30.12.2013 Views

Download thesis - Curtin University

Download thesis - Curtin University

Download thesis - Curtin University

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Department of Spatial Sciences<br />

Improvement of Geodetic Parameter Estimation in Coastal Regions<br />

from Satellite Radar Altimetry<br />

Xiaoli Deng<br />

This <strong>thesis</strong> is presented for the Degree of<br />

Doctor of Philosophy<br />

of<br />

<strong>Curtin</strong> <strong>University</strong> of Technology<br />

August 2003


i<br />

Declaration<br />

This <strong>thesis</strong> contains no material which has been accepted for the award of any other degree<br />

or diploma in any university.<br />

To the best of my knowledge and belief this <strong>thesis</strong> contains no material previously published<br />

by any other person except where due acknowledgment has been made.<br />

Signature:<br />

………………………………………….<br />

Date:<br />

………………………...


ii<br />

ABSTRACT<br />

Altimeter-derived sea surface heights (SSH) are most probably in error in coastal<br />

regions due, in part, to the complex nature of echoes returned from rapidly varying<br />

coastal topographic surfaces (both land and sea). This dissertation presents improved<br />

altimeter-derived SSH results in coastal regions using the waveform retracking<br />

technique, which reprocesses the waveform data through a ‘coastal retracking<br />

system’.<br />

The system, based upon a systematic and comprehensive analysis of satellite radar<br />

altimeter waveforms in the Australian coast, provides an efficient means of<br />

improving altimeter-derived SSH data, not from a single retracker, but instead from<br />

several retrackers depending on the altimeter waveform characteristics. Central to the<br />

system is the use of two retracking techniques: the modified iterative least squares<br />

fitting and the threshold retracking algorithms. To overcome the ‘noise’ in<br />

waveforms caused by fading noise, the fitting algorithm has been developed to<br />

include a weighted iterative scheme. The retrackers adopted in the system include<br />

five fitting models and the threshold method with varying threshold levels. A<br />

waveform classification procedure has also been developed, which enables the<br />

waveform to be sorted and retracked by an appropriate retracker.<br />

Twenty-Hertz waveform data from the ERS-2 and POSEIDON satellites have been<br />

used to estimate a broad contaminated distance of ~10 km offshore the Australian<br />

coast. Two cycles of ERS-2 waveform data are, then, reprocessed using the coastal<br />

retracking system to obtain the improved SSH estimates. Using the AUSGeoid98<br />

geoid grid as a partly independent ground reference, the coastal retracking system<br />

has been able to reduce the contaminated distance to ~5 km. Improved ERS-1 SSH<br />

data from waveform retracking has also been obtained in another coastal region of<br />

Taiwan. Two altimeter-based gravity anomaly grids were created near Taiwan using<br />

the SSH data before and after retracking. Compared with ship-track gravity data,<br />

results show that the accuracy of the gravity anomalies has been improved from<br />

±13.9 mgal before retracking to ±9.9 mgal after retracking.


iii<br />

ACKNOWLEDGEMENTS<br />

I wish to extend my warmest and sincerest thanks to my supervisor, Professor Will<br />

Featherstone. Will has always been patient, helpful and supportive. I have learned a<br />

lot about the research methodology and geodesy from him. To him, I am indebted<br />

not just for his scientific supervision, but also for his understanding, help and<br />

friendship.<br />

I am most grateful to my associate supervisor, Professor Cheinway Hwang (National<br />

Chiao Tung <strong>University</strong>, Taiwan), for giving me the opportunity to work in his Space<br />

Geodesy Laboratory for a month, and for his help with many aspects of this research.<br />

I am grateful also to my other associate supervisors, Dr Jon Kirby and Professor<br />

Phillipa Berry (De Montfort <strong>University</strong>, UK), for their guidance during much of this<br />

research.<br />

My thanks go to all of the friendly team members, both staff and students, in the<br />

Western Australian Centre for Geodesy for their help and advice, particularly to<br />

people who worked with me in the lab: Troy Forward, Simon Holmes, Minghai Jia,<br />

Sten Claessens and Ireneusz Baran. To all the staff in the Department office, both<br />

past and present, I also extend my thanks.<br />

My thanks also go to Dr Michael Kuhn, Dr Mark Stewart and Dr Ruey-Gang Chang<br />

for their helpful comments towards this dissertation, and to Dr Graham Quartly<br />

(Southampton Oceanography Centre, UK), Dr Ronald Brooks (NASA, USA) and Dr<br />

George Hayne (NASA, USA) for providing some references and useful discussions.<br />

This research was funded by the International Postgraduate Research Scholarship<br />

(IPRS) and <strong>Curtin</strong> <strong>University</strong> Postgraduate Scholarship (CUPS). In-kind support was<br />

received from AVISO and ESA, by many of kindly providing POSEIDON and ERS-<br />

1/2 waveform data, respectively.<br />

Finally, I would like to thank my family for the constant love and support, in<br />

particular, my daughter, Lulu, with whom I shared happiness in the past two and half<br />

years. Without this love, patience and support, I am quite sure that this dissertation<br />

would never have been finished.


iv<br />

TABLE OF CONTENTS<br />

ABSTRACT.................................................................................................................ii<br />

ACKNOWLEDGEMENTS ........................................................................................iii<br />

TABLE OF CONTENTS............................................................................................ iv<br />

LIST OF FIGURES .................................................................................................... ix<br />

LIST OF TABLES ...................................................................................................xvii<br />

ACRONYMS AND ABBREVIATIONS ................................................................. xix<br />

1. INTRODUCTION ............................................................................................... 1<br />

1. 1 Applications of Satellite Radar Waveforms................................................. 1<br />

1.1.1 Description of the Satellite Radar Altimeter Waveforms .................... 1<br />

1.1.2 General Applications of Waveform Data............................................. 4<br />

1. 2 Problems in Coastal Areas ........................................................................... 6<br />

1.2.1 Topographic Effects on Waveforms .................................................... 7<br />

1.2.2 Topographic Effects on Geophysical Corrections ............................. 10<br />

1. 3 Justifications for This Study ...................................................................... 11<br />

1.3.1 Previous Work.................................................................................... 11<br />

1.3.2 Significance of the Improvement of Altimeter Data in Coastal<br />

Regions ............................................................................................................ 12<br />

1. 4 The Study Areas......................................................................................... 15<br />

1. 5 Thesis Outline ............................................................................................ 16<br />

2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES..... 18<br />

2. 1 Introduction................................................................................................ 18<br />

2. 2 General Information on Construction of Altimeter Measurements ........... 19<br />

2.2.1 Fundamental Radar Measurement Principles..................................... 19<br />

2.2.2 The Returned Waveform.................................................................... 21<br />

2.2.3 The Range and Sea Surface Height.................................................... 25<br />

2. 3 The Altimeter Waveform Data Used ......................................................... 26<br />

2.3.1 ERS-1 Waveform Data ...................................................................... 26<br />

2.3.2 ERS-2 Waveform Data ...................................................................... 27<br />

2.3.3 POSEIDON Waveform Data ............................................................. 28<br />

2. 4 External Input Data .................................................................................... 29<br />

2.4.1 The AUSGeoid98 Geoid Model of Australia..................................... 30


v<br />

2.4.2 Sea Surface Topography .................................................................... 31<br />

2.4.3 Marine Gravity Data around Taiwan ................................................. 34<br />

2.4.4 The Australian Digital Elevation Model (DEM) ............................... 35<br />

2. 5 Summary .................................................................................................... 36<br />

3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS............... 37<br />

3. 1 Introduction................................................................................................ 37<br />

3. 2 Convolutional Representation of Ocean-Returned Waveforms................. 38<br />

3.2.1 The Brown Model .............................................................................. 38<br />

3.2.2 Average Flat Surface Impulse Response Function ............................ 40<br />

3.2.3 Surface Elevation PTR of the Scattering Distribution ....................... 41<br />

3.2.4 Radar System PTR ............................................................................. 42<br />

3.2.5 Solution of the Radar Returns............................................................ 42<br />

3. 3 Modified Convolutional Representation of the Ocean Return Waveform. 43<br />

3.3.1 Effects of the Earth’s Curvature......................................................... 44<br />

3.3.2 Simplification of the Bessel Function................................................ 44<br />

3.3.3 Approximate Expression of the Non-Gaussian Scattering Surface ... 46<br />

3.3.4 Approximate Expression of the System PTR .................................... 49<br />

3.3.5 A Modified Solution of the Radar Returns Developed in This Study50<br />

3. 4 Ocean Waveform Retracking..................................................................... 55<br />

3.4.1 Least Squares Fitting Procedure ........................................................ 56<br />

3.4.2 Deconvolution Method ...................................................................... 58<br />

3. 5 Ice-Sheet Waveform Retracking................................................................ 60<br />

3.5.1 Fitting Algorithm - β-Parameter Retracking...................................... 62<br />

3.5.2 Off-Centre of Gravity (OCOG) Retracking Algorithm ..................... 65<br />

3.5.3 Threshold Retracking Algorithm ....................................................... 66<br />

3.5.4 Surface/Volume Scattering Retracking Algorithm ............................ 67<br />

3. 6 Land Waveform Retracking....................................................................... 69<br />

3. 7 Brief Introduction to Coastal Waveform Retracking................................. 71<br />

3. 8 Summary .................................................................................................... 72<br />

4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS .. 74<br />

4. 1 Introduction................................................................................................ 74<br />

4. 2 Background ................................................................................................ 75<br />

4.2.1 ERS-2 and POSEIDON Altimeter Footprints.................................... 76


vi<br />

4.2.2 ERS-2 and POSEIDON Altimeter Waveforms.................................. 76<br />

4. 3 Processing Methodology for Determination of Contamination Area ........ 78<br />

4.3.1 Threshold Retracking......................................................................... 78<br />

4.3.2 Mean Power of the Waveform ........................................................... 80<br />

4. 4 Data and Editing......................................................................................... 81<br />

4.4.1 ERS-2 ................................................................................................. 81<br />

4.4.2 POSEIDON........................................................................................ 83<br />

4. 5 Determination of Contamination Area....................................................... 84<br />

4.5.1 Using 50% of Threshold Retracking Points of ERS-2....................... 84<br />

4.5.2 Using the Mean Waveform and its Standard Deviation..................... 89<br />

4.5.3 Using Statistics of the 50% Retracking Point .................................... 94<br />

4. 6 Waveform Contamination Analysis........................................................... 97<br />

4.6.1 Waveform Contamination for Water-to-Land Ground Tracks .......... 97<br />

4.6.2 Waveform Contamination for Land-to-Water Ground Tracks ........ 101<br />

4.6.3 Contamination Caused by the Altimeter’s Operation ...................... 107<br />

4.6.4 Summary of Waveform Contamination in Coasts ........................... 109<br />

4. 7 An Experiment of the Land Effects on Waveforms Using the DEM Data....<br />

.................................................................................................................. 110<br />

4.7.1 Concept of Land Effects Using the DEM Model............................. 111<br />

4.7.2 A Preliminary Test of Land Effects Using the DEM Model ........... 114<br />

4. 8 Summary .................................................................................................. 116<br />

5. COASTAL WAVEFORM RETRACKING SYSTEM: DESIGN AND<br />

IMPLEMENTATION.............................................................................................. 119<br />

5. 1 Introduction.............................................................................................. 119<br />

5. 2 Coastal Waveform Retracking System Design........................................ 120<br />

5. 3 Extraction of the Geodetic Parameter: the Sea Surface Height ............... 122<br />

5. 4 Analysis of Waveform Shapes................................................................. 123<br />

5.4.1 Determination of the Peaks in the Waveform.................................. 123<br />

5.4.2 Waveform Classification.................................................................. 124<br />

5. 5 Fitting Functions ...................................................................................... 127<br />

5.5.1 The Ocean Model............................................................................. 127<br />

5.5.2 Five- and Nine-β-Parameter Models................................................ 131<br />

5.5.3 Comparison between the Ocean and Five-β-Parameter Models...... 131


vii<br />

5. 6 Iterative Nonlinear Fitting Procedure ...................................................... 136<br />

5.6.1 Solutions of the Waveform Retracking Fitting Function................. 137<br />

5.6.2 Linearisation of the Waveform Retracking Fitting Function........... 138<br />

5.6.3 Initial Estimates of the Unknown Model Parameters ...................... 139<br />

5.6.4 Data Weight and Outlier Detection Scheme .................................... 142<br />

5.6.5 Iterative Procedure ........................................................................... 144<br />

5. 7 Threshold Retracking............................................................................... 145<br />

5.7.1 Threshold Retracking Algorithm ..................................................... 145<br />

5.7.2 Selection of the Threshold Level ..................................................... 146<br />

5.7.3 Results of Selecting the Threshold Levels....................................... 149<br />

5. 8 An Assessment of Biases among Waveform Retracking Algorithms ..... 155<br />

5.8.1 Analysis Methods and Data ............................................................. 155<br />

5.8.2 Retracking Results and Analysis...................................................... 157<br />

5.8.3 Biases among Retracking Algorithms.............................................. 160<br />

5.8.4 Biases between the SSH Data before and after Retracking ............. 164<br />

5. 9 Summary .................................................................................................. 167<br />

6. WAVEFORM RETRACKING APPLICATION TO THE AUSTRALIAN<br />

COAST..................................................................................................................... 169<br />

6. 1 Introduction.............................................................................................. 169<br />

6. 2 Waveform Classification in Australian Coastal Regions......................... 170<br />

6.2.1 Typical Waveform Shapes ............................................................... 170<br />

6.2.2 Results of Waveform Categories ..................................................... 176<br />

6. 3 Waveform Retracking Results and Discussion........................................ 177<br />

6.3.1 Data Editing ..................................................................................... 177<br />

6.3.2 Waveform Retracking Using Different Algorithms......................... 177<br />

6.3.3 Waveform Retracking Using Different Fitting Functions ............... 179<br />

6. 4 Analysis of the Single-Track Retracked Results...................................... 182<br />

6. 5 Analysis of the SSH Data Before and After Retracking.......................... 185<br />

6. 6 Collinear Analysis.................................................................................... 187<br />

6.6.1 Methodology .................................................................................... 187<br />

6.6.2 Discussion of Results ....................................................................... 188<br />

6. 7 Evaluation of SSH Data Using Geoid Heights ........................................ 190<br />

6.7.1 Comparison with the AUSGeoid98 Geoid Model ........................... 190


viii<br />

6.7.2 Results of the Comparison between SSH Data and Geoid Heights. 191<br />

6. 8 Summary .................................................................................................. 193<br />

7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES NEAR<br />

TAIWAN FROM WAVEFORM RETRACKING .................................................. 196<br />

7. 1 Introduction.............................................................................................. 196<br />

7. 2 Data and Area........................................................................................... 198<br />

7. 3 Waveform Retracking .............................................................................. 200<br />

7.3.1 Retracking Scheme........................................................................... 201<br />

7.3.2 Examples of Waveform Retracking Results .................................... 202<br />

7. 4 The SSH Extracted from Waveforms ...................................................... 205<br />

7.4.1 Corrections ....................................................................................... 205<br />

7.4.2 The Weight Function ....................................................................... 211<br />

7.4.3 Data Filtering and Editing................................................................ 211<br />

7. 5 Marine Gravity Anomalies from Satellite Altimetry ............................... 212<br />

7.5.1 Along-track Deflections of the Vertical........................................... 216<br />

7.5.2 Averaging Gradients ........................................................................ 216<br />

7.5.3 Gravity Anomaly Computation........................................................ 219<br />

7.5.4 Comparisons between Marine and Ship Gravity Anomalies ........... 219<br />

7. 6 Summary .................................................................................................. 224<br />

8. CONCLUSIONS AND RECOMMENDATIONS .......................................... 226<br />

8. 1 Summary of Dissertation.......................................................................... 226<br />

8. 2 Specific Conclusions................................................................................ 228<br />

8.2.1 The Quantitative Estimation of the Contaminated Distance............ 228<br />

8.2.2 Practical Implementation of the Coastal Waveform Retracking System<br />

.......................................................................................................... 230<br />

8.2.3 Development and Validation of a Coastal Retracking System ........ 232<br />

8.2.4 Other Conclusions............................................................................ 235<br />

8. 3 Recommendations for Future Work......................................................... 235<br />

8.3.1 Extraction of Ocean Returns from the Mixed Waveform Shape ..... 236<br />

8.3.2 Improvement of the Correction Algorithms near Land.................... 236<br />

8.3.3 Improvement of the Altimeter Operation......................................... 237<br />

8.3.4 Geodetic Application ....................................................................... 237<br />

REFERENCES......................................................................................................... 238


ix<br />

LIST OF FIGURES<br />

Figure 1.1 Schematic altimeter mean return waveform over sea surfaces................... 3<br />

Figure 1.2 An observed ERS-2 waveform over deep oceans. (1 bin/gate = 0.45<br />

m and tracking gate = 32.5)....................................................................... 3<br />

Figure 1.3 Altimeter simultaneously illuminates land and ocean near the<br />

coastline..................................................................................................... 8<br />

Figure 1.4 Differences between the ERS-2 SSH and AUSGeiod98 geoid height<br />

along four collinear repeat tracks from cycles 42 and 43 at the<br />

Southern Australian coast, showing the metre-level uncertainty of<br />

SSHs closer to the coastline. ..................................................................... 9<br />

Figure 2.1 Basic measurement principle of the altimeter-derived SSH (From<br />

http://www.jason.oceanobs.com/html/portail/galerie<br />

/banque_img_welcome_uk.php3). .......................................................... 18<br />

Figure 2.2 A schematic geometrical description of the interaction of a pulse and<br />

the scattering surface, and the build up of a return waveform over<br />

the duration of a pulse. Note that, an average returned waveform<br />

from 50 typical individual returns is depicted, while a single typical<br />

return shows more noise. (PLF is the pulse-limited footprint). .............. 22<br />

Figure 2.3 Quasi-time-independent SST (contour interval is 10 cm) from Levitus<br />

et al. (1997). ............................................................................................ 33<br />

Figure 2.4 Coverage of ship-track gravity observations around Taiwan................... 34<br />

Figure 3.1 Values of d ( ξ = 0 ), showing that d


x<br />

Figure 3.6 Nine- β -parameter model, fitting a double-ramp function to two<br />

leading edges of the waveform samples.................................................. 64<br />

Figure 3.7 Schematic description of the OCOG method applied to an observed<br />

waveform. Using the OCOG method to calculate the amplitude (A)<br />

and width (W) of the rectangle, and the position of the 50% of<br />

amplitude interpolated at the leading edge as the estimate of<br />

expected tracking gate............................................................................. 66<br />

Figure 4.1 An example of the observed POSEIDON waveform over deep oceans<br />

(tracking gate = 31.5). ............................................................................. 77<br />

Figure 4.2 ERS-2 ground tracks (35-day repeat orbit, from March to April 1999)<br />

to a distance of 350 km from Australian shoreline (Lambert<br />

projection). .............................................................................................. 82<br />

Figure 4.3 The ocean-ice-mode flag in ERS-2 data supplied close to the<br />

Australian shoreline compared to the Wessel (2000) coastline,<br />

showing that most flags around Australia are on the land (Lambert<br />

projection). .............................................................................................. 82<br />

Figure 4.4 POSEIDON ground tracks (10-day repeat orbit, from January 1998 to<br />

January 1999) to a distance of 350 km from Australian shoreline<br />

(Lambert projection). .............................................................................. 83<br />

Figure 4.5 One cycle of ERS-2 contaminated waveforms (highlighted by circles)<br />

along ground tracks in the southeast coasts of Australia. ....................... 86<br />

Figure 4.6 Distributions of 50% retracking points plotted versus DS759.2 ocean<br />

depths for contaminated waveforms in areas of 105°≤λ


xi<br />

Figure 4.9 Mean power of the waveform (left) and the standard deviation of the<br />

mean waveform (right) for ERS-2 in five 10 km-wide bands from<br />

the Australian coastline........................................................................... 91<br />

Figure 4.10 Mean power of the waveform (left) and the standard deviation of the<br />

mean waveform (right) for POSEIDON cycle 197 in five 10 kmwide<br />

bands from the Australian coastline. .............................................. 92<br />

Figure 4.11 Mean power of the waveform (left) and the standard deviation of the<br />

mean waveform (right) for ERS-2 in five 2 km-wide bands from the<br />

Australian coastline................................................................................. 93<br />

Figure 4.12 Mean power of the waveform (left) and the standard deviation of the<br />

mean waveform (right) for POSEIDON cycle 197 in five 2 km-wide<br />

bands from the Australian coastline........................................................ 94<br />

Figure 4.13 Fifty percent retracking point frequency distributions for 5 cycles of<br />

Poseidon waveform data (left) and 1 cycle of ERS-2 waveform data<br />

(right). They are shown as the percentage of observations in each bin<br />

for five 2 km-wide bands around the Australian coastline. .................... 95<br />

Figure 4.14 The 20 Hz SSH profile of pass 21792 (ascending track, cycle 43 of<br />

ERS-2), approaching the eastern Australian shoreline from the<br />

Tasman Sea. The number beside the SSH samples corresponds to the<br />

waveform in Figure 4.15. ........................................................................ 98<br />

Figure 4.15 20 Hz waveforms of pass 21792 (ascending track, cycle 43 of ERS-<br />

2), approaching Australian eastern shoreline in the Tasman Sea. The<br />

number in the top left-hand corner is related to that in Figure 4.14........ 98<br />

Figure 4.16 The 20 Hz SSH profile of pass 21450 (ascending track, cycle 43 of<br />

ERS-2), approaching the western Australian shoreline from the<br />

Indian Ocean. The number beside the SSH samples corresponds to<br />

the waveform in Figure 4.17. ................................................................ 100<br />

Figure 4.17 The 20 Hz waveforms of pass 21450 (ascending track, cycle 43 of<br />

ERS-2), approaching Australian western shoreline in the Indian<br />

Ocean. The number is related to that in Figure 4.16............................. 100<br />

Figure 4.18 The 20 Hz SSH profile of pass 21636 (ascending track, cycle 43 of<br />

ERS-2), receding from the northwestern Australian shoreline to the<br />

Indian Ocean. ........................................................................................ 103


Figure 4.19 20 Hz waveforms of pass 21636 at the 1st second (~0-7 km alongtrack<br />

distance) from the shoreline (ascending track, cycle 43 of ERS-<br />

2), receding from the northwestern Australian shoreline to the Indian<br />

Ocean..................................................................................................... 103<br />

Figure 4.20 20 Hz waveforms of pass 21636 at the 2nd second (~7-14 km alongtrack<br />

distance) from the shoreline (ascending track, cycle 43 of ERS-<br />

2), receding from northwestern Australian shoreline to the Indian<br />

Ocean..................................................................................................... 104<br />

Figure 4.21 20 Hz waveforms of pass 21636 at the 3rd second (~14-21 km<br />

along-track distance) from the shoreline (ascending track, cycle 43<br />

of ERS-2), receding from the northwestern Australian shoreline to<br />

the Indian Ocean. .................................................................................. 104<br />

Figure 4.22 The 20 Hz SSH profile of pass 21886 (descending track, cycle 43 of<br />

ERS-2), receding from the southern Australian shoreline to the<br />

Southern Ocean. .................................................................................... 105<br />

Figure 4.23 20 Hz waveforms of pass 21886 at the 1st second (0-7 km alongtrack<br />

distance) from the shoreline (descending track, cycle 43 of<br />

ERS-2), receding from the southern Australian shoreline to the<br />

Southern Ocean. .................................................................................... 106<br />

Figure 4.24 20 Hz waveforms of pass 21886 at the 2nd second from the<br />

shoreline (ascending track, cycle 43 of ERS-2), receding Australian<br />

southern shoreline to the Southern Ocean............................................. 106<br />

Figure 4.25 20 Hz waveforms of pass 21886 at the 3rd second from the<br />

shoreline (ascending track, cycle 43 of ERS-2), receding Australian<br />

southern shoreline to the Southern Ocean............................................. 107<br />

Figure 4.26 A schematic geometrical description of the radar return from a point<br />

on the land surface at off-nadir angle ξ and associated colatitude<br />

angle φ where the land elevation above the mean sea level is H .<br />

The distance from the altimeter to the land surface is<br />

xii<br />

'<br />

R<br />

φ<br />

. ................... 113<br />

Figure 4.27 Retracking gate estimates corresponding to the land elevation. The<br />

search radius is 2.7 km (top) and 3.5 km (bottom). .............................. 115<br />

Figure 5.1 Block diagram of the coastal retracking system..................................... 121<br />

Figure 5.2 Typical ERS-2 waveform categories in Australian coasts ..................... 126


Figure 5.3 Variation of the leading edge’s slope with varying SWH (0 - 20 m),<br />

xiii<br />

modelling waveforms using the ocean model ( ξ = 0 °). ....................... 129<br />

Figure 5.4 Effect of the off-nadir angle (in degrees) on the amplitude, leading<br />

edge, and especially the slope of the trailing edge, modelling<br />

waveforms using the ocean model (SWH = 4 m). ................................ 130<br />

Figure 5.5 Variations of the trailing edge’s slope with varying parameter β 5 ,<br />

modelling waveforms using the five-β-parameter parameter function<br />

(linear trailing edge, SWH = 4 m, β<br />

21<br />

= 500 counts, β 5 = β<br />

51)........... 134<br />

Figure 5.6 Variations of the trailing edge’s slope with varying parameter β 5 ,<br />

modelling waveforms using the five-β-parameter parameter function<br />

(exponential decayed trailing edge, SWH = 4 m, β<br />

21<br />

= 500 counts,<br />

β 5 = β 51<br />

). ............................................................................................. 135<br />

Figure 5.7 Mean waveforms and the standard deviation in two 5 km-wide bands<br />

of 0-5 km (top) and 5-10 km (bottom) from the Australian coastline.. 148<br />

Figure 5.8 Ten 15º×15º coastal areas around Australia, showing the area number<br />

(Lambert projection). ............................................................................ 150<br />

Figure 5.9 Mean differences between the SSH data after retracking and<br />

AUSGeoid98 geoid heights in ten areas 5 km from the Australian<br />

coastline................................................................................................. 152<br />

Figure 5.10 Standard deviations of the mean differences in ten areas 5 km from<br />

the Australian coastline......................................................................... 152<br />

Figure 5.11 Mean differences between the SSH data after retracking and<br />

AUSGeoid98 geoid heights in ten areas 5-10 km from the Australian<br />

coastline................................................................................................. 154<br />

Figure 5.12 Standard deviations of the mean differences in ten areas 5-10 km<br />

from the Australian coastline. ............................................................... 154<br />

Figure 5.13 Geoid height and SSH profiles along ground tracks 21085 (top) and<br />

21364 (bottom) (ERS-2, cycle 42) before and after retracking using<br />

the ocean model, the five-parameter model and the 50% threshold<br />

level in coastal regions.......................................................................... 158<br />

Figure 5.14 Retracked and unretracked SSH profiles over open oceans, showing<br />

the biases among them. ......................................................................... 161


xiv<br />

Figure 5.15 Histogram of the differences between the ocean model and fiveparameter<br />

model.................................................................................... 162<br />

Figure 5.16 Histograms of the differences between the 50% threshold retracking<br />

technique and the ocean model as well as the five-parameter model. .. 162<br />

Figure 5.17 Histograms of the differences between unretracked SSH data and<br />

retracked SSH data using the ocean model and the five-parameter<br />

model..................................................................................................... 164<br />

Figure 5.18 Histogram of the differences between untracked SSH data and<br />

retracked SSH dada using the 50% threshold retracking technique...... 165<br />

Figure 6.1 Part of ERS-2 ground tracks off the northwest Australian coast,<br />

showing the orbit number of each ground track. The highlighted<br />

diamonds along tracks will be shown in detail waveform shapes later<br />

in this Section........................................................................................ 171<br />

Figure 6.2 Typical ocean waveforms from pass 21815 (cf. Figure 6.1). ................. 173<br />

Figure 6.3 Waveforms recorded when the ground track 21815 is approaching an<br />

island at latitude ~ -20.5º and leaving the island to water at ~ -20.9º.<br />

The data gap is due to the island (cf. Figure 6.1).................................. 173<br />

Figure 6.4 Waveforms leaving land to water (pass 21636 in Figure 6.1). The<br />

high-peaked specular responses are probably due to off-nadir<br />

brighter calm water. The waveform shifting (~ -22.25°


xv<br />

Figure 6.9 STD (positive values) of the collinear 20 Hz SSH differences of<br />

cycles 42 and 43 before and after retracking in six 5 km wide<br />

distance bands, showing improvement from SSH data after<br />

retracking............................................................................................... 189<br />

Figure 6.10 Mean differences of geoid heights (AUSGeoid98) and SSH data<br />

before and after retracking in six 5 km wide distance bands (cycle<br />

42). ........................................................................................................ 191<br />

Figure 6.11 STD (positive values) of the mean difference before and after<br />

retracking in six 5 km wide distance bands, showing improvement<br />

on SSH data after retracking (cycle 42). ............................................... 192<br />

Figure 6.12 Mean differences between geoid heights (AUSGeoid98) and SSH<br />

data before and after retracking in six 5 km wide distance bands<br />

(cycle 43)............................................................................................... 192<br />

Figure 6.13 STD (positive values) of the mean difference before and after<br />

retracking in six 5 km wide distance bands, showing improvement<br />

on SSH data after retracking (cycle 43). ............................................... 192<br />

Figure 7.1 Distribution of ERS-1 ground tracks ...................................................... 199<br />

Figure 7.2 The histogram of dh (the SSH difference between the five-parameter<br />

model and the ocean model (mean -0.7 cm; STD 13.5 cm; 3070<br />

observations). ........................................................................................ 202<br />

Figure 7.3 Three-ground tracks in the study area .................................................... 203<br />

Figure 7.4 SSHs from ground track A1 (a), D1 (b), and D2 (c) in Figure 7.3<br />

before and after retracking using the retrackers of the five-parameter<br />

and ocean models. ................................................................................. 204<br />

Figure 7.5 SSH profiles and parts of corrections along the track A1. ..................... 208<br />

Figure 7.6 SSH profiles and parts of corrections along the track D1. ..................... 209<br />

Figure 7.7 SSH profiles and parts of corrections along the track D2. ..................... 210<br />

Figure 7.8 Untracked SSH profiles, showing the filtered SSH data (dark curves)<br />

and unfiltered SSH data (light curves) along the tracks A1 (a), D1 (b)<br />

and D2 (c) in Figure 7.3. ....................................................................... 213<br />

Figure 7.9 Retracked SSH profiles, showing filtered SSH data (dark curves) and<br />

unfiltered SSH data (light curves) along the tracks A1 (a), D1 (b)<br />

and D2 (c) in Figure 7.3. ....................................................................... 214


xvi<br />

Figure 7.10 Flow diagram for the SSH extraction from waveforms and recovery<br />

of gravity anomalies.............................................................................. 220<br />

Figure 7.11 Gravity anomalies from altimeter data before retracking (units in<br />

mgal) ..................................................................................................... 221<br />

Figure 7.12 Gravity anomalies from altimeter data after retracking (units in mgal)<br />

............................................................................................................... 221<br />

Figure 7.13 Differences between altimeter-derived (before retracking) and shiptrack<br />

gravity anomalies ......................................................................... 223<br />

Figure 7.14 Differences between altimeter-derived (after retracking) and shiptrack<br />

gravity anomalies ......................................................................... 223


xvii<br />

LIST OF TABLES<br />

Table 2.1 Altimeter Operating Parameters................................................................. 20<br />

Table 2.2 Corrections applied to ERS-1 range measurements and their likely<br />

error......................................................................................................... 25<br />

Table 2.3 The repeat cycle characteristics of ERS-1. ................................................ 26<br />

Table 2.4 The reference ellipsoid characteristics of ERS-1/2 and T/P. ..................... 29<br />

Table 4.1 Contaminated 1 Hz ocean depths and 20 Hz distances from the nearest<br />

shoreline in different ............................................................................... 85<br />

Table 4.2 Differences between 50% retracking points and the tracking gate (bin<br />

31.5) for 5 cycles (January 1998 to January 1999) of Poseidon<br />

waveforms (each bin has the range distance of ~0.469 m). .................... 96<br />

Table 4.3 Differences between 50% retracking points and the tracker point (bin<br />

32.5) for 1 cycle (March to April 1999) of ERS-2 waveforms (each<br />

bin with the range width of ~0.454m)..................................................... 96<br />

Table 4.4 Descriptive statistics of the differences of retracking gate estimates<br />

before and after shifting rightward waveform samples in the range<br />

window (20 Hz waveforms from pass 21886 at the 2nd second from<br />

the coastline, 1 gate≈0.4542 m) ............................................................ 108<br />

Table 5.1 ασ² Values for ERS-1/2 (ocean mode) and TOPEX ............................... 132<br />

Table 5.2 Criteria of the absolute differences between two iterations (the ocean<br />

model for ERS-1/2) ............................................................................... 145<br />

Table 5.3 Along-track geoid and SSH gradients (track 21085)............................... 159<br />

Table 5.4 Along-track geoid and SSH gradients (track 21364)............................... 160<br />

Table 5.5 Biases between different retracking techniques from fitting a quadratic<br />

function to the ERS-2 SSH (cycle 42) differences over an ocean area<br />

(-30°S


xviii<br />

Table 5.8 Biases between the ERS-2 (cycle 43) SSHs before and after retracking<br />

over an ocean area (-30°S


xix<br />

ACRONYMS AND ABBREVIATIONS<br />

AGC<br />

Automatic gain control<br />

AGSO Australian Geological Survey Organisation, now Geoscience<br />

Australia (GA)<br />

ALT.WAP<br />

AVISO<br />

AWF<br />

DEM<br />

ECMWF<br />

EMB<br />

EODC<br />

ESA<br />

FFT<br />

GDR<br />

GPS<br />

IAG<br />

MLE<br />

OCOG<br />

PLF<br />

PTF<br />

PTR<br />

RMS<br />

SKB<br />

SSB<br />

SSH<br />

Altimeter waveform products<br />

Archiving, Validation and Interpretation of Satellite Oceanographic<br />

team<br />

AIDA waveform format<br />

Digital elevation model<br />

European Centre for Medium-Range Weather Forecasts<br />

Electromagnetic bias<br />

Earth Observation Data Centre, UK<br />

European Space Agency<br />

Fast Fourier transform<br />

Geophysical data record<br />

Global Positioning System<br />

International Association of Geodesy<br />

Maximum likelihood estimation<br />

Off centre of gravity method<br />

Pulse-limited footprint<br />

Probability density function<br />

Radar system point target response<br />

Root-mean-square<br />

Skewness bias<br />

Sea state bias<br />

Sea surface height


xx<br />

SST<br />

STD<br />

SWH<br />

T/P<br />

UK-PAF<br />

Sea surface topography<br />

Standard deviation<br />

Significant wave height<br />

TOPEX/POSEIDON<br />

United Kingdom Processing and Archiving Facility


Chapter 1. INTRODUCTION 1<br />

1. INTRODUCTION<br />

This dissertation is concerned with the intersection of two topics of geodetic satellite<br />

radar altimetry in coastal regions: satellite radar altimeter waveform retracking; and<br />

some applications of the waveform retracking. This Chapter will introduce both the<br />

background and the justification for the dissertation.<br />

1. 1 Applications of Satellite Radar Waveforms<br />

Radar altimeters have been flown on a number of satellites (GEOS-3, SEASAT,<br />

GEOSAT, TOPEX/POSEIDON (T/P), ERS-1 and 2, JASON-1, and ENVISAT) and<br />

been found to give useful information about the oceans, such as the large-scale<br />

movement of water and their total mass and volume (e.g., Marth et al., 1993; Chelton<br />

et al., 2001). The basic altimetric measurement is the vertical distance or range<br />

between the satellite and the at-nadir instantaneous sea surface. This range<br />

measurement, when combined with precise knowledge of the satellite orbit,<br />

propagation effects, and the Earth’s geopotential field, allows the determination of<br />

sea surface slopes and hence surface currents. The altimeter range measurements are<br />

determined from the time interval between the time that the radar pulse is transmitted<br />

and the time that the return reflected from the (assumed at-nadir) mean sea surface is<br />

received. Satellite radar altimeters transmit pulses at a certain frequency (e.g.,<br />

1,020 Hz for ERS-2 and 1,700 Hz for POSEIDON). The on-satellite processor<br />

averages the reflected radar returns to give the waveform that relates to the shape of<br />

the sea surface, which in turn contains information on currents, tides and ocean<br />

bottom features related to oceanography and solid Earth geophysics (e.g., Tokmakian<br />

et al., 1994)<br />

1.1.1 Description of the Satellite Radar Altimeter Waveforms<br />

Pulsed-limited radar altimeters produce a measurement of power as a function of<br />

time (e.g., Marth et al., 1993; Quartly and Srokosz, 2001; Chelton, 2001). From this<br />

temporal profile of received power, which is referred to as the altimeter waveform<br />

(described in detail in Section 2.2.2), the range to the at-nadir sea surface can be<br />

determined over oceans. In addition to the range, the waveform also indicates the


Chapter 1. INTRODUCTION 2<br />

reflectivity (estimated from the amount of power in the reflected pulse) and the largescale<br />

roughness of the scattering surface (estimated from the slope of the leading<br />

edge of the waveform). Of these, only the range will be considered in this<br />

dissertation as a geodetic parameter with a view to developing an improved range<br />

determination in coastal regions.<br />

The range is measured by the altimeter indirectly through a direct measurement of<br />

the two-way travel time between the transmission and reception of a radar pulse. The<br />

observed time delay can be converted into a range measurement as long as the<br />

propagation velocity is known. A determination of this velocity, with corrections for<br />

the instrument (e.g., the Doppler shift, antenna gain pattern, and automatic gain<br />

control (AGC) attenuation), the atmospheric refraction (e.g., wet and dry troposphere,<br />

the ionosphere), the sea-state bias, and geophysical adjustments (e.g., ocean tides and<br />

atmospheric pressure loading), is an essential part of radar altimetry (Chelton et al.,<br />

2001).<br />

The profile of the backscattered power (i.e., the waveform) is a function of range<br />

measured by the altimeter. It is the convolution of three terms based on specular<br />

reflection from the sea surface: the system’s point target response (PTR), the impulse<br />

response of the smooth spherical scattering surface, and the ocean-surface height<br />

distribution (e.g., Brown, 1977; Hayne, 1980; Rodriguez, 1988; Jensen, 1999). If the<br />

ocean waves on the sea surface are assumed to be linear, the corresponding statistics<br />

of surface elevation and slopes are Gaussian. This means that the ocean-surface<br />

height distribution is assumed symmetric about some mean value. Therefore, the<br />

altimeter waveform is an odd function relative to the midpoint (i.e., half-power point)<br />

on the leading edge of the waveform, and the range to the at-nadir sea surface<br />

corresponds to this midpoint of the leading edge. The Brown (1977) model is thus<br />

the general function used by the on-satellite data processor (Chelton et al., 2001).<br />

Figures 1.1 and 1.2 show a schematic altimeter waveform based on the above<br />

theoretical description and an observed ERS-2 waveform, respectively. As can be<br />

seen, the waveform consists mainly of three parts (Figure 1.1): the thermal noise,<br />

leading edge and the trailing edge (see Section 2.2.2 for the details), and recorded in<br />

64 range bins or gates (Figure 1.2). From Figures 1.1 and 1.2, there are differences


Chapter 1. INTRODUCTION 3<br />

between the modelled and actual waveforms. The observed waveform shows wraparound<br />

noise at the first and the last five bins, and noise along the trailing edge.<br />

Noise also affects the whole waveform. Due to these differences, it is hard to define a<br />

tracker point that is related to the mid-point on the leading edge of the waveform<br />

tracked by an altimeter. Therefore, it is replaced by a tracking gate, which is<br />

designed before the satellite is launched and located at bin 32.5 for ERS-1/2. The<br />

onboard tracking algorithm (e.g., Brown model) tries to centre the leading edge at the<br />

position of this gate. The ERS leading edge usually contains 3-4 range bins (Figure<br />

1.2).<br />

Returned Power<br />

Thermal Noise<br />

Leading Edge<br />

Trailing Edge<br />

Time or Bins<br />

Figure 1.1 Schematic altimeter mean return waveform over sea surfaces.<br />

1400<br />

1200<br />

1000<br />

Power (counts)<br />

800<br />

600<br />

400<br />

200<br />

0<br />

0 8 16 24 32 40 48 56 64<br />

Bins or Gates<br />

Figure 1.2 An observed ERS-2 waveform over deep oceans.<br />

(1 bin/gate = 0.45 m and tracking gate = 32.5).


Chapter 1. INTRODUCTION 4<br />

1.1.2 General Applications of Waveform Data<br />

The above-mentioned satellite radar altimetric tracking theory has allowed good<br />

estimates of geophysical and oceanographic parameters to be obtained over open<br />

ocean surfaces (e.g., Chelton et al., 2001; Hayne, 1980). The waveform shape based<br />

on the convolution form (e.g., Brown, 1977), especially the sharply rising leading<br />

edge of the waveform (see Figure 1.2), is the basis for the precise range estimation.<br />

Analysing the waveforms provides estimates of oceanographic parameters, which<br />

include the two-way travel time of the pulse (to give the range), the slope of the<br />

leading edge of the returned waveform (to give the significant wave height), and the<br />

backscattered power σ 0 (to relate empirically to the wind speed). These estimates<br />

are obtained by the on-satellite processing of the radar return based on the tracking<br />

algorithm of the Brown (1977) ocean statistics model (e.g., Tokmakian et al., 1994;<br />

Marth et al., 1993; Fernandes, 2003; Chelton et al., 2001). Thus, waveforms are the<br />

fundamental measurements of a radar altimeter.<br />

Although the waveform recorded in 20 Hz format is a basic measurement of the<br />

radar altimetry, the sea surface height (SSH) derived from the range and satellite<br />

altitude above an ellipsoid reference is generally used in geodetic and oceanographic<br />

investigations. Usually, the SSH is averaged to 1 Hz for a typical oceanographic<br />

application (e.g., ocean circulation) and 2 Hz (or a higher rate) for geodetic and<br />

geophysical applications (e.g., gravity anomalies) over an along-track distance per<br />

second (e.g., ~7 km for T/P). However, the 20 Hz waveform data have also found<br />

important applications both in oceanography and geophysics by applying a data postprocessing<br />

technique known as waveform retracking, which extracts the geophysical<br />

parameters from the waveform data using empirical, semi-empirical or physically<br />

based algorithms (e.g., Quartly and Srokosz, 2001; Chelton et al., 2001).<br />

Over oceans, the modified convolutional representation allows for non-linear wave<br />

parameters, such as skewness of the surface distribution, to be derived from the<br />

shape of the waveform (e.g., Barrick and Lipa, 1985; Hayne, 1980; Rodriguez, 1988;<br />

Tokmakian et al., 1994). The effects of atmospheric liquid water (both clouds and<br />

rain) on the waveform has been investigated using the waveform data over open<br />

oceans by Walsh et al. (1984), Quarlty et al. (1998), and Tournadre (1998).


Chapter 1. INTRODUCTION 5<br />

Satellite radar altimeters have principally been launched to study ocean variability<br />

and the marine geoid using measurements from repeated orbits, and map ocean<br />

gravity using measurements from geodetic missions (e.g., GEOSAT and ERS-1<br />

and 2). However, since Robin (1966, cited in Wingham et al., 1993) first recognised<br />

the potential of satellite altimetry for ice-sheet topography mapping, it has also been<br />

shown that many applications exist over non-ocean surfaces, such as digital terrain<br />

modelling and polar ice sheet topographic mapping in the field of ice dynamics<br />

(Berry et al., 1998; Bamber et al., 1997; Bamber, 1994; Brenner et al., 1997; Davis,<br />

1995; Partington et al., 1991; Remy et al., 1989; Remy and Minster, 1993; Frey and<br />

Brenner, 1990). Over non-ocean surfaces, the situation is more complex, less<br />

predictable, and the nature of return echoes varies widely as a function of the surface<br />

type. The data post processing involves range estimate refinement procedures (i.e.,<br />

waveform retracking) and a slope correction to provide the correct range for the<br />

corresponding ground location.<br />

Waveform retracking over non-ocean surfaces aims to determine the offset between<br />

the tracking gate derived from the on-satellite software and a known, fixed,<br />

instrument-independent position on the leading edge of the waveform and correct the<br />

satellite range measurement accordingly (e.g., Martin, et al., 1983; Davis, 1995;<br />

Bamber, 1994; Zwally, 1996). The slope-induced error correction aims to reduce the<br />

altimeter range between satellite and the closest point to that at nadir (e.g., Brenner et<br />

al., 1983; Cooper, 1989). Therefore, several non-linear range estimation algorithms<br />

have been developed mainly over ice since 1983, such as the β-parameter retracking<br />

algorithm (Martin et al., 1983), the threshold retracking method (Wingham et al.,<br />

1986), and the surface/volume scattering-retracking algorithm (Davis, 1993).<br />

These algorithms provide an important monitoring ‘tool’ for accurate estimates of<br />

growth or shrinkage of the ice sheets (e.g., the Greenland and Antarctic ice sheets).<br />

Wingham et al. (1998) indicate that the elevation of the Antarctic ice sheet fell by<br />

0.9±0.5 cm/year between 1992 and 1996. Information on the characteristics and<br />

changes of the firn (compacted snow), including surface roughness and volume<br />

scattering (e.g., Partington et al., 1989; Davis, 1993; Legresy and Remy, 1998), and<br />

information on the ice-sheet topography, including the surface melt streams and


Chapter 1. INTRODUCTION 6<br />

slopes (e.g., Phillips, 1998, cited in Zwally and Brenner, 2001) can also be derived<br />

from radar altimetry.<br />

Altimetry over land is even more complicated owing to rapidly varying topography.<br />

The echo waveform over land comprises reflections from an essentially static, nonhomogeneous<br />

surface. It contains various types of cover and each shows different<br />

characteristics. Some of the methods used over ice surfaces could be applied after<br />

proper modification. For example, Bamber et al. (1997) use a modified threshold<br />

method, with an optimal power threshold. Berry et al. (1998) develop an expert<br />

system to retrack waveforms over land from the ERS-1 mission by selecting one of<br />

the retrackers, which were designed to optimise the determination of the individual<br />

range correction. The ‘ice-mode’ operation of the ERS-1/2 enables the track to<br />

maintain lock over land on ~80% of surfaces (Berry, 2000a). Using a variety of<br />

different tracker algorithms, Berry (2000b) has determined land elevations, achieving<br />

a good repeatability between different passes. This has in turn revealed the<br />

inconsistencies in various digital elevation models (DEMs) obtained from<br />

conventional mapping techniques (e.g., Hilton et al., 2002).<br />

1. 2 Problems in Coastal Areas<br />

Modern satellite radar altimeters can measure the instantaneous SSH to a precision of<br />

approximately 5 cm in the open oceans (e.g., Shum et al., 1998; Chelton et al., 2001).<br />

However, in coastal regions, the waveform measurements are affected by the noisier<br />

radar returns from the (generally rougher) coastal sea states and simultaneous returns<br />

from the land (e.g., Brooks and Lockwood, 1990; Brooks et al., 1997; Nerem, 1995;<br />

Mantripp, 1996), and by less reliable geophysical, wet delay, orbit and instrument<br />

corrections (e.g., Shum, 1998; Andersen and Knudsen, 2000; Chelton et al., 2001;<br />

Fernandes et al., 2003; Quartly and Srokosz, 2001). Poorly modelled tides also affect<br />

the data owing to the shallow depth of the ocean and the irregular shape of the<br />

shoreline. This prevents the altimeter data from providing valuable information on<br />

the geoid shape, tides, the wind field, current characteristics, and dynamic the sea<br />

surface topography in coastal regions.


Chapter 1. INTRODUCTION 7<br />

1.2.1 Topographic Effects on Waveforms<br />

The shape of the radar return waveform can also be significantly affected in coastal<br />

regions by varying coastal topography, such as cliffs, shallow water tides, or the<br />

solid ground that is illuminated together with the sea surface. It is well known that<br />

the altimeter range may be estimated poorly by on-board tracking software in coastal<br />

regions (e.g., Nerem, 1995; Deng et al., 2002; Mantripp, 1996). As stated, oceanic<br />

waveform models are given under assumptions of scattering surface statistical<br />

homogeneity. However, this spatial uniformity of the topographic statistics and<br />

microwave properties of the surface cannot be met in coastal regions (Mantripp,<br />

1996; Quartly and Srokosz, 2001).<br />

In the proximity of the coastline, the altimeter simultaneously views both scattering<br />

from water and land surface (Figure 1.3). These two scattering surfaces may have<br />

different elevations, but the shortest range to the altimeter depends on the distance to<br />

the shoreline and the slope and reflectivity of the land (Brooks, et al., 1997; Brooks,<br />

2002). The nadir range between the altimeter and ocean surfaces may be the same as<br />

the slant range from land to the altimeter within the altimeter footprint. But as the<br />

satellite gets closer to the coastline, the slant range to the altimeter reflected from<br />

higher (or brighter) land may be shorter than that from the nadir oceans. Since the<br />

altimeter measures the shortest range to the reflecting surfaces, the land return within<br />

the footprint can contribute to the received power by the altimeter which then affects<br />

the returned waveform. The received power from different reflecting surfaces is<br />

dependent upon their respective backscattering coefficient and the angle at which<br />

they are being illuminated. These waveform features can be difficult to interpret<br />

using the Brown (1977) model. As a result, waveforms will be distorted by land<br />

return effects in coastal regions (Chapter 4).<br />

In addition to the land return effects, calm (usually brighter) water surfaces (both<br />

enclosed sea and inland water surfaces) near the coastline, such as bays and estuaries,<br />

make the returned power higher and narrower, thus contaminating the altimeter<br />

waveform as well. It will be shown in this research that land topography affects the<br />

waveform though the radar altimeter’s operation when crossing the coastline<br />

(Section 4.6). The altimeter changes its tracking mode when flying land-to-water or


Chapter 1. INTRODUCTION 8<br />

water-to-land, but data or signal reacquisition may take a second or more of time. As<br />

a result, land return effects on waveforms will last longer near the coastline. The<br />

effects of this performance on the waveform should also be considered when<br />

processing altimeter waveform data in coastal regions.<br />

Land<br />

Ocean<br />

Figure 1.3 Altimeter simultaneously illuminates land and ocean near the coastline<br />

The altimeter-derived SSHs are affected by the above factors. The difference<br />

between the SSH after applying all geophysical and environmental corrections and<br />

AUSGeoid98 geoid height along four ERS-2 collinear ground tracks from cycles 42<br />

and 43 is plotted in Figure 1.4. It shows that there is a large uncertainty in the<br />

altimeter-derived SSH data with respect to the geoid when approaching the coastline.<br />

The magnitude of the difference increases closer to the coastline. Since altimeterderived<br />

SSHs should provide the same geoid structure, but different noise<br />

components from the repeat observations, this large data scatter with respect to the<br />

geoid clearly shows the general problem that exists with the altimeter range data


Chapter 1.INTRODUCTION 9<br />

6<br />

5<br />

Cycle42<br />

Cycle43<br />

SSH -Geoid (m)<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0.06 3.07 6.37 9.42 11.62 14.25 16.76 19.30 22.04 24.89 27.81<br />

-1<br />

Distance from the coastline (km)<br />

15<br />

Cycle43<br />

Cycle42<br />

10<br />

SSH - Geoid (m)<br />

5<br />

0<br />

0.19 3.40 6.60 9.82 12.76 15.56 18.47 21.48 24.57 27.70<br />

-5<br />

-10<br />

Distance from the coastline (km)<br />

6<br />

5<br />

Cycle42<br />

Cycle43<br />

SSH - Geoid (m)<br />

4<br />

3<br />

2<br />

1<br />

0<br />

10.60 13.89 16.25 17.65 19.51 21.71 24.15 26.77 29.53<br />

Distance from the coastline (km)<br />

15<br />

10<br />

Cycle42<br />

Cycle43<br />

SSH - Geoid (m)<br />

5<br />

0<br />

0.07 3.17 6.38 9.64 12.38 15.22 18.16 21.20 24.30 27.45<br />

-5<br />

-10<br />

Distance from the coastline (km)<br />

Figure 1.4 Differences between the ERS-2 SSH and AUSGeiod98 geoid height along<br />

four collinear repeat tracks from cycles 42 and 43 at the Southern Australian coast,<br />

showing the metre-level uncertainty of SSHs closer to the coastline.


Chapter 1.INTRODUCTION 10<br />

from the on-satellite tracking algorithm in coastal regions. Therefore, waveforms can<br />

be contaminated by the land returns and sea states in coastal regions. This is the<br />

definition of the term contamination used in this dissertation.<br />

1.2.2 Topographic Effects on Geophysical Corrections<br />

In the vicinity of the coastline, land topography also affects the modelling of some of<br />

the geophysical corrections applied to the range measurement (e.g., Anderson and<br />

Knudsen, 2000; Shum et al., 1998; Fernandes et al., 2003). Of all corrections, the<br />

effects of wet tropospheric delay and the ocean tides on the altimeter data are most<br />

critical (e.g., Anzenhofer et al., 2000; Chelton, 2001). The result is that the accuracy<br />

of the altimeter-derived SSH will also degrade with the degradation of these<br />

geophysical corrections.<br />

For ERS-1/2 altimeters, the wet tropospheric corrections are derived from the actual<br />

measurements made by the onboard microwave radiometer (e.g., Anzenhofer et al.,<br />

2000; Fernandes et al., 2003). The corrections contained in the waveform data<br />

products are also calculated using an estimate of the total integrated water vapour<br />

and surface air temperature at the sub-satellite point using data supplied by the global<br />

forecast analysis in the European Centre for Medium-Range Weather Forecasts<br />

(ECMWF) (NRSC, 1995). However, the radiometer does not work correctly near the<br />

land and cannot switch on or off when the satellite flies across the coastline, causing<br />

some useful data to be unavailable and errors in the range measurements in coastal<br />

regions (e.g., Fernandes et al., 2003; Anzenhofer et al., 2000).<br />

Due to the spatial complexity of tides in shallow water, the global ocean tide models<br />

are still not accurate enough over shallow seas for detailed oceanographic studies.<br />

The error resulting from using global ocean tidal models can be at the decimetre<br />

level in coastal regions (Anzenhofer et al., 2000; Shum et al., 1998). For both<br />

ERS-1/2 altimeter waveform data and geophysical data records (GDRs), the ocean<br />

tide corrections are provided in a format of 1 Hz and based on the Schwiderski (1983,<br />

cited in NRSC, 1995) model rather than on the recently improved tidal models<br />

(NRSC, 1995). Since the ERS-1/2 satellite orbit is sun-synchronous and the choice of<br />

35 days for its main repeat period leads to severe aliasing of the most important tidal<br />

constituents, the use of the Schwiderski ocean tidal model is appropriate for the


Chapter 1.INTRODUCTION 11<br />

ERS-1/2 data. The problem is that the model cannot provide an adequate resolution<br />

for shallow water tidal constituents. This is because tides over shallow waters<br />

become highly complicated and vary rapidly in amplitude and phase. In these areas,<br />

tides are highly dependent on the bathymetry, the shape of the continental shelf, and<br />

the regional tidal regime (Andersen, 1999). In addition, shorter wavelength tidal<br />

features near sharply topographical changes are not represented in the current<br />

dynamical and empirical global tidal models (Tierney et al., 1998).<br />

Although relatively less research has been reported, the sea state bias (SSB)<br />

correction, which is a function of the significant wave height (SWH), is<br />

unfortunately affected by the land topography near the coastline. Land affecting the<br />

SWH and thus the SSB correction for GEOSAT altimeter has been reported by<br />

Fernandes et al. (2003). It will be shown in Chapter 7 that the erroneous SSB<br />

corrections have also been found for ERS-1 range measurements in the Taiwan Strait<br />

approaching the coastline.<br />

1. 3 Justifications for This Study<br />

1.3.1 Previous Work<br />

As mentioned above, improved determination of the SSH data through waveform<br />

retracking in coastal regions is essential because the coastal topography and the sea<br />

states contaminate the altimeter waveform (cf. Section 3.2.1). However, the potential<br />

of the altimeter waveform data in coastal regions has been the subject of<br />

comparatively little study. So far, considering the existing retracking algorithms used<br />

for coastal waveforms from the open literature, they are those developed for dealing<br />

with the waveform data over ice-sheet surfaces (cf., Anzenhofer et al., 2000; Brooks<br />

et al., 1997).<br />

One earlier study by Brooks and Lockwood (1990) indicates that the presence of land<br />

within a radar altimeter’s footprint affects the altimeter’s tracking performance and<br />

thus contaminates the altimeter waveform data. They analysed GEOSAT waveform<br />

data over the Tuamotu Archipelago in the southern Pacific Ocean to determine the<br />

contaminated distance (1.0-6.2 km) from the islands. Their research pointed out that<br />

the island can locally affect the SSH by more than ±4 m. An edit criterion based on a


Chapter 1.INTRODUCTION 12<br />

combination of waveform data, automatic gain control (AGC) and SWH was<br />

suggested, but waveform retracking was not involved. Brooks et al. (1997) retracked<br />

TOPEX waveforms along 18 satellite ground passes in Pacific Rim coastal zones<br />

using the threshold retracking technique (cf. Sections 3.5.2 and 3.7). Anzenhofer et al.<br />

(2000) present a retracking system in coastal regions, which consists of only a single<br />

retracker and a fitting algorithm is employed. The five-β-parameter fitting function<br />

with either a linear trailing edge or exponential decay trailing edge (cf. Section 3.5.1)<br />

is used by Anzenhofer et al. (ibid.) to retrack ERS-1 waveforms.<br />

Using only a single retracker, however, limits the precision of the recovered SSH<br />

data because of various levels of complexity of the waveform data in coastal regions<br />

(see Sections 5.8.2 and 6.4). Therefore, developing an alternative approach to<br />

improving the precision of recovered SSH data in these areas is necessary, which<br />

must be able to cope with diverse waveforms and estimate precise SSHs from these<br />

reprocessed data. This is one of the justifications for this dissertation.<br />

1.3.2 Significance of the Improvement of Altimeter Data in Coastal Regions<br />

Over open oceans, satellite altimetry is one of the most effective techniques that<br />

provides global and accurate homogeneous coverage of the shape of the sea surface.<br />

The shape of the sea surface is the only physical variable measured from space that is<br />

directly and simply connected to large scale movement of water and the total mass<br />

and volume of the ocean. In coastal regions, although there are generally other data<br />

sources available, such as tide gauges, airborne and shipborne gravity data sets,<br />

altimetry is still a significant technique that is able to provide reliable and uniform<br />

surface information. Since the launch of the SEASAT satellite, there has been a<br />

demand for combining the uniform altimetric coverage of data with other data<br />

sources in order to investigate the oceanic and climatic environment. As stated,<br />

however, the precision of the altimeter data is degraded because of the coastal<br />

topography (both land and sea).<br />

As new satellite missions, such as GFO-1, ENVISAT, JASON-1, and CRYOSAT<br />

will contribute more to the existing altimeter data sets of SEASAT, GEOSAT,<br />

ERS-1/2 and T/P, applications of altimetry in coastal regions will continue to grow.<br />

The coastal areas represent eight percent of the ocean surface. Most importantly,


Chapter 1.INTRODUCTION 13<br />

economic and engineering activities are mainly concentrated in these areas, such as<br />

fisheries and offshore oil drilling (Menard et al., 2000). Altimeter data are also<br />

important for understanding coastal tides, interactions between coastal currents and<br />

open ocean circulation, mesoscale and coastal variability, geoid modelling, and<br />

marine gravity anomaly determination. However, this could only be achieved by<br />

specialised improvement of altimeter measurements (both the data themselves and<br />

the corrections) in coastal regions and accurate data sampling through merging, in an<br />

optimal way, different altimeter data sets such as T/P and ERS-1/2, or JASON-1 and<br />

ENVISAT in the future.<br />

After declassification of the GEOSAT geodetic mission, cross-track resolution of the<br />

satellite altimeter profiles has been improved considerably. In particular, the crosstrack<br />

resolution is improved further because of other satellite altimeter missions,<br />

especially the ERS geodetic missions. Consequently, the cross-track resolution has a<br />

trend exceeding the along-track resolution (~7 km). ERS and GEOSAT geodetic<br />

missions, for instance, have cross track spacings of ~8 km and ~5 km at the equator,<br />

respectively (Chelton et al., 2001). Thus, it becomes necessary to keep and improve<br />

20 Hz data for refining the along-track resolution not only in open oceans, but also in<br />

coastal regions.<br />

When compared ERS-1 with GEOSAT altimeters, the ERS-1 range is poorly tracked<br />

by the altimeter. This is because ERS-1 was not a geodetic satellite and had its<br />

geodetic mission added at the end of its life. The prime use of its data was to monitor<br />

the trailing edge of the waveform and then to monitor the air-sea interactions. The<br />

number of the gates along the leading edge is usually 3-4 sample bins. This makes<br />

difficult for the on-satellite tracking algorithm to precisely define the waveform<br />

shape and thus the range to the surface. Instead, GEOSAT has twice the number of<br />

gates along the leading edge to estimate the range more precisely (Fairhead et al.,<br />

2001). Fairhead et al. (2001) conclude that the ERS-1 ranges have been poorly<br />

estimated by the on-satellite tracking algorithm.<br />

In addition, ocean surface waves are a fundamental limitation to the recovery of the<br />

gravity field from altimeter SSH profiles. For instance, to recover the gravity field at<br />

an accuracy of 1 mgal at 20-km full wavelength requires the ocean surface height


Chapter 1.INTRODUCTION 14<br />

change over a 10-km horizontal distance to a precision of 1 cm (Sandwell, 2003;<br />

Fairhead et al., 2001). However, standard on-board data processing of altimeter data<br />

provides a SSH precision of only ~5 cm at this length scale (Chelton et al., 2001;<br />

Sandwell, 2003). The results from Fairhead et al. (2001) show that reprocessing<br />

ERS-1 waveforms significantly reduces the along-track noise particularly in the 50-<br />

10 km wavelength range, thus improving spatial resolution. This also suggests the<br />

improvement of geodetic parameter estimation from waveform retracking is<br />

necessary over both coastal regions and open oceans.<br />

The present theory of altimetry, developed to describe scattering from the ocean<br />

surface or to retrack the return waveform from the non-ocean surface, does not deal<br />

properly with the waveforms in coastal areas. Data post-processing over topographic<br />

surfaces of ice and land have been attended to since 1983 (e.g., Martin et al., 1983;<br />

Partington et al., 1989; Wingham et al., 1986; Davis, 1995; Zwally, 1996; Berry,<br />

2000a), but there is still a comparable lack of post-processing in the coastal areas<br />

(see Section 1.3.1 and Chapter 3). Therefore, in order to fully realise the potential of<br />

the altimetry in coastal areas, it is necessary to improve the altimeter measurements<br />

and obtain a better understanding of the SSH estimates derived from satellite radar<br />

altimetry. Accurate measurements in coastal regions would quantify one important<br />

source of the SSH and thus enable better understanding of the various geophysical<br />

and oceanographical features in coastal regions. This requires more sophisticated<br />

correction models and data processing techniques than in the open oceans and on<br />

land.<br />

Finally, the International Association of Geodesy (IAG) has created a special study<br />

group (3.186) for studying altimetry data processing for gravity, geoid and sea<br />

surface topography determination (http://space.cv.nctu.edu.tw/IAG/main.html). The<br />

subject of this study has been listed as one of the research problems in the group. The<br />

Australian Research Council International Researcher Exchange Scheme (grant<br />

number: X00001267) also supported a project related to this study. Therefore, the<br />

results will contribute significantly not only to altimeter applications, but also to the<br />

altimetric data processing theory and practice. Essentially, satellite altimeter tracking<br />

will realise its potential over the whole Earth’s surface that is covered by the


Chapter 1.INTRODUCTION 15<br />

altimeter orbits, rather than ignoring eight percent of the Earth’s surface in coastal<br />

regions.<br />

1. 4 The Study Areas<br />

The study areas used to demonstrate the retracking system are the Australian coastal<br />

regions and the Taiwan Strait. They are associated with different coastal<br />

characteristics.<br />

As well as for geographical convenience to the author, Australia is chosen for this<br />

study for the following reasons. Firstly, it is surrounded by part of three large,<br />

interconnected ocean basins of the Southern Hemisphere: the Pacific, Indian and<br />

Southern Oceans. Secondly, the Australian coastline (including Tasmania), is<br />

approximately 36,700 kilometres in length (Zann, 1990, cited in Resource<br />

Assessment Commission, 1993), which is one of the longest coastlines in the world<br />

for a single country. Finally, Australia’s coastal zone exhibits the following broad<br />

range of topographical features: coastal plains, hills and mountains, beaches, cliffs,<br />

estuaries, oceans, coral reefs, and islands. Thus, it is assumed that the performance of<br />

altimetry around Australia is reasonably representative of what can be expected in<br />

other parts of the world.<br />

Although the Australian coast provides an almost perfect area for the investigation of<br />

waveform retracking, it is not an area for the application of recovery of gravity<br />

anomalies. In 1992, the Australian Geological Survey Organisation (AGSO, now<br />

Geoscience Australia) released a gravity database that includes denser ship-track<br />

gravity observations around Australia bound by 108°E ≤ λ ≤ 162°E and<br />

8°S ≤ ϕ ≤48°S. However, these AGSO marine gravity data were not provided in a<br />

format such that the crossover adjustment can be conducted to reduce some serious<br />

biases among overlapping tracks (cf. Featherstone, 2003). They were also not<br />

crossover adjusted before supply (ibid.). Therefore, they cannot offer any definitive<br />

indication on the improvements of SSH data through the accuracy of the satellite<br />

altimeter-derived gravity anomalies in the Australian coastal region.<br />

Instead, another study area chosen is the narrow Taiwan Strait (roughly ~110 km<br />

wide), which links the East China Sea to the South China Sea. This area is chosen


Chapter 1.INTRODUCTION 16<br />

because firstly there are some on-going research activities, detailed and reliable shiptrack<br />

gravity anomalies, and bathymetric data available in the area (e.g., Hwang and<br />

Wang, 2002). Secondly, it is a shallow water region with high tidal amplitudes,<br />

which allows the evaluation of the retracking system in a different coastal area.<br />

Thirdly, ~2.5 years of ERS-1 waveform data including all phases provide denser<br />

measurement coverage over the area. This will benefit the recovery of the gravity<br />

field there.<br />

1. 5 Thesis Outline<br />

Chapters 1 and 2 introduce the background of this dissertation and the geodetic<br />

principles of satellite altimetry. The altimeter data products, waveform data sets and<br />

some external input data sources used in this study will be summarised and described<br />

briefly in Chapter 2. In Chapter 3, the existing algorithms developed to retrack<br />

altimeter waveforms over ocean surfaces and non-ocean surfaces will be reviewed<br />

and discussed in the context of coastal retracking. The ocean model, which will be<br />

used in this study, will be deduced, and a detailed qualitative analysis of the model<br />

will be carried out.<br />

Chapter 4 quantifies the contaminated distance around Australian coast using 20 Hz<br />

waveform data of ERS-2 (one cycle, 35-day repeated orbit) and POSEIDON (five<br />

cycles, 10-day repeated orbit). Several examples from ERS-2 waveforms along<br />

satellite ground tracks near the coastline are analysed in detail to understand how<br />

waveforms are contaminated and what the characteristics of contaminated<br />

waveforms are at the Australian coast (Section 4.6). A preliminary test that estimates<br />

the land effects on waveforms in the vicinity of land will be given in Section 4.7.<br />

Chapter 5 develops a coastal retracking system that is based upon a detailed preanalysis<br />

of the coastal waveforms given in Chapter 4. The retracking algorithms used<br />

will be the fitting and threshold methods. Issues involving the fitting algorithms are<br />

fitting functions of an ocean and β-parameter models, the least squares iterative<br />

procedure, linearisation of the parameters, determination of the initial estimates of<br />

parameters, and the weight scheme. A method will be introduced that uses iterative<br />

weights to detect the outliers in the waveforms ensuring the effective convergence of<br />

the fitting procedure. For the threshold retracking algorithm, a method will be


Chapter 1.INTRODUCTION 17<br />

described that selects an appropriate threshold level for coastal waveforms. Finally, a<br />

coastal retracking system will be developed, which makes it possible to improve the<br />

SSH precision from reprocessing the altimeter waveforms, correcting the range<br />

measurements, and extracting the precise SSH data from corrected ranges.<br />

In Chapter 6, the coastal retracking system developed in Chapter 5 is applied to the<br />

Australian coast (0-30 km from the coastline) using two cycles (42 and 43, March to<br />

May 1999) of ERS-2 20 Hz waveform data. Results of waveform classification and<br />

waveform retracking will be presented and discussed in Sections 6.2 and 6.3,<br />

respectively. The internal validation of the quality of the retracked altimeter data will<br />

be conducted by analysis of the single track SSH data in Section 6.4, comparisons<br />

between SSH data before and after retracking in Section 6.5, and a collinear analysis<br />

in Section 6.6. Using an external (partly) independent reference of the AUSGeoid98<br />

geoid model, the results of retracking and comparisons with the unretracked and<br />

external control data will be discussed in Sections 6.4 and 6.7 to demonstrate the<br />

improvements from waveform retracking.<br />

Chapter 7 investigates the improved accuracy of the altimeter-derived coastal marine<br />

gravity field from SSH data after waveform retracking in the coastal region of the<br />

Taiwan Strait. Using ~2.5 years of ERS-1 20 Hz waveform data (


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 18<br />

2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES<br />

2. 1 Introduction<br />

Satellite altimetry employs a nadir-pointing, high-resolution radar to measure the<br />

range to the ocean’s surface with an accuracy of a few centimetres. The range to the<br />

at-nadir sea surface is measured by the altimeter directly through a measurement of<br />

the time difference between the transmission of a radar pulse and the reception of the<br />

echo. If the altitude of the satellite above a reference ellipsoid is known, the SSH<br />

above the reference ellipsoid (Figure 2.1), which is the measurement expected from<br />

the satellite altimetry and corresponds to the geophysical characteristics of the sea<br />

surfaces, can be calculated by subtracting the range from the satellite altitude. In<br />

order to precisely estimate the SSH, it is necessary to know the principle of the<br />

altimeter’s operation, estimates of the range and then the SSH, and the error sources<br />

for the determination of the SSH. These will be described next.<br />

Figure 2.1 Basic measurement principle of the altimeter-derived SSH (From<br />

http://www.jason.oceanobs.com/html/portail/galerie<br />

/banque_img_welcome_uk.php3).


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 19<br />

In addition, Section 2. 3 will briefly summarise the waveform data sets used in this<br />

study. Some external input data sources will be described in Section 2. 4.<br />

2. 2 General Information on Construction of Altimeter Measurements<br />

The altimeter range measurement allows the determination of the SSH related to a<br />

reference ellipsoid and hence oceanographic, geophysical and geodetic applications,<br />

such as the determination of gravity anomalies, mesoscale eddies, Rossby waves and<br />

the surface currents under an assumption of geostrophy (Chelton et al., 2001). Unlike<br />

synthetic aperture radar, radar altimeters do not form an image observation, but make<br />

point measurements at or close to the nadir of the satellite as it follows its orbit. The<br />

coverage of the ground tracks is built up over time as the orbit arcs cover the Earth’s<br />

surface. In addition, the shape of the return pulse has been found to give information<br />

on SWH, while the backscattered power has been used to estimate the surface wind<br />

speed (ibid.).<br />

2.2.1 Fundamental Radar Measurement Principles<br />

The altimeter-observed time delay can be converted into a range to the at-nadir sea<br />

surface as long as the propagation velocity is known as<br />

ct<br />

R = (2.1)<br />

2<br />

where R is the range, c ≈ 3×10 8 m s -1 is the free-space speed of light and t is the<br />

two-way travel time of the radar pulse. The range resolution is given by<br />

c<br />

∆ R ≥<br />

τ<br />

(2.2)<br />

2<br />

where τ is the pulse length. The resolution of one discrete measurement of range<br />

depends on the altimeter’s operating parameters (Table 2.1). For instance, the<br />

resolution is approximately 0.45 m for the ERS-1/2 altimeter operating in an ‘ocean<br />

mode’ with a pulse length τ = 3.03 ns. In order to measure dynamic ocean signals,


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 20<br />

Table 2.1 Altimeter Operating Parameters<br />

Altimeter<br />

Launch<br />

date<br />

Altitude<br />

(km)<br />

Inclination<br />

(deg)<br />

Beamwidth<br />

(deg)<br />

Frequency<br />

(GHz)<br />

PRF<br />

(Hz)<br />

No.<br />

of<br />

gate<br />

s<br />

Tracking<br />

gate<br />

Sampling time<br />

(ns)<br />

Waveform<br />

frequency<br />

(Hz)<br />

No of<br />

waveforms<br />

in average<br />

Seasat 27/06/1978 800 108 1.6 13.5 1020 60 30.5 3.125 10 100<br />

Geosat 12/03/1985 800 108 2.1 13.5 1020 60 30.5 3.125 10 100<br />

ERS-1 17/07/1991 784 98 1.3 13.8 1020 64 32.5 3.03 (ocean mode) 20 50<br />

12.12 (ice mode)<br />

ERS-2 21/04/1995 784 98 1.3 13.8 1020 64 32.5 3.03 (ocean mode) 20 50<br />

12.12 (ice mode)<br />

TOPEX (Ku) 10/08/1992 1334 66 1.0 13.6 4500 128 32.5 3.125 10 2×228<br />

TOPEX (C) 10/08/1992 1334 66 2.7 5.3 1200 128 35.5 3.125 5 4×60<br />

POSEIDON 10/08/1992 1334 66 1.1 13.65 1700 60 29.5 3.125 20 86<br />

GFO 10/02/1998 800 108 1.6 13.5 1020 128 32.5 3.125 10 100<br />

Jason-1 (Ku) 07/12/2001 1334 66 1.3 13.6 1800 104 32.5 3.125 20 90<br />

Jason-1 (C) 07/12/2001 1334 66 3.4 5.3 300 104 32.5 3.125 20 15<br />

Envisat (Ku) 01/03/2002 784 98 1.3 13.6 1800 128 3.125 18 100<br />

Envisat (S) 01/03/2002 784 98 5.6 3.2 450 64 6.25 18 25


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 21<br />

the range and timing precisions are required to a few centimetres (e.g., ~4 cm) and<br />

less than 0.2 ns, respectively (Mantripp, 1996; Chelton et al., 2001). This is achieved<br />

by averaging the estimates from several individual pulses (cf. the rightmost column<br />

in Table 2.1).<br />

2.2.2 The Returned Waveform<br />

A schematic description of the interaction of a pulse-limited radar altimeter with a<br />

diffuse, horizontal and planar sea surface is illustrated in Figure 2.2. Initially<br />

( 0 t < t0<br />

< ), a pulse of electromagnetic energy is transmitted from an on-satellite<br />

altimeter antenna, propagating in a spherical wavefront. The area of pulse that can be<br />

received on ground is defined by antenna beam-width ( θ<br />

A<br />

). When the wavefront<br />

encounters the nearest crests of those ocean waves directly beneath the satellite (i.e.,<br />

at nadir) at t = t0<br />

, it illuminates a point and a reflected signal begins to return to the<br />

altimeter. As time and the pulse progress, the wavefront reaches the surface at points<br />

further from the nadir point. This increases the area of interaction between pulse and<br />

surface, and the illuminated point spreads out rapidly to form a disc (assuming a<br />

locally flat surface). Properly reflected facets within this disc scatter energy back to<br />

the altimeter during the period of t<br />

0<br />

< t < t1<br />

(Figure 2.2). Once the rear of the pulse<br />

reaches the lowest trough at nadir, the region of interaction between the surface and<br />

pulse forms an annular ring of increasing diameter, narrowing width, and the<br />

constant mean area. The backscattered energy reaches its maximum at the point of<br />

transition to an annular ring (i.e., t = t ). Thereafter ( t > t<br />

1<br />

1<br />

), the backscattered energy<br />

begins to decay owing to the finite antenna beam width and the fewer proper<br />

reflected facets available at larger off-nadir angles. The returned power is recorded<br />

over the duration of the pulse, building up a returned waveform (see the bottom in<br />

Figure 2.2) with a rapidly rising leading edge and long decay of the trailing edge.<br />

In order to reduce the fading noise arising from coherency among individual<br />

waveforms, returned waveforms are actually an average of a number of pulses, 50 in<br />

the case of ERS-1/2 (Mantripp, 1996; Quartly et al., 2001), forming theoretically a<br />

average return altimeter waveform over ocean surface shown in Figure 1.1.


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 22<br />

SIDE VIEW<br />

transmitted<br />

pulse<br />

R<br />

θ A<br />

PLF<br />

PLAN VIEW<br />

(Illuminated area)_<br />

t 0<br />

t 1<br />

t 2<br />

RETURN WAVEFORM<br />

leading edge<br />

trailing edge<br />

average return<br />

typical return<br />

t 0 t 1 t 2<br />

time<br />

Figure 2.2 A schematic geometrical description of the interaction of a pulse and the<br />

scattering surface, and the build up of a return waveform over the duration of a pulse.<br />

Note that, an average returned waveform from 50 typical individual returns is<br />

depicted, while a single typical return shows more noise. (PLF is the pulse-limited<br />

footprint).<br />

This averaged returned waveform (Figure 1.1) is a time series of the mean returned<br />

power recorded by a satellite altimeter, and is referred to as the convolution of three<br />

terms in the time domain (e.g., Brown, 1977; Hayne, 1980; Chelton et al., 2001).<br />

Over open oceans, it can be described by the Brown (1977) model (see Chapter 3). A<br />

modelled or ideal average return altimeter waveform over ocean surface is shown in<br />

Figure 1.1. It consists mainly of three parts:<br />

(a)<br />

The thermal noise contains the thermal noise power generated by the altimeter<br />

prior to the first return of a signal from the scattering surfaces. Its effect is to<br />

add a constant power level to the return waveform. It is to be noted that the


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 23<br />

(b)<br />

(c)<br />

observed waveform of ERS-1/2 does not show the thermal noise because of the<br />

operating feature of the altimeter (see Section 4.4).<br />

The leading edge contains the return power from the scattering surfaces within<br />

the pulse-limited footprint (see below for explanation), which involves in the<br />

information about the SWH and range between the satellite altimeter and the<br />

mean sea surface at the nadir.<br />

The trailing edge consists of the return power from the scattering surface<br />

outside the pulse-limited footprint, which can be approximated by a straight<br />

line whose slope depends on the altimeter antenna pattern and the off-nadir<br />

angle (see Chapter 5).<br />

To date, all spaceborne radar altimeters use the pulse-limited (i.e., only the earliest<br />

returns reflected from nadir are recorded) operation to measure the range to the<br />

surface. The pulse-limited operation itself defines one of the diameters of the<br />

measurement footprint (Martin et al., 1983; Quartly et al., 2001). The footprint is the<br />

area of the reflecting surface which is illuminated by the altimeter, and its size<br />

depends on the satellite altitude, the width of the range window and the features of<br />

the reflecting surface (see Section 4.2.1). The pulse-limited footprint is the maximum<br />

circular area from which backscatted power can be simultaneously received (Brooks<br />

et al., 1978). The spreading of the pulse over the pulse-limited footprint corresponds<br />

to the risetime of the leading edge of the return waveform, while the range to the<br />

nadir sea surface measured by the altimeter corresponds to the midpoint at the<br />

leading edge of the waveform.<br />

In practice, a pulse compression technique (Marth et al., 1993; Chelton et al., 2001)<br />

is used to obtain a high signal-to-noise ratio and an acceptable demand on the<br />

satellite’s power system. The radar altimeter emits long-duration chirps, which are<br />

linearly frequency-modulated pulses, then the received return signal is dechirped and<br />

mixed with a deramping chirp. This mixed signal is then filtered and digitised in the<br />

frequency domain where there is a one-to-one correspondence between frequency<br />

and two-way travel time. Individual waveform samples result finally from a fast<br />

Fourier transform (FFT) procedure. The resultant signal is a function of the<br />

difference frequency, and its properties depend not only on the particular part of the<br />

nadir scattering surface, but also the characteristics of the filter frequency used in the


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 24<br />

chirp generation and the deramping processing. In this case, the power spectral<br />

density estimates at discrete frequencies (64 bins for the ERS and POSEIDON) are<br />

obtained and have a similar shape to that shown in Figure 1.1, with only the<br />

difference for the two domains that the abscissa is frequency or two-way travel time<br />

(Chelton et al., 2001). The further details on the actual implementation can be found<br />

in Marth et al. (1993), Mantripp (1996), and Chelton et al. (2001).<br />

Radar altimeters operate as closed-loop regulators in both the time-delay (range<br />

tracking) and radiometric (AGC) domains (Marth et al., 1993; Jensen, 1999). Thus,<br />

the on-satellite radar processor attempts to hold the reference position of the<br />

altimeter waveform at a fixed position and with fixed amplitude so that successive<br />

waveforms may be integrated or averaged without any loss of measurement<br />

resolution. This is conducted by adjusting the timing of the data sampling and by the<br />

attenuators within the receiver (ibid.). As a result, the processor must constantly<br />

measure the varied range to the surface and reflectivity of the scattering surface. It<br />

will be found in Section 5.4.1 (Figure 5.3) that the waveform’s leading edges are<br />

centred on the midpoint, and this remains true regardless of the surface roughness. It<br />

is this stable behaviour that makes the AGC and range tracking algorithm work.<br />

The range to the surface can vary from several metres over ocean surfaces to several<br />

kilometres over non-ocean surfaces within the diameter of the footprint (Mantripp,<br />

1996; Chelton et al., 2001). However, in the vertical dimension the altimeter can<br />

receive the returns only within a ‘range window’ specified by the width of the<br />

frequency spectrum. For instance, it is ~28 m and ~112 m for ERS-1/2 in the ocean<br />

mode (1 range bin equals ~0.45 m) and ice mode (1 range bin equals ~1.82 m),<br />

respectively, and ~30 m for POSEIDON over oceans (1 range bin equals ~0.47 m).<br />

The onboard tracker continuously adjusts the range window to keep the leading edge<br />

of the waveform at a specified position at the centre of the range window. This<br />

position is known as the ‘tracking gate’ (or tracking point), which is designed prior to<br />

the launch of the satellite and is the range bins 32.5 for ERS-1/2 and 31.5 for<br />

POSEIDON. When the surface elevation varies too rapidly for the tracker to respond,<br />

the altimeter is said to have ‘lost lock’. Usually, the altimeter cannot track surfaces<br />

where the mean slope is greater than about one degree (Barrick and Lipa, 1985;<br />

Mantripp, 1996).


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 25<br />

2.2.3 The Range and Sea Surface Height<br />

If the altimeter tracker and AGC circuitry worked perfectly, the position of the<br />

tracking gate would always be centred halfway up the return waveform leading edge<br />

at the tracking gate. Therefore, the tracking gate would itself be an accurate measure<br />

of the two-way travel time. In order to turn this measurement into the range from the<br />

electromagnetic mean sea surface, and, ultimately, a surface elevation (e.g., SSH),<br />

several corrections need to be applied to the data during ground processing. These<br />

include the instrumental and satellite corrections (e.g., software processing delays<br />

and Doppler shift), propagation corrections (e.g., dry and wet tropospheric delay and<br />

ionospheric delay), and geophysical corrections (e.g., tides and inverse barometer<br />

effect). Table 2.2 lists some corrections applied and the correction error to ERS-1<br />

range measurements (cf. Cudlip et al., 1994).<br />

Table 2.2 Corrections applied to ERS-1 range measurements and their likely error.<br />

Error Source Correction magnitude Error<br />

Doppler shift ±33 cm ±0.3 cm<br />

Centre of gravity offset 0-90 cm ±0.1 cm<br />

Ionospheric delay 0-100 cm ±5 cm<br />

Tropospheric delay-dry 1.7-2.5 m ±1 cm<br />

Tropospheric delay-wet 0-50 cm ±5 cm<br />

Ocean Tide ±50 cm a ±10 cm<br />

Earth Tide ±30 cm ±1 cm<br />

Ocean Loading Tide ±10 cm


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 26<br />

measurements, the SSH can be estimated by subtracting the corrected range from the<br />

altitude.<br />

2. 3 The Altimeter Waveform Data Used<br />

Three types of waveform data products from ERS-1, ERS-2, and POSEIDON have<br />

been used in this study. They are provided in the specific data format themselves by<br />

the European Space Agency (ESA) for ERS-1/2 and the Archiving, Validation and<br />

Interpretation of Satellite Oceanographic (AVISO) team for POSEIDON.<br />

2.3.1 ERS-1 Waveform Data<br />

The European Remote Sensing satellite ERS-1 was launched on 17 July 1991 by<br />

ESA. The ERS-1 measurements effectively consist of a combination of several<br />

different missions (Table 2.3).<br />

Phase<br />

Table 2.3 The repeat cycle characteristics of ERS-1.<br />

Repeat cycle<br />

(days)<br />

Number of orbits<br />

per cycle<br />

Ground-track<br />

spacing at the<br />

equator (km)<br />

ice 3 43 931<br />

Multi-discipline 35 501 80<br />

Geodetic 168 2411 16<br />

The ERS-1 Ku-band radar altimeter (RA) is one of the remote sensors carried onboard<br />

the satellite. It emits 1020 pulses per second and the on-board processor<br />

averages the returns in groups of 50 to produce 20 Hz waveforms (e.g., Quartly et al.,<br />

2001). The ERS-1 radar altimeter waveform is provided in a set of power signals<br />

with respect to time at 64 sample bins (cf. Table 2.1 and Figure 1.2). It is noted that<br />

ERS-1 geodetic mission contains two cycles of measurements and its second cycle of<br />

orbit drifts away 8 km from the first cycle of orbit on the equator, thus making a<br />

denser data coverage for geodetic applications.<br />

ERS-1 altimeter waveform products (ALT.WAP) are supplied by the United<br />

Kingdom Processing and Archiving Facility (UK-PAF) on behalf of ESA. The


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 27<br />

ALT.WAP product contains all the information telemetered in the altimeter source<br />

packets together with the corrections, calibration and orbit data required to further<br />

process the data into the GDRs. The data are structured following the CEOS format,<br />

which are unformatted, binary, direct access and with fixed record lengths (NRSC,<br />

1995).<br />

Because of the common format of the CEOS for all on-board remote sensing data<br />

(Mansley, 1996, cited in Anzenhofer et al., 2000), the original ALT.WAP data in the<br />

CEOS format are compressed to remove the un-used information for the altimeter,<br />

generating compressed waveform products in the AIDA waveform format (AWF).<br />

The intermediate ERS-1 waveform written in AWF contains only relevant<br />

information to produce the altimeter GDRs, leading a data reduction from 1 GByte to<br />

about 550 MByte for each 3-day data set (cf., Anzenhofer et al., 2000). They are<br />

stored on the CD-ROM and contain 1 Hz of geophysical and environmental<br />

0<br />

corrections, and 20 Hz of observing time, range, location, orbit altitude, σ (radar<br />

backscattering cross section per unit area), SWH, and waveform data in 64 sample<br />

bins. The exact details of the AWF format can be found in, e.g., Anzenhofer et al.<br />

(2000).<br />

Nearly 2.5 years of ERS-1 waveform data are used in this research, which include<br />

the data from all three phases. The data are supplied by the Ohio State <strong>University</strong>,<br />

USA, but they were originally provided by UK-PAF. Unfortunately, the waveform<br />

data used in this study are not provided in a format such that the orbit number can be<br />

obtained to identify individual ground tracks. This presents a problem for averaging<br />

repeated SSH to reduce the time variability and data noise when computing gravity<br />

fields described later in Section 7.5.2. Therefore, a case-specific averaging method<br />

will be developed to circumvent this problem.<br />

2.3.2 ERS-2 Waveform Data<br />

The ERS-2 satellite was launched on 21 April 1995 into a high inclination orbit<br />

(98.5°), which provides a near-global coverage. Its mission consists of a combination<br />

of the multi-discipline and tandem phases with both 35-day orbits. ERS-2 operates in<br />

two tracking modes: ocean mode and ice mode. The on-satellite tracking algorithm


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 28<br />

used for the ocean mode is known as the Sub-optimal Maximum Likelihood<br />

Estimator (SMLE), while the off-centre of gravity (OCOG) algorithm is used by the<br />

ice mode (Scott et al., 1994).<br />

The ERS-2 waveform data used in this study are taken from CD-ROMs supplied by<br />

De Montfort <strong>University</strong>, U.K., which are the ALT.WAP products supplied originally<br />

by UK-PAF on behalf of ESA. The original data were read, compressed and stored<br />

on the CD-ROM at De Montfort <strong>University</strong>. Each ERS-2 waveform record consists<br />

of 1 Hz and 20 Hz data. Geophysical and environmental corrections are supplied at 1<br />

Hz, and observations, such as the range, waveform longitude and latitude, altitude,<br />

σ<br />

0<br />

, and 64 waveform samples, are supplied at 20 Hz. Two cycles of waveform data<br />

(35-day repeat orbit, March to May 1999) are used in the study.<br />

A problem found in the ERS-2 waveform data edited by De Montfort <strong>University</strong> is<br />

that most of the Doppler range corrections exceed the minimum (-50 cm) and<br />

maximum (50 cm) acceptable values (cf., NRSC, 1995). Therefore, this correction is<br />

not applied to the altimeter range measurement. As stated, the returned signal is<br />

analysed in the frequency domain. Thus, any change in the frequency of the returns<br />

causes an error in the estimation of the range to the at-nadir sea surface. The Doppler<br />

shift changes the relative velocity between the altimeter and the sea surface, thus<br />

causing an error in the range estimate. The range correction for the Doppler shift is<br />

typically ±33 cm for ERS data (cf., Mantripp, 1996). Without applying this<br />

correction will cause the error to the range measurement, which is ~±0.3 cm from<br />

Table 2.1. However, comparing with the required range accuracy of a few<br />

centimetres (e.g., ~4 cm), the effect of this correction is small and can be neglected.<br />

2.3.3 POSEIDON Waveform Data<br />

The POSEIDON altimeter is a solid-state radar altimeter carried by the<br />

TOPEX/POSEIDON (T/P) satellite, which was launched on 10 August 1992 into a<br />

66-degree inclination orbit. It emits pulses at 1700 Hz and averages the returns in<br />

groups of 86 every 53 ms. The waveform is sampled in 64 bins of 3.125 ns width,<br />

with the tracking gate between bin 31 and 32. Waveform samples at the first two and<br />

last two (aliased) bins have been removed from the data products provided by the


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 29<br />

AVISO team. Since TOPEX and POSEIDON altimeters share an antenna on board<br />

the satellite, the POSEIDON altimeter is operated for about 10 percent of the mission<br />

time. The POSEIDON data used in this study comprise five cycles (January 1998 to<br />

January 1999) of 20 Hz waveform data (ten-day repeat orbit) taken from the CD-<br />

ROMs supplied by AVISO.<br />

Table 2.4 The reference ellipsoid characteristics of ERS-1/2 and T/P.<br />

Items ERS-1/2 T/P<br />

Equatorial radius (m) 6378137 6378136.3<br />

Flattening 1/298.257223563 1/298.257<br />

The POSEIDON waveform product is structured as the classical T/P GDRs in terms<br />

of cycle, pass files, headers and format. Each file is a fixed-length unformatted<br />

record and contains a header. All file headers are ASCII and all other data are VAX<br />

binary integers (AVISO/Altimeter, 1998). It consists of ten-day repeat cycles of data,<br />

which were separated into ascending and descending passes. A POSEIDON<br />

waveform record includes 1 Hz elementary data records of a pass-file and 20 Hz data<br />

of the waveform measurements (ibid.). The data are referenced to different ellipsoids<br />

(Table 2.4), so it is necessary to convert the data to a consistent reference (e.g.,<br />

GRS80) if they are used together.<br />

2. 4 External Input Data<br />

The external input data sets used in this research for providing necessary information<br />

for data editing and the result valuation include:<br />

(1) GSHHS (0.2 km resolution) shoreline model (Wessel and Smith, 1996)<br />

– provide information of the coastline’s location;<br />

(2) The DS759.2 (5′×5′ resolution) ocean depth model (Dunbar, 2000) and the<br />

Australian bathymetric model (30" resolution) (Buchanan, 1991) – identify<br />

whether radar returns reflect from land or water, and for result analysis;<br />

(3) Australian DEM (9"×9" resolution, version 2) (Hutchinson et al., 2001)<br />

– provide land elevations above local mean sea level;


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 30<br />

(4) A water/land grid (0.5′×0.5′ resolution) derived from the Generic Mapping<br />

Tools (GMT) high-resolution shoreline data (Wessel, 2000) – combined with<br />

the bathymetric models to flag data accurately;<br />

(5) AUSGeoid98 geoid grid (2′×2′ resolution) (Featherstone et al., 2001) – provide<br />

“ground truth” for validating the algorithms;<br />

(6) Sea surface topography (SST) model (Levitus et al., 2002) – provide the SST<br />

correction for estimating the gravity anomalies; and<br />

(7) Marine gravity data around Taiwan (Hwang and Wang, 2002) – provide the<br />

ground reference for the comparison between the gravity anomalies derived<br />

from SSHs before and after retracking.<br />

Some of these data will be described in detail below.<br />

2.4.1 The AUSGeoid98 Geoid Model of Australia<br />

The principal objective of the retracking system is to estimate the tracking error<br />

caused by the coastal waveform contamination and correct the altimeter range<br />

measurements accordingly. The analysis of the system discussed in Chapters 5, 6 and<br />

7 has arisen because of the necessity of knowing how good the retracking system is.<br />

To assess the retracking system, objective and independent ground references must<br />

be used. In this study, the AUSGeoid98 gravimetric geoid model of Australia<br />

(Featherstone et al., 2001) will be used as one of these independent ground<br />

references. The altimeter-derived SSH before and after retracking will be compared<br />

statistically with the geoid heights interpolated at the observation location from<br />

AUSGeoid98 (Chapter 5 and 6). This model is chosen just because it is the most<br />

recent and precise model readily available around Australia coastal regions at the<br />

time of this study. It is acknowledged that other reference sources of local geoid<br />

models (e.g., AUSGeoid91 and AUSGeoid93 models) or tide gauge data exist.<br />

However, the AUSGeoid98 includes denser and more types of data than other global<br />

or local models and uses arguably improved computational techniques (cf.,<br />

Featherstone et al., 2001), thus precisely representing the geoid features in Australia.<br />

The AUSGeoid98 geoid model was computed using data from the EGM96 global<br />

geopotential model, the 1996 release of the Australian gravity database (both over


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 31<br />

land and in marine areas), a national-wide digital elevation model (DEM), and<br />

satellite altimeter-derived marine gravity anomalies. The geoid height has a<br />

resolution of two minutes in longitude and latitude with respect to the GRS80<br />

ellipsoid. Comparisons of AUSGeoid98 with GPS and Australian Height Datum<br />

(AHD) heights across the continent indicate an RMS agreement of ±0.364 m (ibid.).<br />

It is also acknowledged that the altimeter-derived gravity anomalies, the global two<br />

minute grid of Sandwell et al. (1995), were used in the computation of the<br />

AUSGeoid98 model. However, the ship-track gravity data are also combined with<br />

the altimeter-derived gravity data using the ‘draping’ technique (cf., Kirby and<br />

Forsberg, 1997; Featherstone et al., 2001). This technique involves least squares<br />

collocation, as well as a 200 km high-pass filter. The satellite gravity grid is first<br />

carried out by this filter and the difference between the filtered altimeter grid and the<br />

ship-track gravity anomalies is calculated. Next, a difference grid is generated using<br />

least squares collocation. Then, this grid is added back to the filtered altimeter grid<br />

where no ship-track gravity data existed. Thus, because of the use of the ‘draping’<br />

technique, the ship-track gravity data contribute mainly to the geoid signal in areas<br />

which are close to the Australian coast. Therefore, the AUSGeoid98 can still be a<br />

quasi-independent reference for the purpose of assessing the SSH data before and<br />

after retracking.<br />

The problem that must be noted is that many ship-track gravity data used in the<br />

AUSGeoid98 model have not been cross-over analysed and adjusted (Featherstone et<br />

al., 2001; Featherstone, 2002). Some serious biases among overlapping tracks exist.<br />

Therefore, these ship-track gravity data should be used with some caution. This is<br />

also the reason that the marine gravity data around Taiwan are used in this <strong>thesis</strong> (cf.<br />

Section 2.4.3).<br />

2.4.2 Sea Surface Topography<br />

In order to estimate the gravity anomalies and geoid height from altimeter-derived<br />

SSH, both the time-dependent and time-independent (or explicitly quasi-timeindependent)<br />

sea surface topography (SST) must be removed (Hwang et al., 2002b;<br />

Chelton et al., 2001). The SST is the dynamic sea surface height associated with<br />

geostrophic surface currents (Chelton et al., 2001). The long-term mean SST from a


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 32<br />

reference surface can be computed from a compilation of ship observations of<br />

temperature and salinity profiles (ibid.). For instance, a 1°×1° mean SST grid related<br />

to 3000 dB computed from 100 years of ship observations of temperature and salinity<br />

profiles and smoothed by Levitus and Boyer (1994, cited in Chelton et al., 2001) can<br />

be found in Chelton et al. (Figure 52, p.88, 2001). Another Levitus SST version<br />

(1997) is provided by NOAA with the monthly averages on a 1°×1° grid for 12<br />

months (cf, Hwang et al., 2002b), which is related to a reference sea surface of<br />

1000 m depth level (Levitus et al., 2002). In this study, the quasi-time-independent<br />

SST values adopted is from a 1°×1° grid averaged from these 12 month SST values<br />

by Hwang et al. (2002b). Figure 2.3 shows the averaged Levitus (1997) SST, which<br />

appears to be the similar profile to the version of Levitus (1994) SST, but the<br />

different values due to the different references (cf, Chelton et al., Figure 52, p.88,<br />

2001).<br />

This SST grid (Figure 2.3) will be used to interpolate the required SST values at the<br />

altimeter data point around the Taiwan Strait in Chapter 7. It is applied because of<br />

the existence of the Kuroshio Current passing east of Taiwan and flowing along the<br />

eastern boundary of the East China Sea near the research area (118°E ≤ λ ≤ 123°E<br />

and 22°N ≤ ϕ ≤ 27°N). The Kuroshio Current is a strong western boundary surface<br />

current (Le Traon and Morrow, 2001). It causes the larger SSH gradients (>1 mm/km<br />

from Figure 2.3), so that the effect of SST cannot be neglected in the area around<br />

Taiwan (Hwang, 2003).<br />

However, this SST grid will not be applied to the altimeter data at the Australian<br />

coast in this study. Firstly, from the SST data released by Levitus (1997) to create<br />

this quasi-time-independent mean SST (Figure 2.3), there were no data sources near<br />

the Australian coast. The SST values at this area are based on an interpolated<br />

procedure from a 1°×1° grid (Hwang, 2003), thus causing additional errors to the<br />

altimeter-derived SSH there. Secondly, the ocean currents near the Australian coast,<br />

such as the Leeuwin Current, are weak when compared with the Kuroshio Current<br />

(Le Traon and Morrow, 2001; Hwang, 2003), and do not cause large SSH gradients<br />

(~0.2 mm/km). Thus, the SST effects on the SSH can be ignored at this area.


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 33<br />

Figure 2.3 Quasi-time-independent SST (contour interval is 10 cm) from Levitus et<br />

al. (1997).<br />

Finally, it is known that the SSH can be expressed as (Chelton et al., 2001)<br />

SSH h + h +<br />

=<br />

g d<br />

C + ε<br />

(2.3)<br />

where h g is the geoid height related to the reference ellipsoid, h d is the SST height, C<br />

is the range correction, and ε is the measurement error. However, it must be noted<br />

that the absolute SST values cannot be obtained from an existing SST model,<br />

because all existing SST versions are related to some reference surface (cf, Chelton<br />

et al., 2001; Levitus et al., 1997). From Equation (2.3), if the altimeter-derived SSH<br />

is used as an absolute measurement relative to the reference ellipsoid, nonuniqueness<br />

will be caused by applying the SSTs from different references.<br />

Due to these reasons, the SST was not applied to the altimeter-derived SSH data at<br />

the Australian coast. The comparison between the SSH and the AUSGeoid98 geoid<br />

height focuses on the observation of their profile variation rather than their difference.<br />

The standard deviation of the difference between them will be as an indicator of the<br />

SSH accuracy rather than the mean difference of them (Chapters 5 and 6).


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 34<br />

2.4.3 Marine Gravity Data around Taiwan<br />

To assess the accuracy of gravity anomalies derived from retracked satellite altimeter<br />

waveform data, the ship-track gravity data provided by Hwang and Wang (2002a)<br />

around Taiwan are used in this study (Figure 2.4). These marine gravity data contain<br />

observations conducted from more than 50 marine geophysical cruises during the<br />

period from 1965 to 1996. They are obtained from the gravity databases at Oxford<br />

<strong>University</strong>, the National Geophysical Data Centre data set of the National Oceanic<br />

and Atmospheric Administration (NOAA), and several overseas universities and<br />

agencies (Hsu et al., 1998; Hwang and Wang, 2002a).<br />

Figure 2.4 Coverage of ship-track gravity observations around Taiwan<br />

Some basic data cleaning procedures had been applied to these ship-borne gravity<br />

data, such as removing observations during turning the ship and crossover<br />

adjustment (c.f. Hsu et al., 1998; Hwang and Wang, 2002a). Most gravity data near<br />

Taiwan were provided by five geophysical cruises conducted in 1996 and GPS<br />

navigation was used to obtain high-quality positions and Eötvös corrections. Other


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 35<br />

ship-borne gravity data show lower precision because the Transit navigation system<br />

was used. Therefore, the five cruises were used as control data and others were<br />

adjusted with respect to them. The mean difference and standard deviation of the<br />

crossover points of total shipboard gravimetric data are 6.3 mgal and ±11.2 mgal (see<br />

Hsu et al., 1998). The free-air gravity anomalies were recomputed on the GRS80<br />

ellipsoid.<br />

Through the use of GPS positioning techniques and these careful data adjustments,<br />

the gravity data around Taiwan form a ‘ground control’ on the accuracy of the<br />

satellite altimeter-derived gravity anomalies. Gravity anomalies at 397 points along<br />

the ship tracks are used in comparisons (Chapter 7), which are taken in the area<br />

118°E ≤ λ ≤ 122°E and 23°N ≤ ϕ ≤ 26°N west of Taiwan to be consistent with the<br />

altimeter data (Figure 7.1).<br />

2.4.4 The Australian Digital Elevation Model (DEM)<br />

The shape of the waveforms in coastal regions have been found to differ from those<br />

recorded over open oceans from the previous studies (e.g., Brooks, 1997) and this<br />

study (Chapters 4 and 6). One of the causes of this waveform contamination is that<br />

the waveform may be mixed with the land and ocean returns when the altimeter<br />

simultaneously illuminates land and water. The DEM can provide information about<br />

the terrain elevation and topography on land within the altimeter footprint near the<br />

coastal zone, and thus be used for analysis of the land effects on the waveforms<br />

(Section 4.7).<br />

A DEM is a representation of the terrain using averaged elevation information. The<br />

recent upgraded (version 2) 9 second Australian DEM is a grid of elevation points<br />

covering the whole of Australia with a grid spacing of 9 seconds in longitude and<br />

latitude (approximately 250 m) in the Geocentric Datum of Australia 1994 (GDA94)<br />

horizontal reference frame (Hutchinson et al., 2001). The source data used to create<br />

the 9 second Australian DEM consists of:<br />

(1) revised data sets (spot heights, linear watercourse features, Australian coastline,<br />

and coastal inlets) from Australian GEODATA digital topography products


Chapter 2. SATELLITE ALTIMETER MEASUREMENTS AND DATA SOURCES 36<br />

(including TOPO-250K Relief theme, Drainage layer of TOPO-250K, and<br />

COAST-100K Coastline and State Borders);<br />

(2) trigonometric data points (converted to the GDA94 coordinate system) from<br />

the National Geodetic Data Base;<br />

(3) radar altimeter point elevation data (300 observations) for Lake Eyre; and<br />

(4) additional data sets (spot heights, stream line data, sink point data, cliff line<br />

data and associated contour line data) from digital 1:100 000 scale mapping.<br />

They are supplied by National Mapping Division of Geoscience Australia<br />

(formerly AUSLIG) (ibid.).<br />

Errors in the DEM depend mainly on the resolution of the DEM and the roughness of<br />

the terrain (i.e., the slope). Theoretical estimates and tests of the 9 second DEM<br />

against trigonometric data distributed evenly across the continent indicate that the<br />

standard elevation error of the DEM varies from ~±7.5 m to ~±20 m for most of the<br />

continent (ibid.). Errors are larger in highland areas with steep and complex terrain,<br />

where the largest errors can exceed 200 m (ibid.).<br />

2. 5 Summary<br />

This Chapter has introduced the basic fundamental working principle of a radar<br />

altimeter, the altimeter range measurements and the required sea surface information<br />

(i.e., SSH), and data products. Some external input data sources used in this study,<br />

such as the AUSGeoid98 geoid model, the Australian 9" DEM and the ship-track<br />

gravity around Taiwan, were also described in this Chapter.


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 37<br />

3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS<br />

3. 1 Introduction<br />

Radar altimeter return waveforms are the basic measurement for observing<br />

geophysical parameters of the Earth surfaces, in particular, the ocean. They give the<br />

range between the satellite and the mean sea surface at nadir (via pulse transit time),<br />

the SWH (via return waveform shape characteristics), and the roughness of the sea<br />

surface (via received energy, from which wind speed is empirically inferred).<br />

Retracking is a procedure of waveform data post-processing to improve parameter<br />

estimates over those given by GDRs. These parameters contain range corrections due<br />

to the estimation algorithm and the limited computational time on-board the satellite,<br />

and unmodelled ocean surface effects, such as the surface skewness. Waveform<br />

retracking has different emphases for different applications depending on the<br />

reflecting surfaces.<br />

Over ocean surfaces, altimeter retracking algorithms aim to reduce the following<br />

errors: short-wavelength random noise, and more importantly, potential longwavelength<br />

errors caused by SWH biases, range biases, the altimeter antennamispointing<br />

angle, or unmodelled non-Gaussian ocean surface parameters, such as<br />

the surface skewness and kurtosis (e.g., Hayne and Hancock III, 1990; Hayne et al.,<br />

1994; Hayne, 1980; Rodriguez, 1988; Rodriguez and Martin, 1994b). Over nonocean<br />

surfaces, the on-board satellite altimeter tracker cannot follow the sharp<br />

change of the surface topography (i.e., land/ice), causing an obvious departure<br />

between the midpoint of the leading edge of the waveform and the predesigned<br />

altimeter tracking gate. This leads to an error in the telemetered range measurements<br />

made by the altimeter. Waveform retracking, therefore, aims to produce accurate<br />

surface elevation measurements over ice sheets and land (e.g., Martin et al., 1983;<br />

Ridley and Partington, 1988; Davis, 1993a; Berry et al., 1998). The algorithms<br />

developed to retrack altimeter waveforms over ocean surfaces and non-ocean<br />

surfaces will be discussed in Sections 3.4, 3.5 and 3.6.<br />

In coastal regions, waveforms are contaminated by the land topography (see<br />

Chapters 1, 2 and 4), which exhibits a variety of terrains including the coastline,


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 38<br />

mountains, islands, lowlands, cliffs, reefs, and estuaries, as well as the sea states,<br />

such as standing and very calm water. To the author’s knowledge, all previously<br />

published retracking efforts involve retracking coastal waveforms using existing<br />

algorithms developed for ice sheets (e.g., Anzenhofer et al., 2000; Brook et al., 1997).<br />

A description of these methods will be given in Section 3.7.<br />

3. 2 Convolutional Representation of Ocean-Returned Waveforms<br />

The shape of the radar return waveform represents the observations measured by a<br />

satellite radar altimeter as a function of two-way travel time, which is related to the<br />

scattering characteristics of the reflecting surface. Moor and Williams (1957, cited in<br />

Brown, 1977) first proposed the convolutional representation of the return waveform.<br />

They demonstrated that for a ‘rough’ (i.e., incoherent) scattering surface, the average<br />

return power (as a function of delay time) can be expressed as a convolution of the<br />

transmitted power waveform envelope with a quantity involving<br />

0<br />

σ (radar<br />

backscattering cross section per unit scattering area), the antenna gain, and the range<br />

from the altimeter to any point on the Earth’s surface. Furthermore, when<br />

considering the vertical distribution of the surface height and radar receiver effects,<br />

the average return power (as a function of the delay time) is a convolution of three<br />

terms: the smooth spherical surface, the probability density function of the height of<br />

the specular points, and the radar system’s point target response (e.g., Brown, 1977;<br />

Hayne, 1980; Rodriguez, 1988).<br />

3.2.1 The Brown Model<br />

The mathematical model used to derive the average altimeter returned waveform<br />

from incoherent surface scattering has been based upon physical optics theory,<br />

whereby the surface is treated as a set of specular facets with a given height and<br />

slope probability density distribution. The time-series of the mean returned power<br />

waveform P (t)<br />

measured by a satellite altimeter is expressed as a convolution of<br />

three terms in the time domain as (Brown, 1977)<br />

P( t)<br />

= P ( t)<br />

∗ q ( t)<br />

∗ P ( t)<br />

(3.1)<br />

fs<br />

s<br />

PTR


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 39<br />

where t is the time measured at the satellite receiver such that t = 0 corresponds to<br />

the range to the mean sea level at nadir, P fs<br />

(t)<br />

is the average impulse response from<br />

a flat surface, q s<br />

(t)<br />

is the surface elevation probability density function (PDF) of<br />

specular points within the altimeter footprint, and P PTR<br />

(t)<br />

is the radar system point<br />

target response (PTR). Each of the three terms in Equation (3.1) will be presented<br />

separately later in Sections 3.2.2, 3.2.3 and 3.2.4.<br />

In order to give an analytical representation of the convolution in Equation (3.1),<br />

Brown (1977) first presented a simplified expression of the terms for near-normal<br />

incidence under the following assumptions, which are common to all satellite radar<br />

altimeter systems. The basic assumptions that are inherent in the convolution model<br />

of near-normal incidence rough surface backscatter (Barrick, 1972; Brown, 1977;<br />

Fedor et al., 1979) are as follows:<br />

(1) The root-mean-square (RMS) elevation variations are greater than the radar<br />

wavelength;<br />

(2) The ocean surface is statistically homogeneous and stationary over the number<br />

of pulses needed to determine a mean waveform;<br />

(3) The radar cross section is constant over the illuminated area;<br />

(4) There is no backscatter of the electromagnetic energy from beneath the surface;<br />

(5) The illuminated surface always contains a statistically significant number of<br />

specular points;<br />

(6) The surface elevation and surface slope density distributions are statistically<br />

independent; and<br />

(7) Multiple scattering among various parts of the surface is neglected.<br />

These assumptions are generally valid for an altimeter operating over open ocean<br />

surfaces. However, they are not generally the case over non-ocean surfaces, and<br />

hence an alternative approach must be sought (e.g., Mantripp, 1996). In coastal<br />

regions, areas of land can be simultaneously illuminated with water, thus causing<br />

land returns to the altimeter. Therefore, it is not always the case that the above<br />

assumptions are valid in the coastal zone (ibid.).


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 40<br />

3.2.2 Average Flat Surface Impulse Response Function<br />

Using a Gaussian approximation of the antenna gain, Brown (1977) presents P fs<br />

(t)<br />

over the illuminated area of the Earth’s surface as<br />

where<br />

P fs<br />

1/ 2<br />

( t)<br />

= Aexp(<br />

−α ' t)<br />

I ( β ' t ) U ( )<br />

(3.2)<br />

0<br />

t<br />

4c<br />

α'<br />

= cos(2ξ<br />

)<br />

(3.3)<br />

γ h<br />

a<br />

4<br />

β ' =<br />

γ<br />

a<br />

⎛ c<br />

⎜<br />

⎝ h<br />

⎟<br />

⎠<br />

⎞<br />

1/ 2<br />

sin(2ξ<br />

)<br />

(3.4)<br />

sin 2 θ<br />

γ<br />

a<br />

=<br />

ln 4<br />

(3.5)<br />

2<br />

In Equation (3.2), U (t)<br />

is a unit step function, I ( t<br />

1/ ')<br />

is the modified Bessel<br />

0<br />

β<br />

function (see Section 3.3.2), A is an amplitude scaling term containing the off-nadir<br />

pointing angle ξ and several other constants, such as the radar wavelength, the ocean<br />

reflectivity and the radar antenna gain (Brown, 1977). In Equations (3.3) and (3.4),<br />

h is the satellite altitude above the reference ellipsoid surface, and γ<br />

a<br />

is an antenna<br />

beamwidth parameter defined by Brown (1977). In Equation (3.5), θ is the antenna<br />

beamwidth.<br />

The derivation of Equations (3.3) and (3.4) assumes that the altimeter antenna pattern<br />

is approximately a Gaussian function, otherwise the closed form of Equation (3.2)<br />

cannot be derived from an integration of P fs<br />

(t)<br />

(see Brown, 1977, Equation 2).<br />

According to Brown (1977), this approximation is generally valid out to the point on<br />

the antenna pattern for which there is no appreciable contribution to the<br />

backscattered power.<br />

In Equation (3.2), A can be divided into two parts as<br />

A = A0<br />

Aξ<br />

, where 0<br />

A is an<br />

amplitude scaling term involving only constants of the radar parameters and the<br />

ocean reflectivity, and A ξ<br />

is a term related to the off-nadir pointing angle ξ . They<br />

can be written as


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 41<br />

where G<br />

0<br />

, λ and<br />

A<br />

0<br />

2 2 0<br />

0 λ cσ<br />

( 0)<br />

2 3<br />

( 4π<br />

) L h<br />

G<br />

= (3.6)<br />

4<br />

p<br />

⎛ 4 ⎞<br />

= ⎜ −<br />

2<br />

A<br />

ξ<br />

exp sin ξ ⎟<br />

(3.7)<br />

⎝ γ ⎠<br />

L<br />

p<br />

are radar system parameters (Brown, 1977).<br />

3.2.3 Surface Elevation PTR of the Scattering Distribution<br />

The instantaneous illumination pattern of the altimeter depends on the specific wave<br />

height field at time t within the footprint. Because of the random nature of the sea<br />

surface elevation distribution, it is more instructive to consider the PTR of specular<br />

points for the determination of the waveform. As a first-order approximation, the<br />

distribution of the sea surface elevation is Gaussian (Chelton et al., 2001). The scale<br />

of the wave-height distribution is described by the SWH, which is defined to be the<br />

average crest-to-trough height of the 1/3 highest waves. Brown (1977) uses a<br />

Gaussian specular point PDF to describe the statistics of the linear surface elevations.<br />

It is given as<br />

q s<br />

( ζ )<br />

2<br />

1 ⎛ ζ ⎞<br />

exp<br />

⎜ −<br />

⎟<br />

1 / 2<br />

(3.8)<br />

(2π<br />

) σ<br />

s ⎝ 2σ<br />

⎠<br />

= s<br />

where ζ is the surface elevation (positive upward) above the mean level of the<br />

specular points, σ s<br />

is the RMS surface elevation of the specular points related to<br />

SWH by SWH = 4(2 / c)<br />

σ . The PDF in Equation (3.8) is expressed in the space<br />

s<br />

domain in distance units (metres), while the ocean surface elevation density function<br />

q s<br />

(t) is written in the altimeter’s time domain in Equation (3.1). To coincide with<br />

other two terms in Equation (3.1), a variable<br />

t = −2ζ / c must be used to convert<br />

q s(ζ ) from the space domain to the two-way range time domain in time units<br />

(nanoseconds). The relevant standard deviation can be written as σ = 2σ c .<br />

t s<br />

/


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 42<br />

3.2.4 Radar System PTR<br />

The radar system PTR is primarily the transmitted radar pulse shape. For an idealised,<br />

linear, frequency-modulated altimeter radar pulse, the radar system PTR is given by<br />

(Ulaby et al., 1982, cited in Rodriguez, 1988)<br />

P PTR<br />

[(<br />

at / 2)( T − t )]<br />

2<br />

sin<br />

( t)<br />

− T ≤ t ≤ T<br />

(3.9)<br />

=<br />

2<br />

( at / 2)<br />

where T is the radar transmitted pulse duration (or the compressed pulse length if<br />

pulse compression is employed) and a is a constant which depends on the radar<br />

bandwidth. Thus, P PTR<br />

(t)<br />

is a symmetric expression of the impulse response about<br />

t = 0 . However, using the PTR expressed by Equation (3.9), it is not possible, in<br />

general, to obtain the analytic function of Equation (3.1). Since the width of the PTR<br />

of the short pulse radar altimeters is of the order of 20 ns (e.g., Brown, 1977),<br />

Equation (3.9) can be simplified to a Gaussian function as<br />

P<br />

PTR<br />

2<br />

⎡ ⎤<br />

⎢<br />

1 ⎛ ⎞<br />

⎜<br />

t<br />

( t)<br />

≈ η − ⎟ ⎥<br />

p<br />

PT<br />

exp<br />

⎢ 2<br />

(3.10)<br />

⎥<br />

⎣ ⎝σ<br />

p ⎠ ⎦<br />

where η<br />

P<br />

is the pulse compression ratio, P T<br />

is the peak transmitted power, and<br />

is a measure of the pulse width<br />

σ<br />

P<br />

σ<br />

P<br />

= 0. 425T<br />

(3.11)<br />

in which 0.425 is an approximation of the compressed pulse shape by a Gaussian<br />

function (Marth et al., 1993).<br />

3.2.5 Solution of the Radar Returns<br />

Using Equations (3.2), (3.8) and (3.10), Brown (1977) shows that the convolution of<br />

Equation (3.1) can be reduced to the following expression<br />

P(<br />

t)<br />

≈ P<br />

fs<br />

( t)<br />

∞ ∞<br />

∫∫<br />

0 −∞<br />

⎛ c ⎞<br />

⎜ ⎟P<br />

⎝ 2 ⎠<br />

PTR<br />

⎛ c<br />

( t −τ<br />

) qs⎜<br />

⎝ 2<br />

( τ − ˆ τ ) dτ<br />

d ˆ τ<br />

⎟ ⎠<br />

⎞<br />

(3.12)


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 43<br />

1 ⎡ ⎛ t ⎞ ⎤<br />

≈ ηP<br />

PT<br />

Pfs(<br />

t)<br />

σ<br />

P ⎢erf<br />

⎜ ⎟ + 1<br />

2<br />

⎥<br />

⎣ ⎝ 2σ<br />

⎠ ⎦<br />

where erf (x)<br />

is the error function, and σ is the composite risetime given by<br />

2 2 ⎛ 2<br />

= σ p<br />

+ ⎜ s<br />

2<br />

σ ⎞<br />

⎟<br />

⎝ c σ<br />

⎠<br />

(3.13)<br />

In Equation (3.12), the first term, P fs<br />

(t)<br />

, includes the effects of the antenna<br />

beamwidth and the off-nadir pointing angle. The flat surface emphasises that an<br />

incoherent surface scattering process is assumed. The second term, P PTR<br />

(t)<br />

, contains<br />

the effects of the receiver bandwidth on the transmitted pulse. The third term, q s<br />

(t)<br />

,<br />

contains the sea-state effects. It is assumed that waves on the sea surface are linear<br />

(i.e., an incoherent surface) and therefore the corresponding statistics of surface<br />

elevations and slopes can be expressed by a Gaussian function.<br />

Equation (3.12) is the fundamental model of the returned waveform over ocean<br />

surfaces, named the Brown model. It is important because it clearly shows that the<br />

simplified expression of the altimeter return waveform can be applied to extract<br />

geophysical parameters over oceans from the observed return waveform. However, it<br />

should be noted that some effects are neglected in the Brown model, such as the<br />

curvature of the Earth’s surface, approximation of the Bessel function, and the<br />

nonlinearity of the actual surface waves. These remain an open issue, which has been<br />

addressed by other researchers, such as Rodriguez (1988), Hayne (1980), Challenor<br />

and Srokosz (1989), and Rodriguez and Martin (1994a), and will be summarised in<br />

the following Sections.<br />

3. 3 Modified Convolutional Representation of the Ocean Return Waveform<br />

The Brown model has been modified, e.g., by Hayne (1980) and Rodriguez (1988),<br />

to present general expressions for the radar mean return waveform from non-<br />

Gaussian surfaces. This has allowed more parameters to be estimated from the radar<br />

return, such as the skewness of the sea surface (e.g., Lipa and Barrick, 1981;<br />

Rodriguez, 1988; Rodriguez and Martin, 1994b; Challenor and Srokosz, 1989). This<br />

Section will summarise the modifications made to the Brown model by these authors.


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 44<br />

Then, an analytic expression for the radar ocean return will be derived from some of<br />

these modified expressions for the purposes of this dissertation.<br />

3.3.1 Effects of the Earth’s Curvature<br />

The effects of the Earth’s curvature are neglected in Equations (3.3) and (3.4), but<br />

the off-nadir angle for a given pulse delay time t is slightly smaller than that in the<br />

flat-Earth approximation. Considering its effect, Rodriguez (1988) shows that the<br />

corrected functions α and β related to Equations (3.3) and (3.4) are<br />

4c<br />

1<br />

α = cos(2ξ<br />

)<br />

(3.14)<br />

γ h (1 + h / R)<br />

4 ⎛<br />

β = ⎜<br />

γ ⎝<br />

c<br />

h<br />

1 ⎞<br />

⎟<br />

(1 + h / R)<br />

⎠<br />

1/ 2<br />

sin(2ξ<br />

)<br />

(3.15)<br />

where R ≈ 6371005 m is the spherical radius of the Earth (e.g., GRS80; Morita,<br />

1980). The flat-Earth approximation in Equations (3.3) and (3.4) is not a realistic<br />

case. For the ERS-2 and TOPEX orbit heights of about 785 km and 1336 km,<br />

respectively, this flat-Earth approximation leads to a 5.8% and 9.1% reduction of the<br />

illuminated area of the flat-Earth surface. Thus, the average radar impulse response<br />

from a smooth spherical surface is given by (Rodriguez, 1988)<br />

P s<br />

1/ 2<br />

( t)<br />

= Aexp(<br />

−α t)<br />

I ( β t ) U ( )<br />

(3.16)<br />

0<br />

t<br />

Equation (3.16) includes effects due to the Earth’s curvature which, in general,<br />

cannot be ignored (Hayne and Hancock III, 1990).<br />

3.3.2 Simplification of the Bessel Function<br />

The convolution in Equation (3.1) does not allow for easy analytical integration<br />

because of the presence of the modified Bessel function (Equation 3.2). It can be<br />

seen from Equations (3.2) and (3.4) that the off-nadir angle ξ and return time t are<br />

two variables in the Bessel function. The off-nadir angle is ideally equal to zero, but<br />

in practice the antenna does not always point along the nadir and the maximum<br />

possible pointing error is 0.5 degrees (e.g., Challenor and Srokosz, 1989; Brown,


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 45<br />

1977). In addition, the return time is also a small value compared to the leading edge<br />

rise time (Rodriguez, 1988). For such small off-nadir angles and the case of shortpulse<br />

radar altimeters, Hayne (1980) expands I 0<br />

in the small-argument (assumed<br />

convergent) series expansion (Jeffrey, 1995):<br />

n<br />

2<br />

2<br />

1<br />

0<br />

( ) ∑ ∞ ⎛ z ⎞ ⎛ ⎞<br />

I z =<br />

⎜ ⎜ ⎟<br />

= 0 4<br />

⎟<br />

(3.17)<br />

n ⎝ ⎠ ⎝ n!<br />

⎠<br />

which allows for the term-by-term integration in the convolution model in Equation<br />

(3.1) (Hayne, 1980). Rodriguez (1988) expands the Bessel function presented in the<br />

smooth surface impulse response using Equation (3.17), but keeps only terms of<br />

order<br />

2<br />

β t and lower. Then, ( 1/ 2<br />

)<br />

I β is replaced by exp( 2 t / 4)<br />

0<br />

t<br />

β to give a simpler<br />

expression for analytical integration. Thus, considering the curvature of the Earth and<br />

the above approximation for I<br />

0<br />

, Equation (3.2) can be expressed as the average radar<br />

impulse response from a smooth sphere (Rodriguez, 1988):<br />

P s<br />

2<br />

⎡ ⎛ β ⎞ ⎤<br />

( t)<br />

≈ Aexp⎢−<br />

⎜α − t⎥U<br />

( t)<br />

⎣ 4<br />

⎟<br />

(3.18)<br />

⎝ ⎠ ⎦<br />

The error caused by this approximation was analysed by Rodriguez (1988) based<br />

upon various off-nadir angles and return times, for SEASAT, GEOSAT and TOPEX.<br />

The results show that the error is less than one percent for all times when the offnadir<br />

angles is approximately less than half the altimeter beamwidth (see Table 2.1)<br />

(Rodriguez, 1988).<br />

The antenna off-nadir angle is also called the off-nadir pointing error, which<br />

represents the displacement of the boresight axis of the altimeter antenna from the<br />

nadir. Because of peak power limitations and the high operating altitude, spaceborne<br />

radar altimeters generally use high gain antennas. A direct consequence of the high<br />

gain requirement is a narrow beamwidth, typically less than 3° (Brown, 1977). With<br />

current satellite attitude control systems, the beamwidth is 1.34° for ERS-1/2 and<br />

1.1° for TOPEX (Ku-band, see Table 2.1). Thus, the off-nadir angle approximated as<br />

half the antenna beamwidth cannot be larger than 1°. In addition, it is highly unlikely<br />

that the pointing error could exceed 1° without causing loss of tracking (cf. Barrick


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 46<br />

and Lipa, 1985). Therefore, Equation (3.18) is a reasonable approximation based on<br />

the typical magnitude of the off-nadir angle.<br />

3.3.3 Approximate Expression of the Non-Gaussian Scattering Surface<br />

As stated, the shape of the returned waveform is determined by the distribution of<br />

specular surface scatterers rather than by the actual sea surface height distribution<br />

within the footprint. The differences between the distributions of the scatterers and<br />

the sea surface height lead to biases between the altimeter-measured scattering<br />

surface and the actual mean sea surface, which must be corrected for the altimeter<br />

range measurements (Chelton et al., 2001; Rodriguez, 1988). The range to the sea<br />

surface is estimated from the half-power point of the leading edge of the waveform,<br />

which corresponds to the return from the median height of the specular scatterers<br />

referred to as the electromagnetic (EM) sea level (ibid.).<br />

The difference between the true mean sea level and the mean EM surface in the<br />

footprint is called the sea state bias (SSB), which comprises the electromagnetic bias<br />

(EMB) and the skewness bias (SKB). Owing to the non-Gaussian nature of the sea<br />

surface, the wave troughs are brighter than the peaks for radar wavelengths, arising<br />

in a greater backscattered power per unit surface area from wave troughs than from<br />

wave crests. The result is that the mean EM surface measured by altimeters is biased<br />

lower than the true mean sea surface toward wave troughs. This effect is known as<br />

the EMB. From previous studies (cf. Chelton et al., 2001), the EMB has been<br />

observed to increase monotonically with increasing the SWH. The SKB is the error<br />

in tracking the mean height of the specular scatterers, which is the height difference<br />

between the mean scattering surface and the median scattering surface (ibid.). The<br />

SKB is generally proportional to the SWH, but it is not related in any simple manner<br />

to the SWH or any other geophysical quantity that can be inferred from altimeter<br />

data.<br />

The specular point PDF q s<br />

(t)<br />

in Equation (3.1) represents the probability density<br />

distribution of the specular scatterers within the altimeter footprint. It thus includes<br />

the effects of the SSB. It is evident that the SSB in the surface elevation distribution,<br />

through its effect on the specular point PDF, affects the ability of the on-satellite


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 47<br />

tracking algorithm to determine the exact range to the at-nadir mean sea level. The<br />

Brown model (Equation 3.12) uses a Gaussian specular point PDF to describe the<br />

statistics of the ‘linear’ ocean wave heights. As stated, the ‘linear’ ocean surface is<br />

defined as an incoherent surface (i.e., not including the SSB) in this study. To a firstorder<br />

approximation, the height of the waves and slope of the waves are assumed<br />

zero-mean, Gaussian, uncorrelated random variables. However, higher order effects,<br />

including correlation between heights and slopes, as well as skewness, indicate that<br />

the surface height and slope are not strictly Gaussian (Barrick and Lipa, 1985). For<br />

example, the results obtained by Srokosz (1986) and Rodriguez (1988) regarding the<br />

effects of the SKB on the range to the sea surface show that the SKB causes an error<br />

of 0.8 cm for a skewness of 0.1 and the SWH of 2 m. This error increases to ~4 cm<br />

for the SWH of 10 m.<br />

Because of the effects from both the EMB and SKB on the shape of the leading edge<br />

of the radar returned waveform, as well as the prospect of extracting information<br />

about the sea state of a non-linear scattering surface, numerous studies have<br />

attempted to estimate the SKB (e.g., Lipa and Barrick, 1981; Parsons, 1979;<br />

Rodriguez, 1988; Barrick and Lipa, 1985; Hayne, 1980), and the EMB (e.g.,<br />

Rodriguez, 1988) using a modified expression for the specular point PDF.<br />

Longuet-Higgins (1966, cited in Challenor, 1989) obtained the PDF of surface height<br />

using an approximation of the Gram-Charlier series (cf. Jeffrey, 1995) for the nonlinear<br />

surface:<br />

⎫ ⎛ ⎞<br />

⎨<br />

⎧ 2<br />

1 λ<br />

= + s<br />

η<br />

qs ( ζ ) 1 H<br />

3( η) ⎬exp⎜<br />

− ⎟<br />

2π<br />

σ ⎩ 6 ⎭ ⎝ 2<br />

(3.19)<br />

s<br />

⎠<br />

where<br />

ξ<br />

η = (3.20)<br />

σ s<br />

and<br />

H<br />

n<br />

are the Hermite polynomials (e.g., Jeffrey, 1995) of degree n with


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 48<br />

H<br />

n<br />

n<br />

n 2 d<br />

2<br />

( x)<br />

= ( −1)<br />

exp( x ) exp( −x<br />

)<br />

(3.21)<br />

n<br />

dx<br />

Equation (3.19) introduces the height skewness parameter ( λ<br />

s<br />

) of the sea surface,<br />

which is a measure of the non-linearity of the wavefield, and is a dimensionless<br />

quantity (Hayne and Hancock III, 1990). Barrick and Lipa (1985) and Srokosz (1986)<br />

extend the result in Equation (3.19) to obtain the joint PDF of the elevation and slope,<br />

which contains another non-linear parameter of the cross-skewness (δ ) as<br />

2<br />

1 ⎧ λ<br />

⎛ ⎞<br />

s<br />

δ ⎫ η<br />

qs ( ζ ) = ⎨1<br />

+ H<br />

3( η) − H1( η) ⎬exp⎜<br />

− ⎟.<br />

(3.22)<br />

2π<br />

σ ⎩ 6 2 ⎭ ⎝ 2<br />

s<br />

⎠<br />

In addition, Hayne (1980) uses a four-term series as the expression of radar-observed<br />

surface height PDF from the laboratory measurements of the surface elevation<br />

density function for a wind-generated wavefield, which contains Hermite<br />

polynomials of H<br />

3<br />

, H<br />

4<br />

, and H<br />

6<br />

(see Hayne, 1980, Equation 12). Two non-linear<br />

parameters of skewness and kurtosis are introduced (discussed below).<br />

The skewness parameter contributes significantly to the SSB (cf. AVISO/Altimeter,<br />

1996, p.45; Chelton, 2001), while the cross-skewness term is thought to be related to<br />

the EMB and thus cause a shift of the electromagnetic surface with respect to the true<br />

mean sea surface (Rodriguez, 1988). A rewritten form of the PDF by Rodriguez<br />

(1988) makes the effect of the EMB more apparent:<br />

⎫ ⎛ ⎞<br />

⎨<br />

⎧ 2<br />

1 = + s<br />

( ) ⎬ ⎜ −<br />

R<br />

q ( ζ λ<br />

η<br />

s<br />

) 1 H<br />

3<br />

ηR<br />

exp ⎟<br />

2π<br />

σ ⎩ 6 ⎭ ⎝ 2<br />

s<br />

⎠<br />

(3.23)<br />

where<br />

η<br />

R<br />

( ζ − ζ γσ / 2)<br />

+ 0 s<br />

= (3.24)<br />

σ<br />

s<br />

where γ is the parameter related to the surface shift partly caused by the EMB and<br />

ζ<br />

0<br />

is the altimeter tracker error. The expression in Equation (3.24) indicates that the<br />

effect of γ is to shift the whole surface by the amount of − γσ / 2)<br />

, which is the<br />

(<br />

s


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 49<br />

magnitude of the EMB. This means that the total altimeter height estimation error,<br />

i.e., the shift of the altimeter-observed sea surface with respect to the true mean sea<br />

surface, consists of the altimeter tracker error ( ζ<br />

0<br />

) and the shift caused by the EMB.<br />

Therefore, Equation (3.23) differs from Equations (3.19) and (3.22).<br />

The SKB can be estimated from the above expressions of the specular point PDF<br />

using the altimeter waveform data (e.g., Hayne, 1980; Rodriguez, 1988; Lipa and<br />

Barrick, 1981; Barrick and Lipa, 1985). Using retracked TOPEX data after removing<br />

the EMB, Rodriguez and Martin (1994) estimate that the dimensionless skewness is<br />

very small and ranges from 0.01-0.12.<br />

Theoretically, determining the EMB requires knowledge of the specular point PDF.<br />

Unfortunately, the current theoretical understanding of the EMB is not sufficiently<br />

well developed to estimate the EMB from a purely theoretical basis or from altimeter<br />

data themselves without additional geophysical information or assumptions (Chelton,<br />

2001; Rodriguez, 1988). There are two possible reasons for this. The first is because<br />

two shifts of the EMB and SKB are introduced in the specular point PDF, and they<br />

are not independent to all orders of approximation. Thus, they cannot be<br />

distinguished. The second is that the EMB depends on the characteristics of the<br />

wavefields, while only the SWH can be unambiguously extracted directly from the<br />

altimeter waveform. Therefore, in practice, empirical estimation of the EMB is a<br />

more appropriate method (Chelton, 2001).<br />

3.3.4 Approximate Expression of the System PTR<br />

The altimeter-transmitted pulses have a shape that is described by the radar system<br />

PTR. However, the PTR also includes effects due to the bandwidth of the receiver<br />

(Hayne, 1980). It is generally expressed by Equations (3.9) and (3.10). In the Brown<br />

model (Equation 3.12), the PTR is expressed by a Gaussian function (Equation 3.10).<br />

In reality, the PTRs are not exactly Gaussian (Hayne, 1980; Rodriguez, 1988).<br />

Rodriguez (1988) compares measured PTRs of SEASAT and GEOSAT with an<br />

idealised PTR ( sinc<br />

2 x)<br />

. The results (Rodriguez, 1988) show that the radar system<br />

PTR presents a lack of symmetry and large side lobes, suggesting thus the PTR is not


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 50<br />

2<br />

sinc x or a Gaussian function. The effect of the Gaussian-approximated PTR will<br />

change the shape of the return waveform, especially in the early parts of the leading<br />

edge. To obtain realistic non-Gaussian ocean parameters, Rodriguez (1988)<br />

concludes that the measured PTR must be used in the estimation process to obtain<br />

realistic sea state parameters.<br />

The modified PTR has been developed by Hayne (1980), and Rodriguez and Martin<br />

(1994). Hayne (ibid.) expresses the PTR using a nearly Gaussian function (a fourterm<br />

series). Rodriguez and Martin (1994) expand the PTR as a series of Gaussian<br />

functions. They both appear to improve the estimates of the non-Gaussian ocean<br />

surfaces (Chelton et al., 2001).<br />

3.3.5 A Modified Solution of the Radar Returns Developed in This Study<br />

Following Brown (1977), modified convolutional representations of the waveform<br />

for non-Gaussian ocean surfaces have been given by Hayne (1980), Rodriguez<br />

(1988), Rodriguez and Chapman (1989), Challenor and Srokosz (1989), and Barrick<br />

and Lipa (1985). The main difference among the convolutional solutions is that<br />

researchers take different modified expressions of the surface-height PDF, the radar<br />

system PTR, and the Bessel function. Following this idea, and according to the<br />

specific application of this study, a modified solution of the radar returns developed<br />

by the author will be introduced in this Subsection.<br />

The model introduced in this study is slightly different from the Brown model and<br />

the modified models (e.g., the Hayne (1980) model). The main differences are as<br />

follows:<br />

(1) As stated, the current theory is not sufficiently well developed to accurately<br />

estimate the EMB using only altimeter data. This term will be neglected from<br />

the specular point PDF. Instead, Equation (3.19) is employed as the expression<br />

of the specular point PDF.<br />

(2) Equation (3.18) is used as the function describing the smooth surface response,<br />

because it takes account of the effects of the Earth’s curvature. It is also<br />

computationally simple, but without significant loss of accuracy (Rodriguez,


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 51<br />

1988), provided that the Bessel function has been expanded appropriately<br />

(ibid.).<br />

(3) Since this study does not focus on estimating non-linear parameters of the non-<br />

Gaussian ocean surface, the Gaussian function will still be assumed to be an<br />

adequate expression of the radar system PTR for the purpose of this study.<br />

Using a variable<br />

t = −2ζ / c , Equation (3.19) can be rewritten in the time domain as:<br />

q s<br />

( t)<br />

=<br />

1 ⎧ λs<br />

⎨1<br />

+<br />

2π<br />

σ ⎩ 6<br />

s<br />

2<br />

3 ⎫ ⎛ η ⎞<br />

( − 3 ) exp⎜<br />

−<br />

t<br />

η η<br />

⎟<br />

t<br />

t<br />

⎬<br />

⎭<br />

⎝<br />

2<br />

⎠<br />

(3.25)<br />

in which<br />

c<br />

ηt<br />

= − t<br />

(3.26)<br />

2σ<br />

s<br />

Equation (3.25) is the surface height PDF in the time domain. It can be reduced to<br />

the linear case when the skewness parameter λ<br />

s<br />

is omitted, i.e., by setting λ<br />

s<br />

=0.<br />

The radar system PTR used is assumed to be a Gaussian function, given by<br />

P<br />

PTR<br />

( t)<br />

≈<br />

1<br />

⎡<br />

1 ⎛ ⎞<br />

⎢ ⎜<br />

t<br />

exp − ⎟<br />

2π<br />

σ ⎢ 2<br />

p<br />

⎣ ⎝σ<br />

p ⎠<br />

2<br />

⎤<br />

⎥<br />

⎥<br />

⎦<br />

(3.27)<br />

As stated, for typical short pulse radar altimeters, the width of the PTR is of the order<br />

20 ns or less (Brown, 1977). This approximation of the PTR is found to be adequate<br />

for these widths by Brown (1977).<br />

From Equations (3.18), (3.25), and (3.27), describing the three terms in the<br />

convolutional representation, a modification of Equation (3.1) will be given. The<br />

convolution of the specular point PDF and the system point PTR can be written as<br />

(Bracewell, 2000)


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 52<br />

B(<br />

t)<br />

= qs(<br />

t)<br />

∗ PPTR(<br />

t)<br />

1 ⎡ 1 ⎛ t ⎞<br />

= exp⎢−<br />

⎜ ⎟<br />

2π<br />

σ ⎢⎣<br />

2 ⎝σ<br />

⎠<br />

2<br />

⎤⎡<br />

1<br />

⎥⎢1<br />

+ λcsH<br />

⎥⎦<br />

⎣ 6<br />

3<br />

⎛ t ⎞⎤<br />

⎜ ⎟⎥<br />

⎝σ<br />

⎠⎦<br />

(3.28)<br />

in which λ cs<br />

and σ are the composite skewness and risetime (i.e., the information<br />

relative to the sea surface roughness through the SWH) given by<br />

2 2 ⎛ 2<br />

= σ p<br />

+ ⎜ s<br />

2<br />

σ ⎞<br />

⎟<br />

⎝ c σ<br />

⎠<br />

(3.29)<br />

3<br />

⎛σ<br />

s ⎞<br />

λcs = λs⎜<br />

⎟<br />

(3.30)<br />

⎝ σ ⎠<br />

Then, the waveform can be obtained as:<br />

P(<br />

t)<br />

= P ( t)<br />

∗ B(<br />

t)<br />

=<br />

s<br />

∫ ∞ −∞<br />

= A<br />

P ( x)<br />

B(<br />

t − x)<br />

dx<br />

A<br />

s<br />

exp<br />

0 ξ<br />

τ<br />

[ − d(<br />

+ d / 2) ](<br />

C1<br />

+ C2<br />

) + PN<br />

(3.31)<br />

where<br />

1 ⎡ ⎛ τ ⎞ ⎤<br />

C<br />

1<br />

= ⎢erf<br />

⎜ ⎟ + 1⎥<br />

(3.32)<br />

2 ⎣ ⎝ 2 ⎠ ⎦<br />

3<br />

2<br />

λ ⎧<br />

⎫ ⎛ ⎞<br />

cs<br />

⎡ ⎛ τ ⎞ ⎤ d 1<br />

2<br />

2 −τ<br />

C2<br />

= ⎨⎢erf<br />

⎜ ⎟+ 1⎥<br />

+ ( 1−<br />

3d<br />

− 3d<br />

⋅τ<br />

−τ<br />

) ⎬exp⎜<br />

⎟<br />

(3.33)<br />

6 ⎩⎣<br />

⎝ 2 ⎠ ⎦ 2 2π<br />

⎭ ⎝ 2 ⎠<br />

with<br />

t − t<br />

τ = σ<br />

0<br />

− d<br />

(3.34)<br />

2<br />

⎛ β ⎞<br />

d = ⎜α<br />

⎟<br />

−<br />

σ<br />

(3.35)<br />

⎝ 4 ⎠<br />

where P<br />

N<br />

is an additive noise term that represents the altimeter’s thermal noise, the<br />

time shift t 0<br />

represents the mean level of the specular points on the ocean surface at


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 53<br />

nadir, and σ represents the standard deviation (in delay time units) of the specular<br />

point about this mean level. Equation (3.31) is an extension of the expression<br />

obtained by Brown (1977) to include the effects of the skewness and Earth curvature.<br />

It is also different from the Hayne (1980) model, which includes another non-linear<br />

kurtosis term to estimate parameters for the non-Gaussian surface.<br />

The effects of Equations (3.32) to (3.35) on the waveform can be discussed<br />

qualitatively. The function C 1<br />

is the predominant expression of the leading edge of<br />

the return waveform, while the function C 2<br />

represents the effects of the skewness<br />

and the off-nadir angle on the leading edge. Because of small magnitude of the<br />

skewness 0.01< λ


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 54<br />

significantly affects the waveform in the leading edge when τ ≤ 2 . It has no effect<br />

on the regions of the thermal noise and trailing edge of the waveform (Figure 3.3).<br />

0.12<br />

0.10<br />

ERS<br />

POSEIDON<br />

TOPEX<br />

Values of d<br />

0.08<br />

0.05<br />

0.03<br />

0.00<br />

0.0 5.0 10.0 15.0 20.0<br />

SWH (metres)<br />

Figure 3.1 Values of d ( ξ = 0 ), showing that d


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 55<br />

0.04<br />

0.02<br />

Values of C2<br />

0.00<br />

−0.02<br />

−0.04<br />

−0.06<br />

ERS<br />

POSEIDON<br />

TOPEX<br />

−0.08<br />

0.0 20.0 40.0 60.0<br />

Bin or Gate<br />

Figure 3.3 Values of C<br />

2<br />

( d = 0 , λ<br />

cs<br />

= -1, and SWH = 4 m) showing the small<br />

magnitude of effects on the leading edge of the waveform, and nearly overlapped<br />

curves for ERS-1/2 and TOPEX.<br />

As can be seen from Figure 3.3, for the times near the tracking gate when τ ≤ 1, the<br />

effect of the skewness is to lower the power, and to raise it when the leading edge<br />

starts to rise or near the plateau of the waveform.<br />

Equation (3.31) can be considered as a fundamental model of the ocean return<br />

waveforms. From the above qualitative and quantitative analysis, it is found that the<br />

effects of non-Gaussian ocean surface through skewness λ cs<br />

on the waveforms are<br />

very small (~0.01-0.12). In this case, when retracking waveform to estimate the<br />

range correction, the skewness term can be ignored without significantly affecting<br />

the results in the complete convolutional form. In Chapter 5, Equation (3.31) will be<br />

used to obtain a linear analytical function for retracking waveforms in coastal regions.<br />

3. 4 Ocean Waveform Retracking<br />

In general, a satellite radar altimeter’s on-board range estimation is limited by the onboard<br />

computational capability and by real-time data processing constraints. As the


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 56<br />

orbit errors are reduced by using geometric, dynamic and combined methods (e.g.,<br />

T/P orbit determination), it becomes increasingly desirable to also reduce the errors<br />

due to the on-board range estimation algorithm. In addition to the on-satellite range<br />

estimates, the satellite also transmits the waveform measured by the radar altimeter<br />

to the ground. This waveform contains information about the scattering surface, but<br />

is retracked later to get improved ranges.<br />

The convolutional representation of the radar return waveform, its shape, and the<br />

precise position of the waveform in the range window allow for the retrieval of more<br />

accurate estimates of the range and the SWH, as well as additional parameters such<br />

as the antenna off-nadir angle and the skewness of the surface elevation distribution<br />

using more sophisticated algorithms. Hence, it has long been recognised that<br />

reprocessing altimeter waveform data will lead to improved geophysical estimates<br />

using waveform retracking over ocean surfaces (Brenner et al., 1993; Barrick and<br />

Lipa, 1985; Rodriguez and Martin, 1994a; Rodriguez, 1988).<br />

Ocean waveform retracking is based upon the convolutional representation of the<br />

waveform, such as the Brown model (Equation 3.12), or the modified model (e.g.,<br />

Equation 3.31) for non-linear ocean surfaces. The retracking algorithms used to<br />

estimate the geophysical parameters over ocean surfaces can mainly be categorised<br />

into two classes of least squares estimation and deconvolution, as follows.<br />

3.4.1 Least Squares Fitting Procedure<br />

As stated, an analytic representation of the waveform can be obtained by formulating<br />

a convolution of three terms. The most commonly used algorithm is to fit a<br />

convolution model of the waveform (e.g., Equation 3.31) to the measured waveforms.<br />

An iterative least squares fitting procedure (Hayne and Hancock III, 1990; Hayne<br />

and Hancock III, 1982) or a maximum likelihood estimation (MLE) algorithm<br />

(Challenor and Srokosz, 1989; Parsons, 1979; Rodriguez and Martin, 1994a;<br />

Tokmakian et al., 1994) is performed to extract the parameter estimates.<br />

Hayne and Hancock III (1990) use an iterative least squares fitting procedure to<br />

retrack GEOSAT waveform data. In their algorithm, Equations (3.16), (3.19) and<br />

(3.27) express three terms in the convolution Equation (3.1). The model of the


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 57<br />

convolutional solution is fitted to the waveforms to produce six independent<br />

estimates. These are:<br />

(1) Waveform amplitude, A ;<br />

(2) The time shift of the range-tracker-positioned waveform tracking gate relative<br />

to the true tracking gate of the waveform, t 0<br />

;<br />

(3) SWH;<br />

(4) Noise baseline, P<br />

n<br />

, (i.e., thermal noise);<br />

(5) Skewness of the radar-observed surface elevation PDF, λ<br />

cs<br />

; and<br />

(6) Off-nadir pointing angle, ξ .<br />

The waveform has noise. It can be seen in the trailing edge, but is more difficult to<br />

identify in the leading edge. The parameters affected by the noise in the tailing edge<br />

will degrade the geodetic parameter estimates in the leading edge if the least squares<br />

fitting procedure is applied to an individual waveform. The fitting method in this<br />

case may appear to give statistically small error but does not provide a consistent<br />

accuracy to the individual parameters. Therefore, to reduce the noise on the return<br />

waveforms, effective fitting is usually carried out using waveform averages in<br />

several seconds rather than individual waveforms (e.g., Challenor and Srokosz,<br />

1989). In the procedure of Hayne and Hancock III (1990), ten-second average<br />

waveforms are used. Since the closed form solutions of the theoretical waveform<br />

cannot be deduced from realistic forms of the radar PTR, the fitting algorithm suffers<br />

from the drawback that the fitting function is analytic only if the radar system PTR is<br />

Gaussian (Brown, 1977; Challenor and Strkosz, 1989), near-Gaussian (Hayne, 1980),<br />

or approximated by a set of Gaussians (Rodriguez and Martin, 1994a).<br />

A typical example of using the MLE algorithm is from the UK Earth Observation<br />

Data Centre (EODC). EODC retracks ERS-1 data to re-estimate all parameters<br />

calculated by the on-board algorithm on the ground using the MLE algorithm (e.g.,<br />

Tokmakian, 1994). The convolution function includes unknown parameters A , t 0<br />

,<br />

σ , ξ , and non-linear wave parameters λ cs<br />

and δ . The MLE algorithm has been<br />

tested by EODC with three forms of the convolution functions as follows:<br />

(1) Non-linear model containing both non-linear wave parameters λ<br />

cs<br />

and δ .


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 58<br />

(2) Non-linear model containing only one of non-linear wave parameters λ<br />

cs<br />

or δ .<br />

(3) Linear model without λ<br />

cs<br />

and δ .<br />

Their results show that the MLE algorithm is successful with the linear model for the<br />

estimation of the geophysical parameters, but cannot accurately estimate all<br />

parameters if the non-linear model is used. From EODC’s tests, it has been shown<br />

that the 20 Hz individual waveform data does not produce as accurate results as<br />

when averaging the waveforms to 1 Hz (ibid.). This means an averaging procedure is<br />

essential as well.<br />

Theoretically, the MLE algorithm is more accurate than least squares (Chelton et al.,<br />

2001) and gives asymptotically minimum variance unbiased estimates of the<br />

parameters (Challenor and Srokosz, 1989). However, in practice, because of<br />

unreliability of the statistical weights due to unmodelled errors in the waveforms<br />

(Chelton, 2001) and further complications introduced by numerics of the model<br />

(Tokmakian et al., 1994), MLE estimates are not significantly better than those<br />

estimated by least squares. Therefore, nonlinear least-squares fitting procedure<br />

appears to still be an adequate method of waveform retracking.<br />

3.4.2 Deconvolution Method<br />

The second method estimates parameters from the specular point PDF obtained by<br />

deconvolution, which was introduced by Priester and Miller (1979), Lipa and Barrick<br />

(1981), and refined by Rodriguea and Chapman (1989). The deconvolution method<br />

uses a fact that the average altimeter return waveform is the convolution of the ocean<br />

surface specular point PDF, which is an unknown function, and two functions of the<br />

radar PTR and the smooth surface response, which are known and depend only on<br />

the radar design parameters. By taking the fast Fourier transform (FFT) of the slope<br />

function, the specular point PDF can be obtained from dividing the Fourier transform<br />

of the data by the Fourier transform of the convolution of the smooth surface<br />

response and the radar PTR. The specular PDF will then allow estimation of three<br />

independent sea-surface-state parameters (e.g., Lipa and Barrick, 1981; Barrick and<br />

Lipa, 1985; Rodriguez, 1988; Rodriguez and Chapman, 1989). These are:<br />

(1) the ocean surface standard deviation;


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 59<br />

(2) ocean surface skewness; and<br />

(3) shift of the altimeter tracking gate with respect to the mean electromagnetic<br />

surface measured by an altimeter.<br />

For the convenience of representation, altimeter return power Equation (3.1) can be<br />

thought of as a matrix equation<br />

y = Mx<br />

(3.36)<br />

where y is the vector of return power and M is the convolution given by<br />

M = P ( t)<br />

∗ P ( t)<br />

(3.37)<br />

fs<br />

PTR<br />

and x is the specular point PDF vector<br />

x = q(t)<br />

(3.38)<br />

Thus, the specular point PDF can be obtained by inverting Equation (3.36).<br />

In its most direct implementation, Lipa and Barrick (1981) suggest that<br />

deconvolution can be performed on the slope of the leading edge of the waveform,<br />

rather than the full waveform. This suggestion means that the radar antenna pattern<br />

and the off-nadir angle can be ignored, and the slope of the mean waveform is given<br />

by the convolution of the specular point PDF and the radar PTR. Thus, data can be<br />

neglected in the trailing edge of the waveform, where the noise level is higher<br />

(Section 3.3.5), and the returns depend weakly on the non-Gaussian ocean surface.<br />

This algorithm has the advantages that it is computationally simple and fast, and it is<br />

not necessary to assume any functional form for the radar specular point PDF.<br />

However, it is impossible in general to ignore the radar antenna pattern and the offnadir<br />

angle for altimeters, such as GEOSAT and TOPEX, because it introduces<br />

errors which are of the same order as the quantities to be estimated (cf. Rodriguez,<br />

1988). Another disadvantage of the slope deconvolution is that the algorithm is<br />

highly sensitive to high-frequency noise in the waveform. This leads the<br />

deconvolution of Lipa and Barrick (1981) being carried out only for waveforms that


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 60<br />

had been averaged over a long time period (e.g., 10 seconds) for which an acceptably<br />

low noise level was obtained.<br />

In order to overcome these problems, Rodriguez (1988) and Rodriguez and Chapman<br />

(1989) introduce a "corrected slope convolution model" as<br />

2<br />

1 ⎡ ⎛ β ⎞ ⎤<br />

M = ⎢P′<br />

( t)<br />

+<br />

⎜α<br />

−<br />

⎟P(<br />

t)<br />

⎥<br />

(3.39)<br />

A ⎣ ⎝ 4 ⎠ ⎦<br />

where P′ (t)<br />

is the derivative of the return power with respect to the return time. This<br />

model differs from the expression derived by Lipa and Barrick (1981, Equation 3),<br />

because the second term on the right-hand side represents the radar antenna pattern<br />

and off-nadir angles. Since the return waveform is also a function of the thermal<br />

noise and the off-nadir pointing angle, they must be estimated separately when using<br />

Equation (3.39).<br />

Both deconvolution algorithms presented by Lipa and Barrick (1981) and Rodriguez<br />

and Chapman (1989) have advantages and disadvantages. Rodriguez and Chapman<br />

(ibid.) investigate the use of both methods. They recommend that a spectral filtering<br />

procedure be used to reduce the high-frequency noise in the estimated parameters.<br />

3. 5 Ice-Sheet Waveform Retracking<br />

Ice-sheet surfaces are characterised by the existence of topography, often with a<br />

significant surface slope with respect to the ocean. Over such topographic surfaces,<br />

the altimeter is generally unable to follow abrupt altitude changes and small-scale<br />

slopes, thus losing signal lock. To maintain the satellite track over large surface<br />

slopes, ERS-1/2 altimeters operate in ice mode over ice sheets at a reduced range<br />

resolution or increased sampling time (12.12 ns) compared with the ocean mode<br />

(3.03 ns). The form of the return is determined by the range to the different reflectors,<br />

their scattering properties, and their position within the antenna beamwidth. This<br />

results in a return waveform of generally unpredictable shape, which changes rapidly<br />

and is offset from the tracking gate designed for the altimeter over oceans. The<br />

present theory of altimetry, developed to describe scattering from the ocean surface<br />

such as the Brown (1977) model, does not deal properly with the geometry of ice-


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 61<br />

sheet surfaces. Therefore, onboard range measurements in altimeter records do not<br />

always correspond to the reflecting surface at nadir.<br />

retracking gate<br />

tracking gate<br />

Returned Power (counts)<br />

offset<br />

32.5<br />

Time or Bins<br />

Figure 3.4 The concept of the observed waveform retracking for ERS-1/2, in which<br />

( retracking gate − tracking ) × 0. 45<br />

offset = gate (m).<br />

Estimation of the true ranges to the reflecting surfaces can be obtained using<br />

waveform retracking techniques. Waveform retracking determines the offset<br />

(Figure 3.4) of the actual tracking gate, which is related to the half-power point on<br />

the leading edge of the waveform, from the pre-designed tracking gate, and corrects<br />

the range calculated by the on-board algorithm accordingly. Retracking algorithms<br />

developed over ice sheets can be categorised into four classes:<br />

(1) Fitting algorithm: β -parameter retracking (Martin et al., 1983);<br />

(2) The off centre of gravity (OCOG) technique (Wingham et al., 1986);<br />

(3) Threshold retracking (e.g., Bamber, 1994; Davis, 1997; Ridley and Partington,<br />

1988; Partington et al., 1989; Partington et al., 1991; Femenias et al., 1993;<br />

Zwally, 1996; Remy and Minster, 1993); and<br />

(4) A surface/volume scattering retracking (Davis, 1993a).


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 62<br />

In addition, Brooks et al. (1978) assume that the actual tracking point on the leading<br />

edge of the waveform is related to the half value of the maximum return power. This<br />

retracking method is, in fact, the same as the 50% threshold retracking. Wingham<br />

(1995b) develops a method for determining the average height of a large topographic<br />

ice sheet from altimeter observations. This method is extended to cover the ice-sheet<br />

geometry, as well as including the effects of the penetration of the surface by the<br />

radar waves, and thus differs from all above methods. A Volterra type of integral<br />

equation is derived (Wingham, 1995a), which allows the uniqueness of the solution<br />

for the average ice-sheet height to be estimated simply. However, a discrete<br />

representation is not given, so that practical implementation is not yet available.<br />

3.5.1 Fitting Algorithm - β-Parameter Retracking<br />

Martin et al. (1983) developed the first retracking algorithm for processing altimeter<br />

return waveforms over continental ice sheets. The algorithm was used to retrack all<br />

SEASAT radar altimeter waveforms to obtain corrected surface elevation estimates.<br />

This algorithm fits a 5- or 9-parameter function to the altimeter waveform, which is<br />

based partly on Brown’s surface scattering model (discussed in Section 5.4.3). The 5-<br />

parameter function is used to fit single-ramp returns (Figure 3.5), while the 9-<br />

parameter function is used to fit double-ramp returns (Figure 3.6) (Martin et al.,<br />

1983). This retracking algorithm is also known as β-parameter retracking or the<br />

NASA algorithm (e.g., Davis, 1995). It is a surface-scattering model that can deal<br />

with complex waveforms reflected from one or two scattering surfaces over ice<br />

sheets. The Ice Altimetry Group at NASA’s Goddard Space Flight Centre (GSFC)<br />

has developed a series of retracking algorithms for ice-sheet waveforms based upon<br />

Martin’s functions (Zwally, 1996). The general parameter function fitting the radar<br />

returns is given as (Martin et al., 1983; Zwally, 1996):<br />

n<br />

⎛ t − β<br />

3i<br />

⎞<br />

y(<br />

t)<br />

= β + ∑ +<br />

⎜<br />

⎟<br />

1<br />

β<br />

2i<br />

(1 β<br />

5iQi<br />

) P<br />

(3.40)<br />

i=<br />

1 ⎝ β<br />

4i<br />

⎠<br />

where


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 63<br />

Q<br />

i<br />

⎧ 0<br />

= ⎨<br />

⎩t<br />

− ( β<br />

3i<br />

+ 0.5β<br />

)<br />

4i<br />

for t < β + 0.5β<br />

3i<br />

3i<br />

4i<br />

for t ≥ β + 0.5β<br />

4i<br />

(3.41)<br />

⎛ −<br />

∫ −∞<br />

⎟ ⎞<br />

= x 2<br />

1 q<br />

P(<br />

x)<br />

exp<br />

⎜ dq<br />

2π ⎝ 2 ⎠<br />

(3.42)<br />

where n = 1 or 2 is the number of the ramp in the waveform range window which<br />

corresponds to single or double reflecting surfaces, respectively (Figures 3.5 and 3.6).<br />

Double ramps indicate that two distinct, nearly equidistant surfaces are observed by<br />

the altimeter. The unknown parameters are as follows.<br />

(1) β<br />

1: the thermal noise level of the return waveform.<br />

(2) β<br />

2i<br />

: return signal amplitude.<br />

(3) β<br />

3i<br />

: the mid-point on the leading edge of the waveform, which is chosen as the<br />

correct range point in the retracking procedure.<br />

(4) β<br />

4i<br />

: the return waveform risetime.<br />

(5) β<br />

5i<br />

: the slope of the trailing edge.<br />

When the linear trailing edge is replaced by an exponential decay term (cf., Zwally,<br />

1996), Equation (3.40) is adapted to give<br />

n<br />

⎛ t − β<br />

3i<br />

⎞<br />

y(<br />

t)<br />

= β + ∑ −<br />

⎜<br />

⎟<br />

1<br />

β<br />

2i<br />

exp( β<br />

5iQi<br />

) P<br />

(3.43)<br />

i=<br />

1 ⎝ β<br />

4i<br />

⎠<br />

where<br />

Q<br />

i<br />

⎧0<br />

= ⎨<br />

⎩t<br />

− ( β<br />

3i<br />

+ 0.5β<br />

)<br />

4i<br />

for<br />

for<br />

t < β<br />

3i<br />

t ≥ β<br />

3i<br />

− 2β<br />

− 2β<br />

4i<br />

4i<br />

(3.44)<br />

This function is used to fit the waveform with a fast-decaying trailing edge over seaice<br />

or ice sheets, which are caused by the beam attenuation (cf. Zwally, 1996). It can<br />

simulate the antennae attenuation as the pulse expands on the surface beyond the<br />

pulse-limited footprint. This function is designed to fit the fast-decaying ice mode<br />

return waveforms (ibid.). Another fitting function developed at GSFC is to fit a<br />

simple linear trailing edge to the first ramp and the exponential decay-trailing edge to


Returned Power (counts)<br />

Returned Power (counts)<br />

Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 64<br />

retracking gate<br />

tracking gate<br />

β<br />

51<br />

(trailing edge slope)<br />

β<br />

11<br />

β<br />

21<br />

β<br />

41<br />

(risetime)<br />

β<br />

31<br />

32.5<br />

Time or Bins<br />

Figure 3.5 Five- β -parameter model, fitting a single ramp function to a signal<br />

leading edge of the waveform samples.<br />

tracking<br />

gate<br />

1st retracking<br />

gate<br />

2nd retracking<br />

gate<br />

β<br />

51<br />

(1st trailing<br />

edge slope)<br />

β42<br />

(2nd risetime)<br />

β52<br />

(2nd trailing<br />

edge slope)<br />

β<br />

1<br />

β<br />

41<br />

(1st risetime)<br />

β<br />

21<br />

β<br />

22<br />

β<br />

31<br />

β<br />

32<br />

32.5<br />

Time or Bins<br />

Figure 3.6 Nine- β -parameter model, fitting a double-ramp function to two leading<br />

edges of the waveform samples.


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 65<br />

the second ramp, for which detailed information can be also found in Zwally (1996).<br />

The β parameter retrackers will be employed in the coastal zone for this project<br />

(Chapter 5).<br />

3.5.2 Off-Centre of Gravity (OCOG) Retracking Algorithm<br />

The OCOG retracking method was developed by Wingham et al. (1986), which uses<br />

full waveform samples (excluding the samples in the bins at the beginning and end<br />

due to the wrap-around error that occurs within the digital single processing). It is<br />

based on the definition of a rectangle about the effective centre of gravity (COG) of<br />

the waveform and the amplitude (A) and width (W); see Figure 3.7. The area of the<br />

rectangle equals that of the waveform. The height of the COG is defined as half the<br />

amplitude of the centre of gravity of the waveform. To reduce the effect of lowamplitude<br />

samples in front of the leading edge, the squares of the sample values are<br />

used in the computation (Wingham et al., 1986). The equation used to compute A is<br />

given by<br />

A =<br />

64−n<br />

a<br />

∑<br />

i=<br />

1+<br />

n<br />

a<br />

4<br />

i<br />

64−n<br />

a<br />

∑<br />

P ( t)<br />

P ( t)<br />

(3.45)<br />

i=<br />

1+<br />

n<br />

a<br />

2<br />

i<br />

where P i<br />

(t)<br />

is the value of waveform sample at the i th bin, n a<br />

is the number of<br />

aliased bins at the beginning and end of the waveform (Figure 3.7) depending on the<br />

altimeter (typically n<br />

a<br />

= 4 for ERS-2 and n<br />

a<br />

= 0 for TOPEX/POSEIDON). The range<br />

window is i = 1, 2, ···, 64 sample bins. The waveform width, W, is then obtained by<br />

the relation<br />

∑<br />

AW = P<br />

i<br />

(t)<br />

with the result<br />

W<br />

⎛<br />

= ⎜<br />

⎝<br />

64−na<br />

2<br />

∑ Pi<br />

i=<br />

1+<br />

na<br />

⎞<br />

( t)<br />

⎟<br />

⎠<br />

2<br />

64−na<br />

4<br />

∑ Pi<br />

i=<br />

1+<br />

na<br />

( t)<br />

(3.46)<br />

the location of the COG is formed by<br />

COG =<br />

64−n<br />

a<br />

∑<br />

i=<br />

1+<br />

n<br />

iP ( t)<br />

a<br />

2<br />

i<br />

64−n<br />

a<br />

∑<br />

i=<br />

1+<br />

n<br />

P ( t)<br />

a<br />

2<br />

i<br />

(3.47)


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 66<br />

Finally, the location of the half-power point or mid-point on the leading edge of the<br />

waveform, LEG , is given by<br />

W<br />

LEG = COG −<br />

2<br />

(3.48)<br />

W<br />

Returned Power<br />

50% retracking point estimate<br />

COG<br />

A<br />

Time or Bins<br />

Figure 3.7 Schematic description of the OCOG method applied to an observed<br />

waveform. Using the OCOG method to calculate the amplitude (A) and width (W) of<br />

the rectangle, and the position of the 50% of amplitude interpolated at the leading<br />

edge as the estimate of expected tracking gate.<br />

3.5.3 Threshold Retracking Algorithm<br />

The OCOG algorithm is simple to implement, though it is purely statistical and is not<br />

based upon any physical model of the reflecting surfaces. Moreover, it is sensitive to<br />

the waveform shape affected by surface undulations and off-nadir pointing, because<br />

it uses the full samples in waveform bins. However, when dealing with the returns at<br />

which the slope of the leading edge is smaller (i.e., longer pulse rise times), an<br />

erroneous estimate is produced because the LEG cannot be precisely positioned at<br />

the mid-point of the leading edge of the waveform (e.g., Partington et al., 1989).<br />

To improve the estimation, an empirical method of threshold retracking was<br />

developed and is used by ESA to process altimeter data from the ERS satellite<br />

missions (Davis, 1995). It has also been adopted by NASA/GSFC as an alternative<br />

retracking algorithm for the production of ice-sheet altimeter data sets (Davis, 1997).


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 67<br />

This algorithm is based upon the dimensions of the rectangle (i.e., amplitude, width,<br />

and hence the ‘centre of gravity’) computed using the OCOG method. The threshold<br />

value is then referenced to the amplitude (Bamber, 1994) or the maximum waveform<br />

sample (e.g., Davis, 1997; Zwally, 1996) estimate of the rectangle at 25%, 50%, and<br />

75% of the waveform amplitude. The retracking gate estimate is determined by<br />

linearly interpolating between adjacent samples of a threshold crossing at a steep part<br />

of the leading edge slope of the waveform. This algorithm keeps the same<br />

advantages as the OCOG, but it can determine a more accurate tracking gate position<br />

than the OCOG does (Partington et al., 1989). A disadvantage is that, like the OCOG<br />

method, it is not based on a physical model.<br />

The selection of an optimum threshold level is important when applying this method<br />

to waveforms, because the range to the surface is determined from it. Davis (1995;<br />

1997) tests different threshold levels (10%, 20%, and 50% of the waveform<br />

amplitude) for the purposes of measuring ice-sheet elevation change. The<br />

comparison is performed on SEASAT, GEOSAT-GM, and GEOSAT-ERM<br />

crossover data. It was found that a 10% threshold level can produce more repeatable<br />

elevation estimates, a 20% threshold level is suitable for obtaining the true ice-sheet<br />

elevation in only an average sense, and a 50% threshold level is appropriate only<br />

when surface scattering dominates the return waveform shape.<br />

Thus far, no account has been taken of the thermal noise effect (described in Section<br />

3.4.1) in Equations (3.45), (3.46) and (3.47). However, since the magnitude of the<br />

thermal noise varies with location and time in a given waveform dataset, these<br />

variations cause the leading-edge retracking point to vary if the threshold level is<br />

referenced only to the amplitude in which the thermal noise is included. To<br />

overcome this problem, the mean thermal noise estimated by averaging over earlier<br />

gates 6-10 should first be subtracted from the amplitude (Zwally, 1996).<br />

3.5.4 Surface/Volume Scattering Retracking Algorithm<br />

In general, the leading edge of altimeter waveforms is characterised by a rapid rise in<br />

return power and continuing to increase to the plateau (Figures 2.3 or 3.7). Most<br />

leading edges over ice sheets, however, show first a rapid rise in the return power.<br />

Then, they continue to rise but at a reduced rate, indicating the influence from


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 68<br />

subsurface volume scattering (e.g., Ridley and Partington, 1988; Davis, 1993a). The<br />

reason is that the radar pulse can significantly penetrate the snow over most of the ice<br />

sheet, resulting in elevation estimation errors of up to a few metres (e.g., Partington<br />

et al., 1991) if seeking to determine the height of these surfaces. Therefore, when<br />

retracking is carried out using the β-parameter retracking algorithm, there is an<br />

implicit assumption that the ice sheet surface scatters the radiation in a similar<br />

manner to the ocean surface.<br />

To take into account radar penetration over ice sheets, Partington et al. (1991) use a<br />

threshold retracking method but the position of the retracking gate is selected from<br />

an average waveform derived from all SEASAT waveforms from the region of<br />

interest. To a first order, this method overcomes the penetration problem (ibid.).<br />

Using Gaussian approximations for the altimeter’s antenna pattern and transmitted<br />

pulse shape (i.e., the radar system PTR), Davis and Moore (1993b) derive a closedform<br />

analytical solution for the received power due to volume scattering. Because of<br />

this closed-form, they combine the volume-scattering model with the Brown surfacescattering<br />

model (Equation 3.12) to develop an algorithm capable of retracking<br />

altimeter waveforms over ice sheets. The combined surface and volume-scattering<br />

model is (Davis and Moore, 1993b)<br />

Am<br />

⎡ K ⎤<br />

SV ( n)<br />

= DC + ⎢S(<br />

n)<br />

+ V ( n)<br />

⎥<br />

S<br />

2 ⎣ S1<br />

⎦<br />

(3.49)<br />

in which DC is the thermal noise baseline, A m<br />

is the maximum amplitude of the<br />

model waveform, and S (n)<br />

is an adjusted form of the Brown model (cf. Equation<br />

3.12) given by<br />

1 ⎡ ⎛ t − t ⎡−<br />

⎤<br />

0 ⎞⎤<br />

4c<br />

S( n)<br />

= ⎢1<br />

+ erf ⎜ ⎟⎥exp⎢<br />

( t − t0)<br />

2<br />

⎥<br />

⎣ ⎝ 2σ ⎠⎦<br />

⎣ γ h ⎦<br />

(3.50)<br />

The derivation of Equation (3.50) from Equation (3.12) assumes that the off-nadir<br />

angle ξ = 0 , and the modified Bessel function is expanded to order zero ( n = 0 in<br />

Equation 3.17). The volume-scattering model, V (n)<br />

, is given by


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 69<br />

2 2 2<br />

[ β k c − 2k<br />

c ( t − t )]<br />

2 2<br />

⎛ c β c(<br />

t t0<br />

) ⎞<br />

τ<br />

−<br />

V ( n)<br />

= exp⎜<br />

⎟<br />

− − exp<br />

2 4<br />

2<br />

e s e s 0<br />

4h<br />

c<br />

h<br />

τ<br />

⎝ β β<br />

c ⎠<br />

(3.51)<br />

where constants S 1<br />

and S<br />

2<br />

are normalising factors that ensure the quantities V (n)<br />

and S ( n)<br />

+ V ( n)<br />

range from zero to one, K represents the correct proportion of<br />

volume scattering, c s<br />

is the speed of light in snow, β c<br />

is a constant related to the<br />

antenna beamwidth, β τ<br />

is a constant that determines the 3 dB width of the<br />

transmitted pulse, and k e<br />

is the extinction coefficient of snow. This is a fitting<br />

function, and there are six unknown parameters in the model ( DC , σ , t 0<br />

,<br />

A<br />

m<br />

, K<br />

and k<br />

e<br />

). This model will not be used for the coastal zone retracking in this study,<br />

because there is no known penetration problem present at the coast.<br />

3. 6 Land Waveform Retracking<br />

In spite of the fact that the satellite radar altimeter was developed to measure the<br />

ocean surface, and the on-board algorithms used to control the receiver gain and<br />

height tracker were not designed to cope with terrain signals, substantial quantities of<br />

seemingly useless data were obtained over land. However, SEASAT was the first<br />

altimeter to show its mapping capability over land (and ice-sheet) surfaces. Over<br />

about three months of its lifetime, it collected data over 34% of the Earth’s land<br />

surface (Tapley et al., 1987, cited in Guzkowska et al., 1988). GEOSAT data were<br />

also collected over land both in the 17-day ERM orbit and during period of the GM<br />

(Brenner et al., 1990). Moreover, with the advent of ERS-1/2, the ice mode with a<br />

longer sampling time on the altimeter and different orbit repeat phases provided far<br />

better spatial coverage of data (~80% of the Earth’s surface) for topographic<br />

purposes than previous satellite missions (Berry, 2000).<br />

Waveform data from different satellite altimeters collected over land surfaces have<br />

been analysed by Guzkowska et al. (1988), Brenner et al. (1990) and Berry (2000a).<br />

Guzkowska et al. (1988) categorise the SEASAT waveform data into seven classes,<br />

in which each type of data shows unique characteristics corresponding to the<br />

reflecting surface. Berry (2000) also presents a selection of 10 classes of waveforms<br />

with different characteristics for ERS-1 over land. These results indicate that specific


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 70<br />

problems have been associated with the attempt to retrieve an accurate range to<br />

surface over land. One of them, which is the same as that encountered over ice-sheet<br />

surfaces, is that the on-board satellite processor cannot calculate an accurate range to<br />

surface, because the altimeter is in general unable to centre the return reflected from<br />

the rapidly varying topographic surface in the range window. Thus, in order to work<br />

with these data, waveform retracking must be employed.<br />

Land waveform retracking is mainly conducted via two approaches. The first is to<br />

retrack waveforms using an expert system depending on the specific waveform<br />

categories, which was developed by Berry et al. (1998), Berry (2000a) and Berry et<br />

al. (2000b). ERS-1/2 geodetic missions, plus their operation in the ice mode, resulted<br />

in the acquisition of large and effective volumes of radar returns over land surfaces.<br />

Berry et al. (1998) generate orthometric heights from ERS-1 geodetic mission dataset<br />

using retracked waveform data from an expert systems approach. Data from the<br />

entire ERS-1 geodetic mission has been reprocessed to generate a Regional Altimeter<br />

Result database, and the data are then used to generate orthometric heights.<br />

Land topography, in particular topography with a non-uniform slope, and the nonuniform<br />

reflectivity of the land are main courses that lead to the complicated return<br />

waveform shapes over land. The expert system, on the other hand, contains several<br />

retrackers (Berry, 2000a, 2001). Each retracker is specifically designed and<br />

developed for a particular type of return waveform, for which ten retrackers have<br />

been created for a land waveform-retracking system (Berry, 2001). Some of them are<br />

(Berry et al., 1998):<br />

(1) Spline retracker, which is more effective with pre-peaked waveforms by fitting<br />

bicubic spline functions to the leading edge of the waveform;<br />

(2) Narrow waveforms, for high power specular returns from inland;<br />

(3) Bor retracker, for waveforms showing slope information;<br />

(4) Bog retracker, for Bor-type waveforms; and<br />

(5) Default retracker, i.e., threshold retracker, for waveforms that cannot be<br />

retracked by other retrackers.<br />

Two steps have been taken when reprocessing waveform data from land. The first<br />

step is to sort the waveforms based on their shape and retain all waveforms which


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 71<br />

show a leading edge regardless of complexity of the whole waveform shape. Next,<br />

choosing the optimised retracker to estimate the range to the land surface at the nadir<br />

based upon the individual waveform shapes. More detailed information about the<br />

expert system of land waveform retracking can be found in Berry et al. (1998) and<br />

Berry (2000).<br />

The second approach to land waveform retracking is achieved by retracking all<br />

waveforms using a single retracker, typically the OCOG technique (e.g., Bamber et<br />

al, 1997, cited in Berry, 2000) or one optimised for double-peaked waveforms (eg.,<br />

Brenner et al., 1997, cited in Berry, 2000). Before the retracker is applied to<br />

waveform data, they are edited and filtered so that the single retracker can succeed<br />

for a larger percentage of data. Shape corrections are then applied to individual<br />

retracked ranges to surface and combined to derive mean or median height values.<br />

Results from this approach represent the land topography with a spatial resolution of<br />

15' or 5' (Berry, 2000).<br />

The mathematical methods for these have not been presented because 1) they will not<br />

be used in this dissertation, and 2) moreover, they are not presented by authors.<br />

3. 7 Brief Introduction to Coastal Waveform Retracking<br />

The potential of the altimeter waveform data to be used in coastal regions has been<br />

the subject of comparatively little study (Anzenhofer et al., 2000; Brooks and<br />

Lockwood, 1990; Brooks et al., 1997). So far, considering the existing retracking<br />

algorithms, two approaches have been applied to retracking coastal waveforms,<br />

which are the fitting procedure and threshold technique. The first is to fit an<br />

analytical waveform model to the observed waveform data for determining the more<br />

accurate range estimate. The fitting function used from open literature<br />

(e.g., Anzenhofer et al., 2000) is the five-β-parameter function Equations (3.40) and<br />

(3.43), choosing either the linear trailing edge or the exponential trailing edge. When<br />

using the function fitting algorithm, usually n = 1 in Equations (3.40) and (3.43), an<br />

iterative least squares procedure is used.<br />

An alternative method used by Brooks et al. (1997) in coastal regions is a slightly<br />

modified threshold retracking technique, which tracks the ocean return from the


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 72<br />

hybrid waveforms of water and land close to the coastline. This retracking involves a<br />

three-stage process:<br />

(1) Returns are observed using a visual examination method to identify the ocean<br />

and land returns.<br />

(2) A threshold level is selected for each individual pass such that the threshold is<br />

higher than both the thermal noise level and the minor waveform leakage in the<br />

early gates, and is sufficiently low to intercept the ocean return part in the<br />

waveform. Then, the actual tracking gate position of the ocean return is<br />

obtained by converting the intercepted amplitude on the leading edge of the<br />

waveform to the gate number.<br />

(3) The position of the expected tracking gate is ascertained by averaging midpoint<br />

positions on the leading edge of the waveforms from several seconds of openocean<br />

waveforms in the same area, pass and date. The difference between this<br />

expected tracking gate and the actual tracking gate is computed and then<br />

converted to the range correction in metres. This range correction is applied to<br />

the range measurement from the on-board satellite-tracking algorithm.<br />

This method is appropriate for retracking the mixed waveforms in coastal regions.<br />

However, it cannot be conducted automatically (Brooks, et al., 1997). The method<br />

developed to select the threshold level by the author in Chapter 5 will solve this<br />

problem.<br />

3. 8 Summary<br />

An outline of waveform retracking algorithms, both for ocean and non-ocean<br />

surfaces, has been presented in this Chapter. Starting from the basic convolutional<br />

representation of the waveform, the Brown model is introduced. Then, it is extended<br />

to a form that includes additional corrections (i.e., skewness and Earth’s curvature)<br />

so that it can be used to estimate more parameters over non-linear ocean surfaces.<br />

Finally, a model which will be used in this study is deduced. A detailed qualitative,<br />

and some quantitative, analysis is carried out on this model.<br />

Over non-ocean surfaces, waveform-retracking algorithms have been developed for<br />

obtaining reasonably accurate elevation estimates from altimeter data over ice-sheet


Chapter 3. RADAR ALTIMETER WAVEFORM RETRACKING METHODS 73<br />

and land surfaces. Over ice-sheets, retracking algorithms are mainly based upon<br />

either a linear surface scattering model (i.e., the Brown model) or the threshold<br />

method. In addition, a volume and surface scattering model extends the Brown<br />

model to deal with the returns from snow-covered ice sheets. Over land surfaces,<br />

existing retracking algorithms are either those used over ice sheets or a retracking<br />

system. Because a physically-based model is impossible over land due to the more<br />

rugged and rapidly varying surface, an expert system that depends on the waveform<br />

shapes is currently the only apparently effective retracking method.<br />

It is important to note that improvement of the altimeter data in coastal regions has<br />

been the subject of comparatively little study, although many previous studies have<br />

reported the data are in error in these regions (e.g., Mantripp, 1996; Brooks and<br />

Lockwood, 1990; Hwang, 1998; Featherstone, 2003). The existing retracking<br />

algorithms are also either the β-parameter fitting procedure or the threshold<br />

algorithm for ice-sheet surfaces. This suggests that it is necessary to investigate<br />

further the coastal retracking algorithms, which is the objective of this dissertation.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 74<br />

4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS<br />

4. 1 Introduction<br />

In order to avoid contaminated data in coastal regions, Nerem (1995) and Shum et al.<br />

(1997) in their studies of analysis of global mean sea level and ocean tidal models<br />

over open oceans, simply eliminate altimeter SSH data in areas with water depths<br />

less than 200 m. Shum et al. (1998) show that it is necessary to improve correction<br />

algorithms for data near coastal regions and semi-enclosed seas. Brooks et al. (1990;<br />

1997) investigate land effects on GEOSAT and TOPEX radar altimeter<br />

measurements in the Pacific Rim coastal zone, and provide a slightly modified<br />

threshold retracking algorithm for the data user to rectify and recover the sea surface<br />

topography in close proximity to the land. Although some of these efforts have been<br />

reported previously in the open literature, empirical evidence is not yet published to<br />

show how far away from the shoreline altimeter measurements are poor.<br />

This Chapter uses two approaches of the threshold retracking (Section 3.5.3) and<br />

mean waveform analysis of one cycle of ERS-2 and five cycles of POSEIDON 20<br />

Hz waveform data to quantify the distance from the Australian shoreline that<br />

altimeter measurements are contaminated. This distance is obtained by using the 50<br />

percent threshold retracking point estimates and analysing mean waveform and its<br />

standard deviation across a zone extending out to 350 km offshore from Australia. In<br />

this Chapter, the modelling (and retracking) of the altimeter waveform will not be<br />

addressed. Instead, only the distance from the coastline in which sea surface heights<br />

are poorly estimated by the on-satellite tracking algorithm will be examined.<br />

Quantified distance results will be presented in Section 4.5. In addition, several<br />

examples from waveforms along satellite ground tracks near the coastline are taken<br />

to analyse in detail how waveforms are contaminated and what the characteristics of<br />

contaminated waveforms are in coastal regions (Section 4.6). A preliminary test that<br />

estimates the land effects on waveforms in the vicinity of land will be given in<br />

Section 4.7.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 75<br />

4. 2 Background<br />

When the altimeter pulse interacts with the Earth’s surface, the shape of a returned<br />

waveform depends on the topography of the reflecting surface. Over open oceans,<br />

the altimeter tracker operates as designed and thus uses the Brown (1977) ocean<br />

statistics model (Section 3.2). The on-satellite tracking algorithm can well centre the<br />

ocean return leading edge at a ‘tracking gate’, which is a pre-designed position at the<br />

range window corresponding to the range to the nadir mean sea surface. Over land<br />

and ice sheets, however, the reflection of a satellite radar altimeter signal differs from<br />

the Brown model owing to the non-Gaussian-reflecting surface, where waveforms<br />

are different to the expected/ideal waveform shape over oceans are generally<br />

received (Sections 3.5 and 3.6). This leads to an error in the on-satellite range<br />

estimates that must be corrected through retracking techniques. As stated, retracking<br />

altimeter data aims to estimate the departure of the midpoint on the waveform’s<br />

leading edge from the pre-designed tracking gate, and thus correct the on-satellite<br />

range measurements accordingly (Martin et al., 1983; Bamber, 1994; Berry et al.,<br />

1998).<br />

In coastal regions, the situation differs from open oceans mainly because of the close<br />

proximity to land and the complex spatial-temporal variation of coastal water<br />

surfaces. As a satellite ground track approaches, recedes or runs parallel to the<br />

coastline, even though the altimeter’s nadir point is over the sea, the altimeter may<br />

also track the off-nadir returns reflected from the higher land and any brighter still<br />

inland water. This causes a problem in that the altimeter cannot follow the apparent<br />

accelerations in the surface shape. As a result, the waveforms recorded on board the<br />

satellite will become misaligned with respect to the expected tracking gate (i.e.,<br />

contaminated). Therefore, a postprocessing or retracking is necessary as well on the<br />

on-satellite range calculation (Brooks and Lockwood, 1990; Mantripp, 1996; Quartly<br />

et al., 2001), which will be covered in Chapters 5, 6 and 7. In this Chapter, the spatial<br />

extent of this contamination is quantified so that efficiency can be achieved by<br />

learning over what range to retrack coastal data.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 76<br />

4.2.1 ERS-2 and POSEIDON Altimeter Footprints<br />

The altimeter-illuminated ground region (i.e., along-track footprint) from which most<br />

power returns for each 20 Hz measurement is typically oval in shape and its size<br />

depends on the satellite altitude, surface slope and roughness, and the width of<br />

waveform gates used to compute its diameter (e.g., Chelton et al., 2001, p.36). Two<br />

types of altimeter along-track footprint are important, which are the middle gate and<br />

the automatic gain control (AGC) along-track footprints. The middle gate is the<br />

average of 2, 4, 8 or 16 waveform samples centred on the leading edge of the ocean<br />

return, and the middle-gate along-track footprint is defined by the half-width of the<br />

middle-gate after the tracking gate. For the sea surface with SWH=0 m (i.e., a flat sea<br />

surface) and SWH=15 m, the size of the middle-gate along-track footprint varies<br />

from 8.5 km to 15 km for ERS-2 and 8.5 km to 16 km for POSEIDON. The AGC<br />

gate is the average of waveform samples 17 through 48 centred on the leading edge<br />

of the ocean return. Similarly, the AGC along-track footprint is computed from 15.5<br />

AGC gates after the tracking gate. Its diameter for POSEIDON is about 13.9 km for<br />

a flat sea surface, increasing to about 16.9 km for SWH=15 m (Chelton et al., 2001,<br />

p. 36 and Figure 22).<br />

The on-satellite tracking algorithm compares the returned signal from the middle<br />

gate to that from the AGC gate to estimate the range to nadir sea surface. Since the<br />

AGC footprint is larger than the middle gate footprint, a sharp change in scattering<br />

surface will be first sensed by the AGC gate before the middle gate is affected. In<br />

addition, the footprint diameter estimated by all waveform gates after the tracking<br />

gate is much larger than the above-mentioned sizes, such as 22 km for TOPEX<br />

mission (Brooks et al., 1997). Therefore, it is necessary to consider a larger footprint<br />

when investigating waveform contamination in coasts. However, the exact distance<br />

from the coast at which this occurs is unknown. Hence, this is a secondary<br />

justification for this Chapter.<br />

4.2.2 ERS-2 and POSEIDON Altimeter Waveforms<br />

Figures 1.2 and 4.2 show typical examples of observed ERS-2 and POSEIDON<br />

waveforms over oceans, respectively. As can be seen, they are somewhat different<br />

from the ideal waveform (Figure 1.1), due mainly to reasons of the ‘wrap-around’


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 77<br />

200<br />

150<br />

Power (counts)<br />

100<br />

50<br />

0<br />

0 8 16 24 32 40 48 56 64<br />

Bins or Gates<br />

Figure 4.1 An example of the observed POSEIDON waveform over deep oceans<br />

(tracking gate = 31.5).<br />

error and ‘fading’ (Rayleigh) noise (cf. Partington et al., 1991; Quartly et al., 2001).<br />

The former causes a reduction in power in the last few bins and a commensurate rise<br />

at the beginning of the waveform, and the later causes an undulating trailing edge in<br />

the waveform. The noise affects the leading edge as well, but it is more difficult to<br />

identify in the leading edge and is clearly seen in the trailing edge. In addition,<br />

Figure 1.2 does not show the thermal noise because the ERS-2 altimeter artificially<br />

suppresses this noise (cf. Quartly et al., 2001).<br />

Since altimeter-observed waveforms are different from the Brown model (Figure 1.1),<br />

it is hard to correctly define the tracker point to be tracked by altimeter. Therefore, it<br />

is displaced in the observed waveform by the tracking gate. After subtracting the<br />

noise baseline (i.e., the thermal noise level) from each waveform gate, the tracking<br />

gate is defined by a gate (or bin) number that corresponds to the midpoint on the<br />

leading edge of the returned waveform. On-satellite algorithms try to centre the ramp<br />

of the waveform at the tracking gate, but this generally succeeds only over open<br />

oceans. The tracking gate is specified prior to launch as part of the overall design of<br />

the radar system and corresponds to the mean surface range over the ocean (Brown,


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 78<br />

1977). The tracking gate is chosen to align between waveform sample bins 31-32 for<br />

POSEIDON and bins 32-33 for ERS-2 (e.g., Quartly et al., 2001). The parameters<br />

linked to the reflecting surface can be estimated from the detailed shape of the<br />

leading edge, trailing edge and precise position of the waveform in the sampled<br />

window relative to the altimeter tracking gate location.<br />

4. 3 Processing Methodology for Determination of Contamination Area<br />

Whenever land is within the altimeter footprint, there is a likelihood that the land<br />

return contaminates the waveforms, depending on the height and type of terrain<br />

along the shoreline. In order to quantify the contaminated distance from the shoreline,<br />

it is necessary to calculate waveform anomalies with respect to the ideal waveform<br />

shape. Fitting an ocean-like function to each individual waveform then calculating<br />

the standard deviation of the differences is a reasonable measure of determining if<br />

the waveform contains any departure from the ocean-type response. However, it is<br />

not a good way of calculating a better sea surface height in coastal regions because<br />

the model does not fit the actual return. Therefore, two processing techniques are<br />

used in this Chapter: threshold retracking and arithmetic averaging. The<br />

contaminated area is determined by computing the tracking gate estimates and<br />

analysing the shape of the mean waveform, instead of only focussing upon the<br />

individual waveform shapes. Waveform retracking at the Australian coast will be<br />

covered in Chapter 6.<br />

4.3.1 Threshold Retracking<br />

The threshold retracking technique (see Section 3.5.3) is employed in this Chapter<br />

due to its perceived advantage of sensitivity to the surface topography (Partington et<br />

al., 1989). To estimate a contaminated boundary for coastal waveforms, it is essential<br />

to distinguish whether the returned waveforms are affected by coastal topography<br />

(land and sea-states). Unlike the returned waveform over land or ice, where the<br />

waveform shape is significantly different from that over the ocean, most waveforms<br />

affected by land in coastal regions usually show a combination of water and land<br />

(Brooks et al., 1997). Therefore, if the actual position of the mid-point on the leading


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 79<br />

edge of the ocean waveform can be determined accurately, the purpose of estimating<br />

a contaminated boundary for coastal waveforms will be achieved.<br />

As stated, assuming that the reflecting surface approximates a uniform, spherically<br />

smooth, and diffuse surface with a symmetrical distribution of small scale slopes, the<br />

range from the satellite to the mean surface at nadir corresponds to the location of the<br />

half-power point (i.e., the midpoint) on the leading edge of the waveform (Brown,<br />

1977). These assumptions of statistical homogeneity are usually satisfied over open<br />

ocean surfaces. Thus, the 50% threshold value can be used as the accurate estimate<br />

of the mid-point on the leading edge of the ocean waveform. Since the 50% retracked<br />

position should agree with the tracking gate for ocean waveforms, it may be different<br />

from the tracking gate for the waveform affected by non-ocean surfaces. This can, in<br />

turn, be used to assess the contamination of the waveform in coastal regions. This is<br />

the principle on which the identification of contaminated altimeter data is based in<br />

this study.<br />

When applying the threshold-retracking algorithm, the formulae used to compute A,<br />

W and COG are given by Equations (3.45), (3.46) and (3.47). The 50% threshold<br />

value of the amplitude of the rectangle is used as the estimate of the mid-point on<br />

leading edge of the waveform in this Chapter. It is linearly interpolated between the<br />

bins adjacent to the threshold value at the leading edge of the waveform. Then, its<br />

position at the range window (i.e., the bin or gate number) can be computed from this<br />

estimate. This position is then compared with the expected tracking gate (bins 32-33<br />

for ERS-2) to determine the contaminated distance.<br />

For the ERS-2 altimeter, there is no need to account for the thermal noise effect.<br />

However, attention must be paid to it for the POSEIDON altimeter waveform data.<br />

To overcome this problem, the mean thermal noise estimated by averaging over gates<br />

5-10 is subtracted from the amplitude; the threshold-retracking scheme adopted here<br />

is referenced to 50% of the amplitude above the mean thermal noise level. This<br />

approach has been applied to the POSEIDON data used in this study.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 80<br />

4.3.2 Mean Power of the Waveform<br />

In order to provide a statistically meaningful assessment of the contaminated area, an<br />

averaging procedure was carried out around the Australian coast. This aims to<br />

calculate a broad contaminated area by analysing shape variations between the mean<br />

coastal waveform and the theoretical waveform due to the Brown model. It is likely<br />

that the variation in scattering surfaces within the coastal footprint of the altimeter<br />

creates the error in the individual returned waveforms. Averaging can reduce this<br />

‘noise’ and leave a mean waveform, which should represent the typical shape of the<br />

waveform over the reflecting surfaces (cf. Partington et al., 1989; Quartly et al.,<br />

2001). The mean waveform generated by this averaging technique is then compared<br />

with the theoretical or ideal waveform to estimate the general differences and hence<br />

the contaminated distance from the coastline.<br />

There are a number of possibilities for the variation of the mean waveform shape.<br />

Firstly, the altimeter measurement is both a waveform and a gain measured by the<br />

AGC. The signal processor maintains the value of the AGC gate at a specified value<br />

by adjusting the attenuator in the receiver. Although the AGC applies this variable<br />

attenuation to the return signal, the mean waveform shape will still vary with sharp<br />

changes in the scattering surface. Secondly, the statistical homogeneity property of<br />

the scattering surfaces within coastal areas is changed and probably invalid. Land<br />

returns may contribute to the mean waveform, thus causing a variation. If this is the<br />

case, the variation in different areas can provide the indication of the contaminated<br />

distance from the coast. Finally, the SWH and the satellite off-nadir pointing angle<br />

may cause the variations in the slope of the leading edge and in the trailing edge of<br />

the mean waveform. However, mean waveforms with different slopes and trailing<br />

edges should cross their leading edges at the same position as the tracking gate over<br />

oceans (Brenner et al., 1993; Hayne, 1980). In this case, the contaminated distance<br />

can be obtained from combining the position of the retracking point with the<br />

variation of the mean waveform. Therefore, using the variation of the mean<br />

waveform shape and its standard deviation in different coastal regions should give a<br />

useful indication of the areas in which the waveform data may be contaminated.<br />

For each waveform-sampled bin, the mean power of waveform M (i)<br />

and its<br />

standard deviation STD (i)<br />

are calculated by


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 81<br />

N<br />

1<br />

M<br />

i<br />

( t)<br />

= ∑ Pij<br />

( t)<br />

(4.1)<br />

N<br />

j=<br />

1<br />

STD ( t)<br />

=<br />

i<br />

n<br />

∑<br />

j=<br />

1<br />

( P ( t)<br />

− M<br />

ij<br />

n −1<br />

i<br />

( t))<br />

2<br />

(4.2)<br />

where n is the number of waveforms used in the averaging procedure. Averaging<br />

was carried out over different distance bands with increasing distance from the<br />

Australian coast. These distance bands were used to investigate broad variations in<br />

waveform shape over the spatial-scale of the coastal region and over the time-scale<br />

of a cycle.<br />

4. 4 Data and Editing<br />

4.4.1 ERS-2<br />

One cycle (March to April 1999) of ERS-2 35-day repeat orbit individual waveforms<br />

was used in this study (Section 2.3.2). The geographical distribution of the<br />

observations out to 350 km around the Australian coast is displayed in Figure 4.2.<br />

Each ERS-2 waveform record consists of 1 Hz and 20 Hz data. Geophysical and<br />

environmental corrections are supplied at 1 Hz, and observations, such as the range,<br />

0<br />

location, σ , and waveforms, are supplied at 20 Hz (Section 2.4.2).<br />

Since the purpose of this Chapter is to analyse waveform contamination over coastal<br />

regions, all land returns must be exactly edited out when near the shoreline. For<br />

ERS-2, a flag is set to record the working mode changes based on the location of the<br />

sub-satellite point. When the satellite is over the land, it is flagged as in ice mode.<br />

Therefore, it might be possible to estimate the location of shoreline by observing the<br />

changes of the flag. However, the locations near the Australian coastline are usually<br />

flagged incorrectly (Figure 4.3). This means the flag supplied with the data is not<br />

appropriate for determining the location of the coastline.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 82<br />

110˚<br />

120˚<br />

130˚<br />

140˚<br />

150˚<br />

160˚<br />

0˚<br />

0˚<br />

-10˚<br />

-10˚<br />

-20˚<br />

-20˚<br />

-30˚<br />

-30˚<br />

-40˚<br />

-40˚<br />

110˚<br />

120˚<br />

130˚<br />

140˚<br />

150˚<br />

160˚<br />

Figure 4.2 ERS-2 ground tracks (35-day repeat orbit, from March to April 1999) to a<br />

distance of 350 km from Australian shoreline (Lambert projection).<br />

0˚<br />

110˚<br />

120˚<br />

130˚<br />

140˚<br />

150˚<br />

160˚<br />

0˚<br />

-10˚<br />

-10˚<br />

-20˚<br />

-20˚<br />

-30˚<br />

-30˚<br />

-40˚<br />

-40˚<br />

110˚<br />

120˚<br />

130˚<br />

140˚<br />

150˚<br />

160˚<br />

Figure 4.3 The ocean-ice-mode flag in ERS-2 data supplied close to the Australian<br />

shoreline compared to the Wessel (2000) coastline, showing that most flags around<br />

Australia are on the land (Lambert projection).


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 83<br />

Instead, to identify if the sub-satellite point is over ocean or land near the coastal<br />

regions, the DS759.2 (5′×5′ resolution) ocean depth model (Dunbar, 2000) and a<br />

water/land grid (0.5′×0.5′ resolution) derived from GMT high-resolution shoreline<br />

data (Wessel, 2000) have been used together, instead of using the ERS-2 flags (see<br />

Section 2.4).<br />

4.4.2 POSEIDON<br />

The geographical distribution of the POSEIDON coastal observations around<br />

Australia’s coasts are shown in Figure 4.4. A POSEIDON waveform record consists<br />

of 1 Hz and 20 Hz data. However, all data are supplied at 1 Hz except for 20 Hz<br />

waveform measurements (AVISO/Altimeter, 1998). The land/water flag for<br />

POSEIDON is provided by the AVISO in the waveform products. It is directly used<br />

in this study due to its fairly good agreement with the actual shoreline (Section 2.4.3).<br />

110˚<br />

120˚<br />

130˚<br />

140˚<br />

150˚<br />

160˚<br />

0˚<br />

0˚<br />

-10˚<br />

-10˚<br />

-20˚<br />

-20˚<br />

-30˚<br />

-30˚<br />

-40˚<br />

-40˚<br />

110˚<br />

120˚<br />

130˚<br />

140˚<br />

150˚<br />

160˚<br />

Figure 4.4 POSEIDON ground tracks (10-day repeat orbit, from January 1998 to<br />

January 1999) to a distance of 350 km from Australian shoreline (Lambert<br />

projection).


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 84<br />

4. 5 Determination of Contamination Area<br />

Before calculating the retracking point and averaging the waveforms, geophysical<br />

and environmental corrections were applied to all altimeter range measurements. The<br />

GSHHS shoreline model (Wessel and Smith, 1996) has been employed for<br />

determining the position of the shoreline with the exception of the POSEIDON data.<br />

The distance referred to below is the shortest distance (on the sphere) between the<br />

point observed on the satellite ground track and the shoreline point from the GSHHS.<br />

Since 1 Hz range measurement is usually taken by averaging 20 waveforms, the<br />

whole 20 Hz data record is kept in the calculation, though some waveforms may be<br />

returned from only the land in coastal regions. The purpose is to analyse how the 1<br />

Hz range data averaged from 20 Hz data are contaminated in coastal regions, since<br />

these 1 Hz data are used by most users from GDRs.<br />

4.5.1 Using 50% of Threshold Retracking Points of ERS-2<br />

The Australian coastal area is subdivided into 10 areas of 15°×15°, which will be<br />

used in Sections 5.6 and 5.8. Within these areas, if the value of the 50% threshold<br />

retracking point is between bins 31-33, the mid-point is considered to be at the<br />

expected location, otherwise the data are considered to be contaminated by land or<br />

coastal sea surface states. The criterion of bins 31-33 is referenced to the statistics of<br />

50% retracking points over ocean areas (i.e., 50-350 km offshore Australia), where<br />

96% of the values of 50% retracking points lie within bins 31-33. This proportion<br />

agrees with that in open oceans (cf. Quartly, 2001). This criterion is used to analyse<br />

the impact from land or coastal scattering surfaces on the individual waveform when<br />

a satellite ground track is within a few kilometres of the shoreline. The calculations<br />

were carried out using ERS-2 data and the results summarised in Table 4.1. Figure<br />

4.5 shows contaminated waveforms along satellite ground tracks in a small, but<br />

representative, coastal area (135° ≤ λ < 150°, -45° ≤ ϕ < -30°) based on these<br />

calculations.<br />

It can be seen from Table 4.1 that although the contaminated waveforms comprise a<br />

small percentage (from 0.2% to 2.9%, depending on different sub-areas), they are all<br />

in close proximity to the coastline. The more contaminated data and the longest<br />

contaminated distances occur off the north Australian coast and off the south


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 85<br />

Table 4.1 Contaminated 1 Hz ocean depths and 20 Hz distances from the nearest shoreline in different<br />

Australian coastal regions for the ERS-2 waveform data (20Hz, March to April 1999).<br />

a<br />

Areas<br />

Max.<br />

Distance a<br />

(km)<br />

Min.<br />

Distance<br />

(km)<br />

Mean<br />

Distance<br />

(km)<br />

Max.<br />

Ocean<br />

Depth b<br />

(m)<br />

% of<br />

Contaminated<br />

Waveforms c<br />

Coastal Types d<br />

120°≤λ


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 86<br />

Figure 4.5 One cycle of ERS-2 contaminated waveforms (highlighted by circles)<br />

along ground tracks in the southeast coasts of Australia.<br />

Australian coast, near Tasmania. There are two reasons for this. The first is the<br />

complex shoreline in these coastal regions. The second is that there are more satellite<br />

ground tracks in north and south of Australia because of the inclination of the<br />

satellite orbit with respect to the Australia landmass. The eastern and western<br />

Australian coasts show less contaminated data and a shorter contaminated distance.<br />

Overall however, the shortest and the longest contaminated distances from the<br />

coastline are 0.02 km and 21.76 km, respectively; and the mean contaminated<br />

distance is 6.85 km.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 87<br />

It can be seen from Figure 4.2 that some ERS-2 ground tracks are not available for<br />

this cycle near the Australian coast. This is another (lesser) reason why a small<br />

percentage of waveforms appears to be affected to the east and west. Thus, it is<br />

necessary to estimate a broad boundary around Australia, which contains the most of<br />

contaminated data, by using waveform data from both ERS-2 and POSEIDON<br />

altimeters.<br />

The ocean depths interpolated from an Australian 30" bathymetric model (Buchanan,<br />

1991) for these contaminated data range from 12.2 m to 361.5 m (Table 4.1). From<br />

Table 4.1, it is not evident that the contaminated distance correlates with the ocean<br />

depth, while 50% retracking points show some correlation with ocean depths (Figure<br />

4.6). Also, correlation with depth varies from one area to another. Waveforms over<br />

areas where the ocean depth is less than 50 m (Figure 4.6a) and<br />

100 m (Figure 4.6c) appear to be highly contaminated, while some contaminated data<br />

are found in the deeper (up to ~300 m) ocean (Figure 4.6b). Therefore, if shallow<br />

water data close to the shoreline with water depth less than 200 m are simply<br />

removed (e.g., Nerem, 1995; Shum et al., 1998), some valuable data are probably<br />

omitted. Moreover, some contaminated data may still be kept in the data set in<br />

coastal regions while they should be removed.<br />

The rightmost column in Table 4.1 describes the general nature of the coastal terrain<br />

(cf. Thom, 1984; Kelleher et al., 2002). According to Brooks et al. (1997) and<br />

Brooks (2002), the contaminated distance should be correlated with the coastal<br />

topography, in which the correlation is not only related to the height of the coastal<br />

terrain, but also the slope of the terrain.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 88<br />

50% retracking points<br />

64<br />

7a<br />

48<br />

32<br />

16<br />

0<br />

-400 -350 -300 -250 -200 -150 -100 -50 0<br />

Ocean Depth (m)<br />

50% Retracking Points<br />

64<br />

48<br />

7b<br />

32<br />

16<br />

0<br />

-400 -350 -300 -250 -200 -150 -100 -50 0<br />

Ocean Depth (m)<br />

50% Retracking Points<br />

64<br />

7c<br />

48<br />

32<br />

16<br />

0<br />

-400 -350 -300 -250 -200 -150 -100 -50 0<br />

Ocean Depth (m)<br />

Figure 4.6 Distributions of 50% retracking points plotted versus DS759.2 ocean<br />

depths for contaminated waveforms in areas of 105°≤λ


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 89<br />

4.5.2 Using the Mean Waveform and its Standard Deviation<br />

The mean waveforms, and the standard deviation of these means, were computed<br />

over three distance bands in coastal regions, which range in distance of 0-350 km, 0-<br />

50 km, and 0-10 km off the Australian coast. Each has been divided into several<br />

annuli with intervals of 50 km, 10 km and 2 km, respectively. The time span of the<br />

data used to average is a cycle (35 days for ERS-2 and 10 days for POSEIDON). For<br />

POSEIDON, similar results were obtained from five cycles. Therefore, only the<br />

results from cycle 197, which are representative of all the data examined, are shown<br />

here.<br />

(a) 0-350 km off the Shoreline<br />

Figures 4.7 and 4.8 show averaged waveforms and standard deviations, respectively,<br />

in the distance-bands of 0-350 km with the interval of 50 km around the Australian<br />

coast for ERS-2 and POSEIDON. It is important to note that the coastal effects on<br />

the waveforms come from both Australia and Indonesia when the data are taken to<br />

the north of Australia. Therefore, the data thought to be affected by land near<br />

Indonesia have not been used to calculate or plot the mean waveform. Despite the<br />

different number of measurements in each distance band, the mean waveforms are<br />

generally in agreement with the theoretical model (Figure 1.1). The slope of the<br />

leading edge of the ERS-2 and POSEIDON mean waveforms is similar to that of the<br />

theoretical model. However, it can be seen that the trailing edges show some slight<br />

departure feature from the theoretical model (Figure 1.1) caused by fading noise and<br />

the wrap-around feature for both altimeters (cf. Figure 1.2 and Figure 4.1).<br />

The standard deviations of individual waveforms with respect to the mean waveform<br />

for ERS-2 show a peak in the waveform bins 38 and 39 and two ramps at the leading<br />

edge. Since this occurs in all distance bands, it seems to be a systematic anomaly.<br />

The reason for this should be investigated from more ERS-2 altimeter waveform data.<br />

Compared with ERS-2, the standard deviation of the POSEIDON data shows a<br />

smoother shape (Figure 4.8). Figures 4.7 and 4.8 show a large standard deviation in<br />

the distance band of 0-50 km compared with other standard deviations, implying that<br />

there is more contamination in this area. This will be studied next.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 90<br />

Mean<br />

Wavwform<br />

(counts)<br />

1200<br />

900<br />

600<br />

300<br />

0<br />

64<br />

55<br />

46<br />

37<br />

Bins<br />

28<br />

19<br />

10<br />

1<br />

0- 50 km<br />

50-100 km<br />

100-150 km<br />

150-200 km<br />

200-250 km<br />

250-300 km<br />

300-350 km<br />

STD (counts)<br />

1200<br />

900<br />

600<br />

300<br />

0<br />

64<br />

55<br />

46<br />

37<br />

Bins<br />

28<br />

19<br />

10<br />

1<br />

0- 50 km<br />

50-100 km<br />

100-150 km<br />

150-200 km<br />

200-250 km<br />

250-300 km<br />

300-350 km<br />

Figure 4.7 The mean waveform (top) and the standard deviation of the mean<br />

waveform (bottom) for 1 cycle of ERS-2 data in seven 50 km-wide<br />

bands around the Australia’s coastline.<br />

Mean<br />

Waveform<br />

(counts)<br />

200<br />

150<br />

100<br />

50<br />

0<br />

55<br />

46<br />

Bins<br />

37<br />

28<br />

19<br />

10<br />

1<br />

0- 50 km<br />

50-100 km<br />

100-150 km<br />

150-200 km<br />

200-250 km<br />

250-300 km<br />

300-350 km<br />

STD (counts)<br />

200<br />

150<br />

100<br />

50<br />

0<br />

55<br />

46<br />

Bins<br />

37<br />

28<br />

19<br />

10<br />

1<br />

0- 50 km<br />

50-100 km<br />

100-150 km<br />

150-200 km<br />

200-250 km<br />

250-300 km<br />

300-350 km<br />

Figure 4.8 The mean waveform (top) and the standard deviation of the mean<br />

waveform (bottom) for POSEIDON cycle 197 in seven 50 km-wide<br />

bands around the Australia’s coastline.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 91<br />

(b) 0-50 km off the Shoreline<br />

Mean and standard deviations in the distance bands of 0-50 km with an interval of 10<br />

km are shown in Figures 4.10 and 4.11 for ERS-2 and POSEIDON, respectively. For<br />

both altimeters, the leading edges of the mean waveforms in the bands of 10-20 km,<br />

20-30 km, 30-40 km and 40-50 km have a similar slope of leading edge, but this<br />

becomes less in band 0-10 km compared with the others.<br />

From Figure 4.9, the ERS-2 standard deviations for 30-40 km and 40-50 km agree<br />

well with these beyond 50 km (cf. Figures 4.8 and 4.9), while other standard<br />

deviations also show similar shapes, even if there are some larger values around the<br />

leading edge in other distance bands. The standard deviations for the 0-10 km band<br />

are the largest, with a peak value of 1288 counts. From Figure 4.10 for POSEIDON,<br />

the large standard deviation is evident in the 0-10 km band, especially close to the<br />

tracking gate near the peak of the plot.<br />

Mean Power of Waveform (counts)<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

Bins<br />

STD (counts)<br />

1600<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

Bins<br />

0−10 km<br />

N=9370<br />

10−20 km<br />

N=14202<br />

20−30 km<br />

N=14998<br />

30−40 km<br />

N=13995<br />

40−50 km<br />

N=12514<br />

Figure 4.9 Mean power of the waveform (left) and the standard deviation of the mean<br />

waveform (right) for ERS-2 in five 10 km-wide bands from the Australian coastline.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 92<br />

Mean Power of Waveform (counts)<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

Bins<br />

STD (counts)<br />

60<br />

40<br />

20<br />

0<br />

0 10 20 30 40 50 60<br />

60<br />

40<br />

20<br />

0<br />

0 10 20 30 40 50 60<br />

60<br />

40<br />

20<br />

0<br />

0 10 20 30 40 50 60<br />

60<br />

40<br />

20<br />

0<br />

0 10 20 30 40 50 60<br />

60<br />

40<br />

20<br />

0<br />

0 10 20 30 40 50 60<br />

Bins<br />

0−10 km<br />

N=4400<br />

10−20 km<br />

N=5160<br />

20−30 km<br />

N=3660<br />

30−40 km<br />

N=4620<br />

40−50 km<br />

N=4060<br />

Figure 4.10 Mean power of the waveform (left) and the standard deviation of the<br />

mean waveform (right) for POSEIDON cycle 197 in five 10 km-wide bands from the<br />

Australian coastline.<br />

(c) 0-10 km off the Shoreline<br />

Figures 4.11 and 4.12 show mean waveforms and their standard deviations 0-10 km<br />

off the shoreline with an interval of 2 km for ERS-2 and POSEIDON, respectively. It<br />

can be seen from Figure 4.11 that mean ERS-2 waveforms depart from the<br />

theoretical waveform in distance bands of 0-2 and 2-4 km. In the band of 8-10 km,<br />

the mean waveform agrees well with the theoretical or ocean waveform and with<br />

others at greater distances (cf. Figures 4.7, 4.9 and 4.11). It is evident that the<br />

standard deviation is large in gates for all distance bands. The variation of the<br />

standard deviation in the 0-2 km and 2-4 km distance bands is very large, up to about<br />

3000 counts in the peak.<br />

From Figure 4.12, it is obvious that nearly the same conclusion for the mean<br />

POSEIDON waveforms as for ERS-2 can be obtained. However, POSEIDON’s<br />

standard deviation does not show the obvious large values that the ERS-2 does in the<br />

4-6 km and 6-8 km bands. The largest standard deviation is also in the distance band


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 93<br />

0-2 km and the next largest is in the 2-4 km band. The standard deviation of the<br />

mean waveform increases with decreasing distance.<br />

From Figures 4.11 and 4.12, it can be seen that the larger standard deviation from the<br />

bands of 0-2 km and 2-4 km is dominant, which implies that this is the area of main<br />

contamination of coastal waveforms, but some lesser contamination also occurs in<br />

the other bands as well.<br />

The above results suggest that it is possible to estimate a typical contaminated area<br />

around coasts by simply analysing the variations of the mean waveform from the<br />

ideal and the standard deviation of the mean. Using this approach around Australia<br />

shows that the waveform data will mainly be contaminated when the ground track is<br />

within a distance of about ~4 km from the coast, while there is some smaller level of<br />

contamination to about 4-8 km for POSEIDON and 4-10 km for ERS-2.<br />

Mean Power of Waveform (counts)<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

Bins<br />

STD (counts)<br />

3000<br />

2000<br />

1000<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

1200<br />

800<br />

400<br />

0<br />

0 10 20 30 40 50 60<br />

Bins<br />

0−2 km<br />

N=1098<br />

2−4 km<br />

N=1495<br />

4−6 km<br />

N=1826<br />

6−8 km<br />

N=2323<br />

8−10 km<br />

N=2505<br />

Figure 4.11 Mean power of the waveform (left) and the standard deviation of the<br />

mean waveform (right) for ERS-2 in five 2 km-wide bands from the Australian<br />

coastline.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 94<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

0−2 km<br />

N=560<br />

Mean Power of Waveform (counts)<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

STD (counts)<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

2−4 km<br />

N=740<br />

4−6 km<br />

N=840<br />

6−8 km<br />

N=1000<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

Bins<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

Bins<br />

8−10 km<br />

N=1260<br />

Figure 4.12 Mean power of the waveform (left) and the standard deviation of the<br />

mean waveform (right) for POSEIDON cycle 197 in five 2 km-wide bands from the<br />

Australian coastline.<br />

4.5.3 Using Statistics of the 50% Retracking Point<br />

Statistics of the 50% threshold retracking points have been computed using both<br />

POSEIDON and ERS-2 waveform data in a coastal area of 10 km from the<br />

Australian shoreline. This distance is subdivided into 5 bands with an interval of<br />

2 km. As stated, if there is no effect from land and coastal sea states on the returned<br />

waveform, the threshold value of 50% retracking point should be very close to the<br />

pre-designed tracking gate when it is over oceans. Five cycles of POSEIDON<br />

waveform data and 1 cycle of ERS-2 waveform data are used to compute the 50%<br />

threshold retracking point in the above bands.<br />

Figure 4.13 shows the 50% retracking point distribution for these waveform data. For<br />

POSEIDON, the 50% retracking points comprise 15%, 50% and 59% in bins 30-32<br />

for bands of 0-2 km, 6-8 km and 8-10 km, respectively. For ERS-2, they are 20%,<br />

45% and 45% in bins 31-33 for bands of 0-2 km, 6-8 km and 8-10 km, respectively.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 95<br />

POSEIDON has a higher percentage than that of ERS-2 near the tracking gate with<br />

increasing distance. It was found from the calculation that 94% of POSEIDON<br />

waveforms represent their mid-point of leading edge very close to the tracking gate<br />

in bins 30-32 when beyond 10-20 km of the shoreline, while 84% of ERS-2 data<br />

show in bins 31-33 for the same distance.<br />

% of observations<br />

Poseidon<br />

60<br />

40<br />

0−2 km N=2814<br />

20<br />

0<br />

0<br />

60<br />

10 20 30 40 50 60<br />

40<br />

2−4 km N=3349<br />

20<br />

0<br />

0<br />

60<br />

10 20 30 40 50 60<br />

4−6 km N=4207<br />

40<br />

20<br />

0<br />

0<br />

60<br />

10 20 30 40 50 60<br />

40<br />

6−8 km N=4895<br />

20<br />

0<br />

0<br />

60<br />

10 20 30 40 50 60<br />

40<br />

8−10 km N=6187<br />

20<br />

0<br />

0 10 20 30 40 50 60<br />

Bins<br />

% of observations<br />

ERS−2<br />

60<br />

40<br />

0−2 km N=1075<br />

20<br />

0<br />

0<br />

60<br />

10 20 30 40 50 60<br />

40<br />

2−4 km N=1457<br />

20<br />

0<br />

0<br />

60<br />

10 20 30 40 50 60<br />

40<br />

4−6 km N=1830<br />

20<br />

0<br />

0<br />

60<br />

10 20 30 40 50 60<br />

6−8 km N=2316<br />

40<br />

20<br />

0<br />

0<br />

60<br />

10 20 30 40 50 60<br />

40<br />

8−10 km N=2504<br />

20<br />

0<br />

0 10 20 30 40 50 60<br />

Bins<br />

Figure 4.13 Fifty percent retracking point frequency distributions for 5 cycles of<br />

Poseidon waveform data (left) and 1 cycle of ERS-2 waveform data (right). They are<br />

shown as the percentage of observations in each bin for five 2 km-wide bands around<br />

the Australian coastline.<br />

Tables 4.2 and 4.3 show the statistics of the differences between 50% retracking<br />

point and the tracking gate for POSEIDON and ERS-2, respectively. The 50%<br />

retracking points are obviously offset from the tracking gate by 2.5±5.9 bins for<br />

POSEIDON and by 2.2±6.3 bins for ERS-2 in the band of 0-2 km. These offsets lead<br />

to a mean range difference of about 1m for both altimeters. POSEIDON’s mean


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 96<br />

range difference drops below 10 cm in the 6-8 km band, whereas ERS-2 does not<br />

decrease to below 10 cm until 10-20 km from the shoreline.<br />

Table 4.2 Differences between 50% retracking points and the tracking gate (bin 31.5)<br />

for 5 cycles (January 1998 to January 1999) of Poseidon waveforms (each bin has<br />

the range distance of ~0.469 m).<br />

Band<br />

(km)<br />

Mean<br />

(bins)<br />

Range difference<br />

(cm)<br />

RMS<br />

(bins)<br />

STD<br />

(bins)<br />

Number of<br />

20 Hz Points<br />

0 – 2 2.46 115 ±5.94 ±5.40 2814<br />

2 – 4 1.14 53 ±3.72 ±3.54 3349<br />

4 – 6 -0.38 -18 ±2.49 ±2.46 4207<br />

6 – 8 -0.15 7 ±1.67 ±1.66 4895<br />

8 - 10 0.10 5 ±0.96 ±0.96 6187<br />

10 - 20 0.14 6 ±0.72 ±0.74 23820<br />

Table 4.3 Differences between 50% retracking points and the tracker point (bin 32.5)<br />

for 1 cycle (March to April 1999) of ERS-2 waveforms (each bin with the range<br />

width of ~0.454m).<br />

Band<br />

(km)<br />

Mean<br />

(bins)<br />

Range difference<br />

(cm)<br />

RMS<br />

(bins)<br />

STD<br />

(bins)<br />

Number of<br />

20 Hz Points<br />

0 – 2 2.20 100 ±6.32 ±5.93 1075<br />

2 – 4 0.27 12 ±5.82 ±5.81 1457<br />

4 – 6 -0.47 -21 ±4.74 ±4.72 1830<br />

6 – 8 -0.29 -13 ±3.27 ±3.25 2316<br />

8 - 10 -0.23 -10 ±2.56 ±2.55 2504<br />

10 - 20 -0.14 -6 ±1.94 ±1.95 14224<br />

From Figure 4.13 and the values in Tables 4.2 and 4.3, this 50% retracking analysis<br />

indicates that a distance band of ~8 km and ~10 km are evidently the contaminated<br />

area for POSEIDON and ERS-2, respectively. The effect even extends to ~10-20 km<br />

for some ERS-2 data. Thus, there is a clear difference between POSEIDON and<br />

ERS-2 results from this analysis. It seems that contaminated distance for POSEIDON<br />

is shorter than that for ERS-2.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 97<br />

Finally, from reviewing the waveform products, it was found that ERS-2 provides<br />

20 Hz ground locations for 20 Hz ground waveforms, while POSEIDON provides<br />

1 Hz locations for 20 Hz waveforms. Thus, if a location for POSEIDON is given in a<br />

band, the waveforms 3 km either side of the band also contribute to the calculation in<br />

the band. This means that the results in Table 4.2 are actually computed for a larger<br />

band than 2 km, which is a plausible explanation for the different contaminated<br />

distances, as well as the different sized footprint and altimetric characteristics (also<br />

see Section 4.2.1).<br />

4. 6 Waveform Contamination Analysis<br />

The waveform contamination from near the coastline differs among the altimeter<br />

ground tracks, depending on whether the groundtrack is crossing the shoreline from<br />

water to land or from land to water (e.g., Brooks et al., 1997). In this Section,<br />

examples of ERS-2 altimeter ground tracks at the Australian coast (two water-to-land<br />

and two land-to-water) are analysed to understand the effects of the coastal<br />

topography on waveforms. Both SSH profiles (without retracking) and relevant<br />

waveforms are used for this analysis. Unlike the previous Section, distances in this<br />

Section are defined as the along-track distance to the coastline. The analysis in this<br />

Section focuses on the waveform itself. A comparison of the altimeter-derived SSH<br />

with the geoid heights will be given in Sections 1.2.1 and 5.8.2 to support the results<br />

obtained in this Section.<br />

4.6.1 Waveform Contamination for Water-to-Land Ground Tracks<br />

(a) Tasman Sea (Eastern Australia)<br />

As an ERS-2 altimeter groundtrack from cycle 43 (ascending pass 21792)<br />

approaches land northwestward (Figure 4.14) in the Tasman Sea, the 20 Hz SSH<br />

profile initially shows small but smooth undulation, and then rises greatly in the five<br />

late SSH samples near the shoreline. These relevant 20 successive waveforms are<br />

plotted via the bins (or gates) in Figure 4.15, which correspond to a ~7 km alongtrack<br />

distance. The locations of waveforms in Figure 4.15 correspond to the last 20<br />

SSH measurements numbered in Figure 4.14.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 98<br />

Land - 20<br />

SSH (m)<br />

30<br />

Tasman Sea<br />

- 15<br />

- 10<br />

- 5<br />

-30˚ 18'<br />

153˚ 06' 153˚ 09' 153˚ 12'<br />

- 1<br />

-30˚ 21'<br />

-30˚ 24'<br />

Figure 4.14 The 20 Hz SSH profile of pass 21792 (ascending track, cycle 43 of<br />

ERS-2), approaching the eastern Australian shoreline from the Tasman Sea. The<br />

number beside the SSH samples corresponds to the waveform in Figure 4.15.<br />

1000<br />

1<br />

1000<br />

2<br />

1000<br />

3<br />

1000<br />

4<br />

Counts<br />

500<br />

500<br />

500<br />

500<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1000<br />

500<br />

5<br />

1000<br />

500<br />

6<br />

1500<br />

1000<br />

500<br />

7<br />

1000<br />

500<br />

8<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1000<br />

500<br />

9<br />

1000<br />

500<br />

10<br />

1000<br />

500<br />

11<br />

1000<br />

500<br />

12<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1000<br />

500<br />

13<br />

2000<br />

1000<br />

14<br />

1500<br />

1000<br />

500<br />

15<br />

2000<br />

1000<br />

16<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

2000<br />

1000<br />

17<br />

2000<br />

1000<br />

18<br />

4000<br />

2000<br />

19<br />

8000<br />

6000<br />

4000<br />

2000<br />

20<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

Figure 4.15 20 Hz waveforms of pass 21792 (ascending track, cycle 43 of ERS-2),<br />

approaching Australian eastern shoreline in the Tasman Sea. The number in the top<br />

left-hand corner is related to that in Figure 4.14.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 99<br />

Waveform one in Figure 4.15 occurs when the groundtrack is ~5.4 km from land,<br />

and the land reflection does not yet appear. The land reflections appear clearly first in<br />

the gate (or bin) near gate 45 in the waveform 14 (Figure 4.15), then moving<br />

gradually closer to the tracking gate 32.5 in the successive waveforms 15-20. In the<br />

16th waveform (Figure 4.15), the ocean return is discernible near the tracking gate,<br />

while the land returns begin near the tracking gate. As the groundtrack approaches<br />

the land, the ocean return remains invisible on the leading edge and appears in the<br />

late gates, while the land returns are around the tracking gate in the range window.<br />

Waveforms 1 to 13 (Figure 4.15) basically show the ocean waveform characteristics<br />

with a clear single ramp. From waveforms 14 to 18, a single peak can be observed,<br />

but the shape is different from the open-ocean waveform. The last two (19 and 20)<br />

have a narrow width and significant peaks of the power, showing sharply rising and<br />

descending shape. Beginning with the waveform 16, the altimeter tracker appears to<br />

be responding more to the off-nadir higher and brighter land return than the at-nadir<br />

water return. As a result, the measured range is too short, and calculated at-nadir sea<br />

surface heights corresponding to waveforms 16-20 will be biased too high (Figure<br />

4.14).<br />

(b) Indian Ocean (Western Australia)<br />

Another example of an ERS-2 altimeter groundtrack (cycle 43 and ascending pass<br />

21450) approaching the land is shown in Figure 4.16, where the coast is dominated<br />

by sand dunes (Thom, 1984). In the area where the groundtrack intersects the<br />

shoreline, the land elevation in this area is between ~0-65 m above mean sea level<br />

from the Australian 9" DEM (version 2, Hutchinson, 2001, described in Section<br />

2.5.4). The SSH profile (related to a ~80 km along-track distance) shows no apparent<br />

change in Figure 4.16 when the groundtrack is offshore further than ~3.2 km (before<br />

the SSH sample 10). As the groundtrack gets closer to the shoreline after the tenth<br />

SSH samples, the SSH begins to rise.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 100<br />

SSH (m)<br />

-25<br />

-26˚ 24'<br />

20 -<br />

Land<br />

10 -<br />

-26˚ 27'<br />

1 -<br />

-26˚ 30'<br />

-26˚ 33'<br />

-26˚ 36'<br />

Indian<br />

Ocean<br />

-26˚ 39'<br />

113˚ 18' 113˚ 21'<br />

Figure 4.16 The 20 Hz SSH profile of pass 21450 (ascending track, cycle 43 of<br />

ERS-2), approaching the western Australian shoreline from the Indian Ocean. The<br />

number beside the SSH samples corresponds to the waveform in Figure 4.17.<br />

Counts<br />

10000<br />

5000<br />

1<br />

1000<br />

500<br />

2<br />

1000<br />

500<br />

3<br />

1000<br />

500<br />

4<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1500<br />

1000<br />

500<br />

5<br />

1000<br />

500<br />

6<br />

1000<br />

500<br />

7<br />

1000<br />

500<br />

8<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1500<br />

1000<br />

500<br />

9<br />

1500 10<br />

1000<br />

500<br />

1000<br />

500<br />

11<br />

1500<br />

1000<br />

500<br />

12<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1000<br />

500<br />

13<br />

1500<br />

1000<br />

500<br />

14<br />

1000 15<br />

500<br />

1000<br />

500<br />

16<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1000<br />

500<br />

17<br />

2000<br />

1000<br />

18<br />

4000<br />

2000<br />

19<br />

4000<br />

2000<br />

20<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

Figure 4.17 The 20 Hz waveforms of pass 21450 (ascending track, cycle 43 of ERS-<br />

2), approaching Australian western shoreline in the Indian Ocean. The number is<br />

related to that in Figure 4.16.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 101<br />

Figure 4.17 shows 1-second (20 Hz) waveforms related to the last 20 SSHs in<br />

Figure 4.16. The distance to the coastline from the location of the first waveform is<br />

~3.2 km. The first waveform shows a specular type with a high peak of ~10,000<br />

counts, implying that it is reflected from calm or standing water. As a result, the<br />

relative SSH (the first one in Figure 4.17) is measured too short, because the range is<br />

too long. The following waveforms (two to nine) show ocean-like shapes, but do not<br />

contain obvious land returns until the tenth waveform. The higher coastal land does<br />

not show significant effects on the waveforms. This is because the range window of<br />

64 gates is only related to a difference between surface elevations no larger than<br />

29 m for ERS-1/2. Therefore, the land topography more than ~30 m with respect to<br />

mean sea level around the groundtrack at this area does not appear as obvious land<br />

returns in the waveform range window.<br />

However, from waveforms 10 to 17 (Figure 4.18), the land returns appear to have<br />

affected the altimeter tracker response, and the waveforms have moved slightly to the<br />

left with respect to the tracking gate in the range window. In the last three waveforms<br />

(18-20) in Figure 4.17, the ocean returns have moved to the later gates and the land<br />

returns moved close to the tracking gate. This means that the shorter range<br />

measurement is obtained from the land returns reflected from higher surfaces. As a<br />

result, the calculated SSH will be too high.<br />

4.6.2 Waveform Contamination for Land-to-Water Ground Tracks<br />

(c) Indian Ocean (Northwestern Australia)<br />

Figure 4.18 shows an ERS-2 20 Hz SSH profile from an ascending pass 21636 (cycle<br />

43), which is going from land to water northwestward. This is a tidal plain coast area<br />

(Thom, 1984). The land-to-water crossover occurs at the 16th SSH in Figure 4.18. At<br />

this time, the terrain is higher than the ocean with respect to the radar pulse front, and<br />

the land reflections appear dominantly in the waveforms so that the SSH results are,<br />

probably, calculated from the land returns. As the satellite goes further from the<br />

shoreline, because of the off-nadir angle, the altimeter still measures the closest slant<br />

range to the higher land. As a result, in this case the effects of land on the calculated<br />

SSHs last nearly three seconds (~21 km along-track distance).


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 102<br />

The three-second successive waveforms, which correspond to the SSHs from 1-60 in<br />

Figure 4.18, have been plotted via the bins or gates in Figures 4.20, 4.21 and 4.22.<br />

Most waveforms seem to be specular due to still water in Figure 4.20. From the first<br />

to fifth waveforms, still water produces waveforms with higher peaks and narrow<br />

width. Lower power signals, which begin around the 40th gate from the sixth to 19th<br />

waveforms, are ocean returns as the groundtrack crosses the shoreline from the land<br />

to water.<br />

Beginning with waveform 20 in Figure 4.19, the initial radar returns are from the sea<br />

surface and the land returns move towards the late gates. Between waveforms 28 to<br />

33 in Figure 4.20, the off-nadir terrain and at-nadir sea surfaces are approximately<br />

equidistant from the altimeter, producing waveforms with double ramps. By<br />

waveform 34 when the groundtrack is 3.6 km to the shoreline (Figure 4.20), the land<br />

returns have disappeared.<br />

Although the sea surface returns are dominant in waveforms 34 through 40 in Figure<br />

4.20 and from 41 to 54 in Figure 4.21, these waveforms are misaligned with respect<br />

to the pre-designed tracking gate (32.5) such that the measured ranges are too long,<br />

and the calculated SSHs are too low (see Figure 4.18). In the rest of waveforms from<br />

55 to 60 in Figure 4.21, the half power point shifts slightly to the left with respect to<br />

the tracking gate. After three seconds (about ~21 km along the track or ~6 km from<br />

shore), waveforms become centred in the range window (i.e., at the tracking gate).<br />

(d) Southern Ocean (Southern Australia)<br />

The 20 Hz SSH profile from cycle 43 descending pass 21886 traverses<br />

southwestward across the shoreline from land to the Southern Ocean is shown in<br />

Figure 4.22. There is low land elevation along the coastline (Hutchinson et al., 2001).<br />

The SSH profile drops fast after the groundtrack enters the water, and then rises<br />

gradually. It changes to the normal ocean waveform and thus normal SSH after<br />

~2.5 seconds (~17 km from the shoreline) in this example.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 103<br />

SSH (m)<br />

-20<br />

Indian Ocean<br />

-22˚ 03'<br />

-22˚ 06'<br />

-22˚ 09'<br />

-22˚ 12'<br />

60 -<br />

-22˚ 15'<br />

50 -<br />

-22˚ 18'<br />

40 -<br />

30 -<br />

20 -<br />

10 -<br />

1 -<br />

Land<br />

114˚ 24' 114˚ 27' 114˚ 30'<br />

Figure 4.18 The 20 Hz SSH profile of pass 21636 (ascending track, cycle 43 of<br />

ERS-2), receding from the northwestern Australian shoreline to the Indian Ocean.<br />

x 10 4<br />

Counts<br />

15000<br />

10000<br />

5000<br />

1<br />

2<br />

1<br />

2<br />

2000<br />

1000<br />

3<br />

15000<br />

10000<br />

5000<br />

4<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

3000<br />

2000<br />

1000<br />

5<br />

1000<br />

500<br />

6<br />

1500 7<br />

1000<br />

500<br />

1500<br />

1000<br />

500<br />

8<br />

Counts<br />

2000<br />

1000<br />

0<br />

0 16 32 48 64<br />

9<br />

2000<br />

1000<br />

0<br />

0 16 32 48 64<br />

10<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

0 16 32 48 64<br />

11<br />

6000<br />

4000<br />

2000<br />

0<br />

0 16 32 48 64<br />

12<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

6000<br />

4000<br />

2000<br />

13<br />

3000 14<br />

2000<br />

1000<br />

2000<br />

1000<br />

15<br />

1000<br />

500<br />

16<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1000<br />

500<br />

17<br />

1000<br />

500<br />

18<br />

1500<br />

1000<br />

500<br />

19<br />

1000<br />

500<br />

20<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

Figure 4.19 20 Hz waveforms of pass 21636 at the 1st second (~0-7 km along-track<br />

distance) from the shoreline (ascending track, cycle 43 of ERS-2), receding from the<br />

northwestern Australian shoreline to the Indian Ocean.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 104<br />

Counts<br />

1000<br />

500<br />

21<br />

1500<br />

1000<br />

500<br />

22<br />

1500 23<br />

1000<br />

500<br />

1500 24<br />

1000<br />

500<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

2000<br />

1000<br />

25<br />

2000 26<br />

1000<br />

1500 27<br />

1000<br />

500<br />

1500 28<br />

1000<br />

500<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1500<br />

1000<br />

500<br />

29<br />

1500<br />

1000<br />

500<br />

30<br />

1000 31<br />

500<br />

1500<br />

1000<br />

500<br />

32<br />

Counts<br />

1500<br />

1000<br />

0<br />

0 16 32 48 64<br />

500<br />

33<br />

1000<br />

0<br />

0 16 32 48 64<br />

500<br />

34<br />

0<br />

0 16 32 48 64<br />

1000 35<br />

500<br />

1000<br />

0<br />

0 16 32 48 64<br />

500<br />

36<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

800<br />

600<br />

400<br />

200<br />

37<br />

0<br />

0 16 32 48 64<br />

Bins<br />

1000<br />

500<br />

38<br />

0<br />

0 16 32 48 64<br />

Bins<br />

800<br />

600<br />

400<br />

200<br />

39<br />

0<br />

0 16 32 48 64<br />

Bins<br />

1000<br />

500<br />

40<br />

0<br />

0 16 32 48 64<br />

Bins<br />

Figure 4.20 20 Hz waveforms of pass 21636 at the 2nd second (~7-14 km alongtrack<br />

distance) from the shoreline (ascending track, cycle 43 of ERS-2), receding<br />

from northwestern Australian shoreline to the Indian Ocean.<br />

Counts<br />

800 41<br />

600<br />

400<br />

200<br />

0<br />

0 16 32 48 64<br />

1000 42<br />

500<br />

0<br />

0 16 32 48 64<br />

1000 43<br />

500<br />

0<br />

0 16 32 48 64<br />

1500<br />

1000<br />

500<br />

44<br />

0<br />

0 16 32 48 64<br />

1000<br />

45<br />

1000 46<br />

1000<br />

47<br />

1000 48<br />

Counts<br />

500<br />

500<br />

500<br />

500<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1000<br />

500<br />

49<br />

1000 50<br />

500<br />

1000 51<br />

500<br />

1000<br />

500<br />

52<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1000<br />

500<br />

53<br />

1000<br />

500<br />

54<br />

1000<br />

500<br />

55<br />

1000<br />

500<br />

56<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1000<br />

500<br />

57<br />

1000<br />

500<br />

58<br />

1000<br />

500<br />

59<br />

1000<br />

500<br />

60<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

Figure 4.21 20 Hz waveforms of pass 21636 at the 3rd second (~14-21 km alongtrack<br />

distance) from the shoreline (ascending track, cycle 43 of ERS-2), receding<br />

from the northwestern Australian shoreline to the Indian Ocean.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 105<br />

When the groundtrack exits the land, the first three seconds of 60 waveforms related<br />

to the SSH profile have been plotted in Figures 4.24, 4.25 and 4.26. The first five<br />

waveforms in Figure 4.23 show mostly the response to the land reflections. After<br />

waveform six, which is at a distance of ~2.1 km from the shoreline, the leading edges<br />

of the waveforms shift to the earlier gates. This leftward shift continues to the<br />

waveform 25 in Figure 4.24. After that, the waveforms move back gradually towards<br />

the alignment of the ocean waveform through the rest of waveforms in Figure 4.24<br />

and all waveforms in Figure 4.25. Waveforms after 54 in Figure 4.25, when the<br />

ground track is more than ~17.9 km from shore, become stably aligned with respect<br />

to the tracking gate, appearing to be normal open-ocean returns.<br />

Waveforms in this example do not show obvious high power and the land returns<br />

except for the first six waveforms in Figure 4.23, because there is a beach area here<br />

without the steep slope and sharp changes of the topography (Thom, 1984). However,<br />

apparent waveform contamination still lasts about three seconds (~21 km along-track<br />

distance). This will be discussed in Section 4.6.4<br />

Land<br />

-15<br />

- 20<br />

- 10<br />

- 1<br />

-32˚ 18'<br />

SSH (m)<br />

- 60<br />

- 50<br />

- 40<br />

- 30<br />

-32˚ 21'<br />

-32˚ 24'<br />

Southern Ocean<br />

126˚ 45' 126˚ 48'<br />

-32˚ 27'<br />

-32˚ 30'<br />

-32˚ 33'<br />

Figure 4.22 The 20 Hz SSH profile of pass 21886 (descending track, cycle 43 of<br />

ERS-2), receding from the southern Australian shoreline to the Southern Ocean.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 106<br />

4000<br />

1<br />

4000 2<br />

4000<br />

3<br />

3000 4<br />

Counts<br />

2000<br />

2000<br />

2000<br />

2000<br />

1000<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

4000<br />

5<br />

3000 6<br />

2000 7<br />

2000 8<br />

Counts<br />

2000<br />

2000<br />

1000<br />

1000<br />

1000<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1500<br />

1000<br />

500<br />

9<br />

1000<br />

500<br />

10<br />

1000 11<br />

500<br />

1000<br />

500<br />

12<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

800<br />

600<br />

400<br />

200<br />

13<br />

0<br />

0 16 32 48 64<br />

1000<br />

500<br />

14<br />

0<br />

0 16 32 48 64<br />

600 15<br />

400<br />

200<br />

0<br />

0 16 32 48 64<br />

800<br />

600<br />

400<br />

200<br />

16<br />

0<br />

0 16 32 48 64<br />

Counts<br />

800<br />

600<br />

400<br />

200<br />

17<br />

600<br />

400<br />

200<br />

18<br />

600<br />

400<br />

200<br />

19<br />

600<br />

400<br />

200<br />

20<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

Figure 4.23 20 Hz waveforms of pass 21886 at the 1st second (0-7 km along-track<br />

distance) from the shoreline (descending track, cycle 43 of ERS-2), receding from<br />

the southern Australian shoreline to the Southern Ocean.<br />

Counts<br />

800<br />

600<br />

400<br />

200<br />

21<br />

800<br />

600<br />

400<br />

200<br />

22<br />

600<br />

400<br />

200<br />

23<br />

800<br />

600<br />

400<br />

200<br />

24<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

800<br />

600<br />

400<br />

200<br />

25<br />

800 26<br />

600<br />

400<br />

200<br />

800<br />

600<br />

400<br />

200<br />

27<br />

800<br />

600<br />

400<br />

200<br />

28<br />

Counts<br />

0<br />

0 16 32 48 64<br />

800<br />

600<br />

400<br />

200<br />

29<br />

0<br />

0 16 32 48 64<br />

1000<br />

0<br />

0 16 32 48 64<br />

500<br />

30<br />

0<br />

0 16 32 48 64<br />

1000<br />

0<br />

0 16 32 48 64<br />

500<br />

31<br />

0<br />

0 16 32 48 64<br />

1000<br />

0<br />

0 16 32 48 64<br />

500<br />

32<br />

0<br />

0 16 32 48 64<br />

Counts<br />

800<br />

600<br />

400<br />

200<br />

33<br />

0<br />

0 16 32 48 64<br />

1000<br />

500<br />

34<br />

0<br />

0 16 32 48 64<br />

1000 35<br />

500<br />

0<br />

0 16 32 48 64<br />

1000 36<br />

500<br />

0<br />

0 16 32 48 64<br />

1000<br />

37<br />

1000<br />

38<br />

1000<br />

39<br />

1000<br />

40<br />

Counts<br />

500<br />

500<br />

500<br />

500<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

Figure 4.24 20 Hz waveforms of pass 21886 at the 2nd second from the shoreline<br />

(ascending track, cycle 43 of ERS-2), receding Australian southern shoreline to the<br />

Southern Ocean.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 107<br />

1000<br />

41<br />

1000 42<br />

1000 43<br />

1000 44<br />

Counts<br />

500<br />

500<br />

500<br />

500<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1000<br />

500<br />

45<br />

1000 46<br />

500<br />

1000<br />

500<br />

47<br />

1000<br />

500<br />

48<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

1000<br />

49<br />

1000 50<br />

1000<br />

51<br />

1000<br />

52<br />

Counts<br />

500<br />

500<br />

500<br />

500<br />

Counts<br />

1000<br />

0<br />

0 16 32 48 64<br />

500<br />

53<br />

1500<br />

1000<br />

0<br />

0 16 32 48 64<br />

500<br />

54<br />

1000<br />

0<br />

0 16 32 48 64<br />

500<br />

55<br />

0<br />

0 16 32 48 64<br />

1000 56<br />

500<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

0<br />

0 16 32 48 64<br />

Counts<br />

1000<br />

500<br />

57<br />

1000<br />

500<br />

58<br />

1000<br />

500<br />

59<br />

1000<br />

500<br />

60<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

0<br />

0 16 32 48 64<br />

Bins<br />

Figure 4.25 20 Hz waveforms of pass 21886 at the 3rd second from the shoreline<br />

(ascending track, cycle 43 of ERS-2), receding Australian southern shoreline to the<br />

Southern Ocean.<br />

4.6.3 Contamination Caused by the Altimeter’s Operation<br />

As can be seen from Figures 4.19 to 4.26, after obvious returns from the land or<br />

inland coastal still water disappear from the range window, waveforms continue to<br />

take a second or more of time to align to the tracking gate. The corresponding alongtrack<br />

distance can be longer to ~21 km from the shoreline, which cannot be<br />

explained by the land effects within the altimeter footprint. The reasons for this are<br />

not clear, but it may be linked to the operation of the altimeter. The altimeter works<br />

often in the normal track mode over sea surfaces, while this is not the case over nonocean<br />

surfaces. After a water-to-land or land-to-water transition, the altimeter usually<br />

takes a second or more of time to reacquire the signal over the changed surface<br />

(Brooks et al., 1997; Strawbridge and Laxon, 1994, cited in Scott et al., 1994).<br />

In addition, the ocean waveform should show a linear trailing edge (see Section 5.4.3)<br />

based upon the theoretical waveform model. This means that the ocean waveform<br />

samples can be shifted rightwards in the range window without a change of the<br />

relevant position of the retracking gate. To demonstrate this, 20 waveforms at a


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 108<br />

second time span along descending groundtrack 21886 related to waveforms 21-40 in<br />

Figure 4.24 are chosen. This group of waveforms is characterised to have retracking<br />

gates from 23.3-27.3, which all depart from the pre-designed tracking gate (32.5).<br />

These waveforms are shifted rightwards with respect to the centre of the range<br />

window by 0, 5, 10, 15, 20, or 25 gates (corresponding to the range of 0 m, 2.27 m,<br />

4.54 m, 6.81 m, 9.08 m, or 11.36 m, respectively). Using the least squares fitting<br />

algorithm (Chapter 5), an ocean model (see Section 5.4.1) is fitted to each waveform<br />

to estimate the retracking gate. Differences between the retracking estimates before<br />

and after shift are computed. Descriptive statistics of the differences are then<br />

calculated and listed in Table 4.4.<br />

Table 4.4 Descriptive statistics of the differences of retracking gate estimates before<br />

and after shifting rightward waveform samples in the range window (20 Hz<br />

waveforms from pass 21886 at the 2nd second from the coastline, 1 gate≈0.4542 m)<br />

Shift<br />

(gates)<br />

Number of<br />

waveforms<br />

Min<br />

(gates)<br />

Max<br />

(gates)<br />

Mean<br />

(gates)<br />

STD<br />

(gates)<br />

5 20 -0.10 0.03 -0.02 0.03<br />

10 20 -0.13 0.07 -0.02 0.05<br />

15 19 -0.22 0.09 -0.01 0.08<br />

20 19 -0.20 0.24 0.01 0.09<br />

25 19 -0.26 0.23 -0.01 0.13<br />

As can be seen from Table 4.4, when shifting waveforms 5 and 10 gates rightward in<br />

the range window, the retracking gate estimates have a good agreement with those<br />

before shifting. When shifting 15, 20, or 25 gates, one waveform fails to obtain the<br />

retracking gate estimate, while others still agree well with the estimate before<br />

shifting. The STD increases with increasing the shifting gates. Because of the fading<br />

noise, the maximum range error caused by this shifting is ~6 cm. The fitting<br />

algorithm keeps working until shifting the position of the leading edge rightward in<br />

the range window to gate 57. This testing result indicates that the retracking gate can<br />

be successfully estimated by the fitting procedure.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 109<br />

Therefore, these test results support that the original tracking mode still continues for<br />

some time, when the returns reflected from the surface changed present in the range<br />

window. The results also demonstrate that the altimeter’s operating feature is one of<br />

waveform contamination sources in coasts, and this operating feature influences<br />

mainly waveforms along the ground tracks that leave land to water (cf. Figures 4.19<br />

and 4.23). It has been noted from previous discussions (Sections 4.6.1 and 4.6.2) that<br />

this problem does not appear to happen to the data from water to land ground tracks<br />

where the altimeter probably remains in the normal ocean mode till getting very<br />

close to coast. These data are, thus, less contaminated.<br />

This result also implies a tremendous influence on the use of untracked SSH data in<br />

coastal regions. Only the data from water to land ground tracks could be used in the<br />

area vicinity (e.g., ~5 km) of the coastline without waveform retracking<br />

4.6.4 Summary of Waveform Contamination in Coasts<br />

Considering the groundtrack approaching land, as it gets closer to the coasts, the land<br />

returns within the altimeter footprint begin to appear in the waveform range window.<br />

As long as the range to the nadir ocean surface is shorter than the range to the offnadir<br />

land surfaces, the ocean returns will appear earlier in the range window.<br />

However, as the land returns move into the AGC gate, the altimeter tracker tends to<br />

track the off-nadir terrain, leading the calculated at-nadir SSH too high. The affected<br />

distance is about 7 km along the groundtrack from the coastline for ERS-1/2.<br />

On the other hand, for ground tracks leaving land to water, the altimeter tracker<br />

continues to track the off-nadir land surfaces and responds more to the off-nadir<br />

shorter range at the time when the ground track is just receding from land to water.<br />

When the altimeter ground track goes further into the ocean, land returns move back<br />

to the late gates. Waveforms are, meanwhile, shifted left first and then right with<br />

respect to the centre of the range window (e.g., Figures 4.21 and 4.22). This shift can<br />

last three seconds or more of time, indicating a longer along-track contaminated<br />

distance of ~21 km (cf. Section 4.6.2).<br />

If considering only the shape variations of the waveform in the proximity of the<br />

coasts, the following three typical characteristics can be summarised:


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 110<br />

(1) The land returns appear in the range window together with the ocean surface<br />

scattering. Because of the different surface reflectivity and elevations, the<br />

power superposition of the ocean returns and land returns changes both the<br />

waveform shape and magnitude of the amplitude with increasing the<br />

importance of the land return in the range window. This, in turn, causes the<br />

gate location of the mean-surface retracking point to vary from 50% of the<br />

amplitude (i.e., the half-power point) as the land return signal begins to<br />

increasingly dominate the waveform shape. The ocean returns cannot be<br />

recovered when the land returns predominate in the waveform.<br />

(2) Waveforms shift towards the left or right as a function of along-track distance<br />

with respect to the tracking gate in the range window, but do not significantly<br />

change the magnitude of the waveform’s amplitude. Land returns are not<br />

obviously visible, and the contamination cannot be explained by the land<br />

effects, because the distance to the shoreline may be larger than the radius of<br />

the footprint. In this case, waveform contamination results from the altimeteroperating<br />

feature.<br />

(3) Inland standing water causes very high peaks in the waveform. This type of<br />

specular waveform appears with the large magnitude of the amplitude and the<br />

narrow width of power. These should not be retracked unless the information<br />

on the inland water is required.<br />

4. 7 An Experiment of the Land Effects on Waveforms Using the DEM Data<br />

A basic feature of the waveform is that it can be considered as the cumulative energy<br />

returned from each illuminated surface scatterer, sampled by a series of gates in the<br />

range window (Femenias et al., 1993; Martin et al., 1983). Although most coastal<br />

waveforms have an ocean-like shape with a single ramp, there are a percentage of<br />

waveforms (~20%) that show non-ocean-like shapes, particularly over areas closer to<br />

land (Section 6.2). These waveforms consist of ocean and land returns, hence leading<br />

to a mixed waveform shape. There are two possibilities for these waveforms. The<br />

first is that the ocean return could be recovered from these observed waveforms<br />

using waveform retracking. The alternative is that it might be impossible to recover<br />

the ocean return if it is obscured by the land return when the ground track is very


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 111<br />

close to the coastline, say, ~0-2.5 km, and range to the ocean surface is no longer<br />

recoverable.<br />

In most cases of waveform retracking, the relation between the derived surface and<br />

the actual surface are unknown, as independent measures of the ground truth are not<br />

generally available. A digital elevation model (DEM), however, is a representation of<br />

the terrain using the mean elevation information. Moreover, it includes the coastline<br />

information including the main outline of the land, bays, the outer edge of mangrove,<br />

swamps, closed-off narrow inlets, and watercourses at or near their mouths<br />

(Hutchinson et al., 2001). Therefore, a high-resolution DEM can provide the land<br />

elevation for dealing with mixed waveforms, though the disadvantage is that it is<br />

generated by the land-averaged elevation at each cell. Other alternatives are the<br />

optical satellite remote sensing, topographic maps or air photography, but these are<br />

time consuming. Thus, a high-resolution DEM is still a useful external data source.<br />

Because the well-developed ocean surface scattering model (e.g., Brown, 1977) is<br />

deduced under the assumptions of the surface statistical homogeneity (see Section<br />

3.2), these assumptions are not valid over non-ocean surfaces. Thus, to use the<br />

independent DEM data, some other methods must be developed. The idea is that if<br />

the land returns contained in the waveform can be determined separately using an<br />

independent data source, they might be removed from the shape-mixed waveform to<br />

leave the ocean return alone. Thus, the ocean surface can be recovered. To achieve it,<br />

however, detailed information about the topography along the coastline is necessary.<br />

The topography should include the land and inland water (e.g., lakes and rivers). Part<br />

of the effort, which has been performed and designed in this study, will be presented<br />

in this Section.<br />

4.7.1 Concept of Land Effects Using the DEM Model<br />

As stated in Chapter 2, the altimeter’s beam-limited footprint is the area within<br />

which the beam attenuation is 3 dB or less. It has a larger footprint size than that of<br />

the pulse-limited footprint (see Section 4.2.1). It is in this larger footprint size that<br />

the land returns have higher probability of appearing in the range window. Therefore,<br />

when discussing the land effects on the waveform, it is necessary to account for the<br />

beam-limited footprint rather than the pulse-limited footprint.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 112<br />

Consider two different reflected surfaces of land and water within the beam-limited<br />

footprint from which the power backscatters to the altimeter. The land elevation<br />

above local instantaneous sea level is H . The coordinate of this location can be<br />

defined by the radial distance from the altimeter to the point on the sea surface, the<br />

off-nadir angle ξ , the azimuth angle χ about the axis defined by the line between<br />

the altimeter and the nadir point. Because of the approximate spherical geometry of<br />

the Earth, it is convenient to replace the angle ξ with the colatitude angle φ<br />

subtended by rays from the centre of the Earth to the nadir point and the off-nadir<br />

'<br />

point H ( χ , ξ,<br />

t)<br />

= H ( χ,<br />

φ,<br />

t)<br />

(see Figure 4.26). Define Rφ ( χ,<br />

t)<br />

to be the radial<br />

distance from the altimeter to H ( χ , φ,<br />

t)<br />

. The reflected power is received from this<br />

specular reflector at the two-way travel time t if H ( χ , φ,<br />

t)<br />

falls within the footprint.<br />

To the lowest order for the small off-nadir angles ξ relevant to the altimeter, the<br />

'<br />

distance Rφ ( χ,<br />

t)<br />

is related to the distance R from the altimeter to the mean sea level<br />

at the angle φ by<br />

R ( χ,<br />

φ,<br />

t)<br />

≈ R − H ( χ,<br />

φ,<br />

t)<br />

(4.3)<br />

'<br />

φ<br />

A more precise relation is easily obtained from the geometry in Figure 4.26, but this<br />

'<br />

approximate solution is adequate for the present purpose. In addition, Rφ ( χ,<br />

t)<br />

can<br />

be also related to the satellite altitude alt(t) and the geoid height N( χ , φ , t)<br />

above a<br />

reference ellipsoid as<br />

R ( χ,<br />

φ,<br />

t)<br />

≈ alt(<br />

t)<br />

− H ( χ,<br />

φ,<br />

t)<br />

− N(<br />

χ,<br />

φ,<br />

t)<br />

(4.4)<br />

'<br />

φ<br />

Thus, on the one hand, the range to the land surface (<br />

'<br />

R<br />

φ<br />

) can be estimated by<br />

Equation (4.4), On the other hand, it can be obtained from the altimeter-measured<br />

range ( χ , φ , t)<br />

to the surface and the retracking correction dr 2 (Equation 5.7, see<br />

R s<br />

Section 5.4 for details) estimated by the nine-parameter function (Equation 3.41) as<br />

R ( χ,<br />

φ,<br />

t)<br />

= R<br />

'<br />

φ<br />

s<br />

( χ , φ , t)<br />

+ dr2<br />

= R ( χ , φ , t)<br />

+ G<br />

s<br />

2M<br />

( β − g )<br />

32<br />

0<br />

(4.5)


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 113<br />

alt<br />

R<br />

ξ<br />

R'φ<br />

Rφ<br />

H<br />

Re<br />

N<br />

mean sea<br />

level<br />

reference<br />

ellipsoid<br />

φ<br />

Figure 4.26 A schematic geometrical description of the radar return from a point on<br />

the land surface at off-nadir angle ξ and associated colatitude angle φ where the<br />

land elevation above the mean sea level is H . The distance from the altimeter to the<br />

'<br />

land surface is R<br />

φ<br />

.<br />

where β<br />

32<br />

is the retracking gate estimate in units of waveform gates (related to the<br />

second ramp),<br />

G 2 M<br />

= t p<br />

⋅ c / 2 is the conversion from numbered gates to metres, t p<br />

is<br />

the pulse width, and g<br />

0<br />

is the location in gates of the altimeter tracking gate. From<br />

Equations (4.4) and (4.5), β<br />

32<br />

can be related to the land elevation H ( χ , φ,<br />

t)<br />

by<br />

β = a + bH ( χ , φ , )<br />

(4.6)<br />

32<br />

t<br />

where<br />

a<br />

alt( t)<br />

− N(<br />

χ , φ , t)<br />

− R ( χ , φ , t)<br />

s<br />

= g<br />

0<br />

+<br />

(4.7)<br />

G2M


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 114<br />

and<br />

= 1/<br />

(4.8)<br />

b G2M<br />

Equation (4.6) describes the relationship between the retracking gate estimate and the<br />

land elevation. At a certain time t , β<br />

32<br />

is only linearly related to the land elevation<br />

H ( χ , φ,<br />

t)<br />

at the reflector’s location. Therefore, if the location of β 32<br />

at the range<br />

window can be determined by the independent DEM data source, it can be used as a<br />

constrained condition or initial value of the nine-parameter function. This is because<br />

an appropriate initial value or constraint will make the fitting algorithm converge<br />

rapidly and (probably) accurately. However, it is also clear from Equation (4.6) that<br />

several surface reflectors at the time t within the satellite footprint can have the same<br />

slant range to the satellite. Thus, β<br />

32<br />

is not a unique function of the land elevation.<br />

This can be also presented from the simple test shown below<br />

4.7.2 A Preliminary Test of Land Effects Using the DEM Model<br />

When testing this method, the following steps have been taken for each observing<br />

location ( λ<br />

0<br />

,ϕ<br />

0<br />

) along satellite ground tracks near land.<br />

(1) The algorithm estimates the shortest spherical distance s to the coastline and<br />

extracts the location ( λ , ϕ ) at the point using a shoreline model, at which the<br />

c<br />

c<br />

ground track intercepts the coastline.<br />

(2) A search radius centred at the location ( λ<br />

c<br />

, ϕ<br />

c<br />

) is then determined according to<br />

the value of s and the size of the satellite footprint, which can range from<br />

2-8 km. The land elevations and geoid heights within the search area are<br />

extracted at the point of each elevation cell using the Australian 9" DEM<br />

(Hutchinson, 2001) and AUSGeoid98 (Featherstone et al., 2001) geoid grid<br />

(2´×2´), respectively.<br />

(3) Retracking gate estimates β<br />

32<br />

( χ i<br />

, φ i<br />

, t)<br />

are obtained using Equation (4.6),<br />

where i = 1, 2, ···, n is the number of the 9" cells within the search radius.<br />

Because of the width of the range window, the maximum difference of the


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 115<br />

elevations must be less than 28 m. Thus, elevations higher than 28 m were not<br />

used in the computation.<br />

Figure 4.27 shows two plots of β<br />

32<br />

( χ i<br />

, φ i<br />

, t)<br />

via the land elevation H ( χ<br />

i<br />

, φi<br />

, t)<br />

. The<br />

discontinuous retracking gate estimates in Figure 4.27 are caused by the non-uniform<br />

elevation distribution within the search radius. As can be seen, all land elevations<br />

within the search radius contribute to the return power, thus causing the problem of<br />

non-uniqueness. This is a typical problem for non-ocean surfaces, because several<br />

targets with the same slant range to the altimeter exist within the footprint.<br />

Retracking gate eatimates<br />

(bins)<br />

60<br />

55<br />

50<br />

45<br />

40<br />

35<br />

30<br />

0 1 2 3 4 5 6 7 8<br />

Land elevations (m)<br />

Retracking gate estimates<br />

(bins)<br />

60<br />

55<br />

50<br />

45<br />

40<br />

35<br />

30<br />

0 1 2 3 4 5 6 7 8<br />

Land elevations (m)<br />

Figure 4.27 Retracking gate estimates corresponding to the land elevation. The<br />

search radius is 2.7 km (top) and 3.5 km (bottom).<br />

Thus, although Equation (4.6) is simple, it is difficult to effectively implement.<br />

Theoretically, the altimeter starts to track off-nadir higher and brighter land when the


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 116<br />

distance between the satellite and the land surface is shorter than that to the sea<br />

surface. In practice however, of these land reflectors illuminated by the satellite, it is<br />

hard to determine, based on only the DEM, which one importantly contributes to the<br />

relevant mid-point on the second ramp of the waveform. Also, information about the<br />

surface reflectivity of land and water is important, because the radar returns are<br />

directly related to the nature of the scattering surface illuminated by the transmitted<br />

electromagnetic single. However, the DEM cannot provide the water information on<br />

land.<br />

To overcome these problems, a more detailed function based directly on the radar<br />

equation is necessary. Wingham’s method (1995) may be a way to solve the<br />

problem. However, the DEM contains only land elevation and no water information<br />

over land. This is another difficulty when trying to create the function for the land<br />

return, because the water reflector, even a layer of very shallow water (e.g., after<br />

rain), can make the return power much higher (Berry, 2002). In addition, the DEM<br />

represents the average elevation of a cell (cf. Hutchinson, 2001), while returns reflect<br />

from the facets with different elevations. Another condition that needs to be thought<br />

about is the land slope. Brooks (2002) has found that the land slope affects the<br />

returns reflected from the land near the coasts as well. Some higher land may scatter<br />

forward the radar pulse so that it does not show land return on the waveform. Since<br />

20 Hz waveform data require a spatial resolution of ~350 m, the high-resolution<br />

DEM and detailed topography information are necessary to estimate land returns.<br />

Because of the time limitation of this study, this can only be a topic of future<br />

research.<br />

4. 8 Summary<br />

Five cycles of POSEIDON and one cycle of ERS-2 20 Hz waveform data have been<br />

used to quantify a broad, contaminated boundary around the Australian coast. This<br />

contamination is assumed to have been caused by backscatter from the land and the<br />

more variable coastal sea surface states, coupled with the footprint size of the radars<br />

and the incorrect positioning of the groundtrack.<br />

Since bins 31-33 for the ERS-2 altimeter and bins 30-32 for the POSEIDON<br />

altimeter are the expected bins for the tracking gate over oceans, three major


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 117<br />

conclusions can be drawn from this study. First, the ERS-2 altimetry shows<br />

differences in contamination as a function of the sub-area along the Australian<br />

shoreline. The maximum contaminated distances vary from ~8 km to ~22 km,<br />

depending upon locations and features of the shoreline. Second, both altimeters show<br />

differences in contamination as a function of the distance from the shoreline. Finally,<br />

the contaminated distance for POSEIDON is less than that for ERS-2. A mean<br />

contaminated distance of ~8 km for POSEIDON and ~10 km for ERS-2 can be<br />

observed using the 50% of threshold retracking points in the whole Australian<br />

coastal region.<br />

The mean waveforms and standard deviations of the mean in different distance bands<br />

provide an indication of the contamination influencing the shape and variability of<br />

the returned waveforms in coastal regions. They can provide a reasonable<br />

explanation for the observed variations of the waveform in the proximity of coasts. A<br />

typical contaminated boundary can be ascertained from Australian shoreline to<br />

~8 km and ~10 km for POSEIDON and ERS-2, respectively. Beyond 8 or 10 km (to<br />

350 km), the mean waveform shapes for both altimeters match the mean waveform<br />

shape observed over ocean surfaces, though they do not agree exactly with the<br />

theoretical waveform shape (Section 4.5).<br />

The different reflector characteristics lead to different types of waveforms. In<br />

addition, the contamination distance depends on how the satellite ground track<br />

crosses the shoreline (i.e., approaching or leaving the land). In general, the<br />

waveforms along the altimeter groundtrack leaving land to water suffer longer<br />

contaminated distances than those along the track approaching land from water. In<br />

both cases, the ocean surface sometimes can no longer be recovered from<br />

contaminated waveforms obscured significantly by the land return, when the ground<br />

track is much closer to the coastline, say, ~0-2.5 km, because the land returns<br />

become predominant in the range window.<br />

A preliminary test has been performed to calculate the retracking gate estimate<br />

related to the land returns in the range window using an external independent data<br />

source of the DEM and AUSGeoid98 models in coastal regions. Analysing results<br />

show that non-uniqueness of the estimate over land near coasts is still a main<br />

problem that may need to be solved in the future.


Chapter 4. WAVEFORM CONTAMINATION CLOSE TO AUSTRALIAN COASTS 118<br />

Contamination leads to errors in the on-board calculation of the range measurements.<br />

It is recommended from the results in this Chapter that such contaminated<br />

measurements must be detected prior to their being included in geodetic and<br />

oceanographic solutions in coasts. Until then, altimeter SSH or range data should be<br />

interpreted with some caution for distances less than, say, ~22 km from a coastline,<br />

and discarded altogether for distances less than 4 km.<br />

This Chapter utilises different methods (e.g., statistic and threshold analysis, and<br />

using additional DEM data source) to show that the land topography does influence<br />

the waveform shape and thus the data. This can also be achieved using some other<br />

methods, such as to plot the distance from the coastline against near shore land<br />

height (within 0.5 km of coast) to see if there is a correlation between them.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 119<br />

5. COASTAL WAVEFORM RETRACKING SYSTEM: DESIGN AND<br />

IMPLEMENTATION<br />

5. 1 Introduction<br />

The techniques used to retrack ocean and non-ocean waveforms have been discussed<br />

in Chapter 3. They can also be applied to the coastal waveform (e.g., Anzenhofer et<br />

al., 2000; Deng et al., 2001). During waveform data processing and analysis, it has<br />

been found in this study that fitting the waveform to a well-developed analytical<br />

function (i.e., the Brown (1977) ocean model) is capable of retracking most of the<br />

individual altimeter waveforms (~99% of data, see Table 6.1 in Section 6.2.2) over<br />

open oceans. Similarly, using a 50% threshold level can give reasonable estimates of<br />

the absolute range to the at-nadir sea surface. Using only a single retracker in coasts,<br />

however, limits the precision of the recovered SSH due mainly to the disagreement<br />

of the contaminated waveform with the theoretical/ideal waveform.<br />

Therefore, it has become clear that no single retracking algorithm can always deal<br />

with the diverse waveforms in coastal regions. A coastal waveform retracking system<br />

which consists of different retrackers (e.g., the parametric fitting and the threshold<br />

algorithms), thus, is necessary when attempting to extract the precise SSH from<br />

retracked range measurements.<br />

This Chapter presents the implementation of both fitting and threshold retracking<br />

algorithms, and develops a coastal retracking system that is based upon a detailed<br />

pre-analysis of the coastal waveforms. Issues involved in the algorithm development<br />

are fitting functions, the iterative least squares procedure, linearisation of the<br />

parameters, determination of the initial estimates of parameters, and the weight<br />

scheme. A new method will be introduced that uses iterative weights to detect the<br />

outliers in the waveforms ensuring the effective convergence of the fitting procedure.<br />

For the threshold retracking algorithm, a method will be described that selects an<br />

appropriate threshold level for coastal waveforms. Finally, a coastal retracking<br />

system will be developed, which makes it possible to improve the SSH precision by<br />

reprocessing the altimeter waveforms, correcting the range measurements, and<br />

extracting precise SSH data from these corrected ranges.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 120<br />

5. 2 Coastal Waveform Retracking System Design<br />

A comprehensive analysis of waveform shapes at the Australian coast (Chapter 4)<br />

led to the need to develop a coastal waveform retracking system. The primary idea of<br />

the retracking system is that retrackers must cover most of the coastal waveforms,<br />

and they can complement each other. Since the SSH data can only be recovered from<br />

ocean returns, it is impossible for the system to reprocess the waveforms dominated<br />

by the land return near land to extract SSHs. This means that it is not essential to<br />

have as many retrackers as used over land, where much more diverse waveforms are<br />

found (cf. Berry, 2000). However, it is significant that the system can correctly<br />

categorise the waveforms, so that they can be retracked using an adequate retracker.<br />

This concern generated an additional waveform classification.<br />

Ideally, a physically realistic model should be used to retrack coastal waveforms,<br />

because the waveform measured by a radar altimeter links not only to the geometric<br />

surface of the ocean, but also to oceanic geophysical features. An appropriate model<br />

will provide a correct understanding of the relationship between the geophysical<br />

parameters of interest and the radar return from coastal regions.<br />

With this consideration, the ocean model without non-linear wave parameters<br />

(presented in Section 5.5.1) is first used to retrack ocean-like waveforms. To deal<br />

with narrow, high-power waveforms reflected from calm coastal or inland water<br />

surfaces, the threshold retracking technique with 50% threshold level is used as well.<br />

It was found from an analysis of retracked results that both retrackers could not deal<br />

with some waveforms in close proximity to the land. Fitting an ocean model to nonocean-like<br />

waveforms causes failure of the fitting algorithm. This led to the use of<br />

other retrackers of the five- and nine-β-parameter fitting functions because of their<br />

ability to retrack non-ocean waveforms (presented in Section 5.5.3). The 50%<br />

threshold level cannot always give improved results, particularly in the area closer to<br />

the coastline. To validate the threshold level of the retracking algorithm, a statistical<br />

analysis has been performed to suggest the quality of the retracked SSH data (Section<br />

5.7.2). The evaluation turned out a detailed selection of the threshold levels.<br />

When retracking individual waveforms in coasts using all possible retrackers, results<br />

show improved SSHs, but biases exist among different algorithms. Thus, an


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 121<br />

Figure 5.1 Block diagram of the coastal retracking system


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 122<br />

algorithm assessment has been implemented to analyse and estimate these biases. All<br />

these investigations motivate the development of the coastal retracking system.<br />

Finally, a coastal retracking system has been developed to retrack altimeter<br />

waveforms in Australian coasts, which involves several external input data,<br />

waveform data editing, waveform classification, fitting algorithms and threshold<br />

techniques. The flow diagram in Figure 5.1 illustrates how altimeter waveforms are<br />

reprocessed to give the corrected SSH data. The procedure follows the main steps as:<br />

(1) Data editing - edit out the waveforms that do not show the leading edge in the<br />

range window, retain waveforms that show a leading edge regardless of<br />

complexity of the whole waveform;<br />

(2) Waveform classification - sort waveforms based on different shapes, and<br />

indicate which retracker will be used for each;<br />

(3) Fitting algorithms - make initial estimates of the parameters, determine<br />

observational weights, and run iteration procedure;<br />

(4) Threshold retracking - retrack waveforms with the narrow width and high peak,<br />

as well as the waveforms that failed with the fitting algorithms; and<br />

(5) Corrections - include the biases between each method, range correction, and<br />

the geophysical and environmental corrections.<br />

The description of each step will be presented in this Chapter and the ancillary input<br />

data sets needed by the system are described in Section 2.5. An important factor in<br />

the system is to accurately classify each return waveform so that it can be performed<br />

by an appropriate retracker. Retracking results and analysis at the Australian coast<br />

will be shown in Chapter 6.<br />

5. 3 Extraction of the Geodetic Parameter: the Sea Surface Height<br />

After retracking, the retracked range is then<br />

Retracked Range = Observed Range + Retracking Range Correction<br />

Since the retracking system consists of fitting algorithms and threshold retracker, the<br />

biases between them (discussed later in Section 5. 8) must be applied to the range<br />

measurements. Biases can be calculated by retracking the same cycle of ocean


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 123<br />

waveform data. The corrected range in this study is based upon the retracked range<br />

from the fitting algorithm so that the bias is applied to the range which is retracked<br />

using the threshold retracker by<br />

Corrected Range = Retracked Range + Bias<br />

In addition, calculating an accurate SSH from the altimetric measurement involves<br />

accounting for not only the waveform range corrections, which are discussed in this<br />

study, but also correcting the range for atmospheric propagation delay and applying<br />

appropriate ocean-tide correction to remove the time-varying tidal effect. After these<br />

corrections are applied to the range, the SSH can then be calculated from the satellite<br />

altitude and corrected range as<br />

Corrected SSH = Orbit - Corrected Altimeter Range - Corrections.<br />

Up to now, the corrected SSH data are used at 20 Hz. They can produce 1 Hz or 2 Hz<br />

data products for geodetic applications.<br />

5. 4 Analysis of Waveform Shapes<br />

5.4.1 Determination of the Peaks in the Waveform<br />

The peaks in the waveform (cf. Figure 5.2) can be estimated from an analysis of the<br />

second derivative of the waveform. The first derivative, W ′(t)<br />

, is approximated by<br />

the first forward differences with respect to the gate number<br />

W ( tk<br />

+ 1)<br />

−W<br />

( tk<br />

)<br />

W '( t)<br />

=<br />

∆t<br />

(5.1)<br />

where W t ) is the waveform sample at the time t , ∆ t is the time interval between<br />

( k<br />

waveforms t<br />

k + 1<br />

and t<br />

k<br />

, and k = 1L64<br />

(depending on the altimeter) is the gate<br />

number of the waveform.<br />

Using results of contiguous first forward differences, the second derivative of the<br />

waveform, W ′′(t<br />

) , can be approximated by second forward differences as


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 124<br />

W ′(<br />

t W ′<br />

k + 1)<br />

− ( tk<br />

)<br />

W ′ ( t)<br />

=<br />

∆t<br />

(5.2)<br />

If there are n p peaks (n p = 6 in this study) in the waveform, the gate locations of all<br />

possible peaks L max (i) (i = 1, 2, ···, n), then, can be determined, where the second<br />

derivatives cross zero from negative to positive, and vice versa for all minimum<br />

locations L min (i).<br />

Finally, the peak amplitudes, W amp (i), can be found from the waveform<br />

corresponding to the gates of L max (i). All peaks are checked to determine whether<br />

they are larger than an input criterion, which is obtained prior to the retracking<br />

procedure by analysing ocean waveform amplitude features. Peaks less than this<br />

criterion are not significant. In addition, if the interval between two peaks is too<br />

small, only the larger one is selected as the peak. After editing, the remaining peak(s)<br />

will be used to classify the waveforms. The maximum number of waveform peaks<br />

(ERS-1/2) in Australian coastal regions is found to be three. Results of a detailed<br />

analysis will be given later in Chapter 6.<br />

5.4.2 Waveform Classification<br />

The shape of the radar altimeter waveform depends strongly on the reflecting surface<br />

properties. Small amplitude roughness and slopes of the surface will cause a<br />

broadening of the leading edge of the return. Multiple reflecting surfaces within the<br />

altimeter footprint will cause the waveform to be the superposition of the different<br />

returns from each surface elevation, thus forming the multi-peaks. Over these<br />

multiple reflecting surfaces near the land, the waveform shape cannot follow the<br />

Brown (1977) model.<br />

Waveform classification is based upon a wide ranging analysis of the waveform<br />

shapes at the Australian coast. After observing a number of ERS-2 coastal<br />

waveforms, they can be sorted into nine categories (Figure 5.2) depending on the<br />

waveform shape features, such as amplitude, width, and measure of peaks. Of these<br />

waveform features, the peaks can be ascertained using the method described in<br />

Section 5.4.1, and waveform amplitude and width can be determined using the


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 125<br />

OCOG method (Wingham et al., 1986). A detailed waveform analysis by<br />

geographical area will be given in Chapter 6.<br />

The shape characteristics of the waveform revealed in Figure 5.2 are generally<br />

associated with a particular type of surface. Of these waveforms in coastal regions,<br />

most of them (~80%) present a shape of ocean-like single ramp without the<br />

significant peak in the range window (Figure 5.2a), though the waveform profiles<br />

may shift left or right with respect to the predesigned tracking gate. This type of<br />

return has been found in this study over ocean surfaces at any distance up to 300 km<br />

from the coastline. Waveforms shown in Figure 5.2b to Figure 5.2g are usually found<br />

when the altimeter simultaneously illuminates both water and land surfaces, where<br />

the elevation difference between both surfaces within the pulse-limited footprint is<br />

less than ~28 m (due to the scale of the range window). Because of two or more<br />

elevation surfaces in the footprint, such waveforms occur ~0-10 km from the<br />

coastline (Chapter 3). The narrow-peaked waveform (Figure 5.2h) occurs over the<br />

standing water near coastline, such as bays, gulfs, estuaries, and harbours, as well as<br />

inland lakes. A distinct feature of this type of waveform is its extremely high<br />

amplitude (typically ~8400 counts in average from Section 6.2.2).<br />

Therefore, waveform categories in coastal regions indicate the specific reasons why a<br />

retracking system is appropriate in coasts. The primary factor is that many waveform<br />

shapes observed in coastal regions are not encountered in data sets over open oceans.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 126<br />

(a) No Significant Peak<br />

(b) Single Pre-peak<br />

Power (counts)<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

1<br />

6<br />

11<br />

16<br />

21<br />

26<br />

31<br />

36<br />

41<br />

46<br />

51<br />

56<br />

61<br />

B i n s<br />

Power (counts)<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

1<br />

6<br />

11<br />

16<br />

21<br />

26<br />

31<br />

36<br />

41<br />

46<br />

51<br />

56<br />

61<br />

B i n s<br />

Power (counts)<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

(c) Single Post-peak<br />

Power (counts)<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

(d) Single Middle-peak<br />

1<br />

6<br />

11<br />

16<br />

21<br />

26<br />

31<br />

36<br />

41<br />

46<br />

51<br />

56<br />

61<br />

1<br />

6<br />

11<br />

16<br />

21<br />

26<br />

31<br />

36<br />

41<br />

46<br />

51<br />

56<br />

61<br />

B i n s<br />

B i n s<br />

Power (counts)<br />

4000<br />

3000<br />

2000<br />

1000<br />

0<br />

(e) Double Pre-peaks<br />

Power (counts)<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

(f) Double Post-peaks<br />

1<br />

6<br />

11<br />

16<br />

21<br />

26<br />

31<br />

36<br />

41<br />

46<br />

51<br />

56<br />

61<br />

1<br />

6<br />

11<br />

16<br />

21<br />

26<br />

31<br />

36<br />

41<br />

46<br />

51<br />

56<br />

61<br />

B i n s<br />

B i n s<br />

(g) Multi Peaks (h) Sharp Peak<br />

Power (counts)<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

Power (counts)<br />

20000<br />

15000<br />

10000<br />

5000<br />

0<br />

1<br />

6<br />

11<br />

16<br />

21<br />

26<br />

31<br />

36<br />

41<br />

46<br />

51<br />

56<br />

61<br />

1<br />

6<br />

11<br />

16<br />

21<br />

26<br />

31<br />

36<br />

41<br />

46<br />

51<br />

56<br />

61<br />

B i n s<br />

B i n s<br />

Figure 5.2 Typical ERS-2 waveform categories in Australian coasts


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 127<br />

5. 5 Fitting Functions<br />

To use fitting algorithms, an analytic expression of coastal waveforms is essential,<br />

which describes the interaction of the radar pulse with the scattering surface. As<br />

stated, estimating non-linear wave parameters is not the aim of this study. In fact, it<br />

is impossible to accurately estimate non-linear wave parameters using contaminated<br />

waveforms in coasts. Therefore, the fitting functions used in this study aim at the<br />

determination of the range correction through the revised estimate of the tracking<br />

gate. In the proposed system, an ocean model and the β-parameter models will be<br />

used as fitting functions, described next.<br />

5.5.1 The Ocean Model<br />

In Chapter 3, a modified expression (Equation 3.31) for the waveform over non-<br />

Gaussian ocean surface was presented. The sea surface skewness λ cs<br />

in Equation<br />

(3.33) usually takes a small magnitude. Rodriguez and Martin (1994) estimate the<br />

skewness using TOPEX Ku-band waveforms (cycles 3-27) as between 0 - 0.11 for<br />

both ascending and descending passes, which is obtained by averaging estimates<br />

over all cycles. Therefore, it can be safely neglected in the waveform fitting function<br />

without the loss of the retracking accuracy. For λ = 0 (i.e., neglecting skewness),<br />

Equation (3.31) is the linear surface case and the result can be reduced to a form<br />

similar to the Brown (1977) model as<br />

cs<br />

1 ⎡ ⎛ τ ⎞ ⎤ ⎡ ⎛ d ⎞⎤<br />

P( t)<br />

= PN + A0<br />

Aξ ⎢erf<br />

⎜ ⎟ + 1⎥<br />

exp⎢−<br />

d⎜τ<br />

+ ⎟<br />

2<br />

⎥<br />

(5.3)<br />

⎣ ⎝ 2 ⎠ ⎦ ⎣ ⎝ 2 ⎠⎦<br />

in which five fitting parameters are:<br />

(1) The thermal noise level P<br />

N<br />

,<br />

(2) The amplitude scaling term A 0<br />

,<br />

(3) The tracking-gate-related time t 0<br />

(i.e., the time corresponding to the mean sea<br />

surface),<br />

(4) The risetime σ , and<br />

(5) The off-nadir angle ξ (represented by Aξ<br />

Equation 3.7).


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 128<br />

Equation (5.3) will be the model used in this study for the purpose of waveform<br />

retracking. For simplicity of presentation, it will be called the revised ocean model<br />

hereafter.<br />

After waveform retracking, the retracking range correction, dr , is given as<br />

dr = G ⋅ t / t − )<br />

(5.4)<br />

2M<br />

(<br />

0 p<br />

g<br />

0<br />

where<br />

G 2 M<br />

= t p<br />

⋅ c / 2 is the conversion from numbered gates to metres, t p<br />

is the<br />

pulse width, and<br />

g<br />

o<br />

is the location in gates of the altimeter tracking gate.<br />

The revised ocean model (Equation 5.3) describes the waveform reflected from the<br />

linear Gaussian ocean surface. Compared to the Brown model (Equation 3.12), the<br />

differences between them are that the ocean model includes the effects of the Earth’s<br />

curvature and the modified Bessel function in Equation (3.2) has been replaced by an<br />

approximated expression.<br />

It is recalled briefly here about the series expansion of the Bessel function (Equation<br />

3.17). In practice, it has been expanded to different orders (Parsons, 1979; cf. Davis,<br />

1993; Hayne, 1980; Rodriguez, 1988). When retracking waveforms to determine the<br />

range correction using the Brown (1977) model, it is generally expanded to zero<br />

order (Parsons, 1979; e.g., Davis, 1993). This means that I ( β t ) 1 , and the<br />

0<br />

=<br />

Bessel function does not change its value with varying pulse return times and<br />

corresponds to a zero off-nadir angle.<br />

The approximation used in Equation (5.3) comes from Rodriguez (1988), who<br />

2<br />

expands the Bessel function to first order and keeps terms of order β t and lower,<br />

then replaces series by exp( β<br />

2 t / 4)<br />

. The error introduced from this approximation is<br />

less than 1% (Rodriguez, 1988). This approximation is more reasonable for actual<br />

waveforms, because it takes account of the non-zero off-nadir angle and different<br />

pulse return times.<br />

In Equation (5.3), the sea surface roughness (through the composite risetime) and the<br />

off-nadir angle are two main factors that affect the shape of the waveform. Mean<br />

ocean-return waveforms in Figure 5.3 show the family of the curves, which result


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 129<br />

from varying the RMS wave height (SWH from 0 - 20 m with a 2 m interval). As can<br />

be seen from Figure 5.3, varying SWH changes the rise time (or the slope) of the<br />

leading edge of the waveform. The rougher the sea state, the longer the rise time (i.e.,<br />

lower gradient) of the waveform’s leading edge.<br />

550<br />

500<br />

450<br />

SWH = 0 m<br />

400<br />

Power (counts)<br />

350<br />

300<br />

250<br />

200<br />

SWH = 20 m<br />

150<br />

100<br />

50<br />

0<br />

0 10 20 30 40 50 60<br />

Bins<br />

Figure 5.3 Variation of the leading edge’s slope with varying SWH (0 - 20 m),<br />

modelling waveforms using the ocean model ( ξ = 0 °).<br />

In addition to the surface roughness, another factor that can affect the waveform rise<br />

time is the surface slope or altimeter antenna off-nadir pointing angle. Figure 5.4<br />

shows the effects of the off-nadir angle on the waveform, for which SWH = 4 m is a<br />

typical situation of the sea surface and the numbers give the values of the off-nadir<br />

angles. As can be seen from Figure 5.4, there are three effects:<br />

(1) The off-nadir angle ξ changes not only the slope of the trailing edge, but also<br />

the slope of the leading edge. Both slopes of the leading and trailing edges<br />

decrease with increasing ξ . However, the location (or range to the mean sea<br />

surface) of the mid-point on the leading edge of the waveform does not change<br />

with varying ξ .


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 130<br />

500.0<br />

Power (counts)<br />

400.0<br />

300.0<br />

200.0<br />

ξ=0.0<br />

ξ=0.2<br />

ξ=0.4<br />

ξ=0.6<br />

100.0<br />

ξ=0.8<br />

ξ=1.0<br />

0.0<br />

0 8 16 24 32 40 48 56 64 72 80<br />

Bins<br />

Figure 5.4 Effect of the off-nadir angle (in degrees) on the amplitude, leading edge,<br />

and especially the slope of the trailing edge, modelling waveforms using the ocean<br />

model (SWH = 4 m).<br />

(2) The magnitude of the amplitude of the waveform is affected as well. Since an<br />

off-nadir angle changes the decay of the trailing edge, it will affect the value of<br />

the AGC gate and hence the AGC scaling of the waveform.<br />

(3) The slope of the trailing edge changes approximately linearly with off-nadir<br />

angle (Barrick and Lipa, 1985). The trailing edge’s slope changes its direction<br />

at ξ ≈ 0.6° (Figure 5.4). The reason is that as the surface slope increases from<br />

zero, the pulse-limited footprint moves from the centre of the beam and the<br />

leading edge part of the waveform (both leading edge and amplitude) is<br />

attenuated by the antenna gain function. At steeper angles ( ξ >1°), the<br />

attenuation becomes so great that the AGC loop can no longer compensate the<br />

loss in signal strength (Martin et al, 1983; Brrick and Lipa, 1985). As a result,<br />

loss of tracking will occur.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 131<br />

5.5.2 Five- and Nine-β-Parameter Models<br />

The β-parameter models (Equations 3.40 and 3.43) are also employed in the<br />

retracking system because of their capability of dealing with complex waveforms<br />

reflected from single or multiple scattering surfaces. They are developed based upon<br />

a modified version of the Brown (1977) ocean return function (Parsons, 1979), but<br />

add an additional slope parameter after the ramp (discussed later in Section 5.5.3). A<br />

description of the five- and nine-β-parameter models can be found in Section 3.5.1.<br />

The retracking corrections dr can be computed by the following formulas.<br />

(a) The Five-β-Parameter Model<br />

dr = G 2 M<br />

⋅ ( β − g<br />

0<br />

)<br />

(5.5)<br />

where β<br />

30<br />

is the retracking gate estimate in units of waveform gates.<br />

30<br />

(b) The Nine-β-Parameter Model<br />

dr = G 2 M<br />

⋅ ( β − g )<br />

(5.6)<br />

1<br />

30 0<br />

dr = G 2 M<br />

⋅ ( β − g )<br />

(5.7)<br />

2<br />

60 0<br />

where β<br />

60<br />

is the retracking gate estimate in units of waveform gates (related to the<br />

second ramp), dr 1 is the correction related to the range to at-nadir mean sea level,<br />

and dr 2 is the correction related to the range to the off-nadir (assumed) land surface.<br />

5.5.3 Comparison between the Ocean and Five-β-Parameter Models<br />

Both ocean and five-β-parameter models are fundamentally based upon the Brown<br />

(1977) model. Comparison between them will give qualitative guidance to selecting<br />

an adequate fitting function for the retracking procedure. Therefore, an assessment of<br />

these functions will be discussed in this Section. In order to compare them directly, it<br />

is necessary to simplify the ocean model (Equation 5.3). By assuming a zero offnadir<br />

angle (i.e., A = 0 ), Equation (5.3) can be rewritten as<br />

ξ


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 132<br />

2<br />

⎡<br />

⎤<br />

2<br />

1 ⎛ t − t ⎞ ⎡ ⎛<br />

⎞⎤<br />

0<br />

− ασ<br />

ασ<br />

P(<br />

t)<br />

= P + ⎢ ⎜<br />

⎟<br />

+ 1⎥<br />

exp⎢−<br />

⎜ − −<br />

⎟<br />

N<br />

PA<br />

erf<br />

α t t0<br />

⎥<br />

(5.8)<br />

2 ⎢⎣<br />

⎝ 2σ<br />

⎠ ⎥⎦<br />

⎣ ⎝ 2 ⎠⎦<br />

2<br />

in which P A<br />

= A0<br />

is the amplitude scaling term. The term ασ in Equation (5.8) can<br />

be computed by substituting the altimeter parameters from Table 2.1. The computed<br />

2<br />

values of ασ as a function of SWH are listed in Table 5.1. It can be seen, assuming<br />

2<br />

that a typical situation of the ocean surface (i.e., SWH = 4 m), the values of ασ are<br />

approximately 0.16 ns for both ERS-1/2 and TOPEX. Thus, this term may be safely<br />

neglected for the purpose of the waveform retracking, and Equation (5.8) can be<br />

reduced further to<br />

Table 5.1 ασ² Values for ERS-1/2 (ocean mode) and TOPEX<br />

SWH (m) ERS-1/2 (ns) TOPEX (ns)<br />

0.5 0.01 0.01<br />

2.0 0.05 0.05<br />

4.0 0.16 0.16<br />

6.0 0.35 0.35<br />

8.0 0.62 0.61<br />

10.0 0.96 0.95<br />

12.0 1.38 1.36<br />

14.0 1.88 1.85<br />

16.0 2.45 2.41<br />

18.0 3.10 3.05<br />

20.0 3.82 3.76<br />

P<br />

1<br />

2<br />

⎡<br />

⎢<br />

⎣<br />

⎛ t − t<br />

[ − ( t − )]<br />

0<br />

( t)<br />

= PN<br />

+ PA<br />

erf ⎜ ⎟ + 1 exp α t0<br />

⎝<br />

⎛ t − t<br />

⎝ σ<br />

⎞<br />

2σ<br />

⎠<br />

⎤<br />

⎥<br />

⎦<br />

0<br />

= PN<br />

+ PA<br />

P⎜<br />

⎟exp<br />

α<br />

⎞<br />

⎠<br />

[ − ( t − t )]<br />

0<br />

(5.9)<br />

where P (x)<br />

is the normal probability distribution (Equation 3.42).


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 133<br />

Comparing Equation (5.9) with Equations (3.40) and (3.43) [set i = 1 ], the only<br />

difference between the two functions is in the trailing edge of the waveform. In the<br />

region of the trailing edge, the normal probability distribution P (x)<br />

is equal to unity<br />

for SWHs less than 10 m. Thus, in this region, Equations (5.9), (3.40), and (3.43) can<br />

be written as P t<br />

(t)<br />

, y t<br />

(t)<br />

and y te<br />

(t)<br />

as<br />

[ − ( t − )]<br />

P ( t)<br />

= P + P exp α t<br />

(5.10)<br />

t<br />

y t<br />

N<br />

A<br />

1<br />

+ β<br />

21<br />

1<br />

( β )<br />

51<br />

1<br />

0<br />

( t)<br />

= β + Q<br />

(5.11)<br />

and<br />

y t<br />

1<br />

+ β<br />

21<br />

exp<br />

( − β )<br />

( t)<br />

= β<br />

Q<br />

(5.12)<br />

51<br />

1<br />

From the linear ocean surface (Equation 5.11), β<br />

51<br />

is the slope of the trailing edge.<br />

In addition, Equations (5.10) and (5.12) can be rewritten more simply as<br />

( P ( t)<br />

P ) = [ ln( P ) +α t ] −α<br />

t<br />

ln (5.13)<br />

t<br />

−<br />

N<br />

A 0<br />

and<br />

ln<br />

( y t<br />

( t)<br />

β1 ) = ln( β<br />

21<br />

) − β<br />

51Q1<br />

− (5.14)<br />

where α and β<br />

51<br />

are also the slope of the trailing edge of the ocean model and the<br />

five-β-parameter model with a exponential decay trailing edge, respectively.<br />

The slope of the trailing edge in the ocean model is related to the physical parameters<br />

of the radar antenna pattern, the Earth’s radius, and the off-nadir angle, according to<br />

4c<br />

1<br />

α =<br />

cos(2ξ<br />

)<br />

γ h (1 + h / R)<br />

(5.15)<br />

However, it can be seen from Equations (5.11) and (5.14) that the slope of the<br />

trailing edge depends on only a non-physical (or empirical) parameter, β 51<br />

. In this<br />

sense, the five-β-parameter model is derived partly from the ocean model (Equation<br />

5.3) or the Brown (1977) model (Equation 3.12) in the case of a linear ocean surface.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 134<br />

The slope of the waveform trailing is presented using an empirical parameter rather<br />

than the antenna off-nadir angle in the waveform. Therefore, the five-β-parameter<br />

model may be defined as a semi-empirical surface-scattering model.<br />

Variations of the slope of the waveform’s trailing edge modelled using the five-βparameter<br />

function (with linear trailing edge) for SWH = 4 m, β 21<br />

= 100 (counts),<br />

and − 0.03<br />

≤ β<br />

51<br />

≤ 0. 03 (0.01 interval) are shown in Figure 5.5. From this, two<br />

effects are evident:<br />

1000.0<br />

β5=0.03<br />

800.0<br />

β5=0.02<br />

Power (counts)<br />

600.0<br />

400.0<br />

β5=0.01<br />

β5=0.0<br />

β5=−0.01<br />

200.0<br />

β5=−0.02<br />

0.0<br />

β5=−0.03<br />

0 8 16 24 32 40 48 56 64 72 80<br />

Bins<br />

Figure 5.5 Variations of the trailing edge’s slope with varying parameter β 5 ,<br />

modelling waveforms using the five-β-parameter parameter function (linear trailing<br />

edge, SWH = 4 m, β<br />

21<br />

= 500 counts, β 5 = β<br />

51).<br />

(1) As the value of β 51<br />

increases, the power of the waveform increases, i.e., the<br />

values of β 51<br />

affect the amplitude. Similarly as for the ocean model, the<br />

position of the mid-point on the leading edge of the waveform is not influenced<br />

by variations of the slope in the trailing edge.<br />

(2) The slope of the trailing edge increases with increasing values of β 51<br />

. The<br />

descending trailing edge of the waveform changes to a rising trailing edge after<br />

β<br />

51= 0.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 135<br />

The results in Figure 5.5 show a common feature with Figure 5.4, in which the<br />

trailing edge region is nearly linear. However, β<br />

51<br />

can vary to a greater extent than<br />

the slope of the ocean model. This makes it possible that the five-β-parameter can fit<br />

more complex waveform shapes over non-ocean surfaces.<br />

Figure 5.6 shows modelling ERS-2 waveforms using the five-β-parameter function<br />

(Equation 3.43, with an exponential decayed trailing edge) for SWH = 4 m, β 21<br />

=<br />

100 counts, and − 0.03<br />

≤ β<br />

51<br />

≤ 0. 03 (0.01 interval). From Figure 5.6:<br />

(1) Although β<br />

51<br />

mainly affects the slope of the trailing edge and the amplitude of<br />

the waveform, it slightly changes the slope of the leading edge. However, this<br />

change does not vary the mid-point location on the leading edge.<br />

(2) Unlike the effects in the case of the linear trailing edge, the exponential decay<br />

trailing edge decreases both the slope and the amplitude with increasing β 51<br />

.<br />

Again, the slope of the trailing edge changes its direction after β<br />

51= 0.<br />

1500.0<br />

β5=−0.03<br />

Power (counts)<br />

1000.0<br />

500.0<br />

β5=−0.02<br />

β5=−0.01<br />

β5=0.0<br />

β5=0.01<br />

β5=0.02<br />

β5=0.03<br />

0.0<br />

0 8 16 24 32 40 48 56 64 72 80<br />

Bins<br />

Figure 5.6 Variations of the trailing edge’s slope with varying parameter β 5 ,<br />

modelling waveforms using the five-β-parameter parameter function (exponential<br />

decayed trailing edge, SWH = 4 m, β<br />

21<br />

= 500 counts, β 5 = β<br />

51).


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 136<br />

In addition, if the above analysis is performed on Equation (5.3), one can obtain the<br />

trailing edge part as<br />

2<br />

⎡ ⎛ ασ ⎞⎤<br />

P(<br />

t)<br />

= P + exp⎢−<br />

⎜ − −<br />

⎟<br />

N<br />

A α t t0<br />

⎥<br />

(5.16)<br />

⎣ ⎝ 2 ⎠⎦<br />

and<br />

ln<br />

[ P(<br />

t)<br />

− P ]<br />

N<br />

⎡ ⎛<br />

= ln( A)<br />

+ ⎢−α<br />

⎜t<br />

− t<br />

⎣ ⎝<br />

⎡ ⎛<br />

= ⎢ln(<br />

A)<br />

+ α<br />

⎜t<br />

⎣ ⎝<br />

0<br />

0<br />

2<br />

ασ ⎞⎤<br />

−<br />

⎟⎥<br />

2 ⎠⎦<br />

2<br />

ασ ⎞⎤<br />

+<br />

⎟⎥<br />

−αt<br />

2 ⎠⎦<br />

(5.17)<br />

in which<br />

A<br />

0<br />

= A A , and α is the slope of the trailing edge of the ocean waveform.<br />

ξ<br />

Since only ξ is a variable in the function α (ξ ) related to an altimeter<br />

(Equation 5.15), the trailing edge’s slope varies with variation of ξ . This is the same<br />

result as that discussed in Section 5.5.1, in which ξ has been found to be a dominant<br />

factor of the trailing edge’s slope.<br />

By comparison, it is evident that different models have different slopes of the trailing<br />

edge and variations in the slope of the trailing edge do not change the position of the<br />

mid-point on the leading edge of the waveform. In coastal regions, however,<br />

waveforms show diverse shapes with different slopes (see Figure 5.2) of the trailing<br />

edge. An inappropriate slope description may degrade the parameter estimates of the<br />

mid-point (i.e., retracking gate) on the leading edge or cause a failure of the fitting<br />

procedure. The above analysis is significant for understanding the differences among<br />

the fitting functions. More importantly, it is the different slope characters of the<br />

trailing edges that make these functions complement each other for retracking<br />

different waveform shapes in coastal regions.<br />

5. 6 Iterative Nonlinear Fitting Procedure<br />

No matter whether or not the convolution functions contain non-linear wave<br />

parameters, the solution of the fitting functions is characterised by a non-linear<br />

model. Thus, there is no straightforward method to optimally fit the model to the


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 137<br />

return waveforms and extract the model parameters. This is further complicated due<br />

to the noisy and shape-mixed features of measured waveforms in coastal regions. All<br />

fitting models describe the mean (or idealised) power, whereas real waveforms<br />

measured at the altimeter fluctuate about the mean because of the fading noise,<br />

variations in the surface geometry, and inhomogeneities near the land. Therefore, the<br />

determination of model parameters is based on an iterative least squares approach<br />

and an appropriate weighting scheme.<br />

5.6.1 Solutions of the Waveform Retracking Fitting Function<br />

In general, non-linear models in the standard statistical techniques are linearised by<br />

using approximate values via a first-order Taylor series expansion of the unknown<br />

parameters. The corrections to the approximate values become the unknown<br />

parameters and solutions for the corrections are iterated in order to improve the<br />

approximate values. Taking y i<br />

as a function of the estimated parameter<br />

k = 1,<br />

2, K,m<br />

, it can be written as:<br />

x<br />

k<br />

,<br />

y<br />

=<br />

( x )<br />

i<br />

F i<br />

m×1<br />

(5.18)<br />

where<br />

i = 1,<br />

2, L,<br />

n is the bin-number of the waveform samples in range window<br />

(e.g., n = 64 for ERS-1/2). Assuming x = x 0<br />

+ δ x , y<br />

i<br />

is approximated at x 0<br />

in the<br />

parameter-space by the Taylor series expansion to the first order as<br />

∧<br />

y ≈ y = F x ) + F′<br />

( x ) δ x<br />

i<br />

i<br />

i<br />

(<br />

0 i 0<br />

(5.19)<br />

where F x ) is an approximation computed by the fitting function using the<br />

i<br />

( 0<br />

approximated values of the unknowns, and<br />

F′<br />

( x<br />

i<br />

0<br />

)<br />

⎛<br />

⎜ ∂y<br />

=<br />

⎜ ∂x1<br />

⎝<br />

∂y<br />

∂x<br />

L<br />

∂y<br />

∂x<br />

x 2<br />

0 x<br />

m<br />

0<br />

x0<br />

⎞<br />

⎟<br />

⎟<br />

⎠<br />

(5.20)<br />

is a matrix populated by the partial derivatives of the functional model with respect<br />

to the unknowns. From Equations (5.19) and (5.20), the function is formed as


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 138<br />

V = Aδ x − L<br />

(5.21)<br />

where A = F′<br />

x ) is the design matrix, L = l − F x ) is the vector of the n<br />

( 0<br />

observations determined by the measured waveforms and the first-order<br />

approximations at x<br />

0<br />

. Equation (5.21) can be solved by the least squares approach:<br />

( 0<br />

T −1<br />

T<br />

δ x = ( A PA)<br />

( A PL)<br />

(5.22)<br />

where P is the weight matrix of the observations. The standard deviations of the<br />

estimated parameters are given as<br />

2 T −1<br />

Q x<br />

= σ<br />

0<br />

( A PA)<br />

,<br />

(5.23)<br />

σ<br />

2<br />

0<br />

=<br />

T<br />

V PV<br />

n − m<br />

.<br />

(5.24)<br />

The initial approximation of x<br />

0<br />

is corrected to produce revised estimates<br />

x = x 0<br />

+ δ x . Then, the x replaces the original x 0<br />

as new a priori values and the<br />

iterative procedure is repeated N times until a defined relative error criterion is<br />

satisfied, or a maximum number of iterations is reached.<br />

5.6.2 Linearisation of the Waveform Retracking Fitting Function<br />

Having expressed the return power as analytic functions, the linearisation of these<br />

functions can be explicitly implemented by computing the partial derivatives of each<br />

parameter. Taking the revised ocean model (Equation 5.3) as an example, the five<br />

partial derivatives of the modelled waveform with respect to the parameters are given<br />

by<br />

∂P<br />

∂<br />

P N<br />

= 1<br />

(5.25)<br />

∂P =<br />

∂A<br />

P<br />

0<br />

A 0<br />

(5.26)


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 139<br />

∂P<br />

∂t<br />

0<br />

=<br />

A0<br />

⎧ 2<br />

2 ∂u<br />

∂v<br />

Aξ<br />

exp( −v)<br />

⎨ exp( −u<br />

) − [ erf ( u)<br />

+ 1]<br />

2 ⎩ π ∂t0<br />

∂t0<br />

⎫<br />

⎬<br />

⎭<br />

(5.27)<br />

∂P<br />

=<br />

∂σ<br />

A<br />

2<br />

0<br />

⎧<br />

Aξ<br />

exp( −v)<br />

⎨<br />

⎩<br />

2<br />

exp( −u<br />

π<br />

2<br />

∂u<br />

) −<br />

∂σ<br />

[ erf ( u)<br />

+ 1]<br />

∂v<br />

⎫<br />

⎬<br />

∂σ<br />

⎭<br />

(5.28)<br />

∂P<br />

2<br />

∂ξ<br />

=<br />

A0<br />

2<br />

+<br />

⎪⎧<br />

1<br />

Aξ<br />

exp( −v)<br />

⎨<br />

⎪⎩ A<br />

2<br />

exp( −u<br />

π<br />

2<br />

ξ<br />

∂u<br />

)<br />

2<br />

∂ξ<br />

[ erf ( u)<br />

+ 1]<br />

−<br />

∂Aξ<br />

2<br />

∂ξ<br />

[ erf ( u)<br />

+ 1]<br />

∂v<br />

2<br />

∂ξ<br />

⎫<br />

⎬<br />

⎭<br />

(5.29)<br />

where<br />

u =<br />

τ<br />

2<br />

(5.30)<br />

⎛ d ⎞<br />

v = d⎜τ<br />

+ ⎟<br />

⎝ 2 ⎠<br />

(5.31)<br />

It should be noted that the derivative with respect to ξ 2 has been computed rather<br />

than the derivative with respect to ξ in Equation (5.29). The reason for this is that<br />

the derivative with respect to ξ could be zero if ξ → 0<br />

, and hence the off-nadir<br />

angle cannot be estimated. Thus, assuming that the two sinusoidal terms in Equations<br />

(3.14) and (3.15) can be approximated to the first order for small angles, ξ can be<br />

2<br />

replaced by ξ to avoid the problem. An additional problem with the introduction of<br />

2<br />

ξ is that the estimate of ξ 2 could contain negative values due to the presence of<br />

noise. However, this effect can be neglected in the study, as the aim is to extract the<br />

retracking gate estimate rather than the off-nadir angle. For research that is interested<br />

only in the off-nadir estimate, other approaches can be used (cf. Barrick and Lipa,<br />

1985).<br />

5.6.3 Initial Estimates of the Unknown Model Parameters<br />

High-quality initial estimates of the parameters are important to reduce the number<br />

of iterations and guarantee a convergence in the least squares estimation (and hence


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 140<br />

computational time) that the non-linear least squares procedure takes to converge. If<br />

the initial parameters cannot be estimated precisely, the parameter corrections after<br />

one iteration might be very large, causing the algorithm to diverge rapidly or to<br />

converge on the incorrect result. However, there is no predefined method that<br />

determines the optimum a priori values for the altimeter waveforms. Therefore, to<br />

generate initial parameter estimates, different schemes were tested and evaluated<br />

based on the speed of convergence, the mean-squared error, and ability to fit a wide<br />

variety of waveforms. In accordance with the testing results of these three factors, the<br />

OCOG method can also provide additional information when making initial<br />

estimates of the parameters (see below).<br />

(1) The Ocean Model<br />

Before ascertaining the initial parameter estimates, the initial estimate of the thermal<br />

noise (or noise baseline),<br />

0<br />

P<br />

N<br />

, is determined by calculating the average amplitude of<br />

the waveform samples over five of earlier gates (e.g., gates 6-10 for POSEIDON)<br />

P<br />

1<br />

10<br />

0<br />

N<br />

= ∑W i<br />

5 i=<br />

6<br />

(5.32)<br />

where<br />

W<br />

i<br />

is the i -th measured waveform sample. This value is then subtracted from<br />

each waveform sample and the OCOG method is applied to the waveforms to<br />

calculate relative parameters of the rectangle (described in Section 3.5.2). The results<br />

from the OCOG method are used to determine the initial approximations of the<br />

amplitude term (<br />

0<br />

A<br />

0<br />

) and the time corresponding to the location of the mid-point at<br />

the range bins ( t 0 0<br />

). The waveform amplitude ( A ) is set to be the initial estimate of<br />

0<br />

A<br />

0<br />

, and t 0 0<br />

can be obtained from the gate number of the leading edge ( LEG ) from<br />

the OCOG by<br />

t<br />

0<br />

0<br />

= LEG ⋅T<br />

(5.33)<br />

where T is the duration of the radar pulse ( T = 3. 03ns for ERS-1/2 and T = 3. 125ns<br />

for TOPEX/POSEIDON).


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 141<br />

The initial approximation of the RMS surface roughness (i.e., the composite<br />

0<br />

risetime), σ<br />

cr<br />

, can be estimated by substituting SWH = 2cσ<br />

s<br />

into Equation 3.13 to<br />

give<br />

σ<br />

0<br />

cr<br />

⎡<br />

= ⎢<br />

⎢⎣<br />

( f t<br />

T )<br />

2<br />

2<br />

1/ 2<br />

⎛ SWH ⎞<br />

2 ⎥ ⎥ ⎤<br />

+ ⎜ ⎟<br />

⎝ c ⎠ ⎦<br />

(5.34)<br />

where factor f t<br />

is an approximation (0.425 for T/P and 0.526 for ERS-1/2) of the<br />

compressed pulse shape by a Gaussian function (Marth et al., 1993; Batoula et al.,<br />

1999), SWH = 4<br />

m, which is the common case over ocean surfaces. The initial<br />

2<br />

approximation of the (small) off-nadir angle squared, ξ , can be set to zero.<br />

(2) The Five-β-parameter Model<br />

The initial approximations for the five- β-parameter model are determined using the<br />

method from Anzenhofer et al. (2001), but modified slightly to account for the<br />

thermal noise. The initial estimate of the parameter β<br />

1<br />

is set using Equation (5.32).<br />

The maximum amplitude of the waveform after removing the thermal noise (item 2)<br />

is defined as the initial estimate of the parameter β<br />

2<br />

. The gate number of the leading<br />

edge from the OCOG is taken for the initial approximation of the parameter β<br />

3<br />

. The<br />

initial estimate of the waveform risetime β<br />

4<br />

is set empirically to 1.3 gates, and the<br />

slope of the trailing edge β<br />

5<br />

can be set to zero.<br />

(3) The Nine-β-parameter Model<br />

Measured waveform shapes related to this function usually show double ramps<br />

(Figure 3.6). Of these parameters, the initial estimate of the thermal noise can be set<br />

using Equation (5.32). The initial estimate for the location of the first ramp (usually<br />

related to the mean sea surface) is determined by locating the first sample gate,<br />

whose slope exceeds a threshold value (e.g., 50 counts/gate for ERS-1/2). Once this<br />

threshold level is exceeded, the next five gates are checked to find the maximum<br />

slope. The gate number corresponding to this maximum slope is used as the initial<br />

approximation for the location of the first ramp. The same method can be used to<br />

determine the initial estimate for the second ramp location. Other initial estimates


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 142<br />

can be defined using the similar method described above for the five-β-parameter<br />

model.<br />

5.6.4 Data Weight and Outlier Detection Scheme<br />

Before fitting, a priori weights are determined using the method from Anzenhofer et<br />

al. (2000). It is acknowledged that the heavier weight emphasis should be put on<br />

fitting to the ramp locations which define the retracking correction (e.g., Zwally,<br />

1996; Brenner et al., 1993; Anzenhofer et al., 2000). Using the parameters estimated<br />

from the OCOG retracker, the location of the leading edge, tp , can be estimated.<br />

Then a priori weights of the ERS-1/2 waveform can be determined empirically for<br />

the ocean model and the five-parameter model as<br />

0<br />

p i<br />

= 200 for t ≤ tp -1<br />

(5.35)<br />

= 100 for tp-1<br />

< t ≤ tp + 2<br />

0<br />

p i<br />

(5.36)<br />

0<br />

p i<br />

= 50 for tp + 2 < t ≤ 64<br />

(5.37)<br />

This is a reasonable weight scheme because it accounts for the characteristics and<br />

distribution of the useful information in the waveform. The heaviest weights are put<br />

on the waveform samples in the first plateau (i.e., thermal noise region) before the<br />

leading edge so that they will be thought to be relatively accurate measurements in<br />

the adjustment. The second heaviest weights are applied to the samples of the leading<br />

edge itself. The lightest weights are given to the waveform samples in the trailing<br />

edge, making them contribute less to the solutions (cf. Figure 3.5).<br />

In general, the data in the trailing edge show obvious undulations caused by the<br />

‘fading’ noise, which results in the incoherent superposition of signals from different<br />

reflecting facets (Partington, 1991; Marth et al., 1993; Quartly and Srokosz, 2001).<br />

The noise in the trailing edge greatly influences the iterative fitting procedure and<br />

even accurate results cannot be estimated (Tokmakian et al., 1994). To solve the<br />

problem, two approaches have been used. The first is to average waveforms in the<br />

time span of several seconds, and then to fit a model to this averaged waveform (e.g.,


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 143<br />

Tokmakian et al., 1994; Hayne and Hancock III, 1990). The second is to lightly<br />

weight the waveform samples in the trailing edge (Anzenhofer et al., 2000; Zwally,<br />

1996).<br />

However, the averaging procedure is not always appropriate in coastal regions,<br />

because ocean waveforms will be distorted by contaminated waveforms within the<br />

averaging period. Therefore, the second method is used in the study. According to<br />

Anzenhofer’s (2000) weight scheme, the data after the ramp have been<br />

downweighted in the solution (see Equation 5.37). However, it is found in this study<br />

that some unexpected undulations in the trailing edge significantly affect the solution.<br />

To overcome this problem, an outlier detection approach from an iterative weight<br />

scheme is used in this study.<br />

In this iterative procedure, the initial least squares adjustment is conducted using the<br />

0<br />

a priori values ( β<br />

k 0<br />

) and a priori weights ( p<br />

i<br />

). After the first adjustment, a weight<br />

function<br />

w<br />

i<br />

is given as (Li, 1988)<br />

⎧ 1<br />

⎪ for vi<br />

> 0.7σ<br />

0<br />

vi<br />

w<br />

i<br />

= ⎨<br />

(5.38)<br />

⎪<br />

1<br />

for vi<br />

≤ 0.7σ<br />

0<br />

⎪⎩<br />

0.7σ<br />

0<br />

2<br />

where v<br />

i<br />

and σ<br />

0<br />

, which is the estimation of the unit weight variance, are computed<br />

by Equations (5.21) and (5.24), respectively. The next step then is to determine the<br />

weight (<br />

j+1<br />

p<br />

i<br />

) for the next adjustment using wi<br />

as follows:<br />

p =<br />

j+1<br />

i<br />

p<br />

j<br />

i<br />

w<br />

i<br />

(5.39)<br />

where j means the j-th iteration in the adjustment. In the meantime, estimates of the<br />

β parameters take the place of the originals and repeat the entire cycle to produce<br />

improved estimates of the β parameters. In this way, the observations considered to<br />

contain the outlier will be downweighted but still be kept in the dataset, while nonoutlier<br />

observations will maintain full weight (though downweighted with respect to<br />

the other data in the range window) after the iterative cycle.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 144<br />

5.6.5 Iterative Procedure<br />

The procedure taken here is similar to that used by Anzenhofer et al. (2000), Zwally<br />

(1996) and Davis (1993). The parameter estimates are obtained by an iterative<br />

procedure as follows:<br />

(1) Estimate initial values of the model parameters and the weights of the<br />

measured waveforms.<br />

(2) Compute the new values for the least squares approach using the first-order<br />

Taylor expansion and measured waveforms (Equation 5.21).<br />

(3) Compute a correction to the current estimates of the unknown parameters<br />

(Equation 5.22), and the standard deviations of the parameters (Equation 5.23).<br />

(4) Update the parameter estimates for the new initial estimates.<br />

(5) Repeat steps 2-4 until some exit criterion in Table 5.2 is satisfied.<br />

From the flowchart of the fitting algorithm (Figure 5.1), after corrected parameter<br />

estimates are obtained from Equation (5.22), several exit criteria are applied to<br />

determine whether the fitting iteration continues.<br />

(1) The first exit criterion is based upon an absolute difference specified for each<br />

unknown parameter between ( i +1)<br />

and i iterations. Table 5.2 lists the<br />

absolute difference values for the ocean model parameters. If the absolute<br />

difference of all parameter estimates from two successive iterations is less than<br />

these criteria, the fitting iteration terminates.<br />

(2) The second exit criterion is based upon the relative change in the standard<br />

2<br />

deviation of the unit weight, σ<br />

0<br />

. If the relative change of the σ<br />

0<br />

is less than<br />

2% from one iteration to the next, the fitting loop is ended.<br />

(3) The final exit criterion is the maximum number of the iterations. The<br />

maximum number in this study is set to 20 to avoid infinite looping situation in<br />

the case of divergence. In practice, a flag which records the status of the fitting<br />

loop will be returned with the parameter estimates. The parameter estimates are<br />

retained if the algorithm exits before 20 iterations and the flag does not show<br />

iteration failure, while they are discarded if the other criteria fail in 20<br />

iterations or the loop ends at the maximum iterative number.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 145<br />

In addition, for each iteration step, some constraints are applied, such as amplitude<br />

parameter estimate must be larger than zero.<br />

Table 5.2 Criteria of the absolute differences between two iterations (the ocean<br />

model for ERS-1/2)<br />

Parameters Absolute Differences Units<br />

P<br />

N<br />


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 146<br />

Equation (5.32) using five unaliased waveform samples before doing retracking.<br />

Once the estimate of P<br />

N<br />

is obtained, the threshold level, T<br />

l<br />

, is then set to<br />

T = P + T<br />

(5.40)<br />

l<br />

N<br />

LO<br />

where T<br />

LO<br />

is the threshold level determined using the selection scheme (d) in Section<br />

5.7.2. The retracking location on the leading edge of the waveform (or at the range<br />

window),<br />

g<br />

r<br />

, is linearly interpolated between the bins adjacent to T<br />

l<br />

using<br />

=<br />

T<br />

−1+<br />

−W<br />

l k −1<br />

g<br />

r<br />

g<br />

k<br />

(5.41)<br />

Wk<br />

−Wk<br />

−1<br />

where<br />

g<br />

k<br />

is the location of the first gate exceeding T l<br />

. The g r<br />

is set to ( g<br />

k<br />

−1)<br />

in<br />

the case of W k<br />

W<br />

1<br />

in Equation (5.41). Then, the range correction can be obtained<br />

= k −<br />

from a bin-to-metre conversion factor timing the gate difference between this<br />

estimate and the designed tracking gate as<br />

dr = G ⋅ ( g −<br />

0<br />

)<br />

(5.42)<br />

2M<br />

r<br />

g<br />

where g<br />

0<br />

is the expected tracking gate in units of waveform gates.<br />

5.7.2 Selection of the Threshold Level<br />

As stated, Brooks et al. (1997) select the threshold level for mixed waveform shapes<br />

in coastal regions based upon a visual examination method. These are two concerns<br />

related to this selection. The first is that threshold values are higher than both the<br />

noise and the minor waveform leakage in the early gates. The second is that the<br />

values are sufficiently lower than the sought-after ocean returns. Brooks et al. (ibid.)<br />

determine different threshold values for 18 TOPEX ground tracks near land using<br />

both concerns. Then, they are used to retrack the coastal waveform along the same<br />

pass. This method is good for selecting the threshold level because it takes account<br />

for the strength of the power backscattered from different reflecting surfaces.<br />

However, because an automatic selection of the threshold level is not available, this<br />

restrains the technique from retracking a wide range of the waveforms in coastal<br />

regions.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 147<br />

Considering the shape variations of the waveform in the proximity of the coasts<br />

analysed in Section 4.6, the selection of the threshold level used in the coastal<br />

retracking system differs from these previously published methods used over ice<br />

sheets (Zwally, 1996; Davis, 1995; Davis, 1997) and coastal regions (Brooks et al.,<br />

1997). Four threshold schemes are investigated before arriving at the final threshold<br />

level.<br />

(a) Fixed Counts<br />

The threshold level of the fixed count is defined here as the 50% of the meanwaveform<br />

amplitude in counts. For each satellite ground pass, the mean waveform is<br />

formed from one or two seconds of purely ocean waveforms prior to the coastal<br />

waveforms along it. The power in counts corresponding to the 50% of the meanwaveform<br />

amplitude is the fixed count for the pass. This fixed count is, then, used as<br />

the threshold level to retrack the waveforms along the same pass into coastal regions.<br />

The gate location corresponding to this value is linearly interpolated between the<br />

bins adjacent to the crossing of the fixed counts on the leading edge of the coastal<br />

waveform, being the estimate of the retracking gate. This scheme, in fact, determines<br />

the retracking gate estimates using the 50% of the ocean-waveform amplitude, but it<br />

assumes that the amplitude of the ocean returns does not change with respect to the<br />

reflecting surfaces.<br />

(b) 50% of the Waveform Amplitude<br />

Over open oceans where the waveform is completely dominated by surface scattering,<br />

the half-power point (i.e., the 50% threshold retracking point) best represents the<br />

expected tracking-gate estimate which corresponds to the range to the at-nadir<br />

surface within the altimeter pulse-limited footprint. Therefore, the second selecting<br />

scheme is to choose the 50% threshold point as the threshold level.<br />

(c) 30% and 50% of the Waveform Amplitude<br />

Factors that affect the waveform’s shape (or amplitude) are primarily the off-nadir<br />

angle or the surface slope and the surface geometry. Compared with brighter and<br />

higher land, where returns usually show higher power, ocean returns that appear in<br />

the range window tend to be lower power. Therefore, in the case of the mixed<br />

waveform of land and water, an appropriate low threshold level is suitable such that<br />

retracking can locate the half-power point of the leading edge of ocean returns.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 148<br />

Mean Power and STD (counts)<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

Mean waveform<br />

STD<br />

1 4 7 10 13 16 19 22 25 28 31 34 37 40 43 46 49 52 55 58 61 64<br />

Bins<br />

Mean Power and STD (counts)<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

Mean waveform<br />

STD<br />

1 4 7 10 13 16 19 22 25 28 31 34 37 40 43 46 49 52 55 58 61 64<br />

Bins<br />

Figure 5.7 Mean waveforms and the standard deviation in two 5 km-wide bands of<br />

0-5 km (top) and 5-10 km (bottom) from the Australian coastline.<br />

Therefore, in this third scheme, two threshold levels are selected depending on the<br />

distance between the observing point along the satellite groundtrack and the coast.<br />

The 30% threshold level is only selected for the waveforms within 5 km distance<br />

from the coast, while a threshold level of the 50% of the waveform amplitude is<br />

chosen after 5 km from the coastline. The distance of 5 km is determined based upon<br />

an analysis of the mean waveforms (Section 4.5.2). To determine the proper<br />

threshold level within different distance bands from the coast, the average waveform<br />

is formed using one cycle of ERS-2 20 Hz waveforms from the coastal regions of


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 149<br />

interest within six 5 km-wide bands from the Australian coastline. Figure 5.7 shows<br />

the mean waveforms and STD in two 5 km-wide bands of 0-5 km and 5-10 km from<br />

the coastline. Beyond 10 km from the coastline, the mean waveforms show no<br />

significant change to the 5-10 km band.<br />

The position of the threshold crossing corresponds to the mid-point of the leading<br />

edge. The 30% of the waveform amplitude (or 20% of the amplitude in some areas,<br />

discussed below) is selected to be the threshold level for retracking waveforms<br />

within 0-5 km from the coasts. When the distance from the coast is larger than 5 km,<br />

the ocean scattering becomes predominant in the waveform returns so that the 50%<br />

of the waveform amplitude can be used as an estimate of the threshold level.<br />

(d) Varying Threshold Level<br />

Although the 30% threshold level should effectively be a low-power strength<br />

retracker in coastal regions for the mixed waveforms of land and water, it might<br />

cause two problems. The first one is that waveforms without contamination will be<br />

retracked inaccurately by this lower-level retracker. The second is that changing the<br />

threshold level at 5 km distance from the coast might cause a step jump to the<br />

retracked range measurements. To avoid both problems, a scheme of varying<br />

threshold level is developed.<br />

Different threshold levels are set depending on whether or not the waveform is<br />

contaminated. If the waveform classification indicates a distorted waveform, the<br />

threshold level is set to 30% (or 20%) of the waveform amplitude, otherwise the 50%<br />

of waveform amplitude is set. Compared with the waveform shape, the distance from<br />

the coast is not significant in this criterion. It links the threshold level to the<br />

contaminated waveform, avoiding the step change that might occur in corrected<br />

range measurements from the threshold level (c).<br />

5.7.3 Results of Selecting the Threshold Levels<br />

In order to ascertain the appropriate threshold level, one cycle of 20 Hz ERS-2<br />

waveform data (March to April 1999) and ten (15º×15º) areas around the Australian<br />

coast are used (Figure 5.8). The AUSGeoid98 (2'×2') geoid grid is used as an<br />

external ‘ground truth’ for the result comparisons (see also Section 2.5.1). Because


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 150<br />

the waveform contamination strongly depends on both the distance from the coast<br />

and land surface topography along the coastline, computations are conducted in six 5<br />

km-wide bands around the coasts. Waveforms have been threshold retracked using<br />

the above schemes, respectively. The range correction calculated from waveform<br />

retracking and geophysical corrections from the data products are then applied to the<br />

measured range. The retracked SSH data are then obtained from the satellite altitudes<br />

and corrected range measurements.<br />

105˚<br />

0˚<br />

120˚<br />

135˚<br />

150˚<br />

165˚<br />

0˚<br />

1 2<br />

-15˚<br />

3<br />

4 5<br />

6<br />

-15˚<br />

-30˚<br />

-30˚<br />

7<br />

8 9<br />

10<br />

-45˚<br />

105˚<br />

120˚<br />

135˚<br />

150˚<br />

-45˚<br />

165˚<br />

Figure 5.8 Ten 15º×15º coastal areas around Australia, showing the area number<br />

(Lambert projection).<br />

The comparison between the instantaneous SSH and the geoid height is an<br />

appropriate tool to investigate the overall quality of the threshold-level selecting<br />

schemes. For each along-track altimeter SSH measurement after retracking, the<br />

gridded AUSGeoid98 geoid height is bispline-interpolated to the longitude and<br />

latitude of the point. The difference between the mean SSH and geoid height can be<br />

obtained by subtracting the interpolated geoid height from the corresponding SSH.<br />

Then, the descriptive statistics are computed from this difference. The instantaneous<br />

SSH, of course, is not equivalent to the geoid (see Section 2.5.2). The difference


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 151<br />

between them includes the time variation (e.g., tides) and dynamic SST. However,<br />

averaging can remove the time variation and so the remaining difference is between<br />

the mean SSH and the geoid height. Since the precisely regional geoid model<br />

AUSGeoid98 is used, the mean difference can be a good approximation of the<br />

dynamic SST, which is of the order of ~1-2 m (e.g., Chelton et al., 2001). Therefore,<br />

the verification of any improvement in the SSH will only be seen in the standard<br />

deviations of the mean difference, with a smaller value indicating an improvement in<br />

the determination of the threshold level.<br />

Figures 5.9 and 5.10 show mean differences of the SSH and geoid heights and<br />

standard deviations in a distance out to 5 km around Australia’s coasts, respectively.<br />

Before retracking, the magnitudes of mean differences in ten areas are less than 2 m,<br />

but the standard deviations (STD) are large in most of the areas. In area one, where<br />

the land topography involves rocks and small barrier reefs in the north Australian<br />

coasts, the STD is even larger (±19.93 m). The STDs are small (±0.47 m and ±0.49<br />

m) in south-western and south-eastern areas seven and ten, where the coastal relief<br />

comprises rocks, cliffs and large barrier reefs (cf. Thom, 1984).<br />

After threshold retracking using scheme (b), the STD decreases in six areas.<br />

However, it increases in the other four areas compared with the STD before<br />

retracking. This suggests that the 50% threshold level is not an appropriate selection<br />

for the waveforms within 5 km from the coast. The reason for this is that the mostly<br />

varying land topography and standing water within this distance can heavily<br />

contaminate the waveforms, so that many waveform shapes do not follow the ocean<br />

model. After retracking using schemes (c) and (d), the STD shows improvements in<br />

seven areas. Of them, the results from the scheme (d) show the most significant<br />

improvement than any others. In contrast, scheme (a) gives the least improvement,<br />

and even worse results in most areas after retracking compared with the results<br />

before retracking. For areas 5, 8 and 10, an appropriate threshold level of 20% of the<br />

waveform amplitude has been found from the tests.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 152<br />

2.5<br />

2<br />

unretracked<br />

50% threshold level<br />

30% threshold level<br />

varying threshold level<br />

fixed counts<br />

1.5<br />

Mean Differences (m)<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

1 2 3 4 5 6 7 8 9 10<br />

Areas<br />

Figure 5.9 Mean differences between the SSH data after retracking and AUSGeoid98<br />

geoid heights in ten areas 5 km from the Australian coastline.<br />

25<br />

20<br />

unretracked<br />

50% threshold level<br />

30% threshold level<br />

varying threshold level<br />

fixed counts<br />

15<br />

STD (m)<br />

10<br />

5<br />

0<br />

1 2 3 4 5 6 7 8 9 10<br />

Areas<br />

Figure 5.10 Standard deviations of the mean differences in ten areas 5 km from the<br />

Australian coastline.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 153<br />

Figures 5.11 and 5.12 show mean differences of the SSH and geoid heights and<br />

standard deviations in a distance band of 5-10 km from Australia’s coasts,<br />

respectively. As can be seen from the ordinate scale in Figures 5.11 and 5.12, both<br />

mean differences and the STDs decrease with increasing the distance from the coasts.<br />

It is evident that the SSH is improved after retracking no matter which threshold<br />

level is used. The schemes (b) and (d) show similar results in most areas. This<br />

implies that less waveform contamination occurs in these areas, which gives the<br />

same result as that shown in Chapter 4. The results from scheme (a) present<br />

improvements in this distance band as well. However, the results from Figures 5.11<br />

and 5.12 suggest that scheme (d) is the best threshold level.<br />

The mean differences and the STD in four other 5 km-wide distance bands (10-30<br />

km from the coast) were computed and analysed. When the distance between the<br />

ground track and coastline is far more than 10 km, schemes (b) and (d) generate a<br />

very good agreement for both the mean and STD from the results after retracking.<br />

Results suggest that less waveform contamination occurs in these areas and the<br />

threshold level can be set to 50% of the waveform amplitude.<br />

During testing, it became evident that the varying threshold level can be used to<br />

effectively threshold retrack most waveforms in coastal regions, since it takes into<br />

account the coastal waveform characteristics. The count-fixed level (scheme a)<br />

cannot follow the variations of the scattering surfaces. It assumes that the retracking<br />

amplitude of the ocean waveform does not change with varying the surface<br />

roughness or topography. However, this is not the actual case of the waveform even<br />

over ocean surfaces. The 50% threshold retracking point is set too high to catch the<br />

ocean returns for the contaminated waveforms in coasts. Setting the 30% threshold<br />

level within 5 km from the coast and the 50% threshold after 5 km (scheme c)<br />

presents an improved result. However, changing the threshold level suddenly at a<br />

distance of 5 km causes a step change in the range measurements after retracking. In<br />

contrast, the varying threshold level cannot only generate improved SSH data, but<br />

also avoids this step change.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 154<br />

1.4<br />

1.2<br />

unretracked<br />

50% threshold level<br />

varying threshold level<br />

fixed counts<br />

1<br />

Mean Differences (m)<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

−0.2<br />

1 2 3 4 5 6 7 8 9 10<br />

Areas<br />

Figure 5.11 Mean differences between the SSH data after retracking and<br />

AUSGeoid98 geoid heights in ten areas 5-10 km from the Australian coastline.<br />

12<br />

10<br />

unretracked<br />

50% threshold level<br />

varying threshold level<br />

fixed counts<br />

8<br />

STD (m)<br />

6<br />

4<br />

2<br />

0<br />

1 2 3 4 5 6 7 8 9 10<br />

Areas<br />

Figure 5.12 Standard deviations of the mean differences in ten areas 5-10 km from<br />

the Australian coastline.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 155<br />

It should be pointed out that the majority of coastal altimeter waveforms exhibit an<br />

ocean-like shape with a single ramp (see Chapter 6). Only a small percentage of them<br />

show a significant peak or a combination of ocean and land returns. Consequently,<br />

the use of the 30% and 50% threshold levels over large areas of the coastal regions is<br />

more reasonable from these results. In the test, it has also been found that only 30%<br />

of threshold level sometimes might not be sufficiently low to intercept the soughtafter<br />

ocean returns from the waveforms with double ramps. The 20% retracking point<br />

has been found in this investigation to be the best threshold level for areas five, eight,<br />

and ten (cf. Figure 5.8).<br />

5. 8 An Assessment of Biases among Waveform Retracking Algorithms<br />

In this study, waveform retracking is carried out in coastal regions. Thus, the first<br />

concern in this Section is whether or not the retracked data in coasts matches the<br />

unretracked data in open oceans, and if not, to find the difference (or bias) between<br />

them. In fact, biases have been found in this investigation and in a previous study<br />

(Anzenhofer et al., 2000). In the results of Anzenhofer et al. (2000), a bias<br />

(~66.4±10.7 cm) has shown between the ERS-1 GDR unretracked range and the<br />

retracked range using the five-β-parameter function.<br />

In addition, there is no ‘universal’ retracking algorithm that can deal with all types of<br />

the waveform data (Berry et al., 1998; Martin et al., 1983; e.g., Berry, 2000).<br />

Therefore, a bias question is: if different retrackers are used to retrack individual<br />

waveform data, is there any bias exiting between among retrackers? The second<br />

concern is to estimate if there is a bias among different retracking algorithms, and to<br />

quantify and correct for it.<br />

5.8.1 Analysis Methods and Data<br />

Both the iterative least squares method (Section 5.6.5) and threshold retracking<br />

methods (Section 5.7.1) are used to assess the performance of different retracking<br />

algorithms in coastal regions through waveform retracking. These retrackers include<br />

the 50% and 30% threshold levels, the ocean model and five-β-parameter model. The<br />

five- β -parameter function with an exponential trailing edge is not used, because the


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 156<br />

results of Anzenhofer et al. (2000) indicate that the bias between the five- β -<br />

parameter functions with the linear and exponential trailing edges is insignificant<br />

(~1.3±10.8 cm). For convenience of presentation, hereafter the five- β -parameter<br />

model is called the five-parameter model.<br />

The data used are cycles 42 and 43 of ERS-2 20 Hz waveforms over an ocean area<br />

around Australia out to 350 km from the coastline. Contaminated waveforms in<br />

coasts (obtained in Chapter 4) are retracked first to investigate how retracking<br />

improves the range and thus the SSH, and to find biases. Then, waveforms in an<br />

ocean area of 50-350 km off the coastline are retracked to estimate biases between<br />

both algorithms and SSHs before and after retracking. Waveforms are retracked<br />

using each retracker to estimate the range correction and then the retracked SSH.<br />

Retracked SSHs are used to compute the SSH differences between different<br />

retrackers at each geographical location. The bias is finally estimated by fitting a<br />

cubic polynomial function to the SSH differences.<br />

Both before and after waveform retracking, it is likely that there is still a small<br />

percentage (~4% in this study) of SSH data that may present as outlier measurements.<br />

Therefore, the Z-score is employed as an outlier detection method when computing<br />

the bias<br />

where<br />

∆SSH i<br />

− ∆SSH<br />

Z<br />

i<br />

=<br />

(5.1)<br />

STD<br />

∆ SSH is the difference of the SSH data at the same geographical position<br />

(assumed being normal distribution), STD is the standard deviation of the mean<br />

∆ SSH , Z<br />

i<br />

is the Z-score and<br />

the variable<br />

Z<br />

i<br />

< 3 is selected. The value of<br />

Z<br />

i<br />

> 3 indicates that<br />

SSH<br />

i<br />

is not within 3 STD of the sample mean, and may be considered<br />

as an outlier and deleted.<br />

The above analysis method is similar to the collinear analysis technique, which is<br />

usually applied to repeated altimeter data to assess the mesoscale variability of the<br />

sea surface (e.g., Cheney and Marsh, 1983; Sandwell and McAdoo, 1990), average<br />

repeated SSHs to estimate a mean sea level and its variations (Nerem, 1995; Leben et<br />

al., 1990; e.g., Deng et al., 1997), or compute gravity anomalies (e.g., McAdoo,


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 157<br />

1990). Since the inter-comparison here is conducted between different retracking<br />

algorithms using the same data set, the analysis is also different from the collinear<br />

technique, which involves the time variation and the small separation of the same<br />

ground track (< 1 km) occurring between different repeat orbits. This makes it more<br />

effective to estimate the biases among the different retracking methods.<br />

5.8.2 Retracking Results and Analysis<br />

All contaminated ERS-2 waveform (20 Hz, cycles 42 and 43, from March to May<br />

1999) estimated in Chapter 4 at the Australian coast have been retracked using the<br />

fitting and threshold algorithms. Figure 5.13 shows two typical examples of the<br />

retracking results, plotting the SSH (20 Hz) profile versus along-track distances. In<br />

order to evaluate how the waveform is contaminated near the coast and improved by<br />

the retracking techniques, the geoid height has been bispline interpolated at the each<br />

location along the track using the AUSGeoid98 grid. The geoid height profiles are<br />

also plotted versus along track distances in Figure 5.13. The profiles at the top of<br />

Figure 5.13 represent a ground track 21085 approaching land from water, while those<br />

at the bottom of the Figure is another ground track 21364 leaving land to water.<br />

As stated in Sections 2.5.2 and 5.7.3, it is acknowledged that the geoid height does<br />

not equal the SSH measured by the altimeter. There is a dynamic SST between them,<br />

which can be of the order of ~1-2 m. This difference is evidenced in Figure 5.13.<br />

However, the absolute SST cannot be determined at this time due to the uncertainty<br />

in the geoid undulation (cf. Chelton et al., 2001). In this case, applying a SST<br />

correction may cause an additional error to the SSH, which might be larger than the<br />

errors of other corrections (Hwang, 2003). Therefore, the SST is not applied to the<br />

SSH data in Australian coastal regions (Chapters 5 and 6) when an absolute height<br />

comparison is conducted. The difference between the SSH and geoid height is not<br />

taken as an indicator for the purpose of comparison or analysis. Instead, the profile’s<br />

trends of the SSH and geoid or the along-track slopes will be considered as a more<br />

appropriate indicator to examine the waveform contamination and the improvement<br />

of the retracking procedure.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 158<br />

−18.00<br />

−20.00<br />

−22.00<br />

−24.00<br />

SSH (m)<br />

−26.00<br />

−28.00<br />

−30.00<br />

−32.00<br />

Geoid height<br />

Unretracked<br />

Ocean model<br />

Five−parameter model<br />

50% threshold level<br />

−34.00<br />

0.0 5.0 10.0 15.0 20.0 25.0<br />

Along−track distance from the coastline (km)<br />

−8.00<br />

−10.00<br />

SSH (m)<br />

−12.00<br />

−14.00<br />

−16.00<br />

Geoid height<br />

Unretracked<br />

Ocean model<br />

Five−parameter model<br />

50% threshold level<br />

−18.00<br />

0.0 5.0 10.0 15.0 20.0 25.0<br />

Along−track distance from the coastline (km)<br />

Figure 5.13 Geoid height and SSH profiles along ground tracks 21085 (top) and<br />

21364 (bottom) (ERS-2, cycle 42) before and after retracking using the ocean model,<br />

the five-parameter model and the 50% threshold level in coastal regions.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 159<br />

As can be seen from Figure 5.13, the typical near-land effects on unretracked SSH<br />

profiles begins at the along-track distance of ~20 km and ~17.5 km from the<br />

coastline for the track 21085 (top picture) and track 21364 (bottom picture),<br />

respectively. When approaching the coastline, unretracked SSHs decrease and then<br />

increase gradually (see also Section 4.6). The difference of the minimum SSH height<br />

and the normal SSH height can be large, up to ~ -5 m within ~10 km for both tracks.<br />

In contrast to the change of unretracked SSH profile, the geoid heights show a linear<br />

trend within the whole distance.<br />

By comparison of the geoid heights with unretracked SSHs, it is evident that<br />

unretracked SSHs are affected by waveform contamination in coastal regions.<br />

In addition, according to the shape of the unretracked SSH profiles, the along-track<br />

SSH and geoid gradients are computed and listed in Tables 5.3 and 5.4. The results<br />

of the track 21085 (Table 5.3) show that along-track geoid gradients computed from<br />

the AUSGeoid98 grid are small, from 9.2 ppm to 10.9 ppm when the track is<br />

0-7.5 km and 7.6-19.5 km from the coastline. Corresponding to the same distances,<br />

SSH gradients computed from unretracked SSHs are 975.6 ppm and -509.2 ppm,<br />

respectively. They are much larger than the maximum geoid gradient value of ~150<br />

ppm in Australia (Johnston and Featherstone, 1998; Friedlieb et al., 1997; cf.<br />

Featherstone et al., 2001). After 19.5 km from the coastline, geoid gradients<br />

computed using AUSGeoid98 geoid grid and SSH gradients computed using<br />

unretracked SSHs show small values of 12.5 ppm and -6.4 ppm, respectively. These<br />

results, thus, suggest as well that unretracked SSHs are in error within an along-track<br />

distance of 0-19.5 km from the coastline.<br />

Table 5.3 Along-track geoid and SSH gradients (track 21085)<br />

Gradients<br />

Names<br />

(ppm)<br />

ds = 0.0 - 7.5 km ds = 7.6 - 19.5 km ds = 19.8 - 26.0 km<br />

Geoid height 9.2 10.9 12.5<br />

Unretracked SSH 975.6 -509.2 -6.4<br />

Analysing the along-track SSH and geoid gradients listed in Table 5.4 for the track<br />

21364 can also find that unretracked SSHs are affected within a distance of 0-20 km


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 160<br />

from the coastline. After 20 km, geoid gradients computed from AUSGeoid98 geoid<br />

grid and SSH gradients computed from unretracked SSHs take small values of<br />

8.5 ppm and -21.4 ppm, respectively.<br />

Table 5.4 Along-track geoid and SSH gradients (track 21364)<br />

Names<br />

Gradients (ppm)<br />

ds = 0.0 - 4.7 km ds = 4.8 - 17.5 km ds = 17.5 - 23.5 km<br />

Geoid height 14.2 16.4 8.5<br />

Unretracked SSH 803.3 -363.8 -21.4<br />

It is also apparent from Figure 5.13 that the retracking techniques, irrespective of the<br />

method used, have extended the SSH profile several kilometres shoreward. For the<br />

track 21085 (top picture in Figure 5.13), retracking improves the SSHs over a<br />

distance of ~5-20 km from the coastline. A similar improvement of the SSH can be<br />

found from the track 21364 (bottom figure in Figure 5.13), where the SSH profile<br />

has been extended shoreward ~14 km along track. Although the analysis is only for<br />

two tracks, they are good representative of the ground tracks closer to the coastline.<br />

5.8.3 Biases among Retracking Algorithms<br />

Apart from the improvement in SSHs after retracking, it is evident that there is the<br />

bias among different retracking algorithms, in particular between the fitting<br />

algorithm and the threshold retracking (50% threshold level) algorithm (Figure 5.13).<br />

In order to investigate the biases among different algorithms and even between the<br />

SSH data before and after retracking, waveform data over a open ocean area<br />

(50-350 km from the coastline, -30°S


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 161<br />

A typical example of the SSH profiles before and after retracking is plotted as a<br />

function of the latitude in Figure 5.14. It is evident that the differences of the SSH<br />

are caused by the bias. The histograms of the differences ( ∆ SSH ) between the SSH<br />

data retracked by different methods are shown in Figures 5.15 and 5.16. From<br />

Figure 5.15, the differences of the ocean model and the five-parameter model are<br />

small, between -5 cm and +5 cm (see also Table 5.5 and 5.6) and nearly unidentified,<br />

though the differences do not exactly follow a normal distribution. The results from<br />

the differences between 50% threshold retracking and two fitting functions<br />

(Figure 5.16) show, however, an obvious bias.<br />

Figure 5.14 Retracked and unretracked SSH profiles over open oceans, showing the<br />

biases among them.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 162<br />

Figure 5.15 Histogram of the differences between the ocean model and fiveparameter<br />

model.<br />

Figure 5.16 Histograms of the differences between the 50% threshold retracking<br />

technique and the ocean model as well as the five-parameter model.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 163<br />

Using cycles 42 and 43 of ERS-2 waveform data, the biases have been estimated and<br />

the results are listed in Tables 5.5 and 5.6, respectively. The bias between two fitting<br />

models is small and almost statistically insignificant (Table 5.5). This is because they<br />

are both derived from the Brown model. However, it is significant to find the bias<br />

values of ~57 cm and ~56 cm between the 50% threshold algorithm and the fiveparameter<br />

model and the ocean model, respectively (Table 5.6). These values<br />

indicate that the range retracked by the fitting model is ~57 cm or ~56 cm longer<br />

than that retracked by 50% threshold retracking algorithm.<br />

Table 5.5 Biases between different retracking techniques from fitting a quadratic<br />

function to the ERS-2 SSH (cycle 42) differences over an ocean area (-30°S


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 164<br />

account of noise in the waveform (especially the trailing edge) and how the noise is<br />

affecting the leading edge from one waveform to the next, while the fitting algorithm<br />

is still affected by the noise in the trailing edge. The limited 3-4 sample bins of the<br />

leading edge are also restricted the precision of the estimates for both algorithms.<br />

Thus, the bias problem results from the noise effect in the fitting algorithm and too<br />

simplistic assumptions in the threshold retracking. The biases will be studied further<br />

in the next section.<br />

5.8.4 Biases between the SSH Data before and after Retracking<br />

From Figure 5.13, it is also apparent that there is a difference between the SSH data<br />

before and after retracking. Figures 5.17 and 5.18 show histograms of differences of<br />

the SSH data before and after retracking using different algorithms. It is evident that<br />

the differences of the SSHs before and after retracking appear to follow a normal<br />

distribution, but their means obviously are not equal to zero and indicate the biases.<br />

Figure 5.17 Histograms of the differences between unretracked SSH data and<br />

retracked SSH data using the ocean model and the five-parameter model.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 165<br />

Figure 5.18 Histogram of the differences between untracked SSH data and retracked<br />

SSH dada using the 50% threshold retracking technique.<br />

The biases have also been estimated and the results from cycles 42 and 43 are listed<br />

in Table 5.7 and Table 5.8, respectively. The differences computed from all<br />

retracking algorithms indicate that there are biases between retracked and<br />

unretracked the SSH data.<br />

From Table 5.7 and Table 5.8, a bias of ~30 cm exists between retracked SSH data<br />

(by 50% threshold retracking) and the unretracked SSH data, indicating the measured<br />

range is too long. The bias of ~27 cm between the SSH data retracked by fitting<br />

approaches and the unretracked SSH data suggests the measured range is too short.<br />

Theoretically, it is not necessary to retrack waveforms over open oceans as the onboard<br />

tracking algorithm is based on the Brown model and the assumptions of the<br />

homogeneous statistics of the model are generally satisfied over ocean surfaces.<br />

However, a bias still found between the SSH data retracked by the ocean model and<br />

the unretracked SSH data in this study. Therefore, the subsequent analysis here aims<br />

at understanding the reasons that cause the biases.


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 166<br />

Table 5.7 Biases between the ERS-2 (cycle 42) SSHs before and after retracking<br />

over an ocean area (-30°S


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 167<br />

samples in the range window. Since a sample or gate corresponds to a distance of<br />

~0.4542 m (ERS-1/2), it should be possible that the error is caused by the retracking<br />

procedure itself.<br />

From the above analysis, if one is interested in using data closer to the coastline and<br />

retracking the coastal waveform data, the bias between the SSH data before and after<br />

retracking should be removed when combining the retracked and unretracked data.<br />

Three approaches can be used:<br />

• Accept the already-applied instrument corrections and use the range provided<br />

by the GDR products over open oceans. The corrected range in coastal regions<br />

is obtained by applying the range correction estimated from retracking<br />

procedure to the measured range from the waveform products. Then, these two<br />

types of range measurements can be used together without the bias between<br />

them.<br />

• Use the range measurements from waveform products, retrack all waveforms<br />

(both over open ocean and coastal regions), and then apply the range<br />

corrections obtained from the retracking procedure to all measurements.<br />

• Retrack only the waveform data in coastal regions, but estimate an absolute<br />

bias between the altimeter-derived SSH (without retracking) and independent<br />

sea surface measured at the tide gauge. The bias, then, is used to correct the<br />

SSH over open oceans.<br />

5. 9 Summary<br />

A coastal retracking system has been developed. Central to the system is the use of<br />

two retracking techniques to derive an improved geodetic parameter, namely SSH,<br />

which include both the iterative least squares fitting algorithm and the threshold<br />

retracking algorithm. Retrackers adopted in the system are the ocean fitting model,<br />

five-parameter fitting models, nine-parameter fitting models, and the varying<br />

threshold levels, in which five- and nine-parameter models are used for the<br />

waveforms with both the linear and exponential trailing edges.<br />

By quantitative comparison of the ocean model and the five-parameter model, it has<br />

been found that the slope of the trailing edge modelled by the five-parameter model<br />

has a larger span of variation than the slope modelled by the ocean model. This


Chapter 5. COASTAL WAVEFORM RETRACKING SYSTEM 168<br />

advantage, in turn, gives the five-parameter model greater capability of fitting nonocean-like<br />

waveforms in coasts than the ocean model.<br />

A waveform classification procedure is given in this Chapter, which enables the<br />

waveform to be sorted and retracked by an appropriate retracker. To overcome the<br />

‘noise’ in waveforms caused by fading noise and sources, a weight iterative scheme<br />

has been developed.<br />

Threshold retracking can be an improved method in coastal regions, if an appropriate<br />

threshold level is selected for the retracking. Comparison of retracked SSH data with<br />

available ground truth of the AUSGeoid98 geoid heights confirms that the 50%<br />

threshold level is the best level for open-ocean waveform, but not an adequate level<br />

for the contaminated ocean waveforms in coastal regions. Instead, a varying<br />

threshold level, which is the 30% threshold level for contaminated waveforms and<br />

50% threshold level for uncontaminated waveforms, can make a good agreement<br />

between the retracked SSH data and the geoid heights.<br />

Biases exist both between retracked SSH data from different retracking algorithms<br />

and SSH data before and after retracking. Biases have been analysed and estimated<br />

by retracking ocean waveforms in this Chapter using two cycles of ERS-2 waveform<br />

data readily available at the time of this study. This finding suggests that biases must<br />

be estimated before retracking and applied to the data to obtain consistent results<br />

after retracking.


Chapter 6. WAVEFORM RETRACKING APPLICATION 169<br />

6. WAVEFORM RETRACKING APPLICATION TO THE AUSTRALIAN<br />

COAST<br />

6. 1 Introduction<br />

The principal objective of this Chapter is to determine to what extent the SSH data<br />

can be improved by the retracking system designed and developed in Chapter 5 in<br />

coastal regions. The quantitative distance estimation and detailed analysis of<br />

waveform contamination for the Australian coastal regions has been addressed in<br />

Chapter 4. As part of the development and implementation of the retracking system,<br />

it is important to apply the system to the coastal regions and to assess the quality of<br />

the retracked altimeter-derived SSH data. Therefore, the coastal retracking system<br />

will be tested over a small part of the Australian coast (0-30 km from the coastline,<br />

Figure 4.3) using two cycles (42 and 43, March to May 1999) of ERS-2 20 Hz<br />

waveform data. To obtain a general idea of the impact of retracking on the altimeterderived<br />

SSH, the quality checks of the retracking results consist of the comparisons<br />

with external data source and validation of internal results. The external independent<br />

reference used is the AUSGeoid98 geoid model. The results of retracking and<br />

comparisons with unretracked and external control data will be discussed in this<br />

Chapter.<br />

Following the steps of the retracking block diagram in Figure 5.1, results of<br />

waveform classification and waveform retracking will be presented and discussed in<br />

Sections 6.2 and 6.3, respectively. The internal validation of the quality of the<br />

retracked altimeter data will be conducted by validation of the single track SSH data<br />

(Section 6.4), comparisons between SSH data before and after retracking (Section<br />

6.5) and a collinear analysis (Section 6.6). SSH data before and after retracking will<br />

be compared with the geoid heights interpolated from the AUSGeoid98 model along<br />

ground tracks for the external comparison (Sections 6.7).<br />

In addition, validating the retracked data will involve using ~2.5 years of ERS-1<br />

20 Hz waveform data from different phases in the Taiwan Strait. Retracked SSH data<br />

will be used to estimate the gravity anomalies. The results will then be compared


Chapter 6. WAVEFORM RETRACKING APPLICATION 170<br />

with the external ground control of the ship-track gravity anomalies. The results will<br />

be given in Chapter 7.<br />

6. 2 Waveform Classification in Australian Coastal Regions<br />

As shown in Chapter 4, most of contaminated waveforms in coastal regions present<br />

an ocean-like single ramp. However, the returns also present some other non-oceanlike<br />

shapes, such as a superposition of the energy reflected from each scattering<br />

surface within the satellite footprint, which represent each distinct elevation when<br />

approaching the shoreline. Therefore, sorting data based on waveform shapes is<br />

essential before retracking them. In this Section, the contaminated waveform data<br />

estimated in Chapter 4 will be classified by using the methods presented in Section<br />

5.3.<br />

6.2.1 Typical Waveform Shapes<br />

Contaminated waveforms shown in Chapter 4 around the Australian coast have been<br />

categorised. A detailed example of waveform categorisation given here is from two<br />

ascending (21636 and 21679) and one descending (21815) ground tracks in a<br />

15º×15º area (105°≤λ


Chapter 6. WAVEFORM RETRACKING APPLICATION 171<br />

108˚<br />

110˚<br />

112˚<br />

114˚<br />

116˚<br />

118˚<br />

120˚<br />

-16˚ -16˚<br />

-18˚ 21815<br />

-18˚<br />

21679<br />

-20˚ -20˚<br />

-22˚ -22˚<br />

Indian<br />

Ocean<br />

21636<br />

-24˚ -24˚<br />

-26˚ -26˚<br />

-28˚ -28˚<br />

-30˚ -30˚<br />

108˚<br />

110˚<br />

112˚<br />

114˚<br />

116˚<br />

118˚<br />

120˚<br />

Figure 6.1 Part of ERS-2 ground tracks off the northwest Australian coast, showing<br />

the orbit number of each ground track. The highlighted diamonds along tracks will<br />

be shown in detail waveform shapes later in this Section.<br />

Figure 6.3 also shows waveforms from the pass 21815 but crossing an island<br />

(cf. Figure 6.1). The blank area (~ -20.9°


Chapter 6. WAVEFORM RETRACKING APPLICATION 172<br />

reflected from both land and standing water, and finally change to the typical ocean<br />

shapes at around -22.1° latitude. The waveform shifting in the range window is<br />

obviously observed in Figure 6.4, which is probably due to the operation manner of<br />

the altimeter discussed in Section 4.6.3, and thus causing a longer contaminated<br />

distance for this track.<br />

Figure 6.5 shows waveforms from an ascending ground track 21679 (cf. Figure 6.1)<br />

passing a coastal area of Shark Bay (113°≤λ


Chapter 6. WAVEFORM RETRACKING APPLICATION 173<br />

1000 2000<br />

Counts<br />

-18.60 -18.55 -18.50 -18.45 -18.40 -18.35 -18.30 -18.25<br />

Latitude (degrees)<br />

16 32 48 64<br />

Bins<br />

Figure 6.2 Typical ocean waveforms from pass 21815 (cf. Figure 6.1).<br />

Counts<br />

2500 5000<br />

-21.0 -20.9 -20.8 -20.7 -20.6 -20.5 -20.4 -20.3 -20.2<br />

Latitude (degrees)<br />

16 32 48 64<br />

Bins<br />

Figure 6.3 Waveforms recorded when the ground track 21815 is approaching an<br />

island at latitude ~ -20.5º and leaving the island to water at ~ -20.9º. The data gap is<br />

due to the island (cf. Figure 6.1).


Chapter 6. WAVEFORM RETRACKING APPLICATION 174<br />

Counts<br />

2000 4000 6000 8000 10000<br />

-22.35 -22.30 -22.25 -22.20 -22.15 -22.10<br />

Latitude (degrees)<br />

16 32 48 64<br />

Bins<br />

Figure 6.4 Waveforms leaving land to water (pass 21636 in Figure 6.1). The highpeaked<br />

specular responses are probably due to off-nadir brighter calm water. The<br />

waveform shifting (~ -22.25°


Chapter 6. WAVEFORM RETRACKING APPLICATION 175<br />

Counts<br />

5000 10000 15000<br />

64<br />

48<br />

32<br />

Bins<br />

16<br />

-26.0 -25.9 -25.8 -25.7 -25.6 -25.5 -25.4 -25.3 -25.2 -25.1<br />

Latitude (degrees)<br />

Figure 6.5 Typical ERS-2 radar altimeter contaminated waveforms (pass 21679)<br />

across water, land, and calm/still sea surfaces, showing high-peaked waveform<br />

shapes reflected from the calm water surfaces, mixed shapes of land and water, and<br />

ocean-like shapes.


Chapter 6. WAVEFORM RETRACKING APPLICATION 176<br />

6.2.2 Results of Waveform Categories<br />

The detailed categories of contaminated waveforms in ten Australian coastal regions<br />

(see Figure 5.8) have been conducted and the general statistics of the waveform<br />

categories for these ten coastal areas are listed in Table 6.1. The following<br />

conclusions can be drawn from Table 6.1.<br />

Table 6.1 Statistics of ERS-2 (cycle 42) contaminated waveform shapes in ten<br />

Australian coastal areas (cf. Figure 5.8)<br />

Waveform<br />

categories<br />

% of<br />

waveforms<br />

Mean distance<br />

from coastline<br />

(km)<br />

Mean width of<br />

waveforms<br />

(bins)<br />

Mean peak of<br />

waveforms<br />

(counts)<br />

Ocean * 99.97 168.04 25.8 1304.1<br />

Ocean-like 80.19 10.50 24.9 1380.3<br />

Single pre-peaked 4.64 2.31 10.7 2785.1<br />

Single mid-peaked 0.54 2.54 9.8 2615.0<br />

Single post-peaked 1.35 3.12 11.8 3027.0<br />

Double pre-peaked 0.55 3.08 14.8 2280.9<br />

Double post-peaked 0.73 2.15 16.1 2271.9<br />

Multi-peaked 0.18 2.30 13.0 2117.5<br />

Sharp/High-peaked 4.35 1.73 3.3 8439.5<br />

Unusable 7.51 6.03 34.7 1234.6<br />

* Ocean waveforms over open oceans 50-350 km from the coastline, with 50%<br />

threshold retracking point estimates between bins/gates 31-33.<br />

1. Most contaminated waveforms (80.19%) have ocean-like shapes. The mean<br />

width is similar to that of ocean waveforms, whereas the mean peak value is<br />

slightly larger than ocean waveforms.<br />

2. Of other shapes, the highest peak of the power is the high-peaked waveform<br />

and the next is the single post-peaked waveform, but mean peak values of all<br />

other shapes are larger than the mean peak of the ocean-like waveform.<br />

3. The high-peaked waveform has the narrowest mean waveform width (3.3 bins).<br />

The second narrowest width (9.8 bins) belongs to the single mid-peaked<br />

waveform. Other non-ocean-like waveform widths are still much smaller than<br />

the ocean-like waveform.


Chapter 6. WAVEFORM RETRACKING APPLICATION 177<br />

4. Although ~80% of contaminated waveforms in coastal regions take an oceanlike<br />

shape, other shapes of waveforms are found much closer to the coastline<br />

from this data set (1.73 - 3.12 km in average), in particular the high-peaked<br />

waveform (1.73 km from the coastline in average). This fact indicates further<br />

that using only the ocean model or five-parameter model cannot re-process all<br />

waveforms in coastal regions. Other retrackers are more significant in such<br />

areas to remove errors caused by waveform contamination.<br />

In addition, from a waveform categorisation in this study, high-peaked waveforms<br />

have been found to take a larger percentage of contaminated waveforms (from 5.4%<br />

to 6.9%) in north-west Australian coastal areas 1 and 2 because of the rugged<br />

coastline, complex sea states and inland/standing water.<br />

6. 3 Waveform Retracking Results and Discussion<br />

Given the above observations, 20 Hz waveforms from cycles 42 and 43 offshore<br />

Australia are re-processed using the coastal retracking system developed in Chapter 5<br />

out to a distance of 30 km. Choosing 0-30 km is based on the observation in Chapter<br />

4, where the maximum contaminated distance can occur out to ~22 km from the<br />

coastline.<br />

6.3.1 Data Editing<br />

The first step of retracking is data editing to check whether or not the return presents<br />

a leading edge (i.e., ramp) in the range window. This is conducted through the simple<br />

heuristic assessment method used by Scott et al. (1994). The maximum 8.2% of noramp<br />

(i.e., unusable) waveforms are filtered out from this step (see Table 6.2 and<br />

Table 6.3). The second step is to classify each return waveform so that it can be<br />

retracked by an appropriate retracker.<br />

6.3.2 Waveform Retracking Using Different Algorithms<br />

Next, a specific retracker based upon the classification is applied to each waveform,<br />

which is either a fitting or threshold retracking algorithm. When using the iterative<br />

procedure, an adequate fitting function must be selected and the procedure is


Chapter 6. WAVEFORM RETRACKING APPLICATION 178<br />

repeated until a pre-defined existing criterion is satisfied (see Section 5.5.5). Since a<br />

linear solution to a non-linear problem is used, sometimes the fitting algorithm does<br />

not work (Zwally, 1996; Davis, 1995). Therefore, the threshold-retracking algorithm<br />

is used to replace the iterative least squares fit to compute the retracking correction if<br />

and when the fitting procedure fails. The threshold retracking technique thus works<br />

for both the high-peaked waveforms and those fail to the fitting algorithm. The<br />

system will select automatically the threshold level, which is 30% or 50% of the<br />

waveform amplitude, for each waveform depending on its shape. As stated in Section<br />

5.6.3, the 50% threshold level is used to retrack the ocean-like waveforms, while<br />

other shapes of waveforms will be retracked by the 30% threshold level.<br />

The percentages of the waveforms retracked by fitting and threshold algorithms are<br />

listed in Tables 6.2 and 6.3. It can be seen that 77.7% - 94.9% (cycle 42) and 83.5% -<br />

96.2% (cycle 43) of waveforms in this area can be retracked by the iterative least<br />

squares fitting procedure. The rest of the waveforms (the last column) still require<br />

threshold retracking. They are 4.8% - 18.1% of waveforms for cycle 42 and 3.8% -<br />

15.6% of waveforms for cycle 43, respectively. The percentages of using different<br />

retrackers vary from area to area depending on the reflecting surface topography.<br />

Table 6.2 Waveform data status and percentage of the waveforms retracked using<br />

fitting and threshold retracking algorithms in Australian coastal regions (cycle 42)<br />

Area<br />

No. of<br />

Unusual<br />

Fit<br />

Threshold<br />

number<br />

Areas<br />

data<br />

data<br />

%<br />

%<br />

%<br />

1 120°≤λ


Chapter 6. WAVEFORM RETRACKING APPLICATION 179<br />

Table 6.3 Waveform data status and percentage of the waveforms retracked using<br />

fitting and threshold retracking algorithms in Australian coastal regions (cycle 43)<br />

Area<br />

No. of<br />

Unusual<br />

Fit<br />

Threshold<br />

number<br />

Areas<br />

data<br />

data<br />

%<br />

%<br />

%<br />

1 120°≤λ


Chapter 6. WAVEFORM RETRACKING APPLICATION 180<br />

Table 6.4 Percentages of ERS-2 waveforms fitted using different functions in<br />

Australian coastal regions (cycle 42).<br />

Area<br />

No. of fit<br />

Ocean<br />

E-five-<br />

Nine-<br />

number<br />

Areas<br />

data<br />

model<br />

parameter a<br />

parameter b<br />

%<br />

%<br />

%<br />

1 120°≤λ


Chapter 6. WAVEFORM RETRACKING APPLICATION 181<br />

Figure 6.6 Typical coastal waveforms fitted using the following models:<br />

(a): the five-parameter model with an exponential trailing edge,<br />

(b) and (c): the ocean model,<br />

(d) and (f): the nine-parameter model with linear trailing edge,<br />

(e): the nine-parameter model with an exponential trailing edge.<br />

waveforms are retracked by the fitting algorithm, while only 2% of waveforms need<br />

to be threshold retracked. The percentage of fitting algorithm increases with<br />

increasing distance from the coastline, confirming that the waveform shapes agree<br />

well with the typical ocean waveform after 5 km offshore. This was also seen in<br />

Chapter 4.<br />

Figure 6.6 shows six typical coastal waveforms and the functional fits to them. It also<br />

shows that different fitting functions work well to deal with different shapes of<br />

waveforms in coastal regions. As stated, most of waveforms are single peaked (or<br />

ramp) pulses. The ocean model fits these waveforms very well as shown in Figure<br />

6.6(b) and (c). Some waveforms near land show higher power and double ramps,<br />

which need to be fitted by the five-parameter function with an exponential trailing<br />

edge as shown in Figure 6.6(a), or the nine-parameter function with either a linear


Chapter 6. WAVEFORM RETRACKING APPLICATION 182<br />

trailing edge (cf. Section 3.5) in Figure 6.6(d) and (e) or an exponential trailing edge<br />

(Figure 6.6(e), cf. Section 3.5).<br />

6. 4 Analysis of the Single-Track Retracked Results<br />

After retracking, two types of corrections from atmospheric propagation delays and<br />

geophysical models supplied with the waveform data by NRSC (1995) are applied to<br />

both retracked and unretracked range measurements to obtain the SSH. Since the EM<br />

bias correction caused by ocean surface waves is not supplied with the waveform<br />

data products used in this study, it is not applied to unretracked range measurements.<br />

However, the retracking procedure removes the need for this correction (Brenner et<br />

al., 1993; Hayne, 2002). The geophysical and environmental corrections applied are<br />

dry and wet tropospheric range corrections, ionospheric correction, and tide<br />

correction. The tide correction is the sum of the ocean tide from the Schwiderski<br />

model and the ocean loading tide from the Ray-Sanchez model (NRSC, 1995). The<br />

Doppler range correction is not applied to the range measurement, because most<br />

values were found to exceed the error criteria of ±55 cm given by NRSC (1995) as<br />

stated in Section 2.4.2.<br />

As mentioned in Section 5.8.3, the relative biases among different retracking<br />

algorithms are determined by retracking the waveforms over open oceans. They are<br />

then applied to the range measurements retracked by the threshold retracking<br />

algorithm. Finally, the altimeter-derived SSH is based on a reference of the ocean<br />

model. This aims to remove the biases caused by reprocessing the waveform data<br />

using different algorithms. As the determined biases depend strongly on the<br />

algorithm, they can also be determined by other methods. For example, the method<br />

used by Dong et al. (2002) may be employed for this purpose, which validates the<br />

absolute bias between the altimeter-derived sea level and that derived by the tide<br />

gauge data using the local geoid model, SST model, and data measured at the tide<br />

gauge equipped with Global Positioning System (GPS) receivers. By using a similar<br />

method, it is possible to estimate the absolute biases between different retracking<br />

algorithms, but this remains an open issue for future work because of data<br />

availability limitations at the time of this study.


Chapter 6. WAVEFORM RETRACKING APPLICATION 183<br />

To compare the results from the retracking system with those retracked by a single<br />

retracker, two ground tracks (21085 and 21364) are chosen, which are the same<br />

tracks as these presented in Figure 5.13 (Section 5.8.2). The SSH (20 Hz) and geoid<br />

height profiles (AUSGeoid98) are also plotted via the along-track distance from the<br />

coastline in Figure 6.7. As stated in Sections 2.5.2 and 5.8.2, this aims to compare<br />

the profile trends of the SSH and geoid height for investigating the contamination<br />

caused by coastal sea states.<br />

To compare results with those in Figure 5.13 (Section 5.8.2), the pictures have the<br />

same X- and Y-scales as those in Figure 5.13. It is evident when comparing Figure<br />

6.7 with Figure 5.13 that erroneous SSHs caused by the fitting algorithm when<br />

approaching the coastline (0-5 km) in Figure 5.13 have been removed in Figure 6.7.<br />

The bias between fitting and threshold retracking algorithms is also removed after<br />

applying the bias correction determined in Section 5.8 (cf. Figures 6.7 and 5.13).<br />

Because of use of the retracking system, more SSH data can be derived from<br />

corrected range measurements than the SSH data from a single fitting retracker, thus<br />

keeping the useful high-frequency singles in coastal regions. The numbers of SSH<br />

data obtained from the system and a single retracker (the ocean model) are N system<br />

retracking = 78 and N single retracker = 59 for track 21085 and N system retracking = 73 and N single<br />

retracker = 63 for track 21364.<br />

As can be seen from Figures 5.13 and 6.7, both single retracker and the retracking<br />

system can improve the SSH profile shoreward several kilometres. More importantly,<br />

the retracking system comprising several retrackers can improve further the SSHs<br />

shoreward ~2.5 km from ~5 km to ~2.5 km for both tracks. However, it is important<br />

to note that the SSH profiles cannot be closed completely to the coastline because of<br />

land-dominated returns discussed in Chapter 4. In the case of tracks 21085 and<br />

21363, there are still an along-track distance of ~2.5 km that cannot be recorded by<br />

waveform retracking.


Chapter 6. WAVEFORM RETRACKING APPLICATION 184<br />

−18.00<br />

−20.00<br />

−22.00<br />

−24.00<br />

SSH (m)<br />

−26.00<br />

−28.00<br />

−30.00<br />

−32.00<br />

Geoid height<br />

Unretracked<br />

Retracked<br />

−34.00<br />

0.0 5.0 10.0 15.0 20.0 25.0<br />

Along−track distance from the coastline (km)<br />

−8.00<br />

−10.00<br />

SSH (m)<br />

−12.00<br />

−14.00<br />

−16.00<br />

Geoid height<br />

Unretracked<br />

Retracked<br />

−18.00<br />

0.0 5.0 10.0 15.0 20.0 25.0<br />

Along−track distance from the coastline (km)<br />

Figure 6.7 Geoid height and SSH profiles (ERS-2, cycle 42) along ground track<br />

21085 (top) and 21364 (bottom) before and after retracking using the coastal<br />

retracking system, showing the same profiles as those in Figure 5.13.


Chapter 6. WAVEFORM RETRACKING APPLICATION 185<br />

Comparing the SSH profiles before and after retracking in Figure 6.7, the SSH<br />

profiles after retracking show a good agreement with the geoid-height profile from<br />

~2.5 km to ~25 km to the coastline. However, it is also apparent that the retracking<br />

has added some measurement noise. Compared with unretracked SSHs, the retracked<br />

SSH profile becomes nosier. The noise in the retracked SSH profile might be caused<br />

by, firstly, the shorter wavelength signal, which suggests that on-board tracking<br />

algorithm cannot follow well temporal variations of the scattering surface; secondly,<br />

unmodelled non-linear parameters in this study (e.g., the skewness). Thus, a low-pass<br />

filtering procedure is still essential when using 20 Hz SSH data after retracking for<br />

the geophysical application in coastal regions. The search radius of the filter depends<br />

on the applications, say 5-7 km for the recovery of gravity anomalies.<br />

6. 5 Analysis of the SSH Data Before and After Retracking<br />

The comparison is performed by calculating the difference dh between the retracked<br />

SSH h r and unretracked SSH h g at the same geographic location along each ground<br />

track. This difference (dh=h r − h g ) represents the sum of the bias of SSH data before<br />

and after retracking, and tracking error that can be removed by waveform retracking.<br />

The mean difference and standard deviation (STD) are calculated at six<br />

5 km wide distance bands. Choosing different distance bands allows one to see how<br />

close to the coastline the altimeter-observed ranges are improved by the retracking<br />

procedure. Results are listed in Tables 6.6 and 6.7 for cycles 42 and 43, respectively.<br />

It should be noted that the bias between SSH before and after retracking (described<br />

in Section 5.8.4) is not applied to the individual SSH data. Instead, the biases can be<br />

applied to the mean difference between the SSHs before and after retracking in<br />

Tables 6.6 and 6.7 to analyse the error removed by the retracking procedure.<br />

Therefore, results in Tables 6.6 and 6.7 contain the biases, which are -27.8±12.4 cm<br />

(cycle 42) and -27.5±10.8 cm (cycle 43) (ibid.). It is recalled here that the bias could<br />

be estimated and removed by using the methods mentioned in Section 5.8.<br />

Tables 6.6 and 6.7 show that STD of the SSH differences decreases with increasing<br />

the offshore distance. The STD decreases to less than 1 m for distances greater than<br />

15 km from the coastline. Because the comparison is conducted at the same along-


Chapter 6. WAVEFORM RETRACKING APPLICATION 186<br />

track point and same time, the satellite orbit error and any other random errors can be<br />

removed by taking the difference of the retracked and unretracked SSH. Therefore,<br />

the large values of the STD indicate, on one hand, that the significant error occurs<br />

due to waveform contamination near the coastline. On the other hand, the retracking<br />

algorithm itself introduces the additional error in the SSH data near land, because the<br />

ocean returns are difficult to recover from land-dominated waveforms.<br />

Table 6.6 Descriptive statistics of ERS-2 20 Hz SSH differences before and after<br />

retracking (dh = h r – h g , cycle 42).<br />

Distance<br />

to<br />

coastline<br />

(km)<br />

Number<br />

of<br />

SSH<br />

Mean<br />

difference<br />

(m)<br />

Max<br />

difference<br />

of SSH<br />

(m)<br />

Min.<br />

difference<br />

of SSH<br />

(m)<br />

STD<br />

±(m)<br />

0-5 5074 -0.79 9.25 -9.99 2.35<br />

5-10 6517 -0.20 9.43 -9.82 1.31<br />

10-15 7416 -0.24 6.62 -8.15 0.99<br />

15-20 7313 -0.25 7.43 -6.89 0.71<br />

20-25 7728 -0.30 7.71 -8.38 0.57<br />

25-30 7496 -0.28 5.69 -9.13 0.57<br />

Table 6.7 Descriptive statistics of ERS-2 20 Hz SSH differences before and after<br />

retracking (dh = h r – h g , cycle 43).<br />

Distance<br />

to<br />

coastline<br />

(km)<br />

Number<br />

of<br />

SSH<br />

Mean<br />

difference<br />

(m)<br />

Max<br />

difference<br />

of SSH<br />

(m)<br />

Min.<br />

difference<br />

of SSH<br />

(m)<br />

STD<br />

±(m)<br />

0-5 6218 -0.79 8.41 -9.99 2.43<br />

5-10 8924 -0.14 9.89 -9.82 1.39<br />

10-15 9820 -0.31 6.63 -9.47 1.02<br />

15-20 9850 -0.30 8.11 -9.91 0.84<br />

20-25 10140 -0.31 4.71 -5.70 0.65<br />

25-30 9660 -0.32 7.05 -6.31 0.56<br />

If considering only the amount of the bias, the important amount of error removed by<br />

waveform retracking is -51.2 cm and 7.8 cm for cycle 42 in distance bands of 0-5 km<br />

and 5-10 km, respectively, while the results do not show a significant amount of


Chapter 6. WAVEFORM RETRACKING APPLICATION 187<br />

error in the rest of the four distance bands. Similar results can be found for cycle 43,<br />

for which the amount of error in distance bands of 0-5 km and 5-10 km is -51.5 cm<br />

and 13.5 cm, respectively. Analysis of the results suggests that waveform retracking<br />

can effectively remove the altimeter tracking error in the vicinity of coastal regions.<br />

When the satellite ground track goes further to the open ocean, the tracking error<br />

becomes less significant. In the case of this study, the amount of tracking error is<br />

~-3 cm. This agrees well with the result (-2.37 cm) of Brenner et al. (1993).<br />

6. 6 Collinear Analysis<br />

From the analysis in Sections 6.4 and 6.5, it is evident that retracking can remove<br />

errors caused by altimeter on-board tracking algorithm due to coastal sea states and<br />

land topography. In this Section, the collinear differences, which are computed using<br />

altimeter-derived SSH profiles obtained along identical ground tracks, are employed<br />

to estimate how the retracking procedure improves the altimeter data in coastal<br />

regions. The differences are also used to qualify altimeter data before and after<br />

retracking.<br />

6.6.1 Methodology<br />

Over open oceans, ERS-2 SSH profiles from different cycles contain primarily geoid<br />

and SST information, as well as the random measurement error, which includes<br />

errors introduced by satellite orbit error, altimeter range measurement error, and<br />

temporal variability of the ocean surface, such as tides (Cheney et al., 1983). In<br />

coastal regions, however, waveform contamination causes an additional error. The<br />

relative collinear offset between the profiles is<br />

dH<br />

= H 2<br />

− H1<br />

+ ε , where the SSH<br />

measurements at the same point, H<br />

1<br />

and H<br />

2<br />

, occur at different times of cycles t 1<br />

and t 2<br />

, and ε is a random measurement error.<br />

The collinear differences remove the long-wavelength geoid, the quasi-timeindependent<br />

SST and variations of the ocean surface, representing thus the amount of<br />

orbit error, altimeter range measurement error, and the temporal variations of the<br />

surface over open oceans. As stated, the random error includes the tracking error<br />

caused by waveform contamination for unretracked data in coastal regions. After


Chapter 6. WAVEFORM RETRACKING APPLICATION 188<br />

removing the satellite orbit error, SSH data from repeat cycles can also be used to<br />

determine the mean sea surface (e.g., Deng et al., 1997; Nerem, 1995; Wang and<br />

Rapp, 1991). Thus, the STD of the difference dH here is used as a measure of the<br />

quality of a retracking algorithm’s range estimate. If errors caused by waveform<br />

contamination can be removed by retracking algorithms, this will result in smaller<br />

STDs. By comparing the STD of the collinear difference dH before and after<br />

retracking, it can be determined whether or not the retracked SSHs are improved by<br />

the retracking system.<br />

The requirement of the collinear analysis is that data from different cycles must be<br />

reduced to a reference orbit, so that the comparison can be conducted on the same<br />

basis (Cheney et al., 1983; Wang and Rapp, 1991). In reality, the satellite ground<br />

tracks do not repeat exactly due to orbital variations caused by gravitational<br />

influences and drag effects. The cross-track deviation is ~1 km for most satellites (cf.<br />

Brenner et al., 1990; Wang et al., 1991). Therefore, 1-second normal points are<br />

usually generated (ibid.), and the geoid gradient correction is applied to data (Wang<br />

and Rapp, 1991) for the reduction of the observed data to the reference orbit to<br />

obtain exactly ‘collinear’ data. The cross-track geoid gradient correction is applied to<br />

correct the effects of the geoid slope caused by non-repeated tracks (ibid.).<br />

6.6.2 Discussion of Results<br />

In this study, 20 Hz SSH data are retained to see possible high-frequency signals in<br />

coastal regions. Since cycle 43 contains more orbits and thus more complete data<br />

coverage, it is chosen as a reference. SSH data from cycle 42 are then reduced to<br />

relevant points along ground tracks of cycle 43 using the method developed to<br />

compute the collinear difference between SSHs from different cycles by Deng et al.<br />

(1996). The cross-track geoid gradient correction is not applied to the data here,<br />

because data came from two adjacent cycles and the effects of the geoid slope can be<br />

neglected.<br />

The descriptive statistics of the collinear SSH differences (20 Hz) computed from<br />

cycles 42 and 43 have been calculated both before and after retracking. Results have<br />

been plotted via six 5-km-wide distance bands in Figures 6.8 and 6.9. There are high<br />

values of STD for dH using both unretracked and retracked data. The reasons may


Chapter 6. WAVEFORM RETRACKING APPLICATION 189<br />

be the satellite orbit error and temporal variability of the ocean surface. As can be<br />

seen from Figure 6.9, the STDs of dH after retracking are smaller than STDs before<br />

retracking in all distance bands, in particular 0-15 km from the coastline. It is clear<br />

from this comparison that retracking improves the precision of altimeter range<br />

estimates in all distance bands to a different extent, in particular in two distance<br />

bands of 0-5 km and 10-15 km. After 15 km from the coastline and in 5-10 km<br />

distance, the STDs after retracking are still slightly smaller than STD values before<br />

retracking, indicating improvement from waveform retracking as well.<br />

Before Retracking<br />

After retracking<br />

Mean differences (m)<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

-0.2<br />

0-5 5-10 10-15 15-20 20-25 25-30<br />

Distances from the coastline (km)<br />

Figure 6.8 Mean differences of the collinear SSH data of cycles 42 and 43 before and<br />

after retracking in six 5 km wide distance bands.<br />

Before retracking<br />

After retracking<br />

STD (m)<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0-5 5-10 10-15 15-20 20-25 25-30<br />

Distances from the coastline (km)<br />

Figure 6.9 STD (positive values) of the collinear 20 Hz SSH differences of cycles 42<br />

and 43 before and after retracking in six 5 km wide distance bands, showing<br />

improvement from SSH data after retracking.


Chapter 6. WAVEFORM RETRACKING APPLICATION 190<br />

6. 7 Evaluation of SSH Data Using Geoid Heights<br />

6.7.1 Comparison with the AUSGeoid98 Geoid Model<br />

The collinear analysis of SSH data along repeating tracks can determine the internal<br />

consistency (i.e., precision) of altimeter measurements. In order to assess the<br />

accuracy of the retracked data and to determine how the altimeter retracked SSH<br />

profiles represent the true sea surface, an independently surveyed reference surface,<br />

referenced to the same ellipsoid, is required. The surface used here is the<br />

AUSGeoid98 2′×2′ geoid grid (Featherstone et al., 2001). This geoid height model,<br />

actually, is not an absolutely independent ground truth, because it uses altimeterderived<br />

free-air gravity anomalies offshore (ibid.). However, as stated in Section<br />

2.5.1, altimeter-derived gravity and ship track gravity data were used for<br />

computation of the geoid in a ‘draping’ method such that the dense ship-track gravity<br />

contribute significantly to the grid signal near the Australian coast, especially<br />

0-30 km from the coastline, while the altimeter-derived gravity data contribute<br />

mainly to the area which is far away from the Australian coast and without the shiptrack<br />

data (cf. Kirby and Forsberg, 1998; Featherstone et al., 2003, Figure 1).<br />

Therefore, the AUSGeoid98 geoid model can be thought of as a quasi-independent<br />

reference.<br />

To first order, the altimeter-observed mean sea surface represents the geoid<br />

(Mantripp, 1996). This means that comparison of the geoid height and SSH can<br />

provide assessment of the quality of retracked SSH data. It is also recalled that the<br />

SST is not removed from the SSH data around the Australian coast due to the reasons<br />

stated in Section 2.5.2. As stated in Section 5.7.3, since the quasi-independent<br />

reference AUSGeoid98 is used, again, the verification of any improvement in the<br />

SSH will only be seen in the standard deviations of the mean difference, with a<br />

smaller value indicating an improvement in the determination of the altimeterderived<br />

SSH.<br />

For each along-track altimeter-derived SSH both before and after retracking, the<br />

gridded geoid height is bispline interpolated with longitude and latitude of the<br />

location. Then it is subtracted from the corresponding SSH to give the difference<br />

between the SSH and geoid height. The statistics of differences is used to quantify


Chapter 6. WAVEFORM RETRACKING APPLICATION 191<br />

the quality of the retracking system. Again, the STD is used to measure how the<br />

waveform retracking system works rather than the mean difference between the<br />

collinear differences.<br />

6.7.2 Results of the Comparison between SSH Data and Geoid Heights<br />

The descriptive statistics of differences between SSH data (20 Hz) before and after<br />

retracking and gridded geoid height has been computed. For convenience of<br />

comparison, mean differences and the STD of the differences have been plotted via<br />

six 5-km-wide distance bands in Figures 6.10 and 6.11 for cycle 42, and Figures 6.12<br />

and 6.13 for cycle 43, in which the STD is presented only using the positive value.<br />

Again, the STDs show large values for both SSH data before and after retracking.<br />

The reasons may be the temporal variations of the ocean surface and other incorrect<br />

corrections, such as ocean tides and wet tropospheric range corrections, but retracked<br />

SSH data show a higher precision than unretracked data.<br />

Before retracking<br />

After retracking<br />

Mean difference (m)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0-5 5-10 10-15 15-20 20-25 25-30<br />

Distances from the coastline (km)<br />

Figure 6.10 Mean differences of geoid heights (AUSGeoid98) and SSH data before<br />

and after retracking in six 5 km wide distance bands (cycle 42).<br />

For cycle 42, Figures 6.10 and 6.11 indicate that both mean difference and STD<br />

decrease when using retracked SSH data. The values of STD drop down to ~1 m and<br />

the significant improvement occur in distances 0-15 km to the coastline. After 15 km<br />

from the coastline, there is no significant decrease in the value of the STD of<br />

differences. From Figures 6.12 and 6.13 for cycle 43, similar results can be obtained.<br />

These results suggest that more precise SSH profile can be obtained from retracked<br />

SSH data, especially in distance of 0-15 km from the coastline.


Chapter 6. WAVEFORM RETRACKING APPLICATION 192<br />

Before retracking<br />

After retracking<br />

STD (m)<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0-5 5-10 10-15 15-20 20-25 25-30<br />

Distances from the coastline (km)<br />

Figure 6.11 STD (positive values) of the mean difference before and after retracking<br />

in six 5 km wide distance bands, showing improvement on SSH data after retracking<br />

(cycle 42).<br />

Before retracking<br />

After retracking<br />

Mean differences (m)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0-5 5-10 10-15 15-20 20-25 25-30<br />

Distances from the coastline (km)<br />

Figure 6.12 Mean differences between geoid heights (AUSGeoid98) and SSH data<br />

before and after retracking in six 5 km wide distance bands (cycle 43).<br />

Before retracking<br />

After retracking<br />

STD (m)<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0-5 5-10 10-15 15-20 20-25 25-30<br />

Distances from the coastline (km)<br />

Figure 6.13 STD (positive values) of the mean difference before and after retracking<br />

in six 5 km wide distance bands, showing improvement on SSH data after retracking<br />

(cycle 43).


Chapter 6. WAVEFORM RETRACKING APPLICATION 193<br />

Figures 6.11 and 6.13 show that the STD of differences after retracking drops down<br />

to ±1 m when the distance to the coastline is larger than 10 km for cycle 42 and 5 km<br />

for cycle 43, respectively. However, the STD before retracking decreases to less than<br />

±1 m when the distance to the coastline is more than 20 km for both cycles. This<br />

indicates clearly that the retracked SSH can be determined accurately to the same<br />

level as SSHs over open oceans after offshore 10 km or even 5 km (e.g., cycle 43),<br />

while unretracked SSH data can have the same accuracy after only 15 km from the<br />

coastline. This means that these improved SSHs after retracking in coastal regions<br />

can be used with sufficient accuracy to compute coastal geoidal undulations, coastal<br />

ocean circulations, gravity anomalies and ocean tides. The results from Figures 6.11<br />

and 6.13 also show a better accuracy at retracked SSH data in Australian coastal<br />

regions than unretracked SSH data. This demonstrates that the coastal waveform<br />

retracking system developed in this study has improved and expended altimeterobserved<br />

measurements ~10 km shoreward in generally in Australian coastal regions<br />

comparing with unretracked SSH data when using AUSGeoid98 as a quasiindependent<br />

reference.<br />

6. 8 Summary<br />

Regional investigations around the Australian coast have demonstrated that the<br />

waveform retracking system developed in Chapter 5 can effectively reprocess 20 Hz<br />

waveform data and remove most errors caused by waveform contamination in these<br />

areas. Two cycles (42 and 43, March to May 1999) of ERS-2 20 Hz waveform data<br />

sets offshore 30 km around the Australia’s coast have been used to test the coastal<br />

retracking system developed in this study. The whole coastal region is divided into<br />

10 15°×15° sub areas. SSH data both before and after retracking are derived from<br />

uncorrected and corrected range measurements for the purpose of comparison and<br />

validation. Altimeter-derived SSH data used are 20 Hz without applying any filtering<br />

procedure to avoid any loss of high frequency signals. Using these SSH data and the<br />

AUSGeoid98 model, several analyses have been conducted in this Chapter.<br />

The results of waveform classification show that most coastal waveforms (80.19%)<br />

have ocean-like shapes with a single ramp but without significant peaks. Of the<br />

remaining ~12.34% of waveforms, predominant waveforms have single pre-peaked


Chapter 6. WAVEFORM RETRACKING APPLICATION 194<br />

(4.65%) and high-peaked (4.35%) waveform shapes. Others, such as single midpeaked<br />

and post-peaked, double pre-peaked and post-peaked, and multi-peaked<br />

waveforms, are from 0.73% to 1.35%.<br />

The categorised waveform is then retracked by an appropriate retracker depending on<br />

its shape. Retracking results at the Australian coast show that 77.7% - 96.2% of<br />

waveforms can be retracked using least squares iterative fitting algorithm, while<br />

3.8% - 18.1% of waveforms are needed to be retracked by the threshold retracking<br />

algorithm. The percentages of using different retrackers vary from area to area. Of<br />

these waveforms retracked by fitting algorithm, over 98.3% of waveforms are fitted<br />

using the ocean model and the rest of waveforms are fitted by the five-parameter<br />

model with an exponential trailing edge and the nine-parameter model with either a<br />

linear trailing edge or an exponential trailing edge. It has been found that waveforms<br />

retracked by the threshold retracking algorithm decrease from 35% in a distance of<br />

0-5 km from the coastline to 4% in a distance of 5-10 km from the coastline. After<br />

10 km offshore, only 2% of waveforms need to be threshold retracked.<br />

By comparison of along-track altimeter-derived SSH data before and after retracking,<br />

the largest errors caused by waveform contamination are found in a distance band of<br />

0-5 km, which is ~-51.2 cm (cycle 42) and ~-51.5 cm (cycle 43). Errors of 7.8 cm<br />

and 13.5 cm are found in the distance band of 5-10 km from the coastline. Beyond 10<br />

km offshore, the error has been found to be less significant (~-3 cm).<br />

Results from single track analysis indicate that the coastal retracking system can deal<br />

with more waveforms to derive thus more SSH data than a single retracker (e.g., the<br />

ocean model). More importantly, the system can make further the SSH profile closer<br />

to the coastline ~2.5-5 km. This is an improvement of the single retracker, at which<br />

the SSH profile presents without showing effects of waveform contamination after<br />

5 km from the coastline.<br />

In general, by comparing altimeter-derived along-track SSH data before and after<br />

retracking with AUSGeoid98 geoid heights, results show that the altimeter-derived<br />

SSH profiles can be extended ~5-10 km shoreward by the use of a coastal retracking<br />

system and range corrections. These improved SSHs show a same accuracy level as<br />

those over open oceans, thus indicating a great potential to apply them with sufficient


Chapter 6. WAVEFORM RETRACKING APPLICATION 195<br />

accuracy to altimetric geodetic applications in coastal regions. However, SSH<br />

profiles at a distance of 0-5 km to the coastline cannot be recovered by the waveform<br />

retracking procedure at the same precision level as SSH data after 5 km from the<br />

coastline.<br />

The problem found from this research is that, in general, the retracked SSHs become<br />

nosier compared to the unretracked data. Therefore, a low-pass filtering process is<br />

necessary before applying these retracked data.


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 196<br />

7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES NEAR<br />

TAIWAN FROM WAVEFORM RETRACKING<br />

7. 1 Introduction<br />

Satellite altimetry has provided the most valuable of data sets for the recovery of the<br />

marine gravity field. Numerous local and global marine gravity anomalies have been<br />

created using a variety of successful techniques. Papers reporting current results<br />

include those of Hwang and Parsons (1995), Sandwell and Smith (1997), Andersen<br />

and Knudsen (1998), Tziavos et al. (1998), Hwang (1998), Andersen et al. (2001),<br />

and Hwang et al. (2002b). The precision and spatial resolution of marine gravity<br />

anomalies derived from satellite altimetry have also been significantly improved in<br />

open oceans over the past 25 years (Tapley and Kim, 2001; Rapp and Yi, 1997).<br />

Using the high-density data collected from multi-satellite missions, the precision of<br />

the derived global gravity fields is reported to have a range of ~±3 mgal to<br />

~±14 mgal from the comparisons of ship-track gravity and altimeter-derived gravity<br />

measurements (e.g., Tapley and Kim, 2001; Hwang et al., 2001).<br />

However, the precision of altimeter-derived gravity anomalies in coastal regions is<br />

not at the same level as those in open oceans. The differences between predicted<br />

anomalies from altimeter data and ship-track gravity data tend to become larger in<br />

coastal zones (e.g., Hwang et al., 1998b; Andersen and Knudersen, 2000;<br />

Featherstone, 2002; Hwang et al., 2002b). Hwang et al. (2002b) estimate a global<br />

gravity anomaly grid using Seasat, GEOSAT, ERS1-2, TOPEX/POSEIDON<br />

altimeter data and report a precision of ~±13 mgal in the Fiji Plateau and<br />

Mediterranean Sea, where the SSH data are likely affected by land and islands.<br />

Featherstone (2002) investigates the differences around the Australian coastal<br />

regions between ship-track gravity data and the most recent altimeter-derived gravity<br />

grids of GMGA97 (Hwang et al., 1998b), KMS01 (Andersen et al., 2001), and<br />

Sandwell’s version 9.2 (Sandwell and Smith, 1997). The result shows that larger<br />

differences (>100 mgal in magnitude) occur in areas closer to the coastal regions,<br />

which are consistent for all altimeter grids (see Figures 11-14 of Featherstone, 2002).


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 197<br />

The reasons for these larger differences observed near the coastal regions are<br />

probably as follows.<br />

• Land topography and rugged sea states contaminate coastal altimeter waveform<br />

measurements, causing SSHs to be in error. For example, TOPEX/POSEIDON<br />

altimeter waveform data are affected by land to ~0-34.8 km from the coastline<br />

in some coastal regions (Brooks et al., 1997); ERS-2 altimeter waveform data<br />

can be contaminated to a distance range of ~0-22 km off the Australian<br />

coastline (cf. Chapter 4).<br />

• Altimeter measurements cannot be properly corrected due to inaccurate<br />

geophysical corrections such as shallow-sea tides and the atmospheric delay.<br />

The recent analyses in accuracy of these corrections are given, e.g., by Shum et<br />

al. (1998), Andersen and Knudsen (2000), and Hwang et al. (1998b). In<br />

addition, waveform contamination introduces errors to the SWH for the SSB<br />

correction, which will be discussed later.<br />

• There are still large biases in ship-track gravity data due to improper or<br />

impossible data cleaning (i.e., crossover analysis, poor navigation in position,<br />

and Eötvös corrections) so that they cannot always offer definitive control on<br />

the accuracy of the altimeter-derived gravity anomalies.<br />

• Lack of altimeter data over the land near the coastline presents a problem that<br />

the gravity computation over coastal waters will be theoretically inaccurate.<br />

This is because the gravity recovery procedure requires uniform data around<br />

the computation point (see Eq. 15 of Hwang et al., 1998b), but this may not be<br />

the case when the point is close to the coastline.<br />

• The FFT technique requires the data to be on a regular grid. Incomplete data<br />

with gaps in coastal regions will cause the edge effect (i.e., Gibbs phenomenon)<br />

in recovery of gravity anomalies.<br />

In considering the first reason above, the error caused by altimeter waveform<br />

contamination in coastal regions can be mostly removed using waveform-retracking<br />

(e.g., Brooks et al., 1997; Deng et al., 2001). This Chapter focuses on improving the<br />

accuracy of altimeter-derived gravity anomalies through retracking ERS-1 altimeter<br />

waveform data in the Taiwan Strait over an area bound by 119°E ≤ λ ≤ 122°E and<br />

23°N ≤ ϕ ≤ 26°N. For the third cause above, there are some well-adjusted ship-track


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 198<br />

gravity data in the study area (see Section 2.5.3) which can be used to conduct an<br />

analysis between them and altimeter-derived gravity anomalies (not withstanding the<br />

errors in these data; mentioned above).<br />

The Taiwan Strait is characterised by shallow water (ocean depth less than 200 m),<br />

small islands, and rugged coastlines. Thus, it is assumed reasonably representative of<br />

the altimeter waveform features that may be encountered in most coastal regions.<br />

Nearly 2.5 years of ERS-1 altimeter waveform data over the area include the<br />

measurements from the ice, multi-disciplinary and geodetic phases, producing denser<br />

data coverage for the recovery of the gravity anomalies. These are the reasons that<br />

this area is chosen.<br />

The data and area signatures, waveform retracking methods and SSH extraction will<br />

be described in Sections 7.2, 7.3 and 7.4, respectively. In order to assess how the<br />

retracking procedure improves the accuracy of the altimeter-derived gravity<br />

anomalies, two SSH sets (i.e., before and after retracking) are extracted from<br />

waveform data using the same method except for the retracking procedure.<br />

Retracking is carried out over a larger area including the China Seas (i.e., the East,<br />

South, Yellow Seas and Bohai). The description in Section 7.5 will relate to the<br />

recovery of gravity anomalies from altimeter-derived data both before and after<br />

waveform retracking from the inverse Vening Meinesz formula (Hwang, 1998a)<br />

using the 1-D FFT technique. Comparisons of recovered anomalies from estimates<br />

by these two data sets to ship-track gravity data will be described. It should be<br />

acknowledged here that although these ship-track gravity data may contain errors,<br />

they are the only external data that can be used to indirectly verify the tracking<br />

algorithm’s differences.<br />

7. 2 Data and Area<br />

As stated in Section 2.4.1, ~2.5 years of 20 Hz ERS-1 waveform data from AWF<br />

products are used in a coastal region close to China and Taiwan bounded by<br />

110°E ≤ λ ≤ 129°E and 20°N ≤ ϕ ≤ 42°N. Figure 7.1 shows the geographical<br />

distribution of the observations, where the study area for the gravity anomalies in this<br />

Chapter is represented by a small rectangle (118°E ≤ λ ≤ 123°E and


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 199<br />

22°N ≤ ϕ ≤ 27°N). Among the ERS-1 repeat cycles, the geodetic mission (GM)<br />

enables the acquisition of a high density of altimeter measurements and thus<br />

improves the spatial resolution of the altimeter-derived SSH and hence the<br />

determination of marine gravity anomalies.<br />

Figure 7.1 Distribution of ERS-1 ground tracks<br />

As the original purpose of the data used in this area was to investigate the shallow<br />

water tides, observations with the ocean depth deeper than 200 m were omitted,<br />

making no data in the eastern area of Taiwan (Figure 7.1). This presents a problem<br />

for the recovery of the gravity anomaly in this part of the study area, where the final<br />

marine gravity field will be dominated by the reference gravity field and cannot show<br />

the high frequency components (i.e., ~20 km wavelength) because of the lack of data.<br />

The study area is a narrow strait (roughly ~110 km wide) between Taiwan and China<br />

linking the East China Sea to the South China Sea (named the Taiwan Strait). The<br />

computation area is 1° larger than the final gridded gravity field (119°E ≤ λ ≤ 122°E


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 200<br />

and 23°N ≤ ϕ ≤ 26°N) at the edges to avoid the Gibbs effect of the FFT technique.<br />

Table 7.1 shows a summary of altimeter observations in the area.<br />

Table 7.1 Summary of ERS-1 observation cycles bounded by 118°E ≤ λ ≤ 123°E and<br />

22°N ≤ ϕ ≤ 27°N<br />

Repeat cycle<br />

Time period<br />

Number<br />

Ascending<br />

Descending<br />

(days)<br />

of cycles<br />

pass<br />

pass<br />

3 23/12/1993 - 10/04/1994 36 0 1<br />

35 13/07/1993 - 23/12/1993<br />

21/03/1995 - 02/06/1996<br />

5<br />

13<br />

8 6<br />

168 04/10/1994 - 24/03/1994 1 35 24<br />

Although only ERS-1 altimeter data are used, an adequate density (4×4 minutes) of<br />

measurements is the main characteristic of the data set readily available at the time of<br />

this research. From Table 7.1, the 3-day repeat cycle shows only one descending<br />

ground track, but the 168-day repeat cycle presents 35 ascending and 24 descending<br />

tracks. Together with eight ascending and six descending tracks from the 35-day<br />

repeat cycle, satellite ground tracks from three types of repeat cycles create a denser<br />

coverage of data than that using only single repeat cycle of data over this area,<br />

increasing greatly the spatial resolution (Figure 7.1). This will contribute most to the<br />

high-frequency parts of the coastal marine gravity field.<br />

To assess the accuracy of the gravity anomaly from satellite altimeter, the ship-track<br />

gravity data at 397 points described in Section 2.5.3 around Taiwan are used in this<br />

study (Figure 2.5).<br />

7. 3 Waveform Retracking<br />

To retrack the return waveforms in this area, the iterative least squares procedure<br />

(Section 5.5) is used to fit an analytic function representing the expected return to the


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 201<br />

waveform. The forms of the analytic function used in this study are the fiveparameter<br />

model (Equation 3.40 and n =1) and the ocean model (Equation 5.3).<br />

7.3.1 Retracking Scheme<br />

Waveform contamination occurs mostly in the vicinity of land and islands (see<br />

Chapter 4). Over ocean surfaces, the contamination occurs only over regions where<br />

the rugged sea state or the off-nadir satellite attitude excursions affect the shape of<br />

the return (Brenner et al., 1993). However, research conducted by Fairhead et al.<br />

(2001) and Sandwell (2003) indicates that waveform retracking is required over all<br />

oceans to improve the resolution and quality of the marine gravity field. Therefore,<br />

all 2.5 years of ERS-1 20 Hz waveforms shown in Figure 7.1 are retracked. Before<br />

retracking, an edit criterion (Section 5.2) was applied to remove waveforms in which<br />

there is no leading edge in the waveform sample window, which removed ~2% of the<br />

data.<br />

When fitting, a priori values and weights are determined using the method presented<br />

in Section 5.5.4. Retracking results are obtained by retracking the waveform over<br />

ocean surfaces and in coastal regions using the fitting functions of either the ocean<br />

model or five-parameter model. Each waveform is first classified into ocean-like and<br />

non-ocean shapes so that it can be retracked by the appropriate functional fit. The<br />

iterative procedure is repeated until the unknown parameter estimate ( t 0<br />

or β<br />

3<br />

)<br />

converges to within 0.1% or<br />

−4<br />

σ<br />

0<br />

−σ<br />

0<br />

< 10 . It converges in typically 2-4<br />

i<br />

i −1<br />

iterations. In addition, the five-parameter function replaces the ocean model to<br />

compute the retracking correction when the ocean model fails.<br />

The retracking procedure is conducted by using either the five-parameter model or<br />

the ocean model. Retracking was first performed using both models. A collinear<br />

analysis was applied to one cycle of 35-day repeat of 1 Hz SSH data (from July to<br />

August 1993) to investigate whether there is a serious bias between both methods.<br />

The difference ( dh ) between the retracked SSH data calculated from the ocean<br />

model and five-parameter model was computed and analysed. Figure 7.2 shows a<br />

histogram of dh in which the mean SSH difference is -0.7±13.5 cm (3070<br />

observations). Of the residuals, 95% are less than 10 cm and 83% are less than 5 cm,


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 202<br />

indicating a good agreement between both methods and a similar result as ERS-2<br />

(see Section 5.8). This result also means that the combination of both ocean and fiveparameter<br />

models can maintain a consistent measurement noise level across the<br />

entirety of the sea surface profile. Since the ocean model physically represents the<br />

returns over oceans, this is the reason that both models are finally chosen to be the<br />

main retracking functions in this study.<br />

Frequency of data<br />

800<br />

600<br />

400<br />

200<br />

0<br />

-1.45<br />

-1.27<br />

-1.08<br />

-0.90<br />

-0.71<br />

-0.52<br />

-0.33<br />

-0.14<br />

0.05<br />

dh (m)<br />

0.23<br />

0.42<br />

0.61<br />

0.80<br />

0.99<br />

1.18<br />

1.37<br />

Figure 7.2 The histogram of dh (the SSH difference between the five-parameter<br />

model and the ocean model (mean -0.7 cm; STD 13.5 cm; 3070 observations).<br />

7.3.2 Examples of Waveform Retracking Results<br />

Figure 7.3 shows one ascending (north-westwards) and one descending ground track<br />

(south-westwards) in the study area. The ascending track (A1) goes from water to<br />

land. The descending track (D1) recedes from land to water. It is parallel to the<br />

coastline at ~37°N latitude, and then intercepts the land at ~32.4°N latitude. The<br />

track then re-commences as D2 and goes to the Taiwan Strait. The 20 Hz waveforms<br />

have been retracked using the methods mentioned above and compressed to 1 Hz<br />

SSH data; the before/after retracking 1 Hz SSH results, plotted versus latitude, are<br />

shown in Figure 7.4 for these three tracks. The profiles from top to bottom of<br />

Figure 7.4 show the SSH from tracks A1, D1 and D2. The SSH derived from the<br />

original (i.e., not retracked) and retracked (using the five-parameter and ocean<br />

models) data are shown. It is apparent that the retracking has removed some<br />

measurement noise near the coastal regions caused by land topography and<br />

rugged/rough sea states.


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 203<br />

Figure 7.3 Three-ground tracks in the study area<br />

As for the SSH derived from waveforms before retracking, Figure 7.4a does not<br />

show the obvious land effect on the unretracked SSH data when the ascending track<br />

(A1) approaches land at ~23.5°N latitude, while the contamination which may be<br />

caused by rough sea states occurs from ~21.8°N to ~22.4°N latitudes. Descending<br />

track (D1, Figure 7.4b) shows obvious sharp changes in unretracked SSH results<br />

when the track gets close to the land. Similar sharp changes can also be found in<br />

unretracked SSH data along the descending track (D2, Figure 7.4c). It is evident in<br />

Figure 7.4 that the retracking, no matter which method, has removed the ‘error’<br />

(unexpected sharp changes) in the SSH and the highly variable measured sea surface<br />

is replaced by the smoothed profile. The SSH data from retracked waveforms using<br />

the five-parameter model or ocean model show similar results (Figure 7.4).


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 204<br />

Sea Surface Height (metre)<br />

14.0<br />

12.0<br />

10.0<br />

8.0<br />

Five−parameter<br />

Ocean + five−parameter<br />

without retracking<br />

6.0<br />

21.7 22.2 22.7 23.2<br />

Latitude (degree)<br />

18.0<br />

Sea Surface Height (metre)<br />

16.0<br />

14.0<br />

12.0<br />

10.0<br />

8.0<br />

Five−parameter<br />

Ocean + five−parameter<br />

without retracking<br />

6.0<br />

32.0 34.0 36.0 38.0 40.0<br />

Latitude (degree)<br />

sea Surface Height (metre)<br />

21.0<br />

20.0<br />

19.0<br />

18.0<br />

17.0<br />

16.0<br />

15.0<br />

Five−parameter<br />

Ocean + five−parameter<br />

without retracking<br />

14.0<br />

22.0 23.0 24.0 25.0 26.0<br />

Latitude (degree)<br />

Figure 7.4 SSHs from ground track A1 (a), D1 (b), and D2 (c) in Figure 7.3 before<br />

and after retracking using the retrackers of the five-parameter and ocean models.


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 205<br />

The results in Figure 7.4 show that it is necessary to retrack all waveforms in this<br />

area due to its characteristics of shallow water, rugged coastlines and rough sea states.<br />

It also can be seen from Figure 7.4 that the retracking methods used can extend the<br />

reliability of the measurements shoreward. From this analysis, there is a small bias<br />

between the SSH before and after retracking (cf. Figure 7.4). A similar problem has<br />

also been found when using ERS-2 waveform data around Australian coastal regions<br />

(Section 5.8.4), where a bias of ~27 cm is estimated from two cycles of waveform<br />

data. Since the along-track geoid slope will be used in this study to calculate the<br />

gravity anomaly, the differentiation operation can effectively remove this bias. Thus,<br />

its effect on the SSH can be reduced. However, it cannot be neglected when<br />

analysing the absolute temporal variation in SSHs or mixing the original data (before<br />

retracking) and retracked data.<br />

7. 4 The SSH Extracted from Waveforms<br />

To assess the results from the above retracking, the SSH is extracted using two<br />

approaches. The first is to create the 1 Hz SSH data set from the 20 Hz data source<br />

without waveform retracking, while the other is to retrack the 20 Hz waveform and<br />

then to ‘compress’ the retracked SSH to the 1 Hz SSH set. No matter which method<br />

is used, a quadratic function is fitted to 20 Hz SSHs and Pope’s (1976) tau-test is<br />

applied to detect the erroneous SSH data. Final 1 Hz SSHs are calculated at the midpoint<br />

of the fitting function after removing the outlying observations.<br />

7.4.1 Corrections<br />

To obtain the data set of the SSH, geophysical and environmental corrections are<br />

applied to the range measurement. The corrections for atmospheric propagation<br />

delays are obtained directly from the AWF waveform product. The AWF product did<br />

not supply the inverse barometer correction, which is related to the hydrostatic<br />

response of the sea surface with increasing or decreasing atmospheric pressure.<br />

Instead, it was computed using the surface atmospheric pressure which is available<br />

indirectly via the dry tropospheric correction obtained from the AWF. The tidal<br />

correction applied is the superposition of the elastic ocean tides (i.e., the ocean tide<br />

and loading tide) derived from CRS4.0 global ocean tide model (Eanes, 1999), and


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 206<br />

the solid Earth tide supplied by the AWF. The vertical (radial) velocity of the<br />

satellite causes a frequency Doppler shift which affects the time delay measurement<br />

and thus the range. For each elementary measurement, the Doppler correction is<br />

added to the altimeter range. Table 7.2 lists the variability of the corrections applied<br />

in the area and their maximum and minimum magnitude. From Table 7.2, the change<br />

of the SSB was found to cause an error in the SSH data in the study area.<br />

Table 7.2 The variability of the range corrections to deliver SSH in the study area<br />

(cf. Figure 7.1)<br />

Correction Min (m) Max (m)<br />

Dry tropospheric delay 2.27 2.30<br />

Wet tropospheric delay 0.12 0.40<br />

Ionospheric delay 0.05 0.09<br />

Inverted barometer 0.05 0.17<br />

SSB -1.23 -0.02<br />

Elastic Ocean tides -2.15 0.22<br />

Solid Earth tides -0.10 0.03<br />

Doppler shift -0.01 0.01<br />

As stated in Section 3.3.3, the SSB is the difference between the apparent sea level<br />

measured by an altimeter and the true mean sea level. Theoretical understanding of<br />

the SSB, particularly the EMB, remains limited, so the SSB correction is calculated<br />

from the SWH and the wind speed using an empirical model derived from analysis of<br />

altimeter data itself (AVISO/Altimeter, 1998; Chelton et al., 2001). The SWH, as<br />

computed by the on-satellite processor (MacArthur, 1980, cited in Martin et al.,<br />

1983), is a measure of the ocean wave amplitude. Over the coastal regions near the<br />

land, the SWH computed on-satellite is not useful due to the effects of land<br />

topography and generally more rough sea states. Therefore, the SSB may no longer<br />

be correct in these areas.


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 207<br />

The SSB’s effect on the SSH is analysed through SSH profiles along the three<br />

ground tracks in Figure 7.3. SSH profiles along tracks are plotted based upon<br />

whether they are retracked or unretracked, applied by all geophysical corrections<br />

including SSB or not. Besides the SSB, parts of corrections of dry tropospheric delay<br />

and elastic ocean tides are chosen to plot along each track (from Figures 7.5 to 7.7),<br />

because they also show the larger variety in magnitude (Table 7.2). As can be seen,<br />

the SSB shows an obvious step change in this coastal region, but it is not expected to<br />

change significantly and suddenly over a few kilometres unless a rougher sea state is<br />

present.<br />

Compared to the SSB, other corrections of dry tropospheric delay and ocean tides<br />

show a relatively smooth change. As a result, the SSH profiles are shifted upward a<br />

value of the sum of corrections after they are applied to the range measurements.<br />

However, since the SSB should be subtracted from the SSH, it doubles its effect on<br />

the SSH when the step change occurs. Therefore, the SSB is not applied to the range<br />

measurements in this study. This may leave an error for both unretracked and<br />

retracked SSH data. However, as the retracking procedure partly removes the ocean<br />

wave influence on the waveform, the need of the range measurement for the SSB has<br />

been reduced. This is the reason that the retracked SSH shows the smoothing profiles<br />

in Figures 7.5, 7.6 and 7.7. In addition to the SSB, there are some sharp changes to<br />

be found in SSH profiles without retracking even in ocean areas far away from the<br />

coastline (see, Figures 7.5 and 7.6), while waveform retracking, again, obviously<br />

removes these unexpected changes. This implies that the SSH cannot be precisely<br />

computed from the on-satellite tracking algorithm in this coastal and semi-enclosed<br />

sea.


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 208<br />

14.0<br />

Sea Surface Height (metre)<br />

10.0<br />

6.0<br />

2.0<br />

−2.0<br />

21.7 22.2 22.7 23.2<br />

Latitude (degree)<br />

Figure 7.5 SSH profiles and parts of corrections along the track A1.


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 209<br />

18.0<br />

Sea Surface Height (metre)<br />

14.0<br />

10.0<br />

6.0<br />

2.0<br />

−2.0<br />

32.0 34.0 36.0 38.0 40.0<br />

Latitude (degree)<br />

Figure 7.6 SSH profiles and parts of corrections along the track D1.


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 210<br />

22.0<br />

Sea Surface Height (metre)<br />

18.0<br />

14.0<br />

10.0<br />

6.0<br />

2.0<br />

−2.0<br />

22.0 23.0 24.0 25.0<br />

Latitude (degree)<br />

Figure 7.7 SSH profiles and parts of corrections along the track D2.


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 211<br />

7.4.2 The Weight Function<br />

After waveform retracking, both retracked and unretracked 20 Hz SSH data derived<br />

from waveforms along ground tracks are created to 1 Hz SSH sets. Before the<br />

calculation of the gravity anomaly using these SSH data, a procedure of data filtering<br />

is necessary to delete possible outliers. The 1 Hz data are then separately smoothed<br />

and edited by a low-pass filter to remove the noise in the measurements and ensure<br />

the accuracy of altimeter-derived gravity anomalies. Because of the advantage of the<br />

data filter, a Gaussian gridding kernel is chosen as the weight function for removing<br />

outliers in SSH data and computing the average gradients along ascending and<br />

descending tracks (stated below). For a point ( ϕ , λ)<br />

from the observation ds<br />

kilometres, the formula is<br />

2<br />

⎡ 1 ⎛ ds ⎞ ⎤<br />

w(<br />

ϕ , λ)<br />

= exp⎢−<br />

⎜ ⎟ ⎥<br />

ds ≤ R s<br />

. (7.1)<br />

⎢⎣<br />

2 ⎝ σ ⎠ ⎥⎦<br />

where R s<br />

= D / 2 is the search radius of the weight function, D is the search<br />

window size, and σ = D / 6 . The value of D is important for this function due to the<br />

degree of smoothness increasing with it. From Equation (7.1), an influenced point<br />

from the observation will gain a small weight near the edge of the search window no<br />

matter how large the size is. Thus, a large value of D will over smooth the filtered<br />

results, while a small value of D will still leave unreasonable results. In this study,<br />

D =70 km is chosen for the along track SSH smoothing. This value ensures that there<br />

are about 10 seconds [time] of SSH data to be included in each filtering procedure<br />

for obtaining a robust result. For calculating along-track average gradients on a 2'×2'<br />

grid, the value of D is chosen to be the diagonal of the cell (~5 km) so that all SSH<br />

data can be searched (see Section 7.5.2).<br />

7.4.3 Data Filtering and Editing<br />

Data filtering will smooth out any high-frequency signal that is larger than the<br />

proposed resolution to avoid data aliasing. SSH data filtering is a moving average<br />

procedure, which uses a weight function (Equation 7.1, D = 70 km) and slides along<br />

the track by 1 second of data (i.e., 20 SSHs) for each calculation. This procedure is


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 212<br />

carried out at all observation points, resulting in a filtered SSH set. The final SSH at<br />

( ϕ , λ)<br />

is given as:<br />

SSH ( ϕ , λ)<br />

= ∑(<br />

SSH )<br />

i<br />

wi<br />

∑ wi<br />

.<br />

(7.2)<br />

i<br />

i<br />

where<br />

w represents the value of the weight at the point ( ϕ , λ)<br />

for the particular i th<br />

i<br />

observation within the distance D .<br />

Data that disagree by more than three times of STD are flagged as outliers and<br />

replaced by the filtered values. In addition, when the number of SSH data within D<br />

are less than 4, the smoothing procedure is not conducted, and all data are deleted.<br />

Approximately 1% of the unretracked and 2% of retracked initial data sets are<br />

eliminated using this filter, implying that retracked data show more short-wavelength<br />

signal than the unretracked data. Figures 7.8 and 7.9 show filtered (dark curves) and<br />

unfiltered (light curves) SSH results before and after retracking along ground tracks<br />

in Figure 7.3 through the above Gaussan smooth procedure, respectively. As can be<br />

seen, the Gaussian weight function removes the measurement error in observations.<br />

7. 5 Marine Gravity Anomalies from Satellite Altimetry<br />

There have been several approaches developed to compute gravity anomalies from<br />

satellite altimeter data over the past two decades. The four main types of the<br />

philosophies and computational procedures have been summarised by Featherstone<br />

(2002). Each of them has advantages and disadvantages. Featherstone (ibid.) shows<br />

that there are some significant differences among marine gravity anomalies<br />

computed using different methods from satellite altimetry around Australia, and<br />

these tend to become larger in coastal areas. Therefore, it is still an open question to<br />

say which one is better than others.<br />

One of them is the approach described by Hwang (1998a), which is used in this study<br />

due to its numerically efficient computation and no requirement of crossover<br />

adjustment (e.g., Sandwell and Smith, 1997; Hwang, 1998a; Andersen and Knudsen,<br />

1998). In this approach, the gravity anomaly is calculated from the along-track<br />

deflections of the vertical (or along-track SSH slopes) derived from satellite altimeter


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 213<br />

Figure 7.8 Untracked SSH profiles, showing the filtered SSH data (dark curves) and<br />

unfiltered SSH data (light curves) along the tracks A1 (a), D1 (b) and D2 (c) in<br />

Figure 7.3.


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 214<br />

Figure 7.9 Retracked SSH profiles, showing filtered SSH data (dark curves) and<br />

unfiltered SSH data (light curves) along the tracks A1 (a), D1 (b) and D2 (c) in<br />

Figure 7.3.


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 215<br />

data using the inverse Vening Meinesz formula (Hwang, 1998a). The use of alongtrack<br />

geoid slope data has been proved to be one of effective techniques used in the<br />

calculation of gravity anomalies from satellite altimeter data, as the differentiation<br />

operation effectively reduces the effect of the long wavelength radial orbit error (e.g.,<br />

Sandwell, 1984; Hwang and Parsons, 1995; Sandwell, 1997).<br />

Before calculating along-track deflections of the vertical, the quasi time-independent<br />

estimates of dynamic ocean topography must be removed from the altimeter-derived<br />

SSHs. Such removal is conceptually required because the ocean surface is not an<br />

equipotential surface, but the geoid is. The remove-compute-restore technique is<br />

used, in which the deflections of the vertical (i.e., ξ north-south and η east-west)<br />

implied by EGM96 (Lemoine et al., 1998) global geopotential model to spherical<br />

harmonic degree 360 is used as a reference field, and it is removed from the raw<br />

deflections of the vertical to yield a residual deflections of the vertical set. These<br />

north and east component of deflections of the vertical are then interpolated onto a<br />

regular grid using least squares collocation (Moritz, 1980) after a procedure of<br />

determining outliers in the data (cf. Hwang et al., 2001). The inverse Vening<br />

Meinesz formula (Hwang, 1998a) is employed to convert these gridded north-south<br />

and west-east components to gravity anomalies via the 1-D FFT technique. The final<br />

gravity anomalies are obtained by adding back the eliminated 360 degree EGM96<br />

reference field.<br />

The approach has been successfully applied to recover marine gravity anomalies<br />

from multi-satellite altimeter data (e.g., SEASAT, GEOSAT, ERS-1 and<br />

TOPEX/POSEIDON) both at global and regional scales (Hwang, 1998a; Hwang et<br />

al., 2002b). The advantage of this computational procedure is that data are carefully<br />

prepared using the gridding and filtering techniques so that the data can be extracted<br />

efficiently. It is noted that ERS-1 altimeter data close to the coastline (~20 km from<br />

the coastline line) had not been used by Hwang (2001) to avoid data contamination<br />

in coastal regions. In this study, however, data used are only from the ERS-1<br />

altimeter in a coastal region. This presents a necessity for this study that some slight<br />

modifications (see Section 7.5.1) should be conducted in the computing procedure to<br />

use all potential retracked SSHs in coastal regions when recovering the gravity<br />

anomalies through the inverse Vening Meinesz formula.


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 216<br />

7.5.1 Along-track Deflections of the Vertical<br />

Before computing the along-track deflections of the vertical, in order to obtain<br />

geoidal height from SSHs, the value of the quasi-time independent SST interpolated<br />

from the model of Levitus et al. (1997) is subtracted from the SSH. In addition, the<br />

reference gravity field of the EGM96 to harmonic degree 360 is also removed from<br />

the geoidal height, meaning the along-track deflections of the vertical (generated<br />

from below Equation 7.3) are residual deflections of the vertical or residual geoidgradients.<br />

Taking the first horizontal derivatives of the altimeter-derived SSHs along track<br />

generates the along-track deflections of the vertical (DOV), which is the negative<br />

gradient of the geoid (Sandwell, 1992; Olgiati et al., 1995; and Hwang et al., 1998b).<br />

The along-track gradient, ε , is calculated at the midpoint between two successive<br />

observations by the formula<br />

ε<br />

k<br />

=<br />

N<br />

k +1<br />

− N<br />

∆s<br />

k<br />

(7.3)<br />

where<br />

N<br />

k<br />

is the geoidal height at the observation k , and ∆ s is the distance between<br />

observations k + 1 and k . In addition, the following Equation is less noisy for the<br />

along-track DOV:<br />

ε<br />

N<br />

1<br />

− N<br />

2∆s<br />

k + k −1<br />

k<br />

=<br />

(7.4)<br />

7.5.2 Averaging Gradients<br />

Usually, before computing the along-track geoid slope, the SSHs from repeat<br />

missions are reduced to a reference orbit (see Section 6.6) and then averaged them to<br />

remove the time-variant ocean signals and measurement noise (e.g., Hwang and<br />

Parsons, 1995; Hwang et al., 2002a; Rapp and Yi, 1997; Sandwell and Smith, 1997).<br />

However, lack of a reference orbit makes it difficult to directly average geoid heights<br />

for ERS-1/35 repeat data. Therefore, a new method has been developed to average<br />

along-track geoid gradients from repeat cycles in this coastal region. This procedure<br />

averages the geoid slope rather than the geoid height, thus avoiding the introduction


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 217<br />

of offsets at data gaps (cf. McAdoo, 1990; Kirby, 1996) and reducing the timevariant<br />

signals.<br />

The data used contain 3-day and 35-day repeat missions, and geodetic mission. Of<br />

these missions, the repeat mission repeats ground tracks to within 1 km, while the<br />

ground track of ERS-1/GM is unpredictable (with respect to the 35-day repeat track).<br />

For the repeat missions, the same pass (ascending or descending ground track) from<br />

different cycles does not repeat exactly due to some causes, such as orbital variations<br />

caused by gravitational influence and drag effects. Orbital manoeuvres have been<br />

used in the satellite to maintain a ground track within some specified linear tolerance<br />

of a nominal or reference track (Wang and Rapp, 1991). For most satellite altimeters<br />

such as GEOSAT and T/P, the aim is to achieve cross-track deviations on the order<br />

of 1 km (Wang and Rapp, 1991). This is the same case as the ERS satellite. Taking<br />

this data characteristic in mind, ideally the averaging method should retain the alongtrack<br />

resolution, but also average the data into the spaces between tracks from<br />

different repeat cycles so that a complete along-track gradient set can be generated.<br />

In this new averaging procedure, residual geoid heights at any location are first<br />

sorted and found in a 2′×2′ cell. A weighted mean geoid gradient, ε , is then<br />

computed at this cell using a Gaussian weight function (7.1). The search radius<br />

(~2.5 km) is chosen to be half the diagonal of the cell so that all SSHs in a cell can be<br />

searched. In addition, the standard deviation of the mean geoid gradient is calculated<br />

in the cell by the formula<br />

∑<br />

( ε − ε )<br />

i<br />

2<br />

i<br />

σ<br />

k<br />

=<br />

(7.5)<br />

n −1<br />

where n is the number of the sea surface slopes in the cell. When n =<br />

1, it is<br />

impossible to calculate σ<br />

k<br />

from (7.5). In this case, σ<br />

k<br />

is given a value of 0.21 µrad<br />

which is based upon an empirical noise level of ERS-1 (cf. Hwang et al., 2001). This<br />

usually happens to the mean slopes formed from the geodetic phase.<br />

After averaging gradients and before forming the regular grids for the use of 1-D<br />

FFT algorithm, a similar outlier-removing procedure to Hwang et al. (2002b) is<br />

applied to the mean gradients to remove residual erroneous data. Unlike Hwang’s


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 218<br />

scheme, where a 4′×4′ cell was used due to a high density of the ground coverage<br />

from multi-satellite altimeter data, the size of the cell is increased to 8′×8′. The<br />

reasons for this are as follows. First, the data used comes only from the ERS-1<br />

altimetry, thus a small cell cannot provide enough data points (some 4′×4′ cells may<br />

contain only one data point from the GM phase) and produce larger degrees of<br />

freedom to make the result more reliable, leading to improper detection of outliers.<br />

Second, since the study area is a coastal region, a small cell will cause all data to be<br />

eliminated when close to the coastline. However, it should be acknowledged that a<br />

larger cell might oversmooth the data.<br />

Averaging is carried out separately on the ascending and descending groundtracks.<br />

Table 7.3 lists the descriptive statistics results from averaging gradients in this way<br />

and outlier detection. Before detecting outliers, the mean value of the averaged<br />

slopes (before retracking, both ascending and descending) is 0.13 µrad and their<br />

standard deviation is ±3.01 µrad, while the mean of the slopes (after retracking, both<br />

ascending and descending) is 0.25 µrad and their standard deviation is ±2.61 µrad.<br />

After detecting outliers, the standard deviations reduce to ±2.34 µrad and ±1.82 µrad<br />

before and after retracking, respectively. There are 15% and 13% of outliers which<br />

are detected from geoid gradient data sets before and after waveform retracking,<br />

respectively. From Table 7.3, the number of final residual geoid gradients before<br />

retracking is more than that after retracking. Again, this shows that the retracking<br />

procedure is effective.<br />

Table 7.3 Descriptive statistics of the averaged residual geoid gradients before and<br />

after the outlier detection on a 8′×8′ cell.<br />

Before detecting outliers After detecting outliers<br />

Methods Mean<br />

(µrad)<br />

STD<br />

(µrad)<br />

Number<br />

of slopes<br />

Mean<br />

(µrad)<br />

STD<br />

(µrad)<br />

Number<br />

of slopes<br />

Before retracking 0.13 ±3.01 5949 -0.01 ±2.34 5177<br />

After retracking 0.25 ±2.61 5806 0.14 ±1.82 4953


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 219<br />

7.5.3 Gravity Anomaly Computation<br />

After removing outliers, the along-track geoid slopes are converted to regular grids<br />

of east (η) and north (ξ) vertical deflection using least-squares collocation. The<br />

covariance functions used here are taken from Hwang and Parsons (1995). Then, the<br />

calculations for the determination of residual gravity anomalies are carried out using<br />

the inverse Vening Meinesz formula Equation (23) of Hwang (1998a), and the<br />

innermost zone effects is estimated by Equation (47) of Hwang (ibid.) using 1-D FFT<br />

and the remove-restore technique. As stated, a 1° border is used to avoid Gibbs edge<br />

effects. The flow diagram in Figure 7.10 illustrates how altimeter waveform<br />

observations are processed and converted to gravity anomalies. The procedure<br />

follows the steps described in Section 7.4 for extracting the SSHs and estimating the<br />

gravity anomalies in this Section. Two gravity-anomaly grids (2′×2′) are generated<br />

by this procedure: a grid without waveform retracking; and a grid with waveform<br />

retracking. Figures 7.11 and 7.12 show the equivalent free air anomalies relative to<br />

GRS80 normal gravity estimated by unretracked and retracked ERS-1 SSH data.<br />

7.5.4 Comparisons between Marine and Ship Gravity Anomalies<br />

The accuracy of the gridded free air anomaly in Figures 7.11 and 7.12 can be<br />

assessed by comparing the ship-track gravity data with anomaly fields generated<br />

using altimeter data. The comparison is made along ship tracks in the Taiwan Strait<br />

in the vicinity of Taiwan (Section 2.4.3), where the ship data available is shown in<br />

Figure 2.5 and the altimeter ground tracks are shown in Figure 7.1. Each of the<br />

gridded anomaly data sets derived from altimeter SSH data before and after<br />

retracking is interpolated using a bispline function to the position of the ship-derived<br />

gravity anomaly. The number of the ship-track gravity data used in the comparisons<br />

in this area is 397 points. The descriptive statistics of the differences are shown in<br />

Table 7.4. Considering the retracked results in the comparison, it can be seen that the<br />

solution is better than that which uses unretracked SSHs in the anomaly estimation.<br />

The standard deviation of the difference (altimeter-derived anomalies minus ship<br />

anomalies) of ±13.93 mgal before retracking is larger than that of ±9.92 mgal after<br />

retracking. Therefore, the precision of the gridded altimeter-derived gravity anomaly<br />

has been improved by ~4 mgal.


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 220<br />

Figure 7.10 Flow diagram for the SSH extraction from waveforms and recovery of<br />

gravity anomalies


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 221<br />

119˚<br />

26˚<br />

120˚<br />

121˚<br />

122˚<br />

26˚<br />

183.79<br />

89.96<br />

25˚<br />

25˚<br />

66.44<br />

49.49<br />

34.99<br />

m<br />

gal<br />

21.45<br />

7.90<br />

24˚<br />

24˚<br />

-6.59<br />

-23.55<br />

-47.07<br />

-147.56<br />

23˚<br />

119˚<br />

120˚<br />

121˚<br />

23˚<br />

122˚<br />

Figure 7.11 Gravity anomalies from altimeter data before retracking (units in mgal)<br />

119˚<br />

26˚<br />

120˚<br />

121˚<br />

122˚<br />

26˚<br />

187.70<br />

91.47<br />

25˚<br />

25˚<br />

67.90<br />

50.91<br />

36.39<br />

m<br />

gal<br />

22.82<br />

9.25<br />

24˚<br />

24˚<br />

-5.27<br />

-22.27<br />

-45.83<br />

-143.78<br />

23˚<br />

119˚<br />

120˚<br />

121˚<br />

23˚<br />

122˚<br />

Figure 7.12 Gravity anomalies from altimeter data after retracking (units in mgal)


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 222<br />

It is acknowledged that there is a low accuracy from Table 7.4 for both altimeterderived<br />

gravity anomalies before and after retracking with respect to other studies,<br />

e.g., Rapp and Yi (1997). As stated in Section 2.5.3, however, the accuracy of the<br />

ship-track gravity data after the crossover adjustment is ±11.2 mgal, although these<br />

data has been subjected to a data cleaning processes. This means that there is still the<br />

error existing in the ship-track gravity data. Therefore, the error of the ship-track<br />

gravity data should be a possible reason that causes larger STDs in Table 7.4.<br />

Figures 7.13 and 7.14 show the different results along ship tracks using unretracked<br />

and retracked altimeter data, respectively. As can be seen from Figure 7.13,<br />

numerous larger differences with magnitudes greater than +15 mgal occur in any<br />

ship track not being limited to near the coast. As stated, the ERS-1 waveform data<br />

are contaminated in this area not only in coastal regions but also in the water area far<br />

from the shoreline (cf. Figures 7.5 and 7.6). The larger differences here are more<br />

likely to have been subjected to some form of waveform contamination. This<br />

becomes more evident when comparing Figure 7.13 with Figure 7.14 (after<br />

retracking), where larger differences (>15 mgal in magnitude) appear along the ship<br />

tracks that are very much closer to Taiwan’s coast. These larger differences near the<br />

coastline may be due either to the limited distance that waveform retracking can<br />

improve or the Gibbs edge effect at the coastline.<br />

Table 7.4 Descriptive statistics of the differences between altimeter-derived grids and<br />

the 397 ship-track free-air gravity anomalies (units in mgal)<br />

Grid Max Min Mean STD RMS<br />

Before retracking 32.71 -41.62 -4.86 ±13.93 ±14.74<br />

After retracking 19.82 -41.08 -6.09 ±9.92 ±11.63


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 223<br />

119˚<br />

26˚<br />

120˚<br />

121˚<br />

122˚<br />

26˚<br />

25˚<br />

25˚<br />

35<br />

25<br />

15<br />

5<br />

-5<br />

m<br />

gal<br />

24˚<br />

24˚<br />

-15<br />

-25<br />

-35<br />

-45<br />

23˚<br />

119˚<br />

120˚<br />

121˚<br />

23˚<br />

122˚<br />

Figure 7.13 Differences between altimeter-derived (before retracking) and ship-track<br />

gravity anomalies<br />

119˚<br />

26˚<br />

120˚<br />

121˚<br />

122˚<br />

26˚<br />

25˚<br />

25˚<br />

35<br />

25<br />

15<br />

5<br />

-5<br />

m<br />

gal<br />

24˚<br />

24˚<br />

-15<br />

-25<br />

-35<br />

-45<br />

23˚<br />

119˚<br />

120˚<br />

121˚<br />

23˚<br />

122˚<br />

Figure 7.14 Differences between altimeter-derived (after retracking) and ship-track<br />

gravity anomalies


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 224<br />

In interpreting the anomaly difference here, one should recall that different altimeter<br />

SSH data sets are extracted from the same waveform source except for the retracking<br />

procedure. All other calculations are carried out using the same procedure. Therefore,<br />

the results shown in Table 7.4 reflect the improvements only from waveform<br />

retracking in coastal regions.<br />

7. 6 Summary<br />

Gravity anomaly predictions were made from ~2.5 years of ERS-1 SSH data sets<br />

before and after retracking using the inverse Vening Meinesz formula and the results<br />

were compared with ship-track gravity data in the Taiwan Strait. The comparison of<br />

results demonstrated that waveform retracking has improved the quality of the SSH<br />

in coastal regions near the China Sea and Taiwan Strait, thus improving the accuracy<br />

of the altimeter-derived gravity field. The agreements between the predicted and<br />

ship-track gravity anomalies are ±13.9 mgal before retracking and ±9.9 mgal after<br />

retracking.<br />

The analysis of extracting the SSHs from altimeter waveform data suggests that the<br />

SSB correction is inaccurate when it is computed using the empirical model in the<br />

coastal region, due mainly to the incorrect SWH estimated from the on-satellite<br />

processor. Waveform retracking can remove the need for the ocean wave influence<br />

and thus reduce the SSB error, while without or using inaccurate SSB corrections<br />

causes the error to the unretracked SSH data. In addition, the method of averaging<br />

residual geoid gradients developed for the AWF data product in this study has been<br />

demonstrated as being effective. When applied this averaging method to the<br />

unretracked SSHs, an accuracy of ~ ±14 mgal can be obtained for the altimeterderived<br />

gravity anomalies, which is a similar level to that in Hwang et al. (2001).<br />

Results from this Chapter show a great potential of providing accurate gravity field<br />

information in coastal regions by using satellite altimeter SSH data after waveform<br />

retracking. For further research, to overcome the problem of altimeter data<br />

discontinuity very near to the coastline, combining the uniform coverage of the<br />

altimeter data with ship-borne gravity and land gravity data will provide an accurate<br />

high-resolution gravity field in coastal regions. It is also necessary to improve the un-


Chapter 7. IMPROVED COASTAL MARINE GRAVITY ANOMALIES 225<br />

modelled or poorly modelled geophysical corrections (e.g., SSB, wet tropospheric<br />

delay, sallow-water tides). However, this was not a specific topic of this research.


Chapter 8. CONCLUSIONS AND RECOMMENDATIONS 226<br />

8. CONCLUSIONS AND RECOMMENDATIONS<br />

8. 1 Summary of Dissertation<br />

Satellite radar altimetry is able to provide globally homogeneous and repeated sea<br />

surface information with very high accuracy and resolution. However, the accuracy<br />

of the satellite altimeter-based SSH degrades in coastal regions due to altimeter<br />

waveform contamination, inaccurate geophysical corrections, and enhanced ocean<br />

variability. Because of lower precision of the data in coastal areas, improvement of<br />

the quality of altimeter data is an important issue before using them for geodetic and<br />

oceanographic applications. When considering only the improvement of the altimeter<br />

data themselves, the approach used is to retrack waveforms to estimate the range<br />

correction, and then to derive the SSH in coastal regions. This approach has shown<br />

the altimeter’s potential in coastal regions. However, previous research on retracking<br />

procedures reprocess the coastal waveform using either merely an existing retracker,<br />

which is used to retrack data over non-ocean surfaces, or a non-automatic method<br />

that is hard to retrack a wide range of altimeter data near land. Compared with<br />

previous work, this is the first time that this research has carried out the waveform<br />

contamination in such a detailed way in a wide range of coastal regions.<br />

In an effort to precisely recover the altimeter-derived SSH data to the greatest extent<br />

close to the coastline, a dedicated coastal waveform retracking system has been<br />

proposed in this <strong>thesis</strong>. This system is developed based upon a systematic analysis of<br />

altimeter radar waveforms in Australian coastal regions. It has the advantage of<br />

containing several retrackers required to reprocess the waveform data automatically<br />

depending on the shapes of the waveform, while offering the precise SSH data sets<br />

derived from corrected range measurements that are estimated through adequately<br />

modelling the altimeter waveform at each along-track location in coastal regions.<br />

A coastal retracking system has been designed and developed (Chapter 5). Central to<br />

the system is the use of two retracking techniques to derive accurate geophysical<br />

parameters, which include both the iterative least squares fitting algorithm and the<br />

threshold retracking algorithm. Retrackers adopted in the system are the ocean fitting<br />

model, five-parameter fitting models, nine-parameter fitting models, and the varying


Chapter 8. CONCLUSIONS AND RECOMMENDATIONS 227<br />

threshold levels, in which five- and nine-parameter models are applied to the<br />

waveforms with both linear and exponential trailing edges. In order to obtain<br />

optimum and consistent estimates of the range correction, detailed investigations<br />

have been performed to test and select appropriate retrackers. A waveform<br />

classification procedure is also developed, which enables the waveform to be sorted<br />

and retracked by an appropriate retracker. To dear with the fading ‘noise’ in the<br />

waveforms, a weight iterative scheme has been developed.<br />

Data used are 20 Hz waveforms from two cycles of ERS-2 altimeter mission, five<br />

cycles of POSEIDON mission, and nearly 2.5 years of ERS-1 multi-missions. The<br />

AUSGeoid98 geoid model and ship-track gravity anomalies in the Taiwan Strait<br />

have been used as (partly) independent ground truth. In conjunction with the<br />

shoreline model, an Australian bathymetric grid, and a water/land grid, these data<br />

sets have been used in this research for the purpose of developing and implementing<br />

this system.<br />

The validation of the retracking algorithms has been tested using these data sources<br />

in both the Australian coast and the Taiwan Strait. Results from a number of<br />

comparisons and tests will be specifically concluded in Section 8.2. Generally<br />

speaking, regional investigations around the Australian coast have demonstrated that<br />

a waveform retracking system can effectively remove errors caused by waveform<br />

contamination in the area due to the presence of land, coastal sea states (e.g. inland<br />

or still water), and even the altimeter-operating feature. This study indicates that the<br />

altimeter-derived SSH data along a ground track, as supplied in GDR products, can<br />

be extended 5-15 km kilometres shoreward by the use of a coastal retracking system<br />

developed in this study.<br />

To investigate how retracked SSH data can improve the accuracy of altimeterderived<br />

coastal marine gravity fields in coastal regions, an application has been<br />

conducted using the ERS-1 20 Hz waveform data (


Chapter 8. CONCLUSIONS AND RECOMMENDATIONS 228<br />

averaging geoid gradients has been developed for the data used due to lack of the<br />

reference orbit information. Two altimeter-based gravity anomaly grids are created<br />

using the altimeter data before and after retracking. Results are then compared with<br />

the ship-track gravity data. Compared with the result of ±13.9 mgal before retracking,<br />

an improved accuracy of ±9.9 mgal is obtained after retracking. In addition, it has<br />

been found that the SSB correction is inaccurate in coastal regions because of the<br />

coastal waveform contamination.<br />

8. 2 Specific Conclusions<br />

This research has investigated the use of a coastal retracking procedure to reprocess<br />

satellite radar altimeter waveform data for extracting improved SSH data sets in<br />

coastal regions. The investigations conducted involve a quantitative estimation of the<br />

waveform-contaminated distance, the development and calibration of the coastal<br />

waveform retracking system, a comprehensive analysis of waveform data, the<br />

implementation of the retracking system, and applications of the system in Australian<br />

coastal regions and the Taiwan Strait. The key conclusions of the research are made<br />

below.<br />

8.2.1 The Quantitative Estimation of the Contaminated Distance<br />

Five cycles of POSEIDON and one cycle of ERS-2 20 Hz waveform data offshore<br />

0-350 km have been used to quantify a broad, contaminated boundary around the<br />

Australian coast (Chapter 4). This contamination is assumed to have been caused by<br />

backscatter from the land and the more variable coastal sea surface states, coupled<br />

with the footprint size of the radar altimeters, the direction that the satellite ground<br />

track crosses the coastline, and the operating feature of the radar altimeter. In<br />

addition, a preliminary experiment of land effects on the waveform has been<br />

performed using waveforms, the Australian 9"×9" DEM and the AUSGeoid98 2'×2'<br />

geoid model.<br />

Using the mean waveforms and standard deviations of the mean in different distance<br />

bands (Section 4.5.2), a typical contaminated boundary can be ascertained from the<br />

Australian shoreline to ~8 km and ~10 km for POSEIDON and ERS-2, respectively.<br />

Beyond 8 km or 10 km (to 350 km), the mean waveform shapes for both altimeters


Chapter 8. CONCLUSIONS AND RECOMMENDATIONS 229<br />

match the mean waveform shape observed over ocean surfaces, though they do not<br />

agree exactly with the theoretical waveform due to the fading noise and the wraparound<br />

error.<br />

Using the 50% threshold retracking point as the estimate of the pre-design tracking<br />

gate (Sections 4.5.2 and 4.5.3), three main conclusions can be drawn. First, the<br />

ERS-2 altimetry shows differences in contamination as a function of the sub-area<br />

along the Australian shoreline. The maximum contaminated distances vary from<br />

~8 km to ~22 km, depending upon locations and features of the shoreline. Second,<br />

both POSEIDON and ERS-2 altimeters show differences in contamination as a<br />

function of the distance from the shoreline, while ocean depths do not show an<br />

important correlation with the contaminated distance (Section 4.5.1). Finally, the<br />

contaminated distance for POSEIDON is less than that for ERS-2. A mean<br />

contaminated distance of ~8 km for POSEIDON and ~10 km for ERS-2 can also be<br />

observed using the 50% of threshold retracking points in the whole Australian<br />

coastal region.<br />

The ERS-2 altimeter also shows distance differences (Section 4.6.4) in waveform<br />

contamination depending on different manners of the ground track crossing the<br />

shoreline (i.e., approaching or leaving the land). In general, the waveforms along the<br />

altimeter ground track leaving land to water suffer longer contaminated distance than<br />

those along the track approaching land from water. The contaminated distance for a<br />

land-approaching track is ~7 km along the ground track for ERS-2. On the other<br />

hand for a satellite ground track leaving land to water, the along-track contaminated<br />

distance can be as long as ~21 km from the coastline. In both cases, the ocean<br />

surface sometimes can be no longer recoverable from contaminated waveforms<br />

obscured significantly by the land return, when the ground track is very closer to the<br />

coastline, say ~0-2.5 km.<br />

The reasons that cause the distance differences of waveform contamination can be<br />

the specific reflected topography (both land and sea) in coastal regions. From a<br />

ERS-2 waveform shifting test (Section 4.6.3), the result shows that altimeter’s<br />

operating feature should be one of reasons that cause the longer distance of distorted<br />

waveforms along-track leaving land to water. This type of distorted waveform


Chapter 8. CONCLUSIONS AND RECOMMENDATIONS 230<br />

usually shows usually the shift forward and backward with respect to the expected<br />

tracking gate in the range window, but do not significantly change the magnitude of<br />

the waveform’s amplitude. Land returns are not obviously visible in these<br />

contaminated waveforms, because the distance between the observing location and<br />

the land is usually beyond the satellite footprint. Using TOPEX waveform data,<br />

Brooks et al. (1997) find also the same problem that longer contaminated distances<br />

of the waveform occur upon the descending tracks. Thus, the result of the shifting<br />

test in this research can apply to not only the ERS-2 altimeter, but also TOPEX.<br />

Contamination leads to errors in the on-board calculation of the range measurements<br />

so that the accuracy of altimeter range measurements in such areas is lower and thus<br />

presents the problem for extending the use of altimeter data to the shoreline. It is<br />

recommended from the quantitative contaminated distance in this study that the SSH<br />

data within the distance provided by the GDR products are not reliable and must be<br />

detected prior to their being included in geodetic and oceanographic solutions in<br />

coastal regions. It is also suggested that waveform data reprocessing is essential to<br />

improve altimeter-derived SSH data in these areas. For most GDRs users without<br />

waveform products, it is recommended that altimeter SSH or range data should be<br />

interpreted with some caution for distances less than say ~22 km from a coastline,<br />

and discarded altogether for distances less than 5 km.<br />

8.2.2 Practical Implementation of the Coastal Waveform Retracking System<br />

Regionally comprehensive investigations of waveform characteristics and validation<br />

of the coastal retracking system around the Australian coast have demonstrated that<br />

diverse waveform shapes exist, and the retracking of 20 Hz waveform data can<br />

effectively remove most errors caused by waveform contamination in these areas.<br />

Two cycles of ERS-2 20 Hz waveform data sets offshore 30 km around Australian<br />

coastal regions have been used for implementation of the coastal retracking system<br />

developed in this study. The AUSGeoid98 model was again used as a ground truth<br />

for the purpose of comparison. In an effort to improve the altimeter-derived SSH<br />

results, the following conclusions can be made:<br />

(1) Results of waveform classification show that most coastal waveforms (80.19%)<br />

have ocean-like shapes with a single ramp but without significant peaks. Of the


Chapter 8. CONCLUSIONS AND RECOMMENDATIONS 231<br />

remaining ~20% of waveforms, predominant waveforms have single prepeaked<br />

(4.65%) and high-peaked (4.35%) waveform shapes. Others, such as<br />

single mid-peaked and post-peaked, double pre-peaked and post-peaked, and<br />

multi-peaked waveforms, take the percentages of 0.73% - 1.35%. Although<br />

there is only a small percentage of non-ocean-like waveforms, the statistical<br />

results show that these waveforms are much closer to the coastline (1.73-3.12<br />

km in average), indicating that a retracking system containing several<br />

retrackers is necessary.<br />

(2) Retracking results in Australian coastal regions show that 77.7% - 96.2% of<br />

waveforms can be retracked using a least squares iterative fitting algorithm,<br />

while 3.8% - 18.1% of waveforms have to be retracked by the threshold<br />

retracking algorithm. The waveforms show differences in percentages of using<br />

different retrackers as a function of the sub-area along ground tracks. Of these<br />

waveforms retracked by the fitting algorithm, over 98.3% of waveforms are<br />

fitted using the ocean model and the rest of waveforms are fitted by the fiveparameter<br />

model with an exponential trailing edge and the nine-parameter<br />

model with either a linear trailing edge or an exponential trailing edge. It has<br />

been found that the use of the threshold retracker decreases with increasing<br />

offshore distance from 35% in an offshore distance of 0-5 km to 4% in a<br />

offshore distance of 5-10 km. After 10 km offshore, only 2% of waveforms are<br />

threshold retracked, which agrees well with the estimation of the contaminated<br />

distance in Chapter 4 and indicates that the method developed to categorise<br />

coastal waveforms is effective.<br />

(3) Results from single track analysis indicate that the coastal retracking system<br />

has a capability of dealing with diverse waveforms in coastal regions. This<br />

advantage means that the system derives more accurate SSH data than a single<br />

retracker. By comparing altimeter-derived along-track SSH data before and<br />

after retracking with AUSGeoid98 geoid heights, results show that the<br />

altimeter-derived SSH profiles can be extended ~5-10 km shoreward by the use<br />

of coastal retracking system and range corrections. However, SSH profiles at a<br />

distance of 0-5 km to the coastline cannot be recovered by waveform<br />

retracking procedure at the same precision level as SSH data after 5 km from


Chapter 8. CONCLUSIONS AND RECOMMENDATIONS 232<br />

the coastline. This is most probably due to predominant land returns in the<br />

waveform range window and the incorrect geophysical corrections in these<br />

near land areas.<br />

In summary, the effect of a coastal waveform retracking system, which contains a set<br />

of methods, computing programs and retrackers for reprocessing waveform data, has<br />

been demonstrated by the improvement observed in Australian coastal regions. When<br />

using two cycles of ERS-2 20 Hz waveform data and the AUSGeoid98 geoid grid as<br />

an independent ground reference, results after retracking provide a typical<br />

improvement of ~5 km on the offshore distance. Compared with the contaminated<br />

distance of ~10 km before retracking, the use of the retracking system is able to<br />

reduce this contaminated distance to ~5 km.<br />

8.2.3 Development and Validation of a Coastal Retracking System<br />

The development and validation of the coastal retracking system includes design and<br />

investigation of algorithms, selection of the threshold levels, estimation of biases<br />

between different retrackers and the SSH data before and after retracking, as well as<br />

the tests on the effectiveness of the system.<br />

(d) Retracker Investigation<br />

An ocean return model has been deduced and is a modified convolutional solution of<br />

the radar returns based upon the Brown (1977) model. This ocean model without the<br />

non-linear ocean wave parameter has been used as a main retracker in this research<br />

due to its clearly physical description of the ocean surface. By quantitative<br />

comparison to the five-parameter model, it has found that the slope of the trailing<br />

edge modelled by the five-parameter model has a larger span of variation than the<br />

slope modelled by the ocean model. This advantage makes, in turn, the fiveparameter<br />

model hold a greater capability of fitting non-ocean-like or irregular<br />

waveforms in coastal regions than the ocean model. Thus, the five-parameter model<br />

replaces the ocean model for these irregular shapes of waveforms in the<br />

implementation of the system.<br />

Selection of the threshold level has been performed by comparing an available<br />

ground truth of the AUSGeoid98 geoid heights with the SSH data before and after


Chapter 8. CONCLUSIONS AND RECOMMENDATIONS 233<br />

threshold retracking, in ten 15°×15° sub-areas around the Australian coastal regions<br />

and six 5 km wide distance bands. Results confirm that the 50% threshold level is the<br />

best level for open-ocean waveform, but not an adequate level for the contaminated<br />

ocean waveforms in coastal regions. Instead, a varying threshold level, which is the<br />

30% threshold level for contaminated waveforms and 50% threshold level for<br />

uncontaminated waveforms, can give a good agreement between the retracked SSH<br />

data and the geoid heights. This result of the threshold level broadly agrees with that<br />

used by Brooks et al. (1997).<br />

(b) Bias Evaluation<br />

Biases exist both between retracked SSH data from different retracking algorithms<br />

and SSH data before and after retracking. Biases have been analysed and estimated<br />

by retracking ocean waveforms in Section 5.8 using two cycles of ERS-2 waveforms<br />

and can be concluded as follows, which suggest that biases must be estimated before<br />

retracking and applied to the data to obtain consistent results after retracking.<br />

(1) Bias between the SSH data retracked by the ocean and five-parameter fitting<br />

algorithms is −0.85±1.35 cm, which is not statistically significant.<br />

(2) Bias between the SSH data retracked by the 50% threshold retracker and the<br />

fitting algorithms (both the ocean and five-parameter models) is<br />

+56.35±5.84 cm.<br />

(3) Biases between altimeter-derived SSH data before and after retracking are<br />

+27.58±8.78 cm for fitting algorithms and −30±3.58 cm for the 50% threshold<br />

retracking algorithm.<br />

It should be pointed out that the differences between the standard deviations are<br />

partly due to the different numbers of the waveform data used for the comparison.<br />

This means that it is not reliable to determine which retracker is better though a<br />

direct comparison of the STDs. Thus, an independently external ocean surface<br />

reference is necessary to estimate absolute biases precisely. However, as any<br />

gravimetric determination of the geoid is deficient in the zero- and first-degree terms<br />

(Heiskanen and Moritz, 1967; Featherstone et al., 1996), the geoid heights<br />

interpolated from AUSGeoid98 are not taken as a reference to ascertain the absolute<br />

bias between the SSH before and after retracking.


Chapter 8. CONCLUSIONS AND RECOMMENDATIONS 234<br />

For the biases existing between the SSH data before and after retracking, however,<br />

there is alternative approach that can be adopted to remove them. Because the bias is<br />

caused by the instrument and the on-board tracking algorithm (Hayne and Hancock<br />

III, 1990; Chelton, 2001; Hayne, 2002), it can be corrected from the instrument<br />

correction. The correction can be computed by a ground data processing system or<br />

by waveform retracking. The recent research of Dong et al. (2002) show that the<br />

absolute bias can also be determined using altimeter data and GPS levelling at tide<br />

gauges.<br />

(c) Algorithm Validation<br />

To test how effective the retracker is, retracking using only a single retracker has<br />

been conducted in two distinct coastal areas of Australia and the Taiwan Strait. The<br />

AUSGeoid98 geoid heights and the ship track gravity anomalies are used as external<br />

input data sets of the ground truth.<br />

When comparing the geoid heights with the SSH data in Australian coastal regions<br />

(Section 5.8.2), the emphasis is on the shapes of the height profiles and the changes<br />

of the along-track slope. Results from single track comparison show that the SSH<br />

gradients in the vicinity of the land before retracking are large varying from<br />

~363.8-975.6 mm/km over offshore along-track distances of 0-19.5 km, being much<br />

larger than the maximum geoid gradient value of ~150 mm/km in Australia (cf.,<br />

Johnston and Featherstone, 1998; Friedlieb et al., 1997; Featherstone et al., 2001).<br />

After retracking, the SSH gradients decrease by ~16.0 mm/km over offshore alongtrack<br />

distance of 4.7 km − 19.5 km, agreeing well with the geoid gradient<br />

(~16.4 mm/km) at the same distance. Results indicate strongly that retracking, no<br />

matter which retracker is used, can improve the altimeter-derived SSH in coastal<br />

regions.<br />

Using ~2.5 years of ERS-1 SSH data sets before and after retracking, gravity<br />

anomaly predictions in the Taiwan Strait are made from the inverse Vening Meinesz<br />

formula (Hwang, 1998). Retrackers used are the fitting algorithm including the ocean<br />

and five-parameter models. Results are then compared with limited ship-track gravity<br />

data. The comparison of results demonstrated that waveform retracking has improved<br />

the quality of the SSH in coastal regions near the China Sea and Taiwan Strait, thus


Chapter 8. CONCLUSIONS AND RECOMMENDATIONS 235<br />

improving the accuracy of the altimeter-derived gravity field. Improved residual<br />

geoid gradients and later on gridded gravity anomaly estimation are obtained in this<br />

coastal region when the SSHs after retracking are used instead of unretracked SSH<br />

data. The STD of agreements between the predicted and ship-track gravity anomalies<br />

are ±13.9 mgal before retracking and ±9.9 mgal after retracking. This shows a great<br />

potential of providing accurate gravity field information in coastal regions by using<br />

satellite altimeter SSH data after waveform retracking.<br />

8.2.4 Other Conclusions<br />

A preliminary test has been performed to calculate the retracking gate estimate<br />

related to the land returns in the range window using external independent data<br />

source of the DEM model in coastal regions. Analysing results show that nonuniqueness<br />

of the estimate over land near coastal regions is still a main problem that<br />

should be solved in the future.<br />

The analysis on extracting the SSHs from altimeter waveform data suggests that the<br />

SSB correction is inaccurate when it is computed using the empirical model in the<br />

coastal region due mainly to the incorrect SWH estimated from the on-satellite<br />

processor. Waveform retracking can remove the need for the ocean wave influence<br />

and thus reduce the SSB error, while an error is caused to the unretracked SSH data<br />

by the need of this correction no matter whether or not the SSB is applied. In<br />

addition, the method of averaging residual geoid gradients developed for the AWF<br />

data product in this study has been demonstrated being effective, because a similar<br />

accuracy level (~±14 mgal) as Hwang et al. (2001) is obtained using unretracked<br />

altimeter data.<br />

8. 3 Recommendations for Future Work<br />

The improvement of the altimeter SSH data in Australian waters and the Taiwan<br />

Strait has demonstrated that the coastal waveform retracking system developed in<br />

this research has the potential of being applied to other coastal regions, as well as<br />

other satellite altimeter waveform data sets. With higher precision of the altimeter<br />

measurements over open oceans, demand for improving the precision of the data in<br />

coastal regions will become increasingly desirable to realise the global coverage of


Chapter 8. CONCLUSIONS AND RECOMMENDATIONS 236<br />

the altimetric missions. Therefore, recommendations and further work will be given<br />

in this Section.<br />

8.3.1 Extraction of Ocean Returns from the Mixed Waveform Shape<br />

Current retracking algorithms including those in the coastal system cannot recover<br />

the ocean returns that appear in the late gates behind the land returns in the range<br />

window when the ground track is very much closer to the coastline (Brooks et al.,<br />

1997). However, if the land return in the waveform can be estimated and removed, it<br />

should be possible to recover the ocean return. As stated, this can be achieved only<br />

when other external and high-resolution reflected surface topography are available.<br />

With new satellite altimetry missions, improved DEM created from satelliteobserved<br />

data, and detailed topography information, developing new waveform<br />

retrackers for this type of waveform will become an area requiring further research<br />

effort.<br />

8.3.2 Improvement of the Correction Algorithms near Land<br />

This research has demonstrated that waveform contamination can be removed in a<br />

certain offshore distance using waveform retracking techniques. However, waveform<br />

retracking can improve only the range measurements. The incorrect geophysical and<br />

atmospheric corrections still affect the precision of the altimeter-derived SSH data<br />

sets, which are the direct measurements in geodetic applications. Improvement of the<br />

corrections in coastal regions has received attention from several researchers, such as<br />

Shum et al. (1998) and Anzenhofer et al. (2000). To fulfil the satellite altimetric<br />

potential in coastal regions, developing geophysical and environmental correction<br />

algorithms will be another future research field.<br />

As concluded in this research, the SSB correction causes a large error in the SSH<br />

data near the coastline. In general, the SSH data before used for geodetic applications<br />

must apply an instrument correction to remove the effects of the SSB. However, how<br />

can the SSH be improved by such incorrect SSB correction in coastal regions?<br />

Fortunately, retracking procedures also removes the need for the SSB correction.<br />

Therefore, retracking in a wide range of coastal regions will also be one for future<br />

research.


Chapter 8. CONCLUSIONS AND RECOMMENDATIONS 237<br />

8.3.3 Improvement of the Altimeter Operation<br />

From this research, the operation of the altimeter affects the waveform, in particular<br />

waveforms along ground tracks leaving land to water. This result suggests that it is<br />

essential to improve the on-board tracking algorithm based on this operating feature<br />

or the design of the altimeter operation.<br />

8.3.4 Geodetic Application<br />

For further research of determining gravity anomalies in coastal regions, and to<br />

overcome the problem of altimeter data discontinuity very near to the coastline,<br />

combining the uniform coverage of the altimeter data with shipborne gravity and<br />

land gravity data will provide a very accurate high-resolution gravity field in coastal<br />

regions.


238<br />

REFERENCES<br />

Andersen, O. B. and Knudsen, P. (1998) Global Marine Gravity from the ERS-1 and<br />

Geosat Geodetic Mission Altimetry, Journal of Geophysical Research, Vol.<br />

103, pp. 8129-8137.<br />

Andersen, O. B. and Knudsen, P. (2000) The Role of Satellite Altimetry in Gravity<br />

Field Modelling in Coastal Areas, Physics and Chemistry of the Earth (A), Vol.<br />

25, No. 1, pp. 17-24.<br />

Anzenhofer, M., Shum, C. K. and Renstch, M. (2000) Coastal Altimetry and<br />

Applications, The Ohio State <strong>University</strong>, Columbus, USA, 36 pp.<br />

Andersen, O. B., Knudsen, P. and Trimmer, R. (2001) The KMS2001 Global Mean<br />

Sea Surface and Gravity Field, Proceedings of the IAG2001 Scientific<br />

Assembly, Budapest, Hungary, September [CD-ROM].<br />

AVISO/Altimeter (1996) AVISO User Handbook: Merged TOPEX/POSEIDON<br />

Products, AVI-NT-020101-CN, Edition 3.0, Toulouse, France, 196 pp.<br />

AVISO/Altimeter (1998) Preliminary AVISO User Handbook: POSEIDON<br />

Waveform Product, AVI-NT-02-555-CN, Edition 1.0, Toulouse, France, 101 pp.<br />

Bamber, J. L. (1994) A Digital Elevation Model of the Antarctic Ice Sheet Derived<br />

from ERS-1 Altimeter Data and Comparison with Terrestrial Measurements,<br />

Annals of Glaciology, Vol. 20, pp. 48-54.<br />

Bamber, J. L. (1994) Ice Sheet Altimeter Processing Scheme, International Journal<br />

of Remote Sensing, Vol. 15, No. 4, pp. 925-938.<br />

Bamber, J. L., Ekholm, S. and Krabill, W. (2001) A new, high-resolution digital<br />

elevation model of Greenland fully validated with airborne laser altimeter<br />

data, Journal of Geophysical Research - Solid Earth, Vol. 106, No. B4, pp.<br />

6733-6746.<br />

Barrick, D. E. and Lipa, B.J. (1985) Analysis and Interpretation of Altimeter Sea<br />

Echo, Advances in Geophysics, Vol. 27, pp. 61-100.<br />

Berry, P. A. M., Sanders, R. F., Leenmans, C. and Bron, E. (1998) Generating<br />

Orthometric Heights from the ERS-1 Altimeter Geodetic Mission Dataset:<br />

Results From an Expert Systems Approach, in: Geodesy on the Move, Schward,<br />

K. (ed.), Springer Verlag, Berlin, Germany, pp. 113-118.<br />

Berry, P. A. M. (2000a) Topography from Land Radar Altimeter Data: Possibilities<br />

and Restrictions, Physics and Chemistry of the Earth (A), Vol. 25, No. 1, pp.<br />

81-88.


239<br />

Berry, P. A. M., Hoogerboord, J.E. and Pinnock, R.A. (2000b) Identification of<br />

Common Error Signatures in Global Digital Elevation Models Based on<br />

Satellite Altimeter Reference Data, Physics and Chemistry of the Earth (A),<br />

Vol. 25, No. 1, pp. 95-99.<br />

Berry, P. A. M. (2001) Personal Communication, Faculty of Computing and<br />

Engineering, De Montfort <strong>University</strong>, U.K.<br />

Berry, P. A. M. (2002) Personal Communication, Faculty of Computing and<br />

Engineering, De Montfort <strong>University</strong>, U.K.<br />

Bracewell, R. N. (2000) The Fourier Transform and Its Applications (Third Edition),<br />

McGraw Hill, Boston, USA, 597 pp.<br />

Brenner, A. C., Bindschadler, R. A., Thomas, R. H. and Zwally, H. J. (1983) Slope-<br />

Induced Errors in Radar Altimetry over Continental Ice Sheets, Journal of<br />

Geophysical Research, Vol. 88, No. C3, pp. 1617-1623.<br />

Brenner, A. C., Frey, H. V. and Zwally, H. J. (1990) Comparisons between GEOSAT<br />

and SEASAT Tracking over Non-Ocean Surfaces, Geophysical Research<br />

Letters, Vol. 17, No. 10, pp. 1537-1540.<br />

Brenner, A. C., Koblinsky, C. J. and Zwally, H. J. (1993) Postprocessing of Satellite<br />

Return Signals for Improved Sea Surface Topography Accuracy, Journal of<br />

Geophysical Research, Vol. 98, No. C1, pp. 933-944.<br />

Brenner, A., Frey, H., Dimarzio, J., and Tsaoussi, L. (1997) Topography over South<br />

America from ERS, http://earth.esa.int/symposia/papers/frey/ [last accessed: 8<br />

March 2004].<br />

Brooks, R. L., Campbell, W. J., Ramseier, R. O., Stanley, H. R. and Zwally, H. J.<br />

(1978) Ice Sheet Topography by Satellite Altimetry, Nature, Vol. 274, No.<br />

5671, pp. 539-543.<br />

Brooks, R. L. and Lockwood, D. W. (1990) Effects of Islands in the Geosat Footprint,<br />

Journal of Geophysical Research, Vol. 95, No. C3, pp. 2849-2855.<br />

Brooks, R. L., Lockwood, D. W. and Lee, J. E. (1997) Land Effects on TOPEX<br />

Radar Altimeter Measurements on Pacific Rim Coastal Zones,<br />

http://topex.wff.nasa.gov/docs.html [last accessed: 8 March 2004].<br />

Brooks, R. L. (2002) Personal Communication, Computer Sciences Corporation,<br />

Wallops Island, VA 23337, USA.<br />

Brown, G. S. (1977) The Average Impulse Response of a Rough Surface and Its<br />

Applications, IEEE Transactions on Antennas and Propagation, Vol. AP-25,<br />

No. 1, pp. 67-74.<br />

Buchanan, C. (1991) Bathymetric 30 Arc Second Grid of the Australian Region,<br />

Geoscience Australia, Canberra, Australia [CD-ROM].


240<br />

Challenor, P. G. and Srokosz, M. A. (1989) The Extraction of Geophysical<br />

Parameters from Radar Altimeter Return from a Nonlinear Sea Surface, in:<br />

Mathematics in Remote Sensing, Brooks, S. R. (ed.), Oxford <strong>University</strong> Press,<br />

Oxford, UK, pp. 254-268.<br />

Chelton, D. B., Ries, J.C., Haines, B. J., Fu, L.-L. and Callahan, P.S. (2001) Satellite<br />

Altimetry, in: Satellite Altimetry and Earth Sciences: A Handbook of<br />

Techniques and Applications, Fu, L.-L. and Cazenave, A. (eds.), Academic<br />

Press, San Diego, USA, pp. 1-132.<br />

Cheney, R. E. and Marsh, J. G. (1983) Global Mesoscale Variability from Collinear<br />

Tracks of SEASAT Altimeter Data, Journal of Geophysical Research, Vol. 88,<br />

pp. 4343-4354.<br />

Cooper, A. P. R. (1989) Slope Correction by Relocation for Satellite Radar Altimetry,<br />

International Geoscience and Remote Sensing Symposium, Vancouver, Canada,<br />

July, pp. 2730-2733.<br />

Cudlip, W. and Milnes, M. (1994) Overview of Altimeter Data Processing at the<br />

EODC, International Journal of Remote Sensing, Vol. 15, No. 4, pp. 871-887.<br />

Davis, C. H. (1993a) A Surface and Volume Scattering Retracking Algorithm for Ice<br />

Sheet Satellite Altimetry, IEEE Transactions on Geoscience and Remote<br />

Sensing, Vol. 31, No. 4, pp. 811-818.<br />

Davis, C. H. and Moore, R. K. (1993b) A Combined Surface- and Volume-Scattering<br />

Model for Ice-Sheet Radar Altimetry, Journal of Glaciology, Vol. 39, No. 133,<br />

pp. 675-686.<br />

Davis, C. H. (1995) Growth of the Greenland Ice Sheet: A Performance Assessment<br />

of Altimeter Retracking Algorithms, IEEE Transactions on Geoscience and<br />

Remote Sensing, Vol. 33, No. 5, pp. 1108-1116.<br />

Davis, C. H. (1997) A Robust Threshold Retracking Algorithm for Measuring Ice-<br />

Sheet Surface Elevation Change from Satellite Radar Altimeter, IEEE<br />

Transactions on Geoscience and Remote Sensing, Vol. 35, No. 4, pp. 974-979.<br />

Deng, X., Chao, D. B., Li, J. C. and Chen, J. Y. (1996) A Preliminary Process of<br />

TOPEX/POSEIDON Data in the China Sea and Neighbour Sea, ACTA<br />

Geodaetica et Cartographica Sinica, Vol. 3, pp. 226-321.<br />

Deng, X., Chao, D.B. and Li, J. C. (1997) A Preliminary Estimate of the Mean Sea<br />

Level Variations in the China Sea from TOPEX/POSEIDON Data,<br />

Proceedings of the IAG International Conference of Crustal Movement and<br />

Hazard Reduction in East Asia and South-East Asia, Wuhan, China, November,<br />

pp. 509-516.<br />

Deng, X., Featherstone, W. E., Hwang, C. and Berry, P. A. M. (2001) Retracking<br />

ERS-2 Radar Altimeter Waveform Data in Australia's Coastal Zone,


241<br />

Proceedings of the IAG 2001 Scientific Assembly, Budapest, Hungary,<br />

September [CD-ROM].<br />

Deng, X., Featherstone, W. E., Hwang, C. and Berry, P .A. M. (2002) Estimation of<br />

Contamination of ERS-2 and POSEIDON Satellite Radar Altimetry Close to<br />

the Coasts of Australia, Marine Geodesy, Vol. 25, No. 4, pp. 249-271.<br />

Dong, X., Moore, P. and Bingley, R. (2002) Absolute Calibration of the<br />

TOPEX/POSEIDON Altimeter Using UK Tide Gauges, GPS, and Precise,<br />

Local Geoid-Differences, Marine Geodesy, Vol. 25, pp. 189-204.<br />

Dunbar, P. (2000) DS759.2 – Terrain Base, Global 5-minute Ocean Depth and Land<br />

Elevation, from NGDC, http://dss.ucar.edu/datasets/ds759.2/ [last accessed: 8<br />

March 2004].<br />

Eanes, R. (1999) Improved Ocean Tide Model from Satellite Altimetry, Fall Meeting<br />

1999, American Geophysical Union, San Francisco, USA.<br />

Fairhead, J. D., Green, C.M. and Odegard, M. E. (2001) Satellite-Derived Gravity<br />

Having an Impact on Marine Exploration, The Leading Edge. Vol. 20, pp. 873-<br />

876.<br />

Featherstone, W. E., Alexander, K. and Sideris, M. G. (1996) Gravimetric Geoid<br />

Refinement Using High Resolution Gravity and Terrain Data, Geomatics<br />

Research Australasia, No. 64, pp. 75-99.<br />

Featherstone, W. E., Kirby, J. F., Kearsley, A. H. W., Gilliland, J. R., Johnston, G.<br />

M., Steed, J., Forsberg, R. and Sideris, M. G. (2001) The AUSGeoid98 Geoid<br />

Model of Australia: Data Treatment, Computations and Comparisons with<br />

GPS-Levelling Data, Journal of Geodesy, Vol. 75, pp. 313-330.<br />

Featherstone, W. E. (2003) Comparison of Different Satellite Altimeter-Derived<br />

Gravity Anomaly Grid with Ship-Borne Gravity Data around Australia, in:<br />

Gravity and Geoid 2002, Tziavos, I.N. (ed.), Department of Surveying and<br />

Geodesy, Aristotle <strong>University</strong> of Thessaloniki, pp. 326-331.<br />

Fedor, L. S., Godbey, T. W., Gower, J .F. R., Guptill, R., Hayne, G. S., Rufenace, C.<br />

L. and Walsh, E. J. (1979) Satellite Altimeter Measurements of Sea States - An<br />

Algorithm Comparison, Journal of Geophysical Research, Vol. 84, No. B8, pp.<br />

3991-4001.<br />

Femenias, P., Remy, F., Raizonville, R. and Minster, J. F. (1993) Analysis of<br />

Satellite-Altimeter Height Measurements above Continental Ice Sheets,<br />

Journal of Glaciology, Vol. 39, No. 133, pp. 591-600.<br />

Fernandes, M. J., Bastos, L. and Antunes, M. (2003) Coastal Satellite Altimetry -<br />

Methods for Data Recovery and Validation, in: Gravity and Geoid 2002,<br />

Tziavos, I.N. (ed.), Department of Surveying and Geodesy, Aristotle<br />

<strong>University</strong> of Thessaloniki, pp. 302-307.


Frey, H. and Brenner, A. C. (1990) Australian Topography From Seasat Overland<br />

Altimetry, Geophysical Research Letters, Vol. 17, No. 10, pp. 1533-1536.<br />

242<br />

Friedlieb, O. J., Featherstone, W. E. and Dentith, M. C. (1997) A WGS84-AHD<br />

Profile over the Darling Fault, Western Australia, Geomatics Research<br />

Australasia, Vol. 67, pp. 17-32.<br />

Guzkowska, M. A. J., Rapley, C. G., Cudlip, W. and Mason, I. M. (1988) Satellite<br />

Radar Altimetry over Arid Regions, Proceedings of IGARSS'88 Symposium,<br />

Edinburgh, Scotland, September, pp. 659- 662.<br />

Hayne, G. S. (1980) Radar Altimeter Mean Return Waveform from Near-Normal-<br />

Incidence Ocean Surface Scattering, IEEE Transactions on Antennas and<br />

Propagation, Vol. AP-28, No. 5, pp. 687-692.<br />

Hayne, G. S. and Hancock III, D. W. (1982) Sea-State Related Altitude Errors in the<br />

Seasat Altimeter, Journal of Geophysical Research, Vol. 87, No. C5, pp. 3227-<br />

3231.<br />

Hayne, G. S. and Hancock III, D. W. (1990) Corrections for the Effects of<br />

Significant Wave Height and Attitude on Geosat Radar Altimeter<br />

Measurements, Journal of Geophysical Research, Vol. 95, No. C3, pp. 2837-<br />

2842.<br />

Hayne, G. S., Hancock III, D. W. and Purdy, C. L. (1994) The Corrections for<br />

Significant Wave Height and Attitude Effects in the TOPEX Radar Altimeter,<br />

Journal of Geophysical Research, Vol. 99, No. C12, pp. 24941-24955.<br />

Hayne, G. S. (2002) Personal Communication, NASA Goddard Space Flight Centre,<br />

Wallops Island, Virginia, USA.<br />

Heiskanen, W. A. and Moritz, H. (1980) Physical Geodesy, W.H. Freeman, San<br />

Francisco, USA, 364 pp.<br />

Hilton, R. D., Featherstone, W. E., Berry, P. A. M., Johnston, C.P.D. and Kirby, J.F.<br />

(2003) Comparison of Digital Elevation Models (DEMs) over Australia and<br />

External Validation Using ERS-1 Satellite Radar Altimetry, Australian<br />

Journal of Earth Sciences, Vol. 50, No. 2, pp. 157-168.<br />

Hsu, S., Liu, C., Shyu, C. Liu, S., Sibue, J., Lallemand, S., Wang, C. and Reed D.<br />

(1998) New Gravity and Magnetic Anomaly Maps in the Taiwan-Luzon<br />

Region and Their Preliminary Interpretation, TAO, Vol. 9, No. 3, pp. 509-532.<br />

Hutchinson, M., Stein, J. and Stein, J. (2001) Upgrade of the 9 Second Australian<br />

Digital Elevation Model, Centre for Resource and Environmental Studies,<br />

Australian National <strong>University</strong>, Canberra, http://cres.anu.edu.au/dem/ [last<br />

accessed: 8 March 2004].<br />

Hwang, C. and Parsons, B. (1995) Gravity Anomalies Derived from Seasat, Geosat,<br />

ERS-1 and TOPEX/POSEIDON Altimetry and Ship Gravity: A Case Study


243<br />

over the Reykjanes Ridge, Geophysical Journal International, Vol. 122, pp.<br />

551-568.<br />

Hwang, C. (1998a) Inverse Vening Meinesz Formula and Deflection-Geoid Formula:<br />

Applications to the Predictions of Gravity and Geoid over the South China Sea,<br />

Journal of Geodesy, Vol. 72, pp. 304-312.<br />

Hwang, C., Kao, E. and Parsons, B. (1998b) Global Derivation of Marine Gravity<br />

Anomalies from Seasat, Geosat, ERS-1 and TOPEX/POSEIDON Altimeter<br />

Data, Geophysical Journal International, Vol. 134, pp. 449-459.<br />

Hwang, C. and Wang, C. (2002a) New Free-Air Gravity Anomaly Grid of Taiwan,<br />

Journal of Surveying Engineering, Vol. 44, No. 2, pp. 1-22 (in Chinese).<br />

Hwang, C., Hsu, H. and Jang, R. (2002b) Global Mean Sea Surface and Marine<br />

Gravity Anomaly from Multi-Satellite Altimetry: Applications of Deflection-<br />

Geoid and Inverse Vening Meinesz Formulae, Journal of Geodesy, Vol. 76, pp.<br />

407-418.<br />

Jeffrey, A. (1995) Handbook of Mathematical Formulas and Integrals, Academic<br />

Press, New York, USA, 410 pp.<br />

Jensen, J. R. (1999) Radar Altimeter Gate Tracking: Theory and Extension, IEEE<br />

Transactions on Geoscience and Remote Sensing, Vol. 37, No. 2, pp. 651-658.<br />

Johnston, G. M., and Featherstone, W. E. (1998) AUSGeoid98: A New Gravimetric<br />

Geoid for Australia, Australia's National Mapping Agency, Canberra, Australia,<br />

http://www.ga.gov.au/nmd/geodesy/ausgeoid/docs/iemsgary.pdf [last accessed:<br />

8 March 2004].<br />

JPL (1993a) TOPEX/POSEIDON Project, TOPEX Ground System Software<br />

Interface Specification, Volume 2: Design (SIS-2), Altimeter Sensor Data<br />

Record (SDR)-Alt SDR Data, 633-751-23-001, Rev.C, Jet Propulsion<br />

Laboratory, California Institute of Technology, Pasedena, USA, 77 pp.<br />

JPL (1993b) TOPEX/POSEIDON Project, TOPEX Ground System, Software<br />

Interface Specifications Volume 2: Design (SIS-2), Altimeter Sensor Data<br />

Record (SDR)-Alt SDR, 633-752-23-021, Rev.B, Jpl Propulsion Laboratory,<br />

California Institute of Technology, Pasedena, USA, 38 pp.<br />

Kelleher, G., Bleakley, G., Walls, K. and Dingwall, P. (2002) Marine Region 18:<br />

Australia/New Zealand,<br />

http://www.ea.gov.au/coasts/mpa/nrsmpa/global/volume4/pubs/v4chap18.pdf<br />

[last accessed: 8 March 2004].<br />

Kirby, J. F. (1996) The Development and Application of a New Algorithm for Ocean<br />

Geoid Recovery, PhD Thesis, <strong>University</strong> of Edinburgh, UK, 156 pp.<br />

Kirby, J. F. and Forsberg, R. (1997) A Comparison of Techniques for the Integration<br />

of Satellite Altimeter and Surface Gravity Data for Geoid Determination, in:


244<br />

Geodesy on the Move, Forsberg, R., Feissel, M. and Dietrich, R. (eds.),<br />

Springer-Verlag, Berlin, Germany, pp. 207-212.<br />

Leben, R. R., Born, G. H., Thompson, J. D. and Fox, C. A. (1990) Mean Sea Surface<br />

Variability of the Gulf of Mexico Using Geosat Altimeter Data, Journal of<br />

Geophysical Research, Vol. 95, pp. 3025-3032.<br />

Legresy, B. and Remy, F. (1998) Using the Temporal Variability of Satellite Radar<br />

Altimetric Observations to Map Surface Properties of the Antarctic Ice Sheet,<br />

Journal of Glaciology, Vol. 44, No. 147, pp. 197-206.<br />

Lemoine, F. G., Kenyon, S.C., Factor, J.K., Trimmer, R.G., Pavlis, N.K., Chinn,<br />

D.S., Cox, C.M., Klosko, S.M., Luthcke, S.B., Torrence, M.H., Wang, Y.M.,<br />

Williamson, R.G., Pavlis, E.C., Rapp, R.H., Olson, T.R. (1998) The<br />

Development of the Joint NASA GSFC and the National Imagery and Mapping<br />

Agency (NIMA) Geopotential Model EGM96, NASA/TP-1998-206861, NASA<br />

Technical Paper, USA, 575 pp.<br />

Le Traon, P. Y. and Morrow, R. (2001) Ocean Currents and Eddies, in: Satellite<br />

Altimetry and Earth Sciences: A Handbook of Techniques and Applications, Fu,<br />

L.-L. and Cazenave, A. (eds.), Academic Press, San Diego, pp. 171-216.<br />

Levitus, S., Monterey, G. I. and Boyer, T. (1997) Seasonal Variability of Dynamic<br />

Height and Its Fourier Analysis, NOAA NESDIS Atlas 15, Government Printing<br />

Office, Washington, D.C., USA.<br />

Levitus, S., Monterey, G. I. and Boyer, T. (2002) Seasonal Variability of Dynamic<br />

Height and Its Fourier Analysis, http:/www.nodc.noaa.gov/OC5/dyndoc.html<br />

[last accessed: 8 March 2004].<br />

Li, D. R. (1988) The Error and Reliability Theory, Surveying and Mapping Publisher,<br />

Beijing, China, 331 pp.<br />

Lipa, B. J. and Barrick, D. E. (1981) Ocean Surface Height-Slope Probability<br />

Density Function from SEASAT Altimeter Echo, Journal of Geophysical<br />

Research, Vol. 86, No. C11, pp. 10921-10930.<br />

McAdoo, D. C. (1990) Gravity Field of the Southeast Central Pacific from Geosat<br />

Exact Repeat Mission Data, Journal of Geophysical Research, Vol. 97, No. B3,<br />

pp. 3274-3260.<br />

Mantripp, D. (1996) Radar Altimetry, in: The Determination of Geophysical<br />

Parameters from Space, Fancey, N. E., Gardiner, I.D. and Gardiner, R.A.<br />

(eds.), Institute of Physics Publishing, London, UK, pp. 119-171.<br />

Marth, P. C., Jensen, J. R., Kilgus, C. C., Perschy, J. A., MacArthur, J. L., Hancock,<br />

D. W., Hayne, G. S., Purdy, C. L., Rossi, L. C. and Koblinsky, C. J. (1993)<br />

Prelaunch Performance of the NASA Altimeter for the TOPEX/POSEIDON<br />

Project, IEEE Transactions on Geoscience and Remote Sensing, Vol. 31, No. 2,<br />

pp. 315-332.


245<br />

Martin, T. V., Zwally, H. J., Brenner, A. C. and Bindschadler, R. A. (1983) Analysis<br />

and Retracking of Continental Ice Sheet Radar Altimeter Waveforms, Journal<br />

of Geophysical Research, Vol. 88, No. C3, pp. 1608-1616.<br />

Menard, Y. and Fu, L. L. (2000) Jason-1, on the Tracks of TOPEX/POSEIDON,<br />

AVISO Altimetry Newsletter, Vol. 7, pp. 3-8.<br />

Moritz, H. (1980) Advanced Physical Geodesy, Wichmann, Karlsruhe, 500 pp.<br />

Olgiati, A., Balmino, G., Sarrailh, M. and Green, C. M. (1995) Gravity Anomalies<br />

from Satellite Altimetry: Comparison between Computation via Geoid Heights<br />

and via Deflections of the Vertical, Bulletin Geodesique, Vol. 69, No. 4, pp.<br />

252-260.<br />

Nerem, R. S. (1995) Measuring Global Mean Sea Level Variations Using<br />

TOPEX/POSEIDON Altimeter Data, Journal of Geophysical Research, Vol.<br />

100, No. C12, pp. 25135-25151.<br />

NRSC (1995) Altimeter Waveform Product ALT.WAP Compact User Guide, PF-<br />

UG-NRL-AL-0001, The United Kingdom Processing and Archiving Facility,<br />

UK, 32 pp.<br />

Parsons, C. L. (1979) An Assessment of GEOS-3 Waveheight Measurements, in:<br />

Ocean Wave Climate, Earle, M. D. and Malahoff, A. (eds.), Plenum Press,<br />

New York, USA, pp. 235-251.<br />

Partington, K. C., Ridley, J. K. and Rapley, C. G. (1989) Observations of Surface<br />

Properties of the Ice Sheets by Satellite Radar Altimetry, Journal of Glaciology,<br />

Vol. 35, No. 120, pp. 267-275.<br />

Partington, K. C., Cudlip, W. and Rapley, C. G. (1991) An Assessment of the<br />

Capability of the Satellite Radar Altimeter for Measuring Ice Sheet<br />

Topographic Change, International Journal of Remote Sensing, Vol. 12, No. 3,<br />

pp. 585-609.<br />

Pope, A. J. (1976) The Statistics of Residual and the Detection of Outliers, Technical<br />

Report, TR-NOS-65-NGS-1, National Ocean Survey, Rockville.<br />

Priester, R. W. and Miller, L. S. (1979) Estimation of Significant Wave Height and<br />

Wave Height Density Function Using Satellite Altimeter Data, Journal of<br />

Geophysical Research, Vol. 84, No. B8, pp. 4021-4026.<br />

Quartly, G. D., Srokosz, M. A. and Guymer, T. H. (1998) Understanding the Effects<br />

of Rain on Radar Altimeter Waveforms, Journal of Geophysical Research, Vol.<br />

22, No. 11, pp. 1567-1570.<br />

Quartly, G. D. and Srokosz, M. A. (2001) Analyzing Altimeter Artifacts: Statistical<br />

Properties of Ocean Waveforms, Journal of Atmospheric and Oceanic<br />

Technology, Vol. 18, pp. 2074-2091.


246<br />

Rapp, R. H. and Yi, Y. (1997) Role of Ocean Variability and Dynamic Ocean<br />

Topography in the Recovery of the Mean Sea Surface and Gravity Anomalies<br />

from Satellite Altimeter Data, Journal of Geodesy, Vol. 71, pp. 617-629.<br />

Remy, F. and Minster, J. F. (1993) Precise Altimetric Topography in Ice-sheet Flow<br />

Studies, Annals of Glaciology, Vol. 17, pp. 195-200.<br />

Remy, F., Mazzega, P., Houry, S., Brossier, C. and Minster, J. F. (1989) Mapping of<br />

the Topography of Continental Ice by Inversion of Satellite-Altimeter Data,<br />

Journal of Glaciology, Vol. 35, No. 119, pp. 98-107.<br />

Resource Assessment Commission (1993) Coastal Zone Inquiry: Final Report,<br />

Australian Government Publishing Service, Canberra, Australia, 635 pp.<br />

Ridley, J. K. and Partington, K. C. (1988) A Model of Satellite Radar Altimeter<br />

Return from Ice Sheets, International Journal of Remote Sensing, Vol. 9, No. 4,<br />

pp. 601-624.<br />

Rodriguez, E. (1988) Altimetry for Non-Gaussian Oceans: Height Biases and<br />

Estimation of Parameters, Journal of Geophysical Research, Vol. 93, No. C11,<br />

pp. 14107-14120.<br />

Rodriguez, E. and Chapman, B. (1989) Extracting Ocean Surface Information from<br />

Altimeter Returns: The Deconvolution Method, Journal of Geophysical<br />

Research, Vol. 94, No. C7, pp. 9761-9778.<br />

Rodriguez, E. and Martin, M. (1994a) Assessment of the TOPEX Altimeter<br />

Performance Using Waveform Retracking, Journal of Geophysical Research,<br />

Vol. 99, No. C12, pp. 24957-24969.<br />

Rodriguez, E. and Martin, M. (1994b) Estimation of the Electromagnetic Bias from<br />

Retracked TOPEX Data, Journal of Geophysical Research, Vol. 99, No. C12,<br />

pp. 24971-24979.<br />

Sandwell, D. T. (1984) A Detailed View of the South Pacific Geoid from Satellite<br />

Altimetry, Journal of Geophysical Research, Vol. 102, pp. 10039-10054.<br />

Sandwell, D. T. (1992) Antactic Marine Gravity Field from High-Density Satellite<br />

Alitmetry, Geophysical Journal International, Vol. 109, pp. 437-448.<br />

Sandwell, D. T. and Smith, W. H. F. (1995) Gravity Anomaly Profiles from ERS-1,<br />

Topex and Geosat Altimetry, EOS, Transaction of the American Geophysical<br />

Union, Vol. 76, No. 17, pp. S89.<br />

Sandwell, D. T. and Smith, W. H. F. (1997) Marine Gravity Anomaly from Geosat<br />

and ERS-1 Satellite Altimetry, Journal of Geophysical Research, Vol. 102, pp.<br />

10039-10054.


247<br />

Sandwell, D.T. (2003) Significant Improvement In Global Marine Gravity By Re-<br />

Tracking Altimeter Waveforms, paper presented to the IUGG General<br />

Assembly, Sapporo, Japan, July [CD-ROM].<br />

Scott, R. F., Baker, S. G., Birkett, C. M., Cudlip, W., Laxon, S. W., Mantripp, D. R.,<br />

Mansley, J. I., Morley, J. G., Rapley, C. G., Ridley, J. K., Strawbridge, F. and<br />

Wingham, D. J. (1994) A Comparison of The Performance of the Ice and<br />

Ocean Tracking Modes of the ERS-1 Radar Altimeter over Non-Ocean Surface,<br />

Geophysical Research Letters, Vol. 21, No. 7, pp. 553-556.<br />

Shum, C. K., Woodworth, P., Andersen, O., Egbert, G., Francis, O., King, C., Klosko,<br />

S., Provost, Le. C., Li, X., Molines, J., Parke, M., Ray, R., Schlax, M.,<br />

Stammer, D., Tierney, C., Vincent, P. and Wunsch, C. (1997) Accuracy<br />

Assessment of Recent Ocean Tide Models, Journal of Geophysical Research,<br />

Vol. 102, No. C11, pp. 25173-25194.<br />

Shum, C. K., Parke, M. E., Schutz, B. E., Abusali, P. A. M., Gutierrez, R., Pekker, T.,<br />

Tapley, B. D., Martinez Benjamin, J. J., Blaha, J., Jacobs, G., Jeffress, G. A.,<br />

Schaudt, K. J. and Wingham, D. J. (1998) Improvement of<br />

TOPEX/POSEIDON Altimeter Data for Global Change Studies and Coastal<br />

Applications, AVISO Altimetry Newsletter, Vol. 6, pp. 102-103.<br />

Soussi, B. and Zanife, O.-Z. (1999) ERS-2 Radar Altimeter off Pointing Impact Study<br />

Task1 - Impact of the Mispointing on the Altimeters measurements by<br />

Simulation,<br />

http://www.ifremer.fr/cersat/FICHES/DIVERS/CLS/PDF/E_002F1.PDF<br />

[last accessed: 8 March 2004].<br />

Srokosz, M. A. (1986) On the Joint Distribution of Surface Elevation and Slopes for<br />

a Nonlinear Random Sea, with an Application to Radar Altimetry, Journal of<br />

Geophysical Research, Vol. 91, No. C1, pp. 995-1006.<br />

Tapley, B. D. and Kim, M. C. (2001) Applications to Geodesy, in: Satellite Altimetry<br />

and Earth Sciences, Fu, L.-L. and Cazenave, A. (eds.), Academic Press, New<br />

York, USA, pp. 371-406.<br />

Thom, B. G. (1984) Geomorphic Research on the Coast of Australia: A Preview, in:<br />

Coastal Geomorphology in Australia, Thom, B. G. (eds.), Academic Press<br />

Australia, Sydney, pp. 1-15.<br />

Tierney, C., Parke, M. E. and Born, G. H. (1998) An Investigation of Ocean Tides<br />

Derived from Along-Track Altimetry, Journal of Geophysical Research, Vol.<br />

103, No. C5, pp. 10273-10287.<br />

Tokmakian, R. T., Challenor, P. G., Guymer, T. H. and Srokosz, M. A. (1994) The<br />

U.K. EODC ERS-1 Altimeter Oceans Processing Scheme, International<br />

Journal of Remote Sensing, Vol. 15, No. 4, pp. 939-962.


248<br />

Tziavos, I. N., Forsberg, R. and Sideris, M. G. (1998) Marine Gravity Field<br />

Modelling Using Ship-Borne Gravity and Geodetic Mission Altimetry Data,<br />

Geomatics Research Australasia, Vol. 69, pp. 1-18.<br />

Wang, Y.-M. and Rapp, Richard H. (1991) Geoid Gradients for Geosat and<br />

Topex/Poseidon Repeat Ground Tracks, 408, The Ohio State <strong>University</strong>,<br />

Columbus, USA, 26 pp.<br />

Wessel, P. and Smith, W. H. F. (1996) A Global, Self-Consistent, Hierarchical,<br />

High-Resolution Shoreline Database, Journal of Geophysical Research, Vol.<br />

101, No. B4, pp. 8741-8743.<br />

Wessel, P. (2000) GMT - Generic Mapping Tools, http://gmt.soest.hawaii.edu/ [last<br />

accessed: 8 March 2004].<br />

Wingham, D. J., Rapley, C. G. and Griffiths, H. (1986) New Techniques in Satellite<br />

Tracking Systems, Proceedings of IGARSS' 88 Symposium, Zurich,<br />

Switzerland, September, pp. 1339-1344.<br />

Wingham, D. J., Rapley, C. G. and Morley, J. G. (1993) Improved Resolution Ice<br />

Sheet Mapping with Satellite Altimeters, EOS, Transaction of the American<br />

Geophysical Union, Vol. 74, No. 10, pp. 113-116.<br />

Wingham, D. J. (1995a) The Limiting Resolution of Ice-Sheet Elevations Derived<br />

from Pulse-Limited Satellite Altimetry, Journal of Glaciology, Vol. 41, No.<br />

138, pp. 413-422.<br />

Wingham, D. J. (1995b) A Method for Determining the Average Height of a Large<br />

Topographic Ice Sheet from Observations of the Echo Received by a Satellite<br />

Altimeter, Journal of Glaciology, Vol. 41, No. 137, pp. 125-141.<br />

Wingham, D. J., Ridout, A.J., Scharroo, R., Arthern, R. and Shum, C. K. (1998)<br />

Antarctic Elevation Change from 1992 to 1996, Science, Vol. 282, No. 5388,<br />

pp. 369-580.<br />

Zwally, H. J. (1996) GSFC Retracking Algorithms, http://icesat4.gsfc.nasa.gov/ [last<br />

accessed: 8 March 2004].<br />

Zwally, H. J. and Brenner, A. C. (2001) Ice Sheet Dynamics and Mass Balance, in:<br />

Satellite Altimetry and Earth Sciences: A Handbook of Techniques and<br />

Applications, Fu, L.-L. and Cazenave, A. (eds.), Academic Press, San Diego,<br />

USA, pp. 351-370.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!