, and (14.17)
u(x, 0) = f (x), 0 £ x £ l (14.18)
where u is the concentration and K is the coefficient of diffusion.
We invoke the finite difference method to solve Eq. (14.16) subject to conditions,
Eqs. (14.17) and (14.18). For this, we define two mesh constants (mesh intervals) h and k

such that N M
is an integer (refer Figure 14.1).
I
The grid points in two dimension can then be defined as (xi, tj), where xi = i h for i =
0, 1,… m and tj = jk for j = 0, 1,…. Using the forward and central difference approximation
for the time and space derivatives, one gets for a grid point (xi, tj) not lying on the boundary
(i.e. i ¹ 0, i ¹ m, j ¹ 0) the following expressions:

V Y U 
 V Y U
V YJ  U K
 V YJ U K
 V YJ  U K

,
J K J K
 (14.19)
L I

tj+1

tj

Xi–1 Xi Xi+1 X

FIGURE 14.1 The mesh structure for solving the one-dimensional diffusion equation.

The above expression has a truncation error of the order O(k + h2). Rearranging Eq. (14.19),
one gets the expression for u at the next time by denoting the approximation of u(xi, tj) with
wij as follows:
366 u BASICS OF ATMOSPHERIC SCIENCE

Ë  ,L Û ,L <X
X  ÌÍ  I ÜÝ X   X > (14.20)
J K  J K
I  J  K J  K

The initial condition u(x, 0) = f (x) for 0 £ x £ l implies that wi,0 = f (xi) for 0 £ x £ l implies
that wi+1 = f (xi) for each i = 0, 1,… m. One can use the above values of wi,0 in Eq. (14.20) to
calculate wi,1 for i =1, 2,… m – 1. Also, one can use the boundary conditions u(0, t) = 0 and
u(l, t) = 0 as w0, j = wm,j = 0. For j = 1, the above boundary conditions become w0,1 = wm,1 = 0.
Using this and the initial conditions wi,0 = f(xi) for each i = 0, 1,…, m in Eq. (14.20) all the
values of wi,1 for all i can be found out. Repeating the above procedure, we can find wi,2, wi,3,
… wi,m–1, and so on. The above method is known as the explicit forward difference method,
since the unknown wi,j+1 is expressed explicitly in terms of known quantities wi,j and wi+1,j.

Computational Stability
In the finite difference method, the derivatives are replaced by differences and it is hoped that
in the limit of small values of the mesh constants (mesh intervals) the solution of the finite
difference equations will converge to the solution of the differential equation. However, the
above does not hold always in reality and the nature of the solution depends critically on the
computational stability of the difference equation. In this section, we will provide the stability
criteria for the one-dimensional diffusion equation.
Let D and N denote the exact solution of the difference equation and the numerical
solution obtained for a computer with finite accuracy for a one-dimensional diffusion
equation. Let the difference between N and D called the round-off error denoted by e. Since
the numerical solution N must identically satisfy Eq. (14.20) and since N = D + e,

Ë  ,L Û ,L < %
%  F  ÌÍ  I ÜÝ % F
  F  K % F  K > (14.21)
J K J K  J K J K
I  J  K J J  K J

Since D happens to be the exact solution of the difference equation, D will identically satisfy
Eq. (14.20) and hence, one finds that the error e also satisfies the difference equation

Ë  ,L Û ,L <F
FJ K 
 ÌÍ  I ÜÝ F
 J K

I   J  K
F J  K > (14.22)

The difference scheme is computationally stable if the errors remain bounded as the solution
progresses from a particular time t to the next time t + k. Essentially, this requires that
FJ K 
…

(14.23)
FJ K

The errors at any given time can be assumed to vary randomly with respect to x and can be
expressed by a Fourier series as
N 
BU FJLO Y
F Y U
Ç(14.24) F
O
where kn is the wave number. Also, the errors are assumed to grow exponentially with time.
Since the one-dimensional diffusion equation is a linear equation and the principle of
NUMERICAL MODELLING OF THE ATMOSPHERE u 367

superposition holds, the behaviour of each wave (each term) of the series in Eq. (14.24) will
exhibit the same behaviour as the series itself. For convenience, we will consider just one
term in the series of Eq. (14.24) to represent the error.
JLO Y
F N Y U
F BU F (14.25)

Substituting Eq. (14.25) in Eq. (14.20) and dividing by F BU F LO Y one gets J

F BL

,L <F JLO I
 F JLO I
 > (14.26)
I
F LO I  F  JLOI
J
LO I   DPT LO I

Using DPT LO I
BOE TJO  in Eq. (14.26), one gets
  

BL  ,L LO I
F  
TJO  (14.27)
I 
The left-hand side of Eq. (14.23) gives the amplification factor and is given by
F J K 
F (14.28)
BL

FJ K

From Eqs. (14.28), (14.27) and (14.23), one gets

FJ K   ,L L I
]F ]  TJO … (14.29)
 BL  O

FJ K I


BL 
For Eq. (14.29) to hold, 
… (14.30)
I 
Equation (14.30) gives the stability requirement for the solution of the forward difference
equation. It is clear that the mesh increment in time (time step) k has to be chosen in such a
manner that Eq. (14.30) is satisfied. If k is chosen above this critical value, the error will
grow without any bound and will completely swamp the solution. A more detailed discussion
of computational stability is presented in Appendix 6.

Elliptic Equations
Consider the two-dimensional Poisson equation defined on a domain D as given below:

˜  V Y Z
˜  V Y Z

³  V Y Z
 G Y Z
(14.31)
˜Y  ˜Z 
subject to boundary conditions.
u(x, y) = g(x, y) on GD
where D is defined as,
a < x < b, c < y < d (14.32)
368 u BASICS OF ATMOSPHERIC SCIENCE

GD is the boundary of domain D and assume that both f and g are continuous functions.
CB ED
Choose two mesh constants h and k such that I BOE L for some integers
O N
m and n (refer Figure 14.2). This divides the interval [a, b] into n equal parts of width h and
interval [c, d] into m equal parts of width k. The above describes a grid on the domain D
with grid points (xi, yj), where xi = a + ih for i = 0, 1,…n and yj = c + jk, where j = 0,
1,…m. Using the centered finite difference formula to approximate second derivatives,
Eq. (14.31) becomes
V Y  Z
 V Y Z
 V Y  Z
V YJ Z K 
 V YJ Z K
 V YJ Z K 


J K J K J K


I L
G YJ Z K
 0 I
 0 L 
(14.33)
for i =1, 2,…, n – 1 and j = 1, 2,…, m – 1 and the boundary conditions are:
u(x0, yj) = g(x0, yj); u(xn, yj) = g(xn, yj) for j = 0, 1,…, m
u(xi, y0) = g(xi, y0); u(xi, ym) = g(xi, ym) for i = 1, 2,…, n – 1 (14.34)
In central difference form the above Eq. (14.33) becomes by approximating u(xi, yj) with wi, j
as

ËÈ I Ø  Û È IØ

 Ì É Ù  Ü X  X   X 
 É Ù <X  X  >  I G Y Z
(14.35)
Ê Ú Ê LÚ
ÍÌ L
J K J  K J  K J K J K J K
ÜÝ
for i = 1, 2,…, n – 1 and j = 1, 2, …, m – 1, subject to boundary conditions:
w0,j = g(x0, yj); wn, j = g(xn, yj) for j = 0, 1,…, m
wi,0 = g(x0, y0); wi,m = g(xi, ym) for i = 1, 2, …, n – 1. (14.36)

FIGURE 14.2 The mesh structure for solving the two-dimensional Poisson equation.
NUMERICAL MODELLING OF THE ATMOSPHERE u 369

Using the boundary conditions Eq. (14.36) in the system given by Eq. (14.35), we get a
linear system of size (n – 1) (m – 1) by (n – 1) (m – 1) with wi, j being the unknowns at all
the interior mesh points. The above linear system can be solved by matrix methods or
through relaxation methods.

Hyperbolic Equations
Consider a one-dimensional wave equation
˜  V Y Z
˜  V Y Z

B   Y   U !  (14.37)
˜U  ˜Y 
with conditions
u(0, t) = u(l, t) = 0; u(x, 0) = f(x), and

˜V Y 

= g (x) (14.38)
˜U
Like the earlier set of parabolic and elliptic equations, a mesh of grid points are defined such
that xi = i h for i = 0, 1,…m and tj = j k for j = 0, 1,…(refer Figure 14.3). Again the central
finite difference formula are used to approximate the second derivative

FIGURE 14.3 The mesh structure for solving the one-dimensional wave equation.

V Y U 
 V Y U
 V Y U 
Ë V Y  U
 V Y U
 V Y  U
Û
 B Ì (14.39)
J K J K J K J K J K J K

  Ü 
L Í I Ý
for i = 1, 2,…, n and j = 1, 2,….The boundary conditions are:
u(x0, tj) = u(xn, tj) = 0 for j = 0, 1,….
V Y U
 V Y U

u(xi, t0) = G Y H Y J    ž O (14.40)


J J


GPS
L
J J
370 u BASICS OF ATMOSPHERIC SCIENCE

In central difference form the above Eq. (14.39) becomes by approximating u(xi, tj) with
wi,j as
BL
X    M 
X  M  X  X
X  XIFSF M (14.41)
J K  J K J  K J  K J K
I
for i = 1, 2…n and for j = 1, 2… subject to boundary conditions:
w0, j = wn,j = 0 for j = 0,1,…, m;
X X
X G Y
 J  J 
H Y
 GPS J   ž O  (14.42)
J  J
L J

Equation (14.41) shows that the unknown wi,j+1 is calculated explicitly from the knowledge of
wi–1,j, wi,j, wi+1,j, and wi,j–1 and a repetition of the above procedure will result in marching the
solution in the t direction.

14.2 MODERN NUMERICAL WEATHER PREDICTION


Numerical Weather Prediction is the science of predicting the weather, i.e. numerical weather
prediction predicts the future state of the atmosphere by integrating the governing equations
of atmospheric motions numerically from the knowledge of the present state of the
atmosphere. Numerical Weather Prediction has emerged and grown very rapidly in the last
two decades with improvements in the model physics, better numerical methods and
availability of powerful supercomputers. The following subsections provide brief description
of the overview and various components of numerical weather prediction.

14.2.1 Overview
The science of dynamic meteorology provides the underlying basis and the methodology for
modern numerical weather prediction (NWP). The chief objective of dynamical prediction is
to forecast the future state of the atmosphere from knowledge of the present state by
considering the weather prediction problem as an initial value problem. In addition to the
present state of the atmosphere, one also requires the knowledge of the governing equations
of atmospheric motion. These equations are expressed in mathematical form as a system of
partial differential equations and they embody the physical laws of conservation of
momentum, mass and energy. NWP concentrates on the following two problems: firstly to
diagnose the current state of the atmosphere and secondly to model numerically the evolution
of the atmosphere with time.
In NWP, the initial conditions are obtained by analyzing the present state of the
atmosphere. Meteorological observations collected through conventional and non-conventional
observation platforms are incorporated into the atmospheric model using data assimilation, to
provide the initial conditions for the numerical forecast. The atmospheric numerical models
are based on the laws of conservation of momentum, mass and energy. Many physical
processes in the atmosphere called sub-grid scale processes occur at a scale much too small
to be directly resolved by the atmospheric numerical model. The effects of these sub-grid
processes such as radiation, convection, boundary layer, etc. have to be parameterized in the
numerical models.
NUMERICAL MODELLING OF THE ATMOSPHERE u 371

14.2.2 Observations
Hundreds of thousands of atmospheric observations such as pressure, wind speed, wind
direction, temperature and humidity are taken every day throughout the globe. The nature of
the meteorological observation platforms taking the observations, however, vary considerably.
These comprise the in situ surface observations over land and ocean, (the latter from buoys
and ships), and the upper air observations over land using radiosondes/rawinsondes. Further,
the observation from wind profilers, aircraft, satellite and radars provide important
information over data-sparse regions which can supplement the conventional observations.
Each observational platform has its own attributes such as the sensor resolution, systematic
and statistical errors associated with the measurement as well as the location and time of the
observation. For effective use of all the above meteorological observations in a NWP model,
it is necessary to monitor the availability of these observations, adopt quality control
measures to check their quality and reliability and finally to process all these observations
into a form that can be readily used. Unfortunately, the above observations collected daily
through both conventional and non-conventional observational platforms are far from
adequate to determine the current state of the atmosphere satisfactorily. Further, there are
huge data gaps over the oceans. Hence, the NWP is typically an under-determined initial
value problem.

14.3 DATA ASSIMILATION


The numerical models to be employed in numerical weather prediction require the accurate
and complete specifications of the three-dimensional structure of the initial state of the
atmosphere. In addition to conventional surface and upper air observations, additional non-
conventional observations of the atmosphere from satellites, radars, wind profilers and other
remote sensing systems are also available. However, the amount of observations is still very
much lower compared to the number of degrees of freedom of the model. Furthermore, the
observations are also not regularly distributed in space and time. Hence, there is a need for an
efficient data assimilation technique to utilize these irregular observations to provide the
initial conditions for the numerical model. The following subsections provide an overview and
the various facets of data assimilation.

14.3.1 Overview
The meteorological observations collected are non-uniformly distributed in space and time.
Spatial interpolation methods are utilized to put the non-uniformly distributed observation
onto a regular latitude–longitude grid. The difficulty lies not in the spatial interpolation of
observation, but in the fact that the data available are not adequate to provide the initial
conditions, which characterize the current (initial) state of the atmosphere. To cite an example,
modern global models have a typical resolution of 1° ´ 1° and twenty levels in the vertical.
Although present day modern global models have much higher horizontal and vertical
resolutions, the above horizontal resolution is used to simply show that the numbers of grid
cells are much larger than observations. These models have at least four prognostic variables
(zonal and meridional components of wind, temperature and humidity) and a surface pressure
372 u BASICS OF ATMOSPHERIC SCIENCE

for each vertical column. Hence, the number of grid cells equal 360 * 180 * 20 = 1296000,
while the number of prognostic variables amount to 360 * 180 * 81 = 5248800. Even
allowing for a time window of ± 3 hours, the number of observations is only of the order of
10–100 thousand only. This clearly indicates that the number of observations is typically
about two orders of magnitude less than the number of degrees of freedom of the model.
In order to reduce the under-determination of the initial value problem, it is clear that
additional information (data) is necessary. This additional information is called background or
first-guess. In the initial stages of the development of NWP, climatology (i.e. long-term
average information) was used as the background. However, with the general improvements
of forecasts, a short-range forecast can be chosen as the background in the operational data
assimilation system. Such a procedure of combining background fields with observation
allows the information to be accumulated in time into the model state and to propagate to all
the variables of the model. This is the basic idea of data assimilation.
Talagrand has proposed the following definition of data assimilation:
Assimilation of meteorological or oceanographical observations can be described
as the process through which all the available information are used in order to
estimate as accurately as possible the state of the atmospheric or oceanic flow.
Traditionally, data assimilation is made up of two procedures: objective analysis of
observations and data assimilation.

14.3.2 Objective Analysis


Objective analysis is an important component of the data assimilation. In the objective
analysis all the meteorological observations, throughout the globe, which are collected at a
given time (say 00 UTC) from the irregularly-spaced conventional and non-conventional
observational platforms are checked for quality and accuracy. Further, in the objective
analysis, these meteorological observations are used to produce an analysis of the
meteorological fields over a regular latitude–longitude grid at standard pressure levels using a
suitable interpolation scheme. The earliest objective analysis schemes were based on empirical
approaches such as successive correction method (SCM) and nudging. Subsequently, objective
analysis schemes based on statistical estimation theory using both sequential and variational
approaches were employed. Also, multivariate statistical methods such as optimal
interpolation (OI), as well as variational methods such as three-dimensional variational (3D-Var),
four-dimensional variational (4D-Var) approaches are increasingly being used for objective
analysis.
In the SCM method, the background (or first-guess) fields serve as the first estimate.
The above estimate is successively corrected iteratively after providing for suitable weights. In
the nudging approach, a term is added to the prognostic equations that nudges the solution
towards the observation, the latter interpolated to the model grid. The nudging approach is
typically used to assimilate small-scale observations and is generally not used for large-scale
assimilation.
The statistical interpolation schemes not only combine the background information of
the atmosphere (first-guess) with the observations, but also utilize statistical information about
the errors in the first-guess and in the observations. In the statistical interpolation scheme, the
NUMERICAL MODELLING OF THE ATMOSPHERE u 373

analysis is obtained by adding to the background the observational innovations, the latter
weighted by the optimal weights. The innovations are just the difference between the
observations and the background (or first-guess). The optimal weights are just the background
error covariance multiplied by the inverse of the total error variance. The sum of the
background error covariance and the observational error covariance is known as the total
error covariance.
In the least square method, the analysis error is minimized by determining the optimal
weights using the least square approach. Again, in the OI method, the optimal weight matrix
is found that minimizes the analysis error covariance. In the variational approaches, a cost
function is defined that is proportional to the distance of the analysis with the background and
to the observations. The above cost function is then minimized to obtain the analysis. In the
4D-Var, the cost function includes within itself the distance to observations over a time
interval, known as the assimilation window. The difference between the OI and the 3D-Var
approaches is only in the method of solution. In OI method, the weights are calculated for
each grid cell, while in the 3D-Var, the minimization of the cost function is performed directly
to obtain the analysis.

14.3.3 Initialization
It may so happen that the objectively-analyzed data are inconsistent. The above inconsistency,
say, in the balance between the mass and the wind fields may appear as noise. Such noise
manifests as a spuriously large, gravity wave when the inconsistent objectively-analyzed data
are used as the initial condition in a numerical weather forecast. The above presence of the
spurious large gravity wave in the initial condition can degrade or spoil all chances of a
reasonable numerical weather forecast. The data initialization procedure ensures that the
gravity wave noise is under control.
The importance of data initialization can be readily appreciated from the following
example. Consider the synoptic scale motion in mid-latitudes. It is known that for the above
motions, the wind and the pressure fields are in approximately geostrophic balance. That is,
the acceleration following the motion, which is very small in magnitude, is simply the
difference between the two, near equal and large terms, representing the Coriolis and the
pressure gradient forces. While the geopotential field can be determined from observations
fairly accurately, the wind fields have an observational error of 10% to 20%. The above error
in the observed winds will contribute to imbalances between the wind and the pressure fields
undermining the accuracy of the objective analysis. The initial imbalance between the mass
and the wind fields will lead to spurious accelerations. These accelerations will manifest as
large amplitude gravity wave oscillations as the wind flow tends to adjust from the initial
unbalanced state to a quasi-geostrophic balance state. However, the observed synoptic scale
motions do not include such large amplitude gravity waves. Hence, the presence of such
gravity waves in the fields can be considered as undesirable “noise”. Unless the above gravity
wave noise is controlled, this will lead to incorrect initial velocity and initial pressure
tendencies, which will ultimately degrade the forecast. The initialization process will modify
the objectively-analyzed data to ensure that the gravity wave noise is minimized.
One possible approach to overcome the noise problem is to disregard the observed
winds and obtain a derived wind field from the observed geopotential fields using the
374 u BASICS OF ATMOSPHERIC SCIENCE

objective analysis procedure. One of the simplest schemes to derive the wind field consistent
with the geopotential fields is to invoke the gradient wind balance and to derive the wind
fields from such a balance. The difficulty with the above approach is that the radius of
curvature needs to be determined for the trajectories of the air parcel. A better approach is to
obtain the balance fields by assuming at the initial time, that the initial wind is non-divergent
and also that the local time variation of the divergence component of the wind vanishes. The
above assumption leads to a direct relationship of the non-divergent wind with the
geopotential fields at the initial time through the balance equation. It turns out that the
observed baroclinic motions are not usually non-divergent and in such cases the divergent part
of the motion cannot be summarily disregarded. Further, for a given geopotential field, the
above balance equation provides for a horizontal wind field which is not unique.
The earlier and traditional approach to data assimilation using the separate processes of
objective analysis and initialization does work reasonably well in regions of adequate data
coverage. However, there are difficulties with the above approach over oceanic regions and
over southern hemisphere. This has led the major forecasting centres worldwide to adopt
continuous assimilation procedures in which the components of objective analysis and
initialization are combined with a continuous cycle of data assimilation known as four-
dimensional data assimilation.

14.3.4 Data Assimilation Cycle


It has been earlier mentioned in this chapter that the number of degrees of freedom of a
typical global model is about two orders of magnitude larger than the total number of
observations, available at any given time. This required the use of additional information
(background or first-guess) to prepare the initial condition for a numerical weather forecast.
Due to overall general improvements of forecasts, a short-range forecast can be chosen as the
first-guess in the operational data assimilation systems or data assimilation cycles. Figure 14.4
shows schematically the various components of a typical data assimilation cycle used in most
forecast centres worldwide.

Observations (+/–3 h)

Global analysis and balancing

Background or first-guess

Initial conditions

6 h forecast

Global forecast model

Operational forecast
FIGURE 14.4 Schematic plot indicating the various components of a typical data
assimilation cycle
NUMERICAL MODELLING OF THE ATMOSPHERE u 375

Most of the forecasting centres typically use a 6-hour assimilation cycle performed four
times a day. In the 6-hour data assimilation cycle for a global model, the first-guess or the
background field is a model 6-hour forecast (say, a forecast from 12 UTC to 18 UTC). The
first-guess fields and the observations are then combined to obtain the analysis using either
sequential or variational assimilation approaches (OI method, 3D-Var or 4D-Var). In the above
stage of the analysis cycle, all available observations within a time window of ± 3 hours (i.e.
from 15 UTC to 21 UTC) of the analysis time are used to update the analysis defined by the
first-guess. The analyzed data are then used in a 6 hourly forecast (18 UTC to 00 UTC of the
next day), which is then used as the background or first-guess for the 00 UTC analysis time
of the next day. The above-mentioned data assimilation cycle is then repeated. The model
forecast plays an important role in achieving four-dimensional data assimilation during the
data assimilation (analysis) cycle. Over regions of adequate and rich data coverage, the
analysis gains from the information contained in the observation. However, over regions of
poor data coverage, the forecast benefits from the information upstream, i.e. the forecasts are
able to transport information from data-rich to data-sparse areas.

14.4 SPECTRAL AND FINITE ELEMENT METHODS


In addition to the method of finite differences, there exist other general methods such as
spectral method and finite element method to solve a system of partial/ordinary differential
equations. Both the spectral methods and the finite element methods are considered superior
to the finite difference method as far as the accuracy of the numerical solution is considered.
Both the spectral methods and the finite element methods are together known as Galerkin
methods. The following subsections introduce briefly the Galerkin method and discussed
briefly the applications of the spectral method and the finite element method.

14.4.1 Galerkin Method


The method of finite differences has been introduced earlier in this chapter to solve partial
differential equations. In this method, the dependent variables (or unknowns) are specified at
certain grid points in space and time and the derivatives are replaced by differences using the
Taylor series expansion. There is another very important method of solving partial differential
equations known as the Galerkin method. In the Galerkin method, the dependent variables
(unknowns) are represented as a summation of functions, called the basis functions that have
a prescribed spatial structure. Usually, the coefficients associated with each basis function are
a function of time. The above Galerkin representation ensures that partial differential
equations are transformed into a set of ordinary differential equations for the coefficients.
These ordinary differential equations are solved using the finite differences in time. The two
most popular Galerkin methods are the spectral method and the finite element method. While
the spectral method uses orthogonal functions as basis functions, the finite element method
uses functions that are zero except in a limited region where they are low-order polynomials.
While the spectral method has been extensively used in meteorology for a number of years,
the method of finite elements is being applied to meteorological problems in recent times only.
376 u BASICS OF ATMOSPHERIC SCIENCE

Time-independent Example
The Galerkin procedure as applied to time-independent problems can be illustrated with the
following equation:
L(u) = f(x) (14.43)
where u is a dependent variable, f (x) is a specified forcing function and L is a differential
operator. Let the above equation be solved in the domain a £ x £ b, and assume that
appropriate boundary conditions (say, u at x = a and x = b) are provided. Consider a series of
linearly-independent functions jj (x) as the basis functions and approximate the dependent
variables u(x) as a finite series as given below
/
V Y
# Ç V K K K Y
(14.44)
K 

where uj is the coefficient for the jth basis functions. The error involved in satisfying the
differential Eq. (14.43) with the sum of the finite N terms in Eq. (14.44) is

È/ Ø
F/ -É Ç V KK K Ù  G Y
(14.45)
ÊK  Ú
The Galerkin procedure requires that the error be orthogonal to each basis function in the
following sense:
C

Ô F/ KJ EY  J  ž / (14.46)
B

The final form of the equation in the Galerkin procedure is obtained by substituting
Eq. (14.45) into Eq. (14.46)
C È / Ø C

Ô KJ - É Ç V KK K Ù EY  Ô KJ G Y
EY
ÊK  Ú
 J  ž / (14.47)
B B

As is clear from Eq. (14.47), the original differential Eq. (14.43) reduces to a system of N
algebraic equations that relate the unknown coefficients uj to the transforms of the forcing
function. The above procedure is very general and can be easily extended to more dependent
and independent variables.

Time-dependent Example
The Galerkin procedure can be easily extended to time-dependent problems as shown below.
Consider the following time-dependent differential equation
˜V
 - V
 (14.48)
˜U
NUMERICAL MODELLING OF THE ATMOSPHERE u 377

where the differential operator may be nonlinear. As before approximate, u(x, t) with a finite
series of N terms as
/
V Y U
# Ç V K U
K K Y
(14.49)
K 
where the coefficients uj (t) are functions of time and jj (x) are the basis functions. As
mentioned before, in the Galerkin procedure, it is more convenient to treat the time-
dependence term using finite differences in time. The final form of the equation is obtained
by substituting Eq. (14.49) into Eq. (14.48), multiplying by ji (x) and integrating over the
domain from a to b to get
C C È Ø
/ EV K /
Ç EU ÔB
KJK K EY  Ô KJ - É Ç V K K K Ù EY  J  ž / (14.50)
K  B ÊK  Ú
Equation (14.50) gives a system of N coupled ordinary differential equations in the
coefficients uj (t). By introducing finite differences in time for the time derivatives,
Eq. (14.50) can in principle be solved.

14.4.2 Spectral Method


The spectral method will be employed to solve the equation of the following form:

E V
G Y
  … Y … Q (14.51)
EY 
with boundary conditions u(0) = u(p) = 0 (14.52)
The following basis functions are chosen for solving the above equation using the spectral
method:
jj = sin jx, j = 1, …, N (14.53)
jj are chosen in Eq. (14.53) as they are orthogonal on the interval 0 £ x £ p and also satisfy
both the boundary conditions, Eq. (14.52). Using Eq. (14.53) as the basis function

È/ Ø /
- É Ç V KK K Ù Ç  K 
V K K K (14.54)
ÊK  Ú K 

and Eq. (14.47) becomes


/ Q Q
Ç 
K V K Ô KJK K EY Ô KJ G Y
EY J  ž / (14.55)
K   

Q Q
 Q
Since Ô TJO JY TJO KYEY

Ô <DPT J  K
Y  DPT J  K
Y >EY

E JK
(14.56)
 
378 u BASICS OF ATMOSPHERIC SCIENCE

where dij is the Kronekar delta with dij = 1 if i = j and 0 otherwise. Using Eq. (14.56) which
is nothing but the orthogonality condition in Eq. (14.55), one gets
Q

V K
o
J

Q ÔK K
G EY (14.57)


It is chosen from Eq. (14.57) that each coefficient uj is proportional to the finite Fourier
transform of the forcing term and can be calculated and hence the solution u(x) can be
obtained.

14.4.3 Finite Element Method


In this section, the finite element method will be employed to solve the same differential
equation, Eq. (14.51) subject to boundary condition, Eq. (14.52). For the finite element
method, the interval 0 £ x £ p is divided into N + 1 segments such that (N + 1) Dx = p.
Hence, the basis functions are chosen as tent-shaped, piecewise linear function as shown in
Figure 14.5.

1
jj(x)

0
0 ( j – 1)Dx j Dx ( j + 1)Dx p = (N + 1)Dx
x

FIGURE 14.5 Schematic of a tent-shaped, piecewise linear function used in finite element
method.

The point x = j Dx is known as the model point and the basis function jj has a value 1
at the model point. The basis function decreases linearly to zero at x = (j ± 1) Dx and is zero
everywhere else. From Figure 14.5, it is clear that jj(x) can be defined as follows:
0, when x > ( j + 1) Dx or x < ( j – 1) Dx Þ
Ñ
jj = [x – ( j – 1) Dx]/Dx, when ( j – 1) Dx £ x £ j Dx ß (14.58)
Ñ
= [( j + 1) Dx – x]/Dx, when j Dx £ x £ ( j + 1) Dx à
Due to the above definition of the basis function, the boundary conditions, Eq. (14.52) are
automatically satisfied. It is to be noted that the coefficient uj is actually the value of the
function u(x) at x = j Dx, since j ( jDx) = 1 and j (i Dx) = 0 for i ¹ j. Equation (14.47) then
becomes
/ Q
E K K Q
Ç VK ÔK K EY 
 Ô K K G Y
EY  (14.59)
K   
NUMERICAL MODELLING OF THE ATMOSPHERE u 379

Integrating the first term of the left-hand side of Eq. (14.59) by parts, one gets
/ Ë E È K J EK K EK J EK K Ø Û
Q Q
Ç V K Ô ÌÌ EY ÉÊ EY  EY EY ÙÚ ÜÜ EY  Ô K K G Y
EY  (14.60)
K   Í Ý 

The first term in the left-hand side of Eq. (14.60) vanishes since all the j’s are zero at x = 0
and x = p. The final form of Galerkin equation becomes
/ Q
Ë EK EK K Û Q
Ç V K Ô Ì EYJ EY Ý
Ü EY Ô K K G Y
EY J  " / (14.61)
K   Í 

Differentiating Eq. (14.58) yields the following:


0, when x > ( j + 1)Dx or x < ( j – 1)Dx Þ
Ñ
EK K
ÑÑ
= 1/Dx, when ( j – 1) Dx £ x £ j Dx ß (14.62)
EY Ñ
= –1/Dx, when j Dx £ x £ ( j + 1) Dx Ñ
Ñà
Using Eq. (14.62), the left-hand side of Eq. (14.61) becomes
V 'Y  V 'Y  V 'Y
J
(14.63)
J J

'Y 
as only three terms in the summation are different from zero. To evaluate the right-hand side
of Eq. (14.61), approximate f (x) in terms of the basis function
/
G Y
# Ç G KK K (14.64)
K 
Hence, the integral in the right-hand side of Eq. (14.61) becomes
/ Q /  'Y
J 

Ç K Ô KJ K KG EY Ç GK
Ô KJK K EY (14.65)
K   K   'Y
J 

Using x = x – i Dx, the above integral can be expanded into three integrals of the form
  'Y
Y Y  'Y
Y  'Y
 Y 'Y  Y

o GJ  Ô 'Y 
EY   GJ Ô 'Y 
EY  GJ  Ô 'Y 
EY (14.66)
'Y 'Y 

Equation (14.66) reduces to


G J  G  G J 
(14.67)
J


Hence, the final form of the finite element expansion becomes
V   V  V  G    G  G 
(14.68)
J J J J J J


'Y 
380 u BASICS OF ATMOSPHERIC SCIENCE

Comparing Eq. (14.68) with the centred finite difference form indicates that both the finite
element and the finite difference equations are the same except that in the former the forcing
function appears as a weighted average. If the forcing term is sinusoidal in form, the finite
element solution is much more accurate for the shorter wavelength as compared to the finite
difference method.

14.5 CHALLENGES IN WEATHER AND CLIMATE FORECASTS


There are countless challenges in the field of weather and climate forecasting using numerical
models. The foremost of the challenges confronting the weather forecasting researcher is to
extend the forecast range by employing advanced modelling research and methodology and
utilizing advanced data assimilation techniques with increased non-conventional observations.
Among the challenges in climate modelling and forecast is the answer to the following
question. To what extent can the climate be predicted and to determine to what extent human
influences affect climate? The following subsections briefly outline weather forecasting from
a historical perspective, ensemble forecasting and climate forecasting.

14.5.1 Weather Forecasting—A Historical Perspective


Weather forecasting can be defined as the application of science and technology to predict the
state of the atmosphere at a given location for a future time. The application of technology
(development of sensors, automated collection of meteorological observations, dissemination
of observations and the computing technology resources to run the numerical weather forecast
model) and science (use of models employing better physics, data assimilation schemes,
model verification and model diagnostics) together has immensely contributed to improved
weather forecasting.
Villhelm Bjerknes was the first to advocate in 1904 that the weather forecasting has to
be considered as an initial value problem of mathematical physics. Bjerknes suggested that
starting from the observed atmospheric state, the governing equations of atmospheric motions
be integrated forward in time to yield the weather forecast. Lewis Fry Richardson in 1922
actually carried out the first numerical weather forecast by direct hand calculation.
Richardson’s first attempts at numerical weather forecast failed badly since he predicted a
huge 145 hPa rise in pressure over a time period of 6 hours, while the observed pressure
remained more or less the same during the above time duration. Subsequently, detailed
analysis of Richardson’s results indicated that the failure was due to non-application of
smoothing technique to the data that resulted in the appearance of unphysical surges in the
pressure field. Furthermore, Richardson’s results also suffered due to the violation of the
stability requirement (refer to CFL condition in Appendix 7). Once the suitable technique and
the stability requirement were utilized, it turned out that Richardson’s weather forecast was
indeed reasonably accurate.
The first successful numerical weather forecast was performed in 1950 using the ENIAC
digital computer. The members of the above successful team were John Von Neumann,
Jule Charney and Ragnor Fjortoft. The first successful numerical weather forecast utilized the
barotropic vorticity equation—an equation simple enough and requiring lower memory and
NUMERICAL MODELLING OF THE ATMOSPHERE u 381

computing resources. In 1955, Norman Philip developed a two-layer hemispheric quasi-


geostrophic atmospheric model, which is considered the first among the atmospheric general
circulation models (AGCMs).
During the late 1950s and the beginning of 1960s, three separate groups ventured to
develop multilevel three-dimensional AGCMs based on the primitive equations. The first
group was part of the United States Weather Bureau, while the other two were the parts of the
University of California, Los Angeles, and the Lawrence Livermore National Laboratories. By
the middle of 1960s, National Centre for Atmospheric Research (NCAR), Geophysical Fluid
Dynamics Laboratory (GFDL), and the U.K. Meteorological Office (UKMO) had initiated
work on AGCM. Subsequently, several groups emerged throughout the world that developed
and improved AGCMs for global circulation studies such as European Centre for Medium
Range Weather Forecast (ECMWF). Global weather forecast models have horizontal
resolutions of 50 km–100 km, while regional mesoscale models have horizontal resolutions of
10 km–50 km. Further; storm scale models have much higher horizontal resolutions
(1 km–10 km). Current models have vertical levels ranging from 20 to 50 levels extending
from the surface to the stratosphere or even up to mesosphere in some models.
Despite the increased computing power in recent times and the associated increase in the
horizontal and vertical resolutions in atmospheric models, it is clear that several important
atmospheric processes cannot be explicitly resolved with the present models. The effects of
such sub-grid scale processes need to be parameterized in atmospheric models, i.e. the effects
of sub-grid scale processes are formulated in terms of the resolved fields. Furthermore, in
addition to sub-grid scale processes, some of the processes in the atmosphere are not
understood fully, and hence require fitting to observed phenomena; i.e. they need to be
parameterized. Some of the important atmospheric processes that are parameterized are
associated with radiation, cumulus convection and boundary layer. The nature of the
parameterization used in an atmospheric model has an important effect on the model forecast,
especially at larger time scales and is a challenging research area in weather forecasts.
If an important physical (sub-grid scale) process, which is observed in the real
atmosphere, is not parameterized, the above process may still manifest in the model
integrations as representing the resolved scales. The above behaviour known as aliasing can
seriously degrade the weather forecast. Also, when an atmospheric process (supposedly grid-
scale process) having scales not much smaller than the grid size occurs in the atmosphere, the
above can lead to the following difficulty—the resolved scales and the unresolved scales to be
parameterized are not very well separated. If the scales are not very well separated, the effects
of the unresolved scales on the larger resolved scales are difficult to parameterize as the
unresolved scales, which are not much smaller than grid size, get aliased onto the shortest
waves present in the solution. Sea-breeze circulations in models of horizontal resolutions of
50 km–100 km and cumulus parameterization in models having horizontal resolution of
10 km are two examples of unresolved scales, which are not well separated from the
corresponding resolved scales (grid size).
Very short-range forecasts are defined as forecasts up to 12 hours, while short-range
forecasts refer to forecasts between 12–72 hours. The medium-range forecasts are between
3-7 days, while extended-range forecasts refer to forecasts of about two weeks. Further,
monthly and seasonal forecasts are possible and are being performed in recent times for
382 u BASICS OF ATMOSPHERIC SCIENCE

periods ranging from one month to a season ahead of time. Over the years, there has been a
steady improvement in the accuracy and the utility of the weather forecast. Advances in
observing systems, computer technology and the increased understanding of the physics of the
atmosphere have all contributed towards improved weather forecasts at all time ranges.
Very-short range forecasts, in recent times have shown improved skill, especially for
prediction of the evolution and the movement of large- and medium-scale weather systems.
However, the utility and the accuracy of these forecasts decrease as the scale of the weather
system decreases and the time range of the forecasts increases. Forecasting the evolution and
movement of small-scale systems and intense phenomena such as tornadoes, hailstorms and
flash floods are, however, difficult. This difficulty in forecasting small-scale systems is due to
inadequate observations, insufficient computational ability and incorrect understanding of the
physical processes that are responsible for these events. However, there has been a general
improvement of forecasting small-scale systems when these systems are associated with
terrain, land use characteristics and also land–sea contrasts. Also, the lead-time warnings of
intense systems such as tornadoes have increased in recent times.
Short-range forecasts are showing increased accuracy due to improved observing
systems and use of efficient data assimilation schemes. The accuracy of forecasts of the
evolution and movement of large-scale systems and their associated precipitation patterns are
indeed good. There is also improved model estimation of the quantitative precipitation for
larger lead times, today, when compared to the performance a decade ago. However, the
forecasts of the precipitation patterns associated with smaller scale systems such as mesoscale
convective systems, fronts and thunderstorms are still quite difficult for most models.
Improved observing systems, high-resolution models and increased understanding of the
underlying physics may minimize the above difficulties substantially.
Medium-range forecasts have improved greatly over the last few decades. Advance
forecasts up to three days corresponding to major synoptic systems are regularly being done
today. Further, skill of the five-day precipitation forecasts has improved significantly. Similar
improvements in the three-day temperature forecasts are routinely observed, although the skill
associated with the temperature forecast, seven days in advance is somewhat poor.
In the extended range, the predictability of day-to-day weather for periods beyond a
week is very small. Hence in the extended range, operational centres examine the mean
temperature of 6–10 days and the occurrence of precipitation departure from normal. There is
significant improvement in the accuracy of the mean temperature of 5 days and precipitation
forecasts. It is hoped that improvements in the observing systems, development of better
models and utilization of statistical techniques may provide for better skill of the mean
temperature and mean precipitation fields for the time scale beyond a week in the extended
range. Over the last one decade, monthly and seasonal forecasts of temperature and
precipitation are being utilized in agricultural and energy sectors. It is envisaged that use of
improved models and application of sophisticated statistical methods will lead eventually to a
situation where the utility of the monthly and seasonal forecasts can be enhanced.
Lorenz showed in the early 1960’s that the atmosphere has a finite limit of predictability
of weather of about two weeks. He, in the above study, demonstrated that the forecast skill of
the atmospheric models depends not only on the accuracy of the initial conditions and
accurate treatment of physical processes, but also on the instabilities of the atmospheric flow.
NUMERICAL MODELLING OF THE ATMOSPHERE u 383

Hence, the loss of skill associated with the growth of errors due to instabilities depends on the
evolution of the atmosphere. That is, the weather forecast at some times can be accurate up to
a week, while at other times, the errors grow very fast reducing the forecast skill to just three
days. The above clearly reveals that one needs to account for the stochastic nature of the
atmospheric evolution. Epstein, in 1969, first proposed the concept of stochastic-dynamical
forecasting. Although the above idea did show promise for low-order models, it turned out to
be unfeasible for modern numerical weather prediction models with very large number of
degrees of freedom.

14.5.2 Ensemble Forecasting


It is seen that small changes in the initial conditions lead to large divergence in the results of
a numerical weather forecast after a few days indicating a definite limit to deterministic
predictability. The concept of ensemble forecasting—the atmospheric model is integrated with
a limited number of ensemble members under different initial conditions, was first proposed
by Leith in 1974. By analyzing the spread of the results of the different ensemble member’s
integration, one can determine and assign a measure of confidence to any forecast result.
Typically, an ensemble will consist of about 30 to 50 member forecasts.
Figure 14.6 shows the divergence of an ensemble of forecasts of temperature at a given
model point. The mean (shown as dotted line) and the spread s (the root mean square
distance of the ensemble members from the unperturbed control forecast) are also shown in
Figure 14.6. All the members of the ensemble begin at about the same initial condition, but
after a day or so they diverge from one another due to the chaotic nature of the atmosphere.
For ensemble forecasting, we need to integrate the models a large number of times, each time
with a different initial condition, i.e. choose a different initial perturbation for each member of
the ensemble. Hence, usually a low-resolution model is used for ensemble forecast. Earlier
studies have indicated that the advantages of employing a high-resolution model, vis-à-vis, a
lower-resolution model are mostly lost after about five days of integration. Hence, ensemble
forecasting methods are usually utilized for forecasts of five days or more.

FIGURE 14.6 Divergence of an ensemble of forecasts of temperature at a model grid point. In the
figure, the dotted line shows the mean and s indicates the spread of the ensemble
members from the unperturbed control
384 u BASICS OF ATMOSPHERIC SCIENCE

Ensemble forecasting has essentially the following objectives:


(i) To improve the weather forecast through ensemble averaging,
(ii) To provide a measure of reliability of the weather forecast from the knowledge of
the spread of the ensembles, and
(iii) To provide a quantitative basis for weather forecasting.
A good ensemble is one in which the true evolution of the atmosphere is not different
from the members of the ensemble, while for the bad ensemble, the true evolution of the
atmosphere is very different from the members of the ensemble. Even if the ensemble is bad,
it is still useful as it brings out the deficiency in the forecasting system.
It is important to choose the initial perturbation for each member of the ensemble in
such a way that the ensemble be a representation of the actual error probability density
function (PDF). Since the number of ensemble members is finite and there are a very large
number of possible uncertainties, it is debatable whether the ensemble forecasting can truly
represent the development of the initial uncertainties. Hence, it is important to choose an
ensemble, which forms a best approximation to the set of possible forecasts. An important
requirement for the above is that the ensemble should be chosen so that it traces the direction
of maximal growth on the error probability density function. The direction of maximal growth
on the error PDF will correspond to the initial conditions to which the numerical weather
forecast model is most sensitive.
Broadly, two methods exist in ensemble forecasting that determine the manner in which
the initial perturbations are generated. While the first method (denoted as Monte Carlo
forecasting) uses random initial perturbations, the second method (to which the following
methods known as breeding and singular vector), the perturbations depend very much on the
dynamics of the underlying flow.
In the Monte Carlo forecasting, the random initial perturbations sampling the estimated
analysis errors are added to the initial conditions. Studies with Monte Carlo forecasting have
shown that the random initial perturbations do not grow as fast as the actual analysis errors.
Further, the Monte Carlo forecasting does not contain “finite-size” growing errors, which are
very much present in the actual analysis. Also, the estimate of the analysis error covariance
did not change with time, indicating that the effects of the “errors-of-the-day” were not being
incorporated in the Monte Carlo forecasting method. The limitations of the Monte Carlo
forecasting are that it uses random initial perturbation and these random initial errors do not
give rise to large divergence among ensemble members. Further, the perturbations in the
Monte Carlo forecasting will be in the stable directions in phase space and will ultimately get
damped out by the model. Also, as the spread among the ensemble members will be small and
the modelled atmosphere will hence be unrealistically predictable in the Monte Carlo
forecasting. Due to the above reasons, Monte Carlo forecasting method is considered as an
inefficient approach to ensemble forecasting.
It is evident that the initial errors over the data sparse regions give rise to the largest
divergence among ensemble members. These errors, which are non-random in nature, need to
be identified and be included in the ensemble. That is, the ensemble must contain the
direction of the maximum error growth. Two methods, singular vector and breeding exist to
identify the directions of maximum error growth.
NUMERICAL MODELLING OF THE ATMOSPHERE u 385

Singular vectors essentially indicate the directions of the greatest instability in the
initial uncertainty and are presently used in the European Centre for Medium-Range Weather
Forecast (ECMWF) for ensemble forecasting. The early stage of the development of the error
growth is governed by linear dynamics and the timescale over which the linear dynamics is
dominant lasts for about 2–3 days. In this method, a linearization of the nonlinear atmospheric
model is done for small perturbation, the latter representing the initial linear error growth.
Using matrix algebra, the singular vector, which provides the direction in phase space in
which the error PDF has maximum growth, is found. In the singular vector method, the
ensemble members are defined with initial conditions, which correspond to the singular vector
directions, the first member having the direction of the largest error growth, the second
member having the direction of the second largest error growth and so on.
In the breeding method, which is presently being used in National Centre for
Environmental Prediction (NCEP), the following procedure is adopted. Given an atmospheric
flow, which is evolving a breeding cycle is initiated by selecting a random initial perturbation
with amplitude of the order of a typical measurement uncertainty. It is to be noted that the
random initial perturbation is introduced only once in the breeding cycle. The numerical
weather prediction model is integrated twice, from the control and from the perturbed initial
conditions (i.e. without and with the perturbation). At fixed intervals of time (say, every
6 hours), the control forecast is subtracted from the perturbed forecast. The difference
between the two output forecasts is then scaled down so that it has the same amplitude as the
initial perturbation and then added to the corresponding new analysis or model state. Studies
have shown that the perturbations generated in the breeding cycle (after an initial transient
period of 3–4 days) had a large growth rate. The breeding method causes each successive
perturbation vector to rotate closer to the significant error directions (i.e. in the direction of
maximum error growth) since the method breeds or chooses the fastest growing modes. In
this breeding method after several iterations, the growth rate of the perturbations approaches
the actual forecast error growth and hence the direction of the maximal growth can be
determined.

14.5.3 Climate Forecasting


Climate is the average and variations of weather over long periods of time. While, climate is
usually defined as average weather in a narrow sense, it can be defined rigorously as the
“statistical description in terms of the mean and variability of the relevant quantities such as
precipitation, surface temperature and surface wind over a period of time varying from
months to thousands or millions of years”. Intergovernmental Panel on Climate Change
(IPCC), has proposed the above definition of climate. The World Meteorological Organization
(WMO) has suggested a period of thirty years to be an appropriate period for the climate
system. “Climate is what you expect, weather is what you get”—a popular phrase sums up the
chief difference between climate and weather. Fluctuations in the climate system arise
naturally from interactions between the atmosphere, the ocean, the land and the part of the
earth surface, which is covered with ice. Further, fluctuations in the climate system may arise
from variations in the sun’s intensity and the changes in the earth’s energy balance due to
volcanic eruptions. During the last (twentieth) century, changes in the radiative forcing have
386 u BASICS OF ATMOSPHERIC SCIENCE

resulted in increased amounts of greenhouse gases and trace constituents. Greenhouse gases
are those which absorb outgoing long-wave radiation from the earth, while being transparent
to incoming short-wave radiation causing a net increase in the average temperature over the
whole globe.
Climate models may be simply defined as the numerical representation of the various
components of the earth’s climate system. Further, climate models use quantitative methods
(manifested as a set of mathematical equations) to simulate the interactions of the various
components of the earth’s climate system such as the atmosphere, the oceans, land surface
and ice. Climate models are typically expressed as a system of partial differential equations
derived from the basic laws of physics, fluid motion, and incorporating the various chemical
processes which occur in the climate system. The horizontal resolution of the present climate
models is of the order of several hundred kilometres and has about 20 to 30 levels in the
vertical. The above resolutions in both the horizontal and vertical directions are applicable for
both the atmospheric and the oceanic components of the climate model. The accuracy and
utility of the climate models is greatly determined by our ability to represent the physical and
chemical processes occurring in the atmosphere and the ocean as well as the interactive
representations of all the components of the climate system.
Climate models can vary from very simple zero- and one-dimensional models, to models
of intermediate complexity such as two-dimensional axially symmetric and vertical mean
models, and then to the complex three-dimensional general circulation models (GCMs). An
example of a zero-dimensional climate model is one where the prescription of solar constant
and the average earth temperature enables one to calculate the effective earth emissivity of the
long-wave radiation emitted to space. The energy balance models (EBMs) and the radiative–
convective models (RCMs) are examples of one-dimensional climate models. Averaging the
full three-dimensional models, in the vertical or in the zonal direction, leads to a two-
dimensional vertical mean and the zonally-symmetric climate models.
In addition to classification of climate models based on their dimensions, a classification
of deterministic models into explicit dynamical models and statistical dynamical models is
also usually done. While explicit three-dimensional models are based on the dynamical
equations of conservation laws of mass, momentum and energy, the statistical dynamical
models are based on time average equations in which the effects of large-scale transient
eddies are parameterized with respect to the mean quantities.
EBMs are one-dimensional models (variation in latitude only) based on the first law of
thermodynamics and provide steady state solution of the energy equation. In EBMs, the value
of the solar constant is prescribed, while the albedo of the earth surface is parameterized in
terms of the surface temperature. Further, EBMs also incorporate parameterized heat fluxes
and do not include any dynamics involving the atmosphere and the oceans. EBMs determine
the temperature changes at the earth surface assuming that the net radiative flux is zero.
EBMs, also at times, adopt a statistical dynamical approach.
RCMs are one-dimensional models (variation in the vertical direction only) based on the
thermodynamic energy equation and have a convective adjustment scheme. The RCM
incorporates the upwelling and downwelling radiative transfer through different layers of the
atmosphere that absorb and emit infrared radiation. Further, RCM also accounts for the
upward transport of heat by convection and evaluates the vertical profile of temperature in the
NUMERICAL MODELLING OF THE ATMOSPHERE u 387

atmosphere. The RCM can be used to study the effects of varying greenhouse gas
concentration on the effective emissivity and hence on the surface temperature of the earth.
Energy balance models do provide for the mean climatic information that is devoid of
statistical sampling fluctuations. In most weather models including general circulation models
(GCMs), there is inherent noise called weather noise attributed to the internal dynamics. The
presence of such weather noise limits the use of such models to be utilized for studies
involving prediction of climate change as the weather noise obscures true climate change. A
solution to the above would be to devise a model (called the statistical dynamical model)
which would only compute the statistical properties of the atmosphere and not the day-to-day
fluctuations of the atmosphere.
The two-dimensional statistical dynamical models (SDMs) contain parameterization of
the transports of heat and momentum in addition to the thermodynamic energy equation and
the equations for the mean zonal and meridional motions. Belonging to this group are the
zonally symmetric SDM (varying in latitude and vertical) and the vertical mean SDM (varying
in latitude and longitude). The two-dimensional zonally-symmetric SDMs are intended to
simulate the zonal mean structure of the atmosphere as manifested in the mean meridional
circulation and its influence on the temperature field. The two-dimensional vertical mean
SDM incorporates the land–sea contrast and is intended to investigate the longitudinal
variation of the climate at the earth surface or in the vertical mean. The limitations of both the
two-dimensional SDMs are related to the difficulty in formulating physically consistent
parameterization for the large-scale eddy processes.
The most general climate model—the GCM is time-dependent global model and can be
used for weather forecasting, understanding climate and its changes. A general circulation
model (also called a global climate model) is defined by a system of partial differential
equations and is used to numerically simulate changes in climate due to slow changes in, say,
physical parameters such as greenhouse gas concentrations, and to the slow changes in
boundary conditions such as solar constant. Generally, GCMs are run for years on end and the
utility of a GCM is judged by the quality of the statistics (means and variability) of climate.
The GCMs can be an atmospheric GCM (AGCM), an oceanic GCM (OGCM) or an
atmosphere–ocean coupled GCM (AOGCM). Inclusion of a sea ice model or a model for
evapotranspiration over land in an AOGCM leads to a full climate model.
An AGCM models the atmosphere by prescribing the sea surface temperature and
studying the simulated atmospheric motion. Typically, an AGCM does incorporate a land
surface model (LSM). Detailed parameterization for convection, radiation, land surface
processes, albedo, hydrology and cloud cover are available in an AGCM. The spectral method
or the method of finite differences is employed to solve the system of equations in an AGCM.
Typical AGCM resolution in the horizontal varies from 1° to 5° in latitude, longitude and has
about 20 levels in the vertical. An AGCM may incorporate processes involving atmospheric
chemistry also.
The simplest ocean numerical model called the swamp model provides for unlimited
source of water vapour to the atmosphere without storing or transporting heat and other
quantities. An improvement over the swamp model called mixed layer model is available in
which the ocean is represented as an ocean surface mixed layer. In this mixed layer model,
the ocean can store heat and also provide for water vapour to the atmosphere. A further
388 u BASICS OF ATMOSPHERIC SCIENCE

improvement of the ocean mixed layer models of intermediate complexity can be designed by
including some parts of the dynamics of the oceans and incorporating a few vertical layers so
that some of the large-scale ocean processes may be resolved while the remaining unresolved
ocean processes may be parameterized. Finally, a fully developed OGCM may be utilized to
simulate the general circulation in the world oceans. The OGCM is run with prescribed fluxes
from the atmosphere and may or may not contain a sea ice model. The OGCM may have
slightly higher resolution (about 150 km in horizontal) than an AGCM, but usually has similar
number of vertical levels. The observed eddies in the ocean are much smaller in horizontal
extent (a few hundreds of kilometres) and about an order of magnitude smaller than the
atmosphere. This would require that an OGCM ideally should have horizontal resolutions of
the order of 10 km.
The development of AOGCMs are very much essential since the atmosphere and the
ocean are strongly linked through the exchanges of momentum, heat and mass. AOGCMs
have the advantage that neither the fluxes across the interface of the ocean surface nor the sea
surface temperature has to be prescribed from the outside. The early stages of AOGCM
development required an ad hoc process of flux correction to obtain a stable climate. The
process of flux correction is effectively a linearization of the climate about its present state
and was thought necessary in the early stages of the AOGCM development so that models
simulating climate changes can start from initial conditions, which were reasonable. In recent
times, there is an increasing trend of not employing flux correction in AOGCMs. This is
because the flux correction process at best is an ad hoc process, meant to take care of model
deficiencies, and would not be required if the physical and chemical (both individual as well
as interactive) processes are modelled satisfactorily.
It is important that the simulation of climate model be systematically compared and
validated with observations. Further, it is necessary that climate models, in addition to
simulating the mean observed features, also reproduce the observed transient (fluctuating)
behaviour of the climate system. Presently global models are utilized to study the effects of
increasing concentration of the greenhouse gases on future global climates. GCMs are also
used to simulate the effects of volcanic eruptions on the atmosphere–ocean circulation. GCMs
are also utilized to simulate past climate periods, which represent abnormally warm and cold
climate conditions, respectively. GCMs can also be used to investigate the effects of variation
of the solar constant on the climate system. Also, studies investigating the effect of changes in
the atmospheric heat balance on the hydrological cycle can be undertaken using the GCMs.

REVIEW QUESTIONS

1. Mention the important and different numerical methods used in numerical weather
prediction.
2. What is data assimilation?
3. Are partial differential equations inherently more difficult to solve as compared to ordinary
differential equations? If so, why?
4. What is meant by computational instability?
5. What are relaxation methods?
NUMERICAL MODELLING OF THE ATMOSPHERE u 389

6. How are the partial differential equations classified into elliptic, parabolic and hyperbolic
equations?
7. State the difference between explicit and implicit finite difference schemes.
8. What are Galerkhin methods?
9. What is ensemble forecasting?
10. What is breeding method?
11. State the various types of climate models.
12. What are the advantages of Galerkhin methods over finite difference methods?
13. What is Atmospheric General Circulation Model?
14. What is Ocean General Circulation Model?
15. What are Coupled Ocean Atmospheric General Circulation Models?
16. What is objective analysis?
17. What is meant by initialization?
18. What are the components of a numerical weather prediction system?
19. Why the spectral methods are widely used in general circulation models?
15 Chaos in the
Atmosphere

Considering the importance of accurate weather prediction, it is no wonder that in addition to


greater efforts towards procedures involving the numerical modelling of the earth’s
atmosphere, a significant part of the research has been devoted to understanding the problem
of weather prediction from a more fundamental point of view. The work of Edward Lorenz in
1960s provided the initial impetus for work of this kind.
In 1821, the celebrated French mathematician, Pierre-Simon de Laplace wrote:
...An intelligence that at a given instance was acquainted with all the forces by
which nature is animated and with the state of the bodies of which it is composed
would—if it were vast enough to submit these data to analysis—embrace in the
same formula the movements of the largest bodies in the universe and those of the
lightest atoms: Nothing would be uncertain for such an intelligence, and the future
like the past would be present to its eyes.
The above view of Laplace of predictability augurs well for all deterministic systems,
and had wide support among the scientific community, until the work of Lorenz introduced
the fundamental characteristic of chaotic systems.
The following is Edward Lorenz’s account in his own words of what happened on a
particular day in 1961 which led to the discovery of chaos:
At one point I decided to repeat some of the computations in order to examine
what was happening in greater detail. I stopped the computer, typed in a line of
numbers that it had printed out a while earlier, and set it running again. I went
down the hall for a cup of coffee and returned after about an hour, during which
time the computer had simulated about two months of weather. The numbers being
printed out were nothing like the old ones. I immediately suspected a weak
vacuum tube or some other computer trouble, which was not uncommon, but
before calling for service I decided to see just where the mistake had occurred,
knowing that this could speed up the servicing process. Instead of a sudden break,
390
CHAOS IN THE ATMOSPHERE u 391

I found that the new values at first repeated the old ones, but soon afterward
differed by one and then several units in the last decimal place, and then began to
differ in the next to last place and then in the place before that. In fact, the
differences more or less doubled in size every four days or so, until all the
resemblance with the original output disappeared somewhere in the second month.
This was enough to tell me what had happened: the numbers that I had typed in
were not the exact original numbers, but were rounded-off values that had
appeared in the original printout. The initial round-off errors were the culprits;
they were steadily amplifying until they dominated the solution. In today’s
terminology, there was chaos.
The science of chaotic systems—systems which exhibit sensitive dependence on the
initial conditions is best illustrated using the forced nonlinear pendulum. Another interesting
example of a chaotic mechanical system is a system consisting of two balls that move in
response to gravity, the so-called bouncing ball problem. The new and complex field of chaos
requires new tools for the description of chaotic system. Poincare sections and Lyapunov
exponents are utilized extensively to characterize the chaotic system. An important and
interesting aspect of chaotic system concerns the nature of the path (route) to chaotic
behaviour from supposedly non-chaotic behaviour. Period doubling behaviour is one of the
commonly exhibited routes to chaotic behaviour.
The concepts and tools to investigate chaotic systems such as mentioned above could be
duly extended to study systems, which are relatively closer to the atmosphere. The Lorenz
model (attractor) is an important chaotic system, which investigates the Rayleigh-Benard
convection problem—an example of fluid in a container whose top and bottom surfaces are
maintained at different temperatures. Finally, there is a need to consider the atmosphere as a
chaotic system and describe the associated limits of deterministic predictability of the atmosphere.
Section 15.1 illustrates the important example of the forced nonlinear pendulum as a
chaotic system. The concepts of Poincare section and Lyapunov exponents are then described
in Section 15.2. While the period doubling and route to chaotic behaviour are outlined in
Section 15.3, the bouncing ball problem as a chaotic system is illustrated in Section 15.4. The
Lorenz attractor is introduced in Section 15.5, while the last subsection 15.6 discusses the
associated limits of deterministic predictability considering the atmosphere as a chaotic system.

15.1 ILLUSTRATIVE EXAMPLE OF CHAOS—


FORCED PENDULUM
Assuming small displacements, the governing equation for a simple pendulum of mass m
connected by a massless string of length l to a rigid support is

E R H

 R (15.1)
EU M
In the above equation, q is the angular displacement (in radians), g is the acceleration due to
gravity and t is the time. In this section, the following modifications to the above simple
picture are assumed: (i) the amplitude of oscillations is no longer small and hence sin q is no
392 u BASICS OF ATMOSPHERIC SCIENCE

longer approximated to q, (ii) the effects of damping are included, and the frictional force is
assumed proportional to the velocity, and (iii) incorporation of an external driving force acting
on the pendulum. It is assumed, for convenience, that the external forcings have a form,
which is sinusoidal with time and having an amplitude AD and an angular frequency wD. The
above external driving force may correspond to a situation where the pendulum mass has an
electric charge and is subjected to an oscillatory electric field. With the above-mentioned
modifications, Eq. (15.1) becomes

E R H ER

 TJO R  R  "% TJO X % U
(15.2)
EU M EU
where q is the damping parameter. Equation (15.2) can be very easily solved numerically once
ER
the initial conditions on q and are known.
EU
The results for q as a function of time obtained by solving Eq. (15.2) are shown in
Figures 15.1(a), 15.1(b) and 15.1(c) for various values of the forcing amplitude. The
parameter values used in obtaining the figures are q = 0.5, l = g = 9.8, wD = 2/3, time step
ER
Dt = 0.04, while the initial conditions are q (0) = 0 and 
= 2.0. With no external
EU
forcing (AD = 0), (refer Figure 15.1(a)), it is obvious that the pendulum with non-zero
damping would come to a state of rest after a few oscillations. For moderate forcing
(AD = 0.5) (refer Figure 15.1(b)), one finds the response to be made up of two regimes. The
first regime is characterized by the decay of the initial transient. This is followed by the
second regime in which the pendulum settles into a steady oscillation that oscillates at the
external driving frequency wD. A balance between the energy dissipated by the damping and
the energy added by the external driving force essentially determines the amplitude of the
steady oscillation, the so-called second regime.

0.1 0.8
AD = 0 AD = 0.5
0.6
0.05
0.4
q (radians)

q (radians)

0 0.2
0
–0.05 –0.2
–0.4
–0.1
–0.6
–0.15 –0.8
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
Time (s) Time (s)

(a) Time series of q for the forced (b) Time series of q for the forced pen-
pendulum for no forcing (AD = 0). dulum for moderate forcing (AD = 0.5).

FIGURE 15.1 Contd.


CHAOS IN THE ATMOSPHERE u 393

(c) Time series of q for the forced pendulum for large forcing (AD = 1.2).

In Figures 15.1(a), (b) and (c), q is restricted to lie between –p and + p. The values of parameters are
dR
q = 0.5, l = g = 9.8, wD = 2/3, Dt = 0.04. The initial conditions are q (0) = 0 and (0) 2.
dt

FIGURE 15.1

The nature of the solution of Eq. (15.2) changes markedly when the driving amplitude
AD assumes a rather high value of 1.2 (refer Figure 15.1(c)). The sharp changes in q
(around ± p) as seen in the figures for the case of AD = 1.2 are primarily due to resetting of
the angle q to ensure that q lies within the range of – p to + p. It is immediately clear that the
pendulum does not settle into any sort of steady oscillations for the case of AD = 1.2. That is,
at large value of the driving force, the pendulum does not repeat its behaviour, i.e. does not
exhibit any steady state or regular behaviour. It might appear that if one waits long enough
(i.e. for very large values of time), the pendulum would settle into some sort of regular
behaviour. However, it turns out that this is not the case. The above behaviour is perplexing
considering the fact that we are dealing here with a deterministic system. It is known from the
theory of differential equations, that once the initial conditions are given, the solution for
Eq. (15.2) is completely determined for all future times. However, it is to be noted that the
pendulum behaviour at high drive is not all-random behaviour and is an example of a chaotic
behaviour. However, it is obvious that we are dealing with a system, which is deterministic,
and at the same time exhibiting irregular behaviour.

15.2 POINCARE SECTION AND LYAPUNOV EXPONENTS


Although the deterministic nature appears irreconcilable with the irregular behaviour exhibited
by chaotic systems, it is possible to reconcile the apparent contradiction as follows. Let us
examine the nature of the stability of the solution corresponding to our pendulum problem by
envisaging two identical pendulums with exactly the same length and damping parameter
values. Imagine that both the pendulums are set in motion at the same time with same
external driving force. The only difference is a very small change in the initial conditions, say
ER
for the first pendulum at time t = 0 is 2.0 radians s–1, while the same for the second
EU
394 u BASICS OF ATMOSPHERIC SCIENCE

pendulum at initial time differs by 10–3 radians s–1 to the first pendulum. The results for
Dq = | q1 – q2 | (difference in q for the two pendulums) as a function of time for the two
driving amplitudes (AD = 0.5 and 1.2) are shown in Figures 15.2(a) and 15.2(b). Notice the
regular dips in Dq as seen in Figure 15.2(a) for moderate drive amplitude of AD = 0.5. These
dips in Dqÿ occur when one of the pendulums reaches a turning point, following which Dq will
vanish, since the two pendulums will cross one another. Disregarding the sharp dips in Dq for
AD = 0.5, it is clear that Dq exhibits a steady and a reasonably rapid fall with time. That is,
the motions of the two pendulums become more and more close to one another since Dq
approaches zero with increase of time. This means that the pendulum motion is predictable
for moderate (and obviously for small and zero) forcing, i.e. knowledge of the motion of the
first pendulum enables one to predict the future motion of the second pendulum, despite not
knowing the initial conditions of the second pendulum.

10
AD = 1.2
1

0.1
Dq (radians)

0.01

0.001

0.0001

1e-05

1e-06
0 20 40 60 80 100 120 140 180
Time (s)

(a) Dq depicted in logarithmic scale versus (b) Dq depicted in logarithmic scale versus
time for moderate forcing (AD = 0.5). time for large forcing (AD = 1.2) in the
forced pendulum problem.
FIGURE 15.2

The situation for the high drive (refer Figure 15.2(b) for AD = 1.2) shows that Dq
(depicted in logarithmic scale) increases very rapidly and irregularly with time. That is, the
motions of the two pendulums diverge from one another with increase in time. If one repeats
the above calculation (keeping the parameter values to be the same) for a range of different
initial values of q, however, keeping Dq fixed (say 10–3 radians) at the initial time and average
the Dq over the entire set of different runs and plot the averaged Dq with time, the behaviour
is similar to the mean line for AD = 1.2 case. The mean line, drawn to the logarithmic scale in
Dq, scales as log (Dq) ~ lt implying
Dq ~ elt (15.3)
where the parameter l is known as the Lyapunov exponent. The above results indicate that at
high forcing, pendulums, which start with nearly but not exactly the same initial conditions
will follow trajectories (give rise to solutions) that diverge in an exponential manner. It is now
possible to appreciate the apparently irregular behaviour of the pendulum at high drive despite
the system being deterministic. That is, a system while obeying deterministic laws of physics
may still exhibit irregular and unpredictable behaviour due to an extreme sensitivity to the
CHAOS IN THE ATMOSPHERE u 395

initial conditions. This in essence is what is meant by chaotic behaviour. From the above
discussions, it follows that a chaotic regime is characterized by l > 0, while for the non-
chaotic regime l < 0.
While it might appear that all the predictions are hopelessly impossible in the chaotic
regime, it turns out that one can make some accurate prediction on q even in the chaotic
regime. For this, one needs to plot the results in the so-called phase space plot, i.e. instead of
plotting the time series of q, one needs to plot the angular velocity w = dq /dt as a function of q.
Figures 15.3(a) and 15.3(b) show the phase space plots for moderate and high drive
except that dq /dt and q at initial time equal 0 and 0.2 respectively. For moderate forcing
(AD = 0.5), (refer Figure 15.3(a), the phase space plot, except for the initial transients, which
depends on the initial conditions, exhibits a regular orbit, which is independent of the initial
conditions. For high drive (AD = 1.2) (refer Figure 15.3(b)), the phase space plot exhibits
several orbits that are nearly closed and which last for one or two cycles only. While the
phase space plot shown in Figure 15.3(b) for the high drive case is not a simple one, it does
exhibit a pattern with significant structure when displayed using the so-called Poincare
section.

1 2.5
AD = 0.5 AD = 1.2
2
1.5
0.5
1
w(radians/s)

w (radians/s)

0.5
0 0
–0.5
–1
–0.5
–1.5
–2
–1 –2.5
–1 –0.5 0 0.5 1 –4 –3 –2 –1 0 1 2 3 4
q (radians) q (radians)
(a) Phase space representation of the forced (b) Phase space representation of the forced
pendulum for moderate forcing (AD = 0.5). pendulum for large forcing (AD = 1.2).

FIGURE 15.3

Figure 15.4 shows the phase space plot for the high drive case where the solution q and
ER
are plotted only at those times that are in phase with the driving force. That is, the phase
EU
space plot of Figure 15.4 is plotted only for those times for which wD = np, where n is an
integer. Such a phase space plot is called a Poincare section. Since the numerical solution is
available only in multiples of fixed time step Dt, the above condition is modified to the following:

OQ 'U
U 
X% 
The idea that systems appear a lot simpler when expressed using the Poincare section
can be appreciated from an analogy with the stroboscope. Imagine that one wishes to read the
label of a vinyl record as it is rotated in the record player at high speed. It is a matter of
396 u BASICS OF ATMOSPHERIC SCIENCE

common experience that it is impossible to read the label as the record rotates at high speed.
However, it turns out that it is possible for the human eye to read the label of the rotating
record if the record is illuminated with a light source or stroboscope which is turned on and
off at the same frequency of the record player. In the above case, the human eye will receive
input only when the record has a particular orientation and hence the eye will be able to read
the label in the record without any difficulty. The above analogy reveals that objects (systems)
look much simpler when they are observed at a rate (frequency) that matches the frequency
inherent in the object (system). The above function of the stroboscope has led us to examine
the phase space plot only at times that matches the external drive frequency and this is exactly
the idea behind plotting the Poincare section. As in the vinyl record example, one should
expect significant structure of the chaotic system with its associated simplified phase space
plot when examined using the Poincare section. As seen in Figure 15.4, the Poincare section
for the pendulum in high drive does exhibit significant structure and is very much simpler as
compared to Figures 15.3(a) and 15.3(b). Further, it turns out that except for the initial
transients, the phase space plot is the same for a wide-range of initial conditions. That is,
despite not being able to predict the behaviour of q (t) at all future times in the chaotic
ER
regime, one is in a position to expect that the system will possess values of q and which
EU
will put it on this surface of points, as in Figure 15.4. This means that the motion of the
pendulum as revealed in the Poincare section is such as to be drawn along this surface of
points. Such a surface is known as attractor. It is to be noted that attractors are also present in
non-chaotic regimes. In the case of moderate forcing (AD = 0.5), the associated Poincare
section will reduce to a single point since at any particular point of the external drive cycle,
one would always find the same values of q and dq /dt. While in the non-chaotic regime, the
attractors have simple form; they have a very complicated structure in the chaotic regime.
Chaotic attractors have a fractal structure and are referred as strange attractors. If one were
to construct a high-resolution plot of Figure 15.4 and confine one’s attention to the region
q > 2 radians, one notices that there is more structure in the attractor than is obvious in
Figure 15.4. It turns out that the closer one looks, the more structure one finds. All chaotic
attractors have the above feature since all chaotic attractors are fractal objects.
2.5
AD = 1.2
2

1.5
w(radians/s)

0.5

–0.5
–4 –3 –2 –1 0 1 2 3 4
q(radians)
FIGURE 15.4 Poincare section for the forced pendulum for large forcing (AD = 1.2).
CHAOS IN THE ATMOSPHERE u 397

A better understanding and appreciation of the nature of the strange attractors can be
obtained from the following analogy. Imagine a small town of ten thousand people. Assume
that to cater to the above population, there is a need for one shopping mall. Assume that the
town does indeed have a shopping mall. Imagine a situation where a new company opens its
new plant on the outskirts of the town, creating an opening of additional jobs for ten thousand
more people. If the same town expands rapidly to accommodate twenty thousand people and
in the process opens an additional shopping mall, then the equilibrium is maintained. Such
equilibrium can be called an attractor. Now, envisage a situation that instead of adding ten
thousand to the original population of ten thousand, some three thousand people move away
from the town leaving behind seven thousand people only. Assume that the existing shopping
mall can only exist if it can cater to some eight thousand regular customers. Since only seven
thousand people remain, the shopping mall runs into heavy losses and ultimately closes down.
Meanwhile, the demand for a shopping mall increases and a second and different company
goes ahead and builds a new shopping mall. The new owners believe that a new shopping
mall will attract new people, not only from that town, but also from other neighbouring
towns. It turns out that the new shopping mall does attract new customers. However, let us
assume that many more people in the town had decided to leave the town and their plans are
not going to be changed just because of the appearance of a new shopping mall. The owners
try to break even and fail and after some time close the shopping mall. Demand again rises. A
third party opens a new shopping mall, which attracts new customers, but not enough to bring
in the profits. The shopping mall closes again. The above kind of situation also corresponds
to some kind of equilibrium, the so-called dynamic equilibrium. A dynamical kind of
equilibrium is called a strange attractor. The difference between the attractor and the strange
attractor is that while the attractor represents a state to which the system will ultimately settle,
the strange attractor represents a behaviour in which the system runs from one situation to
another situation without ever settling down.

15.3 PERIOD DOUBLING AND ROUTE TO CHAOS


As discussed in the earlier section, the forced pendulum exhibits chaotic behaviour at high
drives and regular oscillatory motion at moderate drives. The above change in behaviour at
high drive raises important questions on the transition from regular oscillatory to chaotic
behaviour. Further investigation reveals that the forced pendulum exhibits transition to chaotic
behaviour at several different values of the external forcing amplitude. It turns out that the
transition to chaotic behaviour is more clearly brought out at higher driving amplitudes and
hence we shall investigate the motion of the pendulum at these higher driving amplitudes.
Figures 15.5(a), 15.5(b) and 15.5(c) depict the numerical results for q as a function of
time for the following three values of the forcing amplitude (AD = 1.35, 1.44 and 1.465), with
the other parameter values and the initial values of q and dq /dt unchanged. Again, the results
of q in the figures are restricted to lie within the range of – p to + p. For a forcing amplitude
of AD = 1.35, Figure 15.5(a) clearly reveals that except for the sharp resetting of q near – p
and +p, the forced pendulum exhibits steady oscillations with a period close to 3p radians.
398 u BASICS OF ATMOSPHERIC SCIENCE

4 4
AD = 1.35 AD = 1.44
3 3
2 2
q (radians)

q (radians)
1 1
0 0
–1 –1
–2 –2
–3 –3
–4 –4
0 20 40 60 80 100 0 20 40 60 80 100
Time (s) Time (s)

(a) Time series of q for the forced (b) Time series of q for the forced
pendulum with forcing (AD = 1.35). pendulum with forcing (AD = 1.44).

4
AD = 1.465
3
2
q (radians)

1
0
–1
–2
–3
–4
0 20 40 60 80 100
Time (s)

(c) Time series of q for the forced pendulum


problem with forcing (AD = 1.465).

FIGURE 15.5 In the above figures q is restricted to lie between – p and p.


Since the external forcing frequency wD is the time period of the external driving force


Q
equals Q  This essentially reveals that for a forcing amplitude AD = 1.35, the pendulum
X%
oscillates at the same frequency as the external driving force. Such a behaviour is termed as
Period-1 behaviour. For the forcing amplitude of AD = 1.44, Figure 15.5(b) shows a
behaviour which is still periodic. However, it turns out that in this case, the period of the
pendulum equals twice the period of the external driving force. It can be readily seen from
Figure 15.5(b) that for AD = 1.44, the maximum of q alternates between q ~ 1.9 and qÿ ~ 2.6.
The above behaviour is named as Period-2 behaviour. It is well known that the response of a
nonlinear system excited by a single frequency external source is characterized by
components, which are integral multiples of the driving frequency (so-called harmonics)
CHAOS IN THE ATMOSPHERE u 399

together with a component, which equals the external driving frequency. However, the
period-2 behaviour seen in Figure 15.5(b) for AD = 1.44 has a nonlinear response
corresponding to a frequency which is half of the external driving frequency or twice the time
period of the external driving period. For the forcing amplitude of AD = 1.465, Figure 15.5(c)
shows that the motion of the pendulum exhibits a time period which is four times the external
driving period. In other words, the maximum of q alternates, other than the sharp resetting
near q = – p and +p, between the following four values, q ~ 1.69, 2.82, 1.52 and 2.72. The
above behaviour is called as the Period-4 behaviour. Again, we have a situation where the
nonlinear system exhibits a response, which contain subharmonic frequency (one fourth of the
external driving frequency).
It is clear from the above discussions that the response of the forced pendulum exhibits
period doubling behaviour at amplitudes close to AD = 1.44 and AD = 1.465. The question,
which naturally arises, is the following. Will the period doubling behaviour continue with
increased AD values and if so what this means to transition to chaotic behaviour?
A convenient way to answer the above question is to utilize a bifurcation diagram, as shown
in Figure 15.6.

FIGURE 15.6 Bifurcation diagram for the forced pendulum problem.

The above bifurcation diagram for q as a function of various driving amplitudes is


plotted as follows. Choose a particular drive amplitude, say, AD = 1.35. By neglecting the
initial transients (say corresponding to some two to three hundred external time periods), plot
the value of q at times that are in phase with the external driving force. For regular period-1
behaviour, the q value remains a single value determined by the external drive amplitude.
Repeat the above for several drive amplitudes. The transition from period-1 behaviour to
period-2 behaviour can be immediately discerned using the bifurcation diagram and it takes
place at a drive amplitude of AD = 1.424. Further, the transition between period-2 and period-4
behaviour occurs at a drive amplitude of AD = 1.459. As seen in Figure 15.6, the above period
doubling process continues although the resolution of Figure 15.6 is inadequate in manifesting
the period-8 behaviour and higher period behaviour. The inference is that the period doubling
cascading process will end ultimately in a transition to chaotic behaviour.
400 u BASICS OF ATMOSPHERIC SCIENCE

The above discussions provided some qualitative basis for understanding the transition
from regular non-chaotic to chaotic behaviour for the forced pendulum example. The
question naturally arises whether the above-mentioned period doubling transition to chaotic
behaviour is observed in other chaotic systems. Further, are there other ways in which the
transition to chaotic behaviour can occur? It turns out that many systems exhibit chaotic
behaviour. However, in most chaotic systems, there appears to be only a few ways in which
the transition from regular to chaotic behaviour occurs and one of the ways is the period
doubling scenario. It turns out that many systems exhibit the period doubling route to chaos
and such systems do exhibit universal properties. From Figure 15.6, it is clear that the
spacing between the period doubling transition becomes progressively smaller as the order of
the transition increases, i.e. the period-2 region extends from AD » 1.424 to 1.459, while the
period-4 region is much smaller extending only from AD » 1.459 to 1.476. If An is the driving
amplitude at which the transition to period-2n transition occurs, then the progressive shrinkage
of the periodic window can be described by a parameter dn, defined as
"O  "O 
EO (15.4)
"O  "O
where dn as n becomes large approaches a constant d called the Feigenbaum. It turns out that
all systems which exhibit period doubling route to chaos show universal properties in the
sense that all such systems possess the same value of d ~ 4.669.

15.4 BOUNCING BALL PROBLEM


In this example, the system consists of two balls that move in response to gravity in the
vertical direction only. The balls during their motion can collide with one another. Further, the
bottom ball can also collide with the ground. For convenience, it is assumed that all the
collisions are completely elastic. The governing equations of motion of the balls with masses
m1 and m2 (subscript 1 refers to the bottom ball, while 2 refers to the top ball) is given by

E  Y
H (15.5)
EU 
E  Y
H (15.6)
EU 
where x1 and x2 are the vertical positions of balls 1 and 2 from the ground. The above two
equations can be written as a system of four first-order ordinary differential equations by
defining the velocities v1 and v2 of the two balls. Again, knowing the initial position and the
initial velocities of the two balls, the above problem can be very easily solved. The elastic
collision of ball 1 with the ground is handled by simply altering v1 to – v1 when x1 £ 0. The
elastic collision of the two balls occur when x2 < x1. The expressions for the final velocities
of the two balls undergoing elastic collision with one another is given by
È N  N Ø
v1f = É W  ÈÉ N ØÙ W (15.7)
Ê N  N Ú
Ù Ê N  N Ú
J J
CHAOS IN THE ATMOSPHERE u 401

È N Ø
v2f = É W  ÈÉ N  N ØÙ W (15.8)
Ê N  N Ú
Ù Ê N  N Ú
J J

where the subscripts i and f refer to the velocities before and after the elastic collision
between the two balls.
The results for the position of the second ball (the upper one) as a function of time are
shown in Figures 15.7(a), 15.7(b) and 15.7(c) for the following three cases:
(i) The balls have the same mass,
(ii) The upper ball has twice the mass of the lower ball, and
(iii) The mass of the upper ball is nine times the mass of the lower ball.
In all the above cases, the mass of the lower ball m1 is taken as unity. Also, the initial
positions of the upper and lower balls are taken as 3 and 1, respectively, while the initial
velocities of both the balls are assumed to be unity. The time step chosen is Dt = 0.0003. For
convenience, all the above quantities are assumed to be unit less.

4 4
m 2 = m1 m2 = 2m1
3.5 3.5
3 3
2.5 2.5
x2

x2

2 2
1.5 1.5
1 1
0.5 0.5
0 0
0 5 10 15 20 0 5 10 15 20
Time (s) Time (s)
(a) The position of the upper ball as a (b) The position of the upper ball as a
function of time for m2 = m1. function of time for m2 = 2m1.

4
m2 = 9m1
3.5
3
2.5
x2

2
1.5
1
0.5
0
0 5 10 15 20
Time (s)
(c) The position of the upper ball as a
function of time for m2 = 9m1.

FIGURE 15.7
402 u BASICS OF ATMOSPHERIC SCIENCE

It is clear from Figure 15.7(a) that in the first case (both the balls have the same mass),
the motion is periodic. The upper ball from Figure 15.7(a), for the first case, appears to
approach the ground at heights, which do not quite reach the same level. The above is due to
the discreteness of the time step. However, it is clear that the upper ball reaches the same
maximum height (for the first case), and also this maximum height is attained at the same
evenly spaced time intervals. For the maximum height, the discreteness of the time step has
smaller impact since the velocity of the ball is smaller close to the maximum height. The
results for the second case (upper ball twice as massive as the lower ball) are markedly
different as can be seen from Figure 15.7(b). In this case, the position of the upper ball varies
in an irregular way with time and hence the resultant motion is no longer periodic. The
maximum heights reached by the upper ball in the second case are no longer the same, while
the times at which these maximum heights are achieved are no longer evenly spaced. For the
third case, (upper ball is nine times as massive as the lower ball), the results in Figure 15.7(c)
indicate a nearly periodic behaviour, but not exactly periodic behaviour.
Figures 15.8(a), 15.8(b) and 15.8(c) depict the phase space plots for the above three
cases where the velocity of the upper ball is plotted as a function of the position of the upper

(a) Phase space plots corresponding to (b) Phase space plots corresponding to
the position and the velocity of the the position and the velocity of the
upper ball for m2 = m1 upper ball for m2 = 2m1.

(c) Phase space plots corresponding to the position


and the velocity of the upper ball for m2 = 9m1. The
above plots are drawn only at those times when the
lower ball hits the ground.

FIGURE 15.8
CHAOS IN THE ATMOSPHERE u 403

ball at those times when the lower ball hits the ground. The above is just one example of
constructing the Poincare section. Another alternate construction of the Poincare section
would correspond to the plotting of the velocity of the lower ball as a function of its position
at those times when the position of the upper ball is at its maximum height. It is expected that
in the first case, which gave rise to regular periodic motion, the attractor should have a very
simple form and this indeed is seen in Figure 15.8(a). For the second case, (refer Figure 15.8(b)),
however, the attractor in the phase space has a very complicated structure, indicating that the
system is in a chaotic regime. Further, for the third case, (refer Figure 15.8(c)), the attractor
has an interesting structure which is intermediate between the first and the second cases.

15.5 LORENZ ATTRACTOR


The discussions up to now in this chapter, while not directly related to the atmosphere or even
to fluid motions illustrated the various facets of chaos in simple mechanical systems such as
the forced pendulum and the bouncing ball problems. In this section, the well-known Lorenz
system as a chaotic system is introduced. The Lorenz system is in many ways closer to the
atmosphere than the earlier examples considered. Lorenz in 1959 developed a low-order
atmospheric model—a model whose evolution was described by 12 variables. The 12-variable
model was forced by external heating and had damping terms to handle dissipation. Lorenz
set out to find a non-periodic solution and by assuming the external heating to be a function
of both latitude and longitude, he was able to finally obtain a non-periodic solution.
To convince himself, that the solution did not have any periodicities, Lorenz ran the above
12-variable model for several simulated years. Lorenz printed the evolution of the 12-variable
model with three significant digits. Since Lorenz wanted to repeat part of the integration in
more detail, he preferred to start the integrations from an intermediate stage, by punching in
the values of the variables from the print out of the earlier run with three significant digits.
Lorenz, to his surprise found that the new solution, started from an intermediate stage, was
completely different from the original solution. After carefully comparing the two solutions,
Lorenz found that the two solutions did coincide at the beginning of the new run. However,
after a few days of integration the last digit of the two runs started differing from one another
and after about two months of integrations there were no resemblance whatsoever between the
two runs. The above provides briefly the historical picture of the discovery of “chaos”. Later,
in 1963 Lorenz made further approximations and put forward a three-variable model, known
as the Lorenz model, to study numerically the well-known Rayleigh-Benard problem. The
above problem essentially consists of a fluid in a container whose upper and lower surfaces
are held at different temperatures. Despite, its apparent simplicity, the above fluid model does
resemble the real atmosphere in the sense that the latter is also a fluid, which is heated from
below. Lorenz three-variable model was essentially arrived at after several assumptions and
approximations to the original Navier Stokes equations and they are as follows:
EY
= s (y – x) (15.9)
EU
EZ
= – xz + rx – y (15.10)
EU
404 u BASICS OF ATMOSPHERIC SCIENCE

E[
= xy – bz (15.11)
EU
The Lorenz variables x, y and z are related to the temperature, density and velocity of the
fluid. While the parameter r is a measure of the temperature difference across the fluid
container, the parameter s refers to the Prandtl number (ratio of the coefficient of viscosity to
the coefficient of thermal conductivity) of the fluid, and the parameter b refers to the ratio of
the width to the height of the container. Although, the Lorenz equations have no resemblance
to the full and modified Navier Stokes equations as applied to a rotating earth, Lorenz
believed, quite rightly, that any behaviour exhibited by these equations (Eqs. (15.9)–(15.11))
must also be observed in the full weather equations. In the early 1960s, when the computing
resources were very much limited, Lorenz thought it prudent to numerically solve Eqs. (15.9) –(15.11)
rather than attempting to solve the full weather equations. Before discussing the results of the
numerical solution of the Lorenz’s equations, it is pertinent to note that the Lorenz equations
are dissipative in nature. The above can be easily observed from the divergence of the flow
˜Y ˜Z ˜[
   T  C  
(15.12)
˜ Y ˜Z ˜[
Equation (15.12) clearly indicates the dissipative nature of the Lorenz equations in that an
initial volume V will contract with time to Ve–[s +b+1] t. Further, the Lorenz system is nonlinear
(contains products of the dependent variables) and is also autonomous (the coefficients are
time independent). To solve numerically the Lorenz equations, one requires the values of the
parameters s, r and b as well as the initial conditions on the three dependent variables
x, y and z, respectively. The results of the numerical solution of the Lorenz equations for the
variable z as a function of time are shown in Figure 15.9 for various values of the parameter
r(r = 5, 10 and 25). The above results are obtained by assigning the values of the parameters

FIGURE 15.9 Variation of the variable z as a function of time for different values (r = 5, 10 and 25)
of the temperature difference across the fluid container.
CHAOS IN THE ATMOSPHERE u 405

s = 10, b = 8/3. Further, the initial conditions used in obtaining the above results are x = 1,
y = z = 0 at the initial time. The time step in the numerical algorithm used has a value of 0.0001.
It is clear from Figure 15.9 that the fluid system, except for the initial transients which
die away, exhibits a steady convective motion for the case of r = 5. The warm fluid produced
at the lower surface rising and the cooler fluid returning from the upper surface characterizes
the fluid motion. This is what one would expect when the lower surface is slightly warmer
than the upper surface. The result as seen from Figure 15.9 for the slightly stronger forcing
(r = 10) is again characterized by the establishment of the steady convective fluid motion,
after the initial transients decay away. For the r = 10 case, the initial transients take a little
longer to decay as compared to the r = 5 case. Further, the solution at r = 10 case,
corresponds to a relatively stronger steady convective circulation as compared to the r = 5
case. The nature of the fluid behaviour, however, changes markedly for the r = 25 case, as
can be seen from Figure 15.9. For the case of r = 25, the initial transients are more or less
periodic, but are followed by highly irregular or chaotic behaviour at later times. From the
above discussions, it is clear that the response of the Lorenz system is characterized by steady
convective circulation (non-chaotic regime) for low and moderate forcing (r = 5 and 10) and
a highly irregular and unpredictable behaviour at large forcing (r = 25). Detailed investigation
have revealed that the transition from the steady convective behaviour to chaotic behaviour
occurs at r = 24.74.
Similar to the forced pendulum and the bouncing ball problems, one can construct phase
space plots and Poincare sections for the Lorenz system. The phase space plots are
constructed by envisaging that the variables x, y and z represent coordinates in some abstract
space and the solution of the Lorenz equations at a given time can be represented as a point
in the above abstract space. Further, the evolution of the solution of the Lorenz equations will
be represented by a trajectory in this space. The phase space plots can be constructed by
simply obtaining the projection of this trajectory onto, say, the xz plane by plotting the values
of z as a function of x. Such a phase space plot is shown in Figure 15.10 for the case of

45
r = 25
40

35

30

25
z

20

15

10

5
0
–20 –15 –10 –5 0 5 10 15 20
x
FIGURE 15.10 Phase space plot of z versus x for the case r = 25 of the Lorenz model.
406 u BASICS OF ATMOSPHERIC SCIENCE

r = 25. It is clear from Figure 15.10 that the Lorenz system undergoes approximately periodic
and regular oscillation characterized by roughly circular orbits on one side of the line x = 0.
The trajectories of the Lorenz system as seen from Figure 15.10, then move to the opposite
side of the line x = 0 and undergo a new series of periodic or regular oscillations. If one were
to plot the phase space plots by projecting the trajectories on either the yz plane or the xy
plane, a similar result as seen in Figure 15.10 is obtained. The above phase space plot of the
Lorenz system does reveal some underlying regularity. However, one has to construct the
Poincare section for the Lorenz system to unravel any underlying attractor in the system. The
Poincare system can be easily constructed for the Lorenz system by simply plotting the places,
say the values of z and y, when the trajectory intersects the yz plane.
Figure 15.11(a) shows the results of the plots of z as a function of y, when x = 0 (when
the trajectory cross the yz plane) for the case of r = 25. Figure 15.11(b) is again the Poincare
section for the case of r = 25, with z plotted against x, with the points plotted only when y = 0.
Despite, the irregular and chaotic behaviour for the case of r = 25, as seen in Figure 15.9, the
Poincare section (refer Figures 15.11(a) and 15.11(b)), shows the existence of a highly regular
behaviour in the phase space trajectory of the Lorenz system. The above Poincare section
defines an attractor surface in the phase space and it can be shown that the above attractor
surface is independent to the choice of the initial conditions.

40 45
r = 25 r = 25
35 40
+

30
+
+
++++++
+ 35
++++
30
25
25
z

20
z

20
15
15
10 10
5 5
0 0
–10 –5 0 5 10 –15 –10 –5 0 5 10 15
y x
(a) Phase space plots of z versus y with (b) Phase space plots of z versus x with
points plotted only when x = 0, for the points plotted only when y = 0, for the
case r = 25 of the Lorenz model. case r = 25 of the Lorenz model.

FIGURE 15.11

The Lorenz model also exhibits the period doubling route to chaos. In addition, the Lorenz
model also exhibits a different route to chaos called the “intermittency route to chaos”.
Figure 15.12 shows the results of the Lorenz model for z as a function of time for two
values of r (r = 160 and r = 166.5), respectively. Again the values of the parameter s and b,
and the initial conditions remain the same as before. Figure 15.12 shows that, except for the
initial transients, for the case of r = 160, the solution of the Lorenz equation exhibits periodic
oscillations. While the above pattern of the waveform of these oscillations are not very
simple, they are at least regular, stable and persist for all times. The case of r = 160 typically
corresponds to period-1 behaviour. If one were to examine the behaviour of the Lorenz system
CHAOS IN THE ATMOSPHERE u 407

for smaller r, one would find the period doubling scenario and the situation ultimately leading
to chaos. Instead, if one observes the nature of the solution for larger r (r = 166.5), as seen
from Figure 15.12, one observes that the motion is still approximately periodic for many
cycles of the oscillation. However, these periodic oscillations are suddenly and abruptly
interrupted by deviations of the periodic oscillations, which can be called as “chaotic
interludes”. The behaviour of the Lorenz system at r = 166.5 is definitely irregular, but only
barely so. With a slightly larger value of r, the chaotic interludes occur more and more often.
With a further increase of r, the chaotic interludes will dominate and a stage will be reached
where these interludes would ultimately swamp the underlying periodic behaviour. The above
scenario is called “the intermittency route to chaos”.

600

500 r = 166.5

400

300
z

200

100
r = 160
0
0 5 10 15 20 25 30
Time (s)

FIGURE 15.12 Variation of the variable z as a function of time for r = 160 and r = 166.5 of the
Lorenz model.

15.6 LIMITS OF DETERMINISTIC PREDICTABILITY


The atmospheric system is much more complex than the simple fluid system considered by
Lorenz in the three-variable model. Since the above three-variable model was arrived at from
the full weather equations, any behaviour (chaotic or otherwise) of the Lorenz system should
also be observed in the full weather equations. From the chaotic behaviour exhibited by the
Lorenz system, one can conclude that the atmosphere is indeed a chaotic system.
A consequence of the above is that small initial errors associated with the estimate of the
current state of the atmosphere can grow very rapidly and have a decisive impact on the
forecast of the weather. Due to the limited number of observations (70% of the surface area is
covered by oceans and these are primarily data sparse regions) available and the non-uniform
distribution of these observations over the earth surface, there is always some amount of
uncertainty in the estimate of the current state of the atmosphere. The above growth of initial
408 u BASICS OF ATMOSPHERIC SCIENCE

errors limits the deterministic predictability of the atmosphere to about two weeks. Lorenz
proposed the above estimate of two weeks in 1963 from studies, which investigated the
doubling time of small errors obtained from identical twin experiments. Studies have also
shown that small errors in resolvable, coarse, synoptic scales double in about two to three
days. However, once the errors become large, their growth rates decrease. Further, the small
errors in the finer irresolvable cloud scales double very fast, in a matter of hours or less.
Since one is not concerned with the forecasting of the finer scales, the above fast doubling
time of the finer scales poses no special difficulties. However, the errors in the finer scales
after attaining somewhat appreciable size tend to produce error in the resolvable scale through
nonlinear interactions. Hence, it is clear that the rate of error growth determines the limit of
deterministic predictability. Even though the two-week limit gives an estimate of the average
predictability, the actual predictability is somewhat variable and is determined by the nature of
the atmospheric instabilities present. While, there are periods where the atmospheric
predictability is much lower than the average, there are also periods where the predictability
remains fairly skillful, for say, 15 days. It is known that the time scales of atmospheric
instabilities depend on the spatial scales with smaller scale instabilities growing very much
faster than that associated with larger scales. The above discussion was mainly confined to the
mid-latitudes and these limits of predictability over mid-latitude regions are primarily
determined by the synoptic scale baroclinic instabilities, which dominate the dynamics of
these regions.
For the tropics, the situation is markedly different since the dynamics of convective
instabilities and barotropic instabilities are the most dominant in these regions. Also, the
evolution of the synoptic waves is much more strongly modulated in the tropics by convective
rainfall than is the situation for the mid-latitudes. One of the major problems in tropical
forecasting is the fact that most global atmospheric models are somewhat inadequate in
realistically parameterizing sub-grid scale processes, such as cumulus convection, which are
among the most dominant mechanism in the tropics. Despite the fact that the error growth rate
due to internal dynamics in the tropics is less than that of the mid-latitudes, the error growth
rate due to the model inadequacies is very much greater in the tropics than the mid-latitudes.
This effectively reduces the useful skill forecasts in tropics to about 3–5 days, while the same
for the mid-latitude forecasts remain at about 7 days. Even though the tropics have a lower
limit of atmospheric weather predictability, they have a larger predictability at long-time scales
(of the order of a month to season), since they are much more responsive to the sea surface
temperature anomalies than the mid-latitudes.

REVIEW QUESTIONS

1. What is chaos?
2. What is the difference in the response of the forced pendulum to moderate and large
forcing?
3. What are Lyapunov exponents?
4. What are Poincare sections?
5. What is period doubling route to chaos?
CHAOS IN THE ATMOSPHERE u 409

6. What is a strange attractor?


7. What is Feigenbaum constant?
8. Name any natural system which exhibits chaotic phenomena.
9. What do the Lorenz’s equations physically model?
10. Is the Lorenz system a dissipating system?
11. What is the intermittent route to chaos?
12. What is the significance of the Lorenz attractor to the atmosphere?
13. Will the bouncing ball problem exhibit chaos if the two-ball system is extended to the three-
ball system?
14. Define “deterministic limit of predictability of the atmosphere”. What is its value?
15. Is the skill in deterministic prediction higher or lower in mid-latitudes as compared to the
tropics?
16. What is the typical doubling time of the error growth of the resolvable scales in the mid-
latitudes?
17. Who discovered chaos?
18. Can a linear system exhibit chaotic behaviour?
Useful Universal
Appendix 1 Physical Constants

Acceleration due to gravity at mean sea level g = 9.81 m s–2


Gravitational constant G = 6.673 ´ 10–11 N m2 kg–2
Mean radius of the earth a = 6.371 ´ 106 m
The earth’s mean angular speed of rotation w = 7.29 ´ 10–5 s–1
Universal gas constant R* = 8.314 ´ 103 J K–1 kmol–1
Specific gas constant for dry air Rsp = 287 J K–1 kg–1
Specific gas constant for water vapour Ry = 461 J K–1 kg–1
Avogadro number NA = 6.02 ´ 1026 kmol–1
Planck constant h = 6.63 ´ 10–34 J s
Boltzmann constant k = 1.38 ´ 10–23 J K–1
Speed of light c = 3 ´ 108 m s–1
Stefan Boltzmann constant s = 5.67 ´ 10–8 W m–2 K–4
Solar constant Fs = 1370 W m–2
Mean distance between the sun and the earth 1.5 ´ 1011 m
Mean radius of the sun 6.96 ´ 108 m
Standard sea level pressure p0 = 1013.25 hPa
Standard sea level temperature T0 = 288.15 K
Standard sea level density r0 = 1.225 kg m–3
Molar mass of dry air M = 28.97 kg kmol–1
Molecular weight of water mv = 18.016 kg kmol–1
Density of dry air at standard temperature and pressure (STP) r = 1.29 kg m–3
Density of liquid water at STP rw = 1000 kg m–3
Density of ice at STP ri = 917 kg m–3
411
412 u APPENDIX 1

Specific heat capacity of dry air at STP at constant pressure cp = 1004 J K–1kg–1
Specific heat capacity of dry air at STP at constant volume cv = 717 J K–1kg–1
Specific heat capacity of water vapour at 0°C:
at constant pressure 1850 J K–1kg–1
at constant volume 1390 J K–1kg–1
Specific heat capacity of liquid water at 0°C 4217 J K–1kg–1
Specific heat capacity of ice at 0°C 2106 J K–1kg–1
Specific latent heat of vaporization at 0°C Lv = 2.5 ´ 106 J kg–1
Specific latent heat of vaporization at 100°C L = 2.26 ´ 106 J kg–1
Specific latent heat of fusion at 0°C Lf = 0.33 ´ 106 J kg–1
Specific latent heat of sublimation at 0°C Ls = 2.83 ´ 106 J kg–1
Appendix 2 Vector Identities

1. The dot product is defined as B ¹ C B C DPT ²B C


2. The cross product is defined as B – C ËÍ B C TJO ²B C ÛÝ O is the area of the parallelo-
gram spanned by the vectors a and b. Here n is a unit vector perpendicular to the
plane containing the vectors a and b. The direction of the unit vector is given by the
right-hand rule where one simply points the forefinger of the right-hand in the
direction of a and the middle finger in the direction of b, the resulting unit vector n is
coming out of the thumb. It is to be noted that a ´ b = –b ´ a. Unlike the dot product,
the cross product is special to three dimensions.
3. The triple product B ¹ C – D
has the value of the determinant of the matrix having
a, b and c as row vectors. One has the following identity involving the triple product
B ¹ C – D
C ¹ D – B
D ¹ B – C

4. The triple cross product is defined as


B – C – D
 D – C
– B B ¹ D
C  B ¹ C
D
5. Multiple product identity is written as
B – C
¹ D – E
B ¹ D
C ¹ E
o B ¹ E
C ¹ D

6. The gradient is a vector operation, which operates on a scalar function to produce a


vector whose magnitude is given by the maximum rate of change of the function at the
point of the gradient. The direction of the gradient is pointed in the direction of that
maximum rate of change. In rectangular coordinates the gradient of function f (x, y, z)
is defined as
Ë ˜ ˜ ˜Û
³G ÌJ K L Ü G
Í ˜Y ˜Z ˜[ Ý

413
414 u APPENDIX 2

7. The divergence of a vector field, v = iv1 + jv2 + kv3, in rectangular coordinates is


defined as the scalar product (dot product) of the del operator and the vector. The
divergence of the vector field is indicated as

Ë ˜ ˜ ˜Û ˜W ˜W ˜W


EJW W ³¹W ÌJ  K  L Ü ¹ <JW  KW  LW >  
Í ˜ Y ˜ Z ˜ [Ý ˜Y ˜Z ˜[
The divergence of a vector (if negative) quantifies the tendency of neighbouring
vectors to point towards one another or (if positive) quantifies the tendency of
neighbouring vectors to point away from one another.
8. The curl of a vector field, v = iv1 + jv2 + kv3, in rectangular coordinates is defined as
the vector product (cross product) of the del operator and the vector. The curl of the
vector field is indicated as

Ë ˜W ˜W Û Ë ˜W ˜W Û Ë ˜W ˜W Û


³–W Ì  ÜJ  Ì  Ü K  Ì ˜ Y  ˜Z Ü L
Í ˜ Z ˜ [ Ý Í ˜ [ ˜ Y Ý Í Ý
Unlike the gradient and the divergence operators, the curl operator is special to three-
dimensions.
9. The Laplacian Ñ2f is defined as the divergence of the gradient of a function, i.e.

˜ G
³ G ³ ¹ ³G ǘ YL

L

10. The divergence of the curl is equal to zero, i.e. ³ ¹ ³ – W




11. The curl of the gradient is equal to zero, i.e. Ñ ´ (Ñf ) = 0


12. Vector identities involving the gradient operator:
(i) Ñ( f + g) = Ñf + Ñg
(ii) Ñ(cf ) = cÑf, for any constant c
(iii) Ñ( fg) = f Ñg + gÑf
È G Ø H ³G  G ³H

(iv) ³ É Ù
Ê HÚ H
(v) Ñ(F × G) = F ´ (Ñ ´ G) – (Ñ ´ F) ´ G + (G × Ñ)F + (F × Ñ)G
13. Vector identities involving the divergence operator
(i) Ñ × (F + G) = Ñ × F + Ñ × G
(ii) Ñ × (cF) = c Ñ × F, for any constant c
(iii) Ñ × ( f F) = f Ñ × F + F × Ñf
(iv) Ñ × (F ´ G) = G × (Ñ ´ F) – F × (Ñ ´ G)
14. Vector identities involving the curl operator:
(i) Ñ ´ (F + G) = Ñ ´ F + Ñ ´ G
(ii) Ñ ´ (cF) = cÑ ´ F, for any constant c
(iii) Ñ ´ ( f F ) = f Ñ ´ F + Ñf ´ F
(iv) Ñ ´ (F ´ G) = F(Ñ × G) – (Ñ × F)G + (G × Ñ)F – (F × Ñ)G
APPENDIX 2 u 415

15. Vector identities involving the Laplacian operator:


(i) Ñ2 ( f + g) = Ñ2 f + Ñ2g
(ii) Ñ2 (cf ) = cÑ2f, for any constant c
(iii) Ñ2 ( fg) = f Ñ2g + 2Ñf. Ñg + gÑ2 f
16. Gauss divergence theorem states that the volume integral of the divergence of a vector
function is equal to the integral over the surface of the component normal to the
surface, i.e. vÔ <³ ¹ W>E7 ÔÔ W ¹ E"
4
17. Stokes theorem states that the area integral of the curl of a vector function is equal to
the line integral of the field around the boundary of the area, i.e.

ÔÔ <³ – W> ¹ E" vÔ W ¹ EM


4

18. The gradient operator in cylindrical polar coordinates (x = r cos q, y = r sin q, z = z) is


given by:
˜G  ˜G ˜G
³G FS  FR  F[
˜S S ˜R ˜[
19. The divergence operator in cylindrical polar coordinates is given by:
˜WS WS  ˜WR ˜W[
³¹W   
˜S S S ˜R ˜[
20. The Laplacian operator in cylindrical polar coordinates is given by:

˜G  ˜G  ˜ G ˜G
³ G   
˜S  S ˜S S  ˜R  ˜[ 
21. The gradient operator in spherical polar coordinates (x = r sin q cos j, y = r sin q sin j,
z = z cos q) is given by:
˜G  ˜G ˜G
³G FS  FR  F[
˜S S ˜R ˜[
22. The divergence operator in spherical polar coordinates is given by:

 ˜   ˜WR DPU R  ˜WG


³¹W 
S WS
  WR 
S ˜S S ˜R S S TJO R ˜G
23. The Laplacian operator in spherical polar coordinates is given by:

˜  G  ˜G  ˜ G DPU R ˜G  ˜ G
³ G 
   
    
˜S S ˜S S ˜R S ˜R S TJO R ˜G
Appendix 3 Atmospheric Ozone

Ozone is generally formed in the stratosphere through natural chemical reactions involving
ultraviolet solar radiation and atmospheric oxygen. Solar radiation in the wavelength range
(0.1 mm < l < 0.2 mm) is virtually completely absorbed in the photodissociation reaction
O2 + hn = 2O (A3.1)
The atomic oxygen so produced by the above reaction is a major atmospheric
constituent at levels higher than 100 km. At lower levels, even though atomic oxygen is a
trace constituent, it plays a very important role in the formation of another important trace
substance called ozone. Ozone is formed when atomic oxygen combines with molecular
oxygen, in the presence of a third molecule M, required to carry the excess energy released in
the following reaction:
O2 + O + M = O3 + M (A3.2)
The probability of the above reaction as embodied in Eq. (A3.2) increases in proportion
to the square of the density of the gas. At very low densities, a free oxygen atom can exist
almost indefinitely, while at higher densities its probable lifetime is distinctly short. Due to the
above density dependence, the atomic oxygen is a stable species in the upper mesosphere and
thermosphere, while in stratosphere its lifetime is short as it combines very rapidly to form
ozone. Ultraviolet radiation with wavelengths longer than 0.2 mm is not strongly absorbed by
the photodissociation of oxygen and is, thus, able to penetrate lower into the atmosphere
where it comes across ozone and is duly absorbed in the photodissociation reaction
O3 + hn = O2 + O (A3.3)
The free oxygen atom which results in the above reaction quickly recombines with
molecular oxygen to form another ozone molecule according to Eq. (A3.2). Hence, when
Eqs. (A3.2) and (A3.3) take place in sequence, one after the other, there is no net chemical
change but only an absorption of radiation and a consequent heat input. A sequence of
reactions embodied in Eqs. (A3.2) and (A3.3), repeated many times, the generation of a single
416
APPENDIX 3 u 417

oxygen atom in Eq. (A3.1) ultimately leads to the absorption of many photons of radiation.
Due to the high absorptivity of ozone at l < 0.31 mm, and the rapidity with which the
dissociated ozone molecules are replaced through Eq. (A3.2), even the small trace amounts of
ozone present in the stratosphere are capable of absorbing virtually all the solar radiation in
the range 0.2 mm < l < 0.31 mm. The maximum absorption of solar radiation throughout most
of the wavelength range takes place at a height of 50 km, where there is a large input of
energy per unit mass. The above absorption at 50 km is responsible for the existence of the
stratopause (region of temperature maximum) in the atmosphere. However, the height of
maximum ozone concentration is located at a much lower height of 25 km. Some
stratospheric ozone is transported to the troposphere and can influence ozone amounts near
the earth surface, especially in remote unpolluted regions of the world. Ozone also occurs in
very small amounts at ground level. Ozone can be produced near the earth surface through
chemical reactions between naturally occurring gases and gases that form pollution sources,
i.e. through a reaction between sunlight and volatile organic compounds (VOCs), and nitrogen
oxides (NOx), some of which are produced by human activities. It is to be noted that ground-
level ozone is a component of urban smog—a serious air pollutant.
While the stratospheric ozone blocks harmful incoming solar radiation by absorbing all
radiations with wavelengths less than 0.31 mm all life on earth has adapted to this filtered
solar radiation, the ground level ozone, in contrast, is simply a pollutant. While the ground
level ozone can absorb some incoming solar radiation however, it cannot make up for
stratospheric ozone loss. The amount of ozone above a location on the earth varies naturally
with latitude, season, and from day-to-day. Under normal circumstances, the ozone layer is
thickest over the poles and thinnest around the equator. Also in most situations, the ozone
layer is normally thicker in winter and early spring, and it can typically vary naturally by
about 25% between January and July. Weather conditions can also cause considerable daily
variations. Ozone depletion occurs when the natural balance between the production and
destruction of stratospheric ozone is destroyed in favour of destruction. Observations of an
antarctic ozone hole and atmospheric records indicating seasonal declines in global ozone
levels provide strong evidence that global ozone depletion is indeed occurring. The term
ozone hole refers to a large and rapid decrease in the abundance of ozone molecules, in
Antarctic and not to the complete absence of ozone. The Antarctic ozone hole occurs during
the southern spring between September and November. Based on data collected since the
1950s, the scientists have determined that the ozone levels were relatively stable until the late
1970s. Severe depletion of ozone over the Antarctic has been occurring since 1979 and a
general downturn in global ozone levels has been observed since the early 1980s. Global
ozone levels have generally declined at an average of about 3% between 1979 and 1991. The
above rate of decline is about three times faster than that recorded in the 1970s. Although
natural phenomena can cause temporary ozone loss, chlorine and bromine released from
synthetic compounds are now accepted as the main cause of this depletion. Ozone-depleting
substances (known as ODS) are effective ozone-depleters for two main reasons. The first
reason is that they are not much reactive, chemically speaking, which means that they can
survive long enough in the atmosphere to drift up into the stratosphere. The second reason is
that they aid the natural reactions that destroy ozone. Unlike most chemicals released into the
atmosphere at the earth’s surface, the ODS are not “washed” back to earth by rain or
418 u APPENDIX 3

destroyed in reactions with other chemicals. They simply do not break down in the lower
atmosphere and they can remain in the atmosphere from 20 to 120 years or more. Once the
ODS reach the stratosphere, the ultraviolet radiation in the range of 0.2 mm < l < 0.28 mm
breaks up these molecules into chlorine (from CFCs, methyl chloroform, carbon tetrachloride)
or bromine (from halons, methyl bromide) which, in turn, break up ozone. Both chlorine and
bromine atoms activate and speed up the ozone destruction reactions without getting altered
or destroyed themselves. Hence, a single chlorine atom can destroy up to 100000 ozone
molecules before it finally forms a stable compound and diffuses out of the stratosphere. The
ozone depletion can only be stopped if the concentrations of ozone-destroying chemicals are
reduced, and the natural balance between ozone creation and destruction restored. However,
this might require the complete elimination of CFCs, halons, carbon tetrachloride, methyl
chloroform, HCFCs, and methyl bromide. Some studies in 1991 have indicated that even
with the current global schedule to eliminate ozone-destroying substances, the ozone layer
would not return to normal (pre-1980 levels) until the middle of the 21st century.
Equations of Motion in
Appendix 4 Spherical Coordinates

In this appendix, the governing equations of atmospheric motion in spherical coordinates as


given in Eqs. (6.60)–(6.62) is derived from the Navier Stokes equation in a rotating frame of
reference in Eq. (6.26) together with the expression for the acceleration due to gravity in Eq.
(6.33). In terms of the unit vectors i, j and k (refer Figure 6.9), the velocity vector is defined
as u = ui + vj + wk and the angular velocity vector can be expressed as
w = w ( j cos j + k sin j), where j is the latitude. If one neglects the small variations in the
earth’s rotation, the unit vector representing the angular velocity of the rotation of earth is
given by w /w = j cos j + k sin j is fixed in space
Since, the unit triad vectors i, j, and k change with time, following the motion, the
expression for the otal derivative of say the zonal component of the velocity u is given by,
EV EV EW EX EJ EK EL
J  K L V W L (A4.1)
EU EJ EU EU EU EU EU
The total derivatives of the unit triad vectors can be calculated as follows. Since k = r / r,
where r is the position vector and r is the magnitude of the position vector, one gets
EL E S  ES S ES
 (A4.2)
EU EU S S EU S  EU
ES ES
Since V BOE X Eq. (A4.2) becomes
EU EU
EL V LX 
 VJ  WK
(A4.3)
EU S S S
Since the unit vector w = j cos j + k sin j is constant, the following relation holds:
E EK W TJO K EL W DPT K
 Z DPT K  K  TJO K  L (A4.4)
EU EJ S EU S

419
420 u APPENDIX 4

EK W
since  Substituting of Eq. (A4.3) into Eq. (A4.4), one gets
EU S
EK V UBO K W
 J  L (A4.5)
EU S S
Since by orthogonality of unit vectors, i = j ´ k, one obtains
EJ E K EL
–L  K– (A4.6)
EU EU EU

Substituting from Eqs. (A4.2) and (A4.5) into Eq. (A4.6), and using the orthogonality of unit
vectors, one gets
EJ V UBO K V
K L (A4.7)
EU S S
The Coriolis term in Eq. (6.26) can be written as
Z – V X X DPT K  W TJO K
J  X V TJO K K  X V DPT K L (A4.8)
The pressure gradient force term in Eq. (6.26) can be written in terms of the small
incremental distances dx = r cos j dl in the eastward direction, dy = r dj in the northward
direction and dz = dr in the vertical direction, as follows:
˜Q ˜Q ˜Q
³Q J  K L (A4.9)
˜Y ˜Z ˜[
Combining the centrifugal force and the Newtonian gravitational force into the acceleration
due to gravity, Eq. (6.26) can be written as
EV EW EX V W XË
<V UBO K K  VL >  <V UBO K J  WL >  V VJ  WKÛÝ
S Í Ô
J  K L 
EU EU EU S S
 Î ˜Q ˜Q ˜Q Þ
X X DPT K  W TJO K
J  X V TJO K K  X V DPT K L  Ï J  K L ß  HL ' (A4.10)
S Ð ˜ Y ˜Z ˜ [ à
I
where F = F (x)i + F (y)j + F (z)k is the frictional force per unit mass given by ³ V
S
Rearranging terms in i, j, and k, one gets the Eqs. (6.60)–(6.62).
Appendix 5 Relaxation Methods

Elliptic partial differential equations naturaly arise in science and engineering; of which the
well-known examples are the Poisson and the Helmholtz equations. The above equations in
two dimensions assume the following form:
Ñ2G = F(x, y) [Poisson equation] (A5.1)
2
and Ñ G – HG = F(x, y) [Helmholtz equation] (A5.2)
where F(x, y) is a known forcing function, H is a known positive coefficient and G is the
dependent variable. In order to solve equations (A5.1) and (A5.2) over a region, it is
necessary to know either the dependent variable G or its normal derivative, or a combination
of G and its normal derivative on a curve C enclosing the region. The first boundary
condition is known as Dirichlet condition, the second is known as Neumann condition, while
the third is known as the mixed condition. Relaxation methods are iterative methods to solve
elliptic partial differential equations. In this appendix, the different relaxation methods are
introduced to solve the Poisson equation in two dimensions over a rectangular region covered
by a grid of equally spaced points seperated by a distance d. The simplest finite difference
approximation for the Laplacian ∇ 2 at some point (i, j) for solving Eq. (A5.1) is given as

(J  K  (J  K  (J K   (J K   (JK ³ ' (J K
³(J K (A5.3)
E E
Substituting Eq. (A5.3) in the Poisson equation Eq. (A5.1), one gets
³' (J K G J K

(A5.4)
where f = Fd 2. Assume an initial “guess” of G is made and let Gmi, j represent the mth
estimate. The residual Rmi, j is then defined as
(JN K  (JN K  (JN K   (JN K   (JN K  GJ K 3JN K (A5.5)

421
422 u APPENDIX 5

The objective of the subsequent iteration is to reduce the residual to some acceptably small
value even though the exact solution with Ri,j = 0 everywhere over the region cannot be
achieved. Given the mth estimate, Gmi,j, an improved estimate Gm+1 i, j may be obtained which
will temporarily reduce the residual to zero at (i, j) and as follows:

(JN K  (JN K  (JN K   (JN K    (JN K  GJ K  (A5.6)


Subtract Eq. (A5.6) from Eq. (A5.5) and solving for Gm+1
i, j gives

(JN K (JN K   3JN K


(A5.7)
It can be shown that reducing the residual a given (i, j) point to zero, increases the residual at
the four surrounding points by an amount exactly equal to the correction to (JNK  When the

(m + 1)th approximation is calculated over the entire grid using Eq. (A5.7), the method is
called simultaneous relaxation. The above method, does yield convergence although it may
converge slowly.
Improvements to the simultaneous relaxation method to ensure rapid convergence can be
easily achieved. One easy way of achieving faster convergence is to use the most recent new
estimates of G at preceding points as they are obtained, i.e. use (JN K and (JN K instead of
(JN K and (JNK  in calculating the residual Rmi,j for determination of (JN K in Eq. (A5.7). The

above modification ensures faster convergence and is known as sequential relaxation. It is to


be noted that the residuals have the same sign over at least several grid points. Hence, it is
beneficial to over relax, i.e. add a larger correction than is given in Eq. (A5.7) resulting in the
following relation:
(JN K (JN K  B 3JN K
 … B …  (A5.8)
where a is the over-relaxation coefficient. The above method (Eq. A5.8) provides for faster
convergence and is called as successive over-relaxation (SOR) method.
Appendix 6 Von Neumann Stability Analysis

For illustrating the Von Neumann stability analysis, we utilize the one-dimensional heat
conduction equation which is of the parabolic type, as follows:

˜5
˜5
B (A6.1)
˜U
˜Y 
The explicit finite-difference representation of the above equation is

5 O   5 O 5 O   5 O  5 O 


J J
B J J


J
(A6.2)
'U ' Y

The numerical solution of Eq. (A6.1) is influenced by two sources of error, namely the
discretization error and the round-off error. The discretization error (T ) is the difference
between the exact analytical solution (A) of the patial differential equation, say, Eq. (A6.1)
and the exact round off free solution (D) of the corresponding difference equation, i.e.
Eq. (A6.2). The round-off error (e) is the numerical error introduced after a repetitive number
of calculations in which the digital computer is constantly rounding the numbers due to the
finite representation of numbers. The round-off error can be thought of as the difference
between the numerical solution of the difference equation (N) from a computer with a finite
accuracy and the exact round-off free solution (D) of the corresponding difference equation.
Since the numerical solution N must satisfy the difference equation and also that N = D + e,
one can substitute N for T in Eq. (A6.2), resulting in

% O  F O  % O  F O
J J J J
% O  F O  % O  F O  % O  F O

J J J J J J
(A6.3)
B'U ' Y

423
424 u APPENDIX 6

Since by definition, D is an exact solution of the difference equation, D must satisfy


Eq. (A6.2), i.e.
O 
%  %O %   % O  % o

O O

J J J J J
(A6.4)
B 'U ' Y

Subtracting Eq. (A6.4) from Eq. (A6.3), one gets


O 
FJ  FO O
F J   F O  F O

(A6.5)
J J J

B' U ' Y

From Eq. (A6.5) one sees that the error identically satisfies the difference equation. Since
errors are inevitable in any computer solution, we state that a solution is stable if the errors
remain bounded with integration, i.e. if F O BOE F O  are the error at time steps (n) and (n + 1)
J J

at a grid point (i), then for the solution to be stable, one requires that
O 
FJ
… (A6.6)
FJ
O

We shall assume that the round-off error be represented as a Fourier series in x and also
assume that the time variation is represented by an exponential variation with time (since
errors tend to grow or diminish exponentially with time) as follows:
/ 
F Y U
Ç FYQ BU
FYQ<JL N Y> (A6.7)
N 

È Q Ø
where a is a constant, LN ÉÊ ÙÚ N where m = 1, 2, 3… and L is the length of the domain.
-
The smallest allowable wavelength is 2Dx corresponding to m = N/2, while the largest
allowable wavelength is L corresponding to m =1. Since the original differential equatuion is
linear and since the round-off error satisfies the same difference equation, substitution of
Eq. (A6.7) into the difference equation (A6.5) will lead to the following conclusion; the
behaviour of each term of the series will be same as the series itself. Hence, it is convenient
to consider just one term of the series and write
F N Y U
FYQ BU
FYQ<JL N Y > (A6.8)
Substituting Eq. (A6.8) into Eq. (A6.5), one gets
B U 'U
BU JLN Y  'Y

F F N F F N
JL Y BU JL Y
<F F  F BU FJLN Y  F BU FJLN Y 'Y
>
(A6.9)
B'U 'Y

Dividing both sides of Eq. (A6.9) by F F N and rearranging, and substituting trignometric
BU JL Y

identities, one gets


B 'U L 'Y
F B'   TJO
U
(A6.10)
N

'Y
 
APPENDIX 6 u 425

It is to be noted that
O B U 'U

F N
L Y
F F
J
J B'U
F (A6.11)
F F N
O BU JL Y
F J
Hence, the following condition must be satisfied for the solution to be stable:
O 
F J B'U B 'U L 'Y
O
F  
TJO  N … (A6.12)
F J 'Y

The modulus of the right-hand side of Eq. (A6.10) is called the amplification factor G and
evaluating the inequality of Eq. (A6.12), i.e. G £ 1, we have two possible situations which
must be satisfied simultaneously, namely
B'U L 'Y
F B'U  
TJO … N
(A6.13)
'Y


B 'U L 'Y
and  
TJO N •   (A6.14)
'Y

The inequality Eq. (A6.13) results in
B 'U L 'Y

TJO N •  (A6.15)
'Y

B'U
Since is always positive, the condition expressed in Eq. (A6.15) always holds. From
'Y

Eq. (A6.14), one gets
B'U  LN 'Y

TJO … (A6.16)
'Y


For the above condition given by Eq. (A6.16), to hold, one requires
B'U 
… (A6.17)
'Y
 
For the given explicit finite difference solution of the one-dimensional heat conduction
equation to be stable, the time step Dt must be chosen such that the condition given by
Eq. (A6.17) holds. The above condition severely restricts the time step allowed for the
explicit finite difference scheme and hence a better option would be to go in for a fully
implicit or a semi-implicit scheme for ensuring that the solution remains computationally
stable.
Fortran Computer Program
for Numerical Solution of the
Appendix 7 Barotropic Vorticity Equation

The barotropic vorticity equation was the first model solved with the electronic computer in
the late 1940s and is applicable for the dynamics of both the atmosphere and oceans. The
barotropic vorticity equation is given by

˜[ ˜ ˜
=  VZ [  G
 WZ [  G
(A7.1)
˜U ˜Y ˜Z
˜Z ˜Z ˜G
where z = ³Z  VZ  W Z C (A7.2)
˜Z ˜Y ˜Z

˜Z ˜³Z ˜Z ˜³Z ˜Z
Setting F(x, y, t) =  C (A7.3)
˜Y ˜Z ˜Z ˜Y ˜Y

˜Z Ø
Equation (A7.1) can be rewritten as ³ ÈÉ  ' Y Z U
(A7.4)
Ê ˜U ÙÚ


If F(x, y, t) is known in two-dimensional space, then ¶y /¶t can be found from Eq. (A7.4) at
each point of that two-dimensional space. By using ¶y /¶t and y at some time, a new value
of yÿ can be obtained by a suitable time integration scheme. The finite difference form of Eq.
(A7.4) is given by
Ë  È ˜Z Ø Û
Ì ³ ÉÊ ˜U ÙÚ Ü ' J K
 (A7.5)
Í Ý J K

Ë È ˜Z Ø Û Z  Z  Z  K Z   Z
̳ É
J K J K J J K J K

where

ÙÜ (A7.6)
Í Ê ˜U Ú Ý

J K
E

426
APPENDIX 7 u 427

where x = i d; i = 0, 1, …M and y = j d; j = 0, 1, …. N, d is the grid spacing and


' < Z  K  Z  K 
³ Z


  Z  K  Z  K 
³ Z



E
J K  J J J K J J J K

 Z J  K  Z J  K 
³

Z

J K
 Z J  K  Z J  K 
³

Z

J K
>

C
Z  K Z  K
(A7.7)
E
J J

The finite difference form of Fi,j as seen in Eq. (A7.7) conserves both energy and average
˜Z
vorticity. By making an initial guess of and using Eqs. (A7.6) and (A7.7), Eq. (A7.5)
˜U
˜Z
can be solved using the relaxation method to obtain  The streamfunction can then be
˜U
obtained at a later time by using the central difference approximation

˜Z Z  '  Z ' U U U U

 (A7.8)
˜U  'U
where Dt is the time step. It is important to note that the stability condition (known as
Courant-Friedrichs and Levy condition) need to be satisfied, i.e.
'
…  where M is the
. U
E
maximum of either VZ PS WZ within the region.

c FORTRAN PROGRAM FOR BAROTROPIC VORTICITY EQUATION


DIMENSION PSI(21,9), SLP(21,9), F(21,9), GPT(21,9), R(21,9)
DIMENSION PSI1(21,9), GPT1(21,9), PSI2(21,9), SL(21,9)
REAL K
open (unit = 11,file =’outputstream.dat’)
IMAX = 21
JMAX = 9
D = 1.0E + 5
DT = 90.
TDT = 90.
BETA = 0.16E-10
T = 0.0
TEND = 900.
LMAX = 11
PI = 4.*ATAN (1.)
c A = 1.E + 6
A = 1.E + 0
TPI = PI*PI
DI = TPI/FLOAT (IMAX-1); DJ=TPI/FLOAT (JMAX-1)
C INITIAL VALUES OF STREAMFUNCTION
DO 10 I = 1, IMAX
DO 10 J = 1, JMAX
428 u APPENDIX 7

PSI1 (I, J) = A*SIN ((FLOAT (I)-0.5)*DI)*SIN ((FLOAT (J)-0.5)*DJ)


10 CONTINUE
DO 15 I = 1, IMAX
DO 15 J = 1, JMAX
PSI (I, J) = PSI1 (I,J)
15 CONTINUE
L=0
40 L=L+1
WRITE (11, 24) T
24 FORMAT (1X, F10.2)
DO 20 I = 2, IMAX-1
DO 20 J = 2, JMAX-1
SLP(I, J) = (PSI1(I + 1, J) + PSI1(I, J + 1)+PSI1(I – 1, J) + PSI1(I, J – 1) – 4.*PSI1
1 (I, J))/ (D*D)
20 CONTINUE
DO 30 I = 2, IMAX-1
DO 30 J = 2, JMAX-1
F1 = (PSI1(I + 1, J + 1) – PSI1(I – 1, J + 1))*SLP(I, J + 1)
F2 = (PSI1(I + 1, J – 1) – PSI1(I – 1, J – 1))*SLP(I, J – 1)
F3 = (PSI1(I + 1, J + 1) – PSI1(I + 1, J – 1))*SLP(I + 1, J)
F4 = (PSI1(I – 1, J + 1) – PSI1(I – 1, J – 1))*SLP(I – 1, J)
F (I, J) = (0.25/ (D*D))*(F1 – F2 – F3 + F4) + (0.5*BETA/D)*
1 (PSI1 (I + 1, J) – PSI1 (I – 1, J))
30 CONTINUE
C SOLUTION BY SEQUENTIAL RELAXATION METHOD
C INITIAL GUESS PSI TENDENCY ASSUMED ZERO EVERYWHERE
DO 50 I = 1, IMAX
DO 50 J = 1, JMAX
GPT (I, J) = 0.0
50 CONTINUE
LL = 0
90 LL = LL + 1
DO 60 I = 2, IMAX-1
DO 60 J = 2, JMAX-1
R(I, J) = F(I, J) + (GPT(I + 1, J) + GPT(I, J + 1) + GPT(I – 1, J)+GPT(I, J – 1)-4.*
1 GPT (I, J))/ (D*D )
GPT1(I + 1, J) = GPT(I, J) + D*D* R(I, J)/4.
60 CONTINUE
C CHECK FOR CONVERGENCE
C FIND THE LARGEST ABS. DIFFERENCE OF GPT BETWEEN ITERATIONS
BIG = 1.0E-3
DO 70 I = 2, IMAX-1
DO 70 J = 2, JMAX-1
SL (I, J) = (GPT1 (I, J)-GPT (I, J))
IF (ABS (SL (I, J)). GE. BIG) THEN
BIG = ABS (SL (I, J))
ENDIF
70 CONTINUE
IF (BIG.GT.1.0E-3) THEN
DO 80 I = 2, IMAX-1
APPENDIX 7 u 429

DO 80 J = 2, JMAX-1
GPT (I, J) = GPT1 (I, J)
80 CONTINUE
IF (LL.GT.1000) THEN
GO TO 150
ENDIF
23 FORMAT (1X, E10.2, 1X, I7)
GO TO 90
ENDIF
C CONVERGENCE ATTAINED
C CYCLIC EAST WEST BOUNDARY CONDITION
DO 100 J = 2, JMAX-1
GPT1 (1, J) = GPT1 (IMAX-1, J)
GPT1 (IMAX, J) = GPT1 (2, J)
100 CONTINUE
C TIME MARCHING
IF (L.EQ.1) THEN
DO 110 I = 1, IMAX
DO 110 J = 1, JMAX
PSI1 (I, J) = PSI1 (I, J) + DT*GPT1 (I, J)
110 CONTINUE
WRITE (11, 21) ((PSI1 (I, J), I = 1, IMAX), J = 1, JMAX)
T = T + DT
GO TO 40
ENDIF
IF ((L.NE.1).AND. (L.LT.LMAX)) THEN
DO 120 I = 1, IMAX
DO 120 J = 1, JMAX
PSI2 (I, J) = PSI (I, J) + (2.*DT)*GPT1 (I, J)
120 CONTINUE
T = T + DT
DO 140 I = 1, IMAX
DO 140 J = 1, JMAX
PSI (I, J) = PSI1 (I, J)
PSI1 (I, J) = PSI2 (I, J)
140 CONTINUE
WRITE (11, 21) ((PSI1 (I, J), I = 1, IMAX), J = 1, JMAX)
GO TO 40
21 FORMAT (7(2X, E9.2))
ENDIF
IF (L.EQ.INT (TEND/DT)) STOP
150 END
Fortran Computer Program for
Numerical Solution of the
Appendix 8 Shallow Water Equation

The basic model equations which provide for the shallow water equations are presented in
this section. The following shallow water equations do include the Coriolis term and have
employed potential vorticity in the advection term. The shallow water equations are:
˜V
 I7  GW  ˜) =0 (A8.1)
˜U ˜Y
˜W ˜)
 I6  GV  =0 (A8.2)
˜U ˜Z
˜1 ˜ 6 ˜7
  =0 (A8.3)
˜U ˜Y ˜Z
where P = gh, u and v are the east-west and north-south velocity components, while

  Ë ˜ W ˜V Û
6 1V 7 1W ) < V   W  > BOE I Ì  Ü (A8.4)
 1 Í ˜ Y ˜Z Ý
The following derivatives and averaging properties are used in the finite difference form:

"< Y  'Y
Z >  "< Y  'Y
Z >
K M K M
dx A = (A8.5)
'Y
"< Y Z  'Z
>  "< Y Z  'Z
>
K M K M
dy A = (A8.6)
'Z

A–x = \ "< Y  'Y 
Z >  "< Y  'Y 
Z >^ (A8.7)

K M K M


A–y = \ "< Y Z  'Z 
>  "< Y Z  'Z 
>^ (A8.8)

K M K M

430
APPENDIX 8 u 431

Using Eqs. (A8.5) to (A8.8) in Eqs. (A8.1) to (A8.3), one gets


˜V
 I  Z 7  YZ  G W YZ  E Y ) = 0 (A8.9)
˜U
˜W
 I  Y6  YZ  G V YZ  E Z ) = 0 (A8.10)
˜U
˜1
 E Y6  E Z7 = 0 (A8.11)
˜U
C FORTRAN PROGRAM FOR SHALLOW WATER EQUATION IN F PLANE
PARAMETER (IMAX=32, JMAX=16)
DIMENSION U (IMAX, JMAX), V (IMAX, JMAX), UNEW (IMAX, JMAX), VNEW
1(IMAX,JMAX), P(IMAX,JMAX),PNEW(IMAX,JMAX),UOLD(IMAX,JMAX),VOLD
2(IMAX,JMAX),POLD(IMAX,JMAX),CU(IMAX,JMAX),CV(IMAX,JMAX),Z
3(IMAX, JMAX), H (IMAX, JMAX), PSI (IMAX, JMAX)
OPEN (UNIT=12, FILE=’outputshallow.dat’)
A =1.E + 6
DT = 90.
TIME = 0.0
DX = 1.0E + 5
DY = 1.0E + 5
ALPHA = 0.001
NMAX = 100
PI = 4.*ATAN (1.)
TPI = PI*PI
DI = TPI/FLOAT (IMAX-1)
DJ = TPI/FLOAT (JMAX-1)
F = 2.*7.29E-5*SIN (40.*PI/180.)
C INITIAL VALUES OF STREAMFUNCTION
DO 10 I = 1, IMAX
DO 10 J = 1, JMAX
PSI (I, J) = A*SIN ((FLOAT (I)-0.5)*DI)* SIN ((FLOAT (J)-0.5)*DJ)
10 CONTINUE
21 FORMAT (7(1X, E10.3))
C INITIALIZE VELOCITIES
DO 20 I = 1, IMAX-1
DO 20 J = 1, JMAX-1
U (I + 1, J) = – (PSI (I + 1, J + 1) – PSI (I + 1, J))/DY
V (I, J + 1) = (PSI (I + 1, J + 1) – PSI (I, J + 1))/DX
20 CONTINUE
C PERIODIC CONTINUATION IN X AND Y DIRECTIONS
DO 30 J = 1, JMAX-1
U (1, J) = U (IMAX, J)
V (IMAX, J + 1) = V (1, J + 1)
30 CONTINUE
DO 40 I = 1, IMAX-1
U (I + 1, JMAX) = U (I + 1, 1)
V (I, 1) = V (I, JMAX)
40 CONTINUE
U (1, JMAX) = U (IMAX, 1)
432 u APPENDIX 8

V (IMAX, 1) = V (1, JMAX)


DO 50 I = 1, IMAX
DO 50 J = 1, JMAX
UOLD (I, J) = U (I, J)
VOLD (I, J) = V (I, J)
POLD (I, J) = 50000.
P (I, J) = 50000.
50 CONTINUE
N=0
130 N=N+1
C COMPUTE Pu, Pv, and H
DO 60 I = 1, IMAX-1
DO 60 J = 1, JMAX-1
CU (I + 1, J) = 0.5* (P (I + 1, J) + P (I, J))* U (I + 1, J)
CV (I, J + 1) = 0.5* (P (I, J + 1) + P (I, J))* V (I, J + 1)
Z (I + 1, J + 1) = ((4./DX)*(V(I + 1, J + 1) – V(I, J + 1)) – (4./DY)*
1 (U(I + 1, J + 1) – U(I + 1, J)))/(P(I, J) + P(I + 1, J) + P(I + 1, J + 1) + P(I, J + 1))
H(I, J) = P(I, J) + 0.25*(U(I + 1,J)*U(I + 1,J) + U(I, J)*U(I, J)+
1 V (I, J + 1)*V (I, J + 1) + V (I, J)*V (I, J))
60 CONTINUE

C PERIODIC CONTINUATION IN X AND Y DIRECTIONS


DO 70 J = 1, JMAX-1
CU (1, J) = CU (IMAX, J)
CV (IMAX, J + 1) = CV (1, J + 1)
Z (1, J + 1) = Z (IMAX, J + 1)
H (IMAX, J) = H (1, J)
70 CONTINUE
DO 80 I = 1, IMAX-1
CU (I + 1, JMAX) = CU (I + 1, 1)
CV (I, 1) = CV (I, JMAX)
Z (I + 1, 1) = Z (I + 1, JMAX)
H (I, JMAX) = H (I, 1)
80 CONTINUE
CU (1, JMAX) = CU (IMAX, 1)
CV (IMAX, 1) = CV (1, JMAX)
Z (I+1, 1) = Z (IMAX, JMAX)
H (IMAX, JMAX) = H (1, 1)
C TIME INTEGRATION OF U, V and P EQUATIONS
DO 90 I = 1, IMAX-1
DO 90 J = 1, JMAX-1
UNEW(I + 1, J) = UOLD(I + 1, J) + (TDT/8.)*(Z (I + 1, J + 1) + Z (I + 1, J))*(CV (I + 1, J + 1)
1 + CV (I,J+1)+CV(I,J)+CV(I+1,J))+(TDT/4.)*F*(V (I+1, J+1) +V (I, J+1) +
2 V (I, J) +V (I+1, J))-(TDT/DX)*(H (I+1, J)-H (I, J))
VNEW (I, J+1) = VOLD (I, J+1) - (TDT/8.)*(Z (I+1, J+1) +Z (I, J+1))*(CU (I+1, J+1) +CU
1 (I,J+1)+CU(I,J)+CU(I+1,J))-(TDT/4.)*F*(U (I+1, J+1) +U (I, J+1) +
2 U (I, J) +U (I+1, J))-(TDT/DY)*(H (I, J+1)-H (I, J))
PNEW (I, J) = POLD (I, J) - (TDT/DX)*(CU (I+1, J)-CU (I, J)) - (TDT/DY)*(
1 CV(I,J+1)-CV(I,J))
90 CONTINUE
C PERIODIC CONTINUATION IN X AND Y DIRECTIONS
APPENDIX 8 u 433

DO 100 J = 1, JMAX-1
UNEW (1, J) = UNEW (IMAX, J)
VNEW (IMAX, J + 1) = VNEW (1, J + 1)
PNEW (IMAX, J) = PNEW (1, J)
100 CONTINUE
DO 110 I = 1, IMAX-1
UNEW (I + 1, JMAX) = UNEW (I + 1, 1)
VNEW (I, 1) = VNEW (I, JMAX)
PNEW (I, JMAX) = PNEW (I, 1)
110 CONTINUE
UNEW (1, JMAX) = UNEW (IMAX, 1)
VNEW (IMAX, 1) = VNEW (1, JMAX)
PNEW (IMAX, JMAX) = PNEW (1, 1)
IF (N.LE.NMAX) THEN
TIME = TIME + DT
ELSE
STOP
GO TO 170
ENDIF
IF (N.LE.1) THEN
TDT = TDT + TDT
DO 120 I = 1, IMAX
DO 120 J = 1, JMAX
UOLD (I, J) = U (I, J)
VOLD (I, J) = V (I, J)
POLD (I, J) = P (I, J)
U (I, J) = UNEW (I, J)
V (I, J) = VNEW (I, J)
P (I, J) = PNEW (I, J)
120 CONTINUE
WRITE (12, 21) ((U (I, J), I=1, IMAX), J=1, JMAX)
WRITE (12, 21) ((V (I, J), I=1, IMAX), J=1, JMAX)
WRITE (12, 21) ((P (I, J), I=1, IMAX), J=1, JMAX)
GO TO 130
ELSE
C ASSELIN FILTER
DO 140 I = 1, IMAX-1
DO 140 J = 1, JMAX-1
UOLD(I, J) = U(I, J) – ALPHA*(UNEW(I, J) – 2.*U(I, J) + UOLD(I, J))
VOLD(I, J) = V(I, J) – ALPHA*(VNEW(I, J) – 2.*V(I, J) + VOLD(I, J))
POLD(I, J) = P(I, J) – ALPHA*(PNEW(I,J) – 2.*P(I, J) + POLD(I, J))
U (I, J) = UNEW (I, J)
V (I, J) = VNEW (I, J)
P (I, J) = PNEW (I, J)
140 CONTINUE
C PERIODIC CONTINUATION IN X AND Y DIRECTIONS
DO 150 J = 1, JMAX-1
UOLD (IMAX, J) = UOLD (1, J)
VOLD (IMAX, J) = VOLD (1, J)
POLD (IMAX, J) = POLD (1, J)
U (IMAX, J) = U (1, J)
434 u APPENDIX 8

V (IMAX, J) = V (1, J)
P (IMAX, J) = P (1, J)
150 CONTINUE
DO 160 I = 1, IMAX-1
UOLD (I, JMAX) = UOLD (I, 1)
VOLD (I, JMAX) = VOLD (I, 1)
POLD (I, JMAX) = POLD (I, 1)
U (I, JMAX) = U (I, 1)
V (I, JMAX) = V (I, 1)
P (I, JMAX) = P (I, 1)
160 CONTINUE
UOLD (IMAX, JMAX) = UOLD (I, 1)
VOLD (IMAX, JMAX) = VOLD (I, 1)
POLD (IMAX, JMAX) = POLD (I, 1)
U (IMAX, JMAX) = U (I, 1)
V (IMAX, JMAX) = V (I, 1)
P (IMAX, JMAX) = P (1, 1)
WRITE (12, 21) ((U (I, J), I = 1, IMAX), J = 1, JMAX)
WRITE (12, 21) ((V (I, J), I = 1, IMAX), J = 1, JMAX)
WRITE (12, 21) ((P (I, J), I = 1, IMAX), J = 1, JMAX)
GO TO 130
ENDIF
170 END
Fortran Computer Program
for Numerical Solution of the
Appendix 9 Forced Damped Pendulum

The basic model equation which provides for the forced damped pendulum is

E R H ER

 TJO R  R  "% TJO X % U
(A9.1)
EU M EU
The parameter values used in the program are: q = 0.5, l = g = 9.8, wD = 2/3, time step
Dt = 0.094, i.e. Dt = 3*p /100, AD = 1.2, while the initial conditions are: q (0) = 0.0 and
ER

 The method employed is the well-known fourth-order Runge-Kutta algorithm for
EU
solving a system of first-order ordinary differential equation initial value problem.

C PROGRAM FOR FORCED DAMPED PENDULUM


C
PARAMETER (N = 1000, L = 100)
DIMENSION Y (2, N)
COMMON /CONST/ Q, B, W
PI = 4.0*ATAN (1.0)
H = 3.0*PI/L
Q = 0.5
B = 1.2
W = 2.0/3.0
Y (1, 1) = 0.0
Y (2, 1) = 2.0
OPEN (UNIT = 30, FILE = ‘OUTPUTPENDULUM.DAT’)
C
C Using the Fourth-order Runge-Kutta algorithm to integrate the equation
C
DO 10 I = 1, N – 1
T = H*I
Y1 = Y (1, I)

435
436 u APPENDIX 9

Y2 = Y (2, I)
DK11 = H*G1 (Y1, Y2,T)
DK21 = H*G2 (Y1, Y2,T)
DK12 = H*G1 ((Y1 + DK11/2.0), (Y2 + DK21/2.0), (T + H/2.0))
DK22 = H*G2 ((Y1 + DK11/2.0), (Y2 + DK21/2.0), (T + H/2.0))
DK13 = H*G1 ((Y1 + DK12/2.0), (Y2 + DK22/2.0), (T + H/2.0))
DK23 = H*G2 ((Y1 + DK12/2.0), (Y2 + DK22/2.0), (T + H/2.0))
DK14 = H*G1 ((Y1 + DK13), (Y2 + DK23), (T + H))
DK24 = H*G2 ((Y1 + DK13), (Y2 + DK23), (T + H))
Y (1,I + 1) = Y(1, I) + (DK11 + 2.0*(DK12 + DK13) + DK14)/6.0
Y (2,I + 1) = Y(2, I)+(DK21 + 2.0*(DK22 + DK23) + DK24)/6.0
C
C Ensuring that theta lies between [-p,ÿp]
C
Y (1,I+1) = Y(1,I+1)-2.0*PI*NINT(Y(1,I+1)/(2.0*PI))
10 CONTINUE
WRITE (30, 20) (H*I, Y (1, I), Y (2, I), I=1, N)
STOP
20 FORMAT (3F16.8)
END
C
FUNCTION G1 (Y1, Y2, T)
COMMON /CONST/ Q, B, W
G1 = Y2
RETURN
END
C
FUNCTION G2 (Y1, Y2, T)
COMMON /CONST/ Q, B, W
G2 = –Q*Y2 – SIN (Y1) + B*COS (W*T)
RETURN
END
Fortran Computer Program
for Numerical Solution
Appendix 10 of the Lorenz System

The basic model equations for the Lorenz system are:


EY
= s (y – x) (A10.1)
EU
EZ
= – xz + rx – y (A10.2)
EU
E[
= xy – bz (A10.3)
EU
The parameter values used in the program are: s = 10, b = 8/3, r = 25, time step Dt = 0.01
while the initial conditions are: x(0) = 1.0, y(0)= z(0)= 0. The method employed is the well-
known fourth-order Runge-Kutta algorithm for solving a system of first-order ordinary
differential equation initial value problem.

C FORTRAN PROGRAM FOR THE LORENZ SYSTEM


C
PARAMETER (N = 1000, L = 100)
DIMENSION Y (3, N)
COMMON/CONST/SIGMA, B, R
PI = 4.0*ATAN (1.0)
H = 1.0e-2
SIGMA = 10.0
B = 8./3.
R = 25.0
Y (1, 1) = 1.0
Y (2, 1) = 0.0
Y (3, 1) = 0.0
OPEN (UNIT=30, FILE= ‘OUTPUTLORENTZ.DAT’)
C
C Using the Fourth-order Runge-Kutta algorithm to integrate the equation
C
DO 10 I = 1, N – 1
437
438 u APPENDIX 10

T = H*I
Y1 = Y (1, I)
Y2 = Y (2, I)
Y3 = Y (3, I)
C
DK11 = H*G1 (Y1, Y2, Y3, T)
DK21 = H*G2 (Y1, Y2, Y3, T)
DK31 = H*G3 (Y1, Y2, Y3, T)
C
DK12 = H*G1 ((Y1 + DK11/2.0), (Y2 + DK21/2.0),
1 (Y3 + DK31/2.0), (T + H/2.0))
DK22 = H*G2 ((Y1 + DK11/2.0), (Y2 + DK21/2.0),
1 (Y3 + DK31/2.0), (T + H/2.0))
DK32 = H*G3 ((Y1 + DK11/2.0), (Y2 + DK21/2.0),
1 (Y3 + DK31/2.0), (T + H/2.0))
C
DK13 = H*G1 ((Y1 + DK12/2.0), (Y2 + DK22/2.0),
1 (Y3 + DK32/2.0), (T + H/2.0))
DK23 = H*G2 ((Y1 + DK12/2.0), (Y2 + DK22/2.0),
1 (Y3 + DK32/2.0), (T + H/2.0))
DK33 = H*G3 ((Y1 + DK12/2.0), (Y2 + DK22/2.0),
1 (Y3 + DK32/2.0), (T + H/2.0))
C
DK14 = H*G1 ((Y1 + DK13), (Y2 + DK23),
1 (Y3 + DK33), (T + H))
DK24 = H*G2 ((Y1 + DK13), (Y2 + DK23),
1 (Y3 + DK33),(T + H))
DK34 = H*G3 ((Y1 + DK13), (Y2 + DK23),
1 (Y3 + DK33),(T + H))
C
Y (1, I + 1) = Y (1, I) + (DK11 + 2.0*(DK12 + DK13) + DK14)/6.0
Y (2, I + 1) = Y (2, I) + (DK21 + 2.0*(DK22 + DK23) + DK24)/6.0
Y (3, I + 1) = Y (3, I) + (DK31 + 2.0*(DK32 + DK33) + DK34)/6.0
10 CONTINUE
C
WRITE (30,20) (H*I,Y(1,I),Y(2,I),Y(3,I),I=1,N)
STOP
20 FORMAT (4F16.8)
END
C
FUNCTION G1 (Y1, Y2, Y3, T)
COMMON /CONST/ SIGMA, B, R
G1 = SIGMA*(Y2 – Y1)
RETURN
END
C
FUNCTION G2 (Y1, Y2, Y3, T)
COMMON /CONST/ SIGMA, B, R
G2 = –Y1*Y3 + R*Y1 – Y2
RETURN
END
C
FUNCTION G3 (Y1, Y2, Y3, T)
COMMON /CONST/ SIGMA, B, R
G3 = Y1*Y2-B*Y3
RETURN
END
Bibliography

Aguado, E. and J.E. Burt (2007): Understanding Weather and Climate, 4th ed., Pearson/
Prentice-Hall.
Andrews, D.G. (2000): An Introduction to Atmospheric Physics, Cambridge University Press.
Andrews, D.G., J.R. Holton and C.B. Leovy (1987): Middle Atmospheric Dynamics, Academic Press.
Batchelor, G. (1967): An Introduction to Fluid Dynamics, Cambridge University Press.
Blacakadar, A.K. (1997): Turbulence and Diffusion in the Atmosphere, Springer.
Brown, R. (1999): Fluid Mechanics of the Atmosphere, Academic Press.
Byers, H.R. (1959): General Meteorology, McGraw Hill.
Chamberlain, J.W. and D.M. Hunten (1987): Theory of Planetary Atmosphere, 2nd ed.,
Academic Press.
Dutton, J.A. (1976): The Ceaseless Wind: An Introduction to the Theory of Atmospheric Motion,
McGraw Hill.
Danielson, E.W., J. Levin and E. Abrams (2003): Meteorology, 2nd ed., McGraw Hill.
Gill, A.E. (1982): Atmosphere-Ocean Dynamics, Academic Press.
Goody, R.M. (1995): Principles of Atmospheric Physics and Chemistry, Oxford University Press.
Goody, R.M. and Y.L. Yung (1989): Atmospheric Radiation, 2nd ed., Oxford University Press.
Haltiner, G.J. and R.T. Williams (1980): Numerical Weather Prediction and Dynamic
Meteorology, 2nd ed., Wiley.
Hartmann, D.L. (1994): Global Physical Climatology, Academic Press.
Holton, J.R. (1992): An Introduction to Dynamic Meteorology, 3rd ed., Academic Press.
Houghton, J.T. (2006): The Physics of Atmospheres, 3rd ed., Cambridge University Press.
Houghton, J.T., F.W. Taylor and C.D. Rodgers (1984): Remote Sounding of the Atmospheres,
Cambridge University Press.
439
440 u BIBLIOGRAPHY

Houze, R.A. (1993): Cloud Dynamics, Academic Press.


Jacobson, M.Z. (1999): Fundamentals of Atmospheric Modeling, Cambridge University Press.
Lindzen, R.S. (1990): Dynamics in Atmospheric Physics, Cambridge University Press.
Liou, K. (1980): An Introduction to Atmospheric Radiation, Academic Press.
Lorenz, E.N. (1995): The Essence of Chaos, Washington University Press.
McIlveen, R. (1991): Fundamentals of Weather and Climate, Chapman and Hall.
McIntosh, D.H. and A.S. Thom (1983): Essentials of Meteorology, Taylor and Francis.
Pedlosky, J. (1987): Geophysical Fluid Dynamics, 2nd ed., Springer-Verlag,
Peixoto, J.P. and A.H. Oort (1992): Physics of Climate, American Institute of Physics.
Rogers, R.R. and M.K. Yau (1989): A Short Course in Cloud Physics, 3rd ed., Pergamon.
Salby, M.L. (1996): Fundamentals of Atmospheric Physics, Academic Press.
Sellers, W.D. (1965): Physical Climatology, The University of Chicago Press.
Stephens, G.L. (1994): Remote Sensing of the Lower Atmosphere: An Introduction, Oxford
University Press.
Stull, R.B. (1988): An Introduction to Boundary Layer Meteorology, Kluwer.
Stull, R.B. (2000): Meteorology for Scientists and Engineers, Brooks/Cole.
Tennekes, H. and J.L. Lumley (1972): A First Course in Turbulence, MIT Press.
Trenberth, K.E. (Ed.) (1992): Climate System Modeling, Cambridge University Press.
Tritton, D.J. (1988): Physical Fluid Dynamics, 2nd ed., Oxford University Press.
Visconti, G. (2001): Fundamentals of Physics and Chemistry of the Atmosphere, Springer-Verlag.
Wallace, J.M. and P.V. Hobbs (2005): Atmospheric Science, An Introductory Survey, 2nd ed.,
Academic Press.
Washington, W.M. and C.L. Parkinson (1986): An Introduction to Three-dimensional Climate
Modeling, Oxford University Press.
Index

Absolute Atmospheric
coordinate system, 192 boundary layer, 214
humidity, 47 general circulation model, 387
vorticity, 204, 242 observations, 16
Absorptivity, 85–86 pressure, 24
Acceleration Available potential energy, 336–337
centripetal, 153
Coriolis, 153
Acoustic wave, 239 Backing wind, 178
Adiabatic Baroclinic atmosphere, 241
lapse rate, 53 Barograph, 25
process, 46 Barometer
speed of sound, 250 aneroid, 24
Advection of planetary vorticity, 304 mercury-in-glass, 24
Aerosols, 105–109 Barotropic atmosphere, 193, 241
Air mass, 269–270 Beer’s law, 87–88
Air parcel, 59 Bergeron process, 128–129
Air temperature, 18 b-effect, 305
Aitken nuclei, 106–108 b-plane, 168–169
Albedo, 103 Bifurcation, 399
Altocumulus, 134 Bjerknes’ circulation theorem, 193–194
Altostratus, 134 Black body radiation, 78–81
Aneroid, 24 Body force, 154
Angular momentum, 353–357 Boundary layer (see Atmospheric boundary layer)
Annulus, 358 Boyle’s law, 36
Anticyclonic flow, 173 Breeding method, 385
Atmosphere
baroclinic, 241
barotrophic, 193, 241 CAPE (Convective available potential energy),
heat engine, 343 136
middle, 10 Carnot cycle, 70
441
442 u INDEX

Cartesian coordinate system, 150 Continuum, 13


CCL (Convective condensation level), 135 Convection, cumulus parameterization, 381
CCN (Cloud condensation nuclei), 117–118 Convective
Cell circulation, 405
ferrell, 12, 352 clouds, 135
Hadley, 12, 352 instability, 63
polar, 12, 353 Convergence to solution, 366
Central differencing, 362 Convergence zone
Centrifugal force, 156, 160 ITCZ (Intertropical Convergent Zone), 12, 289
Centripetal acceleration, 153 OCZ (Oceanic Convergent Zone), 289
Chaos TCZ (Tropical Convergent Zone), 289
bouncing ball problem, 400–403 Coordinates
forced pendulum, 391–393 Cartesian, 150
limits of deterministic predictability, 407–408 isobaric, 196–201
Lorenz attractor, 403–407 natural, 170–172
Lyapunov exponent, 394 spherical, 165–167
period doubling route, 397–400 Coriolis force, 156–160
Poincare section, 395–396 Coriolis parameter, 159
Charles’ law, 36–37 Cumulonimbus, 134
Circulation Cumulus, 134
definition, 190–191 Curvature,
general radius of, 171
angular momentum balance, 353–357 vorticity, 196
longitudinally dependent flow, 349 Cyclonic flow, 173
requirement of theory, 350 Cyclostrophic flow, 173
Hadley, 351
secondary, 229–233
theorems, 192–193 Dalton’s law, 37
Walker, 317, 347 Data assimilation, 371–375
Cirrocumulus, 134 Deformation, 190
Cirrostratus, 134 Derivative, total, local and convective, 161–162
Cirrus, 134 Diffusion equation, 363
Claushius Clapeyron equation, 71–72 Discretization error, 422
Climate forecast, 380, 385–388 Dishpan experiments, 358–359
Cloud Divergence, 190
formation mechanism, 132–133 equation, 203
seeding, 137–138 Droplet growth in warm clouds, 118–126
type, 133 Dry air, equation of state, 37–38
Cold front, 272 Dynamic coefficient of viscosity, 215
Computational stability, 366–367, 422–425
Conservation laws, 149
conservation of energy, 178–179 Eddy-turbulent, 220
conservation of mass, 162–165 Eddy,
conservation of momentum, 165–167 exchange coefficient, 224–225
Continuity equation stress, 224
Eulerian approach, 162–163 Ekman layer, 227–229
isobaric coordinates, 199–200 El Nino, 317
Lagrangian approach, 164–165 Electric field, 7
Continuous collision model, 124–126 Elliptic equation, 367–369
INDEX u 443

Emagram, 55 Fronts
Emissivity, 85–86 cold, 272
Energy occluded, 273
conservation of energy, 178–179 stationary, 272–273
internal, 42–44, 334–335 warm, 270–272
potential, 334–335
Ensemble forecasting, 383–385
ENSO (El Nino Southern Oscillation), 316–323 Galerkin method, 375–377
ENSO—Indian monsoon, 323–324 Gas constant
Enthalpy, 44 specific, 38
Entrainment, 65–67 universal, 37
Entropy, 67–69 General circulation (see Circulation, general)
Equatorial Geopotential, 50
inertia gravity wave, 256–260 height, 50
Kelvin wave, 261–262 Geostrophic wind, 169
mixed Rossby gravity wave, 260–261 Gradient wind, 174–175
Rossby wave, 256–260 Gravitational force, 160–161
EQUINOO, 324 Gravity, 160–161
Equivalent potential temperature, 57–59 Gravity waves
Eulerian description, 149 inertia, 256–258
Explicit time differencing, 365–366 internal, 250–252
Extensive, 36 shallow water, 244–246
Extratropical cyclones, 273–275 Group velocity, 241

Finite difference, 362–363 Hadley circulation, 351


Finite element method, 378–380 Hail, 130–131
First law of thermodynamics, 41–42, 178–179 Heat, 41
Flow engine 69–70
cyclostrophic, 173 latent, 49
geostrophic, 169 low, 298
gradient, 174–175 specific, 44–45
inertial, 172–173 Homogeneous nucleation of
Forced pendulum, 391–393 ice particles 127–128
Forces water vapour, 111–114
body, 154 Homogenous atmosphere, 51
centrifugal, 156, 160 Hurricane (see Tropical cyclone)
Coriolis, 156–160 Hydrostatic equilibrium, 49–50
gravitational, 160–161 Hypsometric equation, 51–53
gravity, 160–161
pressure gradient, 154–155
surface, 154 Indian ocean dipole, 323
viscous, 217–219 Inertia gravity wave, 256–258
Forecasting Inertial
climate, 380, 385–388 circle, 173
weather, 380–385 flow, 173
Four-dimensional variational assimilation (4D-Var), frame of reference, 150
372 period, 173
f-plane, 168 Intensive, 36
Friction velocity, 225 Interannual variability of monsoon, 291–292
444 u INDEX

Intermittency route to chaos, 496–497 Mixing length, 223–225


Internal Mixing ratio, 47
energy, 42–44, 334–335 Momentum equation
gravity wave, 250–252 atmospheric boundary layer, 221–223
Intertropical convergent zone, 12, 289 in rotating coordinates, 150–153
Intraseasonal variability of monsoon, 290 scale analysis, 167–168
Ionosphere, 6 spherical coordinates 165–167
Irradiance, 84–85 Monsoons–India
Isallobaric wind, 170 compressibility, rotation and moisture
Isobaric coordinate system, 196–201 effects, 288–289
continuity equation, 199–200 differential heating of land and sea, 287–288
divergence equation, 203 disturbances
geostrophic wind, 200 heat low, 298
horizontal equation of motion, 200 low level jet, 300
thermal wind, 200–201 mascarene high, 300
vorticity equation, 203 mid tropospheric cyclone, 297
monsoon depression 295–297
monsoon trough, 299
Jet stream, 275–278 offshore trough vortex, 298
onset vortex, 297
southern hemisphere equatorial trough,
Kelvin circulation theorem, 192 300
Kelvin wave, 261–262 tibetan anticyclone, 299–300
Kinematic viscosity coefficient, 219 tropical easterly jet, 299
Kinetic energy, 338–342 interannual variability, 291–292
Kirchhoff’s law, 87 intraseasonal variability, 290
Kohler curve, 114–116

Natural coordinate system, 170–172


La Nina, 320–322 Neutral stability, 61
Lagrangian approach, 149 Numerical approximation
continuity equation, 164–165 for one-dimensional diffusion equation,
Land and sea breeze, 194–195 365–366
Lapse rate of temperature for two-dimensional Poisson equation, 367–368
dry adiabatic atmosphere, 53 for one-dimensional wave equation, 369–370
homogeneous atmosphere, 51 Numerical modeling, 360–361
Laws of conservation (see Conservation laws) Numerical weather prediction, 370–375
Linear perturbation, 240
Local derivative, 161
Logarithmic wind profile, 227–229 Orographic waves, 246–248
Longitudinal waves, 248–250 Oscillation
Lorenz attractor, 403–407 Brunt Vaisala, 61, 247
intraseasonal, 290

Mass–conservation, 162–165
Meridional circulation, 351–353 Pendulum day, 173
Mesopause, 10 Perturbation method, 240
Mesoscale, 346 Phase speed, 240
Mesosphere, 10 Planck’s law, 79–81
Middle atmosphere, 10 Planetary vorticity, 204
INDEX u 445

Poisson equation, 367 Shear vorticity, 195–196


Polar front, 268 Shear, vertical, 176–178
Potential energy, 334–335 Shearing stress, 215–217
available, 336–337 Skew T-log p diagram, 56
Potential temperature, 46 Solenoidal term
equivalent, 57–59 in circulation theorem, 192–193
Potential vorticity, 204–207 in vorticity equation, 201–203
Precipitation, measurement, 25–26 Sound waves, 248–250
Predictability, 407–408 Sources and sinks of aerosols, 108–109
Pressure Southern oscillation, 317
definition, 4 Specific heat, 44, 45
gradient force, 154–155 Spectral
measurement, 24–25 irradiance, 84
variation in vertical, 4–5 method, 377–378
Psuedoadiabatic process, 56–57 radiance, 82–83
Spherical coordinates, 165–167
Spin down, 229–232
Radiance, 83 Squall line, 315
Radiation, spectrum, 76–78 Stability
Radius of curvature, 171 computational 366–367, 422–425
Rayleigh-Benard convection, 403 parcel, 59–63
Reference frame slice method, 63–65
inertial, 150 Stocks theorem, 192
noninertial, 150 Stratopause, 10
Reflectivity, 87 Stratosphere, 10
Relative vorticity, 201 Streamfunction, 243
Remote temperature sounding, 94–100 Stuve diagram, 56
Reynolds number, 221 Surface
Reynolds stress, 223 force, 154
Rossby gravity wave, 256–261 layer, 225–226
Rossby number, 159–160 Synoptic scale, 346
Rossby wave, 241–244
Rotating coordinates, 150–153
Roughness length, 226 Temperature
equivalent potential, 57–59
horizontal distribution, 11
Saturated adiabatic lapse rate, 37 measurement, 18–21
Saturation potential, 46
mixing ratio, 47 sea surface, 317–323
specific humidity, 47–48 vertical distribution, 8–11
vapour pressure, 47 virtual, 49
Scale analysis of equations of motion, 167–168 Tephigram, 55
Scattering of solar radiation, 91–92 Thermal wind, 176–178
Mie scattering, 92 Thermocline, 318
Rayleigh scattering, 92 Thermodynamic
Sea breeze, 194–195 diagram, 55–56
Second law of thermodynamics, 67–69 energy equation, 41–42, 178–179
Secondary circulation, 229–233 Thermosphere, 10
Shallow water Thunderstorm, 312–315
equations, 252–256 Tilting term, 201–203
gravity waves, 244–246 Tornado, 316
446 u INDEX

Total planetary, 204


derivative, 161–162 potential, 204–207
potential energy, 334–335 shear, 196
Tropical circulation, 280–281
Tropical cyclones
climatology, 302–303 Walker circulation, 317, 347
Wave equation, 369–370
eye and eye wall, 311–312
Waves
factors responsible for formation, 301–302
acoustic, 239
life cycle, 306–308
equatorial Kelvin, 261–262
movement, 303–306
gravity inertia, 256–258
structure, 308–310
gravity internal, 250–252
Tropopause, 10
gravity shallow water, 244–246
Troposphere, 10
orography, 246–248
Turbulent flow, 219
Rossby, 241–244
equation of mean motion, 221–223
Rossby gravity, 256–261
sound, 248–250
Vapour pressure, 47 Wind
Vertical coordinate, pressure, 196–201 cyclostrophic, 173
Viscosity geostrophic, 169
dynamic, 215 gradient, 174–175
kinematic, 219 isallobaric, 170
Viscous force, 217–219 profile–logarithmic, 227–229
von Karman constant, 226 thermal, 176–178
Vorticity
absolute, 204, 242
curvature, 196 Zonal
equation, 201–203 temperature observed, 347–348.
natural coordinate, 195–196 wind observed, 347–348

Вам также может понравиться