Satellite Oceanography An Introduction For Oceanographers and R
Satellite Oceanography An Introduction For Oceanographers and R
MARINE SCIENCE
Series Editor: Dr. TOM ALLAN,
Institute of Oceanographic Sciences, Wormley,
Surrey
SATELLITE OCEANOGRAPHY
an introduction for oceanographers and
remote-sensing scientists
I. S. ROBINSON,
Lecturer in Physical Oceanography, Department of
Oceanography, University of Southampton
This broad-ranging book bridges the void between
brief review papers on oceanographic applications
of remote sensing and detailed technical reports
about sensors and satellites. The text fills this gap
by providing marine scientists with a broad general
introduction to the subject from first principles
and a thorough explanation of the different tech¬
niques for applying satellite data in oceanography.
It will also give sensor technologists and remote
sensing specialists an oceanographer's perspective of
satellite remote sensing in ocean study. It provides
a view across all types of satellite remote sensing
of the ocean, visible, infrared and microwave
frequencies, and both active and passive sensors.
Pitching the academic level at postgraduate/
senior undergraduate level, and assuming a general
scientific training. Dr. Robinson introduces the
subject from first principles, discussing central
ideas and theories. A section on fundamental
principles follows in a manner accessible to scien¬
tists from all disciplines, with no mathematical
competence assumed beyond that of any numerate
scientist. The next section explores the areas of
application using particular remote sensing tech¬
niques. Included here are certain mechanisms
which the non-physical scientists may not find
simple to grasp: in such cases, the author offers c|ip(iip(iiociio CiiO<^
assistance to the physical understanding by referring
to suitable texts. This is a unique book which
opens an exciting new area and will inform practi¬ Occidental
sing scientists, giving an overview of all types of
College
oceanographic satellite remote sensing. There is
enough information for specialist readers to
understand oceanographic remote sensing pro¬
cesses and their applications, and sufficient breadth
to inform the non-specialist where remote sensing
makes a contribution to marine science. LIBRARY
I. S. ROBINSON, M.A.,Ph.D.
Lecturer in Physical Oceanography
Department of Oceanography
University of Southampton
Distributors:
Australia, New Zealand, South-east Asia:
Jacaranda-Wiley Ltd., Jacaranda Press,
JOHN WILEY & SONS INC.,
G.P.O. Box 859, Brisbane, Queensland 40001, Australia
Canada:
JOHN WILEY & SONS CANADA LIMITED
22 Worcester Road, Rexdale, Ontario, Canada.
Europe, Africa:
JOHN WILEY & SONS LIMITED
Baffins Lane, Chichester, West Sussex, England.
North and South America and the rest of the world:
Halsted Press: a division of
JOHN WILEY & SONS
605 Third Avenue, New York, N.Y. 10016, U.S.A.
COPYRIGHT NOTICE -
All Rights Reserved. No part of this publication may be reproduced, stored in a retrieval
system, or transmitted, in any form or by any means, electronic, mechanical, photocopying,
recording or otherwise, without the permission of Ellis Horwood Limited, Market Cross
House, Cooper Street, Chichester, West Sussex, England.
Table of Contents
Foreword.11
Preface . 13
3.3.3 Fronts.72
3.3.4 Upwelling.74
3.3.5 Mixed-layer dynamics.75
3.3.6 Coastal and estuarine dynamical and topographic
phenomena.76
3.4 Ocean waves.77
3.5 Space and time scales of oceanographic phenomena, in relation to
satellite sampling.84
7.6.3 Upwelling.237
7.6.4 Ocean fronts.237
7.6.5 The Mediterranean Sea.240
7.6.6 Shallow-sea fronts.240
7.6.7 Coastal-fringe fronts.244
7.6.8 Commerical fisheries.246
References.428
Index 444
Foreword
The properties processes and populations of the ocean have always been of
scientific interest and are becoming of greater and greater practical importance.
The traditional uses of the ocean, as a support for ships, as a source of food, as
a sink for waste, become increasingly significant as the population of the world
increases: the concern about climatic change, and the possible effect of man’s
activities, has underlined the major part the ocean plays in the coupled air, ice,
water system.
The ocean itself is a complicated system with movement and variability
on all time and space scales — yet the oceanographers’ traditional instrument,
the ship, gives neither a time series at a fixed point nor a synoptic chart at fixed
time. This, as the author of this book points out, has steered oceanographers
towards problems that can be attacked on the basis of detailed observations
in a vertical column of the ocean. The relatively recent arrays of instrumented
moorings and drifting buoys are so expensive that they can cover only a small
area of the ocean surface.
One technique, adumbrated by some enthusiasts about twenty years ago,
is now seen to have a vast potential for observing the near-surface layer of
the ocean on a global scale - satellite observations are becoming quantitative
and increasingly accurate: they are being used to extend the scope of in situ
observations so as to measure surface variability, to study motions in relation to
space rather than to time. They have already contributed to our knowledge
of the general circulation of the ocean, to meso-scale eddies and upwelling, to
small-scale frontal systems, to internal and to surface waves.
There are two major problems in exploiting remote sensing ofthe ocean. One
is the need to deal with the vastly increased number of observations available;
we must acquire the ability to cope with large data-sets rather than individual
observations. The other is that since the remotely sensed fields are essentially
confined to the sea surface there is the need to relate them to the internal,
three-dimensional structure; there will be a continuing need for measurements
within the ocean.
12 Foreword
To make the most effective use of this rapidly evolving satellite observing
system oceanographers must learn more about the techniques of remote sensing,
so that they can interpret the observations confidently, while the specialist in
remote sensing must learn more about the ocean and its problems.
The author of this book has seen the need for a body of information
common to oceanographers and specialists in remote sensing technology — I am
confident that his account of the problems and the promise will go far to bridge
the gap between the two communities and so will speed the day when satellite
oceanography is fully integrated into marine science.
This book has grown out of my own developing interest in the use of data from
earth-orbiting satellites in the study of dynamical oceanography. As I sought
to understand the techniques employed in remote sensing, and the physical
processes underlying those techniques, I became frustrated at the difficulty of
tracking down information and finding it presented in a form readily accessible
to a physical oceanographer who is not a remote-sensing specialist. There
seemed to be a gap between the brief review papers which did not penetrate far
enough and the technical reports on sensors and satellites in which it was easy to
get lost in a wealth of material, much of it irrelevant to the oceanographer who
is trying to develop satellite remote sensing as a marine science research tool.
At the same time it became apparent that whilst a lot of space technology
was being offered in support of ocean applications, there was not always
evidence that the remote-sensing specialists fully understood the scientific
problem areas being studied by oceanographers.
Faced with the prospect of teaching a course in satellite oceanography
to students of oceanography, and finding no suitable course text, I decided to
attempt to write it myself, if only to encourage others to improve on it.
This book is the result, and it is written for the three kinds of people I
' have mentioned. Firstly, it is for practising oceanographers, to acquaint them
with an exciting new observational tool which is now available to them and
which is beginning to revolutionise branches of marine science. It is both a
background reader providing a broad general introduction and a primer to lead
oceanographers aspiring to use remote sensing through the first principles of
the subject and on into examples of ocean applications, and to leave them
equipped to grapple with the most recent scientific literature. It is also written
in order that sensor technologists and remote-sensing specialists can gain an
oceanographer’s view of the relevance of satellite remote sensing to the study of
the ocean. It therefore includes some background information on those aspects
of ocean science which are benefiting most from this new technique. Thirdly,
14 Preface
as well as being a book for practising scientists, it has been written with post¬
graduate and senior undergraduate students of both oceanography and remote
sensing in mind.
I have therefore tried to pitch the academic level at around that of the
senior undergraduate/postgraduate, assuming a general scientific training and
ability to handle scientific concepts,but aiming to introduce from first principles
those ideas and theories which are central to the subject of the book. The first
section on fundamental principles should be readily accessible to scientists from
all subject disciplines, and no mathematical competence is assumed beyond that
expected of any numerate scientist. The second section, which explores areas
of oceanographic application using particular remote-sensing techniques, is
also intended to be accessible to all scientists. However, such is the complexity
of some of the mechanisms (for example, synthetic aperture radar) that the
non-physical scientist will find it harder to grasp than the physicist, engineer,
or applied mathematician; but I have tried so far as practicable to make the
text self-contained even for the biological oceanographer, marine chemist, or
sedimentologist.
Where help may be needed with physical concepts I have tried to refer to
a suitable text. I have also referenced the key papers which thoroughly describe
the techniques which I am introducing, and mentioned examples of marine
applications in the oceanographic and remote-sensing literature. The reference
list is not an exhaustive bibHography, but should provide an adequate source
of material for a literature search, and a useful springboard for diving into a
detailed review of a particular subject.
I am indebted to many people who have given me help along the way
with the writing and preparation. The book took shape in my mind during a
NATO-funded study visit to North America in 1982 with encouragement from
Pahl Le Blond, Bill Emery, and Jim Simpson. It developed with some guidance
from Tom Allan whose advice as series editor I was fortunate to have at every
stage of the writing. Fellow staff and students at Southampton Department
of Oceanography have taken a lively interest in the work, and colleagues have
kindly commented on parts of the manuscript, notably Simon Boxall, Henry
Charnock, John Daniels, Trevor Guymer, Paul Mather, and Neil Wells. Report
material I had previously written with J. O. Thomas for Oxford Computer
Services was made available for use in the book. Image originals were helpfully
suppUed by K. S. Baker, P. Bayhss, H. Gordon, B. Holt, R. Legeckis, J. G. Marsh,
N. MacFarlane, R. Pingree, R. K. Raney, and R. C. Smith. Permission to copy
other illustrations is acknowledged in the figure captions. I must also thank
those whose labour contributed to the typescript and figures: Elsie Foss, Jenny
Mallinson, Graham Johnson, and especially Jean Watson whose patient days and
hours at the word processor have made the book possible. Finally I am, grateful
to my family for letting me spend time on the book outside my normal working
hours, and most of aU to my wife, Diane, whose encouragement made me believe
Preface 15
I could start writing, and whose patient love and support has enabled me to
complete the task.
Southampton
July 1984
To DIANE
CHAPTER 1
The launch of Seasat in 1978 saw the first satellite specifically designed
for and dedicated to ocean surveillance. To overcome the cloud cover problem,
most of the sensors operated in the microwave part of the electromagnetic
spectrum, in which clouds are transparent. Information was recovered about
surface temperature, surface roughness, sea surface height, and ocean wave
parameters, for a period of three months until an untimely power failure
terminated the satellite’s useful life (see Allan 1983).
In the same year an experimental satellite in the Nimbus series carried a
visible wavelength scanner designed to observe ocean colour — the Coastal Zone
Colour Scanner — which has supplied a wealth of synoptic views of the colour
of large areas of the world’s oceans, at a spatial resolution of better than 1 km.
The same satellite has observed sea surface temperature by microwave
measurement.
Now, and in the coming few years, satellites are being built and plarmed
which will exploit and improve the observational methods already developed to
survey the colour, temperature, surface height, and surface roughness parameters
of the sea. Such is the momentum of space technology development that ocean¬
viewing sensors are likely to be produced irrespective of the degree of coopera¬
tion and collaboration between space technologists and ocean scientists. It is
therefore of importance that oceanographers not only learn how to make full
use of the ocean data reaching us from space, but take an active part in the
development of new space technologies. Indeed, it is hoped that this book will
encourage further the dialogue between the space scientists who promote the
new technologies and the oceanographers who will apply them, by introducing
the techniques of satellite remote sensing to oceanographers and by providing
the satellite technologist with a survey of oceanographic applications.
have been used for many years by conventional oceanographers using sound
waves to penetrate the ocean and view the sea bed, or to observe suspended
sediment or else biological activity in the deep scattering layer. More recently,
acoustic remote-sensing techniques have been developed to detect the thermo-
haline structure of the oceans by acoustic tomography (Flatte et al. 1979) or to
measure velocities by the doppler shift of backscattered sound pulses. There can
be no fundamental objection, therefore, to extending these remote-sensing
principles to the viewing of the sea from space by electromagnetic means through
the intervening layer of atmosphere.
A more fundamental objection to satellite oceanography lies in the fact that
the electromagnetic radiation used by the sensors barely penetrates the sea
surface. Apart from the visible wavelength sensors which may penetrate to
30 m depth, it is effectively true that satellites observe only the surface skin
of the ocean, and are limited to a two-dimensional view. This is a major limita¬
tion since the vertical structure of the ocean is of very great significance to
the oceanographer, whether he is concerned with current velocities, salinity,
temperature and density, distribution of chemical forms, distribution of primary
productivity, or fish populations. If he is on board a research vessel with a
suitable winch and cable, vertical samphng to great depths does not present
insuperable difficulties.
Furthermore, the satellite’s orbit characteristics limit its time-sampling
capabilities at a given location to discrete overpasses once or twice every day, or
every several days, except in the case of the geostationary satellites which can be
parked high above the equator. In comparison with the continuously samphng
or integrating sensor mounted on a research vessel or buoy, the sateUite sensor
begins to appear less effective, particularly if it is further Hmited by cloud cover
which in some locations may be present for nearly 100% of the time.
For both these reasons, there has tended to be a pessimistic view prevalent
amongst many oceanographers that sateUite remote sensing has little of signi¬
ficant value to offer oceanography, with the exception of the use of microwaves
to provide aU-weather sensing of surface phenomena such as height, roughness,
and wave statistics which do not have a depth dependence at all. This is
obviously the task which satelUte remote-sensing performs best, and accounts
for the choice of microwave sensors for Seasat, and the proposed ERS-1.
However, at least three reasons can be advanced to argue that it is extremely
short-sighted to suppose that satellite remote-sensing has no contribution
to make to the study of other ocean parameters which do have a strong depth
dependence or a high-frequency temporal variabiUty which cannot be detected
from a space platform. In the first place, whatever variabUity occurs at depth,
it is the surface parameters of temperature and velocity and of concentrations
of salt and dissolved gases etc., which control the exchange of energy and matter
between the ocean and atmosphere and therefore control the global ocean/
atmosphere heat engine. Thus whilst the satellite may be able to sample at only
Sec. 1.3] The value of satellite data to oceanography 21
one depth level, it does so- at the most important level — the surface. Indeed,
MacIntyre (1977), in seeking to draw attention to the importance of the surface
skin layer of the ocean, points out that on a logarithmic scale the first millimetre
below the surface represents the top half of the ocean! This ‘top half’ is what
is sampled by the satellite. In this, the last quarter of the 20th century, there is
a serious effort being made by both oceanographers and meteorologists to under¬
stand the factors controlling climate, and its variabiHty. Remote sensing of the
surface water parameters promises to contribute significantly in this important
field of scientific study.
The second reason why it is inappropriate to reject remote-sensing tech¬
niques because of hmitations in depth sampUng and high-frequency temporal
sampling capabilities is that to do so is to ignore the positive advantages of
satellite-gathered data — notably its two-dimensional synoptic view and high
spatial resolution (in the case of scanning sensors) and its ability to provide a
low-frequency time series over long periods ranging from weeks to years, even at
isolated oceanic locations. To achieve the synoptic view is impossible by other
means, and the long-period sampHng capabilities would be extremely expensive
by other means. It would require many research vessels to obtain a truly synoptic
survey of a large region, and even then the spatial sampling would be very coarse.
Moreover, since tides with a 121^-hr period are present over most of the ocean, it
is not possible to stretch ‘synoptic’ beyond a period of about 1 hour. For this
reason, the colour and infrared scanning sensors have great potential value, even
if they record only the occasional cloud-free image at irrjegular intervals, for they
supply a synoptic view of the horizontal spatial structure which is impossible
to achieve by other means, and which in fact has not been seen before. In this
novelty of finding new structures and patterns for the first time lies much of
the excitement and enthusiasm shown by other oceanographers for the use of
satelhte data. All too often this enthusiasm hardly persists beyond the adorn¬
ment of the laboratory and office wall with colourful satellite pictures! The
reasons for this may lie in the methodology of science itself. The very questions
that we ask, and the theories we formulate, are only scientifically valid if they
are capable of being tested against data. Therefore conventional, pre-satellite
oceanographic science has tended to develop theories and ask questions con¬
cerning the depth variability and high-frequency time variability of the ocean -
questions which we can answer and theories which can be tested against the sort
of data which ship and buoy sampling methods are good at collecting, that is,
time series and depth profiles. Questions about horizontal spatial variability
and synoptic structures have been kept more in the background. However, now
that we have the means to observe spatial structures, it is important that the
theoretical framework of oceanographic science should develop to start asking
more questions and to formulate more hypotheses concerning spatial structure.
Only as the theoretical base of the science develops will the full value and
significance of synoptic satellite data fields be reahsed.
22 Oceanography from space? [Ch. 1
There are many textbooks of remote-sensing now available, but they are
generally concerned with the application of data collected over land, and are
occupied as much with airborne remote-sensing techniques of air photography
and photogrammetry as with satellite surveillance. The purpose of this book is to
introduce to marine scientists the fundamentals of satellite remote-sensing over
the ocean. Its bias is that of the oceanographer, and the intention throughout
is to present satellite remote-sensing techniques in their context as tools for
oceanographic appUcation. Notwithstanding this, it should also be of interest to
the professional remote-sensing scientist and sensor technologist because remote¬
sensing methods are evaluated in relation to those tasks which oceanographers
genuinely wish to perform, as distinct from those which the space hardware
designer would like him to perform!
The subsequent chapters are divided into two main sections. The first (A) is
a broad review of the fundamentals of satellite remote-sensing, from the oceano¬
grapher s viewpoint. Within this section Chapter 2 provides an overview of
‘the possibilities in space’ - the satellites and orbits available, the different types
of sensors and their capabilities, and the way in which the data are transmitted
to earth and disseminated to applications scientists. Chapter 3 follows with an
introductory survey of ‘the possibilities for oceanography’. This seeks to place
remote sensing in the broad context of ocean science, with a review of the types
of phenomena which oceanographers are studying, a reminder of their length-
and time-scales, and the sampling strategy necessary to measure them, followed
Sec. 1.4] The scope of the book 23
FUNDAMENTALS OF REMOTE
SENSING IN OCEANOGRAPHY
CHAPTER 2
o > 00
1
*n
O'
O'
00
O'
t a>
o 0 OS O' ON Os 0 00
On
a>
H z os
t-H
1-H 00
ON
r-H w> -2 s
cd
Ce:^
CO
< > t
<u <D <u
m S o CO
c/3 d d d
CL os o
O Z d
0 d d
il> <N 04 6
w 04 os SO so so CO
H 04 «-H ON
W
IS oi 04 m
S s a 0 0 0
00
00 g oo
H c^ 00 5 c
o^ ^ S o
< 00
CO Lj ^
c/3 o o c d c
< O OS O cd i<5“
s
W <N o ^ a: ^
(N
c/i 00 !S <c <=>
a>
Satellites for oceanography — general descriptions
(N GO
>
00
r- r-
c/3 Os 00
00
d
o Pi CZl
D *
M E S
pa 52 o
o t o e *3 5 N
S O os cn c/i ^
E
CN 04
o
OS Zl <L>
r- oo
Os
2® oo
■*-»
< On ON
o ic.
O ^ > ■ *
00
d
o E
r o
z o Z C c * .52
o
p<
a: ^
c^ ;3
O ' t ro 04
o CO t r-- m
oi (N <N 00
00
Z Oo 00
H Z Z Z
<N
00 00
H Os 00 d
E
< ^ *
o 2
CO
c/i CO
Q
>> * lo E "O S H
o Functional, but in standby mode, November 1983.
Z lO
E
c U-)
Table 2.1
oo (N fO
r- 00 00
OS
in Os
o 2 00 ^ 00
(U
H t--
r- ^
r- 0)
ON C 2^
00
* Still operational in November 1983.
<L)
o d E CO
H C
IT)
ON
.2 o CO
3 SO «-H ’O o
< a I e
CO ^ tI ^ (N ^ SO
Q ro (N * ro
(N OJ t lO t
Z
< (N
C
C3 op
0) g
D*
‘S .ac/5 Q,
§ a> ^ O
<U Vh ^ d d
I 2 =
(U ■3 J2 o
ft g
^ (D
c/5 00 cd
cd E o o
e .d d
X bO
.yCL 3CT" CL
'n ^
cd
o
2 -S >> D
0) CL H •-
GO CU CL
O < 1>
28 The possibilities in space — space hardware and data transmission [Ch. 2
AHITUDE CONTROL
^SUBSYSTEM PACKAGE
ORBIT DIRECTION
SUN SENSOR
DETECTOR
INERTIAL
MEASUREMENT
UNIT
INSTRUMENT
MOUNTING
PLATFORM
SUNSHADE
INSTRUMENT
MOUNTING
PLATFORM
HYDRAZINE
TANK (21
ADVANCED
REACTION VERY HIGH
SYSTEM RESOLUTION
SUPPORT RADIOMETER
STRUCTURE
STRATOSPHERIC
BATTERY SOUNDING UNIT
MODULES 14)
Sec. 2.1]
(c)
TRANET beacon
antenna
Telemetry,
r tracking &
; command Sensor
module
6.1 m
SAR antenna
SASS
antennas
Laser retroreflector
SAR data link VIRR \
(d) antenna '-TTAC antenna
Fig. 2.1 - Sketch of some satellites used for oceanographic purposes (not
necessarily to the same scale), (a) Landsat 1 (b) TIROS-N (c) Nimbus 7
(d) Seasat
30 The possibilities in space — space hardware and data transmission [Ch. 2
solar cells, backed up by batteries to store energy for when the satellite is on the
night-time side of the earth. Typical power requirements on operational satellites
are given in Table 2.1. In general, sensors with larger power demands must be
mounted on a larger bus which is able to support the larger solar cells which are
required.
There are relatively strict temperature ranges under which the sensors and
other electronic equipment will operate accurately, and care must be taken with
the design of the satellite to prevent overheating. Heat generated by electronic
equipment, or absorbed from incident radiation, must be matched by long¬
wave radiation emission into space, which is dependent on the choice of surface
coating. Active control of the temperature can be achieved by varying the net
long-wave radiation, through attitude control, or by the operation of shutters to
increase or decrease the area of radiation surfaces pointing towards cold space.
The satellites’ own environmental control systems therefore need to be sophis¬
ticated and reliable for a long operational life, and designs have evolved and
improved with experience. Within series of satellites the basic design is the same
although the sensors carried and the communications system may differ between
satellites. Amonst oceanographically important satellites, the Landsat and
Nimbus platforms are of the same basic design, whilst the TIROS/NOAA
meteorological satellites form a different series which includes the Defence
Meteorological Satellite Programme (DMSP). The GOES-Meteosat vehicles
have very different design criteria, being geostationary. Seasat was yet another
design, but based on the earlier Agena system. There are clearly advantages in
being able to test new sensors on a proven vehicle, but this is not always possible,
and when ERS-1 is launched later in this decade most of the satellite subsystems
as well as the sensors will be completely new designs.
One of the subsystems which differs from series to series is the Aspect
Control. A space vehicle’s orientation may need to be controlled for a number of
reasons. Polar orbiting satellites such as Landsat, Nimbus, and NOAA require a
slow spin so that the same face always points down towards the centre of the
earth. In contrast, the GOES vehicles in geostationary orbit spin once every six-
hundred milliseconds on an axis parallel to the lines of longitude at the earth’s
equator to provide a built-in east/west scan for the sensors. Precise spin rates can
be maintained by alteration of the mass distribution of the satellite and hence its
moment of inertia. Other methods of attitude control include inertial systems
which torque with respect to the earth’s magnetic field, and reaction jets, but these
have a limited endurance. Jet thrusters are required to produce orbit corrections,
or alterations if the mission demands it as in the case of Seasat. The more inher¬
ently stable the platform is, and the greater the natural damping of unwanted oscil¬
lations and rotations, the longer its useful life will be. Atmospheric drag would
reduce the operational life, and necessitate frequent course corrections for an op¬
erational satellite, but fortunately at altitudes above 300 km it is very small, and
lifetimes of several years are achieved provided no major system faults develop.
Sec. 2.1] Space platforms and their orbits 31
From Newtonian dynamics, the period T for the satellite to travel round
the orbit is
T = (2-2)
where G is the constant of gravitation and M the mass of the earth, and
GAf= 3.98603 X 10^'* m^s“^. The instantaneous rate at which the satellite
describes its orbit is
For a circular orbit centred on the earth, e = 0 a = r. In this case the horizontal
speed of the satelhte is
Vo = iGM/af\ (2.4)
In terms of the height h above the earth (Fig. 2.3) and the acceleration due to
gravity at the earth’s surface g = GM/R^, where R = 6378 km is the earth’s
mean equatorial radius, (2.4) may be written
During launch the rocket must be fired to obtain a trajectory such that at the
desired height h of the satellite, its speed is Fq, assuming that a circular orbit is
required.
satellite orbit
The apparent satellite orbit, as viewed from the surface of the earth, may
appear very different because the earth is rotating on its axis beneath the celestial
shell once every day, from west to east. For earth observation, two types of orbit
are most useful.
A geostutionaiy orbit is achieved if the satellite orbits in the same
direction as the earth with a period of one day. If it is positioned in a circular
orbit above the equator it becomes stationary relative to the earth and there¬
fore always views the same area of the earth’s surface. From (2.2) with T as
1 day = 86 400 seconds, we obtain a = 42 290 km, and therefore for a
geostationary orbit /2 = a — /? = 35910 km.
A much lower orbit has a much shorter period. Satellites at a height of
The possibilities in space — space hardware and data transmission [Ch. 2
between 500 and 2000 km are normally placed in a nearly circular polar orbit
which carries them approximately over the poles with a period of about 1 to
2 hours. As the earth rotates under this orbit the satellite effectively scans from
north to south over one face and south to north across the other face of the
earth, several times each day, achieving much greater surface coverage than if it
were in a non-polar orbit.
The simple expressions presented above to represent orbital dymanics are
based on the assumption of spherical symmetry and zero drag. This is not the
case in practice, and perturbations therefore occur in the elliptical orbit. The
drag of the atmosphere on low orbits has been briefly mentioned and will result
in the acceleration and eventual burn-up of vehicles in low orbit. However, for
the operational life of ocean-monitoring polar orbiters and geostationary
satellites, the effect can be neglected, as can the effect of solar wind and radia¬
tion unless the vehicle has a very low density and a large surface area. The
principal deviation from a pure elliptical orbit is due to non-symmetrical gravity
forces owing to the irregular figure and mass distribution of the earth. Solar and
lunar gravitation is not important, but the equatorial bulge of the earth produces
the major effect. Orbit dynamics with these extra influences are discussed in
more detail by King-Hele (1959, 1964). It is shown that the equatorial bulge
causes a slight change to the period, and results in the perigee of an elliptical
orbit changing position with time, that is, dw/dt 0. Of most interest in our
context is the precessional mechanism which causes the orbital plane itself to
rotate, that is, df2/dr 0. In fact
o o o
00 VO o 04
‘O
<N CN
On
d
VO
VO 04 CO
CO
CO o o o
W 00 00 o VO
DS o^ VO
(N
ON On
VO d rT o
CO
o o^
< CO
g O o o
O 00 o VO
H On 00 o-
< (N VO d o
04
H CO
IZl
O O o O
w
ro o oo
(S CO o-
o CO <N d CM
VO CM
CO
CM O o
U o o
H
< (N o
.ss CO fS VO CM
CO
0) O
'M u
o o o
cd H
W _
o
Tj-
o
o
<N VO CM
CO
H
<C 00 ^ ^ -H CO
o
< ^
■S
to 00 C--
o o 00 o
wd < 00 o o C =«
zo W
C--
CO
CO
J3
&
p lo CO
o CO o o»
CQ
IH «I CO
CO
CN
CO
»o ON
o
^ o CO
s ON ON
SO r--
o
o CN
o
■ o-
a> CO m °
u VO CM
(N <N 00
o 00 oo ON o o
< r-
>1 D
<2 O
z o
z <N (21) CM
VO
</)
O) o O
<N
00
00
oi r-
00 ON
X
a> u
-I-* CO
« z ON
t/3 §z (N
00
CM m °
t S/B = southbound overpass, N/B = northbound overpass.
CO <N <N o
I r- r- ON
fS z
ri p
CO
VO
04 (On
VO «
VO O'
05
oo
o o OO 00 C/2g
ON
CQ <
r~“
ON
H J 00
O 00
Co lo CO s®. ^
H o ON
ai < 00
CN
ON ON O
^ g CO
C/D
< Q
w CO
2:
Z < VO
00
«s
o CO
OO ON c/5 g
On On
o CO
CN
r-
00
O
00 CQ
lo CO
00
o 00 o cog
On ON CO
<N
(/3
c
4S g 0) o >»
cd
•w 0) Cd
o I § §•»- 'O
XJ c
(U
CO <u .2.
c?^
E .S
cd
3
(D >»
l-H
.
>. D ^ O B -2
cr M'S .3 )-< *s a
3 E « s
■*cd->
(1> "cd S c 4_. 0) 5i) cd ,
X)
O
C/D
S 4 .S "
z
c o
<u
II
O u- c
2 "E
o t-H
S «
S (U
>—j
C/D ^ o (X
o 3
z C/5
36 The possibilities in space - space hardware and data transmission [Ch. 2
satellites. With a period of about 103.2 minutes there are almost 14 orbits a day,
and therefore 14 southward sweeps across the earth in daylight. To calculate the
distance between successive ground tracks we simply calculate the longitude
through which the sun has moved in 103.2 minutes, that is, about 25.8 degrees.
(Were the orbit not sun-synchronous, then the calculation would be more
complicated, taking into account both the precession of the orbit and the fact
that the earth rotates through nearly 361° per day relative to the stars. For the
sun-synchronous orbit these two effects cancel exactly.) This corresponds to a
spacing at the equator of about 2865 km. After 14 orbits, the ground track has
traversed 361.43 degrees of longitude, so that orbit 15, or the first on the second
day, is shifted 1.43° to the west of orbit 1 on the first day, that is, 159.4 km to
the west at the equator. After 18 days (that is, 251 orbits) the ground track
starts to repeat itself, so that day 19 will be nominally the same as dayl. In
practice, small orbit perturbations due to non-sphericity, and course correction
by the gas-reaction jet system, result in small deviations of a few kilometres
from exact repeatability, but over a year the repeat coverage over the 18-day
cycle is accurate to within about 30 km.
The sampling rates of the remotely-sensed data depend not only on the
orbit but also on the swath-width of the sensor. Thus for the multispectral
scanner on Landsats 1 to 3, with a swath of 185 km, one day’s daytime observa¬
tions will only cover 14 strips 185 km wide (Fig. 2.8). It will take 18 days to
fill in all the gaps, and the sampling period is therefore 18 days. If a particular
location falls in the small overlap zone between strips (which increases in width
with latitude) then sampling will occur on two successive days in every 18.
DAY 2. ORBIT A
Landsat 4 illustrates how a slight change in orbit pattern can alter the
samphng characteristics. It will be seen from Table 2.2 that it is rather lower
Sec. 2.1] Space platforms and their orbits 39
than the previous Landsat series, and therefore has a shorter orbital period.
Successive orbits are separated by 24.72° (that is, 2752 km at the equator), and
on the second day the tracks lie about halfway between those of the previous
day (Fig. 2.9). The repeat period is now 16 days (233 orbits), and it will be seen
from Fig. 2.9 that locations in the overlap zone will be resampled after 7 days
(on the western edge of the swath) or 9 days (on the eastern margin).
Fig. 2.9 - Orbit repeat cycle pattern of Landsat 4. Orbit numbers are shown,
counting from orbit 0 on day 1 at the eastern edge of the diagram.
Fig. 2.11 also indicates which parts of the spectrum are used by remote¬
sensing sensors on satellites. The choice of bands is governed firstly by the
atmospheric transmission spectrum, and secondly by the application. Fig. 2.12
illustrates the shape of the emission spectra for blackbody radiation emitted
from a body due to thermal excitation. The source at a temperature of 6000 K
corresponds to the sun, and at 300 K to the earth’s surface. Thus if features of
the land or sea are to be observed by the reflection of incident solar radiation
in the same way as the human eye observes, then the frequency range of high-
energy solar radiation should be used, between 100 nm and 100 qm, peaking in
the visible range. Alternatively, if the self-emission of radiation by the sea is to
be the means of remote sensing, sensors should be used for the 3 to 40 qm wave-
WAVE DESCRIPTION BAND NAME REMOTE SENSING APPLICATION
42
ZH‘AON3nD3aJ
UJ‘H10N3T3AVM
The possibilities in space — space hardware and data transmission
[Ch. 2
Fig. 2.11 — The electromagnetic spectrum, showing some band definitions and typical remote-sensing applications
Sec. 2.2] Sensors on satellites 43
m
O
z
<
2
in
>
o
o
m
o
<
_i
m
length range. However, not all the parts of these ranges are useful, since the
atmosphere will not transmit them, as illustrated by the typical transmission
spectrum of the atmosphere drawn in Fig. 2.13. The atmosphere (that is, both
the air itself and water vapour and other aerosols) absorbs most radiation with
a wavelength less than 400 nm and much above about 1 fim. There is a definite
window at the visible wavelength range, and also at select bands within the
infrared. These bands have therefore been exploited for remote sensing, leading
to the visible/near infrared and thermal infrared families of sensors (Fig. 2.11).
Between around 20 /xm up to around 10 m wavelengths there is more absorption,
but above this (that is, for frequencies less than around 100 GHz) there is very
httle absorption. Although there is little reflected or emitted energy at such high
frequencies, these radar bands are exploited by the active microwave sensors
which create their own radiation with which to illuminate the target, and then
observe the nature of the reflected signal. These sensors are known as active
devices as distinct from the passive IR and visible wavelength sensors which rely
on naturally occurring radiation. According to their configuration and applica¬
tion, these active microwave devices are known as altimeters, scatterometers, or
synthetic aperture radars. In addition to active microwave devices, passive
microwave radiometers have also been developed to observe the low level of
thermally emitted microwave radiation.
The possibilities in space — space hardware and data transmission [Ch. 2
aoNviiii^SNvai
Sec. 2.2] Sensors on satellites 45
f A more precise definition of radiant flux quantities, units, and symbols will be found in
section 6.2 in the context of visible light, although the principles apply across the whole
of the e.m. spectrum.
46 The possibilities in space — space hardware and data transmission [Ch. 2
surprising that its electronic counterpart has been developed. This is the multi-
spectral radiometer in which the incoming radiation is measured within several
discrete spectral windows in order to describe its spectral composition or colour.
The spectral windows depend on both the spectral response of the electronic
sensing device, which may be tuned to a particular narrow frequency band, and
the spectral characteristics of any filters used. Whilst the term ‘colour’ is given
to the information obtained, it is not likely to correspond exactly to colour as
registered by the human eye, since the wavebands are likely to be different.
Multispectral sensors are also not limited to three bands, and some may have
many more. In the ocean-viewing context the colour or ‘spectral signature’ of
the radiation as measured by multispectral radiometers is used in two ways.
Firstly, it carries useful information about environmental parameters at sea, and
this is particularly so in the visible/near infrared when the energy is reflected.
Secondly, it provides a very useful way of distinguishing between variability
due to changes in the ocean parameters and variability due to changes in the
atmosphere through which the radiation travels from the sea to the sensor.
This is the principal benefit of multispectral data in the thermal infrared and
microwave regions of the spectrum.
Passive sensors are limited to the measurements of radiance and colour, and
possibly polarisation although the only ocean-monitoring technique so far
developed to exploit the latter operates in the microwave. Since passive sensors
depend on either solar illumination or emitted radiation from the sea, their
energy source generates continuous incoherent electromagnetic radiation, and
there is therefore no chance of obtaining phase or wave speed information. Since
active sensors illuminate their target (the sea) with their own pulse of electro¬
magnetic radiation, there is scope to measure not only the amplitude but also
the phase of the reflected signal, and the travel time of the pulse, both of which
can be made to yield further information about the sea surface when suitably
analysed. All the active sensors so far developed for satellites operate in the
microwave part of the spectrum (0.1 to 1000 GHz), because of the atmospheric
transparency there. The altimeter measures the travel time of a radar pulse to
estimate the height of the sea surface, whilst the difference between the shape
of the outward and return pulses can be interpreted in terms of the sea surface
wave height. The synthetic aperture radar (SAR) uses the ampHtude of the
return signal as a measure of the radar reflectance (and hence roughness) of
the sea surface, but it is able to gain a very high spatial resolution by using the
travel time of the return pulse as a measure of the distance between the point
being illuminated and the satellite subtrack, and by using the shift in frequency
of the return pulse compared with the emitted as a measure of doppler shift, and
hence of position in the along track direction (see Chapter 12).
Another property of electromagnetic radiation which is capable of being
measured is the polarisation, if any, of the electric and magnetic vector. Because
a rough sea will reflect incident radar energy in differing amounts according to
Sec. 2.2] Sensors on satellites 47
radiation so easily, and the IFOV is likely to be circular and less precisely defined,
except in the case of the SAR where the phase and frequency information of the
received signal is used to effectively synthesise the small discrete field of view
and hence to achieve a fine spatial resolution.
Some sensors have a wide IFOV and simply make one downward looking
observation at periodic intervals during the overflight of the satellite,
to give an average value of radiation over a wide area of earth surface, typically
50 km X 50 km or more centred at the sateUite sub-point. The altimeter on
Seasat is an example of this type of field of view. Other sensors, particularly
those operating in the visible and infrared wavelengths, are sufficiently sensitive
to resolve the small amounts of energy to be found in much narrower fields of
view. Thus the AVHRR (Advanced Very High Resolution Radiometer) on the
NOAA satelhtes which measures in the visible and infrared, or the CZCS (Coastal
Zone Colour Scanner) on Nimbus 7, measuring in the visible and near infrared,
have IFOV of about 1 km square or 800 metres square respectively. The visible
wavelength MSS (MultiSpectral Scanner) on Landsat has an IFOV of about
80 metres sqauare, while the thematic mapper (TM) on Landsat has an IFOV
as small as 30 metres square. To produce areal coverage greater than a single
swath of 1 IFOV width around the globe on each satellite orbit, the sensors are
made to scan across the satellite tracking directions. For polar orbits the scan
timing is arranged so that subsequent scans occur after the satellite has travelled
a distance approximately equal to the IFOV projection on the satelhte subtrack
(Fig. 2.15). The radiometer is sampled at a rate so that subsequent IFOVs just
overlap along the scan direction, and there are typically 2000 to 4000 samples
along a scan line. In this way the whole area within a wide swath is efficiently
SATELLITE
track
SCAN DIRECTION -^
~~ _ . - --
~~~ —
- 1''°J1-SENS.N^SC^AN R^TURN)
—-——
1 IFOV '
1
sampled once only. The scanning is usually achieved by a rotating mirror, and
may be performed in one direction only or backwards and forwards across the
swath depending on the sensor design. In the case of the proposed Along Track
Scanning Radiometer on the European ERS 1 satellite, a mirror will rotate about
an oblique axis to describe curved scan hnes, crossing beneath the satellite in one
direction and returning back across the swath at an angle 60° in front of or
behind the satellite. In the case of the geostationary satellites (for example
Meteosat) the whole satellite rotates about its axis to achieve scanning parallel
to lines of latitude, whilst the sensor’s field of view is rotated north-south to
point at different latitudes on each satelhte rotation, thus achieving a coverage
of the whole face of the globe as visible from that location in space.
The details of individual sensor geometries will be examined in the applica¬
tions chapters, but some general comments concerning spatial resolution can be
made here. It should be noted that the spatial resolution which is possible for a
given sensor is dictated by the IFOV and not by the sampling rate. It is the
latter which normally dictates the pixel (picture element) spacing, but this is
not the same as the spatial resolution. This is illustrated in Fig. 2.16. In each
case the IFOV and hence the true spatial resolution is the same. However,
whilst there is efficient matching of sampling range and IFOV size in (a), in (b)
the low sampling rates lead to much larger pixel spacing, but since the IFOV is
now smaller than the pixels there are significant areas of ground unseen by the
satellite. In (c) the sampling is more frequent than it need be, with overlapping
IFOVs. Although the pixels are much smaller this does not mean that the
spatial resolution is improved from (a), since objects in the overlap region will
contribute to more than one pixel, thus blurring their position. Careful numerical
analysis of the data, bearing in mind the point spread function of the sensor,
may be able to achieve higher resolution in case (c), but in terms of simply
reproducing an image from the data recorded for individual pixels, case (c)will
not give more spatial detail than (a).
The way in which the sensor integrates the incoming radiation over the
IFOV may be a distinct advantage in oceanographic applications, where the
average conditions over an area are often of more interest than the specific
conditions at a point as measured by a ship, particularly if there is significant
variability over short length scales. On the other hand this can be a disadvantage
if there is a single or a few discrete anomalously bright points in the field of
view, for example, a bright white object such as a ship (or even a sun reflection
from a small but suitably angled reflecting surface on a ship) in the the field of
view of a visible wavelength sensor, or a corner-cube reflector on a buoy in the
IFOV of an active radar sensor. In these cases the radiant energy recorded for
that pixel may be so anomalously high that it must be rejected as a false record
for the general sea conditions, and a data point is lost. Even worse, the data
value may be higher than appropriate for the actual sea conditions, but not
sufficiently high to be rejected, leading to interpretation error.
50 The possibilities in space — space hardware and data transmission [Ch. 2
IFOV
PIXEL
IFOV
^-
PIXEL
IFOV
PIXEL
Finally, mention must be made of the way the ground resolution varies with
position along the scan line. Fig. 2.17 illustrates this. Because resolution and the
IFOV are based on angles, the metric ground resolution varies with the distance
from the satellite and the look angle relative to the local vertical. On scanners
such as the AVHRR and CZCS which cover a swath of several thousand kilo-
SENSOR
metres this is exacerbated by the curve of the earth too. Thus the resolution
length scale at the extremes of the scan may be several times longer than that
at the satelhte sub-point. Furthermore, if the IFOV is square, the projection
from the sea surface will be square at the sub-point but increasingly rectangular
towards the extremes of the scan line, the longer axis being along the scan line
direction. This degradation of resolution with obUque viewing is particularly to
be borne in mind because although the pixels are equally distorted in the scan
direction, in the satellite track direction they remain equally spaced, or on scales
where earth curvature effects are important they even become closer towards the
scan extremes. Thus there is considerable overlap of IFOVs from one scan line
to the next, which is not immediately apparent from the geometrical distortion
of the image data which will be discussed in Chapter 5.
o
c/3 c/3 X
C3 <L) 00 00
o •+-»
C/D cd x; 3 X X
u. 00 >> 00 00
7d cd o
o -4-»
Cd
*Sh nd
>>
H
high-resolution scanner
XI I (O ' o <u
«_ 00 ix 00
tJO c/3 (U cd <D Cd c
narrow-swath
<D c p
C <D c > cd
O ’o. I 0) cd > cd o
o o o o
a C X5 c/5 o o
cd 173 cd
&0 c/3 O k- cd
<L> <u
a ’> ■4-» c/3 (1>
cd cd
cd
t-H > cd
Satellite sensors used for oceanographic applications
g o
CL.
'o
Cd O. <u
cd C
C .3^
o o
i 2 T3
C/D £
I ^
(U (U a>
passive
a> (U
? > > > > > >
o n) cd cd cd
■< Cl. n) Cl. CX a- O.
<D Pi
>
CTj
near IR
<u Pi ^ jj Pi ^
c:: a> S
visible
& 173 03 3 cd cd
o 1- P ^
c/3 cd
c
C a
X ii c
cd g
« 5> bH
(U
£ 2 §
<= x:
^ p ^ X > P «
73 00
LANDSAT
H z t
<i> c/3 <; I M3
c/3
> W c/3 c/3 D s
O) O < C/D c/3 o <
<
CQ
CJ
3 O W Pi S X
o on w w
H z
o
cd
C/D
Table 2.3
00
Multi-spectral scanner
a> 00
CX .3 oo o p .3
>> c a.
c £ a Ou
Cd "o
O <l> o <u WH a S
<u o c/5 a> ^ cd (D (U
P
s
cd a > Pi
p
oN Pcda 1
• -t
o o
a
^
C ^ J <D —H O cd
o o
00 i-i C S-J cd 2
c cd 3 o ^
CJ o > o Cd
T3 C/D O •*-»
cd
C/D < o U 0)
Pi x
O a
>L Pi
cd c H Di Pi C/3 Pi
MSS
•? o hJ C/D CJ s
(D ok-l H K N
l-H CJ
X) < U E
.£3
<d
Sec. 2.3] Data retrieval 53
(U s
00 e
o 00
'O
.2
O
>-1 <u d>
X
0)
2
d d>
T3 o
d) o
4=1
-+-»
O W) H-*
a
<30
<D •^H O l-H
c Cd cd a cd T3 4-» 4-» Cd
<D <U rj .a Cd Cd
t-H O 4-/ d>
> « s s ^ o d) .■^ o
*5i)
G d) 00
cO > 'f > 4-J 1-H cd
2
:3
Id cfl 'a
(1> 4^
3
^
s =« XI O Cd
4->
d) <-H
o o
o ct; a 00 a c/5 CO
Cd CO 3h >
O) a a c« d> 00 l-H o
C 2
on "eS
o Jn
<1> ^
-4-^ o Oh H-J O o O t-H
2 o
2h Uh d) Q o
e O o Cd o
_ 0^ C lo 3h
cd cd d>
o o o .g & “ Oh d «1>
§ “ 1 ^ 1-H
2 00
CO cd
>.
Td 00
Ch a> o d
« O > o cd CO o *cii
H at Vi
<D cd d o 0) CO /-s
o o O
• 1-H CO CO
o
w
O 4-4 4->
00
a o
<u o d) 0:5
> > > I—<
CO cd cd d) ^
3 & 3 52
o o O
'E^ 2
*5 S
H
H <
< H -t- H -t-
< 7 < 7 < g
CO CO
^ m !Z1 I t/5 I O W
C c/5 <* C/5 <« O cd
w w oi s § H O
c/5 W c/3 W c/3 g PJo
S<1>
00
o »-
c a “
o D
1-H 4/ 7 d) 4->
o t-H O
o d) S
4-* CD O <2 a
■> o O
cio ^ a ^
.a °
s
Oh
cd
i-H
cd
s •*0 cd
o
CO o •TS a d) C ^
0) cd
> cd d
-4-»
<l) a cd
c cd 2a o
_p 00
t-l Q
o .a o
<u >.
(/3
C/!)
o
c
^
-a
'3> -S
9^
oi Cd 2
o C
H 5 S 05 05 TJa>
> 05
S
(/)
CQ < c/^ CO
O
0!^: c/^ w S o
C/5 O c/5 C/D o
t-H
a Qh
o
54 The possibilities in space — space hardware and data transmission [Ch. 2
position and orientation, and its own environmental parameters and the state
of its power supply. This ‘housekeeping’ information must also be incorporated
into the transmission back to earth. With several sensors on board most satellites,
a fairly complex multiplexing system is required to feed the various pieces of
information into a continuous data stream.
However, it is the scanning sensors which largely dictate the capacity of the
communication downlink to earth. As sensors have been developed, so the data
transmission rates have increased, as indicated in Fig. 2.18. Higher bit rates
demand higher frequency transmission links. Thus the MSS on Landsats 1
to 3 had a data rate of 15.06 Mbits s”^ and used S-band communications, that
is, a carrier frequency of around 2 to 5 GHz, whilst the NOAA satellites carrying
the AVHRR have a lower data rate of 665.4 kbits s“^ (because the scan lines are
further apart than for Landsat), and a carrier frequency of 99 kHz can be
used. The lower frequency transmission and lower data rate make it easier to
receive the data from the NOAA satellites, and high-quaUty data can be received
at relatively moderate costs, as exempHfied by the receiving station built at the
University of Dundee, Scotland. Bayhss (1981a,b) provides a useful guide to
the size of antennae required for a given power of signal in order to achieve a
sufficiently high signal-to-noise ratio for the bit error rate (BER) to be accept¬
ably small. If the satelhte can be accurately tracked by the antenna as it passes
overhead, then the antenna size can be reduced.
Fig. 2.18 - The logarithmic growth of data transmission over two decades of
satellite remote-sensing of the earth (after R.W. H. Stewart)
for which a continuing sequence of satellites is envisaged and which are designed
to supply data for regular commercial and research appHcations. For both these
series a set of ground stations is established worldwide, and data dissemination
networks are set up. In principle it should be possible to obtain image data from
any part of the world by request if on-board recording facilities are available,
but in practice it is normally easiest to obtain data received at a local receiving
station. Even then, with notable exceptions such as the Dundee University
receiving station, it may take several weeks to obtain images, and it is to be
hoped that near-real-time data dissemination will soon become more widely
available. Image data may come in the form of black-and-white photographic
plates, or in digital form on computer-compatible magnetic tapes (CCTs),
capable of being read on standard computer tape-drives in contrast to the higher-
density digital tapes on which the data is stored at the receiving station. If
a satellite is designated operational, then it is assumed that many of the users
will prefer to have the data in as advanced a state of processing as possible.
Landsat data for example, are usually destriped (see Chapter 5), and in
future it is to be expected that atmospheric corrections and even sea surface
temperature calibrations will be applied to infrared images. This may be useful
for certain applications, but for research use it is often preferable, though not
always possible, to be supplied with the data as transmitted by the satellite,
since the ground station processing may unwittingly introduce errors as well as
corrections, depending on the use being made of the data. At the time of writing,
the cost of data from operational satellites is under review, and it is being
proposed that full economic costs be charged, although it is very difficult to
establish just what these should include (for example, the development costs, the
launch costs, the running costs of the satellite, or merely the cost of maintaining
the data archives). It is further suggested that US environmental satellites could
be sold to and operated by private companies rather than government agencies.
Such a development would undoubtedly lead to further price rises which would
inhibit research use. This would be a pity since there is still scope for the
development of further applications in the marine environment, in addition
to those already identified as justifying the commercialisation of the satellite
programme.
Other satellites such as Nimbus 7 and Seasat are looked on as research
and development programmes. For these a data dissemination network is not
necessarily set up for open access,but it is normal to establish panels of scientific
investigators who have a first right to analyse and interpret information from
the satellite. As nations other than the USA develop their space programmes
it remains to be seen what arrangements will be made to make data widely
available. In Europe, the European Space Agency has established the Earthnet
Organization to disseminate data from some of the American satellites, as
illustrated in Fig. 2.19. Some or all of the ground stations are also expected to
receive ERS-1 data.
58
LANDSAT HCMM NIMBUS-7 SEASAT
The possibilities in space — space hardware and data transmission
£
CO
0
0
o
c
£
E
n
0
0
[Ch. 2
Fig. 2.19 - The Earthnet structure for dissemination of some satellite data in Europe
Sec. 2.3] Data retrieval 59
One final point worth discussing in this chapter is the question of the
archiving of the data. For land coverage, such as the MSS on Landsat, it is
reasonable to assume that apart from specific requests for many repeat images,
it is only necessary to archive a few cloud-free scenes of each given area, corres¬
ponding perhaps to different seasons of the year. Beyond that the indiscriminate
archiving of every image received from the satellite can be considered to be an
unnecessary expense. For oceanographic appUcation the perspective is different.
Unlike the land, the sea is constantly in motion, and jthe images received by
satellite over the ocean, from whatever sensor, are likely to be constantly
changing. Since research use of the data may not commence until some years
after the satellite s launch, particularly as oceanographers are only slowly
discovering the potential value of remote sensing, there is a much stronger case
for archiving all the data received, so that spatial and temporal variability can
both be studied. With the vast quantities of information retrieved by remote
sensing this may turn out to be very costly, and oceanographers will have to
decide whether the cost of indiscriminate archiving is justified in the light of
the potential value of the applications to be discussed later in this book.
CHAPTER 3
Having seen what hardware is available in space, we take a general view in this
chapter of the variety of ways in which oceanographers are able to use satellite
remote sensing as an observational tool for their science. We first consider what
oceanographic parameters can be measured using the different types of sensor.
As a reminder of the environmental background to satellite oceanography we
then take a brief look at the range of problems and areas of research being
tackled by oceanographers which may benefit from the input of remotely-
sensed data. In particular, it is useful to examine the space and time scales of
oceanographic phenomena in order to identify the spectral windows through
which they should be observed, and to compare this with what is available from
satellite remote sensing. It is then possible to consider how remote sensing from
space not only can make observations relevant to existing research frameworks
but also promises to make possible the opening of new areas of oceanographic
study.
1/3 ^ o
rX
c CD
^ (D ;:3
.2 #v cd
4)
o
cd
b£)
G
»-i a>
+-»
cd 55
o
cd S c 3S 1-2
a ^ ^ C*-. ^ S 3 E
surface roughness
"a
a
c« ^ i::
c/5 S <D
*-* C 4> 4) CD 7d
o
l-< _ c/5 O- c/3 ag G C •S 3
CQ 2 E '■ E > ■*s ^
4= DG
c/3 0
s cd O ^ £ E o ^ c
.-M o bO bX) Vi Cl
G G 0) r-
as &£)- 4> O 0
tH G G •-
W) Vh'' g ^ “ J3 T3
CD ■4x' M
a (-4
4= —
o d W) cd " TD o CD OD bf) cd
& o ^ g- cS <D O G O CD G C
c
oj t-i O o cd .3 0 .0
&
o
O -2
Scg G >> • GJ Vh
kH
o o O 0) G
G OD CD
fl o C 3 C .G
cd c/3 CD <D
cd ><<2 cd 5^ cd c/3
0) Vi
-4-» G .JG
a; ^ -G
^ o ’G
o
<D 5
cd
3
c«
O Ss
W O C
bO
G
o CD • C/l
c
(U
.fcs
a CD
'a O CD
c CD c/3
*C/5
with doppler information
CQ <D i^\ >-> -S Vi
sc3 “3 52 'O 0)
c/3
0
o;
a< CtJ ocd 2J
c«
i s G
"g
cx
.2
3 ■§
>» C14 ^
C/5
■ •
c d> 4)
>
cd CI4 G G- OD
O 2 E^ o E kH G
>>
Cd x>
^00 c/3
1) o G kH *3
o X>
O T3 CD
O ^ P Q> CD
Cl.
QD
kH
3 ^
>> « a c w
^ R 'C cd C4-I £ S
■?
a U S
c
a>
rP .2 S Cd C SQ^ »G X
!y5
0 <3
C/D
(X
^cd ^ E OD
^ D
G CI4 c G
0
4) (D ^
o tH
<D 'E ^
cd rt)
CD O
,cd O G IH
G
"S) *5 S
7d Cd
(D
G
ID G
Wi S
E 3 C kH
as
CD M E G
t/i
as H-*
e/5 £ ^
H-J "G
o o (
cd
c/3
4>
goes C<
SCAT
a, C/D U
ALT
SAR
SAR
H
E S
cd
X ^ u <
$ K CO
1^
High-resolution
viewing
imaging radar
<D
Qh
(D
+-» oC 2
I—I <0
bO
>* CD 4> .s
■4-J
E
dar
E CD
0 cd
.2 ■E E
i3 o E 0 G
1 G' G
kH
c cd .ta •-
CD H ^ =S G
CO H 2
•G
cd G
0^
00
cd O
cu <
Sec. 3.1] The oceanographic capabilities of satellite sensors 63
text to systems which have already been developed and flown on satellites. All
of these are microwave devices which can penetrate cloud cover, and rely on the
backscattering of radar waves from the sea surface to convey information about
that surface back to the satellite.
By recording the return time for a pulse directed towards the satellite
sub-point, the altimeter measures the height of the sea surface relative to its
own position, and if this can be fixed then it is possible to determine the
absolute height of the sea surface. Oceanographic appHcations of this informa¬
tion include the study of tides and ocean circulation, and revealing details
of ocean bathymetry. In addition, the deformation of the microwave pulse on
reflection carries information about the significant wave height of the ocean
waves.
The scatterometer emits an oblique pulse of microwave radiation, and
measures the intensity of the backscattered return energy. Since this depends on
the surface roughness at length scales comparable to the radiation wavelength,
it carries information about the sea state over the illuminated area. Depend¬
ing on the particular wavelength used, this can be interpreted in terms of the
near-surface wind magnitude, the surface-wind stress, or the energy in the
surface-wave field. By viewing the same piece of sea surface from different
directions a measure of wind or wave direction can be achieved.
The synthetic aperture radar, SAR, operates in essentially the same way as
the scatterometer, but by measuring the timing and phase of the backscattered
signal as well as its amphtude sophisticated processing is able to reproduce an
image of the backscattering strength of the surface (that is, its roughness as seen
by the radar). It achieves very high spatial detail, with a resolution of order tens
of metres in contrast to the average return from an area of order 50 km X 50 km
achieved by the scatterometer. The roughness it measures is due to small waves
and ripples of a few centimetres in length. The capacity of SAR to image
anything of oceanographic interest depends on the extent to which oceano¬
graphic features are reflected in the surface roughness patterns. This has turned
out to be a most exciting aspect of satellite oceanography. Before SARs were
flown on satellites it was hoped that they would be able to image swell patterns
(that is, regular surface waves with lengths of over 100 m), ship wakes, and
perhaps slicks due to sea-surface contamination by hydrocarbons or floating
debris. In fact the Seasat SAR proved to be capable of revealing much more,
including internal waves and bottom topographical features. Because oceano¬
graphers had never had a radar bird’s-eye-view of the ocean before, it had
hardly occurred to them that dynamical or topographic phenomena centred
many tens of metres below the surface could have a surface roughness signature
capable of being imaged by a SAR. The discovery that this does happen under
certain circumstances opens up new applications possibilities for active radar
sensors, as well as raising many questions concerning the mechanisms by which
the phenomena are imaged.
64 The possibilities for oceanography [Ch. 3
We shall consider this further in Chapters 10 and 12. First we must examine
the science of oceanography as a whole to see where these remotely-sensed
observations can make a useful contribution.
scales and which control many other chemical and biological oceanographic
processes. We exclude here a discussion of surface-wave phenomena, important
to the analysis of active radar sensor data, but this will be treated in greater
detail in the next section.
Mean circulation
This is the ocean circulation that oceanographers have traditionally sought to
describe. As an example. Fig. 3.1 shows the mean circulation thought to occur
in the upper layers of the world ocean in July. The more energetic parts are well
documented from observations, whereas the weaker gyres are more conjectural.
The main features are a series of gyres which extend longitudinally across the
entire ocean, and which vary with latitude in their sense of rotation. The rotation
sense is correlated with the latitudinal shear of the prevailing wind stress, and
this circulation is therefore described as the wind-driven circulation. The circula¬
tion at depth is considerably different, and depends on the distribution of mass
due to temperature and salinity variations — hence the deep-ocean circulation
is considered to be a thermohaline circulation. Since remote sensing only
penetrates the surface layer, it is the wind-driven circulation of the oceans which
is most relevant in the context of satellite oceanography.
Sec. 3.3]
Dynamical oceanographic processes
Fig. 3.1 — The general circulation of the surface layer of the world ocean in July
67
68 The possibilities for oceanography [Ch. 3
Fig. 3.2 — The general circulation of the North Atlantic. Flow rates indicated in
10‘ m’s-' (from (TOPEX 1981))
There are clear dynamical reasons why this should be so, relating to the
Coriolis force due to the earth’s rotation, and its variation with latitude.
Stommel (1958) gives a general explanation in terms of the balance of vorticity
in the ocean. Because it is a boundary current, the path of the western flow
is controlled to a certain extent by topography. Thus the Florida current flows
energetically from the Gulf of Mexico through the Florida straits, with a velocity
of the order of 2 m s”\ and bends northwards to follow the American coast as
the Gulf Stream.
The ocean gyres tend to be composed of distinct water masses, with
different temperature, salinity, and chemical constituents, and often distinct
biological populations too. These aspects are capable of being studied by visible
and thermal radiation remote sensors, whilst the dynamics of the gyres may be
detected by the sea-surface slopes which accompany them, measurable using
satellite altimeters as discussed in Chapter 9.
Sec. 3.3] Dynamical oceanographic processes 69
The high energy of variability at a period of one year, over a broad range of
length scales, indicates the importance of annual fluctuations of the general
circulation. A large amount of variability is found at length scales between
50 km and 500 km, with a peak around 100 km and 100 days period, cor-
70 The possibilities for oceanography [Ch. 3
where D is the depth of the layer of less dense water overlaying the deeper
ocean, with a density contrast of Ap, and / is the Coriolis parameter. The
Rossby radius, Lr, governs the size of the smallest scale of motion which can
be maintained in geostrophic balance (see, for example, Pedlosky (1979)).
It turns out that over most of the ocean where the supposed mean circula¬
tion is weak, the eddy field is much more energetic. Thus it is very hard to
determine exactly what the mean circulation is. Moreover, there may well be a
contribution to the mean flux of mass, momentum, energy, salt, heat, and other
chemical tracers, which is not found by simple time averaging of the flux but is
due to nonhnear interactions within the eddies (Rhines 1977). However, neither
the contribution of the eddies to the mean fluxes nor the distribution of energy
within the eddy spectrum is yet well understood. Even the origin of the eddies is
not fully clear. Mesoscale eddies could result from baroclinic instability (which
is related to the density stratification), from topographic generation (by inter¬
action of the mean currents with bottom topography), and from wind forcing.
One mechanism of their generation occurs when the western boundary current
reaches higher latitudes and becomes detached from the coast. Meanders occur
in the detached jet which breaks up into rings. For example, in the Atlantic, the
Gulf Stream rings entrap warmer (Sargasso) water from the south-east side of
the Gulf Stream and transport it to the north and west, whilst cold rings do the
opposite with water from the Continental slope (Fig. 3.4). These rings may then
Fig. 3.4 - The formation of a cyclonic eddy or ring from a Gulf Stream meander.
The left sketch shows the meander developing, and the right sketch shows the ring
of cooler slope water detached from the Gulf Stream and isolated within the
warmer Sargasso water. The solid line represents the 15° C isotherm at 200 m
depth. The dashed hne is the approximate limit of the Gulf Stream on the
Sargasso side (after (Parker 1971))
Sec. 3.3] Dynamical oceanographic processes 71
wind up into the mesoscale eddies. Much of the energy probably shifts to shorter
length and time scales until eventually it is dissipated as heat in the turbulent
eddies at scales less than 1 cm. The study of the turbulent energy cascade in
the ocean circulation and the extent to which energy may also be transported
to larger length and time scales is of considerable interest to oceanographers at
present.
3.3.2 Tides
Tides are the other major dynamical process governing water movements in the
oceans besides the general circulation. Forced by the gravitational attraction
of the moon and the sun, the tides are identified by their regular periodicity
covering a spectrum of discrete frequencies (tidal harmonics) corresponding to
the frequencies in the gravitational forcing due to the orbital parameters of
the earth, moon, and sun. Principally, tidal motions are horizontal, although the
most obvious effect of the tides is the periodic rise and fall of sea level. The
dominant frequencies are the lunar semidiurnal (period 12.42 hours), the solar
semidiurnal (period 12 hour), the lunar diurnal frequency (25.8 hour period),
and the luni-solar diurnal harmonic (23.9 hour). These can account for the main
features apparent in typical time series of observed tidal heights, such as those
illustrated in Fig. 3.5. These include the normal semidiurnal tide of period 12.42
hours (a lunar period because the lunar influence is strongest), amplitude-
modulated over the fortnightly spring/neap cycle by the addition of the weaker
solar 12-hour cycle. In addition, there is a diurnal inequality of the semidiurnal
tides (one tide each day is higher than the other), due to the diurnal frequencies
present. This can be a dominant effect resulting in primarily diurnal tides if the
dynamic response of the ocean is locally stronger for diurnal than semidiurnal
forcing (see the lower curve of Fig. 3.5). The ocean as a whole appears to show a
greater response to the semidiurnal forcing than the diurnal. There are longer-
IMMINQHAM
Fig. 3.5 - Time series of tidal elevations at (a) Immingham, UK, showing a typical
semidiurnal tide, and (b) Kuala Baram, Malaysia, showing a combination of a
strong diurnal and weak semidiurnal tide. (Tidal data provided by I.O.S., Bidston
Observatory)
72 The possibilities for oceanography [Ch.3
3.3.3 Fronts
Turning from ocean-scale phenomena to smaller-sized dynamical features, we
first consider fronts. As in the atmosphere where the terminology was first used,
fronts in the ocean represent a boundary between two distinct water mass types,
in which the temperature and/or salinity, and hence the density, have strong
horizontal gradients. Simple hydrostatic considerations indicate that such a
situation cannot occur in a stationary fluid unless the interface between the two
water masses is horizontal. If it is inchned, horizontal pressure gradients exist,
which must either produce acceleration of the fluid or be in equiUbrium with the
Coriolis force. If a front is recognisable as a quasi-stable feature, it is assumed
that it is in this condition of geostrophic balance, in which case one water mass
must be moving relative to the other in a horizontal direction parallel to the
interfacial plane. Fig. 3.6 illustrates this schematically.
surface expression
Fig. 3.6 - Schematic of ocean front between water masses of densities pj and pj
(Pi > Pj). Velocity directions are appropriate for N. Hemisphere
Sec. 3.3] Dynamical oceanographic processes 73
Fronts are found in several situations in the ocean. The Gulf Stream itself
may be considered to be a large-scale front separating the cooler continental
slope water from the warmer Sargasso water. Within the mesoscale eddy systems,
convergences may occur which will steepen the horizontal density gradients
leading to frontogenesis. Such fronts exist on length scales of a few kilometres
across the front and tens of kilometres along it. Fronts of this nature can
develop instabilities along their length which finally lead to their break-up.
In shallow seas a different type of front has been observed (see, for example
Simpson, Hughes, & Morris (1977)), controlled by tidal energy and solar heating.
In summer the solar heating of the ocean heats the surface layers of the sea
which become less dense and tend to lie stably above the deeper, cooler, and
denser water. Wind mixing of the surface layer deepens and strengthens the
thermochne, but in shallow tidal seas the turbulence associated with tidal flow
over the rough sea bed is often sufficiently vigorous to break down this stable
stratification and mix the water column to a uniform density from surface to
sea bed. The boundary between the stratified and well-mixed sea is characterised
by a front, where the thermocUne rises to the surface, shown schematically in
Fig. 3.7. Such fronts are found around the British Isles during the summer
months. Their location appears to be governed by the parameter hjU^, where U
is the surface tidal stream amplitude at Springtide and h is the sea depth. For
small values of hjU^, in shallow water with strong tidal streams, the sea is
vertically well mixed. Stratification occurs at high values of the parameter, and
the frontal region occurs where h/U^ is about 70, {U in m s“\ h in m).
SURFACE EXPRESSION
cooling, in different depths of sea, or from fresh water runoff from an estuary.
It is not clear whether such fronts are maintained solely by advection or whether
they are in dynamic geostrophic balance as are the other types of front discussed
above.
Fronts are important to oceanographers in several ways. Dynamically, they
represent a step in the turbulent energy cascade from large-eddy scales down to
billow-turbulence scales of a few metres. They are of considerable importance to
the productivity of the sea since they tend to bring together cooler but nutrient-
rich water, with warmer less rich water. The combination of temperature and
nutrients arising from cross-frontal mixing leads to greatly enhanced productivity
of plankton, attended by increased fish stocks. Because they represent boundaries
between water masses, fronts may also define the limits of patches of pollutant
in the sea. There is often a surface convergence associated with fronts due to
secondary circulation in the vertical plane at right angles to the geostrophic flow
of the front. This causes an accumulation of surface debris, and may trap surface
pollutants, natural surface films, oil slicks, etc.
For a variety of reasons, therefore, it is valuable to be able to identify frontal
regions. Some fronts have a strong thermal signature and are readily located on
satellite infrared imagery as discussed in Chapter 7. Others may have a signature
in the visible wavelength spectrum, due to the enhanced productivity, the
different colours of the two water masses, and surface debris and slicks. There is
also the possibility of identifying fronts with synthetic aperture radars due to a
change of surface roughness at the front, or because the surface wave pattern
changes abruptly across the front.
3.3.4 Upwelling
Upwelling is the term used by oceanographers to describe the situation where
cool, but nutrient-rich water from the lower layers of the ocean, too dark for
plant Life to grow, is raised towards the surface. Brought into the light, such
waters become very fertile and rich feeding ground for fish. For upwelling
to occur, there must be a divergence of the surface currents, and this usually
occurs as a result of the wind field causing surface water drift in the presence of
topographic constraints. For example a longshore wind, flowing in the Northern
Hemisphere with a coastline to the left of the direction of travel, sets up
a geostrophic flow away from the coast, and upwelling tends to occur. This
upwelling is a regular occurrence off the North-West African coast, and off the
Peruvian coast, where wind conditions are suitable. Nonetheless, upwelling may
well be intermittent, depending both on the weather and, to a certain extent,
on the ocean-wide thermohaline circulation. Because of the importance to
commercial fisheries, much oceanographic research has been performed in order
to understand and predict upwelling, and remote sensing from satellites is
Sec. 3.3] Dynamical oceanographic processes 75
-5 0 5 0 10 0 10 20
E 500
x:
a
O)
o 1000
1500
Surface waves are found in the ocean with periods ranging from a few tenths
of a second to around twenty seconds. There are of course wavelike motions
occurring in the sea on much longer timescales of minutes, hours, or days, but
these are associated with the dynamical phenomena such as tides or mesoscale
eddies mentioned in the previous section. Within the range of surface waves it
is convenient to make the distinction between capillary waves at the short
millimetre wavelength end of the spectrum, and swell, which are waves of several
hundred metres length. The middle range of wavelengths between centimetres
and tens of metres are generally termed wind-waves. Surface-wave phenomena
have received considerable scientific study, and there are several useful texts
available on the subject. Kinsman (1965) provides a broad introduction from
first principles, Phillips (1977) presents the mathematical basis of the wave
interaction processes which are important in interpreting remotely-sensed
surface roughness data from active radar devices, whilst Le Blond & Mysak
(1978) offer a thorough survey of recent research work.
The mechanism of water-wave propagation can be modelled by a relatively
simple mathematical theory, if an assumption of small amplitude waves is made.
Being linear, this theory enables a sinusoidal wave profile to be assumed, since
any more complex profile can in principle be synthesised by the addition of
different wavelengths. Gravity is the controlling force which causes wave
propagation once the surface is displaced from its rest position. It can be shown
that the frequency o rads s“^ and the wavenumber k rads m“^ (wave oscillation
period T and wavelength X are given by
27r 2n
T = — and X = —
a k
= gk, or (3.1)
for waves longer than a few centimetres, provided that the water depth is at least
half the wavelength. The phase speed c, at which troughs and crests travel, is
defined as a X
^ k T
1/2 1/2
g g\
(3.2)
_k_ _2tt_
Consequently the longer waves, with longer periods, travel faster than shorter
wavelengths, and dispersion of the waves occurs. In shallow water, where the
wavelength is 20 times the depth, the propagation is constrained by the depth
h, in which case
Sec. 3.4] Ocean waves 79
= gh (3.3)
In this case, all wavelengths travel at the same speed (nondispersive), but the
shallower the water the slower the waves, leading to refraction of waves over
shallow sandbars or sloping beaches. These waves, controlled by the presence of
the sea bed, are often referred to as ‘long gravity waves’.
Fig. 3.9 shows the relationship between the period and the wavelength of
surface gravity waves, and Fig. 3.10 the phase velocity variation with wavelength.
In each case the parabola represents the case of deep-water propagation, obeying
equations (3.1) and (3.2). The other curves correspond to the influence of the
sea bed at different depths. Where they become straight lines, the motion can be
considered to be that of long waves, obeying equations (3.3) and (3.4). The
phase speed, c, of (3.2) and Fig. 3.10 represents the speed at which troughs and
crests appear to travel normal to their own alignments, and as such is of some
importance to the imaging of surface waves by remote-sensing techniques.
However, it is a parameter of less fundamental significance dynamically than the
group velocity C. This is the velocity of energy or information propagation. In a
wavefield of gradually varying wavenumber or frequency it is the speed at which
a group of waves of the same frequency will travel. Within the group the troughs
and crests will normally travel at a different speed from the group itself because
in general the phase speed and the group velocity are different. It can be shown
that the group velocity is C = bo/dk. For surface gravity waves, therefore,
C =
1/2
, ,
0
1
whereas for the nondispersive long waves the phase and group velocity are the
same, C = c =
The practical significance of the group velocity is illustrated when consider¬
ing the rate at which wave energy radiates out from a storm centre. The longest
waves propagate away fastest, followed by the shorter waves, but at the group
velocity rather than the phase speed. The waves of 8 to 20 second period
are attenuated less by dissipation processes than the shorter waves, and can
propagate thousands of kilometres across the ocean as swell, characterised by
long, nearly-parallel crests of almost sinusoidal waveform. These are the waves
which are often revealed by imaging radars. The distance and time from the
storm which generated them have been shown to correspond exactly to the
group velocity of the given wavelength. An observer will note that the wave¬
length of swell from a particular storm will gradually decrease, since the later
waves must have a lower group velocity than the earlier ones.
80 The possibilities for oceanography [Ch.3
Fig. 3.9 - Variation of period with wavelength for surface gravity waves in water
of different depths
Fig. 3.10 - Variation of phase speed of gravity waves with wavelength, for
different depths, h, of water
Sec. 3.4] Ocean waves 81
The information in Figs 3.9 and 3.10 is not the complete story concerning
water waves. At short wavelengths surface tension contributes to the restoring
forces which control the wavelike propagation of a surface disturbance. At
wavelengths less than about 1 cm, surface tension dominates, and gravity effects
are negligible. The resulting wave is called a capillary wave and is characterised
by the dispersion relation = yk^ where y is the surface tension divided by
the water density. The corresponding phase and group velocities are c —
and C = 3l2(ykyi\
In this case the waves are dispersive, but the shorter waves travel faster than
the longer ones, and the group velocity is faster than the phase speed. The case
of ripples of a few centimetres wavelength is of particular interest to radar
remote sensing since this is often the scale which interacts strongly with active
radars having a similar electromagnetic wavelength. Expanding the region near
the origin of Fig. 3.10 results in a phase-speed versus wavelength curve of the
shape shown in Fig. 3.11. There is a minimum wave speed if
T5
O
O
0.
CO
a>
0)
(0
x;
0.
I'ig. 3.11 - Variation of phase speed with wavelength for short wavelength
(capillary-gravity) waves
82 The possibilities for oceanography [Ch. 3
Although the water viscosity has little influence over water waves, other
dissipative processes, including surface tension effects and nonlinear wave
interactions, cause short wavelengths to be damped out rapidly. Capillary waves
are therefore short-lived, but being generated by wind stress, they significantly
roughen the surface as ‘felt’ by the wind, and have a considerable influence on
the transfer of momentum between the wind and the waves. The mechanisms
by which waves are generated by the wind are still not fully understood, and
it may well be that the development of radar remote-sensing techniques is able
to offer new ways of studying the problem. A gentle breeze of less than 5 m s“^
has very little effect on the sea surface except to roughen the surface with
ripples. A fresh breeze of 10 m s"^ is able to produce quite a rough sea, and
higher winds produce correspondingly rougher seas. In this situation, the waves
do not have a regular sinusoidal shape with a given wavelength and amplitude.
Instead, the fluctuations of wave-height are random, and an aroused sea must be
described by its energy spectrum.
Fig. 3.12 illustrates typical wave-energy spectra appropriate to different
wind conditions. The energy spectrum represents the amount of wave energy
present in a small bandwidth of frequencies, and it is obtained by spectral
analysis of a time series of surface-height measurements at a point. One way
of interpreting it is to imagine the waves in the aroused sea to be dispersed into
the component individual sinusoidal waves of different wavelengths, as the wave
field propagates away from the storm area. The energy of waves at a particular
wavenumber/frequency will correspond to the energy indicated in the spectrum
at that frequency. The amplitude of the sinusoidal waves is proportional to the
square root of the energy. Fig. 3.12 indicates clearly how higher winds are able
to produce more wave energy. The spectral distribution of this energy is worth
noting. At the high frequency end corresponding to the shortest ripples, the
wavefield is quickly saturated and is unable to contain more energy as the wind
rises. Instead, the energy peak moves to lower frequencies corresponding to
larger waves, and it takes a strong storm to generate the longest swell waves.
Fig. 3.12 can be used as the basis of wave-prediction algorithms. It corres¬
ponds to a fully-aroused sea, and only appHes if the given wind strength has been
blowing for a sufficiently long time (the duration), and over a sufficient length
of sea (the fetch) upwind of the observing position. If not, the wave conditions
are said to be duration-Hmited or fetch-limited. Whilst the wave-energy spectrum
contains much information about the wavefield, it has no information about the
direction of propagation of the different wave components. It is very difficult to
measure the directional wave-energy spectrum (that is, the distribution of energy
in different wavenumbers in different directions) by conventional methods.
The abihty of satellite-borne radars to sample the wavefield in two horizontal
directions will therefore make a significant contribution to studying wave
directionahty. A wave spectrum requires a certain amount of interpretation, and
a much simpler measure of surface roughness is given by the significant wave
Sec. 3.4] Ocean waves 83
Fig. 3.12 — Typical surface wave amplitude spectra at two wind speeds (adapted
from Hasselmann et al. (1973))
height. It is defined as the average of the highest one-third of the ‘wave height’
measured between adjacent lower and upper turning points on a time series
record of surface height. It is therefore given the symbol and has proved to
be a useful oceanographic parameter because it corresponds quite closely to the
visual estimates of wave height made in weather reports from ships at sea. It can
be related to the wave-energy spectrum, and it is therefore possible to construct
hij-i wave-height prediction curves such as that shown in Fig. 3.13. The construc¬
tion and improvement of such predictions which tend to be location-specific
provides significant help to the shipping and offshore engineering industries. One
problem for remote-sensing techniques is to make sure that even though the
radars may be measuring roughness at length scales corresponding to the high-
frequency end of the spectrum, they are able to accurately estimate hxj'^ which
corresponds principally to waves at frequencies near to the spectral peak. These
and many other aspects of radar remote sensing of surface waves will be dealt
with further in Chapters 10 to 13.
84 The possibilities for oceanography [Ch. 3
WIND
PERIOD. 8
Typical Possible
Surface horizontal Horizontal remote-sensing
displacement water length surface
amplitude velocity scale Time scale signature of the
m m s"‘ km phenomenon
Large ocean gyres 0.5 0.01 3000 one to many surface slope
years colour
4.1 INTRODUCTION
In this chapter we take a brief look at the scientific methodology which is
necessary if satellite remote sensing is to be used for making valid measurements
of oceanographic parameters. Many other books on remote sensing deal with the
topic of fundamental principles in some detail (for example, Barrett & Curtiss
(1976), Lillesand & Kiefer (1979), Reeves (1975), Sabins (1978)), but they
are mostly concerned with land applications and have a particular bias towards
visible and near-visible wavelength imagery. Our concern here is to point out
the ways in which the methodology of remote sensing of the sea differs from
the quite well established approach developed for land applications. Some tasks
are easier over sea, whilst others are much more difficult.
Bearing in mind the point of view of the oceanographer who is sceptical of
remote sensing ever producing much more than a ‘pretty picture’ of the ocean,
this chapter introduces the supporting measurements and processing tasks which
must accompany satellite-derived data if useful quantitative information about
ocean parameters is to be derived. When presented with a clear and colour¬
ful computer-produced image based on satellite data, there is a tendency for
the remote-sensing enthusiast to jump immediately to unjustified conclusions,
interpreting as oceanographic features patterns which may be artefacts of the
sensing mechanism, the atmospheric interference, or the processing technique.
Conversely, the cautious experimental oceanographer may declare that any
apparently interesting features are nothing more than such artefacts. The true
‘satellite oceanographer’, if we may use the title, follows a careful path between
these extremes. On the one hand he must be sceptical and avoid reaching con¬
clusions which are not justified, but at the same time he is motivated by the
expectation that the remotely-sensed data received by the satellite contain
at least some information about the sea, even if corrupted by various sources
of error. He will also be searching particularly for the type of information which
conventional measurement techniques cannot observe. The scientific approach in
90 Principles of remote sensing of the sea [Ch.4
that can be done is to use the most recent calibration (in-flight or pre-flight).
Even if this is slightly in error, it will be an absolute error, applying equally to
the whole scene of a scanned image. The relative accuracy, that is, when com¬
paring the values of two points on the same scene, will not be affected by slow
sensor drift, but will depend on the resolution sensitivity of the sensor. Sensor
drift creates worse problems when comparison is made between two images of
the same sea area received several months apart.
The conversion of the sensor output into a digital value for transmitting to
earth is generally a reliable procedure with modern microelectronics, and can be
monitored by running a standard voltage ramp through the system periodically.
Once in digital form, the signal can be transmitted with high accuracy, and
effectively no further errors need be introduced into the digital data. It is
possible to lose sensor radiometric resolution in the digitisation processes. If the
transmission is hmited to 8-bit numbers, then a range of only 256 digitisation
levels is permitted. Ideally this should be arranged so that a digitisation interval
corresponds to the sensor sensitivity at each incident radiation level. If the
digitisation interval is too large, then the true sensitivity of which the sensor is
capable is not reflected in the digital data. If the digitisation interval is smaller
than the ■ sensitivity, then digital levels are being wasted in communicating an
apparent radiometric resolution which is not present in the sensor output.
Furthermore, users presented with such data may try to interpret variability
of the integer values which is purely an artefact of the digitising circuits, and
powerful image processors may be made to enhance patterns which have no
oceanographic significance whatsoever. Care should always be taken, therefore,
not only to check the calibration of the sensor but also its sensitivity in relation
to the digitisation interval present in the data.
useful signal. Radiation leaving the sea surface reaches the sensor, containing
information about the sea which is of value to the oceanographer. Ray 2
represents the radiation leaving the sea which is absorbed by the atmosphere
en route, whilst ray 3 is that which is scattered by the atmosphere out of the
sensor field of vision. The sum of 1,2, and 3 represents that energy which should
be received by the sensor if a complete correlation is to be made between the
received signal and the ocean environmental parameter under investigation. The
absence of 2 and 3 reaching the satellite must therefore be allowed for in some
form of atmospheric correction to the received data. In addition, rays 4,5, and
6 reach the sensor without having left the sea surface in the field of view, and
therefore constitute extraneous ‘noise’ on top of the signal, so far as the ocean
scientist is concerned. Ray 4 is that energy emitted by the constituents of the
atmosphere. Ray 5 is energy reflected by scattering into the field of vision of
the sensor, and ray 6 is a special case of scattering into the field of vision —
energy which has previously left the sea surface but from outside the field of
view. Ray 1 may of course represent energy which has been forward scattered,
provided it originally left the sea surface within the field of view.
Fig. 4.1 - Atmospheric pathways of e.m. radiation between the sea and the
satellite sensor
94 Principles of remote sensing of the sea [Ch.4
for each observation, or for each complete image scene. It is simpler and less
time-consuming to construct a universal atmospheric correction based on an
average model of atmospheric effects. This is the case with the routine analysis
of sea-surface temperature data from the AVHRR on the NOAA satellites. This
has the effect of removing the greater part of the atmospheric error, but cannot
allow for spatial and temporal variability in atmospheric aerosol and water
vapour composition, resulting in a level of uncertainty which can often be
tolerated provided its hmiting magnitude is known. Such an approach depends
on a broad database of meteorological observations from which to construct an
average atmosphere, and is invalidated when an exceptional atmospheric condi¬
tion is encountered. Such was the case in 1982 when the Mexican El Quichon
volcano produced unusually high levels of stratospheric dust in a low-latitude
belt around the world.
A fourth strategy, not unhke the second, multispectral, approach, is to
mount an atmospheric sounding sensor on the same satellite as an oceanographic
sensor for which an accurate atmospheric correction is required. A microwave
sounder is able to resolve separate components within the atmosphere and thus
provide a means of correcting an IR radiometer for thermal IR absorption
and emission in the atmosphere. It is also proposed that the microwave channels
of the Along Track Scanning Radiometer (ATSR) to be flown on ERS-1 will
provide a means for modelling more accurately the atmospheric corrections to
be apphed to the altimeter on ERS-1. The problem in this case differs from
those discussed in the context of Fig. 4.1, since the interpretation of altimeter
data requires a knowledge of the actual speed of microwave radiation through a
variable atmosphere.
The ATSR will be described in Chapter 14, but its concept is an example
of another possible atmospheric correction strategy. Fig. 4.2 illustrates the way
in which an oblique view of the earth’s surface necessitates looking through a
much longer path length of atmosphere than for nadir viewing. This of course
must be taken into account when making atmospheric corrections of images
scanned by large-swath-width scanners such as the CZCS or the AVHRR. This
extra problem has been turned to good effect in the ATSR design. By viewing
Fig. 4.3 - Distortion due to oblique viewing of the curved earth surface, (a) The
projection of the square IFOV on the earth’s surface, (b) The distortion of
a square on the ground as viewed from space, plotted in rectangular pixel
coordinates without allowing for oblique viewing and curvature effects
Sec. 4.5] Geophysical calibration 99
sea truth collected within a few hours of the overpass may suffice for calibration
purposes, although there will always be more confidence placed in the results the
closer the sea truth can be timed to coincide with the overpass.
The provision of facilities for synoptic sea data collection presents severe
logistic difficulties not encountered by land remote-sensing workers. It is much
more costly to have to use a boat to reach a sampling point than simply driving
there on land. If the measurement demands that a scientifically-equipped
research vessel be used, then the cost escalates. If the sensor being calibrated is
an all-weather microwave device, then the only limitation on obtaining synoptic
sea and satellite data is that the weather should not be too bad to prevent the
sea observations being made. If visible or infrared sensors are being calibrated,
then cloud-free conditions are also required for there to be any useful satellite
data for comparison with the sea truth. Waiting for the suitable combination of
clear skies but not too strong a wind can be an expensive exercise with a fully-
equipped and staffed research vessel. For a satellite such as Landsat with a
16- or 18-day repeat cycle, only one or two clear weather passes a year may be
gained for some locations. An alternative is to moor buoys to collect sea data
over a period of weeks or months, in the expectation that at least some samples
will be obtained synoptic with a cloud-free overpass. An instrumented buoy
programme is itself costly, however, and it is evident that a dedicated sea-truth¬
gathering exercise for satellite calibration purposes requires a significant
commitment of resources and research effort, however it is performed.
Another means of obtaining synoptic sea truth is to make use of oceano¬
graphic data which happened to be collected within the field of view of the
satellite sensor at or near the time of the overpass, as part of an unrelated
oceanographic research programme. This might be termed a ‘serendipity
method’! It can include measurements made from ships of opportunity, and
may combine meteorological records made from commercial shipping, as well as
scientific observations from research vessels and instrumented buoys. It almost
invariably has the drawback that the oceanographic parameter has not been
measured in quite the way that would have been chosen for a ‘sea truth’ exercise.
For example, sea surface temperature may have been measured from the
engine-intake water, at a few metres below the surface, whereas ideally the skin
temperature should be measured for calibrating IR sensors. This suggests a fourth
approach to sea truth collection which is to mount instrumented packages on
ships of opportunity which make regular crossings of certain seas and oceans.
The instruments have to be simple and robust, but can be designed to measure
the particular parameters of relevance to the task of satellite data calibration and
validation.
Another related problem in the comparison of sea and satellite data is
that of spatial sampling. To obtain as broad a database as possible for calibra¬
tion algorithms, it is desirable to obtain sea data over a wide range of possible
values corresponding to a wide range of satellite data values. This requires sea
102 Principles of remote sensing of the sea [Ch.4
pixel. One simple example is the skin temperature measured in the presence of
large waves, where it has been noted that whilst the underlying sea surface
temperature may be uniform, the actual surface skin temperature may vary by
as much as a few tenths of a degree Kelvin between trough and crest, a length
scale of tens to hundreds of metres. To avoid the problem one must make
statistical allowance for small-scale variability, if it is known, in the calibra¬
tion and comparison procedure, or seek to measure an average value at sea
by subsampUng within a pixel. The only practical method for doing this is by
performing continuous sampHng along a transect and making sure that the
sampling rate is very much more spatially dense than the resolution of the
satellite sensor being cahbrated.
It is worth emphasising here that the spatial-averaging capabilities of
satellite sensors is of significant advantage to the oceanographer in many applica¬
tions. When seeking to observe the variability of large-scale processes it is a
problem if measurements of the phenomena are corrupted by local variability
due to small scale processes. The satellite sensor conveniently avoids this problem
so far as sub-pixel variabihty is concerned, whilst variability at medium-length
scales larger than the pixel size can be smoothed out because of the complete
sampling of the area achieved by imaging sensors. This enables large-scale
variability to be revealed in circumstances where ship measurements might not
have been able to distinguish it from the noise of the smaller-scale processes.
So far in this section we have considered data collection at sea in the
context of the geophysical cahbration of satellite data. There is another related
use, and that is the testing of atmospheric correction algorithms. Since these
effectively deduce the water-leaving radiation characteristics from radiances
measured at the satellite, they require sea-level observations of the radiation
for their validation. Such radiation measurements at ground level can also serve
to demonstrate whether a remote sensor is capable of detecting certain sea
parameters. For example, in coastal waters the CZCS is believed to be capable of
revealing chlorophyll and suspended sediment patterns in the sea. The imagery
is often highly contaminated by atmospheric effects, and it is therefore difficult
to estabhsh a correlation between satellite colour signature and the sea-water
parameters. If colour measurements at sea level can demonstrate such a
correlation, then it is worth improving the atmospheric correction until use¬
ful information can be extracted from the satellite data. If colour measurements
at sea level do not correlate with the water-quality parameters, then there is
clearly no point in looking for such a correlation in the satellite data.
For both these reasons, radiation measurements at sea, viewing the ocean
from above as the satellite does but without any intervening atmosphere, have
an important part to play in the quantitative interpretation of satellite data.
Unfortunately these are not measurements traditionally made by oceanographers.
Such optical measurements as have been made have usually been subsurface.
Only recently with the development of satellite oceanography has work seriously
104 Principles of remote sensing of the sea [Ch.4
begun which deploys radiometers and radiance meters from the mastheads or
the booms of research vessels or platforms. The spatial sampling and spatial-
averaging problems mentioned above present obstacles to the interpretation of
the data. Aircraft carrying analogue satellite sensors can take a wider view, but
then the atmospheric problems are introduced once more, in a less tractable
form than for the satellite since the variable height of the aircraft and its varying
tilt and pitch continuously alter the path length of the radiation. There is clearly
scope for the further development of radiation measurements at sea, not only in
the visible and infrared, but also in the microwave, seeking to understand how
radar backscatter is related to sea-surface roughness.
Finally, having mentioned in section 4.3 that the positional registration of
satellite data is essential if it is to be useful to the oceanographer, it should be
remembered that accurate positional registration of sea-sampled data is also
vital if it is to be correlated with satellite data. The errors in position-fixing at
sea away from land are comparable with those of the positional registration of
imagery, although the increasing number of navigational satellites in orbit will
improve navigation for ships and buoys equipped to use them. The question of
position fixing is the limiting factor in the use of instrument packages on ships
of opportunity, and ideally requires regular positional updates from the ship’s
navigational equipment to be logged along with the oceanographic data. The
discrepancies in sea and satellite data correlation which arise from errors in
positional registration of either or both will depend on the sensor sensitivity, the
spatial variability of the parameter in the region surrounding the measurement,
and also on the size of the IFOV. For example, if it is known that the SST is
uniform (that is, constant within the sensor’s radiometric resolution after
digitisation) for 10 km in all directions, then position fixing to within 10 km
would suffice. If, on the other hand, the sampling point is close to a front,
it is very important that the sampling point and the corresponding pixel with
which it is identified are located on the same side of the front. However, even
with a highly sensitive sensor, and strong spatial gradients, there is little value
in improving positional accuracy of sea-truth measurements to less than the
IFOV length scale.
It is difficult to generalise any further on the question of comparability
between ‘sea truth’ and satellite data. Each sensor and situation must be
considered individually, bearing in mind the points that have been made in this
chapter. Calibration errors are due to inadequacies in both the satellite remote
sensing and the sea-sampling techniques. It is worth remarking that satellite
remote sensing of the ocean has developed in some areas to the point at which
the limiting accuracy is that of the sea measurement rather than the satelhte. For
example, significant wave height could be measured with the Seasat altimeter
to a precision as good as could be achieved by measurements at sea. However,
there is certainly no question of the satellite sensor replacing the instrumented
buoy or research vessel. On the contrary, the way towards real scientific progress
Sec. 4.6] Oceanographic sampling for ‘sea truth’ 105
in improving the accuracy of remote sensing lies not only in better space instru¬
ments, but in improved oceanographic techniques to match them. Quite apart
from the temporal and depth resolution which can only be achieved by measure¬
ments at sea, the remote sensor can only achieve its potential performance in
spatial sampling of the ocean if it is backed up by a well-coordinated programme
of careful acquisition of sea truth.
CHAPTER 5
ticated image processing can generate it! Yet such is the power of modern
computer techniques to enhance the smallest variability, smooth the result, and
enhance again to produce a clear, crisp, multicoloured pattern, that the uncritical
user of this equipment can easily convince himself that he has stumbled on
a genuine oceanographic feature. In fact he has merely used his imagination in
much the same way that we can ‘see’ all sorts of interesting pictures, shapes,
or even a map of Australia if we gaze long enough into a sky full of cumulus
cloud patterns! Whilst there is a place in satelhte oceanography for looking
particularly at pattern shapes and the length scales of variability, there are also
dangers in merely generating pretty pictures. The safeguard against uncritical
conclusions is to keep track of the numbers involved, both the raw digital values
from the sensor and their calibration in terms of oceanographic parameters.
For example, a thermal ‘front’ may appear by enhancement out of what was
previously an apparently uniform piece of sea. If we then find out that the
‘front’ has in fact a temperature step of only 0.2 deg K, when the thermal
resolution of the IR sensor is known to be 0.2 deg K and the noise another
0.2 deg K, we must conclude that the front is in our imagination and not really
there. Thus it is important to understand exactly what the image-processing
software is doing to the numbers from the original satellite data set, so that
the correct brightness value of a pixel on the screen can always be related to
a physical parameter.
Band 1 Image
Band 1 image
Fig. 5.2 - Image display organisation of data from the CCT of Fig. 5.1 to produce
a correctly-oriented image, assuming a left to right scan, for (a) satellite travelling
from south to north, and (b) satellite travelling from north to south
Sec.5.2] Digital image data from satellites 111
whole of band 1 is sequentially filed, line by line, before band 2, etc., as shown in
Fig. 5.3(a), or LSB when the first sample on each line is stored line by line, then
the second sample and so on as in Fig.5.3(b). It will be noted that the latter
case in particular is not suited to the incorporation of line by line calibration
data.
(a)
Samples, Lines, Bands
Band 1 Band 2 Band 3
Sampla 1 32 32000 31 32 31000 32000 31 32 31000 32000
Lina 1 LI LI L2 L2 L2000 L2000 LI LI L2000 L2000
(b)
Lines, Samples, Bands
Band Band 3
Band 1
*7
3ample 1 31 31 31 32 52000 32000 31 31 32000 32000
Line 1 L2 Liooe L2000 LI Liooe L2000 LI L2 Liooe L2000
Fig. 5.3 - Ordering of image data on a CCT in the case of (a) Sample, Line, Band
organisation, and (b) Line, Sample, Band organisation
The data values are normally stored as 8-bit binary codes, that is, as integer
values between 0 and 255, although sometimes (for example, for AVHRR
data) 10-bit words are used enabling an integer range between 0 and 1023. As
discussed in Chapters 2 and 4, the integer value corresponds to different levels
of electromagnetic radiation, and it is sensible to assume that a higher number
corresponds to a higher radiance level. This is true for visible and microwave
data, but for the thermal infrared the convention is the reverse of this. Thus
as the sea temperature rises, the infrared radiation at the satellite increases, but
the digital value returned to earth decreases. The reason for this is that a simple
grey-level visual reconstruction, using Ughter shading for higher numbers and
darker shading for lower, causes cloud tops to appear whiter than land or sea,
which assists in meteorological interpretation. The result is an image in which
the warmer parts are darker, which suits better our conceptual grasp of black
objects being warmer than white ones. Particular care must therefore be taken
when dealing with multispectral data from sensors which have both thermal IR
and visible/near-IR channels, since the data is a mixture of both conventions.
The calibration information accompanying the earth-view data will define
whether the calibration gradient is positive or negative, and hence define which
convention is being used.
112 Principles of image processing [Ch. 5
Assuming 8-bit data values are being recorded, the typical image dataset
described in Fig. 5.1, apart from the calibration and housekeeping data, will
have 2000 X 2000 X 3 X 8 = 96,000,000 or 96 M bits of information. In terms
of disk storage on the computer, with 1 byte = 8 bits, a 12 M byte disk capacity is
required for the three-channel image. In practice many scanners produce larger
datasets, with longer scan lines and more channels. With so many numbers to
be processed, dedicated image-processing computers are the most satisfactory
means of handing satellite image data, although it is still possible to extract a
lot of valuable information, and to perform calibration processes using ordinary
mainframe computers with sufficiently large disk storage capacity.
(ii) by moving a cursor over a digitising tablet and selecting instructions from
a menu on the tablet,
(iii) by selecting commands from a menu on the display screen or the VDU
screen, using the digitising tablet cursor, a joystick, or a trackball, or
(iv) by a combination of these methods,
depending on the proprietary system used. It is customary to be able to enter a
mode of operation whereby the trackball or joystick can be used to zoom in
upon a detailed view of part of the image and to roam over the entire image
in this magnified perspective. Various image-enhancement processes can be
performed, with interactive intervention to produce the clearest image which
best emphasises the feature under study. However, there is a danger in these
interactive processes that quantitative and critical science can degenerate into
merely producing colourful pictures, and the user can lose track of the real data
values in the pre-written software provided with the computing system. Those
computer algorithms which operate in order to correct the image geometrically,
and those which use multispectral information to correct atmospherically
and to calibrate the data in terihs of oceanographic parameters, tend to require
more complex arithmetic processes. These operations are therefore much
slower, having to be performed on the main CPU rather than in the image
processor itself. However, as a new generation of array-processor-driven machines
becomes available, it is likely that even these tasks will be accomplished almost
instantaneously, that is, taking a few seconds rather than many minutes.
Once an image has been processed, it is useful to be able to store the final
result for future use. Floppy-disk storage offers a convenient and cheap method
at present, but whilst CCTs provide a suitable method for transferring data from
supplier to user, and from one user machine to another, floppy disks do not have
an acceptable standard format, and are not advisable for data transfer between
users or different computers, unless the floppy-disk drives and associated driving
software are identical.
79 m between scan lines, by selecting every 6th or 3rd sample and every 4th
or 2nd line, an approximate correction is automatically apphed to allow for the
non-square pixels.
This approximate form of geometric processing results in loss of information
from the rejected pixels. A more satisfactory geometric correction technique
requires the resampling of the image according to a mathematical transforma¬
tion, utilising all the data values in the original image to contribute to the data
values in the final image. Fig. 5.5 illustrates two possible cases. In (a) the final
pixel size is similar to that of the original pixels, and in (b) the final pixels cover
a much larger area than the originals. The reverse of (b) could be envisaged, with
the final pixels much smaller than the originals, but in that case it is not possible
to find enough information in the original data to make every new pixel value
independent of its neighbour, and resolution cannot be improved by that means.
In Fig. 5.5 resampling can be performed most simply, but less accurately, by
‘nearest-neighbour substitution’ in which a given pixel, for example, that with
centre at A, is assigned the value of the pixel on the original grid whose centre
is closest to A, that is, a. An improved bilinear interpolation procedure is to
assume a hnear gradient of the pixel values between pixel centres on the original
grid, so that resampling for B will utilise the values at pixels b, c, d, and e,
effectively fitting the parameter value to a twisted plane through b, c, d, and e,
and locating the resampled value at B on that plane. By using higher-order
polynomials, or spline curves to fit the parameter surface on the original grid, it
is possible to obtain even better sampling on the new grid, incorporating the
values of the neighbouring 9 or 16 old grid centres. Such processing can be costly
in computer time, and introduces a smoothing or filtering effect,which increases
with the number of original pixels used to contribute to each new pixel value.
If smoothing or filtering is to be performed anyway, it may turn out to be
cheaper in computer time to use the nearest-neighbour resampling and then
apply a standard filter, rather than use a complex geometrical resampHng
polynomial which has to be re-evaluated for many blocks of pixels across
the image. Given the availability of several different resampling algorithms on
proprietary systems, the user must make a judgement based on the application
concerning which one to use.
The geometrical transformation itself, that is, determining the shape of
the new pixel grid pattern relative to the old, can be achieved by two distinct
methods, as discussed in section 4.9. In the first, ground control points on the
original image are identified and their pixel indices (row and column number)
noted. These pixel coordinates are then linked to a set of new pixel coordinates
based on the map projection of the imaged area. A general warping algorithm,
available with most proprietary image processors, then fits the old points to
the new locations. By treating the rest of the image as if it were a rubber sheet,
capable of being stretched and rotated unevenly across its surface, all the old
pixel coordinates are located relative to the new, so that resampling of the actual
116 Principles of image processing [Ch. 5
values can be performed. The more ground control points there are, the more
likely is the map to be a satisfactory one. Such a generalised warping process is
very time-consuming, when million or so data points have to be re-evaluated.
It is somewhat quicker if a simple rotation or linear stretch can be applied to
the whole scene, to bring it onto a map base, but this is only feasible if the
underlying distortions of the satellite perspective have been removed.
(a)
Fig. 5.5 - Typical examples of geometrical resampling grids: (a) similar size,
(b) resampled grid much coarser than original
identified as being N pixels away from the pixel corresponding to the sub-point
(normally the central pixel in a scan-line), then ^ = N8^. We need to find a
as a function of (3, where a is the distance from the satellite sub-point to the
pixel found at viewing angle /3.
Now a = i? i// = /? (tt — j3 — 7).
(h+R) sinj3
From the sine rule: siny =
R
|(i?+/z) sin|31
thus a = R TT — p— sin
Using this, the scan line can either be replotted with differential spacing
between pixels, or resampled.
Fig. 5.6 — Geometric sketch for relating along-scan distance ‘a’ to along-scan
viewing direction /3
118 Principles of image processing [Ch. 5
The identification of latitude and longitude along the scan line is rather
more complicated. We assume that the latitude and longitude of the scan-line
centre (the satellite sub-point) can be obtained from a knowledge of the satellite
orbit at a given time. The direction of the scan line is assumed to be at right
angles to the travel direction, and this can be determined from a knowledge
of the orbit inclination and the instantaneous latitude. Fig. 5.7 illustrates the
problem. It is treated analytically by considering the satellite orbit path to be
the ‘equator’ of a satellite coordinate system, denoted by a prime ('), in which
the scan lines correspond to lines of longitude. Thus point P in the scan line
sampled when the satellite was at S has true latitude (pp and longitude Xp and
in the ‘satellite co-ordinates’ 0p and X'p. X'p = X'g, and if X' is measured from
where the track crosses the true equator at X, then
sin 05
X' sm -1
sin /
Fig. 5.7— Sketch to illustrate the relationship between geographic latitude and
longitude [(pp,
\p] of point P and the ‘satellite coordinates’ [0p, X'p] based on
the satellite orbit as ‘equator’
cosisinf—I-Xq) cos/—isin^c
\2 / \rI + cosXo sin / sin
sin/
where Xq is the true longitude of X.
Thus the actual latitude and longitude of a point can be defined in terms of
its distance a from the satellite sub-point which is at latitude 0^ on an orbit
of inclination / which crossed the equator at longitude Xq. Note that Xo is
the equator-crossing latitude if the earth had not rotated whilst the satellite
travelled from X to S. A time correction must therefore be applied to the actual
equator-crossing latitude.
The above analysis offers the framework of a theoretical pixel-location
procedure, which should be developed individually for given satellite orbits. Care
must be taken in defining the directions of angles related to the travel direction
around the orbit. Positioning near the poles may be better achieved by trans¬
forming to a different set of earth coordinates which does not have a singularity
at the pole. Once the pixels have been located in terms of (0p, Xp) the whole
image can be resampled either to provide uniform sampling across 0 and X, or
to achieve square pixels on whatever map projection is being used. At this
stage, the new image geometry is still subject to errors arising from inaccurate
knowledge of the satellite orbit. However, a small linear translation and/or
rotation of the new image, based on just a few ground control points, should
be sufficient to register the image accurately.
Such a binary image can be used in two ways. One is to ensure that certain
image-enhancement techniques operate on the sea pixels only. The mask acts as a
label attached to those pixels to be included or excluded from image-processing
commands. The detail of how this is implemented vary with different processing
systems, but in one form or another the facility is to be found in all well-
developed imaging software. The other use is in the preparation of slides or hard¬
copy records of imagery. If two masks are prepared, one of the visible land and
the other of cloud, it is visually helpful to overlay the sea-processed image with a
mask of green or black for the land and white for the cloud (if these colours
do not create ambiguity with colours already used in the sea processing). The
observer can thus readily locate the image by the coastline pattern, and quickly
appreciate the masking effects of cloud so that attention can be concentrated on
the open-sea parts of the image with which the oceanographic user is concerned.
The binary masking image can normally be created from the original image.
Cloud is usually apparent on both visible and thermal IR imagery as pixels with
much higher digital values than the sea, because on visible images the clouds
reflect more sunlight than the sea, and in IR data they are generally cooler than
the sea, emitting less radiation and hence appearing lighter because of the
thermal IR densitometric convention. Thus by assigning all pixels with values
over a certain threshold as ‘on’, the cloud mask is created. The land mask can
cause problems since in visible wavelength imagery, depending on the channel,
the land and sea may have similar brightness, and for thermal IR data the land
may be warmer or cooler than the land, depending on season, time of day or
night, etc. In general the sea will be cooler than land in summer and warmer in
winter, and whilst sea temperature varies very little at night, the land cools down
significantly in the night. There are often situations, therefore, when land and
sea have similar temperatures. One way round this is available when a satellite
delivers both a thermal IR and a visible image of the same area. It is quite likely
that when the thermal IR image cannot produce a suitable land mask, the visible
channel will, or vice versa. In practice the most reliable channel for the produc¬
tion of land masks is in the near-IR if available, where there is very little solar
reflection by the sea, but significant amounts from the land. Band 7 of Landsat
MSS, for instance, or channel 2 of the AVHRR, makes a useful coastline
delineator. Provided that the different channels have been collected by the same
sensor, the pixels of one channel coincide with the pixels of another, and the
masks made from one channel are available for use with all the other channels.
If it is not possible to create a mask from the image itself, then it might be
necessary to create one from a map, using TV input to the imaging system, or
drawing on a digitising tablet. This is rather laborious, however, and presupposes
further that the image has been very accurately position located, something
which is not so important when the coastline shape is already directly evident on
the image.
Sec. 5.6] Smoothing, filtering, and noise reduction 121
■ pixel at value 21
□ pixel at value 10
(b)
(a)
(0
19.2
0
<D o
1 a CO
-1
ro 00
1 1 r- I 1 1 1 I 1
Fig. 5.8 — Digitisation effects in an image, (a) A noise-free linear ramp of values
between 19.0 on the left and 21.0 on the right, as it might appear when sampled
with a digitising interval of 1.0. (b) The effect on (a) of a smoothing filter and
rounding down, (c) The effect of smoothing and contour plotting of (a)
of the filter is most easily conceived as moving the centre square of the matrix
over the image, pixel by pixel, and placing in the new (filtered) image dataset
for that pixel a number corresponding to the operation specified in the matrix.
Thus in the 3X3 moving-average filter of Fig.5.9(a) the pixel value is replaced
by the average value of the pixel and its eight immediate neighbours. Such a
filter would give a considerable smoothing effect to Fig.5.8(a), and might result
in Fig. 5.8(b), where the averaged value has been rounded down to the nearest
digital value. If noise still persisted, a larger-area filter, 5X5 or 7 X 7, could
be used, although the larger an area covered by the filter the more likely it
is to start smoothing out steep gradients of ocean parameters which are real.
A compromise can be achieved by using a weighted average, which allows the
pixels near the centre of the filter to influence the output value more than those
at the perimeter. A hypothetical example is illustrated in Fig. 5.9(b). It should
be noted that in performing the filtering operation the computer must use real
rather than integer arithmetic. If instead of rounding down to produce an
integer digital image, the real numbers are stored, it is possible to plot contours
as in Fig. 5.8(c). However, whilst such detailed contouring may give an illusion
of high parameter resolution, it cannot in fact contain more information than
was present in the original data. Thus, if patterns begin to emerge in such
contours they must be regarded as artefacts of the filtering process unless they
include a contour interval of at least the digitisation interval.
Images from the Landsat MSS have a characteristic 6-line stripe effect,
produced because the sensor scans six lines together with six separate, parallel
sensors with slightly different characteristics. De-striping algorithms are available
to remove this stripiness by a statistical technique, but it is found that whilst
this is adequate over land, such algorithms may introduce more stripe noise into
over-sea parts of an image. For oceanographic applications, it has been found
that a 6 X 8 moving average filter is the most satisfactory way of removing the
stripes and generally smoothing the data before further atmospheric correction
or geophysical calibration processing is performed.
Filters can have uses other than smoothing, and a variety of different types
are described in advanced image-processing texts and proprietary software
manuals. One application is edge-enhancement (that is, highlighting sudden
changes in the parameter values as would occur at a coastline in a near-IR image,
at a front on a thermal IR image, or where a surface slick, or wind shelter,
produces a much smoother patch of sea surface on a radar image). Another use
is to emphasise regions of increased parameter gradients either in a particular
direction, or in general. Other techniques such as Sobel and Pseudoplastic
filters can create some fascinating visual effects, but it is not clear how much
apphcation they have to satellite oceanography. The available image-processing
software tends to be dominated by land applications, and it should not be
assumed that it can necessarily contribute much to oceanography, although there
is room for further experimentation.
124 Principles of image processing [Ch. 5
the fuU range of pixel values available. The result of this in a monochromatic
image is to make the sea look an almost uniform grey. If there are 256 digital
levels available, and these are distributed evenly over a grey scale from black at 0
to white at 255, a difference of one or two in the pixel value results in a grey¬
scale difference which is too small to be resolved by eye. Processing algorithms
known as contrast stretches are able to enhance the contrast in one part of the
image, at the expense of contrast in another part.
Contrast-stretch techniques are best understood in relation to the image
histograms which in themselves can carry some useful oceanographic informa¬
tion. Fig. 5.10 shows a typical histogram for a single-band visible wavelength
image. It will be seen that apart from a significant number of zero pixels, which
may be data dropout points, and a peak at 255 corresponding to cloud, most of
the pixels lie in the range 50 to 150. The cumulative histogram, showing the
number of pixels having a value less than or equal to a given level, shown in
Fig. 5.11, gives a clue to how an automatic contrast stretch can be applied. The
shape of the cumulative histogram is fitted between 0 and 255, and the resulting
curve (Fig. 5.12) is used to re-value each pixel. This is achieved on the computer
by creating a look-up table (LUT) to which each original pixel value is referred
before displaying it on the screen with its new value. It is easy to see that the
cumulative histogram for the modified image will be a straight-line ramp through
the origin (Fig. 5.13), and the actual histogram will be uniform (Fig.5.14). This
has the effect of spreading all the variabihty in image density evenly across the
range of grey levels, and it is therefore called a histogram-equalisation stretch.
It is clear from Fig. 5.12 that the range between 50 and 150 on the original
image which contained most of the pixels has now been stretched out to fill the
values between about 30 and 220, whilst the other parts of the full range have
been correspondingly compressed.
Fig. 5.10 - Typical image histogram Fig. 5.11 - Cumulative histogram of Fig. 5.10
126 Principles of image processing [Ch. 5
Fig. 5.12 - Calibration curve for histogram Fig. 5.13 — Cumulative histogram of
equalisation contrast stretch. image of Fig. 5.10, recaUbrated using the
curve in Fig. 5.12
their new position along the grey scale, and make an approximate estimate of
original parameter values. The other drawback of the automatic histogram-
equalisation stretch is that it may enhance a lot of variability on the image in
which the oceanographer has no interest. Thus, cloudy parts of the image are
enhanced to show different cloud shading, or land parts appear much clearer,
whilst if open sea covers only a small part of the whole image the automatic
contrast stretch may in fact compress the sea variability into a smaller range of
values. This can be avoided by masking out the land and cloud as discussed in
section 5.5 and applying the equalisation techniques to the histogram of the sea
pixels only.
In contrast to the nonlinear histogram-equalisation stretch, the linear
stretch applied to a selected part of the parameter range is slightly more time-
consuming to apply, may not result in such a good contrast being achieved over
the whole image, but is nevertheless more useful in most oceanographic applica¬
tions. Fig. 5.15 shows a typical example of the recalibration curve for a linear
stretch. In practice it is possible with proprietary software to make up a variety
of different stretches, to suit different applications, but the linear one has
the widest use in oceanography. The range between /jnin /^ax is stretched
to fill the whole 0 — 255 range. Outside and all pixels are set to 0 or
255 respectively, and therefore any variability in this range is lost. The art of
applying the linear stretch is to keep /max~-^min as small as possible, but to set
^min and to include all the parameter values for sea pixels. In practice,
Iinax
land (darkest pixels on a thermal IR image), the cold clouds, and the sea with an
intermediate range of temperatures. On the basis of such a histogram it would be
safe to choose and /^ax shown, to enhance the sea-surface temperature
variability patterns. Unfortunately, not all histograms are as clear-cut as Fig. 5.16.
The land and sea may be at similar temperatures, in which case masking using
another channel may be necessary. Moreover, there is usually a significant
number of pixels with values between the main groups, which can be explained
as those which contain both cloud and sea, land and sea, or land and cloud
within the IFOV. These may make it difficult to choose appropriate limits
for a hnear-contrast stretch. They also, incidentally, introduce errors in the
estimation of sea-surface temperature, as will be discussed further in Chapter 7.
Fig. 5.16 — Typical histogram of a thermal IR image of the British Isles taken in
summer daytime when the land is significantly warmer than the sea. Note the
distinctly separate land, sea, and cloud bands in the distribution
h
-
Ii +12+ h
h
/j + /2 4- It,
h
^3 =
I1 + I2+ I3
This approach has only limited value in oceanography, having been applied only
to Landsat images of coastal and estuarine waters.
Another technique which is used extensively in land application to
emphasise spatial structure in images is principal-components analysis. This
applies a linear radiometric transformation to N spectral bands to produce N
transformed orthogonal image datasets in such a way that the variability in the
image of the first transformed band (the first principal component) is maximised.
The second transformed band maximises the remaining variability, and so on.
This sometimes has the effect in ocean images of enhancing the land —sea —cloud
contrast in the principal component, since these features give the greatest vari¬
ability in radiance across the image, whilst the oceanographic features of interest
appear in the second or third components. Because it is a computer-controlled
transformation, which separates the user from the original satellite data values
in a complex manner, the quantitative interpretation of an image treated by
principal-components analysis is obscured, and its use in satellite oceanography
appears to be limited. Indiscriminate application of the technique is certainly
not recommended. A more thorough description of the principal-components
analysis technique can be found in the standard works on image-processing
mentioned at the beginning of the chapter.
/'.MTrase’"
' «■■ ‘.i '• <i ^r-T-t.4 ‘tf. * 'a -O 'i»- *•.. ‘
OCEANOGRAPHIC APPLICATIONS
OF SA TELLITE REMOTE SENSING
CHAPTER 6
6.1 INTRODUCTION
6.1.1 Scope
Of all the techniques used in ocean remote sensing, the observation of ocean
colour from satellites is perhaps the most easily understood in concept. For
as long as men have sailed the sea they have noted its characteristic colour
and sought to navigate, to fish, to chart its depth, or to seek out its riches by this
means. In the twentieth century the science of optical oceanography has grown
up in order to assist the study of water quality, and in particular the measure¬
ment of the primary productivity of planktonic organisms, which can be detected
by underwater optical observations. Most optical oceanography has concentrated
on subsurface measurements of scattering, absorption, attenuation, and their
variation with spectral wavelength. Apart from such easily-observed phenomena
as ‘red tides’ associated with certain plankton blooms, shipboard observations
of water colour from above the sea surface have not contributed very much to
marine science. However, since the use of colour photographs or scanned images
from aircraft and satellites has provided a wider spatial perspective to define
larger-scale structures in ocean colour, oceanographers have begun to see the
scientific, quantitative, possibilities in studying the colour of the ocean as viewed
from high above the surface.
In this, the first of the chapters devoted to individual techniques of
oceanographic remote sensing, we look in some detail at those sensors which
operate in the visible part of the electromagnetic spectrum. The principles of
atmospheric and underwater optics are introduced before considering the way
in which scanned ocean-colour data are processed, atmospherically corrected,
and analysed for oceanographic interpretation. Now that satellite ocean-colour
data have been received for more than a decade, there is a variety of application
areas to describe, making contributions to both the commercial and research
interests of marine science.
It is appropriate to consider visible-wave length remote sensing first, not
only because it was historically the first method developed to study the ocean
[Sec. 6.1] Introduction 135
from space but also because it is the only technique which penetrates beneath
the sub-millimetre surface skin and ‘sees’ directly into the surface layers of the
sea down to a depth of several metres.
For further reading on satellite observations of ocean colour, the 26 papers
on the subject in Gower (1981) give a broad coverage of recent research, and
reviews have recently been written by the author (Robinson 1983) and by
Gordon & Morel (1983).
Radiometric sensitivity
(noise equivalent reflectance %) 0.57 0.57 0.65 0.70
Fig. 6.1 — Nominal spectral responses of the Landsat MSS visible and near-infrared
channels (from NASA Landsat User's Handbook)
monitors. The scanner operates in the way described in Chapter 2, but because
the total field of view scanned is so small (less than 12 degrees) the scan mirror
oscillates rather than spinning right round. (On Landsats 1 to 3 the oscillation
period was 33 ms.) Another feature of the MSS is that in order to gain the high
resolution of 79 m scan-line width, without leaving large gaps between successive
scan lines as the satellite speeds on its way, it is necessary to scan 6 hnes simul¬
taneously, with 6 separate parallel sensors. Of course it is impossible for the
radiometers and filters to be exactly identical, resulting in a tendency for 6-line
stripiness to appear on the images reconstructed from MSS data. The other
pecuharity of the Landsat MSS is the way in which the scan samples are
obtained. Although the IFOV is nominally a square of 80 m sides at the satellite
sub-point, the radiometer is sampled while the mirror scans along the line at a
rate equivalent to a distance of 56 m at the sub-point. Fig. 6.2 illustrates the
overlap which consequently occurs, and the result is that the nominal pixel size
when reconstructing the image should be 56 X 79 m. This rectangular rather
than square pixel shape must be allowed for by appropriate image-processing
techniques. A display system which is always used for Landsat images can be
adjusted to generate rectangular pixels itself. If square pixels are displayed on
the screen a simple way of approximately removing the distortion is by sampling
every 3rd scan sample, but every 2nd line. If a full-resolution image is required
without loss of data, using a square pixel display, the whole image dataset must
be resampled using the techniques described in Chapter 5.
Sec. 6.1] Introduction 137
PIXEL
IFOV WIDTH
79m
Fig. 6.2 - Nominal pixel size of Landsat MSS compared with its IFOV
Since the swath width is 185 km, the data stream is framed into ‘scenes’
which are also 185 km long in the satellite track direction, but with 10% overlap.
Each scene therefore has 2340 scan Hnes, and 3240 pixels per hne, that is,
7 581600 pixels per channel, or over 30-miUion separate radiance observations.
Such high resolution, resulting in so much data, can be an embarassment to the
oceanographer who may be interested in much larger length scales of ocean
variability than the 80 m which is still often too coarse for land appHcations. For
this reason, and also since positional registration of a scene is difficult without
landmarks, Landsat is in fact of much less value over the open ocean than in
coastal and estuarine regions where the 80 m resolution can be exploited and
where coastal topography facilitates position fixing. The archives of Landsat
imagery have mostly been compiled with respect to the requirements of land-
use applications, and therefore there are few scenes available which do not
have some land on them, that is, there are few open-ocean scenes available.
A further serious drawback for oceanographic applications is that because of
the narrow swath width of 185 km and only a small overlap with the next day’s
adjacent overpass, the repeat period for imaging the same point is in general
18 days. Given the problems of cloud cover in many coastal and sea areas,
the frequency of useful scenes may be reduced to one or two per year — of
apparently little value in the study of continuously-variable dynamical processes.
Landsat MSS data therefore tend to be viewed by oceanographers as
‘second best’ compared with the observations gained from satellites and sensors
designed primarily for ocean applications. Nonetheless, nearly a decade of
archived Landsat data covering estuaries and coastal seas is now available, much
of it as yet unviewed and very little of it having been processed for atmos¬
pheric and geometric corrections. Despite the drawbacks mentioned above there
is considerable scope for extracting oceanographically-useful data from this
still-growing archive. Some of the marine application areas which have begun to
be exploited are discussed in section 6.4.
Landsat 4 marked a new generation of Landsat technology, and although it
carries an MSS essentially the same as the earlier ones it also carries the first of
138 Visible wavelength ‘ocean-colour’ sensors [Ch. 6
O
to o o
<N fN
(N o
NO lO X X
11 to
(N o
£ E
O o o
1-H fN rs
to
CO
CN NO o
1 to
cs (N fS
00
o
<N
Table 6.2 Characteristics of the Landsat Thematic Mapper (TM)
to
VO
1 to
to cs
to
o
ON
o NO
1 to
NO <N
83 X 10* bps
ON
NO O
m O
o NO
E o m
o
1 to
o
X lo <N X
CO
NO
tN
E +1 00 VO E
o o
o ro m
o
VO
o NO
1 to
(N <N
to
d
(S
to
d V£)
1 to
to CN
OCI
•o o
.p I-t
D.
T3 Oh
o 2
C
cd C w
i-t
*->
o 0) E
O
Oh o cd
G
P3
E
:t G(U E(U .in
Ocd
CA T3cQ
Ij) ot/i o
c a c c ccd o
o ao a> ■S O)
I
ccd
II
C/3 1 > G
cd p
U 3 o o o o
60 *s D a* cd
o D a> (D 6 E
Data rate
(l>
E c/}
3 5 .ac/)
•a .B o .2 ’o c <«-■
c c p 60 o
*3
E
o o o
I-t
X
ed
Ca> 6 .S
n Z 2 O s H-] 2 Oh
Sec. 6.1] Introduction 139
s
3.
lO
cvi
VO
T
IT)
o
o
00 On ON
lO cn ro
o (N (N
t/5 o
u
N
QJ
u
00
C +1 ro 00
g o (N
u
VO
<-i
3 'O
cd
(N
1636 km
O
u
+1
VO NO B 00 0ON 00
a> 00 (N in m X
m VO
fi o a^
o c4 NO CN (N
S +1
iri
lO
00
N
O (N
"a
-w 00
(/)
n
o O
u +1 NO
(N
a> O m
.a NO
i-i +1
V
■<-»
ro
(3 ■vt
JS
u
so
(when not tilted in track direction)
.S .s
cd
n CU!) tiO
H
><
cd a>
s c
33
cd
o C C
.q cd
o
<D cd S c/5
O 3^ C o Ui
c c *> th!
M-H O
o 4-. Oh
a.2 o
cd
C/5
C cd cd
c/5 > 3)
o c C4 43
I
b
cd c
T3 c:
o s ]3) C q TJ
O C o cd cd
Z nJ V-> u*
cd o c/5 '$
Ol Cd c 00
T3
c
4) Vh
3
-<-»
cd
I o
d
3 :3
oc o
q t-M
X
cd
o
d
43
"S
CQ C/5 Z < o s Z V3
Sec. 6.2] Optical theory relevant to ocean-colour viewing from space 141
Customary
Quantity Definition symbol SI unit
Radiant energy Q J
(joule)
Radiant flux or power </) W
dr
(watt)
Radiant energy density w J m'^
dV
(V = volume)
d0
Radiant flux density at a surface
dA
Irradiance (landing on surface) E Wm-2
Table 6.4 lists the quantities, their units, and their normal symbols, which
are used in the measurement of visible-wavelength electromagnetic radiation. As
presented they are total quantities integrated across a broad spectral range whose
limits are assumed to be implicit in the context of their use. Spcctful (juuntitics
can be represented by restricting each to a narrow wavelength band. This is
normally denoted by the use of subscript X, and the units must be altered by the
142 Visible wavelength ‘ocean-colour’ sensors [Ch.6
AREA, A
Sea Surface
area A
T T \ ' ''
<t>u
Fig. 6.4 — Definition sketch; downwelling and upwelling iiradiance, and
A
y 7! flux 6
the ocean surface as an extended light source. Fig. 6.6 illustrates the definition
of L, which is the radiant flux per unit solid angle in a given direction, per unit
projected source area in that direction. The units are therefore W sr“\ By
introducing the source area into the definition of radiance, the concept of the
brightness of an object is represented. Thus two sources of different size which
emit the same total flux isotropically will, at great distances, appear to have the
same radiant intensity, but viewed from nearby the larger source will have the
smaller radiance, corresponding to the fact that it is not as bright as the smaller
source. A Lambertian source is one in which the radiance L is independent
of the angle from which it is viewed. The assumption is normally made in the
viewing of ocean colour from space that the sea is a Lambertian source in
terms of the upwelling of backscattered light from below the surface. For a
Lambertian plane surface, L = Fu/rr.
K{\) --^
E{\,z) dz
where z is the depth within the medium. K is specified as downwelling or
upwelling according to the use of or E^ in the expression.
5= /“$;,i?(X)dX. (6.2)
0
Fig. 6.7 shows a typical sensor response function. The stated bandwidth of the
sensor, (X —Xi), is normally calculated such that
2
<I)(Xi to X ) =
2 (6.4)
^peak
but consideration of (6.1) to (6.3) reveals that assumption (6.4) is not con¬
sistent. Fig. 6.8(a) illustrates that when is smoothly varying with X, the
estimation of $ in the stated sensor range Xi, to X using (6.4) is entirely
2
assumed response
Turning now to the geometric properties of the sensor, Fig. 6.9 illustrates
the limiting ray paths from a small sea-area element 6.A within the IFOV of
the sensor. The sensor has an aperture ^4^ and is distant from the sea surface.
It is viewing at an angle d to the surface normal. If the surface radiance is L in
direction 6, then given the definition of L, the radiant flux from 6.A which is
intercepted by the sensor aperture is
where is the solid angle subtended by the sensor aperture as viewed from
the sea surface. Although not necessary, it is perhaps easier to understand (6.5)
by considering dA to be very much smaller than Ag. The total flux intercepted
Sec. 6.2] Optical theory relevant to ocean-colour viewing from space 147
by the sensor from the whole of its IFOV, which has area A, is obtained by
integrating the contributions of all d^l. The angle d will change very slightly for
each of these, but for a sensor with a narrow angular field of view the variation
of 6 will be negligible. The solid angle subtended by the sensor will be constant,
and L may change very slightly with 6, but on the assumption that the surface
is Lambertian, L will be constant. Hence the total flux is
Now A, being the IFOV, will depend on the solid angular field of view of the
sensor, a^, the distance from the satellite, and the surface inclination to the
viewing direction, that is.
A COS0
(6.7)
(6.9)
(i) are rays which are scattered towards the sensor by the atmosphere after
previous atmospheric scattering;
(j) are rays which have emerged from the water outside the IFOV and sub¬
sequently been scattered into the field of view. They do not contribute to Z-vv>
which by definition is the brightness of a particular area of sea;
(k) are rays scattered by the astmosphere towards the sensor, which have been
reflected from the sea surface outside the IFOV and therefore do not contribute
to Lp
(h), (i), (j), and (k) all contribute to L^, the atmospheric path radiance.
Thus if Ls is the radiance received by the sensor, it is made up of
water and turbid water. The reflected signal is small in this example because the
sun elevation was low at the time of observation (36°) and therefore only sky
glitter contributed. It can be seen that even in turbid water the maximum
percentage of the satellite signal which contains information about the sea water
is 35%, at 550 nm wavelength (yellow light). It decreases gradually towards the
blue end and rapidly towards the red end of the spectrum.
Z-p and T both depend on the scattering and absorption properties of the
atmosphere. Sturm (1981a) discusses these in some detail. Here we note that T
at a given wavelength of light, and over a path length / in the direction denoted
by distance coordinate s (see Fig. 6.11) is given by
K and hence r and T are due to the collision of photons of light energy
with air molecules and with aerosols. The latter are small solid or liquid particles
which may have been Hfted by wind from the earth’s surface, or sublimated
and condensed gases from the atmosphere. Since the molecular composition of
the atmosphere is generally uniform and well known, it is convenient to treat
molecular effects separately from aerosol effects which are variable in space and
time. The optical thickness of molecular effects alone, TmCX), is due largely to
scattering by air molecules, and depends very little on absorption except for
ozone which absorbs light around 600 nm wavelength. Since molecular scatterers
are small compared with the wavelengths of visible light, Rayleigh scattering
theory applies, and an expression for rm(X), (sometimes known as the Rayleigh
optical thickness) which is sufficiently accurate for remote-sensing applications,
has been derived:
p(0)-p(z)
rm(\,z)= 0.00879 (6.11)
Pn(0)
where the wavelength X is given in pm and p(z) is the atmospheric pressure
at height (z). Pn(0) is the pressure at sea level of a standard atmosphere of
temperature 15° C.
The ozone optical thickness, ro(X), due to ozone absorption, can also be
calculated on the basis of observations of ozone distribution in the atmosphere
(see Sturm (1981a) or McClatchey et al. (1972)).
It is not possible to calculate the aerosol optical thickness so readily. Mie
scattering theory applies in this case, since the particles are much larger than
molecular sizes. Sturm (1981a, 1983) discusses methods of calculating which
require a knowledge of the local meteorological visibility as a measure of the
amount of aerosol haze present, but this does not promise a reliable method
for atmospheric correction of satellite data. Some research indicates that the
wavelength dependence of is given by
rA(X,z) = A{z)\~^
B
rA(Xi) X2
(6.12)
rA(A2) Xi
-^R + • (6.13)
T'L^ now includes all photons which have penetrated the sea surface,
even though they may have emerged outside the field of view and been scattered
into the sensor by the atmosphere. It is appropriate for them to be counted as
‘signal’ rather than ‘noise’ because they do contain ocean information. However,
in order to include them in this way, it is necessary to use the diffuse trans¬
mittance T' rather than the direct beam transmittance T. This tacitly assumes a
Lambertian surface, and also that does not vary significantly between the
IFOV and the adjacent pixels which make a small contribution to T'L^.
The strategy of atmospheric correction is therefore to obtain Za and Zr,
whence Z'Z^ can be determined and then Z^ if Z'can be estimated. Now Zr,
the Rayleigh path radiance, can be calculated quite accurately, and Sturm
(1981a,b) gives the rather complex expressions necessary to achieve this. He also
presents an approximate expression for T', but the accuracy of this is not as
Sec. 6.2] Optical theory relevant to ocean-colour viewing from space 153
important as the correct solution of the path radiance, given the relative magni¬
tude of the terms indicated in Table 6.5. It should be noted that the important
oceanographic aim is to measure the variation of across the sea surface as
accurately as possible, and therefore whilst Lr and must also be measured
accurately, particularly the latter which is Hkely to be spatially variable, an error
in T' will not alter the estimated spatial variability of very much.
Fig. 6.12 - Optical pathways to sensor, grouped differently from Fig. 6.10 (from
Robinson (1983))
~ *^(^2, ^l)•
(6.14)
^a(^2)'^a(^2)
where (6.15 b)
<*^a(^i)’'a(^i)
Wavelength
R=E,/E,
In the case of a flat surface, optical theory shows that R can be related to
Iw by
i (6.16)
where n is the refractive index of water and p is the Fresnel reflectivity of the
water—air interface which causes some of F^ to be reflected back down into
the water. 6' is the subsurface direction of a ray which emerges in the direction B
(the viewing angle of the satellite), and therefore n sin0 = sin0 (see Fig. 6.15).
If the surface is rough, then (6.16) will not be strictly valid since the rays of
hght leaving in the direction of 6 will make a different angle with the surface
156 Visible wavelength ‘ocean-colour’ sensors [Ch.6
normal. Provided that 6 is not close to the critical angle of about 48°, however,
a roughened surface will not introduce much error. If the surface is very rough,
of course, whitecaps and foam appear, and the whole approach is invalidated.
Assuming the ocean is a perfect Lambertian surface, Q should be direction-
independent and equal to a value of n. However, Austin (1980) has suggested
from observations at sea that Q should in fact be approximately 5.0 for near¬
nadir observations. Given Ly, from the satellite sensor, R can then be estimated
using (6.16), by calculating from the known sun illumination at the top of
the atmosphere, and estimating the atmospheric transmittance. Strictly, account
should be taken of the dependence of the refractive index on sahnity and
temperature, but given the much larger errors of estimating atmospheric inter¬
ference this is probably not worth considering. Since Q and are only weakly
wavelength-dependent, it often more satisfactory, and easier, to cancel them out
in the reflectance spectral ratio:
b,10^m’
Fig. 6.16 - Absorption (a) and scattering (b) of pure sea-water (after Robinson
(1983))
400 700
X,nm
with wavelength. Adding Fig. 6.18 to the absorption curve of Fig. 6.16 explains
the name given to yellow substance, since the least absorption is in the yellow,
middle part of the spectrum. Yellow substance not associated with local
phytoplankton productivity is thought to derive from land drainage (Hojerslev
1980), and is found in strong concentrations in enclosed seas such as the Baltic.
If yellow substance can be detected from space, it might therefore serve as an
indicator for the plumes of river discharge and the dispersal into a shallow sea
of fresh-water runoff.
The other, and most diverse, contribution to ocean colour is the presence
of suspended particulate material not related to phytoplankton. This may be
resuspended bottom sediment, riverborne sediment, eroded coastal and beach
material, or due to the dumping of sewage-sludge waste or dredging spoil. Its
spectral characteristics are as diverse as its composition in terms of natural
material colour and size distribution, and there appears to be no consensus in
the literature regarding a standard spectrum shape. Sathyendranath & Morel
(1983), in their review of the theoretical and experimental studies in the litera¬
ture of the optical properties of suspended sediment, are able to point to
Sec. 6.2] Optical theory relevant to ocean-colour viewing from space 161
Fig. 6.19 - Typical reflectance spectra for case 1 water. The arrow shows which
curves correspond to increasing chlorophyll (and related suspended particulate)
concentration. The dashed line is the dear-water spectrum (from Robinson
(1983))
162 Visible wavelength ‘ocean-colour’ sensors [Ch. 6
absorption. Because the spectral shape does not change much with sediment
load, the water colour does not change, and a sediment algorithm based on
spectral ratios appears less hopeful. Instead, an algorithm relating suspended load
to the absolute reflectance in a single waveband is more appropriate. One excep¬
tion to this occurs in those situations where the sediment is highly coloured, and
the strength of the colour is likely to be correlated with the amount of sediment.
In this case the reflectance spectra would be rather different from Fig. 6.20,
with a peak corresponding to the colour of the sediment.
Fig. 6.21 shows reflectance spectra for case 2 waters in which yellow
substance dominates. Like the chlorophyll spectra in Fig. 6.19, the measured
concentration leads to increased absorption and a decrease in blue reflectance.
Consequently a calibration algorithm based on spectral ratios should be more
promising in this case, although there is less of a well-defined hinge point for
normahsing the spectral ratios.
400 700
X,nm
SENSOR SUN
Fig. 6.22 - Specular reflection from a wavy surface. For a given solar and sensor
direction as shown in the upper diagram, direct reflection only occurs for a small
range of surface slopes, illustrated in the centre diagram. The result is a highlight
on the image corresponding to the location of such slopes, as illustrated in the
lower diagram
The contribution due to sun glint depends on the size of 6f', and the extent
to which fo is achieved within the sea area being viewed. 5f' depends on the
solid angle subtended by both the sun’s disk and the satellite sensor aperture
at the sea surface, fo depends on the zenith angle of both the satellite and the
sun. Whether fo is achieved depends on the sea state. If the sun and satellite
are both overhead, fo = 0 and a calm sea will give total specular reflection. If it
is roughened, the sun glint will be decreased because much of the sunlight will
166 Visible wavelength ‘ocean-colour’ sensors [Ch.6
soon be reflected away from the sensor, but the area of sea over which the sun
ghtter pattern occurs will be increased because off-nadir parts of the image
will now have facets producing direct reflection of the sun. Thus for all view¬
ing directions except the one which produces direct specular reflection in calm
conditions, increased surface roughness will tend to increase the chance of some
specular reflection occurring. Most satellite orbits are planned to avoid the
occurrence of sun glitter, by ensuring that fo for the given satellite/sun geometry
is larger than is likely to be encountered at sea. If fo is in a range which may be
encountered in a sufficiently rough sea, then it is necessary to apply corrections
based on the sea-surface roughness, or the wind speed, in order to remove the
extra radiance due to specular reflectance which is measured by the satellite.
Bernstein (1982a) illustrates the use of such a correction in the context of
reflected infrared radiation.
ramp
Fig. 6.23 - Schematic of data flow in the Coastal Zone Colour Scanner
Sec. 6.3] Calibration and application of CZCS data 167
Os VO C^
Tt VO VO OS
<N CO <N
t-H
a oq <N Tj-
4> a>
O o o O
cn ^ d d d d
CO '2
'4-^ lO s
•S *o
so
CO
Os
<N
CO VO
VO <N r- »-H
CO CN
o o o o
d d d d
o VO CO CN
v ^^ o 00 VO CO
o. c CO o VO
o Q oq
o o
«—1
o
o
o
e
o
^ s d d d d
.w VO ^
es
IM <
CO
r'
rs
o
r--
CO
00
CO VO 1-H
W CO (N 1-H
o o O o
o d d d d
(U
c VO VO 00
.3 CO
Os
CO
CO
(N
Qq CO (N o
2 o o o o
1 00 d d d
lo d
O ON
ON CO
lo •<
VO ;2: r- r-- 00
s C os CO CO Tf
c
o fa VO
CO CNJ 1-H
« 5 o o o o
Vi ^ d d d d
U b.
N .2d
1-H Os Tj-
W 13 o C-- VO 00
<u '—' CO VO Os
Si ^ oq (N
o
(S o
o
O o
<*- .S Tj* "*3^ d d d d
O C3 00 CO
.—V ®® ^ <
(N 2
2> « VO
00 CO CO
VO
VO 1—:
e3 X
o
CO
o
(N
o
1—1
O
C d d d d
JO
ll CO (N VO VO
o VO
OS
VO
CO
os
OS
CO
1-H
cs
CO
oq CO VO 1-H CN
■*-* o o o O o
XI
d d d d d
2f
S 1
J3 0) <N CO r-- VO VO
•M Cu U-) o VO CO 1-H
1-H 1-H
O CO <N 1-H OS
C o o O o O
_o d d d d d
.3
c3 t-.
> X) (U CN CO VO
C CO (U <D <u 0) CD
3 q> c c G C c
NO
C ^ c c c c c
o 03 C3 CT3 d ca
NO X X X X X
X ^
<U o o u U u
JO
«
H /
168 Visible wavelength ‘ocean-colour’ sensors [Ch. 6
Cg, Cc, and Q are transmitted in the data stream. V} is also independently
coded in the extra calibration data inserted in each line. Hence it is possible to
monitor the amphfier/digitiser characteristics, Oy and where
(6.21)
and
(6.22)
fc (6.23)
(Cc-by)
that IS, fls = -.
Qy Lq
~ ~~ (6.24)
Cly Gy G^
or Ls = ^cQ + ^c
1
where Ac and (6.25)
fly flj Uy Ug
It has been found that and Be do not change significantly from one scan line
to the next, and therefore it is customary to evaluate A^. and once only for a
scene and apply (6.25) to all pixels in the image. There are slow fluctuations in
Ac and B^., as Table 6.6 illustrates. A general sensitivity loss of the sensor with
time has also been noted and possible corrections for it are discussed by Sturm
(1983). This may be due to degradation of the on-board calibration lamps, in
which case the above calibration technique cannot be regarded as absolute, and
it becomes difficult to compare images which have been observed several months
or years apart. Absolute comparisons then require assumptions to be made
about the nonvariability of Lyy in certain clear-water areas.
^a(^2) _ ^
COa(^i)
and in (6.12) the Angstrom exponent, B is zero. Fig. 6.24 offers a possible
way of iterating for a better estimate of S, e, and hence r a which is required to
improve the estimate of T'. This iterative procedure works if there are several
areas of dark-water pixels, which must be well away from cloud and land, and
should also be fairly close to the satellite subtrack (small viewing zenith angle 0).
It must also be oceanographically reasonable to assume that these are clear water,
with very low turbidity because the concentrations of sediment, chlorophyll,
and yellow substances are all low. In this case the water-leaving radiances
in channels 2 and 3 can also be assumed to obey the formulae stated in
Table 6.7(D). The estimates of S and e made from these groups of pixels are
then applied to the rest of the scene. This allows a pixel-by-pixel atmospheric
correction, but assumes that the aerosol does not change its optical properties
significantly over the scene, other than its optical thickness.
Another route to S is to measure Z-vv,\ for all four channels at various
locations synoptic with the overpass. The correction algorithm scheme in
Fig. 6.24 is then entered both from the top and through the box at the bottom
left which leads to an estimate of e for the ground-observation pixels which is
assumed to hold over the whole image.
Pixel-by-pixel correction is only possible if one of the bands can be assumed
to carry information about Z-a without anyLw contribution. In relatively clear,
deep-sea water Z-w,670 0 rnay be a reasonable assumption, but in a plankton
bloom, or in turbid coastal waters, this is not the case. If Z,w,670 is small com¬
pared with Za,670 fh® iterative loop on the right can be used. If case 1 waters
are being studied then a reliable algorithm is available to predict Z-w,670 froiT^
a knowledge of Z-vv ii^ channels 1 to 3. In turbid coastal waters dominated by
170 Visible wavelength ‘ocean-colour’ sensors [Ch. 6
s
o
u
oO
a>
a
a>
•si
c/5
00
5U
1 N
•S
O- o
2 E
o. o
o ^
T3 ""
C
3
o!U
oVh
CL,
I ,
CN
.s?
Sec.6.3] Calibration and application of CZCS data 171
e(X„X.) = (^j .
e(443,670) = [-)
Sturm (1983) points out that the iteration should only proceed from this
point provided V443 ^ V520 - V550
and 0.3 < V443, V520, V550 < 3.0.
(F) Use an average value of /-a,670 from clear-water pixels to calculate
_ 0.26 COSgLA,670
7’o,(67O,0,0o)
and then use e(X, 670) from (E) to obtain
'^A,x — '^A,610^(\^70)
(G) For case 1 waters, Smith & Wilson (1981) have shown that,
. 1.661
^ss,443 ^ss,443
= 12.06
•^ss,670 \-^ss,550/
The state of the art is therefore promising for the open ocean, but still
unclear for turbid coastal waters. The principal difficulty lies in the fact that
channel 4 is at 670 nm and for many waters Z-w,670 0- W^re it 150 nm greater,
then only the most turbid fluid mud suspensions would produce a water-leaving
radiance at that wavelength, and it would serve as an independent channel
for atmospheric effects. It would be harder to estimate S(X) over the wider
wavelength interval, but that would be a lesser problem than is presently encoun¬
tered in atmospherically correcting CZCS data in coastal areas. In principle,
channel 5 of the CZCS (700—800 nm) should be useful for atmospheric correc¬
tion, but its poor radiometric sensitivity and its broad bandwidth limit its value
to approximate but fast atmospheric filtering algorithms.
-1.71
■^ss,443
C- 1.13 used for C <l .5 idg \ ^
_-^ss,550_
(6.27)
-2.439
■^ss,520
C = 3.326 used for C> 1.5 //g 1“*
_-^SS, 550_
where the subsurface radiance 4s rather than the water-leaving radiance Lw has
been specified. As in (6.16) we have
174 Visible wavelength ‘ocean-colour’ sensors [Ch. 6
L ss (6.28)
1-P(0')
where the refractive index n and surface reflectivity p are weakly wavelength-
dependent, and p is also dependent on the view angle, and must therefore be
evaluated separately for each point along a scan line. It will be noted that
^ss,443 is used only for low-chlorophyll concentrations because at higher values
the absorption becomes so great and Z,w,443 correspondingly so small that its
accuracy in a ratio algorithm is not adequate.
Equation (6.27) is not the only algorithm to be developed, and Sturm
(1983) lists a variety of others. There is significant variability, and in general the
coefficient A (in (6.26)) is lower for chlorophyll algorithms in coastal waters,
because of yellow-substance absorption which reinforces the chlorophyll effects
on the spectrum shape. In case 1 water, an accuracy of log C within is
claimed with the NASA algorithm (Clark 1981) appUed universally.
Algorithms for total suspended sediment in case 1 waters are quoted by
Sturm as:
-0.88
■^ss,443
S = 0.4 mg 1-1
-■^ss,550_
-1.09
•^ss,443
5 = 0.33 mg 1-1 (6.29)
-^ss,520.
-4.38
-^ss,520
S = 0.76 mg 1-1
j^ss.SSO-
The apparent success achieved with CZCS data over case 1 waters can be
misleading if taken out of context. The only reason why a good sediment
algorithm can be achieved with spectral reflectance ratios is that the particulate
load is controlled by the phytoplankton population which has a chlorophyll
colour signature. Hence the variables C, S, and A:(X) are not independent in
case 1 waters. In case 2 waters, no such correlation can be found between
Sec. 6.3] Calibration and application of CZCS data 175
synoptic oceanography. They reveal new aspects of the spatial and temporal
distribution of chlorophyll and sediment in detail previously unobtainable from
ship-sampling methods. If oceanographers, in their use of such images, are to
move beyond the descriptive phase, they need to formulate the questions that
will lead to new theoretical understanding of the factors controUing the observed
distributions. One promising field of study is the way physical processes -
advection by residual flows and upweUing, mixing across fronts, and stirring by
tides or storms etc. - control the biological processes. An example of this is the
study of Gulf Stream rings by Gordon et al. (1982) illustrated in Fig. 6.26.
■.ir'Y
& Baker (1982), Smith et al. (1982)). Total productivity must be integrated
throughout the whole water column. The major but not necessarily the entire
contribution to the total comes from the upper well-lit few metres which control
the satellite colour data. It is therefore necessary to construct a model of the
vertical structure of productivity in order to extrapolate from satellite observa¬
tions to total productivity. This is an area requiring further research along the
hnes of Platt & Herman (1982). Previous estimates of global and regional
productivity have been based on ship observations which, given the spatial
variability revealed by the CZCS, may turn out to be unrepresentative.
Another area of study which has already benefited significantly from the
CZCS imagery is the identification of productive regions in the vicinity of
oceanic or shallow-sea fronts. With a CZCS image it is possible readily to deter¬
mine the location of productive patches relative to the position of a thermal
front or other physical feature. Whilst ship sampling can offer a view of the
vertical structure of such patches, it cannot provide the same perspective as the
satellite in both horizontal directions. Fig. 6.27 is a good example. In this image
of the English Channel regions of enhanced productivity are revealed by multi¬
channel processing after atmospheric correction. A productive patch, probably
a dinoflagellate bloom, stands out where it would be expected along the Ushant
Front at the transition between tidally mixed water to the south and east and
stratified water to the north and west. The front allows the nutrients from the
lower layer on the stratified side, where it is too deep and dark to be used up by
plant growth, to be mixed towards the surface on both sides of the front, where
it is able to support primary production. A second productive region in Fig. 6.27
marks the line of the continental shelf edge. Although the water is stratified
here, and in general the nutrients are found in the lower, unproductive layer, the
presence of the shelf break gives rise to mixing processes which locally transport
nutrients into the upper layer where they can support growth. The most prob¬
able cause of the mixing is the generation of internal waves by tidal flow onto
and off the shelf. The shelf-break patch appears unusually strong because it is
probably due to the highly reflective species of Coccolithophores.
In the biological oceanographic applications of the CZCS so far reported,
satellite observations have been used to extend the usefulness of, but not to
replace, shipbome measurements. One near-real-time apphcation which requires
good communication from the receiving station to the oceanographer at sea is
to use the satellite image to locate the most active areas of productivity (for
example, an algal bloom). A research vessel can then be directed to those
places where physical and biological parameters require detailed vertical
profile sampling for a period of days to weeks. During this time the wider area
in which the particular process is set can be regularly surveyed by the satellite.
Holligan et al. (1983) have demonstrated the effectiveness of combined ship and
satellite surveying of a dinoflagellate bloom in the English Channel such as that
illustrated in Fig. 6.27.
Fig. 6.27 - Image of the Celtic Sea and N. Bay of Biscay processed to enhance
CZCS data recorded on 22 June 1981 to reveal areas of high productivity. The
areas where productivity is greatest are at the shelf break (A) and along the
Ushant Front (B). Both of these are regions where mixing processes cause
nutrients to be brought into the weU-lit surface layer so that productivity can be
supported. Note the streamer S indicating a southward transport of the shelf
break bloom, probably caused by a pair of eddies in the Bay of Biscay (see Pingree
(1984)). The white straight lines across the image are due to data dropout in the
transmission or reception. (Image provided by R. Pingree, lOS)
species of fish in particular colours of water, since the colour is strongly cor¬
related with the origin, temperature, and nutrient levels of the water masses in
this sea area.
Fig. 6.29 — CZCS channel 2 Image of the English Channel recorded on 26 March
1981. The more turbid water along the English coast is swept further south by
headland eddies such as that off Portland BiU (P). Note also the fine structure
lined up east-west with the tidal streams (e.g. at a) which is not yet fully
explained, and the banded feature (b) to the west of Jersey referred to in the
text. (Image supplied by and reproduced with permission of Dundee University
Satellite Receiving Station)
182 Visible wavelength ‘ocean-colour’ sensors [Ch. 6
sensors, there is every reason to expand our theoretical framework to include all
aspects of spatial variability. So far, the testing of models of spatial variability
has largely been limited to advective systems in which the length scales could be
inferred from a point-measured time series, but the satellite scanner has removed
that limitation. It is worth noting also that the measurement of the length scales
of dynamical processes from images is not dependent on the development of an
accurate calibration algorithm. Provided that spatial variability of atmospheric
effects is removed, accurate length scales can be obtained even when there is
only a crude calibration available, in just the same way as frequency peaks can
be readily obtained from a time series whose amplitude calibration is suspect.
As more CZCS imagery becomes available for study, it is to be expected
that many new dynamical features such as those on Figs 6.29 and 6.30 will be
found, and as a result we can expect exciting developments in physical oceano¬
graphy in the next decade. Indeed, further close scrutiny of Fig. 6.30 will reveal
a long narrow feature over the centre of the English Channel which corresponds
exactly with the presence of the Hurd Deep — a long narrow trench where
the depth is 110 m compared with the 70 m surrounding it. How it comes to be
revealed in a CZCS image is at present an open question whose answer may lead
us into more areas of new oceanographic understanding.
Because of the infrequent repeat cycle, synoptic sea data have also been difficult
to obtain, and it is much harder to test the accuracy of calibrations.
Most attempts have been made to relate the satellite radiances to the
suspended sohds concentration. Algorithms fall into two broad categories,
those which use colour (multispectral information) and those which use
intensity (single-band information). Since, in general, non-chlorophyll-related
suspended solids do not have a strong colour signature, it is more useful to
seek simple relations between S and the reflectance in a single band. Often
band 5 has been the most useful since it is less contaminated by atmospheric
effects. For low concentrations (<60 mg T^) a linear relation is satisfactory,
whilst over a wider range of sediment loads the reflectance is better correlated
with log S (Munday & Alfoldi (1979)).
The most successful use of colour correlation with S has come from the
use of chromaticity analysis, as described in section 5.9. A good example is
presented by Alfoldi & Munday (1978), and is illustrated in Fig. 6.31. Radiance
values for Landsat scenes of the Bay of Fundy, Canada, were mapped on to
a two-colour chromaticity transform plane, and S (measured simultaneously
with the overpass) was correlated with the position along the resulting locus.
Atmospheric contamination was treated by shifting the whole locus to a
normalised position. The success of this technique in studying the Bay of Fundy
is probably due to the highly-coloured red sediments which are input by local
rivers, resulting in a strong colour signature to the concentration. This causes
Fig. 6.31 Radiance values from Landsat MSS channels 4 and 5 viewing Minas
Basin in the Bay of Fundy, Nova Scotia, plotted on a Chromaticity Transform
plane. There is a correlation between the suspended solid concentration measured
synopticaUy in situ and the position along the chromaticity locus of the
corresponding pixels (adapted from Alfoldi & Munday (1978))
Sec. 6.4] Oceanographic uses of Landsat MSS data 187
the locus in the transform plane to lie at an angle to both axes, so that better
resolution of sediment values is obtained by correlation of sediment with the
position along the locus than by using the position along either axis.
MacFarlane’s haze-corrected data for the Solent estuarine area in the south
of England was also calibrated against synoptic sediment data (MacFarlane &,
Robinson (1984)). As well as correcting for atmospheric effects, the radiance
values which came from four separate overpasses have been normalised by
eo-i
0 i 1-1-1-1 ■■ 1
20 40 60
scaling with the factor l/cos0o to allow for varying sun illumination from scene
to scene (where do is the sun zenith angle). Fig. 6.32 shows the resulting cor¬
relation that can be achieved between a satellite measurement and the suspended
sediment measured gravimetrically in samples collected within the top metre of
the water column. Figs 6.32 (a) and (b) show the poor correlation between the
sediment and the raw satellite data (the MSS digital counts in channels 4 and 5);
(c) and (d) show the good correlation possible after atmospheric correction and
normalisation for solar illumination; (c) is a correlation with the channel 5
reflectance; and (d) is for the chromaticity-transformed channel 4 data. In both
cases, the atmospheric correction and illumination normaUsation has brought
the groups of data points from different overpass dates onto or close to a single
straight line. The conclusion may therefore be drawn that given sufficient field-
measured sediment data synoptic with an overpass, and by applying fairly
lengthy processing operations to the data, it is possible to achieve a multidate,
site-specific sediment calibration algorithm for Landsat MSS. Because we are
usually deahng with case 2 waters, and the sediment optical properties can
be expected to vary from estuary, to estuary, a universal algorithm in terms
of sediment seems unlikely. However, it may be possible to develop a universal
algorithm which relates atmospherically-corrected, illumination-normalised
Landsat data to an optical field measurement of turbidity. It would then
be possible to achieve caHbration of previously-received Landsat data in terms
of the backscattering properties of the water. This could be interpreted in
terms of the local type of sediment by field measurements made completely
independently of further satellite overpasses. If the appUcation merely called for
optical turbidity information there would be no need to visit the estuary at all.
Although such a universal algorithm seems promising, it is yet to be developed.
Fig. 6.33 - Landsat MSS image of R^, the subsurface reflectance ratio in
channel 5, for the west Solent, UK. Atmospheric correction based on channel 7
data and an 8 X 6 smoothing filter have been applied. Note the banded structure
between Cowes and the mouth of Southampton Water, which is ahgned with
strong tidal streams flowing westwards through the Solent at this state of the tide,
VA hours after high water. The image was recorded on 23 Feb. 1979. (Image
processed by N. MacFarlane at Space Dept., RAE Farnborough, UK)
It is important when analysing images such as these to make sure that the
patterns observed are genuinely of suspended sediment distribution. If it is not
possible to make an inspection at sea, then it can often be helpful to look at
several images. Those features which are permanent, or are consistent with tidal
conditions, are likely to have a marine rather than an atmospheric explana¬
tion. It is harder to be certain that it is not a surface-roughness pattern which is
being observed. Like the sediment, this could line up with tidal flows, revealing
streaks of faster-flowing water which against a headwind could produce a region
190 Visible wavelength ‘ocean-colour’ sensors [Ch.6
of whitecapping. Since the sediment patterns being sought give rise to very small
differences in radiance, it is not unreasonable to expect that whitecaps could
produce a comparable radiance change. The availability of sea data, even if not
synoptic, will help to resolve the issue. Great care is required in interpreting
images of remote areas when no first-hand knowledge of the local oceanography
is available.
If the interpreter is reasonably certain that sediment patterns are being
imaged, there remain two problem areas to be overcome before the satellite data
can be used for studying sediment dynamics. The first is that the satellite only
responds to sediment loads in the one to five metres or so near the surface,
and less for very turbid water. Yet the potential appHcations of the data in civil
engineering studies of sedimentation, deposition and scouring, and sediment
transport, are all much more concerned with either bedload transport or the
concentration of near-bottom suspended sediment. There is a need to gain .much
more knowledge and understanding of the vertical distribution of suspended
sediment throughout the water column before the use of Landsat data can be
fully integrated into offshore and estuarine sedimentation research programmes.
The other problem is that with such infrequent cloud-free Landsat imagery,
it is not usually possible to obtain time sequences to detect even seasonal pattern
changes. Of course, given a knowledge of the tidal flow patterns in an estuary, a
single image revealing a synoptic sediment distribution can be readily interpreted
in terms of sources and sinks of sediment, direction of sediment movement,
etc., bearing in mind the limitations of a near-surface view as discussed in the
previous paragraph. There are ways, however, of gaining a measure of time-
sequence information. One is by using the overlap region to obtain passes on
adjacent days (possible with Landsats 1-3 but not with the orbit of Landsat 4).
Indeed, at higher latitude with a greater overlap it is occasionally possible to
obtain images on three successive days, as demonstrated by Horstman & Hardtke
(1981). Another approach in a tidaUy-dominated estuary is to arrange the avail¬
able archived Landsat images in sequence according to the stage of the tide
within the 1214-hour semidiurnal cycle at the time of the overpass, irrespective
of the actual time or date of the image. This has been shown to make a fairly
consistent interpretation possible in the Solent (Robinson & Srisaengthong
(1981)). For example, the patterns shown in Fig. 6.33 are repeated almost
exactly on other images viewed at the same time of between 1 and 2 hours
after local high water, but months or years apart in real time. In contrast,
other images for other states of the tide show a marked difference in pattern^
consistent with different tidal flow conditions.
sake, nor is accurate calibration essential, the spatial patterns being the important
information. On many Landsat images of the coastal zone it is possible to see
the plumes of discharging rivers and large sewage outfalls, and on occasions
the plumes of sludge dumped by barges at sea. Whilst satellite surveillance does
not have such a major role to play in the pollution monitoring of estuaries as
does airborne remote sensing, it is still possible to learn a lot about the dispersal
of discharged matter by studying several Landsat images from different states of
the tide.
Sometimes the contribution to pollution control can be more subtle,
as in the work of Klemas (1980). This examined the position of small frontal
structures in Delaware Bay which had a definite colour signature on the Landsat
imagery, and which also were influential in pollutant dispersal. It is in its ability
to offer a synoptic view of estuaries that Landsat has considerable potential
in coastal oceanography. Coastal areas can have such a complex combination of
eddies, fronts, and plumes that it is sometimes difficult to interpret observations
from ships, whereas the satellite image can place these in context.
As with the CZCS, Landsat MSS can also reveal interesting biological
features such as the plankton bloom viewed in the English Channel on Landsat
images (Le Fevre et al. (1982)). Further study of why this should have such
a strong sediment-hke signature revealed that it was due to a species of
Coccolithophores with calcite scales which outlasted the life of the organism and
produced a dense suspension of particulate material following a bloom (Holligan
et a/. (1983)).
Another interesting use of Landsat imagery apparently viewing another
plankton bloom, this time in the North Atlantic, is the work of Gower et al.
(1980). They analysed the spatial variability of the radiance in the Landsat data
and used this to comment on the turbulent eddy length scales of the sea area.
This is an example of the innovative approach to exploiting satellite data in
dynamical oceanography which was discussed in section 6.3.5.
where a chart is not available, or for checking the vaUdity of existing but old
charts. The fact that Landsat will not penetrate much below 20 m in clear water
is not too great a drawback, since most vessels that will use the charts for naviga¬
tional purposes have a draught less than 20 m. Fleming and Le Lievre (1977)
were able to chart features off the Labrador coast which could be subsequently
verified by a survey vessel.
In many locations, the bathymetric application is hindered by the presence
of sediment. By using both bands 4 and 5, and even 6, it ought to be possible
to begin to distinguish between the backscatter due to sediment and that due to
the sea bed, provided that suspended sediment is uniform with depth and has
a uniform reflectance over the spectrum. Then in deep water only the sediment
will show up, uniformly in all bands. But as shallow water is approached the
reflectance in band 4 will increase more rapidly than in band 5. Cracknell et al.
(1982) attempted a quantitative study of bathymetry in the presence of
sediment for the Tay estuary, but with inconclusive results.
daily shortwave energy budget over the ocean. Essentially, the method simply
detects whether clouds are present or not within each pixel, and uses an appro¬
priate model to calculate from known atmospheric absorption effects how much
solar shortwave energy is reaching the sea. By measuring the brightness of cloud
tops with the satellite sensor, it is possible to make a crude estimate of the
absorbing properties of the cloud and improve the estimate of how much short¬
wave energy penetrates through the cloud. The insolation is modelled every
hour with the hourly updates of the scanned image, enabling a daily total to be
calculated. Therefore the satellite data is used in a novel way towards producing
a chmatology of shortwave energy into the sea.
A further development of this will be to use spectral information from the
CZCS to estimate the depth of energy penetration into the sea, and thence
to model more accurately the thermal structure of the upper ocean and its
horizontal variability. Simpson & Dickey (1981) and Dickey & Simpson (1983)
have already shown how such information could be put to valuable oceano¬
graphic and climatological use. As oceanographers become more familiar with
ocean-colour data from satellites it is likely that many other applications will
come to light.
CHAPTER 7
7.1 INTRODUCTION
The monitoring of sea-surface temperature (SST) from earth-orbiting infrared
radiometers is the technique of marine remote sensing which has had the widest
impact on oceanographic science. For example, the displaying of an IR image
of a frontal or eddy structure is quite commonplace within a scientific paper
describing observational or theoretical studies of a dynamical feature. Amongst
the reasons for this must be included:
(a) The very good correlation, even without complex atmospheric correction or
calibration algorithms, between light and dark shades of grey on an IR image
and cooler and warmer water, giving a valid qualitative view of a thermal feature
in the ocean.
(b) The operational nature of the meteorological polar-orbiting satellites in the
tiros/noAA series and the Geostationary Meteorological Satellites, which has
ensured a regular and continuous supply of data from the infrared radiometers
which are carried on them, primarily for meteorological purposes but with a
ready capability for ocean surveillance given cloud-free conditions.
(c) The wide dissemination of the IR data, due to the Automatic Picture
Transmission (APT) facility on both series of satellites carrying IR radiometers,
and also the comparative ease of receiving the full resolution data at receiving
stations around the world. For example, British oceanography has benefited
greatly from quick and easy access to IR data through the Dundee University
Satellite Receiving Station, an example of how high-quality data reception
can be achieved at low cost (Bayliss 1981a,b). This is only possible with the
relatively low data rates required by the current generation of IR sensors, and
high-resolution imaging will require higher-frequency data transmission which
demands more sophisticated and expensive ground stations.
The principal source of infrared data for sea-surface temperature measure¬
ment is the TIROS/NOAA series of polar-orbiting meteorological satellites.
Tables 2.2 and 2.3 list the general characteristics of these satellites in com¬
parison with others, and details of the sensor operating wavelengths and image
[Sec. 7.1] Introduction 195
size are presented in Table 7.1. The Advanced Very High Resolution Radiometer
(AVHRR) was first flown on TIROS-N in 1978. Its four spectral channels, two
visible and two infrared, and its digital data transmission with 10-bit binary
word size (1024 levels) were a distinct improvement on the VHRR flown on
the previous NOAA 5 satellite which had just one visible and one IR channel and
an analogue data transmission with about 100 radiance levels. An AVHRR was
flown on NOAA 6, whilst NOAA7 carried an AVHRR/2, which has an extra
channel in the infrared. Being an operational meteorological satellite programme,
there is always at least one functioning satellite in orbit. The original operational
system had two satellites working, in sun-synchronous orbits separated by about
90° (approximately 6 hours) of longitude, so that one would give a morning and
the other an afternoon daytime overpass at each location. The trimming of space
programme budgets may mean that satellites will operate only singly in future,
with the sensor possibly alternating between AVHRR and AVHRR/2.
Spin rate
100 rpm
Time to sample one line 30 ms
Time interval between adjacent lines 600 ms
Radiometer telescope rotation between lines 0.125 mrad
Total north-south scan angle 18°
Total time to produce one image 25 min
Image repeat time 30 min
t Scan option b
Sec. 7.1] Introduction 197
Table 7.2(b) Spectral characteristics of the GOES visible and infrared spin scan
radiometer VISSR
As weU as the visible and the IR radiometers, both series of satellites carry
atmospheric sounders which measure the radiance at wavelengths where there
is no clear atmospheric window through which to view the earth. These provide
information about the atmospheric composition which can help in the atmos¬
pheric correction of the SST data. On the NOAA series this is the 20-channel
High-resolution IR Spectrometer (HIRS), with a ground resolution of 17.4 km.
On Meteosat there is a single-channel water-vapour (WV) radiometer, whilst
GOES 4 and 5 carry the VISSR Atmospheric Sounder (VAS).
Other satellites which have carried infrared radiometers capable of yielding
sea-surface temperature data are listed in Table 7.3. The HCMR is particularly
useful for measuring diurnal temperature fluctuations, and Deschamps & Frouin
(1984) present an oceanographic appHcation. However, some of the instruments
in Table 7.3 have been beset with problems, or have not provided such high-
resolution or drift-free data as was expected, and in the rest of this chapter we
shall draw most examples from the highly successful AVHRR programme.
198 Sea-surface temperature from infrared scanning radiometers [Ch. 7
G
O
T3ci> &
eg
Wj
U-)
cd S _2G X
£ a. G
o s
-G V3 ^ 00 4-* -4-i
O ’-H CN G cn
<U tn ^ G .G
wc^ OT3
cd o^ ^ C 2 o CN
s (U
C/5 r- £ i ON
t t
^
O cd
>>
X*
I •r*
^>7 cd
o ^
d e e
(N
tT)^
a> ^
o
l-l s s
a o 5
D.
oc .S
o
s x od
s
cd QJ
cd
G
O c/3
C/3
c/3
s
a
X o PC
’O .2 ^ X
C3 T3 IZ)
(N X g >
it) G
00
3 cd
W-) X X
m 2 c <
X) X G cd 5/3
S On c/5 •o
X
.2 l-l X
o z
U
N
CN cd .2 o
z U
00
^ o
oJ &o
(U G* V3
cd
u 6D
s
o X »o
s .s
cd
o
TO r4 CN *cd >
IZ)
G OD
(D in > ?
cd cd t<_,
T3 o X t T3 X CO ^ O
G r-- • in o
cd • CN
X
H
d o
Si
s cd ^
Q ca
•4-*
00 X G X
'G (N CO O o
G 00 CN o
in c/3 G
cd cd
T3 o X t X 00 •G
cd G On
cn
(D l-l
CA cd
X cn
d
CO
G
o
Xcd
<N
u ^4 X
a OX) 00 X
o .G -G
o
cd
a G Cd cd (U
G
cd
G >-i u O
G
cd
CO
G CO
E c
s >.
4-> <H
o
/-s in
• <N
O
00
(N
O O
O O
I ^
’
5 ^ cd o ^ 1—( 4-* <n X
>.
620
g.2 G G Cd
cd cd 60 t t X X
o .52 S.o
o o E g
(O.f
10.5
X* "cd ^
00
(D o o S3 S
0) lo X
i X
S 2
X 2 c0)
73 JrJ
a>
0^ cd
G C
G/
u u .SP-'S
e/3 a>
X X (U
Q
>
o
t-’
pS G
o
z
Cd E *•4 X
H <D 4. X
o g x"
G C/3 CO 4-»
c/3
G cd ^G a> •G
X M x"
O G G T3 c
4-'
i-H cd 13 G
G cd
a X O H G X E
0) OX) c/3 < G
G 1 O E
cd ‘(U 0) W IH cd o
C/5 ffi C/5 1 Z a CO U
Sec.7.2] The physics of infrared radiation 199
Cl
(7.1)
\Mexp(C2/Xr) - 1]
is the spectral exitance (sometimes called emittance), that is, the radiant flux
density of radiation per unit bandwidth centred at leaving unit area of surface,
irrespective of direction (the equivalent of irradiance but for energy leaving
rather than falling on a surface).
X is the wavelength in metres.
Cl and C are constants with the values;
2
Cl = 3.74 X 10-^®Wm^
C = 1.44 X 10-2 nr degK
2
M^oT^ (7.3)
r = Cj (7.4)
where C = 2897 fxm K“^
3
In the context of remote sensing of the sea and land surface, we note that
the peak of the emission spectrum for a perfect emitter at 300 K is around
10 /im, and sensors operating near this wavelength will be most suited for
measuring the temperature of the surface. From (7.1) and Fig.7.1 it is apparent
that even at the wavelength of the sea-surface emission peak, the direct solar
radiation will still be substantially greater than the earth radiation. At night this
is not a problem because there is no solar radiation, but in the daytime reflected
solar radiation must be allowed for. Experiments have shown that a surface
will absorb a proportion of incident IR radiation equal to its emissivity, e. For
the sea surface, e is about 0.98, and varies very little with wavelength, tempera¬
ture, or surface roughness (in the wavelength range of interest between and3
As discussed in section 6.2.2 a satellite sensor measures the radiance, that is,
the radiant flux per unit solid angle per unit projected area per unit bandwidth.
For infrared emission, the surface can be considered to be Lambertian with
uniform radiance in all directions, so that:
Lx = Mjn. (7.5)
WAVELENGTH^m
Fig. 7.4 — Relative spectral response for (a) the 3.1 channel, and (b) the
11 pm channel, of the AVHRR on TIROS N. (From Deschamps & Phulpin
(1980))
204 Sea-surface temperature from infrared scanning radiometers [Ch. 7
sensor bands centred at 3.7 /rm, 11 /Jin, and 12 nm. On balance, the higher wave¬
length window is the most useful because it corresponds to the peak of the
thermal emission spectrum, whilst the 3.7 /rm window is able to provide a useful
correction channel, particularly for night-time imagery not troubled by solar
reflection.
a given level of noise. The cooler the sensor, the smaller the Brownian thermal
noise. With a sensor cooled to 105 K (requiring a carefully designed heatsink
radiating out to cold space), the AVHRR achieves sensor noise levels of around
0.05 degK in the very short integration time of less than 25 fxs permitted during
the scanning process.
In the analogue/digital converter, the least count is an increment in energy
rather than temperature, and therefore the digitisation resolution in equivalent
temperatures varies both with the temperature of the sea surface being viewed,
and with the wavelength channel. For typical sea temperatures between 270
and 300 K, the resolution is 0.13 — 0.10 degK for channels 4 and 5, and
0.17—0.05 degK for channel 3. The sum of the sensor noise and the digitising
error is specified for channels 4 and 5 of the AVHRR at around 0.12 degK. In
practice, Bernstein (1982a) has shown that a NEAT of around 0.05—0.07 degK
appears to be achieved. Similar noise levels are specified for channel 3.
The black-body cavity for calibration of the AVHRR is heated to around
290 K, as measured by four platinum-resistance thermometers. Errors arise from
the assumed emissivityof the black-body, the existence of temperature gradients
across it (a problem for the AVHRR), and the accuracy and resolution of the
thermometers. The total temperature uncertainty in the radiometer estimation
of sea-surface temperature, arising from errors in the in-flight calibration using
the black-body, is quoted by the makers as 0.31 degK.
The significance of these errors depends on the application of the data. The
study of spatial temperature structure in the sea requires accurate measurement
of temperature difference, whilst absolute calibration is less important. NEAT
is therefore the limiting parameter in this case. On the other hand, a measure
of average sea-surface temperature over an area of many pixels (for example, for
a climate database) can reduce NEAT over many pixels, but in this case the
requirement of comparability with the data from other orbits, other satellite
sensors, and other methods of measuring the sea temperature, requires the
absolute calibration to be accurate. Averaging over several scan lines (and hence
several black-body calibrations) will reduce some of the calibration error, but
any bias in the calibration cannot be removed without recourse to sea-truth
comparisons.
these gives T^, the black-body reference temperature. For a given image, X^, X^,
and are averaged over many scan lines to reduce error before performing the
following calibration to estimate the apparent temperature of each pixel.
A separate calibration is performed for each waveband. Within each wave¬
band, the actual sensor response is governed by a normalised wavelength response
curve. This is measured before the flight and is available from the satellite manu¬
facturers (NOAA/NESS 1978, 1979) in the form of n discrete values of
for / = 1 to n. The X; are separated by AX so that X] and X„ encompass the total
response of the sensor in a given waveband, and
.2 ^(X,-)AX = 1
I = 1
(see Fig. 7.4). The emitted radiance from the black-body reference, as detected
by the sensor at waveband X, is then calculated as
n Ci^\(Xr)AX
1
(7.6)
’ ( =
X^[exp[C2/X/rr] -IJtt'
Theoretically, the radiance of the deep-space view should be similarly calculated
from the assumption that the temperature of deep space is 4 K. However, to
bypass a small nonlinearity in the digitising process, a correction is incorporated
into the deep-space radiance which is supplied along with the other calibration
constants as
With two radiance values, and two corresponding digital numbers in the
data stream, a linear, two-point, calibration can be constructed, with gradient
2s,A. ^T,X
(7.7)
and intercept
A — x X. (7.8)
Then the radiance measured in each pixel within waveband X can be obtained
from
L'x= GxiDN)x+Ix (7.9)
where the significance of L'x is that it is the radiance as recorded by the satellite
in the given spectral band defined by $x. Thus if the spectral radiance actually
reaching the sensor is L(X),
Each value of L'x for each pixel can be interpreted in terms of a brightness
temperature T^, the temperature of the perfect emitter which would produce
that radiance if there were no intervening atmosphere. This is related to by
Sec. 7.3] The near-surface thermal structure in the ocean 207
,, « Ci0;,^(X,)AX
L\ — 2- (7.11)
which is not readily inverted to obtain from L'x- Instead, (7.11) is used to
construct a curve of the form
deeper part of what was the winter mixed layer by a strong temperature gradient
called the seasonal thermocline. The summer mixed layer may be only about
20 m deep after a calm spell of sunny weather. The actual depth of the seasonal
thermocline and the temperature of the upper mixed layer are dependent on
the recent weather history of a location, and vary considerably in time and
space. Indeed, the temperature profile may have a step-like structure down to
the depth of the main thermocline, the deeper steps corresponding to strong
wind-mixing events a few weeks previously on which have developed a new
seasonal thermocline, broken up by subsequent but weaker storms. Under these
circumstances, without temperature profiles available from ships or buoys it is
not possible to deduce from a satellite-measured sea-surface temperature map
how deep the surface layer extends. In winter it can be expected to reach to
the main thermocline, but in summer the upper mixed layer depth cannot be
specified.
These structures found between 20 and 1000 m limit the depth of sea
where temperature is the same as that of the near-surface water, and therefore
limit the general usefulness of satellite IR sensors. The development of a diurnal
thermocline is even more troublesome for the interpretation of the sea-surface
temperature within the top few centimetres as representative of temperature
in the upper mixed layer as a whole. The process is similar in principle to the
development of the seasonal thermocline. As the day progresses past noon, solar
heating causes the surface temperature to increase. Because the short-wave
solar radiation is progressively reduced by absorption as it penetrates the water
column, the water temperature rises most near the surface, and a strong gradient
of temperature may occur in the top metre or so of the water column, raising the
surface temperature locally above the upper mixed layer temperature (Fig. 7.5).
If there is significant wind-mixing, the effect is very weak because all the solar
energy is mixed well down throughout the upper mixed layer or contributes
to the slow growth of the seasonal thermocline. On a calm day, however, the
surface may be up to a degree K or more warmer than the mixed layer, whose
upper limit is now not found at the surface but at a depth of a metre or so.
Under these conditions, the satellite measuring the temperature of the top of
the diurnal thermocline cannot detect the true mixed-layer temperature. Yet it
is this rather than the temperature of the top few centimetres which is important
oceanographically in studies of the heat flux of the upper ocean. Moreover, it
is quite possible that the diurnal thermocline can mask a horizontal structure
present in the mixed-layer temperature. It is very hard to predict the diurnal
thermocline strength, and it can be very patchy over length scales as short as
tens to hundreds of metres.
The problems are greatest on sunny afternoons on relatively calm days.
Radiometric measurements of sea-surface temperature under these conditions
may differ significantly, by up to a degree, from the‘bulk’sea-surface tempera¬
ture representative of the top one or two metres, as measured from a ship
210 Sea-surface temperature from infrared scarming radiometers [Ch. 7
15.0 16.0
(that is, at about 10 cm depth) is found in the top millimetre. At night the
solar radiation is zero, and the net upward heat transfer through the surface is
given by
2n = Qir + Gg + 2h + Ql
where
Gir = infrared radiation emitted by the sea surface (positive)
Gg ~ downwelling infrared radiation from the sky, cloud, and atmosphere
(negative)
Gh = sensible heat flux
Ql — latent heat flux (normally positive).
Under most circumstance is positive, and the temperature deviation
Sr is negative (that is, the surface is cooler than the water a few millimetres
below). During the day the solar radiation is largely at or near the visible wave¬
lengths. This penetrates below the sea surface as electromagnetic energy and,
apart from a small amount of absorption in the skin layer, contributes little
negative heat flux to offset 67 is therefore normally negative during both
day and night, and observations suggest a typical value is between —0.1 degK
and —0.5 degK.
Under special conditions where Qn is negative, 5T has been observed to be
positive, but this is exceptional. The physical processes controlling the surface
skin temperature are illustrated schematically in Fig.7.6.
The subject has attracted a small but steady trickle of interest from physical
oceanographers, and has been reviewed in some detail by Katsaros (1980), and
previously by Saunders (1973). It is clearly a problem for the measurement of
sea-surface temperature by infrared radiometry from above, but has hitherto
been considered too small to worry about in comparison with other errors
encountered in IR radiometry. However, a recent review of the skin-temperature
deviation in relation to present and projected accuracies of sea-surface tempera¬
ture retrieval from satellite sensors (Robinson, Wells, & Charnock (1984)) points
out that developments in remote-sensing technology have caught up with the
effect. It now constitutes a potentially significant source of error in SST
algorithms.
Table 7.4 lists some of the observations made of the skin effect. It is very
difficult to measure accurately, and some of the laboratory or sheltered experi¬
mental conditions which give rise to large temperature deviations may not be
typical of open-sea conditions. As shown in Fig. 7.6, the temperature deviation
6r is defined as the difference between the temperature of the sea at the sea —air
interface, and the so-called ‘bucket temperature’ or bulk SST which effectively
is the temperature of water taken from the top metre of the sea and mixed well.
In a well-stirred sea, there is unlikely to be a significant temperature gradient
between about 1 cm and 1 m depth, and ST will be concentrated in the top
millimetre. If a diurnal thermocline develops there is likely to be a strong
212 Sea-surface temperature from infrared scanning radiometers [Ch. 7
NichoUs (1979) --0.5 ± 0.3°C at sea low to moderate wind and waves
Paulson & Parker (1972) -1.14 to laboratory no waves
-1.81°C
Ewing & McAlister (1962) -0.6°C at sea low winds, night-time
HiU(1972) (i) ~-1.0°C laboratory low wind, no surface film
(ii) -2.0°C laboratory low wind, surface film present
(iii) ~-0.1°C laboratory higher wind conditions
o
no wind
thermal gradient within the first metre of water, the temperature decreasing
with depth, and the temperature changes across this thermocline must be dis¬
tinguished from the skin effect 8T when observations of 57’ are being made, and
if corrections for it are eventually to be applied.
Until now, no attempt has been made to predict the size of the skin-
temperature deviation in order to convert radiometric observations of SST into
the equivalent bulk SST. However, with 8T values of up to 1 degK observed
in the open sea, there is clearly a need to develop a better understanding of
the effect. Theoretical analysis (Saunders (1967b)) suggests that a possible
approach to prediction is to relate the effect to 2n> the surface heat flux,
and U* the friction velocity of the wind at the surface. Assuming that the
convective processes occurring just below the skin layer (Fig. 7.6) are due to
wind-driven turbulence rather than convective instability (which occurs in flat
calm conditions) it is suggested that
S-p — (7.13)
kU*
where k is the molecular conductivity of the sea water, v is the kinematic
viscosity, and X* is a constant to be determined. Estimates of X*. by various
workers are listed in Table 7.5. A fair scatter of values is noted, so that even if
further research confirms the basic form of (7.13) as a prediction equation, it is
clear that other environmental parameters contribute to the skin effect and
must be incorporated in the value of X* These include the size and wavelength
of surface waves, the presence of slicks and the effect of surface tension, and
the existence of whitecaps or other evidence of the complete destruction of the
thermal skin by the physical breaking of the surface. (It is interesting to note
that the thermal skin deviation seems to reappear witin 10 seconds of the broken
surface reforming.)
(7.15)
(7 17)
216 Sea-surface temperature from infrared scanning radiometers [Ch. 7
, 1
= (l-e)^xCos0sr(X,0s;O,Ps) - r(X,0;Ps,O) (7.18)
TT
where E\A% the solar irradiance at the top of the atmosphere on a plane normal
to the solar direction which is at zenith angle Sg. The downwelling solar irradiance
at the sea is therefore
£'xCos0sr(X,0s; O.Ps),
^(X, 0; p,Ps)
^ra,\(^) = (1 - e) t(X,0; Ps,0) ^a(p) dp.
dp
(7.19)
Here, rather than calculate the diffuse reflectance of all the downwelling atmos¬
pheric radiation, it has been assumed that the plane reflection of the emission
along a column of atmosphere with zenith angle d will give a similar and more
readily calculated result.
In principle, knowing the atmospheric temperature profile and the distribu¬
tion of r with pressure, which depends on a given distribution of water vapour,
carbon dioxide, and ozone in the atmosphere, equations (7.14)-(7.19) can be
solved to yield Such a strategy is unpractical because it is not possible to
know /^(p) and t accurately enough, because of their temporal and spatial
variability. Ideally, it would be convenient if they could be measured with a
multi-channel IR atmospheric sounder such as the HIRS, but such an instru¬
ment does not necessarily yield a unique temperature profile, and high-accuracy
atmospheric sounding requires a prior knowledge of sea-surface temperature,
the very parameter we are trying to determine after atmospheric correction!
Whilst they are useful for meteorological purposes, such sounders do not yet
promise to provide an accurate means of direct atmospheric correction for SST
retrieval algorithms.
again the difference between the two brightness temperatures can yield a
measure of the atmospheric effect.
Where equations (7.14) to (7.19) are useful is in the simulation of typical
satellite-observed radiances given a variety of different, but typical, atmospheric
profiles. By modelling the different responses of the separate sensor channels
to different sea-surface temperatures, a simulated dataset of brightness tempera¬
tures corresponding to true sea-surface temperatures can be generated, from
which empirical algorithms can be extracted. This enables the accuracy of SST
retrieval from a proposed sensor to be statistically explored before the satellite is
launched or even built.
For a satellite in orbit, the most satisfactory SST atmospheric-correction
algorithm is derived from a real dataset of satellite and synoptic sea measure¬
ments. The important consideration is that the points chosen for calibration
purposes should represent a wide range of latitudes, seasons, and weather
conditions, so that all typical atmospheric conditions are allowed to influence
the dataset. If only one type of atmospheric profile were encountered, the
calibration algorithm would probably be more accurate for that particular
profile, but of less value under other conditions. Since a universal algorithm is
sought for all locations, sensors, and atmospheric conditions, in order to provide
an operational product for wide dissemination, the operational algorithms
now available are based on as wide a dataset as possible. The dataset should also
consistently represent a range of different viewing angles, if the algorithm is to
be applicable to all parts of a scanned image, at least within certain limiting
angles. Beyond about 9 = 50° the accuracy of atmospheric correction becomes
much poorer.
Table 7.6 lists the algorithms which were introduced by NOAA/NESS for
the correction of NOAA 7 AVHRR data, effective from 14 September 1982
(McClain (1982)). Since radiance values can be uniquely interpreted as bright¬
ness temperatures, there is nothing to be gained from correcting the radiances,
which conceptually might seem to be the right approach. Instead, the correction
algorithm employs the brightness temperatures for each channel, which can be
calculated from the satellite data using the method described in section7.2.4,
and delivers an atmospherically-corrected estimate of SST. In Table7.6, r^fcis
the corrected SST in degrees C, and T^.j, Tn, and are the satellite observed
brightness temperatures in degrees K from the three different channels. The
algorithms were based on a dataset of nearly one-hundred SST observations from
drifting buoys. These continuously monitor the temperature within the top
metre of the ocean, and when a cloud-free satellite overpass occurs the relevant
sea temperature can be selected from the buoy record.
Because the atmospheric contributions are different at night from the day¬
time (since there is no reflected solar radiation at night), it is necessary to use
different algorithms for night and day. The former tend to be more accurate,
since the day-time corrections should strictly include a dependence on the sun
218 Sea-surface temperature from infrared scaiming radiometers [Ch. 7
zenith angle. Table 7.6 also lists the accuracies achieved with the night-time
algorithms, as tested against a set of 94 drifting-buoy temperatures which were
independent of the dataset from which the algorithms were constructed.
DAY-TIME
Split-window rsfc(4/5) = 1.0351 Tn + 3.046 (Tn - 7^12) - 283.93
NIGHT-TIME
Split-window rsfc(4/5) = 1.0527 Tn + 2.6272 (Th-T^) - 288.23
Triple-window (3/4/5) = 1.0239 Tn + 0.9936 - T^) - 21%Ae
Dual-window rsfc(3/4) = 1.0063 Tn + 1.4544 (73.7- Tu) - 272.47
Table 7.6 (b) Accuracy test of above equations, using an independent set of
N drifting-buoy observations
N 94 94 94
Bias 0.06 deg C 0.14 degC 0.16 deg (
Scatter 0.60 0.64 0.72
Root-mean-square 0.61 0.66 0.73
deviation
or ^2 ^2 ^3.7 “f ^2 ^11 ■“ 0
must be satisfied within certain limits.
In the end it will never be possible completely to eliminate cloud contamina¬
tion by very small clouds, but any resulting errors will be contained within
acceptable bounds.
A different approach can be used when high accuracy of SST measurement
is being achieved by averaging the satellite observations over areas of, say,
50 X 50 pbcels. Averaging can be performed when it is believed that there are no
significant horizontal sea-temperature gradients within the area. It leads to high
accuracy because sensor noise errors are greatly reduced by averaging. However,
a few pixels contaminated by cloud can seriously distort the average. One way
of avoiding this problem is to examine the histograms of the radiance counts,
the brightness temperature, or the atmospherically-corrected temperature distri¬
bution within the 2500 pixels. Fig. 7.7(a) illustrates a histogram in cloud free
conditions. There is a Gaussian spike centered at what is taken to be the true
radiance count appropriate to the average SST and atmospheric transmission
conditions throughout the 50 X 50 km area. Fig.7.7(b) shows the case in which
cloud is partially obscuring some pixels and completely obscuring a few. The
Gaussian spike now has a long cool tail due to the cloud-contaminated pixels,
and the true SST value can be obtained not by averaging the whole histogram,
but by fitting a Gaussian spike of 2500 values to the warm end and determin¬
ing where the centre is located. With appropriate computer algorithms this can
be done automatically, although care must be taken to exclude cases of 100%
cloud contamination. It is claimed that this approach can cope with 80% cloud
cover, provided that 50 X 50 pixels are used. For smaller sample areas, the
technique becomes progressively less reliable.
ARHRR/2 ATSR/M
(observed) (predicted)
degC degC
buoy and ship observations. Following the eruption of the Mexican El Chichon
volcano in 1982, large quantities of dust were injected into the stratosphere in
equatorial latitudes, rendering the atmospheric-correction algorithm invalid and
enforcing a temporary discontinuance of SST information at low latitudes.
Now that the use of satellites to measure SST is becoming regarded as
established operational practice, it is worthwhile to consider exactly what
temperature is required by oceanographers and meteorologists. The true
sea-surface temperature is the skin temperature which is measured by the radio¬
meter, but hitherto oceanographers have used either the bucket temperature, or
even the engine-intake temperature to provide an SST database. In practice
the requested SST parameter will vary in definition according to the application.
Thus in studies of the mixed-layer and the modelling of its depth, the bulk SST
at about 1 m depth is probably most valuable, since it tends to eliminate the
very shallow diurnal thermocline which has little effect on the total heat content
of the mixed-layer. On the other hand air —sea interaction is concerned with the
sea-air temperature difference which drives the heat flow by sensible and latent
heat transfer between ocean and atmosphere. Thus the actual skin tempera¬
ture might be considered appropriate in this context. However, recalling the
discussion of section 7.3.3, it appears that the thermal skin is itself a response
to the heat flux, and that the temperature-difference driving the heat flow
is that between the sea temperature a few centimetres below the surface, and
the atmospheric temperature just above the surface. 2ir and will be slightly
modified by the temperature deviation across the surface skin, but it seems
conceptually correct to use the temperature below the skin as the most
appropriate definition of SST in the context of air-sea interaction studies. In
addition it should be remembered that the empirical formulae which have been
developed for parameterising sea —air fluxes are based on the bulk sea and
air properties and ignore the skin.
It can therefore be concluded that the provision of an SST dataset based on
the surface-skin temperature measured from satellites is less useful for climate
modelling and air —sea interaction studies than if a correction has been made
to the radiometric temperature to allow for the presence of a skin-temperature
deviation. As we have seen, the MCSST is such a database. Although it does not
set out to allow for the skin effect, it does so implicity by matching satellite
observations to buoy measurements. It is also important for the long-term
continuity of climate datasets that the satellite-based SST data are essentially
related to the same parameter as the previous data from ships, particularly if the
satellite records come to be thought of as superseding and replacing ship records.
It is important for climatologists of future generations that along with the
SST data is archived the details of the calibration, so that it is possible to deter¬
mine whether shifts of mean SST or changes in the variability of SST which
may emerge from the data for the decade of the 1980 s are a real occurrence
or merely an artefact of the changing methods of collection of the data. The
224 Sea-surface temperature from infrared scanning radiometers [Ch. 7
7.5.1 Climatology
Studies of the earth’s climate require an understanding of the complex
interactions between the ocean and the atmosphere. Fundamental to this is
the temperature of the ocean surface through which energy is interchanged.
Small changes in the sea-surface temperatures on the one hand may represent a
significant change in heat energy storage within the ocean, and on the other
hand may have significant impact on the atmospheric flow and consequent
weather patterns.
Present SST databases used in climatology have a horizontal resolution
from 1° X 1° to 5° X 5° and are averaged over periods of 5 to 30 days. The major
difficulty with the present database is that it is dependent on the monitoring of
the ocean by merchant ships, which produce measurements with rms errors of
0.5 deg K and with a spatial coverage limited by the major trade routes. Satellites
therefore offer a good opportunity to produce measurements comparable to
and eventually more accurate than merchant ships but with less temporal and
spatial aliasing. However, it is important that the satellite database does not
have a systematic bias in its errors as a result of atmospheric contamination, or
the surface skin effect, nor episodic deviations due to anomalous atmospheric
contamination as occurred following the El Chichon volcanic eruption referred
to above.
Whilst it may not prove too difficult in the near future to improve on the
accuracy of present ship-of-opportunity sampling, it is worth considering what
the ideal requirements are in different applications of a climatological database
of SST. The following three application areas are examples of the use to which
a climatological SST database may be put.
Sec.7.5] The potential use of SST data from satellites 225
influence on the location of the early development of tropical cyclones and may
affect the subsequent track of the cyclone. Modelling studies in mid-latitudes
have shown that warm anomalies may have a significant influence on the develop¬
ment of depressions in the N.E. Pacific Ocean and in the Tasman Sea. For these
phenomena an absolute SST accuracy of 0.5 degK should be sufficient, with
spatial information on scales of 50 to 100 km.
Information regarding the influence of SST on subsynoptic scales of motion
is small, though it is known that the distribution of sea fog is very sensitive to
SST, and mesoscale convective systems may be influenced by SST. To study
this further, SST would be required at a very high spatial resolution (1 km), with
accuracy of better than 1 degK.
In all these studies emphasis will be on the spatial variations of SST over the
timescale of the process or feature, rather than absolute accuracy. For instance,
continental-shelf tidal fronts have temperature variations of a few tenths of a
degree over length scales of 10 km. Mesoscale eddies, however, will often have
temperature anomalies of ±2 degK on space scales of 50 to 200 km. Coastal up-
welling regions will have temperature variations of up to 5 degK over distances
of 50 km. Large-scale dynamical motions in extratropical latitudes may be more
difficult to detect since the vertical velocity is relatively small and only small
displacements of the main thermocline occur. However, in equatorial latitudes,
because of the increase in speed of the long waves and the shallowness of the
thermocline, a strong influence on the SST may result.
H OD
c/3 O
00
•o lo CO
<N O JD lO IT)
o
cO o o o d O o
t-< W)
3
o
o
cO
CO
u Cd)
0)
E
•3
T3
a> tn <N fS ^
.id o
I I dod
3
Summary of SST requirements for different application areas
*'/
o a> >. B >»
C ‘S U<
cO
bi
CO
bH
CO
b^
CO
.2 c M t-H JS
CO CO *3
i>i.
^
CO
a- s.
^ 03 03 03 03
o'^
DO —
>. >>§ ^3
5?
I
ft>
t-i
>» >> >, >» VO di
O in <N •n
I S I
1 B
.S CO
cn
A A O I
2 3
— O
O
E E E o E E E E E E
.2-B S o phd
^.3 3 o o o o in o o rH
c/5D. O
JO
0^
O'
aj
bN
o
cs
o
in
o
<N B
1 o
(N
o •—(
a>
CO S _
M CM 52 CO CO
a> 13 lo
CO
fO
2 c c
> x> x> X) ,3 •“ .2 .2
o .2
o 3
cO
*Sb ^
43 O
S '5b '5b
2 u <u
CO O a a o 03
a> O o OC OS
Table 7.8
O O
O
CO
Sri 3
<l>
a> E ^ 03
X «3 oj
5 c/3
C/3 &<
bt
Q>
Vi C^ ^ 3
•3 -H O &
^
CO
CO
*3 E 3
“ c :3 ii
C
O
o .^5 O
X3 O
“ .2
M w) c o .a
S
8 .2 3
"cO
o o *3 o ^ cO -
o O
;3 ^ O «> 2 2 8s
a .S 3 1o
a lu bo. >
cr CO 3
cj 3 "5
O 3
5
—
3
< I •§ S*s
^
_ 03. u
o 3I w
'^■2 ^ O
S fc -S
5 2 <4-.
CJ
■S § a
H 3 H co 03 S, =3
00 o o 2' o-o
CO 03
oo CO ^ o Q Q
230 Sea-surface temperature from infrared scanning radiometers [Ch. 7
eo^N^ MARCH 1982
Fig.7.8 — Global maps of (a) SST and (b) SST-anomaly, for March 1982, derived from the NOAA-NESDIS satellite data archives. The contours
in (a) have been drawn at 2 deg C intervals for clarity at this small reproduction size, but are supplied at 1 deg C intervals on the original. (Data
supplied by R. Legeckis.)
232 Sea-surface temperature from infrared scanning radiometers [Ch. 7
Fig. 7.9 - SST archived data available for January 1977 from (a) ship and buoy
measurements and (b) satellite infrared radiometer observations (from Miyakoda
& Rosati (1982), published by the American Geophysical Union)
< -H S
< 5 ^
^ c o
■3
c .ti >»
_g o -O
^ J=
x> o
•o C
(U
•o q> 0/5
o
o o
• • ^ on
«j TS 2^
^ TJ *5
^ “ 2:
-go
• I C9
R. Legeckis of NOAA/NESDIS.)
2 SI
o cd •♦■^
« ^ O
CQ
SC f3 ^
!>
«
8.3£
u-3-
O
C/3 (l>
a> is S
60 3 n>
go's
.is 60
^■§1
I c o
C cQ
S J> J2
•c
^ E -3
'S
- “ E
S^ “
So; 5
r- 3
^00
' ££
03"^
»-i >—> o •w
r^ »o
c ?
pH O U
Sec.7.6] Examples of the application of satellite SST data 235
warmer offshore waters. The Gulf Stream front meanders, and to the north of
it is a clearly defined warm-core eddy in the cooler water. There is also evidence
of a cold-core eddy which has broken away and moved into the warmer
Mid-Atlantic water.
Using earlier VHRR images of a similar cold-core eddy, Spence & Legeckis
(1981) were able to extract dynamical information without needing to correct
the data atmospherically. They analysed a sequence of images, two between 9
and 12 March 1977 and five between 8 and 16 April, which were sufficiently
cloud-free for a cold-core eddy to be identified and followed. After digitisation
of the analogue data, and geometrical rectification, the eddy could be identified
on computer printout by tracing the locus of the steepest temperature gradient.
From this, the orientation of the eddy, which was elliptical, could be seen to
rotate in a counterclockwise direction at a rate of about 20° per day (Fig. 7.11),
equivalent to 5% of the local Corolis frequency f. This could be compared with
analytical theory of the dynamics of baroclinic eddies. Using satellite data, this
is a relatively simple measurement to make, and is achieved without the need for
careful temperature calibration, whereas to make the same synoptic observations
from ship or buoy sampling in order to explore the dynamical theory of eddies
would be much more time-consuming. Of course the acquisition of ship data
to provide vertical soundings through the eddy would enable the analysis to
be taken further than if no ship data were obtained at all. The conclusion to be
drawn is that when ship and satellite data can be combined to complement
each other, with the satellite data supplying horizontal synoptic spatial informa¬
tion, there is considerable scope for improving our understanding of dynamical
processes in the ocean. On a cautionary note, it should be pointed out that
eddies may exist below the sea surface without having a thermal surface signa¬
ture, and the absence of eddy structures on an IR sea-surface image need not
imply that there are no dynamical features present in the water below the
surface layer.
•37°N
36°N
Fig. 7.11— The location of the steepest temperature gradient of a cold-core ring
in the Atlantic Ocean similar to that illustrated in Fig. 7.10. This was determined
from infrared images observed by the VHRR on NOAA-5. A sequence of posi¬
tions between 8 and 16 April is shown, revealing an apparent counterclockwise
rotation of the ellipse shape at a rate of about 20° per day. (From Spence &
Legeckis (1981), copyrighted by the American Geophysical Union.)
Fig.7.12 - Meteosat infrared channel 2 image at 11.55 UT on 17 February 1978.
This image has been enhanced to illustrate the range of sea-surface temperatures
off the coast of northwest Africa. UpweEing of cooler subsurface water along the
coast is marked by a hght grey tone at A. (Image supplied by ESA.)
Sec.7.6] Examples of the application of satellite SST data 237
7.6.3 Upwelling
Fig. 7.12 is a Meteosat IR image from February 1978 showing the west coast of
Africa off Mauritania and Senegal. It reveals very clearly the upwelling that
occurs along this coastline as the trade winds blow to produce offshore surface
drift. It is most marked immediately south of Cape Verde and northwards
towards C. Blanc. Domain et al. (1980) used a sequence of Meteosat data during
the spring and summer of 1980 to chart the progress of the thermal evolution.
Warmer surface water gradually moves northward, and the upwelling area
recedes north of C. Blanc by July or August. This area has been well studied
by numerous ship experiments, and the satellite observations are in accord with
the expected patterns of sea-surface temperature. The satellite, however, is able
to identify more specifically the location of the upwelling zone and the timing
of its movement, and this has proved to be of some benefit to local fishing
operations.
Being close to the equator this area lends itself to study by Meteosat which
views Cape Verde with a zenith angle of no more than about 15°, achieving
nearly maximum resolution. With data-sampling every half hour, a technique
of data-processing was adopted which dealt with all the data from one day and
selected for each pixel the most persistent value. In this way, transient clouds
passing over the region could be eliminated to leave the permanent land and
sea temperature features — a useful benefit of the frequent sampling-rate
of Meteosat which offsets to some extent the disadvantage of the large pixel
size (4 km X 5 km). Meteosat data is not so readily calibrated or atmos¬
pherically corrected, and reliance has to be placed on in situ measurements
to calibrate individual scenes, but this does not prevent the upwelling patterns
being determined.
Given a year’s images of the area between January and December 1979
from the AVHRR archives at the Dundee Satellite Receiving Station, between 30
and 40 images were obtained with a part of the front visible through gaps in the
clouds, and about half this number had essentially clear views of the front which
enabled its position to be accurately located. Several of these images came in
pairs an hour apart on the same day, given the overlap between successive swaths
of the TIROS/NOAA satellites. There were also three or four periods when two
or three images within two or three days were cloud-free. Thus it was possible
to study the long-period seasonal movement of the front (which appeared to
be seasonally stationary) and the short-period meanders illustrated in Fig. 7.14
which shows the line of the front for several different images during the year.
In this study it was possible to match the evidence of the satellite data to the
archives of ship data. What the satellite imagery is able to add to previous
understanding is a synoptic view of the front at given times, and its variability.
Thus, whilst averaged ship data might indicate that the front weakens into a
Fig. 7.14 - Composite of lines drawn to represent the steepest thermal gradients
marking the edge of the Iceland—Faroes front on about 40 clear-weather IR
images recorded between November 1978 and July 1980. The front position is
subjectively determined and depends to some extent on the photographic contrast
stretch employed on images such as Fig. 7.13. On most of the images only a
portion of the front was visible through the clouds.
240 Sea-surface temperature from infrared scanning radiometers [Ch. 7
more gradual gradient towards the east, the satellite data show that in fact the
front remains relatively strong, but meanders and moves its location more at
the eastern end than the western. A hydrographic station at the mean frontal
position will sometimes lie to the north and sometimes to the south of it, so
that if hydrographic data are time-averaged it will tend to smooth out the front.
Again we see the synoptic spatial view of the satellite being the particular
contribution which is able to bring new information into what was already quite
a well-understood oceanographic feature.
Careful analysis of the images has also revealed that the warm-water
intrusion seen on Fig. 7.13 about 180 km east of the Icelandic coast, where it
extends northwards into the cold area, is a semi-permanent feature. Although
it is masked by turbulent meanders which appear to travel along the front
and break off into eddies at this point, there is usually a tongue of warm water
reaching 50 to 100 km northwards at this location, which corresponds closely
with a deep notch in the Iceland—Faroe ridge over which the cold water
is known to overflow southwards. Further research is needed to examine the
hypothesis that the existence of the warm northward intrusion may be related
to the occurrence of subsurface southward overflow. If this was the case the
disappearance from an image of a clear warm intrusion might signify that the
intermittent overflow had temporarily ceased. Such an idea is speculative, but
illustrates the type of research possibilities opened up by satellite IR data.
It would certainly be interesting if it turned out that a sensor limited very
definitely to measuring a surface parameter was able to indicate the occurrence
of a process 300 m deep.
The Iceland—Faroes front is one example of many such features visible on
IR imagery from around the world as reviewed by Legeckis (1979).
HiUON
<
h
(/)
«
Ui
k
Examples of the application of satellite SST data
Fig. 7.15 — Mean distribution of the more persistent surface thermal features in the Mediterranean Sea observed on satellite infrared image datasets
241
for the period 1977—1979, showing the inferred circulation. (From Gallagher etal. (1981).)
Fig. 7.16 — NOAA-6 AVHRR channel 4 image of the British Isles recorded at
1900 GMT on 16 May 1980, showing clearly the presence of shelf-sea tidal-
mixing fronts between the weE-mixed water column (cooler, lighter shading) and
stratified water (warm surface, darker shading). (Image supplied by the Dundee
University Satellite Receiving Station.)
Sec.7.6] Examples of the application of satellite SST data 243
the tidal-mixing front found in shallow tidal seas (Simpson & Pingree (1978)).
Fig. 7.16 shows the existence of these fronts around the UK, where they are
found between about May and September. They are prominent in the English
Channel south-west of Cornwall, in the Celtic Sea south-east of Waterford, and
stretching across to South Wales in the Irish Sea south-west of the Isle of Man,
and running north-north-east from the coast of N. Ireland.
The way in which these fronts occur at the boundary between stratified
water (having a warmer surface temperature) and well-mixed water which is
stirred by tidally driven turbulence, has been described in section 3.3.3. Fig. 3.7
illustrates the temperature profile schematically and shows why the front has
an SST signature. The surface is cooler in the weU-mixed region, essentially
because the solar heating has to heat the whole water column and not just down
Fig. 7.17 - Map of contours of the parameter log,,,(/z/u’) in the Irish Sea,
English Channel, and Celtic Sea, based on tidal stream observations. The contour,
log,o(/t/«®) = 1.85, is found in approximately the same position as the shelf sea
fronts in Fig. 7.16. (From Simpson (1981).)
244 Sea-surface temperature from infrared scanning radiometers [Ch.7
to the thermocline as on the stratified side. The frontal region has a certain
amount of upwelling associated with it, and is usually characterised by enhanced
productivity. Its location is therefore of much interest to biological oceano¬
graphers who have used IR images to guide the location of their cruise-sampling
tracks for sampling of productive patches.
The physical processes of tidal fronts are now understood, and it can be
shown that the occurrence of a front depends on the magnitude of the tidal
energy dissipation by bottom friction, distributed throughout the depth. Thus
/z/u^, where h is the local depth and u is the tidal-stream amplitude, is an appro¬
priate parameter controlling whether stratification is present or not. A certain
value of hlu^ is seen to mark the transition, and hence a corresponding contour
on a map of should define the position of a front. Fig. 7.17 shows such a
contour map, and the agreement between the contour h/u^ = 70 m“^ s^ (that is,
\ogio hj= 1.85), and the actual frontal positions in Fig. 7.16 demonstrates
how well the theory is able to model the actual process in the sea.
Satellite imagery has been used to study the meandering and spring—neap
excursions of these fronts, and to examine whether seasonal movement occurs.
Clearly, if only horizontal movement is being studied, the satellite data provides
a much cheaper and easier way of proceeding than field surveys at sea. Satellite
IR imagery from all over the world is now revealing similar fronts to be present
in continental shelf seas where the depth is sufficiently small, and the tidal
streams are strong enough, to break down the seasonal thermocline.
It is worth noting too that the front is equally visible on certain CZCS
images, indicating that it has a colour signature which persists when the thermal
signature disappears in spring. This is a reminder that thermal imagery reveal¬
ing SST patterns can only be an indication of dynamical processes so long as a
temperature gradient exists to act as a tracer. It also points out the difference
between two classes of thermally-imaged phenomena, those such as the coastal-
fringe front in which physical processes are occurring which happen to be
revealed by temperature gradients, but which could exist without having a
thermal signature, and those like the tidal-mixing front in which the tempera¬
ture structure and heat balance is an essential part of the physical process. It is
intriguing to pose the question whether some of the turbulent eddies which can
be seen at fronts on images such as Fig.7.13 are dynamically related to the front
and therefore fall into the second category, or whether they are evidence of a
turbulence structure which occurs over the whole of the ocean and just happens
to be imaged in this case because the strong temperature gradient provides a
tracer capability.
species, for others the temperature range is vital. These charts, which can be
communicated to vessels at sea, can be of genuine assistance to commercial
fishing fleets in helping them to find the shoals of fish they are seeking.
CHAPTER 8
2hc'^€(9,4>)
X® [e\p(hc/XkT) — 1]
where
Mx= the spectral radiance at wavelength X, in units of power per unit solid
angle per unit projected area per unit wavelength interval
/z = Planck’s constant = 6.626 X 10”^ J s
c = speed of light = 2.9979 X10® m s"^
k = Boltzman’s Constant = 1.38 X 10“^ J
T = the temperature of the emitter in degrees K
and e(9, 0) is the emissivity of the surface in the direction (6, 0). 9, 0 and the
elemental solid angle dl2 are defined in Fig. 8.1.
_ 2^ e(d,<t>)
[exp(hf/kT) — 1]
hf hf
exp 1+
kT kT
2kf^Te{e,(j>) 2kTe{d, (p)
Thus 5/(0, 0) = (8.1)
X^
Ae=- (8.2)
f G(e,(P)d9.
HIT
(A helpful introduction to the theory of microwave antennae in the
remote-sensing context will be found in Chapter 9 of Reeves (1975).)
Sec. 8.1] The physical principles of passive microwave radiometry 251
B(ei)
The antenna also receives energy from one polarisation state only, that
is, intercepting only half the possible energy flux. Thus the total power
received over a bandwidth between / and /+ A/; at an antenna illuminated by
a brightness Bf (6, (p) of unpolarised radiation, is
1 f
B = -A, Bj-(0,<p)G(e,<P)dndf. (8.3)
2 j/ Jatt
Now from (8.1), if the emitter in direction (0,<p) relative to the antenna
(Fig. 8.2) has a temperature T(0, <p),we can write (8.3) as
kA^Af
P = -^i^Jid,<P)e{d,<t>)G{d,<l>)da (8.4)
However, T^{d, (/>), will normally vary with direction, owing to the variability of
temperature in the atmosphere and across the sea surface being viewed. We
therefore express the power received by the antenna as
P^kT^Af (8.5)
z
+ sec0 / rair(z)0£(z)exp[—sec0 / a(z)dz]dz (8.7)
-d-
Sec. 8.1] The physical principles of passive microwave radiometry 253
FREQUENCY (GHz)
Fig. 8.4 - Contributions to antenna temperature from absorption by oxygen,
water vapour, and the galactic background (After Swift (1980))
The emissivity e and the power reflection coefficient R (as used in (8.7))
are simply related by
^H.v = 1 ~ -^H,V
where the subscripts H and V denote horizontal and vertical polarisation states
which need to be considered, since the antenna measurement is of one polarisa¬
tion state only. For a plane surface, R is dependent on the incident (and
reflected) radiation zenith angle 6, and the complex relative dielectric constant
of water, e:
cos9 — v^— sin’'^0 ^
Rh =-;-■; (8.8)
[cosd + y/e — sin^_
Fig. 8.5 illustrates the general shape of the variation of and eywith
viewing angle, although the exact position of the Brewster angle at which ey
peaks will vary with the value of e.
The complex dielectric constant for sea-water has been expressed by Lane
«fe Saxton (1952) as
, (^s-O . o
e = +-i —— (8.10)
Infeo
Salinity (%o) 0 10
120 r—
12 18 24 30
SEA SURFACE TEMPERATURE Tg (®C )
Figs 8.6 and 8.7 illustrate the resulting variation of brightness temperature
for different SST and salinity conditions at two separate microwave frequencies.
It is apparent that the brightness temperature is very much lower than the actual
SST. At the lower frequency (Fig. 8.6), is much more responsive to salinity
than to SST, and at high salinities the brightness temperature actually decreases
as SST increases. At higher frequencies the salinity-dependence becomes less
(Fig. 8.7), and Tg is a much better measure of SST.
The surface emissivity can also be significantly changed by the presence of
a thin layer of another material on the surface, such as an oil slick, or a film
of organic material, a raft of foam generated by breaking waves and wind action,
or fresh-water ice. Generally, the dielectric constant of the layer will be less
than for sea-water, and the emissivity consequently greater. However, because of
multiple reflections at the air-film and film-sea interfaces, standing waves can
258 Passive microwave radiometers [Ch. 8
be set up, resulting in e having a strong dependence on the layer thickness. For
a given frequency, as the layer thickness changes, e goes through alternately
high and low turning-points, which may differ by a factor of 2. This provides
for the possibility of oil-slick detection from aircraft using microwave
radiometers, but many different frequencies are required to identify the slick
thickness given this complicated e variability with thickness.
The other factor affecting surface emissivity of the sea is the surface
corrugation and roughness due to wind-driven waves. If the surface roughness
length scale can be assumed large compared with the electromagnetic wave¬
lengths, then the resultant emissivity due to an ensemble of different plane
reflecting facets at different angles can be calculated by (8.8) and (8.9), pro¬
vided that the sea-surface slope statistics are known (see Chapter 10). Such
calculations (for example, Stogryn (1967)), using Cox & Munk’s (1954) relation¬
ship between surface-wind speed and sea-surface slope statistics, show the
interesting result that whereas the emissivity of horizontally-polarised radia¬
tion is variable with wind speed at all viewing angles, the vertically-polarised
emissivity appears to be constant at all wind speeds and surface roughness
when viewed at an angle of 50°. Although the actual emissivity is affected also
by backscattering and diffraction from surface-roughness length scales com¬
parable with the microwave wavelengths, so that the model of reflection used
by Stogryn is not entirely valid, in practice there remains a view angle around
50° at which €v is surface-roughness invariant.
Since the wind speed also controls the production of foam, there is a more
complicated relationship between wind speed and emissivity than is suggested
purely by the calculations of surface-roughness statistics.
the thermal skin temperature deviation effect. This is rather academic at present,
when microwave radiometers are not capable of resolving SST accurately enough
to make the thermal skin worth worrying about, but as improvements are made
in future, it could well be another useful feature of microwave radiometry.
The geometry of the offset reflector is such that the footprint is oval rather
than circular in shape. The various frequency bands and their corresponding
footprints are listed in Table 8.1. The scan period is 4.096 s, which is sufficiently
fast to achieve a continuous coverage of the ground within the swath even for
the smallest footprint at 37 GHz, which has to be sampled more often than the
higher frequencies with larger footprints. The five frequencies are sampled
for horizontal and vertical polarisation, resulting in ten separate data channels
containing geophysical information. A considerable amount of data processing
is required to present the channels of data in terms of antenna temperature on a
grid pattern covering the swath. Further details of the procedures are presented
by Njoku (1980). Assessements of the early performance of the SMMR on
Seasat and Nimbus-7 are to be found respectively in Lipes (1982) and Gloersen
(1981).
Axi$ of Rotation
Off*at
Reflactor
distribution over the oceans (for example, Staelin et al. (1976)). Without more
than one channel in the 1—20 GHz range the radiometers could not properly
separate sea-surface temperature from other contributions to the antenna
temperature, and applications were limited to identifying features with a
strong emissivity contrast, such as ice boundaries, or with a strong effect on
atmospheric transmission properties, such as rain and snow.
I DI V 1 S t G = I 0O KM
Fig. 8.12 — Grid cell configuration for the Seasat SMMR (four cells in the swath
width W) (From Njoku (1980))
268 Passive microwave radiometers [Ch.8
(1) Sea surface temperature 6.6V, 6.6H plus corrections 150 X 150
(SST) for water vapour (3) and
cloud (4)
(2) Surface wind 10.7H plus SST (1), water 85 X 85
vapour (3) and cloud (4)
corrections
(3) Water vapour 18V, 18H,21V, 21H 54 X 54
(4) Cloud liquid water 18V, 18H, 21V, 21H, 37V, 37H 54 X 54
(5) Rain rate 18H,37H 54 X 54
noted, however, that these comparisons are made with satellite data obtained in
the absence of sunglint, significant rain, or radio-frequency interference, all of
which seriously degrade the retrieval algorithms. Another limitation is caused
by the side-lobe correction which assumes that the side lobe views the sea at
a similar temperature to the main beam. If it views the land, serious errors
occur, and this restricts the use of the SMMR for SST retrieval at this accuracy
to locations at least 600 km away from a land mass.
Fig. 8.13 — Data-processing scheme for the Seasat and Nimbus-7 SMMRs
(Adapted from Wilheit et al. (1980))
adjacent grid square. These are used in the nonlinear environmental model to
predict the brightness temperatures which they would be expected to produce at
the satellite. Comparison with the actually observed values shows if the predic¬
tion was valid. If not, a differential correction is applied to alter the original
estimates of the geophysical parameters, and the iteration procedure continues
until the satellite-observed radiation in all ten channels can be predicted
sufficiently accurately.
Fig. 8.14 — Nonlinear iterative algorithm scheme for recovering the environmental
parameters from the SMMR (From Thomas (1981))
c o T3
o a <u hs
<U
e <u
<u 3
> cr 1> 5d cr ’T)
o o a> a (L> Cd
<D
'4-»
c Cd o
CO S S cd
C o
C
o s g C .2 C
o
.2 g g
.2 o a>
c o *■4-* <D
o <u C 00
Table 8.4 Comparison of IR and microwave measurements of sea-surface temperature
!U o- - ^ 4Z
cd
'o
s o cd
Sn -
o - ^ ca £? !U 00
c C
O w ^ Cd o
cd «« ii H-1 cd ■$
(U
•4-)
cd
(-1 nU
cq a> O ^ Cu cd ’>
> cd ^ § o 5 o :2
a "S > a> x:
s M c C 4:5 "cd
o
achieved
^ »-i fl)
c/3
ca C.UJ t-. ^ o <u 00 CD
1.5 degK
o o c
1 10 Ui c/3 o o -4->
2 degK
c/3 Cd <D
oq d T3
Ui
>% ^ 2 W S 'o
4-1 rz3 cd
z< ro
d
0 Cd Cd
>
c/3 cd
>
c/3
Cd
oq 0 vu W Q
+ +
C Q 0 0
.2
.2
X> 3 V-*
0 u
u 0 0 a>
0-1 c/3 CL X 'O
e <U (U
3 Ui
-Td E -4->
C/3
o -4-^
-S
Cd Cd
0)
c cd
o -£ ^ .2
X
So o O
6 ta
<D
C 00
c: .2
X) & e
0
13
00 .2
SZ
c Ui u ’-M
t; S § ^ cd 0
on-board target
cd 00
s ca -o 'TU o (i>
<u
Ui n:3
CO a>
.9 C
<L» o
(U
4-J s
cd
*2 00
0
c >
<D
00 Ui o o cd 00
cd Ui
Cd 0)
Cd
CL ^
^ I
Cd
U4
Ui X> •2 > s
CO
.2
'O
0 2
O
00 o cd
Ui ■$
0.6 degK
C/3 0 .2i
0.1 degK
N Ui
Cd c -2 o o cd <1>
Cd
cd g & a> O
Ui Xl
0
0 *>
CCl
«—(
Ui
-4—• a- E X5 Ui G ca -4-J
/< ro
1
1
^
fTD ?l
(I 0
0
D <L>
c
cd
S .S L—1
Ui
:3
0 cd
<D
«j Z
c cd
u CO
1 oq
1
-- + -I- + + +
C H
Presently achievable sensitivity
tZ)
_o on
Presently achievable absolute
(D
-4—>
a> .2
O
z^d
e a>
o
Cd 0
kH C/D C
<D 0) cU
c/3
0
cd
b-i o f2
c
-4-*
Um c
c
C4—1 c^ 0
oc Ui
.s 0
C/3
S'
'u
u
D
u->
0
0
0
.2
cd
-4-^ X C u
u Cd
0 Cum
U-*
t3 X
00 0 0
0 Ui cd
accuracy
<u Ui '0
>> a> 00
XI >v Q./
-4-* ■4-M <L> X U CD
ZJ
-4-> s [> c/3
u-»
Cd
CL
c/3
00
C
u-<
o ’S00 U-> ■4-» cd
Cloud
*C/3 0
cd 0 00 00 U-4 *0
• '?
Un cd
cd c
s Cd
«D
E
■4-M
Cd
cx
0
CO
X
0
S c/5 w C/5 <C 00 > <
272 Passive microwave radiometers [Ch.8
cloudy areas, it would seem that on balance the superior resolution of the IR
devices at present outweighs their cloud limitation. Very little can be done,
however, about the IR cloud problem, whereas there is room for significant
improvement in the microwave radiometer performance through technological
development, larger antennae, array imaging, etc. Furthermore, given the micro-
wave radiometer’s diversity in measuring other parameters besides SST and
its value for ice studies, it is possible that in a decade or so it will have been
developed to a design which can compete with IR radiometers.
ture because of diurnal thermocline and surface skin effects), or both. The maps
must be constructed by interpolating between many point observations, whilst
applying smoothing criteria so that one single measurement which is seriously
in error will not grossly distort the map. This is more of a problem with ship
data from many vessels, than with the satellite data which should have self-
consistency. The satellite data are also regular and have a fuller coverage than
the ship data, and although the ground resolution (150 km X 150 km cells)
seems rather poor, it is at least as good as can be achieved by routine reports
from ships on passage. Consequently, when the database from which they were
constructed was compared with the resulting maps of SST, the SMMR data were
found to have a scatter of 1.5°C relative to the map, whereas the ship data
scatter was 2.1 C. This suggests that the regular, uniform coverage of a single
sensor, although its inherent accuracy is only around 1°C, is able to produce a
better map than the haphazard sampling from ships.
Fig. 8.15 illustrates the monthly SST-anomaly maps produced by satellite
and ship observation for July to September 1978. They show a general
agreement, but the satellite maps are more complete than the ship maps. By
constructing 8-day coverage SST maps from the SMMR data (which would
be almost impossible with the less regular ship data) and presenting them in a
time sequence (Fig. 8.16), it has become possible to observe fluctuations in
the intensity of the cold equatorial tongue which stretches across the Pacific,
and meanders in the strong meridional SST gradient just north of the equator.
The wavelength and speed of the meanders can be estimated by presenting the
time-variation of a particular isotherm, as shown in Fig. 8.17, and individual
meanders can be tracked over 3000 km. In these interesting observations,
Bernstein & Morris have selected the type of information which the SMMR is
best able to supply. Comparison with Fig. 7.9 reveals that IR radiometers would
be prevented by clouds from obtaining equatorial SST maps with an 8-day
sampling frequency, and it appears that microwave radiometry from space is the
only means of directly examining the dynamics of the equatorial ocean at length
scales of around 2° latitude and above, and timescales of about 10 days.
Jul 7 - Aug 4,1978 7-Aug 4.1978
274
Passive microwave radiometers
Fig. 8.15 - Monthly maps of sea-surface temperature anomaly in the North Pacific Ocean for July, August, and September 1978. (a) Based on
[Ch.8
Seasat SMMR data, (b) Based on ship-observed data. (From Bernstein & Morris (1983) copyrighted by the American Geophysical Union)
Sec.8.5] Applications of SMMR data 275
(a) (e)
(b) (f)
20*N
10°N-
Fig. 8.16 - Time sequence of 8-day composited Seasat SMMR SST maps for the
Eastern Tropical Pacific. There are 6-day steps between the maps. (All periods in
1978):
(a) 6 13 July (b) 12 ^19 July (c) 18-25 July
(d) 24-^ 31 July (e) 30 July ^ 6 Aug (f) 5-12 Aug
(g) 11 ->18 Aug (h) 17-24 Aug (i) 23 — 30 Aug
(j) 29 Aug ->■ 5 Sep (k) 4-11 Sep (1) 10-17 Sep
(m) 16 ^ 23 Sep (n) 22 - 29 Sep (o) 29 Sep — 5 Oct
(Adapted from Bernstein & Morris (1983))
Sec. 8.5] Applications of SMMR data 277
12-19 July
18-25 July
24-31 July
30 July-6 Aug
5-12 Aug
11-18 Aug
17-24 Aug
23-30 Aug
29 Aug-5 SapI
4-11 Sapi
10-17 Sapi
16-23 Sapi
22-29 Sapi
28 Sapi - 5 Oel
Fig. 8.17 — Composite plot of the 24°C sea-surface isotherm from the fifteen
maps of Fig. 8.16, illustrating the westward progression of wave-like features
along the dashed-line trajectories. (From Bernstein & Morris (1983) copyrighted
by the American Geophysical Union)
CHAPTER 9
Satellite altimetry of
sea-surface topography
9.1 INTRODUCTION
9.1.1 An exciting new oceanographic instrument
Judging by the number of scientific papers which have appeared since the three-
month flight of Seasat in 1978, the observations made by the radar altimeter
on board have attracted the attention of many ocean scientists. This is not
surprising when the achievement of the Seasat altimeter is considered. It was
able to measure the distance between the satellite and the sea surface below it to
within 10 cm, no mean feat when the distance is about 800 km, an accuracy
of one in eight million! When it is realised that the altimeter offers oceano¬
graphers the chance of doing something they have dreamt of for a century,
namely, measuring the absolute slope of the sea surface, the important implica¬
tions of the altimetric achievement become apparent. Add to this the ability of
the altimeter to accurately measure the significant wave height and to estimate
the surface wind speed at the same time, and we have the reasons why the major
space agencies are each planning to fly another altimeter within the 1980 s.
The satellite-borne radar altimeter promises a great deal for dynamical
oceanography, but even the accuracy achieved on Seasat was barely sufficient
to be of real value, particularly because the flight was so short. It will take the
next generation of satellite altimeters to measure distances within a few centi¬
metres before the technique’s exciting potential in oceanography is fully realised.
In this chapter we examine the principles of satellite altimetry and the
various sources of error which must be allowed for. We look at the different
components of the altimetric signal and how the oceanographic part can be
extracted. We explore the rhechanisms of dynamical oceanography which make
possible the interpretation of sea-surface height and slopes in terms of ocean
currents on a variety of length and time scales. Finally we look at other oceano¬
graphic applications of altimetry, including the measurement of ocean tides and
the estimation of ocean bathymetry. We leave until Chapter 11 the consideration
of another separate use of the altimeter to measure significant wave height and
to estimate surface-wind speed. Satellite altimeter data have also been applied
[Sec. 9.1] Introduction 279
enthusiastically to the study of the geoid and in the observation of the snow-
and ice-covered polar regions. Except insofar as these applications overlap with
oceanography they lie outside the scope of this book.
Scientific understanding of how to analyse and apply the observations
of satellite altimetry is rapidly developing. This chapter is intended as an
introduction to the principles of the technique, and not a review of the latest
developments in the recent literature. Many of the key papers have been
published together in collected form, and to explore the field further the reader
is directed towards the final eleven papers in Gower (1981), the first sixteen
papers of Bernstein (1982 c), the first fourteen papers of Kirwan et al. (1983),
and part 4 of Allan (1983).
SATELLITE ORBIT
regarded as noise to be removed from the geoid signal. Until the geoid is better
defined, marine dynamicists must find ways of extracting oceanographically
useful information from altimeter data which is not distorted by unknown geoid
variability. As in other areas where geodesy impinges on oceanographic science,
such as the measurement of mean sea level at tide gauge stations levelled into a
geodetic network, caution and clarity of thought are required to avoid unsound
assumptions in which the geodesist and oceanographer each rely on the other to
provide an absolute reference level. Otherwise what is believed to be an absolute
measurement will turn out on closer inspection to be entirely arbitrary, and
variable.
Before considering these geodetic problems, we first examine the process of
measuring the distance between the satellite and the sea surface.
c
UJ
TIME (n«)
Fig. 9.2 - Smooth curves fitted through averaged altimeter pulse shapes corres¬
ponding to two different sea states. SWH = significant wave height. (From
Walsh, UUana, & Yaplee (1978))
of the pulse reaches the ground, and then levels out. The timing algorithm fits
a curve to the leading edge of the return signal up to its maximum, and hence
the effective area whose height is sampled is the illuminated circle just before it
becomes an annulus. For a calm sea and a pulse length of fp, this has a radius
of ''a ~ (2 but for a rough sea with significant wave height hij-^ its area
is increased to
Fig. 9.3 - Interaction of a pulse of duration t with a smooth sea surface. The
lower part shows the annular shape of the illuminated region, and to the right is
shown the strength of the returned signal. (Based on Walsh, Uliana, & Yap lee
(1978))
284 Satellite altimetry of sea-surface topography [Ch.9
Ah = VitjCm h{n-\)
where h is the true height and n is the refractive index of the medium
(n = c/cni)- In the estimation of the error Ah, it is sufficiently accurate to use
the approximate height of the satellite (for example its nominal orbit height).
In practice the refractive index of the medium varies with height, in which
case the error is calculated as
Ah = /J'(n-l)dz . (9.1)
where the constant a = 80.5 m^ s"^, N is the number of free electrons per unit
volume, and / is the radar frequency. Thus for the ionosphere, from (9.1) and
(9.2),
Now N varies between day and night and between winter and summer
hemispheres, being less at night and in summer. It increases with the solar
sunspot maximum. Consequently it cannot be directly predicted. It can be
calculated from ground observations looking at its effect on radiation coming
through the ionosphere, and this approach was used to make estimates of Ahj
for the Seasat altimeter (LoreU et al. (1982)). However, ground observations
are widely spaced, and the necessary extrapolations between them can lead to
gross errors in the estimate of N. Fortunately, the influence of the term in
the error reduces the magnitude of Ahi to a maximum of 20 cm at 13.5 GHz,
and for Seasat the correction could be made to within 3 to 5 cm. For higher
accuracy it would be necessary to have more ionospheric observations or else to
exploit the /"^ factor in the correction by using a multifrequency altimeter.
In this approach the difference in travel times at different frequencies could be
used to estimate the transmission error.
In the troposphere the refractive index is influenced by the dry gases
Sec. 9.2] Distance measurement with a radar altimeter 285
present in the atmosphere and by water vapour. This leads to one of the major
altimetric errors, of the order of 2.5 m, which has to be calculated if the
altimeter measurements are to be of any value. The methods used for calculating
the Seasat tropospheric errors are given by Tapley et al. (1982 a).
The dry atmosphere correction, in metres, can be obtained from the
known vertical distribution of gases in the atmosphere, using
where 0 is the latitude of the subsatellite point and po is the surface air pressure
in Pascals. Thus A/Z(j is typically of the order of 2.3 m, but using global inter¬
polations of the surface pressure field from meteorological observations it is
possible to estimate A/z^j to an accuracy of within 1 to 2 cm.
The wet atmospheric correction A/z^ can be calculated from
the area of the effective footprint. The rougher the sea, the harder it is for
the logic algorithm which tracks the slope of the leading edge of the pulse
to operate accurately, and a tracker bias occurs. In addition, the shape of ocean
waves is not sinusoidal. Troughs tend to be flattened and crests steepened, so
that the radar return for the troughs is stronger than from the crests, leading
to an electromagnetic bias which emphasises the troughs, and overestimates
the distance between the satellite and the mean sea level. The overall effect
of surface roughness is known as sea-state bias and was estimated for Seasat
to be 7% of the significant wave height, which should be subtracted from the
calculated satellite-to-ocean distance (Born et al. (1982)). The electromagnetic
bias which will be common to all altimeters operating at 13.5 GHz appeared
to be between 1.5% and 2% of and the tracker error, specific to Seasat,
was 5% to 5.5%. Even when a 7% bias correction is applied, there remains a
scatter of around 3% of /21/3 in the observations, as checked from repeat orbits
over the same piece of sea under different wave conditions.
Given these various sources of error it is remarkable that the Seasat orbit
determinationf resulted in errors of only 1 to 2 m, but this still makes the orbit
tracking a principal source of error for satellite altimetry. One way of objectively
determining the error in orbit calculations is to compare the altimeter measure¬
ments at points where the ground tracks cross on different orbits. The experience
of Seasat has led to an expectation that it will be possible in future to reduce
the radial orbit error to 10 cm provided that an improved geopotential model
is developed (which will require further measurement of the earth’s gravity
distribution by satellites), provided that the satellite is dense and/or in a higher
orbit than Seasat to minimise drag and pressure forces, and provided that the
computation methodology is refined.
It should be noted that the present error level of around 1.5 m applies to
the whole-earth coverage of all orbits. Because the error varies over length scales
of order 10000 km or more along the track, then its effect can be minimised
to effectively less than 10 cm over arc lengths of 1000 km or less. This can
be achieved either by looking only for relative height differences, or by fitting a
Unear error trend between two ground-control points spaced close together,
where this is possible.
1 Noise 5.0
Altimeter
\ Bias 10 2.0
Sea state Waveheight and 7 ± 2% of /2i/3 2% of hy^ 500-1000
related bias tracker biases
Troposphere Mass of air 240 0.7 1000
Troposphere Water vapour 10-40 3.0 50-500
Ionosphere Free electrons 2-20 3.0 50-10000
Liquid water Clouds, rain 10-100 30-50
1 ' Gravity
Drag
10 km
300
140
30.0
40000
10000
Orbital error <
Solar radiation
1 Station location
300
100
30.0
10.0
10000
10 000
Timing Data time tag 5.0 20000
f See ‘SEASAT ephemeris analysis’, special issue of J. Astronaut. Sci. 27 (no. 4), 1980.
Sec. 9.3] Establishing a datum 289
the altimeter data, and therefore cannot be used for its correction. Consequently
it is not surprising that most interest has been shown in Seasat altimeter data not
by dynamical oceanographers but by geodesists, determining the marine geoid
(Brown etal. (1983)).
If tidal corrections are applied, the geoid does not change with time,
and it is therefore possible with repeat orbits to detect time variability in sea-
surface topography from the differences between different overpasses of the
same transect of ocean. At present, however, there is no way of confidently
distinguishing the residual long-period average sea-surface topography from the
spatial variability of the geoid. The only prospect of being able to achieve an
independent measure of the geoid to an accuracy of a few centimetres over
length scales of a few hundred kilometres is by using a dedicated gravitational
satellite system in which the distance between two satellites in low orbit would
be accurately measured (Douglas et al. (1980)). Without this, future successors
to the Seasat altimeter will continue to provide information only on time-varying
aspects of ocean dynamics.
fv = - — (9.6)
p ox
fu =-. (9.7)
P
These are the east and north pressure balances respectively. Here p is the local
fluid pressure, u and v are the horizontal velocities in the x and y directions, p
is the sea-water density, and / is the local Coriolis parameter, defined as
/ = 2^2sin0
where 0 is the latitude, and = 7.272 X 10"^ rad s“^ is the earth rotation rate.
Sec. 9.4] Application of altimetry to the study of ocean currents 291
Equations (9.6) and (9.7) express the physical reality that water flowing
horizontally experiences a horizontal force perpendicular to the direction of
flow, tending to move it to the right of its direction of flow in the northern
hemisphere, and to the left in the southern hemisphere (where / is negative).
For currents at the sea surface, the horizontal pressure gradient is propor¬
tional to the sea-surface slope measured relative to the equipotential surface
(that is, the geoid). If the sea-surface height is f relative to the geoid, (9.6) and
(9.7) become
fv = g — (9.8)
bx
fu = -g — . (9.9)
by
Equations (9.8) and (9.9) show how sea-surface topography measured from a
satellite altimeter can be directly related to ocean currents. Simple inspection
will show that a current of magnitude 1 m s”^ at 43°N (where / = 10"'* s"‘)
will have an associated sea-surface slope of about 10“^, or 1 cm in 1 km. Ocean
circulation velocities are normally of the order of a few cm s"\ with a slope of
less than 1 mm in 1 km, or 10 cm in 100 km. This is the reason why such a
high-accuracy geoid is required for satellite altimetry. Clearly, these typical
geostrophic slopes are much smaller than some geoid slopes.
Nevertheless, their observation is of great importance to oceanographers, for
the following reason. Until now, apart from a few costly deep-sea current-meter
moorings, oceanographers have been unable directly to measure ocean currents
far away from land with any absolute accuracy, since they cannot provide a
stationary platform to which the measured speed can be related. They have,
however, been able to measure the vertical distribution of density in the ocean
from temperature and salinity soundings. This mass distribution is related to the
pressure field by the hydrostatic balance equation
bp
^ = -pg. (9.10)
bz
If the pressure field dictated by the mass distribution results in the horizontal
pressure gradient being different at different depths, then from (9.6) and (9.7)
the horizontal geostrophic velocity must vary with depth. This is expresed by
the ‘thermal wind equations’ (named from the meteorological analogue), one
form of which is:
_^(pv) _
(9.11)
bz f
^(pu) _ £^ (9.12)
bz f by
292 Satellite altimetry of sea-surface topography [Ch.9
294 [Ch.
-15.
S 0.0
QJ
or
2.5
£
2.0
C3
LU
I >-5
Li_J
CXI
^ l.o
cr
CD
0.0
longitude
Fig. 9.4 - Analysis of eight collinear altimeter profiles collected along track 2
of Fig. 9.6, by Seasat over the Gulf of Mexico, between September 17 and
October 8, 1978.
(a) Sea height relative to the reference ellipsoid.
(b) Profiles replotted after subtraction of the group mean.
(c) Profiles after removal of orbit error through tilt and bias adjustment, overlaid
(lower), and displaced vertically at 3-day intervals (upper).
(d) Root-mean-square variability about the mean.
(From Thompson, Born, & Maul (1983) published by the American Geophysical
Union)
Sec. 9.4] Application of altimetry to the study of ocean currents 295
LONGITUDE
Fig. 9.5 - As for Fig. 9.4, but for track 3 on Fig. 9.6. (From Thompson, Born, &
Maul (1983) published by the American Geophysical Union)
296 Satellite altimetry of sea-surface topogrq)hy [Ch.9
30*N
25*N
20*N
Fig. 9.6 - Tracks of Seasat over the Gulf of Mexico, showing the location (tracks
2 and 3) of the altimeter transects illustrated in Figs 9.4 and 9.5. Also shown is a
typical (but not permanent) current pattern in the Gulf, in which an anticyclonic
eddy is about to separate from the Loop Current. H and L correspond to raising
and lowering of the sea surface and subsurface isothermal surfaces in geostrophic
balance with the current system. (Adapted from Thompson, Born & Maul (1983))
Menard (1983) has developed the techniques for mapping the spatial
distribution of the time variability of sea-surface topography, based on the three-
day repeat orbits. The deviation of the surface height from the mean of the
8 or 9 orbits are calculated at each 0.1° of latitude along the orbit track to
obtain the surface deviation for each (ith) pass at each location. Fig. 9.7
shows a map of surface-height variability in the North Atlantic, drawn by hand
from the values of a^(Af') at each calculation point, where the variance is:
aMAf'l =
n
Menard also calculated the variable velocity field Av„ in the direction n, normal
to the satellite track s, with the geostrophic relation:
Sec. 9.4]
Application of altimetry to the study of ocean currents
Fig. 9.7 — Sketch of sea-surface topography variability in the N.W. Atlantic, constructed from the Seasat altimeter 3-day repeat data along the
track lines shown. The interpretation of the data is subjective away from the track lines. (From Menard (1983), copyrighted by the American
297
Geophysical Union)
298 Satellite altimetry of sea-surface topography [Ch. 9
(U
O
•M
00
o G d^
O
-i ^
o “ J3
3:
o
s
. ^
<U a>
tG > •O
G
B ^
dj ^
•M (1> 00
a c C
D
C/3 B °
*0
% »o O
d> CO
r*
G
(D
^1'
«-7
>v >>
73 ^
E S O
o
*00
g ^
CO o cO
o a> o >
Bo^
. O 0)
.■az:5
o
B o
■^ o _ S
l-t IH V3
CO
> .B
cO
<w
G>
.2P O
*S
4=
cO
<u
V)
'B
CJ c
o
a 2^
<u CO Xi
B B ■=«
CO
"eo
Xi - s
.S
--S '-G
O o G
^ 2
>> ^
ON o 00
ON CO *Q
.s
.Op Cu.
d> IH
I-I 00
O O o O O o
(O CM CM (O
I t
300 Satellite altimetry of sea-surface topography [Ch.9
was calculated.
n
On the assumption of an isotropic turbulent eddy field, that is,
o\Av„) = o\Av,)
cycles per day (diurnal and semidiurnal tides), and the satellite in a repeat orbit
has a regular return time of a few days, some aliasing is inevitable. Fig. 9.10
illustrates the effect of aliasing. The observed time series appears to have a
much longer period than the signal being sampled. If the sampling interval is an
exact multiple of the signal period, the signal is aliased to zero frequency, and
no information about the tide at that frequency can be extracted. Thus a sun-
synchronous orbit which repeated its track exactly over an exact multiple of
24 hours would alias the Solar S2 and Si constituents to zero frequency, and the
lunar M2 tide would appear as a 14-day period constituent.
phase agreement, to within one degree, but less satisfactory amplitude agreement
(about 47% low). This is probably typical of the expected performance of
sateUite altimeters for tidal observations. It will be easier to obtain more accurate
phase than amplitude data. The Cartwright & Alcock results also bear this out.
0.5° GRID, 1 M CONTOUR
Ae = 6378137m, 1/f= 298.257
REF. NO. 83402 Sec. 9.5]
Fig. 9.13 — Spatial distribution of the ground track over the N. Atlantic for a
10-day orbit repeat pattern.
Fig. 9.14 - Spatial distribution of the ground track over the N. Atlantic for a
3-day orbit repeat pattern
Sec. 9.7] The selection of satellite orbits for altimeters 309
Fig. 9.15 illustrates how the precession rate is affected by the inclination
and the height of the satellite. The constraints of atmospheric drag (at low
altitude) and lack of ocean coverage (at low inclination) are included, and the
dotted shading indicates the precession rates at which tidal aliasing is worst.
Two possible regions for an ocean-altimeter orbit are indicated. The one around
64 inclination gives less tidal aliasing problems than the one around 72°, but
poorer global coverage.
aliasing
Fig. 9.15 — Orbit inclinations plotted against the resulting precession rates for
different heights of orbit, showing the limits within which a suitable altimeter
orbit is found. The dots indicate that tidal aliasing is a problem. The numbered
regions represent the parameter combinations of promising orbits.
reflected energy, but since even the roughest sea is unlikely to have slopes tilted
at more than 20°-25° from the horizontal, specular reflection is important only
for viewing angles between 0 and about 15°.
For viewing angles greater than this, a return signal is only achieved by
scattering at the surface, and this depends on the surface roughness. Fig. 10.1
illustrates different degrees of surface roughness resulting in (a) purely specular
reflection, (b) some scattering but a dominant reflection in the specular direc-
(c)
Fig. 10.1 - Radio-wave scattering from (a) a smooth surface (specular reflection)
(b) an intermediate surface, and (c) a rough surface
Sec.10.2] Radar reflection from the sea surface 313
tion, and (c) complete scattering. The definition of roughness varies with radar
wavelength Xr and viewing angle 6. A smooth surface (a) can be considered to
exist when the vertical height of surface roughness irregularities is
(10.3)
2 sm0
For sea waves whose crests are at an angle 0 to the radar line of sight (Fig. 10.3),
the criterion is
XRSin0
(10.4)
2 sin0
Fig 10.3 - Bragg scattering geometry; surface-wave direction at angle </> to plane
of radar waves
In the context of ocean remote sensing, Xr and 6 are fixed within narrow
bands, whilst the sea surface contains a whole spectrum of different wavelengths
as discussed in section 3.4. This results in the radar responding most strongly to
those waves for which
XRsin0
2sin0 ’
and a is therefore an approximate measure of the amplitude of such waves.
The choice of radar wavelength is seen to be crucial in determining the
sea-surface roughness which can be detected, both in controlling the height of
roughness which can be detected (through (10.2)) and also the ocean wave¬
length, if any, which will dominate the backscatter return (10.4). For Seasat,
the L-band SAR had Xr — 230 mm, 6 ~ 20°, and therefore would observe
as ‘rough’ any surface for which 40 mm. The scattering cross-section was
most strongly influenced by waves whose apparent wavelength along the line of
sight was about 300 mm.
Sec. 10.3] Wind-generated surface-wave roughness 315
In an imaging radar, there are several processes which can contribute to the
apparent spatial variation of backscatter cross-section. These include the layover
effect, when the surface is tilted along the line of sight, and the various modula¬
tions which occur because of the interaction of different scales of roughness.
The swell waves being imaged have much longer wavelengths than the short
gravity waves which cause the Bragg resonance, and there are various possible
ways in which the longer waves can modulate the Bragg scattering by the shorter
waves. These will be discussed in more detail in Chapter 12.
For obhque viewing of rough surfaces, shadowing will occur, in which
part of the surface is hidden from the view of the sensor by another part. This is
a further complication which may need to be incorporated into any model of
surface scattering which is constructed.
The whole subject of interaction between ocean and radar waves is reviewed
by Valenzuela (1978 b).
10.2.4 Polarisation
As well as the amplitude of the reflected radiation, other properties are also
changed by reflection, and potentially can be analysed to yield information
about the reflecting surface. Doppler frequency shift and phase information will
be considered further in the chapter on SAR (Chapter 12). Another property is
the polarisation of the radio waves. The transmitted signal is normally horizon¬
tally or vertically polarised. On a single reflection, the polarisation is not changed
very much, but if multiple reflections occur, the polarisation orientation can be
significantly altered. In this case, a radar system employing HV polarisation
(Horizontal polarisation of the emitted energy, and reception of a vertically
polarised return) would encounter a significant return. This occurs when viewing
land surfaces, and useful information can be gained to distinguish between, for
example, grasslands and forest.
Over the ocean, very Httle multiple reflection occurs, and ocean-viewing
sensors are normally HH or W. There is little to be gained by examining the
cross-polarisation signals, but the polarisation ratio P = CTvv/*^hh> the ratio
between scattering cross-section for vertically and horizontally polarised waves,
can contain oceanographic information. P is usually large, because vertically
polarised radar reflects more strongly than horizontally polarised, but P varies
with angle of incidence and radar frequency. It has been found to vary with
wind speed. Valenzuela (1978a) discusses further the possibilities of using
polarisation information for remote sensing purposes.
follows that it is the part of the surface wavelength spectrum between about 5
and 500 mm which is important for the creation of radar surface roughness.
Reference to Figs 3.9, 3.11, and 3.12 reminds us that these are high-frequency
waves, with a period much less than 1 s, for which surface tension is usually an
important if not the dominant restoring force. They are also located in the high-
frequency tail of typical ocean-wave spectra, and the energy associated with
them is negligible in comparison with the rest of the energy in the main part of
the spectmm.
For this reason, they have been of very little interest to oceanographers
hitherto, except insofar as they influence the surface-roughness height in the
parameterisation of the turbulent atmosphere—ocean boundary layers. Typical
oceanographic wave-measuring instruments have been designed to measure the
main part of the spectrum, and are unable to resolve such high-frequency, short-
wavelength waves. The shape of this end of the spectrum is therefore not well
documented, and not very well understood. Indeed, most of the information
we have about high-frequency surface waves is based on radar measurements.
If the radar surface-wave interaction problem is to be objectively tackled, new
techniques of high-frequency wave measurement by independent methods, such
as lasers, must be developed.
Until such observations become available, it is being assumed that the high-
frequency tail of the spectrum is a continuation of the —5 power law which is
a very good model of the observed shape of the frequency spectrum up to a
frequency of 12 rads s~* (14-second period). In this region, the spectral energy
density is inversely proportional to the fifth power of frequency. Phillips (1977)
describes this as the saturation range of the spectrum, because it represents
a steady state even when the peak area of the spectrum at lower frequencies
is evolving in response to wind forcing. The assumption is made that once the
wind is stronger than a few cm s"*, sufficient to roughen the surface with high-
frequency ripples, the ripples reach a high enough amplitude for energy to be
lost from this part of the spectrum at the same rate as it is fed in by the wind.
The energy is lost by dissipation processes, and by nonlinear wave—wave inter¬
action it is cascaded to higher and lower frequencies. In this way the high-
freqency tail remains the same whilst the low-frequency peak of the spectrum
develops towards lower frequencies, depending on the strength of the wind.
The hypothesis that this theory holds into the Bragg resonant wavelength,
capillary-wave part of the spectum, still remains to be experimentally tested.
If it does, then it presents a problem for the principles of scatterometry which
assume, with empirical justification, that the radar backscatter increases propor¬
tionately with increased wind speed. If the saturation hypothesis holds, and
Bragg resonance is the only backscatter mechanism, it would be expected that
the radar backscatter would soon reach saturation like the wave energy in the
tail of the spectrum. That this does not happen may mean that mechanisms
other than Bragg scattering contribute to the radar backscatter cross-section. As
Sec. 10.4] Surface films and slicks 317
surface debris
Fig. 10 4 - Surface convergence thickens surface films and forms a line of surface
debris
Short gravity waves and capillary waves are damped by the dynamic
elasticity of the water surface, that is, by changes in surface tension which occur
when the surface is stretched or compressed. This has the effect of extracting
energy from those waves which depend wholly or partly on surface tension to
provide the restoring force necessary for wave propagation. If a surface film is
present, the surface tension is lower than it would be in the absence of the film,
and stretching and compression of the film due to the presence of waves provides
the dynamic elasticity which enhances the wave damping. Thus capillary waves
and short gravity waves are always damped in the presence of surface films. If
the surface film is spatially patchy, varying in thickness, or lined up in slicks due
to surface convergence, it is to be expected that the capiUary/gravity wave energy
will reflect that patchiness, being greatest where there is no surface film.
The rate of damping is not, however, simply related to the thickness of the
film. Fig. 10.5 shows the variation in damping coefficient with the film pressure
(that is, the reduction in surface tension due to the presence of a surface film).
Fig. 10.5 - Rate of damping of capillary waves (wavelength 0.52 cm) on sea-water
as a function of film pressure (the decrease in surface tension due to a surface
film). (After Garrett (1967))
Sec. 10.5] Dynamical causes of sea-surface roughness patterns 319
the wind, even if it is only a relatively light breeze, the sea may become quite
rough. If the tidal current varies from place to place, so will the sea-surface
roughness. This potentially provides a mechanism for defining the distribution
of tidal currents, provided that the wind is not so strong that the whole sea
becomes uniformly rough.
Fig. 10.6 - Surface convergence produces wave steepening and possibly breaking,
enhancing surface roughness
waves. A ship steering a curved course, or changing speed, does not produce
such a clear pattern, although Stoker (1957) gives solutions for other types of
ship tracks. It should be remembered that if the sea is rough, the ship-wave
pattern will be lost, and not be capable of being imaged.
Fig. 10.7 - Surface wave pattern generated by a ship moving at a constant velocity
further research in this area. The Synthetic Aperture Radar in particular has
revealed a wealth of surface-roughness patterns in the images recovered from the
short Seasat flight. We shall return to these examples in Chapter 12, but of
course the subject of radar interaction with sea-surface roughness which has
been introduced in this chapter has relevance to a variety of different active
radar remote-sensing instruments.
CHAPTER 11
The altimeter as a
surface-roughness sensor
when the trailing edge of the pulse has reached the bottom of the wave troughs at
the nadir point that the illuminated area will equal the area of the illuminated
annulus in the flat case, and thereafter remain constant (on average). In contrast
with the flat case, where the reflection of an emitted step pulse grows to its full
strength within the time of the pulse duration, it takes many times this duration
for the pulse to grow to full strength in the rough-sea case, provided that the pulse
width is small compared to the wave height.
Fig. 11.1 — Altimeter pulse incident on a rough sea surface showing extended size
of pulse limited circle. (From Walsh, Uliana, & Yaplee (1978))
The higher the waves the longer it will take for the return signal to grow, so
that the slope of the leading edge of the reflected pulse is inversely proportional
to the significant wave height. To obtain a meaningful measure of wave height
from the slope of the return pulse, it must be possible to measure the pulse-return
times and to sample the growth of the pulse with sufficient precision. Also, the
pulse width (a pulse must be only 1 ns long if its effective length in space is to
be about 0.3 m) must be small compared with the wave height. If it is not, the
leading-edge slope of the reflected pulse will be hard to distinguish from the slope
obtained from reflection at a smooth flat surface.
So far we have assumed that radar power reflected back to the satellite
is proportional to the area of surface illuminated at any instant. Given a rough
surface, there will in general be a random distribution of sea-surface slopes
at each point where the microwaves reflect from a portion of the sea surface.
Since the altimeter depends on direct specular reflection, it will receive only
those reflections for which the surface facet is oriented normally to the incident
radiation. For a single return pulse the random distribution of surface facets
results in a signal which is dominated by noise (Fig. 11.2). It requires the addition
of n return pulses to reduce the noise by a factor of When 1000 pulses
are examined, the resulting standard deviation is only 3%, as illustrated in the
lowest curve of Fig. 11.2.
As mentioned in Chapter 9, the greater the significant wave height, the greater
is the area effectively sampled for altimetric measurements, and the altimetric
measuring system must take account of the fact that over this wider area the
326 The altimeter as a surface-roughness sensor [Ch. 11
average distance from the satellite to the mean sea level is slightly greater than
under calm conditions when the sampled area is a smaller circle at nadir. A more
significant error is introduced into the altimetry by a rough-sea surface because
of the skewed surface-slope distribution. Because the sea waves tend to be slightly
trochoidal in profile, with peaked crests and flatter troughs, the surface-area
elements with zero slope are found preferentially in the trough regions, whilst the
greatest slopes are found nearer the crests. Since the strongest radar reflections
come from the zero-slope areas, there is a tendency for the strongest signal to
come from surfaces below the mean level, and this tends to give an overestimate
of the distance to mean sea level when many samples are averaged from a rough
sea, producing a sea-state-related error in the altimetry.
Fig. 11.2 — Examples of a single and averaged return pulses received by a radar
altimeter after scattering from many wave facets on the sea surface. (From
Townsend, McGoogan, & Walsh (1981))
achieved slightly before the true timing of the pulse return from mean sea level.
This is allowed for in the altitude tracker algorithm which positions the time
corresponding to the mean level. In Fig. 11.3, the three curves have been centred
about this point. On Seasat the wave-height algorithms operated on the values
of 60 gates, and so the profiles are centred with the return corresponding to
the mean level fixed between gates 30 and 31. For Seasat the height tracker
algorithm allowed for the plateau-droop by locating the mean sea level assumed
return at a time when the power had a value of 60/53 of the mean of the 60
gates surrounding it.
Fig. 11.3 — Seasat altimeter pulse shapes for three different wave-height condi¬
tions, and fitted solution curves. The data are averaged over 10 000 pulses (10 s),
and have been corrected for gain biases in the sampling gates
’^li?(0)|2 , -tan^0
o 0 = —=— sec^0 exp — (11.2)
s L s _
where /?(0) is the Fresnel reflectance of the air—sea interface at normal
incidence, and has a value between —2.08 and —2.37 dB at the 13.9 GHz
frequency used by the Seasat and GEOS altimeters, depending on water
temperature and salinity.
Now the relationship between the mean-square slope of the sea surface
and the wind speed has been known for some time, following interest in optical
glitter patterns (Cox & Munk (1954)). The empirical relation is
front at 6.25 ns spacing, the Seasat altimeter pulse was only 3.2 ns wide and
there were 60 gates at 3.125 ns spacing to sample the return pulse.
Several algorithms were developed to interpret the GEOS-3 data (reviewed
by Fedor et al. (1979)), and intercomparisons between these and synoptic sea
truth led to the following conclusions.
At low sea states, hy-i < 4 m, and the standard deviation
S.D.(;zi/3) < 0.75 m
At medium sea states 4.1 < hy-^ <8.0,
S.D.(/2i/3) < 0.5 m.
Insufficient sea truth was available for calibration at high sea states.
These results required averaging of up to 1600 pulses (over 16 s), and resulted in
a sea-state measurement averaged over 105 km length of ground track.
These achievements approach the limits of resolution of buoy and ship
measurements of waves, and the Seasat results turned out to be even better.
With a short lifetime of three months, there was not a great deal of synoptic
sea truth available for Seasat calibration tests. Fedor & Brown (1982) made com¬
parison between 51 buoy measurements and nearby Seasat overflights off the
North Atlantic and North Pacific coasts of the USA and in the Gulf of Mexico,
‘nearby’ in this context being within 80 km and hours of coincidence. There
was a mean difference of 0.07 m and a standard deviation of 0.29 m within the
available range of observed wave height, 0.5 m < hy^ < 5 m, when an optimum
algorithm was used. The comparisons are shown in Fig. 11.4. The on-board
processor tended to give wave heights biased 0.5 m higher for /z 1/3 > 2 m.
There was a fortunate coincidence between the lifespan of Seasat and the
Joint Air-Sea Interaction experiment in the North-East Atlantic. Webb (1981)
made comparisons between the altimeter measurement and a pitch—roll buoy
and found their ratio to be 0.96 ± 0.04, with a standard deviation of 0.1. The
eight samples were all at low wave heights, less than 2—3 m. A much larger
dataset of visual observations of wave height was compared by Queffeulou et al.
(1981), and although a standard deviation of 87 cm resulted, this was probably
due more to the inadequacies of the ship-observed data. No significant bias was
noted, even though wave heights up to 8 m were sampled. The overall conclusion
to be drawn is that the altimeter was able to measure significant wave height to
a precision and accuracy at least as good as in situ instruments, with the
added advantage for many applications of supplying an areal average. To achieve
the low standard deviations quoted above, the wave height was obtained from
a 10-second average of 10 000 pulses, corresponding to a ground-track length
of 67 km. With such a short lifetime, exhaustive confirmation and fine-tuning
improvements of the wave height algorithm could not be achieved, but there is
no doubt that the radar altimeter is proven as a powerful and accurate means
of measuring significant wave height.
Sec. 11.4] Perfonnance of wind-speed algorithms 331
Fig. 11.4 — Scatter plot comparing significant wave height estimates from the
NOAA buoy network and Ocean Weather Station PAPA with those inferred from
the Seasat altimeter using the Fedor algorithm. (After Fedor & Brown (1982),
copyrighted by the American Geophysical Union)
Fig. 11.5 - Scatter plot comparing buoy-measured wind speeds with those
inferred from Seasat Altimeter measurements using the Brown algorithm. (After
Fedor & Brown (1982), copyrighted by the American Geophysical Union)
earth at a suitable point in the orbit. Wave-climate data does not require a very
high frequency of coverage, and the repeat-orbit frequencies of 3 to 10 days
suggested in Chapter 9 as being appropriate for altimetry can provide adequate
sampling for most purposes. What is important is that the data should continue
to be available over a period of years. With a microwave sensor, the weather does
not interrupt the data series, but unfortunately the failure of Seasat cut short
the database that was beginning to be established. The task will be continued
when the next satellite altimeter is launched.
The results from Seasat provide an enlightening foretaste of the sort of
results which will be achieved with future satellites. Chelton, Hussey, & Parke
(1981) combined all the available data into an average wave-height map for the
world during July, August, and September 1978 (Fig. 11.6a), and also produced a
wind-speed map (Fig. 11.6b), from the altimeter data. Given the sparse coverage
of the southern hemisphere by shipping routes, and because the ships which
are present tend to avoid storms, this map must be the first realistic indication
of storm-wave conditions in the southern winter, albeit for only one winter
season. Waves with an average height of over 5.5 m can be seen to occur in
the south Indian Ocean over a large region. High waves occur as well in the path
of the East African Jet wind system, in the Arabian Sea. These and many
other interesting features of the wave-climate map open up exciting new areas
for oceanographers to study. It is truly remarkable that without ever having
to venture near these inhospitable areas of rough sea we are able to chart them
with all the confidence that comes from the successfully-tested wave-height
algorithms.
Mognard et al. (1983) have taken a closer look at the southern hemisphere
by plotting monthly maps of wave height and wind speed. Figs 11.7 and 11.8.
These clearly show how a region of maximum wind and waves is found in the
South Atlantic in July, but migrates eastward to the Indian Ocean and across
to the South Pacific during August and September. It is interesting to note how
the change to a 3-day repeat orbit with less dense spatial sampling has resulted
in smoother contours for September, since less information is contained in the
map.
Because wind and wave height are measured together, it is possible to infer
the sweU distribution from the satellite data using the following procedure. The
wave energy is calculated from wave-height data. The wave energy due to local
wind generation is also calculated from wave-prediction models, using the
satellite-measured wind field, assuming that a fully-developed sea has arisen in
equilibrium with the wind forcing. When this is subtracted from the observed
wave energy the residual energy must be that which has propagated into the
region from outside, that is, sweU-wave energy. Hence the swell height can be
mapped as in Fig. 11.9. This is a minimum estimate, since the locally wind-
driven wave energy may have been overestimated by assuming a fully-developed
sea.
334 The altimeter as a surface-roughness sensor [Ch. 11
(b)
Fig. 11.6 - Seasat altimeter data for 7 July-10 October 1978. (a) Wave heights;
(b) Scalar winds. (From Chelton, Hussey & Parke (1981))
Sec. 11.5]
SIGNIFICANT WAVE HEIGHT (m), SIGNIFICANT WAVE HEIGHT (m), SIGNIFICANT WAVE HEIGHT (m)
Applications of satellite-altimeter-derived wave height
Fig. 11.7 - Seasat altimeter-derived maps of significant wave height in the Southern Ocean for (a) July, (b) August, and (c) September-Octo-
ber 1978. Contours are at 1 m intervals, and the shaded region is over 5 m. (From Mognard et al. (1983), published by the American Geo¬
335
physical Union)
336
The altimeter as a surface-roughness sensor
Fig. 11.8 - Sesat altimeter-derived maps of wind speed over the Southern Ocean for (a) July, (b) August, and (c) September - October 1978.
Contour intervals are 2 ms"'. The shaded region is greater than 10 m s"‘. (From Mognard et al. (1983), published by the American Geo¬
physical Union)
[Ch. 11
Se c. 11.5 ]
Applications of satellite - altimeter - derived wave height
Fig. 11.9 - Seasat altimeter-derived maps of swell height in the Southern Ocean for (a) July, (b) August, and (c) September-October 1978.
Contour intervals are 1 m. Heights greater than 4 m are shaded. (From Mognard et al. (1983), published by the American Geophysical Union)
337
338 The altimeter as a surface-roughness sensor [Ch. 11
Mognard (1983) has extended this interesting method for estimating the
swell climate to cover the whole world for the Seasat lifetime (Fig. 11.10).
Furthermore, for the time when the 3-day repeat orbit was in operation,
3-day maps of swell height have been produced for the North Atlantic Ocean
(Fig. 11.11). From these it is possible in conjunction with a knowledge of storm
centres based on the wind- and wave-height data, to study the way in which the
swell propagates out from storms. The promise of similar interesting and use¬
ful oceanographic data from further altimeters suggests that these methods can
be applied in the future to give us an understanding of the continuously-evolving
wave climate.
It is probable that oceanographers will also be able to use repeated mapping
of ^1/3 to test hypotheses concerning the generation, propagation, and decay of
ocean waves, in order to be able to improve swell and wave forecasts based on
wind forecasts. Queffeulou et al. (1981) provide an example of this approach,
by comparing the satellite-observed wave height with that predicted by a
wave-forecast model over the North Atlantic.
Fig. 11.10 - Field of mean significant swell heights in metres deduced from Seasat altimeter measurements for the period 7 July -10 October
1978 (From Mognard (1983))
340 The altimeter as a surface-roughness sensor [Ch.ll
Fig. 11.11
Sec. 11.5 ] Applications of satellite-altimeter-derived wave height 341
Fig. 11.11
342 The altimeter as a surface-roughness sensor [Ch.ll
Fig. 11.11
Sec. 11.5] Applications of satellite-altimeter-derived wave height 343
Fig. 11.11 - North Atlantic swell maps deduced from the Seasat altimeter
measurements for the periods (a) Sep. 19—21, (b) Sep. 22—24, (c) Sep. 25—27,
(d) Sep. 28-30, (e) Oct. 1-3, (f) Oct. 4-6, and (g) Oct. 7-9, 1978. (From
Mognard (1983))
CHAPTER 12
from visible and infrared wavelengths to microwaves which can penetrate the
cloud and give all-weather views of the ocean. This has been the rationale behind
the development of imaging radars, and in particular the synthetic aperture radar
(SAR) which has already proved capable of spatial resolution down to 25 m.
An active radar device is necessary to obtain such resolution, and therefore
what is actually being imaged is primarily the roughness of the sea surface.
Assuming a Bragg resonance type of backscattering process, given oblique
surface viewing, it is sea-surface perturbations on a length scale comparable to
the radar wavelength which appear as roughness on radar images. Herein lies the
excitement of working with SAR images of the ocean surface. As mentioned in
Chapter 10, roughness of the sea surface is not a concept with which oceano¬
graphers have much experience, except in the context of ocean waves, and even
there little attention has hitherto been paid to the very short wavelengths which
produce radar roughness. Yet the SAR images which were obtained first from
aircraft, then from Seasat and latterly from the Shuttle Imaging Radar (SIR),
have revealed not only surface waves and swell but other oceanographic processes
which have a (previously unrecognised) surface-roughness signature.
It is beyond the scope of this book to go into great detail concerning
instrument design, but whilst it has been possible to discuss radiometers or even
the altimeter in terms of the broad concepts of their operation, without being
concerned about instrumental operating details, it is necessary to describe the
SAR more thoroughly. This is because the imaging mechanism and its interaction
with oceanographic parameters is closely related to the engineering design of the
instrument. It is helpful to split up the SAR imaging mechanism into three
distinct, though interrelated, processes. The first is the fundamental principle
of the synthetic aperture radar, as applied to stationary targets such as the
land surface. Second is the interaction between the basic SAR operation and
a surface such as the sea which is itself moving. This becomes closely related to
the question of how waves are imaged by the SAR. Third is the problem of
understanding how other oceanographic phenomena are capable of being imaged
by SAR through their influence on the surface roughness and movement. All
three aspects of SAR imaging are discussed in this chapter. It is important
for the oceanographer to understand fundamentally how the radar operates even
if he need not concern himself with the hardware details, and it is equally
necessary for the radar scientist to appreciate how oceanographic processes come
to have a SAR signature, if both are to develop the use of SAR as the powerful
oceanographic tool which it is capable of becoming.
known as squint mode. The analysis is rather more complex for squint mode,
but the principle is the same. The distance from the satellite to the ground in the
viewing direction is called the slant range. Its projection onto the sea surface
is called the ground range. The direction parallel to the satellite track is known
as the azimuth direction.
As (12.1)
2sin0 ’
6 being the angle between the slant-range direction and the vertical.
The resolution of a normal antenna is dictated by its aperture width. Dr.
Standard antenna theory gives the beamwidth (Fig. 12.2) as (3 = Xr/Dr in the
Sec. 12.2] Principles of SAR operation 347
far-field away from the antenna. Thus the resolution at range i? is p = /?/XZ)r,
in the plane normal to the range direction. For oblique viewing this becomes
larger in the ground range direction. Typically, radars will operate with a wave¬
length between 50 mm and 1 m. Even with a low satellite orbit having a range R
of about 750 km, a ground resolution of 100 m will require an antenna aperture
of between 375 m and 7.5 km, depending on the wavelength chosen! This is
clearly impracticable with present space technology, and an alternative approach
must be sought which makes use of the phase of the coherent signal to carry
more information and allows a smaller antenna to be used.
a very short space of time when a short pulse of very high frequency and high
power occurs. As this is a linear process, this filter can be applied after the chirp
has been reflected from the ground and received at the antenna, to create the
signal which would have been returned if the short pulse itself had been emitted.
This process in the SAR context is termed range compression of the signal.
fraquancy
From this point of view, range resolution depends on the ability to sample
sufficiently often for the frequency shifts not to be aliased by the sampling rate.
The resolution ceU is much smaller than (which would be the resolu¬
tion if pulse compression were not performed), because points such as C in
Fig. 12.5 can be identified clearly by a separate reference signal, centred at an
appropriate time in the return signal time series.
satellite
Fig. 12.6 — Definition of range cells, neglecting curvature of constant range lines
aperture is constructive. The same is not true of a point such as ‘d’, which is
much closer to one end of the aperture than the other. The change of phase
across the aperture of the reflection from ‘d’ therefore causes destructive inter¬
ference when the signal is integrated, and consequently does not contribute to
the received signal. Points ‘b’ and ‘c’ may or may not contribute to the signal,
depending on the phase geometry. This governs the effective beam-width and
hence the resolution. The larger the aperture, the closer to ‘a’is the point from
which the reflected signal ceases to contribute to the received signal because of a
sufficiently large phase variation between the two ends of the aperture.
Fig. 12.7 — Phase pattern of radar waves reflected from points opposite the centre
(upper diagram) and extremity (lower diagram) of a radar antenna. Note the large
phase change along the antenna in the latter case
The synthetic aperture uses this principle, but instead of integrating along
the length of a long antenna at a single time the reflections from all the targets of
one single pulse, it builds up this result gradually. Thus as it passes‘t’ (Fig. 12.8)
it emits a pulse and receives the reflection from ‘a’, ‘b’, ‘c’,‘d’,‘e’, etc. It does
the same as it passes ‘u’, ‘v’, ‘w’, etc., and stores the results. Finally, at the data-
processing stage, the reflected signal from ‘a’ received at‘t’, ‘u’, ‘v’, etc. is added
up to synthesise the same effect as the automatic integration along a very wide
real-aperture antenna. The same is done for the reflected signals from‘bVc’,‘d’,
etc., so that the result is a measure of the different values of radar backscatter
cross section at ‘a’, ‘b’, ‘c’, etc.
Sec.12.2] Principles of SAR operation 353
satellite ground
scattering
track
This begs the question of how it is possible to preserve separately the phase
information of the contributions from individual points ‘a’, ‘b’, ‘c’, etc. to the
signal received when the sensor is at a particular point‘t’. The key to this lies in
the steady motion of the radar itself in the azimuth direction, which produces
a doppler shift in the frequency of the returned signal. Assuming the targets
are all stationary, the target which is broadside on to the satellite will reflect the
incident signal with no doppler shift. The targets forward of the satellite have a
relative velocity towards the satellite, and will tend to increase the frequency of
the reflected signal, imparting a positive doppler shift. Those targets already
passed by the satellite will produce a negative doppler shift. For example, when
the satellite is at ‘w’, the signal from ‘d’ will be at the emitted frequency, the
signal from ‘e’ will be at a slightly higher frequency, and those from ‘f ’ and ‘g’
higher stiU, because the relative velocity between target and satellite has a bigger
component parallel to the radio wave travel directions at ‘g’, than at ‘f ’ or ‘e’.
At ‘d’, the relative velocity is zero, whilst for ‘c’, ‘b’, and ‘a’ it is in the opposite
direction and there is a frequency reduction in the reflected signal. Consequently,
as the radar approaches a target, the contribution of that point to the reflected
signal is first at a higher frequency than the emitted signal and gradually reduces
in a linear way till it is less than the emitted signal when the satellite has passed
by. At any given time, for a given range, the contributions from each point along
the azimuth at that range can be identified uniquely by its doppler-shifted
frequency. This enables the summation of different contributions from different
target points to be performed in the aperture synthesis.
354 Synthetic aperture radar [Ch. 12
Fig. 12.9 - Frequency of signal received after reflection at ‘d’ (as defined in
Fig. 12.8) while sensor travels from position ‘t’ to ‘z’
Hence p^. = •Or/2, which is at first sight a surprising result. The resolution
is independent of the range and of the radar wavelength, and improves with
a smaller aperture! The reason is that the construction of the synthetic aperture
can proceed, and information about a target can be gained, so long as the target
is illuminated by the real beam. The wider the real beam is (by having a smaller
real aperture) the more information can be communicated about a particular
target. An increase of R increases the time a target is illuminated, and this
offsets the deleterious effects of a more distant target. The reduction in size
of aperture Dr is limited by the need to generate a sufficiently powerful signal,
but clearly a very high resolution is theoretically possible.
For the azimuth-compression process to work, there is a limitation imposed
on the spacing of the successive pulses. The temporal sampling rate must be at
least twice the highest doppler frequency shift in the return signal, if the doppler
history of each point is to be unambiguously recognisable without aliasing in the
processing stage. This means there must be at least one pulse in the time it takes
the real aperture to travel half its length. On the other hand the pulse-repetition
frequency (PRF) cannot be too great, bearing in mind that for a reasonably wide
swath in the ground-range direction the returned pulse cannot be very short,
and pulses must not overlap. The constraint is therefore.
2V V c
— - — <PRF<- (12.2)
Dr Pa 21Psin0
where W is the swath width and V is the platform velocity, again ignoring the
effect of earth and orbit curvature for simplicity. Eqn (12.2) is a severe constraint
on the design of practical SARs, and requires a compromise between improved
azimuth resolution and a wider swath width.
356 Synthetic aperture radar [Ch.l2
___j
Si{t) = ai(xi, ri)exp 1 (12.3)
1
/CO
c
1
where Oy (xi, ri) is the complex amplitude of the echo, containing the phase
and am.plitude of the reflection and carrying information about the backscatter
cross-section at (xi, ry) which is the desired end product, co is the angular
frequency of the radar, and so cj = 2jrc/\R. Now,
The
English
Channel
Sj
’■^OTV
358 Synthetic aperture radar [Ch. 12
Azimuth image shift occurs because of the doppler shift introduced into
the reflected signal from a target moving in the range direction. The shift is too
small to be detected by the range-compression process, but becomes apparent in
the azimuth-compression process. It means that every reflection from a target
moving away from the sensor with speed Wr has a frequency reduction of
2mr/Xr in addition to the linearly-decreasing frequency shift due to the
satellite’s motion. In the aperture-synthesis procedure, this results in the target
not appearing in its proper place, but since it has the same doppler history as a
point encountered earlier by the satellite, the location of the moving target is
shifted in the negative azimuth direction. Similarly, a target moving towards the
360 Synthetic aperture radar [Ch. 12
Ax J/2
PaV^
(12.7)
RV„
This can give rise to interesting image phenomena when viewing over land.
Trains or road vehicles moving in the range direction appear to be shifted to one
side of the track or road on which they are travelling. It is even apparent in
ocean imaging. Fig. 12.14 shows a ship radar reflection clearly displaced from its
wake and its wave pattern.
Range walk occurs when Mr is sufficiently large to cause the target to pass
through one or more range-resolution cells within the time taken for the beam-
width to pass over it (that is, within the time the target is contributing to
aperture synthesis). The consequence is that full correlation will not occur for
that target within one range cell. Not only will its presence be smeared across the
range cells it has entered, as with an ordinary optical image, but the reduced
correlation within one range cell will cause it to smear in azimuth also. In terms
of the ground-range resolution cell size Pq, and the azimuth resolution p^, we
note that the integration time of a SAR designed to achieve pj^^ is
\R XR
In this time the target must move a distance less than Pq, that is.
2PgPa^
I WrI < (12.8)
XR
for range walk to be avoided.
Amplitude reduction follows from the above interactions reducing the
strength of the signal returned to the satellite, and this further reduces the
ability of the SAR to detect moving targets.
Target movement in the azimuth direction does not have such a striking
effect, but contributes further to image degradation. The effect of is to
change the relative velocity between the satellite and the target, and this there-
Fig. 12.14 - Seasat SAR image in the English Channel, showing a ship image
(white) displaced from the image of the wake by the effect of its own ipotion on
the SAR processing. Note both the dark ship wake (a smoother surface) and the
characteristic ship-wave pattern. (Reproduced from Raney (1983) by permission
of Marconi Ltd)
fore changes the rate at which the reflected radar frequency decreases as the
target is passed by the satellite. If positive, the frequency decreases more
slowly, and if negative it changes more rapidly. In either case the doppler
history of the reflections from the target are not the same as the reference signal,
and correlation is impaired. The result is termed azimuth defocusing, and the
consequence after processing is for the image to be blurred. An appropriate
criterion for this to be negligible is
2VPa^
«A<
(12.9)
362 Synthetic aperture radar [Ch.l2
Exactly the same effect occurs if there is steady acceleration in the range
direction — as well as a straightforward azimuth shift, defocusing also occurs.
It is clear that the earth’s own rotation must introduce errors due to the above
target motions, but since these are regular, and uniform over an image, it is
possible to make allowances for them, principally by making adjustments to
the reference signal against which the data is correlated. The reference signal can
in fact be thought of as the response of a perfect stationary reflector located at
the point on the earth’s surface whose radar cross-section is being investigated.
A more detailed discussion of the problems of moving targets will be found in
Tomiyasu (1978).
surface roughness by this means, but there is no clear evidence of this yet
available. The subject can be examined further in Barber (1983).
Concerning the hypothesis that it is short gravity waves which produce the
roughness that backscatters the radar waves through Bragg resonance, it is
necessary to assume that the roughness field does not substantially change within
the integration time of aperture synthesis. This is typically between 0.1 and
4 seconds, depending on the radar frequency. It does not matter that several
individual surface ripple troughs and crests may pass a given point in this time,
since what the radar responds to is the presence of a given degree of roughness
within a particular pixel. What is important is that the interaction time of the
whole field of short gravity ripples should be greater than the integration time,
that is, that the ripple field as a whole does not decay, or is not generated, within
the integration time. Tucker (1983) points out that the viscous decay time of
10 cm wavelength waves is around 80 s, of 4 cm waves it is 13 s, of 2 cm waves
it is 3 s, and for 1 cm waves 0.8 s. Thus, provided that the radar requires a Bragg
resonance wavelength of 4 cm or more, the ripples should persist at a nearly
constant ampHtude during aperture synthesis. Under these conditions SAR
images of the sea surface can be regarded as capable of recording meaningful
information about surface roughness. Meanwhile the long swell waves being
imaged have a period of typically 8-20 s, and so although their orbital motions
may influence the SAR, so long as the integration time is around 1 s it is reason¬
able to assume that their troughs and crests stay approximately in the same pbcel
during this time.
There are three distinct types of wave-imaging mechanism proposed:
hydrodynamic interactions, electromagnetic interactions (sometimes called tilt
modulation), and motion effects (velocity bunching and azimuth image smear).
The first of these could be viewed by radar, visible light, or any other method
of observing surface roughness. The second is related specifically to side-looking
radar and Bragg resonance backscattering, but it is not unique to synthetic
aperture radars. The third is intimately related to the aperture synthesis
procedure. We shall consider each process separately in order to gain some
insight into the different mechanisms, but an adequate theory must eventually
consider all the different mechanisms together.
The foundation of any theoretical model is to describe the swell-wave modu¬
lation of the radar backscattering cross-section o in terms of a complex variable,
the Modulation Transfer Function R(ks), where kj is the ocean wavenumber
vector. Thus
(12.10)
Oo is the background value of a in the absence of the waves. The wave-modula¬
tion effect is represented by the product of /?(ks) and Z(ks) which is the
Fourier transform of the surface elevation associated with the longer waves. The
role of the modulation transfer function is simply illustrated in the case of a
Sec. 12.3] SAR imaging of ocean waves 365
case of a SAR, waves propagating in the azimuth direction should not be imaged
by tilt modulation.
Fig. 12.17 — The variation with incidence angle of the dimensionless tilt-modula¬
tion transfer function, for X-band radar, for horizontal (HH) and vertical (W)
polarisation. (After Alpers (1983))
368 Synthetic aperture radar [Ch. 12
Fig. 12.18 — Motion of Bragg scattering ripples on the surface of swell waves due
to (a) the phase speed of the ripples relative to the surface, (b) advection by the
surface component of the swell wave orbital velocity, and (c) the vertical rise
and fall of the swell-wave surface
The effect on the image of target velocity and acceleration in the range
direction has been discussed in section 12.2.5. Target velocity causes azimuth-
image shift. The rising face of a wave is therefore shifted on the image in the
positive azimuth direction, and the falling face in the negative azimuth direction.
For swell waves where troughs and crests lie parallel to the azimuth direction,
this has no imaging effect, but if the direction of troughs and crests has a
component parallel to the ground range direction (that is, propagating in the
azimuth direction), then Fig. 12.19 illustrates how the waves can be imaged. For
the case shown where the waves are propagating in the positive azimuth direc¬
tion, there is a concentration of scattering facets, and hence a brighter image,
above the wave troughs, and a reduction over the crests. For waves propagating
in the negative azimuth direction, the bright and dark on the image would be
Sec.12.3] SAR imaging of ocean waves 369
Fig. 12.19 - Azimuth image shift (velocity bunching) and azimuth-image smear
due to the motion of a single-frequency wave train. (After Alpers (1983))
AZIMUTH DIRECTION
--WAVE DIRECTION
Fig. 12.20 — Nonlinear azimuth image shift. Note that the sequence of points
along the image plane is now different from that along the ocean plane, in
contrast to the linear process shown in Fig. 12.19
R du,
^vb ~
Vdxo'
where Uj is the average velocity of surface facets in the range direction during the
SAR integration time, and xq is the azimuthal coordinate in the ocean-surface
plane (N.B. for waves propagating in the range direction, Cyb = 0). Alpers &
Rufenach (1979) suggest that the velocity-bunching mapping is linear only so
long as I Cvbl < 0.3.
The implications of this for the minimum amplitude fg of swell of a certain
wavelength which will result in nonlinear velocity-bunching are illustrated in
Fig. 12.21. fs has been scaled by a geometric factor
where <f) is the angle between the swell wavenumber vector and the azimuth
direction, and 9 is the radar-viewing zenith angle (as in Fig. 10.3). For
azimuthally-propagating waves, this simplifies to a factor of cos0. For satellite-
borne SARs, R/V is large, and there is clearly a severe limitation on the size of
waves which can be imaged by velocity bunching. In wave-fields with a broad
wavenumber spectrum, nonlinear velocity bunching tends to shift the spectral
peak towards lower azimuthal wavenumbers, because it is able to image better
the longer waves travelling in an azimuthal direction. It is also implied that if the
linearity criterion is not met, imaging by other types of modulation will also be
severely degraded.
128
(SEASAT)
SWELL WAVELENGTH.m
Another motion effect arises from the acceleration of the surface during the
integration time, which produces azimuth image smear. Fig. 12.19 illustrates the
relative amount of smear for different facets, governed by the vertical accelera-
Sec. 12.3] SAR imaging of ocean waves 371
tion field of the swell wave. Whilst it is principally a degrading effect, the
correlation between smear and the wave phase implies that a suitable refocusing
of the aperture-synthesis process could potentially enhance the image to improve
the wave-phase contrasts of light and dark tones. For Seasat, azimuth smear had
an influence on image quality comparable with velocity bunching, but for a SAR
with a shorter integration time its effect is significantly reduced.
The parameter values for the Seasat-SAR are given in Table 12.1. It is
instructive to discover what these values imply for some of the dependent
quantities discussed already in this chapter. The given view angle of around 20°
and the radar wavelength of 23.5 cm lead to Bragg resonance with waves around
30 cm wavelength, normally described as ripples, but not to be confused with
capillary waves, since 30 cm waves are strongly gravity dependent and are
influenced very little by surface-tension effects.
Inspection wiU show that the pulse length, R.F. bandwidth, pulse-repetition
frequency, resolution ceU size, and swath width satisfy the interrelated
constraints discussed in section 12.2.3, notably (12.2). This is illustrated in
Fig. 12.22. Here it is revealed that the SAR data-recording system must cope
with an added complication. To achieve a pulse-repetition rate so that two pulses
are emitted in the time for the the satellite to travel half its aperture length,
eight pulses must be emitted before the return from the first pulse is received
back at the sensor. The slant range has to be matched to the pulse-repetition
rate so that the emitted and return pulses fit together efficiently.
With a real aperture of about 11m, the SAR is theoretically capable of an
azimuth resolution of 514 6 m. In practice this becomes 25 m when four
independent looks are used to reduce speckle.
"O
ctj
(D
O O
00 ’
Week 4: JuM7-23
XO
eg
13 1
<u
Q
H
o
^ ft
Week 3: Jul 10-17 v
-C r
a^
.22 -o
c C
S
a:
z .t:
o >-•
o
C/3 C
cu eg
s
c ^
c
o ^
a S'
.
a> Z
eg
Week 2: Jul 3-9
^ > >
eg r )
*J w fl.
eg
*13
OC
<
bO
ft
C/2 W
a:
Station coverage
o P.
op 5 <wc
eg 00
<D fteg ^
a(
>
o A (U-e
o
(D o
0) cd-c
a ’k‘^
E cdfj
o ft
U O
eg tu.
bo^Ci
rs
„
ft ft1
cgie
• a> H**
Sf o
2 li-
Sec.12.4] Observations of ocean waves with Seasat SAR 375
The SAR operated only when data could be transmitted in real time to one
of five receiving stations at Oakhanger, England; Shoe Cove, Newfoundland;
Merritt Island, Florida; Goldstone, California; and Fairbanks, Alaska. Fig. 12.23
shows the coverage of these stations in the Northern Hemisphere, and also
indicates the ground tracks of all the swaths collected in nearly 500 separate
passes during the 98 days between 4 July and 10 October 1978 on which the
SAR was operated. More detailed descriptions of the SAR data coverage and
examples of the imagery can be found in Beal et al. (1981), Fu & Holt (1982),
and Vesecky & Stewart (1982).
(a) (b)
Fig. 12.26 - Comparison of dominant ocean-wave length (a), and direction (b),
as measured by Seasat SAR and surface pitch-roU buoys. (After Vesecky &
Stewart (1982) copyrighted by the American Geophysical Union)
Fig. 12.27 - SAR image of the Pacific Ocean, 400 km southwest of Vancouver
Island (Seasat revolution 1349), 29 Sep. 1978. Swell waves show up in the brighter
patches where the wind of about 2.5 m s”' is able to roughen the surface suffi¬
ciently to allow the modulation mechanism to occur. Elsewhere the sea is smooth
and appears dark to the SAR. The direction of radar illumination is towards the
top of the image, which is W.N.W.
Fig. 12.28 - Swell-wave pattern around the Island of Foula, 70 km north west of
Fair Isle in the N.E. Atlantic. The image area is about 30 km square
directional information, and the ease with which a SAR image can be Fourier
analysed digitally and optically offers a means of readily obtaining directional
wave spectra.
Fig. 12.29 is an example of the image wavenumber spectrum obtainable
from SAR. It is a polar plot, indicating wave direction, with wavenumber increas¬
ing radially (that is, the longest waves are found closer to the centre). This
particular example shows that there are two distinct wave trains present, since
two distinct peaks of image variability can be detected, one corresponding to
waves of length 177 m in a direction of 290°, and the other, smaller peak cor-
Spacecraft
velocity vector
[!□ IZD □□ ES H
0-2 2-4 4-6 6-8 8-10
Imaye enerijy
density (linear scalel
responding to waves of length 93 m and direction 270°. We note that the polar
plot is symmetrical, and there is a 180° ambiguity in direction. Generally this
is easily resolved by assuming that waves are travelling in the direction of the
recent surface wind, if known, and are most likely to be travelling from the open
ocean towards the coast, if there is a coastline nearby. In this case, just north¬
east of Cape Hatteras on the U.S. east coast, there was meteorological and
oceangraphic evidence (Beal (1981)) to confirm that the SAR spectrum was
representing the principal features of the actual wave field.
However, there are serious shortcomings in obtaining wave spectra from the
SAR, and much research needs to be performed before SAR-measured spectra
can be confidently used for oceanographic interpretation without supporting
sea measurements. It is tempting to assume that a directional spectrum such as
Fig. 12.29 is the same as a directional wave-energy spectrum, but there remain
questions concerning all aspects of the information content: wavelength, wave
direction, and wave amplitude. This is to be expected, considering the complexi¬
ties of the SAR wave-imaging process which have already been discussed. The
spatial resolution of the SAR severely limits the high-wavenumber resolution,
whilst it is still not clear whether the SAR spectrum should correspond to the
wave-height or the wave-slope spectrum, since the different imaging mechanisms
rely on both to varying extents. Fig. 12.30 shows a comparison between an
omnidirectional SAR image spectrum and wave-slope and wave-height spectra
obtained from a pitch-roll buoy measuring at the same time as a Seasat overflight
during the JASIN experiment (Vesecky& Stewart (1982)). At short wavelengths,
there is better correspondence between the SAR data and the wave slope, but in
the more important peak area of the spectra, we see that the SAR peak is shifted
to a longer wavelength than the wave-height peak, which is itself at longer wave¬
lengths than the slope spectrum. This accords with our general expectations that,
being able to image the longer waves more readily, the SAR spectrum is biased
to lower wavenumbers. However, the shape of the SAR spectrum is broader than
the ground-measured spectrum, which could well be the result of nonlinearity
in the modulation process shifting some spectral energy to higher wavenumbers,
and thus reducing the steepness of the spectral slope at wavenumbers above the
peak values.
The spectra displayed in Figs 12.29 and 12.30 were of wave systems
propagating primarily in the range direction. Are azimuth waves observed as
clearly in spectra? The work of Beal et al. (1983), examining the variation in
SAR image spectra along a single swath where the changing character of the
waves was known or could be readily assumed from the topography and weather,
reached the conclusion that as the waves turned from range to azimuth orienta¬
tion, the SAR was less able to measure them, and consequently the directional
spectra became distorted. Not only was there a bias towards lower wavenumbers
but also the spectral peaks were biased towards the range direction. Thus while
it is still possible to gain some useful oceanographic insights into the variability
Sec.12.4] Observations of ocean waves with Seasat SAR 383
of the wave field and its spectrum over the length scales of mesoscale dynamical
ocean processes, it is best to regard the information as qualitative rather than
quantitative. It requires further theoretical study of the wave-imaging process,
and further experience from synoptic satellite and surface-wave measurements,
before it will be possible to define a nonlinear transfer function which relates the
ocean spectra to the SAR image spectra.
The same is true if wave amplitude is to be determined from the amplitudes
of the image spectra. The problem is greater here because the density of the
image depends on the ripple roughness, and hence the local wind speed, as well
as the swell-wave amplitude. However, an appropriate measure of hij^ may be
obtained by comparing the height of the spectral peak to the low-wavenumber
noise. The peak-to-background ratio (PBR) is indicated as the signal-to-noise
jiatio on Fig. 12.30, and it appears to be possible to relate this empirically to the
significant wave height, to an accuracy of ±0.75 m.
WAVENUMBER,m
Fig. 12.30 - Comparison between a SAR image spectrum and the wave-slope and
wave-height spectra of coincident data from a pitch-roll buoy in the JASlN area,
4 Aug.1978. The SAR image and wave-slope spectra have been normalised so that
their peaks have the same height as the wave-height spectrum. (From Vesecky
& Stewart (1982), copyrighted by the American Geophysical Union)
(c) The wavelength between individual light and dark bands is typically between
several hundred metres and several kilometres, and usually decreases from the
leading wave in a packet to the trailing edge.
(d) The separate groups of waves are typically tens to a hundred kilometres
apart.
(e) The crests (or surface manifestations of a constant-phase line) are usually
tens to hundreds of kilometres long, and very often the lengths of crests (as
revealed on images) decrease towards the rear of the wave group.
(f) These waves appear either as dark in a light background (presumably under
rough-sea conditions), as light in a dark background (calmer conditions), or as
dark and light bands in the intermediate case, suggesting that internal waves can
be imaged over a broad range of wind conditions.
2
ghiki^{p2-Pi) (12.12)
COiw
Pi “b Pi ky^hi coth kiy^h2
if the lower layer is deep. If the lower layer is also shallow compared to the
wavelength this simplifies to:
ghih2iP2~ Pi)kiw (12.13)
COiw^ —
P2^1+ Pl^2
ghih2{P2~ Pi)km (12.14)
or <^iw — / ■*■
P2hi+ P\hi
when the frequency is too low for the Coriolis parameter / to be ignored (that
is, for wave periods of order half an hour or more).
These dispersion-relation models provide an approximate framework for
extracting useful dynamical information from the spatial measurements of
internal waves which can be made from SAR images, as discussed later in
section 12.5.4.
Fig. 12.32 - Packets of internal waves revealed by a SAR image off the coast of
Portugal. Some correlation can be seen between the internal wave-crest directions
and the bottom contours shown. Swell waves are also visible at a smaller scale,
and brighter patches occur where they break near the beach. (From Allan (1983))
388 Synthetic aperture radar [Ch. 12
INTERNAL WAVE
DIRECTION
surface convergence slicK
\
For the simplest case (12.13), the waves are non dispersive, and the phase
and group velocities are both equal to:
ghih2{p2
Cm — Qw “ (12.15)
P2 ^1+ Pl^2
The surface velocity (that is, the wave-orbital velocity) associated with an
internal wave, at a point where the thermocline is displaced a distance fiw above
its rest position is:
_ fiw<^iw
“iw - ^ (12.16)
hi
Whilst linear internal-wave theory goes some way towards explaining the
observed phenomena, if the waves are strong enough to generate a surface
signature, then it is likely that in many cases they are of such a large amplitude
as to be nonlinear. Because the density gradient along which they propagate may
be very small, very little energy is required to generate a large-amplitude wave.
In this case, the Korteweg de Vries equation must be solved, and an appropriate
solution is a solitary wave or a train of internal solitons (Osborne & Burch
(1980)). Fig. 12.34 illustrates the internal soliton. Its profile is given by the dis¬
placement f js of the thermocline above its mean level, in terms of the maximum
displacement fo below the mean level, as:
= -fosech^ (12.17)
fo(^2 hi)
The speed of the soliton is c - Ciw (12.18)
2 hih2
Sec.12.5] Internal waves imaged by SAR 389
where cjw is the speed of linear waves given in (12.15) and the characteristic
horizontal length scale of the solition given by
V2
4 hi hi
Lc —
Fig. 12.34 — An internal solitary wave, with fluid particle velocities shown by
the arrows
Although they are nonlinear solutions, solitons can pass through each other
without change, and a train of internal solitons can exist as shown in Fig. 12.35,
/ I \
Their spacing is discussed in detail by Osborne & Burch (1980), but the
significant result of the theory in connection with the SAR images is that the
leading soliton of a group must be of greatest amplitude, and subsequent solitons
390 Synthetic aperture radar [Ch.l2
become both smaller and more closely spaced. This behaviour resembles the
features observed on many of the images, where a packet of internal waves has
a gradual decrease in wavelength towards the back, and the front crests are
longer than the rear ones, suggesting they are probably of greater amplitude.
The wavelengths encountered in the packets imply that the period of the
internal wavelike motion is between a few minutes and an hour. The spacing
of the separate packets of waves can be accounted for by the hypothesis that
the waves are generated by tidal flow over topography. Very plausible theories
have been advanced by Lee & Beardsley (1974) and Maxworthy (1979). They
suggest that tidal flow over a topographic feature sets up a natural standing wave
of large amplitude. When the tidal flow slackens and reverses, the downward
displacement of the thermocline propagates away as a wave, and if it is of
sufficient amplitude, the nonlinear soliton theory predicts that it will split
into a train of solitons. A related suggestion by Osborne & Burch (1980) is that
an internal solitary wave may be produced by tidal interaction a long way away
and propagate into a region of shoaling topography where the new depth
constraints cause the solitary wave to split up into a train of solitons. Either
way, it is reasonable to suppose that the time interval between generation of
separate packets is the tidal period, normally 12]6h, but sometimes 24—25 h
where diurnal tides dominate.
vigorous in the case of a large-amplitude internal soliton, and the surface velocity
sufficiently large, as to generate surface ripples by turbulence. In this case any
wind-driven waves or swell already present, might steepen to cause breaking and
white-capping which would have an enhanced roughness signature. A graphic
description of this type of effect, which they describe as a ‘rip’, is given by
Osborne & Burch (1980).
Whilst possible mechanisms clearly exist for imaging internal motions by
surface roughness, there is still a need to find out more about the details of them,
such as the phase relation between the internal-wave crests and the rougher/
smoother patterns, or the wind conditions under which they will be visible, or
the ripple size which is generated.
Given the horizontally variable surface velocity (12.14), the possibility
arises of velocity bunching being an imaging mechanism. The wavelengths
are probably too long for this, and the internal waves have been observed both
in the range and the azimuth direction, but a complete study of the imaging
process will have to consider what effects the surface motions will have on the
SAR. Until the mechanism is fully understood, it remains reasonable to assume
a one-to-one correspondence between light and dark wavelike patterns on the
image, and internal wave crests and troughs, or vice versa. It appears that internal
waves are readily seen by satellite SAR because of the advantages of an elevated
viewing point from which to survey their spatial structure, and because of the
ability of radar to detect subtle changes in surface roughness more readily than
visible wavelength sensors can.
Fig. 12.36 - A portion of the image shown in Fig. 12.12. The letters correspond
to sandbanks; S.F., South Falls; S. Sandettie; O.R., Outer Ruytingen; W.D.,West
Dyck. The tidal current was flowing towards the south-west, with a speed at the
surface of around 0.6 to 1.0 m s"*. The wind speed was about 25 km/h ►
J,
y 4:>^-$f
Hti.y 3
^ ^ *■
4' **-c3v
€
«S^'*S^'^“I**
K--/ >
^I 4**'*
-A:'|jA3t.<’V
■H.
394 Synthetic aperture radar [Ch. 12]
Microwave scatterometers
13.1 INTRODUCTION
In this, the final chapter to be devoted to a particular type of satellite sensor, we
look at the radar scatterometer. This is the name usually given to an oblique-
viewing active microwave device, which measures the backscattered radar energy
from a fairly broad sea-surface area illuminated by a long pulse of energy
at a particular frequency. No account is taken of the timing, the phase, or
the coherency of the return signal in comparison with the emitted pulse. The
amplitude of the return signal is interpreted empirically as a measure of the sea-
surface roughness. Depending on the frequency of the radar, and hence the wave¬
length of the Bragg resonant surface roughness, as discussed in section 10.2.3,
the magnitude of the return can be related to either the surface-wind speed
and its related stress, or the surface-wave field, leading to the descriptions of
‘wind-scatterometers’ and ‘wave-scatterometers .
Most practical experience of satellite scatterometry to date has been gained
from the Seasat scatterometer, although a scatterometer was flown on Skylab
in 1973 and several aircraft experiments were flown to prove the scatterometer
concept’ prior to Seasat’s launch. The Seasat-A Satellite Scatterometer (SASS)
operated at a frequency of 14.6 GHz (wavelength approximately 2 cm). Thus
Bragg resonant backscattering occurred for waves of about 3 cm length. These
are surface capillary-gravity ripples which are sensitive to the surface-win
stress, and tend to line up with their crests perpendicular to the wind direction.
Experiments have shown that the backscattering cross-section measured at the
SASS wavelength is strongly correlated with the wind-stress vector, and y
comparing the backscatter for the same area of sea viewed from two different
directions at right angles to each other it is possible to deduce the wind direction.
Although surface-wind stress is the most useful parameter for oceanographers
and meteorologists (since it provides a boundary condition for models of ocean
and atmosphere), and although in principle the wind stress is more clo^se y
coupled to surface roughness than is the wind speed, it was m fact surface-
wind speed for which algorithms were developed to interpret the SASS da a.
The reason for this was that to produce a valid empirical algorithm to calibrate
396 Microwave scatterometers [Ch. 13
(13.2)
Sec.13.2] Wind scatterometiy 397
where Pa is the air density and Co.h is the dimensionless drag coefficient,
appropriate for calculating stress from a wind velocity measured at height h.
This gives an easily-calculated approximate estimate of since from many
observations it has been established that
for wind speed measured at 10 m. However, this does not cater for measurements
made at different heights, and does not allow for the variation of due to the
stability of the air column, which alters the wind-velocity profile. To cater for
this, the Monin-Obukhov equation is used to define the velocity profile;
U*
U{z) - Us (13.4)
K
Here Ug is the sea-surface velocity which is normally small compared to the wind
speed a few metres above. K is vonKarman’s constant (normally assigned a value
of 0.4). zq is a parameter known as the roughness height, which for a solid
surface is dictated by the height of surface-roughness perturbations, but is not
directly related to wave height because the sea surface is in motion. Over the sea,
an accepted empirical expression for zq is
0.0156 U*^
Zq = -m. (13.5)
g
i// is a correction to the logarithmic profile to allow for the stability of the atmos¬
phere. L is the Monin-Obukhov length, and represents the height above which
stability effects are important. In terms of the appropriate bulk coefficients, L
is given by
L =
KgAT
where 7^ is the atmospheric temperature at the same reference height as that
used for U and Cp. AT is the air-sea temperature difference, AT =
degK. Typically, L is between 30 and 100 m. The form of i// is debatable, as
indeed are many of the empirical formulae used in this whole field of study, but
typical form is,
Z—Zo z —Zo
for stable atmospheres, L> 0, i// = 4.7
L L
Z — Zq Z—Zq
for unstable atmospheres, L<0, \p = 3.8
L L
Thus to obtain U* accurately from a measured wind speed at a particular
height, it is necessary to use (13.4) with an estimate of Zq from (13.5), using an
anticipated value for U* which may need subsequent correction. Whilst the
forms of the equations used in air-sea boundary layer theory have a sound
398 Microwave scatterometers [Ch. 13
The fact that a fairly good fit can be achieved demonstrates that a scatterometer
should be capable of measuring wind stress. The steeper the gradient on the
graph (that is, the greater the exponent a), the more sensitive the radar is to
wind-speed variations. Thus the result shown in Fig. 13.2 (also based on data
Sec. 13.2] Wind scatterometry 399
Fig. 13.1 - The variation of the normalised radar backscatter cross-section with
the friction velocity m* of the wind over the sea surface, for a horizontally-
polarised 13.9 GHz scatterometer viewing at a zenith angle of 0 = 40° in the
upwind direction. The fitted line is
logio(a°)= -48.59027 + 19.9658 logioU*dB.
(Based on Jones & Schroeder (1978))
2 Q Frequency.QHz
A 13.9
V 13.3
• 8.9
oc ■ □ 4.6
♦ 1.3
■ 0.4
1.0 - ♦
»_ _1
oLjl-
0.01 0.1 1.0
Fig. 13.2 - Variation of the wind-speed exponent a with the Bragg resonant
wavenumber (1/ X®)* upwind viewing and vertical polari^tion. Different points
for the same radar frequency correspond to different viewing angles. (After Jones
& Schroeder (1978))
400 Microwave scatterometers [Ch. 13
-14 . a
e • • •
-18 + a •.K +
+ a e 4.+ *
cr\ dB 4 .• %
••4"
♦ 1-4 4
f 4
-22
a vertical
polarization
4- horizontal
-26
From these and other observations, which can be studied in more detail in
Jones & Schroeder (1978) and Moore & Fung (1979), the conclusion is drawn
that a radar at about 14 GHz, with a viewing-angle range centred at 40°, should
be the most satisfactory device for measuring the speed and direction of the
wind stress, and hence the wind speed.
Fig. 13.4 - The Seasat scatterometer illumination pattern and swath definition,
showing the constant doppler lines which define the resolution cells along the
beams
402 Microwave scatterometers [Ch. 13
antenna beam/
Fig. 13.5 — Instantaneous doppler cell defined by the antenna pattern beam
width and the doppler filter response. To the bottom right the resolution cell is
shown to be constructed by integrating the doppler-cell response over time whilst
the satellite (and hence the instantaneously defined cell) moves forward over the
ground
1 4V IV 3V 2V
2 4H IH 3H 2H
3 4V 4H 3V 3H
4 IV IH 2V 2H
5 4V 4V 3V 3V
6 IV IV 2V 2V
7 4H 4H 3H 3H
8 IH IH 2H 2H
404 Microwave scatterometers [Ch.l3
the same doppler filters were used for each beam. If the doppler patterns were
symmetrical, the 12 range cells in the main beam would lie at the same distance
from the satellite track for both fore and aft beams. However, the rotation of
the earth underneath the satellite orbit caused an extra doppler shift, different
for fore and aft views, which produced an asymmetry in the ground pattern
of constant doppler lines. The effect was most marked at the equator where the
earth’s surface speed is greatest. This resulted in a non-matching of cells from
the fore and aft beams, illustrated in Fig. 13.6, which had to be corrected for in
the data processing.
Fig. 13.6 - An example of the resolution cell locations for beams 1 and 2 on
Seasat revolution 331 over Hawaii. Only a few of the cells have been drawn in,
to make clear the differences between the forward- and rear-looking beams, due
to the asymmetric doppler effect of the earth’s rotation. The swath for beam 2
is about twice as wide as for beam 1 at this near-equatorial latitude
a° = (t/, X, e) (13.6)
in which the only variables affecting are the wind speed U, and direction x
relative to the pointing direction of the radar beam, the radar beam incidence 6
at the sea surface, and e the polarisation state (H —H or V—V) of the measure¬
ment. This assumption is based on the observational programme mentioned in
section 13.2.2, but the empirically-derived form of (13.6) is still being improved
as more data from land-based, airborne, and satellite scatterometers are gathered.
As the database is widened, it may prove necessary to include a dependence on
further variables, such as the wind fetch or the amplitude of the low-frequency
swell wave end of the gravity-wave spectrum.
The model function selected as the basis of the SASS wind-vector algorithm
was one of the form
logio(a°)- G
log 10 (13.8)
H
This cannot be solved by substituting for G and H from the empirical model
tabulations, because Xi and Xa are unknown, but it is known that:
X2 = Xi ~ Htt for viewing on the right of the satellite track and
X2 = Xi + Htt for viewing on the left.
Sec.13.3] Experience with the Seasat scatterometer 407
Fig. 13.7 — Backscatter as a function of incidence angle for the SASS-1 model
function for winds of 23.6 and 4.6 m s“'. The different responses are shown for
upwind (U), downwind (D) and crosswind (C) azimuths, at both horizontal and
vertical polarisations. (From Pierson (1983))
Instead, (13.8) can be plotted as a function of x for the two parameter sets
(a®i, di, ei) and (o°2, ^2. ^2) corresponding to the radar returns from the fore
and aft views of the same patch of sea surface. Fig. 13.8(a) shows typical curves
for the two functions If'i(x) and 1^2 (x) where x is the wind direction measured
clockwise from the radar-pointing azimuth (N.B. the azimuth directions for
and W2 differ by 90°). If, instead, the curves are plotted so that x is now
the wind direction relative to a common direction, in this case the azimuth
of the forward antenna (Fig. 13.8(b)), W\ remains the same but W2 is shifted
90° to the right (for radar beams to the right of the direction of travel). Since
both curves apply to the same small area of sea, the wind magnitude and direc¬
tion are identified by the intersection of the two curves, assuming that the wind
vector has not changed between the two views.
Unfortunately, the curves normally intersect four times, although occasion¬
ally there are only two intersections. The wind speed is approximately the same
for the four possible solutions, leaving four possible directions, the true one and
three aliases. In principle, it should be possible to remove these by adding two
further curves, corresponding to the response at the polarisation states opposite
to the ones already used. This is the purpose of operating modes 3 and 4.
408 Microwave scatterometers [Ch. 13
(b)
Fig. 13.8 — Typical curves of wind speed versus wind direction appropriate to
a particular radar return for a given incidence angle and polarization state.
corresponds to beam 1 and W2 to beam 2. (a) Each curve is plotted against the
wind direction measured relative to the azimuth of the radar beam in question,
(b) Each curve is plotted against the wind direction measured relative to a fixed,
common direction, in this case the azimuth of beam 1. The actual wind direction
must be found at one of the intersections of and W^, that is, x= 40°, 140°,
220°, or 320°. One of these is the true direction, and the other three are aliases
Fig. 13.9 shows three typical families of curves for one polarisation state (solid
lines) and the slightly different curves for the opposite polarisation (dashed
lines). The aliases are removed, leaving a unique solution. However, the curves
in Figs 13.8 and 13.9 assume that the measurements of 0° are noise free.
In practice the curves are not actually plotted, and the solution is not obtained
graphically, but a least-squares fitting method is used (described by Jones etal.
(1982)) which minimises a difference function based on (13.7), summed over all
the available doppler cells falling within the grid square being studied. Given the
noise in the data, the solution is not so clearly defined as the noise-free curves
imply, and it has not yet been possible to achieve a unique solution even when
using both polarisations. The least squares algorithm normally offers four
solutions, but the noise in the data increases the errors in determining the four
Sec.13.3] Experience with the Seasat scatterometer 409
candidate directions. Since the noise has the effect of shifting the curves of
Figs 13.8 and 13.9 up or down, it causes the worst wind errors in candidate
directions close to 0°, 90°, 180°, and 270°. Indeed, in some cases the noise could
result in the curves not crossing at all, leading to no solution being found. When
four solutions are offered, the interpretation and application of the SASS data
must begin with an initial analysis to determine which is the most likely of the
candidate directions to be correct.
Fig. 13.9 — Three examples of the same type of plot as Fig. 13.8(b), for x = 0°,
40°, and 80°. In this case, however, two extra curves are included (drawn as
dashed lines) corresponding to the opposite polarisation states on beams 1 and ^
to those represented by the solid lines. The four curves cointersect at only one
point, giving a unique wind speed and direction solution. (From Pierson (1983))
410 Microwave scatterometers [Ch. 13
Fig. 13.10 - An example of mapped SASS wind vectors (including aliases) over
the JASIN area. (From Wurtele et al. (1982), copyrighted by the American
Geophysical Union)
Sec.13.3] Experience with the Seasat scatterometer 411
question of determining the wind direction will remain until all the data are
analysed. They are presently available in two forms. The sensor data record
(SDR) contains the parameters a° 9, and e for given radar azimuths. The
geophysical data record (GDR) contains the wind speed and direction for all
possible solutions (the true and the aliased). A short period of data (about
two weeks) has had the aliases removed.
The most reliable method of removing the aliases, but the most laborious,
is to plot the wind streamlines by hand with reference to the surface-pressure
charts and wind observations from shipping, if available for the area. The data
can be plotted automatically in terms of the four wind-vector possibilities, as
illustrated in Fig. 13.10, but from there the process is subjective, and its success
depends on the skdl and experience of the analyst in the drawing of meteoro-
Fig. 13.11 - Schematic depicting the solutions produced by the SASS wind
algorithm corresponding to five possible surface-wind vectors, drawn in rela¬
tion to the SASS beam pattern on the ground. (From Wurtele et al. (1982),
copyrighted by the American Geophysical Union)
412
+
V
•f
+
+
•V
V +
+
-h
+
+
-V
4
\
/
-
r
■4
4r
4-
-f-
-h
-f-
•V
Microwave scatterometers
4
-v \ \
■v ■V
-F
■V
-
+
-V-
Fig. 13.12 - Typical SASS wind vectors (and aliases) corresponding to idealised surface wind-velocity patterns, (a) Cyclonic vortex (northern
hemisphere); (b) Col, or hyperbolic point separating pairs of cyclonic and anticyclonic circulations; (c) Convergence line; (d) Frontal zone.
[Ch. 13
logical charts. Wurtele et al. (1982) have rationalised so far as possible the
subjective choices and decisions which are made in the process. Although
the data are noisy, some help in interpretation can be gained from noting the
SASS wind vectors and aliases which are predicted from given wind vectors in
noise-free conditions. Fig. 13.11 shows these to be related in a systematic way to
the direction of the satellite. Winds at 45° to the track (that is, blowing along one
of the radar azimuths) produce only a single alias at 180° to the true direction,
whilst a wind in the satellite direction produces 3 aliases perpendicular to each
other and the true direction. Wurtele et al. (1982) also present the SASS
vector-field patterns which would be associated with idealised streamline
patterns typical of certain synoptic wind conditions. These are reproduced in
Fig. 13.12. As with any manual interpretation process, an experienced analyst
begins to recognise in the SASS vector plots the patterns associated with given
synoptic meteorological features. However, to achieve a successful interpretation
it is necessary to work with a total field which is at least as large as the typical
synoptic scales of the wind field. A single SASS overpass, as drawn in Fig. 13.10,
would by itself be insufficient to achieve a confident removal of directional
aliases.
414 [Ch. 13
SASS
Fig. 13.13 — Comparison of the Seasat scatterometer winds derived using the
Wentz algorithm-with the surface winds measured at sea in the JASIN experiment.
The latter are based on 60-minute means of automatically-logged ship and buoy
data. Each point represents the mean of aU comparisons on a given overpass. The
anomalous point 1 is referred to in the text. (From Guymer (1983 a))
Sec.13.4] Applications of wind scatterometry 415
Overall, it can be concluded that the SASS data should be able to yield
the surface wind to an accuracy of better than ±1.7 ms“^ in speed and ±17°
in direction, assuming that directional ambiguity has already been removed
successfully. These statistics are based on comparing datasets in which the wind
speed never exceeded 16 m s“\ However, there was opportunity to make a few
isolated comparisons with storms and gale force winds (Jones et al. (1982)),
and in these there was a tendency for the SASS to underestimate wind speeds.
This could be due to problems of poor atmospheric correction in the rain-bands
associated with the storms, inaccuracies in the spatial registration of cells which
show up because the storm-wind field changes over a relatively short length
scale, or inadequacies in the SASS model function for a® at high wind speeds.
Another means of checking the SASS speed data, but not the direction,
was against other Seasat wind-speed measurements from the altimeter and the
SMMR. Wentz, Cardone, & Fedor (1982) found a generally good agreement,
particularly for nadir viewing, and oblique viewing yielded a standard deviation
of only 1.42 m s"^ between SMMR and SASS predictions from 329 comparisons
for winds up to 20 m s"^ SMMR and SASS estimates of hurricane winds were
also in good agreement.
pass the problem by modelling the wind field, but in this case there is just as
much need to know the realistic wind-stress field in order to validate the models.
When surface-wind stress vectors can be prescribed over the ocean from satellite
scatterometers, the way will be open to use the models to explore dynamical
processes which occur on a variety of length and time scales. With global cover¬
age every few days, synoptic coverage over a swath hundreds of kilometres
wide, and the ability to continue supplying this data over a period of years,
the satellite scatterometer should be able to cope with most of the demands
for boundary data made on it by the modellers.
The types of process whose study will benefit from satellite-derived wind-
stress vectors range from localised upwelling to global ocean circulations.
Localised upwelling and downwelling are related to surface horizontal velocity
divergence or convergence. Fig. 13.14 shows schematically how this can be
influenced by the wind stress. If for simplicity we assume that the sea-surface
horizontal currents in the ocean are in geostrophic balance with the wind
stress, then if there is a shear in the wind stress, there will be a corresponding
convergence or divergence in the surface current. Continuity then demands
downwelling or upwelling. Hence the shear, or more precisely the curl, of the
IdTy dTx\
wind stress 1-1 should be directly proportional to the upwelling. The
\dx dyI
ability of a satellite scatterometer to provide the direction as well as the magni¬
tude of the wind stress means that the wind-stress curl can be accurately defined,
making it possible to study the effects of the resulting surface convergence/
divergence on localised mixed-layer and thermocline dynamics. This mechanism
is probably a generation mechanism for quasi-geostrophic wavelike motions
which propagate in the upper ocean and are controlled by both the Coriolis
force due to earth rotation and the buoyancy forces due to density stratification.
With length scales of tens to hundreds of kilometres and periods of tens to
hundreds of days, their modelling would benefit from a knowledge of the
varying wind-stress curl over such length and time scales.
On a larger length scale, the modelling of the major ocean currents requires
the input of wind data. Some currents are strongly wind-dependent. For
example, the Somali current in the Western Indian Ocean increases its flow
during the year as the Indian Ocean Monsoon wind changes. Even for well-
studied flows such as the Gulf Stream, there is still uncertainty as to the
exact relationship between its meandering path and the location of the line
of zero wind-stress curl over the North Atlantic Ocean. Another ocean-wide
phenomenon is the complex ocean—atmosphere interaction which occurs in the
Equatorial Pacific, in which an anomaly in the wind-stress field on the Western
side can trigger off a wave which propagates eastwards and ultimately causes
a significant rise in the sea temperature off Ecuador and Peru, with disastrous
consequences for the local fisheries and a marked change in the coastal weather
patterns. El Nino, as it is called, has attracted a great deal of scientific attention
recently. If the mechanisms which produce it can be adequately modelled it may
be possible to predict the consequences several months ahead. Since wind stress
controls one of the important links in the chain of processes which are thought
to govern this ocean—atmosphere feedback loop, both its modelling and the
ultimate prediction of the phenomenon will benefit greatly from a detailed
knowledge of the time-evolving surface wind stress field.
Finally, as for the SMMR and altimeter estimates of sea state, the wind-
stress data which the SASS has collected for the southern hemisphere (although
mostly not yet processed) represents a very great improvement in data coverage
for the large areas of sea which are rarely traversed by shipping and even less
often sampled for oceanographic and meteorological parameters. Studies of
the circumpolar current and other dynamical ocean processes in the Antarctic,
which may have great bearing on the biological productivity of the area, and on
the formation of bottom water which spreads northwards at great depth, will be
assisted a great deal by the SASS data, and even more when a future satellite
scatterometer is able to supply a longer dataset.
into wave scatterometry has been linked to the development of the SASS and
the SAR. A lot of the recent development has come from two United States
experiments - the Marineland and the West Coast Experiments (Shemdin
(1980 a, b)).
A single-frequency wave scatterometer can be used as a two-scale wave
probe by illuminating an area of sea which is small compared with the length of
long ocean-gravity waves. If the modulation of the Bragg scattering waves by the
long waves is represented in a known (calculated or measured) modulation
transfer function, it is possible to recover information about the spectrum of the
long waves by analysing the doppler shifts of the returned radar energy. The
doppler shift is induced by the orbital velocity of the long waves. Directional
information can be obtained by illuminating an area which is short in one
direction and much longer than gravity wavelengths in the other. Since the wave
velocities parallel to the long direction will average out over the length of foot¬
print, only those orbital velocities normal to the long direction will be detected.
In a dual-frequency scatterometer, two closely-spaced and coherently-
related radar frequencies /i, /a, are used to illuminate an area of sea much larger
than the long gravity wavelength. The two frequencies may be emitted together,
or one after the other in a short time interval during which the sea-surface
conditions will not change. By adding together the two returns, a signal is
obtained whose overall energy is related to the amplitude of the Bragg scatter¬
ing waves. However, because the two frequencies are coherently related, a beat
frequency (A/ = /i — /2) can be detected, and the amplitude of this spectral
line which stands out above the rest of the background of the return radar
spectrum depends on the modulation pattern of the Bragg ripples which is itself
in resonance with the beat frequency. If /i and f2 are chosen so that the Bragg
resonance wavelength of the beat frequency falls within the wavelength range of
long gravity (swell) waves, then the amplitude of the beat-frequency spectral
line can be related through the modulation transfer function to the amplitude
of the long waves at the interference wavelength. The theory is developed fully
in Alpers & Hasselmann (1978).
CHAPTER 14
14.1.3 TOPEX
Another ocean-oriented satellite in an advanced programme planning stage
which may receive funding to enable it to fly within the next few years is the
US/French TOPEX mission. The name TOPEX stands for Ocean Topography
Experiment, and at the centre of the mission is a satellite designed particularly
for radar altimetry over the sea surface. The significant difference from the
Seasat and ERS-1 projects will be that the satellite design and orbit will
be tailored to the demands of measuring sea-surface topography, as discussed in
section 9.7, rather than having to compromise between those and the conflict¬
ing demands of radar measurement of sea state. Consequently the proposed con¬
figuration is for a non-sun-synchronous orbit to avoid tidal aliasing, at a height
of 1334 km to reduce atmospheric drag, with an inclination of 63.4° to obtain
optimum orbit-intersection angles at low- to mid-latitudes. It will therefore
exclude the polar regions, and be of less value for providing regular sea-state
statistics, but the design goal is to measure the satellite height above the sea
surface to an accuracy of two centimetres. To allow for ionospheric transmission-
speed effects, a two-wavelength radar instrument will be used. A 10-day repeat
frozen orbit is envisaged as the basic mode of operation, but with the flexibility
to change the ground-track coverage for special purposes.
Five years is therefore the minimum envisaged duration for the experiment,
in order to determine the longer-period changes in circulation such as the inter¬
annual variability. By the end of this time, the chosen orbit configuration should
make it possible to extract the low- to mid-latitude global-variability field of
sea-surface topography on time scales between 20 days and five years, and
length scales from 30 km to the ocean basin widths. To use this knowledge
of the sea-surface topography to develop a scientific understanding of the
wind-driven ocean circulation, it will also be desirable to know the global wind¬
forcing over the oceans during the lifetime of the mission. Ideally, therefore,
a wind scatterometer should be flown on a separate satellite, or possibly on
TOPEX itself. At the same time, the more in situ ship and buoy measurements
of ocean circulation parameters that can be made during the mission, the more it
will be possible to relate the surface topography (that is, the barotropic motions)
to the vertical structure (the baroclinic motions). In this respect TOPEX wiU
form part of the proposed World Ocean Circulation Experiment (WOCE) whose
aim is to determine the heat and freshwater flow between the ocean basins over
a five-year period between 1988 and 1993.
Whilst topographic variability should be determined with TOPEX to an
accuracy of a few centimetres, provided that accurate orbit-determination is
achieved, it will not be possible to obtain a measurement of the mean sea-surface
topography unless the geoid is known accurately to within a few centimetres
over length scales as short as a few tens of kilometres. Whilst TOPEX will
itself contribute significantly to an improved knowledge of the geoid, it cannot
provide an independent geoidal measurement with which to calibrate mean-
ocean topography. This must come from independent gravity measurements,
achievable on a global scale only by satellite systems such as GRAVSAT. An
improved knowledge of the gravity field will also assist in orbit determination,
and reduce the demand for large numbers of ground-tracking stations. Thus,
the usefulness of TOPEX and other altimeters such as that proposed for the
Canadian RADARSAT, is ultimately dependent on gravity-measuring satellites.
This increases the total cost of such exercises, but it should be noted that
the gravity experiment need only occupy a few months of the total life-span
of-the TOPEX mission.
latter are observing properties of the water itself (temperature, turbidity, colour,
etc.). There will be many instances where there is no correlation at all between
the patterns in images of the same sea area from an imaging radar and from a
radiometer. But where there is obvious correlation, then that in itself is evidence
of some dynamical process which links the water property under surveillance
to the surface roughness. Some possible effects like this have been observed
already — the convergence associated with a front often results in a change
in surface roughness across it, or an accumulation of surface material producing
a smooth slick at the front. Thus the front acquires a radar signature as well
as a thermal and/or colour signature. There may well be other effects like
this waiting to be discovered, which may give oceanographers further insights
into dynamical processes occurring near the ocean surface. ERS-1, carrying
a scanning IR radiometer as well as an active microwave device capable of
producing radar images, offers promise in this direction.
Whilst this approach of comparing images from different sensors is relatively
new for oceanographic applications, it has been used more for land applications,
and is finding increasing value in polar research and the study of sea ice.
14.4 CONCLUSION
This book commenced by addressing itself to the question of whether satellite
oceanography could be isolated as a subject for study in its own right. Whilst
a new technique is being developed, there is good reason for studying it on its
own as a separate branch of ocean science, but books like this will not have
done a service to oceanography if they merely encourage the growth of a breed
of ‘satellite oceanographers’. The aim of this book has been to introduce the
techniques of remote sensing to oceanographers in general. It has also attempted
to give remote-sensing specialists a flavour of those branches of oceanography in
which satellites can be useful, and to show both groups of scientists how remote
sensing is being applied to the advancement of oceanographic science. Ultimately,
satellite remote sensing should become just another observational technique
available to the oceanographer, and one which is understood to some degree
by all oceanographers in just the same way that even the most desk-bound
theoretical oceanographer appreciates the importance, and the limitations, of
observations made at sea.
At present, however, many oceanographers are not familiar with the type
of ocean observations and measurements which can be made using satellites. Of
those who have taken an active interest in remote sensing a large proportion have
concentrated on the development and calibration of sensors and the validation
of data. Relatively few have yet managed to make creative scientific advances in
oceanography which could not have been achieved without observations from
space.
The challenge to those pursuing research in satellite oceanography, both
oceanographers and remote-sensing scientists, is to explore theories and concepts
Sec. 14.4] Conclusion 427
and to develop applications which would not have been possible using only con¬
ventional oceanographic observations. These will include looking at windows in
the space—time spectrum of oceanographic processes which require the synoptic
spatial resolution and long-period sampling capabilities of the satellite sensors.
They will also make use of the different types of sea observation (for example,
radar surface roughness) and the spatially-averaged data which are obtained using
satellites. The most fruitful areas of research are likely to be those in which con¬
ventionally-gathered data and satellite observations are used to complement
each other in order to reveal a fuller perspective of oceanographic processes than
either of them is individually capable of providing. Given a coordinated deploy¬
ment of the exciting range of satellite sensor types discussed in this book, and
a continuation of existing ship and buoy sampling programmes, the stage is set
for a steady unfolding of new scientific understanding of the ocean over the next
decade.
References
Bernstein, R. L., & Morris, J. H. (1983). Tropical and Mid-latitude North Pacific
sea surface temperature variability from the Seasat SMMR. J. Geophys.
Res. 88C 1877-1891
Blume, H-J. C., Kendall, B. M., & Fedors, J. C. (1978). Measurement of ocean
temperature and salinity via microwave radiometry. Boundary-layer
Meteorol. 13 295-308
Boermer, W. M., Jordan, A. K., & Kay, I.W. (eds) (1981). Special issue on inverse
methods in electromagnetics. Trans. I.E.E.E. Antennae and Propagation,
AP-29 (2) 185-417
Born, G. H., Wilkerson, J. C., & Lame, D. B. (eds) (1979). Seasat Gulf of Alaska
Workshop Report. Vols 1 and 2. Rep. 622—101. Jet Propulsion Lab.
Pasadena, California
Born, G. H., Richards, M. A., & Rosborough, G. W. (1982). An empirical deter¬
mination of the effects of sea state bias on Seasat altimetry. J. Geophys.
Res. %1C 3221-3226
Bracalente, E. M., Boggs, D. H., Grantham, W. L., & Sweet, J. L. (1980). The
SASS scattering coefficient Oq algorithm. I.E.E.E. J. Oceanic Eng. OE-5
145-154
Brown, G. S. (1979). Estimation of surface wind speeds using satellite-borne
radar measurements at normal incidence. J. Geophys. Res. 84B 3974—3978
Brown, R. D. (1983). M2 Ocean tide at Cobb Seamount from Seasat altimeter
data. J. Geophys. Res. 88C 1637—1646
Brown, R. D., Kahn, W. D., McAdoo, D. C., & Himwich,W; E. (1983). Roughness
of the Marine Geoid from Seasat altimetry. J. Geophys. Res. 88C
1531-1540
Bukata, R. R, Bruton, J. E., Jerome, J. H., Jain, S. C., & Zwick, H. H. (1981).
Optical water quality model of Lake Ontario. 2: Determination of
chlorophyll a and suspended mineral concentrations of natural waters from
submersible and low altitude optical sensors. Applied Optics. 20 1704—1714
Bullard, R. K. (1983a). Land into sea does not go. In: Cracknell, A. P. (ed.).
Remote sensing applications in marine science and technology, Dordrecht,
D.Reidel,p. 359-372
Bullard, R. K. (1983b). Detection of marine contours from Landsat film and
tape. In: Cracknell, A. P. (ed.). Remote sensing applications in marine
science and technology, Dordrecht, D. Reidel, p. 373—381
Bunker, A. F., Charnock, H., & Goldsmith, R. A. (1982). A note on the heat
balance of the Mediterranean and Red Seas. Jour. Marine Research 40
(supplement) 73—84
Campbell, W.J., Gloersen, R, Nordberg,W., & Wilheit,T. T. (1974). Dynamics and
morphology of Beaufort Sea ice determined from satellites, aircraft and
drifting stations. Proc. COSPAR symp. on approaches to earth survey
problems through use of space techniques. Constanz, F. R. G., May 1973.
p. 311-327
References 431
Campbell, W. J., Ramseier, R. 0., Zwally, H. J., & Gloersen, R (1980). Arctic
sea-ice variations from time-lapse passive microwave imagery. Boundary-
layer Meteorology. 18 99—106
Cartwright, D. E. (1983). Detection of large-scale ocean circulation and tides.
Phil. Trans. Roy. Soc. Lond. A309 361—370
Cartwright, D. E. & Alcock, G. A. (1981). On the precision of sea surface
elevations and slopes from Seasat altimetry of the Northeast Atlantic
Ocean. In: Gower, J. F. R. (ed.). Oceanography from space, New York and
London, Plenum Press, p. 885—895
Cartwright, D. E. & Alcock, G. A. (1983). Altimeter measurements of ocean
topography. Chapter 19. In: Allan, T. D. (ed.). Satellite microwave remote
sensing, Chichester, Ellis Norwood, p. 309—319
Cartwright, D. E., Edden, A. C., Spencer, R., & Vassie, J.M. (1980). The tides of
the northeast Atlantic Ocean. Phil. Trans. Roy. Soc. Lond. A298 87—139
Castleman, K. R. (1979). Digital image processing. Englewood Cliffs, N. J.,
Prentice-HaU
Charnock, H. «fe Pollard R. T. (1983). Results of the JASIN project. London,
The Royal Society, 229 pp
Chelton, D. B., Hussey, K. J., & Parke, M. E. (1981). Global satellite measure¬
ments of water vapour, wind speed and wave height. Nature 294 529—532
Cheney, R. E. (1982). Comparison data for Seasat altimetry in the western North
Atlantic. J. Geophys. Res. 87C 3247—3253
Cheney, R. E., Marsh, J. G., & Beckley, B. D. (1983). Global mesoscale variability
from collinear tracks of Seasat altimeter data. J. Geophys. Res. 88C
4343-4354
Citeau, J. & Domain, F. (1981). A short review of an oceanographic use of
Meteosat data at ORSTOM remote sensing service. In: Application of
remote sensing data on the continental shelf. Proceedings of an EARSeL-
ESA Symposium. ESA SP-167, Paris, European Space Agency, p. 145-156
Clark, D. K. (1981). Phytoplankton pigment algorithms for the Nimbus-7 CZCS.
In: Gower, J. F. R. (ed.). Oceanography from space. New York, Plenum
Press, p.221—231
Colwell, R. N. (ed.) (1984). Manual of remote sensing, (2nd edn). Falls Church,
Virginia, American Society of Photogrammetry, 2440 pp
Cox, C. S. (1974). Refraction and reflection of light at the sea surface. In: Jerlov,
N. G. & Steemann Nielsen, E. (eds). Optical aspects of oceanography,
London and New York, Academic Press, p. 51—75
Cox, C. & Munk, W. (1954). Measurements of the roughness of the sea surface
from photographs of the sun’s glitter. J. Opt. Soc. Am. 44 838—850
Cracknell, A. R, MacFarlane, N., McMillan, K., Charlton, J. A., McManus, J., &
Ulbricht, K. A. (1982). Remote sensing in Scotland using data received from
satellites. A study of the Tay Estuary region using Landsat multispectral
scanning imagery. Int. J. Remote Sensing 3 113—137
432 References
Cutting, E., Born, G. H., & Frautnich, J. C. (1978). Orbit analysis for Seasat-A.
J. Astronautical Sci. 26 315—342
Deschamps, P.Y. & Frouin, R. (1984). Large diurnal heating of the sea surface
observed by the HCMR experiment. J. Physical Oceanogr. 14 177—184
Deschamps, P. Y. & Phulpin, T. (1980). Atmospheric correction of infrared
measurements of sea surface temperature using channels at 3.7, 11 and
12 Atm. Boundary-layer Meteorology 131—143
Dickey, T. D. & Simpson, J. J. (1983). The influence of optical water type on the
diurnal response of the upper ocean. Tellus 35 142—154
Dixon, T. H., Naraghi, M., McNutt, M. K., & Smith, S. N. (1983). Bathymetric
prediction from Seasat altimeter data. J. Geophys. Res. 88C 1563—1571
Domain, F., Citeau, J., & Noel, J. (1980). Sea surface temperatures studied by
Meteosat data along the coast of Senegal and Mauritania. In: Coastal and
Marine Applications of Remote Sensing, Proc. 6 th Annual Conf of Remote
Sensing Society, 1979, Dundee (Cracknell, A.P. (ed.)),The Remote Sensing
Society, p. 59—67
Douglas, D. C., Goad, C., Morrison, C., & Foster, F. (1980). A determination of
the geopotential from satellite-to-satellite tracking data. J. Geophys. Res.
85B 5471-5480
Duntley, S. Q. (1965). Oceanography from manned satellites by means of visible
light. In: Ewing, G.C. (ed.). Oceanography from space. Woods Hole Oceano¬
graphic Institution, p. 39—46
Dyer, K. R. (1973). Estuaries, a physical introduction. Chichester, J. Wiley
Ewing, G. C. (ed.) (1965). Oceanography from space. Ref. No. 65—10. Woods
Hole Oceanographic Institution, Mass., 469 pp
Ewing, G. & McAlister, E. D. (1960). On the thermal boundary-layer of the
ocean. Science, New York. 131 1374—1376
Fedor, L. S. & Brown, G. S. (1982). Waveheight and wind speed measurements
from the Seasat radar altimeter. J. Geophys. Res. 87C 3254—3260
Fedor, L. S., Godbey, T. W., Gower, J.F. R., Guptill, R., Hayne, G. S., Rufenach,
C. L., & Walsh, E. J. (1979). Satellite altimeter measurements of sea state —
an algorithm comparison. J. Geophys. Res. 84B 3991—4001
Flatte, S. M., Dashen, R. D., Munk, W. H., Watson, K. M., & Zachariasen, R.
(1979). Sound transmission through a fluctuating ocean. London,
Cambridge University Press
Fleming, E. A. & Le Lievre, D.D. (1977). The use of Landsat imagery to locate
uncharted coastal features on the Labrador Coast. Proc. 11th Int. Symp. on
Remote Sensing of Environment. Michigan, April 1977, p. 775—782
Francis, C. R., Thomas, D. P., & Windsor, E. P. L. (1983). The evaluation of
SMMR retrieval algorithms in satellite microwave remote sensing. In: Allan,
T. D. (ed.). Satellite microwave remote sensing, Chichester, Ellis Horwood,
p. 481-498
Fu, L.-L. & Holt, B. (1982). Seasat views oceans and sea ice with synthetic
References 433
London, D. Reidel
Gower, J. F. R (ed.) (1981). Oceanography from space. New York and London,
Plenum Press, p. 978
Gower, J. F. R., Denman, K. L., & Holyer, R. J. (1980). Phytoplankton patchiness
indicates the fluctuation spectrum of mesoscale oceanic structure. Nature
288 157-159
Grantham, W. L., Bracalente, E.M., Jones, W. L., & Johnson, J.W. (1977). The
Seasat-A satellite scatterometer. I.E.E.E. J. Oceanic. Eng. OE-2 200—206
Grassl, H. (1976). The dependence of the measured cool skin of the ocean on
wind stress and total heat flux. Boundary-layer Meteorology 10 465—474
Guymer, T. H. (1983a). Validation and applications of SASS over JASIN. In:
AUan, T. D. (ed.). Satellite microwave remote sensing, Chichester, Ellis
Horwood, p. 87—104
Guymer. T. H. (1983b). A review of Seasat scatterometer data. Phil. Trans.
Roy. Soc. London A309 399—414
Hammond, D. L., MenneUa, R. A., & Walsh, E. J. (1977). Short pulse radar used
to measure sea surface wind speed and S.W.H. I.E.E.E. Trans. Ant. Prop.
AP-25 61-67
Hansen, B. & Meincke, J. (1979). Eddies and meanders in the Iceland-Faroe
ridge area. Deep Sea Res. 26 1067—1082
Hasse, L. (1971). The sea surface temperature deviation and the heat flow at
the sea-air interface. Boundary-layerMeteorology 1 368—379
HasseImann, K. et al. (1973). Measurements of wind-wave growth and swell
decay during the Joint North Sea Wave Project (JONSWAP). Erganzungsheft
zur Deutschen Hydrographischen Zeitscrift. Reiche A8 Nr 12, 95 pp
Hasse Imann, K., Raney, R.K., Plant, W.J., Alpers,W., Shuchman, R. A., Lyzenga,
D. R., Rufenach, C. L., & Tucker, M. J. (1984). Theory of SAR ocean wave
imagery: A MARSEN view. J. Geophys. Res. (In press)
Hill, R. H. (1972). Laboratory measurement of heat transfer and thermal
structure near an air-water interface. J. Phys. Oceanography 2 190 — 198
Hojerslev, N. K. (1974). Inherent and apparent optical propteries of the Baltic.
Rep. Inst. Physical Oceanography, University of Copenhagen, No. 23,70 pp
Hojerslev, N. K. (1980). On the origin of yellow substance in the marine environ¬
ment. Rep. Inst. Physical Oceanography, University of Copenhagen, 42
39-56
Holligan, P. M., Viollier, M., Dupouy, C., & Aiken, J. (1983). Satellite studies on
the distributions of chlorophyll and dynoflagellate blooms in the western
English Channel. Continental Shelf Research 2 81—96
Horstmann, U. & Hardtke, P. G. (1981). Transport processes of suspended
matter, including phytoplankton, studied from Landsat images of the
Southwestern Baltic Sea. In: Gower, J. F. R. (ed.). Oceanography from
space. New York and London, Plenum Press, p. 429—438
Hovis, W. A., Clark, D.K., Anderson, F., Austin, R.W., Wilson, W.H., Baker, E.T.,
References 435
Ball, D., Gordon, H. R., Mueller, J. L., El-Sayed, S. Z., Sturm, B., Wrigley,
R. C., & Yentsch, C. S. (1980). Nimbus-7 Coastal Zone Colour Scanner:
system description and initial imagery. Science 210 60—63
Hughes, B. A. & Gower, J. F. R. (1983). SAR imagery and surface truth com¬
parisons of internal waves in Georgia Strait, British Colombia, Canada. J.
Geophys. Res. 88C 1809—1824
Hughes, B. A. & Grant, H. L. (1978). The effect of internal waves on surface
wind waves, 1. Experimental measurements. /. Geophys. Res. 83C 443—454
Huhnerfuss, H., Walter, W., & Kruspe, G. (1977). On the variability of surface
tension with mean wind speed. J. Phys. Ocean.l 567—571
Isimaru, A. (1978). Wave propagation and scattering in random media. Academic
Press, New York
Jerlov, N.G. (1976). Marine optics, Amsterdam, Elsevier
Johnson, R. W. (1975). Quantitative sediment mapping from remotely-sensed
multi-spectral data. In: Shahrokhi, S. (ed.). Remote sensing of earth
resources. Vol. IV, University of Tenessee,Tullahoma,Tenn., p. 565—576
Johnson, J. W., Williams, L. A., Bracalente, E. M., Beck, F. B., & Grantham, W. L.
(1980). Seasat-A satellite scatterometer instrument evaluation. I.E.E.E. J.
Oceanic Eng. OE-5 138—144
Jones, W. L. & Schroeder, L. C. (1978). Radar backscatter from the ocean:
dependence on surface friction velocity. Boundary-layer Meteorology 13
133-149
Jones, W. L., Schroeder, L. C., Boggs, D. H., Bracalente, E. M., Brown, R. A.,
Dome, G. J., Pierson, W. J., & Wentz, F. J. (1982). The Seasat-A satellite
scatterometer: the geophysical evaluation of remotely sensed wind vectors
over the ocean. J. Geophys. Res. 87C 3297—3317
Katsaros, K. B. (1977). The sea surface temperature deviation at very low wind
speeds: is there a limit? Tellus 29 229—239
Katsaros, K. B. (1980). The aqueous thermal boundary layer. Boundary-layer
Meteorology 18 107—127
Katsaros, K. B., Taylor, P. K., Alishouse, J. C., & Lipes, R. G. (1981). Quality
of Seasat SMMR atmospheric water determinations. In: Gower, J. F. R.
(ed.). Oceanography from space. New York and London, Plenum Press,
p. 691-706
King-Hele, D. (1959). The effect of the earth’s oblateness on the orbit of a near
satellite. Proc. Roy. Soc. London 247A 49-72
King-Hele, D. (1964). Theory of satellites in an atmosphere. London,
Butterworth, 165 pp
King-Hele, D. (1976). The shape of the earth. Science 192 1293-1300
Kinsman, B. (1965). Wind waves. Prentice-Hall, 676 pp
Kirwan, A. D., Ahrens, T. J., & Born, G. H. (eds) (1983). Seasat special issue II:
Scientific results. Reprinted from J. Geophys. Res. 88 (C3), Washington,
American Geophysical Union, p. 1529—1952
436 References
Klein, L. A. & Swift, C.T. (1977). An improved model for the dielectric constant
of sea water at microwave frequencies. I.E.E.E. Trans. Antennae and
Propagation AP-25 104—111
Klemas, V. (1980). Remote sensing of coastal fronts and their effects on oil
dispersion. Int. J. Remote Sensing 1 11—28
Klemas, V., Bartlett, D., Philpot, W., & Rogers, R. (1974). Coastal and estuarine
studies with ERTS-1 and Skylab. Remote SensingEnvir. 3 153—174
Krauss, E. B. (ed.) (1977). Modelling and prediction of the upper layers of the
ocean. Pergamon, Oxford
Kropotkin, M. A., Verbitskiy, V. A., Sheveleva, T. Y., & Tarashkevich, V. N.
(1978). Radiation temperature of a water surface with an oil slick.
Oceanology 18 730—731
La Fond, E. C. (1962). Internal waves. In: Hill, M. (ed.). The sea. Vol. 1,
Interscience, New York, p. 731—751
Lamb, H. (1936). Hydrodynamics. Cambridge University Press
Lane, J. A. & Saxton, J. A. (1952). Dielectric dispersion in pure polar liquids at
very high radio frequencies. Ill The effect of electrolytes in solution. Proc.
Roy. Soc. London 214 531—545
Large, W. G. & Pond, S. (1981). Open ocean momentum flux measurements in
moderate to strong winds. J. Phys. Oceanogr. 11 324 — 336
LeBlond,P. A. & Mysak, L. A. (1978). Waves in the ocean. Elsevier, 602 pp
Lee, C.Y. & Beardsley, R. C. (1974). The generation of long nonlinear internal
waves in a weakly stratified shear flow. J. Geophys. Res. 79 453—462
Le Fevre, J., Viollier, M., Le Corre, P., Dupouy, C., & Grail, J. R. (1982). Remote
sensing observations of biological material by Landsat along a tidal thermal
front and their relevancy to the available field data. Estuarine, coastal and
shelf science 16 37—50
Legeckis, R. (1979). A survey of world wide SST fronts detected by environ¬
mental satellites. J. Geophys. Res. 83C 4501—4522
Lillesand,T. M. & Kiefer, R.W. (1979). Remote sensing and image interpretation.
Chichester, J. Wiley, 612 pp
Lintz, J. & Simonett, D. S. (1976). Remote sensing of environment. Addison-
Wesley Publishing Co., Advanced Book Program, Reading, Mass.
Lipes, R. G. (1982). Description of Seasat radiometer status and results. J.
Geophys. Res. 87C 3385-3395
Lodge, D.W. S. (1981). The Seasat-1 synthetic aperture radar: introduction, data
reception and processing. In: Cracknell, A. P. (ed.). Remote sensing in
meteorology, oceanography and hydrology, Chichester, Ellis Horwood,
p. 335-356
Longuet-Higgins, G. S. & Stewart, R. W. (1964). Radiation stresses in water
waves, a physical discussion with applications. Deep Sea Res. 11 529-562
Lorell, J., Colquitt, E., & Anderle, R. J. (1982). Ionospheric correction for
Seasat altimeter height measurement. J. Geophys. Res. 87C 3207-3212
References 437
In: Beal, R. C., De Leonibus, P. S., & Katz, I. (eds), Spaceborne synthetic
Aperture radar for oceanography. Baltimore and London, Johns Hopkins
University Press, p.75—86
Rouse, L. J. & Coleman, J. M. (1976). Circulation observations in the Louisiana
Bight using Landsat imagery. Remote Sensing of Environment 5 55—66
Sabins, F.F. (1978). Remote sensing: principles and interpretation. San Francisco,
Freeman
Sathyendranath, S. & Morel, A. (1983). Light emerging from the sea - interpre¬
tation and uses in remote sensing. In: Cracknell, A.P. (ed.), Remote sensing
applications in marine science and technology, Dordrecht, D. Reidel,
p. 323-358
Saunders, P. M. (1967 a). Aerial measurement of sea surface temperature in the
infrared. J. Geophys. Res. 12 4109—4117
Saunders P. M. (1967b). The temperature at the ocean—air interface. J. Atmos.
Sci. 24 269-273
Saunders, P. M. (1973). The skin temperature of the ocean — a review. Mem.
Soc. Roy. des Sci. de Liege. Ser. 4 93—98
Schooley, A. H. (1977). Temperature of ocean skin related to cloud shadows.
J. Phys. Oceanogr. 1 486—487
Schroeder, L. C., Boggs, D. H., Dome, G., Halberstam, 1. M., Jones, W. L., Pierson,
W. J., & Wentz, F. J. (1982). The relationship between wind vector and
normalized radar cross section used to derive Seasat-A satellite scattero-
meter winds. J. Geophys. Res. 87C 3318—3336
Scott, J. C. (1972). The influence of surface-active contamination on the initia¬
tion of wind-waves. J. Fluid Mech. 56 591—606
Shemdin, O. H. (1980a). The West Coast experiment: an overview. EOS. 61
649-651
Shemdin, 0. H. (1980b). The marineland experiment: an overview. EOS 61
625-626
Simpson, J. H. (1981). Sea surface fronts and temperatures. In: Cracknell, A.P.
(ed.). Remote sensing in meteorology, oceanography and hydrology,
Chichester, Ellis Horwood, p. 295—311
Simpson, J. H., Hughes, D.G.,& Morris, N. C.G. (1977). The relation of seasonal
stratification to tidal mixing on the continental shelf. In: Angel, M. (ed.),
A voyage of discovery, Oxford, Pergamon Press, p. 327—340
Simpson, J. H. & Pingree, R. D. (1978). Shallow sea fronts produced by tidal
stirring. In: Bowman, M. J. & Esaias, W. E. (eds). Oceanic fronts in coastal
processes. New York, Springer-Verlag, p. 29
Simpson, J. J. & Dickey, T. D. (1981). The relationship between downward
irradiance and upper ocean structure. J. Phys. Oceanogr. 11 309—323
Simpson, J. J. & Paulson, C. A. (1980). Small scale sea-surface temperature
structure. /. Phys. Oceanogr. 10 399—410
Singh, S. M. & Warren, D. E. (1983). Sea surface temperatures from infrared
References 441
p. 267-280
Sturm, B. (1983). Selected topics of Coastal Zone Colour Scanner evaluation.
In: CrackneU, A.P. (ed.), Remote sensing applications in marine science and
technology, Dordrecht, D. Reidel, p. 137—168
Swift, C. T. (1980). Passive microwave remote sensing. Boundary-layer
Meteorology 18 25—54
Tapley, B. D., Born, G.H., & Parke, M. E. (1982a). The Seasat altimeter data and
its accuracy assessment. J. Geophys. Res. 87C 3179—3188
Tapley, B. D., Lundberg, J. B., & Born, G. H. (1982b). The Seasat wet tropo¬
spheric range correction. J. Geophys. Res. 87C 3213—3220
Taylor. P. K. (1983). The scanning multichannel microwave radiometer — an
assessment. In; Allan, T. D. (ed.). Satellite microwave remote sensing,
Chichester, Ellis Horwood, p. 463—480
Thomas, D. P. (1981). Microwave radiometry and applications. In: Cracknel!,
A. P. (ed.). Remote sensing in meteorology, oceanography and hydrology,
Chichester, Ellis Horwood, p. 357—369
Thompson, J. D., Born, G. H., &Maul, G.A. (1983). Collinear track altimetry in
the Gulf of Mexico from Seasat measurements, models and surface truth.
J. Geophys. Res. 88C 1625—1636
Tomiyasu, K. (1978). Tutorial review of synthetic-aperture radar (SAR) with
applications to imaging of the ocean surface. Proc. I.E.E.E. 66 563—583
TOPEX (1981). Satellite altimetric measurements of the ocean. Report of the
TOPEX Science Working Group. NASA. Jet Propulsion Laboratory,
California Institute of Technology, California
Townsend, W. F., McGoogan, J. T., & Walsh, E. J. (1981). Satellite radar alti¬
meters — present and future oceanographic capabilities. In; Gower, J. F. R.
(ed.). Oceanography from space. New York and London, Plenum Press,
p. 625-636
Trask, R. P. & Briscoe, M. G. (1983). Detection of Massachusetts Bay internal
waves by the synthetic aperture radar on Seasat. J. Geophys. Res. 88C
1789-1799
Tucker, M. J. (1983). Observations of ocean waves. Phil. Trans. Roy. Soc.
London A309 371—380
Uslenghi, P. L. E; (ed.) (1978). Electromagnetic scattering. Academic Press,
New York
Valenzuela, G. R. (1978a). Scattering of electromagnetic waves from the ocean.
In; Lund,T. (ed.), Surveillance of environmental pollution and resources by
electromagnetic waves. Hingham, D. Reidel, p. 199—226
Valenzuela, G. R. (1978b). Theories for the interaction of electromagnetic and
oceanic waves - a review. Boundary-layer Meteorology 13 61-85
Valenzuela, G. R. & Wright, J. W. (1979). The modulation of short gravity¬
capillary waves by longer scale periodic flows. A higher order theory. Radio
Science 14 1099—1110
References 443
'W? ^ I*-'
- MtK ‘i*'** , ' - ^ <|0 '.jr,; ^
551.46«ij||^jan„„raohV: an.i/Rpbinsonj
REMOTE SENSING IN METEOROLOGY, OCEANOGRAPHY AND HYDROLOGY
Editor; A. P. CRACKNELL, Carnegie Laboratory of Physics, University of Dundee
"timely and valuable ... an excellent book" — J. O. Thomas, Imperial College of Science and Technology,
University of London, in International Journal of Remote Sensing.
"a book with a unique history . . , undeniably handsome" — E, C. Barrett, University of Bristol, in Journal oi'<
Climatology.
ENVIRONMENTAL AERODYNAMICS
R. S. SCORER, Professor of Theoretical Mechanics. Imperial College of Science and
Technology, University of London
"authoritatively written by an eminent expert... excellently illustrated" - Choice (USA).
"authoritative and vary readable ... will become a wall read text for those interested in a quantitative
explanation to hydrodynamical problems of the great outdoors" — Garry Hunt in New Scientist.
published by distributed by : j
ELLIS NORWOOD LIMITED HALSTED PRESS a division of \
Publishers Chichester JOHN WILEY & SONS
New York Chichester Brisbane Toronto