Understanding Synthetic Aperture Radar Images PDF
Understanding Synthetic Aperture Radar Images PDF
Chris Oliver
Shaun Quegan
SciTEGH
PUBLISHING, INC.
© 2004 by SciTech Publishing, Inc.
Raleigh, NC 27613
This is a corrected reprinting of the 1988 edition originally published by Artech House:
Boston.
All rights reserved. No part of this book may be reproduced or used in any form whatsoever
without written permission from the publisher.
10 9 8 7 6 5 4 3 2 1
ISBN: 1-891121-31-6
SciTech books may be purchased at quantity discounts for educational, business, or sales
promotional use. For information contact the publisher:
Oliver, Chris
Understanding synthetic aperture radar images / Chris Oliver, Shaun Quegan.
Includes bibliographical references and index.
ISBN 1-891121-31-6
1. Synthetic aperture radar. I. Quegan, Shaun, 1949- II. Title.
TK6592.S95043 2004 97-41709
621.3848—dc21 CIP
To Betty and Sue, who understand
Foreword
Over the past few years, the development of spaceborne and airborne SAR
systems for the observation of the Earths surface has experienced a rapid
growth. The ERS-I, JERS-I, ERS-2, and RADARSAT satellite systems have
provided and continue to provide us with long time-series of high quality SAR
data. At the same time, advanced airborne SAR systems exist in many countries
of the world and provide multiple frequency and polarization observations for
both civilian and military use. The SIR-C/X-SAR shuttle missions have widely
demonstrated the advantages of having such multiple parameter systems in
space.
Despite the enormous amount of available data, little effort has been
expended in making use of the data in applications and scientific disciplines. To
achieve this, both the physical content of the data and the information content
of the SAR images, as well as the bridge between the two sources of information,
must be understood by the users of Earth observation data.
While several early volumes in radar remote sensing covered either the
physical content of SAR data, or the basics of SAR image formation and SAR
image statistical properties, there is a strong need for an integrated book that
brings together both the basic properties of SAR images related to image
formation and those related to scene properties.
In this context, this book appears very timely to fill the above mentioned
gap. It provides the readers with a synoptic understanding of SAR image
fundamental properties; these are a prerequisite for the development of models
and methods used in the interpretation and exploitation of SAR data for various
applications.
This book deals with most aspects of SAR images on land surfaces and
contains the most up-to-date coverage of the subject, including the underlying
principles, recent developments, and future directions for research activities in
this area. The subject is treated with the breadth and depth required for use as
a reference text for graduate students, engineers, and application scientists. At
the same time, the many illustrations of the underlying theoretical principles,
contained in the different chapters, will certainly inspire the development of
new tools and stimulate new practical applications of SAR images.
The authors have been leading authorities in the field for many years.
Thus, the clear, thorough, and objective writing of the volume is a reflection of
the authors' extensive experience as scientists and teachers. In addition, the
authors have demonstrated that scientific texts can also be written in beautiful
English language.
I am confident that this book will quickly become a well-read text for
those interested in SAR images, and thus will contribute significantly to expand
the effective use of SAR data in the observation and monitoring of the Earths
environment.
Thuy Le Toan
Toulouse, September 1997
Preface
The authors have been collaborating on SAR research since 1982, when Shaun
Quegan joined Marconi Research Centre, Great Baddow, and Chris Oliver was at
the Royal Signals and Radar Establishment, Malvern (later to become part of the
Defence Research Agency (DRA), which is now the Defence Evaluation and Re-
search Agency (DERA)). At that time the status of SAR as a means of producing
images of the Earths surface was very different from the present. Only the short-
lived Seasat mission had provided spaceborne data, and few airborne systems were
in operation. Images were not widely available (although the DRA X-band SAR
provided us with a data source) and in most cases suffered from various types of
defect. Image quality issues often overshadowed the task of information extrac-
tion so that, in common with many other researchers, we invested considerable
time and effort on methods of data correction. The success of this work (such as
the signal-based motion compensation schemes described in Chapter 3) in com-
bination with significant progress in system engineering and calibration has en-
sured that close to ideal quality SAR data is now to be expected from modern
systems. We have also seen the development of new techniques, such as polarimet-
ric and interferometric SAR, and the advent of a diversity of SAR data providers,
including spaceborne systems (ERS-I and 2, JERS, and Radarsat) and flexible
multifrequency polarimetric systems carried on aircraft or the Space Shuttle.
As the accent has changed from data correction, the information con-
tained in SAR images has come to occupy its rightful place as the purpose of
data acquisition. The investigation of methods for information extraction has
formed the second and continuing strand in our long collaboration. In this we
have been much helped by the complementarity of our interests. Chris Oliver
has primarily been concerned with the development of image-interpretation
tools (usually for high resolution SAR in a military context), while Shaun
Quegan has concentrated more on their impact on remote sensing applications.
Necessarily, there has been significant overlap in our work. This combination of
interests means that we can go some way towards crossing the cultural divides
that hamper SAR development, in particular those that separate military from
civil applications and applications from algorithm development.
Military applications are primarily concerned with detecting and recogniz-
ing targets, which usually demands imagery of the highest resolution. Civil
applications, on the other hand, may require information about many diverse
aspects of the environment, normally at lower resolution. The two have com-
mon ground in the need to understand the properties of distributed scatterers:
in the military case because this constitutes the background clutter against
which target detection takes place; in the civil case because this clutter is often
the primary object of interest. More generally, both types of application have an
interest in scene understanding. However, military needs, where cost may not
have been a dominating issue, have given rise to sophisticated techniques that
are potentially of considerable value to the civil field. These have become
accessible due to the reduction in the cost of computing. Such is true of many
of the developments described in this book.
The divide between remote sensing applications and algorithm develop-
ment is symptomatic of the fact that SAR imaging presents many opportunities
for formulating mathematical problems that can be pursued in an almost
abstract way. Linking the problem and its solution to what is really needed by
an application may be much harder; not enough attention has been paid to this
issue. This does not mean that a rigorous approach should be avoided in the
development of image-interpretation tools. On the contrary, such tools must be
well grounded in the principles of optimization or conservation if they are to be
of any general value. Ad hoc methods rarely prove robust. We tried to adopt
such a rigorous approach throughout this book, but always with concern for
validation of a technique and its relevance to applications.
In our attempt to bridge these cultural divides, the process of information
extraction is viewed as a whole, from sensor parameters and their impact on
image properties, through data models describing image statistics and the char-
acteristics of the terrain, to the output information required by the application.
This approach is tempered by our knowledge of the scattering process, based on
physical understanding, models, and experimental findings. Such a complete
view of what makes up the image is essential in identifying which properties of
the scene affect the data. From this we can recognize those features (if any) that
carry the information required by the application. The role of image analysis is
to provide an optimum representation of these features. To achieve this, we
formulate a series of models that represent our best attempt to encapsulate the
properties of the imaging process and those of the scene. Optimality can then
often be defined in a framework such as that provided by a Bayesian analysis,
although other criteria are possible. Always, the effectiveness of these optimal
solutions must be judged by the user, whose requirements should impinge on
every stage in the design of image-interpretation tools. Put briefly, we believe
that these tools should be firmly based in an understanding of the nature of the
SAR image; that their performance must be quantified and the best tools
recognized; and that these best tools should be made available to the image
interpreter, who is the final arbiter of their efficacy.
Although this sets out our philosophy, the reader may be helped by a more
specific guide to the main structural components of the book. Fundamental
aspects of how SAR images are formed, their basic characteristics arising from
the imaging process, and their statistical properties are covered in Chapters 2,
4, and 5. These chapters underlie much of the later material. Chapter 3 is
concerned with the recovery of nearly ideal images from imperfect systems; since
its emphasis is on the SAR processing, it can be omitted by those readers
interested only in working from the given data. Its purpose is to illustrate that
current techniques should ensure that the data are of the highest quality. Image
analysis tools for scene understanding in single-channel data form a central core
of the book and are developed in Chapters 6 to 9. Chapter 10 deals with target
detection in an overtly military context. The purpose of Chapters 11 and 12 is
to cast much of the content of Chapters 5 to 9 into the forms required to handle
multitemporal, multifrequency, polarimetric, and interferometric data and to
explore the much enhanced range of information available when such mul-
tichannel data are available. Real examples and contact with applications are a
recurrent theme throughout, but Chapter 13 becomes more closely oriented
with applications because it considers image classification. This is a central
concern in remote sensing. It provides a focal point where image analysis,
physical understanding, scene properties, and user requirements all interact.
Finally, in Chapter 14 we present our view of the current status and
prospects for further development. This ultimately depends on the extent to
which the material of this book finds its way into applications. It will be obvious
that currently available techniques show considerable promise and that in many
cases preferred approaches can be identified. We hope that readers are persuaded
to exploit the most powerful image analysis tools available, leading to more
effective use of the data and thus allowing the analyst to concentrate more time
on the real task of image interpretation. By the ensuing interplay of algorithms
and applications, the shortcomings of our current understanding and methods
will be exposed. This is the engine that will drive further development of
techniques for extracting information from SAR images.
A c k n o w l e d g m e n t s
This book reflects many years of SAR research during which the authors have
had the good fortune to work with, learn from, and disagree with each other
and numerous associates at our own and other research establishments. It has
been an exciting time. On many occasions we seem to have been forced back to
re-examine our fundamental assumptions, often in heated discussion. One of
the first questions we asked each other before this undertaking was, "Can you
take criticism?" The give and take involved in being joint authors has only been
possible because we thought (correctly) that the answer was "Yes." This book
represents the fruits of this often painful but ultimately rewarding interaction.
We hope our efforts are better for it.
Chris would like to extend his grateful thanks to a large number of
friends and colleagues in the DRA who have contributed to the research pro-
gram. Chris Baker, Dave Belcher, Dave Blacknell, Alan Blake, Ian Finley, An-
drew Home, Richard White, and Jim Wood shared in the research into
autofocus and signal-based motion compensation for SAR imaging. Dave
Blacknell, Alan Blake, Eric Jakeman, Robert Tough, Keith Ward, and Richard
White all contributed to the understanding of data models. Steve Luttrell
provided exciting collaboration on super-resolution. Dave Blacknell, Alan
Blake, and Richard White are all involved in despeckling filters and segmen-
tation methods. Pierfrancesco Lombardo enjoyed a productive year at the
DRA working on texture analysis before returning to the University of Rome,
La Sapienza.
InfoSAR has been engaged in the development of real-time SAR
autofocus processing and image-understanding techniques for the DRA. Chris
would like to acknowledge the contributions of Mike Delves and Rod Cook,
who are involved with the entire program; Keith Beckett and Gordon Pryde,
who developed the real-time SAR processor and autofocus software; and Ian
McConnell and Dave Stewart, who made significant progress in implementing
and developing the image-understanding algorithms. Indeed, many of the algo-
rithms described in this book, with subsequent development, are available in
the InfoPACK software package from InfoSAR.1
Chris would like to acknowledge the considerable investment made by the
DRA and the Ministry of Defence, in terms of staff, equipment, and funding
for the SAR research team. He is also grateful for the encouragement to under-
take a review of the whole topic over the last two years that was initially reported
in DRA Technical Reports Nos. DRA/LS2/TR96016 to 96021. Many illumi-
nating discussions with Dave Blacknell and Richard White assisted in their
preparation. Chapters 3 and 5 to 10 in this book are largely based on this
material and are British Crown Copyright 1997/DERA (with the exception of
Figures 6.12(a) and 8.1 and Table 10.2). So also are Figures 1.1, 4.1, 4.9,
12.11 (d—j), and 12.12. These are published with the permission of the controller
of Her British Majesty's Stationary Office.
Past and present members of Shaun Quegan's SAR Research Group within
the Sheffield Centre for Earth Observation Science (SCEOS) contributed enor-
mously to the overall research program and in putting this book together. Kevin
Grover not only did most of the work on change detection in Amazonia but was
willing to visit SCEOS long after his Ph.D. was completed in order to put the
images in good order for Chapter 12. Jiong Jiong Yu is helping to extend the
theoretical treatment of change detection and its applications. Ron Caves played
a central role in developing algorithms for segmenting multidimensional SAR
data and in producing quantitative measures of segmentation performance.
Mike Preston moved this work forward on both fronts. Ian Rhodes, Coomaren
Vencatasawmy, and Fraser Hatfield carried out much of the analysis of the
MacEurope data; I am grateful to Coomaren for supplying Figure 11.6, which
is taken from his M.Sc. dissertation. Mark Williams did much to improve our
understanding of the relationship between backscattering models and image
texture.
As far as the preparation of this book itself is concerned, we are grateful for
many important contributions. From DRA, Alan Blake provided the image
simulation and analysis software used throughout Chapters 5 to 9, Tony Currie
1. Many images from this book, as well as other examples, can be found at the InfoSAR
website: http://www.infosar.co.uk/.
the interferometric image shown in Figure 4.9, and Mark Williams the simulation
in Figure 5.8. Les Novak, of MIT Lincoln Laboratories, provided Table 10.2,
which is taken from one of his many informative papers on target detection and
recognition. A crucial input was that of Ian McConnell and Dave Stewart of N A
Software who processed all the examples of reconstruction filters and segmenta-
tion algorithms illustrated in Chapters 6 to 8 as well as Figures 12.11 and 12.12.
This involved a great deal of effort and was essential to the value of those chapters.
Special thanks must go to Mike Preston and Jiong Jiong Yu of SCEOS for
producing many of the figures and displaying remarkable patience when Shaun
realized he had asked for the wrong thing. Mike, in particular, devoted much time
to solving the numerous problems that arose in getting images to Chris in the
right form. Members of SCEOS were also willing to liaise with colleagues else-
where who were kind enough to allow access to data and images. Geoff Cook-
martin took care of the data provided by ESA-ESTEC that forms Figures 13.1
and 13.2. Jiong Jiong Yu did most of the hard work involved in transferring data
from CESBIO and DCRS, while Mike Preston took the lead with ESA-ESRIN.
Shaun's unreserved thanks must go to Mari Bullock, who, as SCEOS
secretary, not only typed his parts of the manuscript but also took care of so
many of the detailed tasks involved in putting all the material together. Her
initiative, care, and sheer devotion to getting the job done played a large part in
making it possible. She also managed to run SCEOS very effectively while
Shaun made himself incommunicado in order to write (which tells you some-
thing about how the organization really works).
A symptom of the coherence and sustained nature of the research program
from which this book springs is the sharing of staff between SCEOS, DRA, and
N A Software. DRA provided support for Dave Blacknell and Mark Williams
while they worked in Sheffield; both are now at DRA, as is Kevin Grover. Ron
Caves spent two years with N A Software before returning to Sheffield, and Ian
McConnell joined N A Software from Sheffield.
The extent of the SAR research program within SCEOS would have been
impossible without the continuing support of the UK Science and Engineering
Research Council and later the National Environmental Research Council,
when they took over responsibility for Earth Observation.
In preparing this manuscript we had tremendous help from colleagues in
other research groups who have been willing to give data, advice, and time when
we sought it, providing an excellent and refreshing example of the sense of
community within the field.
From CESBIO, Toulouse, we must particularly thank Thuy Le Toan for
advice that vastly improved Chapter 13 (though any blemishes are our respon-
sibility) and for providing data. Florence Ribbes and Oscar Casamian did an
excellent and difficult job in rendering color images into informative black-and-
white forms for our purposes. Their efforts became Figures 13.6 and 13.8. At a
deeper level, Thuy's insistence on understanding what we see in SAR images was
a continual spur to the spirit; she knows measurement is easier than knowing
what we have measured.
The Danish Centerfor Remote Sensing (DCRS), Department for Electromag-
netic Systems (EMI) of the Technical University of Denmark (DTU) provided help
in two very substantial ways. First, they provided the EMISAR images; our
thanks go to Soren Madsen for permission to use them and Erik Lintz for
comments on the text while providing that permission. Second, much of the
analysis of the EMISAR images was carried out by Jesper Schou of DCRS
during a four-month individual course at Sheffield. Thanks must go to his
supervisor, Henning Skriver, for instigating what turned out to be a very fruitful
visit and to Jesper, for the quality of his work at Sheffield and his willingness to
help in preparing material after his return to Denmark.
Colleagues from DLR who helped us include Richard Bamler and Man-
fred Zink. Richard supplied material on SAR processing and copies of unpub-
lished work done with Keith Raney (now of Applied Physics Laboratory, Johns
Hopkins University) on the linear system model for SAR imaging. Manfred gave
advice on SAR calibration.
We have several reasons to thank the European Space Agency. First and
foremost they supplied data. All the ERS images shown were provided by ESA.
Those from the Tapajos site in Amazonia were given to us under the auspices of
the TREES program organized by the Joint Research Centre of the European
Union, Ispra, whose support we would like to acknowledge. All other ERS images
were provided in the framework of the First ERS Announcement of Opportunity.
In addition, ESA provided access to data from two airborne campaigns. The
SAREX images used in Chapter 8 were collected in the framework of the ESA
South American Radar Experiment, while the NASA/JPL AirSAR data informing
Chapters 11 to 13 were acquired during the 1992 MacEurope campaign. In the
latter case, ESA was a formal partner of NASA and we would like to thank both
organizations. Other contributions to this book include Figures 13.1, 13.2, and
13.7, which are reproduced from ESA publications.
Individual ESA scientists also helped us greatly. Maurice Borgeaud of
ESA-ESTEC made available the figures and data that form Figures 13.1 and
13.2. Henri Laur, Giorgio Emiliani, and Jose Sanchez of ESA-ESRIN provided
the imagery and the calibration target analysis shown in Figure 2.9. In both
cases, this was not straightforward and we gratefully acknowledge the time and
effort expended on our behalf.
Most of the work we describe using data from the Tapajos site would not
have been possible without the support of colleagues in the Instituto Nacional
Pesquisas Espaciais (INPE), Sao Jose dos Campos, Brazil. Special thanks must go
to Corina Yanasse for many years of advice, discussion, and help. We are
indebted to Sidnei Sant'Anna, Alejandro Frery, and Corina for permission to
use Figure 13.7. This opportunity to display information on forest types using
texture measured from the Radarsat satellite arose only a week before comple-
tion of the manuscript during a visit to Sheffield by Corina and Sidnei, and we
were delighted by it. More generally, effective collaboration with INPE was
greatly aided by discussion and material help from Luciano Dutra and Thelma
Krug.
Shaun must give particular thanks to Rob Brooks, Ralph Cordey, and Paul
Saich of Marconi Research Centre, Great Baddow. Rob introduced me to the
fascinating field of SAR research. Ralph has been a continual source of critical
appraisal, encouragement, and financial support and he and Paul have been
welcome collaborators on several projects. MRC must also be thanked for
seconding Dave Blacknell to Sheffield for the duration of his doctoral studies.
This book would never have been completed (almost) on schedule with-
out the continual involvement of Susanna Taggart at Artech House in London.
She managed to make her demands in such a pleasant manner that it was
impossible to take offense.
When he started writing this book Chris had no conception of the
demands it would make on his time and energy. He especially wishes to thank
his wife Betty for her understanding and patience over the whole period, but
particularly the last few hectic weeks. In contrast, Shaun knew perfectly well
what was involved and went like a lamb to the slaughter. He has no words that
can do justice to the forbearance of his wife Sue.
About the Authors
After earning a degree in physics from Worcester College, Oxford, Chris Oliver
conducted his doctoral research into low-energy nuclear structure at Liverpool
University. In 1967, he joined the photon statistics group at the Royal Signals and
Radar Establishment (RSRE), Malvern [later the Defence Research Agency (DRA)
and now the Defence Evaluation and Research Agency (DERA)]. He was one of
the joint inventors of the technique of single-clipped photon-correlation spec-
troscopy exploited in the Malvern Correlator, for which he won the MacRobert
Award for Engineering Achievement in 1977.
Chris Oliver led a section at the DERA (Malvern) undertaking long-term
research on the extraction and processing of information from SAR images since
1981; from 1991 until 2000, he took on a consultancy role to free him from man-
agement responsibilities. He was appointed Visiting Professor of Physics at
King's College London in 1987. In 1996, he was invited to become a Visiting Pro-
fessor in Electronic Engineering at University College, London. He has also been
a Visiting Professor at "La Sapiensa," the University of Rome, in 1999 and 2001.
Chris Oliver has published in excess of 100 papers and holds 7 patents,
many in both Europe and the United States in addition to the United Kingdom.
His academy status was recognized within the DERA, where he was an Individ-
ual Merit Deputy Chief Scientific Officer. In 2000, he was appointed a Com-
mander of the British Empire (CBE) in Her Britannic Majesty the Queen's
Birthday Honors as recognition of his contribution to radar. Since his retire-
ment, Chris has set up InfoSAR Ltd. (www.infosar.co.uk), which offers consul-
tancy and training in SAR exploitation. The company has produced a SAR image
interpretation software suite, InfoPACK, which exploits and extends the princi-
ples described in this book and offers significantly greater sensitivity than any
others.
Shaun Quegan received his BA in mathematics in 1970 and M.Sc. in mathe-
matical statistics in 1972, both from the University of Warwick. After teaching
for several years (and running a large mathematics department) an ever-growing
interest in physics led him to undertake research into large-scale modeling of the
ionosphere and upper atmosphere at the University of Sheffield, leading to a
Ph.D. in 1982. He then joined Marconi Research Centre as a research scientist,
becoming Remote Sensing Applications Group Chief in 1984. This was a
fortunate opportunity to work with an excellent team involved in SAR research,
from which many fruitful long-term collaborations sprung, including that with
Chris Oliver. The offer of a post at the University of Sheffield in 1986 provided
an chance to build an academic SAR research group, whose activities have
flourished. In 1993 he was awarded a professorship, in the same year helping to
instigate the Sheffield Centre for Earth Observation Science and becoming its
first director. In this role he plays a central part in coordinating and developing
the wide range of remote sensing skills in the University and in bringing them
to bear on environmental science and application problems.
Contents
1. Introduction .................................................................... 1
This page has been reformatted by Knovel to provide easier navigation. vii
viii Contents
Introduction
In the last few years, high quality images of the Earth produced by synthetic
aperture radar (SAR) systems carried on a variety of airborne and spaceborne
platforms have become increasingly available. Two major leaps forward were
provided by the launch of the ERS-1 satellite by the European Space Agency
in 1991 and the advent of very flexible airborne systems carrying multifre-
quency polarimetric SARs, of which the NASA/JPL AirSAR provides perhaps
the single most influential example. These systems ushered in a new era of
civilian radar remote sensing because of their emphasis on SAR as a measure-
ment device, with great attention being paid to data quality and calibration.
This emphasis continues to play a major part in the development, deployment,
and application of current systems.
ERS-I was the first in a series of orbital SARs planned to have long
lifetimes and semioperational capabilities. The JERS-I, ERS-2, and Radarsat
satellite systems are currently in orbit, with ENVISAT planned for launch in
1999. By providing a long time series of accurate measurements of the backscat-
tering coefficient, these satellites allow dynamic processes to be observed over
most of the Earths surface, with impacts in many areas, such as vegetation
mapping and monitoring, hydrology, sea-ice mapping, and geology. The unique
capability of SAR to exploit signal phase in interferometry has given rise to
completely new tools for glaciology and the study of tectonic activity.
Because of the constraints imposed by deploying a radar in space, these
systems are simple, in the sense of using single frequencies and polarizations
with modest resolution. By contrast, airborne systems have been able to dem-
onstrate the advantages of having multiple frequencies and polarizations avail-
able. These advantages were further demonstrated, but from space, by the
SIR-C/X-SAR mission of the Space Shuttle. In addition, the AirSAR system
indicated that longer wavelength radars can have a special role in Earth ob-
servation due to their ability to penetrate vegetation canopies and interact
with structural elements of trees and the underlying soil. Much longer wave-
length systems are now in operation and promise to provide effective methods
for reconnaissance and remote sensing over heavily vegetated areas.
While civilian systems have concentrated on radiometric accuracy and
investigation of natural targets, the priority of military systems is the detection
and recognition of man-made targets (often vehicles) against a clutter back-
ground. The need for rapid reconnaissance has placed considerable emphasis
on airborne systems that can be deployed on demand. By its nature, the
military recognition task usually demands resolution better than Im and sys-
tems that can operate at long range, for survivability. These two requirements
enforce long synthetic apertures that, because airborne systems are preferred,
have needed the development of sophisticated motion-compensation (MOCO)
techniques.
While an enormous amount of effort has been expended on systems to
acquire SAR data, comparatively little has been expended on making the best
use of those data. Two lines of attack are needed to achieve this. One is the
investigation of the physical content of the data, by means of experiment,
observation, and modeling. The other is to examine the actual properties of the
data in light of what is known about the physics, the general properties of the
world, and the applications of interest, to identify and extract optimal estimates
of the information-bearing components of the data. This bridging role between
what we know about the data and how we best get at the information they
contain is the subject of this book.
Radar systems are capable of producing very high quality images of the
Earth, as demonstrated by the high-resolution image shown in Figure 1.1. In
order for this imagery to have value it must be interpreted so as to yield
information about the region under study. An image analyst soon learns to
recognize features such as trees, hedges, fields (with internal structure and
texture), shadows, and buildings, which encompass a range of natural and
man-made objects. However, it is a time-consuming task to extract the informa-
tion, particularly when large areas must be examined. Furthermore, there is no
guarantee of consistency between analysts or measure of the performance they
achieve. These limitations motivate the search for automatic algorithms to
derive the relevant information more quickly and reproducibly, or, in some
circumstances, more sensitively.
The need for automatic or semiautomatic methods becomes even more
pressing when polarimetric, multifrequency, and/or multitemporal images are
Figure 1.1 High-resolution (< 1m) DRA X-band SAR image of typical rural scene. (British
Crown Copyright 1997/DERA.)
Note that low-level algorithms are not sufficient; local details have to be
incorporated into some overall form of understanding. This demands methods
for overcoming defects in the detailed picture, for example, by joining edges
across gaps or merging regions where appropriate. On the other hand, sophisti-
cated high-level techniques are of no value if the underlying tools are incapable
of bringing out the information with sufficient sensitivity. The aim of this book
is to develop and relate the tools needed at both these levels. It is comprised of
the following six principal components:
Principles of SAR
Image Formation
2.1 Introduction
The basic geometry of a SAR is shown in Figure 2.1. A platform moving with
velocity K at altitude h carries a side-looking radar antenna that illuminates the
Earths surface with pulses of electromagnetic radiation. The direction of travel
of the platform is known as the azimuth direction; distance from the radar track
is measured in the range direction. For simplicity, we deal only with the case
where the antenna is oriented parallel to the flight line, not squinted.
Typically the antenna is rectangular with dimensions da X de, where a and
e denote azimuth and elevation, respectively. These dimensions are significant
because they determine the area illuminated by the antenna. As a rule of thumb,
the radiation from an antenna of length d spreads out over an angle
where K is the radar wavelength. This relation determines both the azimuth and
elevation beam widths, \\fa and \\fe; only the former is marked on Figure 2.1.
Wavelength is related to the radar carrier frequency^ (in Hz) by
^f0 = c (2.2)
where c is the speed of light. Although we will use (2.1) when describing basic
properties of SAR images, other definitions of beamwidth are often used. More
complete descriptions of the two-dimensional antenna pattern are also important
because high-quality SAR imaging must account for the spatial variation of the il-
lumination pattern [2,4]. Later sections in this chapter deal with these issues.
While the radar beam sweeps over a fixed scatterer X, the distance R
between the scatterer and the platform will vary symmetrically about its mini-
mum value R0. For the case of interest to us, the scatterer will be imaged at an
azimuth position determined by this broadside position of closest approach and
at a range determined by R0. The corresponding ground range R (see Section
2.3) is also marked on Figure 2.1.
Important parameters of the pulses that modulate the carrier signal are the
pulse length T^, the bandwidth B, and the pulse repetition frequency (PRF). Values
of these parameters for the ERS-I satellite SAR and representative values for an
airborne SAR are given in Table 2.1. Also recorded in the table is the polarization
configuration of the systems. This design choice affects the nature of the radar
Figure 2.1 The basic flat-Earth geometry of an airborne SAR.
Table 2.1
System Parameters for Satellite and Airborne SAR Systems
(a) (b)
Figure 2.2 (a) The spherical geometry appropriate to spaceborne SAR and (b) illustration of a
spreading pulse of duration 7p; also indicated is the elevation antenna pattern.
area received at a point on the ground depends on which part of the beam is
illuminating that point. The square law decrease of energy with range further
affects the power at the ground. Similar considerations apply to the received
signal. For the monostatic geometry of most interest to us, the transmit and
receive antenna patterns will normally be the same (but this may not be the case
for systems using active phased array antennas).
The radiated pulse spreads out as a growing shell of thickness cip. A point
object whose distance from the radar is R0 will therefore return a signal of
duration T\ that will arrive back at the platform with the time delay
Tj=IR0Zc (2.3)
must hold. As we shall see, the PRF cannot be chosen freely but must exceed a
lower limit imposed by the azimuth processing. A slight complication in the case
of spaceborne radar, where ranges are very long, is that the reception period does
not correspond to the previous pulse. For example, the reception period of the
ERS-I radar following the /Vth pulse corresponds to the (TV— 9)th preceding
pulse [4].
For a SAR, two different range measurements are commonly used. Cor-
responding to the slant range R0 is the distance from the subplatform track on
the Earths surface, known as the ground range Rg (see Figure 2.2(a)). The
ground range may also be referenced to some other locus, such as that of the
beam center or of the first near range pixel. Note that the ground range requires
a reference surface or projection for its measurement. We treat only the simple
case of a perfectly spherical Earth (of radius RE = 6,370 km). Other important
quantities marked on Figure 2.2(a) are the platform altitude h and the local
incidence angle 0.
Conversion between variables is straightforward by means of the relations
Rg=REl (2.5c)
7 = 0- a (2.5d)
r>
sin a = —sinO (2.5e)
RE + h
where the approximation in (2.5a) is for 7 small. The slant range R0 is inferred
directly from the time delay using (2.3), but given any one of R0, Rg) 0, a, or 7,
the other four can be found. For airborne platforms, h <^ RE, in which case
(2.5a,b) reduce to the flat-Earth approximations
R
R0 = -L- (2.6a)
sin0
Rl = h2 + Rj (2.6b)
rt=\/B (2.7)
rs = cllB (2.8)
The ground range resolution r is the change in ground range associated with a
slant range change of rs. Using (2.5a-c) yields
V,^'- ** - 1 (29)
r ^ ( ^ +/;)sin7 sin 9
so the ground range resolution varies nonlinearly across the swath. This can have
important consequences for image properties, particularly for spaceborne or
short-range airborne systems, as will be seen later in this chapter.
To illustrate these relations, we shall use parameters from ERS-I, as
summarized in Table 2.1. The time delays for near and far range are selected so
that the incidence angle 0 varies from 19.3 to 26.4 degrees between near and far
range. This gives a total ground swath of 99.3 km; the equivalent slant swath-
width is 38.6 km. Since the transmitted signal bandwidth is 15.5 MHz, the slant
range resolution is fixed at 9.68m, but the ground range resolution degrades
from 21.8m at far range to 29.3m at near range.
For short-range airborne systems, the resolution change across the swath
can be much greater. As an example, a system operating with incidence angles
between 15 and 60 degrees would have ground range resolutions at far range
that are a factor 0.3 of the near range resolution. In image terms, such changes
in resolution can have serious consequences. Features that are clearly distin-
guishable at far range can become nearly invisible at near range. Within distrib-
uted targets, such as agricultural fields, there are fewer measurements available
to estimate target properties at near range. In this airborne example, fields at
near range would contain only 30% as many pixels as at far range, despite having
the same size on the ground.
These effects are less significant for long-range airborne systems, such as
those of most importance in military applications, which operate over a range
of incidence angles much nearer grazing angle. For example, if an aircraft is
flying at an altitude of 10 km the incidence angle at a slant range of 100 km is
84.3 degrees. If this is near range and the ground swath is 10 km wide, the
incidence angle at far range is 84.8 degrees, corresponding to a change in ground
range resolution of—0.1% across the swath. The difference between slant range
and ground range resolution in this case is less than 0.5%.
When displaying SAR data, there is therefore a choice between handling
slant range or ground range data. In slant range, each pixel corresponds to the
actual SAR measurement (subject to any calibration corrections made in the
processing), but the geometry is distorted relative to a map projection. Alterna-
tively, the data can be resampled to ground range so that the geometry is correct
in the map projection. Assuming a sensible interpolation scheme (this is not
always the case [7]) the interpolated measurements will have the correct values,
but considerable spatial correlation may have been introduced between pixels.
This correlation will be higher at near range than at far range simply as a
consequence of there being fewer actual measurements per unit length at near
range.
Data from ERS-I illustrate these effects well. These data are available
both in slant range complex (SLC) format, with a slant range spacing of 7.9m,
and in the standard precision image (PRI) format, which is resampled to a
ground spacing of 12.5m. The geometry of these two types of data is shown in
Figure 2.3, where the same region on the ground is depicted as derived from
(a) SLC data, (b) SLC data in which pixels have been averaged in blocks of five
along azimuth, and (c) PRI data. Azimuth here runs from the top to the
bottom of the page. The SLC pixel spacing is 4m in azimuth, so the range and
azimuth pixel dimensions are very different on the ground, leading to the
obvious distortion. The PRI data are geometrically correct, while the azimuth
averaged data are close to correct geometry. (Figure 2.3(b) is discussed further
in Section 4.9.)
The spacing in the PRI data is 57.3% of the ground range solution at far
range and 42.7% at near range. The higher degree of correlation at near range
is not visually obvious but can be seen in the range correlation coefficients
shown in Figure 2.4. The four panels in this figure show intensity correlation
coefficients (see Section 4.6) calculated from slant range data at near (Figure
2.4(a)) and far (Figure 2.4(b)) range and the corresponding quantities measured
from PRI data (Figure 2.4(c,d)). For slant range, the system-induced correlation
is independent of location in the image, having values around 0.25 at lag one
and effectively zero thereafter. For the PRI data, at near range the correlation
coefficient at lag 1 exceeds 0.6 and is near 0.2 at lag 2; at far range, the
corresponding value at lag 1 is ~0.5 and the lag-2 correlation is indistinguish-
able from the noise.
(O)
(a)
Figure 2.3 ERS-1 images of an agricultural region derived from (a) SLC data, (b) SLC data
averaged (in intensity) in blocks of 5 pixels along azimuth, and (c) PRI data. All
images are shown in amplitude (see Chapter 4). (Copyright ESA, 1992.)
Corr. Coeff.
Corr. Coeff.
(a) (b)
Corr. Coeff.
Corr. Coeff.
(C) (d)
Figure 2.4 Intensity correlation coefficients calculated in the range direction for ERS-I data
from an agricultural area: (a) SLC, near range; (b) SLC, far range; (c) PRI, near
range; and (d) PRI, far range.
In many SAR systems the transmitted waveform p(t) is of the form
where co0 is the carrier frequency of the radar expressed as radians/s" 1 . The phase
of the signal is therefore
and the instantaneous frequency (given by the time derivative of the phase) is
Since the frequency changes linearly with time, this sort of signal is
referred to as linear frequency modulation (FM) or chirp, with an FM rate of
(3/TT H Z / S " 1 . Plots of the phase, the real part, and the instantaneous frequency
of the complex envelope exp(— ifit2) against distance are shown in Figure
2.5, where time has been converted to distance units using the relation x =
ct. The total frequency sweep or bandwidth is seen to be
(a)
r(x)[v]
(b)
f(x) [cycles m2]
(C)
Figure 2.5 (a) Plots of phase, (b) real part of signal, and (c) frequency of signal for the complex
envelope of a chirp signal with bandwidth B = 100 MHz and pulse length Tp = 10~6
sec. Time (t\ is converted to distance (x) units by the relation x = ct.
where
Since imaging radars are designed so that the time-bandwidth product BTp is
large, a good approximate solution is rt ~ XlB, which is the basis for (2.7).
The ratio of the resolution after processing to the original pulse length is
known as the compression ratio. Using (2.7) we see that
We will often refer to this as the ideal form of the SAR PSF. Plots of the pulses
described by (2.14b) and the ideal pulse (2.16) are given in Figure 2.6, where,
as previously, we have used the parameters B = 100 MHz and Tp = 10~6 sec.
We have also converted to distance units [m] by the replacement x = ct and
scaled each plot by XITp.
The effects of matched filtering when there is more than one target present
are illustrated by Figure 2.7. The top panel shows the real part of the complex
envelope of the returned signal, while the bottom panel plots the amplitude of
the output signal after filtering. In this simulation, the scattering was taken to
be from two point objects of amplitudes 1 and 2 at positions x = — 5 and x =
5. Although they are invisible in the uncompressed envelope, they are clearly
Figure 2.6 Plots of the exact (dashed curve) and ideal (solid curve) compressed chirp pulse
for B = 100 MHz and ip = 10~6 sec, with conversion to distance units using
x = ct
resolved at the right positions and with approximately the right magnitudes after
the processing.
This treatment has ignored two important processing aspects. First, the
point target response shown in Figure 2.6 has significant sidelobes. These are
often reduced by applying a weighting to the correlation operation described by
(2.14a). This tends to degrade the resolution from the ideal value of XlB.
Second, in modern digital systems the compressed signal is sampled (at the
Nyquist rate or above), so each pulse yields a vector of measurements. Sub-
sequent operations on the compressed pulse are therefore inherently digital.
Note also that it is not mandatory to use a frequency-modulated pulse to
gain good range resolution. Short simple pulses can be used instead, but very
(a)
(b)
Figure 2.7 (a) The real part of the uncompressed signal envelope when there are two scat-
terers present and (b) the output signal after pulse compression. The scatterers
are of amplitude 2 at x = 5 and 1 at x = -5.
The pulse compression in the range direction is often carried out on a pulse-by-
pulse basis and is a standard form of processing used in many sorts of radars.
The particular technology of SAR (and the source of much of its complexity)
manifests itself in the azimuth direction.
After range processing on a single pulse, the measurement in each range
gate contains a contribution from each scatterer in an azimuthally extended and
curved strip whose width is controlled by the azimuth beamwidth. We ignore
the range curvature and treat the strip as straight (see Figure 2.1); for a full
treatment see [1,4]. A real aperture radar simply outputs the power measured at
each range to produce an image whose azimuth resolution would be R^a, where
R0 is the slant range at the position of closest approach and \\ta is the azimuth
beamwidth. A single point scatterer at range R0 would effectively be smeared
across a distance R0^a in the azimuth direction. Using Table 2.1, a quick
calculation indicates an associated resolution of the order 5 to 6 km for ERS-I.
Even for an airborne X-band (3-cm wavelength) radar using a 1-m antenna at
a range of only 15 km, the azimuth resolution is 450m. These values greatly
exceed the available range resolution and would significantly hamper the value
of radar for reconnaissance and mapping.
The essence of the SAR processing is the exploitation of the nearly quad-
ratic range variation of a point scatterer as the beam passes over it. This variation
is easily calculated for the airborne flat-Earth geometry shown in Figure 2.1.
(The corresponding treatment for an orbital SAR is given in Appendix 2A.) The
slant range R0 defines a line on the ground (see Figure 2.1). For a scatterer Xon
this line we use a coordinate system whose X-axis points along the direction of
travel and whose origin is fixed at X If the azimuth coordinate is x, it is easy to
see that the distance R between the platform and Xsatisfies R2 = R01 + x2. For
a narrow beam, X is only illuminated when x <^ R0, in which case
R - R0 + — (2.17)
2R0
^x) = _ ^ _ 2 ^ i (2J8)
X XR0
^ = - ^ (2.19)
dx KR0
Applying results for the chirp pulse from Section 2.4 we therefore have a spatial
bandwidth given by
where we have used the relation \\fa = X/da, and an associated spatial resolution
ra = ^ L (2.21a)
ra=~^-^- (2.21b)
' RE + h 2
where RE is the Earths radius, and h is the platform altitude (see Appendix 2A).
This critical result (2.21) shows that the SAR azimuth resolution is half
the real antenna dimension in the azimuth direction (and better than this in the
orbital case) and independent of range and wavelength. For ERS-I, we therefore
expect azimuth resolutions of the order of 4.45m, while aircraft can achieve
resolutions of the order 0.5m. Notice that these resolutions assume a fixed
pointing direction of the antenna, broadside to the platform track, so the SAR
is operating in strip map mode. A higher resolution is obtainable if the beam is
steered (normally electronically) to keep a target in the beam for a larger time
and thus to form a longer synthetic antenna. This is known as spotlight mode
SAR and is capable of improving the azimuth resolution considerably, at the
expense of spatial coverage [3].
The azimuth processing is often described by the term SAR focusing, and
we will refer to the coefficient of the quadratic phase term (3 as the focusing
parameter. SAR focusing is equivalent to synthesizing a large antenna, of length
Ds given by (2.20), from all the pulses impinging on a given scatterer as the beam
passes over it. Since v|i^ is proportional to X, this means that Ds is linearly
proportional to both the slant range and the wavelength. Hence, increasing the
range or wavelength gives a longer effective antenna, which is the fundamental
reason why SAR produces resolution independent of either range or wavelength.
There is, however, an associated cost. The data are sampled in the azimuth
direction, and the Nyquist condition means that the sample spacing must not
exceed the resolution length [2,5]. This constraint imposes a lower limit on the
PRF, as already noted in Section 2.3. Hence, the number of samples that must
be stored and processed is directly proportional to both range and wavelength.
The above formulation of the azimuth processing in terms of distance
illustrates that the motion of the platform is not fundamental, only that samples
are gathered coherently along the synthetic aperture. However, for a moving
platform, it is common to make use of equivalent temporal expressions, by
means of the relation x = Vt. These are summarized in Table 2.2. Of particular
note is the expression
/ a = f f = ~ [Hz] (2.22)
2TT at KR0
Table 2.2
Azimuth Processing Parameters (Flat Earth)
The prior simplified treatment dealt with the response of an isolated point target
mainly as a means of discussing system resolution. However, radiometric prop-
erties of the image are also important, and we now consider how the radar
reflectivity of the scene is translated into image brightness. In doing so, we will
provide a more general and flexible description of the SAR processing than that
given in Section 2.5.
For a point target, the most general description of its scattering behavior
at a single frequency is provided by the complex scattering amplitude, Spq, which
we will also refer to as complex reflectivity. This quantifies the scattering into
polarization state p of an incident plane wave with polarization q through the
relation [14]
Polarimetric data are treated in Chapter 11. In the early chapters of this book,
where we are mainly dealing with single-channel data, we will omit the sub-
scripts and write the complex scattering amplitude and RCS simply as 5 and o\
Consider now a point scatterer with complex reflectivity S at azimuth
position x'and slant range R when the platform is at azimuth position x. If a
pulse is emitted at time t= 0, after pulse compression the received signal is
Here K' (R) accounts for the elevation antenna pattern, processing gain during
the range compression, and the range spreading loss, which is proportional to
Rr2; while the two-way amplitude azimuth antenna pattern is a(§). The term
hr(t — 2RJc) describes the time delay in the range PSF (see (2.14)) and hence
allows the target to be positioned in the range direction. The terms K' (R), hr(t
— 2RJc), and a(x/R) vary much more slowly with range than the phase of the
exponential term AmRIX. We therefore make the simplification that in these
three terms, R can be replaced by R0, the range of closest approach.
The range R has a variation with along-track separation x — x described
by (2.17) and can be written
R= R0+^(x-xf
AlT
where (3 is the focusing parameter. Expressions for (3 in the spaceborne and
airborne cases are given by (2A.1) or Table 2.2, respectively. With the further
assumption of negligible range migration (or its correction in the processing) we
can write the received signal appropriate to range R0 as
where
= K(R0)S(x)*k(x) (2.27)
where
We can regard k(x) as a prefilter containing two effects, the beam weighting given
by the polar diagram a(<$) of the antenna in the azimuth direction and the range
variation or Doppler shiftterm caused by relative motion of the platform and the
scatterer [8,16,17].
In treating S{x) as continuous we move from point to distributed targets.
A key property of such targets is that the phase of the scattered field has a
correlation length that is much shorter than the resolution of the radar meas-
urement. At scales comparable with this correlation length the target has a
well-defined RCS and we can define a differential backscattering coefficient a° as
where A^4 is the elemental area or facet over which the phase is coherent
(effectively constant), Pi is the incident power per unit area, and Ps is the power
per unit area scattered from the facet, observed at range R. However, since
different facets contribute independent phases, the observed electric field is
dominated by interference effects, giving rise to the phenomenon known as
speckle (see Chapter 4). For distributed targets, the quantity of interest is the
average value of a ° within a pixel; from (2.29), this requires normalization by
the illuminated area (see Section 4.6 for a fuller discussion). A target is described
as uniform if a° is constant within it.
The range-compressed raw data E(x) is digital, being built up sequentially
at each range by gathering pulse returns and saving them in memory. As we have
seen, the azimuth processing is a correlation with this stored data and, hence,
can be represented as a second filtering operation, giving the output
where
The quantity ha(x) is known as the azimuth PSF. Typically the SAR processing
filter l(x) will contain an amplitude weighting term W(x/R0), but it will always
have as its main purpose the removal of the quadratic phase term [8]; that is,
l(x) will have the form
Hence,
To relate (2.30) and (2.33) to earlier results, note that a unit strength point
target can be represented as S(x) = 8(x), where b(x) is the Dirac delta function,
so
For a beamshape that is uniform over the beamwidth \\fa and zero else-
where and the same processor weighting, we can set a($) = W(<$>) = rect(4>/i|; J .
Then the point target response ha(x) gives the same result as (2.14) except for
the range dependent term 13,RQ). The associated values for the FM rate and
aperture length are 2/XR0 and R0\\fa, as given in Table 2.2.
To complete the system description, we must include system noise. The
model for the azimuth processing at range R (see Figure 2.8) then becomes
The range-dependent system terms found in (2.26b) are absorbed into the
coefficient C(R) whose magnitude is given by
^ - ^ i
where PT is the peak transmitted power, GR is the one-way power gain of the
antenna pattern at the specified range, L allows for system losses, and GP is the
processing gain due to the range compression. The system noise appearing in
each range gate after the range compression is unaffected by the antenna pattern
or the Doppler shifts present in signals scattered from real targets and so appears
in the system diagram after the prefiltering by k(x). It would be expected to be
white, zero-mean, and Gaussian. However, because the noise in the image has
passed through the range compression, it will usually exhibit range dependence
because of corrections carried out in the processing.
Although the voltage equation (2.34) is applicable to any form of target,
for single-channel observations of distributed targets it is more useful to deal
with the backscattered power. As will be explained in Chapter 4, the mean
power in the received signal from a uniform distributed target with backscatter-
ing coefficient a ° can then be written
where the two-dimensional SAR PSF, h(x, R), can be separated into the slant
range and azimuth PSFs as
This implies that the PSF is independent of position, which is the design aim
of most SARs. For a less well-designed system, the PSF may vary slowly with
position, but (2.38) still provides a useful and meaningful local approximation.
Equations (2.14) and (2.33) show that the azimuth and range PSFs can
be derived theoretically for known beam patterns and processor weighting and
(2.38) implies that the complete SAR PSF is known if the range and azimuth
PSF are known separately. In practice, uncertainties in the real antenna pattern
and effects in the processing motivate investigation of the performance of the
system by direct measurements of the (local) PSF using pointlike calibration
devices. Figure 2.9 shows an example of such a measurement together with cuts
through the peak response in both range and azimuth. This figure was provided
by ESA-ESRIN, Frascati, and is based on measurements by the ERS-I satellite
from a calibration target [18]. The sample spacing is 7.9m in slant range and
4m in azimuth.
2.7 Calibration
Applications of SAR data can exploit two aspects of SAR images. On the one
hand, we have image structure making up the features we see. This is the critical
issue in SAR as a mapping or target detection device. In this case, resolution and
geometric accuracy are the important concerns, at least for single images;
(a)
Amplitude
(b)
Figure 2.9 Measured ERS-1 PSF using a calibration target shown (a) in the PRI amplitude
image (Copyright ESA, 1995); (b) two-dimensional plot of the amplitude of the SLC
PSF and its corresponding three-dimensional representation; (c) profiles through
the peak response in the range and azimuth dimensions, interpolated by a factor
of 2.
Range Profile
(C)
dB
Sample
Azimuth Profile
dB
Sample
radiometry and phase integrity only become important when feature detection
relies on methods that combine information from different images. On the
other hand, SAR is a measuring instrument providing estimates of the complex
radar reflectivity or backscattering coefficient of the scene. Here radiometry and
phase preservation are critical. Without accurate estimates, there is no basis for
meaningful comparisons of target properties measured at different positions,
incidence angles, polarizations, frequencies, or times.
The process of converting measured voltages in the in-phase (I) and
quadrature (Q) channels of the radar into accurate estimates of complex reflec-
tivity normally requires very careful calibration. A full treatment of this subject
is beyond the scope of this book, but an excellent review article and reference
list is provided in [19]. However, the essence of the problem is contained in
(2.34) and (2.36), where it can be seen that the effects of noise, azimuth
processing (as described by ha(x)), and the gain term C (which contains range
antenna pattern, propagation, and range compression terms; see (2.35)) must
all be accounted for in estimating the complex reflectivity or backscattering
coefficient from the measured voltages.
For single-channel data, spatially dependent image effects associated with
the gain term C include:
The first of these is easy to correct, since the range is known from the
timing, and any worthwhile processor would remove this effect. The latter
effects are both often obvious, particularly in airborne images. They are fre-
quently removed by ad hoc methods assuming thatCT°should show no system-
atic variation with range or azimuth throughout the image of interest [20,21].
In doing so, real physical effects are lost (such as variation of cr° with incidence
angle or genuine variation in scene content), and such methods should only be
considered as a last resort. They can provide a form of relative calibration within
a single image but are of little value if quantitative values for a° are desired.
A much more satisfactory approach to removing antenna pattern effects is
to make use of accurate estimates of the range antenna pattern and the elevation
pointing angle of the beam center, if they are available. Acquiring this informa-
tion often relies on measurements using targets of known RCS spaced across the
swath. The accurate deployment of such devices is a time-consuming task, but
failure to provide sufficient information to allow the correction of antenna
effects has been a repeated cause of problems in extracting meaningful informa-
tion from SAR data [22].
Given that the system gain term C can be accurately derived at each
position in the image, the effects of processing need also to be considered. In
(2.36) we see that processing acts as a multiplicative term affecting the estimate
of a° from the measured power; this is often referred to as processor gain [17].
As noted in [19], the SAR processor may perform many functions in addition
to the basic SAR correlation (e.g., slant range to ground range conversion, look
extraction, and antenna pattern correction) and errors in any of these may affect
the processor gain. Different processors will normally give different outputs even
on the same data. Probably the most comprehensive and successful treatment of
this problem has been that applied to data from the ERS-I satellite. Relatively
simple corrections taking account of the processing strategies at different proc-
essing centers are available and have been shown to produce great consistency
in the measurements [23].
Detailed structure is important when imaging point objects, and this is
dependent not just on the processor gain (see (2.36)) but on the PSF itself (see
(2.34)). This distinction is important, since the latter is much more sensitive to
errors in the processing than the former. As a result, the appropriate method by
which to measure a depends on the quality of the SAR focusing. For well-
focused data, a can be measured directly from the peak response of the point
target. However, errors in the focusing (which can occur particularly as a result
of unknown or uncorrected platform motions; see Chapter 3) have motivated
the development of methods based on the integrated point target response.
These are comparatively insensitive to the accuracy of the focusing parameter
[24,25] and are therefore often preferred, particularly when dealing with air-
borne data. When high-performance autofocus algorithms are used (see Chap-
ter 3), this is a less telling consideration.
In many cases, precise measurements of a for an unknown point target are
not of interest. Instead, point targets with known a are used to infer the
quantitative behavior of the system term Cha{x) in (2.34). It can be seen,
however, from (2.36) that, for measurements of cr°, the important quantity is
ICI2 times the energy in the SAR PSF. Hence, where calibration is carried out
for the purposes of measuring the properties of distributed targets, integrated
energy methods are to be preferred.
The interplay of point and distributed targets is also important when faced
with the extra problems associated with calibrating polarimetric data. This will
be discussed in Chapter 11.
2.8 Summary
This chapter dealt with a number of key concepts in SAR image formation. The
geometry of SAR images creates possible problems for display and for data
analysis. The natural coordinate system of a strip-map SAR is in terms of time
delay and the azimuth position of closest approach (zero Doppler frequency).
This implies a distortion compared to a map projection that gets worse in a
nonlinear fashion toward near range as incidence angles decrease. Resampling
to a map projection leads to varying resolution and interpixel correlation across
the swath, whose effects may be very significant.
A central concept is the SAR PSF, which describes how a point scatterer is
imaged by the SAR and how energy due to the point target becomes spread into
surrounding pixels. In the ideal case, its properties are simply linked to basic
system parameters. The PSF provides the basis for defining SAR resolution. We
have here adopted fairly simple measurements of resolution, and many others
are possible, but any such measure is essentially a way of describing the width
of the PSF. Other manifestations of the PSF are sidelobes and ambiguities,
which provide known and quantifiable effects on image quality. In addition,
there are image defects that are hard to classify but are often visually obvious,
particularly in airborne images. These include azimuth banding (which can be
caused by antenna yaw or varying azimuth pixel size), interference by alien
radars, and fringes caused by multipath effects, due to reflections off the plat-
form. Such effects can often be very difficult to deal with satisfactorily and are
not considered further (but see [26]).
The SAR PSF is fundamental in the linear system description of SAR
imaging, which provides the basis for relating the imaging of point scatterers to
that of distributed scatterers and for dealing with system noise. The PSF also
plays a key role in the use of SAR as a measuring instrument attempting to
provide accurate estimates of the RCS (a) for point scatterers or the backscat-
tering coefficient (<r°) for distributed scatterers because accurate estimates rely
on image calibration, which is often provided by point scatterers of known RCS
deployed in a scene. To characterize targets comprising a few point scatterers,
typical of man-made objects, the PSF is seen to be the important quantity, while
for distributed targets the energy in the PSF is more useful.
References
[1] Curlander, J. C , and R. N. McDonough, Synthetic Aperture Radar: Systems and Signal
Processing, New York: J. Wiley & Sons, 1991.
[2] Ulaby, E T., R. K. Moore, and A. K. Fung, Microwave Remote Sensing: Active and Passive,
VoIs. 1-2, Reading, MA: Addison-Wesley, 1981, 1982.
[3] Carrara W. G., R. S. Goodman, and R. M. Majewski, Spotlight Synthetic Aperture Radar:
Signal Processing Algorithms, Norwood, MA: Artech House, 1995.
[4] Bamler, R., and B. Schattler, "SAR Data Acquisition and Image Formation," SAR Geocod-
ing: Data and Systems, G. Schreier (ed.), Karlsruhe: Wichmann, 1993, pp. 53—102.
[5] Kovaly, J. J., "Radar Techniques for Planetary Mapping with an Orbiting Vehicle," Synthetic
Aperture Radar, J. J. Kovaly (ed.), Dedham, MA: Artech House, 1976, pp. 32-54.
[6] Elachi, C., Spaceborne Radar Remote Sensing: Applications and Techniques, New York: IEEE
Press, 1988.
[7] Quegan, S., "Interpolation and Sampling in SAR Images," IEEE Trans, on Geoscience and
Remote Sensing Vol. 28, 1990, pp. 641-646.
[8] Brown, W. M., and C. J. Palermo, Random Processes, Communications and Radar, New
York: McGraw-Hill, 1969.
[9] Rihaczek, A. W , Principles of High Resolution Radar, New York: McGraw-Hill, 1969.
[10] Papoulis, A., Signal Analysis, New York: McGraw-Hill, 1977.
[11] Woodward, P M., Probability and Information Theory, with Applications to Radar, Oxford:
Pergamon Press, 1953.
[12] Raney, R. K., "Considerations for SAR Image Quantification Unique to Orbital Systems,"
IEEE Trans, on Geoscience and Remote Sensing, Vol. 29, 1991, pp. 754-760.
[13] Laur, H., and G. M. Doherty, "ERS-I SAR Calibration: History and Results," CEOS
Workshop on SAR Calibration, 1992.
[14] Van ZyI, J. J., and F. T. Ulaby, "Scattering Matrix Representation for Simple Targets," Radar
Polarimetry for Geoscience Applications, E T. Ulaby and C. Elachi (eds.), Norwood, MA:
Artech House, 1990, pp. 17-52.
[15] Jin, M. Y., and C. Wu, "A SAR Correlation Algorithm Which Accommodates Large Range
Migration," IEEE Trans, on Geoscience and Remote Sensing, Vol. 22, pp. 592—597.
[16] Harger, R. O., Synthetic Aperture Radar Systems, Theory and Design, New York: Academic
Press, 1970.
[17] Raney, R. K., and R. Bamler, "Comments on SAR Signal and Noise Equations," Proc.
IGARSS 94, Pasadena, 1994, pp. 298-300.
[18] Woode, A. D., Y-L. Desnos, and H. Jackson, "The Development and First Results from
the ESTEC ERS-I Active Radar Calibration Unit," IEEE Trans, on Geoscience and Remote
Sensing, Vol. 30, 1992, pp. 1122-1130.
[19] Freeman, A., "SAR Calibration: an Overview," IEEE Trans, on Geoscience and Remote
Sensing, Vol. 30, 1992, pp. 1107-1121.
[20] Quegan, S., C. C. F. Yanasse, H. de Groof, P. N. Churchill, and A. J. Sieber, "The
Radiometric Quality of AgriSAR Data," Int. J. Remote Sensing, Vol. 12, pp. 277-302.
[21] Begin, D., Q. H. J. Gwyn, and F. Bonn, "Radiometric Correction of SAR Images: a New
Correction Algorithm," Int. J. Remote Sensing, Vol. 8, 1987, pp. 385-398.
[22] Yanasse, C. C. E, S. Quegan, and R. J. Martin, "Inferences on Spatial and Temporal
Variability of the Backscatter from Growing Crops Using AgriSAR Data," Int. J. Remote
Sensing, Vol. 13, 1992, pp. 493-507.
[23] Laur, H., P. Bally, P Meadows, J. Sanchez, B. Schaettler, and E. Lopinto, "ERS SAR
Calibration: Derivation of the Backscattering Coefficient CT° in ESA ERS SAR PRI Prod-
ucts," ESA Document ES-TN-RS-VM-HL09, Issue 2, Rev. 4, 1997.
[24] Gray, A. L., P. W Vachon, C. E. Livingstone, and T. I. Lukowski, "Synthetic Aperture Radar
Calibration Using Reference Reflectors," IEEE Trans, on Geoscience and Remote Sensing, Vol.
28, 1991, pp. 374-383.
[25] Ulander, L. M. H., "Accuracy of Using Point Targets for SAR Calibration," IEEE Trans.
Aerosp. Electron. Syst., Vol. 27, 1991, pp. 139-148.
[26] Massonet, D., "Radar Image Quality White Paper," CEOS SAR Cal/Val Working Group,
1991.
given that X is broadside to the satellite at time 0. The distance R between the
satellite and X is
R cosyS^
* 2R,(RE+h)
m_^_^RE{RE+h)cos,W {2A2)
XX R0
2TT dt X R0
1. In most other places in this book we use a coordinate system for which x is in the azimuth
direction.
width on the ground is approximately R0^a and that the velocity of the beam
on the ground is V— RE£l cos 7. Hence the illumination time is
M, = _^L (2A.4)
' V RED, cosy
We can now apply the results for the chirp pulse from Section 2.4. The
azimuth (Doppler) bandwidth is
with associated time resolution HB. In this time the beam moves a ground
distance VlB, so the azimuth resolution on the ground is
RE
r = cosy — (2A.6)
1
* RE+h I^
Using the approximation \\ta = \lda and noting that 7 is small, we find
ra = - A - cos Y ^ - - - ^ S - i (2A.7)
" 7?£ + A 2 RE + h 2
These results should be compared with (2.17) to (2.22) and Table 2.2.
3
3.1 Introduction
• Radiometric distortion;
• Image defocus;
• Geometrical distortion;
• Radiometric distortion;
• Increased sidelobe levels.
3.2 Defocus
A striking example of image defocus is shown in Figure 3.1. This image was
obtained with the DRA airborne X-band SAR at a range of about 40 km using
the nominal value for the focus parameter with no conventional M O C O
applied. The physical reason for the defocus is the occurrence of unknown
fluctuations in the microwave path length between the antenna and the scene.
One possible cause is a variation of atmospheric refractive index [5,6], which
can be significant for high-resolution systems with long atmospheric paths. An
important feature of atmospheric effects is that it is not enough to know the
Figure 3.1 Defocused image formed by processing data with the expected focus parameter.
sensor track perfectly; only a measurement of the actual microwave path will
suffice. A second cause of fluctuations in path length is the failure of the sensor
to follow a defined trajectory, to within a small fraction of a wavelength, over
the entire synthetic aperture. We here analyze the implications of these un-
known path length fluctuations and compare means for correcting for them by
autofocus processing. In practice, there is no need to distinguish the actual
mechanism, since both give rise to apparent range fluctuations. However, the
following discussion will be couched in terms of unknown trajectory variation
since this is usually the dominant contribution.
Defocus stems from the erroneous assumption that the antenna phase
center traversed a straight path at known velocity during image formation. Any
departure from this defined track degrades the reconstruction. In Figure 3.2 we
illustrate the impact of such an irregular track on the difference between the
actual distance from a scatterer O and that if there were no unknown motions.
As shown in Chapter 2, focusing the SAR involves knowing this distance
accurately and correcting for the associated phase delay.
Suppose that an object, O, is imaged by a sensor traveling along the actual
track AC from origin A (O, 0) at time 0 to position C (x, y) with range R at
time t and broadside range R0, as shown in Figure 3.2. Approximate displace-
ments x and y, can be obtained by expansion up to the order of acceleration in
each direction such that x ~ (V + AV)t + aj2/2 and y ~ Vrt + art2/2
where Vand Vr are along-track and radial velocities, ax and ar the corresponding
accelerations, and A Kan along-track velocity mismatch. The expected track is
AB, with the sensor ending at position B (x0, 0) where X0 = Vt, with range R.
The difference between the expected and actual ranges would be given by Ai?
Figure 3.2 A comparison of path length for expected and actual tracks.
p v2 v
From (3.1), the instantaneous frequency, or phase gradient, would be given by
In sideways-looking SAR the image is formed at the position where this fre-
quency is zero. Thus a nonzero value of the Doppler centroid term, a, shifts the
image in azimuth, resulting in azimuthal distortion. If the axis of this synthetic
aperture is not aligned with the real antenna there will also be a resultant
modulation in the power in the SAR image. Thus, changes in a yield both
geometric and radiometric distortion. However, a nonzero value of a does not
degrade the focus directly. This is caused by an error in the quadratic focus term,
so Ap T^ 0. Note that indirect degradation can be caused if the full bandwidth
is not retained.
In the following sections we will consider methods for overcoming these
deviations from the expected trajectory. In the first stage, autofocus processing
is used to determine a correction to the quadratic focus term. In the second, this
is associated with a time-dependent phase correction that resamples the phase
of the received signal to match that of the reference. We will demonstrate that
this approach is capable of yielding focused, geometrically undistorted imagery.
sin — — pAfA T
A{k) = exp[nr(aMAT-kkT + (W2AT2)] LA/A71M Jj {U)
s i p A r
TO" ^ Jj
Fit
Figure 3.3 A comparison of estimated phase gradient, co(>), with best fit and expected
reference as function of time.
where M is the number of samples of separation A T in the aperture T. The
second part describes an envelope that has a peak when k = (3Af2A 7/2, from
which the focus parameter can be derived.
As already noted, PPP, PG, and PD are coherent techniques that operate
directly on the complex received signal. Both MLR and CO, on the other
hand, depend on the intensity of the compressed image and are therefore
incoherent methods. Following Chapter 2, azimuthal compression is performed
by crosscorrelating the received signal with a reference of the form
Eref(j) = exp[—f(a/A7" + (3/ 2 Ar 2 )] over aperture T. This differs from the
unperturbed reference in adopting unknown values of the Doppler centroid
and focus parameters, a and (3. The compressed image intensity is then formed
by taking the square modulus of the received field. Multilook registration and
contrast optimization exploit different aspects of this compressed image.
The MLR technique detects the difference between the actual and speci-
fied focus parameters in two (or more) subapertures. As shown in Figure 3.4,
the gradient of the reference is determined by its specified focus parameter, (30.
However, if the actual focus parameter is (3, the compressed image will not form
at the origin but will be shifted by A tl2 in each subaperture, where A(3 ~
2(3A^/ T. The second subaperture yields an equivalent shift for the same scatterer.
The displacement between the two image positions (= A ^) is directly measur-
able and is caused by a change in the Doppler frequency between the subaper-
tures, where At = Aa/2(3 from (3.3), so the mismatch in the focus parameter
Actual
Expected
Figure 3.4 Effect of focus parameter on image formation position for two-look MLR.
is related to the change in the Doppler centroid by A (3 ~ Aa/ T over the
aperture T.
In the CO method, the focus parameter (3 is varied for the complete
aperture Twith consequences illustrated in Figure 3.5. If the data have the same
focus parameter as the reference, that is, (30, the compressed peak for any object
in the scene has high contrast, as shown. At other values of (3 the image is
defocused with a lower contrast peak. Thus, CO is based on the simple obser-
vation that focusing the image yields features that are both narrower and of
greater intensity. The algorithm estimates the optimum value of the focus
parameter by iterating over a set of reference functions with varying values of (3.
PPP is restricted since it operates only on dominant scatterers. The other
four autofocus processes are performed on several (e.g., 32) different range gates.
This improves the accuracy of the measured focus parameter since individual
estimates are independent.
Once autofocus is implemented, the blurred image in Figure 3.1 is im-
proved as shown in Figure 3.6. In order to achieve this result, the scene was
divided into eight strips along track that were focused independently (using
CO) and the image displayed using the estimated (3 for each strip. It is apparent
that defocus effects to the right of the original scene have been removed while
focus on the left of the scene has been improved. Note that the horizontal line
about a quarter of the way up each image is a range marker, not a feature of the
low contrast
high contrast
low contrast
Figure 3.5 Effect of varying focus parameter on the compressed image in CO.
Figure 3.6 Strip autofocused version of Figure 3.1 after using the CO autofocus method.
actual image. This autofocus process is signal-based; the effect of the unknown
motion on image quality is used to correct itself.
(8p 2 ) r « ^ (3.5)
SCR x r 5
for an aperture of duration T Thus, all five methods have the same form for
their statistical uncertainty, differing only in the value of c, which is 40 for PPP,
80 for PGA and PD, 96 for MLR, and 90 for CO.
This result applies to a single look within the aperture. However, there are
T0I T uncorrelated looks that can each be exploited for autofocusing over an
aperture T If the algorithm uses all available looks, the mean-square error in (3
for aperture 7" would be reduced by Tl T0 to [24]
(8p 2 ) r - (3.6)
X /T
SCR X T4
where c takes the values given previously. This is a factor b5 worse than for a
point target and indicates the importance of using point targets for autofocus
wherever possible. This conclusion applies equally to all autofocus techniques
and so does not differentiate between them from either theoretical or practical
considerations.
2 r/Va4
(8p )rr - (3.8)
° SCR X Tl
(Sp 2 L - (3.9)
K X SCR X T\
Variations in both a and (3 not only affect image defocus, as discussed in the pre-
vious section, but also give rise to geometric and radiometric distortion in the
SAR image. In this section we demonstrate how all three effects can be related to a
mismatched focus parameter. Initially we consider azimuthal geometric distor-
tion. To illustrate the problem, Figure 3.7 shows a strip SAR image of length
7.2 km and swathwidth 4.7 km at a range of 37 km, with antenna length da —
1.8m and X — 0.031m. Under these conditions the illuminated azimuth beam-
width at the center of the swath is 678m, so the strip subtends 10.6 full apertures.
The image has undergone autofocusing that was applied in strips of 192m, so
there are about 3.5 strips per full aperture. The local focus appears good, but over-
laying the scene with map data registered at the center of the strip reveals that the
image is distorted in azimuth by hundreds of meters toward the edges.
t
Ax, (*) = J &V(t'W (3.10)
o
Position (km)
Figure 3.8 Plot of the fractional focus parameter variations derived during the first pass of
strip autofocus processing.
match of 10 ms" 1 . The change between focused imagery on the left of Figure 3.1
and the defocused output on the right could be attributed to an azimuthal
acceleration of 10 ms~ 2 or a change in radial acceleration of 0.1 ms~ 2 . The former
is unrealistic, so it is reasonable to interpret errors in (3 as predominantly due to
radial acceleration in this instance. Indeed, if a SAR sensor includes conventional
M O C O and INS that reduce unknown motions to effectively the same order in
azimuthal and along-track directions, this result suggests that the azimuthal
velocity contribution to image distortion can be ignored. Accordingly, in the
following discussion we attribute all changes in focus parameter to radial motion,
leading to an azimuthal distortion of [20,25—27]
w.vjt&iv+tm. ( 3. 12 )
Equation (3.11) showed that azimuthal distortion was the product of the
range and the pointing angle offset, — VJ V. Assuming that the real antenna was
stabilized to be orthogonal to the assumed track, this would result in reduced
energy in the compressed peak. The ratio of this offset angle to the real beam-
width at time t is then given by
,M = H j M ^ + W,
Rak{ fi VX
for aperature length da. Assuming a rectangular antenna weighting for simpli-
city, the power in the compressed image would then be reduced by a factor
P= psimre/iy (3J6)
V ire J
The autofocus methods in Section 3.2 typically reduce the phase error across
the aperture to about TT/20 radians, corresponding to about 1 % degradation in
the 3-dB width of the PSF, which is comparatively insignificant. In this section
we derive the effect of the residual error in the focus parameter 8(3, following
signal-based M O C O , on geometric and radiometric distortion.
(&x>)»2 - V T ^ £ J S = <L 1 c Il
r 3
° p vr 2TT v SCR x yv; r
From distortion considerations, therefore, it is important to maximize the
subaperture duration. Thus, autofocusing on extended clutter and using the
phase gradient estimator in crowded environments is to be avoided.
This prediction is compared with measured distortions in Figure 3.10,
using results obtained by performing independent autofocus and phase correc-
tion of a second SAR image of the same scene. Displacements between objects in
the two scenes were then measured directly, after the scenes had been suitably
registered [I]. Since the scenes suffer from independent sources of error, the rms
error in a single SAR image would be reduced by a factor of V 2 from about 1.5m
between both images to 1. Im in a single image. The residual structure in the plot
indicates that slow fluctuations are not being properly corrected, probably caused
by range-dependent effects. If we ignore these, the rms distortion over the strip
Along-track shift# m
Position, km
would be about half the total, that is, 0.55m. From (3.17), the predicted distor-
tion for da = 1.8m, N7. = 32, and c = 90 (for contrast optimization auto focus),
assuming the full aperture is used (so that T= T0), would be 1.6/v SCR over the
whole image strip. The observed performance indicates an SCR of about 10 dB,
which is consistent with these images. A comparison with Figure 3.7 shows that
phase correction has reduced distortion by about two orders of magnitude. Note
that if it was not possible to use the full aperture for autofocus the residual
distortion would be increased by (T0/ T)312.
The corresponding result for across-track distortion is obtained from
(3.5) and (3.13). The distortion over the aperture T is given by
8j/ ^ V2T2bfi/2$R0, so the rms cumulative distortion over the full aperture
is given by
(8/V/2 - ^ i ! * 5 = A. I
x /T
c T^ (318)
° 2R0 P VT 4ir V S C R X Nr T
Assuming the same conditions as given previously for Figure 3.7, the predicted
across-track distortion over a single aperture would be 0.0013m, leading to a
value of 0.0042m over the full strip, which is negligible compared with the range
resolution.
The residual radiometric distortion depends on the ratio of the residual
angular skew to the real beamwidth. It can be derived from (3.5) and (3.15).
During the aperture time T the fractional pointing angle difference is given by
Se ~ VTda 8(3/(3 i? 0 \, so the rms cumulative relative skew over the full aperture
is given by
T
° R0X (3 2TTR0\ \ SCR X Nr T3
Under the same conditions as for Figure 3.7, the predicted skew is 0.00023
of the full beamwidth over the full aperture and 0.00071 beamwidths over the
strip, which causes a negligible reduction in image intensity. The difference
between the predicted track and the average over the strip is likely to be much
more significant. This is determined by the stability of the sensor platform. If
the trajectory could be defined by combined INS and GPS, the angular uncer-
tainty for the average track might be about 1.5 mrad, corresponding to a
fractional skew of about 0.09 for a 1.8-m antenna, which completely dominates
random effects. From (3.16) we find that even this would only contribute a
radiometric error of about 1%. In fact, this could be further reduced by correct-
ing for the angular difference once the average track had been calculated during
the signal-based M O C O stage.
for small phase errors, 8(f). From (3.1), the phase error at time t within the
aperture due to uncertainty in the focus parameter is given by 8(() ~ 8(312.
Hence, following (3.14), the mean-square residual phase error over A^rangelines
within the aperture t = ± 772 would be given by
772
(8(|)2}r - (8p 2 ) r -i- f t*dt = - (3.21)
X lT N /T
T JTI2 80 X SCR XNr
with c taking the values derived in Section 3.2.3. This residual phase error is
independent of aperture duration, provided that all possible looks are used in
the signal-based M O C O .
For the same operating conditions as for Figure 3.7, (3.21) predicts rms
residual phase errors of 0.059 rads leading to an ISLR of —24.5 dB. Typical peak
sidelobe levels in a SAR system would be comparable with ISLR values about
10 dBs greater. Thus, the random integrated sidelobe power introduced by phase
errors is about an order of magnitude less than the deterministic contribution.
This random ISLR component could be further reduced by averaging over a
larger number of rangelines, provided they contain strong scatterers.
The discussion of residual errors in the previous section has shown the importance
of using the full aperture for autofocus to minimize all forms of image quality deg-
radation. However, if the unknown motion varies on a short time scale, this will
not be possible and we must use shorter apertures whose duration ^represents the
time over which the focus parameter is approximately constant. Even when this
condition is satisfied, the response of the autofocus process will be a function of
the rate of change of the focus parameter. Iterative autofocus, described in Sec-
tion 3.5.1, helps to circumvent the nonuniform frequency response of a given ap-
erture by progressively removing motion contributions until only random
contributions remain. Nevertheless, the most significant limitation to autofocus
processing is the requirement to use short apertures to compensate high-fre-
quency components. As demonstrated in Section 3.4, treating each aperture inde-
pendently can result in appreciable distortion and cumulative phase error for
small apertures. However, in Section 3.5.2 we describe a coherent, iterative, auto-
focus technique that allows us to realize long-term low distortion, corresponding
to the full aperture, while simultaneously correcting higher frequency compo-
nents. Finally, in Section 3.5.3 we describe how a combined conventional and sig-
nal-based M O C O scheme can be implemented offering optimum performance.
(b)
(C)
Figure 3.11 Estimated autofocus correction, 8(3/3, as a function of along-track position after
(a) first, (b) second, and (c) tenth autofocus iteration.
The estimated correction then has the form A|3 + T], where A (3
represents the true focus mismatch and T| the estimation error.
2. The signal is phase-corrected using the focus parameters in (3.14).
This has the effect of removing the true focus mismatch A(3 with a
residual error T].
3. The aperture length is doubled and the process repeated. The residual
focus errors from the previous two apertures are now averaged and
contribute to the correction for the new aperture. The new residual
error in (3 corresponds to the double-length aperture and so is a factor
of 16 smaller than the previous value, from (3.6), assuming that all
possible looks are used for autofocus. The cumulative along-track
distortion and phase error are now determined by this larger aperture.
4. The process of doubling the aperture is repeated until it attains the full
aperture length.
This technique can be applied to the PD, MLR, and C O algorithms under
all circumstances and PPP and PG in a sparse scatterer environment. The final
error in the focus parameter for the full aperture is then as expected from (3.6)
when T = T0. Thus, this approach yields the maximum sensitivity for the full
aperture without compromising high-frequency tracking performance. How-
ever, the rms error in phase and azimuthal position of each sample is determined
by the shortest aperture used.
where Nj(f) is the PSD of the uncertainty introduced in the autofocus process
given by NT{f) ~ ( 8 ^ | ) . The form of the response for CO, PG, and PD
autofocus methods has been shown to be given by [24,30]
mf)
^ ~ r h \ [ l ~ rT^l s i n *f + —fcos *f (3 23)
-
while the response for MLR has a similar shape [24]. For subsequent iterations
the autofocus filter operates on the output of the previous iteration, following
phase correction. Hence
Repeating the process for progressively doubled aperatures allows the final
power spectrum of the sensor accelerations after autofocus and phase correction
to be calculated for a specified sensor acceleration spectrum and autofocus filter
duration. Figure 3.12 plots the predicted residual power spectrum after signal-
based M O C O for the DRA Canberra system. In an earlier study [24] the
approximate acceleration power spectrum for the Canberra SAR was found to
be of the form
which is shown as curve (a) in Figure 3.12. A typical X-band SAR system was
assumed to have a resolution of Im and a range of 100 km. The full aperture
length would be 1.5 km, corresponding to a duration T0 = 7.5 sec for V =
200 ms" 1 . The predicted performance for the combined system can be derived
by supposing that the progressive autofocus and phase correction scheme de-
scribed in Section 3.5.2 was implemented, commencing with the shortest
aperture of 0.47 sec (= T0I16, length 94m) in curve (b) and incrementing in
powers of two up to the full aperture in curve (f).
Increasing aperture duration reduces low-frequency components by
12 dBs for each aperture doubling, as predicted in (3.6). A reduction of about
90 dB in the original uncompensated aircraft acceleration spectrum is expected
at low frequencies. However, at high frequencies, additional spectral contribu-
tions are introduced by the estimation errors in the autofocus process. These
cannot be removed by subsequent iteration since the autofocus response is near
zero at these frequencies. If signal-based M O C O had commenced with a longer
subaperture, however, high-frequency degradation would be significantly re-
duced. For example, the contributions introduced by an aperture of 0.94 sec
would be a factor 12 dBs smaller. Thus, an initial stage in designing such a
signal-based M O C O would begin with an aperture duration that only degraded
the original high-frequency acceleration spectrum slightly.
PSD(f)
Frequency, f
Figure 3.12 Predicted residual acceleration power spectrum after autofocusing: (a) original
PSD, (b) after autofocus with T = 0.47 sec, (c-f) subsequent iterations with
progressive doubling of aperture to T= 7.5 sec.
, x Azimuth (m)
(a)
Figure 3.13 Measured azimuthal PSF at long range while encountering turbulence: (a) uncor-
rected response, (b) with conventional MOCO, and (c) with conventional and
signal-based MOCO.
3.6 Conclusions
In this chapter the basic physical principles governing SAR image formation
using signal-based M O C O techniques have been discussed. A series of algo-
rithms has been developed to rectify SAR image defects caused by variations in
Power (dB)
range due to unknown sensor motion. In each case, analysis establishes the
performance limits imposed by the algorithm and the radar data themselves.
Initially image defocus was corrected using autofocus methods. These
yield a time-dependent focus parameter error that was shown to be the source
of:
, . Azimuth (m)
(c)
Figure 3.13 (continued).
References
[1] Belcher, D. P, R. G. White, C. J. Baker, and C. J. Oliver, "The Residual Distortion of
Phase-Corrected Autofocused SAR images," Int. J. Remote Sensing, Vol. 14, 1993,
pp. 769-781.
[2] Muff, D., A. M. Home, and C. J. Baker, "Spaceborne SAR Autofocus," SPIE Conf. on SAR
image analysis, simulation and modeling, Paris, 1995, pp. 7-17.
[3] Belcher, D. P, A. M. Home, and C. J. Baker, "Spotlight Mode SAR Autofocus," IEE Conf.
SAR 93, Vans, 1993.
[4] Oliver, C. J., "Problems in SAR Image Formation," DRA Tech. Report
DRA/LS2/TR96016, 1997.
[5] Porcello, L. J., "Turbulence Induced Phase Errors in Synthetic Aperture Radar," IEEE
Trans. Aerospace Electron. Systems, Vol. AES-6, 1970, pp. 636-644.
[6] Rondinelli, L. A., and G. W. Zeoli, "Evaluation of the Effect of Random Tropospheric
Propagation Phase Errors on Synthetic Array Performance," Eighth Annual Radar Symp.
Record, University of Michigan, 1962, pp. 235—256.
[7] Carrara, W. G., R. S. Goodman, and R. M. Majewski, Spotlight Synthetic Aperture Radar:
Signal Processing Algorithms, Norwood, MA: Artech House, 1995.
[8] Werness, S. A., W G. Carrara, L. S. Joyce, and D. B. Franczak, "Moving Target Imaging
Algorithm for SAR Data," IEEE Trans. Aerospace Electron. Systems, Vol. 26, 1990,
pp. 57-67.
[9] Eichel, P. H., D. C. Ghiglia, and C. V. Jakowatz, "Speckle Processing Method for Synthetic-
Aperture-Radar Phase Correction," Optics Lett., Vol. 14, 1989, pp. 1-3.
[10] Eichel, PH., and C. V. Jakowatz, "Phase-Gradient Algorithm as an Optimal Estimator of
the Phase Derivative," Optics Lett., Vol. 14, 1989, pp. 1101-1109.
[11] Wahl, D. E., P. H. Eichel, D. C. Ghiglia, and C. V. Jakowatz, "Phase Gradient Auto-
focus—a Robust Tool for High Resolution SAR Phase Correction," IEEE Trans. Aerospace
Electron. Systems, Vol. 30, 1994, pp. 827-834.
[12] Yoji, G. N., "Phase Difference Auto Focusing for Synthetic Aperture Radar Imaging,"
United States Patent No. 4999635, 1991.
[13] Swiger, J. M., "A Digital Autofocus Method for Synthetic Array Spotlight Radar Using Map
Drift," Hughes Aircraft Invention Disclosure, 1976.
[14] Mancill, C. E., and J. M. Swiger, "A Map Drift Autofocus Technique for Correcting Higher
Order SAR Phase Errors," 27th Annual Tri-Service Radar Symp. Record, Monterey, CA,
1981, pp. 391-400.
[15] Vant, M. R., "A Spatially-Variant Autofocus Technique for Synthetic Aperture Radar,"
Radar '82, London: IEE, 1982, pp. 159-163.
[16] Bendor, G. A., and T. W. Gedra, "Single-Pass Fine-Resolution SAR Autofocus," Proc. IEEE
National Aerospace and Electronics Conf. NAECON, Dayton, OH, New York: IEEE, 1983,
pp. 482-488.
[17] Finley, I. E, and J. W. Wood, "An Investigation of Synthetic Aperture Radar Autofocus,"
DRA Memorandum No. 3790, 1985.
[18] Wood, J. W, "The Production of Distortion-Free SAR Imagery," Radar '87, IEE Conf.
Publ. No. 281, London: IEE, 1987, pp. 471-473.
[19] Wood, J. W, "The Removal of Azimuthal Distortion in SAR Images," Int. J. Remote
Sensing, Vol. 9, 1988, pp. 1097-1107.
[20] Oliver, C. J., "Review Article—Synthetic-Aperture Radar Imaging," /. Phys. D: Appl. Phys.,
Vol. 22, 1989, pp. 871-890.
[21] Green, J. E, and C. J. Oliver, "The Limits on Autofocus in SAR," Int. J. Remote Sensing,
Vol. 13, 1992, pp. 2623-2641.
[22] Oliver, C. J., "High-Frequency Limits on SAR Autofocus and Phase Correction," Int. J.
Remote Sensing, Vol. 14, 1993, pp. 495-519.
[23] Blacknell, D., A. P. Blake, C. J. Oliver, and R. G. White, "A Comparison of SAR Multilook
Registration and Contrast Optimisation Algorithms Applied to Real SAR Data," Radar
"92, IEE Conf. Publ. No. 365, London: IEE, 1992, pp. 363-366.
[24] Oliver, C. J., D. Belcher, A. R Blake, and R. G. White, "Algorithms for Focused Linear
Synthetic Aperture Radar Imaging," SPIE Conf. on Algorithms for SAR Imagery, Orlando,
FL, SPIE Proc, Vol. 2230, 1994, pp. 60-71.
[25] Blacknell, D., I. P. Finley, A. Freeman, C. J. Oliver, S. Quegan, I. A. Ward, R. G. White,
and J. W Wood, "Geometrical Accuracy in Airborne SAR Images," IEEE Trans. Aerospace
Electron. Systems, Vol. AES-25, 1989, pp. 241-256.
[26] Finley, I. P., C. J. Oliver, R. G. White, and J. W Wood, "Synthetic Aperture Radar—Auto-
matic Change Detection," UK Patent No GB 2,234,13OB, 1988, and US Patent No.
4,963,877, 1990.
[27] Wood, J. W, R. G. White, and C. J. Oliver, "Distortion-Free SAR Imagery and Change
Detection," Proc. U.S. National Radar Conf, Ann Arbor, MI, 1988, pp. 95-99.
[28] Mims, J. H., and J. L. Farrell, "Synthetic Aperture Imaging with Manoeuvers, "IEEE Trans.
Aerospace Electron. Systems, Vol. AES-8, 1972, pp. 410-418.
[29] Kirk, J. C , "Motion Compensation for Synthetic Aperture Radar, "IEEE Trans. Aerospace
Electron. Systems, Vol. AES-Il, 1975, pp. 338-348.
[30] Blacknell, D., "SAR Motion Compensation Using Autofocus with Implications for Super-
Resolution," Ph.D. Thesis, University of Sheffield, UK, 1990.
4
Fundamental P r o p e r t i e s of S A R Images
4.1 Introduction
Image
Data
Image interpretation Information Applications
feature
techniques
Data World
model model
Physical or
phenomenological
description
Figure 4.2 Flow chart showing the interrelation of measurements, data models, and informa-
tion.
with sensor characteristics, resolution being particularly important. In combi-
nation with a "world model" embodying general types of knowledge about the
scene (e.g., that an agricultural area is likely to contain fields of a single crop),
the data model is exploited to formulate image interpretation techniques that
are optimal in some sense. These provide a modified image in which the
"information" is more clearly portrayed than in the original image. An impor-
tant issue here is one of scale. When the interpretation techniques are local, we
are concerned with the reconstruction of some underlying image property
(parameter); this is essentially image filtering of some form. When the data
model imposes a large-scale description, we are more concerned with image
structuring. These two scales are not clearly distinct and there is continual
interplay between local and global scales in several of the algorithms. However,
whichever scale is chosen, the final determinant of information is in the context
of an application: what properties of the world become detectable or measurable
as a result of the interpretation technique?
While exploiting each model to its limit, we identify where those limits
are and develop the models to encompass progressively more demanding data
properties. Each such development is then again worked through to its limit.
Initially we focus our attention on single-frequency, single-polarization systems
for the sake of simplicity and to bring out some of the main issues as clearly as
possible. Later chapters are more concerned with methods applicable to mul-
tichannel systems. Large amounts of the single-channel theory and approach can
be successfully cast into a multidimensional framework, but new insights and
problems arise, many of which still await a complete solution.
Although we described measurement as a central issue, this needs some
interpretation, dependent on our application. For small target detection in
single-channel data, image calibration is often unnecessary because we are not
concerned with precise measurements of RCS or backscattering coefficient, only
the contrast between the target and its background. Even slow drifts in radiome-
try within the image are often of no importance in this context. Similar consid-
erations apply if our concern is structural mapping. For small target recognition,
especially if multiple channels are to be combined, calibration becomes a far
more pressing concern, although relative calibration may be more important
than absolute calibration. When our application deals with distributed targets,
accurate radiometry is nearly always an issue, with ensuing demands on system
accuracy and post-processing. A major exception, whose importance becomes
clear in Chapter 8, is when the predominant information is carried not by the
single-channel mean backscattering coefficient but by image texture. This image
property is independent of the mean and hence of radiometric calibration,
unless the radiometric distortions occur on scales comparable with the scale
lengths of textured regions.
The notion of information and its relation to measurement also needs
some qualification. Particularly at large scales, we begin to encounter difficulties
in consistency between an exploitable data model and human world knowledge.
From Figure 4.1, it is clear that an observer will automatically classify whole
regions as single entities, labeled perhaps as fields or woodland, without being
concerned about local drifts in radiometry or features within a region (see, for
example, the large area of woodland above the center of the image). Other world
knowledge is also automatic to the observer, such as expecting field boundaries
to be approximately straight over reasonable distances. This large-scale view can
easily come into conflict with results generated by algorithms whose perspective
is much more local. As we will see, algorithms can provide image structure
(segmentation) that a human being may instantly perceive as too detailed (or
omitting important features). However, within the algorithm it will be guaran-
teed that the inferred structure is supported by measurement. The observer can
override measurement when it conflicts with the description thought to be
appropriate. The distinction between information described by measurement
and that derived from knowledge and application needs then becomes acute.
Much of the work described throughout this book attempts to identify this
distinction and lessen it where possible.
TV
%
Ae* = X V (4- 1 )
k=\
(C) (d)
(e) (f)
Figure 4.3 Different representations of the complex SAR image of a rural scene using ERS-I
SLC imagery (Copyright ESA, 1992): (a) in-phase (real) component, (b) quadrature
(imaginary) component, (c) phase, (d) amplitude, (e) intensity, and (f) log intensity.
Note the geometric distortion as in Figure 2.3(a).
Transmit/Receive
Antenna
W*i'*2> = — e x p f - ^ ^ ^ l (4.2)
1 2
TTCT v o- J
with mean value and standard deviation both equal to (J, so that in
this case CV = 1.
5. The log intensity Z) = In /has a Fischer—Tippett distribution [17]
PD(D)=e—^[-e—\ (45)
a V a J
whose mean value and variance are In a — 7 E and IT2/6, respectively. The
symbol 7 E denotes Euler s constant whose approximate value is 0.57722
[18]. This distribution is easily converted to that for a properly normal-
ized dB image, since we can set Dn = 10log 10 / = (10log10 e)D and
hence PD (Dn) = — Pn(Dn/K) where K= 10 log 10 * Notice that for
the log intensity data, the variance is independent o£ the mean value, un-
like the distributions (4.2) to (4.4). Hence, log data is often preferred as
a means of standardizing or stabilizing the variance, which is a common
first step in statistical analysis [19].
(a) (b)
Frequency
Phase Amplitude
(C) (d)
Frequency
Frequency
(e) (f)
Figure 4.5 Comparison of theoretical distributions (solid curves) with observations (histo-
grams) for a visually homogeneous area from ERS-I SLC imagery: (a) in-phase
component, (b) quadrature component, (c) phase, (d) amplitude, (e) intensity, and
(f) natural log.
The implications of this speckle model are important. If our interest is in
single images of distributed targets and, hence, phase provides no information,
we can get rid of it and use only amplitude, log, or intensity data. Hence the
initial chapters in this book are concerned almost exclusively with image data
for which phase has been discarded. Phase becomes important when we turn to
polarimetric and interferometric data in Chapter 11. For such data, pairs of
phase images, each having a uniform distribution, can nonetheless carry infor-
mation in their phase difference (see also Section 4.12 and Figures 4.9 and
4.10). In Chapter 10 we also demonstrate that phase information is critical
when we wish to carry out high-performance imaging of deterministic (man-
made) targets.
The observation that many distributed targets obey the simple speckle
model allows us to provide a partial answer to our question about whether an
image pixel carries information. Considered in isolation, each pixel supplies a
real measurement of backscattered power or amplitude. When the pixel is made
up of many elementary scatterers, the observed power is an estimate of an
underlying RCS whose true value is being masked by interference effects. If the
pixel is part of a uniform distributed target, the information per pixel is low
since the collection of pixels in the target can supply only one useful number,
a. All that greater numbers of pixels do is to reduce the estimation error in a
when averaging pixels. For nonuniform but textured targets, this conclusion
needs only slight modification; more parameters than the mean may be needed
to characterize the region, but typically their number will be few (see Chapter
5). The information content of a pixel only becomes high when a pixel is not
speckled (e.g., a single bright point scatterer) or the distributed target is made
up of objects of large enough size and sufficient contrast to be resolved. A
noteworthy example of the latter is when high-resolution images of urban areas
are acquired.
The interpretation of the parameter a in (4.2) to (4.5) needs a number
of comments. First of all, we saw in Chapter 2 that the geophysical quantity
corresponding to an intensity measurement by the SAR is an RCS (or) for
point targets or average backscattering coefficient (<r°) for distributed targets.
Working in terms of a or a° in standardized units means that measurements
from different sensors (which may have different resolutions, transmitted pow-
ers, and processing gains, for example) can be directly compared. Since we
are here dealing with distributed targets, cr° is the relevant quantity. As dis-
cussed in Section 2.7, converting measured powers to <r° values requires careful
calibration and characterization of the system, and the supplied data may
need some form of preprocessing to derive <r°. However, for the purposes of
mapping, target detection, and texture measurement (see Chapter 8), calibra-
tion may not be necessary as long as radiometric distortions are changing
slowly relative to the features of interest in the scene. There are numerous
examples, particularly of airborne missions, where the calibration has been
dubious, but meaningful information has been extracted from the data
[23,24]. The intensity value at each pixel of a distributed target should there-
fore be regarded as a scaled estimate of a° or, if the area correction has not
been made, to the mean RCS of each pixel. With this proviso and to avoid
convoluted descriptions, we will often use the terms "RCS" or "cross section"
to refer to the mean intensity even when the data have not been properly
calibrated. This viewpoint is adequate for most of the early chapters in this
book, but a more careful approach relating observed intensity to underlying
a° is given in Section 4.5.
A second important point is that we need to distinguish the measured
value at a pixel and the parameter value o~. Figure 4.4 and (4.1) both indicate
that the observed value at each pixel is the resultant of interfering Huygens
wavelets unless a single scatterer completely dominates the return. Hence the
value of a is specific to each pixel; the measured value is just a sample from the
distribution parameterized by a. To make this clearer, note that the mean
intensity at a pixel is given from (4.1) by
Assuming that the amplitude and phase of each scatterer are independent of
each other and that the phases of different scatterers are independent and
uniformly distributed we have
(Z) = I ( A 2 ) (4-6)
k=l
a = z\ + z\ = / (4.7)
that is, the observed intensity. Since the distributions (4.3) to (4.5) are all
transformations of the number pair (zx, Z2), they will all lead to the same
estimate [25]. Hence, in the absence of any further knowledge or assumptions,
the best estimate of a at every pixel is to form the intensity image.
Improved estimates of a at a pixel can be gained by combining measure-
ments at that position. Since a is the mean power, this suggests that the correct
approach, given Z independent measurements, is to average the measurements
in intensity (often called incoherent averaging because all phase information is
discarded). From (4A), this will preserve the mean value cr while reducing the
variance of the measurements by a factor Z to become a/Z. Indeed, it is easy to
show that this provides the maximum likelihood estimator of a given Z inde-
pendent exponentially distributed measurements each with the same mean value
a (see Chapter 6). Other methods (e.g., averaging in amplitude or in log
intensity) provide inferior estimates of a. Note also that averaging the complex
data is of no value, since averaging Z independent complex samples, each of
which is zero-mean Gaussian with variance a/2 in the / and Q channels, leads
to a zero-mean Gaussian variable with variance a/(2Z) in each channel. Thus
there is no speckle reduction and the information-bearing parameter has been
scaled by the factor 1IL.
Independent measurements can be provided by splitting the synthetic
aperture Doppler bandwidth, as described in Section 2.5. The separate images
are referred to as looks, so that this process of averaging in intensity is known
as multilooking, and the resultant image is known as Z-look. The underlying
premise of multilooking is that each look should be estimating the same RCS,
so it is important that the looks should be at the same position, nearly simul-
taneous, and without relative radiometric distortion. They also need to be at
the same frequency and polarization configuration. Nonetheless, marked dif-
ferences between looks can occur because of the slightly different viewing angle
used by each look. For example, the mean RCS over extended regions defined
by field boundaries showed differences exceeding 20 dB between looks in
P-band (68 cm) data produced by the NASA/JPL AirSAR system [26]. In this
case, the angular separation between looks was about 1 degree. Smaller differ-
ences occurred at L-band (24 cm), and differences were not significant at
C-band [26].
A second approach to improving estimates of a at a given position relies
on making assumptions about the spatial properties of a. Perhaps the simplest
such assumption is that the RCS is constant over some neighborhood surround-
ing the pixel of interest. In this case, given L independent pixel values, the
situation is statistically identical to that previously described, where inde-
pendent looks at the same point are formed in the processor. Hence, again the
measurements should be combined by averaging in intensity. Both cases are
therefore referred to as multilooking, but they suppress different types of infor-
mation in order to improve radiometric accuracy. In the first case, angular
variation in RCS is neglected; while in the second, spatial variation is lost. Both
provide equally accurate estimates of a over a uniform target, but in heteroge-
neous regions the two methods may provide different estimates of the mean
RCS at a given position due to the differing widths of the PSE The use of pixel
averaging clearly requires larger data volumes to be supplied but has the advan-
tage that the averaging window can be adapted to adjust to heterogeneity in the
scene. This is exploited in the filtering and segmentation procedures discussed
in Chapters 6 and 7.
Note that in both cases the resolution will have been degraded. In the
first case this is because each look uses only a fraction \lL of the available
bandwidth, so the resolution degrades by a factor L (see Section 2.5). The
second case is more complicated, since the averaging acts as a low-pass filter;
the value assigned to the resolution after the filtering depends on the precise
definition of the resolution being used. Both cases become complicated when
the looks are not independent, such as will occur if the looks are formed with
overlapping Doppler bandwidths in the frequency domain or if pixels are
correlated in the spatial domain [27]. This is far more likely in spatial averag-
ing, since here sampling conditions or interpolation will often ensure that
significant interpixel correlation occurs. We ignore this complication for the
moment, but see Section 4.4.
A fortunate circumstance is that not only is averaging in intensity the
correct way to carry out multilooking for preservation of the information-
bearing parameter, but the resulting measurement has a well-known analytic
PDF (unlike averaging in amplitude or dB, for example). The Z-look average
intensity
I = \th (4-8)
where the Ik are independent variables each exponentially distributed with mean
o~ is known to obey a gamma distribution with order parameter L [16]
(Note that we often denote the average given in (4.8) by / ; the overbar is
omitted here for notational simplicity.) The average intensity has moments
r<£) U J
with special cases (/) = a and var(/) = a2/L. The latter relation motivates the
definition of the equivalent number of looks (ENL) as
where the averages are carried out in intensity over a uniformly distributed
target. The ENL is equivalent to the number of independent intensity values
averaged per pixel. It is often applied not just to describe the original data but
also to characterize the smoothing effects of post-processing operations such as
image filtering. Even for the original data it may be noninteger if the looks being
averaged are correlated.
When resolution is not considered critical, data are often supplied in
multilook form because of the ensuing reduction in data volume. For display
purposes, it may also be preferable to use the square root of multilook intensity,
A = v / , since this has reduced dynamic range. The data then have what is
known as a square root gamma distribution, which is readily derived from (4.9)
using the change of variable relation
P^A) = IAP1(A2)
This yields
The discrete scatterer model for distributed targets implies that all the informa-
tion at each pixel in single-channel data is carried by the mean incoherent power,
a, appropriate to its particular set of scatterers. The observed intensity at each
pixel then has the conditional probability
I = vn (4.15)
Frequency
(C) (d)
Figure 4.6 ERS PRI image displayed in (a) intensity and (b) amplitude. (Copyright ESA, 1992.)
(c) Comparison of three-look gamma distribution with histogram of squared PRI
data from an agricultural field, (d) Corresponding comparison of original PRI data
with square root gamma distribution with L = 3.
also use the symbol n in Chapter 2 and Section 4.6.) Equivalently, the complex
reflectivity at each pixel can be written as
that is, the in-phase and quadrature speckle components are independent zero-
mean Gaussian random variables each with variance 1/2. Notice that we have not
included a deterministic phase term in the formulation of the complex reflectivity
because it would be completely randomized by the uniform phase of the speckle
term. (For coherent image pairs, where phase differences are information bearing,
this model is inadequate; see Chapter 11.) Again, we see that the observed signal
can be regarded as the product of an underlying RCS term and a speckle term.
Similarly, for Z-look data, the decomposition (4.15) is possible, where n is now a
unit mean gamma distributed variable with order parameter L.
This formulation of the SAR image as a deterministic RCS modulating a
random stationary speckle process has led to the description of speckle as
multiplicative noise, even though it is a repeatable interference phenomenon.
We will see in Chapter 11 that, when dealing with multidimensional SAR data,
the meaning of speckle becomes much less obvious, and its treatment as station-
ary signal-independent noise does not seem viable. In this sense, we regard
speckle as a term most useful in the context of single-channel SAR data.
Although the RCS underlying the observed values is essentially determi-
nistic, its magnitude is unknown and normally is most usefully described by a
probabilistic model. Given a set of TV pixels, we therefore need to consider the
joint distribution of the RCS (T = ((T1, . . . , a ^) and the speckle values n = (nv
. . . , n^). If we assume the speckle values are independent of each other and of
RCS, we can write
k=l
PfJI,n) = PMni,...,IN/nN)f[^!hl
*=i nk
Integrating out the speckle terms yields the joint PDF of intensity as
Pa(a) = f[^-ak)
Then
This speckle model for uniform targets underlies much of the existing work on
SAR image analysis and is used extensively in later parts of this book (e.g., in
the "cartoon" model for SAR segmentation in Chapter 7).
The analysis in Sections 4.3 to 4.5 was in terms of the given data and ignored
how the imaging process affects the information carried by the data. Essentially
there are three imaging effects that must be taken into account when trying to
recover the backscattering coefficient cr°, namely: (1) a scaling of <r° due to
propagation, antenna pattern, and processing effects; (2) a bias in a ° due to
system noise; and (3) spatial correlation induced by the processing. To see this,
we start from the linear imaging equation (2.34) that relates the observed
voltage, %(x), to the complex reflectivity, S(x), by
where the terms in the equation are as defined in Chapter 2. (Particular note
must be taken, however, of the symbol R0 for range, in order to avoid confusion
with the ACF notation used below.)
For distributed targets, the rapid spatial variation of phase in S(x) causes
the complex reflectivity to have the same characteristics as white zero-mean
Previous Page
Pa(a) = f[^-ak)
Then
This speckle model for uniform targets underlies much of the existing work on
SAR image analysis and is used extensively in later parts of this book (e.g., in
the "cartoon" model for SAR segmentation in Chapter 7).
The analysis in Sections 4.3 to 4.5 was in terms of the given data and ignored
how the imaging process affects the information carried by the data. Essentially
there are three imaging effects that must be taken into account when trying to
recover the backscattering coefficient cr°, namely: (1) a scaling of <r° due to
propagation, antenna pattern, and processing effects; (2) a bias in a ° due to
system noise; and (3) spatial correlation induced by the processing. To see this,
we start from the linear imaging equation (2.34) that relates the observed
voltage, %(x), to the complex reflectivity, S(x), by
where the terms in the equation are as defined in Chapter 2. (Particular note
must be taken, however, of the symbol R0 for range, in order to avoid confusion
with the ACF notation used below.)
For distributed targets, the rapid spatial variation of phase in S(x) causes
the complex reflectivity to have the same characteristics as white zero-mean
Gaussian noise. Hence, %(x) is also zero-mean and Gaussian. In this case, the
most useful statistical descriptor of % (x) is its spatial ACE Assuming S(x) and
n(x) are wide-sense stationary, so also will be % (x); its ACF is then defined as
where we have assumed that noise and signal are independent. The range
dependence of C is not explicitly shown to avoid any confusion with the ACF
notation. The deterministic ACF of k(x) is denoted R^(x) and is defined by
with a similar definition for R[(x). Rk(0) = I \k(y)\ dy is known as the energy
in k(x).
Since S(x) and n(x) are both white, we can write
and
(See discussion following (2.14); also (2.16) and Table 2.2.) Under
these conditions Ri(x) = A smc(2xl d), where A is a constant, and
where the substitution \\fa = Wda was used. Hence, the ACF of the
returned signal is
and the signal and noise components in the complex image have
identical statistical properties; the only difference is the power in the
two contributions. Signal and noise are therefore inseparable at the
pixel level.
4. Since the data are in fact sampled, of most interest are the correlations
at multiples of the pixel spacing, x0. The complex (field) correlation
coefficient at lag k in the azimuth direction is defined by
+
p W J ^ W > (4 32)
p lJ
- m*y\)
where we used the fact that % (x) is mean-zero and assumed stationarity.
In the ideal case, with sample spacing dj2 (Nyquist sampling), the cor-
relation coefficient would be given by (see (4.30))
^ . ^ - , . W f ,4.34)
var(7)
\C\2<J°R, (0)
CNR = — ^— (4.35)
NR1(O)
Corr. Coeff.
Corr. Coeff.
(a) (b)
Corr. Coeff.
Corr. Coeff.
(C) (d)
Figure 4.7 Intensity correlation coefficients calculated in the azimuth direction for ERS-1 data
from an agricultural area: (a) SLC, near range; (b) SLC, far range; and (c) PRI, near
range; and (d) PRI, far range.
CNR = H ^ L (436)
IN
The CNR varies widely in many SAR scenes because of variations in range and
in <r°. Since ICI is proportional to R0'2 (see (2.35)), the CNR can be signifi-
cantly worse at far range, particularly for airborne systems. Systems are normally
designed with sufficient transmitted power to prevent this from being a major
problem for most targets. However, regions of low a°, such as water bodies
under windless conditions, can often give noise-dominated responses especially
when they occur at far range.
The use of the word clutter to describe the power from a distributed target
is an indication that small target detection is hampered by this component of
the returned signal as well as by the system noise. The extent of this problem is
often described in terms of the square of the peak response from a point target
of RCS a compared to the powers from the clutter and noise background in
which it is embedded. In (4.22), a point target of RCS a and associated phase
i— i—
c|) is represented by S(x) = V a ^ 8(#), giving the response C\cre*bha(x). The
corresponding signal-to-clutter ratio (SCR) is therefore
S C R = * * (4.37)
where we have assumed that the peak response occurs at x = 0. In the ideal case,
this reduces to
fl(0)
SNR = SCR X CNR = ^ ^ ' (4.39)
NR1(O)
The signal-to-clutter relations (4.37) and (4.38) display the critical fact
that detectability of targets in clutter is not affected by the distance from the
radar, which is in marked contrast to the behavior of, for example, normal
surveillance radars where the SCR is inversely proportional to range. However,
since C is inversely proportional to R02, the SNR falls off in proportion to R05.
Nonetheless, in most cases, clutter is the dominant factor affecting target detec-
tion and more general analysis of scene structure. Even in those instances where
noise dominates over clutter, the similarity of the properties of noise and clutter
(see consequence 3 of Section 4.6) allows a unified treatment of both cases.
Note that these expressions for CNR, SCR, and SNR have been developed
by an analogue treatment, but the data are sampled. This entails no modifica-
tion to the expressions but only to how they are calculated, in two ways.
1. The values Rh (0) and Rt (0) correspond to the energies of ha(x) and
l(x). For adequately sampled data with spacing X0, the energy of ha(x)
is given by
i M N
where the intensity at position (i, j) is denoted I1-. Hence, the expected value of
/ is a, but its variance depends on the spatial correlation of the pixels and is
given by
where pIa and pIr are the azimuth and range intensity correlation coefficients.
Note that the correlation coefficients may here include both system effects
and correlation induced by the underlying RCS. Hence the same expression
applies for textured data (see Chapter 5). If the pixels are independent, the
correlation coefficients are zero at all lags except lag 0, so (4.42) reduces to
the familiar result
in which case
W(/)~^(4-A-_2_+_n
MN K M N M N J
ENL = MV JJL
(1 + 2P7 J l ] + 2p /r [l] + 4 P / Jl]p / r [l]) var(7)
which takes the value 3 MN/4 for the three-look ERS-I data, when we use the
measured correlation coefficients noted previously. (Notice that this analysis
assumes a uniform target. If the RCS varies within the target, the spatial
correlation coefficients may be altered, as described in Chapter 5, with the effect
of increasing var(/) and reducing the ENL.)
Spatial correlation also implies that the multilook average (4.41) is no
longer the MLE of the mean intensity. In fact, it can be shown that the MLE
requires complex data and is given by
a - - i - H - ( V S S + ) = —— S 1 Tt 1 S
MN MN
Equations (4.32) and (4.34) have shown that both the complex and intensity
data will normally be correlated due to the SAR processing. However, the
bandwidth of the data is doubled by the squaring operation needed to form
intensities. Hence, preserving the information requires that the spatial sampling
of the intensity data be twice that of the complex data. (Since the data are now
real, the memory requirement is the same.) For this reason, correctly sampled
intensity data always exhibit significant spatial correlation. Clearly p[k] or p£k]
are known if the PSF is known. (Since the same considerations apply to range
and azimuth, we have omitted the corresponding subscript on the correlation
coefficients.) However, uncertainties in the PSF often make it preferable to
measure p[k] from the data. Methods to do this are discussed in the next section;
here we are simply concerned with removing system-induced correlation. Possi-
ble approaches are described in the following subsections.
4.9.1 Subsampling
If the system-induced correlation becomes negligible beyond a certain lag k0, the
data can simply be subsampled at this spacing. This method clearly discards
information and degrades resolution but does not distort the single point PDF.
When used as a precursor to edge or segment detection it may be acceptable
since the objects of interest are large and the image structure found in this way
can be reimposed on the original data. However, single pixels may contain most
of the relevant information when small objects or detailed structure are of
interest, and this approach is then not suitable.
4.9.2 Pre-Averaging
Averaging blocks of nonoverlapping pixels reduces correlation at the expense of
resolution. For example, if the data are averaged in azimuth in blocks of TVto form
. /TV
1\ k=(i-l)N+l
Z(I-WN)Pi[N-k]
X ( i - W / ^ ) P/W
\k\<N
For the ERS-I PRI values shown in Figure 4.7, taking N =" 2 would lead
to a correlation coefficient of 1/6, while with N = 3 this drops to a value of
1/10. Figure 4.8 shows measured azimuthal intensity correlation cofficients
Corr. Coeff.
Corr. Coeff.
(a) (b)
Figure 4.8 Azimuthal intensity correlation coefficients for (a) ERS SLC data and (b) SLC data
that were first averaged in intensity in blocks of 5 pixels along azimuth.
based on (a) original ERS SLC data and (b) SLC data that have been averaged
in intensity in blocks of 5 pixels in azimuth [33] (see Figure 2.3). The reduced
correlation at lag 1 is obvious; the peak at lag 3 (corresponding to lag 15 in the
original data) is not significant.
Using this method of reducing correlation will clearly change the PDF of
the data. For example, the averaged data will have reduced variance (see (4.42)).
Exact expressions for the PDF of averages of correlated intensity data can be
derived using methods described in [16]. These require that the complex ACF
be available; this can be inferred from the ACF of the intensity data, using the
Siegert relationship (4.34), only when the complex ACF is known to be real and
positive. This approach to reducing correlation therefore leads to considerably
increased complication if rigorous analysis is to be applied to the averaged data.
4.9.3 Interpolation
For data sampled at or above the Nyquist rate, the original analogue signal can
be reconstructed from the samples and a new set of samples taken with spacing
chosen to remove correlation or reduce it to negligible levels. This is equivalent
to interpolating a new set of samples whose spacing exceeds that of the original
data. The classical method for carrying out this operation uses the Shan-
non-Whittaker sampling theorem and is based on interpolation using sine
functions. When the spacing of the new samples is related to the old spacing by
an integer multiple of 2~n (e.g., 3/2, 5/4, etc.), this operation can be carried out
efficiently using the fast Fourier transform (FFT). In this method, the data are
first interpolated onto a finer grid and then samples are discarded.
To illustrate this, assume that we wish to increase the sample spacing to
9/8 times the original, in one dimension only. A data sample of length M (which
must be a power of 2 for FFT methods) is Fourier transformed. The transformed
data vector is made eight times longer by appending 7Af zeros to it {zero-
padding), and the inverse transform of this lengthened data set is taken. This
yields data interpolated onto a grid eight times denser than the original. Finally
we discard eight out of every nine samples. Extending this approach to two
dimensions is straightforward, either directly by using two-dimensional FFTs or
by carrying out the interpolation first in one dimension, and then the other.
A full discussion of interpolation methods would be lengthy, involved, and
out of place in this text, but we note the following points:
P > ) = SinC
(f]
and zeros occur at spacing dal2. (For simplicity, we only describe the
azimuth direction.) More usually, this condition will be only approxi-
mately satisfied. Where the ACF varies across an image (as will always
occur, for example, in an image that has been resampled to ground
range, such as the ERS PRI product), only approximate methods are
available unless unequal spacing is adopted, with consequent adverse
geometric effects.
The analysis in Section 4.6 assumed a wide-sense stationary [30] target in order
to define the spatial correlation induced by the system. Here we make a more
careful treatment that has considerable practical consequences for estimating
this correlation. In essence, the imaging equation (4.22) tells us that the complex
image, %(x), is zero-mean and Gaussian [28]; hence, its statistical properties are
entirely described if we know the two-point correlation function
The assumption of stationarity allows us to separate out the RCS term \oj from
the speckle term and to regard the speckle as a correlated unit mean process
(assuming that h{x) is normalized to unit energy) multiplying the mean RCS.
A consequence of (4.45) is that the distribution of the Gaussian process
%(x) is known, so the joint distribution of intensity and phase can be written
down for an arbitrary set of pixels, using the change of variables
%(x) = y[7{x)e^
To obtain the joint PDF of intensity alone requires the phase to be integrated
out, but this rapidly becomes very complicated. Fortunately we have little need
for the multipoint joint distribution of intensity, although extensive use will be
made of two-point statistics. For a pair of pixels the analysis is identical to that
carried out in Chapter 11 when analyzing the joint distributions of correlated
polarimetric channels. Analogous treatments for the spatial and temporal do-
mains will be found in [16,28]. A detailed analysis of the intensity correlation
function and the effects of RCS variations is given in Section 5.6.
From (4.45), it can be seen that the ACF of the complex data is stationary
and proportional to the deterministic ACF of the PSF as long as the mean RCS
is stationary within the target. Similarly, fourth-order moments of the complex
data depend on the stationarity of the second-order moments of RCS (see
Chapter 5). These statements need some comment. The data model used when
we discuss image filtering and segmentation in Chapters 6 and 7 is that at each
pixel in the image there is a well-defined RCS that we are trying to estimate. In
that sense, the RCS is deterministic and speckle due to interference effects is the
only source of fluctuation. While speckle is stationary, RCS varies from place to
place, so ensemble statistics (over many speckle patterns) and spatial statistics
are equivalent only in regions of constant RCS. That being the case, (4.45)
would be true only over such uniform targets.
This becomes of concern when (4.45) is used to estimate Rh(x). It is
obvious from Figure 4.1 that most images are nonstationary at large scales.
When we consider individual scene components, such as fields or woodland
areas in Figure 4.1, it is no longer at all clear whether stationarity holds or not.
In some cases, trends in mean brightness across a field may be visible, but tests
on visually uniform targets often display statistically significant trends [36].
These tests rely on least squares fitting of a polynomial form (for example, a
linear polynomial) to observations. Tests based on distributional fitting were
found to be less sensitive to trends in the mean RCS [36].
This appears to call the usefulness of (4.45) into question. However, it can
be shown that (4.45) is valid if (a) is interpreted as the average RCS across a
region as long as (1) the RCS varies slowly on a scale comparable with the length
over which the PSF takes values significantly different from zero and (2) the test
region is much bigger than the scale length of the PSF. Hence, an estimation of
the complex ACF does not rely on knowing that the test region is stationary.
The intensity moments relevant in texture analysis (Chapter 8) are not so
robust to target heterogeneity. In this case, an important consideration is the
scale of variation in the target compared with the region used for estimation. If
the RCS is assumed to have a correlation length that is comparable to or greater
than the region used for estimation, trends will normally occur across the region
and the estimate will be distorted. If the correlation length is much shorter than
the estimation region, a proper sampling of the distribution of RCS is available
and accurate estimates will be possible. Scale lengths of variation in mean or
texture properties will also be seen to have considerable importance in the
segmentation procedures developed in Chapters 7 and 8.
When dealing with textured regions, methods for testing stationarity have
received little attention as yet, and visual inspection is normally used. The eye
is often very good at detecting lack of stationarity. For example, a human
interpreter would certainly recognize a dark feature running from top to bottom
in the woodland region in Figure 4.1 and so determine that the region is not
stationary. However, as noted, statistical tests may prove stronger than human
perception in some cases. Nonetheless, later chapters in this book will deal
extensively with statistical properties of regions selected to be stationary mainly
on the basis of visual inspection. A target will be referred to as homogeneous if
there are no significant trends in the spatial average properties within the target.
Stronger assumptions (e.g., ergodicity) are irrelevant when only single images
are available at a given time. The equivalence of ensemble and spatial statistics
is only of importance in those sections of this book dealing with simulation in
Chapter 5. Here inferred fluctuations in RCS from observations over a class of
objects (e.g., woodland) are used to provide a probabilistic model from which
to construct test images with similar properties.
The speckle model developed previously is based on (4.1) with the assumption
that TV is very large at every pixel and that the scatterers within each pixel are
statistically identical. However, the same model can also be used to represent
other important situations.
1. One scatterer dominates the return, which is the case of interest for
target detection when a strong scatterer is embedded in weak clutter.
The ensuing PDF of intensity is then a modified Rician distribution
[16]
c y _ Vl + 2 X SCR
1 + SCR
Until now we have been mainly concerned with the formation and properties of
single SAR images, but recent developments in SAR technology have made the
use of multiple SAR images of the same scene increasingly important. An impor-
tant distinction is whether the different images are mutually coherent or not. By
this we mean that the phase relations in the images can be meaningfully com-
pared. We noted in Section 4.3 that the interference effects giving rise to speckle
are noiselike but represent true electromagnetic scattering. In the absence of ac-
tual system noise, an identical sensor viewing an unchanged scene with the same
geometry would measure an identical complex image. Changes in the scene or the
sensor geometry, operating frequency, or polarization will cause the complex im-
age to change. However, in certain important cases the changes give rise to images
that are correlated and in which the phase properties of image pairs contain useful
information. Particularly important are SAR interferometry and SARpolarimetry.
In SAR interferometry, the same scene is viewed from slightly different
directions by a SAR carrying two antennas or by a single-antenna SAR viewing
the scene at different times (repeat-pass interferometry). Under certain condi-
tions [39], the phase difference between the scenes can be used to infer the
height of scatterers within the image so that a digital elevation model for the
scene can be constructed. For repeat-pass interferometry, a requirement is that
the geometrical arrangement of scatterers (at the subpixel level) should be
essentially unchanged between the times of gathering the two images, otherwise
the phase difference between the images will carry no information. Figure 4.9
shows an amplitude (a) and phase difference (b) image produced by the DRA
C-band dual antenna system. Interferometric fringes are very well defined over
most of the phase difference image, except where coherence is lost in the radar
(a)
(b)
(C)
Figure 4.9 Interferometric image of the Brecon Beacons, Wales, produced by the DRA
C-band dual antenna SAR: (a) amplitude image, (b) phase difference image with
gray-scale cycle corresponding to a change in altitude of 150m, and (c) inferred
image of the relief. (British Crown Copyright, 1997/DERA.)
shadow areas (the hill on the right side of the image) and the reservoir (lower
left). Each complete cycle of phase corresponds to an altitude change of 150m
at the center of the image, allowing the relief in the image to be inferred with
an accuracy of ± 5m, as shown in Figure 4.9(c).
In SAR polarimetry, nearly simultaneous images of the scene are formed
with different polarization configurations. Typically the radar transmits and
receives radiation that is linearly polarized in the horizontal (H) and vertical
(V) planes (relative to the plane defined by the wave vector and the normal
to the surface being illuminated), giving rise to four images—HH, HV, VH,
and W , where, for example, HV refers to transmitting horizontally polarized
radiation and receiving the vertically polarized component of the backscatter.
Depending on the physics of the interactions with the target, the phase dif-
ferences of polarimetric images can be information bearing (see Chapter 11).
There are also several other image types not available for a single image. To
illustrate this, in Figure 4.10(a—d), we show L-band amplitude images of a
region in Jutland, Denmark, produced by the HH, HV, VH, and W channels
of the Danish Centre for Remote Sensing (DCRS) EMISAR polarimetric system.
The complex channels can be combined in pairs in a number of ways, but
particularly useful prove to be phase difference (e), correlation coefficient (f)
(see Chapter 11 for a discussion of this parameter), and amplitude ratio ((g)
and (h) show the I H H / W l and IHV/Wl ratio images). The images shown
as Figure 4.10(e,f) were formed by combining the H H and W channels.
Other useful images can be formed by combining all the polarimetric channels
in a variety of ways, as discussed in Chapter 12.
Although system or scene changes may prevent images from being mutu-
ally coherent, much information can be gained from incoherent images of the
same scene. A particularly important example is multitemporal data, which can
be used for change detection. As we shall see in Chapter 12, ratio images like
those in Figure 4.10(g,h) play an important role for such data but provide only
one of numerous possibilities for displaying information when multiple data sets
are available.
Although multidimensional SAR data offer new types of information and
give rise to new image types, the processing involved is essentially no different
from that already described in Sections 2.2 to 2.5. Each image is processed
independently and, in isolation, displays the speckle and correlation properties
explained in Sections 4.3 to 4.6. In particular, the spatial correlation produced
by the system should be identical in the four channels unless the H and V
antenna patterns are substantially different. Clearly, when coherence between
images is an issue, great care must be taken to preserve phase throughout the
processing chain and issues such as system phase noise become relevant. How-
ever, this involves no new fundamental concepts.
(a) (b)
(c) (d)
Figure 4.10 (a-d) HH, HV, VH, and VV L-band amplitude images of an agricultural region in
Jutland, Denmark, produced by the DCRS EMISAR; (e) HH-VV phase difference;
and (f) correlation coefficient images, together with the (g) HH/W and (h) HV/VV
amplitude ratio images.
(g) (t)
Figure 4.10 (continued).
in the normal SAR imaging mode. This book will confine its attention to SAR
processing of the clutter band alone. Readers with an interest in MTI applica-
tions are referred to [40].
4.13 Summary
When trying to infer properties of the world from SAR data, only those
properties that leave noticeable or measurable effects in the image are of interest.
Our first visual response to these effects is in terms of image structure; armed
with experience and our knowledge of the world we can recognize information
carried by shapes, linear features, and objects in the scene. A loose distinction
can be made between small and distributed targets. For the latter, coherent
speckle caused by interference between elementary scatterers is a dominant
aspect of their expression in the image. A well-developed theory and much
experimental evidence indicates that, in single-channel data, speckle can be
treated as a stationary noiselike quantity multiplying the underlying RCS. The
true nature of speckle as a repeatable electromagnetic scattering phenomenon
only becomes manifest when multiple channels can be combined, as in SAR
interferometry or polarimetry.
For distributed targets, the speckle model (and more refined RCS models
developed in Chapter 5) indicates that all the available information is carried by
a small number of parameters that characterize the observed distributions. At
the simplest level, there is only a single parameter, which is the mean intensity.
The viability of this simple speckle model provides a solid basis for the image
analysis methods developed in the following chapters. The imaging process
complicates the extraction of geophysical properties from the SAR measure-
ments because it induces scaling of the backscattering coefficient, a bias due to
system noise and spatial correlation. Post-processing is needed to calibrate the
data. Spatial correlation must be accounted for both in parameter estimation
and in subsequent image analysis techniques. For the latter, it is often necessary
first to decorrelate the data.
Parameter estimation, either locally or over extended targets, plays a cen-
tral role in extracting information from SAR data, which is a reflection of its
function as a bridge between world knowledge (carried both in recognition of
image structure and in the empirical basis for describing distributed targets) and
physical understanding. Single images provide an important introduction to the
issues it raises and give rise to the highly developed theoretical and methodo-
logical treatment of Chapters 5 to 10. The understanding gained in this simpler
case then helps to bring insight to the more complex task of defining and
extracting the information carried by multidimensional SAR data in later
chapters.
References
[1] Ulaby, F. T., R. K. Moore, and A. K. Fung, Microwave Remote Sensing: Active and Passive,
Vol. 3, Norwood, MA: Artech House, 1986.
[2] Tsang, L., J. A. Kong, and R. T. Shin, Theory of Microwave Remote Sensing, New York:
Wiley-Interscience, 1985.
[3] Fung A. K., Microwave Scattering and Emission Models and Their Applications, Norwood
MA: Artech House, 1994.
[4] Kong, J. A. (ed.), Polarimetric Remote Sensing, New York, Amsterdam, and London:
Elsevier, 1990.
[5] Kuga, Y., M. W Whitt, K. C. McDonald, and F. T. Ulaby, "Scattering Models for
Distributed Targets," Radar Polarimetry for Geoscience Applications, F. T. Ulaby and C.
Elachi (eds.), Norwood, MA: Artech House, 1990, pp. 111-190.
[6] Frison, P. L., E. Mougin, and P. Hiernaux, "Observations and Interpretation of Seasonal
ERS-I Wind Scatterometer Data over Northern Sahel (Mali)," Remote Sensing ofEnviron-
ment, submitted 1997.
[7] Oh, Y, K. Sarabandi, and F. T. Ulaby, "An Empirical Model and an Inversion Technique
for Radar Scattering from Bare Soil Surfaces," IEEE Trans. Geosci. Remote Sensing, Vol. 30,
1992, pp. 370-381.
[8] Resnikoff, H. L., The Illusion of Reality, New York: Springer-Verlag, 1989.
[9] Julesz, B., E. N. Gilbert, and J. D. Victor, "Visual Discrimination of Textures with Identical
Third-Order Statistics," Biol. Cybern., Vol. 31, 1978, pp. 137-140.
[10] Julesz, B., H. L. Frisch, E. N. Gilbert, and L. A. Shepp, "Inability of Humans to Discrimi-
nate Between Visual Textures That Agree in Second-Order Statistics—Revisited," Percep-
tion, Vol. 2, 1973, pp. 391-405.
[11] Brillouin, L., Science and Information Theory, New York: Academic Press, 1956.
[12] Whitt, M. W, F. T. Ulaby, and K. Sarabandi, "Polarimetric Scatterometer Systems and
Measurements," Radar Polarimetry for Geoscience Applications, F. T. Ulaby and C. Elachi
(eds.), Norwood, MA: Artech House, 1990, pp. 191-272.
[13] Frisch, U., "Wave Propagation in Random Media," Probabilistic Methods in Applied Mathe-
matics, Vol. 1, A. T. Bharrucha-Reid (ed.), New York: Academic Press, pp.75-198.
[14] Lang, R. H., and J. S. Sidhu, "Electromagnetic Backscattering from a Layer of Vegetation:
a Discrete Approach," IEEE Trans. Geosci. Remote Sensing, Vol. 21, 1983, pp.62—71.
[15] Taylor, J., Scattering Theory, New York: Wiley, 1972.
[16] Goodman, J. W, "Statistical Properties of Laser Speckle Patterns," Laser Speckle and Related
Phenomena, J. C. Dainty (ed.), New York: Springer-Verlag, 1984, pp. 9-75.
[17] Arsenault, H. H., and G. April, "Properties of Speckle Integrated with a Finite Aperture
and Logarithmically Transformed,"/. Opt. Soc. Amer., Vol. 66, 1976, pp. 1160-1163.
[18] Abramowitz, M., and I. Stegun, Handbook of Mathematical Functions, New York: Dover,
1965.
[19] Kendall, M. G., and A. Stuart, The Advanced Theory of Statistics, Vol. 1, London: Charles
Griffin and Co., 1958.
[20] Bush, T. F, and F. T. Ulaby, "Fading Characteristics of Panchromatic Radar Backscatter
from Selected Agricultural Targets," IEEE Trans. Geosci. Electron., Vol. 13, 1975,
pp. 149-157.
[21] Lee, J.-S., "A Simple Speckle Smoothing Algorithm for Synthetic Aperture Radar Images,"
IEEE Trans. Systems, Man and Cybernetics, Vol. 13, 1983, pp. 85-89.
[22] Spiegel, M. R., Statistics, New York: McGraw-Hill, 1972.
[23] Yanasse, C. C. F, S. Quegan, and R. J. Martin, "Inferences on Spatial and Temporal
Variability of the Backscatter from Growing Crops Using AgriSAR Data," Int. ]. Remote
Sensing, Vol. 13, 1992, pp. 493-507.
[24] Beaudoin, A., M. Deshayes, S. Hardy, T. Le Toan, and D. Girou, "Use of Airborne SAR
Data for the Mapping of Shifting Cultivation in French Guiana," Proc. SAREX-92, ESA
WPP-76, 1992, pp. 185-191.
[25] DeGroot, M. H., Probability and Statistics, Reading, MA: Addison-Wesley, 1989.
[26] Quegan, S., "Parameter Estimation for Polarimetric Measurements from Vegetated Areas,"
Proc. Third International Workshop on Radar Polarimetry, Nantes, 1995, pp. 626-635.
[27] Wilson, J. J. W, "The Radiometric Resolution for a SAR Employing Correlated Range and
Azimuth Looks," Marconi Research Center Doc. No. ER-TN-MRCAM-0241, 1986.
[28] Davenport, W. B., Jr., and W. L. Root, An Introduction to the Theory of Random Signals and
Noise, New York: IEEE Press, 1987.
[29] Papoulis, A., Signal Analysis, New York: McGraw-Hill, 1977.
[30] Papoulis, A., Probability, Random Variables and Stochastic Processes, Singapore: McGraw-
Hill, 1984.
[31] Siegert, A. J. E, MIT Rad. Lab. No. 465, 1943.
[32] Raney, R. K., and R. Bamler, "Comments on SAR Signal and Noise Equations," Proc.
IGARSS '94, Pasadena, CA, 1994, pp. 298-300.
[33] Caves, R. G., "Automatic Matching of Features in Synthetic Aperture Radar Data to Digital
Map Data," Ph.D. Thesis, University of Sheffield, UK, 1993.
[34] Keys, R. G., "Cubic Convolution Interpolation for Digital Image Processing," IEEE Trans.
Acoust. Speech, Signal Processing, Vol. 29, 1981, pp. 1153-1160.
[35] Quegan, S., "Interpolation and Sampling in SAR Images," IEEE Trans, on Geoscience and
Remote Sensing, Vol. 28, 1990, pp. 641-646.
[36] Horgan, G., and H. Meena, "An Investigation of the Variability of SAR Data and Its Effect
on Classification," Proc. Final Workshop on MAESTRO/Agriscatt: Radar Techniques for
Forestry and Agricultural Applications, ESA WPP-31, 1992, pp. 149-153.
[37] Schaefer, D. W, and P. N. Pusey, "Statistics of Non-Gaussian Scattered Light," Phys. Rev.
Lett., Vol. 29, 1972, pp. 843-845.
[38] Pusey, P. N., D. W. Schaefer, and D. E. Koppel, "Single-Interval Statistics of Light Scattered
by Identical Independent Scatterers," /. Phys. A: Math. Nucl. Gen., Vol. 7, 1974,
pp. 530-540.
[39] Rodriguez, E., and J. M. Martin, "Theory and Design of Interferometric Synthetic Aper-
ture Radars," IEE Proc. F, Vol. 139, 1992, pp. 147-159.
[40] Klemm, R., and J. Ender, "New Aspects of Airborne MTI," (FGAN - FFM) IEEE Int.
Radar Conf Proc, 1990, pp. 335-340.
5
Data M o d e l s
5.1 Introduction
Speckle itself conveys very little information about a scene other than that
it contains many randomly positioned scattering elements. However, in the
wooded region it is apparent that there are fluctuations in addition to speckle.
Physically, these appear to correspond to strong returns from the crowns of trees
with shadows behind them. If we disregard the deterministic positions of these
light and dark fluctuations, we can treat this type of natural clutter as a noiselike
texture. Note that texture is a consequence of fluctuations in the RCS; a clutter
sample comprised of speckle alone is not considered textured. Texture measures
will take the form of spatial average properties describing the depth of fluctua-
tion of the RCS within the local region. These can be single-point statistics, such
as the probability density function (PDF). More information is contained in
two-point properties such as the power spectrum or ACF. Texture will be
exploited in Chapters 8 and 9, both for characterizing the clutter within a
homogeneous region and identifying boundaries between regions. The former
conveys information that may enable a user to identify cover type, for example,
while the latter carries information about the structure within the scene.
One of the most important classes of information within an image is
contained in its structure. This comprises features such as the boundaries of
large areas of texture, for example, fields or woods, as well as discrete objects,
such as individual trees, hedges, roads, and buildings. The former can be
identified readily by exploiting the texture model. However, a simple noise
model would be an inappropriate description for an urban area that is made up
of discrete objects. The relative positions of objects in man-made clutter are
constrained by rules; for example, houses lie along streets, or trees in orchards
are arranged in rows. It follows that spatial average properties over a region do
not encapsulate the relevant information optimally. In this chapter we shall
confine our attention largely to data models for natural clutter.
Let us now identify measures that can provide discrimination between
such classes of texture. The mean intensities of the field and woodland regions
in Figure 5.1 differ by an order of magnitude in RCS, which would certainly
enable one to differentiate between them. Suppose, however, that the brightness
of the field was the same as the wood, as is often the case in SAR images. A
second discriminant could then be the contrast f = -^/var //(I)), which takes the
values 1.003 and 1.41 for the field and woodland regions in Figure 5.1, respec-
tively. The result for the field region is close to unity, as expected for pure
speckle, while the woodland region has increased fluctuations. Urban areas
typically display even stronger contrast. A third distinguishing characteristic is
the spatial extent of image features. In the field region, only speckle fluctuations
are observed, with a scale corresponding to the PSF. The woodland sample also
includes variation on a scale consistent with tree size and separation. While
many objects in urban areas, such as individual houses, are of similar size to
isolated trees, some buildings can be on a much larger scale. Each of these
general properties might aid image interpretation. The specific measure that
optimizes the information about these properties then depends on the detailed
form of the data model.
This chapter is concerned with introducing realistic models for SAR image
data. The problems inherent in data characterization are considered in Section
5.2, initially restricted to the PDF. A few widely used empirical PDFs are
introduced and compared in Section 5.3. The implications of the product
model, which separates the physically based speckle model from the pheno-
menological RCS model, are described in Section 5.4. This extends the basic
treatment given in Chapter 4. A comparison of predicted with measured prop-
erties for empirical and product models is given in Section 5.5. The product
model has the advantage that it contains correlation properties implicitly, as
outlined in Section 5.4.1. These are applied in an analysis of the effect of system
resolution on measured clutter properties in Section 5.6. This discussion high-
lights the importance of including the interaction of RCS correlation properties
and the PSF in any rigorous clutter analysis. This is an aspect that is omitted
from analysis based on fitting clutter PDFs alone and illustrates some of the
dangers of artificially restricting the scope of the data model. The discussion of
data models then concludes, in Section 5.7, with a summary of circumstances
under which they fail.
Having established a reasonable model for natural clutter textures, we then
describe techniques for simulating textures with these properties in Section 5.8.
Simulation is an important topic in seeking to understand SAR images, for two
primary reasons. First, provided simulated textures encapsulate all the properties
of the data model, a visual comparison with original textures will reveal failures
in that model. For example, a model based on single-point statistics could not
reproduce correlation properties. Furthermore, we will demonstrate that one
can discriminate between textures with different higher order correlation prop-
erties that would not be represented in any model based on the ACF and
single-point statistics alone. Second, once a model is validated it can provide a
number of sample images that differ randomly but are based on an identical
scene. It is clearly not possible to achieve this with real SAR data. There is a
danger that image-interpretation algorithms might be designed to provide ex-
cellent response to a specific example image but fail to provide robust generic
solutions. By allowing a very large number of realizations of the same scene to
be investigated, accurate performance statistics can be compiled.
Historically, SAR data models have been largely directed at describing single-
point statistics through the PDF. Initially, therefore, it is important to consider
whether typical SAR images are actually capable of representation in such simple
terms. In this section we show that such an approach is at best only an
approximation. There are fundamental problems in terms of the nature of the
scene being imaged, the physical properties of the imaging process, and finally
the analysis of measured data. Let us begin by considering the scene itself. On
the coarsest scale the scene is not homogeneous because a map of the area would
define boundaries between regions with different properties, such as fields/
woodland in Figure 5.1 compared with perhaps land/sea or forest/mountains at
lower resolution. Determinism prevails at the limit of fine resolution where it
would be possible, in principle, to represent every scattering element within the
scene and derive a unique image under defined conditions. Representation in
terms of texture can only apply as an approximation on an intermediate scale
when the scatterer spatial distribution appears random.
Image formation also introduces variability through physical differences
in the imaging process. SAR images depend on the system wavelength due to
differing foliage penetration, for example. The response to polarization also
varies, as discussed in Chapter 11. Geometrical effects due to the incidence angle
will influence the scene through processes such as shadowing. Finally, the
resolution of the imaging system will have a strong effect on the appearance of
the clutter because it will average out features on a smaller scale than the
resolution while preserving larger ones. In view of the number of processes that
contribute to differences between images, it is doubtful whether it is ever worth
attempting to define data models with great precision. We need only be con-
cerned with general properties that can be exploited in image interpretation.
Let us next address the issue of the level of complexity needed to represent
the data PDF. We are faced with a compromise between choosing a simple
model (characterized by few degrees of freedom) that fails to reproduce the
observed PDF accurately and a more sophisticated model (with many degrees
of freedom) that fits the data. A region of untextured speckle is characterized by
a single degree of freedom corresponding to the mean intensity, as discussed in
Chapter 4, consistent with the field example in Figure 5.1. However, the
associated negative exponential intensity, or Rayleigh amplitude, distribution is
not capable of representing some natural clutter, such as the woodland sample
in Figure 5.1. A variety of PDFs having two degrees of freedom have been
proposed, such as the Weibull, lognormal, and K distributions. The additional
degree of freedom allows them to represent different contrasts in the data;
contrast has already been identified as a potential texture discriminant. How-
ever, there are occasions where even this two-parameter PDF fails to provide a
reasonable match to the data and it is necessary to introduce more parameters.
We expect this approach to yield a better fit because there are additional degrees
of freedom. However, there is often no physical justification, since it merely
represents a more complicated parameterization. The process could be repeated
indefinitely, using mixture distributions for example, until an exact match to the
measured PDF, within statistical errors, was obtained. The resultant PDF would
not encapsulate any knowledge about the data but would merely provide a
(smoothed) alternative representation. It would provide no indication of how
properties might vary as imaging conditions change, for example.
There is a sense, however, in which some classes of PDF have a better
phenomenological basis than others. Analysis based on these models should
provide a better fit to the observed data with the same number of parameters.
The product model in Section 5.4, which separates speckle and RCS properties,
provides a physically based factorization of clutter that yields an effective com-
pact representation. Other empirical PDFs, such as the lognormal PDF, have
nonphysical properties that restrict their value.
Despite these reservations, let us now consider how to verify the data
model. The first requirement is that the test sample should be homogeneous.
How is this to be tested? One could adopt a nonparametric approach, such as
the Kolmogorov—Smirnov (K-S) test [I]. We will show later that using a test
based on the appropriate data model would be more sensitive. However, this
model cannot be verified until a suitable homogeneous sample is available.
Thus, the issues of model verification and homogeneity testing cannot be
separated. In the absence of a model, the human observer effectively provides
the test of homogeneity, as in the selection of the test regions in Figure 5.1.
A second issue concerns region size. It is important that it should be large
enough to characterize the statistical properties of the region completely; that is,
it must be on a much larger scale than any correlations. This is particularly
important if there are large-scale effects, such as ground undulations, that might
still be represented by a homogeneous texture if a large enough region is selected.
In addition, it needs to contain sufficient independent samples to provide good
statistical accuracy. This requirement for large regions conflicts with the obser-
vation that the real world does not normally consist of large homogeneous areas.
Thus we end up with some compromise choice of region size that approximates
homogeneity over a sufficiently large area for adequate statistical accuracy.
Once the sample has been selected the different data models can be
investigated. The PDF of the sample intensity, for example, can be compared
with the different candidate PDFs. This raises the question of how appropriate
parameters to describe the PDF are to be estimated. We might match some
arbitrary moment of the measured and theoretical PDFs. However, it is desirable
to identify the most sensitive measure for characterizing the data. Since the form
of the candidate PDFs is known, we can try to derive the MLE for optimum
discrimination. We discuss these issues in detail in Chapter 8 when we consider
texture characterization. In any event, the method adopted to estimate parame-
ter values will influence the quality of fit. It is also necessary to select the method
for characterizing the goodness of fit between data and model. Initially we derive
MLEs for the parameters required by each model, followed by a K-S test to
establish if the sample is consistent with that theoretical model. The K-S
hypothesis test is not dependent on any model and is determined by the
maximum difference between the cumulative probability distributions.
Early distributional measurements on radar data were carried out at low resolu-
tion so that objects, such as trees and houses, were much smaller than a
resolution cell. Contributions from RCS fluctuations averaged out so that no
spatial variations or correlation effects were visible. The resultant complex field
PDF was then Gaussian with Rayleigh amplitude and negative exponential
intensity PDFs, respectively, as discussed in Chapter 4.
In order to describe higher resolution data, it was necessary to introduce
a second parameter to characterize image contrast. The log normal and Weibull
are two such distributions that have been widely applied. For example, the
amplitude of low-resolution sea clutter has been fitted to the Rayleigh distribu-
tion, corresponding to pure speckle; while the log normal has been applied at a
higher resolution [2-4]. A wide range of ocean measurements at different
resolutions were shown to be consistent with the Weibull distribution [5—7].
Land clutter was found to be Rayleigh distributed over uniform regions [8] but
log normal over built-up areas [8,9]. The Weibull distribution has also been
applied extensively to land [5,10,11], weather [12], and sea-ice [13] clutter.
These distributions can be fitted to either amplitude or intensity data.
The log normal distribution is given by
where x is a generic observable, and (3 and Kare the mean and variance of In x.
The mean value and nth normalized moment of this distribution are
The log normal distribution predicts zero probability of the observable having
a value zero. As such it is a very poor representation of single-look intensity
speckle. However, it usually provides a better match to amplitude PDFs, par-
ticularly if it is used to represent regions of strong spatial variation such as urban
areas [8,9].
The Weibull distribution is given by
where b is a scaling parameter, and c controls the shape. The mean and nxS\
normalized moment of this distribution are
This model forms the basis of reconstruction filters and segmentation methods
described in Chapters 6 and 7 that are designed to derive specific values of a
associated with an observed value /. Note that, in order to exploit this model, it
is essential that the speckle and RCS fluctuations have very different spatial
scales. Normally we resample the SAR image so that speckle contributes little
correlation between samples, as discussed in Chapter 4. Any RCS fluctuations
on a larger spatial scale can then be separated out for further study.
The next component of the product model is to incorporate an RCS
model that describes the underlying fluctuations. For a given RCS distribution
P((T), the product model asserts that the PDF of the observed intensity is
given by
( Y F
^a) = U - £!^ exp _v2L (5.7)
P
[(Cr)) T(v) [ (a)J
where v is an order parameter, with moments given by
Vv J T(y)
This PDF is completely characterized by two variables: the mean RCS (a) and
the order parameter v. The square of the contrast, that is, the normalized
variance, is a useful quantity to describe such data and is given by
var cr/(a) — 1/v.
Jakeman and Pusey [16] proposed a theoretical model that provided
important insight into the basis for this empirical form of clutter PDF. The
simple speckle model in Chapter 4 showed that the field within a pixel could be
represented as the sum of the contributions from the elementary scatterers
N
within the resolution cell, so % — ^ A. exp\i^>. , as in (4.1). Gaussian speckle
then arises from the random interference of many scatterers and corresponds to
the limit as TV —> °o. However, Jakeman and Pusey investigated the effect of
assuming that the effective number of scatterers TV was itself a random number.
If the population TV is controlled by random processes of birth, death, and
migration, the consequent population statistics would be negative binomial
[17]. Under these conditions the resultant intensity was shown to be K-distrib-
uted [18-20], rather than the negative exponential form expected from the
central limit theorem. This derivation suggests that non-Gaussian clutter statis-
tics can be attributed to the fluctuation in the effective number of scatterers
within a resolution cell (see also Appendix 1 IA).
An equivalent formulation for complex reflectivity could be provided
[20,21] using a Fokker—Planck rate equation to describe the RCS. This can be
identified as the continuum analogue of a population operated on by processes
of birth, death, and migration. The solution to the resulting rate equation yields
a gamma-distributed RCS with an order parameter given by the ratio of migra-
tion to birth rates and a correlation coefficient at lag X given by [21]
where y is the migration rate. While this allows all values of order parameter, it
restricts the correlation function to a negative exponential form since it is
fundamentally a Markov process. However, it is consistent with the choice of a
gamma PDF as a description of the RCS [22,23].
The gamma distribution has the convenient mathematical property that it
can be regarded as the resultant of a Gaussian random walk process with the final
complex reflectivity at a given position expressed as a vector in 2v dimensions,
that is, S = (S1, S2, . . . , S2v), representing independent real and imaginary
contributions. Note that each of the contributions S1, S2, . . . , Siv represents a
one-dimensional Gaussian random walk. The total RCS is then given by cr =
ISI2, which is equivalent to summing v independent RCS contributions. For
simplicity we assume that the correlation properties of each complex reflectivity
component are identical. Any form of valid correlation can be adopted, unlike
the solution to the Fokker-Planck equation. However, it is only strictly applica-
ble for half-integer values of v. One consequence of this formalism is that all the
higher order correlation properties of S and a can be derived by exploiting the
factorization properties of Gaussian processes [23,24].
Alternatively, these can be derived directly for joint Gaussian processes.
For example, the bivariate PDF for correlated RCS samples separated by a
distance X is given by [17,25]
A
' 2"*(i-n,p0)r(v)l (,.(X) J
( a 2 ) - (a)
(((T(O)(T(X)Y) +9 Y2U+ V)
^^=2^(i-Pa(x)y^
((J) Y2(v)
P(I) = \P(I\(i)P((j)dj = ? -^
} U } K}
{ Y(L)Y(V)I(I))
( Y/2~
for Z-look SAR, where Kv_L [•] is the modified Bessel function of order v — L
(Section 9.6 in [26]). Note that this tends to the gamma distribution for Z-look
speckle as v —> °o. The intensity moments are given by
w
i-r(£) vr<v)
Equation (5.14) can be separated into two, formally similar, parts describing the
speckle and cross-section contributions. The normalized variance is given by
v a r / / ( / ) 2 = \/L + l/v + I/Lv, which reduces to v a r / / ( / ) 2 = 1 + 2 / v for
single-look SAR.
The PDF of the amplitude IA = -Jl j is also K-distributed such that
/ \(L+v)/2
P(A) = \P(A\<j)P(<j)ch = - —
KUK)
i T(L)TiV)Ul))
( Y/2~
x AL+V-1KVL 2A\^-\ (5.15)
However, the issue that determines the acceptance of this model is the
extent to which it is consistent with samples of natural clutter. Previous investi-
gation has shown that the K distribution is consistent with a large number of
coherent scattering experiments, over wavelengths from optical to sonar, and
types of scatterer from atmospheric turbulence to natural radar clutter
[14-16,23,24,28-33]. In the radar field the K distribution has been used
extensively to represent both sea clutter [14-16,18,34-36] and land clutter
[22-24,31-33,37-39]. Note that the observed statistics would be modified by
any additive input noise in the radar receiver [40,41].
Having introduced the K, Weibull, and log normal PDFs, we now establish
which provides the best fit to the examples of SAR image data in Figure 5.1.
Following the discussion in Section 5.2, we shall determine ML parameter
values for the candidate distributions and then perform a K-S test for goodness
of fit. The best-fit theoretical PDFs are compared with histograms of the
original data in Figure 5.2. The parameter values and fit probability for the
different distributions to the field region are summarized in Table 5.1 and those
for the woodland region in Table 5.2.
We expect the field sample to correspond to pure speckle with a negative
exponential PDF. From Table 5.1 it is apparent that the log normal PDF provides
a very poor fit whereas the exponential, Weibull, and gamma distributions have a
high probability of being correct. The predicted exponential, Weibull, and
gamma distributions are sufficiently close that they can all be represented by the
single (exponential) line in Figure 5.2(a). The Weibull shape parameter, rand the
gamma order v, are slightly less than 1.0 (which corresponds to pure speckle).
Since the data represent an estimate of the PDF, it would not be expected to have
the exact form. Hence, models with an additional parameter, such as Weibull or
gamma PDFs, would be expected to yield a higher probability of fit to the
estimated histogram, as observed. The distinction between the last three distribu-
tions in Table 5.1 is therefore probably insignificant. However, the incorrect form
of the log normal in Figure 5.2(a), particularly near the origin, is apparent. Fitting
a K distribution to the data was impossible because the estimated PDF actually
had a lower contrast than that predicted by the speckle model alone. This would
be expected for about half the samples taken over regions of constant RCS. When
this condition is observed, it is reasonable to assume that the sample actually
contains a constant RCS such that v = °°.
The K distribution provides an excellent fit to the example of woodland
clutter, as shown by the results of the K-S test in Table 5.2 and the comparison
of PDFs in Figure 5.2(b). All other distributions have a negligible probability of
being correct over this sample size. Physically, the exponential would be ex-
pected to be incorrect since it cannot describe RCS fluctuations with speckle.
P(D
Intensity, I
(a)
Figure 5.2 Comparison of the intensity PDF with theoretical best fits for (a) field and (b) woo
land regions from Figure 5.1. Log plots of data histogram (full); theoretical PDFs fo
(a) log normal (short dashes) and exponential (longer dashes); and (b) K, log norm
Weibull, and gamma (progressively increasing dash length) distributions.
The log normal predicts a probability of zero at zero intensity that conflicts with
the speckle model. The Weibull PDF, on the other hand, yields an incorrectly
large probability for zero intensity. The gamma distribution matches the low
intensities well but predicts probabilities for large intensities that are too high.
One might therefore conclude that only the K distribution should be
retained as a clutter model, with v = °° assumed where the statistics appear
smoother than speckle fluctuations. This would be premature since, while these
results are typical of natural clutter at this resolution for a C-band SAR, many
other experimental comparisons have been conducted at differing wavelengths,
resolutions, and polarizations. The log normal has been found to be appropriate
for urban clutter [8,9] while the difference between Weibull and K PDFs is
sufficiently small that slight changes in clutter properties make the Weibull
P(D
Intensity, I
(b)
Figure 5.2 (continued).
Table 5.1
Summary of ML Fit Parameters for Candidate Distributions and K-S Fit
Probability for the Field Region in Figure 5.1
Another reason for not rejecting log normal and gamma distributions for
natural clutter is that they can be useful as approximations to the K distribution
as sample size becomes small. Thus, in discussing texture edge detection and
segmentation in later chapters, which should be applied to the smallest window
size possible, analytic log normal and gamma approximations can be applied
with only minor degradation, compared with using K distributions that can
only be evaluated numerically.
In the previous section we ignored the effect of the PSF on the observed
intensity; we now deal with this issue. Suppose we image a region of extended
clutter, whose RCS is represented by a correlated gamma-distributed process.
Following Section 4.6 and retaining a one-dimensional treatment for simplicity,
the intensity image at position x, I(x), arising from a scene with complex
scattering amplitude, S(£), at position ^ is given by
O
O O
O
Kx) = \%(xf = J Id^2SH1)S*H2)h{x - Z1W(X - y (5.17)
— OC
O
-O
Upon taking the ensemble average over all speckle realizations, we observe that
the crossterms between the scattered field at different positions average out
because of the random phase of scattering elements. Hence,
O
C
(/)= jdl-(\S(1-f)\Hx-Zf=(V) (5.18)
O
-O
where the PSF is normalized to unit energy and the ensemble average is assumed
to be position-independent. In Chapter 4 we showed that the complex ACF was
effectively independent of fluctuations in the RCS. The intensity ACF is given by
00 00 O
C 00
(Kx)Kx + X)) = J J J \d^di2d^A(S{QS*{QS*{Qs{^))
—O C—0—0
— 0
X h{x - Qh*(x - Z2)h*(x + X- Qh(x + X-Z4) (5.19)
r =0 2
2
{I(x)l(x + X)) = (a) h + JdWx - Qh* (x + X-Z)
[ -OO
00 OC /£. t. \
If there are no RCS fluctuations, that is, v = °°, only the first two terms remain
and (5.21) reduces to the Siegert relation [42], as in Chapter 4, with
for Z-look images. The integral over the n-po'mt correlation function within a
single pixel yields the effect of the PSF on RCS moments. If the RCS is
gamma-distributed, only a rectangular PSF yields a K-distributed output, albeit
with modified order parameter and correlation properties.
Equations (5.21) and (5.23) represent the effect of the PSF on the ACF and
single-point statistics, respectively. Unfortunately no closed analytic form is avail-
able for the filtered PDF itself. In common with other forms of RCS, the intensity
distribution can only be derived numerically, which removes one of the advan-
tages suggested for the gamma-distribution RCS model [43].
We may use (5.21) to calculate the form of intensity ACF we would expect
for the woodland region in Figure 5.1. The two-dimensional correlation coeffi-
cient for a is assumed to be given by
Pa(x,y) =exp-M-M ( 5. 2 4)
* y
where x and y denote lag values in azimuth and range directions, respectively,
and £x and € are the corresponding l/e correlation lengths. This type of
correlation property has been reported for other examples of woodland texture
[24,33]. It is also consistent with the previous Markov model derivation that
allowed arbitrary values of v. In addition, the PSF is assumed, for simplicity, to
have a two-dimensional Gaussian envelope defined by
|^,7)| = exp - 4 ~ 4W W
< 5 - 25 )
I x y
with lie widths of wx and wy in azimuth and range directions. While this is an
oversimplification, it allows the required integrations to be performed more
easily. Appropriate parameter values for the woodland example from Figure 5.1
were estimated from the ACFs of both complex and intensity image leading to
tx = 8.0 pixels, € = 4.5 pixels, wx — 1.36 pixels, and w = 1.13 pixels. The
predicted form for the intensity ACF was then calculated from (5.21) and is
compared with measured data in Figure 5.3. Imaging averages out some of the
contrast, so the order parameter for the RCS has to be selected to yield the
observed ACF value at zero lag. In this case a value of 1.52 was deduced, which
would yield an ACF zero-lag value of 3.32 if there were no averaging effect on
imaging, rather than the value of 2.932, which was actually observed. While this
comparison is only approximate and is based on oversimplistic assumptions, the
consistency indicates the general validity of the product model for natural SAR
clutter, based on a correlated gamma-distributed RCS.
The same set of parameters can also be used to predict higher single-point
moments based on (5.23) [24]. Values for the second to fourth normalized
ACF(r)-l
ACF lag, r
Figure 5.3 Log plot of predicted and measured normalized ACFs for the azimuthal direction of
woodland texture in Figure 5.1.
Table 5.3
Comparison of Different Predicted Moments of the Woodland Intensity Distribution
intensity moments of the woodland texture from Figure 5.1 are shown in
column 2 of Table 5.3. In column 3 we show the moments derived by assuming
that the observed intensity was K-distributed, as actually indicated by the K-S
test. The order parameter could then be estimated directly from the second
moment, since (/ 2 )/(7) = 2(1 + l/v), using (5.14), leading to an effective
value of 2.15. RCS moments, consistent with the predicted ACF for a correlated
RCS in Figure 5.4, are listed in column 4. If these fluctuations were imaged by
a system that has a delta-function PSF so that there was no averaging, the
expected intensity moments would have the form of column 5. The results in
column 6 were obtained using (5.25) to derive the averaged intensity moments
after imaging. Comparing columns 5 and 6 it is apparent that the imaging
process has reduced the moments as expected, yielding results that are reason-
ably consistent with the data.
The success of the correlated gamma-distributed RCS model in predicting
both the ACF and single-point moments of the image intensity demonstrates
that the model provides a valid representation of these samples of natural clutter.
A similar approach has been applied to a variety of radar clutter samples
[24,33,37], yielding similar consistency between prediction and data. This
consideration of the effect of the imaging process demonstrates that it is not
sufficient to provide arbitrary parameterized fits of a K distribution to measured
single-point statistics. The interaction between the RCS fluctuations and the
PSF must be properly estimated. Only this approach can predict the properties
of the image of a given scene as the system PSF is changed.
5.8 Simulation
An important way to demonstrate that the properties of SAR images are well
understood is to simulate them based on the models available. A human ob-
server is often capable of observing differences that tests based on specific
models would not detect and so assessing the validity of the model. Further-
more, once a model has been validated, simulation can be used to generate large
quantities of statistically identical examples with which to develop and test
image-interpretation algorithms. Finally, both structural and textural informa-
tion could be combined in a SAR simulator that could be used for training both
human observers and automatic algorithms. The scene would then be regarded
as consisting of regions of homogeneous texture. The edges of these regions
might either be defined by structural features, such as known boundaries, or
arise from small changes in parameter values. Isolated objects would require a
more deterministic approach. In this section we confine our attention to the
simulation of homogeneous textures with the required properties.
A set of simulated textures is compared with an original SAR image of a
young plantation made up of rows of trees in Figure 5.5, where the original SAR
texture is the second from the left. The simulation technique is designed to
reproduce the mean and ACF of the original sample. Both the RCS and speckle
in each simulation are different realizations with identical parameters. It is
difficult to discriminate between the four examples either visually or using
statistical tests [33]. Simulation generates a texture in which the underlying
Figure 5.5 Comparison of measured SAR texture from a plantation of young trees (second
from left) with three example simulations. Single-look image, 3-m resolution.
where 7(#) is the incomplete gamma function (Section 6.5 in [26]). The corre-
lation coefficient for the gamma-distributed output pCT, which represents the
desired cross section, can be related to that for the negative-exponential input
Py, by [25]
Figure 5.6 Comparison of measured SAR texture from mature woodland (on right) with three
sample simulations. Single-look image, 3-m resolution.
T r J e 4 ~ T \ d y s ^vVJ WTexp["V(T]*" (5 28)
*
V2TT _i L 2 J o T(y)
This defines both .F(CT0), the forward mapping function for y0 —> cr0, and its
inverse i7" 1 Ij 0 ), describing the inversion y0 <— (T0. In principle, there are no
restrictions on the range of order parameter values that can be achieved with this
method, although the numerical inversion process has to be handled carefully.
The relationship between the desired correlation coefficient for the
gamma-distributed variable p a , and that for the underlying Gaussian process py,
can then be derived from the single- and two-point moments of a, defined by
CC
(a»)= \dy(F-\y))nP{y) (5.29)
— cc
where
and
X CO
where
TX \ i [ -Ti-2W2 +
yi\ , ™
p ex
( Ji ' J 2 ) = /, P -r ^ (5-32)
The inverse mapping F~l(y) can be solved numerically from (5.28).
Hence, the correlation coefficient for the correlated gamma variable, defined in
(5.27) can be evaluated from (5.29) to (5.32). This allows the underlying
Gaussian correlation coefficients at all lag values to be defined from the desired
gamma values. However, this correlation function may again not be valid, so an
approximate version that is close to the required form has to be derived. The
desired filter amplitude spectrum then corresponds to the square root of the FT
of this correlation function.
The simulated correlated K-distributed image is then generated as follows:
• The higher order correlation properties are not specified and so will
vary between methods.
• The methods cannot reproduce the directional effect of shadowing
because this is not represented in the ACF.
5.9 Conclusions
In this chapter we introduced a correlated K-distributed noise model that seems
to represent natural clutter well. The discussion of simulation in Section 5.8
reveals that this model is capable of representing both the single-point statistics
and ACF of SAR texture correctly. However, the important physical property of
(a)
(b)
(b)
Figure 5.8 Comparison of a sample of (a) C-band SAR woodland clutter with (b) simple
deterministic simulation. Resolution 1.5m by 1.5m, single-look.
shadowing is not contained within the ACF and cannot be represented by these
noise models. Image interpretation based on texture, described in Chapters 8
and 9, is based on the same model and similarly fails to take account of
shadowing. While many important remote sensing properties, such as tree size
and spacing, will be contained in the ACF, this cannot represent the optimum
description. However, we will demonstrate that much useful information can be
extracted even with an incomplete model.
References
6.1 Introduction
The product model for SAR clutter, introduced in Chapters 4 and 5, postulated
that the observed intensity in single-channel images comprises an underlying
RCS on which speckle is imposed. This speckling process defines the condi-
tional probability of observing an intensity /from RCS a, which characterizes
the Forward Problem. This chapter is concerned with the Inverse Problem of
deriving the RCS, given the observed intensity in the image. Even though
speckle properties are well understood, the ambiguity introduced by speckle
means that there is no unique inversion from / t o a. However, by interpreting
data in terms of a series of increasingly sophisticated models, we can produce
image reconstructions of continually improving quality.
This chapter is concerned with the development of filters to provide a
reconstruction of RCS. Initially, in Section 6.2, we examine the implications of
the speckle model for reconstructed image quality. In Section 6.3 Bayesian
inference is introduced, which provides the basis of most of the filters described.
Section 6.4 exploits the speckle model alone, with no information about RCS.
Once this model has been exhausted, we introduce and exploit a series of further
models, commencing with a data model to describe single-point RCS statistics
(Section 6.5), followed by world models describing scene properties, namely,
neighborhood correlation (Section 6.6) and image structure (Section 6.7). Each
improves the quality of the despeckled image. Most of these algorithms are
adaptive linear processes. However, we demonstrate in Section 6.8 that non-
linear iterative processes are capable of offering still further improvement in
image quality. We next apply the concept of a structure model in an iterative
scheme, analogous to simulated annealing, in Section 6.9.
In passing, it should be noted that a variety of other filters have been
proposed for smoothing speckle while retaining strong RCS fluctuations, such
as the median [1,2] and sigma [3] filters. While these schemes have desirable
aspects, such as computational simplicity, they fail to include the specific prop-
erties of SAR speckle. In consequence, they do not perform well enough in
practice to aid image interpretation significantly.
Though the data models in Chapter 5 were developed for comparatively
high-resolution SAR imagery, many potential applications require algorithms
capable of operating on lower resolution (such as satellite) imagery. In principle,
there is no reason why image-interpretation methods should depend on resolu-
tion; algorithms should be controlled simply by underlying theory and the data
themselves. Many algorithm properties, however, will be most evident with
high-resolution imagery, and developments in this chapter will primarily be
illustrated with the example of 3-m resolution, 2-look DRA X-band airborne
SAR imagery shown in Figure 6.1.
A "good" reconstruction filter demonstrates the following properties:
Figure 6.1 DRA X-band airborne SAR image of a rural scene: 3-m resolution, 2-look.
• Absence of artifacts;
• Radiometric preservation.
The first three properties can be observed visually; the last is better verified
numerically, as described in Section 6.2. Visual inspection can be used to
examine the effectiveness of candidate reconstruction filters in the following
manner:
and
V =
I ' ^T Kr(I + K) + 7B = K^l + I j + I (6.1)
where V}, VJ7, and Vn are the normalized variances of intensity, RCS, and
speckle, respectively. This relationship between V1 and V0. is important because
it is not possible to measure V0. directly; it has to be inferred from V1.
All the reconstruction methods described require estimates of the sample
mean and normalized variance over some window comprising TVpixels, defined
by
? and
• TF X * SD[ST]s \\T XOi - !)2 («)
For Z-look SAR the expected values for intensity images would be given
by r — 1 and SD[V] = y/l/L . Similar expressions could be derived for the ratio
of the pixel amplitude to the square root of the RCS, for which Jr=I and
SD[Vr] = - / / T 2 U V ( P ( Z + 1/2)) - 1 . This treatment is too simple since a ;
is estimated rather than known, but a rigorous analysis for the statistics of the
intensity ratio over uniform regions in segmentation has been reported [4],
When the observed mean value differs significantly from one, it is an indication
of radiometric distortion. If the reconstruction follows the original image too
closely, the SD would be expected to have a lower value than predicted. It would
be larger than predicted if the reconstruction fails to follow genuine RCS
variations. This provides a simple test that can be applied to any form of
reconstruction, including both despeckling filters and segmentation (see the
next chapter). Departures from prediction indicate a failure of the model or of
the algorithm attempting to implement it, without specifying the manner in
which it fails. For example, the reconstruction may provide an incorrect value
for a region of constant RCS or it may fail to register the positions of changes
in RCS. The test can be applied over the entire image or to specific regions.
Applied globally, it suffers from the problem that it does not indicate if the
reconstruction follows speckle fluctuations too closely in some areas but insuf-
ficiently closely elsewhere. In spite of these reservations, the test gives two
measures of the quality of reconstruction and we will apply it as a numerical test
of the algorithms developed in this chapter.
for Z-look SAR. P0. (a) is the a priori PDF that encapsulates prior knowledge
about the RCS. P1(I) = J P(IId)P0(a)da only serves to normalize the ex-
pression and need not be included specifically in most instances. Generally
we wish to provide an estimate of a that represents its most likely value
given an observed / . This is equivalent to maximizing the log likelihood
\ = [nP№(a\I) (6.6)
N N f j\L JL-I r JJ ~
for Z-look SAR, where pixels are assumed independent. The MLE for a is then
given by
which is the average intensity over all the pixels in the window, corresponding
to multilooking. Note that if this is applied to a single pixel the MLE is equal
to the intensity of that pixel, as noted in (4.7). As we will show, different values
for the MLE in subsequent algorithms depend on constraints introduced by
different world models.
The result of applying this algorithm to the original image in Figure 6.1,
using an 11 by 11 window, is shown in Figure 6.2(a). While the method has
reduced the SD of the speckle variations in regions of relatively constant RCS
(a) (b)
(C) (d)
(e) (f)
Figure 6.2 Despeckled versions of Figure 6.1: (a) multilook, (b) MMSE, (c) structural gamma
MAP, (d) CML, (e) CMAP, and (f) Crimmins.
by a factor VW, that is, 11, it has also degraded resolution by the same factor
resulting in the blurring of small objects. Only if features of interest within the
scene occur on a large enough scale can multilook images be effective. An
adaptive approach that matches the size of a window to the scale of objects of
interest would provide optimum despeckling at the required resolution. Where
fluctuations were consistent with a constant RCS the window size can be
increased to improve despeckling as far as possible.
where crMMSE is the MMSE estimate for RCS and k is selected to minimize the
mean-square residual between intensity and fit, yielding [7]
V, V1(I + I/I)
A better estimate for or can be obtained if we have prior knowledge about the
PDF of the RCS. The Bayes rule in (6.4) shows how this a priori PDF can be
used to provide a MAP reconstruction when combined with the likelihood
function. We demonstrated in Chapter 5 that the RCS of natural clutter can, in
many cases, be well represented by a gamma distribution of the form
{ixj r(v) i \x _
where JUL and v are the mean RCS and order parameter, respectively. These
cannot be measured directly and must be estimated from the data. The order
parameter can be obtained using (6.1) and the fact that the normalized variance
of the RCS is given by V0. = 1/v. Hence, estimates for JUL and v are obtained by
passing a window over the original image and setting
TJ/-J-2
So far we have exploited the speckle model in reconstructing the RCS of a single
pixel, combined with the single-point PDF of the RCS in Section 6.5. Sur-
rounding pixels were only involved through estimating the population mean
and variance (or order parameter when a gamma a priori PDF was assumed).
All the reconstructions illustrated in Figure 6.2 so far suffer from visible defects.
Improvement can only result from the introduction of further prior knowledge.
Let us introduce a constraint that relates a pixel intensity to its neigh-
bors via a correlated neighborhood model that selects the smoothest possible
reconstruction consistent with observed intensity variations in a local region.
The conditional probability of observing the intensities within the local neigh-
borhood surrounding a central pixel with RCS a (illustrated in Figure 6.3) can
be expressed through the conditional probability P100(ZIa). In the absence of a
known form for this PDF let us adopt a gamma distribution such that
^(/k) = N ^ e x P r - ^ l (6.16)
VoV T(a) L a J
Figure 6.3 Neighborhood diagram (for m = 9 pixels).
/ \ m-\
; =0
j\L a
=/ TL-I
expr TJ Im-I/ \ Ja-I
exp |~ /
Hence, the MLE for the cross section of the center pixel, a, is given by
L + mal/L .^ .
a
CML = - T T JZ- (6-18)
1 + ma/L
where a and / are estimated over a 3 by 3 neighborhood in the same way
as before. This filter can be written in a form similar to the MMSE result,
namely,
Q m
— ~ + °"CMAP {ma + L + l - v ) - U 0 - mal = 0 (6.21)
Estimates for / and a are obtained over the local neighborhood. The correlated
neighborhood model implies that fluctuations within the neighborhood should
be small so that a should be large. The full range of fluctuations would be
expected over the larger window that is used to estimate JJL and v as in (6.12).
The results of applying this algorithm to the original image of Figure 6.1
are illustrated in Figure 6.2(e). Upon comparing this result with Figure 6.2(d)
we observe that the speckle in uniform regions is further suppressed using the
MAP reconstruction. This illustrates the advantage of including more prior
knowledge where it is available. However, the algorithm has insufficient speckle
reduction performance.
With the exception of gamma MAP reconstruction, the filters considered so far
have not taken account of any structure. However, we have observed in MMSE
and gamma MAP despeckling that the presence of bright objects within the
filter window has a deleterious effect via the introduction of artifacts. Results
with smaller windows, such as CML and CMAP, show that these artifacts can
be reduced at the expense of degraded speckle reduction. It is instructive to
enquire whether a structural approach alone could provide useful despeckling
performance. Crimmins [2] proposed a geometrical filter based on iterative
complementary hulling. This does not incorporate any specific model for
speckle statistics but depends on the shape of features in the scene. The image
is treated with a set of small (4-pixel) templates that are intended to respond to
structural features. At the same time, these templates average over speckle
fluctuations. The algorithm is nonlinear, iterating eight times, and actually
modifying the image 64 times in the process, leading to speckle reduction. Thus,
this filter offers a potential for both adaptive structure enhancement and speckle
reduction and can be regarded as an adaptive multilook despeckling method
where the window over which averaging is performed is tuned to the scene,
albeit imperfectly. This effect is apparent in Figure 6.2(f), which illustrates the
result of applying the algorithm to Figure 6.1. The method has provided
effective speckle reduction in homogeneous regions and generally appears to
give better speckle reduction, structure preservation, and lack of artifacts than
the other methods discussed so far. However, some small-scale structure has
been lost, which is to be expected. This tendency to lose or degrade the shape
of small objects can be a serious disadvantage. Nevertheless, this reconstruction
might be considered to give visually the best reconstruction of those illustrated
in Figure 6.2.
a
MMSE =x+k(x-x) (6.22)
V~ V~ - l/Z/
k = ^ = 1* V \ (6.23)
K vx(i + i / r )
Equations (6.22) and (6.23) have a form_similar to those for the first
iteration. However, the sample estimates x and Vx are now obtained by apply-
ing the filter window to the result of the previous iteration. In addition, the
progressively despeckled data have an effective number of looks Z', that in-
creases as iterations proceed. Z' is estimated from the normalized variance of the
complete image.
An MMSE reconstruction of Figure 6.1, after 8 iterations with an 11 by
11 window, is shown in Figure 6.4(a). Speckle has been smoothed considerably,
but excess noise around strong objects is still apparent, though reduced com-
pared with a single look. The primary drawback of the reconstruction, however,
is the failure to preserve structure in the scene. Indeed, upon continuing to 200
iterations, nearly all structure is averaged out. It appears that MMSE does not
converge to the RCS but continues to smooth out all variation because approxi-
mately half of the measurements of the statistics in the window yield a variance
lower than that expected for speckle and so the algorithm adopts the local
average value, leading to progressively increased smoothing.
(a) (b)
Figure 6.4 Despeckled version of Figure 6.1 after 8 iterations: (a) MMSE and (b) structural
gamma MAP.
intensity image described by the speckle conditional probability P(I \ cr).
Though convergence is slower, there is less chance of progressively increasing
radiometric distortion. A similar approach will be adopted in all the following
iterative reconstruction methods.
As the iterations of the gamma MAP algorithm proceed, we might hope
that x converges to a and the PDF for x converges to (6.11). Hence, the form
of the iterated MAP joint probability is identical to (6.13), leading to the same
form for the gamma MAP solution for the cross section [13], namely,
T>rr2
- ^ - + CT1^p (L - v + 1) - LI = 0 (6.24)
where I and L are the original intensity and number of looks, respectively,
and JUL and v _are estimated from the current iteration, so that
JUL = x and v = 1/Vx . In fact, x does not converge to a; artifacts will be in-
troduced as described earlier that will continue to be reconstructed in the
same manner as genuine RCS variations.
The consequences of applying 8 iterations of the structured gamma MAP
algorithm within an 11 by 11 window are illustrated in Figure 6.4(b). Homo-
geneous regions are clearly smoothed without the loss of structure noted with
the artifacts around features are still clearly visible as expected. Upon extending
to 200 iterations, there is little change in the quality of the reconstruction. Some
additional smoothing of uniform regions is apparent, structure is preserved, and
the artifacts around features are retained. The spatial extent of these artifacts can
be reduced, as mentioned in Section 6.5.1, by reducing window size. As ex-
pected, this degrades the despeckling but can be rectified by iteration. However,
for a 3 by 3 window, although artifacts are reduced, the structure in the scene is
smoothed out to a greater extent than with the larger window. A window
dimension could be selected to provide an acceptable compromise between
artifact reduction and oversmoothing.
1 + ma/L
where I0 and L again correspond to pixel intensity and the number of looks
for the original data and x and a are estimated from the current iteration
over the neighborhood. While (6.25) has the same form as for the first pass
in (6.18), iteration causes the pixel values to become increasingly correlated,
violating the assumptions of the derivation. Thus, m is now the effective
number of independent pixels that can be estimated from the ACF within
the neighborhood.
The result of applying this algorithm for 8 iterations using a 3 by 3
neighborhood is illustrated in Figure 6.5(a). Speckle is considerably reduced
compared with the single-pass result in Figure 6.2(d), though it is still visible,
indicating that more iterations are necessary. Detail is also better preserved
than with the MMSE reconstruction in Figure 6A, and artifacts are much
less obvious than with either MMSE or gamma MAP. However, increasing
to 200 iterations leads to the oversmoothed result in Figure 6.5(c). Notice
that shadows behind the hedge, which were clearly visible after 8 iterations,
have largely disappeared. Thus, this approach does not appear to converge
satisfactorily and must be rejected.
Tjrt-2
— 0 ^ - + o- C M A P (ma + L + l - v ) - LI0- max = 0 (6.26)
Again the original intensity I0, and number of looks L, is used; a, x, and m are
estimated over the small neighborhood for the current iteration, while fx and v
retain their values from the original image, estimated over the larger window.
The general behavior of this algorithm would be expected to be similar to that
for the previous ML version in Section 6.8.3, with the exception of an improve-
ment due to the inclusion of the a priori PDF.
Results for 8 and 200 iterations with a 3 by 3 neighborhood and 11 by 11
window are shown in Figure 6.5(b,d). Based on a comparison with CML after 8
iterations (in Figure 6.5(a)), it is apparent that speckle has been smoothed better
in homogeneous regions and that detail is better preserved. Both are conse-
quences of the increased prior knowledge being exploited. Upon increasing to
Next Page
(a) (b)
(C) (d)
Figure 6.5 Iterated despeckled version of Figure 6.1:8 iterations of (a) CML and (b) CMAP and
200 iterations of (c) CML and (d) CMAR
200 iterations (in Figure 6.5(d)), however, the reconstructions display excessive
oversmoothing. Thus, both the correlated neighborhood models fail to converge
to the RCS as the number of iterations increases and should be rejected.
The results illustrated in Figures 6.4 and 6.5 indicate that the filters described
in Section 6.8 are capable of attaining very good speckle reduction in uniform
Previous Page
(a) (b)
(C) (d)
Figure 6.5 Iterated despeckled version of Figure 6.1:8 iterations of (a) CML and (b) CMAP and
200 iterations of (c) CML and (d) CMAR
200 iterations (in Figure 6.5(d)), however, the reconstructions display excessive
oversmoothing. Thus, both the correlated neighborhood models fail to converge
to the RCS as the number of iterations increases and should be rejected.
The results illustrated in Figures 6.4 and 6.5 indicate that the filters described
in Section 6.8 are capable of attaining very good speckle reduction in uniform
regions. However, their symmetrical response tends to smooth out directional
features. This suggests that the comparatively good result achieved with the
Crimmins filter, shown in Figure 6.2(f), does not stem solely from iteration to
reduce speckle but also from its response to structures within the scene. There-
fore, in this section we introduce structure sensitivity into the correlated neigh-
borhood model by subdividing the neighborhood into a set of templates that
provide directional information [14-16]. Instead of calculating CML or CMAP
solutions for the central pixel using all pixels in the neighborhood, a set of 12
different solutions can be produced corresponding to each of the configurations
illustrated in Figure 6.6. Structure can then be preserved by selecting the most
probable configuration, based on (6.25) or (6.26), respectively. Unlike all the
previous filters, which were aimed at providing local optimization, this new
approach is intended to derive the global MAP solution for RCS, given the
constraints imposed by the different models. As such, it is a member of the
well-known class of global optimization problems. A familiar difficulty with
attempting analytic optimization is that, if analytic gradient descent techniques
are adopted to find the maximum (or minimum), the reconstruction can
become trapped in some local optimum rather than progressing to the global
solution. Well-designed optimization methods try to avoid this difficulty and
derive the global solution. One such generic technique is simulated annealing.
Let us consider how this approach can be combined with the structured neigh-
borhood model to provide a global MAP reconstruction of the RCS.
Figure 6.6 The 12 allowed configurations around the central pixel in the structured correlated
neighborhood model.
the RCS is modeled as a Markov random field that is governed by the Gibbs
distribution describing the conditional probability of the RCS cr, given a noisy
estimate x. Thus,
/>H*) = ^ e x p [ - ^ ] (6.27)
where Z is a normalizing constant; T is a constant termed the distribution
temperature, by analogy with true annealing; and W represents a potential
function that encapsulates the data multipoint statistics over the entire image,
or at least as much as is feasible. Simulated annealing is guaranteed to find the
global minimum of such a system provided that [18]:
T>—^ (6.28)
In(I + $k)
While this scheme is guaranteed to find the values of a that yield the global
minimum of W, both the calculation of the joint probability over the entire image
and randomly perturbing the pixel value are extremely time consuming.
Significant speed improvements can be achieved if the objective function is
defined locally rather than over the complete image [14—16,18—20]. There need
be no restriction on correlation between pixels, though the general form could be
limited to a specific model characterized by variance and correlation length
parameters, for example [20]. Even with this simplification, the algorithm would
still be impractical if every pixel could take an arbitrary value of RCS. One way to
reduce complexity is to restrict the number of possible values that the intensity
can take [18,20]. The output then takes the form of a crude quantization that is
unsuitable for RCS estimation. Early methods also encountered difficulty in
defining the prior knowledge to be exploited. In particular, it was found that
heuristic models for real scenes gave poor results [19]. White [14—16] proposed a
novel simulated annealing scheme that restricted the number of configurations
for which the potential is defined to those illustrated in Figure 6.6 rather than the
total number of possible final states of the entire image. This enables a continuous
valued output to be implemented, suitable for RCS estimation.
White [14-16] also proposed a means for overcoming the second time-con-
suming step. He adopted a deterministic ML solution for the center pixel of each
of the 12 candidate configurations corresponding to the prior correlated neigh-
borhood model. This departure from strict annealing means that there is no
guarantee of ultimate convergence to the global MAP solution, though it speeds
execution considerably. As we will demonstrate, reconstructions are visually satis-
factory and also achieve a good quality measure, as defined in Section 6.2.
Let us now examine how the potential function W in (6.27) can be related
to the correlated neighborhood model. Following White [14-16] we define W
in terms of the log likelihood, X, for the local neighborhood by
Thus, the global MAP solution for X corresponds to the minimum of W. The
structured neighborhood model considers the 12 individual 3-pixel configura-
tions in the neighborhood separately. The original speckle conditional PDF, the
intensity of the center pixel, and the number of looks are retained.
In fact, simulated annealing only approximates the global MAP solution.
It is therefore important to select operating parameters to derive an effective
solution. As noted in (6.27), the algorithm depends on an annealing tempera-
ture to prevent convergence on local minima. The choice of initial temperature
T0 and its change per iteration are crucial to the success of the algorithm. If T0
is too low the solution converges on local minima, whereas if it is too high the
reconstruction tends to be too smooth. Similarly, if ^decreases too rapidly there
is little probability of the 3 pixels in a configuration belonging to the same
distribution and the technique fails. On the other hand, if it decreases too slowly
convergence is delayed and the image is smoothed too much. It must be
appreciated that there is no absolute "best" reconstruction available using this
method. In examples that follow we adopt a cooling schedule based on T0 =
2.5 and (3 = 4, which preserves details of weak targets within the scene.
The conditional probability for configuration j is then given by (6.27)
with Wj = A. •, and the CDF of the configuration probability is given by
I0 +a(Xl +x2)/L
1+ 2a/L
The ML solution with a Gaussian model for both speckle and the local neigh-
borhood fluctuations is given in [14—16]. However, this reconstruction ap-
peared too smooth, so an edge detection stage was introduced, resulting in a
segmentation, rather than reconstruction, algorithm.
The equivalent analysis for the ACMAP algorithm yields a conditional
probability for the Kth. configuration similar to (6.31) except that it includes the
a priori PDF for (J10 similar to that in (6.20). The MAP solution for (J10 similar
to (6.21) and (6.26), is then given by the quadratic expression
VCTicMAp(j0
+ ^ A cMAp(^)(2a + L+ 1-v)-U0- Ct(X1 + * 2 ) = 0
M-
(6.33)
where JUL and v retain their values from the original image, estimated over the
larger window, as in Section 6.8.4.
x U f f e T ^ L ^ W ! ) ] (634)
7 = ^xJV1 (6.35)
/ ^ - 7Q + ( a /^)(*i A/ + x2y) _ I0 + ( 2 a / Z ) V ^
O" ACML (AJ {o.5b)
I+ 2a/L I+ 2a/L
A similar derivation can be performed when the a priori PDF for (JK is
introduced. The MLE for 7 is the same as (6.35), leading to a quadratic
expression for the MAP solution for a of the form
(C) (d)
Figure 6.7 Structured neighborhood annealed reconstructions of Figure 6.1 after 8 iterations:
constant RCS for (a) ACML and (b) ACMAP, and sloping RCS for (c) ACML and (d)
ACMAR
The ACMAP solution for the sloping RCS model, shown in Figure 6.8(d), re-
moves these. Since all reconstructions approximate to a global optimum, given
the slightly differing initial constraints, it is difficult to select one solution in pref-
erence to another. However, it is apparent that the fields have been broken up with
distinct edges between regions where the constant RCS has been assumed in Fig-
ure 6.8(a,b), whereas most of these edges have been removed with the sloping
RCS model in Figure 6.8(c,d). This "terracing" effect is an artifact of the constant
RCS model, not a property of the data.
Generally, there is little to choose between these annealed reconstructions.
All represent a considerable advance visually on any of the techniques described
(a) (b)
(C) (d)
Figure 6.8 Structured neighborhood annealed reconstructions of Figure 6.1 after 200 itera-
tions: constant RCS for (a) ACML and (b) ACMAR and sloping RCS for (c) ACML and
(d) ACMAR
earlier, in terms of both speckle reduction and feature preservation. ACML with
constant RCS or ACMAP with sloping RCS appear to represent uniform
speckle regions marginally better, indicated by the results after 8 iterations. The
terracing artifact, introduced by the constant RCS model, may be significant in
some remote sensing applications where slow variations in RCS are important,
in which case the ACMAP with the sloping RCS model is to be preferred. Any
difference in radiometric preservation cannot be judged visually and can only
be tested quantitatively, as described in the next section.
6.10 Comparison of Reconstruction Algorithms
(C) (d)
Figure 6.9 Ratio of original image to despeckled reconstruction: (a) Crimmins filter and 200
iterations of (b) gamma MAP, (c) CML, and (d) CMAR
the ratio image can vary, demonstrating yet another failure of the reconstruction
to follow the speckle model. In this comparison we have confined our attention
to those algorithms that appear to give a visually satisfactory reconstruction. It
is immediately obvious that all fail to yield uncorrelated residual speckle, with
the ratio images for CML (Figure 6.9(c)), CMAP (Figure 6.9(d)), and the
Crimmins filter (Figure 6.9 (a)) showing the greatest depth of modulation by the
underlying structure while the annealed reconstructions show considerably less.
The ratios for gamma MAP and annealed reconstructions reveal artifacts sur-
rounding any structure in the scene with gamma MAP having the greater spatial
extent due to the window size adopted.
(a) (b)
(C) (d)
Figure 6.10 Ratio of original image to annealed reconstruction after 200 iterations: constant
RCS for (a) ACML and (b) ACMAR and sloping RCS for (c) ACML and (d) ACMAP.
Algorithm Mean SD
result, whereas gamma MAP has converged after a few iterations. Thus, ACML
requires a total execution time an order of magnitude longer than simple gamma
MAP or about three times longer than the structured version.
These timings show that introducing a gamma prior PDF in a MAP recon-
struction increases execution time significantly, for example, gamma MAP vs.
MMSE, CMAP vs. CML, and ACMAP vs. ACML. The corresponding results for
ACML and ACMAP in Figure 6.8 suggest that the most important consideration
is whether to adopt the sloping region model to avoid the terracing artifact. This
tends to slow down execution, except for ACMAP. Since terracing is only visible
after a large number of iterations, it can be ignored in many applications.
6.11 Discussion
(b)
Figure 6.11 High-resolution (3m) DRA SAR image near Bedford: (a) original image and (b)
ACMAP reconstruction using sloping RCS model.
age quality shows little trace of speckle. Clearly this technique presents a very
powerful method for removing speckle. By way of contrast, a 24-m resolution
ERS-I PRI image of the Feltwell area in the United Kingdom and its reconstruc-
tion are shown in Figure 6.12(a,b). Many of the features that dominated the 3-m
resolution image, such as hedges and individual trees, are averaged out, but field
(a)
(b)
Figure 6.12 Low-resolution (24m) ERS-1 SAR image of the Feltwell area: (a) original image
(Copyright ESA, 1992) and (b) ACMAP reconstruction using sloping RCS model.
structure shows up quite clearly. It is apparent, however, that the reconstruction in
Figure 6.12(b) fails to preserve many of the thin dark features in the original im-
age, such as large sections of the drainage channel and various roads, which are
clearly distinguishable to the human observer. The effect is present, though not to
so great an extent, in the higher resolution imagery in Figure 6.11. This failure to
capture real structure can only be overcome at present by adopting a system reso-
lution with enough pixels at the scale of interest for the filters to respond ade-
quately.
While the theoretical framework of this chapter is essential for a proper
understanding of the algorithms, users are principally concerned with selecting
the best algorithm for their application. This choice is crucially dependent on
the application and differs considerably between (say) a military application
concerned with the detection and recognition of small targets and a remote
sensing application attempting to classify crops. The former is primarily con-
cerned with structural integrity for recognition purposes, so that bias is less
damaging than in remote sensing, which requires accurate RCS estimation. This
issue can be highlighted by examining the four small objects visible in the dark
field near the bottom left of Figure 6.1. For a military application it would be
important that these be clearly reconstructed without any reduction in the SCR
in the immediate vicinity. Only iterated ACML and ACMAP with constant or
sloping RCS, shown in Figure 6.8, satisfy this requirement. Thus, the other
algorithms are probably unsuitable for military applications. On the other hand,
most algorithms yield acceptable reconstructions for the RCS of extended
clutter away from structure in the scene, once the bias is corrected. Gamma
MAP is extensively used in this context [11,12].
References
[1] Pratt, W. K., Digital Image Processing New York: Wiley, 1978.
[2] Crimmins, T. R., "Geometric Filter for Reducing Speckle," Appl. Opt., Vol. 24, 1985,
pp. 1438-1443.
[3] Lee, J. S., "A Simple Speckle Smoothing Algorithm for Synthetic Aperture Radar Images,"
IEEE Trans. Syst. Man Cybern., Vol. 13, 1983, pp. 85-89.
[4] Caves, R. G., S. Quegan, and R. G. White, "Quantitative Comparison of the Performance
of SAR Segmentation Algorithms," IEEE Trans. Image Proc, submitted 1996.
[5] Lee, J. S., "Digital Image Enhancement and Noise Filtering by Use of Local Statistics,"
IEEE Trans. Pattern Anal. Mach. Intell,Vol 2, 1980, pp. 165-168.
[6] Frost, V. S., J. A. Stiles, K. S. Shanmugan, and J. C. Holtzman, "A Model for Radar Images
and Its Application to Adaptive Filtering of Multiplicative Noise," IEEE Trans. Pattern Anal.
Machine IntelL, Vol. 4, 1982, pp. 157-166.
[7] Kuan, D. T., A. A. Sawchuk, T. C. Strand, and P. C. Chaval, 'Adaptive Restoration of
Images with Speckle," IEEE Trans. Acoust. Speech Signal Process., Vol. 35, 1987,
pp. 373-383.
[8] Oliver, C. J., "Review Article—Information from SAR Images," / Phys. D: Appl. Phys., Vol.
24, 1991, pp. 1493-1514.
[9] Lopes, A., E. Nezry, R. Touzi, and H. Laur, "Structure Detection and Adaptive Speckle
Filtering in SAR Images," Int. J. Remote Sens., Vol. 14, 1993, pp. 1735-1758.
[10] Lee, J. S., "Refined Filtering of Image Noise Using Local Statistics," Comp. Graph. Image
Proc, Vol. 17, 1981, pp. 24-32.
[11] Le Toan, T., F. Ribbes, L.-F. Wang, N. Floury, K.-H. Ding, J. A. Kong, M. Fujita, and T.
Kuroso, "Rice Crop Mapping and Monitoring Using ERS-I Data Based on Experiment
and Modelling Results," IEEE Trans. Geosci. Remote Sens., Vol. 35, 1997, pp. 41-56.
[12] Leysen, M. M., J. A. Conway, and A. J. Sieber, "Evaluating Multi-Temporal ERS-I SAR
Data for Tropical Forest Mapping: Regional Mapping and Change Detection Applica-
tions," Proc. of the Second ERS-I Symposium, Hamburg, Germany, ESAVoI. SP-361, 1994,
pp. 447-452.
[13] McConnell, L, and C. J. Oliver, "Comparison of Annealing and Iterated Filters for Speckle
Reduction in SAR," Europto Conf. on SAR Image Analysis, Simulation and Modelling,
Taormina, Italy, SPIE Proc, Vol. 2958, 1996, pp. 74-85.
[14] White, R. G., "Simulated Annealing Applied to Discrete Region Segmentation of SAR
Images," DRA Memorandum 4498, 1991.
[15] White, R. G., "A Simulated Annealing Algorithm for SAR and MTI Image Cross-Section
Estimation," Europto Conf. on SAR Data Processing for Remote Sensing, Rome, SPIE Proc,
Vol. 2316, 1994, pp. 137-147.
[16] White, R. G., "A Simulated Annealing Algorithm for Radar Cross-Section Estimation and
Segmentation," SPIE Int. Conf. on Applications of Artificial Neural Networks Vf Orlando, FL,
SPIE Proc, Vol. 2243, 1994, pp. 231-241.
[17] Metropolis, N., A. W Rosenbluth, M. N. Rosenbluth, A. H. Teller, and E. Teller, "Equa-
tions of State Calculations by Fast Computing Machines," /. Chem. Phys., Vol. 21, 1953,
pp.1087-1091.
[18] Geman, S., and D. Geman, "Stochastic Relaxation, Gibbs Distributions and the Bayesian
Restoration of Images," IEEE Trans. Pattern Anal. Mach. Intell, Vol. 6, 1984, pp. 721-741.
[19] Wolberg, G., and T. Pavlidis, "Restoration of Binary Images Using Stochastic Relaxation
with Annealing," Pattern Recog. Iett., Vol. 3, 1985, pp. 375-388.
[20] Kelly, P A., H. Derin, and K. D. Hartt, "Adaptive Segmentation of Speckled Images Using
a Hierarchical Random Field Model," IEEE Trans. Acoust. Speech Signal Process., Vol. 36,
1988, pp. 1628-1641.
[21] Cvijovic, D., and J. Klinowski, "Taboo Search: an Approach to the Multiple Minima
Problem," Science, Vol. 267, 1995, pp. 664-666.
[22] McConnell, L, R. G. White, C. J. Oliver, and R. Cook, "Radar Cross-Section Estimation
of SAR Images," Europto Conf. on SAR Image Analysis, Simulation and Modelling, Paris,
SPIE Proc, Vol. 2584, 1995, pp. 164-175.
7
7.1 Introduction
In the previous chapter the emphasis was directed toward reconstructing the RCS
at each pixel. This was based on exploiting phenomenological models for speckle
and RCS combined with a series of world models that introduced constraints into
the reconstruction. In this chapter the approach is again based on the product and
speckle models but relies on other forms of constraint to derive information. In
Section 7.2 we associate each pixel with a particular class, taken from a restricted
set representing RCS values. We demonstrate the need to average over many pixels
before such classification can be effective, which is only sensible if the boundaries
of the averaging window lie within a region of constant RCS in the image. This
leads to the cartoon model, described in Section 7.3, and its application in
segmentation. Finally, while the principles of these information-extraction tech-
niques may be elegant, most users are primarily concerned with their effectiveness
at fulfilling specific image-interpretation functions. Thus, we will consider quan-
titative image-quality measures for the different algorithms in Section 7.4. In
Section 7.5 a discussion of the principles underlying the exploitation of both
reconstruction and segmentation is presented.
Rather than estimating the RCS, as described in the previous chapter, we might
attempt to classify each pixel in terms of its RCS. Suppose that a restricted set of
possible RCS values is selected and each pixel assigned to the nearest class.
Speckle causes uncertainty in this assignment. An optimized classifier for the
known speckle model can readily be defined, based on the conditional prob-
ability of each RCS class, which is given by (6.5). The log likelihood for Z-look
SAR therefore has the form
Since Z is a known constant for the image, for ML classification we merely need
to identify that class k, that maximizes the quantity —In ak — L/dk for each pixel.
Different cross-section classes should be sufficiently separated so that speckle
does not cause too many misclassifications. For reasonable error rates, this
implies that class spacing should be related to the SD of image intensity, which
is (Jk / V Z .
An example of this form of ML cross-section classification, applied to the
test image in Figure 6.1, is shown in Figure 7.1 (a). The image intensity is
classified into 12 equally spaced predefined states. It is immediately apparent
that the process is dominated by speckle. Upon doubling the spacing so that
only 6 states are considered, the result in Figure 7.1 (b) is obtained. Even with
this coarse quantization, speckle still dominates.
These examples make the point that ML classification based on a single
pixel is rendered ineffective by speckle. Under these conditions, the PDF for a
(a) (b)
Figure 7.1 ML classification of the individual pixels from Figure 6.1 into RCS values: classified
into (a) 12 states and (b) six states, each twice the separation of those in (a).
region of constant RCS extends over a wide range of intensity, so it is mapped into
several different classes. Furthermore, any RCS fluctuations that are present tend
to be dominated by speckle. Since speckle limits pixel classification, it is reason-
able to inquire whether concepts evolved in earlier sections could be relevant. RCS
reconstruction was based on exploiting constraints that neighboring pixels were
similar. Multilook despeckling introduced this in the form of an average over a
window. Though this was unsuccessful in structured regions, it gave satisfactory
results in uniform regions. Averaging over neighboring pixels narrows the inten-
sity PDF so that classification may become possible. However, scenes contain
structures that may not be aligned with the window defined in this averaging
process. It is therefore essential to develop techniques for aligning such averaging
windows with image structure. In the next section we consider how this type of
approach can yield images that are segmented into regions of the same RCS.
Segmentation depends upon the introduction of a cartoon model that asserts im-
ages are made up of regions, separated by edges, in which some parameter such as
the RCS is constant. The number and position of these segments and their mean
values are unknown and must be determined from the data. Instead of attempting
to reconstruct an estimate of RCS for each pixel, segmentation seeks to overcome
speckle by identifying regions of constant RCS. In later sections we show that it
provides a powerful tool for image interpretation for two reasons.
Both these requirements are essential for rapid robust processing. They
will only be met when the segmentation process is driven entirely by the data
and the theoretical properties of SAR images. Only algorithms that meet these
requirements will be discussed here. The examples in this chapter correspond to
2-look high-resolution airborne SAR and lower-resolution 6-look ERS-I SAR
with widely different operating conditions. However, they are processed with
identical algorithms and parameter settings.
In principle, segmentation could be addressed as a process of fitting the
cartoon model to data with all possible positions of edge, number of regions,
and mean value. Such a global fit is feasible in one dimension [I]. However,
attempting to follow the same approach with two-dimensional scenes results in
a vast computation task, so we resort initially to heuristic methods that have the
aim of identifying regions in an image without actually trying every possible
configuration [2].
We can be faced with conflicting requirements in segmentation. It is
sometimes important to preserve resolution as far as possible, particularly in the
military context. Other applications may wish to identify regions that differ by
only a small change in RCS. The interplay between these requirements is evident
when we consider some previous approaches to segmentation. Many were
inadequate for segmenting single-look SAR, so the speckle was first reduced by
smoothing. However, smoothing should not precede segmentation but be in-
corporated into the process [3]. If smoothing does precede segmentation, algo-
rithms can be ineffective either because too small an averaging window leads to
inadequate speckle reduction [4] or because an overlarge window yields de-
graded resolution [5]. The difficulty stems from the adoption of a single-scale
smoothing algorithm rather than an adaptive filter of varying region size [6].
High-intensity differences, such as those from targets, can potentially be de-
tected over small regions; whereas low-intensity differences, important in re-
mote sensing applications, require averaging over large regions to attain
adequate sensitivity. The conflict between the different requirements is resolved
in an adaptive algorithm. This commences with small filters for small-object
detection; filter size then increases progressively to give improved sensitivity, at
the cost of resolution. In this manner any uniform region is detected at the
highest resolution at which sensitivity is adequate.
Three basic classes of segmentation algorithm can be identified that ex-
ploit the cartoon model in different ways, namely: edge detection (discussed in
Section 7.3.1), region merging (Section 7.3.2), and fitting (Section 7.3.5). In
Section 7.3.3, ML edge detectors are derived for inclusion into segmentation
schemes; their implementation is discussed in Section 7.3.4.
NO NO
Output previous segmentation Output previous segmentation
(a) (b)
Figure 7.2 Flow diagram for model-based segmentation algorithms: (a) RGW and (b) merge
using moments.
bright edges are detected with the smallest masks that are used first, that is, at
high resolution. These edges are retained and the pixels excluded from successive
stages of edge detection. Weak edges are subsequently detected with larger
windows. Thus, the algorithm adapts its resolution to signal strength.
The region-growing stage of the algorithm proceeds as follows.
Initially disks are large; subsequently, the size of the disks being added is
reduced and the same process is repeated. Eventually, every pixel is assigned to
a particular region and a complete region map constructed. The full cycle is then
repeated, yielding more sensitive edge detection. Iteration continues until the
overall coefficient of variation (SD/mean) reaches some preset threshold related
to the speckle statistics.
RGW was found to operate more successfully on amplitude rather than
intensity images. The result of applying RGW model-based segmentation to
Figure 7.3(a) is illustrated in Figure 7.3(b). Here, and throughout the chapter,
each segment is shown with each pixel set to the mean intensity of the segment as
calculated from the original pixel values. Figure 7.3 (b) contains a large number of
small regions. These result from the choice of threshold, which is related to the
detection probability and is the only user-selected variable for the process. How-
ever, some segments are probably anomalies arising from flaws in the algorithm.
Section 7.3.3 will describe how sensitivity could be improved by introduc-
ing an ML hypothesis test for the presence of an edge within the window, based
on the speckle and cartoon models. This section will also demonstrate that
RGW segmentation is not optimized for the most accurate choice of edge
position and will propose a corresponding ML solution.
(a) (b)
(C) (d)
Figure 7.3 Comparison of segmentation methods: (a) original image, (b) RGW7 (c) Gaussian
MUM, and (d) optimized MUM.
7.3.2 Region Merging
Region merging exploits the cartoon model more directly and attempts to grow
large uniform regions by merging neighboring small regions (which might
comprise only a single pixel). The crucial issue is whether the decision to merge
is reliable, which depends on the statistical criterion adopted to control merging.
In fact, an identical criterion can be used to describe both splitting and merging
[12,13]. One implementation of merging, merge using moments (MUM)
[14,15], is based on the Student t test. The sequence of operations, illustrated
in Figure 7.2(b), can be summarized as follows.
+ = NA™A+NB™B (J_+_n
NA+NB-2 {NA NJ
where var A, var B, and var{y4 + B} are the variances and NA and NB
are the number of pixels in each region, respectively.
3- Calculate the normalized difference t = A — B\/^v2ir{A + B} and
apply the Student t test for NA + NB — 2 degrees of freedom to the
data. This evaluates the probability that A and B should be merged.
4. Compile a list in order of priority for merging all pairs of adjacent
subregions.
5. Merge the highest priority region and repeat operation for new subregions.
The process is iterated until the false alarm probability for failing to merge
a constant region reaches some preset value. Thus, this algorithm has the
desirable property that operator controls are restricted to a single parameter, as
with RGW. The result of applying MUM to Figure 7.3(a), starting from a 2 by
2 pixel tessellation, is shown in Figure 7.3(c). The effect of the coarse initial
tessellation is evident. However, the number of anomalous small regions seems
to be lower than with RGW. A more rigorous comparison of segmentation
methods can be found in Section 7.7.1.
The Student t test is only rigorous when distinguishing between regions
with Gaussian PDFs, which is clearly inappropriate for speckle. Significant
performance improvement is possible through an ML derivation using the
correct statistics, given in Section 7.3.3.
7.3.3 ML Split/Merge Theory
Both splitting and merging depend on estimating whether two regions belong to
the same population or not. In this section we derive an ML test to determine this,
given SAR speckle statistics. In fact, there is more than one possible criterion for
which the test could be optimized. Either the edge-detection probability in RGW
or the merge probability in MUM is maximized or the edge-position accuracy is
optimized. We derive the ML estimators for both cases and demonstrate the type
of improvement they yield if introduced into existing algorithms.
Let us assume that we retain adaptive window sizes. The cartoon model
implies that a window should contain either a single uniform region or two
uniform regions with an edge at some unknown position within that window.
(Any other possibility will have been eliminated by detecting an edge in an
earlier, smaller window). The joint probability of these two states can be ex-
pressed in terms of the conditional probability P(ZIa) for each. Let us assume
that pixels 1 to k are in region A with RCS <JA, while pixels k + 1 to M are in
region B with RCS (JB. We consider a single dimension for simplicity. The results
for two dimensions can be easily obtained. The joint probability for the whole
window, given single-look intensity Ij at pixel j , is given by
k M
j=\ j=k+\
k M
i r /• i 1 r /•"
ex
=n-«p-- n — p—— (7.3)
j=l (JA I VA]j=k+l V3 L °"5_
The probability that the data are consistent with a merged region of mean
RCS (T0 has the same form with (JA = <JB = (J0. The log likelihood for split and
merged states, ^sp\lt(k) and A.merge, can be derived by taking logs of these joint
probabilities and a log difference estimator defined such that
s
M * ) V W - W
+ M In(T0 + ^ M (7.4)
where the mean intensities are estimated over the appropriate windows. JHie ML
estimates of RCS in each region are (JA=IA,(JB=IB, and (J0 = I0 . Note
that IA and IB are functions of k so that the MLE for edge position k is
obtained by maximizing [15,16]
r
PAARXM-k)= ^> I U " % J (7.6)
[1 +
R(M-k))
where R denotes the true ratio of mean values. If an edge is "detected" when r
is outside the range [tv Z2L the detection probability at position k, given that the
true edge is at this position, is given by
Pd=l-gP(r)dr
1
= !_ X-M) f k Y
k(M - k)Y{k)T(M -k){(M- k)R)
\ r h 1
XUk F\k M-k + 1- - —
[221L' ' ' (M-k)R_
r h Ii
— tk F k M- k + 1- - — [ C7 7 )
121 { }
L' ' ' (M-k)R\\
where 2F\i'] ls t n e Hypergeometric function [18]. The false alarm probability
Pfo corresponds to substituting R=I into (7.7). The threshold values tx and Z2
lie above and below 1 and can be related to the equivalent value of XD, which
can be derived from (7.5). Hence,
When the edge is at the center of the window, k = Mil and these combine to
the same threshold. This corresponds to the ratio detector proposed by Touzi et
al. [19] and Bovik [20], which is given by rT= min[r, l/r]. Equivalent numeri-
cal solutions for detection probability at other edge positions can also be per-
formed [16].
The equivalence of rT and r when the test edge is in the center of the
window is confirmed by a comparison of the detection probabilities for the ML
measure and Touzi ratio as a function of edge position, shown in Figure 7.4. The
two regions in these simulations have means of 1 and 8, and the results were
obtained over 104 samples. At the center both give identical detection prob-
ability, while the Touzi ratio is suboptimum elsewhere. The simulated results
were also compared with predicted values, showing close agreement [16]. Since
the Student t test was employed in the original version of MUM (in Section
7.3.2), we include simulations based on this for comparison. It is apparent that
adopting a Gaussian rather than the correct gamma PDF for speckle leads to
performance degradation.
This derivation accounts for different numbers of pixels in the two regions
and incorporates speckle statistics correctly. It can therefore be used directly in
place of the Student t test in the MUM segmentation scheme.
As mentioned earlier, there are two distinct criteria that can be applied to
edge detection. The first relates to the accuracy with which edge position is
determined and is described by (7.5) and the results of Figure 7.4. The associ-
ated filter geometry would be to fix a window position with respect to the scene
and scan the test edge, which will be called the fixed window, scanned edge
(FWSE) configuration. Figure 7.4 demonstrates that the detection probability
at a given position is maximized when the central edge in this window is aligned
with an edge in the RCS. This suggests that another useful edge-detection
geometry would be to scan a window with a fixed central test edge over the data,
which will be called the scanned window, center edge (SWCE) configuration.
Note that SWCE is equivalent to the Touzi ratio test [19].
However, the results of Figure 7.4 depend on knowing the actual edge
position in advance. In a real application this would not be the case and the
algorithm would determine both the presence and position of any edge. The
FWSE window is placed at some arbitrary position in the scene and the test edge
scanned. SWCE, on the other hand, scans the window over the scene. In each
Pdet
Edge position
Figure 7.4 The dependence of detection probability on edge position. ML, simulation (•) and
prediction (full line); Touzi ratio, simulation ((8)) and prediction (dotted line), Student
tmeasures, simulation only (Oamplitude, •intensity), joined by smooth curves for
clarity.
case, the position of greatest likelihood denotes the edge location. The response
at this position is then compared with an appropriate threshold to detect
whether it is significant, yielding a probability Pd that an edge is detected in the
correct position. It is also important to consider whether an edge is detected
anywhere in the window, not necessarily in the correct position. This is de-
scribed by the sum of all the individual values of Pd over the window and is
denoted by P tot . We would expect this to be maximized by the SWCE test.
However, neither Pd nor Ptot can be derived analytically if the edge position is
not known and we have to resort to simulation to compare the window
configurations [16].
In Figure 7.5 (a) we illustrate a comparison of P d for the two configura-
tions. A 16 by 1 window is used with the ratio of mean values varying between
1 and 16. Offsets of 0, ± 2 , ± 4 , and ± 6 pixels between the true edge position
and the center of the window are simulated for FWSE. Alternatively, SWCE
scans over the same simulated data. FWSE yields higher values of Pd for small
offsets ( < 4), whereas SWCE is preferable for larger offsets. Figure 7.5(b) shows
the corresponding comparison for Ptot. This shows that SWCE always yields
greater overall detection probability, as expected from Figure 7.4.
The choice of which configuration to adopt depends on the application.
SWCE yields a higher probability of detecting edges in the scene. However, they
are less likely to be in the correct position than with FWSE. In order to
maximize both performance measures, a two-stage configuration has been pro-
posed [16]. In the first stage the SWCE configuration is applied to optimize P tot .
The approximate edge position derived from this defines the fixed window
position for FWSE. The edge position is then refined, leading to an improved
value for Pd. Provided that the edge position derived in SWCE is mainly within
the central half of the fixed window for the second stage, this joint test was
shown by simulation [16] to yield optimum results for both performance
measures at the expense of slightly increased execution time.
R
(a)
Figure 7.5 Dependence of (a) P6 and (b) Ptot on configuration, difference in RCS, and original
edge position. Window 16 by 1 pixels; R = 1,..., 16; overall Pfa = 10%; 104 samples.
SWCE configuration (full line,O); FWSE configuration (dashed line, •, offsets 0, 2,
4, and 6 pixels with increasing dash length).
X=XAUnT" (7.9)
7= 1
where Nj and / are the size and estimated RCS of the yth region, respectively.
The segmentation process then corresponds to identifying the MLE for each
region position and RCS. As such it is a standard global optimization problem
PtC
R
(b)
Figure 7.5 (continued).
Another iteration?
YES
NO
Iteration at lower
YES temperature?
NO
Final merge
for given Pfa
In the absence of any shape constraint, this approach will tend to separate
the image into regions of bright and dark speckle if convergence is unconstrained.
This problem is overcome by introducing a curvature penalty term, analogous to
surface tension, into the objective function in (6.29), which penalizes irregular
shapes. This additional contribution to the objective function is defined by
where f{ denotes the segment label for pixel / , N1: is a set of neighboring pixels
(say 8), x is a constant weighting coefficient, and 8 is the Kronecker delta.
Hence, if pixels / and j do not have the same segment label, Ws is incremented
by X- Empirically a value of x — 0.05 seems to offer acceptable edge smoothing
and is used in all the examples of annealed RCS segmentation shown.
We illustrate the effect of this term in Figure 7.7. Figure 7.7(a) shows the
result of applying the simulated annealing algorithm without the curvature
penalty term. The false alarm probability was set at 10~5, which resulted in
1,000 regions being identified. Upon introducing the curvature penalty, the
number of regions was reduced to 832 with a segmentation as shown in Figure
7.7(b). The complicated edge structure in Figure 7.7(a), caused by following the
(a) (b)
Figure 7.7 Comparison of simulated annealing (a) without and (b) with curvature penalty.
bright and dark speckle, is largely removed in Figure 7.7(b). Small objects appear
reasonably well preserved and large areas of constant RCS have been identified,
compared with the other segmentation techniques illustrated in Figure 7.3.
In common with other simulated annealing implementations, there are a
variety of parameters that have to be determined, such as starting temperature,
number of inner loop iterations, cooling schedule, and number of outer loop
iterations. In addition, there are questions over the initial tessellation stage that
determine both the position and number of initial regions. In principle, simu-
lated annealing will migrate regions to those areas of the image that have a larger
density of objects. However, this may involve an excessive number of iterations.
Thus, it is advisable to ensure that the original tessellation assigns segments
suitably with greater density in regions of high contrast, as noted earlier.
(C) (d)
(e) (f)
Figure 7.8 Segmented reconstructions of Figure 7.3(a): (a) RGW (653 segments), (b) RGW
(2,409 segments; also shown as Figure 7.3(b)), (c) optimized MUM (662 segments;
also shown as Figure 7.3(d)), (d) SGAN (674 segments), (e) SGAN (2,389 segments),
and (f) original anneal (5,027 segments).
greatly improved, reconstruction quality. There is no evidence of speckle breaking
uniform regions into segments, as in Figure 7.8(b). Both techniques reconstruct
the hedges and their shadows similarly. In addition, buildings are represented by
a large number of segments. However, there are detailed differences in the recon-
structions of the four test objects at the bottom left. MUM (c) retains the four
expected ones and introduces a fifth, alongside the center of the objects to the
right. SGAN (d) includes this fifth object but fails to preserve the fourth object at
the top left of the original set. This difference arises from the detailed difference
between the heuristics of the local MUM operation and global SGAN. Upon
reducing the threshold level in SGAN so that 2,389 segments are found, as
illustrated in Figure 7.8(e), we observe that these same five objects are found.
However, additional segments are also detected, which may represent genuine
objects or false alarms due to speckle. If Figure 7.8(e) is compared with 7.8(b),
which shows the RGW result with approximately the same number of regions, we
note that SGAN does not generate "flecks" but concentrates segments in regions
with much structure, such as the buildings on the bottom right. Visually, there-
fore, this algorithm yields a more desirable type of segmentation than RGW.
We observe that the original annealed segmentation [23—25], shown in
Figure 7.8(f), contains a large number of segments. Some of these appear to
correspond to bright or dark flecks caused by speckle. This suggests that the
threshold should be raised to reduce the number of segments. However, even
with the original settings, the algorithm fails to reproduce the four objects in
the field on the bottom left of the image. Thus, it yields false segments while
failing to detect genuine objects. It has therefore been superseded by the more
recent version (SGAN) that is optimized for the true speckle statistics.
Without prior knowledge of the true scene content, we cannot make any
absolute distinction between the performance of these algorithms. The compari-
son so far indicates that the original RGW algorithm seems incapable of providing
an adequate reconstruction with the same number of regions as MUM or SGAN,
suggesting that there is a number of misplaced regions in RGW. Incorporating the
optimized split and merge criterion could yield a significant improvement, how-
ever. It is probably not the filter architecture that is inadequate but the edge test.
The corresponding ratios of the original image to the segmented versions
in Figure 7.8(a—d) are shown in Figure 7.9, plotted over the range 0.5 to 1.5. A
visual inspection provides insight into segmentation quality. For a perfect seg-
mentation, the ratio image should appear to be made up of pure speckle. Any
sign of structure indicates a failure of the algorithm. It is apparent that RGW
with the small number of segments, in Figure 7.9(a), has large departures from
pure speckle in regions with structure, such as the hedges and, particularly, the
buildings on the bottom left. Even increasing the number of regions still leaves
strong evidence of unwanted structure (Figure 7.9(b)). These effects are much
(a) (b)
(C) (d)
Figure 7.9 Ratio of original image in Figure 7.3(a) to corresponding segmented versions in
Figure 7.8 (a-d).
if the segments really consist of uniform regions, where TVis the image size, m the
number of segments, and ^ the number of pixels in the/th segment. Note that the
bias in the variance introduced for a small region can be quite large compared with
HL. On the other hand, this contribution makes comparatively little difference to
the sum over the whole image that is dominated by the larger regions. In this ex-
ample, N= 65536, m ~ 660, n- ~ 99, and L —2.2. Substituting these values into
(7.11) suggests that the SD of the intensity ratio should be 0.672 instead of the
value of 0.674 for the corresponding test in Section 6.10.1. Thus, the estimation
correction tends to reduce the observed SD of the ratio slightly. The observed val-
ues of SD are much greater than the uncorrected predictions and so we will ignore
the correction. It would, of course, be important to retain if making a detailed
study of the estimated SD in individual small segments.
Note that a segmentation can fail for two reasons: either the value of a
within a segment is incorrect or an edge between regions of different a is not
recognized or wrongly placed. Both will yield excess values of SD, and it is not
possible to differentiate between these causes of failure using this ratio test.
The numerical values of the image quality measure are listed in Table 7.1.
Segmentation yields a mean ratio of 1 since every pixel in the segmented scene
is assigned the average value over the appropriate segment. Thus, segmentation
preserves radiometric accuracy completely, unlike the reconstruction filters. For
L ~ 2.2 the predicted (uncorrected) SD for pure speckle would be 0.674 for
intensity and 0.345 for amplitude. It is clear that SGAN yields results consistent
with pure speckle; MUM is slightly worse; but RGW requires many more
segments to approach the speckle prediction. Even with 2,409 segments it still
gives a considerably larger fractional discrepancy than MUM. Thus, image
Table 7.1
Ratio Measures for Segmentation of the DRA 3-m SAR Image.
Table 7.2
Execution Time (seconds) for the Three Different Segmentation Algorithms as a Function
of Image Size (Measurements taken on Sun Sparc 10)
Image Size
(b)
Figure 7.10 Annealed segmentation applied to (a) high-resolution DRA image of Figure 6.11(a)
and (b) low-resolution ERS-1 image of Figure 6.12(a).
Reconstruction
or structure?
structure
filters segmentation
reconstruction
Speed or Speed or
quality? quality?
Simple Annealed
Anneal RGW/MUM segmentation
filters
Figure 7.11 Flow diagram for selection of algorithms for RCS exploitation.
• a 0 may not be constant over the field, caused by crop variability, ground
slope change, or varying soil moisture, for example. Many applications
wish to determine whether this is the case. Both RCS reconstruction
and segmentation provide an optimum smoothed reconstruction of a 0
as a function of position.
• The map from which the boundaries are derived may be incorrect. The
edge maps generated by segmentation offer the potential for identifying
such changes and therefore updating the map.
(C) (d)
Figure 7.12 Comparison of the RCS reconstructions achieved when different segmentation
algorithms are applied to Figure 6.12(a): (a) RGW (805 segments), (b) RGW (2,305
segments), (c) MUM (809 segments), and (d) SGAN (827 segments).
ous theory [26] can again be neglected. In common with the earlier segmenta-
tions on higher resolution data (see Table 7.1) SGAN appears closest to consis-
tency with the speckle model.
If the information required is contained in the edge maps, then it is
instructive to inquire whether the different segmentation routines would yield
a similar ordering of effectiveness from the point of view of edge structure. Edge
maps generated by RGW, MUM, and SGAN are shown in Figure 7.13. One
indication of the relative quality can be obtained by following the drainage
Table 7.3
Ratio Measures for ERS1 Feltwell Image
Figure 7.13 Comparison of the edge maps corresponding to the segmentations of Figure 7.12.
channel on the left. Only in SGAN is its presence evident over most of its
course. If we reject RGW, we observe that the details of the edge maps generated
by MUM and SGAN differ considerably, even though they give similar global
quality measures. SGAN appears to contain more small segments. Since the
total number of segments is similar, this indicates that it must also contain more
large ones. The impression is that these images contain more edges than the user
would like. Since these reconstructions are consistent with both speckle and
cartoon models, these excess edges are due to real features in the scene, implying
that a high-level scene description in terms of, for example, field boundaries,
overrides or rejects information present in the image. It is also possible to see
real features in the original image that have been lost in these reconstructions,
in particular some of the long, thin, dark features, such as roads and the drainage
channel. The human observer is obviously capable of exploiting additional
forms of prior knowledge about possible feature types, such as long, thin lines.
No equivalent capability is currently available for segmentation, though an
approach based on simulated annealing has been recently reported [27].
In the absence of ground truth it is impossible to make quantitative
statements about the quality of the edge maps generated by segmentation
routines. This type of consideration can be dealt with by designing a suite of
tests [26,28] to measure the response of different algorithms to important
features for the application. These might be long, thin structures or small
targets, for example. If the user's requirements could be adequately represented,
this test suite could be used to optimize the response of the different algorithms.
This is an area that requires considerable future development if the requirements
of different applications are to be represented in a quantifiable fashion.
References
[1] Delves, L. M., R. T. McQuillan, and C. J. Oliver, "A One-Dimensional Despeckling
Algorithm for SAR Images," Inverse Problems, Vol. 4, 1988, pp. 471-484.
[2] Delves, L. M., R. T. McQuillan, R. Wilkinson, J. B. E. Sandys-Renton, and C. J. Oliver,
"A Two-Dimensional Segmentation Algorithm for SAR Images," Inverse Problems, Vol. 7,
1991, pp. 203-220.
[3] Freitag, B. J., B. Guindon, A. J. Sieber, and D. G. Goodenough, "Enhancement of High
Resolution SAR Imagery by Adaptive Filtering," Int. Geoscience Remote Sensing Symp.,
IGARSS '83, Paris: ESA, 1983.
[4] Oddy, C. J., and A. J. Rye, "Segmentation of SAR Images Using a Local Similarity Rule,"
Pattern Recog. Lett., Vol. 1, 1983, pp. 443-449.
[5] AIi, S. M., and R. E. Burge, "New Automatic Techniques for Smoothing and Segmenting
SAR Images," Signal Process., Vol. 14, 1988, pp. 335-346.
[6] Rosenfeld, A., and M. Thurston, "Edge and Curve Detection for Visual Scene Analysis,"
IEEE Trans. Computers, Vol. 20, 1971, pp. 562-569.
[7] Abdou, I.E., and W. K. Pratt, "Quantitative Design and Evaluation of Enhancement/Thre-
sholding Edge Detectors," Proc. IEEE, Vol. 67, 1979, pp. 753-763.
[8] Frost, V. S., K. S. Shanmugan, and J. C. Holtzman, "Edge Detection for Synthetic Aperture
Radar and Other Noisy Images," Int. Geoscience and Remote Sensing Symp., IGARSS '82,
Paris: ESA, 1982.
[9] Grimsom, W. E. L., and T. Pavlidis, "Discontinuity Detection for Visual Surface Recon-
struction," Comp. Vision Graphics Image Proc, Vol. 30, 1985, pp. 316—330.
[10] White, R. G., "Low Level Segmentation of Noisy Imagery," DRA Memorandum 3900,
1986.
[11] White, R. G., "Change Detection in SAR Imagery," Int. J. Remote Sensing, Vol. 12, 1991,
pp. 339-360.
[12] Horowitz, S. L., and T. Pavlidis, "Picture Segmentation by a Tree Traversal Algorithm," /
ACM, Vol. 17, 1976, pp. 368-388.
[13] Gerbrands, J. J., and E. Backer, "Split and Merge Segmentation of SLAR Imagery: Segmen-
tation Consistency," IEEE 7th Int. Conf. Pattern Recognition, Vol. 1, 1984, pp. 284-286.
[14] Cook, R., and J. B. E. Sandys-Renton, "A Parallel Segmentation Algorithm (Merge Using
Moments) for SAR Images," Appls. of Transputers, Vol. 1, 1991, pp. 311-316.
[15] Cook, R., I. McConnell, and C. J. Oliver, "MUM (Merge Using Moments) Segmentation
for SAR Images," Europto Conf. on SAR Data Processing for Remote Sensing, Rome, SPIE
Proc, Vol. 2316, 1994, pp. 92-103.
[16] Oliver, C. J., D. Blacknell, and R. G. White, "Optimum Edge Detection in SAR," IEE
Proc.Radar Sonar Navig, Vol. 143, 1996, pp. 31-40.
[17] Oliver, C. J., I. McConnell, D. Blacknell, and R. G. White, "Optimum Edge Detection in
SAR," Europto Conf on SAR Image Analysis, Simulation and Modelling, Paris, SPIE Proc,
Vol.2584, 1995, pp. 152-163.
[18] Abramowitz, M., and I. A. Stegun, Handbook of Mathematical Functions, New York: Dover,
1970, Section 15.1.
[19] Touzi, R., A. Lopes, and P. Bousquet, "A Statistical and Geometrical Edge Detector for SAR
Images," IEEE Trans. Geosci. Remote Sens., Vol. 26, 1988, pp. 764-773.
[20] Bovik, A. C , "On Detecting Edges in Speckled Imagery," IEEE Trans. Acoust. Speech Signal
Process., Vol. 36, 1988, pp. 1618-1627.
[21] Cook, R., I. McConnell, D. Stewart, and C. J. Oliver, "Segmentation and Simulated
Annealing," Europto Conf on SAR Image Analysis, Simulation and Modelling II, Taormina,
Italy, SPIE Proc, Vol. 2958, 1996, pp. 30-37.
[22] Geman, S., and D. Geman, "Stochastic Relaxation, Gibbs Distributions and the Bayesian
Restoration of Images," IEEE Trans. Pattern Anal. Mach. IntelL, Vol. 6, 1984, pp. 721-741.
[23] White, R. G., "Simulated Annealing Applied to Discrete Region Segmentation of SAR
Images," DRA Memorandum 4498, 1991.
[24] White, R. G., "A Simulated Annealing Algorithm for SAR and MTI Image Cross-Section
Estimation," Europto Conf on SAR Data Processing for Remote Sensing, Rome, SPIE Proc,
Vol. 2316, 1994, pp. 137-147.
[25] White, R. G., "A Simulated Annealing Algorithm for Radar Cross-Section Estimation and
Segmentation," SPIE Int. Conf. on Applications of Artificial Networks !{Orlando, FL, SPIE
Proc, Vol. 2243, 1994, pp. 231-241.
[26] Caves, R. S., S. Quegan, and R. G. White, "Quantitative Comparison of the Performance
of SAR Segmentation Algorithms," IEEE Trans. Image Proc, submitted 1996.
[27] Hellwich, O., "Line Extraction from Synthetic Aperture Radar Scenes Using a Markov
Random Field Model," Europto Conf. on SAR Image Analysis, Simulation and Modelling II,
Taormina, Italy, SPIE Proc, Vol. 2958, 1996, pp. 107-113.
[28] Delves, L. M., R. Wilkinson, C. J. Oliver, and R. G. White, "Comparing the Performance
of SAR Image Segmentation Algorithms," Int. J. Remote Sensing, Vol. 13, 1992,
pp. 2121-2149.
8
Texture Exploitation
8.1 Introduction
Techniques for removing speckle from SAR images and deriving the RCS were
described in Chapters 6 and 7. Information is then carried at the single-pixel level.
However, this is not the only type of information contained in a SAR image. We
illustrate this in Figure 8.1 with an example of airborne SAR imagery from the
Tapajos region of the Amazon rain forest, obtained with the CCRS C-band
system, as part of the SAREX program [1-3]. These 6-m resolution, 5-look data
were obtained with illumination from the right at an incidence angle of about
60 degrees. A highway runs diagonally from top right to bottom left. The region
to the left of this is comprised of primary forest; while that to the right is made up
of a mixture of primary forest, secondary forest, and clearings with pasture and
crops.
Suppose that the remote sensing task is to distinguish those regions that
correspond to primary forest from secondary forest and clearings. It is apparent
that the radar return in this image falls into two categories. One, corresponding
to secondary forest and clearings, has no fluctuations above those expected for
5-look speckle. This is caused by the fact that vegetation is sufficiently dense and
uniform so that the RCS is effectively constant. The other, corresponding to
primary forest, has appreciable excess fluctuations caused by tall trees penetrat-
ing the canopy and resulting in bright tree crowns with associated shadow.
Visually, it appears that there is no relevant information in the mean RCS. Some
regions of primary forest have a larger RCS than the secondary forest and
clearings; others have a smaller value. This follows from the fact that both
Figure 8.1 SAREX image of part of the Amazon rain forest in the Tapajos region.
returns correspond to wet vegetation canopies with essentially the same water
content, having the same RCS at the 6-cm wavelength used. Thus, the informa-
tion required to discriminate between the two clutter types resides in image
texture. This cannot be estimated from a single pixel but requires a finite
window to characterize local statistics. The situation is further complicated if
the texture is correlated, as we will show in Chapter 9.
We demonstrated in Chapter 5 that clutter textures could be described by
the normalized variance. Thus, single-point texture statistics can be utilized as a
means of characterizing clutter in image interpretation. This chapter is con-
cerned with establishing optimum means for extracting and exploiting this
textural information.
In Section 8.2 we discuss how texture information can be derived without
any knowledge of the data distribution. However, where prior knowledge of the
form of PDF is available, it should be exploited. Model-based texture parameter
estimation is discussed in Section 8.3. ML texture estimators for various forms
of PDF are derived in Section 8.3.1, leading to texture measures that can be used
in texture characterization. Assuming that clutter textures are actually K-distrib-
uted, the associated statistical uncertainty in the estimation of the order parame-
ter for the different measures is derived in Section 8.3.2. The normalized log
measure is then exploited in a texture analysis of rain forest data in Section 8.3.3.
The issue of texture classification into regions of different parameter values is
addressed in Section 8.4. Section 8.5 provides an analysis and discussion of
techniques for optimum texture segmentation. Finally, we compare the quality
and speed of algorithms for texture exploitation in Section 8.6.
—,,n-P-^in-[^t,,] (8..)
where the bars denote ^ample averages. Though both / and/ 2 are unbiased,
since (l) = (I) and (l2) — (I2), the estimated variance var / is biased since
((/) 2 ) ¥= (I)2. In fact, the mean and variance of this estimator are given by
exactly, and
var[^7] = ^((/4)-4(/3)(/)-(P} 2
^ s ii - 1 (8.4)
/2 /2
The uncertainty in this estimator depends on errors in both numerator and
denominator, which are clearly not independent. The expected value of this
estimator can be derived by perturbing both numerator and denominator about
their mean value [6] leading to
which is biased to O(l/N). The variance of the quantity can be derived following
a similar process leading to
j^i]-±.№-turn+<u$-sen ,>
(8
LP J N((lf (if (if (If )
Let us examine the difference between true and self normalizations for a
K-distributed intensity with order parameter v, by substituting appropriate
moments from (5.14). With true normalization the bias is
var[var/| 4
/ ? n
co a/-\
L J ^ l h + - + ^ + ^ (8.8)
4 2 3
(I) NK v v VJ
and a variance of
2
- S B H ) H ) H ) ™
With no RCS variation (v —> ©o) the variance with true normalization has
a bias of \IN compared with 2IN for the self-normalized variance. In the
opposite limit with strong RCS fluctuations (v —> 0), the biases are 21 Nv and
12/Nv2, respectively. This shows that the bias of the self-normalized estimate is
greater than that with true normalization. If we consider the variance of the
estimators, we find that the estimator with true normalization has a variance of
8/'Ny for v —> oo, whereas that for the self-normalized estimator is AIN. In the
opposite limit (v —» O), the variances are 1AAINv3 and SOINv3, respectively.
Thus, the variance of the self-normalized estimator varies monotonically be-
tween a factor of 2 and 1.8 less than that with true normalization. Though
self-normalized estimators introduce more bias, they have smaller uncertainties
and so are generally a more useful quantity to measure. They will be adopted
for all the discussion that follows.
where P(I9 st) is the ith constituent of the mixture, sz is the vector of parameters
for the kh distribution, wi is its weight, and q the number of components in the
mixture. The estimated PDF is fitted, using ML methods based on expectation
maximization [9], to determine appropriate weights, parameter values, and
number of components for the mixture distributions. In principle, increasing
the number of components improves the fit; however, individual parameter
values are less accurately determined.
As an example, we show a three-component mixture distribution fit to the
PDF of the log of a K-distributed variable in Figure 8.2 [10]. The original
distribution is far too asymmetric to be represented by a single Gaussian com-
ponent. However, after three components have been summed the resultant is
close to the original shape. Note that the quality of the representation was
improved considerably by taking the log of the data. It would require many
more components to represent the PDF of the original data.
So far we have not exploited any model for the data in attempting to derive infor-
mation from the texture. However, Section 8.2.3 indicates that an optimum tex-
ture measure is determined by the form of the data. For the remainder of this
chapter we shall assume that a data model is known, characterized by its single-
point PDF. This PDF then determines optimum measures for texture parameters
(in Section 8.3), classification (in Section 8.4), and segmentation (in Section 8.5).
Figure 8.2 Mixture distribution fit (short dashes) to the PDF of the log of a K distribution (full
curve) made up of three Gaussian contributions (long dashes).
x _7Vln(2^)_^-^ ( 8 J 3 )
2 2V
\L = X and V = x* - x2 (8.14)
This Gaussian PDF could represent either the intensity or the amplitude
of the data. Self-normalized texture measures of the form
V1=^L-X (8.15)
and
/>(*) = - J — e J - f i ^ ^ l (8.17)
W2irW [ 2W
where P and W&K the mean and variance of In x. The associated log likelihood
is given by
N lnx
X = -"WtW) -NhTx- ( -®2 (8,8)
2 2W
and the MLEs for (3 and Why
VL = b 2 / - ( b / ) 2 (8.20)
(L+v)
o ( r \ 2 L+V-2 I YT
2
P(I) = - — / Jfv f 2 — (8.21)
T(L)T(V) {ILJ L V M* _
for Z-look SAR, where JUL and v are the mean and order of the intensity PDF,
respectively. In this case the associated log likelihood is
X= A H n ^ -N\nT(v) + ^-\n^
[T(L)] 2 [^.
+ JLIlZl]^j +NLK I2 f^ut (8.22)
V 2 ; I L \ jA, J j
Unfortunately there is no exact analytic form for the partial derivatives of
the last term in (8.22) which prevents a derivation of the MLEs. These have to
be determined by a numerical search over JUL and v to find the largest value of
X; the corresponding values of JUL and v are the MLEs. Since this is very time
consuming, it is useful to provide approximate forms for the K distribution that
have analytic solutions.
and
These reduce to the results for a gamma distribution as L —> °°, when
speckle contributions would be averaged out [15]. Note that the right-hand side
of (8.26) involves a normalized log measure, defined by
U = InI-InI (8.27)
which has been used as a texture measure [7,11]. When (8.26) is inverted it
provides a close approximation to the ML solution for the order parameter for
K distributions with large numbers of looks, and is exact for a gamma distribu-
tion. However, the normalized log in (8.27) provides a poorer approximation to
an ML estimate for v for single-look images. Nevertheless, we shall demonstrate
that adopting this measure yields little performance degradation. It corresponds
to approximating a K distribution by a gamma distribution, as proposed by
Raghavan [16], and has been applied widely in texture analysis [7,12,17—19].
8.3.2 Uncertainty in Order Parameter Estimates
In the previous section we derived ML texture measures V1, VA, V1, and U
corresponding to particular PDFs. In this section we apply the same texture
measures to characterize data that we assume to be K-distributed. Obviously, we
would expect a texture measure arising from a PDF that was close to a K
distribution to yield the greatest statistical accuracy. For example, we would
expect Vp defined in (8.15), to be a poor measure whereas U, defined in (8.27),
and VD defined in (8.20), should represent considerable improvements.
From this point on we will confine our attention to single-look SAR since
it yields maximum resolution and retains all the texture information available
to the sensor. The expected values of the texture estimators can be related to the
order parameter, in the limit of large TV, by evaluating appropriate moments of
the K distribution. Thus,
( ^ ) = TTTT" 1 = 1 + - (8-28)
4vT2
(V)=VL-J= <y) -i (829)
A
() orPfv + IJ
^k = Xf 1 + lYl + «) (8.32)
v N\ vA vJ
v
4NvT2{v)(l + 2i/i|»<°>(v) - tym(v + -i JH
X -i*pL-4-A (8-33)
- ^ = - 6- (8.34)
(2)
v Nvty {v)
l+
Wu - v (835)
v 2N(vtyM(v) - l)
for the relative bias in order parameter from V1, VA, VL, and U, respectively. The
corresponding expressions for relative variance are
^ ^ = £ r (8.39)
v2 N ( I - 1 ^ ) (v))
The uncertainty in the estimated order parameter is greater than the bias,
since the latter is of order UN whereas its standard deviation is of order
1/\lN. It is therefore possible to ignore the effects of bias for simplicity,
particularly for large N. In order to compare the performance of different
measures, a plot of their predicted relative standard deviation is shown in Figure
8.3. These predictions are compared with numerical results for the exact K
distribution. Some simulated results are also included.
The first conclusion that can be drawn is that there is good agreement
between theory and simulation, except for V1 with small values of order parame-
ter where the discrepancy can be partially resolved by introducing the second-
order term [13], with still higher orders needed for v ^ 0.4.
Let us next compare the performance of the different measures. The
numerical ML calculation for the exact K distribution [12], denoted by the full
curve, effectively represents the envelope of the optimum performance of the
approximate analytic measures. It can be seen that Uis best for v < 3, whereas
VA is best for 3 < v < 10. Hence, some combination of Uand VA should yield
a closer approximation to the true ML solution over the range 0.1 < v < 10
[20]. Equation (8.24) suggests that the MLE for a K distribution should
combine both U and V1 (rather than VJ). Joughin et al. [21] used numerical
simulation to show that the variance of the estimate of v is exponentially related
to its mean value. In the limit of large v, the minimum variance for unbiased
SD/mean
Order parameter
estimators is approximately given by the Cramer—Rao bound that has the value
v2/N [13]. The errors in all the measures considered here have this form, with
constants of proportionality taking the value 1.0, 1.63, 11.9, and 2.58 for Vp
F 4 , VL, and U, respectively. Thus, V1 is asymptotically the optimum estimator
for large v where it attains the Cramer—Rao bound. However, it is the worst
estimator in the opposite limit where U appears best.
For large values of v the texture fluctuations are very weak and the
intensity distribution tends to the negative exponential speckle PDF for single-
look imagery. Strong texture is of more interest, corresponding to smaller values
of v, where U or V1 provide reasonable approximations to the true MLE.
Subsequent discussion on texture classification and edge detection will generally
be based on these texture measures.
(a) (b)
(C) (d)
Figure 8.4 Texture analysis of SAREX image from Figure 8.1: (a) original image, (b) normalized
log texture measure output for 8 by 8 pixel window, (c) annealed version of (b) to
reduce speckle, and (d) overlay of derived edges on original image.
8.4(b) [18,19]. Note that in this instance the measure was not inverted to
estimate the order parameter. This texture image is corrupted by the effects of
speckle, which can be reduced considerably using the simulated annealing filter,
described in Chapter 6. After suitable resampling to overcome correlations
introduced by the scanning window, the despeckled texture image is shown in
Figure 8.4(c). Regions corresponding to clearings in the original image appear
dark, with primary forest generally lighter. This indicates that the texture meas-
ure could be applied to discriminate between the different classes in the scene.
Indeed, the distributions of texture measures for primary forest and clearings are
so well separated in this example that it is possible to apply a simple threshold
to the texture measure to distinguish between the two classes. The edges ob-
tained in this manner are overlaid on the original image in Figure 8.4(d). It is
apparent that this provides successful delineation of the region boundaries.
Table 8.1
Comparison of the Average Probability of Correct
Classification for the Three Textures Using the V1, VA,
Vb and U Measures
Texture Measure
V1 VA VL U
Table 8.2
Comparison of the Average Probability of Correct Classification
of the 12 Correlated Textures Using K-S or Likelihood Tests
Based on the Exact K and Approximate Gamma and Log normal PDFs
Likelihood Test
Xf=^[vfhv,-vfW,-lnr(vf)
+ (Vl)K" 0 *] (8 40)
-
where / and In / represent the average intensity and average log of intensity
within the pixels assigned to class q and v is the MLE for the order parameter
derived from (8.26). This expression for the texture likelihood can be incorpo-
rated into an annealing scheme comprised of the following stages:
1. Initialize by tessellating the image into square blocks and assign each
block randomly to some class q out of m possible classes.
2. Calculate / , In / , and v over the pixels belonging to each class and
m
evaluate the total configuration likelihood from X= ^ X .
The penalty for edge curvature is again included to override the effects of
speckle, as described in Section 7.3.5.
The result of classifying the rain forest image in Figure 8.4(a) into five
classes with initial segments of 7 by 7 pixels and a curvature penalty parameter
of 0.01 is illustrated in Figure 8.5(a). Applying a threshold to separate into
primary and secondary forest/clearing classes leads to the boundaries shown as
an overlay in Figure 8.5(b).This demonstrates the potential of combining an-
nealing techniques with ML estimates for texture.
(a) (b)
Figure 8.5 Annealed classification of the Tapajos rain forest image from Figure 8.4(a):
(a) five-class classification and (b) overlay when thresholded into two classes.
The texture estimation and classification examples in Sections 8.3.3 and 8.4.3
suggest that the position of edges within texture can be estimated accurately.
Here we analyze the limiting performance for texture edge detection, and hence
segmentation [23,24], following the approach for RCS segmentation described
in Section 7.10.3.
Edge detection, based on both the noncommittal K-S test and the specific
likelihood test, is discussed in Section 8.5.1. The latter adopts the analytic
gamma PDF approximation for the K distribution. This ML edge detector is
then incorporated into model-based segmentation in Section 8.5.2 and an-
nealed segmentation in Section 8.5.3.
Let us again define the probability that the edge is detected in the correct
position by Pd and that the overall detection probability within the window is
Ptot. Unfortunately, the threshold required for a specific value of false alarm
probability for the FWSE configuration is a function of position, mean, order,
and Pfa [23,24]. No analytic dependence is available, so the threshold has to be
derived from simulation. Once correct values are available, a comparison of K-S
and ML edge detection can be made for both FWSE and SWCE configurations.
This reveals that the ML edge detector yields higher detection probability than
that based on the K-S test, demonstrating that the introduction of a more
specific model leads to better performance. As in RCS segmentation, we find
that the FWSE configuration yields higher values of Pd than SWCE, whereas
SWCE gives larger values of P tot than FWSE.
The theoretical expression for ML edge detection can be incorporated into
segmentation algorithms, such as RGW or MUM, as discussed in Section 8.5.2.
Alternatively, it can be introduced into a global optimization procedure (anneal-
ing) in the form of a cost function, as described in Section 8.5.3, in a similar
manner to that described for intensity segmentation in Chapter 7. In each case
the theoretical framework is the same. However, the first types of segmentation
rely on a model-based approach testing for edges while the annealing approach
considers all possible region states and should be inherently more effective.
8.5.2 Model-Based Segmentation
In this section we examine the effect of modifying the merge criterion in the
MUM algorithm, described in Chapter 7, to apply to texture following the
analysis of Section 8.5.1. Figure 8.6 (a) illustrates the estimated order parameter in
each segment resulting from applying ML order parameter and mean segmenta-
tion simultaneously, based on the likelihood difference measure in (8.41). The
distributions of order parameter for clearing and primary forest regions are suffi-
ciently well separated for a single threshold to discriminate between them, leading
to the regions identified by the overlay in Figure 8.6(b). The method clearly
provides accurately defined edges with a sensitivity of about 2 pixels. There seem
to be more false alarms in the primary forest on the right side of the scene than
were found in Figure 8.4(d). However, this is a real property of the data, not a
failure of the algorithm. It indicates that the data statistics do not map exactly into
the two classes. Theory suggests [22—24] that the potential resolution of texture
segmentation is on a finer scale than the 2 by 2 limit imposed by the initial MUM
segmentation. Therefore, we require a scheme in which a fixed window with a
sliding edge is adopted to refine the edge position once maximum total detection
has been achieved with a scanning window [23,25]. This appears to be compatible
with the two-stage form of RGW suggested in Section 7.3.4.
+ (V. - l ) m 7 ~ - v ; ) (8.42)
The method then follows the procedure described for RCS annealed
segmentation in Section 7.3.5 except that two variables need to be estimated.
The curvature penalty term, introduced in Chapter 7, is again incorporated into
the annealing algorithm. The result of applying annealed texture segmentation
to Figure 8.4(a) is illustrated in Figure 8.6(c). As previously, we fix the number
of regions to be produced by the segmentation. An average region size of 2,000
pixels and a curvature penalty parameter of 0.4 were adopted. When the order
(a) (b)
(C) (d)
Figure 8.6 Joint mean and order parameter texture segmentation of Tapajos rain forest image
from Figure 8.4(a): (a) and (b) MUM and (c) and (d) SGAN. Note that (a) and (c)
display the order parameter in each segment and that (b) and (d) show an overlaid
edge map on the original image.
parameter values in the different segments were thresholded into two classes, the
overlay in Figure 8.6(d) was obtained. These results are similar to those in Figure
8.6(b) while avoiding the 2-pixel quantization of the MUM implementation.
These results can only be regarded as preliminary, though they are obviously
promising. Further research is required into the derivation of false alarm prob-
ability for texture edge detection.
8.6 Discussion
In this chapter a theoretical framework has been developed that enables approxi-
mate ML solutions to be obtained for exploiting texture. It is important to con-
sider the kinds of applications and data types for which texture measures would be
relevant. In exploiting RCS information, the value of RCS has direct physical sig-
nificance. However, the connection between the value of a texture measure and a
physical property of the scene is ambigous. Clearly texture describes the depth of
fluctuation, which depends on both physical RCS and spatial relationships be-
tween objects in the scene. Thus, discriminating between regions of different tex-
ture allows us to distinguish between different scene categories. The difference in
contrast between fields, woods, and built-up areas has already been mentioned as
an example. A texture measure can be used either to classify into regions of differ-
ent textures, which either depend on a prior reference data base or may be learned
from the data, or in a segmentation process that determines uniform regions of
significantly different texture properties. The former approach identifies a texture
sample with the nearest member of a reference set, while the latter provides data-
driven separation into an unknown number of regions. Both may become part of
a higher level classification process that attempts to label the texture (e.g., as
woodland or urban areas) depending on the value of the texture measure.
A flow diagram for texture exploitation is shown in Figure 8.7. The initial
question is whether information is contained in texture or in intensity values. If
the latter, the data should be processed following the RCS exploitation tech-
niques described in Chapters 6 and 7. If the former, we require prior knowledge
of the PDF of the data. This PDF in turn defines MLEs that encapsulate the
information about that texture. We have already demonstrated in Sections 8.4
and 8.5 that either a gamma or log normal distribution provides near-optimum
discrimination. The second issue to resolve is whether classification or segmen-
tation is the most suitable way to proceed in the given application. Classification
has the advantage that data are assigned to the nearest of a fixed number of
classes. Segmentation, on the other hand, is totally data-driven, so its statistical
uncertainty is greater. However, it will not be deceived if an unexpected sample,
inconsistent with the reference set, is encountered.
If the classification route is selected, the log likelihood for each reference
is calculated as in (8.24) (for predefined references) or (8.40) (for data-driven
processing) and the largest value taken to denote the class. Section 8.4.3 dem-
onstrated the power of simulated annealing in data-driven classification.
In comparing segmentation algorithms, issues of image quality and execu-
tion time have to be addressed, as in Chapter 7, with basically the same
algorithm options. We already established that annealing (SGAN) approximates
a global optimum segmentation if the initial conditions are set correctly. At
Intensity No Information
exploitation in texture?
Yes
Clutter ML estimator
PDF
Classify or
segment?
segment
classify
Speed or
quality?
speed quality
Figure 8.7 Flow diagram for selection of algorithms for texture exploitation.
present MUM is the only model-based segmentation algorithm that has been
applied to texture, though we indicated in Section 8.5.2 that a modified version
of RGW should do better. However, both model-based techniques adopt
heuristics that render them suboptimum, so SGAN is potentially the best.
Further development of SGAN is required for both classification and segmen-
tation. The form of the curvature penalty term needs further investigation as
does the "best" choice of average region size. Quality comparison should be used
to verify that SGAN is operating correctly and to quantify the extent to which
MUM (or RGW) is less effective. Unfortunately, image-quality assessment is
more complicated than for RCS segmentation because it is not possible to define
an equivalent simple measure to the ratio adopted there. Probably the best
approach is via a test suite using patterns of simulated textures. The user can
then impose his own cost function for the different types of segmentation error.
Comparative execution times for MUM and SGAN segmentation are
listed in Table 8.3. The MUM results are for an initial 3 by 3 tessellation of the
image; SGAN was set to provide 500 regions. Both methods have execution
Table 8.3
Comparison of Execution Times for MUM and SGAN Texture Segmentation
on Test Image as Function of Size
References
[1] Shimabukuro, Y. E., F. Ahern, and P. F. Hernandez, "Initial Evaluation of Airborne and
ERS-I SAR Data for the Tapajos National Forest, Brazil," Proc. SAREX-92, ESA Vol.
WPP-76, 1993, pp. 87-94.
[2] Quegan, S., and K. D. Grover, "Change Detection and Backscatter Modelling Applied to
Forest Modelling by SAR," Proc. European Symp. on Satellite Remote Sensing II, Paris,
September, 1995, pp. 241-251.
[3] Grover, K. D., and S. Quegan, "Image Quality, Statistical and Textural Properties of
SAREX Data from the Tapajos Test Site," Proc. SAREX-92 Workshop, ESA Vol. WPP-76,
1994, pp. 15-23.
[4] Stephens, M. A., "Use of the Kolmogorov-Smirnov, Cramer-von Mises and Related
Statistics Without Extensive Tables, " / Royal Statistical Soc. B, Vol. 32, 1970, pp. 115-122.
[5] Press, W. H., S. A. Teukolsky, W. T. Vetterling, and B. R Flannery, Numerical Recipes in C,
Cambridge, UK: Cambridge University Press, 1994, Chap. 14.
[6] Oliver, C. J., "The Interpretation and Simulation of Clutter Textures in Coherent Images,"
Inv. Problems, Vol. 2, pp. 481-518.
[7] Lombardo, P., and C. J. Oliver, "Simultaneous Segmentation of Texture Properties of
K-Distributed SAR Images," Europto Conf. on SAR Data Processing for Remote Sensing,
Rome, SPIEProc, Vol. 2316, 1994, pp. 104-114.
[8] Abramowitz, M., and I. A. Stegun, Handbook of Mathematical Functions, New York: Dover,
1970, Chap. 6.
[9] Luttrell, S. P, "An Adaptive Bayesian Network for Low-Level Image Processing," Proc. 3rd
Int. Conf on Artificial Neural Networks, London: IEE, pp. 313-316.
[10] Blacknell, D., "Texture Anomaly Detection in Radar Imagery," Europto Conf on SAR Data
Processing for Remote Sensing, Rome, SPIE Proc, Vol. 2316, 1994, pp. 125-136.
[11] Oliver, C. J., "Optimum Texture Estimators for SAR Clutter," / Phys. D: Appl. Phys, Vol.
26, 1993, pp. 1824-1835.
[12] Blacknell, D., "A Comparison of Parameter Estimators for the K Distribution," IEE Proc.
Radar Sonar Navig, Vol. 141, 1994, pp. 45-52.
[13] Lombardo, P, and C. J. Oliver, "Estimation of Texture Parameters in K-Distributed
Clutter," IEE Proc. Radar Sonar Navig, Vol. 141, 1994, pp. 196-204.
[14] Kreithen, D. E., S. M. Crooks, W. W. Irving, and S. D. Halversen, "Estimation and
Detection Using the Product Model," MIT-Lincoln Lab. Report No. STD-37, 1991.
[15] Harter, H. L., and R. S. Moore, "Maximum Likelihood Estimation of the Parameters of
Gamma and Weibull Populations from Complete and from Censored Samples," Tech-
nometrics, Vol. 7, 1965, pp. 639-643.
[16] Raghavan, R. S., "A Method for Estimating Parameters of K-Distributed Clutter," IEEE
Trans., Vol. AES-27, 1991, pp. 238-246.
[17] Lombardo, P., C. J. Oliver, and R. J. A. Tough, "Effect of Noise on Order Parameter
Estimation for K-Distributed Clutter," IEE Proc. Radar Sonar Navig, Vol. 142, 1995,
pp. 533-540.
[18] Oliver, C. J., A. P. Blake, and R. G. White, "Optimum Texture Analysis of SAR Images,"
SPIE Conf. Algorithms for Synthetic Aperture Radar Imagery, Orlando, FL, SPIE Proc, Vol.
2230, 1994, pp. 389-398.
[19] Oliver, C. J., "Edge Detection in SAR Segmentation," Europto Conf on SAR Data Processing
for Remote Sensing, Rome, SPIE Proc, Vol. 2316, 1994, pp. 80-91.
[20] Jahangir, M., D. Blacknell, and R. G. White, "Accurate Approximation to the Optimum
Parameter Estimate for K-Distributed Clutter," IEE Proc. Radar Sonar Navig, Vol. 143,
1996, pp. 383-390.
[21] Joughin, I. R., D. B. Percival, and D. P. Winebrenner, "Maximum Likelihood Estimation
of K Distribution Parameters for SAR Data," IEEE Trans. Geo. Remote Sens., Vol. 31, 1993,
pp. 989-999.
[22] Oliver, C. J., "Optimum Classification of Correlated SAR Textures," Europto Conf. on SAR
Image Analysis, Simulation and Modelling II, Taormina, Italy, SPIE Proc, Vol. 2958, 1996,
pp. 64-73.
[23] Oliver, C. J., I. McConnell, and D. Stewart, "Optimum Texture Segmentation of SAR
Clutter," Proc. EUSAR 96, Konigswinter, 1996, pp. 81-84.
[24] Oliver, C. J., and P. Lombardo, "Simultaneous Mean and Texture Edge Detection in SAR
Clutter," IEE Proc. Radar Sonar Navig, Vol. 143, 1996, pp. 391-399.
[25] Oliver, C. J., D. Blacknell, I. McConnell, and R. G. White, "Optimum Edge Detection in
SAR," Europto Conf. on SAR Image Analysis, Simulation and Modelling, Paris, SPIE Proc,
Vol.2584, 1995, pp. 152-163.
[26] Caves, R. C , and S. Quegan, "Matching Segmentation Algorithms to ERS-I SAR Appli-
cations," Europto Conf. on SAR Data Processing for Remote Sensing, Rome, SPIE Proc, Vol.
2316, 1994, pp. 148-158.
[27] Cook, R., I. McConnell, D. Stewart, and C. J. Oliver, "Segmentation and Simulated
Annealing," Europto Conf. on SAR Image Analysis, Simulation and Modelling II, Taormina,
Italy, SPIE Proc, Vol. 2958, 1996, pp. 30-37.
9
Correlated Textures
9.1 Introduction
where Xand Kare the lags in each dimension, v the order parameter and € the
correlation length. In Figure 9 . 1 v takes values of 0.5, 1.0, and 2.0 while €
corresponds to 1, 2, 4, and 8 pixels. These textures can be simulated exactly since
they have half-integer order parameters and thus obey the random walk model of
Section 5.8. For this simulation the imaging process did not introduce any further
correlations. Real data would have to be resampled to match this behavior by
removing PSF correlations, as described in Chapter 4.
Figure 9.1 Montage of 128 by 128 pixel regions of simulated correlated textures, shown as
amplitude images. The textures, characterized by (v, €) are labeled from 1 to 12 and
have positions in the montage denoted by:
1(0.5,1) 2(0.5,2) 3(0.5,4) 4(0.5,8)
5(1.0,1) 6(1.0,2) 7(1.0,4) 8(1.0,8)
9(2.0,1) 10(2.0,2) 11(2.0,4) 12(2.0,8)
N (/(0,0)/(X, Y)) f ^
T1[X9Y) ^ ^ ^ >± = 8 X ) 0 o r ^l + I J + r^X.Y) (9.2)
Figure 9.2 Random selection of 32 by 32 pixel regions of the different correlated textures. The
different textures are laid out in positions given by:
5 8 6 12 3 8 10 9 7 4 6
6 11 11 1 3 9 3 5 10 9 6
7 2 6 1 4 11 9 7 12 8 11
12 4 2 10 5 9 4 5 8 1 2
for the process described by (9.1), where 8 X 0 is the Kronecker delta. As shown in
Section 5.6, the contribution at zero lag includes coherent interference terms aris-
ing from individual random scatterers and RCS fluctuations. The RCS correla-
tion properties are contained in the nonzero lag ACF coefficients, ra (X, Y), which
include a constant background of 1. In order to derive information about these we
have to remove the peak at the origin, which is caused by speckle alone.
The statistical uncertainty in the ACF coefficients in (9.2) determines the
error with which order and correlation length are estimated. We would not
expect a least-squares fit to provide MLEs unless the errors had a Gaussian PDF
and were uncorrelated from lag to lag. Early work was confined tofittingthe
intensity ACF [1,4]. However, in Chapter 8 we demonstrated that determining
the order parameter by fitting intensity contrast was much poorer than fitting
either normalized log, variance of log, or amplitude contrast [2]. Similarly, we
would expect that correlation length determination would be improved by
fitting ACF measures that have inherently smaller statistical errors. In Chapter
8 we demonstrated that the normalized log measure approximates the MLE for
order parameter, but there is no corresponding definition of ACF. Therefore, we
adopt the ACF of the log of intensity (which has the variance of the log as its
value at the origin) and the ACF of amplitude to compare with the intensity
ACF. Unfortunately, theoretical derivations of the errors in these ACF coeffi-
cients and the resultant correlation lengths are more cumbersome than those for
the intensity ACF. The amplitude ACF is given by [2]
where
^(i,r) = (i-v№,F)-i)f+1
X 2Fx[v + \,v+\-,V^r0[X,Y)-X)] (9.4)
and 2F\(•) is the Gaussian hypergeometric function (Section 15.1 in [5]). The
log ACF is [2]
r, r(X,Y) = l + bxn8r0 — 5-
6(4)(0)(v) - lnv + ln|JL - 7 E )
W0Hv) - l n v + lnfx)2 , , . ,
\ l ( ! , , ,, ~ T ( U ^ ) - 1 ) (9-5)
(i|i ( 0 ) (p) - I n v + InJUL - 7 E )
where
Lag, k
Figure 9.3 Comparison of a log plot of the predicted intensity (full), amplitude (short dashes),
and log (long dashes) ACFs, r(k) - 1, for texture 4, in which v = 1.0 and € = 4 pixels.
Initially let us consider the determination of the order parameter from
single-point statistics. Expressions for the errors in v for uncorrelated textures
were given in (8.36) to (8.38). Figure 8.3 showed that simulation and prediction
were in good agreement [3,4,6-8]. No general theoretical form for correlated
processes is available, so the behavior has to be deduced from simulation. The
fractional errors in order parameter, as a function of correlation length, are
shown in Table 9.1 [9]. The results for textures with € = 0 can be compared
with the uncorrelated theory from Chapter 8 and show reasonable agreement.
The simulated results for small € show slightly increased errors, with the log
ACF providing the best estimate of order. As correlation lengths increase the
errors increase, showing the effect of correlation between pixels reducing the
number of independent estimates within the sample. The fractional error in v
is then approximately the same for all three ACF estimates. The amplitude ACF
could well be a more appropriate estimate than the log ACF over the typical
range of values encountered with real SAR clutter. It consistently offers better
performance than the ACF of intensity, without the sensitivity to correlation
length and order shown by the log ACF.
Table 9.1
Dependence of Fractional Error in Order Parameter on Correlation Length for Sample Size
of 32 by 32 Pixels
Simulation
Note: Results for ACFs of intensity, amplitude, and log are shown. Simulation was performed
over 103 samples. Predictions for uncorrelated data are included for comparison. The pre-
dictions in brackets refer to the second-order correction for the intensity ACF. Uncertainty
±0.01.
A rigorous derivation of the error in estimated correlation length is ex-
tremely complicated. Previous approximate theory for fitting the intensity ACF
[l,3>4,10] makes assumptions that lead to predictions that are qualitatively
reasonable. Rigorous derivations have been presented for a negative exponential
ACF only; even these require numerical evaluation [2]. Again we can only
determine behavior by simulation, as summarized in Table 9.2 [9]. It is apparent
that the fractional error in correlation length does not show much dependence
on v, €, or the choice of ACF measure. There is evidence that adopting the
amplitude or log ACF yields a slight advantage for small €, though none for
larger ones. Over all v and € considered, the amplitude ACF seems to offer the
smallest error in determining i.
In general, one might conclude that the amplitude ACF provides the most
robust estimator on which to base order parameter and correlation length
fitting. The simulated errors can be regarded as guide to the accuracy with which
any specific value of v and € could be determined for different forms of ACF.
Indeed, in many instances fitting a Gaussian profile to the ACF might be
regarded as a means of determining some characteristic correlation length.
Certainly, where only small windows are used to calculate the ACF and deter-
mine €, errors in ACF coefficients are likely to dominate over the discrepancy
between the true shape and a Gaussian of similar width.
It should be noted that the drawback to fitting either amplitude or log
ACFs for determining correlation length is that the theoretical form of ACF has
to be re-evaluated continually during the fitting process. While this is simple for
Table 9.2
Measure v €= 2 €= 4 €=8
Table 9.3
The Average Probability of Correct Classification of the 12 Test
Textures Using Intensity and Amplitude Contrast and Variance
of Log Measures
(97)
1 2vtu) 2vi{j) J
where |Xr(/) and JUL1C/) are the mean real and imaginary values for the jth
component and Vr(j) and V1(J) are the corresponding variances. The PDFs of
the spectral coefficients derived from simulation suggest that the means are
P(L0)
(a)
Figure 9.4 PDF of likelihood difference estimate between classes 7 and 11 for simulated
texture 7: (a) unnormalized and (b) normalized. Intensity (dashed), amplitude (dot-
ted), and log (full) PDFs compared.
approximately zero and the variance of each component is the same. However,
consideration of the log likelihoods shows that approximating the mean by zero
degrades classification performance appreciably. Instead, it is preferable to as-
sume that the means and variances are identical, in which case (9.7) can be
modified to
where U| L = juur = JUL1 and V = Vx = V1At should be noted that the spectrum has
special cases for the components at (0, 0), (0, nyl2), (nxl2, 0), and (nx/2, nyl2),
where the window dimensions are nx X ny. The joint PDF of the spectrum over
the whole window for texture from class p is then given by
P(L n )
LD
(b)
Pp(at(l),a.(\),...,at(M), O1(Af))
- iA i J (*rU)-*fU)f+MJ)-»,U)f} (99)
-^\hvPU)p[ 2vp(j) J m
where Af = nx X ^ , and the log likelihood of class /> by
Table 9.4
Comparison of Average Probability of
Correct Classification for the 12
Correlated Textures Based on ACF Fit
and Spectrum Classification
0.814 0.946
9.4 ML Correlated Texture Edge Detection
where S(j) = a^[j) + a? (j) and S (j) = 2V (y) are the observed and refer-
ence power spectral densities, respectively. Note that the special cases identified
in the previous section have to be treated separately.
The treatment of edge detection for uncorrelated textures in Section 8.6
can now be extended to include correlation by incorporating the texture spec-
trum into the edge-detection scheme in place of single-point statistics. We adopt
the same window geometry with k pixels in region 1 and M — km region 2.
The log likelihood for each region is obtained from (9.11), and the joint
likelihood that the region is split at position k is then given by [9,12]
where the/ch components in region 1, S1(J) and S (j), denote the observed
power spectrum and reference for class px in region 1 comprising pixels 1 to k,
for example. This form relies on prior definition of the reference spectra. These
are unknown in segmentation, so the spectrum has to be estimated from the
data themselves. In this case the ML values of the spectral components are
identical to their observed values, so S (j) is replaced by S1(^'). Hence, the log
likelihood for a split at position k becomes
This can be compared with the log likelihood that the merged region is
consistent with a single overall spectrum estimate SQ(j), denoted by the sub-
script 0. The log likelihood Xmerge is then derived from (9.13) by summing over
the complete region. The region is split if the log likelihood difference,
M k M
1 1
M*) = V ^ " V * - X ^oO) " X AC/) - XlnS2(/) (9.14)
y=i j=i j=k+i
The first two conclusions are the same as previous results for uncorrelated
texture (in Chapter 8) and differences in mean only (in Chapter 7). Since both
PDF-based and spectrum-based edge detection are derived from MLEs, it is not
immediately clear why the latter is less effective. However, it should be appreci-
ated that the spectral method makes no assumption about the relationship
between different spectral coefficients; each is separately determined from the
data. The PDF technique, on the other hand, assumes that the data are com-
pletely described by a single texture parameter, the order, which is determined
from the data. This representation makes use of more specific prior knowledge
about the texture and so would be expected to yield greater accuracy.
We next comment on the dependence of Pd and P101 on order parameter
for approximately uncorrelated textures. Again, we compare both spectral and
PDF-based performance for both FWSE and SWCE configurations. The results
for Pd lead to the following observations:
9.5 Discussion
References
[1] Oliver, C. J., "Review Article—Information from SAR Images,"/ Pkys- DiAppl. Phys., Vol.
24, 1991, pp. 1493-1514.
[2] Lombardo, P., and C. J. Oliver, "Estimating the Correlation Properties of K-Distributed
SAR Clutter," IEE Proc. Radar Sonar Navig., Vol. 142, 1995, pp. 167-178.
[3] Oliver, C. J., "Clutter Classification Based on a Correlated Noise Model," Inv. Problems,
Vol. 6, 1990, pp. 77-89.
[4] Oliver, C. J., "Parameter Estimation with Correlated Textures," Inv. Problems, Vol. 5, 1989,
pp. 903-914.
[5] Abramowitz, M., and I. A. Stegun, Handbook of Mathematical Functions, New York: Dover,
1970.
[6] Oliver, C. J., "Optimum Texture Estimators for SAR Clutter," /. Ploys. D: Appl. Phys., Vol.
26, 1993, pp. 1824-1835.
[7] Lombardo, P, and C. J. Oliver, "Estimation of Texture Parameters in K-Distributed
Clutter," IEEProc. Radar Sonar Navig., Vol. 141, 1994, pp. 196-204.
[8] Blacknell, D., "A Comparison of Parameter Estimators for the K Distribution," IEE Proc.
Radar Sonar Navig, Vol. 141, 1994, pp. 45-52.
[9] Oliver, C. J., "Optimum Classification of Correlated SAR Textures," Europto Conf. on SAR
Image Analysis, Simulation and Modelling, Taormina, Sicily, SPIE Proc, Vol. 2958, 1996,
pp. 64-73.
[10] Oliver, C. J., "The Sensitivity of Texture Measures for Correlated Radar Clutter," Inv.
Problems, Vol. 5, 1989, pp. 875-901.
[11] Blacknell, D., "Texture Anomaly Detection in Radar Imagery," Europto Conf. on SAR Data
Processing for Remote Sensing, Rome, SPIE Proc, Vol. 2316, 1994, pp. 125-136.
[12] Oliver, C. J., "Correlated Texture Information in SAR," DRA Tech. Memorandum
DRA/LS2/TR96020, 1997.
[13] Oliver, C. J., D. Blacknell, and R. G. White, "Optimum Edge Detection in SAR," IEE
Proc Radar Sonar Navig, Vol. 143, 1996, pp. 31-40.
[14] Oliver, C. J., I. McConnell, D. Blacknell, and R. G. White, "Optimum Edge Detection in
SAR," Europto Conf on SAR Image Analysis, Simulation and Modelling, Paris, SPIE Proc,
Vol.2584, 1995, pp. 152-163.
[15] Oliver, C. J., I. McConnell, and D. Stewart, "Optimum Texture Segmentation of SAR
Clutter," Proc EUSAR 96, Konigswinter, 1996, pp. 81-84.
[16] Oliver, C. J., and P Lombardo, "Simultaneous Mean and Texture Edge Detection in SAR
Clutter," IEEProc Radar Sonar Navig, Vol. 143, 1996, pp. 391-399.
1 0
Target Information
10.1 Introduction
(C) (d)
Figure 10.1 Simulated images of a tanklike object at different resolutions showing the effects
of introducing a weak background with an SCR of 20 dB: (a) 3-m resolution, no
background; (b) 0.75-m resolution, no background; (c) 3-m resolution with back-
ground; and (d) 0.75-m resolution with background.
we are involved with characterizing the target based on the structure of the
return. Both these processes are degraded by the speckle background. Adding
the same background level to the high-resolution image in Figure 10.1(b) leaves
the brighter returns largely unchanged, as shown in Figure 10.1(d), but the
sidelobe structure is again destroyed.
In Section 10.2 we introduce a Bayesian approach to target detection and
show the relationships between different detection criteria. Since targets occupy
only a very small fraction of the pixels within the image, the simultaneous
minimization of false detections and maximization of genuine detections is
extremely important. In Sections 10.3 and 10.4 we discuss how the nature of
target and background properties affects the detection process and the influence
of uncertainties introduced by estimating target and background parameters.
Section 10.5 then addresses further refinements associated with problems en-
countered in constantfalse alarm rate (CFAR) detection. Once targets have been
detected, target recognition is introduced in Section 10.6. The issues of dis-
crimination and classification are addressed in Sections 10.7 and 10.8. In
Section 10.9 we outline an approach to target super-resolution that allows
targets to be reconstructed with increased spatial resolution by introducing prior
knowledge into the imaging process. This can provide a useful processing stage
before classification is attempted.
Let us initially explore a Bayesian approach to target detection [I]. The problem
is one of testing two hypotheses for the data vector x, namely, that a target is
present or not. Detection performance can then be characterized in terms of the
probability of false alarm, Pfa, that the data x will be interpreted as denoting the
presence of a target when one is not present, and the probability of detection,
Pd, that a target will be detected when one is indeed present. The Bayes criterion
for detecting a target states that
where P(TIx), the a posteriori PDF, describes the probability of a target given
the data; P(x\ T), the likelihood function, describes the probability of data x
when a target is present; P(T) is the a priori probability that a target is present;
and P(x) is the probability of obtaining that data. Similarly for the background
hypothesis,
The MAP criterion implies that a target should be considered present when
P(B\x)
from which
P(x|5)
« > , 00.4)
The threshold t is selected to give a defined false alarm probability. The detec-
tion conditions in (10.3) and (10.4) now depend on knowing the correct target
and background likelihood functions.
In practice, only the background PDF is characterized. Under these con-
ditions we are forced to adopt anomaly detection, a suboptimum approach. The
data are assigned to the background if they appear consistent with the known
background distribution, so
OC
Ph=\P{x\B)dx (10.6)
t
where the threshold t can be selected to yield a certain CFAR. The correspond-
ing detection probability for a target is
00
Pd=jP(x\T)dx (10.7)
t
surrounded by a guard ring a few pixels wide to prevent any leakage from the
target into the boundary ring of M pixels, which is used to estimate the average
background level IB, as well as the normalized background variance VB. A
conventional two-parameter CFAR [4] applied to image intensity is defined by
the criterion that "detection" occurs when
1 1 1
^ IT >t (10.8)
Figure 10.2 Diagram of target detection configuration comprising an ROI for a target in the
center, a guard ring to avoid leakage from the target into the background, and a
boundary ring to estimate the background level.
In the simplest implementation of (10.8) we assume that the background
RCS is constant, so for single-look SAR, yjVB — 1, resulting in the one-parame-
ter CFAR condition
OC
Ph = — Jexp - — dl = exp - — (10.10)
^B L °"B J L °"B_
where a B is the background RCS. Hence, the required t for a given false alarm
probability is given by
* = -aBln/>& (10.11)
Pd = exp - — = e x p - i ^ (10.12)
CT R
L TJ
7
B(7J = - — F T T e x P ~ — (iai3)
r
^V°"BJ W L °"B_
and the appropriate false alarm probability is given by the incomplete gamma
function (Section 6.5 in [8])
pfa=rL-M/r» (io.i4)
This enables the threshold level t for the incoherent average / to be derived
numerically for a specific CFAR. Similarly, the detection probability for a
Swerling 1 target filling the ROI would be
Pd=TU,^-)/r(m) (10.15)
A typical result for the dependence of P d on the SCR is illustrated in Figure
10.3 for windows of 1, 4, 9, 16, and 36 pixels. P^ was set to 10~3 throughout.
It is apparent that Pd is considerably increased by incoherent averaging over
many pixels. Thus, it is advantageous to increase resolution as far as possible so
that the image of the target is spread over as many pixels in the ROI as possible,
with consequent reduction in speckle.
This analysis is, however, misleading because, in practice, both target and
background RCS are unknown and have to be estimated from the data, leading
to CFAR loss [5] • In addition, real targets would be unlikely to have uniform
RCS over the window.
SCR
Figure 10.3 Variation of Pd with SCR for negative exponential statistics; dependence on region
size m= 1,4,9,16, and 36 from right to left.
follows the same description as the segmentation theory summarized in Chapter
7, with a log likelihood difference similar to (7.5), such that
where
7 (10J7)
« =4 r ^
M +M
A target is detected when \ D exceeds some threshold t. As described in
Chapter 7, the detection probability as a function of the SCR, 7?, can now be
expressed in terms of the observed ratio r ( = / T / / B ) , with the additional con-
straint that r > 0. Hence,
t
Pd =l-jPr(r\R,m,M)
0
T{M +
=1- " } (Jg-T ^ 1 U + ^ + i; -J-Ll (10.18)
So far, both clutter and target intensity PDFs have been treated as negative
exponential. Any fluctuation in underlying RCS will influence detection that is
based on the RCS alone. An underlying gamma-distributed cross section, lead-
ing to a K-distributed intensity, has been demonstrated to provide a close fit to
Detection probability
SCR
Figure 10.4 Variation of Pd with SCR for negative exponential statistics; the effect of estima-
tion errors. Three configurations are compared with: (a) m = 1, M= 16; (b) m = 4,
M = 28; and (c) m = 16,M= 52. Results for known means are denoted by dashed
curves, those for estimated means by full curves.
( VI+VB)/2 r i—T
p /(VB 1)/2
^ = TvH — " ^B-! 2 P i (10.19)
for single-look SAR, where |xB and vB are the background mean and order,
respectively. In Section 10.4.1 we describe standard CFAR detection, based on
the RCS, for a Swerling 1 target against this type of textured clutter. If the target
ROI subtends several pixels it is possible to characterize higher order properties
of the target region, such as an order parameter, in addition to the average RCS.
In Section 10.4.2 we take advantage of this additional degree of freedom in
proposing a texturelike target model representing the interference between
scattering centers in a single target. We show how this can result in a target
detection algorithm that again uses the Neymann—Pearson criterion with ML
parameter estimates.
10.4.1 Standard CFAR Target Detection
If a target occupies a single pixel, the RCS is the only measurable quantity. For
a K-distributed background, the false alarm probability, from standard CFAR
theory on a single pixel, is given by (10.6) as
K) 2 K 2 (i 2o)
^ =w^i—Y ^\ P] °-
The appropriate value of threshold t can be derived numerically by invert-
ing (10.20) for a given P^. If the threshold is derived assuming negative
exponential rather than K-distributed clutter, as in (10.11), the modified false
alarm rate P^ is given by
P h=
' F H ( " VB I n 7 ^ V X [ 2 V - v B l n ^ ] (10.21)
Figure 10.5 illustrates the effect the K-distributed background has on the
predicted false alarm rate. As vB increases, the background PDF tends toward a
negative exponential so that the two values of false alarm rate become the same.
In the opposite limit of small vB, corresponding to a spiky background distri-
bution, Pfa increases by between one and two orders of magnitude compared
with P^. This demonstrates the crucial importance of adopting a realistic
background clutter model for target-detection theory. Ward [9] demonstrated
the consequences of ignoring background RCS variations in maritime target
detection. Theoretical studies of corrected CFAR algorithms for Weibull
[4,10-14] and K-distributed [15] clutter have been reported.
If a target subtends many pixels, we can estimate the CFAR conditions
over more than the single pixel discussed previously. We have already demon-
strated that it is advantageous to perform incoherent averaging over many pixels.
Unfortunately, there is no analytic relationship between the single-pixel K
distribution and the result of performing incoherent averaging when the cross
section itself varies between pixels. (Note that multilook imaging discussed in
Chapter 4 assumes that the cross section is constant and that only the speckle
fluctuates.) For simplicity we adopt the same gamma distribution approxima-
tion to the K distribution proposed in Chapter 8. The mean of the two PDFs
is the same and the order parameters are related by
Order
Figure 10.5 Dependence of false alarm probability for a single pixel of K-distributed back-
ground, incorrectly assumed to be negative-exponentially distributed, on clutter
order parameter. Results for nominal negative exponential false alarm prob-
abilities of 10~3 (full curve) and 10~4 (dashed curve) are shown.
where v is the order of the K distribution, V1 is that for the approximating gamma
distribution, i|/(°)(') is the digamma function, and 7 E is Eulers constant (Chapter
6 of [8]). Thus, the approximate form for the single-pixel intensity PDF is
/ \ WV., _ r 7~
The false alarm probability is then derived in the same way as (10.14), yielding
P
h =r ™>y>—\/r{™>y) (10.25)
Equation (10.25) can be inverted numerically to yield the appropriate
value of threshold for a given Pfa.
To derive the target-detection probability we retain the Swerling 1 target
model, adopted in Section 10.3, with uniform RCS over the ROI. The threshold
setting, derived by inverting (10.25), is inserted into (10.12) to yield Pd.
The dependence of Pd on region size and background order parameter is
illustrated in Figure 10.6(a,b). In each case Pfa = 10" 3 and the SCR varies from
1.0 to 5.0. In Figure 10.6(a) we show the dependence on region size for a
K-distributed background order parameter of 1.0. If these results are compared
with those in Figure 10.2 (with no background cross-section fluctuations), it is
obvious that additional fluctuations degrade performance. As region size in-
creases, both speckle and cross-section fluctuations are averaged out, improving
Pd as before. In Figure 10.6(b) we illustrate the effect of vB on Pd, for m — 16,
as the SCR varies from 1.0 to 5.0 and P^ = 10~3. This illustrates the increase
in SCR that is required to keep Pd constant as vB decreases and background
texture becomes more spiky.
SCR
(a)
Detection probability
SCR
(b)
Figure 10.6 Variation of Pd with SCR for Swerling 1 target and K-distributed background;
Pfa = 10~3. (a) vB = 1, dependence on m = 1, 4, 9, 16, and 36 pixels from right
to left, (b) m = 16, dependence on vB = , 4, 1, and 0.5 from left to right.
Detection probability
SCR
Figure 10.7 Variation of Pd with SCR for Pfa = 10~3, m = 16, and vB = 1. Dependence on vj
= 0.1,0.2,0.5, and 1 from top to bottom on the left-hand side.
_ i N _
|x - / = — ^L and Inv - i|i(°)(v) = In/ - In/ (10.26)
N j=i
These estimates are then substituted into the expression for the log likelihood
difference, similar to (8.41), leading to an ML log likelihood difference given
by [16]
where the target region is denoted by the subscript T, the background by B, and
merged regions by 0; v denotes the order parameter estimated from (10.26).
Detection occurs when XD exceeds the threshold t. Initially simulation is per-
formed with target and background regions that have the same input properties
to determine the appropriate threshold value for a given P&. Further simulations
with varying mean and order parameters can then be performed based on this
threshold to derive the performance characteristics.
Note that we no longer exploit the prior knowledge that either target or
background has a negative exponential PDF (with v = ). In each case, both the
mean and order parameter are estimatedTor each region.
Initially, in Figure 10.8(a), we vary vB between °° and 0.5, while v T is
maintained at oo} corresponding to a negative exponential PDF. Threshold
settings are determined by simulation over 105 samples, whereas results for Pd
are obtained over 104 samples. These results should be compared with equiva-
lent results for detection of RCS differences when parameter values are known
in advance, shown in Figure 10.6(b). First, it should be noted that including
the effect of texture in target detection now yields the best detection for a given
SCR when the two order parameters differ most. Second, if we consider results
for vB = oo, we see that having to estimate both mean and order has led to
considerable reduction in detection performance compared with the original
situation where these were known a priori.
In the second comparison, in Figure 10.8(b), we set vB = 1.0 but allow
v T to vary between 1.0 and 0.1. This represents the expected situation, with
target fluctuations more marked than background ones. If we compare results
for vB = v T = 1.0 in Figure 10.8(b) with those for vB = v T = °° in Figure
10.8(a), it is clear that reducing both vB and vT increases the uncertainty in the
determination of parameter values and degrades detection. However, as vT falls
below vB in Figure 10.8(b), detection improves considerably, which should be
compared with Figure 10.7 where detection was based on RCS alone. Though
reducing vT increases detection for small SCRs in Figure 10.7, the results for
Detection probability
(a)
Detection probability
SCR
(b)
Figure 10.8 Variation of Pd with SCR for K-distributed target and background with parameters
estimated over m = 16 and M = 52; Pfa = 10~3. (a) V1 = °o; curves correspond to
V =
B °°' 4 ' 1 ' ar|d 0-5 ordered from bottom to top. (b) vB = 1.0; curves correspond
to vj = 1,0.5, 0.2, and 0.1 ordered from bottom to top on the left-hand axis.
larger SCRs are degraded. In the present results, the effect of reducing vT is that
it becomes the dominant factor in detection and the performance is essentially
independent of SCR for small v T . Texture alone provides target discrimination
under these conditions. Thus, introducing the textured target model leads to a
considerable improvement in detection capability.
Note that this method provides an optimized balance between differences in
mean and order parameter as it corresponds to the Neymann—Pearson observer [2].
It uses ML estimates of parameters and overcomes the degradation where detection
is based solely on differences in the RCS, which is all that is possible if the target
merely subtends one pixel. The analysis shows that if the target occupies many
pixels it is advantageous to use both RCS and texture in the detection process. Note
that the analysis only approximates the K-distribution by a best-fit gamma distri-
bution, as described earlier, resulting in slightly suboptimum performance.
A more sophisticated model which identifies the actual positions of strong
scatterers should provide even better detection. Indeed, the goal of such an
approach would be to provide a filter matched to a target with which to scan
the scene. This is tantamount to omitting a distinct target detection stage and
proceeding directly to a recognition mode, which will be addressed later. At
present it should be noted that the textured target model achieves significant
improvement in detection if the target extends over several pixels and shows
strong fluctuations, consistent with a K distribution.
Sections 10.3 and 10.4 dealt with comparatively simple target-detection scenar-
ios. There are two areas in which this simple treatment should be extended:
• Obtaining more accurate estimates of background properties;
• Handling inhomogeneous backgrounds.
(b)
Figure 10.9 Comparison of (a) optical photograph and (b) simulated SAR image of tank.
image evolves with angle the strength and position of the dominant scattering
centers change, due to broadside flashes, obscuration, and interference effects.
However, the general image properties appear consistent with the textured target
model in Section 10.4.2. In order to provide a reference set for such a target,
about 256 separate orientations, each comprising (say) 32 by 32 elements, are
required. However, this only represents horizontal rotations; pitch and roll must
also be taken into account. If we allow for 20 different roll and pitch angles,
Figure 10.10 Variation of simulated SAR image of model tank with orientation over 360 degrees.
then the total data base for a single target class would comprise about 108
elements. Add to this the fact that we require reference sets for all potential
targets (say 50) and the total data base consists of 5 X 109 elements.
Next consider discriminating between the simulated tank, with the refer-
ence set shown in Figure 10.10, and an armored infantry fighting vehicle (AIFV),
with a corresponding set of simulated training data shown in Figure 10.11 [25].
For 100% classification, all the orientations of the tank image must be distinct
from any for the AIFV. Once receiver noise and background speckle are in-
cluded, it is by no means obvious from Figures 10.10 and 10.11 that this
condition can be satisfied. A reduction in resolution with an increase in back-
ground clutter would certainly degrade the ability to distinguish between them.
Figure 10.11 Variation of simulated SAR image of model AIFV with orientation over 360 degrees.
Given the scale of the computation load for real-time ATR based on such
a set of references, it is essential to minimize the number of images to which full
classification is applied.
10.7 Target D i s c r i m i n a t i o n
Image
potential targets
classified
targets
Targets
Table 10.1
List of Features Used in Discrimination Stage of Target Recognition
Size Mass
Diameter •
Ratio
Perimeter
Rotational inertia •
Texture Standard deviation •
Weighted fill ratio
Fractal dimension •
Contrast Peak CFAR
Mean CFAR •
% bright CFAR
Polarization (where applicable) % pure
% pure even •
% bright even
Time dependence Change detection
runway lights. Following target detection using a one-parameter CFAR with Pfa
= 10~3, many detections are visible in Figure 10.13(b). In Figure 10.13(c) we
illustrate the effect of applying a simple discriminant such that only vehiclelike
objects are retained [25], characterized by major and minor axis lengths and total
RCS. The first two tests are equivalent to the diameter and ratio tests in Table
10.1, while the total RCS is related to the mean CFAR statistic. The majority of
unwanted objects, such as the buildings, fences, and runway lights, fail to meet
these criteria; and we are left with a high detection rate on the targets of interest
and a very small number of residual incorrect targets.
Discriminants based on texture in Table 10.1 measure the variation of
RCS within the target region. The standard deviation describes the single-point
statistics and is generally applied to the log of the data [26], which would be
optimal if the target had a lognormal PDF, as discussed in Chapter 8. We also
showed there that the normalized log measure, from which the order parameter
could be derived, would be preferable for a K-distributed target. The weighted
fill ratio measures the fraction of total energy contained in the brightest 5% (say)
of the scatterers within the ROI, and the fractal dimension provides a measure
of the spatial dimensionality of the detected object [29]. Contrast features are
all related to values of the CFAR statistic (see (10.8)) within the ROI. Peak and
mean CFAR statistics are self-explanatory; the percentage bright measure is the
fraction of pixels whose CFAR statistic exceeds some threshold.
Polarimetric features, where available, introduce another dimension into
discrimination and are discussed in more detail in Chapters 11 and 12. Used in
the context of discrimination we are concerned with classifying returns in terms
of the number of reflections they have undergone. If the received fields Ehh, Ehv,
E^ in the different polarization states are measured, then the energy correspond-
ing to odd and even bounce mechanisms is given by [26]
(b)
(C)
Figure 10.13 Example of target detection applied to airborne X-band SAR image: (a) original
multilook SAR image, (b) detections based on assumed negative exponential
background PDF, and (c) discriminated result when a simple size constraint is
placed on targets.
lengths, particularly over a wet or flooded surface, also shows very clear signatures
of double-bounce scattering [30]. This means that it is important to apply this test
in a suitable context. Further discussion of polarization effects can be found in
later chapters.
The percentage pure measure is defined as that fraction of pixels within
the ROI for which some fraction of the energy can be attributed to either the
even- or odd-bounce category. The percentage pure even measure describes the
fraction of pixels for which the energy in the even-bounce channel exceeds some
threshold. The percentage bright even measure corresponds to the fraction of
bright pixels in the image that are mainly even-bounce scatterers.
Once individual features have been measured the results for each can be
combined using a specific criterion, such as a quadratic distance measure of the
form [26]
where x is the vector of features from the detected object; m and Care the mean
and covariance of these features, estimated from training on another set of
targets; and n is the number of features.
Upon comparing the performance of the different measures for a set of
three types of target, namely a tank, a self-propelled gun, and an armored
personnel carrier, Novak et al.[26] found that the best set of discriminants to
use were those indicated by • in the third column of Table 10.1. This includes
two size-based measures, one of which involves the distribution of RCS, which
introduces independent information. The texture features carry independent
information about the spikiness of the distribution (through the standard devia-
tion of the log) and the space-filling property (through the fractal dimension).
The total RCS is included through the mean CFAR measure. A single po-
larimetric measure is also included (the percentage pure even); other polarimet-
ric measures are correlated with this feature and do not provide independent
information. Applying these discrimination tests to fully polarimetric high-reso-
lution data, Novak et al. [26] report that the total false alarm rate was reduced
from one in 1.4 X10 5 , after the polarization whitening filter was applied, to one
in 7X10 5 . False alarms from natural clutter were reduced nearly twice as much
as those from man-made clutter. This clutter rejection was achieved with no
appreciable loss in detection probability for genuine targets.
A different type of approach to discriminating targets from false alarms is
based on their time dependence, as listed in the final row of Table 10.1. Identifying
objects within the scene that have changed since previous images were acquired
allows us to keep track of objects, such as vehicles, which are likely to be most
significant as targets. Target change detection exploits the high degree of geo-
metrical accuracy that can be achieved following the methods described in
Chapter 3, so it is possible to register individual pixels precisely in different
scenes. This enables a direct comparison of detections in each scene in order to
identify changes [31,32]. Figure 10.l4(a,b) shows a region of about 1.5-km square
on consecutive days. Figure 10.l4(c) identifies those targets that were detected in
the first image, but not the second; while Figure 10.l4(d) shows the opposite. The
probability of detecting genuine changes, primarily in the two formations of
vehicles visible in Figure 10.l4(c), is high, while the false alarm probability else-
where in the scene is very low.
Once the detection and discrimination stages have rejected as much clutter as
possible, the final stage of the ATR scheme consists of target classification.
Potentially this could be based on pattern-matching techniques using all the
information in the data. A theoretical framework can be devised similarly to that
used for homogeneous correlated textures in Chapter 9. Section 9.3.2 showed
that the required textural information is contained in the complex spectrum.
For deterministic targets the assumption of homogeneity is no longer applicable
and it is necessary to estimate the likelihood for each relative shift d of reference
and data. Taking the logarithm of the random texture results in the distributions
of the estimated complex spectral coefficients being approximately Gaussian
[33,34]. Even if this assumption is not correct for targets, it should be possible
to treat the spectrum as if it consisted of a Gaussian mixture distribution of
sufficient terms [35]. The reference data then depend on the mean and variance
of the real and imaginary components at each spectral coefficient for each
mixture contribution. For homogeneous textures the mean and variance in the
real and imaginary components were the same. For simplicity, we make the same
assumption for targets so that the likelihood for target class p and shift d, related
to (9.10), is given by
M[
\p{d) = -M\ni;-Jj\\nSp{j+d)
(C) (d)
Figure 10.14 Target change detection: (a) first image, (b) second image, (c) targets detected
in (a) but not (b), and (d) targets detected in (b) but not (a).
where Sp{j) = 2 Vp(j) is the power spectral density of the yth component of the
^?th target class. The overall maximum value of\p(d), then identifies both target
position and class.
It should be noted that (10.30) provides a theoretical framework for ML
target recognition. However, such an approach is time consuming and it is
important to examine compromise solutions offering slightly degraded perform-
ance but reduced cost. Pattern matching for deterministic structures depends on
orientation and displacement, unlike the correlated texture classification ap-
proach described in Section 9.3. Nevertheless, the latter can provide a description
of ML classification. The correct orientation of the rectangular ROI, as described
in the previous section, reduces the alignment problem to one of two-dimensional
translation [36]. A variety of slightly different pattern-matching techniques such
as eigenimage classification [37,38], shift-invariant twodimensional pattern
matching [38], and quadratic distance correlation classification [36,39,40] can
then be applied; these all offer similar performance. Though they are not truly ML
methods, they can be related to these through (10.30). The last term of (10.30),
corresponding to the square modulus of the difference of the data and reference
components, contains a contribution from the crosscorrelation between reference
and data. Thus, it is related to the quadratic distance correlation classifier or
shift-invariant two-dimensional pattern-matching classifier proposed by Novak
et al. [36]. Note, however, that minimum mean-square and correlation measures
are only optimal if the variances of the classes are the same, which is not the case.
A variety of other simplified classifiers have been proposed, but they all
lead to degraded performance. Therefore, these should be rejected in favor of
the two-dimensional pattern-matching methods. Though somewhat subopti-
mal, the minimum mean-square error and crosscorrelation measures have been
refined and realistically implemented. Classification results for the shift-invari-
ant two-dimensional pattern matcher over a restricted number of targets are
shown in Table 10.2 [36]. Very similar results were reported also for the quad-
ratic distance correlation classifier [36] and, more recently, for the eigenimage
approach [38]. The probability of detection was reduced to about 90% after this
classification stage rather than the 100% performance achieved after detection
and discrimination stages [36].
These results show promise but are only preliminary, gathered over a
restricted range of geometries. Including more target classes and imaging con-
Table 10.2
Confusion Matrix for Classification of Tank, APC,
and Self-Propelled Gun Against Clutter for 1 -ft Resolution Polarimetric Data
Tank 100
APC 100
Gun 100
Clutter 2 4 94
Source: [36]. Copyright 1994, reprinted with kind permission from Elsevier Science Ltd., The
Boulevard, Langford Lane, Kidlington 0X5 1GB, UK.
figurations would tend to degrade classification. It is not clear what performance
this factorized approach with a discrimination stage could achieve under these
circumstances. Nevertheless, at present it provides the only realistic option to
bridge the gap between the enormous number of initial detections and final
target recognition.
g = //f+b (10.31)
where g, f, and b are vectors representing the complex image samples, the field
scattered by the target, and additive zero-mean Gaussian background speckle,
respectively; H is a matrix representing the imaging process. The inverse scat-
tered field problem of recovering f from g represents the normal reconstruction
process. The inverse cross-section problem of recovering the samples of a from
g [41,42] is closer to the real requirement but is considerably more complicated.
where W denotes the adjoint of H. The effect of the additive background can
be appreciated by considering a singular value decomposition (SVD) of the
imaging matrix H [43]. Suppose that u ; and v;- are the ixh components of the
orthonormal basis vector in the scattered field and image, respectively. Each
component of ui passes through the imaging system to give an output compo-
nent v- attenuated by X-, the corresponding eigenvalue of the SVD. The least-
squares reconstruction can then be expressed as
where P(f) encapsulates the prior knowledge about the scattered field, which
itself depends on prior knowledge about the RCS. P(glf) is the likelihood
function which incorporates information about the imaging process itself. If the
PDFs are assumed to be Gaussian, the peak value of the a posteriori PDF, /^fIg),
corresponds to the minimum mean-square error estimate of f given g. Thus, the
MAP solution can be adopted as a suitable reconstruction, frec, from the entire
range of possible solutions.
Let us assume a Gaussian a priori PDF for the scattered field [46] given
by
7
^) S * / r \ eXrf"ftC"f 1 (1035)
det (IT C J
where Cs denotes the covariance matrix of the object. If we assume for simplicity
that the two-point statistics described by the covariance matrix are in fact
delta-correlated, then the matrix contains diagonal terms only, corresponding to
a weighting function profile or underlying cross section [47]. The background
clutter is modeled as in Chapter 4 by the Gaussian likelihood PDF defined by
Ag|f) - — ^ - « p [ - ( g - M)fCb->(g - M)] (10.36)
L J
det TT C b
ex
^(%) - . V P[~(f -(JC-L(f -U ] (10.37)
and
C
:L = WC^H + C~l (10.39)
a
^Hg) ^(gk№) (io-4o)
Note that the relationship between g and (T is carried through the scattered
field £ Thus, one means of solving (10.40) is to solve for f given g for various
estimates of O". The global maximum then corresponds to the desired solution.
In order to avoid a global search, a hill-climbing routine can be used to
determine the direction in which to change cr. Unfortunately, P(O1Ig) does not
exhibit Gaussian behavior amenable to linear techniques. However, the behavior
for small changes can be developed in terms of a Taylor expansion [42], depend-
ing on the first derivative, which can be incorporated into an iterative approach
to super-resolution [48].
(C) (d)
(e) (f)
Figure 10.15 Stages in super-resolution reconstruction: (a) original 3-m resolution image with
background, (b) derived background and target cross sections,(c) reconstructed
image, (d) derived background and target cross section for second iteration, (e)
reconstructed image, and (f) original 0.75-m resolution image without back-
ground for comparison.
detections in Figure 10.15(b) have been reconstructed more strongly, as has a
new false detection on the right of Figure 10.15 (d).
The acid test of such a reconstruction is to compare it with the known
target model. Figure 10.15(f) shows a simulated image of a tank with no back-
ground at a resolution four times better in each direction. It seems that three of
the four scattering centers in the super-resolved image (Figure 10.15(e)) have
approximately the correct position and RCS. However, the scatterer on the lower
right is significantly displaced. Reconstructions with different speckle realiza-
tions consistently reproduce the three "correct" scatterers while the fourth varies
appreciably. This illustrates that background speckle prevents further iterations
being useful. While these early results are somewhat crude, a more sophisticated
iterative approach [48] yields similar consequences. In this case, reconstruction is
automatically terminated when the effect of speckle begins to dominate.
10.10 Summary
This chapter has been devoted to deriving information about targets from SAR
images. Three principal stages were described: detection, discrimination, and
classification. In each case the aim was to develop optimum techniques for
extracting the relevant information.
Detection was treated in Bayesian terms based on the Neymann—Pearson
observer (Section 10.2). Following a conventional CFAR analysis for a Swerling
1 target against constant RCS clutter (Section 10.3), we demonstrated:
[ 1 ] Woodward, P. M., Probability and Information Theory with Applications to Radar, New York:
McGraw-Hill, 1953.
[2] Neyman, J., and E. S. Pearson, "The Problem of the Most Efficient Tests of Statistical
Hypothesis," Phil. Trans. Roy. Soc. (London), Vol. A231, 1933, pp. 289-333.
[3] Galati, G., and R. Crescimboni, "Basic Concepts on Detection, Estimation and Optimum
Filtering," Advanced Radar Techniques and Systems, G. Galati (ed.), London: IEE, 1991,
Chap. 1.
[4] Goldstein, G. B., "False-Alarm Regulation in Lognormal and Weibull Clutter," IEEE
Trans., Vol. AES-9, 1973, pp. 84-92.
[5] Musha, T, and M. Sekine, "CFAR Techniques in Clutter," Advanced Radar Techniques and
Systems, G. Galati (ed.), London: IEE, 1991, Chap. 3.
[6] Blacknell, D., and R. J. A. Tough, "Clutter Discrimination in Polarimetric SAR Imagery,"
Europto Conf. on SAR Data Processing for Remote Sensing II, Paris, Proc. SPIE, Vol. 2584,
1995, pp. 179-187.
[7] Marcum, J. L, and P. Swerling, "Studies of Target Detection by Pulsed Radar," IRE Trans.,
Vol. IT-6, 1960, pp. 59-308.
[8] Abramowitz, M., and I. A. Stegun, Handbook of Mathematical Functions, New York: Dover,
1970.
[9] Ward, K. D., "Compound Representation of High Resolution Sea Clutter," Electron. Lett.,
Vol. 17, 1981, pp. 561-565.
[10] Sekine, M., T Musha, Y. Tomiyata, and T. Iraba, "Suppression of Weibull-Distributed
Clutter," Electron. Comm. Japan, Vol. K62-B, 1979, pp. 45-49.
[11] Cole, L. G., and P W Chen, "Constant False Alarm Detector for a Pulse Radar in a
Maritime Environment," Proc. IEEE, NAECON, 1987, pp. 1101-1113.
[12] Clarke, J., and R. S. Peters, "Constant False Alarm Detector Adaptive to Clutter Statistics,"
RSRE Memo. (UK) No. 3150, 1978.
[13] Farina, A., A. Russo, F. Scannapieco, and S. Barbarossa, "Theory of Radar Detection in
Coherent Weibull Clutter," IEE Proc. F, Vol. 134, 1987, pp. 174-190.
[14] Hansen, V. G., "Constant False Alarm Rate Processing in Search Radars in Radar—Present
and Future," IEE Conf. Publ, Vol. 105, 1973, pp. 325-332.
[15] Kreithen, D. E., S. M. Crooks, W W. Irving, and S. D. Halversen, "Estimation and
Detection Using the Product Model," MIT-Lincoln Lab. Report No. STD-37, 1991.
[16] Oliver, C. J., and P. Lombardo, "Simultaneous Mean and Texture Edge Detection in SAR
Clutter," IEE Proc. Radar Sonar Navig, Vol. 143, 1966, pp. 391-399.
[17] Novak, L. M., and M. C. Burl, "Optimal Speckle Reduction in POL-SAR Imagery and Its
Effect on Target Detection," Millimetre Wave and Synthetic Aperture Radar, Proc. SPIE, Vol.
1101, 1989, pp. 84-115.
[18] Novak, L. M., and M. C. Burl, "Optimal Speckle Reduction in Polarimetric SAR Imagery,"
IEEE Trans. Aerospace Electron. Systems, Vol. AES-26, 1990, pp. 293-305.
[19] Novak, L. M., and C. M. Netishen, "Polarimetric Synthetic Aperture Radar Imaging," Int.
J. Imaging Sci. Tech., Vol. 4, 1992, pp. 306-318.
[20] White, R. G., "A Simulated Annealing Algorithm for Radar Cross-Section Estimation and
Segmentation," SPIE Int. Conf. on Applications of Artificial Networks !^Orlando, FL, SPIE
Proc.y Vol. 2243, 1994, pp. 231-241.
[21] Watts, S., "Radar Detection Prediction in Sea Clutter Using the Compound K-Distribu-
tion Model," IEE Proc. F, Vol. 132, 1985, pp. 613-620.
[22] Rohling, H., "Radar CFAR Thresholding in Clutter and Multiple Target Situations," IEEE
Trans. Aerospace Electron. Systems, Vol. AES-19, 1983, pp. 608-621.
[23] Hansen, V G., and J. H. Sawyers, "Detectability Loss Due to Greatest-of-Selection in a
Cell Averaging CFAR," IEEE Trans. Aerospace Electron. Systems, Vol. AES-16, 1980,
pp. 115-118.
[24] Weiss, M., "Analysis of Some Modified Cell Averaging CFAR Processors in Multiple Target
Situations," IEEE Trans. Aerospace Electron. Systems, Vol. AES-18, 1982, pp. 102-114.
[25] Home, A. M., R. G. White, and C. J. Baker, "Extracting Militarily Significant Information
from SAR Imagery," Battlefield Systems International 96, Chertsey, UK, 1996.
[26] Novak, L. M., S. D. Halversen, G. J. Owirka, and M. Hiett, "Effects of Polarisation and
Resolution on the Performance of a SAR Automatic Target Recognition System," Lincoln
Laboratory J., Vol. 6, 1995, pp. 49-67.
[27] Environmental Research Institute of Michigan (ERIM) study under Strategic Target Algo-
rithm Research (STAR) contract.
[28] Kreithen, D. E., and S. D. Halversen, "A Discrimination Algorithm and the Effect of
Resolution," IEEE National Radar Conf., 1993, pp. 128-133.
[29] Butterfield, J., "Fractal Interpolation of Radar Signatures," IEEE/NTC Proc, 1991,
pp. 83-87.
[30] van ZyI, J. J., "Unsupervised Classification of Scattering Behaviour Using Radar Po-
larimetry Data," IEEE Trans. Geoscience Remote Sens., Vol. 27, 1989, pp. 36-45.
[31] Finley, I. P., C. J. Oliver, R. G. White, and J. W. Wood, "Synthetic Aperture Radar—Auto-
matic Change Detection," UK Patent No. GB 2,234,13OB, 1988; US Patent No.
4,963,877, 1990.
[32] Wood, J. W, R. G. White, and C. J. Oliver, "Distortion-Free SAR Imagery and Change
Detection," Proc. US National Radar Conf, Ann Arbor, MI, 1988, pp. 95-99.
[33] Blacknell, D., "Texture Anomaly Detection in Radar Imagery," Europto Conf on SAR Data
Processing for Remote Sensing, Rome, Proc. SPIE, Vol. 2316, 1994, pp. 125-136.
[34] Oliver, C. J., "Optimum Classification of Correlated SAR Textures," Europto Conf on SAR
Image Analysis, Simulation and Modelling I, Taormina, Sicily, SPIE Proc, Vol. 2958, 1996,
pp. 64-73.
[35] Blacknell, D., "A Mixture Distribution Model for Correlated SAR Clutter", Europto Conf.
on SAR Image Analysis, Simulation and Modelling II, Taormina, Sicily, SPIE Proc, Vol. 2958.
1996.
[36] Novak, L. M., G. J. Owirka, and C. M. Netishen, "Radar Target Identification Using
Spatial Matched Filters," Pattern Recognition, Vol. 27, 1994, pp. 607-617.
[37] Turk, M., and A. Pentland, "Eigenfaces for Recognition,"/. Cognitive Neuroscience, Vol. 3,
1991, pp. 71-86.
[38] Novak, L. M., and G. J. Owirka, "Radar Target Identification Using an Eigen-Image
Approach," Proc. IEEE National Radar Conf, Atlanta, GA, 1994, 1996, pp. 129-131.
[39] Mahanalobis, A., A. V Forman, M. Bower, N. Day, and R. Cherry, "A Quadratic Distance
Classifier for Multi-Class SAR ATR Using Correlation Filters," SPIE Conf. on Ultrahigh
Resolution Radar, Los Angeles, SPIEProc, Vol. 1875, 1993, pp. 84-95.
[40] Mahalanobis, A., A. V. Forman, N. Day, M. Bower, and R. Cherry, "Multi-Class SARATR
Using Shift-Invariant Correlation Filters," Pattern Recognition, Vol. 27, 1994, pp. 619-626.
[41] Luttrell, S. P., "A Bayesian Derivation of an Iterative Autofocus/Super-Resolution Algo-
rithm," Inverse Problems, Vol. 6, 1990, pp. 975-996.
[42] Luttrell, S. P, "The Theory of Bayesian Super-Resolution of Coherent Images: a Review,"
Int. J. Remote Sensing, Vol. 12, 1991, pp. 303-314.
[43] Bertero, M., and E. R. Pike, "Resolution in Diffraction Limited Imaging, a Singular Value
Analysis. Part 1: The Case of Coherent Illumination," Opt. Ada, Vol. 29, 1982,
pp. 727-746.
[44] Guglielmi, V, F. Castanie, S. Puechmorel, and P Piau, "Comparison of Super-Resolution
Techniques for SAR Pre-processing," Europto Conf. on SAR Image Analysis, Simulation and
Modelling, Paris, SPIE Proc, Vol. 2584, 1995, pp. 252-263.
[45] Cox, R. P., "Probability, Frequency and Reasonable Expectation," American J. Physics, Vol.
17, 1946, pp. 1-13.
[46] Luttrell, S. P, "Prior Knowledge and Object Reconstruction Using the Best Linear Estimate
Technique," Opt. Ada, Vol. 32, 1985, pp. 703-716.
[47] Luttrell, S. P., and C. J. Oliver, "Prior Knowledge in Synthetic-Aperture Radar Processing,"
/. Phys. D:Appl. Phys., Vol. 19, 1986, pp. 333-356.
[48] Pryde, G. C , L. M. Delves, and S. P. Luttrell, "A Super-Resolution Algorithm for SAR
-Images," Inverse Problems, Vol. 4, 1988, pp. 681-703.
[49] Novak, L. M., G. R. Benitz, G. J. Owirka, and L. A. Bessette, "ATR Performance Using
Enhanced Resolution," SPIE Conf. on Algorithms for Synthetic Aperture Radar Imagery III,
Orlando, FL, SPIE Proc, Vol. 2757, 1996, pp. 332-337.
[50] Oliver, C. J., "The Limits on SAR Resolution Imposed by Autofocus Uncertainty," Int. J.
Remote Sensing, Vol. 14, 1993, pp. 485-494.
1 1
11.1 Introduction
Up to this point, we have been concerned only with the information carried by
a single SAR image. Of increasing importance are SAR systems that provide
multidimensional information via multiple frequencies or polarizations. Such
systems provide a much enhanced capacity for investigating Earth terrain be-
cause different frequencies and polarizations allow the probing of different
scattering mechanisms and different components of the scattering layers. For
example, in forest imaging, modeling suggests that for shorter wavelengths the
backscatter from mature conifers is dominated by direct crown backscattering
at all polarizations but that at long wavelengths the main component of the
return depends on the polarization: trunk-ground interactions dominate at HH,
for W the major contribution to the return is direct crown backscatter, while
the HV return is predominantly due to the primary branches in the crown layer
[1,2].
As for single-channel data, a primary step in extracting information from
such images is to develop a model of the image statistics for distributed scatterers
because in many applications, such as agriculture, forestry, and hydrology, such
targets are the objects of interest. Even when we are concerned with small
scatterers, their detection and classification is crucially affected by their clutter
surroundings. Hence, the main goal of this chapter is to establish statistical
models appropriate to multichannel data, especially polarimetric data. We will
find that much of the structure already developed in earlier parts of this book
transfers naturally into higher dimensions and provides a clear guide as to what
is meant by information. In particular, the Gaussian data model again plays a
central role in the analysis but with an added richness provided by the extra data
channels and the enhanced range of useful parameters.
A key concern when handling multichannel data is the correlation be-
tween channels, and radar polarimetry provides an ideal vehicle to discuss the
issues this raises. As a result, we initially concentrate on this data type before
applying the lessons learnt more generally. The basic concepts of polarimetric
measurement are introduced in Section 11.2. The treatment is at a sufficient
level to appreciate what is described by polarimetric data, but any reader
intending to make serious use of such data would be well advised to consult a
more comprehensive treatment, such as that provided in [3]. Given that po-
larimetric data are inherently multidimensional, an immediate concern is to
elucidate which of the many possible combinations of channels carry informa-
tion and, hence, are useful to display as images. To deal with this we must have
some knowledge of which parameters convey physical information about the
scene and/or describe the statistical distributions characterizing the data. Ac-
cordingly, in Section 11.3 we develop an argument based on collections of point
scatterers that suggests the multidimensional Gaussian model is crucial in de-
scribing polarimetric data, but with modifications caused by fluctuations in the
numbers of scatterers. The Gaussian model described in Section 11.4 leads
naturally to the image types that display the information carried by the po-
larimetric data. The associated single-look distributions are developed in Section
11.5, and their properties are discussed in Section 11.6. Section 11.7 is con-
cerned with parameter estimation and its attendant problems. This leads us to
consider multilook data and provides a natural bridge to the Stokes vector
formulation of scattering in Section 11.8. Texture in polarimetric data is dis-
cussed in Section 11.9. Finally, Section 11.10 discusses how the results of this
chapter are applicable to other forms of multidimensional data, especially inter-
ferometric SAR.
Here we deal only with the most common form of polarimetric SAR, in which
a linearly polarized signal is transmitted and two orthogonal polarizations of the
backscattered signal are measured. When a horizontally polarized wave is trans-
mitted, the signals received in the horizontal (H) and vertical (V) channels
undergo separate SAR processing to produce measurements 5hh and 5vh of the
local copolarized and crosspolarized complex scattering amplitudes. By inter-
leaving H and V polarized transmitted pulses, the corresponding terms Sw and
5hv can also be measured to give the full polarimetric response of a scatterer.
These four measurements allow the response to an arbitrary transmitted polari-
zation to be calculated (for a fixed frequency, incidence angle, resolution, and
time), as follows. Any polarization state of the transmitted wave can be uniquely
described by an electric field vector of the form
E = EJ + E.J (in)
where the subscript i denotes incident and h and v are unit vectors defined by
h = z X k/\z X k and v — h X k. Here z is a unit normal to the Earths
surface and k is a unit vector parallel to the wave vector k. The backscattered
wave observed at a distance R in the far field of the scatterer is then given by
{ }
U J R Uh O U J
where the subscript s denotes scattered. Note that in this expression the coordi-
nate system is throughout defined relative to the direction of the transmitted
wave (the backscatter alignment convention); the transformations needed if the
scattered wave is defined relative to its propagation direction (the forward
scattering alignment convention) are discussed in [3]. Notice also that a single-
channel SAR measures only one of the complex scattering amplitudes Spq, which
is normally one of the copolarized responses 5 hh or Sw.
Implicit in this formulation are the conditions:
• Adequate sampling for the returns from both the H and V transmitted
pulses;
• Preservation of phase coherence between pulses;
• Correct sampling so that all responses are measured at the same posi-
tion;
• Preservation of the internal state of the scatterers between pulses.
While the first three of these conditions are concerned with the system
and can be treated as an engineering problem, the last depends on the physics
of the target. For targets whose internal state is unaltered by the polarization of
the probing wave, reciprocity will hold [4], that is,
Shv=Svh (11.3)
This is expected to be the case for most naturally occurring scatterers. However,
targets can be constructed for which this condition is violated, and (11.2) may
then become meaningless except for pure H and V inputs.
Even when the targets are reciprocal, measured values of 5hv and Syh may
not be equal, because of system imperfections. In fact, polarimetric SAR can be
affected by a number of forms of distortion, in addition to those due to
uncorrected spatial or temporal variations in power or gain discussed in Chapter
2. These include:
Of these the first is the most pernicious because it scrambles the informa-
tion in the different channels. Fortunately, methods to correct for crosstalk are
well developed [5-7] although incomplete because of the inadequate treatment
of PSF effects [8,9]. Channel imbalance can be separated into correction of
phase and amplitude distortions. Calibration targets or internal calibration
tones are needed to remove these effects [10,11]. In this chapter we assume that
all corrections have been made and that targets are reciprocal. The data at each
pixel can then be represented by a three-vector:
°hh °i
S= Shv = S2 (11.4)
V o w7 v°3/
a W r
\ ^ \ N k\e
s= ; - x '- e^k (n-5)
[aM +ibM) k=
\rkMel%M)
S = 7G (11.6)
The observations will be pure Gaussian (so that we can set J ' = 1) only for
number distributions in which (Nk)/(N)k —> 1 as (N) —» <*>.
This result places the multivariate Gaussian distribution at the heart of
polarimetric data analysis, for two reasons.
S = Va m (11.10)
Sk = Jj~ke*kmk (11.12)
Sk =myfcei*k (11.14)
where mis a circular zero-mean Gaussian variable with unit variance, which is
independent oi k. Although apparently similar to (11.10), this decomposition is
of little value for a number of reasons.
S = TG (11.15)
f a
i V<W>L2 V° r i q '3ft3 >
a a
C= V i 2pL2 O" 2 V a 2 a 3 P23 (11.16)
^V°"ia3Pi*3 V°"2°"3P2*3 °"3 ,
where
Q/ = ( v ; ) ai.17)
^ = (N2) a us)
is the backscattering coefficient in channel k and
ft,= ^ ^ (1U9)
(d) («0 (O
(g) (h) 0)
Figure 11.1 Estimated C-band polarimetric parameters from Feltwell, UK, Imaged by the
NASA/JPL AirSAR system: (a) ahh, (b) ahv, (c) a w , (d)| Phh,hv|, (e)| Phhf J , (f)| PhV/
J , (g) ZPhhf hv, (h) ZPhh vv, and (i) Zp h v w . The channel powers apq are indi-
cated by the dB scales. The scaling of the amplitude and phase of the corre-
lation coefficients is from 0 to 1 and from -TT to IT, respectively.
(a) (b) M
(d) (e) (0
(g) (k) 0)
The second of these is related to the fact that for azimuthally symmetric
targets, the cross-polarized correlation coefficients will be zero [20], so the
(a) (b) M
(g) (*0 «
covariance matrix (11.16) contains only five real parameters. This condition
would be expected to hold for the shorter wavelengths if the canopy appears
isotropic, but row effects in the soil could rise to the observed asymmetry due
to the greater penetrating power at longer wavelengths. Additionally, we will
show in Section 11.9 that the longer wavelength data need further parameters
for a full description, because they are not pure Gaussian and exhibit texture.
Hence Figures 11.1 to 11.3 do not tell the full story.
and
These equations relate the covariance matrix CR of the real data vector SR = (av
bv . . . , aMy bj^)1 to the corresponding complex covariance matrix C
The Gaussian PDF is completely defined by its covariance matrix and in
the real case is given by
P S
( *) = I L 1 , e x P H S R C R ' S R ) <1L23)
V(2lT) ICRI
where I Q l is the determinant of CR. It can be shown that I Q l = 2~2Afl Cl2 and
Sj1C^1S11 = 2S1^C-1S, where S+ is the conjugate transpose of S. Hence,
P^=W2 (n-27)
where p^ is the correlation coefficient of intensity between channels i and k
This important connection between the correlation structures of the fields and
intensities, known as the Siegert relationship [22], has already been encountered
in a different guise as (4.34).
It should also be noted that for polarimetric data all fourth-order field
/ \
moments are zero unless they are of the form (SS -SlS*,), otherwise each
corresponding term on the right-hand side of (11.25) contains a factor that is
zero. For moments of this form, we can write
(SiSjS;s;) = C^Cj1+cacjk (11.28)
Of particular importance for polarimetric data are parameters formed from pairs
of channels, for which the PDF is jointly Gaussian with M = 2 (see, for
example, the images formed from the correlation coefficients in Figures 11.1 to
11.3). We will denote these channels as Sk = rke®k with k — 1, 2, and their
complex correlation by
p= ^£^ (11.29)
where a k is the RCS of channel k. Here 1 and 2 can stand for any of hh, hv, or
w.
Pairs of channels can be combined in a variety of ways, but for single-look
data, two combinations have particular importance. These are the complex
Hermitian product S1S2 and the channel ratio S1IS2. Since the phases of these
two quantities are the same, only three real quantities are involved, viz. rx r2, T1Zr2
and O1 — O2. The distributions of these quantities have been discussed by a
number of authors [17,23—25], and we only briefly indicate their derivations. A
first step is to set M — 2 and transform coordinates in (11.24) by
52 = Jo^Esmtye^ (11.30b)
2TT 2 (I - IPI )
4ir2 ( l - F s i n 2 i | i )
where Y = |p| COs(B1 — G2 — Z p ) . Substituting t = tan i|/ and using partial
fractions allows i|/ to be integrated out. The joint distribution P(Q i> 62) *s then
obtained as
7 1 1 7
P(Q) = ^ ^ ^ *> (1 L 3 6 )
2 2
[(7 + Q ) - H P I Q 2 ] '
2
where 7 = Cr2Za1.
For the amplitude product, we return to (11.31) and integrate out B1 and
O2 to yield the joint PDF
*A«-M^4--%L-),,№i) (1,38,
X sm2\\) y Xsin2\\f) \ X J
where X = ^J(J1(J2 (1 — |p| J. The integral over \\f assumes a standard form with
the substitution u = cosec \\f and leads to
The distributions (11.34), (11.36), and (11.39) each depend on two real pa-
rameters. For the phase difference distribution these are the amplitude and phase
of the complex correlation coefficient, IpI and Z p ; for the amplitude ratio they
are IpI and 7 = <J2I®\ while the amplitude product has a distribution involving
IpI and (T1CF2. For single-look data, the phase difference and amplitude ratios are
particularly important, and their properties are discussed in this section. The
significance of the Hermitian product becomes clearest in the context of pa-
rameter estimation and then in its multilook form. This is treated in Section
11.7, although a detailed treatment of the single-look form is given in [24].
Frequency
Figure 11.4 L-band EMISAR phase difference images from an agricultural area: (a) HH-VV and
(b) HH-HV. Observed and theoretical phase difference distributions for a single
winter wheat field of 5,537 pixels in the above images: (c) HH-VV, IpI = 0.67, Z p
= 33 degrees; and (d) HH-HV, IpI = 0.19, Z9 = - 8 9 degrees.
together with the associated theoretical distributions (shown by the solid curves)
using values of p estimated from the field. For the H H - W data, IpI ~ 0.67 and
Z p ~ 33 degrees while the corresponding values for H H - H V are 0.19 and
— 89 degrees, respectively. The good fit between theory and measurement is
clear, as is the dependence of the width of the distribution on IpI.
Quantitative measures of fit are much aided by the fact that the phase
difference distribution has an analytic CDF given by
W = Jl^dK
= /(<(> - Zp)+ /(IT + Zp) -TT < c)j < IT (11.40a)
where
= j _f IPI an* COS-I[-1P| COS *]\ (j i m
2
2*{ Vi-|p| COS^ J
This greatly simplifies the use of the K-S test of fit (see [14], for example).
It is clear that the natural measures of centrality and width for the single-
look phase difference distribution are Z p and IpI, but it is common in the
literature to find phase statistics described in terms of mean and standard
deviation. These quantities can be expressed in terms of p since [24]
|p|si nZp
(G,) = Zp + 2 cos-'dplcosZp) (11.41)
COS2
Vi ~ IPI ^P
and
2 ^i n2
Both have serious flaws as summary statistics that arise from the circular
nature of the phase measurement. These are easily demonstrated by a simple
example. Consider the two phase difference distributions indicated in Figure
11.5. These are identical apart from a TT phase shift, with that on the left
[rad]
[rad]
Figure 11.5 Examples of two phase difference distributions, identical except for a phase shift
of TT radians.
centered on zero and that on the right centered on u radians. Note that both
distributions have mean zero but the variance is much larger for the distribution
on the right. Hence, the mean and variance can provide a very misleading way
of summarizing the distributional shape unless the modal phase difference is
near zero. However, it is not uncommon in polarimetric data for modes well in
excess of IT/2 to occur (e.g., in woodland at longer wavelengths [14]). A
thorough quantitative analysis of this issue is given in [27].
Avoiding the use of mean and variance presents no problem when the
distribution is given by (11.34), but alternative measures are needed for more
general cases (including multilook data; see Section 11.7). Such measures,
described in [28,29], are based on assigning to each observed phase 6d a unit
vector (cos 0d, sin 0d). Taking the average of all these vectors yields a
vector R(cos 0, sin Gj. The mean direction 0 is independent of choice of origin;
the resultant length lies^n the range [0,1], with 1 indicating coincident points.
The quantity V = 1 — R is known as the sample circular variance. It is easy to see
that with these definitions the distributions illustrated in Figure 11.5 would have
the same sample circular variance and have mean direction zero and IT as desired.
Examples of the use of these measures are shown in Figure 11.6 for L-band
AirSAR data for regions of two different cover types. In Figure 11.6(a) the
measured values of A^S^S^) are plotted on a circular dial for regions of wheat;
while Figure 11.6(b) shows_the same quantity for coniferous woodland. The
arrows indicate the vector i?(cos 0, sin Oj and the circles are each of radius 1.
The mean direction is similar for both cover types (26 degrees for wheat and
19.8 degrees for conifer), but there is much greater dispersion in the wheat
measurements. This is reflected in the sample circular variance measures of
0.206 for wheat and 0.003 for conifers.
ters, in this case IpI = 0.67 and 7 = 0.9 for the I H H / W I data and IpI = 0.195,
7 = 0.056 for IHV/WI. It is noticeable in this figure that the IHV/WI ratio (b)
gives better defined structure, but the I H H / W I ratio in (a) picks out a region at
lower left that is not obvious in (b). This probably indicates a crop with vertical
structure, such as standing cereal [31]. The widths of the curves are not as easy to
interpret as for the phase difference because the amplitude ratio distribution gets
narrower as IpI increases and wider as 7 increases [34]. Hence, in the two plots
shown as (c) and (d) there are competing effects (note the difference in scales).
N
1
C = -XS^S^t (11.44)
N k=i
(a) (b)
(c) (d)
F i g u r e 11.7 L - b a n d E M I S A R a m p l i t u d e r a t i o i m a g e s f r o m a n a g r i c u l t u r a l a r e a : (a) I H H / W I a n d
(b) I H V / W i . O b s e r v e d a n d t h e o r e t i c a l a m p l i t u d e r a t i o d i s t r i b u t i o n s f o r a s i n g l e
I H V / W I , IpI - 0 . 1 9 , 7 - 0 . 0 5 6 .
where S^ denotes the £th complex data vector. In other words, the sample
covariance matrix is the MLE of the true covariance matrix. In the case of two
channels, the covariance matrix has the form
a^i-fl^l2 (11.46a)
^2 =77 2 K T (n-46b)
and
The joint PDF of the estimates has the form of a complex Wishart distri-
bution [35] given by
P(S,,S7,A,B) = ^ ^V—
V ! 2 ;
TtT(N)T(N-I)W
I
r » i " Ii
1(T2(J1 + (T1(J2 - 2J(T1(T2 Re(pR*)\
A
-N± i
I (11.47)
,. . v 2N(NZ)N f27VVo:^:2~|p|Zcosfi>-Zpl"
P(Z9 * 2 - ^ 1 exp - J *
V ;
TrAr(TV)(Vo1^) [ A
A
I J
where KN_l(.) is a modified Bessel function (Section 9.6 of [26]). From the joint
distributions (11.47) and (11.48) we can derive the marginal distributions of
three parameters, detailed in the following subsections, which are of particular
interest (see also [24]).
11.7.1 The Multilook Phase Difference Distribution
The first of these is the phase difference, whose marginal distribution has the
form [24,36]
PtJ)J-W I (27V~2>!
*l I 2TT [[(N- I)!] 2 2 2 ^ " »
(2N -I)Y . x i 1 i
l
X -i '— cos-H-Y) + -—J7 + —-^
_(1-K2)^ {\-Y*f\ 2(7V-I)
y T(TV-I) r ( ^ - l - r ) l+(2r + l ) ^ |
2 l j
i r ( ^ - i - r ) r(7v-i) (i-r^r J
? = |p| = - J U = (n.50)
where
1
Fl* h J \- W 1 W Y TjI + a)T(l + bfjl + c) zi
3 2 i
' " " ' T(a)T(b)T(c)k T(l + d)T(l + e) /!
22{N-1) Y2IN) 1 / ;
(U53)
M—i^o ±^
where the approximation is for large TV using Stirling's formula (Section 6.1 of
[26]), so at least 314 samples would be needed to reduce the bias to less than
0.1. This bias can be seen in Figures 11.1 to 11.3, where panels (d—f) show r
with N= 16. Very few pixels take values less than 0.1; for the large areas in
the crosspolarized channels that are essentially featureless, most of the pixels
lie in the range 0.1 to 0.3. Perhaps more important than (11.51) is the a
posteriori PDF Pnp|\r J, which is proportional to the PDF given in (11.51) if IpI
is equally likely to take any value in the range [1, O]. However, it is now
considered as a function of IpI; the constant of proportionality normalizes the
expression to have unit area as it is integrated over the range 0 ^ IpI ^ 1. Plots
of this PDF and a discussion of its use in estimating confidence intervals for IpI
are given in [41,42].
which yields the single-look expression (11.36) when we set L = 1. (Note that
(11.36) gives the amplitude ratio, so a simple change of variables is also in-
volved.) The derivation of this result together with theoretical plots and a
comparison with data can be found in [34,38]. The result is identical to the
single-channel form in (7.6) (with M = 2L, k = Z, and R = (T1Z(T2) if the
channels are uncorrelated, so IpI = 0.
11.8 The Stokes Scattering Operator Formulation of
Polarized Scattering
The previous section showed that, for Gaussian polarimetric data, multilooking
provides the best estimates of the information-bearing parameters. Here we
provide another description of the scattering behavior, originally developed for
optical data, which provides a useful route to forming such data.
Radar systems can make direct phase measurements on the received data,
permitting the scattering matrix to be measured. Essential to this is the avail-
ability of stable local oscillators providing a reference phase over the period
of the measurement. By contrast, in optical data, the time scales for coherence
of the individual wave packets making up the signal are so short that phase
measurements must be inferred. This provides the basic motivation for the
description of polarization state introduced by Rayleigh, using the Stokes vector
[3,43]. The Stokes vector corresponding to a completely polarized wave with
polarimetric field vector (Ey, B^)1 is a four-component real vector given by
F I2 - I F
2
£
F- vl Fh (II.55)
^ 2 Im(^A*) J
The incident and scattered complex field vectors, E1 and E s , are related by
the scattering matrix S according to (see 11.2)
pikR
Es =—SV (11.56)
R
P =—QMF[ (11.57)
r2
where the matrix M is known as the Stokes scattering operator, and Q is the
matrix
A O O O^
Q= ° ' ° ° (.1.58,
0 0 1 0
vO O O -h
For reciprocal scattering, 5 and M are symmetric, and the terms in M are
given by
Mn = ^ J 2 + 2 K v | 2 + | S h h | 2 ) (11.59a)
M12=1A(\SJ-2K\2+\SJ) (11.59e)
M2i=L2M^Sl-S^Sl) (11-590
M34 = - I l m ( 5 v v 5 * h ) (11.591)
^44 = ^ ( l \ v | 2 - R e ( ^ W ) ) (11.59J)
C 22 - A T 3 3 + A f 4 4 (11.60b)
S = TG (11.61)
where v is an order parameter. In this case we can derive the PDF of the
scattering matrix data for the Af-dimensional product model as
= - J - J ^ - r r ^ « - e~^ cJ-S!C^)dT
ir^|c|r(v) Jo \ T2 )
=
^ q ^ )(vStc"ls)"f ^ - - f 2 7 ^ ^ ) (11 63)
-
This is a multidimensional K distribution. In similar fashion, any of the other
PDFs arising from the product model can be derived by integrating over T.
Since in this case the multilook phase difference, amplitude ratio, and sample
coherence are all independent of T, they are also independent of texture. In the
second case, the texture variable fluctuates between the samples being averaged.
Few, if any, analytic results are available in this case. For example, even the PDF
of a sum of samples has no analytic form.
Using (11.8) and the fact that [T1) = 1 implies that C- is an unbiased
estimate of Q7-. For each single channel, the analysis described in Section 8.3
indicates that a reasonable estimate of the order parameter v (see (11.62)) is
given by solving the equation
same order parameter, there are effectively 3./V(correlated) samples. Two possible
approaches to using all the information would be to estimate v for each channel
and average the results or instead to solve [24]
where 6" = I (see (11.46)). The relative merits of these two approaches is at
present unknown.
(b)
(c)
Figure 11.8 Estimated values of v~] calculated overthe same region as in Figures 11.1 to 11.3,
using single-look HH data from the NASA/JPL AirSAR at (a) C-band, (b) L-band,
and (c) P-band.
(barley, grass, and wheat), and root crops (potatoes and sugarbeet) gave rise to
very clear distributional differences, as summarized in Table 11.1. It was found
that the Gaussian model appeared viable for all vegetation types at C-band.
While conifers and root crops remained Gaussian, cereal crops became non-
Gaussian at L-band, in most cases being reasonably well fitted by a K distribu-
tion. At P-band only the conifer forest appeared Gaussian. Root crops were
better fitted by K distributions; cereal crops, although modeled better by K
distributions, in many cases gave a poor fit [14].
Table 11.1
Distributional Forms Observed as a Function of Wavelength and
Cover Type for the Feltwell Region in AirSAR Data
The Gaussian behavior of the C-band data is as expected, since at the low
resolution of the imagery we would expect many independent scatterers per
resolution cell, with the backscatter arising predominantly from volume scatter-
ing in the vegetation canopy. In contrast, no satisfactory physical interpretation
of the results at the longer wavelengths is currently available. It is hard to see
how scattering in the canopy could produce the spatial variability implied by
the measurements: the data were gathered in late July, so in most cases the fields
would exhibit full ground cover with crops at a late stage of development.
Alternatively, we may see a soil effect because significant penetration to the soil
surface and even some penetration into the ground will occur at L-band and
particularly at P-band. However, the soil surface would appear comparatively
smooth at these wavelengths. Hence, it may not contribute greatly to the
backscatter nor be able to give rise to the fluctuations in backscattering coeffi-
cient needed to explain the observed behavior. A possibility is that agricultural
treatment gave rise to long-scale variations in surface slope. However, these are
likely to be greater for the root crops than the cereals, whereas the cereals show
the greatest degree of non-Gaussian behavior. In these circumstances, we should
perhaps regard these data as providing a good example of where image analysis
outstrips our current understanding of the underlying physics.
An attractive aspect of these data is that they allowed us to test whether
the phase difference and amplitude ratio distributions were independent of
whether the data are Gaussian or K-distributed, as predicted by the product
model for texture. To this end, K-S fits were performed for all frequencies and
regions using the CDFs given as (11.40) and (11.43). The results confirmed
the robustness of these distributions, particularly at C- and L-bands; at P-band,
a few of the cereal fields gave poor fits, associated with the departure from the
K distribution [14]. A further aspect of the model that was investigated was
the behavior of the order parameter between channels. (The order parameter
was calculated by inverting (8.31), but it should be noted that the interpreta-
tion of this parameter is clear only for those regions displaying a K distribu-
tion.) According to the product model, the order parameter should be the
same for all channels; this was observed in the C- and L-band data. However,
at P-band, significant differences between the estimated order parameters in
the three channels were observed for cereal crops, especially for the grass fields
[14]. This suggests a breakdown in the simple texture model, which could
occur because at the long wavelengths the assumption of a large mean number
of scatterers is no longer valid or because the different polarizations are effec-
tively not seeing the same scatterers due to different attenuation properties.
The theoretical basis for differences in texture variables between polarizations
is discussed in [12], and observed differences are also described in [46]. In the
latter case, the significance of the differences is unclear, because we currently
do not have a well-developed theory to establish confidence intervals for the
order parameter.
11.11 Summary
Despite the apparent complexity introduced when multichannel data are avail-
able, simple models provide a viable means of describing the information
available from distributed targets. For a single frequency, physical reasoning and
measurement support descriptions based on local Gaussian behavior modulated
by a scalar texture, which can be interpreted as a number fluctuation effect. This
interpretation cannot be pushed too far, because, for example, the differing
penetration depths of H and V polarized waves into media that show preferred
horizontal or vertical structures (e.g., cereals) indicate that not all scatterers
receive the same incident wave. Nonetheless, distributions derived from this
model appear consistent with observations, except in some cases at the longest
wavelengths.
Because such simple models can be used, the information-bearing parame-
ters in the data are readily identified. When the data are Gaussian, there is a
well-developed estimation theory for these parameters. In principle, few diffi-
culties arise in applying this theory. In practice, the complicated expressions for
the sample coherence and phase difference distributions have at present pre-
vented the development of a simple treatment for confidence intervals and
moments, for example. For textured data, the estimation theory runs into the
problems identified in Chapter 8, preventing analytic MLE solutions; only
approximate solutions are currently available.The success of a simple tractable
model for the data provides a sound basis for our investigations of filtering,
classification, and segmentation in later chapters. However, it must be remarked
that the empirical basis for this model is currently very incomplete. Compara-
tively few studies over a limited set of terrain types have attempted to validate
it. We have indicated that problems occur at long wavelengths, but more
investigation is needed to establish the effects of, for example, resolution and
cover type. Nonetheless, we will adopt the product model as our basic descriptor
of single-frequency multichannel data in the following chapters, with data
gathered at different frequencies being considered independent.
References
[1] Le Toan, T., A. Beaudoin, S. Goze, J. A. Kong, C. C. Hsu, H. C. Han, and R. T. Shin,
"Microwave Interaction and Analysis of Polarimetric Data: Retrieval of Forest Biomass,"
Final report to ESA ESTEC on Contract No. 9352/91/NULC(SC), 1992.
[2] Le Toan, T, A. Beaudoin, J. Riom, and D. Guyon, "Relating Forest Biomass to SAR Data,"
IEEE Trans. Geosci. Remote Sensing, Vol. 30, 1992, pp. 403-411.
[3] Ulaby, F. T, and C. Elachi, Radar Polarimetry for Geoscience Applications, Norwood, MA:
Artech House, 1990.
[4] Tsang, L., J. A. Kong, and R. T. Shin, Theory of Microwave Remote Sensing, New York: Wiley
Interscience, 1985.
[5] Quegan, S., "A Unified Algorithm for Phase and Crosstalk Calibration—Theory and
Observations," IEEE Trans. Geosci. Remote Sensing, Vol. 32, 1994, pp. 89-99.
[6] van ZyI, J. J., "Calibration of Polarimetric Radar Images Using Only Image Parameters and
Trihedral Corner Reflector Responses," IEEE Trans. Geosci. Remote Sensing, Vol. 28, 1990,
pp. 337-348.
[7] Freeman, A., J. J. van ZyI, J. D. Klein, H. A. Zebker, and Y. Shen, "Calibration of Stokes
and Scattering Matrix Format Polarimetric SAR Data," IEEE Trans. Geosci. Remote Sensing,
Vol.30, 1992, pp. 531-539.
[8] Quegan, S., and I. Rhodes, "Problems in the Linear Distortion Model for Polarimetric
Calibration," Proc. CEOS Calibration Workshop, ESA WPP-048, 1993, pp. 127-132.
[9] Sarabandi, K., L. E. Pierce, and F. T. Ulaby, "Calibration of a Polarimetric Imaging SAR,"
IEEE Trans. Geosci. Remote Sensing, Vol. 30, 1992, pp. 540-549.
[10] Sheen, D. R., A. Freeman, and E. S. Kasischke, "Phase Calibration of Polarimetric Radar
Images," IEEE Trans. Geosci. Remote Sensing, Vol. 27, 1989, pp. 719-730.
[11] Zebker, H. A., and Y. Lou, "Phase Calibration of Imaging Radar Polarimeter Stokes
Matrices," IEEE Trans. Geosci. Remote Sensing, Vol. 28, 1990, pp. 246-252.
[12] Williams, M. L., S. Quegan, and D. Blacknell, "Intensity Statistics in the Distorted Born
Approximation: Application to C-band Images of Woodland," Waves in Random Media,
1997 (in press).
[13] Oliver, C. J., "A Model for Non-Rayleigh Scattering Statistics," Opt. Acta, Vol. 6, 1984,
pp. 701-722.
[14] Quegan, S., and I. Rhodes, "Statistical Models for Polarimetric Data: Consequences,
Testing and Validity," Int. J. Remote Sensing, Vol. 16, 1995, pp. 1183-1210.
[15] Jakeman, E., "On the Statistics of K-Distributed Noise,"/ Phys. A: Math. Gen., Vol. 13,
1980, pp. 31-48.
[16] Jao, J. K., "Amplitude Distribution of Composite Terrain Radar Clutter and the K Distri-
bution," IEEE Trans. Geosci. Remote Sensing, Vol. 22, 1984, pp. 1049-1062.
[17] Yueh, S. H., J. A. Kong, J. K. Jao, R. T Shin, H. A. Zebker, T. Le Toan, and H. Ottl,
"K-Distribution and Polarimetric Radar Clutter," Polarimetric Remote Sensing, J. A. Kong
(ed.), Amsterdam: Elsevier, 1990.
[18] Held, D. N., W E. Brown, A. Freeman, J. D. Klein, H. Zebker, T. Sato, T Miller, Q.
Nguyen, and Y. Lou, "The NASA/JPL Multifrequency Multipolarisation Airborne SAR
System," Proc. IGARSS '88 Symp., Edinburgh, 1988, pp. 345-349.
[19] Wooding, M., E Lodge, and E. Attema, (eds.), MAC-Europe 91: Final Results Workshop
Proc., ESAWPP-88, 1995.
[20] Nghiem, S. V, S. H. Yueh, R. Kwok, and E K. Li, "Symmetry Properties in Polarimetric
Remote Sensing," Radio ScL, Vol. 27, pp. 693-711.
[21] Papoulis, A., Probability, Random Variables and Stochastic Processes, Singapore: McGraw-
Hill, 1984.
[22] Siegert, A. J. E, MIT Rad. Lab. No. 465, 1943.
[23] Davenport, W. B., and W. L. Root, Random Signals and Noise, New York: McGraw-Hill,
1958.
[24] Tough, R. J. A., D. Blacknell, and S. Quegan, "A Statistical Description of Polarimetric and
Interferometric Synthetic Aperture Radar Data," Proc. Roy. Soc. London A, Vol. 449, 1995,
pp. 567-589.
[25] Middleton, D., Introduction to Statistical Communication Theory, New York: McGraw-Hill,
1960.
[26] Abramowitz, M., and I. Stegun, Handbook of Mathematical Functions, New York: Dover,
1965.
[27] Quegan, S., L. V. Dutra, and K. D. Grover, "Phase Measurements in MAESTRO Po-
larimetric Data from the UK Test Sites," Int. J. Remote Sensing, Vol. 15, 1994,
pp. 2719-2736.
[28] Mardia, K. V, Statistics of Directional Data, New York: Academic Press, 1972.
[29] Fisher, N. L, Statistical Analysis of Circular Data, Cambridge: Cambridge University Press,
1993.
[30] Oh, Y, K. Sarabandi, and F. T. Ulaby, "An Empirical Model and an Inversion Technique
for Radar Scattering from Bare Soil Surfaces," IEEE Trans. Geosci. Remote Sensing, Vol. 30,
1992, pp. 370-381.
[31] Dubois, P. C , J. J. van ZyI, and T. Engman, "Measuring Soil Moisture with Imaging
Radars," IEEE Trans. Geosci Remote Sensing, Vol. 33, 1995, pp. 915-926.
[32] Schreier, G., "Geometrical Properties of SAR Images," SAR Geocoding: Data and Systems,
G. Schreier (ed.), Karlsruhe: Wichmann, 1993, pp. 103-134.
[33] Shi, J., J. Dozier, and H. Rott, "Snow Mapping in Alpine Regions with Synthetic Aperture
Radar," IEEE Trans. Geosci. Remote Sensing, Vol. 32, 1994, pp. 152-158.
[34] Joughin, I. R., D. P. Winebrenner, and D. B. Percival, "Probability Density Functions for
Multilook Polarimetric Signatures," IEEE Trans. Geosci. Remote Sensing, Vol. 32, 1994,
pp. 562-574.
[35] Goodman, N. R., "Statistical Analysis Based on a Certain Multivariate Gaussian Distribu-
tion (an Introduction)," Ann. Math. Stat., Vol. 34, 1963, pp. 152-177.
[36] Tough, R. J. A., "Interferometric Detection of Sea Surface Features," DRA Memorandum
4446, 1991.
[37] Barber, B. C., "The Phase Statistics of a Multichannel Radar Interferometer," Waves in
Random Media, Vol. 3, 1993, pp. 257-266.
[38] Lee, J.-S., K. W. Hoppel, S. A. Mango, and A. R. Miller, "Intensity and Phase Statistics of
Multilook Polarimetric and Interferometric SAR Imagery," IEEE Trans. Geosci. Remote
Sensing, Wichmann, Vol. 32, 1994, pp. 1017-1028.
[39] de Groot, M. H., Probability and Statistics, Reading, MA: Addison-Wesley, 1989.
[40] Touzi, R., and A. Lopes, "Statistics of the Stokes Parameters and of the Complex Coherence
Parameters in One-Look and Multi-Look Speckle Fields," IEEE Trans. Geosci. Remote
Sensing, Vol. 34, 1996, pp. 519-531.
[41] Giani, M., C. Prati, and F. Rocca, "Study on SAR Interferometry and Its Applications,"
ESRIN Contract No. 8928/90/F/BZ, 1992.
[42] Foster, M. R., and J. Guinzy, "The Coefficient of Coherence, Its Estimation and Use in
Geophysical Data Processing," Geophys., Vol. 32, 1967, pp. 602—616.
[43] Born, M., and E. Wolf, Principles of Optics, New York: Pergamon Press, 1965.
[44] Quegan, S., and I. Rhodes, "Statistical Models for Polarimetric Data," Proc. IEE Seminar
on Texture Analysis in Radar and Sonar, IEE Digest No. 1993/207, London, 1993,
pp. 8/1-8/8.
[45] Oliver, C. J., "Optimum Texture Estimators for SAR Clutter," / Phys. D, Vol. 26, 1993,
pp.1824-1835.
[46] Sheen, D. R., and L. P. Johnston, "Statistical and Spatial Properties of Forest Clutter with
Polarimetric Synthetic Aperture Radar," IEEE Trans. Geosci. Remote Sensing, Vol. 30, 1992,
pp. 578-588.
[47] Rodriguez, E., and J. M. Martin, "Theory and Design of Interferometric Synthetic Aper-
ture Radars," IEE Proc.-F, Vol. 139, 1992, pp. 147-159.
[48] Zebker, H. A., and R. M. Goldstein, "Topographic Mapping from Interferometric Syn-
thetic Aperture Radar Observations,"/. Geophys. Res., Vol. 91, 1986, pp. 4993-4999.
[49] Kwok, R., and M. A. Fahnestock, "Ice Sheet Motion and Topography from Radar Inter-
ferometry," IEEE Trans. Geosci. Remote Sensing, Vol. 35, 1996, pp. 189-200.
[50] Wegmuller, U., and C. Werner, "Retrieval of Vegetation Parameters with SAR Inter-
ferometry," IEEE Trans. Geosci. Remote Sensing, Vol. 35, 1996, pp. 18-24.
[51] Massonnet, D., M. Rossi, C. Carmona, F. Adraga, G. Peltzer, K. Feigi, and T. Rabaute,
"The Displacement Field of the Landers Earthquake by SAR Interferometry," Nature, Vol.
364, 1993, pp. 138-142.
[52] Werner, C. L., P. Rosen, S. Hensley, E. Fielding, and E. Chapin, "Detection of Aseismic
Creep Along the San Andreas Fault Near Parkfield, California with ERS-I Radar Inter-
ferometry," Proc. Third ERS Scientific Symp., Florence, Italy, March 18-22, 1997.
[53] Gabriel, A. K., R. M. Goldstein, and H. A. Zebker, "Mapping Small Elevation Changes
over Large Areas: Differential Radar Interferometry," /. Geophys. Res., Vol. 94, 1989,
pp. 9183-9191.
A p p e n d i x 11A N u m b e r Fluctuations a n d t h e G e n e r a l i z e d
Product M o d e l
x(u) = (exp/(u.S R ))
N I f M M
= n ^ ' I " , % COs(^ + < ( , , - af)\) (11A.2)
(M Y (M Y
RI = |J>, % ««K -«,)) [S o, % K - a,)J + sin
MM
xW^fi/ofi! (HA.6)
i+{ (IIA 7)
(M)= 4 -
where y = — > 7——— — = , so
As TV —> °°,(j/) -> - (R2)/4 since i?2 stays finite, and
u<CRu = ( ( u . S R ) 2 )
M
cos 6
= £ (w/* ( ^ ~ ap+ (t))cos(0^ • a * + *))
where the last equality arises from averaging over the uniformly distributed
phase, (f). Hence,
2 1
(R ) = 2U CRU (11A.9)
and
SR = T G R (HA. 11)
x(u) = (exp(/Tu.G R ))
((u.G R ) 2 * + 1 ) = 0 (11A.14)
and
((u.GR)") = 1 A^,y
(2*)! k<\ 2 S
SO
Comparing (1IA. 10) and (1IA. 16), it is immediately clear that any model
based on number fluctuations of scatterers where the mean number of scatterers
is large has an equivalent representation as a product model. The form of the
number fluctuations only affects the texture variable X which must be chosen
to have the moments
12.1 Introduction
S = TG (12.1)
with T a positive real scalar for which (T2) = 1 and G a zero-mean Gaussian
variable specified completely by its complex covariance matrix.
In one dimension, S and G are simply complex scalars and only intensity
data need be considered because phase carries no information about distributed
targets. In this case, the product model has an alternative form given by
This form is useful because, for any distributed target, it decomposes the
data into an RCS term a, and a unit mean speckle term «, that is target-inde-
pendent. Hence, we can think of the data as a process cr affected by stationary
multiplicative noise, n. The RCS, a, is made up of a mean intensity term (since
(/) = (lGI2)) and a texture term T2. The local estimation of a formed the theme
of Chapters 6 and 7, while separation out of the texture was discussed in
Chapter 8.
It was shown in Section 11.3.2 that, in polarimetric data, there is no useful
way to make a similar separation into a stationary "noise" term and an "infor-
mation" term unless all the channels are uncorrelated. Instead, it is more fruitful
to consider estimation of the information-bearing quantities, of which a key
element is the covariance matrix characterizing G. The information carried by
the texture T is not defined a priori unless we assume some prior PDF for T.
Despite this change in perspective, we find that several of the filters developed
in Chapter 6 have polarimetric equivalents, as we now describe.
z =
c (kl^KI^H >Re(^i4)ylrn(zlzl)iRe(S1*;),
associated with any complex three-vector z = (S1, Z2, Z5)1. More generally, if there
are M channels, the corresponding vector requires M2 terms. (For M= 1, zc is
simply the channel power.) By the product model,
Sc = :PG C (12.4)
(c) (d)
Frequency Frequency
(e) (f)
Figure 12.2 (a) The HH-VV phase difference image corresponding to Figure 12.1; (b) the
modulus of the two-look HH-VV correlation coefficient; (c,d) the 16-look images
corresponding to (a,b); (e,f) observed phase difference distributions from a field
of rye for single-look and 16-look data, respectively.
scattering operator rather than from the scattering matrix. This is more general
than the MMSE filter described in [9] using single-look data. The treatment in
[9] is also much complicated because it deals with uncalibrated data. However,
calibration can be carried out in an optimized way as a separate step [10]. It is
also simpler than the filter derived in [11], which is based on a concept of
speckle that, as argued in Section 11.3.2, does not seem appropriate for po-
larimetric data.
A linear estimate of Gc from the data S is given by
G c = A + BS c (12.5)
Hence,
2
/\ Il
The MMSE filter chooses B to minimize \e), where e — Gc — Gc and
I • I is the Euclidean norm. Writing
where G' = G c — \GC/ and S' = Sc — \SC), and differentiating with respect to
the coefficients of B yields a minimum for \ej when
cs(i,j) = (sc(i)sc(j))-(SM[SAJ))
= (T*){CG(i,j) + (Gc(i))(Gc(j))) ~ (GM(GAJ)) (12-12)
where I9 is the 9 X 9 identity matrix. In this form, all the terms can be directly
estimated from the observed data. The mean vector \SC) = \GC) can be estimated
from a window surrounding the pixel of interest. In order to estimate (T"4), note
that
(^)SAJ)) _ ( r 0 (GAi)GAj))
[
(SM(SAJ)) '(GM(GAJ)) '
Since G is Gaussian, the normalized moment on the right-hand side of
(12.14) can be readily calculated using (11.2). It is most convenient to set i —
/ w i t h 1 < / < 3, whereupon it takes the value 1 + XIL for Z-look data. A
problem in using (12.14) was already encountered in Section 6.4.2: finite
sampling effects may cause the estimate of (T4) to be less than 1, despite the fact
that (T4) ^ 1 in the population. This can have nonphysical consequences, such
as leading to negative estimates of the powers in the channels. Since this
occurrence indicates an artifact of the sampling, (T4) should be constrained to
be the maximum of 1 and the value supplied by (12.14).
The one-dimensional version of the MMSE filter reduces to
P(T\S)*P(S\T)P(T)
where f denotes complex conjugate and tr denotes trace. The log likelihood is
then given by
where terms not involving Thave been omitted. Setting 8X/8 T= 0 leads to the
MAP solution for T as
This treatment can be cast into multilook form by noting that the expres-
sion for P{T\S) would be unchanged if one-look Stokes scattering operator data
were available, since P(TIS) = P(TISSt). For data in multilook Stokes form, we
instead have an Z-look average covariance matrix C at each point (SSf is the
one-look form of C). Then P(TlC) ex P (OT)P(T). As before, the actual
covariance matrix at the point is T2 C The multilook data C are known to obey
a complex Wishart distribution [8], given by
CL~Mexp(-Ltr(C-iC)/T2)
P(C T) = i i f- '- (12.19)
V
' K{L,M)T2ML\C\L
where K[L, M) = ifWM-vn T(L)- -T(L -M+ \)L~ML and P[C) refers to the
joint_PDF of Re(Cy) and Im(C1) for 1 < i, j < M. Forming K =
In(T^CI T)P(T)), assuming gamma-distributed T1 as before, and differentiating
with respect to 7"yields
w = Pz (12.21)
(/2)=I^4(WX) (12-24)
jklm
SO
(72)-trUC)2+(trUC))2 (12.26)
and
Note that any orthogonal transformation P gives the same mean and
variance as the original data z, since A is then just the identity matrix. Note also
that,foreach component of w, (l^-l4) — 2\\wp/. Hence, the intensity I wt\2 in
channel / always has normalized variance 2. This follows by setting P to be a
matrix containing a single nonzero row and is independent of texture.
The trace of a matrix is equal to the sum of the eigenvalues; and clearly, if
X is an eigenvalue of AQ then X2 is an eigenvalue of (AQ2. Hence,
</) = X \ - (12.28)
and
For both the Gaussian and product model, Vj is minimized when the ratio
XX?/(XA.,-) is minimum. Lagrange multiplier methods immediately show that
this occurs when all the X- are equal, so A = C~l (up to a scalar multiple). This
implies that the minimum value of normalized variance is given by
for the textured case, which reduces to just HMm the Gaussian case. Experi-
mental verification of this result is reported in [12,14,16].
From (12.21), the covariance matrix of the transformed data w is given by
so that Cw has the same eigenvalues as CA and the eigenvalues of Q, are related
to those of CA by the transformation P. In the particular case of optimum
speckle reduction, CA is the identity matrix and so the eigenvalues are all unity.
Also, from (12.33)
Cw = TM-1Pt = p(pt p) 1 Pt = IM (12.35)
(b)
Figure 12.4 (a) The output from full polarimetric filtering of the scene shown in Figure 12.1 and
(b) the corresponding output using intensity data only.
/ = trfc^c) (12.36)
where C is the multilook average covariance matrix extracted from the Stokes
scattering operator data. As shown in the discussion in the next section, practical
implementation of this filter produces speckle reduction consistent with theory.
The impact of the polarimetric whitening filter on target detection is discussed
Intensity
Pixel
Figure 12.5 Profile of intensity in the range direction through a trihedral corner reflector in an
original HH EMISAR image and after filtering using full polarimetric data or only
intensity data.
in [17] in which it was found that the local estimate of the covariance matrix
required by (12.36) had little effect on target detection and that a fixed value of
C could be used. However, in the dataset used, the correlation coefficients
showed little variation with cover type. More general investigation of this issue
is needed.
A oc Cy1U (12.37)
This last equality is the Siegert relationship (see (4.34) and (11.27)). The
constant of proportionality in (12.37) is arbitrary and can be chosen locally.
For Z-look textured data obeying the product model, the minimizing
vector is again given by (12.37) and the value of the minimum normalized
variance is
7
L (A'tr) 2 x
'
77 0.2932 0.4041
19 0.3383 0.4958
63 0.3187 0.4667
48 0.3314 0.4720
65 0.3116 0.4601
61 0.3689 0.4717
in Figure 12.5. Some resolution is lost, but this appears less marked than for the
full polarimetric filter.
While this treatment was concerned with finding the filter giving mini-
mum normalized variance, it is instructive to consider arbitrary weighted sums
of intensities in the special case where the like and crosspolarized channels are
uncorrelated. Setting (lS-12) = a- for 1 < i < 3 and (S1S3*) — ^a1Cr3 p, the
normalized variance of a weighted sum of intensities for untextured data is then
given by
e A
A 3 = - , , |2, (12.41)
y(eA2 + 1 - |p|2)
and is indicated in Figure 12.6 by the solid curves, separating two regions in
which Vj is reduced and a single region in which Vj is enhanced. Also marked
on the figure is the line along which Vj takes infinite values. The global
minimum has the value
v
**»-2xJ7tf (nA2)
speckle reduction
speckle
enhancement
Figure 12.6 Plot of regions in the (A2, A3) plane where linear filtering of polarimetric intensity
data with uncorrelated like and crosspolarized components gives rise to speckle
reduction or enhancement. Boundaries between these regions (where Vj — L~])
are marked as solid curves. The axes are normalized in terms of eA2 and yA3, and
the global minimum of normalized variance is marked in the first quadrant.
larized channels. In SAR data, the variance is proportional to the mean intensity,
so ordering principal components by variance causes the frequencies that carry
most power to have the larger eigenvalues and thus apparently carry more
information. Finally, the noise concepts applicable to optical data (with additive
noise) do not apply in any simple way to SAR data. A thorough analysis of the
PCT is given in the context of scattering matrix data in [18], where it is shown
that for azimuthally symmetric targets or scenes the PCT effectively attempts to
remove the phase difference between the two copolarized channels, so this
information is lost in the displayed image of each principal component. The
phase information is embedded in the eigenvectors. It is not clear, therefore, that
the PCT performs a useful function in terms of bringing out information.
Similar reservations about its value are raised in [19] when it is applied simply
to the channel powers. In particular, the great sensitivity to channel scaling is
demonstrated (see also [20]).
Despite these negative impressions of the utility of the PCT for SAR data,
it should be noted that it forms a central concept in [21], where the eigenvectors
of the covariance matrix are interpreted as representing different scattering
mechanisms with the eigenvalues corresponding to probabilities. A type of
nonorthogonal transform referred to as a generalized form of the PCT is also
developed in [22], based on the concept of speckle in polarimetric data as a
nonstationary noise. It appears to give degenerate results in the context of the
product model, and there has been no in-depth assessment of its value for image
analysis purposes.
When dealing with pairs of images of the same scene, produced at different
times or by different frequencies or polarizations, it is often necessary to describe
differences between the images. Such differences can be in the structure appar-
ent in the different images (see Figures 11.1 to 11.3) or in the parameters
describing a distributed target present in both images. Particular importance
attaches to images gathered at different times, so differences correspond to
change. In this section (and Section 12.5) we will develop the tools needed to
carry out this task, but defer examples of their use to Section 12.6.
and
var(y) = VOr(Z1) + var(Z2) - 2^JV^i(I1) W2i(l2) P1 (12.44)
where p 7 is the correlation coefficient of the intensity images at that point. For
uncorrelated image pairs, the variance is simply the sum of the variances of each
image pixel. As an example, for optical data with uncorrelated identically
distributed additive noise in the two images, the difference image would have a
mean giving the true difference but with double the noise variance of a single
image. For uncorrelated Z-look SAR intensity images, detailed descriptions of
the distribution and first two moments of d are given in [23,24]. For our
purposes it is sufficient to observe that if the true values of the RCS at a point
in the images are (T1 and (J2, respectively, then d is an unbiased estimator of (J1
— (J2 but
(using (4.10)). This means that simple thresholding of the difference image will
yield bigger errors for a given change in a bright area than a darker area. This is
a familiar problem, encountered in various guises when dealing with speckled
data.
To counter this problem, an alternative approach is to consider the ratio
image, Q = I2II1 (which is equivalent to the difference of the log of the
images). The ratio image has the advantages of being unaffected by multipli-
cative calibration errors (as long as they are the same for both images) and by
slope-induced effects on the backscattering coefficient (such effects are multi-
plicative and identical for images with the same imaging geometry [25]).
For image regions in which there is an underlying complex correlation
with value p between the two times, the distribution of Q is given by (11.54).
When the channels are uncorrelated this takes the simple form
^' T'(I.)(y+Qfl
where 7 = (J2I(Ji is the true change in RCS. This result also follows directly
from (7.6). For a given L and 7, the probability that the ratio exceeds a threshold
T is therefore the detection probability
P(Z)^f 0 I * j 2 Z - * - l [ 7J
A convenient reduction formula for computation is
_ pZ - i\ **(l - Jg)
7
ft ^
y = ^^r— ( 12 - 4 9)
A /=1
where k and / are pixel labels. This estimate has a PDF (a generalized form of
(12.46)) given by (7.6) with the replacements r ^ 7 , 7? -> 7, k-^ L2, and M
—> L1 + L2, with moments
00 Z1 <k
TL ^a
L
(a)
TL Pfa
(b)
T L ^ a
L
(C)
Figure 12.7 Plots of t h e threshold value, TL, needed for detection of a 3-dB change as a
d e t e c t i o n p r o b a b i l i t y : ( a ) P6 = 0 . 9 , ( b ) P6 = 0 . 9 5 , a n d ( c ) P6 = 0.99.
If L1 is small, 7 is significantly biased, since
l
( 7 ) = ml= 7 if A > (12-51)
l
A~
An unbiased estimate is therefore given by 7 c = L[1L1(L1 — 1)7; this estimate
has variance
2
var(7c) = ^ * / " V if A > 2 (12.52)
L2[L1 2)
PE(T) = R > T}
= r(A + L2) ^n2 - n (-i)^-*-1
k
T(L1)T(L2) £X JL1 +L.-k-i
X l + J^ r- (12.53)
Hence, the probabilities that 7 is at least a factor 8U greater than 7 c and S1 less
than 7 c are given by
and
Both of these are independent of 7, so that for chosen confidence values, 8U and
S1 can be determined.
Less accurate but more convenient estimates of the confidence interval for
7 can be made for large values of L1 and L2 by assuming that 7 c is normally
distributed. Then
where F(^) = ,— e t2l2 dt . The lower confidence value is found using a simi-
lar expression, but with (l — S~l) replaced by (8p* — l) in the argument of F(-).
These expressions can also be used to estimate the number of independent
looks required to measure a ratio to a specified level of precision. For example,
taking Lx-L1- L> estimating 7 to within 3 dB with 95% confidence involves
setting 8U = 2, S1 = 0.5, and both probabilities to 0.05. The argument of F in
both cases will then be 1.645, yielding L ^ 23 to put an upper limit on 7 or L
> 7 for the lower limit. If the precision is increased to 1 dB, the required
numbers of independent looks increase to 129 and 82, respectively.
The number of independent pixels required based on the Gaussian ap-
proximation is in fact bigger than is indicated by an exact calculation based
on (12.53). This is illustrated by Figure 12.8, which shows the exact and
approximate probabilities that 7 exceeds 8U7C, based on (12.54a) and (12.55).
In both cases, 7 = 2. Figure 12.8(a) shows the two probabilities as a function
of Z for 8U = 1.259 (1 dB), while Figure 12.8(b) gives the same information
for 8U = 1.995 (3 dB). From Figure 12.8(b) it can be seen that, in fact, only
11 looks are needed for a 3-dB 95% upper confidence limit, not the 23
inferred from the Gaussian approximation. Similarly, Figure 12.8(a) shows
that 102 looks are required for a 1-dB confidence limit, not 129.
Note that these methods allow a quantitative answer to the question of
whether two regions are different: if 1 is not within the confidence interval for
7, then they are statistically different. Equivalently, this question can be ad-
dressed using the normalized ratio
rT - m i n f i ^ 1 ] (12.56)
which deals only with the size of a change, not whether it is an increase or
decrease (see Section 7.3.3 and [23,26]).
The estimates of the intensity ratio discussed in Section 12.4.2 relied on a prior
segmentation in order to define homogeneous areas in two images (or different
areas in the same image). That segmentation can be based on:
• Visual inspection;
• Transfer of a known segmentation from elsewhere (e.g., using a regis-
tered field map);
• Using the data themselves to learn the segmentation.
Previous Page
where F(^) = ,— e t2l2 dt . The lower confidence value is found using a simi-
lar expression, but with (l — S~l) replaced by (8p* — l) in the argument of F(-).
These expressions can also be used to estimate the number of independent
looks required to measure a ratio to a specified level of precision. For example,
taking Lx-L1- L> estimating 7 to within 3 dB with 95% confidence involves
setting 8U = 2, S1 = 0.5, and both probabilities to 0.05. The argument of F in
both cases will then be 1.645, yielding L ^ 23 to put an upper limit on 7 or L
> 7 for the lower limit. If the precision is increased to 1 dB, the required
numbers of independent looks increase to 129 and 82, respectively.
The number of independent pixels required based on the Gaussian ap-
proximation is in fact bigger than is indicated by an exact calculation based
on (12.53). This is illustrated by Figure 12.8, which shows the exact and
approximate probabilities that 7 exceeds 8U7C, based on (12.54a) and (12.55).
In both cases, 7 = 2. Figure 12.8(a) shows the two probabilities as a function
of Z for 8U = 1.259 (1 dB), while Figure 12.8(b) gives the same information
for 8U = 1.995 (3 dB). From Figure 12.8(b) it can be seen that, in fact, only
11 looks are needed for a 3-dB 95% upper confidence limit, not the 23
inferred from the Gaussian approximation. Similarly, Figure 12.8(a) shows
that 102 looks are required for a 1-dB confidence limit, not 129.
Note that these methods allow a quantitative answer to the question of
whether two regions are different: if 1 is not within the confidence interval for
7, then they are statistically different. Equivalently, this question can be ad-
dressed using the normalized ratio
rT - m i n f i ^ 1 ] (12.56)
which deals only with the size of a change, not whether it is an increase or
decrease (see Section 7.3.3 and [23,26]).
The estimates of the intensity ratio discussed in Section 12.4.2 relied on a prior
segmentation in order to define homogeneous areas in two images (or different
areas in the same image). That segmentation can be based on:
• Visual inspection;
• Transfer of a known segmentation from elsewhere (e.g., using a regis-
tered field map);
• Using the data themselves to learn the segmentation.
Note that, in general, all three approaches will produce different answers
and problems. A visual inspection can input much greater world knowledge into
the process, to the extent of ignoring data that conflict with what is expected.
It, however, performs differently in bright and dark areas, despite similar statis-
tical conditions. Transferring a segmentation carries a number of problems: map
data can often be erroneous; different structures on a map may produce similar
responses in the image; and registration and matching of geometries are essen-
tial. Using the data to learn the image structure relies on an image model, which
probability
(a)
probability
number of looks
(b)
Figure 12.8 Approximate and exact probabilities that the true intensity ratio exceeds the
observed unbiased ratio by a factor 8U, as a function of the number of looks (a) 8U
= 1dBand(b)8u = 3dB.
must be complete enough and realistic enough to produce meaningful segmen-
tations that conform with known properties of typical scenes. Resistance to
speckle is an important concern in such methods.
Methods for learning image structure were discussed in Chapter 7 for
single-channel data. However, a glance at Figures 11.1 to 11.3 makes clear that
increased information on scene structure is provided by combining images
acquired with different frequencies, polarizations, or at different times. In this
section we discuss how the methods developed in Chapter 7 can be extended to
multidimensional data.
When multiple channels are available, an immediate concern is the defi-
nition of what is meant by a segmentation and the parameters on which it is to
be based. In the context of single-frequency polarimetric data, the simplest data
model is that segments are statistically homogeneous and completely charac-
terized by a covariance matrix. This is the polarimetric version of the cartoon
model developed in Section 7.3. Two segments are distinct if any of the nine
parameters (three channel powers and three amplitudes and phases of the
correlation coefficients) differs significantly between them. Hence, in principle,
segmentation of polarimetric data should be based on estimates of all nine
parameters. In practice, the sampling properties of the correlation coefficient
terms render them less useful than the channel powers when making decisions
about edges and merging of regions. For example, there are no easily applicable
rigorous tests for when the phases or amplitudes of the correlation coefficients
of two regions are significantly different. We, therefore, consider only algo-
rithms based on the powers in the three channels. MuItitemporal and polarimet-
ric segmentation can, therefore, be treated similarly.
(c) (d)
(e) (f)
(g) (h)
Figure 12.9 (a) L-band HH intensity image acquired by the NASA/JPL AirSAR system;
(b) single-channel segmentation using the RGW algorithm; (c,d) same as for (a,b)
but for the HV channel; (e/f) same as for (a,b) but for the VV channel; (g) segmen-
tation formed by retaining any edge found in (b), (d), or (f); and (h) multidimen-
sional segmentation using modified RGW.
The alternative is to treat the multichannel image as a single entity and
make decisions on the existence and position of an edge using information from
all channels simultaneously. In order to do this, both the edge-detection and
region-growing components of RGW have been modified [29,30]. As described
in Section 7.3.1, for edge detection RGW uses a threshold applied to the
normalized gradient calculated from the two halves of a series of windows of
various sizes. In the multichannel case, the threshold is instead applied to the
rms value of the normalized gradient in all channels. This produces a single edge
map that can be used to control segment growing. Similarly, merging disks uses
an rms measure of the difference between the mean values of disks in each
channel. The average contrast in each channel is calculated after each iteration,
and the algorithm stops when there is an increase in the rms average contrast.
Figure 12.9(h) shows the result of applying this process (again indicating the
mean H H intensity in each segment). Note the clear visual improvement in edge
structure as compared with Figure 12.9(g). Notice also possible loss of informa-
tion; for example, an edge detected only in the HV image in the top right of the
image is lost in the multidimensional segmentation.
The edge map generated by this procedure can be used to test whether a
given segment boundary occurs in all or only some of the channels. For each
channel, the probability that the test edge is due purely to speckle can be calcu-
lated. Those edges for which this probability exceeds a given threshold are thus
ascribed to speckle and considered not to be present in that channel. The result of
applying this operation to each of the intensity images in Figure 12.9(a,c,e) is
shown as Figure 12.10(a—c). Clearly this provides a means of interrogating the
data to establish the contribution to overall structure provided by each channel.
This is most easily illustrated using color overlays (so that edges present in
different combinations of channels can be identified). However, in Figure 12.10
we instead show (a) the edges present in H H alone; (b) the edges in both H H and
W (shown in black), in H H but not W (dark gray), and in W but not H H
(light gray); (c) the edges present in H H , W , and HV (black), in H H or W but
not HV (dark gray), and in HV but not in either H H or W (light gray). We can
see that much of the structure is common to all channels but that there are
significant differences and each channel adds information about the overall field
structure. In addition, use of multichannel segmentation to improve the estimates
of polarimetric or texture parameters is discussed in Section 11.9.3 and [29,30].
(b)
to
Figure 12.10 (a) Edges in the multidimensional segmentation shown in Figure 12.9(h) that are
present in the HH channel, (b) Edges present in both the HH and VV channels
(black), HH but not VV (dark gray), and VV but not HH (light gray), (c) Edges
present in HH, VV, and HV (black); HH or VV but not HV (dark gray), and HV but
not HH or VV (light gray).
may not be strong enough to put the rms measure above threshold. However,
this is by design, otherwise problems of edge jitter would not be eased by
combining channels. The price paid for an image that is easier to interpret is
therefore a potential loss of information that is carried in only one or a small
number of the available channels. A more rigorous approach when the images
are uncorrelated is to extend the annealed segmentation algorithm described in
Section 7.3.5. This is possible because, if the boundaries in the different images
of the scene are expected to lie in the same position, the log likelihoods for each
image for given boundaries can be summed to form the total log likelihood. The
optimization can then proceed as for single-channel data. However, as yet there
is no rigorous criterion for the false alarm probability, so a postprocessing merge
stage to yield a defined P^ is not possible.
The three ERS-I PRI images in Figure 12.11 (a—c) were gathered on April
16, June 9, and August 18, 1992, respectively, from the Feltwell area, UK. The
same display conditions are adopted for each, revealing the markedly different
multitemporal behavior of the backscattering coefficient in different fields during
the year. Crop development and soil moisture changes give rise to images that
have very different structures (see Section 13.2 for a further discussion of this
point). Field definition appears better in the two earlier images due to significant
changes in crop state over the growing period. The August image exhibits gener-
ally low contrast, probably reflecting the fact that many of the fields have already
been harvested by this date so that the image is dominated by soil effects.
Each individual image was first segmented using annealed segmentation
with an average region size chosen to be 50 pixels, corresponding to 1,310
segments [31]. The average intensities within each segment for the three indi-
vidual images are shown in Figure 12.1 l(d—f). Figure 12.11 (d,e) yields reason-
able evidence of field boundaries, whereas the segments in (f) are more
fragmented. The corresponding edge maps are shown in Figure 12.11(g-i).
They seem to have little detailed structure in common except for the drainage
channel on the left of the image, parts of which are visible in the different scenes.
In contrast, the edge map from multidimensional annealed segmentation shown
as Figure 12.11 (j) appears distinctly superior to any of the individual cases
shown as (g-i). The structure of the fields and the river appears much cleaner,
more complete, and better defined in this composite result.
The average intensities within the multidimensional segments for each of
the original images are shown in Figure 12.12(a-c). If Figure 12.12(a) is com-
pared with (b), the very different behavior of the backscattering coefficient of
different fields is dramatically evident. Moreover, we can use the inferred struc-
ture to examine the more subtle differences present in the August image of
Figure 12.12(c). The statistical significance of these spatial and temporal
changes can then be assessed using the analysis developed in Section 12.4. The
(a) (d)
(b) (e)
(C)
(f)
Figure 12.11 (a-c) ERS-1 PRI images of an agricultural area near FeItWeII7 UK, gathered on (a)
April 16, (b) June 9, and (c) August 18, 1992 (Copyright ESA). (d-f) The corre-
sponding one-dimensional annealed segmentations; (g-i) the associated edge
maps; and (j) the edge map resulting from multidimensional annealed segmen-
tation. (British Crown Copyright, 1997/DERA.)
(i)
(9)
(h) (J)
Figure 12.11 (continued).
importance of such tools for monitoring large agricultural areas is made much
clearer in Section 13.2, where we discuss some of the empirical background
relevant to analyzing the multitemporal behavior of a° for crops.
Before closing this section, it should be remarked that in addition to the
currently inadequate treatment of false alarm probabilities noted previously, another
shortcoming is that multidimensional annealed segmentation provides a rigorous
approach only when the channels are independent. Further development is needed
to provide optimal methods when channels are correlated, as in polarimetric data.
(C)
Figure 12.12 Mean intensities within the segments indicated by Figure 12.11(j) for the individ-
ual images shown as Figure 12.11(a—c). (British Crown Copyright, 1997/DERA.)
(b)
Figure 12.13 ERS-I PRI images of a region bordering the Tapajos National forest on (a) July
31 and (b) December 18,1992. (Copyright ESA.)
(a)
(b)
Figure 12.14 (a) Ratio of the images shown as Figure 12.13(a,b) and (b) thresholding of the
ratio image at 1.5 dB.
(c) (d)
Figure 12.15 (a,b) Images formed by 18 X 18 block averaging of the images shown in Figure
12.13; (c) ratio of the images shown in (a,b); and (d) histogram of values in the
ratio image, truncated at - 3 and 4.99 dB.
(C) (d)
(e) (f)
Figure 12.16 Images and histograms corresponding to Figure 12.15(c,d)for postprocessing by:
(a,b) gamma MAP filtering, (c,d) annealed despeckling, and (e,f) multidimen-
sional RGW segmentation.
ing less smoothing. The global MAP result (implemented by annealed despeck-
ling) is shown in Figure 12.l6(c,d), with a comparatively narrow histogram.
Finally, the effect of performing a multidimensional segmentation and ratioing
mean values within segments is shown in Figure 12.l6(e,f). Each spike in the
histogram indicates the ratio values within one or a small number of segments.
To carry out a classification requires a threshold to be chosen, but the lack
of structure in the histograms means that this choice is not clear-cut. Classifica-
tions based on a 1.5-dB threshold are shown as Figure 12.17(a-d) for the four
methods employed, together with the results for a ratio image that has not been
preprocessed (Figure 12.17(e)). Based on the theory in Section 12.4.1, this
threshold corresponds to a false alarm rate of 4.74 X 10~5 in Figure 12.17(a)
when we make allowance for the correlation of samples in the ERS-I data. This
correlation means that the 18 by 18 3-look averages are equivalent to approxi-
mately 257 looks (see Section 4.8, especially the discussion following (4.43)).
The preprocessed images all pick out the main structural blocks of the
data. The blurring produced by the large averaging window is clear in Figure
12.17(a). Annealing preserves the edge structures much better, while segmenta-
tion displays an intermediate behavior. Very noticeable in the gamma-MAP
filtered data is the noisiness of the image, with many random false detections in
the forest areas. By contrast, the other three algorithms produce little random
noise; detections in the forest areas tend to occur as coherent regions.
Also shown as Figure 12.17(f) is a forest/nonforest classification based on
a threshold applied to band 5 of the TM data. Under the assumption that this
is a reliable classification (this is discussed in detail in [4]), we can carry out a
quantitative comparison of the various methods. Table 12.2 shows the classifi-
cation accuracies for the different methods. The first and second columns
indicate the proportions of correctly detected nonforest and forest pixels, respec-
tively. The third column indicates the area of false nonforest detections as a
proportion of the true nonforest area, which in this image forms 22.3% of the
total area.
As expected, the table indicates that ERS-I is not ideal for forest dis-
crimination, with less than 50% of the nonforest areas being correctly de-
tected. Missed detections are clear when comparing Figure 12.17(f) with any
of Figure 12.17(a—d). For example, the nonforest area protruding from the
right side of the square pasture region is not picked out in the radar image
by any of the methods. This reflects the fact that even low vegetation canopies
can have a backscattering coefficient similar to that of primary forest at C-
band. Longer wavelengths are more appropriate to this application, such as
the L-band data provided by JERS-I; a comparison of ERS-I and JERS data
for Tapajos will be found in [4]. Two other comments are relevant here. First,
a more complete temporal coverage of the region may have given better per-
(a) (b)
(c) (d)
(e) (f)
Figure 12.17 (a-d) One-bit images formed by thresholding the images shown as Figures
12.15(c), 12.16(a), 12.16(c), and 12.16(e), respectively; (e) the corresponding
image using the ratio of the original (unpreprocessed) ERS-1 images; and (f) a
forest/nonforest template prepared by thresholding band 5 of a TM image ac-
quired on July 29,1992.
Table 12.2
Classification Accuracies of the Preprocessing Methods
formance. Long time sequences of ERS images gathered over temperate forests,
which also show very stable behavior, allow good forest discrimination using
change-detection techniques. We do not have enough images from Tapajos
to test this. Second, C-band systems are capable of good forest discrimination
in the tropics if they have sufficient resolution (preferably combined with
not too steep an incidence angle). This has been demonstrated in Chapter 8
for airborne data and will also be illustrated using spaceborne data from Ra-
darsat in Chapter 13.
Nonetheless, Table 12.2 does provide a means by which to compare the
performance of the different algorithms. The removal of small areas of change
and the poor definition of edges cause simple averaging to give the lowest
percentage of correct nonforest detections. The highest detection rate is that
which uses the gamma MAP filter, but at the expense of significantly increased
false detections (18%) that also have a highly random noiselike quality, as noted
previously. Overall, annealed despeckling gives the best performance, while
segmentation performs a little worse than annealing, with a slightly lower
detection rate and more false detections.
These results illustrate once more the importance of algorithm selection
discussed in Sections 6.10 and 7.4. In this application, annealing appears to
provide the best performance of the algorithms tested. However, it requires
considerably more computing power than the simplest method based on block
averaging. For large-area estimates, or an initial assessment of overall pattern,
this suggests that averaging may provide the most useful approach. However,
because large windows must be used for averaging to give adequate performance,
this is at the price of lost boundary detail, failure to detect small areas, and
underestimation of nonforest areas. When a higher level of performance is
required, a more sophisticated and powerful method such as annealing must be
used, possibly in a selective manner to regions identified to be of interest by
averaging.
12.7 Concluding Remarks
This chapter has covered a wide range of techniques for extracting information
from multidimensional SAR images. It is very much based on foundations built
in earlier chapters, extending techniques developed for single-channel data to
this more general and demanding environment. Other aspects of multichannel
filtering for which there are no useful one-dimensional analogies (e.g., po-
larimetric matched filtering) are discussed in the next chapter, since these are
closer to classification techniques.
What should be clear here and from Chapter 11 is that the development
in this area is far from complete. Among a variety of areas of weakness, we note
that there has been comparatively little testing of algorithms, objective methods
for algorithm performance are not established, and there are known limitations
in the data model used by the segmentation algorithms in the context of
polarimetric data. In a more general context, the state of development of these
algorithms is such that large-scale assessment of their value to selected applica-
tions is now timely but has not yet been carried out. Section 12.6 indicated an
approach to this problem. In doing so, it raises the difficult issue of separating
the possible contribution of radar images to a given application (what informa-
tion is potentially available?) from the information provided after postprocessing
(has an algorithm destroyed or misrepresented the available information?). This
is perhaps the central issue in methods for image understanding because it
requires us to ensure that our data models are true representations of informa-
tion sources in radar data and that the algorithms correctly embody these
models. There is little doubt that on both counts our current level of under-
standing is not complete.
References
[1] Quegan, S., A. Hendry, and J. Wood, "The Visibility of Linear Features in SAR Images,"
Proc. IGARSS 88, Edinburgh, 1988, pp. 1517-1520.
[2] Greeley, R., and D. G. Blumberg, "Preliminary Analysis of Shuttle Radar Laboratory
(SRL-I) Data to Study Aeolian Features and Processes," IEEE Trans. Geosci. Remote Sensing,
Vol. 33, 1995, pp. 927-933.
[3] Quegan, S., R. G. Caves, K. D. Grover, and R. G. White, "Segmentation and Change
Detection in ERS-I Images over East Anglia," Proc. of the First ERS-I Symp., Cannes, ESA
SP-359, 1993, pp. 617-622.
[4] Grover, K. D., S. Quegan, and C. C. F. Yanasse, "Quantitative Estimation of Tropical Forest
Cover by SAR," IEEE Trans. Geosci. Remote Sensing, in press 1997.
[5] Borgeaud, M., J. Noll, and A. Bellini, "Multitemporal Comparisons of ERS-I and JERS-I
SAR Data for Land Applications," Proc. IGARSS 94, Pasadena, 1994, pp. 1603-1605.
[6] Schmullius, C , J. Nithack, and M. Kern, "Comparison of Multitemporal ERS-I and
Airborne DLR E-SAR Image Data for Crop Monitoring," Proc. Second ERS-I Symp.,
Hamburg, ESA SP-361, 1994, pp. 79-84.
[7] Wooding, M. G., A. D. Zmuda, and G. H. Griffiths, "Crop Discrimination Using
Multitemporal ERS-I SAR Data," Proc. Second ERS-I Symp., Hamburg, ESA SP-361,
1994, pp. 51-56.
[8] Goodman, N. R., "Statistical Analysis Based on a Certain Multivariate Gaussian Distribu-
tion (an Introduction)," Ann. Math. Stat., No. 34, 1963, pp. 152-177.
[9] Goze, S., and A. Lopes, "A MMSE Speckle Filter for Full Resolution SAR Polarimetric
Data,"/ Elect. Waves and Appns., No. 7, 1993, pp. 717-737.
[10] Quegan, S., "A Unified Algorithm for Phase and Crosstalk Calibration—Theory and
Observations," IEEE Trans. Geosci. Remote Sensing, Vol. 32, 1994, pp. 89-99.
[11] Touzi, R., and A. Lopes, "The Principle of Speckle Filtering in Polarimetric SAR Imagery,"
IEEE Trans. Geosci. Remote Sensing, Vol. 32, 1994, pp. 1110-1114.
[12] Schou, J., "Validation of Filters and Statistical Models," Short Course report, University of
Sheffield, 1996.
[13] Lopes, A., E. Nezry, R. Touzi, and H. Laur, "Structure Detection and Adaptive Speckle
Filtering in SAR Images," Int. J. Remote Sensing, Vol. 14, 1993, pp. 1735-1758.
[14] Novak, L. M., and M. C. Burl, "Optimal Speckle Reduction in Polarimetric SAR Imagery,"
IEEE Trans. Aerospace Elect. Sys., Vol. 26, 1990, pp. 293-305.
[15] Bruniquel, J., and A. Lopes, "Multi-variate Optimal Speckle Reduction in SAR Imagery,"
Int. J. Remote Sensing, Vol. 18, 1997, pp. 603-627.
[16] Lee, J., M. R. Grunes, and S. A. Mango, "Speckle Reduction in Multipolarisation, Multi-
frequency SAR Imagery," IEEE Trans. Geosci. Remote Sensing, Vol. 29, 1991, pp. 535-544.
[17] Novak, L. M., M. C. Burl, and W W. Irving, "Optimal Polarimetric Processing for
Enhanced Target Detection," IEEE Trans. Aerospace Elect. Sys., Vol. 29, 1993, pp. 234-244.
[18] Quegan, S., and L. V. Dutra, "SAR Calibration and Principal Component Analysis," Proc.
ThirdAirborne SAR(AIRSAR) Workshop, Jet Propulsion Lab, Pasadena, 1991, pp. 138-146.
[19] Quegan, S., "Application of Principal Components Analysis to Quadpolarised Data,"
GEC-Marconi Research Centre: Studies of Multi-frequency and Multi-polarisation SAR Appli-
cations, Appendix E, MTR 93/73, Contract RAE IB/89, 1993.
[20] Singh, A., and A. Harrison, "Standardised Principal Components," Int. J. Remote Sensing,
Vol. 6, 1985, pp. 883-896.
[21] Cloude, S. R., and E. Pottier, "An Entropy Based Classification Scheme for Land Applica-
tions of Polarimetric SAR," IEEE Trans. Geosci. Remote Sensing, Vol. 35, 1997, pp. 68-78.
[22] Lee, J.-S., and K. W. Hoppel, "Principal Components Transformation of Multifrequency
Polarimetric SAR Imagery," IEEE Trans. Geosci. Remote Sensing, Vol. 30, 1992,
pp. 686-696.
[23] Touzi, R., A. Lopes, and P. Bousquet, "A Statistical and Geometrical Edge Detector for SAR
Images," IEEE Trans. Geosci. Remote Sensing, Vol. 26, 1988, pp. 764-773.
[24] Rignot, E. J. M., and J. J. van ZyI, "Change Detection Techniques for ERS-I SAR Data,"
IEEE Trans. Geosci. Remote Sensing, Vol. 31, 1993, pp. 896-906.
[25] Schreier, G., "Geometrical Properties of SAR Images," SAR Geocoding: Data and Systems,
G. Schreier (ed.), Karlsruhe: Wichmann, 1993, pp. 103-134.
[26] Caves, R. G., "Automatic Matching of Features in Synthetic Aperture Radar Data to Digital
Map Data," Ph.D. Thesis, University of Sheffield, UK, 1993.
[27] White, R. G., Low Level Segmentation of Noisy Lmagery, Tech. Report 3900, Roy. Signals
and Radar Estab., Malvern, 1986.
[28] White, R. G., "Change Detection in SAR Imagery," Int. J. Remote Sensing Vol. 12, 1991,
pp. 339-360.
[29] Caves, R. G., and S. Quegan, "The Role of Segmentation in Multi-channel SAR Image
Analysis," Proc. RSS '95: Remote Sensing in Action, Southampton, Remote Sensing Society,
1995, pp. 1171-1178.
[30] Caves, R. G., and S. Quegan, "Multi-channel SAR Segmentation: Algorithm and Applica-
tions," Proc. European Symp. on Satellite Remote Sensing IL, Paris, 1995, pp. 241—251.
[31] McConnell, L, and D. Stewart, "Multi-dimensional Annealed Segmentation," N.A. Soft-
ware Rpt. No. 10, DRA Contract No. CSM1/072, 1997.
[32] Quegan, S., K. D. Grover, and C. C. E Yanasse, "Use of C Band SAR for Quantitative
Estimation of Tropical Forest Cover—Methods and Limitations," Proc. Lnternational Symp.
on Retrieval of Bio- and Geophysical Parameters from SAR Data for Land Applications,
Toulouse, 1996, pp. 167-178.
[33] Sant'Anna, S. J., C. C. F. Yanasse, and A. C. Frery, "Estudo Comparativo de Alguns
Classificadores Utilizando-se Imagens Radarsat da Regiao de Tapajos," Proc. First Latino-
American Seminar on Radar Remote Sensing: Lmage Processing Techniques, ESA SP-407,
1997, pp. 187-194.
[34] Le Toan, T, A. Beaudoin, J. Riom, and D. Guyon, "Relating Forest Biomass to SAR Data,"
IEEE Trans. Geosci. Remote Sensing, Vol. 30, 1992, pp. 403-411.
[35] Dobson, M. C , F. T. Ulaby, T. Le Toan, E. S. Kasischke, and N. L. Christensen,
"Dependence of Radar Backscatter on Coniferous Forest Biomass," IEEE Trans. Geosci.
Remote Sensing, Vol. 30, 1992, pp. 412-415.
[36] Lecomte, P, and E. P. Attema, "Calibration and Validation of the ERS-I Wind Scatterome-
ter," Proc. of the First ERS-I Symp., Cannes, ESA SP-359, 1993, pp. 19-29.
[37] Le Toan, T, F. Ribbes, N. Floury, and U. R. Wasrin, "Use of ERS-I SAR Data for Forest
Monitoring in South Sumatra," Final Report, TREES Project, 1997.
1 3
Classification of S A R Imagery
13.1 Introduction
1. The two crop types only become separable in this case during the period
from June to August, with a peak separation between signatures of
around 5 to 7 dB. This corresponds to the period between sugarbeet
Winter wheat
10 fields + average + stdev
[dB]
Go
Jan Feb Mar Apr May Jun JuI Aug Sep Oct Nov Dec
1993
, * Flevoland 1993
Sugarbeet
10 fields + average + stdev
a0 [dB]
Jan Feb Mar Apr May Jun JuI Aug Sep Oct Nov Dec
1993
<b) Flevoland1993
Figure 13.1 Time series of field-averaged a° values measured by ERS-1 in Flevoland, the
Netherlands, in 1993: (a) winter wheat and (b) sugarbeet [26].
attaining full coverage and wheat being harvested. In the early part of
the year the signatures are practically the same in both cases, dominated
by soil effects. For sugarbeet, the comparatively rapid development of
full ground cover is marked by a rapid rise in or°, which then remains ef-
fectively constant throughout the rest of the year until harvesting. This
behavior is indicative of volume scattering in the canopy, which rapidly
saturates as the canopy density increases. For wheat, plant development
gives rise to a much wider trough in cro, which can be explained, given
the low-volume fraction occupied by the vegetation, in terms of increas-
ing attenuation of the soil response as the vegetation grows. Values be-
come comparable to those for sugarbeet around September, when the
behavior of both crops is again soil-dominated.
2. Within each crop type there is considerable variation in a° at a fixed
time, perhaps due to slight variations in plant growth stage and (in the
case of wheat) different soil roughness states. For our sample there is
no overlap in the signatures during the critical period.
This example indicates that for these two crop types (more generally, for
large-leafed root crops and cereals) there is reasonable expectation of good
classification performance, as long as the data acquisitions occur at the right
time and enough pixels can be averaged to give sufficient statistical separability.
Note that only data for 1993 are shown here. Other years show similar behavior,
although with time shifts possibly related to weather conditions, location, or
sowing times [26].
When discrimination of crop types with similar structure is desired, the
empirical evidence is less encouraging. Figure 13.2 shows a comparison of time
series for winter wheat and barley. It is obvious that the behavior of the two crop
types is closely related. The best conditions for separating them occur during
the reproductive stage when the difference in the two plant structures results in
different attenuation of the soil response at W polarization. Since this differ-
ence is small, viable classification relies on large fields and accurate calibration
(in order to minimize the uncertainty in cr°). A further important consideration
relevant to many areas (but not Flevoland) is correction for topographic effects.
Slopes of even a few degrees can induce apparent a° changes that are significant
in relation to the small dynamic range covering different crops. While this
macroscale variation in slope is a disturbing factor, the sensitivity of SAR to
microscale structure can be turned to an advantage. In particular, its ability to
detect changes and differences in soil roughness can provide information on
early season soil preparation, from which a broad crop type may be inferred
[29]. Similar inferences can be made from interferometric data [27].
o 0 [dB]
Figure 13.2 Time series of mean field-averaged a° values for winter barley and winter wheat
for the 1993 growing season: (a) Netherlands data and (b) Feltwell, UK, data [26].
bare
ly confier grass potatoes sugarbeet wheat
Figure 13.3 Plots of the HH, HV, and VV backscattering coefficients and the amplitude (circles)
and phase (crosses) of the HH-W correlation coefficient for C-band AirSAR data
gathered over Feltwell, UK.
L Band HH Power (dB)
bare
ly confier grass potatoes sugarbeet wheat
Figure 13.4 As Figure 13.3 but for L-band.
5. There is useful information in the correlation coefficient. At C-band,
phase differences tend to be small, but Figure 11.1 shows that much
of the field pattern can be seen. Conifers and root crops exhibit a stable
phase difference close to 0 degrees (except for one anomaly); whereas
the cereals display a noisy negative phase, around — 10 degrees on
average. Such differences are probably connected with the dominant
vertical orientation of the stalks in the cereal canopies, while the other
vegetation types have no such preferred orientation in their constitu-
ents. Root crops tend to show higher coherence, and in our dataset can
be separated from other vegetation types on this basis (except for one
exceptional sugarbeet field). At L-band the reduced value of coherence
and positive phase difference of conifers are notable, but the variability
within the other vegetation types weakens the value of this trend as a
classifier. (It should be noted, however, that a large positive phase
difference is a very clear indicator of conifer forest at P-band for this
test site [31].) Unlike C-band, the cereals and root crops have similar
average values of coherence, but this is very variable (extremely so for
barley). The correlation phase of the cereals is also very variable but
more stable for the root crops.
where the prior probability of class n occurring is given by P(n) and P(A\n) is
the likelihood that A occurs, given that the measurement is from class n. The
MAP classifier selects that n for which P{n\A) is maximal. This is equivalent to
minimizing (lnP(A\n) + In P(n))> which is often a more convenient expression
to handle. In order to use MAP classification, both the likelihood P(A\n) and
the prior probabilities P(n) must be available. If the P(n) are unknown, they are
often assumed equal for all classes, in which case this approach reduces to ML
classification.
For multichannel SAR data, we have seen that there are three basic models
for P(AI n) in which the underlying complex single-look data:
Only for jointly Gaussian data has the classification theory been well-
developed (although an approach for single-channel textured SAR data has
been proposed in [47]) and we concentrate on this case.
U
P(An) = — ^ \" (13.2)
L
K{L,M)\Cn\
is minimized. (Note that this measure can be negative since we omitted terms
not involving n in In P(AIn)). The expression (13.3) is clearly linear in the
variables Ke(A^ and Im(Ay), which implies that the decision boundaries (that
is, where dn = d^) are hyperplanes in M2 dimensions. The one-dimensional
version of (13.3) is given as (7.1) for gamma distributed RCS values. This
makes clear that the term In(I CJ) corresponds to a measure of the overall spread
of values within the population of class n.
Although representing an optimal classifier, (13.3) relies for its success on
having known values for P(n) and Cn and the fact that Cn is uniquely defined for
class n. In practice, neither condition may hold. Typically P(n) and Cn have to be
inferred from the data, in the latter case either by use of training sets (supervised
classification) or from clusters inferred from the data (unsupervised classifica-
tion). Iterative learning of the Cn is possible and gives performance that has been
reported to improve on that of supervised classification in a limited study of sea
ice classification [49]. The prior probabilities P(n) present more of a problem
because they are rarely known a priori. However, they can have a significant
impact on pixel-based classification. This is demonstrated in [50], where it is
shown that adaptive learning of the prior probabilities can give markedly im-
proved performance. Notice that if the classes of interest are homogeneous
enough to be well represented by a single known covariance matrix, then (13.3)
implies that the prior probabilities P(n) become less important as L increases.
(Equivalently, as L increases, the MAP and MLE classifiers become equivalent.)
Hence, an alternative to learning the prior probabilities is instead to minimize
their impact by making L large. However, for large Z, mixed populations will
often occur within an arbitrary classification window, which indicates that the full
power of the classification is best attained within a segmentation where homoge-
neous regions are first identified, as described in Chapters 7, 8, and 12.
A bigger obstacle to applying (13.3) in practical applications is the vari-
ability of the covariance matrix describing a single desired class. This intrinsic
variability (rather than variability as a consequence of estimating the class
covariance) is obvious from Figures 13.1, 13.3, and 13.4. In some cases this does
not represent a fundamental difference in covariance structure and can be dealt
with using a normalization by the power in one of the channels [51]. This
cannot, however, remove the effects of the highly fluctuating phase differences
observed for some cover types in Figure 13.4(b), for example. Where a given
class n must be described by a distribution of covariance matrices P(Q), the
likelihood function of class n takes the form
The forms of the classifiers for some of these special cases are now
discussed.
(T1 0 V°"ia3P
C= 0 (J2 0 (13.5)
V°W* 0 (J3 ^
For notational simplicity the subscript n appropriate to class n has been omitted
from the parameters (J1 and p. Then
and
When the channels are uncorrelated, the covariance matrix of class n is diagonal
Cn = d i a g [ a l B , (T2n,... ,dMn]
with inverse
C;1 =diag[(r;1,cr;);,...)(r^]
Then
As expected, the decisions are based just on the sample intensities because
any observed correlations between channels are simply due to finite sampling
effects. In such circumstances, we can instead consider only the measured
intensities in our data vector. The joint PDF, therefore, is a product of inde-
pendent Z-look gamma distributions, and the distance measure between classes
is exactly the same as that given by (13.3). An advantage of taking this approach
is that it is applicable when the number of looks in different channels is
different. This can arise, for example, if the observing geometries for a given
region are different between channels because the flight lines of the sensor varied
between observations. In this more general case,
M
fA ^
dn = -\nP(n) + £ d -*- + h a , (13.10)
-i K, J
= z M z =
/ = X K f tr(^zzt)
i
could be formed. Here A — P^P. For a two-class separation, the problem is then
to determine the matrix A giving the maximum mean intensity ratio between
the target class and the clutter class, that is, to maximize rtc = tr(y!Ct)/trC/4Cc)
where Cx and Cc are the covariance matrices of target and clutter classes,
respectively. This is known as polarimetric matched filtering [52,60,61]. As noted
in [60], a possibly more useful approach is to maximize the contrast between the
two classes. In this case, the matrices Amax and Am[n, which give rise to the
maximum and minimum values of rtc, are found; if rtc(max) > ^ ( m i n ) , Amzx
provides the maximum contrast filter; otherwise Ax^1n is selected. Both forms of
filtering can be translated into optimum transmit and receive polarizations
[52,53]. Explicit solution schemes for these problems and examples of the use
of this approach for distributed target discrimination in polarimetric data will
be found in [52,53]. However, the value of the full polarimetric approach has
been called into question by Novak et al. [61] when detecting man-made targets
in clutter. Here it was found that simpler detectors using only intensity infor-
mation provided comparable or better performance than the polarimetric
matched filter (for the same number of transmitted pulses), without requiring
the exact knowledge of target and clutter covariance needed for the polarimetric
matched filter. (By contrast, Novak et al. [62] note the value of the polarimetric
whitening filter in target detection; see Section 12.3.1.) Since target detection is
the most usual example of where a simple binary classification is sufficient, the
real value of such techniques in remote sensing is therefore still unclear.
L band
P band
Figure 13.5 Classification of C-, L-, and P-band AirSAR data by scattering mechanism for an
area consisting of mixed forest and agricultural fields. White denotes an odd
number of reflections, gray an even number, and black corresponds to diffuse
scattering.
signatures are found in several agricultural fields (unfortunately, we do not have
a detailed description of the cover types or region characteristics for these areas),
and diffuse scattering is encountered in many areas other than the forest. (In at
least some cases, these correspond to regions with low shrubs.) These results are
found even though the P-band forest data in particular show the large H H - W
phase differences expected for an even number of reflections.
Recognition of a dominant scattering mechanism within each region also
forms the basis of the classification scheme described in [64]. In this case the
eigenvectors of the covariance matrix are interpreted as evidence of different
scattering processes and the degree of mixing or randomness of the mechanism
is quantified by an entropy measure. Two examples of resulting classifications
will be found in [64]: in one case applied to distinguishing urban, ocean and
vegetated areas, and in the other for discriminating forest from clear cut areas.
However, whether this approach provides new insights into the information
carried by polarimetric data needs further investigation, since it relies on post-
processing of the covariance matrix, which is the primary information source
(and is directly provided by scattering models, for example). Also, in moving to
quantities derived from the covariance matrix, the estimation theory for the
extracted parameters loses the well-known structure available for the covariance
matrix itself (see [48] and Chapter 11).
1. The changes for wetland rice are much larger than the 3 dB observed
for nonforest regions at Tapajos, in principle allowing greater classifi-
cation accuracy.
2. The gamma MAP filter did not perform very well for the forestry
example given in Chapter 12 (see Figure 12.17(b)) but produces
acceptable results here. This may be because no postprocessing was
performed in Figure 12.17(b). In Figure 13.6, however, a last stage
reclassifies isolated pixels to be consistent with their neighbors [65].
3. No extensive ground data or corroborative evidence were available to
validate the classification in Figure 13.6(d), whereas in Section 12.6
TM data was used to provide a reference. (SPOT data were available
over the study area but were unable to provide a useful discrimination
of the growth stage of the rice.) The justification for the classification
is therefore primarily on the basis of the known relation of RCS to rice
growth stage [65].
(c) (d)
Figure 13.6 (a) ERS-I PRI image of the Demak area, Indonesia, for February 16, 1994; (b)
corresponding image for M a r c h 6,1994 (Copyright ESA). (c) Ratio image formed
after preprocessing; (d) image classification: early rice (black); late rice (mid-
gray); nonrice (white) (From: [65]. © 1997 IEEE).
(a)
(b)
Figure 13.7 (a) Radarsat fine-mode image from Tapajos, Brazilian Amazonia (© Canadian
Space Agency/Agence Spatiale Canadiene, 1997). (b) Image in (a) after applica-
tion of 5 by 5 Frost filter, (c) Coefficient of variation in (a) using a 7 by 7 sliding
window, (d) Order parameter in (a) within an 11 by 11 sliding window [69].
(C)
(d)
Figure 13.7 (continued).
Similar principles were used in [37] (which also includes a survey of previous
classification studies) for ERS-I /JERS-I coregistered images, where now the tree
class was further broken down into structural form and leaf type (broad leaf or
needles). Another hierarchical scheme combined AirSAR and SIR-C data to
distinguish agricultural crops and other sorts of land cover using pixel-based
methods [34], with reasonable accuracies obtained.
Texture in SIR-C L-band data was one of the parameters used in [35] to
make a separation between forest and low vegetation. Its value as a discriminator
of forest from nonforest in tropical regions was also clearly demonstrated using
airborne data in Chapter 8. A further striking example of how it may provide the
most important information source in tropical forest monitoring is illustrated by
Figure 13.7 [69]. The four panels in the figure show (a) a fine-mode Radarsat
image (resolution of the order 6m by 8.9m) from the Tapajos region of Brazilian
Amazonia (which is the same region as that shown in Figure 12.13); (b) the image
in (a) after applying a 5 by 5 Frost filter [70]; (c) the coefficient of variation in
image (a) calculated within a 7 by 7 sliding window; (d) the estimated order
parameter in image (a), assuming K-distributed data, calculated within an 11 by
11 sliding window. The data were obtained under the Radarsat Announcement of
Opportunity under a Principal Investigator ship held by Dr. C. C. F. Yanasse of
I N P E . This set of images illustrates a number of important issues. First, the
resolution and incidence angle afforded by the Radarsat fine mode clearly allows
texture to be discriminated from space and as predicted for the Tapajos region in
[45]. Second, as expected from airborne data gathered during the SAREX cam-
paign [44,71], texture is perhaps the most important discriminator for distin-
guishing cover types in tropical regions (and probably other forested regions of the
world) if C-band data are to be used. Third, and of great significance, is that
texture allows us to distinguish regions of regenerating and primary forest. This
can be seen at the upper left of the roughly rectangular area of pasture in the
middle of the image, which is an area of regenerating forest. It displays compara-
tively low values in the coefficient of variation image, which are equivalent to large
order parameters. Hence it is dark in Figure 13.7 (c) and bright in Figure 13.7(d).
We can interpret this as indicating that in the regenerating forest there is a more
uniform canopy than in the primary forest; in the latter, emergent trees create a
much "rougher" canopy topography, leading to greater variation in backscatter
due to radar shadows and highlights.
Figure 13.8 Biomass map of the Nezer test site, les Landes, France, derived from SIR-C data
[72].
a robust classifier because it is consistently higher for soil surfaces than for forest,
whatever the soil moisture or roughness state. It is also insensitive to miscalibra-
tion and slope effects because it is formed by channel ratios, as discussed in
Chapter 11. The accuracy of the forest/nonforest separation is around 87% [72].
The mapping of forest biomass relies on the L-band HV backscattering
coefficient, estimated after applying the Kuan et al. filter [73] over a 9 by 9
window. This parameter is known to be sensitive to biomass in the range of 0 to
65 tonnes/ha [74]. A strong relationship between the two can be established in les
Landes because of the dominance of maritime pine in this area, with ensuing
regularity of tree structure [74]. Training stands indicate an accuracy of about
80% for classification into five biomass ranges (clear cut, 8 to 20 tonnes/ha, 21 to
33 tonnes/ha, 34 to 50 tonnes/ha, and greater than 50 tonnes/ha).
References
[1] Rott, H., P. Skvarca, and T. Nagler, "Rapid Collapse of Northern Larsen Ice Shelf, Antar-
tica," Science, Vol. 271, 1996, pp. 788-792.
[2] New Views of the Earth: Scientific Achievements of ERS-I, ESA SP-1176/1, 1996.
[3] New Views of the Earth: Applications Achievements of ERS-I, ESA SP-1176/11, 1996.
[4] Second ERS Applications Workshop, London, ESA SP-383, 1996.
[5] Special Issue on SIR-C/X-SAR, IEEE Trans. Geosci. Remote Sensing, Vol. 33, 1995.
[6] Elachi, C , Spaceborne Radar Remote Sensing: Applications and Techniques, New York: IEEE
Press, 1987.
[7] Proc. of the First ERS-I Symp., Cannes, ESA SP-359, 1993.
[8] Proc. of the Second ERS-I Symp., Hamburg, ESA SP-361, 1994.
[9] Proc. of the Third ERS-I Symp., Florence, ESA SP-414, 1997.
[10] Le Toan, T., "Scatterometer Measurements on Crop and Soil Surfaces," Proc. ESA-EARSeL
Workshop, Alpbach, ESA SP-166, 1981, pp. 99-110.
[11] Le Toan, T., A. Lopes, and M. Huet, "On the Relationships Between Radar Backscattering
Coefficient and Vegetation Canopy Characteristics," Proc. IGARSS 84 Symp., Strasbourg,
ESASP-215, 1984, pp. 155-160.
[12] Lopes, A., and T. Le Toan, "Effet de Ia Polarisation d'une Onde Electromagnetique dans
l'Attenuation de l'Onde dans un Couvert Vegetal," Proc. Third Int. Symp. on Spectral
Signatures of Objects in Remote Sensing, Les Arcs, ESA SP-247, 1985, pp. 117-122.
[13] Alphonse, and T. Le Toan, "Retrieving Vegetation and Soil Parameters from Radar Meas-
urements," Proc. Fourth International Symp. on Spectral Signatures of Objects in Remote
Sensing, Aussois, ESA SP-287, 1988, pp. 105-111.
[14] Bush, T. E, and E T Ulaby, "Fading Characteristics of Panchromatic Radar Backscatter
from Selected Agricultural Targets," IEEE Trans. Geosci. Electronics, No. 13, 1976,
pp. 149-157.
[15] Ulaby, F. T, "Radar Response to Vegetation," IEEE Trans. Antennas and Propagation, Vol.
23, 1975, pp. 36-45.
[16] Ulaby, F. T , T. F. Bush, and P P. Batlivala, "Radar Response to Vegetation II: 8-18 GHz
Band," IEEE Trans. Antennas and Propagation, No. 23, 1975, pp. 608-618.
[17] Bush, T. E, and F. T. Ulaby, "An Evaluation of Radar as a Crop Classifier," Remote Sensing
of Environment, Vol. 7, 1978, pp. 15-36.
[18] Ulaby, F. T , C. Allen, G. Eger, and E. Kanemasu, "Relating the Radar Backscattering
Coefficient to Leaf Area Index," Remote Sensing of Environment, Vol. 14, 1984,
pp. 113-133.
[19] Hoekman, D. H., L. Krul, and E. P W. Attema, "A Multi-Layer Model for Backscattering
from Vegetation Canopies," Proc. IGARSS 82 Symp., Munich, IEEE Pub., 1982,
pp. 41-47.
[20] Ulaby, F. T., R. K. Moore, and A. K. Fung, Microwave Remote Sensing: Active and Passive,
Vol. 2, Reading, MA: Addison-Wesley, 1982.
[21] Yanasse, C. C. E, S. Quegan, and R. J. Martin, "Inferences on Spatial and Temporal
Variability of the Backscatter from Growing Crops Using AgriSAR Data," Int. J. Remote
Sensing, Vol. 13, 1992, pp. 493-507.
[22] Snoeij, P., and P J. E Swart, "Processing and Calibration of the DUTSCAT Data During
the AGRISCATT Campaigns," Proc. MAESTRO IAGRISCATT Final Workshop: Radar
Techniques for Forestry and Agricultural Applications, ESA WPP-31, 1992, pp. 160-164.
[23] Hawkins, R. K., and L. D. Teany, "SAREX 1992 Data Calibration," Proc. SAREX-92
Workshop: South American Radar Experiment, ESA WPP-76, 1993, pp. 41-53.
[24] Shimada, M., and M. Nakai, "Inflight Evaluation of L Band SAR of Japanese Earth
Resources Satellite-1," Adv. Space Res., Vol. 14, 1994, pp. 231-240.
[25] Rossi, M., B. Rognon, and D. Massonnet, "JERS-I SAR Image Quality and Interferomet-
ric Potential," IEEE Trans. Geosci. Remote Sensing, Vol. 34, 1996, pp. 824-827.
[26] Satellite Radar in Agriculture: Experience with ERS-I, ESA SP-1185, 1995.
[27] Wegmuller, U., and C. Werner, "Retrieval of Vegetation Parameters with SAR Inter-
ferometry," IEEE Trans. Geosci. Remote Sensing, Vol. 35, 1997, pp. 18-24.
[28] Hagberg, J. O., L. M. H. Ulander, and J. Askne, "Repeat-pass SAR Interferometry over
Forested Terrain," IEEE Trans. Geosci. Remote Sensing, Vol. 33, 1995, pp. 331-340.
[29] Lemoine, G., H. de Groof, and H. J. C. Leeuwen, "Monitoring Agricultural Land Prepa-
ration Activities with ERS-I to Complement and Advance Early Season Crop Estimates in
the Framework of the MARS Project,"Second ERS Applications Workshop, London, ESA
SP-383, 1996, pp. 7-12.
[30] Nghiem, S. V, S. H. Yueh, and R. Kwok, "Symmetry Properties in Polarimetric Remote
Sensing," Radio ScL, Vol. 27, 1992, pp. 693-711.
[31] Quegan, S., and I. Rhodes, "Statistical Models for Polarimetric Data: Consequences,
Testing and Validity," Int. J. Remote Sensing, Vol. 16, 1995, pp. 1183-1210.
[32] Cordey, R. A., J. T. Macklin, P. A. Wright, P. J. Saich, S. Quegan, A. Wielogorska, and W
P. Loughlin, "Studies of Multi-frequency and Multi-polarisation SAR Applications," Final
Report on DRA Contract RAE IB/89, GEC-Marconi Research Centre Doct. MTR
93/73A, Vol. 3, 1993.
[33] Freeman, A., J. Villasensor, J. D. Klein, P Hoogeboom, and J. Groot, "On the Use of
Multi-Frequency and Polarimetric Radar Backscatter Features for Classification of Agricul-
tural Crops," Int. J. Remote Sensing, Vol. 15, 1994, pp. 1799-1812.
[34] Ferrazoli, P., S. Paloscia, P. Pampaloni, G. Schiavon, S. Sigismondi, and D. Solimini, "The
Potential of Multifrequency Polarimetric SAR in Assessing Agricultural and Arboreous
Biomass," IEEE Trans. Geosci. Remote Sensing, Vol. 35, 1997, pp. 5-17.
[35] Pierce, L. E., F. T. Ulaby, K. Sarabandi, and M. C. Dobson, "Knowledge-based Classifica-
tion of Polarimetric SAR Images," IEEE Trans. Geosci. Remote Sensing, Vol. 32, 1994,
pp. 1081-1086.
[36] Hess, L. L., J. M. Melack, S. Filoso, and Y. Wang, "Delineation of Inundated Area and
Vegetation Along the Amazon Floodplain with the SIR-C Synthetic Aperture Radar," IEEE
Trans. Geosci. Remote Sensing, Vol. 33, 1996, pp. 896-904.
[37] Dobson, M. C , L. E. Pierce, and E T. Ulaby, "Knowledge-based Land-Cover Classification
Using ERS-I/JERS-I SAR Composites," IEEE Trans. Geosci. Remote Sensing, Vol. 34, 1996,
pp. 83-99.
[38] Rignot, E. J. M., R. Chellappa, and P Dubois, "Unsupervised Classification of Polarimetric
Data Using the Covariance Matrix," IEEE Trans. Geosci. Remote Sensing, Vol. 30, 1992,
pp. 697-705.
[39] Wong, Y.-E, and E. C. Posner, "A New Clustering Algorithm Applicable to Multispectral
and Polarimetric SAR Images," IEEE Trans. Geosci. Remote Sensing, Vol. 31, 1993,
pp. 634-644.
[40] Anys, H., and D.-C. He, "Evaluation of Textural and Multipolarisation Radar Features for
Crop Classification," IEEE Trans. Geosci. Remote Sensing, Vol. 33, 1995, pp. 1170—1181.
[41] Chen, K. S., W. P. Huang, D. H. Tsay, and E Amar, "Classification of Multifrequency
Polarimetric SAR Imagery Using a Dynamic Learning Neural Network, IEEE Trans. Geosci.
Remote Sensing, Vol. 34, 1996, pp. 814-820.
[42] Hara, Y., R. G. Atkins, S. H. Yueh, R. T. Shin, and J. A. Kong, "Application of Neural
Networks to Radar Image Classification," IEEE Trans. Geosci. Remote Sensing, Vol. 32,
1994, pp. 100-109.
[43] Swain, P. H., and S. M. Davis, Remote Sensing: The Quantitative Approach, New York:
McGraw-Hill, 1978.
[44] Beaudoin, A., M. Deshayes, S. Hardy, T. Le Toan, and D. Girou, "Use of Airborne SAR
Data for the Mapping of Shifting Cultivation in French Guiana," Proc. SAREX-92 Work-
shop, ESA WPP-76, 1994, pp. 185-191.
[45] Oliver, C. J., A. Blake, and R. G. White, "Optimum Texture Analysis of SAR Images,"
SPIE Conf. on Algorithms for Synthetic Aperture Radar Imagery, Orlando, FL, SPIE Proc,
Vol. 2230, 1994, pp.389-398.
[46] Williams, M. L., "The Influence of Canopy Shape on SAR Speckle Distributions over
Woodland," Proc. IGARSS 97, Singapore, 1997, pp. 755-757.
[47] Nezry, E., A. Lopes, D. Ducrot-Gambart, C. Nezry, and J.-S. Lee, "Supervised Classifica-
tion of K-Distributed SAR Images of Natural Targets and Probability of Error Estimation,"
IEEE Trans. Geosci. Remote Sensing, Vol. 34, 1996, pp. 1233-1242.
[48] Goodman, N. R., "Statistical Analysis Based on a Certain Multivariate Gaussian Distribu-
tion (an Introduction)," Ann. Math. Stat., Vol. 34, 1963, pp. 152-177.
[49] Lee, J.-S., M. R. Grunes, and R. Kwok, "Classification of Multi-look Polarimetric SAR
Imagery Based on Complex Wishart Distribution," Int. J. Remote Sensing, Vol. 15, 1994,
pp.2299-2311.
[50] van ZyI, J. J., and C. F. Burnette, "Bayesian Classification of Polarimetric SAR Images
Using Adaptive a Priori Probabilities," Int. J. Remote Sensing, Vol. 13, 1992, pp. 835-840.
[51] Yueh, H. A., A. A. Swartz, J. A. Kong, R. T. Shin, and L. M. Novak, "Bayes Classification
of Terrain Cover Using Normalized Polarimetric Data," /. Geophys. Res., Vol. 93, 1988,
pp. 15261-15267.
[52] Kong, J. A., H. A. Yueh, H. H. Lim, R. T Shin, and J. J. van ZyI, "Classification of Earth
Terrain Using Polarimetric Radar Images," Progress in Electromagnetics Research: Polarimetric
Remote Sensing, J. A. Kong (ed.), New York, Amsterdam, and London: Elsevier, 1990,
pp. 327-370.
[53] van ZyI, J. J., H. A. Zebker, and C. Elachi, "Polarimetric SAR Applications," Radar
Polarimetry for Geoscience Applications, F. T. Ulaby and C. Elachi (eds.), Norwood, MA:
Artech House, 1990, pp. 315-360.
[54] Lim, H. H., A. A. Swartz, H. A. Yueh, J. A. Kong, R. T. Shin, and J. J. van ZyI,
"Classification of Earth Terrain Using Polarimetric Synthetic Aperture Radar Images," /.
Geophys. Res., Vol. 94, 1989, pp. 7049-7057.
[55] Rignot, E. J. M., and R. Chellappa, "Segmentation of Polarimetric Synthetic Aperture
Radar Data," IEEE Trans.Image Processing, Vol. 1, 1992, pp. 281-299.
[56] Rignot, E. J. M., C. L. Williams, J. B. Way, and L. A. Viereck, "Mapping of Forest Types
in Alaskan Boreal Forests Using SAR Imagery," IEEE Trans. Geosci. Remote Sensing, Vol. 32,
1994, pp. 1051-1059.
[57] Evans, D. L., T. G. Farr, J. J. van ZyI, and H. A. Zebker, "Radar Polarimetry: Analysis Tools
and Applications," IEEE Trans. Geosci. Remote Sensing, Vol. 26, 1988, pp. 774-789.
[58] Zebker, H. A., J. J. van ZyI, and D. N. Held, "Imaging Radar Polarimetry from Wave
Synthesis,"/ Geophys. Res., Vol. 92, 1987, pp. 683-701.
[59] Touzi, R., S. Goze, T. Le Toan, and A. Lopes, "Polarimetric Discriminators for SAR
Images," IEEE Trans. Geosci. Remote Sensing, Vol. 30, 1992, pp. 973-980.
[60] Swartz, A. A., H. A. Yueh, J. A. Kong, L. M. Novak, and R. T. Shin, "Optimal Polarisations
for Achieving Maximum Contrast in Radar Images," /. Geophys. Res., Vol. 93, 1988,
pp. 15252-15260.
[61] Novak, L. M., M. B. Sechtin, and M. J. Cardullo, "Studies of Target Detection Algorithms
that Use Polarimetric Radar Data," IEEE Trans. Aerosp. Elect. Systems, Vol. 2, 1989,
pp. 150-165.
[62] Novak, L. M., G. J. Owirka, and C M . Netishen, "Radar Target Identification Using
Spatial Matched Filters," Pattern Recognition, Vol. 27, 1994, pp. 607-617.
[63] van ZyI, J. J., "Unsupervised Classification of Scattering Behaviour Using Radar Po-
larimetry Data," IEEE Trans. Geosci. Remote Sensing, Vol. 27, 1989, pp. 36—45.
[64] Cloude, S. R., and E. Pottier, "An Entropy Based Classification Scheme for Land Applica-
tions of Polarimetric SAR," IEEE Trans. Geosci. Remote Sensing, Vol. 35, 1997, pp. 68-78.
[65] Le Toan, T, F. Ribbes, L.-F. Wang, N. Floury, K.-H. Ding, J. A. Kong, M. Fujita, and T.
Kurosu, "Rice Crop Mapping and Monitoring Using ERS-I Data Based on Experiment
and Modelling Results," IEEE Trans. Geosci. Remote Sensing, Vol. 35, 1997, pp. 41-56.
[66] Kurosu, T, M. Fujita, and K. Chiba, "Monitoring of Rice Crop Growth from Space Using
ERS-I C-band SAR," IEEE Trans. Geosci. Remote Sensing, Vol. 33, 1995, pp. 1092-1096.
[67] Bruniquel, J., and A. Lopes, "Multi-variate Optimal Speckle Reduction in SAR Imagery,"
Int. J. Remote Sensing, Vol. 18, 1997, pp. 603-627.
[68] Lopes, A., E. Nezry, R. Touzi, and H. Laur, "Structure Detection and Adaptive Speckle
Filtering in SAR Images," Int. J. Remote Sensing, Vol. 14, 1993, pp. 1735-1758.
[69] Sant'Anna, S. J. S., C. C. F. Yanasse, and A. C. Frery, "Estudo Comparativo de Alguns
Classificadores Utilizando-se Imagens Radarsat da Regiao de Tapajos," Proc. First latino-
American Seminar on Radar Remote Sensing: Image Processing Techniques, Buenos Aires, ESA
S?-407, 1997, pp. 187-194.
[70] Frost, V. S., J. A. Stiles, K. S. Shanmugan, and J. C. Holtzman, "A Model for Radar Images
and Its Application to Adaptive Filtering of Multiplicative Noise," IEEE Trans. Pattern Anal.
Machine IntelL, Vol. 4, 1982, pp. 157-166.
[71] Grover, K. D., and S. Quegan, "Image Quality, Statistical and Textural Properties of
SAREX Data from the Tapajos Test Site," Proc. SAREX-92 Workshop, ESA WPP-76, 1994,
pp. 15-23.
[72] Souyris, J. C , T. Le Toan, N. Floury, C. C. Hsu, and J. A. Kong, "Inversion of Forest
Biomass Using SIR-C/X-SAR Data," IEEE Trans. Geosci. Remote Sensing, submitted 1997.
[73] Kuan, D. T., A. A. Sawchuk, T. C. Strand, and P C. Chaval, "Adaptive Restoration of
Images with Speckle," IEEE Trans. Acoust. Speech Signal Process., Vol. 35, 1987,
pp. 373-383.
[74] Le Toan, T, A. Beaudoin, J. Riom, and D. Guyon, "Relating Forest Biomass to SAR Data,"
IEEE Trans. Geosci Remote Sensing, Vol. 30, 1992, pp. 403-411.
1 4
14.1 Introduction
Perhaps the most remarkable feature of our understanding of SAR images is that
very simple, approximate models can provide good representations of the prop-
erties of the images. Almost all of the developments described in this book flow
from four such models.
From the first of these, we can immediately infer the behavior of single point
targets, or small clusters of targets, with correct representations of the interference
between them. The sampling properties of the data are also immediate. From the
second, in combination with the first, the properties of speckle arise. When there
are very many scatterers in every resolution cell, the complex reflectivity can be
represented as a white Gaussian process, so the complex image is a correlated
Gaussian process where correlation is determined by the PSE All the elementary
properties of distributed and point targets follow and allow us to quantify the role
of targets, distributed scatterers, and system noise in the observed image.
This basic model needs only to include number fluctuations in the scat-
terer populations between resolution cells in order to provide viable repre-
sentations of texture. Both single-point PDFs and the correlation properties of
textured regions are explained consistently within this simple structure. It also
extends readily to deal with the multidimensional SAR images produced by
polarimetry and interferometry.
A crucial but not obvious consequence is that, for most purposes, the
information contained in the scene only becomes available when products of
channels are formed. For a single channel, this means generation of the intensity
image; while for correlated multichannel data we must include quantities such
as phase difference, amplitude ratio, or correlation coefficient. From the com-
plex image itself we can mainly infer only properties of the sensor, such as the
PSF. The principal exception to this is for isolated point targets, where the phase
signature is given by the SAR PSF and can be used as a detector, as well as being
critical in super-resolution.
While the first two models are central in explaining what we see in SAR
images, the third is crucial in recovering the information they contain. Interfer-
ence between scatterers manifests itself as speckle in single-channel images and
more generally in the multivariate Gaussian distribution encountered for higher
dimensional images. In order to deal with these high-frequency fluctuations we
assume that they mask a real-world complex reflectivity that changes on much
slower scales (except at edges and near isolated targets). All the approaches to
filtering, reconstruction, and segmentation described in earlier chapters have at
their root this assumption of a comparatively smooth reflectivity, punctuated by
structural elements in the scene. The fourth model allows us to build this
knowledge about discontinuities into our algorithms and to make them adaptive
to local structure. The validation of these methods takes us back, at least for
single-channel data, just to the properties of one-dimensional speckle.
Expressed in these terms, the physical content of much of the material in
this book is low because we made little use of electromagnetic theory, the
dielectric properties of materials, or the nature of the scattering medium.
Equally, only the most general concepts about the nature of a scene were
introduced. Instead, we principally concerned ourselves with the processes by
which the arrays of numbers corresponding to digital SAR images need to be
treated in order to isolate the information-bearing components of the data. It is
through these components that the links to deeper physical theories can be
made: by characterizing the data, we bring into focus what physical theory needs
to explain. Equally, the structure we see in an image, on which we can bring our
knowledge of the world to bear, relies on the choice of what to display and how
to display it. This means we must deal properly with situations where more than
a single parameter is needed to capture the information in the data, either
because of multidimensionality or texture.
In this sense, we were concerned with providing the bridge that connects
the physics of scattering, through measurement, to investigation of real world
processes, as illustrated in Figure 14.1. Notice that the arrows connecting the
boxes in this figure are two-way, due to the fact that, in some cases, we can
measure an effect (e.g., texture) without having any clear idea of the physical
phenomenon causing it (indeed, we may have several theories, but be unable to
decide which, if any, is correct). An examination of many of the images in this
book reveals numerous effects that we cannot currently explain. Image analysis
then acts as a stimulus to physical understanding. Equally, the need to apply
images places requirements on the performance of image analysis methods in
order for them to be useful. For example, the loss of much of the prominent
drainage channel shown in Figure 6.12(a) in the reconstruction of Figure 6.12(b)
and to a lesser extent in the segmentation shown as Figure 7.10(b) indicates that
neither algorithm is currently reliable in a mapping context at this spatial resolu-
tion. Such failures motivate new and more powerful methods.
In order to make this bridge as safe as possible, we tried to establish firm
foundations and rigorous developments based on data models. However, as with
all attempts to represent complex phenomena by models (and hence make them
susceptible to mathematical treatment), it is critical to assess the extent to which
the models are complete, accurate, tractable, and useful. By surveying these
highly interrelated issues, we will summarize what was achieved, identify fail-
ings, and indicate directions where further efforts are required.
Physical Image
Application
Theory Analysis
Figure 14.1 Image analysis bridges the gap between physical theory and the high-level un-
derstanding required to apply the images.
which form the backbone of most of the analysis in the earlier chapters,
completeness needs to be coupled with economy. We tried to identify a minimal
set of parameters from which all other measurements can be inferred. Striking
success was demonstrated over a variety of systems for the Gaussian model and
the K distribution model. By measuring a covariance matrix and an order
parameter, all the information is captured. However, we also identified situ-
ations where these models begin to break down, including:
• Very high resolution data ( < Im), where mixture distributions or other
types of distribution become necessary to describe the data (see Section
5.7);
• Long wavelength data (50 to 100 cm), where for some cover types the
distribution may not be properly characterized by either of these mod-
els (see Section 11.9);
• Very long wavelength data (tens of meters) provided by systems such as
CARABAS [1] in which the assumption of many comparable scatterers
per resolution cell is likely to be violated;
• Urban areas, where semideterministic structures violate the conditions
leading to the local Gaussian model.
In the first three of these cases, the amount of data that is available and
has been thoroughly examined is very limited. Very high resolution data are
likely to remain of most relevance in the military context, so clutter models for
target detection will drive the developments. The second two data types have
both military and civilian applications but are unlikely to become widely avail-
able in the near future, not least because of problems of interference with other
systems. Where this problem can be circumvented, an important application for
both types is likely to be in forestry. Particularly for the very long wavelengths,
the data distributions will then need better characterization since they may carry
information on the number density of scatterers and hence tree spacing. (This
is currently very speculative.) Urban areas present special problems due to corner
reflector and specular effects but represent possibly important target types for
application of SAR [2].
If we inquire whether the models are complete in the physical sense,
our answer must be no, for a number of reasons. Perhaps the most pertinent
is that the linear imaging model and the speckle model both rely to some
extent on a representation of the world as a collection of discrete scatterers
whose interaction with the incident wave is by single scattering. The product
model for polarimetric data (Appendix 1 IA) effectively made the further as-
sumption that all scatterers see the same field. In scattering theory terms, this
corresponds to the Born approximation, while the inclusion of attenuation
effects can be described by the distorted Born approximation. This simplifi-
cation may seem unreasonable, since it ignores multiple scattering and situ-
ations, such as surface scattering, for which a continuous model of the medium
seems more appropriate. We nonetheless find that the predictions based on
it are often consistent with the data. For example, since different polarizations
penetrate to different depths in forest canopies, measured order parameters
in the H H and W channels could, in principle, be different. The few meas-
urements that address this question seem to indicate otherwise. In theoretical
terms, [3] also indicates that, at least for one formulation of surface scattering,
the continuous medium model effectively reduces to a discrete single scattering
model.
Another gap in our physical description is in texture modeling. At various
points we discussed texture as a number fluctuation effect, a shadowing/high-
light effect, or a random variation in the underlying continuous RCS. The links
between these points of view are only just becoming established [4,5]. This
physical understanding of texture is important in our statistical modeling of the
process, since Section 5.8 indicated that only by inserting such knowledge can
we produce textures corresponding to those observed in woodland. It is also
crucial in fully exploiting texture because it indicates the geophysical parameters
controlling texture. This insight also indicates the sensor and target conditions
under which texture is likely to be observable [6]. However, no such insight is
available for the texture distributions observed at very high resolutions. Indeed,
at these resolutions the world appears to become progressively more complex.
Simple statistical models break down and the links of more complicated models
to physical parameters become tenuous.
At the other end of the spatial scale, all of the algorithms described in this
book acquire their information locally and only build up long-range spatial
correlations by iteration. They contain no a priori knowledge, for example,
about linear structures or corners, despite these having very high information
content in human perception of scenes [7]. This is reflected in the failure, noted
previously, to detect some structural scene elements that are obvious to a human
observer. Some algorithms that can make use of long-range information exist,
such as the Hough transform for detecting extended linear features [8]. How-
ever, their integration into the Bayesian approach adopted in this book does not
seem straightforward. Indeed, it is questionable whether the type of information
inherent in scene structure is susceptible to any form of Bayesian analysis.
Nonetheless, work by Hellwich [9] demonstrated that line finding can be
introduced into a simulated annealing scheme, as mentioned in Chapter 7. Also,
some aspects of shape can be controlled within a Bayesian treatment, for
example, through the curvature penalty involved in annealed segmentation
(Section 7.3.5). The extent to which such methods can fully capture our prior
knowledge of the properties of scenes has yet to be established.
It must be recognized that the need for powerful methods depends on the
application. Very useful information can in some cases be inferred from SAR
images simply by looking at them. A very good example of this is the analysis
of the break-up of the Larsen iceshelf reported in [11], which effectively treated
the series of ERS-I images like a set of photographs. For this purpose, all the
relevant information was carried in the structure (the iceshelf margin, icefloes,
and fixed features) and there was sufficient radiometric contrast to pick these
out. Many of the images in this book contain structural information that is
obvious to a human being and, therefore, may be all that is needed for mapping.
However, once we go beyond treating the SAR image purely as a picture,
the machinery developed in this book becomes relevant. Parameter estimation
(e.g., in agricultural applications) and all aspects of handling multidimensional
data (e.g., EMISAR or AirSAR multifrequency polarimetric data) require proper
knowledge of the underlying distributions. Detecting small or weak targets in a
clutter background or rapid, consistent handling of large amounts of data
requires automatic or semiautomatic image analysis tools. To carry out these
tasks effectively it is essential to understand how information about the scene is
encoded in the data. The development of such understanding was an underlying
theme throughout this book.
However, the acid test of usefulness is whether these developments can aid
applications. There has, in fact, been little large-scale validation of any of the
methods in this book. We can attribute this to a number of factors:
Perhaps the most important issues with which to end this book are an assess-
ment of whether we have gone as far as is necessary along the road we have been
following and whether we can fruitfully try to go further. Our aim throughout
has been to try to identify the information in SAR images and to devise methods
that display it to best advantage. For single-channel data we followed both
reconstruction and segmentation routes, and our best current efforts are epito-
mized by Figures 6.11 and 7.10(a) for high-resolution airborne data and Figures
6.12 and 7.10(b) for lower resolution satellite data. Visually the reconstructed
airborne data are of very high quality but are known to contain radiometric
distortion (see Section 6.10) and not to conform to the speckle model in regions
of the image with small-scale structure (see Figures 6.9 and 6.10). The seg-
mented data, on the other hand, are visually less attractive, largely because of
the steps in brightness forced by the cartoon model. However, they are ra-
diometrically unbiased and conform closely to the speckle model (see Figure
7.9). This suggests that the lack of such conformity in the reconstructions is a
failure of the algorithms, where there is small-scale structure, not a failure of the
speckle model.
Images like Figures 6.11, 6.12, and 7.10 represent the state of the art, and
we need to question whether the techniques that produced them are good enough
and whether there is any scope for them to be bettered. Although we have
described objective tests of algorithm performance and are currently working to
improve these tests (e.g., to evaluate the shape preservation aspects of the segmen-
tation algorithms), the first question can only be answered by assessing how they
perform in applications. As in all aspects of SAR imaging, this interplay between
technique and purpose will drive progress. Shortcomings of the algorithms be-
come immediately obvious from failures to emulate tasks that can be carried out
by a human being. However, it is not necessary (and is almost certainly unachiev-
able) that all aspects of image understanding should be automatically reproduc-
ible; only those relevant to the task are needed. The failure to reconstruct
faithfully the drainage channel in Figure 6.12 or 7.10(b) may not prevent these
images from being used effectively in crop classification, for example.
A Bayesian, MAP treatment implemented through simulated annealing
appears to provide the most powerful approach to image analysis currently
available. For single-channel data, the only way this is going to be improved is
by embedding improved prior knowledge in the process. One form of knowl-
edge known to be omitted is the existence of thin linear features, as was already
noted. For many civilian applications, however, multichannel (especially mul-
titemporal) data is likely to be more important. For polarimetric and inter-
ferometric data, there are inadequacies in the segmentation methods, since they
ignore interchannel correlation and texture. The existing MAP approach with
annealing extends correctly to multitemporal data in which interchannel corre-
lation is negligible, though again texture is not yet represented. The real chal-
lenges that seem likely to arise will be the incorporation, if possible, of new
forms of knowledge into the Bayesian framework, particularly other types of
data (such as map information and optical images).
Finally, the methods and results described here provide a yardstick against
which other algorithms and further improvements must be measured. We tried
to make clear the continual development of the algorithms and why the cur-
rently preferred methods are judged to be the most powerful available. We also
tried to indicate the relative merits of other approaches so that users of SAR
imagery can make an informed choice in their selection of image analysis tools.
In the final analysis, it is the fitness of the tool to the task that decides its real
value.
References
A
Adaptive feature-preserving filters 403
Agricultural monitoring 414
Airborne SAR 17 30 68 418
2-look 198
DRA X-band image 26 80
geometry 13
high-resolution data 460
See also SAR; SAR images
AirSAR system 2 349 352 353
373 393 394 417
data classes 438
polarimetric data 421 435
SIR-C combined with 442
Amplitude modulation (AM) 43 44
Amplitude product 335
Amplitude ratio distribution 335 340
analytic CDF 340
EMISAR images 342
uses 340
See also Phase difference distribution
Annealed CMAP (ACMAP) 186 189
iterated 192
for sloping RCS model 182
See also CMAP
This page has been reformatted by Knovel to provide easier navigation. 465
466
Index terms Links
Annealed CML (ACML) 179 186 221
analysis 180
with constant RCS 182
execution time 189
iterated 192
Kth configuration in 180
with sloping RCS 189
See also CML
Annealed segmentation 208 408
application of 219
comparison 211
flow diagram 210
loop stages 209
multidimensional 395
texture 251
See also Segmentation
Annealed texture classification 247
Antenna
pattern 37 59
stabilization 59
Area-based difference measures 388
Armored infantry fighting vehicle (AIFV) 298
Atmospheric refractive index 45
Autocorrelation function (ACF) 22
amplitude 262 266
coefficients 262 263
of complex data 111
Gaussian 259 263
intensity 140 262 263 266
log 262 274
normalized 141 261
parameter fitting 270
B
Background
distribution 280
effect of 285
Gaussian, speckle 308
K-distributed 287 290
PDF 280
properties 279 294
RCS 282 283
region, false alarm 280
speckled 311
weak 278
Backscattering coefficient 33 401 403
Bandwidth 12
Doppler 94
spatial 27
Bayesian approach 456
classifier 430
MAP treatment 460
reconstruction 7 161 189
target detection 278 279 314
Beam weighting 31
Bessel function 133 335 336 343
expanded 361
zero-order 361
C
Calibration 5 34
errors 386
inadequate 5
internal, tones 322
necessity of 91
targets 322
Cartoon model 197 198 202
Change detection 8
on images from different dates 438
pixel-based 385
target 305 306
See also Target detection
Channels
in-phase (I) 36 93
polarimetric 333 341
quadrature (Q) 36 93
uncorrelated 431
Chirp pulse 21 24
Classification
accuracies 408 432 433
based on scattering mechanism 434
D
Data
characterization 126
covariance matrix 375
empirical distributions 128
intensity 380
interferometric 355
multifrequency 356
multilook textured 351
multitemporal 356
polarimetric 7 30 322 350
421
SAR 34 98 319
Data model 76 123 157
advantages 82
correlated K-distributed 142 151
homogeneity testing 128
interrelation flowchart 82
limitations of 142
parameter dependency 82
product model 123 130
verification 128
Defocus 45
cause of 46
F
False alarm probability 211
dependence of 288
derivation 288
pixel-based detection 387
G
Gain 36 37
Gamma-distributed RCS 6 133 134 140
259
H
Histograms 404
Hough transform 455
Hughes Map Drift technique 48
Huygens wavelets 92
I
Image classification. See Classification
Images. See SAR images
Image texture. See Texture
Imaging
coherent 139 140
consequences of 102
contrast averaging 141
incoherent 140
RCS fluctuations 138
Incoherent averaging 93
Incoherent imaging 140
Inertial Navigation System (INS) 56 57 62
In-phase (I) channel 36 93
Integrated sidelobe ratio (ISLR) 62 68
Intensity
ACF 140 262 263 266
data 380
distribution 140
estimates 169
fluctuations 123
K
K distribution 134 135 138 143
150 151 238 354
approximations to 238
background 287 290
clutter 238 287 289
intensity 233 238
log likelihood 239 246
ML calculation for 242
MLE for 242
moment evaluation 240
multivariate 324
pixels 457
texture 240
uncorrelated noise 270
Knowledge-based classification 437
Kolmogorov-Smirnov (K-S) test 127 128 151 231
246 249 338 354
L
Lagrange multiplier methods 377
L-band amplitude images 116 117 118
Linear systems theory 100
L-look 93 94 160 162
average intensity 94
covariance matrix 428
gamma distributions 431
intensity data 130
log likelihood 196
M
Mapping
of forest biomass 444
forward 149
inverse 149
Marginal distributions 343 344
Match filtering 23
Maximum a posteriori (MAP)
classification, applications of 432
crop classification 432
detection criterion 279 280
estimate 162
N
Neumann series expansion 87
Neural nets 433
Neymann-Pearson criterion 286 294
Noise
effects of 99
Gaussian 100
multiplicative 76 96
speckle and 85
Nyquist condition 28
O
Objects
classification of 79
extended 53
relative positions of 125
spatial density of 54
Organization, this book 4
Q
Quadratic distance correlation classification 307
Quadrature (Q) channel 36 93
R
Radar cross section (RCS) 6 125
background 282 283
classification 195
deterministic 6
distribution sampling 112 304
estimating 93 99 178
fluctuations 102 114 128 138
197 234 426
T
TABOO algorithm 179
Target classification 305 434
confusion matrix for 307
description of 314
eigenimage 307
quadratic distance correlation 307
shift-invariant two-dimensional pattern matching 307
U
Uncorrelated channels 431
Urban areas 454
W
Weibull distribution 127 128 129 136
Wiener filter 310
World models 5 157
Z
Zero-padding 109