7 tesis ppp kongzhe chen

301
UCGE Reports Number 20217 Department of Geomatics Engineering Real-Time Precise Point Positioning, Timing and Atmospheric Sensing (URL: http://www.geomatics.ucalgary.ca/links/GradTheses.html) by Kongzhe Chen April 2005

Upload: josue-gonzalez-garcia

Post on 26-Dec-2015

13 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: 7 Tesis PPP Kongzhe Chen

UCGE Reports Number 20217

Department of Geomatics Engineering

Real-Time Precise Point Positioning, Timing and Atmospheric Sensing

(URL: http://www.geomatics.ucalgary.ca/links/GradTheses.html)

by

Kongzhe Chen

April 2005

Page 2: 7 Tesis PPP Kongzhe Chen

THE UNIVERSITY OF CALGARY

Real-Time Precise Point Positioning, Timing and Atmospheric Sensing

by

KONGZHE CHEN

A DISSERTATION

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF DOCTOR OF PHILOSOPHY

DEPARTMENT OF GEOMATICS ENGINEERING

CALGARY, ALBERTA

April, 2005

© Kongzhe Chen 2005

Page 3: 7 Tesis PPP Kongzhe Chen

iii

ABSTRACT

The availability of precise GPS orbit and clock products has enabled the development of a

novel positioning methodology known as Precise Point Positioning (PPP). Different from

the conventional differential methods, PPP is based on the processing of un-differenced

observations from a single GPS receiver.

The thesis investigates different aspects related to the development of a real-time precise

point positioning system and its application to precise timing and atmospheric sensing. This

includes real-time position determination, receiver clock offset and zenith wet delay

parameter estimation using PPP methodology. The real-time aspect of precise orbit and

clock data distribution has also been analyzed with respect to reliability, timeliness,

convenience and available infrastructure. A comprehensive analysis of the error sources

unique to PPP has been provided along with methods for their modeling and mitigation.

Real-time PPP provides an efficient and flexible method for high precision Precipitable

Water Vapour (PWV) determination. It can significantly improve the timeliness and

temporal resolution of PWV estimates that are essential for severe weather forecast and

operational numerical weather prediction. Methods of real-time PWV estimation using PPP

have been investigated and the major error factors including horizontal gradients, mapping

function selection, antenna phase center variations and elevation cutoff angle have been

analyzed.

Page 4: 7 Tesis PPP Kongzhe Chen

iv

To date, precise point positioning has been implemented using dual-frequency GPS

receivers. Since the majority of GPS receivers in use are single-frequency receivers,

methods of precise point positioning using a single-frequency GPS receiver has been

investigated. A method for the estimation of ionosphere horizontal gradients along with the

zenith delay in single-frequency PPP has been developed. Various comparisons and

analysis have also been conducted using different ionospheric mitigation methods and

under different ionospheric conditions.

Numerical results have indicated that a positioning accuracy of sub-centimetre for static

and sub-decimetre for kinematic positioning is obtainable in real-time using dual-frequency

observations. A sub-nanosecond accuracy has been obtained for the receiver clock offset

estimates. An accuracy of 1 mm for precipitable water vapour has been demonstrated. A

sub-metre to decimetre positioning accuracy has been achieved using PPP with only single-

frequency observations.

Page 5: 7 Tesis PPP Kongzhe Chen

v

ACKNOWLEDGEMENTS

I would like to thank my supervisor, Dr. Yang Gao, for his support, encouragement, and

guidance in the past three years.

I am grateful to the examining committee, Drs. Susan Skone, Abraham Fapojuwo, Swavik

Spiewak and Marcelo Santos, for reading the dissertation and giving valuable suggestions.

In the past three years, I have received a lot of support from many researchers. I would like

to express appreciation to Pierre Heroux who has always been ready to discuss with me

about the processing strategies of IGS analysis centers, Ronald Muellerschoen who has

been providing JPL IGDG orbit and clock products in the past two years and answered me

questions about the use of IGDG products, Paul Collins and Ken MacLeod for providing

the NRCan GPS•C real-time code solution orbit and clock products and post-mission phase

solution products via Internet, Dr. Susan Skone and Natalya Nicholson for providing the

radiometer and barometer measurements, Paul Mrstik and Sarka Friedl for providing the

aircraft datasets, Michael Heflin for providing the JPL near real-time precise GPS orbit and

clock products, Dr. Tonie vanDam who provided me atmospheric loading data for some

selected stations and answered me questions about the effects of atmospheric loading in

PPP, Dr. Hans-Georg Scherneck for providing ocean loading coefficients of some selected

stations used in thesis research, Dr. Neil Ashby who answered me questions about

relativistic effects in GPS positioning and timing, Dr. Yong Hu who helped me to conduct

Page 6: 7 Tesis PPP Kongzhe Chen

vi

the Internet data transmission test between University of Calgary and York University, and

Suen Lee for proofreading the thesis.

I also would like to thank my friends for their help and encouragement. They are Xiaobing

Shen, Zhe Liu, Yufeng Zhang, Zhiyu Chen, Wuji Yang, Baichong Chao, Daocang Wu,

Qingyun Hu, Qiaoping Zhang, Hong Li, Shaokui Ge, Chen Xu, Li Sheng, Xiaoji Niu,

Minxue He, Junjie Liu, Chaochao Wang, Wentao Zhang, Changlin Ma, Jau-Hsiung Wang,

Mohamed Abdel-salam, and Adam Wojciechowski.

I would like to thank my parents, and my wife, for their support and encouragement

throughout my studies.

Page 7: 7 Tesis PPP Kongzhe Chen

vii

DEDICATION

To my mother.

Page 8: 7 Tesis PPP Kongzhe Chen

viii

TABLE OF CONTENTS

APPROVAL PAGE ................................................................................................................ ii

ABSTRACT ........................................................................................................................... iii

ACKNOWLEDGEMENTS ................................................................................................... v

DEDICATION....................................................................................................................... vii

TABLE OF CONTENTS .................................................................................................... viii

LIST OF TABLES ................................................................................................................ xii

LIST OF FIGURES ............................................................................................................. xiv

LIST OF SYMBOLS AND ACRONYMS ....................................................................... xviii

CHAPTER 1: INTRODUCTION........................................................................................ 1

1.1 Background ................................................................................................................ 2 1.2 Objectives and Contributions..................................................................................... 7 1.3 Thesis Outline ............................................................................................................ 9

CHAPTER 2: METHODS OF REAL-TIME PRECISE POINT POSITIONING, TIMING AND ATMOSPHERIC SENSING........................................... 11

2.1 Definition of Real-Time........................................................................................... 11 2.2 Real-Time Precise Point Positioning ....................................................................... 13

2.2.1 Concept of Precise Point Positioning........................................................... 15

2.2.2 Advantages of Precise Point Positioning ..................................................... 18

2.2.3 Challenges of Precise Point Positioning ...................................................... 19 2.3 Real-Time GPS Timing ........................................................................................... 21

2.3.1 Time Scales.................................................................................................. 23

2.3.2 GPS Time Transfer Techniques................................................................... 26

2.3.4 Challenges in GPS Time Transfer ............................................................... 32

2.3.5 Real-Time Timing Using IGDG Products ................................................... 35 2.4 Real-Time Atmospheric Sensing ............................................................................. 37

Page 9: 7 Tesis PPP Kongzhe Chen

ix

2.4.1 Water Vapour Sensing Techniques.............................................................. 37

2.4.2 GPS Meteorology......................................................................................... 40

2.4.3 Strategies for ZTD Estimation Using GPS Observations ............................ 46

2.4.4 ZTD Estimation Using PPP Methodology................................................... 49

CHAPTER 3: ERROR MITIGATION AND MODELING IN PRECISE POINT POSITIONING .......................................................................................... 53

3.1 Dual-Frequency GPS Observables .......................................................................... 53

3.1.1 Dual-Frequency Code and Carrier Phase Measurements ............................ 54

3.1.2 Ionosphere-free Combinations..................................................................... 61 3.2 Error Mitigation for Dual-Frequency Measurements .............................................. 65

3.2.1 Satellite Orbit and Clock Errors................................................................... 66

3.2.2 Ionospheric Effects ...................................................................................... 66

3.2.3 Tropospheric Delay...................................................................................... 71

3.2.4 Relativistic Effects ....................................................................................... 72

3.2.5 Phase Wind-up ............................................................................................. 75

3.2.6 Hardware Delays.......................................................................................... 78

3.2.7 Initial Phase Offsets ..................................................................................... 81

3.2.8 Multipath and Measurement Noise.............................................................. 83

3.2.9 Antenna Phase Center Offset and Variations............................................... 84

3.2.10 Site Displacement Effects ............................................................................ 87 3.3 Modeling for Dual-Frequency Measurements ......................................................... 97

3.3.1 Functional Model ......................................................................................... 97

3.3.2 Stochastic Model........................................................................................ 101

CHAPTER 4: REAL-TIME PRECISE GPS ORBIT AND CLOCK PRODUCTS AND ANALYSIS...................................................................................... 105

4.1 IGS Precise Orbit and Clock Products................................................................... 106 4.2 NRCan Real-Time Precise GPS Products.............................................................. 110

4.2.1 GPS•C Code Solution Products................................................................. 111

4.2.2 GPS•C Phase Solution Products................................................................ 114 4.3 JPL Real-Time Precise GPS Products ................................................................... 117

Page 10: 7 Tesis PPP Kongzhe Chen

x

4.3.1 JPL Near Real-Time Orbit and Clock Products......................................... 118

4.3.2 JPL IGDG Orbit and Clock Products......................................................... 119 4.4 Accuracy Statistics of Real-Time Products ........................................................... 121 4.5 Latency and Age .................................................................................................... 122

4.5.1 Latency and Age of GPS•C Corrections ................................................... 124

4.5.2 Latency and Age of IGDG Corrections ..................................................... 127 4.6 Real-Time Product Distribution Issues.................................................................. 132

4.6.1 Broadcast by Satellite ................................................................................ 134

4.6.2 UDP/IP Multicast....................................................................................... 136 4.7 Real-Time Correction Formats .............................................................................. 139

4.7.1 RTCM-104................................................................................................. 140

4.7.2 RTCA-159.................................................................................................. 141

4.7.3 GPS•C Format ........................................................................................... 141

4.7.4 IGDG Format ............................................................................................. 142

4.7.5 Application of Real-Time Orbit and Clock Products ................................ 143

CHAPTER 5: PRECISE POINT POSITIONING USING SINGLE-FREQUENCY GPS DATA ..................................................................... 145

5.1 Single-Frequency Point Positioning ...................................................................... 145 5.2 Ionospheric Models................................................................................................ 148 5.3 Precise Point Positioning with Ionospheric Delay Estimated................................ 155

5.3.1 Estimating Ionospheric Horizontal Gradients with Un-differenced GPS Measurements ............................................................................................ 157

5.3.2 Ionospheric Mapping Functions ................................................................ 163

5.3.3 Results of Positioning and Ionospheric Delay Estimation......................... 167

CHAPTER 6: NUMERICAL RESULTS AND ANALYSIS ........................................ 178

6.1 Software Development and Parameter Modeling .................................................. 179

6.1.1 P3-RT Software Package ........................................................................... 179

6.1.2 Modeling for Parameters............................................................................ 188 6.2 PPP Using Dual-Frequency Measurements ........................................................... 191

6.2.1 Static PPP Using IGDG Products .............................................................. 191

6.2.2 Kinematic PPP Using IGDG Products....................................................... 194

Page 11: 7 Tesis PPP Kongzhe Chen

xi

6.2.3 PPP Using GPS•C Products ...................................................................... 200

6.2.4 Summary.................................................................................................... 204 6.3 Receiver Clock Offset Estimation Using PPP Methodology................................. 205

6.3.1 Receiver Clock Offset Estimation Using IGDG Products......................... 206

6.3.2 Analysis of Receiver Clock Offset Estimation .......................................... 211 6.4 Atmospheric Sensing Using PPP Technique ......................................................... 212

6.4.1 Comparison with IGS Final Tropospheric Products.................................. 213

6.4.2 Comparison with Radiometer Measurements............................................ 215

6.4.3 Analysis of Real-Time Water Vapour Sensing Results............................. 229 6.5 PPP Using Single-Frequency Measurements......................................................... 233

6.5.1 Positioning at Mid-Latitude Stations ......................................................... 233

6.5.2 Positioning Using Data from Different Ionospheric Regions.................... 238

6.5.3 Positioning Using Kinematic Datasets....................................................... 247

6.5.4 Summary.................................................................................................... 251

CHAPTER 7: CONCLUSIONS AND RECOMMENDATIONS................................. 253

7.1 Conclusions............................................................................................................ 253 7.2 Recommendations.................................................................................................. 257

REFERENCES.................................................................................................................... 259

Page 12: 7 Tesis PPP Kongzhe Chen

xii

LIST OF TABLES

3.1 Measurements from Cross-Correlation Receivers ....................................................... 55

3.2 Measurements from Non-Cross-Correlation Receivers............................................... 56

4.1 IGS Products of GPS Satellite Orbit and Clock (after IGS Website, 2004) .............. 106

4.2 NRCan Real-Time Precise GPS Products (after Heroux, 2003; Collins, 2004)........ 111

4.3 JPL Real-Time Precise GPS Products (after Muellerschoen, 2003; Heflin, 2004) ... 117

5.1 VTEC of Each Satellite at the Ionospheric Pierce Point............................................ 155

5.2 Ap Indices of GPS Week 1251 .................................................................................. 168

5.3 Positioning Accuracy Using Ionospheric Estimation Model with SLM450 ............. 170

5.4 Positioning Accuracy Using Ionospheric Estimation Model with Different Mapping Functions (Unit: m) .....................................................................................173

5.5 Statistics of VTEC Estimation with Different Mapping Functions (Unit: 0.1 TECU).........................................................................................................................173

5.6 Positioning Accuracy Using Klobuchar Model ......................................................... 176

5.7 Positioning Accuracy Using GIM.............................................................................. 176

6.1 Accuracy Statistics of Static Positioning Results (Unit: cm)..................................... 194

6.2 Accuracy Statistics of Kinematic Positioning Results (Unit: cm) ............................. 198

6.3 Accuracy Statistics of Positioning Using GPS•C Products (Unit: m)....................... 204

6.4 Receiver Clock Offset Estimation Accuracy ............................................................. 210

6.5 ZTD Estimation Statistics .......................................................................................... 214

6.6 Statistics of PWV Comparison .................................................................................. 226

6.7 Accuracy Statistics of Different Strategies (Unit: mm)............................................. 227

6.8 Accuracy Statistics of Single-Frequency Point Positioning at S1 (Unit: m) ............. 237

6.9 Station Coordinates.................................................................................................... 239

6.10 Ap Indices in August 2004 ....................................................................................... 239

Page 13: 7 Tesis PPP Kongzhe Chen

xiii

6.11 Accuracy Statistics of Ionospheric Estimation Model for GLPS (Unit: m) ............ 240

6.12 Accuracy Statistics of Ionospheric Estimation Model for S1 (Unit: m).................. 241

6.13 Accuracy Statistics of Ionospheric Estimation Model for FAIR (Unit: m) ............. 241

6.14 Accuracy Statistics of Klobuchar Model for GLPS (Unit: m)................................. 243

6.15 Accuracy Statistics of Klobuchar Model for S1 (Unit: m) ...................................... 243

6.16 Accuracy Statistics of Klobuchar Model for FAIR (Unit: m) ................................. 244

6.17 Accuracy Statistics of GIM for GLPS (Unit: m) ..................................................... 245

6.18 Accuracy Statistics of GIM for S1 (Unit: m)........................................................... 246

6.19 Accuracy Statistics of GIM for FAIR (Unit: m)...................................................... 246

6.20 Single-Frequency Point Positioning with Airborne Dataset (Unit: m)..................... 251

Page 14: 7 Tesis PPP Kongzhe Chen

xiv

LIST OF FIGURES

3.1 Solid Earth Tide Effects for S1 from September 2nd to 7th, 2004 ................................ 90

3.2 Ocean Loading Effects for S1 from September 2nd to 7th, 2004.................................. 92

3.3 Pole Tide Effects for S1 from 2000 to 2004 ................................................................ 94

3.4 Atmospheric Loading Effects for S1 from 1995 to 1999 ............................................ 96

4.1 IGS Tracking Network on December 19th, 2004 (from IGS Website, 2004) ............ 107

4.2 AC Solutions with Respect to the IGS Final Orbit Combination (from IGS ACC Website, 2004) ............................................................................................................108

4.3 AC Solutions with Respect to the IGS Final Clock Combination (from IGS ACC Website, 2004)...................................................................................................108

4.4 CACS Network (after ICD-GPS•C, 2001) ................................................................ 112

4.5 CDGPS Radio (after CDGPS Receiver User's Guide, 2003) ................................... 114

4.6 Stations Sharing Real-Time Data with NRCan (after Collins, 2004)........................ 116

4.7 The Global Coverage with 20 Stations (from Collins, 2004) .................................... 116

4.8 IGDG Real-Time Network in 2003 (from Armatys et al., 2003) .............................. 120

4.9 Latency and Age (after Kee, 1996)............................................................................ 123

4.10 Latencies of GPS•C Orbit Corrections ..................................................................... 125

4.11 Latencies of GPS•C Clock Corrections.................................................................... 125

4.12 Ages of GPS•C Orbit Corrections ............................................................................ 126

4.13 Ages of GPS•C Clock Corrections ........................................................................... 127

4.14 Latencies of IGDG Corrections ................................................................................ 129

4.15 Ages of IGDG Orbits ................................................................................................ 129

4.16 Ages of IGDG Clocks ............................................................................................... 130

4.17 Difference between Orbits Aged 28 s and 0 s........................................................... 131

4.18 Difference between Clocks Aged 5 s and 0 s ........................................................... 132

Page 15: 7 Tesis PPP Kongzhe Chen

xv

4.19 Transmission Time over the Internet ........................................................................ 136

4.20 Packet Loss Rate Using UDP Transport Protocol .................................................... 139

5.1 The Equatorward Increase of TEC at Local Time 16:00........................................... 159

5.2 The West to East Increase of TEC at Local Time 6:00 ............................................. 159

5.3 The East to West Increase of TEC at Local Time 18:00 ........................................... 160

5.4 The VTEC at Ionospheric Pierce Point for Satellites at 30° Elevation Angle against Azimuth at Different Local Time ...................................................................160

5.5 Mapping Functions .................................................................................................... 166

5.6 Mapping Function Difference.................................................................................... 166

5.7 Positioning Errors Using Ionospheric Estimation Model for GPS Week 1251......... 170

5.8 Zenith Ionospheric Delay Estimates for GPS Week 1251......................................... 171

5.9 Ionospheric Gradients in the East Direction .............................................................. 171

5.10 Ionospheric Gradients in the North Direction........................................................... 172

5.11 Vertical TEC Comparison......................................................................................... 172

6.1 P3-RT Interface – Setup............................................................................................. 181

6.2 P3-RT Interface – Kinematic Processing................................................................... 182

6.3 P3-RT Interface – Static Processing .......................................................................... 183

6.4 P3-RT Interface – Trajectory ..................................................................................... 184

6.5 P3-RT Interface – Position Errors.............................................................................. 184

6.6 Real-Time PPP........................................................................................................... 186

6.7 Post-Mission PPP....................................................................................................... 187

6.8 Real-Time Static Positioning Using IGDG Products................................................. 192

6.9 Static Positioning Using IGS Dataset ........................................................................ 193

6.10 Positioning Errors with Vehicle Dataset................................................................... 195

6.11 Vehicle Trajectory on September 30th, 2003 ............................................................ 196

6.12 Positioning with Aircraft Dataset.............................................................................. 197

Page 16: 7 Tesis PPP Kongzhe Chen

xvi

6.13 Aircraft Trajectory on August 28th, 2004.................................................................. 197

6.14 Satellite Geometry for Vehicle Dataset on September 30th, 2003 ............................ 199

6.15 Satellite Geometry for Aircraft Dataset on August 28th, 2004 ................................. 200

6.16 Kinematic Positioning Using GPC•C Code Solution Products................................ 201

6.17 Satellite Geometry of S2 on December 2nd, 2003..................................................... 202

6.18 Kinematic Positioning Using GPC•C Phase Solution Products ............................... 203

6.19 Receiver Clock Offset and ZTD Estimates on June 12th, 2004 ................................ 208

6.20 Receiver Clock Offset and ZTD Estimates on June 13th, 2004 ................................ 209

6.21 Receiver Clock Offset and ZTD Estimates on June 14th, 2004 ................................ 210

6.22 ZTD Estimates Compared with IGS Tropospheric Products.................................... 214

6.23 Radiometer, GPS Antenna and MET3A Instruments ............................................... 217

6.24 Niell Mapping Functions .......................................................................................... 218

6.25 Difference between the Wet and Hydrostatic Mapping Functions........................... 219

6.26 Difference between the Calculated and Measured Pressures ................................... 221

6.27 PWV from GPS and WVR on September 2nd, 2004................................................. 223

6.28 PWV from GPS and WVR on September 3rd, 2004 ................................................. 223

6.29 PWV from GPS and WVR on September 4th, 2004 ................................................. 223

6.30 PWV from GPS and WVR on September 5th, 2004 ................................................. 224

6.31 PWV from GPS and WVR on September 6th, 2004 ................................................. 224

6.32 PWV from GPS and WVR on September 7th, 2004 ................................................. 224

6.33 PWV from GPS and WVR on September 8th, 2004 ................................................. 225

6.34 PWV Comparison between GPS and WVR from September 2nd to 8th, 2004.......... 225

6.35 Positioning Errors Using Ionospheric Estimation Model on December 3rd, 2003 ... 234

6.36 Zenith Ionospheric Delay Estimates on December 3rd, 2003 ................................... 234

6.37 Positioning Errors Using Ionospheric Estimation Model on July 27th, 2004 ........... 236

Page 17: 7 Tesis PPP Kongzhe Chen

xvii

6.38 Zenith Ionospheric Delay Estimates on July 27th, 2004 ........................................... 236

6.39 Satellite Geometry of S1 on July 27th, 2004 ............................................................. 237

6.40 Positioning Using Ionospheric Estimation Model .................................................... 248

6.41 Positioning Using Klobuchar Model with Code Measurements .............................. 249

6.42 Positioning Using Klobuchar Model with Code and Phase Measurements ............. 249

6.43 Positioning Using GIM with Code Measurements ................................................... 250

6.44 Positioning Using GIM with Code and Phase Measurements .................................. 250

Page 18: 7 Tesis PPP Kongzhe Chen

xviii

LIST OF SYMBOLS AND ACRONYMS

Symbols

mT weighted mean temperature of the atmosphere

sT surface temperature

NG tropospheric horizontal gradient in north direction

EG tropospheric horizontal gradient in east direction

gm tropospheric gradient mapping function

a azimuth angle

e elevation angle

hzD tropospheric zenith hydrostatic delay

wzD tropospheric zenith wet delay

hm hydrostatic mapping function for troposphere

wm wet mapping function for troposphere

nG ionospheric horizontal gradient in north direction

eG ionospheric horizontal gradient in east direction

1P P-Code pseudorange measurement on 1L

1C C/A-Code pseudorange measurement

Page 19: 7 Tesis PPP Kongzhe Chen

xix

2P P-Code pseudorange measurement on 2L reported by non-cross-correlation

receivers

'P2 P-Code pseudorange measurement on 2L reported by cross-correlation

receivers

iΦ carrier phase measurement on iL

srρ true geometric range

c speed of light

sdt satellite clock error

rdt receiver clock error

orbd satellite orbit error

tropd tropospheric delay

Li/iond ionospheric delay on iL

gdT group delay differential of satellite

21 P/PDCB differential code bias between 1P and 2P of receiver

11 C/PDCB differential code bias between 1P and 1C

reld relativistic effects

iwδ phase windup on iL

iλ wavelength on iL

iN integer phase ambiguity on iL

Page 20: 7 Tesis PPP Kongzhe Chen

xx

)t(ri 0φ initial phase offset of the receiver on iL

)t(si 0φ initial phase offset of the satellite on iL

Pidm P-code multipath on iL

1Cdm C/A code multipath

imδ carrier phase multipath on iL

sol,rrδ solid earth tides

pol,rrδ pole tide

ocn,rrδ ocean loading

atm,rrδ atmospheric loading

ant,rrδ antenna phase center offset and variations of receiver antenna

Acronyms

AC IGS Analysis Center

ACC IGS Analysis Center Coordinator

ARP Antenna Reference Point

BIPM Bureau International des Poids et Mesures

C/A Coarse Acquisition

CACS Canadian Active Control System

CDGPS Canadian-Wide DGPS Service

CDPD Cellular Digital Packet Data

Page 21: 7 Tesis PPP Kongzhe Chen

xxi

CODE Center for Orbit Determination in Europe

CS Control Segment

CSRS Canadian Spatial Reference System

DCB Differential Code Bias

DGPS Differential GPS

DOP Dilution of Precision

EFEC Earth-Fixed Earth-Centered

ERP Earth Rotation Parameters

GFZ GeoForschungsZentrum

GGN Global GPS Network

GIM Global Ionospheric Model

GPS Global Positioning System

GPS•C GPS Correction Service

GPST GPS Time scale

GRAPHIC Group And Phase Ionospheric Correction

GSD Geodetic Survey Division

GSM Global System for Mobile Communication

ICD Interface Control Document

IFB Inter-Frequency Bias

IERS International Earth Rotation and Reference Systems Service

IGDG Internet-based Global Differential GPS System

IGP Ionospheric Grid Point

Page 22: 7 Tesis PPP Kongzhe Chen

xxii

IGF IGS Final products

IGR IGS Rapid products

IGS International GPS Service

IGST IGS Time scale

IGU IGS Ultra-Rapid products

IMS Integrity Monitor Station

IODE Issue Of Data Ephemeris

IONEX IONosphere map EXchange format

IPP Ionospheric Pierce Point

ITRF International Terrestrial Reference Frame

IWV Integrated Water Vapour

JPL Jet Propulsion Laboratory

MCS Master Control Station

MS Master Station

MSAT Mobile Satellite system

MSLM Modified Single Layer Model

NASA National Aeronautics and Space Administration

NGS National Geodetic Survey

NRCan Natural Resources Canada

USNO United States Naval Observatory

OTF On-The-FLY

P-Code Precise Code

Page 23: 7 Tesis PPP Kongzhe Chen

xxiii

PPP Precise Point Positioning

PWV Precipitable Water Vapour

PRN Pseudo Random Noise

QoS Quality of Service

RCP Right Circularly Polarized

RINEX Receiver Independent Exchange Format

RMS Root Mean Square

RS Reference Station

RTACP Real-Time Active Control Point

RTCA Radio Technical Commission for Aviation

RTCM Radio Technical Commission for Marine

RTG Real-Time GIPSY

RTK Real-Time Kinematic

RTMACS Real-Time Master Active Control Station

RTNT Real-Time Net Transfer

SA Selective Availability

SBL Special Bureau on Loading

SISRE Signal-In-Space Range Error

SLM Single Layer Model

SNR Signal-to-Noise Ratio

SP3 Standard Product # 3

SPP Single Point Positioning

Page 24: 7 Tesis PPP Kongzhe Chen

xxiv

SPS Standard Positioning Service

STD Standard Deviation

SWD Slant Wet Delay

SWV Slant Water Vapour

TAI International Atomic Time

TCP Transmission Control Protocol

TEC Total Electron Content

TWSTT Two-Way Satellite Time-Transfer

UDP User Datagram Protocol

UT Universal Time

UTC Coordinated Universal Time

VACP Virtual Active Control Point

VLBI Very Long Baseline Interferometry

VPN Virtual Private Network

VTEC Vertical Total Electron Content

WAAS Wide Area Augmentation System

WADGPS Wide Area Differential GPS

WVR Water Vapour Radiometer

ZHD Zenith Hydrostatic Delay

ZTD Zenith Tropospheric Delay

ZWD Zenith Wet Delay

Page 25: 7 Tesis PPP Kongzhe Chen

1

CHAPTER 1

INTRODUCTION

Ever since Zumberge et al. (1995; 1997) introduced the concept of carrier phase-based

Precise Point Positioning (PPP), which can achieve centimetre to decimetre level accuracy

using dual-frequency code and phase measurements and precise satellite orbit and clock

products, PPP, whose position determination is based on the processing of un-differenced

GPS observations from a single receiver, has now been widely recognized as a new high-

precision positioning technology using GPS. This is because in addition to the

computationally efficient characteristic, the original reason for which it was introduced

(Zumberge et al., 1995; 1997), people have identified a number of unique features of PPP

including simplified operation, cost-effectiveness, no base stations required and positioning

accuracy comparable to the double-difference approach (Kouba and Heroux, 2001a).

However, PPP so far are still used in post or near real-time applications because of the lack

of real-time precise GPS orbit and clock products and the methodology and software to

support real-time data processing.

On the other hand, point positioning using single-frequency observations has not been

considered capable of providing accurate position solutions at centimetre to decimetre level

similar to PPP with dual-frequency measurements, because of ionospheric effects.

Page 26: 7 Tesis PPP Kongzhe Chen

2

Investigation of models for mitigating ionospheric effects to support precise point positioning

using single-frequency GPS data is highly demanded by numerous applications.

In addition to positioning, PPP is capable of supporting precise timing and atmospheric

sensing (Senior and Ray, 2001b; Chen, 2004). Due to its un-differenced nature in data

processing, the method can offer several unique advantages compared to the use of

differential techniques. Investigating its real-time capability is of great interest to many

applications.

1.1 Background

The fundamental navigation technique for the Global Positioning System (GPS) is to use

one-way ranging information from the GPS satellites that are also broadcasting their

estimated positions (Parkinson, 1996). The initial design goal of the system was to provide a

Single Point Positioning (SPP) service for position determination using a single GPS

receiver. SPP is however subject to the effects of all GPS error sources including satellite

orbit and clock errors in the GPS navigation messages, atmospheric effects, and receiver

related errors, etc. After Selective Availability (SA) was turned off in May 2000, the

dominant errors are the errors in the broadcast satellite orbit and clock and ionospheric

effects (Clynch, 2000). Even with a dual-frequency GPS receiver whereby the ionospheric

effects could be effectively eliminated, SPP can only provide position solutions at the

accuracy of a couple of metres.

Page 27: 7 Tesis PPP Kongzhe Chen

3

To obtain positional accuracy at the decimetre to centimetre level, the Real-Time Kinematic

(RTK) positioning technique has been developed and is now widely used in practice. RTK

positioning requires the combination of simultaneous observations from a minimum of two

GPS receivers, with at least one serving as the base receiver with precisely known

coordinates and the rest as rover receivers whose positions are to be determined. Double

differencing between satellites and receivers can remove common errors among GPS

satellites and receivers or reduce errors that are spatially correlated among satellites and

receivers. If equipped with integer carrier phase ambiguity resolution techniques, RTK is

able to provide centimetre accurate positions. There are several drawbacks, however, that are

associated with the current RTK positioning techniques. One drawback is that base receivers

are required to set up at stations with precisely known coordinates, which is not always

feasible in practice. Another drawback related to the current RTK positioning systems is that

the rover receivers must be within the vicinity of the base receivers, typically less than 20 km

(Kouba and Heroux, 2001a).

With the advent of precise orbit and clock products currently available from a number of

organizations including the International GPS Service (IGS), Natural Resources Canada

(NRCan) and the Jet Propulsion Laboratory (JPL), the two major error sources, namely

satellite orbit errors and clock errors, can be effectively mitigated by the use of precise

satellite orbit and clock data. Point positioning using code measurements and the precise GPS

orbit and clock data is able to provide position solutions at metre level accuracy (Lachapelle

et al. 1994a; Heroux and Kouba, 1995). The phase-based point positioning was first

Page 28: 7 Tesis PPP Kongzhe Chen

4

introduced by Zumberge et al. (1995, 1997), who recognized that by processing dual-

frequency code and phase measurements, point positioning could provide comparable

accuracy to that of double differencing method with a greatly reduced computational burden.

This is known as Precise Point Positioning (PPP) and it differs from the conventional SPP

method in several aspects: the use of both dual-frequency code and phase measurements, the

use of precise GPS orbit and clock products and much higher positioning accuracy at

centimetre to decimetre level. PPP is able to overcome the shortcomings associated with the

current RTK methods in that no base stations are required and it has the potential to provide

positioning accuracy similar to the conventional double difference RTK positioning systems

(Gao and Shen, 2002).

PPP has received increasing interest within the GPS positioning and navigation community

in the past decade. Automated GPS data analysis services using precise GPS products have

been set up by Zumberge (1999). Testing by email and ftp, static positioning accuracy of a

couple of centimetres has been reported (Witchayangkoon and Segantine, 1999).

Witchayangkoon (2000) has also developed a PPP system for post processing, which has

demonstrated kinematic position solutions at sub-metre level accuracy and static position

solutions at sub-decimetre level accuracy. After SA was switched off, it was possible to

interpolate the IGS satellite clocks, currently sampled at 5 minute intervals (Kouba and

Springer, 2001b), to any data rate with interpolation error less than 4 cm (Zumberge and

Gendt, 2000). Most of the processing, so far, has been conducted in post-mission since the

time delay for the necessary precise data can be up to several days. For example, the IGS

Page 29: 7 Tesis PPP Kongzhe Chen

5

final products, which can provide GPS satellite orbit and clock accurate to 5 cm and 0.1 ns

respectively, have a latency of 13 days. Even after real-time precise GPS orbit and clock

products became available from some agencies, such as JPL and NRCan (Muellerschoen et

al., 2000; Collins et al., 2001), the tests with real-time orbit and clock products were mostly

done in a simulated real-time mode, which is to “simulate the processing as if in real-time”

(Muellerschoen et al., 2001). Broadcast over the geostationary satellites (Armatys et al.,

2003), the real-time precise GPS orbit and clock products from JPL have also been reported

to improve real-time positioning accuracy using some commercial receivers, such as

receivers from NavCom (Bisnath et al., 2003). However, the reported accuracy of a couple of

decimetres is much worse than the accuracy of sub-decimetre, which has been obtained in

post-mission or simulated real-time mode (Zumberge et al., 1998; Muellerschoen et al., 2001,

Armatys et al., 2003). This is because several error sources, such as solid earth tides and

ocean loading, have not been corrected in these commercial receivers. The potential of PPP

technique would not be fully exploited if these error sources, which are normally reduced in

differential techniques, have not been corrected (Kouba, 2003). There is a significant need

for a systematic investigation of different aspects related to the development of a real-time

precise point positioning system due to its great promise in positioning accuracy and

operational flexibility, and its great potential in enabling numerous new applications.

The use of ground-based GPS observations to estimate the amount of Precipitable Water

Vapour (PWV) in the troposphere has been investigated extensively since 1992 (Bevis et al.,

1992). Due to the importance of high spatial and temporal resolution, all-weather capability

Page 30: 7 Tesis PPP Kongzhe Chen

6

for severe weather forecast and operational numerical weather prediction, it is essential for

ground-based GPS meteorology to obtain high-precision Zenith Wet Delay (ZWD) estimates,

which are nearly proportional to the PWVs overlying the GPS receiver (Bevis et al., 1992),

from GPS observations with latency as short as possible. The estimation so far, however, is

feasible only in post-mission or near real-time (30 minutes to one hour latency) because of

the time required for collecting and compiling GPS raw data, the latency of the precise GPS

products and the complexity of the data processing (Gendt et al., 2001; Pacione et al., 2002;

Reigber et al., 2002; Ware et al., 2004). The real-time PPP provides a new way to enable

real-time ZWD estimation. Unlike traditional double-difference approach, its time latency

and temporal resolution are only limited by the data rate of GPS receivers, so it can support

real-time GPS meteorology applications.

GPS has long been proven to be a primary tool for precise time dissemination and transfer

(Allan and Weiss, 1980). The potential of GPS timing would not be completely exploited if

only code measurements are used in processing (Schildknecht, et al., 1990). Carrier phase

time transfer has been widely investigated by many researchers (Larson and Levine, 1999;

Schildknecht and Dudle, 2000). The sub-nanosecond accuracy so far was only obtained in

post-mission using IGS precise orbit and/or clock products (Larson et al., 2000). A real-time

PPP system has the potential to realize carrier phase time transfer in real-time at sub-

nanosecond accuracy.

Though most GPS receivers in use are single-frequency receivers, point positioning using

single-frequency measurements has not been considered able to provide positioning accuracy

Page 31: 7 Tesis PPP Kongzhe Chen

7

at centimetre to decimetre level, because of ionospheric effects. Different models have been

developed to mitigate ionospheric effects to provide position solutions at the accuracies of a

couple of metres (Klobuchar, 1996; Ovstedal, 2002; Montenbruck, 2003; Beran et al., 2003).

Real-time sub-metre or even decimetre level point positioning accuracy is highly demanded

by single-frequency GPS users.

1.2 Objectives and Contributions

The major objective of this thesis is to develop methodology for real-time precise point

positioning, timing and atmospheric sensing using real-time precise satellite orbit and clock

products. Specific research tasks are described in the following.

1) To identify and investigate the characteristics of various error sources critical to PPP

and further to develop methods to mitigate them in real-time.

2) To investigate appropriate approaches for real-time precise orbit and clock data

transmission with respect to the latency and age.

3) To study different models to mitigate ionospheric effects for PPP using single-

frequency data.

4) To investigate methods for real-time atmospheric sensing using PPP methodology.

Page 32: 7 Tesis PPP Kongzhe Chen

8

5) To investigate the potential of PPP technique for real-time precise timing by studying

the achievable accuracy of receiver clock offset estimates using real-time GPS orbit

and clock products.

6) To develop an operational real-time PPP software system for real-time positioning,

receiver clock offset and zenith tropospheric delay estimation.

The major contributions of this research are summarized in the following.

1) A real-time precise point positioning software system has been developed capable of

real-time positioning, receiver clock offset and zenith tropospheric delay estimation

using real-time precise orbit and clock products.

2) Techniques for real-time distribution of precise orbit and clock products have been

investigated with respect to latency, reliability, coverage, and cost.

3) Error sources relevant to precise point positioning using dual-frequency or single-

frequency observations have been investigated along with the approaches to mitigate

them in real-time.

4) Issues related to real-time precise water vapour estimation, including mapping

function selection, zenith hydrostatic delay modeling, elevation cut-off angle, and

horizontal gradient estimation, have been investigated. The test results have indicated

that PWV accuracy better than 1 mm is obtainable in real-time.

Page 33: 7 Tesis PPP Kongzhe Chen

9

5) A new timing method has been proposed with the capability of recovering

UTC(USNO) at an accuracy of a few nanoseconds in real-time using a single GPS

receiver.

6) Sub-nanosecond receiver clock offset estimates have been obtained by PPP method

with real-time precise orbit and clock products. The tests have demonstrated the

potential of real-time timing using PPP method.

7) A new model has been proposed to mitigate ionospheric effects for PPP with single-

frequency measurements. Test results show that the model can provide sub-metre or

even decimetre level accuracy in real-time.

8) Real-time tests have been conducted to assess the obtainable accuracy of real-time

PPP methodology using two different real-time precise orbit and clock products

which are useful for industrial applications and product development.

1.3 Thesis Outline

Chapter 2 of this thesis gives a brief overview of precise point positioning, GPS timing, and

GPS atmospheric sensing. The background, challenges, methodologies, and real-time aspects

of these applications are discussed.

Page 34: 7 Tesis PPP Kongzhe Chen

10

Dual-frequency GPS observable equations are presented in Chapter 3. In this Chapter, error

sources that are relevant to PPP data processing are described. The characteristic of these

error sources is analyzed along with the methods to mitigate them.

Chapter 4 presents several types of precise GPS orbit and clock products, including those

from IGS, NRCan, and JPL. The accuracy, latency, and sample interval of these products are

also analyzed in this chapter. Appropriate approaches for real-time precise orbit and clock

data transmission are investigated.

In Chapter 5, different ionospheric models are studied for PPP using single-frequency data

with regard to the obtainable accuracy and timeliness. A new model is proposed with great

promise to perform real-time positioning with an accuracy that is only obtainable by other

models in post-mission with a latency of approximately 11 days.

Chapter 6 analyzes the performance of precise point positioning using dual-frequency and

single-frequency data, receiver clock offset estimation, zenith tropospheric delay and

precipitable water vapour estimation, using real-time precise GPS orbit and clock products.

The real-time software package P3-RT, which has been developed to conduct all data

processing tasks in this thesis, is also described in this chapter.

Finally in Chapter 7, the conclusions and recommendations of this thesis are presented.

Page 35: 7 Tesis PPP Kongzhe Chen

11

CHAPTER 2

METHODS OF REAL-TIME PRECISE POINT POSITIONING,

TIMING AND ATMOSPHERIC SENSING

This chapter will describe the methods of precise point positioning, timing and atmospheric

sensing. The definition of real-time will be described first. The concept of precise point

positioning will be given next along with its advantages and challenges. Timing with carrier

phase measurements will also be investigated. A new timing method will be proposed with

the capability of recovering UTC(USNO) at an accuracy of a few nanoseconds in real-time

using a single GPS receiver. GPS meteorology using precise point positioning methodology

and real-time precise GPS orbit and clock products will be finally presented. Challenges and

methods for real-time precise point positioning, timing and atmospheric sensing are

emphasized.

2.1 Definition of Real-Time

This research is focused on real-time applications of PPP methodology, including

positioning, timing and atmospheric sensing. As defined by Hofmann-Wellenhof et al.,

(2000), a position solution is not real-time, unless it is reported in the field immediately or

Page 36: 7 Tesis PPP Kongzhe Chen

12

while still on the station. This requirement is too strict especially for high dynamic

positioning, because it takes time to process but there is also a less stringent definition. If the

processing time is negligible, the position solutions can still be treated as real-time though

they are obtained with a very short delay (Hofmann-Wellenhof et al., 2000). In this research,

real-time data processing is conducted in a computer with real-time precise orbit and clock

products received over the Internet or from a serial port. As tested with Javad GPS receivers,

which output raw observations at 20 Hz, the solution was reported before the next epoch data

was received. Therefore, the processing time for a single epoch data should be less than 50

ms using PPP technique.

In some applications, such as meteorology, real-time is not a strict terminology. It depends

on both the need and expectation (Gutman and Benjamin, 2001). In their position paper for

the real-time applications and products session, Bar-Sever and Dow (2002) have given the

definition of real-time and near real-time:

“We think that the term ‘real-time’ (rt) is too restrictive, and prefer the more flexible ‘near

real-time’ (nrt) to describe latencies ranging from 0 to 6 hours. The boundary between rt and

nrt may be drawn at the latency below which batch processing and data handling is no

longer practical. The IGS is currently producing hourly RINEX files from a large sub-set of

the network, which feed a number of batch-type processes, including the ultra-rapid orbit

determination. We define, therefore, Real Time processes and application as those that

require sub-hourly latency.”

Page 37: 7 Tesis PPP Kongzhe Chen

13

Sub-hourly latency is a general definition. For different applications the timeliness

requirement varies significantly. Real-time positioning, navigation and timing allow latency

of seconds (Bar-Sever and Dow, 2002). In numerical weather forecasting applications, during

stable weather conditions, Precipitable Water Vapour (PWV) estimates can still be

considered as real-time with latency up to 1 hour. During active weather conditions,

however, data for weather forecast should be updated within a few minutes. If the data is

only updated every half an hour, the information can only be considered near real-time

(Gutman and Benjamin, 2001). In GPS meteorology, the sooner the PWV estimates are

provided, the more useful they are for weather forecast.

In this research, data is processed with the computationally efficient PPP method and real-

time precise GPS orbit and clock products; the results, including the coordinates, the receiver

clock offset and the zenith tropospheric delay estimates, can be provided with very short

latency, i.e. less than one second.

2.2 Real-Time Precise Point Positioning

The reason GPS was developed in the first place is to determine coordinates of a position

with a single GPS receiver. This positioning method is often referred to as absolute

positioning or point positioning (Goad, 1998). Point positioning is however subject to the

effects of all GPS error sources including satellite orbit and clock errors in the GPS

navigation messages, atmospheric effects and receiver related errors, etc. Currently, the

Page 38: 7 Tesis PPP Kongzhe Chen

14

dominant errors are the broadcast satellite orbit and clock errors and ionospheric effects

(Clynch, 2000). Therefore, point positioning can only provide position solutions at the

accuracy of a couple of metres using broadcast ephemeris. To obtain better results,

differential positioning approaches are usually used, which require a minimum of two GPS

receivers, with at least one set up at a station with known coordinates (Kouba and Heroux,

2001a). Differencing procedure between satellites and receivers can remove common errors

among GPS satellites and receivers or reduce errors that are spatially correlated among

satellites and receivers. However, the requirement that base receiver(s) be set up at station(s)

with precisely known coordinates is not always feasible in practice. Another drawback

related to differential positioning systems is that the rover receivers must be within the

vicinity of the base receivers (Kouba and Heroux, 2001a). This is because the cancellation of

spatially correlated errors among the GPS satellites and receivers becomes less effective as

the separation or baseline length between the base and rover stations increases. For large-

scale applications such as aerial survey and mapping, current RTK techniques often become

problematic in practice due to the increased operational cost and complexity.

The availability of precise GPS satellite orbit and clock products has enabled the

development of a novel positioning methodology known as precise point positioning (PPP).

PPP processes un-differenced GPS code and/or carrier phase measurements using precise

orbit and clock products. Because of its high precision and operational flexibility, PPP began

to receive increased attention within the GPS positioning, timing, and atmospheric sensing

communities.

Page 39: 7 Tesis PPP Kongzhe Chen

15

In this section, the concept of precise point positioning is described along with the

advantages and challenges of this novel technique. Methods for a real-time precise point

positioning system are introduced. Details of error mitigation and modeling for real-time

precise point positioning methodology will be investigated in Chapter 3.

2.2.1 Concept of Precise Point Positioning

Although the concept of precise point positioning was introduced by Anderle (1976) to

denote point positioning using precise ephemeris and Doppler satellite observations,

Zumberge et al. (1995) were the first to describe PPP method for centimetre accuracy

applications using GPS carrier phase as the principle observable. There are four milestones in

the development of GPS precise point positioning, each resulting in great improvement in

accuracy or practicability.

Code-based PPP (1980s~1995)

The International GPS Service (IGS) has been providing precise GPS orbit and clock

products since 1994 (Kouba and Heroux, 2001a). Some government agencies even started the

work from 1980s (Lachapelle et al., 1996). However, before 1995, only code observations

were used to conduct point positioning with precise GPS ephemeris. Carrier phase

measurements, if even involved, were only used to smooth code measurements (Hatch,

1982). Point positioning using precise GPS ephemeris and code measurements can improve

the accuracy by an order of 2 with respect to the standard point positioning using broadcast

Page 40: 7 Tesis PPP Kongzhe Chen

16

ephemeris, to a couple of metres from about 100 m with SA on (Heroux et al., 1993;

Lachapelle et al. 1994a; Heroux and Kouba, 1995). However, because only the noisy code

measurements were used, the metre level accuracy cannot satisfy the requirements of

centimetre accurate positioning applications such as geodetic surveys where carrier phase

based double differencing methods with ambiguity resolution must be used.

Phase-based PPP (1995~)

It did not take long before PPP method offered centimetre level positioning accuracy by

processing both code and phase observations. Zumberge at al. (1995, 1997) presented daily

precision of a few millimetres in the horizontal and centimetres in the vertical for static

positioning. Sub-decimetre level accuracy was also obtained using receivers set up on a

vehicle (Zumberge at al., 1998). After introducing carrier phase measurements as primary

observables, the accuracy has been improved by an order from metre level to decimetre level.

PPP with SA off (2000~)

The demise of SA greatly benefited civil GPS users who conduct point positioning using

broadcast GPS ephemeris. It is also significant to precise point positioning users, because the

precise clocks become more predictable (Neilan et al., 2000; Zumberge and Gendt, 2000).

With SA on, PPP can only provide centimetre level position solutions with data that

coincides with times when precise satellite clocks have been estimated (Zumberge at al.,

1998). Most users could only access the 15-minute clock solutions from IGS before 2000,

though higher rate clocks were available from several analysis centers (Neilan et al., 2000).

Page 41: 7 Tesis PPP Kongzhe Chen

17

As investigated by Zumberge at al. (1998), using precise clocks with a sample interval of 30

s, kinematic positioning can achieve approximate 7 cm 3D accuracy with GPS observations

at the same data rate. The accuracy would degrade by a factor of three for higher rate data.

After SA was switched off, the 30 s clocks can be interpolated to process data at any rate

with just about 4 mm degradation in the interpolated clocks, compared to about 8 cm (or 20

times larger) degradation in the SA era (Zumberge and Gendt, 2000). Even using the 5-

minute clocks from IGS (Kouba and Springer, 2001b), the errors introduced by interpolation

are less than 4 cm with SA off. Therefore, kinematic PPP is applicable to any data rate after

May 2, 2000, using IGS products.

Real-Time PPP (2000~)

The latest IGS precise GPS orbit and clock products, which are accurate enough to perform

PPP with centimetre level accuracy, are the observed half of the Ultra-Rapid (IGU) products.

But even the IGU products (observed half) still have a latency of about 3-hour. The latency is

even longer for the more accurate and reliable products, such as IGS Rapid and Final orbits

and clocks. So IGS precise products can only support centimetre level accuracy PPP in post-

mission currently.

The advent of real-time precise GPS orbit and clock products from some agencies has made

real-time PPP practicable, for example, the real-time products from JPL and NRCan

(Muellerschoen et al., 2000; 2001; Collins et al., 2001). However, even using the real-time

products, most data processing was still conducted in post-mission to simulate real-time

processing because of the lack of real-time PPP software (Muellerschoen et al. 2001). In this

Page 42: 7 Tesis PPP Kongzhe Chen

18

research, a software package has been developed to perform real-time PPP processing using

orbit and clock products from both JPL and NRCan.

2.2.2 Advantages of Precise Point Positioning

As discussed above, PPP can provide sub-decimetre level kinematic and centimetre level

static positioning accuracy, which is comparable to double-difference technique with On-

The-FLY (OTF) ambiguity resolution. Compared with double-difference methods, PPP has

the following advantages.

Cost Saving and Simplified Operation

PPP brings not only great flexibility to field operations, but also reduces labour and

equipment cost and simplifies operational logistics by eliminating the need for base stations.

On the other hand, errors will be introduced to the position solutions of the rover stations in

differential techniques if the coordinates of the base stations are not precisely determined

(Heroux et al., 2004).

Globally Consistent Solutions

PPP can provide globally consistent position solutions. PPP solutions are consistent with the

precise GPS orbit and clock products, which are usually based on a globally consistent

reference frame. For example, IGS products are currently based on ITRF2000 which is also

the reference frame used by JPL to estimate real-time orbit and clock products.

Page 43: 7 Tesis PPP Kongzhe Chen

19

Computational Efficiency

PPP is computationally efficient for the analysis of GPS data from large networks (Zumberge

et al., 1997). Even today, with advanced computational technology, this property of PPP is

still very useful for real-time or near real-time applications. For example, in GPS

meteorology, PWV estimates should be provided as soon as possible. Data from a GPS

network can be processed in parallel on multiple computers using the PPP method.

Therefore, the processing time of PPP can be much shorter than that of the traditional double

differencing methods employed to process the same datasets from a network. In this research,

real-time atmospheric sensing will be investigated using the computationally efficient PPP

methodology.

Other Useful Parameters

Absolute receiver clock offset and Zenith Tropospheric Delay (ZTD), which are treated as

nuisances in double differencing techniques, can be estimated in PPP and used for timing and

meteorological applications. In this research, the absolute receiver clock offset will be

estimated in PPP processing using real-time orbit and clock products to investigate the

potential of real-time timing using PPP method.

2.2.3 Challenges of Precise Point Positioning

Currently, PPP also has several challenges which are described in this section.

Page 44: 7 Tesis PPP Kongzhe Chen

20

Availability of Precise GPS Orbit and Clock Products

After SA was turned off, high precision GPS orbits and clocks from IGS or other agencies

can be interpolated to any rate for post-mission PPP processing. But in real-time applications,

users are still seeking real-time precise orbit and clock products. Several types of orbit and

clock products are available in real-time currently, such as products from JPL and NRCan,

but most of them are still at developing or testing stages. Receiving and decoding devices for

real-time products are also under development. In Chapter 4, approaches for real-time

distribution of orbit and clock products will be investigated. The latencies of products

received over the Internet and their impact to positioning will also be discussed. Real-time

software systems are also required to conduct real-time data processing using the real-time

products. A real-time software system has been developed and will be presented in Chapter

6.

Convergence Time

Because of receiver- and satellite-specific phase offsets, namely initial phase offsets, which

are generally float values and unknown, the ambiguities in the un-differenced carrier phase

measurements will no longer be integers (Zumberge et al., 1997). It takes time for the float

ambiguities in PPP to converge. Achievements have been made by some researchers to

pseudo-fix or even fix the float ambiguities and a convergence time of several minutes would

be highly desired (Gao and Shen, 2002).

Page 45: 7 Tesis PPP Kongzhe Chen

21

Error Mitigation

Some error sources, which are cancelled out or partially mitigated in differential processing,

have to be properly modeled or estimated in PPP processing. These sources will be studied in

Chapter 3 to fully exploit the potential of PPP methodology. In order to achieve sub-

decimetre level accuracy, dual-frequency receivers are required to mitigate ionospheric

effects. How to obtain comparable accuracy using single-frequency receivers is a big

challenge for GPS researchers but is significant for GPS community because most GPS

receivers currently in use are single-frequency receivers. Point positioning using single-

frequency observations and precise orbit and clock products has been investigated by

researchers with an accuracy of a couple of metres in post-mission (Ovstedal, 2002;

Montenbruck, 2003). In this research, a new ionospheric model will be proposed, which has

great promise to provide sub-metre level accuracy in real-time using single-frequency

observations and precise GPS orbit and clock products. The details of PPP with single-

frequency measurements are discussed in Chapter 5.

2.3 Real-Time GPS Timing

The clocks operating on the GPS satellites need to be precisely synchronized so that GPS

signals can be used for positioning. To synchronize the clocks, one Master Control Station

(MCS) was set up at Falcon Air Force Base near Colorado Springs, Colorado, which collects

the GPS satellite tracking data from five monitor stations around the world. The time error,

Page 46: 7 Tesis PPP Kongzhe Chen

22

frequency error, and frequency drift for each clock operating on the GPS satellite are thus

estimated. The clock errors are then uploaded to each satellite and broadcast to users in real-

time. The process allows clocks across the constellation to be synchronized to within a few

nanoseconds (Francisco, 1996). The synchronization of the clocks at a level of a few

nanoseconds makes possible for metre level accurate position determination, and it also

brings great promise for timing at a few nanoseconds accuracy. GPS has quickly become a

primary tool for high precise time dissemination and transfer ever since it became operational

(Klepczynski, 1996).

Time scale is the foundation of timing and time transfer. There are several different time

scales relevant to GPS time transfer. Before discussing the time transfer techniques, these

time scales will be described first. Currently, only the noisy GPS code observables are used

for time dissemination and transfer in real-time. The more precise carrier phase

measurements, however, are only employed in post-mission because of the latency of precise

GPS orbit and clock products.

As real-time PPP can determine receiver position up to centimetre level accuracy, the

potential of real-time receiver clock offset estimation at the sub-nanosecond level is evident.

PPP has become a preferred choice for precise timing and time transfer. In this research, the

potential of real-time timing using precise point positioning method will be investigated. A

new timing method is proposed with the capability of recovering UTC(USNO) at an

accuracy of a few nanoseconds in real-time using a single GPS receiver. Receiver clock

Page 47: 7 Tesis PPP Kongzhe Chen

23

offset estimates at sub-nanosecond accuracy using real-time precise orbit and clock products

will be presented in Chapter 6.

2.3.1 Time Scales

There are several time scales related to GPS time transfer, including Universal Time,

Coordinated Universal Time, GPS Time and IGS Time.

Universal Time

Universal time (UT) is a time scale based on the Earth rotation on its axis. UT is not uniform

since the angle velocity of the Earth’s rotation is not constant. The fluctuations are partly

caused by the changes in the polar moment of inertia exerted by tidal deformation and other

mass transports. Other factors include the oscillations of the Earth’s rotational axis itself. The

universal time, after corrected for the polar variations, is denoted by UT1 (Hofmann-

Wellenhof et al., 2000). UT1 is still not uniform because of small changes caused by both

regular seasonal variations and irregular and unpredictable changes in the Earth’s rotation

period (Spilker, 1996b).

Coordinated Universal Time

More precise and uniform time scales have been introduced with the availability of high-

precision atomic clocks. The International Atomic Time (TAI) is an atomic time scale

coordinated by the Bureau International des Poids et Mesures (BIPM). TAI is a uniform and

Page 48: 7 Tesis PPP Kongzhe Chen

24

continuous time scale. Because the universal time scale is still widely used, to keep the TAI

close to the universal time scale UT1, a compromise time scale, Coordinated Universal Time

(UTC), was introduced. The lengths of the seconds used in the generation of TAI and UTC

are the same. UTC differs from TAI by an integral number of seconds. The UTC scale is

adjusted by introducing positive or negative leap seconds to ensure approximate agreement

with UT1 within 0.9 s (Spilker, 1996b). At 0 hour on January 1, 1958, TAI and UT1 were

coincident. By the end of 2004, 32 positive leap seconds were inserted into UTC, therefore,

TAI-UTC=32 s.

UTC is not a directly available clock in real-time. UTC(BIPM) time is normally generated

with about 1 month latency. In order to obtain a real-time estimate of UTC, 50 timing centers

around the world generate their own current estimates of UTC, namely UTC(k) (Allan et al.,

1997). For example, UTC(USNO) is a real-time estimate of UTC kept by the U.S. Naval

Observatory (USNO). All of the timing centers are keeping their UTC(k)s within 100 ns of

UTC. In 1996, most UTC(k)s were kept within 10 ns of UTC(BIPM) (Allan et al., 1997).

Currently, the differences between UTC(k)s and UTC(BIPM) have been further reduced

(Lewandowski and Tisserand, 2004)

GPS Time

The GPS time (GPST) is maintained by a set of atomic clocks operating at GPS control

segment (CS) and space. Thus, the GPS time is also based on atomic time but without the

leap seconds of UTC. The introduction of leap seconds in GPS time would make the P-code

Page 49: 7 Tesis PPP Kongzhe Chen

25

receiver out of lock. Other than the leap second effect, however, GPS time is kept by the GPS

control segment to within 1 µs of UTC(USNO) time (modulo 1 s) (Spilker, 1996b).

GPS time was set to agree with UTC at 0 hour on January 6, 1980. Therefore, if ignoring the

fractional part less than 1 µs, GPST-TAI=-19 s and by the end of 2004, GPST-UTC=13 s.

GPS Time is also traceable back to UTC(USNO) with several nanoseconds uncertainty,

which will be discussed in the following.

IGS Time Scale

Since GPS week 1087 (November 5th, 2000), the IGS has provided two sets of clock

estimates, a rapid combination with a latency of 17 hours, and a final set with a latency of 13

days (Kouba and Springer, 2001b). IGS hopes to keep all the IGS clock information,

including those of satellites and receivers, referenced to a common, consistent timescale. At

the beginning of the clock products, the IGS has used a simple linear alignment of the clocks

to the broadcast GPS time for each separate day (Ray and Senior, 2003). However, the

instability of GPS time, which is only kept to within 1 µs of UTC(USNO) time (modulo 1 s),

is so large that this procedure introduces large day-to-day discontinuities in the time and

frequency of IGS clock products. The discontinuities do not affect position solution using the

clock products, but they definitely degrade the performance of time and frequency

dissemination (Ray and Senior, 2003).

To exploit the geodetic receiver for precise time transfer using IGS precise products, a

workshop, namely “IGS/BIPM Pilot Project to Study Accurate Time and Frequency

Page 50: 7 Tesis PPP Kongzhe Chen

26

Comparisons using GPS Phase and Code Measurements” was organized in December 1997

(Ray and Petit, 1999). The project has not only helped refine the IGS clock products and their

link to UTC, it has also provided new opportunity for the BIPM to improve the time transfer

results. Recently, a final timescale, namely IGST, has been generated with regard to the final

IGS combined clock products (Ray and Senior, 2003). These new time scale can greatly

benefit the users who employ IGS products for time transfer.

2.3.2 GPS Time Transfer Techniques

GPS can be used for precise time transfer in several ways, namely, one-way, common-view,

melting-pot and carrier phase.

One-Way

The simplest GPS time transfer technique is the one-way GPS time transfer, which is also

known as “direct-access” or “passive” GPS time transfer. In one-way GPS time transfer, a

GPS receiver, which is set up on a fixed site whose coordinates have been precisely

determined, can calculate the difference between the local clock driving the receiver and the

GPS time using the pseudorange measurements, and then recover UTC(GPS) using broadcast

GPS navigation messages (ICD-GPS-200C, 2000). The broadcast GPS navigation messages

provide the parameters required to relate the GPS time to UTC(GPS), which is used to denote

the time scale of GPS delivered prediction of UTC(USNO) (Hutsell, et al., 2002). The

coefficients are determined by monitoring the clocks on GPS satellites at the USNO

Page 51: 7 Tesis PPP Kongzhe Chen

27

(Klepczynski, 1996). Therefore, user can even link the clock under test to UTC using a single

GPS receiver with C/A code observations.

This method is easy to implement and it can directly recover UTC(USNO) in real-time.

Moreover, it does not need to communicate with other systems. On the other hand, the

accuracy of the recovered UTC(USNO) is limited, e.g., 20 ns after SA switched off

(Lombardi et al., 2001). The error sources include the satellite orbit and clock errors,

ionospheric and tropospheric errors, multipath effects and measurement noise in code

observations. The performance of this technique is also limited by the uncertainty in the

broadcast inter-frequency bias (Tgd). As defined in ICD-GPS-200C, the group delay

differential between 1L and 2L signals is specified as consisting of random and bias

components. The absolute value of the bias part should not exceed 15 ns and the random

variations should not exceed 3 ns (2σ) (ICD-GPS-200C, 2000). The uncertainty between

UTC(USNO) and UTC(GPS) also affects the accuracy of recovered UTC(USNO).

After SA was turned off, the performance of this technique has been improved about a factor

of 2. Moreover, by using multi-channel GPS receivers, and precise ionospheric products, a

10 ns accuracy is achievable. The use of precise GPS orbits and clocks can improve the

precision of receiver clock estimates, but not all time references of the precise orbit and clock

data are traceable to UTC. For example, the traceability of the time reference of IGS Final

precise orbits and clocks, IGS Time scale, is still under development. The traceability of one-

way GPS time transfer will be discussed in Section 2.3.4.

Page 52: 7 Tesis PPP Kongzhe Chen

28

Common-View

Common-view (CV) provides the direct comparison of two clocks at remote locations. In this

method, by differencing GPS observations at two sites to the same satellite at the same

instant, the difference between the two clocks driving GPS receivers at two sites can be

obtained (Allan and Weiss, 1980). By differencing, some common errors can be removed

completely, such as satellite clock errors, uncertainties of broadcast inter-frequency biases.

Some spatially correlated errors can also be reduced, including satellite orbit errors,

ionospheric and tropospheric errors. Therefore, the accuracy of common-view time

comparison is much better than the accuracy of recovered UTC(USNO) using the one-way

method. A 10 ns accuracy can be obtained using single channel GPS receivers. By using

multi-channel receivers, the accuracy can be improved to better than 5 ns (Lombardi et al.,

2001). The accuracy can be further improved by averaging measurements over a long period.

Moreover, if one of the clocks is directly linking with UTC, then the other clock under test

can recover UTC more accurately than the one-way method.

For example, to maintain the TAI and UTC, the BIPM uses the common view technique

based on GPS C/A code observations from time receivers installed in the time laboratories.

At first, only single-channel receivers were used. Multi-channel receivers were later

introduced to the procedure (Defraigne and Petit, 2003).

Common-view requires communication between two receivers. It is mainly used in post-

mission. The performance of common-view is also affected by satellite orbit errors,

ionospheric and tropospheric errors, especially when conducting intercontinental time

Page 53: 7 Tesis PPP Kongzhe Chen

29

transfer. Precise ionospheric products and precise GPS orbits can be used to improve

accuracy in post-mission.

Melting-Pot

Common-view method has been widely accepted. However, it requires exactly simultaneous

observations at both locations, a requirement which is not easy to fulfill except at

measurement laboratories (Klepczynski, 1996). Actually, if precise orbit and clock products

are used, the measurement uncertainty of common-view technique might be just slightly

smaller than the one-way technique when the receivers are widely separated. Moreover, the

satellites, which are commonly viewed by two receivers, are limited in number and in

tracking time, when the receivers are widely separated. In this case, a more robust technique,

melting-pot, can be used (Klepczynski, 1996).

In melting-pot method, a local clock is also compared with a remote clock. It differs from the

common-view, which only uses satellites tracked by two receivers simultaneously, while for

melting-pot, each receiver tracks satellites all-in-view and calculates offset of the local clock

with respect to a common reference time, such as GPS time or IGS time (depending on the

orbit and clock data). The difference of the offsets obtained from two receivers at the same

instant will be the difference of two clocks under test. This method is more robust than the

common-view method because it utilizes satellites all-in-view at each site and sequentially

there are no gaps in data (Klepczynski, 1996). The melting-pot method is probably slightly

less accurate than the common-view method, but it is more robust and derives the offset from

Page 54: 7 Tesis PPP Kongzhe Chen

30

all observations. In melting-pot time transfer, if one of the clocks is traceable to UTC, then

the other clock can recover UTC as well.

Carrier Phase

Common-view time transfer using single-frequency C/A-code GPS receivers, as now widely

used, is limited mainly by the code multipath effects, which are not zero-mean and cannot be

removed by averaging, and over long distances, by error introduced by ionospheric delays,

which can not be mitigated effectively using single-frequency measurements. Though GPS

satellite orbit errors can be greatly minimized by precise GPS orbit products, IGS ionospheric

products, which are used in post-mission time-transfer, can only provide about 2 TECU

accuracy of vertical total electron content at grid points (Ovstedal, 2002). These factors limit

the accuracy of GPS C/A-code time transfer to a couple of nanoseconds. Even after

multichannel P-code dual-frequency GPS receivers are introduced, the accuracy of traditional

common-view time transfer is still limited by the noise and multipath effects in the code

measurement. The potential of GPS time transfer would not be fully exploited unless both the

phase and code measurements from all satellites in view are involved (Schildknecht and

Springer, 1998). Using geodetic GPS receivers, ionospheric effects can be mitigated

effectively by dual-frequency measurements. Carrier phase multipath and noise is also much

smaller than those of C/A code in magnitude. Moreover, the carrier phase multipath is zero-

mean and thus can be further minimized by averaging (Ray, 2000). Therefore, the accuracy

of time transfer using carrier phase measurements is promising. Different from code

Page 55: 7 Tesis PPP Kongzhe Chen

31

measurement, GPS carrier phase measurement contains an unknown ambiguity, which must

be resolved before it is used for precise time transfer.

High accurate position solutions (millimetre to centimetre) achieved by GPS carrier-phase

observations with ambiguities resolved also indicate the potential of GPS carrier-phase for

time-transfer. Since receiver clock errors are closely related to position errors, the use of

geodetic receivers for precise frequency and time transfer has been proposed a long time ago

(Schildknecht, et al., 1990, Schildknecht and Springer, 1998). The technique using geodetic

receivers for time transfer is often called carrier phase time transfer or geodetic receiver time

transfer. In the past years, this technique has been widely investigated (Larson and Levine,

1999; Schildknecht and Dudle, 2000; Ray et al., 2001; Dach et al., 2002; Defraigne and Petit,

2003). Larson et al. (2000) presented an agreement of 1 ns using IGS final orbits and a few

nanoseconds using IGS predicted orbits with the Two-Way Satellite Time-Transfer

(TWSTT) measurements apart from a constant offset duo to the unknown hardware delays in

both ends.

It is important to note that the accuracy of carrier phase time transfer is also affected by

hardware delay instabilities. Therefore, in order to take full advantage of this promising

technology, the receivers should be calibrated carefully. Also, precise GPS satellite orbit

and/or clock products should be used to exploit the precise carrier phase measurements.

Carrier phase time transfer is often referenced as a novel technique different from one-way

and common-view techniques, but it can be implemented in one-way or common-view mode

(Lombardi et al., 2001).

Page 56: 7 Tesis PPP Kongzhe Chen

32

If carrier phase measurements are involved in one-way time transfer, a PPP time transfer is

applied. As PPP can provide position solutions at centimetre level accuracy, the potential of

PPP timing is evident (Senior and Ray, 2001b). In Section 2.3.5, a PPP timing method will

be proposed with the capability of recovering UTC(USNO) at an accuracy of a few

nanoseconds in real-time using a single GPS receiver.

2.3.4 Challenges in GPS Time Transfer

GPS is now widely accepted as a primary tool for time dissemination and transfer. Several

challenges for this application are described in the following.

Traceability of One-Way Time Transfer

Traceability is defined as an unbroken chain of comparison with stated uncertainties

(Lombardi et al., 2001). In GPS timing, traceability is the capability of linking the clock

under test to UTC. In common-view and melting-pot time transfer, if one of the clocks is

precisely linked to UTC, then the other clocks are also traceable to UTC. In one-way GPS

time transfer, the traceability of the clock under test depends on the time reference of the

orbit and clock data. If broadcast orbit and clock data is used, the clock estimates are then

referenced to GPS time. GPS navigation messages contain the parameters needed to relate

GPS time to a time scale of GPS delivered prediction of UTC(USNO), named UTC(GPS), as

already discussed. The uncertainty between UTC(USNO) and UTC(GPS) is still significant.

Page 57: 7 Tesis PPP Kongzhe Chen

33

As investigated by Hutsell et al. (2002), the RMS of daily UTC(GPS) - UTC(USNO) is 5.84

ns for the period from October 2000 to September 2002. It is 6.32 ns for the period from

January 1999 to September 2000 (Gifford et al., 2000). The RMS difference is 7.84 ns and

6.94 ns in 1997 and 1998, respectively (Rivers and Osborne, 1999). Though the performance

of recovering UTC(USNO) using the parameters in the navigation messages has been

improved year by year, the uncertainty is still significant, especially considering that carrier

phase measurements can be used to estimate the clock offset with an accuracy of sub-

nanosecond.

A new method will be proposed in Section 2.3.5. The method estimates receiver clock in the

same way as one-way time transfer, but it uses dual-frequency code and phase

measurements, and the IGDG precise orbit and clock products from JPL, which make real-

time sub-nanosecond receiver clock estimates possible. Moreover, the use of JPL IGDG

products allows the users to recover UTC(USNO) with a much higher accuracy than using

the broadcast messages because JPL IGDG products are referenced to AMC2, one the

alternate master clocks of UNSO.

Calibration of GPS Receiver

The receiver clock offset estimate from GPS data processing is a virtual internal clock

estimate. How to relate the internal clock estimates to the external hardware clock, which is

used to drive the GPS receiver, is a receiver calibration issue (Schildknecht and Dudle,

2000). The external hardware clock is usually the clock to be tested by time transfer

techniques. The problem is that there are very few geodetic receiver systems which have

Page 58: 7 Tesis PPP Kongzhe Chen

34

been calibrated to accurately relate their internal clocks to the external clock standards (Ray

and Petit, 1999a).

Unlike geodetic position determination, in which any effect that is bias-like will be absorbed

into the clock estimates, all those biases should be separated from the clock estimates for

GPS time transfer (Ray and Senior, 2003). No matter which technique is used, the GPS time

transfer performance is affected by hardware delays, including delay in antennas, cable and

receivers. Frequency transfer may not be as sensitive to these delays as time transfer. The

hardware delay in GPS satellite has been calibrated to ns level in the factory before the

satellite launch (ICD-GPS-200C). In order to use GPS receiver for timing, all instrumental

delays from antenna to receiver should also be calibrated (Petit et al., 2001). As investigated

by many researchers, the uncertainty of the instrumental delays is still at 2 ns level even after

careful calibrations (Petit et al., 2001; Landis and White, 2002). The uncertainty in receiver

calibration is a limiting factor for high precision time transfer using geodetic GPS receivers,

as carrier phase can provide sub-nanosecond level clock offset estimates (Ray and Senior,

2003).

Real-Time Carrier Phase Time Transfer

One-way GPS C/A code time transfer is very easy to implement in real-time. Common-view

GPS time transfer needs communication between receivers. Currently, the common-view

GPS C/A code time transfer is also practicable in near real-time or even real-time with

advanced communication technology. Carrier phase time transfer requires precise GPS orbit

and/or clock products. The tests so far were mostly carried out in post-mission using IGS

Page 59: 7 Tesis PPP Kongzhe Chen

35

final orbit and/or clock products. Predicted GPS orbits have also been used with degraded

performance (Larson et al., 2000). The demand for real-time time transfer is significant from

the industry, such as electric power companies. In this thesis, the research results of precise

receiver clock offset estimation will be presented in Chapter 6 using precise point positioning

method and JPL real-time orbit and clock products. The details of JPL real-time products

from the Internet-based Global Differential GPS System (IGDG) will be described in Chapter

4. PPP using IGDG products also provides a new way to recover UTC(USNO), which will be

described in the next section.

2.3.5 Real-Time Timing Using IGDG Products

In the IGS Analysis Centers (ACs) or other agencies that provide precise GPS orbit and clock

products, one receiver clock equipped with a hydrogen maser external frequency is usually

fixed and used as a time reference. For example, JPL IGDG orbits and clocks are referenced

to the clock of IGS station AMC2 (Muellerschoen, 2003). Because of the importance of

AMC2 station in both GPS and timing communities, estimating receiver clock offset using

PPP method and the IGDG real-time orbit and clock products provides a new way to recover

UTC(USNO) in real-time using a single GPS receiver with a much better performance than

the traditional one-way time transfer.

AMC2 station, which is known as one of the “Alternate Master Clocks (AMCs)” because its

clock is selected as a back-up realization of UTC(USNO), is located at the USNO’s Alternate

Master Clock at Schriever APB in Colorado (Matsakis et al., 1999). Hourly Two-Way

Page 60: 7 Tesis PPP Kongzhe Chen

36

Satellite Time-Transfer (TWSTT) is performed by USNO between the Alternate Master

Clock and the Master Clock which serves as a real-time estimate of UTC, namely

UTC(USNO) as already discussed, to steer the clock of AMC2 to the UTC(USNO) (Senior et

al., 1999). This particular relationship between the AMC2 clock and UTC(USNO) provides a

new method for precise real-time recovery of UTC(USNO) using a single GPS receiver. As

described in Powers (2002), since July 3rd, 2002, all timing calibration biases have been

removed from the pseudorange measurements of AMC2 during the data collection procedure.

Therefore, any clock solution referenced to this station would recover UTC(USNO) with an

accuracy of a few nanoseconds (Powers, 2002).

Because AMC2 clock is fixed by JPL during data processing to estimate IGDG real-time

precise orbit and clock products, the receiver clock offset estimates in precise point

positioning using JPL IGDG products are also referenced to AMC2 clock. As presented in

Section 6.3, receiver clock offset estimates at 100 ps accuracy are obtainable in real-time.

Therefore, precise point positioning using IGDG products have the potential of recovering

UTC(USNO) with an accuracy of a few nanoseconds. Timing using PPP methodology and

IGDG products, which keeps the flexibility of the one-way time transfer, can offer a much

better performance to recover UTC(USNO) in real-time than the traditional one-way time

transfer method, in which the accuracy is limited to a level of about 20 ns (Lombardi et al.,

2001). The capability of this method to recover UTC(USNO) is comparable to the common-

view method which however requires one clock precisely linked to UTC(USNO).

Page 61: 7 Tesis PPP Kongzhe Chen

37

2.4 Real-Time Atmospheric Sensing

The use of GPS observations to sense the water vapour in atmosphere has been extensively

investigated since 1992 (Bevis et al., 1992). Due to importance of high spatial and temporal

resolution, all-weather capabilities of the GPS estimated Precipitable Water Vapour (PWV)

for severe weather forecasting and operational numerical weather prediction, the key issue of

ground-based GPS meteorology is how to get high precision Zenith Wet Delay (ZWD)

estimate, which is nearly proportional to the vertically integrated water vapour overlying the

GPS receiver, from GPS observations with as short latency as possible. The estimation so far,

however, is feasible only in post-mission or near real-time (about 30 minutes to one hour

latency) because of the time spent in collecting and compiling GPS raw data, the latency of

the precise GPS products and the complexity of the data processing (Gendt et al., 2001;

Pacione et al., 2002; Reigber et al., 2002; Ware et al., 2004). In this research, a methodology,

which is capable of providing high precision PWV in real-time using a single GPS receiver,

is described. Key factors that would affect tropospheric delay estimation will also be

addressed including horizontal gradients, mapping function selection, antenna phase center

variations, and elevation cut-off angle.

2.4.1 Water Vapour Sensing Techniques

Water vapour in the atmosphere is quite variable and it has significant impact on the global

climate change and small-scale weather development. Therefore, precisely sensing

Page 62: 7 Tesis PPP Kongzhe Chen

38

atmospheric water vapour plays a vital role in meteorological and climatologic research,

including numerical weather prediction and global climate change study (Dodson et al.,

2001). A variety of techniques have been developed to measure the spatial distribution of

water vapour in the atmosphere, including radiosonde, water vapour radiometer (WVR),

Very Long Baseline Interferometry (VLBI) and GPS. Each approach has its own advantages

and disadvantages.

Radiosondes have been widely used all over the world as a reliable tool for determining the

distribution of water vapour in the atmosphere. Installed on balloons, radiosondes can

provide pressure, temperature and relative humidity measurements through a profile of the

atmosphere to a height of about 30 km (Dodson et al., 2001). The advantages of radiosondes

are that they can provide water vapour profile information. But the operational cost restricts

that they are only launched by most meteorological agencies twice per day at limited stations.

For example, in Alberta, radiosondes or rawinsondes are launched twice per day at Stong

Plain (Skone, 2005). Therefore, the temporal and spatial resolution of radiosonde

measurements is not adequate enough, though they can provide good vertical resolution

(Bevis et al., 1992). Another limitation of radiosondes is that the accuracy of humidity

measurements degrades at cold region and at high altitude. Humidity sensors of radiosondes

do not work normally at temperature below –40°C (Dodson et al., 2001). Despite these

limitations, radiosondes are still widely used by meteorological agencies of many countries.

Another water vapour sensing instrument, the water vapour radiometer can be used in two

ways: ground-based and space-based. Ground-based WVRs measure the background

Page 63: 7 Tesis PPP Kongzhe Chen

39

radiation emitted by atmospheric water vapour and then provide integrated water vapour

(IWV) estimates (Bevis et al., 1992). The advantage of ground-based WVRs is that they can

provide good temporal resolution of water vapour distribution in atmosphere, by providing

IWV estimates with high rate and in real-time. But WVRs are expensive devices, so they

cannot be installed densely enough to provide high spatial resolution on the ground. On the

other hand, the space-based WVRs possess some complementary characteristics of the

ground-based ones. These downward-linking WVRs can provide water vapour measurements

with good spatial resolution. But the temporal resolution of their measurements is limited

because the satellites only cycle the Earth few times per day. The recovery of IWV from

space-based WVRs is very complicated over land because the temperature of the hot

background is extremely difficult to sense (Bevis et al., 1992). It is obvious that

measurements of ground-based and space-based WVRs are complementary. Ground-based

WVRs provide high temporal resolution measurements over land, while space-based WVRs

provide high spatial resolution measurements over ocean. A common limitation of ground-

based and space-based WVRs is that their accuracy is degraded when it is raining or heavily

cloudy (Bevis et al., 1992).

VBLI can also be used for atmospheric water vapour sensing. The idea of VBLI for water

vapour measuring is similar to that of ground-based GPS in that it also provides tropospheric

delay estimates. But it is not practical for meteorological applications because of economic

reasons (Niell et al., 2001).

Page 64: 7 Tesis PPP Kongzhe Chen

40

GPS has long been proven to be a powerful tool for atmospheric sensing (Bevis et al., 1992).

It shows great advantages over traditional techniques, which will be discussed in the

following.

2.4.2 GPS Meteorology

The use of GPS observations to sense the water vapour in atmosphere has been extensively

investigated since 1992 (Bevis et al., 1992). Compared with other atmosphere sensing tools,

such as radiosonde and ground-based and space-based WVRs, the advantages of GPS include

the high temporal and spatial resolution of PWV estimates, low cost, and all-weather

capability.

Like WVRs, GPS can be used in two ways to sense atmospheric water vapour. The first

technique utilizes the ground-based, stationary GPS receivers, which have been largely

deployed in continuously operating networks being constructed around the world in a

regional or global scale (Bevis et al., 1992; Businger et al., 1996). The second technique is

the GPS occultation, which uses GPS receivers operating on low earth orbit satellites (Yuan

et al., 1993; Hajj et al., 2002). Water vapour estimated by GPS can play an important role in

the study of climate changes (Yuan et al., 1993), and is of crucial importance for severe

weather forecast and operational numerical weather prediction (Kuo et al., 1993). Therefore,

GPS water vapour sensing can be used in both meteorological and climatologic applications.

The requirements for these applications are different. The key requirements for meteorology

are accuracy and timeliness. Climate research and climatology require even higher accuracy

Page 65: 7 Tesis PPP Kongzhe Chen

41

than meteorology but they are less strict on timeliness. This means that water vapour

estimates for climatology can be achieved in post-mission using the most accurate GPS orbit

and clock products. In this research, ground-based GPS for meteorological applications will

be investigated.

The fundamental of GPS atmospheric sensing is the tropospheric propagation delays. The

tropospheric delays, which are noise in geodetic research, have become signals in

meteorology (Bevis et al., 1992). The tropospheric delay is caused by the larger refractive

index n (n>1) of atmospheric gases than that of free space (n=1), which slows down the

speed of signal to below its speed in vacuum. The spatially varying refractive index also

curves the signal travelling path. The tropospheric delay can be expressed as (Spilker,

1996c):

( ) gpathactualpathgeopathactualtrop Ndsdsds)s(nd ∆+=−= −

−−

610 (2.1)

where g∆ is the difference between the curved and geometric straight paths, and N is the

refractivity and can be expressed as (Spilker, 1996c):

( ) 6101 ×−= nN (2.2)

The refractivity N is a function of pressure and temperature of the atmosphere as provided

by Thayer (1974):

123

12

11

−−− ++= vvvvdd Z)T/P(kZ)T/P(kZ)T/P(kN (2.3)

Page 66: 7 Tesis PPP Kongzhe Chen

42

where 1k =(77.604±0.014) K/mbar, 2k =(64.79±0.08) K/mbar, 3k =(3.776±0.004)×105

K2/mbar, T is the atmosphere temperature (in degrees Kelvin), dP and vP are the partial

pressure of dry air and water vapour (in mbars). 1−dZ and 1−

vZ are the inverse compressibility

factors for dry air and water vapour. As investigated by Davis et al. (1985), the constants 1k ,

2k and 3k can be calculated to a relative accuracy of about 0.02%.

The total tropospheric delay tropd is normally decomposed into two components, which are

the “hydrostatic delay” and the “wet delay”. Hydrostatic delay depends only on the surface

pressure and contributes to about 90% of the total delay. Wet delay is a function of water

vapour distribution (Saastamoinen, 1972; Davis et al., 1985). Mapping functions can be used

to relate the slant hydrostatic delay at any elevation angle to a Zenith Hydrostatic Delay

(ZHD), and the slant wet delay to a Zenith Wet Delay (ZWD) (Niell, 1996). Zenith

Tropospheric Delay (ZTD) is the sum of ZHD and ZWD. Therefore, if ignoring the

horizontal gradients, which will be discussed in Section 2.4.4, the tropd and ZTD can be

expressed as:

wzhztz

wzwhzhwshstrop

DDd

D)e(mD)e(mDDd

+=

+=+= (2.4)

where hsD and wsD are the slant hydrostatic delay and wet delay, respectively, )e(mh and

)e(mw are the hydrostatic and wet mapping functions, respectively, e is the elevation angle,

tzD , hzD and wzD are the ZTD, ZHD and ZWD, respectively.

Page 67: 7 Tesis PPP Kongzhe Chen

43

Several mapping functions have been developed with small difference at low elevation

angles (Davis et al., 1985; Niell, 1996). The ZHD, which is approximately 2.3 m at sea level,

can be calculated to better than 1 mm given the surface pressure measurement accurate to 0.3

mbar or better, using the following formula (Saastamoinen, 1972):

H.cos.P.

Dhz 0002802002660100227680 0

−−=

φ (2.5)

where 0P is the pressure in mbars; φ is the latitude and H is the height above the geoid (km).

The ZWD can be converted to Precipitable Water Vapour (PWV) overlying a GPS receiver,

using the following formula (Bevis et al., 1994):

ZWDPWV ×Π= (2.6)

where the ZWD is given in the unit of length, the dimensionless constant of proportionality

Π can be roughly given as 0.15 (Bevis et al., 1994).

Similarly, the Slant Wet Delay (SWD) can also be related to the integrated Slant Water

Vapour (SWV) as follows:

SWDSWV ×Π= (2.7)

The SWV can then be used for water vapour tomographic modeling (Skone and Shrestha,

2003).

Page 68: 7 Tesis PPP Kongzhe Chen

44

The value of Π can vary as much as 20% different regions and seasons, but it is possible to

predict it with an RMS relative error of less than 2% given only surface temperature

observations at the site. The following expression can be used to calculate the value of Π

(Bevis et al., 1994):

( )[ ]123

610mkkT/kR mv −+

=Πρ

(2.8)

where ρ is the density of liquid water and vR is the specific gas constant for water vapour;

1k , 2k and 3k are the physical constants from the atmospheric refractivity equation (2.3); m

is the ratio of the molar masses of water vapour and dry air; and mT is the weighted mean

temperature of the atmosphere. mT is defined by Davis et al. (1985) as

( )( )dzT/P

dzT/PT

v

v

m

=

2 (2.9)

As investigated by Bevis et al. (1994), the relative error in Π is close to the relative error in

mT . Different regressions have been developed to relate the value of mT with the surface

temperature sT (Schuler et al., 2001; Mendes et al., 2000b). The relationship between mT and

sT varies with location, altitude, weather and season. In this research, mT is calculated from

surface temperature using the linear relation given by Bevis et al. (1994) as follows:

sm T..T 720270 += (2.10)

Page 69: 7 Tesis PPP Kongzhe Chen

45

Both mT and sT are given in degrees Kelvin. This regression was determined by investigating

data in a 2-year period from 13 stations in the United States. These stations spread from

Fairbanks, Alaska, to West Palm Beach, Florida. It was considered suitable for the stations in

Calgary and with a similar accuracy (Gao et al., 2004). The linear relation was demonstrated

with an RMS relative error of less than 2% for the investigated stations (Bevis et al., 1994).

As shown in Equation 2.4, after the removal of the Zenith Hydrostatic Delay (ZHD), which

can be determined with an accuracy of better than 1 mm with precise surface pressure, the

Zenith Wet Delay (ZWD) is obtained from the GPS derived Zenith Tropospheric Delay

(ZTD). Since ZWD is nearly proportional to the quantity of PWV integrated along the zenith

direction, the total PWV can be extracted from ZWD to a relative accuracy of a few percent,

given only surface temperature measurements at the site. Because the high accuracy and

timeliness requirement of water vapour measurements for numerical weather prediction

(Gutman and Benjamin, 2001), the key issue for GPS meteorology is how to obtain high-

precision ZTD estimates from GPS observations with latency as short as possible. Due to the

time spent in collecting and compiling GPS raw data, the latency of the precise GPS products

and the complexity of the data processing, the estimation so far has been conducted only in a

post-mission or near real-time processing mode. In the following, strategies for ZTD

estimation will be discussed.

Page 70: 7 Tesis PPP Kongzhe Chen

46

2.4.3 Strategies for ZTD Estimation Using GPS Observations

An important factor that allows precise ZTD estimation is the availability of GPS satellite

orbits with sufficient accuracy (Gutman and Benjamin, 2001). Before the advent of precise

orbit products available from organizations such International GPS Service (IGS), ZTD had

to be estimated along with satellite orbits and other station parameters. The requirement of

estimating the satellite orbits makes the approach too complex to be feasible for real-time or

near real-time ZTD estimation (Collins et al., 2002). This method could only be used in post-

mission for meteorological research or climatology. For example, currently the IGS uses this

method to estimate ZTD for IGS stations (Heroux, 2003). But as proposed by Bar-Server

(2004), this method will be replaced by PPP methodology for IGS tropospheric products.

As more and more precise GPS orbit products have become available, the processing

procedure can be greatly simplified. In particular, the availability of ultra-rapid products

(predicted half) from IGS has pushed the investigation of real-time GPS meteorology (Kruse

et al., 1999; Fang et al., 2001; Douša, 2001). Because the satellite clock errors in IGS ultra-

rapid products are still considerable (5 ns in the predicted half), normally the double

differencing approach with integer ambiguity resolution capability has to be used to remove

the satellite clock errors in order to derive more precise tropospheric delay estimates (Rocken

et al., 1997), which however can only provide the relative estimates of tropospheric delays or

Slant Water Vapours (SWVs) for short baselines (Bevis et al, 1992). In order to obtain

absolute tropospheric delays or SWVs, either 1) long baselines must be included to reduce

the correlation of the tropospheric delays between stations (Duan et al., 1996; Alber et al.,

Page 71: 7 Tesis PPP Kongzhe Chen

47

2000), or 2) absolute SWVs from a Water Vapour Radiometer (WVR) collocated at one GPS

station should be available to “lever” the relative SWVs to absolute ones (Rocken et al.,

1995). The latter is known as “WVR-levering”. Though both methods can provide PWV

estimates at an accuracy of 1~2 mm (Tregoning et al., 1998), they have several shortcomings

that will be addressed below, besides the time spent in collecting and compiling GPS raw

data from a regional network.

As to the first method, Rocken et al. (1993) have simulated that the absolute PWV cannot be

recovered to better than 1.5 mm even using perfect orbits and dry delays for baselines up to

1000 km. Tregoning et al. (1998) also demonstrated that PWV solutions, which did not

include baselines longer than 2000 km, had a larger bias and a higher scatter with respect to

the radiosonde measurements. To process data from a network including extremely long

baselines, special efforts also have to be taken to reduce the satellite orbit error effects for

near real-time PWV estimation (Ge et al., 2000). As a result, the complexity and the required

processing time are significantly increased. Since the valid time period of PWV estimates is

relatively short for weather forecasts, a forecaster for instance may look for updated

estimates every few minutes during active weather conditions (Gutman and Benjamin, 2001).

Efficient communication and processing techniques to minimize the latency thus have

become a major concern for meteorological applications using this method (Collins et al.,

2002). A discussion on the required accuracy and timeliness of GPS meteorological data can

be found in Gutman and Benjamin (2001).

Page 72: 7 Tesis PPP Kongzhe Chen

48

As to the second method, the “levering” process requires the use of other devices such as

radiometers. Since radiometers are not “all-weather” instruments, they may not generate

useful data under heavy rain conditions and their accuracy may also be degraded when it is

heavily cloudy or lightly raining (Bevis et al., 1992). This will break the all-weather property

of GPS meteorology.

As already discussed, the availability of precise GPS satellite orbit and clock products has

enabled the development of a novel positioning methodology known as precise point

positioning (Zumberge et al., 1997). Two advantages of PPP are that it is computationally

efficient and some absolute information is retained, such as receiver clock offset and zenith

tropospheric delay. PPP provides a new way to perform real-time absolute ZTD estimation

using real-time precise GPS orbit and clock products. PPP does not require spending time in

collecting and compiling GPS raw data from a regional or global GPS network. GPS raw

data can be processed in parallel in local computers. Only ZTD estimates are transferred to a

processing center. Differing from traditional double-difference approach, the timeliness and

temporal resolution of PPP ZTD estimates are only limited by the data rate of GPS receivers,

so it can support real-time GPS meteorology applications.

Haase et al. (2003) have applied PPP approach in post-mission to estimate meteorological

parameters using precise orbits and clocks with a latency of several hours or days. As

investigated by researchers (Rocken et al., 2003; Wang and Dare, 2004), comparable

accuracy of zenith tropospheric delay estimates has been obtained using both PPP processing

and double-difference processing in post-mission. PPP method was also found in some near

Page 73: 7 Tesis PPP Kongzhe Chen

49

real-time applications to save processing time (Reigber et al., 2002; Gendt et al., 2003). In

these applications, high-quality GPS orbits and clocks were first estimated from a global

network and ZTDs were then estimated using parallel PPP processing based on the precise

orbits and clocks from the first step. This idea is the same as what Zumberge et al. (1995;

1997) applied to save processing time. As real-time precise orbit and clock products have

become available from organizations, such as JPL, with sub-decimetre accuracy for orbits

and sub-nanosecond accuracy for clocks (Muellerschoen et al., 2000), the process can be

further simplified. Given the real-time precise orbit and clock products, precise real-time

zenith tropospheric delay estimation becomes feasible using PPP methodology. Its high

timeliness and temporal resolution are of great importance to severe weather forecasting

(Chen, 2004; Gao et al., 2004).

2.4.4 ZTD Estimation Using PPP Methodology

The observation models used in PPP processing will be described in Chapter 3. Several error

sources, which can be completely eliminated or partially mitigated in double difference

method, must be modeled in un-differenced GPS data processing. These error sources will

also be discussed in detail in Chapter 3. Several issues have to be considered to get precise

ZTD estimates using PPP methodology, including tropospheric horizontal gradients,

mapping function selection, antenna phase center variation, elevation cut-off angle, etc.

The estimation of troposphere horizontal gradients was shown to be beneficial for both GPS

positioning and meteorology in Bar-Sever et al. (1998). Bar-Sever (2004) even proposed to

Page 74: 7 Tesis PPP Kongzhe Chen

50

adopt the PPP method to estimate ZTD along with horizontal gradients for IGS tropospheric

combinations. The following equation has been used to model tropospheric effects in this

research (McCarthy and Petit, 2004):

)]asin(G)acos(G)[e(m

D)e(mD)e(md

ENg

wzwhzhtrop

+

++= (2.11)

where

hzD , wzD , )e(mh , )e(mw are the same variables as in Equation 2.4;

NG , EG are the tropospheric horizontal gradients in north and east directions;

)e(mg is the gradient mapping function and

a , e are the azimuth and elevation angles.

The Saastamoinen model in Equation 2.5 has been applied to determine the zenith

hydrostatic delay.

The following gradient mapping function has been used (Chen and Herring, 1997):

003201

.)etan()esin()e(mg +

= (2.12)

The mapping functions, elevation cut-off angle and antenna phase center variation are also

important issues in tropospheric delay estimation (Fang et al., 1998). Niell Mapping

Functions (NMFs) are used for hydrostatic and wet mapping functions in this research (Niell,

1996), which are among the most accurate mapping functions when observations at

Page 75: 7 Tesis PPP Kongzhe Chen

51

elevations angles below 15° are to be included (Mendes and Langley, 2000a). Fang et al.

(1998) have demonstrated that systematic errors have been reduced in ZTD estimation when

using NMFs. Since NMFs are independent of surface meteorology, the users can benefit

when meteorological data is unavailable or unreliable during processing (Niell, 1996). They

are also useful in tropospheric gradient estimation, in which GPS data at low elevation angles

would be included to get precise ZWD estimates. NMFs are valid at elevations as low as 3°

(Niell, 1996).

In some researches, ZWD and ZHD were estimated as a total ZTD using a single mapping

function, which can be either the hydrostatic mapping function or wet mapping function

(Duan et al., 1996). However, because of the difference in mapping functions, especially at

low elevations (Niell, 1996), estimating total delay using a single mapping function will lead

to a bias in the zenith tropospheric delay estimates, which will directly affect the accuracy of

zenith wet delay estimates. In this research, precise pressure measurements (if available) or

pressure models would be used to model zenith hydrostatic delays which are thus mapped

using the hydrostatic mapping function in PPP processing. Therefore, ZWD, possibly

including un-modeled ZHD, is estimated with wet mapping function. The modeled ZHD and

the estimated ZWD will be output as a total ZTD, which will be subtracted by the ZHD

calculated from precise pressure measurements and result in the true ZWD. The details about

why two mapping functions should be used instead of using a single mapping function will

be discussed in Section 6.4.

Page 76: 7 Tesis PPP Kongzhe Chen

52

Bar-Sever and Kroger (1996) found that an elevation cut-off angle of 15° led to a significant

bias between GPS-based estimates of ZWD and the WVR-based estimates. The bias was

reduced dramatically when a cut-off angle of 7° was used. GPS data at low elevation angles

should also be included to separate gradient components from the azimuthally homogeneous

components (Bar-Sever et al., 1998). As investigated by Rothacher et al. (1997), including

low elevation data can increase the number of observations and obtain a better decorrelation

between the estimated height and tropospheric delay parameters. However, the measurement

noise and multipath effects increase at lower elevation angles. In this research, an elevation

cut-off angle of 7° is therefore used.

Both satellite and receiver antenna phase center variations were found to affect tropospheric

delay estimation (Fang et al. 1998; Schmid and Rothacher, 2002). In this research, the IGS

elevation-dependent phase center models have been used.

Page 77: 7 Tesis PPP Kongzhe Chen

53

CHAPTER 3

ERROR MITIGATION AND MODELING IN

PRECISE POINT POSITIONING

This chapter describes the error mitigation and modeling for precise point positioning using

dual-frequency GPS data. The error sources, which are only relevant to precise point

positioning using single-frequency GPS data, will be investigated later on in Chapter 5. Dual-

frequency GPS measurements will be first described. General error sources, either modeled

or estimated to support precise point positioning, will be investigated. Function models and

stochastic models for precise point positioning will also be presented.

3.1 Dual-Frequency GPS Observables

There are two advantages to use dual-frequency GPS measurements for precise point

positioning. The first advantage is that the first-order ionospheric effects can be removed

completely. The second is that most GPS orbit and clock products, including those in the

broadcast and precise ephemeris, are consistent with the dual-frequency ionosphere-free

combinations. In this section, dual-frequency GPS measurements and ionosphere-free

combinations will be discussed.

Page 78: 7 Tesis PPP Kongzhe Chen

54

3.1.1 Dual-Frequency Code and Carrier Phase Measurements

Currently, for positioning purpose, GPS satellites transmit signals on two frequencies,

namely, Link 1 or 1L , and Link 2 or 2L , which are centered at 1575.42 MHz and 1227.6

MHz, respectively (Spilker, 1996a). The 1L and 2L signals are modulated by pseudorandom

noise (PRN) codes so that GPS receivers can track the signals for positioning. 1L is

modulated by both Coarse/Acquisition code (C/A code) and Precision code (P code). 2L is

only modulated by P code. When Anti-Spoofing (AS) is activated, a secure Y code is used

instead of the P code to modulate 1L and 2L signals. The chip rate of Y code is the same as

that of P code, but the Y code is only provided to authorized U.S. government users (Spilker,

1996a). The AS mode was activated on January 31, 1994 and is still operational on Block II

satellites (Langley, 1998a).

For civil GPS receivers, several techniques have been developed to track the encrypted P

code (Y code) (Langley, 1998b). In the civil GPS community, dual-frequency GPS receivers

are usually categorized into two groups based on the tracking techniques used. The first

group are the cross-correlation receivers such as AOA Rogue/TurboRogue and Trimble 4000

serial receivers. The second group are the non-cross-correlation receivers such as Ashtech Z-

XII and AOA Benchmark/ACT receivers (Ray, 1999b). The measurements provided by these

two types of receivers are different and are described in the following.

The cross-correlation style receivers can output at least four types of code and carrier phase

measurements, which are described in Table 3.1.

Page 79: 7 Tesis PPP Kongzhe Chen

55

Table 3.1 Measurements from Cross-Correlation Receivers

Observables Description

1C C/A code at 1L frequency 'P2 Codeless pseudorange at 2L frequency

1L ( 1C ) C/A-based phase at 1L frequency 'L2 cross-correlated phase at 2L frequency

The reason that they are called cross-correlation receivers is that the 'P2 and 'L2

measurements are formed by cross-correlation technique which measures the cross-correlated

( 2P - 1P ) pseudorange and the ( 2L - 1L ) phase differences. Hence, 'P2 and 'L2 can be

expressed as (Ray, 1999b):

'P2 = 1C + ( 2P - 1P ) (3.1)

'L2 = 1L ( 1C ) + ( 2L ( 2P )- 1L ( 1P )) (3.2)

Therefore, the observables ( 1C and 'P2 or 1L ( 1C ) and 'L2 ) are no longer independent of each

other, which will affect the stochastic modelling when using the P1-P2-CP ionosphere-free

combinations to be discussed in Section 3.1.2.

The non-cross-correlation receivers can provide direct measurements of 1P and 2P without

the use of the Y-code by using methods such as squaring technique (Hofmann-Wellenhof et

al., 2000). They can provide six (or more) range measurements as shown in Table 3.2.

Page 80: 7 Tesis PPP Kongzhe Chen

56

Table 3.2 Measurements from Non-Cross-Correlation Receivers

Observables Description

1C C/A code at 1L frequency

1P Y1-codeless pseudorange at 1L frequency

2P Y2-codeless pseudorange at 2L frequency

1L ( 1C ) C/A-based phase at 1L frequency

1L Y1-codeless phase at 1L frequency

2L Y2-codeless phase at 2L frequency

The C/A based measurements 1C and 1L ( 1C ) are the same as the corresponding

measurements from the cross-correlation receivers, but the measurements on 2L are different

between those two types of receivers. Moreover, the non-cross-correlation receivers can

provide two carrier phase ( 1L ( 1C ) and 1L ) and two code ( 1C and 1P ) measurements on 1L .

Normally, 1L ( 1C ) with a higher signal-to-noise ratio (SNR) is reported in RINEX files (Ray,

1999b). The differences between 1C and 1P are not zero-means and are called ( 1P - 1C )

biases. ( 1P - 1C ) biases are relatively stable in time but their magnitude can be up to ~ 2 ns

(Jefferson et al., 2001, Gao et al., 2001). CODE has been estimating ( 1P - 1C ) biases since

GPS week 1057 (Schaer, 2000).

There are four types of pseudorange measurements that could be reported in a RINEX file.

Which pair of measurements ( 1C or 1P , 'P2 or 2P ) should be used to form the ionosphere-

free combinations in PPP data processing depends on the clock products. Before GPS week

1056 (April 02, 2000), IGS precise clock products are consistent with the 1C / 'P2 code

Page 81: 7 Tesis PPP Kongzhe Chen

57

measurements. Starting from GPS week 1056 (April 2nd, 2000), precise clocks of IGS

products is fully consistent with the 1P / 2P code measurements (Kouba, 2003). The clock

errors in the broadcast ephemeris are also consistent with 1P / 2P code measurements (ICD-

GPS-200C, 2000). The same convention ( 1P / 2P ) is also used by JPL and NRCan to estimate

their real-time orbit and clock products (Heroux, 2003; Muellerschoen, 2003). Therefore,

users who choose 1C and/or 'P2 pseudorange measurements should apply the ( 1P - 1C ) code

biases when they use clock information which is consistent with 1P / 2P .

According to specifications, the phase bias ( 1L ( 1C )- 1L ( 1P )) is a constant (Ray, 1999b).

Therefore, difference between 'L2 and 2L is also a constant. As phase observables are

inherently ambiguous, the difference between 'L2 and 2L would be absorbed into the float

ambiguity in precise point positioning.

Some non-cross-correlation receivers, such as NovAtel OEM4 serials, TRIMBLE 5700,

LEICA CRS1000, etc, only report the higher SNR C/A based measurements. In the RINEX

files, these receivers normally report four types of code and phase measurements, 1C , 2P ,

1L ( 1C ) and 2L . Although they are non-cross-correlation receivers, ( 1P - 1C ) code biases

should be applied to their 1C measurements before they are used to form ionosphere-free

combinations.

As discussed above, there are six (or more) types of measurements which may be reported by

dual-frequency GPS receivers in RINEX files. Based on the current conventions, in which

Page 82: 7 Tesis PPP Kongzhe Chen

58

clock information is consistent with 1P / 2P measurements, the following observation equation

models can be formed:

)P(dm)DCBT(c

dddd)dtdt(cP

PP/Pgd

relL/iontroporbs

rsr

1121

11

ερ

++−+

++++−+= (3.3)

)C(dm)DCBDCBT(c

dddd)dtdt(cC

CC/PP/Pgd

relL/iontroporbs

rsr

111121

11

ερ

++−−+

++++−+= (3.4)

)P(dm)DCBT(c

dddd)dtdt(cP

PP/Pgd

relL/iontroporbs

rsr

2221

22

εγρ

++−+

++++−+= (3.5)

)P(dmDCBc)DCBT(c

dddd)dtdt(c

)PP(CP

'PC/PP/Pgd

relL/iontroporbs

rsr

'

' 21121

2

1212

2εγ

ρ

++⋅−−+

++++−+=

−+=

(3.6)

)( mw)N)t()t((

dddd)dtdt(c

iiiis

ir

ii

relLi/iontroporbs

rsri

Φ++++−+

+−++−+=Φ

εδδφφλ

ρ

00

(3.7)

where,

1P is the P-Code pseudorange measurement on 1L (m);

1C is the C/A-Code pseudorange measurement (m);

2P is the P-Code pseudorange measurement on 2L reported by non-cross-correlation

receivers (m);

Page 83: 7 Tesis PPP Kongzhe Chen

59

'P2 is the P-Code pseudorange measurement on 2L reported by cross-correlation

receivers (m);

iΦ is the carrier phase measurement on iL (m);

srρ is the true geometric range (m);

c is the speed of light (m/s);

sdt is the satellite clock error (s);

rdt is the receiver clock error (s);

orbd is the satellite orbit error (m);

tropd is the tropospheric delay (m);

Li/iond is the ionospheric group delay on iL (m);

gdT is the group delay differential of satellite (s);

21 P/PDCB is the differential code bias between 1P and 2P of receiver (s);

11 C/PDCB is the differential code bias between 1P and 1C (s);

reld is the relativistic effects (m);

iwδ is the phase windup on iL (m);

iλ is the wavelength on iL (m/cycle);

iN is the integer phase ambiguity on iL (cycle);

)t(ri 0φ is the initial phase offset of the receiver on iL (cycle);

)t(si 0φ is the initial phase offset of the satellite on iL (cycle);

Page 84: 7 Tesis PPP Kongzhe Chen

60

Pidm is the P-code multipath on iL (m);

1Cdm is the C/A code multipath (m);

imδ is the carrier phase multipath on iL (m);

)(∗ε is the measurement noise (m);

and

22

2

1

6077

=

=

ffγ (3.8)

To calculate the geometric range srρ , the satellite and receiver positions should refer to the

same reference frame. The satellite and receiver positions defined in the Earth-fixed frame

can be expressed by (Rothacher and Beutler, 2002b):

)t(r)t(r)t(r ssant

ss,e

sse δ+= 0 (3.9)

ant,ratm,rocn,rpol,rsol,rr,e,rre,r rrrrr)t(r)t(r δδδδδ +++++= 0 (3.10)

where

)t(r ss,e 0 is phase center of satellite antenna at emission time st ;

)t(r ssantδ is antenna phase center offset and variations of satellite antenna;

)t(r r,e,r 0 is tide-free position of receiver at reception time rt ;

sol,rrδ is solid earth tides;

Page 85: 7 Tesis PPP Kongzhe Chen

61

pol,rrδ is pole tide;

ocn,rrδ is ocean loading;

atm,rrδ is atmospheric loading and

ant,rrδ is the antenna phase center offset and variations of receiver antenna.

To fully exploit the potential of precise point positioning methodology, the error sources and

tide effects shown in Equations 3.3 to 3.10 should be considered and they will be discussed

in Section 3.2.

3.1.2 Ionosphere-free Combinations

After precise GPS orbit and clock products are applied, the biggest error source remaining in

un-differenced GPS measurements is from the ionosphere. Although the range errors due to

the troposphere may be as big as the ionosphere effects, they are more stable spatially and

temporarily. Dual-frequency users can remove the first-order ionospheric effects by forming

ionosphere-free combinations. If ignoring higher-order ionospheric effects, which will be

investigated in Section 3.2, the first-order ionospheric group delay or carrier phase advance

can be expressed in units of distance as follows (Klobuchar, 1996):

= Ndlf

.d

iLi/ion 2

340 (3.11)

Page 86: 7 Tesis PPP Kongzhe Chen

62

where Ndl is the Total Electron Content (TEC), in unit of el/m2, integrated along the path

from an observer to a GPS satellite, and N is electron density.

In precise point positioning, two types of ionosphere-free combinations can be formed. In the

following, only measurements from non-cross-correlation receivers are discussed unless

otherwise specified. Measurements from cross-correlation receivers, after applying the

differential code biases, should be equivalent to the corresponding measurements from non-

cross-correlation receivers except for the correlation between measurements. The first type of

ionosphere-free combination, as presented in Equations 3.12 to 3.14, is the traditional one

based on dual-frequency measurements using the frequency dependent characteristic of

ionospheric refraction.

)P(dmddd)dtdt(c

ffPfPf

P

IFIFreltroporbs

rsr

IF

ερ +++++−+=

−⋅−⋅= 2

22

1

22

212

1

(3.12)

)( mwFddd)dtdt(c

ffff

IFIFIFIFreltroporbs

rsr

IF

Φ+++++++−+=

−Φ⋅−Φ⋅

εδδρ

22

21

22

212

1

(3.13)

where

22

21

202022101011

ff)N)t()t((cf)N)t()t((cf

Fsrsr

IF −+−−+−

=φφφφ

(3.14)

Page 87: 7 Tesis PPP Kongzhe Chen

63

The float ambiguity IFF in this ionosphere-free combination includes the initial phase offsets

of the receiver and satellite on both 1L and 2L . If we assume that the noises in 1P and 2P are

uncorrelated, the noise level of the code combination can be determined by applying random

error propagation law as follows:

( ) ( ) ( ) ( )1

2

2

2

1 31

11

PPPPIF εεγ

εγ

γε ≈

−+

−= (3.15)

The noise of the phase combination can be determined in a similar way.

The first-order ionospheric effects on code and carrier phase are equal in magnitude with

opposite sign. Therefore, another type of ionosphere-free combination can be formed by

averaging code and carrier phase measurements on the same frequency. This combination

has been first introduced to mitigate ionospheric effects for processing single-frequency

measurements, known as GRAPHIC (Group And Phase Ionospheric Correction) (Yunck,

1996a). It has been introduced into precise point positioning using dual-frequency

measurements by Gao and Shen (2002) as:

222

22

2

11111

10101121

111

)()P(

mdmw

)N)t()t(()DCBT(c

ddd)dtdt(c

PCP

P

srP/Pgd

reltroporbs

rsr

Φ+++++

+−+−

+

+++−+=

Φ+=

εεδδ

φφλ

ρ (3.16)

Page 88: 7 Tesis PPP Kongzhe Chen

64

222

22

2

22222

20202221

222

)()P(

mdmw

)N)t()t(()DCBT(c

ddd)dtdt(c

PCP

P

srP/Pgd

reltroporbs

rsr

Φ+++++

+−+−

+

+++−+=

Φ+=

εεδδ

φφλγρ

(3.17)

Plus the traditional ionosphere-free carrier phase combination given in Equation 3.14, there

will be three ionosphere-free observables for each satellite in each epoch. The above P1-P2-

CP (Gao and Shen, 2002) code-phase combinations will reduce the noise level by half

compared to the original code observations or to about one sixth to the traditional

ionosphere-free code combination, since the phase measurement noise is about two orders

smaller than that of code measurements. Precise point positioning using P1-P2-CP code-

phase combination allows estimating the ambiguities associated with 1L and 2L separately.

However, though the combinations keep the integer nature of the ambiguities, the presence of

hardware delays and initial phase offsets only allow float ambiguities to be estimated as:

2210101121

1

)N)t()t(()DCBT(cF

srP/Pgd

CP

+−+

−=

φφλ (3.18)

2220202221

2

)N)t()t(()DCBT(cF

srP/Pgd

CP

+−+

−=

φφλγ (3.19)

The relationship between the float ambiguities 1CPF , 2CPF and IFF can be expressed as:

22

21

22

212

12ff

)FfFf(F CPCP

IF −⋅−⋅

= (3.20)

Page 89: 7 Tesis PPP Kongzhe Chen

65

If hardware delays and initial phase offsets can be resolved, integer ambiguities are

achievable using the code-phase combinations. Progress has been made by researchers to

estimate the initial phase offsets (Gabor and Neren, 2002). gdT has also been provided in

navigation messages with an accuracy of nanosecond level (ICD-GPS-200C, 2000). On the

other hand, Gao and Shen (2002) have proposed a method known as pseudo-fixing to fix the

float ambiguities (including hardware delays and initial phase offsets) to float values. The

method has been demonstrated to reduce the ambiguity convergence time. Fixing or pseudo-

fixing the ambiguities in PPP is beyond the scope of this thesis. In this thesis, all processing

using dual-frequency measurements is based on the traditional ionosphere-free combinations.

3.2 Error Mitigation for Dual-Frequency Measurements

Precise point positioning, which is aimed at providing centimetre level positioning accuracy,

is subject to the influence of all possible error sources in the un-differenced GPS

observations. Most of those error sources can be eliminated completely or mitigated to a

negligible level through single or double differencing. Some of these error sources can also

be safely neglected in traditional point positioning using only code measurements for less

accurate positioning applications. In the following, major error sources that should be taken

into account in PPP are described and their characteristics and methods to mitigate them are

investigated.

Page 90: 7 Tesis PPP Kongzhe Chen

66

3.2.1 Satellite Orbit and Clock Errors

In double differencing techniques, satellite clock errors can be removed completely. Satellite

orbit errors are also reduced by differencing between receivers. In precise point positioning,

precise orbit and clock products are used, which can reduce satellite orbit and clock errors to

5 cm and 0.1 ns in post-mission or sub-decimetre and sub-nanosecond level in real-time.

Precise orbit and clock products used in this research will be discussed in Chapter 4.

3.2.2 Ionospheric Effects

If precise orbit and clock products are used, the ionospheric effects become the biggest error

source in un-differenced measurements. A comparable magnitude error source, tropospheric

effects, is much easier to model, which will be discussed in the next section.

Ionosphere is a dispersive medium for frequencies greater than 100 MHz. The ionospheric

refraction can be approximated by a series expansion in the reciprocal of frequency 1−f .

Considering that GPS signals are right-hand circularly polarized and omitting the higher

orders, the refractive index of the code (group) and phase can be approximated as follows

(Bassiri and Hajj, 1993):

( )

+++++= BBgr cosYXXcosXYXn θθ 22 121

43

21

1 (3.21)

Page 91: 7 Tesis PPP Kongzhe Chen

67

( )

++−−−= BBph cosYXXcosXYXn θθ 22 121

41

21

21

1 (3.22)

where Bθ is the angle of ray with respect to the Earth’s magnetic field, ( )2f/fX p= and

f/fY g= . pf , gf and f are the plasma, gyro and carrier phase frequencies, respectively.

The gyro frequency gf is typically 1.5 MHz, and the plasma frequency pf is rarely exceeds

20 MHz (Klobuchar, 1996). The measured range can then be expressed as:

= ndsmρ (3.23)

The geometric range between the satellite and receiver srρ can be obtained by setting the

refractive index n = 1. Here, by omitting other error sources, the GPS observables discussed

in Section 3.1 can be rewritten as follows (Bassiri and Hajj, 1993):

41

31

21

1 fr

fs

fq

P sr +++= ρ (3.24)

42

32

22

2 fr

fs

fq

P sr +++= ρ (3.25)

41

31

21

111 32 fr

fs

fq

Nsr −−−+=Φ λρ (3.26)

42

32

22

222 32 fr

fs

fq

Nsr −−−+=Φ λρ (3.27)

Page 92: 7 Tesis PPP Kongzhe Chen

68

where

TEC.dLfq p ×== 34021 2 (3.28)

dLcosNBcdLcosffs BBpg θθ == 02 7527 (3.29)

( )dLcosNB.NdLfr Bg θ20

22 1107442437 +×+= (3.30)

where N is the density of electrons, TEC is the total electron content along the GPS signal

propagation path and 0B is the magnitude of the Earth’s magnetic field. q , s and r

represent the first-, second- and third-orders ionospheric effects, respectively.

As discussed in Section 3.1, for dual-frequency users, the following ionosphere-free linear

combinations can be formed to remove the first-order ionospheric effects, and mitigate part

of the second- and third-orders ionospheric effects as follows (Bassiri and Hajj, 1993):

22

212121

222

21

22

122

21

21

ffr

)ff(ffs

Pff

fP

fff

P

sr

IF

−+

−=

−−

−=

ρ (3.31)

22

212121

22

21

2211

222

21

22

122

21

21

32 ffr

)ff(ffs

ffNcfNcf

fff

fff

sr

IF

++

+−−

+=

Φ−

−Φ−

ρ (3.32)

Page 93: 7 Tesis PPP Kongzhe Chen

69

As investigated by Bassiri and Hajj (1993), the second-order ionospheric group delay can be

calculated by the following equation:

TECEsincosAcosEcossinHR

R.

fs

m'mmm

'm

E

Ei

i

×−×

+×= − θθλ 210612

3

183 (3.33)

where ER is the average Earth radius, H is the ionospheric shell height and was set to 300

km by Bassiri and Hajj (1993), 'mθ is the magnetic colatitude of the sub-ionospheric point,

mA and mE are the azimuth and elevation of satellite in a local magnetic East-North-Vertical

coordinates. mE is the same as the elevation angle in local geodetic coordinates E . The

relationship between mA and the azimuth in local geodetic coordinates A has been given by

Bassiri and Hajj (1993).

The magnitude of the second-order ionospheric group delay, as calculated from Equation

3.33, is of the order of 0.16 and 0.33 mm per TEC Unit (TECU) for 1P and 2P respectively

(Bassiri and Hajj, 1993). The residual range error caused by the second-order ionospheric

group delay in the ionosphere-free code combination is then -0.11 mm per TECU. From

Equation 3.24 to 3.27, we can see that the second-order ionospheric phase advances in 1Φ

and 2Φ are half of the group delays in 1P and 2P in magnitude. The residual range error in

the ionosphere-free phase combination is also half of that in the code combination. The

maximum vertical ionospheric electron content is typically about 100 TECU at a solar

maximum near the geomagnetic equator (Langley, 1997). Therefore, the residual range error

in ionosphere-free code combination can be up to –1.1 cm for a satellite at the zenith. The

Page 94: 7 Tesis PPP Kongzhe Chen

70

value can be about 3 times bigger for a satellite at 5° elevation angle. The precision of PPP

solutions is dominated by the un-modeled errors in the phase observations after the

ambiguity convergence. In the mid-latitude regions during a solar maximum, a typical

ionospheric content is about 40 TECU for a daytime average (Langley, 1997), which

corresponds a 2.2 mm residual range error caused by the second-order ionospheric phase

advances in the phase combination for a satellite at the zenith. Therefore, for the sub-

decimetre level position determination, the second-order ionosphere residuals are not critical,

but they will definitely affect the sub-centimetre level zenith tropospheric delay estimation,

which will be discussed in Chapter 6.

The third-order ionospheric group delay is about one order less than that of the second-order

ionospheric group delay. As investigated by Bassiri and Hajj (1993), for TEC=100 TECU,

the third-order ionospheric group delay is about 0.86 mm and 2.4 mm for 1P and 2P

respectively. The residual range error of the third-order ionospheric group delay in the

ionosphere-free code combination is –0.66 mm. The third-order ionospheric phase advances

in 1Φ and 2Φ are one third of the group delays in 1P and 2P . The residual range error in the

ionosphere-free phase combination is also one third of that in the code combination.

Therefore, the third-order ionospheric effects can be safely neglected in precise point

positioning, time transfer and meteorological applications.

It is obvious that the residual range errors of the higher-order ionospheric effects in the P1-

P2-CP ionosphere-free combinations are even smaller than those in the traditional

combinations. With GPS modernization, the measurements on three or more frequencies can

Page 95: 7 Tesis PPP Kongzhe Chen

71

be used to mitigate the ionospheric effects more effectively than the dual-frequency

combinations. For example, triple-frequency combinations can eliminate the ionospheric

effects up to the second-order (Xu, 2003).

For precise point positioning using single-frequency observations, the broadcast Klobuchar

model or the more accurate Global Ionospheric Model (GIM) can be used. The broadcast

Klobuchar model can mitigate about 50% of the ionospheric effects using coefficients in

navigation messages (Klobuchar, 1996). GIM has the potential to support sub-metre level

positioning using post-mission ionospheric products. Ionospheric effects mitigation for

precise point positioning using single-frequency GPS data will be investigated in Chapter 5.

3.2.3 Tropospheric Delay

The troposphere is the lower part of the Earth’s atmosphere. It extends to a height of less than

9 km over the poles and more than 16 km over the equator (Langley, 1998a). In the

troposphere, the refractive index n (n>1) is larger than that in a free space (n=1), which

causes the speed of radio signals to decrease below its speed in vacuum c. The spatially

varying refractive index also causes the signal path to have a slight curvature with respect to

the geometric straight path. Both of them constitute the tropospheric delay (Bevis et al.,

1992). The tropospheric delays in GPS measurements may be even bigger than the

ionospheric effects during ionospheric quiet periods, especially for satellites at low

elevations. But the tropospheric delays are much more stable and easier to model than the

ionospheric effects (Klobuchar, 1996).

Page 96: 7 Tesis PPP Kongzhe Chen

72

Usually, the total troposphere delay tropd is decomposed into two components, which are the

“hydrostatic delay” and the “wet delay”. The hydrostatic delay depends only on the surface

pressure and contributes to about 90% of the total delay. The wet delay is a function of the

water vapour distribution (Saastamoinen, 1972; Davis et al., 1985), which makes GPS a

promising tool for water vapour sensing. Mapping functions can be used to relate the slant

hydrostatic delay at any elevation to a Zenith Hydrostatic Delay (ZHD), and the slant wet

delay to a Zenith Wet Delay (ZWD). The ZHD, which is approximate 2.3 m at sea level, can

be calculated to better than 1 mm given the surface pressure measurement accurate to 0.3

mbar or better (Saastamoinen, 1972).

However, water vapour is a highly variable component in atmosphere. Because of the spatial

and temporal variability of water vapour, the ZWD can hardly be modeled to better than 1~2

cm even using precise meteorological measurements (Bevis et al., 1992). As discussed in

Section 2.4, ZWD can be estimated to centimetre level accuracy for meteorological

applications. Therefore, in this research, ZHD will be modeled while ZWD will be estimated

when processing dual-frequency GPS measurements. Issues related to ZWD estimation using

un-differenced GPS data have been discussed in Section 2.4.

3.2.4 Relativistic Effects

Relativity affects GPS positioning in several ways. Relativistic effects are relevant for

satellite orbit, satellite signal propagation and satellite and receiver clocks (Hofmann-

Page 97: 7 Tesis PPP Kongzhe Chen

73

Wellenhof et al., 2000). For point positioning users, regardless of whether broadcast or

precise ephemeris data is used, the following relativistic effects should be taken into account.

Relativistic Effects on Satellite Clock

GPS satellite clocks are affected by both special (relative velocity of the satellite) and general

(gravity field of the earth) relativistic effects. Therefore, the fundamental frequency of the

satellite clocks, which was selected as 10.23 MHz, has been slightly tuned by a constant

offset of –0.00457 Hz before satellite launch, based on the assumption of a circular orbit

(Ashby and Spilker, 1996).

Because of the eccentricity of the GPS orbits, satellite clocks show periodic variations. The

GPS community has already developed two conventions to calculate the periodic special

relativity variations (Ashby and Spilker, 1996):

cVX

Esineac

ssclk,rel

22 == µδρ (3.34)

where µ is the earth’s gravitational constant, a is the semi-major axis, e is the eccentricity,

E is the eccentric anomaly, sX and sV are the satellite position and velocity vectors, and c

is the speed of light. Users may select the first convention when using broadcast ephemeris.

On the other hand, the second one is often used for SP3 format ephemeris.

Page 98: 7 Tesis PPP Kongzhe Chen

74

Relativistic Effects on Signal Propagation

In addition to these special relativistic effects on the clock, there is still a small delay of the

GPS signal as it passes through the Earth’s gravity field. The corresponding correction is

given by (McCarthy and Petit, 2004):

−+++

=srsr

srsr

gra,rel RRRR

lnc ρ

ρµδρ2

2 (3.35)

where rR and sR are the geocentric distances of receiver and satellite. This term ranges

between about 12.7 mm and 18.7 mm. The effects can be mostly removed in differential GPS

data processing. But the effects should be taken into account in point positioning using un-

differenced data.

Sagnac Effects

Sagnac effects are caused by the Earth’s rotation during the time of the satellite signal

transmitting to the receivers. It can be expressed in the following form (Ashby and Spilker,

1996).

22 rse

sag,rel

rrc

×⋅

Ω+=δρ (3.36)

where eΩ is the Earth angular rotation rate (WGS-84), 7.2921151467×10-5 rad/s, sr and rr

are the position vectors of satellite and receiver at the instant a signal is transmitted,

respectively, and × denotes a vector cross product.

Page 99: 7 Tesis PPP Kongzhe Chen

75

This term can be tens of metres in value and should be taken into account all the time,

regardless of whether the broadcast or precise ephemeris data is used. Equation 3.36 is

widely applied when using SP3 format ephemeris. When broadcast ephemeris are used to

calculate the satellite coordinates, the Sagnac effects can be simply removed during the

calculation of the corrected longitude of ascending node of satellite as follows:

)tt()tt)(( eeee

.

∆+Ω−−Ω−Ω+Ω=Ω 0 (3.37)

where et is the reference time for ephemeris, 0Ω is the longitude of ascending node of

satellite at et , .

Ω is the rate of right ascension, t∆ is the signal propagation time.

3.2.5 Phase Wind-up

GPS satellites transmit right circularly polarized (RCP) radio waves. The electromagnetic

wave from the GPS satellite can be described as a rotating electric field propagating from the

satellite antenna to the receiver antenna. If both antennas keep static, the measured carrier-

phase would be the geometric angle between the instantaneous electric field at the receiver

antenna and a reference direction on the antenna. If the orientation of the receiver antenna or

satellite antenna changes, the reference direction or the direction of the electric field will also

change and thus the measured phase. Therefore, a carrier-phase measurement depends on the

orientation of the antennas of the satellite and the receiver as well as the direction of the line

of sight between the satellite and the receiver (Wu et al., 1993). The effect introduced to the

Page 100: 7 Tesis PPP Kongzhe Chen

76

phase measurements because of the rotation of antennas is called “phase wind-up” or “phase

wrap-up” (Wu et al., 1993; Tetewsky and Mullen, 1997).

The phase wind-up because of the rotation of receiver antenna can be removed completely by

double differencing or absorbed to the receiver clock offset in point positioning. Thus, the

receiver antenna rotation affects PPP timing results (Tetewsky and Mullen, 1997). In timing

applications, the receiver antennas are usually static. The receiver antenna rotation effects are

not considered in this research.

However, satellite antennas rotate slowly because satellites orient their solar panels towards

the Sun (Kouba, 2003). Unlike phase wind-up due to receiver antenna rotation, the effects of

satellite antenna rotation cannot be removed by double differencing because they depend on

the direction of the line of sight between the satellite and the receiver. As investigated by Wu

et al. (1993), the residual phase shift after double differencing could be as large as half a

cycle depending on the baseline length. Neither can they be absorbed to receiver clock offset

in PPP. In PPP, if they are not corrected, the constant part of the effects would be absorbed

into the float ambiguities, the variations will affect the ambiguity convergence and

subsequently affect the results. Wu et al. (1993) have provided two expressions to mitigate

the effects, which give the same results. One of them is more suitable for computer

programming and presented in the following.

As studied by Wu et al. (1993), the phase wind-up correction is determined by the angle

between the two effective dipoles of the satellite and the receiver, and its past history as

follows:

Page 101: 7 Tesis PPP Kongzhe Chen

77

φπ ∆+=∆Φ N2 (3.38)

where φ∆ is a fractional part of a cycle and can be calculated as follows (Wu et al., 1993):

)D'D/D'D(cos)(sign

⋅=∆ −1ζφ (3.39)

'D

and D

are the effective dipole vectors of the satellite and receiver computed from the

current satellite body coordinate unit vectors ( ',',' zyx ) and the local receiver unit vectors

( zyx ,, ), respectively. They are expressed as follows (Wu et al., 1993):

')'('' ykxkkxD ×−⋅−= (3.40)

ykxkkxD ×+⋅−= )( (3.41)

)'( DDk

×⋅=ζ (3.42)

where k

is the unit vector pointing from the satellite to the receiver.

The integer N is given as follows (Wu et al., 1993):

]/)int[(nN prev πφ 2∆−∆Φ= (3.43)

where intn denotes the nearest integer function and prev∆Φ is the previous value of phase

wind-up correction. Equation 3.43 is given based on the assumption that the data rate is high

enough so that the change in the correction is less than half a cycle between successive

computations (Wu et al., 1993). The value of N can be arbitrarily set to any integer,

Page 102: 7 Tesis PPP Kongzhe Chen

78

normally zero, at the beginning of a phase tracking session. It is obvious that the equations

above just provide the information of the change of phase wind-up from the beginning of a

phase tracking session. The initial phase wind-up angles would be combined into the

ambiguities (Leick, 2004).

The correction from Equation 3.38 is given in phase angle. The corrections in range are then

expressed as ∆Φ= 11 λω , ∆Φ= 22 λω , ∆Φ=−−

22

21

2211

ff

ffIF

λλω for 1Φ , 2Φ and IFΦ ,

respectively. From these expressions, it is obvious that an applicable method to remove phase

wind-up effects is to form the widelane phase combination.

Differing from carrier-phase measurements, the rotation of satellite or receiver antenna does

not affect the code measurements. Therefore, no wind-up corrections are required for the

code measurements (Leick, 2004).

3.2.6 Hardware Delays

The points where satellite signals are generated and transmitted are different. Also the phase

center of the receiver antenna is located at a different point from the signal correlation in the

receiver. Therefore, the signal is delayed by both satellite and receiver hardware. This type of

signal delays are called hardware delays or group delays. The hardware delays are different

for 1P and 2P , and the difference between them is called hardware bias, or group delay

differential ( gdT ) for satellites, and Inter-Frequency Bias (IFB) or Differential Code Bias

Page 103: 7 Tesis PPP Kongzhe Chen

79

(DCB) for receivers, respectively. In addition to the DCB between 1P and 2P , there is also

systematic difference between 1P and 1C , which is satellite dependent, and has been

discussed in Section 3.1.

In the case of double differencing between satellites and receivers, the hardware delays of

both satellites and receivers can be completely removed. In the case of point positioning,

where the hardware delay of the receiver is the same to all satellite measurements, it would

be completely absorbed to the receiver clock offset estimation when single-frequency

observations are used. For point positioning using dual-frequency observations, the receiver

hardware delay would be removed or absorbed to the float ambiguities depending on which

type of ionosphere-free combinations are used.

The group delay differentials ( gdT ) of satellites are provided in the navigation messages. The

corrections would benefit single-frequency users. They are also useful for some applications

in addition to positioning, such as ionospheric modeling. In navigation messages, the satellite

clock information is based on the dual-frequency ionosphere-free code combination (ICD-

GPS-200C, 2000). Therefore, the traditional ionosphere-free combinations are immune from

the hardware delays if the broadcast ephemeris is used.

As described by Kouba (2003), no group delay calibration corrections are applied for the

receiver and satellite ( 2L - 1L ) biases in IGS analyses. Thus, no calibrations are to be applied

when the IGS clock products are used in precise point positioning using the traditional

ionosphere-free combinations.

Page 104: 7 Tesis PPP Kongzhe Chen

80

According to ICD-GPS-200C (2000), users who only use the measurements on 1L frequency

shall modify the satellite clock correction with the following equation:

gdss Tdtdt −=1 (3.44)

where sdt is the satellite clock correction calculated from the clock correction coefficients in

the navigation messages. The user who only uses the measurements on 2L frequency shall

modify the code phase with the equation below:

gdss Tdtdt ⋅−= γ2 (3.45)

The value of gdT can be expressed as follows (ICD-GPS-200C, 2000):

( )2111

ttTgd −

−=

γ (3.46)

where it is the GPS time corresponding to the ith frequency signal transmitted from the

satellite.

Precise point positioning using single-frequency measurements will be discussed Chapter 5.

The accuracy of the broadcast gdT values directly affects the single-frequency point

positioning performance, especially after SA turned off. The ionosphere community has been

estimating the group delay differentials since 1993. In 1998, a world wide cooperative

analysis was initiated to determine new gdT values. The first set of new gdT values based on

JPL’s estimates was uploaded to satellite in April 1999. Currently JPL provides updated

Page 105: 7 Tesis PPP Kongzhe Chen

81

estimates of gdT for uplink every quarter or as needed and also monitors the values daily to

identify any abrupt changes in the gdT values due to configuration changes of the satellites

(Wilson et al., 1999).

Hardware delays of both satellite and receiver also exist in carrier phase observations. The

hardware delays in phase measurements would be removed completely by double

differencing. Because hardware delays are different between the code and phase

measurements (Xu, 2003), in precise point positioning, only parts of the hardware delays in

phase measurements are absorbed to the receiver clock estimate. Because phase observables

are inherently ambiguous, the residuals (not absorbed to clock estimates) of hardware delays

in phase measurements would be absorbed by the ambiguity parameters in point positioning

(Leick, 2004). The residuals, along with the initial phase offsets, destroy the integer nature of

ambiguities in point positioning. The integer parts of the residuals are combined into the

integer ambiguities, while the fractional parts would be combined into the initial phase

offsets which are discussed in the following.

3.2.7 Initial Phase Offsets

As shown in Equation 3.7, the un-differenced phase observable includes two non-zero initial

fractional phase terms. The first term, )t(ri 0φ , is the un-calibrated component of phase delay

in the receiver and would be common to all channels of the receiver. The second term,

)t(si 0φ , exists in the satellite transmitter and is different from satellite to satellite (Blewitt,

Page 106: 7 Tesis PPP Kongzhe Chen

82

1989). They are usually called initial (or fractional) phase offsets of the receiver and satellite

(Teunissen and Kleusberg, 1998; Han et al., 2001). The initial phase offsets are part of the

receiver- and satellite-generated signals. Carrier-phase measurements at a specific receiver

on a specific frequency should contain the same receiver-based initial phase offset. Similarly,

carrier-phase measurements to a specific satellite on a specific frequency should contain the

same satellite-based initial phase offset (Gabor and Nerem, 2002).

These initial phase offsets can be completely removed by double differencing between

satellites and receivers. They can also be treated as constants for each receiver or satellite so

that they can be removed by time differencing between phase measurements to the same

satellite at two consecutive epochs (Han et al., 2001). But initial phase offsets are absorbed

into the ambiguity in un-differenced phase measurement, thus they destroy the integer nature

of the ambiguity. Solving these items can help the ambiguity resolution in PPP thus improve

the accuracy. Currently, there are still no values of initial phase offsets available for satellites

and receivers. Gabor and Nerem (2002) have tried to estimate these items to reconstruct the

integer nature of the ambiguity, which has improved the mean baseline length repeatability

about 40~50 percent using the PPP method.

Because the hardware delays in phase measurements are also satellite and receiver

dependent, the integer parts of the hardware delays are absorbed to the integer parts of the

ambiguities, while the fractional parts are absorbed to the initial phase offsets.

Page 107: 7 Tesis PPP Kongzhe Chen

83

3.2.8 Multipath and Measurement Noise

Multipath is mainly caused by reflecting surfaces near the receiver antenna, such as metallic

objects, ground and water surfaces, or introduced by reflections at the satellite during signal

transmission (Hofmann-Wellenhof et al, 2000). The reflected signals may interfere with the

direct signals, and lead to noisier measurements. They may also confuse the signal tracking

in GPS receivers and lead to biased measurements (Langley, 1997).

The theoretical maximum multipath error that can occur in code measurements is

approximately half the code chip length or 150 m for C/A code ranges and 15 m for P(Y)

code ranges. Typical errors are much lower (generally less than 10 m). The carrier phase

multipath error can be up to about one-quarter of the wavelength, which corresponds to

approximately 4~5 cm for phase measurement on 1L or 2L (Braasch, 1996). A common

method for reducing code multipath effects is to smooth code measurements with carrier

phase measurements (Hatch, 1982). This method also reduces the noises of code

measurements. In PPP processing, multipath effects not only degrade the accuracy but also

delay the ambiguity convergence. A Choke Ring antenna would be necessary in the presence

of multipath.

Measurement noises exist in both code and carrier phase observations. The noise magnitude

in code and phase measurements is significantly different. Code noise, which is generally

less than 1% of the chipping rate (<3 m for C/A code and < 0.3 m for P(Y) code), is much

bigger than the phase noise, which is approximately 2 mm or equivalent to 1% of the

Page 108: 7 Tesis PPP Kongzhe Chen

84

wavelength (Langley, 1998b). Currently, because of the advance in signal processing

technology, the noise level of some receivers is much smaller than the theoretical one. As

given by Hofmann-Wellenhof et al. (2000), the typical noise is 10~300 cm, 10~30 cm and

0.2~5 mm for C/A code, P(Y) code and carrier phase, respectively. The noise level of the

ionosphere-free combinations has been discussed in Section 3.1.

Both multipath error and measurement noise are related to the elevation of the observed

satellite. Normally, they increase while the satellite elevation decreases. The residuals of the

ionospheric and tropospheric effects also show similar characteristics. The characteristics of

these error sources provide an approach for stochastic modeling, which will be discussed in

Section 3.3.

3.2.9 Antenna Phase Center Offset and Variations

Phase center offset and variations are relevant to both GPS satellite antenna and receiver

antenna. Antenna phase center is the point where a signal is transmitted or received.

However, it is not a well-defined physical point and varies with the changing direction of the

incoming or outgoing signals. The antenna phase center can be usually given in the form of a

mean offset from a physically defined point on the antenna, normally, antenna reference

point (ARP) is selected, plus a variation. Rothacher and Beutler (2002b) have described the

offset and variation using the following two terms:

Page 109: 7 Tesis PPP Kongzhe Chen

85

• A mean offset, 0r , which is a vector from the ARP to the location of the mean

phase center (average over a certain elevation range).

• An elevation ( e ) and azimuth (α ) dependent phase center variation ( )α,e∆Φ with

respect to the mean phase center 0r .

Both 0r and ( )α,z∆Φ are frequency dependent and different for 1L and 2L . The total

correction for the observation between the station r and the satellite s due to the receiver

antenna phase center offset and variations can be calculated using the following equation

(Rothacher and Beutler, 2002b):

( )ααδρ ,eer),e( srant ∆Φ+⋅= 0 (3.47)

where sre is the unit vector between the station r and the satellite s.

Receiver antenna phase center offset and variations may introduce up to 10 cm of vertical

positioning error, and sub-centimetre horizontal positioning error, if they are not accounted

for (Rothacher and Beutler, 2002b). The effects introduced by the phase center offset of

satellite antenna may be much bigger than those from receiver antenna because satellite

orbits can be referred to a point far from the phase center. The phase center offset of GPS

satellite antenna can be more than one metre for post-mission precise GPS orbit and clock

products. Unlike orbit and clock information provided by the broadcast ephemeris, which

refer to the phase center of the satellite antenna, precise GPS orbit and clock products could

refer to any point chosen by the product providers (Zhu et al., 2003). For example, the force

Page 110: 7 Tesis PPP Kongzhe Chen

86

models used by the IGS community for satellite orbit modeling refer to the center of mass,

their precise GPS satellite coordinates also refer to the same point in IGS products (Kouba,

2003). Normally, the offsets between the phase center and mass center are calibrated before

the satellite launch. For Block IIR satellites, the phase center and mass center locate at the

same point. The offsets for Block II/IIA satellites are normally provided in the precise

products. Real-time precise orbit and clock products are normally based on the broadcast

ephemeris and also refer to the satellite phase center.

IGS has started to standardize the receiver antenna names since 1992 (Rothacher and Beutler,

2002b). Currently, phase offset and variations information of both GPS satellite antennas and

a subset of GPS receiver antennas have been provided in the IGS antenna phase center offset

tables (Rothacher and Mader, 2002a; Kouba, 2003). The variations have been given in two

formats, elevation-dependent, and elevation- and azimuth-dependent (Rothacher and Beutler,

2002b).

The National Geodetic Survey (NGS) of USA has also calibrated a subset of GPS antennas

(Mader, 1999). The format is similar to IGS elevation-dependent format. NGS calibrates

antennas in both relative and absolute methods, while IGS provides only relative calibrated

values. In both NGS and IGS relative calibration, all antenna phase center offsets and

variations are calibrated with respect to the AOAD/M_T antenna, which is a choke ring

antenna and is the most common antenna type within the global network of the International

GPS Service (IGS).

Page 111: 7 Tesis PPP Kongzhe Chen

87

3.2.10 Site Displacement Effects

The Earth is not a rigid body, and therefore is deformed by the gravitational potential. The

gravitational potential in the vicinity of the Earth is a combination of the tidal gravitational

potential of external bodies, including the Moon, the Sun, and other planets, and the Earth's

own potential. Both the external potential and the Earth's own potential contain time

independent part, which is permanent, and time dependent part, which is periodic. The tidal

potentials cause solid Earth deformations, and thus lead to site displacements, which also

include permanent and periodic parts. Correspondingly, two crusts, namely “mean tide” crust

and “tide free” crust, are introduced by IERS (International Earth Rotation and Reference

Systems Service). Site positions are on the “mean tide” crust if the time dependent part of the

tidal contributions is removed. It is obvious that the permanent deformation is still present in

the “mean tide” crust. By removing the permanent deformation from the “mean tide” crust, a

“tide free” crust is obtained. However, a truly “tide free” crust is unobtainable because the

gravitational potential is present all the time. Instead, the “conventional tide free” crust is

usually used (McCarthy and Petit, 2004). In the following, some models, which are used by

IERS to get the “conventional tide free” terrestrial reference frame, will be presented. The

coordinates after applying the corrections from the models are also called “conventional tide

free”, or simply “tide free”, coordinates (McCarthy and Petit, 2004).

Actually, these tidal corrections are not necessarily applied in PPP data processing all the

time. When they should be applied in data processing depends on where the datasets are

collected. For example, when airborne dataset, which is free from the tidal effects of the earth

Page 112: 7 Tesis PPP Kongzhe Chen

88

(Xu, 2003), is processed using PPP method, no tidal corrections are required to be applied to

get the tide-free coordinates. However, sometimes users would find big discrepancies

between the coordinates obtained from PPP and coordinates from differential techniques in

which the base stations are set on the Earth. This is because the differential software has not

considered the differences in the tidal effects between the base on the Earth and rover in the

air, or the software has not been set to process baseline between stations on the Earth and in

the air.

Solid Earth Tides

Deformations due to solid earth tides are mainly caused by the gravitational attractions acting

on the elastic body of the Earth by the Moon and the Sun. The effects of solid earth tides can

reach about 30 cm in the radial direction and 5 cm in the horizontal direction (Kouba and

Heroux, 2001a). Solid earth tides are represented by the Love number hnm and the Shida

number lnm, which weakly depend on station latitude and tidal frequency (McCarthy and

Petit, 2004).

Models have been provided by IERS to calculate the displacement due to the second-degree

tides and displacement due to the third-degree tides. For a 5 mm precision in height

component, only the second-degree tides are necessary (Kouba and Heroux, 2001a).

The following equation can be used to calculate the second-degree tides (McCarthy and Petit,

2004):

Page 113: 7 Tesis PPP Kongzhe Chen

89

( ) ( ) ( )[ ]

⋅−⋅+

−⋅=∆ =

rrRRrRlrRrhRGM

RGMr jjjj

j je

ej2

2

2

3

23

4

321

23

(3.48)

where,

jGM is the gravitational parameter for the Moon (j = 2) or the Sun (j = 3),

eGM is the gravitational parameter for the Earth,

jR , jR is unit vector from the geocenter to Moon or Sun and the magnitude of that vector,

eR is the Earth's equatorial radius,

r is the unit vector from the geocenter to the station,

2h nominal degree 2 Love number,

2l nominal degree 2 Shida number.

Figure 3.1 shows solid earth tide effects for station S1 on roof of the Engineering Building at

the University of Calgary from September 2nd, 2004 to September 7th, 2004. As shown in the

figure, the effects can be up to 15 cm in height, and 5 cm in north and east.

Page 114: 7 Tesis PPP Kongzhe Chen

90

Figure 3.1 Solid Earth Tide Effects for S1 from September 2nd to 7th, 2004

Ocean Loading

Site displacement due to ocean loading is caused by ocean tides. Ocean tides change the

ocean mass distribution and the water load on the crust, thus cause time-varying

deformations of the Earth (McCarthy and Petit, 2004). The deformations are more significant

for coastal sites than sites far from the sea. The displacement can be several centimetres for

coastal sites in the height component. The following simplified models have been provided

by IERS (McCarthy and Petit, 2004) to calculate the corrections in height, east and north

components for a particular site at time t:

Page 115: 7 Tesis PPP Kongzhe Chen

91

( )hjjjjj

jhj tcosfAh Φ−++=∆ =

µχω11

1

(3.49a)

( )wjjjjj

jwj tcosfAe Φ−++−=∆ =

µχω11

1

(3.49b)

( )sjjjjj

jsj tcosfAn Φ−++−=∆ =

µχω11

1

(3.49c)

The summation over j represents the eleven partial tides including the semidiurnal waves M2,

S2, N2, K2; the diurnal waves K1, O1, P1, Q1; and the long-period waves Mf, Mm, and Ssa. The

nodal parameters, jf and jµ , depend on the longitude of the lunar node and can be

calculated from equations provided by IERS. The astronomical argument jχ at t = 0h can be

computed using subroutine from IERS also. The angular velocities jω at t = 0h are provided

by IERS as constants. hjA , wjA and sjA are the amplitudes in height, west and south

components. hjΦ , wjΦ and sjΦ are the phases in height, west and south components. Both

the amplitudes and phases for particular site can be calculated from some ocean loading

service agencies. The coefficients are applicable for stations within 10 km from the particular

site (McCarthy and Petit, 2004). In kinematic positioning, the coefficients can be interpolated

from precomputed sites to the receiver position at each epoch. In this research, coefficients

provided by Scherneck (2003) are used.

Figure 3.2 shows ocean loading effects for the station S1 from September 2nd to 7th, 2004.

Ocean loading effects are about one order smaller than those of solid earth tides for S1.

Page 116: 7 Tesis PPP Kongzhe Chen

92

Figure 3.2 Ocean Loading Effects for S1 from September 2nd to 7th, 2004

Pole Tide

The pole tide is caused by the centrifugal effect of polar motion. The variation of station

coordinates caused by the pole tide can be up to a couple of centimetres (McCarthy and Petit,

2004). The displacements in the height, east and north components can be calculated by the

following equations (McCarthy and Petit, 2004):

( )λλφ sinmcosmsinh 21232 +−=∆ (3.50a)

( )λλφ cosmsinmcose 2129 −=∆ (3.50b)

Page 117: 7 Tesis PPP Kongzhe Chen

93

( )λλφ sinmcosmcosn 2129 +=∆ (3.50c)

where the corrections are given in millimetres. φ and λ are the latitude and longitude of the

station. 1m and 2m represent the variations of the pole from the mean pole, given in seconds

of arc which can be expressed as follows (McCarthy and Petit, 2004):

pp xxm −=1 (3.51a)

( )pp yym −−=2 (3.51b)

where ( px , py ) are the pole position coordinates in seconds of arc, which are provided by

IGS or some other agencies in real-time or post-mission. ( px , py ) are the mean pole position

coordinates in seconds of arc, which are estimated by IERS in a yearly basic. The mean pole

position coordinates can also be calculated as follows (McCarthy and Petit, 2004):

)t(x)tt()t(x)t(x ppp 000−+= (3.52a)

)t(y)tt()t(y)t(y ppp 000−+= (3.52b)

where )t(x p 0 =0.054, )t(y p 0 = 0.357. px and py are given in seconds of arc.

)t(x p 0 =0.00083 and )t(y p 0

=0.00395, in seconds of arc per year. t is given in year and 0t

is 2000.

Page 118: 7 Tesis PPP Kongzhe Chen

94

Because 1m and 2m vary at most 0.8 second of arc, the maximum height and horizontal

displacements are approximately 25 mm and 7 mm, respectively (McCarthy and Petit, 2004).

Figure 3.3 shows pole tide effects for station S1 from 2000 to 2004. As shown in the figure,

pole tide changes very slowly. In a short period, the effects vary only a small amount. In

position solutions, not considering the pole tide would introduce biases. In timing and

meteorological applications, if station coordinates are estimated along with receiver clock

offset and tropospheric delay, pole tide effects can be absorbed into coordinate estimation

and safely neglected.

Figure 3.3 Pole Tide Effects for S1 from 2000 to 2004

Page 119: 7 Tesis PPP Kongzhe Chen

95

Atmospheric Loading

Atmospheric loading is caused by the variations of atmospheric pressure. The temporal

variations in the distribution of atmospheric mass can cause site displacement up to a couple

of centimetres (McCarthy and Petit, 2004). The site displacement due to atmospheric loading

has been widely investigated (vanDam and Wahr, 1987; vanDam et al., 1994; Sun et al,

1995). Pressure loading effects are larger at high-latitude sites because of more intensive

weather there. Effects are smaller for sites near the sea due to the inverted barometer

response of the sae surface, which is characterized by that changes in sea surface height is

inversely proportional to changes in the surface pressure. The horizontal deformations are

about one-third the amplitude of the vertical deformations (McCarthy and Petit, 2004).

Several methods for computing atmospheric loading corrections have been developed so far.

The geophysical model approach was adopted by the IERS at the 2002 IERS Meeting in

Munich, Germany. Regression coefficients determined from a geophysical model could be

used along with local pressure measurements to calculate the vertical deformations, while the

much smaller horizontal deformations are usually neglected in this model. The regression

coefficients can be determined by fitting local pressure to the vertical deformation predicted

by the geophysical model (McCarthy and Petit, 2004).

The regression coefficients for a subset of IGS stations have been estimated by the Special

Bureau on Loading (SBL), which was established by the IERS in February 2002. The

regression coefficients for a given site are applicable for stations within 10 km from the

given site (vanDam, 2003). In kinematic positioning, the regression coefficients can be

Page 120: 7 Tesis PPP Kongzhe Chen

96

interpolated from precomputed sites to the receiver position at each epoch. In this research,

regression coefficients provided by vanDam (2003) are used. In addition to regression

coefficients, atmospheric pressure measurements are also required to calculate the

deformation.

Figure 3.4 shows atmospheric loading effects for station S1 from 1995 to 1999. The data was

provided by vanDam (2003). The pressure variations were calculated by removing the mean

pressure over the period from the observed pressures. From the figure, we can see

atmospheric pressure can change the station height up to 1 cm with millimetre level changes

in north and east. The changes in station height show strong negative correlation with the

changes in pressure. Regression coefficients are estimated based on the relationship between

changes in station height and atmospheric pressure on the station.

Figure 3.4 Atmospheric Loading Effects for S1 from 1995 to 1999

Page 121: 7 Tesis PPP Kongzhe Chen

97

3.3 Modeling for Dual-Frequency Measurements

In this research, the Kalman filter was used to perform optimal estimation. The principle of

the Kalman filter was first introduced by Kalman in 1960 and has been extensively described

in literature (Gelb, 1974; Brown and Hwang, 1992). In the Kalman filter, the relationship

between the measurements and the parameters is described by the functional (or

mathematical) model, while the noise characteristics of the measurements are described by

the stochastic model.

3.3.1 Functional Model

Functional models should be specified to describe the relationship between observables, i.e.

the ionosphere-free code and phase combinations in this research, and the unknown

parameters including receiver clock offset, tropospheric delay parameters, ambiguities,

and/or receiver coordinates. In the following, the observables and unknowns will be

described.

Observables

Dual-frequency GPS measurements and ionosphere-free combinations have been discussed

in Section 3.1. As already discussed, there are two types of ionosphere-free combinations that

could be used for precise point positioning. Although the C1-P1-CP combinations can be

used to facilitate the fixing of integer ambiguities, the integer nature of the ambiguity is

Page 122: 7 Tesis PPP Kongzhe Chen

98

currently corrupted by the existence of non-zero initial phase offset and hardware delay in

un-differenced phase measurements. In this research, the more widely used traditional

ionosphere-free combinations are used for position determination, receiver clock and

tropospheric delay estimation. Equation 3.12 and 3.13 are rewritten in the following as the

primary observables in the Kalman filter.

)P(dmddd)dtdt(c

ffPfPf

P

IFIFreltroporbs

rsr

IF

ερ +++++−+=

−⋅−⋅= 2

22

1

22

212

1

(3.53)

)( mwFddd)dtdt(c

ffff

IFIFIFIFreltroporbs

rsr

IF

Φ+++++++−+=

−Φ⋅−Φ⋅=Φ

εδδρ

22

21

22

212

1

(3.54)

where

22

21

202022101011

ff)N)t()t((cf)N)t()t((cf

Fsrsr

IF −+−−+−

=φφφφ

(3.55)

The ionosphere-free code combination can be smoothed by the ionosphere-free phase

combination to reduce its noise and multipath (Hatch, 1982):

n)n(

nn

P

))(P(n

nn

PP

nIF

n

i

iIF

n

i

iIF

nIF

nIF

nSM

nIFn

SM

Φ−+Φ

−=

Φ−Φ+−+=

==

−−

1

1

1

11

11

(3.56)

Page 123: 7 Tesis PPP Kongzhe Chen

99

where nSMP and 1−n

SMP are the smoothed code combinations at the nth and the (n-1)th smoothing

epochs, respectively, iIFP and i

IFΦ are the ionosphere-free combinations of code and phase at

the ith smoothing epoch. Phase smoothing is also helpful to shorten the ambiguity

convergence time.

Unknowns

The number of unknown parameters varies for different applications. The receiver clock

offset and ambiguities, which are usually considered as nuisance parameters, should be

estimated all the time. In this research, to mitigate tropospheric delay, the zenith wet delay

will be estimated, instead of estimating the total zenith tropospheric delay that has done in

some researches.

As discussed in Section 2.4, estimating troposphere horizontal gradients can improve both

zenith wet delay estimation and the repeatability of the static site position, but will also

introduce two additional parameters. To estimate horizontal gradients, GPS observations at

low elevation angles must be included to separate gradient components from the azimuthally

homogeneous components. In static positioning and meteorological research, when antenna

is normally set up in good observing condition and multipath-friendly environment, satellites

at low elevation, i.e. down to 7°, can be tracked and used for data processing. In this

research, tropospheric gradients will be estimated in static positioning and meteorological

processings. Estimating tropospheric gradients is also applicable in timing application, as

timing receivers are normally set up in good observing condition. Improved zenith wet delay

Page 124: 7 Tesis PPP Kongzhe Chen

100

estimates can also improve the receiver clock offset estimation due to strong correlation

between them (Hackman and Levine, 2003). In kinematic positioning, when observing

condition varies, the satellites tracked may be less than the number required to obtain a

solution for gradients. Multipath effects may also be strong for satellites at low elevations

under kinematic environments. Therefore, only zenith wet delay will be estimated in

kinematic positioning.

In timing and meteorological applications, a GPS antenna is usually set up at a fixed site. In

post-mission, the antenna coordinate parameters can be resolved first, and then be removed

from the unknowns. But in this research, they will be estimated in real-time or simulated

real-time data processing.

In summary, the state vector for static positioning, timing and atmospheric sensing

applications is:

( )TnENwzr F,,F,F,G,G,D,dt,h,, 21λϕ (3.57)

where ϕ , λ , and h are the latitude, longitude and height of the station. rdt is the receiver

clock offset. wzD , NG and EG are the zenith wet delay, horizontal gradients in north and east

directions. n is the number of satellites included in processing. iF is the 1L and 2L combined

float ambiguity in the ionosphere-free phase combination, described in Equation 3.55.

The state vector for kinematic positioning is:

Page 125: 7 Tesis PPP Kongzhe Chen

101

( )Tnwzr F,,F,F,D,dt,h,, 21λϕ (3.58)

where there are no tropospheric gradient parameters.

3.3.2 Stochastic Model

The stochastic model describes the statistical properties of the observables, including both

the precision of observables and the correlation between the observables. The stochastic

model is generally given in the form of the variance-covariance (vc-) matrix, in which the

diagonal terms are the variances and the off-diagonal terms are the covariances (or

correlations) between the observables. To get optimal estimates, the inverse of the vc-matrix

of the observables serves as the weight matrix in the Kalman filter (Tiberius et al., 1999).

Errors in the stochastic model will not only affect the estimation accuracy but also the

ambiguity convergence.

In GPS positioning, different weighting schemes have been applied for satellites at different

elevation angles. The simplest one is equal weight for all observations. This scheme is not

adequate to describe the actual stochastic relationship among the observables, especially for

observables at low elevation (they are included to estimate the gradient parameters in

meteorological applications) because the noise and multipath effects increase when elevation

angle decreases. Another weighting scheme is based on the signal-to-noise ratio (SNR)

information of observables (Talbot, 1998). The SNR can partly reflect the precision of

observables. The problem of this scheme is that not all receivers will output such

Page 126: 7 Tesis PPP Kongzhe Chen

102

information. Although there are fields in RINEX file reserved for SNR information, it is not

uncommon for there to be no data in those fields in the RINEX files or for the data to be

provided in the GPS receiver manufacturer’s own proprietary format. Given the above, this

scheme is not applicable or would lead to incorrect modeling in some cases, especially in

real-time.

Currently, a widely used scheme is the elevation-dependent weighting (Rothacher et al.,

1997). Measurement noise and multipath error increase significantly for observables at low

elevation. The elevation-dependent measurement noise is induced mainly by the receiver

antenna’s gain pattern. Atmospheric attenuation to GPS signal also contributes to the

elevation-dependent measurement noise (Tiberius et al., 1999). The increased multipath

errors in observations acquired from low elevation satellites are due to reflection property of

the circularly polarized GPS signals (Braasch, 1996). The troposphere and ionosphere

residuals are also elevation-dependent because of the elevation-dependent mapping

functions. In this research, the following weighting function is used:

( ) ( )esinew 2= (3.59)

where e is the elevation angle.

The variance and covariance of the code and phase measurement should be given properly to

fully exploit the precise phase measurements. The measurement noises in code and phase

measurements have been discussed in Section 3.2. The noise levels of 1 m and 1 cm have

been used for code and phase measurements by Zumberge et al. (1998). In this research, the

Page 127: 7 Tesis PPP Kongzhe Chen

103

noise levels of 0.3 m and 2 mm are used for code and phase measurements. For satellite at

elevation e, the variance of ionosphere-free combinations will be determined using the

following equations:

( ) ( ) ( )esinesine PP

PIF

IF 2

2

2

22 1

3σσσ ≈= (3.60)

( ) ( ) ( )esinesine IF

IF 2

2

2

22 1

3 ΦΦ

Φ≈=

σσσ (3.61)

where, 21Pσ and 2

1Φσ are the variances of the code and phase measurements for satellite at

zenith, equal to 9×10-2 m2 and 4×10-6 m2, respectively.

Phase smoothing is used in this research for the reduction of the noise level and multipath

effects in the code measurements. The smoothed code and phase observables become

correlated after the smoothing process. Following the smoothing scheme given in Equation

3.56 and assuming that the code and phase measurements are independent before smoothing,

the variance and covariance for the smoothed ionosphere-free code and the ionosphere-free

phase observables will be determined as follows:

Page 128: 7 Tesis PPP Kongzhe Chen

104

( )

( ) ( )

( )

−−++

=

−−−=Φ

ΦΦ

ΦΦ

=

Φ

Φ

2

2

2

2

2

1

1

1

22

2

2

2

2

1

11

11

01

01

01

01

10000

11111

1

1

nIF

nIF

nIF

nIF

iIF

iIF

nIF

IF

nIF

IF

n

nn

n

n

n

nn

n

n

n

n

nn

nnnn,PCov

n

i

n

iP

P

P

nIF

nSM

σσ

σσσσ

σ

σσ

σ

(3.62)

where 2

iIFP

σ and 2

iIFΦ

σ are the variance of the ionosphere-free combinations of code and

phase at the ith smoothing epoch.

To avoid overweighting the code measurement, an alternative smoothing scheme proposed

by Lachapelle et al. (1986) can be used. They proposed a sliding weight scheme and the

weights of code and phase in the smoothed code measurement would be held fixed after a

selected amount of smoothing epochs.

Page 129: 7 Tesis PPP Kongzhe Chen

105

CHAPTER 4

REAL-TIME PRECISE GPS ORBIT AND CLOCK

PRODUCTS AND ANALYSIS

An important element of precise point positioning is precise GPS orbit and clock products.

Today, the advent of several types of real-time precise GPS orbits and clocks has made real-

time PPP possible. In post-mission, the performance of PPP is only affected by the accuracy

and sample interval of precise GPS products. In real-time, however, the precise GPS products

will cause additional degradations to the accuracy of PPP through their latency and age. The

latency and age of the precise GPS products depend on the data processing strategy,

communication issues, and data format of the products.

The purpose of this chapter is to present the development of real-time precise orbit and clock

products, followed by the latency and age analysis of two types of real-time precise GPS

products. The communication issues and formats will also be discussed as how to reduce the

latency and age. For a comparison purpose, some IGS post-mission products will be

discussed at first.

Page 130: 7 Tesis PPP Kongzhe Chen

106

4.1 IGS Precise Orbit and Clock Products

Since 1994 the International GPS Service (IGS) has been providing precise GPS products to

the scientific community with increased accuracy and timeliness (Kouba and Heroux,

2001a). Today, IGS products with different latencies and accuracies are available to GPS

users. Table 4.1 demonstrates the characteristics of different IGS precise satellite orbit and

clock products with respect to different accuracies and timeliness. Compromises are

necessary between “Accuracy” and “Timeliness” when choosing IGS products. From IGS

Ultra-Rapid (predicted half) to IGS Final, the corresponding accuracy of products increases

as well as the latency.

Table 4.1 IGS Products of GPS Satellite Orbit and Clock (after IGS Website, 2004)

Satellites Orbits / Clocks Accuracy Latency Updates Sample Interval

Orbits ~200 cm 2 hours Broadcast

Clocks ~7 ns real-time daily

2 hours

Orbits ~10 cm 15 minutes IGS Ultra-Rapid (predicted half) Clocks ~5 ns

real-time four times daily 15 minutes

Orbits <5 cm 15 minutes IGS Ultra-Rapid (observed half) Clocks ~0.2 ns

3 hours four times daily 15 minutes

Orbits <5 cm 15 minutes IGS Rapid

Clocks 0.1 ns 17 hours daily

5 minutes

Orbits <5 cm 15 minutes IGS Final

Clocks <0.1 ns ~13 days weekly

5 minutes

IGS orbit and clock products are based on a global tracking network and several contributing

Analysis Centers. The IGS global tracking network consists of more than 300 permanent,

Page 131: 7 Tesis PPP Kongzhe Chen

107

continuously operating GPS stations. This number is increasing yearly as more stations are

added to the network. As of December 19th, 2004, the total number is 381 (IGS website,

2004). Figure 4.1 shows the IGS global tracking network.

Figure 4.1 IGS Tracking Network on December 19th, 2004 (from IGS Website, 2004)

Currently, eight IGS Analysis Centers (AC) contribute daily ultra-rapid, rapid and final GPS

orbit and clock solutions to the IGS combinations (Kouba, 2003). As investigated by Beutler

et al. (1995), the combination procedure can average out the random-like noises produced by

different analysis centers using different approaches and modeling, and typically result in

more robust and precise solutions. Figure 4.2 shows the orbit RMS (mm) of the individual

AC solutions with respect to the IGS Final orbit combination. Figure 4.3 shows the clock

RMS (mm) of the individual AC solutions with respect to the IGS Final clock combination.

For display purposes the RMS values of the summaries in Figures 4.2 and 4.3 are weekly

averaged. Both figures were obtained from the website of IGS Analysis Center Coordinator

(ACC) at GFZ Potsdam.

Page 132: 7 Tesis PPP Kongzhe Chen

108

Figure 4.2 AC Solutions with Respect to the IGS Final Orbit Combination (from IGS ACC

Website, 2004)

Figure 4.3 AC Solutions with Respect to the IGS Final Clock Combination (from IGS ACC

Website, 2004)

Page 133: 7 Tesis PPP Kongzhe Chen

109

From Figures 4.2 and 4.3, we can see that IGS orbit and clock products have been improving

in the past years. This is because more and more stations have been added to the network and

the data processing strategy has been improved over years.

Precise products with different accuracies and latencies can be applied to different

applications. For example, the IGS Final and Rapid products, whose 5-minute clock

sampling allows an interpolation of SA-free satellite clocks well below the decimetre level

(Zumberge and Gendt, 2000), can be used to support sub-centimetre to sub-decimetre level

static and kinematic positioning in post-mission to achieve the highest accuracy (Kouba and

Heroux, 2001a), or used to support climate and timing transfer researches. The IGS Ultra-

Rapid (observed half) products with just slightly degraded accuracy and a latency of 3 hours

can still satisfy near real-time PPP applications. The IGS Ultra-Rapid (predicted half)

products are available in real-time. Although Ultra-Rapid (predicted half) clocks are not

accurate enough for carrier phase-based precise point positioning, the Ultra-Rapid orbits,

which is much more accurate than the broadcast orbits, can be used to improve results of

differential data processing in real-time. For example, they have been widely used in ground-

based GPS meteorology (Fang et al., 2001; Douša, 2001). As for real-time PPP application,

real-time orbit and clock products from organizations such as JPL and NRCan, which will be

discussed in the following, are much better than IGS Ultra-Rapid (predicted half), because of

their much more accurate clocks and shorter sample intervals. To save the bandwidth for data

transfer, the real-time orbit and clock products are usually broadcast in the format of

corrections to the broadcast orbits and clocks. IGS also plans to generate real-time products

Page 134: 7 Tesis PPP Kongzhe Chen

110

based on a real-time tracking network. The Real Time Working Group (RTWG) was

established in 2002 to develop real-time infrastructure and processes for IGS real-time

products. Its current focus is on the development of a prototype architecture which makes use

of data from a number of agencies’ real-time subnetworks (IGS RTWG website, 2004).

Currently, there are about 20 stations which can share real-time GPS data to IGS Analysis

Centers, with additional stations scheduled to join the real-time network (MacLeod, 2004).

The IGS orbit/clock products are consistent with the IGS global reference frame which

conforms to the International Terrestrial Reference Frame (ITRF) (Kouba, 2003). Therefore,

PPP using IGS products offers a simple and direct access to the IGS realization of ITRF.

Since its beginning in 1994, IGS has used six different official realizations of ITRF (ITRF92,

ITRF93, ITRF94, ITRF96, ITRF97 and ITRF2000). IGS products have adopted the IGS

ITRF2000 realization (IGS00) since December 2, 2001 (GPS Week 1143) (Kouba, 2003).

Kouba (2002) has also given the transformation parameters between frames and a program to

do transformation.

4.2 NRCan Real-Time Precise GPS Products

The Geodetic Survey Division (GSD) of Natural Resources Canada (NRCan), an Analysis

Center (AC) of the IGS, has provided two types of precise GPS orbit and clock products,

which are real-time or will be real-time in the future. One is based on code processing using

data from a wide area GPS tracking network in Canada. The second type is based on phase

Page 135: 7 Tesis PPP Kongzhe Chen

111

processing using data from a global GPS tracking network. The parameters of these two

products are described in Table 4.2. Details of the tracking network and the accuracy will be

discussed in the following.

Table 4.2 NRCan Real-Time Precise GPS Products (after Heroux, 2003; Collins, 2004)

Satellite Orbit / Clock Products Accuracy Latency Updates Sample Interval

Orbits ~20 cm 20 s GPS•C (code solution)

Clocks ~2 ns ~5 s 2 s

2 s

Orbits ~10 cm 20 s GPS•C (phase solution)

Clocks ~1 ns ~8 hours 2 s

2 s

4.2.1 GPS••••C Code Solution Products

In Canada, the Canadian Active Control System (CACS) was established to improve the

performance of GPS positioning and to provide a direct access to the Canadian Spatial

Reference System (CSRS). Based on CACS, precise GPS satellite orbits and clocks have

been estimated daily since 1992. In 1996, the accuracies of the precise orbits and clocks are

about 10 cm and 1 ns, respectively. Since 1996, the Geodetic Survey Division (GSD) of

Natural Resources Canada (NRCan) has developed a prototype of real-time GPS Correction

Service (GPS•C) based on the CACS stations (Caissy et al., 1996). Figure 4.4 shows the

CACS network.

Page 136: 7 Tesis PPP Kongzhe Chen

112

Figure 4.4 CACS Network (after ICD-GPS••••C, 2001)

The GPS•C system consists of four components, which are the Real-Time Active Control

Points (RTACPs), Real-Time Master Active Control Station (RTMACS), Virtual Active

Control Points (VACPs) and Integrity Monitoring Stations (IMSs) (ICD-GPS•C, 2001).

The RTMACS is used to control and acquire GPS and meteorological data from all RTACPs,

which are set up for data collection, validation and communication. The data is processed at

the RTMACS to generate corrections to the broadcast orbits and clocks, and ionospheric

vertical delay grids. These corrections are then sent to the VACP and IMS stations via

multicast network. VACPs are the primary distribution points of GPS•C corrections. IMSs

Page 137: 7 Tesis PPP Kongzhe Chen

113

are set up to monitor the differences between the local GPS corrections and the GPS•C

derived corrections (ICD-GPS•C, 2001).

After several years of tests, the GPS•C service was launched on October 14, 2003 (Kassam,

2003). Currently, real-time GPS•C products are broadcast via the MSAT (Mobile Satellite

system). CDGPS radios required for GPS users to receive the GPS•C products over satellite

downlink have also been developed. GPS•C products are encoded in a modified RTCA-159

(MRTCA) format which includes a subset of RTCA-159 messages and GPS•C proprietary

messages. The CDGPS radios can localize the corrections to the standard RTCM-104 format,

which enables single-frequency pseudorange users to enhance their positioning precision

(Kassam, 2003). For the users equipped with dual-frequency receivers, CDGPS radios will

also relay the GPS•C orbit and clock products in the modified RTCA-159 format, allowing

the most demanding GPS•C users to achieve the highest possible accuracy. Before GPS•C

products were broadcast by MSAT geostationary satellites, they have been broadcasted to

users over the Internet using multicast technology (Chen et al., 2002). NRCan plans to keep

this service in the future. When broadcast over the Internet, GPS•C products are

encapsulated in the GPS•C IP Multicast format for transmission (ICD-GPS•C, 2001). Figure

4.5 shows the CDGPS Radio.

Page 138: 7 Tesis PPP Kongzhe Chen

114

Figure 4.5 CDGPS Radio (after CDGPS Receiver User's Guide, 2003)

The GPS•C code solution products comprise corrections to broadcast satellite orbits and

clocks, and ionospheric parameters. When broadcast over the Internet, the maximum update

interval is 2 s for clock corrections, 120 s for orbit corrections, and 300 s for ionospheric

parameters. When broadcast over satellites, the update interval of clock corrections is 2 s,

and 20 s for orbit corrections. The GPS•C corrections are based on the NAD83 reference

frame. In Chapter 6, some results obtained using real-time GPS•C code solution orbit and

clock products will be presented.

4.2.2 GPS••••C Phase Solution Products

The GPS•C products were primarily based on processing code observations from CACS

network. Therefore, the accuracy of the products is limited and only practicable for code

processing. Since 2001, NRCan has been enhancing its real-time GPS•C system with carrier

phase processing. This development is aimed at improving the quality of the clock

Page 139: 7 Tesis PPP Kongzhe Chen

115

corrections to facilitate carrier phase-based precise point positioning (Collins, et al., 2001).

However, the performance of the corrections based on carrier phase processing using data

from CACS network is not satisfactory, especially for users at the boundary of CACS

tracking network (Collins, 2004). This is because it takes time for the ambiguities in phase

solution to converge. Each satellite is tracked by CACS network twice daily. The tracking arc

only consists of several hours of data for each pass. The geometry of each satellite is poor

because each satellite is tracked by stations in a limited region. Therefore, users at the

boundary of CACS tracking network can just use the degraded products from processing the

phase observations with un-converged ambiguities, which are only as accurate as those of

code solution. Therefore, to generate phase solution products, a network with global

coverage is preferred.

In 2004, NRCan tried to acquire real-time data from some stations operated by different

agencies outside Canada, some of which are IGS real-time stations. In the middle of 2004,

data from 20 stations can be acquired in near real-time by NRCan. As investigated by Collins

(2004), these 20 stations can fulfill a global coverage with satellites that can be tracked by a

minimum of 4 stations most of the time. Figure 4.6 shows the 20 stations sharing near real-

time data with NRCan. The stations are plotted in different colours indicating different

agencies operating the stations. Figure 4.7 shows the global coverage using the 20 stations.

The colours indicate the number of stations observing each satellite sub-point.

Page 140: 7 Tesis PPP Kongzhe Chen

116

Figure 4.6 Stations Sharing Real-Time Data with NRCan (after Collins, 2004)

Figure 4.7 The Global Coverage with 20 Stations (from Collins, 2004)

It takes several hours for NRCan to accumulate data from the 20 stations. Therefore, GPS•C

phase solution products based on global network have a latency of about 8 hours currently.

NRCan plans to real-time these phase solution products in the future. NRCan GPS•C phase

Page 141: 7 Tesis PPP Kongzhe Chen

117

solution products keep the same format and the same reference frame as the real-time code

solution products. In Chapter 6, some results obtained using NRCan GPS•C phase solution

orbit and clock products will be presented.

Before the development of GPS•C phase solution products, GPS•C products only consisted

of the real-time GPS•C code solution products. In the remainder of this thesis, the term

GPS•C products are meant to refer to the real-time GPS•C code solution products only

unless otherwise specified.

4.3 JPL Real-Time Precise GPS Products

JPL, another Analysis Center of IGS, has generated two types of real-time or near real-time

products based on the NASA Global GPS Network (GGN), which is operated and maintained

by JPL. Table 4.3 shows details of the products. The accuracy statistics of the products are

provided by Heflin (2004) and Muellerschoen (2003), respectively.

Table 4.3 JPL Real-Time Precise GPS Products (after Muellerschoen, 2003; Heflin, 2004)

Satellite Orbit / Clock Products Accuracy Latency Updates Sample Interval

Orbits ~22 cm Near Real-Time (NRT)

Clocks ~0.7 ns ~3

minutes 15

minutes 5 minutes

Orbits ~18 cm ~28 s Internet-based Global Differential GPS System (IGDG) Clocks ~1 ns

~4 s 1 s 1 s

Page 142: 7 Tesis PPP Kongzhe Chen

118

4.3.1 JPL Near Real-Time Orbit and Clock Products

JPL near real-time orbit and clock products can be downloaded from JPL website (Heflin,

2000), and denoted as JPL NRT products in this thesis. When it was introduced, the latency

of JPL NRT is 15 minutes, which has now been shortened to about 3 minutes (Heflin, 2004).

Therefore, JPL NRT should also be considered a type of real-time products according to the

definitions given by Bar-Sever and Dow (2002). JPL NRT orbits are provided in the format

of SP3, while clocks are formatted in RINEX. The reference frame of the products is ITRF

2000 (Heflin, 2004).

JPL NRT products are very useful for GPS meteorological applications, in which IGU

(predicted half) products are widely used currently. As shown in Tables 4.1 and 4.3, JPL

NRT orbits are slightly worse than IGU (predicted half) orbits, but JPL NRT clocks are much

more accurate than those of IGS Ultra-Rapid (predicted half) products. JPL NRT orbits and

clocks are accurate enough for zenith tropospheric delay estimation using PPP methodology.

To access JPL NRT products, users do not need to maintain continuous communication with

the server. Just one access per 15 minutes is enough, which is more robust than using JPL

IGDG products, which needs continuous communication with the server or geostationary

satellite. In Chapter 6, zenith tropospheric delay estimates using JPL NRT products will be

presented.

Page 143: 7 Tesis PPP Kongzhe Chen

119

4.3.2 JPL IGDG Orbit and Clock Products

Other types of real-time products of JPL’s are the real-time Internet-based Global

Differential GPS products (Muellerschoen et al., 2000), denoted as JPL IGDG products.

Decimetre level kinematic positioning accuracy has already been obtained in some

experiments using the real-time products in a simulated real-time mode (Muellerschoen et al.,

2000; 2001). The reference frame of the products is ITRF 2000 (Muellerschoen, 2003).

The prototype of JPL real-time global differential GPS system has been investigated since

the middle of 1990’s (Yunck et al., 1996b). A software package, Real-Time Net Transfer

(RTNT), was then developed for returning data from the NASA Global GPS Network

(GGN), which is operated and maintained by JPL with some stations ready for Internet

connections. From these stations, 1 Hz data was returned over the Internet. In 2000, 15 of

these stations could return better than 95% of the data in less than 3 s (Muellerschoen et al.,

2000). The GPS orbit and clock products are then estimated using the Real-Time GIPSY

(RTG) software. Positioning accuracy of ~10 cm RMS in the horizontal and < 20 cm RMS in

the vertical has been obtained by Muellerschoen et al. (2000) using the orbit and clock

products in a simulated real-time mode. In 2001, the number of real-time stations was

increased to 25 (Muellerschoen et al., 2001). As shown in Figure 4.8, the number of real-time

stations was increased to more than 60 in 2003 (Armatys et al., 2003). Consequently, the

accuracy and reliability of the orbit and clock products were also increased through

redundancy. Sub-decimetre in horizontal and decimetre in vertical kinematic positioning

accuracy was obtained in a simulated real-time processing (Armatys et al., 2003). But

Page 144: 7 Tesis PPP Kongzhe Chen

120

expanding stations will also add to the complexity of processing which as a result, would

delay the orbit and clock products.

Figure 4.8 IGDG Real-Time Network in 2003 (from Armatys et al., 2003)

In 2001, most of the data from the GGN real-time network arrived at the central processing

center within 2 s of its GPS time-tag. Data with latencies less than 2 s was used in computing

the GPS satellite clocks. Older data, up to 6 s in latency, was used to compute the GPS

satellite orbits since they are more predictable than clocks (Muellerschoen et al., 2001). In

2001, with only 25 real-time stations, the latency of the correction message with an Internet

connection was around 2.5 s, with about 2 s spent to accumulate the global data over the

Internet at the central processing center, and 0.5 s to process the clock solutions. Currently,

with more than 60 stations, the latency with an Internet connection has increased to about 4 s.

Page 145: 7 Tesis PPP Kongzhe Chen

121

JPL IGDG orbit and clock products are also broadcast by the geostationary Inmarsat

satellites (Armatys et al., 2003). The latency through Inmarsat is around 1.5 s longer than the

Internet latency. It takes about 1.5 s to uplink the global correction messages to the

geostationary satellite and downlink to the user (Muellerschoen et al., 2001).

4.4 Accuracy Statistics of Real-Time Products

There are no published data about the accuracies of GPS•C code or phase solution products,

JPL NRT and IGDG products. The data in the Tables 4.2 and 4.3, which was provided by

Heroux (2003), Collins (2004), Heflin (2004) and Muellerschoen (2003), was obtained by

comparisons with other more precise products such as the IGS Rapid products

(Muellerschoen, 2003). The orbit and clock comparisons were performed separately.

Sometimes, it is difficult to separate the errors in orbits and clocks, as they can be correlated

with each other, especially in the case of real-time. As investigated by Parkinson (1996), the

orbit error and clock error of each satellite were shown to be negatively correlated, which

means that they would cancel each other partially. In the real-time case, because orbits

update less frequently than clocks, errors in the orbits, especially errors in the radial

component, are “filtered out” by the clock solution (Muellerschoen, 2003). Different

components of orbit errors affect the range measurements in different ways. Signal-In-Space

Range Error (SISRE) is usually used to calculate the effects of GPS orbit and clock errors to

the range measurements. SISRE can be computed by the following equation (Malys et al.,

1997):

Page 146: 7 Tesis PPP Kongzhe Chen

122

)CA(CLK)-(RSISRE 2 22

491 ++= (4.1)

where R is the radial orbit error, A is the along-track orbit error, C is the cross-track orbit

error and CLK is the clock error.

According to equation (4.1), the accuracy statistics of the orbit and clock products discussed

above does not necessarily present their effects on measurements, especially in the real-time

case when radial orbit errors are almost cancelled out by the fast updated clock corrections.

The good performance of these real-time products, which will be shown in Chapter 6, has

proven this point.

4.5 Latency and Age

In real-time applications, two key factors affect the validity of real-time orbit and clock

corrections. They are the latency and age of the corrections.

Latency is the time taken to collect raw data from reference stations and estimate the

correction parameters plus the time spent for the correction data to reach the users (Kee,

1996). Age is the total delay of the correction data including the above correction data

latency and the interval from the time when the correction data is received to the time when

the correction data is actually applied by the users. Since users have to apply one set of

corrections till a new set arrives, corrections could maintain a constant latency for a period of

Page 147: 7 Tesis PPP Kongzhe Chen

123

time while the age is actually increasing. Figure 4.9 shows the relationship between the

latency and the age.

Figure 4.9 Latency and Age (after Kee, 1996)

Theoretically, the best results would be obtained if the orbit and clock corrections were

applied to the measurements whose GPS time-tag coincides with that of the measurements

used to estimate the corrections. However, in real-time application, non-zero latency and age

are unavoidable because of the time taken to estimate the correction parameters and for

transmission. People usually try to shorten the latency and age of the corrections to minimize

accuracy degradation. After SA was turned off in 2000, the age of the corrections is not as

vital as before regarding its effect on the achievable positioning accuracy. Before SA was

turned off, satellite clock errors changed rapidly and randomly. Normally clock corrections

aged up to one minute may not improve the accuracy. After SA was turned off, errors in the

satellite clocks do not change as quickly as the random shifts generated by SA. Clock

corrections can still be helpful in improving the solution even when aged up to one minute.

But minimizing the age values is still of importance as a method to further improve the

positioning accuracy. Satellite orbit error is more predictable than that of satellite clock. By

Page 148: 7 Tesis PPP Kongzhe Chen

124

using correction-rates along with orbit corrections, the orbit corrections can remain valid for

several minutes.

The latency and age of NRCan GPS•C corrections and JPL IGDG corrections will be

presented in the following as received over the Internet. For satellite access, the latency and

age would be about 1.5 s longer because of a longer correction transmission time.

4.5.1 Latency and Age of GPS••••C Corrections

In this section, latency and age of GPS•C code solution orbit and clock corrections are

presented. In 2002, for test purpose, NRCan GPS•C products were sent to a selected port and

IP address of a computer set up in the University of Calgary over an open Internet connection

using UDP (User Datagram Protocol) multicast. The latency and age presented in the

following are obtained from the test conducted on July 8, 2002.

GPS•C orbit and clock products are encoded in a modified RTCA-159 format. When

broadcast over the Internet, the messages are encapsulated in the GPS•C IP Multicast format

for transmission (ICD-GPS•C, 2001). Orbit corrections and clock corrections are encoded

into different messages with different update intervals. Therefore, the latencies and ages of

orbit corrections and clock corrections are different.

Given in Figures 4.10 and 4.11 are the latencies of orbit corrections and clock corrections,

respectively. Though encoded into different message, the latencies for orbit corrections and

Page 149: 7 Tesis PPP Kongzhe Chen

125

clock corrections are similar, ranged between 5 and 6 s. Most of the latencies are

approximately 5.55 s.

Figure 4.10 Latencies of GPS••••C Orbit Corrections

Figure 4.11 Latencies of GPS••••C Clock Corrections

Page 150: 7 Tesis PPP Kongzhe Chen

126

Figures 4.12 and 4.13 show the ages of orbit corrections and clock corrections for satellite

PRN 21. During the period when the test was conducted, GPS•C clock corrections updated

every 2 s. The update interval for orbit corrections was not constant. Only an upper limit of

120 s is defined. The ages of orbit corrections range between 6 and 60 s. The spiked age of

about 150 s in Figure 4.12 is because one or two orbit messages were lost, or were not sent

out from the server. Because correction-rates were also broadcast, the corrections were still

valid even aged up to 150 s (Chen et al., 2002). Another spiked age of about 30 s occurred

because the orbit corrections were updated within 25 s at that time. As shown in Figure 4.13,

the ages of clock corrections range between 6 and 7 s. The spikes exist because clock

messages were lost or not sent out.

Figure 4.12 Ages of GPS••••C Orbit Corrections

Page 151: 7 Tesis PPP Kongzhe Chen

127

Figure 4.13 Ages of GPS••••C Clock Corrections

As shown in Figures 4.12 and 4.13, there may be some packets lost using UDP multicast

over open Internet. UDP is considered connectionless and it sends and receives datagrams on

a “best effort” basis. The spikes in Figure 4.12 and 4.13 may also be caused by messages

which were expected but were not sent out in time. As discussed below, no message loss was

found when testing with JPL corrections using UDP transport protocol. Section 4.6 will

discuss the correction distribution issues.

4.5.2 Latency and Age of IGDG Corrections

Since the middle of 2003, for research purpose, JPL IGDG orbit and clock products have

been sent to a selected port number and IP address of a computer set up in the University of

Page 152: 7 Tesis PPP Kongzhe Chen

128

Calgary over an open Internet connection. The UDP transport protocol is also used to

distribute the corrections. The latency and age presented in the following were obtained from

the test conducted on June 28, 2004.

IGDG orbit and clock corrections are encoded in a JPL proprietary message format. Orbit

corrections and clock corrections are encoded into the same message, which updates every

second. Each 44 bytes message contains updated centimetre-level clock corrections (fast

clock corrections) for all satellites, orbit corrections (including correction-rates) and metre-

level clock corrections (slow clock corrections) for one satellite (Armatys, 2002). Thus,

completely sending the sequence of all satellites takes about 28 seconds, depending on the

number of available satellites. Therefore, the update interval is 1 s for clock corrections and

about 28 s for orbit corrections. The latencies of orbit corrections and clock corrections are

the same, but the ages are different for orbit corrections and clock corrections.

Figure 4.14 shows the latencies of IGDG corrections. The latencies of IGDG corrections are

very consistent. All of the latencies range between 3.5 and 4.5 s.

Figures 4.15 and 4.16 show the ages of orbit corrections and clock corrections for satellite

PRN 1. For a latencies of 3.5 to 4.5 s, the ages of orbit corrections range between 4 and 32 s

with an update interval of about 28 s. As shown in Figure 4.16, the ages of clock corrections

range between 4 and 5 s. The ages of both orbit and clock corrections were obtained by

testing with a GPS data sample interval of 1 s. No clocks aged over 5 s or orbits aged over 32

s are found in the plots, this is because there was no message lost during the test.

Page 153: 7 Tesis PPP Kongzhe Chen

129

Figure 4.14 Latencies of IGDG Corrections

Figure 4.15 Ages of IGDG Orbits

Page 154: 7 Tesis PPP Kongzhe Chen

130

Figure 4.16 Ages of IGDG Clocks

Figure 4.17 shows the difference between applying an orbit correction aged 28 s along with

orbit correction-rate and a zero-latency orbit correction. From the plot, we can see, most of

the time, the difference is less than 5 mm in each coordinate component. The RMS

differences during the test are 3.897 mm in x component, 3.678 mm in y component and

3.678 mm in z component. Using orbit corrections aged 28 s add less than 1 cm range error

to the user’s measurements. Therefore, with orbit correction-rates also broadcast, an update

interval of 28 s has little impact to the validity of the orbit corrections.

Page 155: 7 Tesis PPP Kongzhe Chen

131

Figure 4.17 Difference between Orbits Aged 28 s and 0 s

Figure 4.18 shows the difference between applying a clock correction aged 5 s and a zero-

latency clock correction. From the plot, we can see, most of the time, that the difference is

less than 10 mm. The RMS difference during the test is 5.482 mm. Therefore, a latency of 5 s

in clock corrections adds less than 1 cm range error to the user’s measurement, which is not

significant considering the sub-nanosecond accuracy of the corrections. The differences are

mainly due to the resolution of the clock correction data, which is 1/128 m or about 7.8125

mm. From the plot, we can see that all differences are multiples of 7.8125 mm. As the

accuracy of IGDG clock corrections is sub-nanosecond level, a resolution of 1/128 m is good

enough. However, if the accuracy of these clock corrections can be improved to the accuracy

of the IGS Final clock products (0.1 ns), the resolution should be also improved as well.

Currently, the effective round-off error of the IGDG clock corrections is 1/2 of the resolution,

or about +/- 3.9 mm.

Page 156: 7 Tesis PPP Kongzhe Chen

132

Figure 4.18 Difference between Clocks Aged 5 s and 0 s

4.6 Real-Time Product Distribution Issues

Several factors affect the correction latency and age, including the time taken to accumulate

GPS measurements from the tracking network over the Internet or satellites, the time taken to

estimate the corrections, and the time taken to distribute the corrections.

The time taken to accumulate GPS measurements varies significantly depending on the

coverage of the network and the communication techniques. As for GPS•C code solution

corrections, only data from 12 stations in Canada is used. The maximum communication

delays are less than 0.5 for land communication link and about 1.5 s for satellite

communication link (Lahaye et al., 1998). Therefore, it takes about 1.5 s for GPS•C to

accumulate GPS measurements. But it takes a little bit longer time for IGDG to accumulate

measurements, which acquires data from over 60 stations in a global network. Plots of certain

stations regarding data latency and number of satellites observed in the last hour can be

Page 157: 7 Tesis PPP Kongzhe Chen

133

found at JPL IGDG website (2004). Because of redundancy in JPL real-time tracking

network, processing can be started before data from all stations is returned. For example,

early-arrived data can be used in computing the GPS clocks. Older data can be used to

compute the GPS orbits since they are more predictable than clocks due to their orbital

dynamics (Muellerschoen, 2001). Some communication techniques can be used to further

shorten the accumulation time. For example, NRCan adopted Virtual Private Network (VPN)

to transmit data between some RTACPs and RTMACS (ICD-GPS•C, 2001). VPN is a

mature technology that can be introduced to shorten transmission time between tracking

stations and processing center. VPN offers a cost-effective, scalable, and manageable way to

create a private network over a public infrastructure such as the Internet or over a service

provider's Frame Relay, ATM (Asynchronous Transfer Mode), wireless network, or even

satellite connections. VPN creates tunnels across an IP network and provides users the same

policies as a private network, including security, quality of service (QoS), manageability, and

reliability.

The time taken to estimate the corrections varies depending on the number of measurements

used. For example, in 2000, it takes 0.5 s to process the clock solution using measurements

from 25 real-time stations (Muellerschoen et al., 2001). It was extended to about 1 s for over

60 stations (Muellerschoen, 2003).

The time taken to distribute the corrections varies depending on the techniques used. Usually,

the corrections can be broadcast by geostationary satellites or distributed over the Internet.

Some WADGPS systems, such as OmniSTAR and NavCom, have adopted the former, while

Page 158: 7 Tesis PPP Kongzhe Chen

134

JPL and NRCan plan to adopt both of them to distribute the IGDG and GPS•C products. The

advantages and limitations of satellite- and Internet-based transmission will be discussed in

the following.

4.6.1 Broadcast by Satellite

Some WADGPS systems broadcast their precise data over geostationary satellites. The main

advantage of satellite-based broadcasting is that the receiving device can be integrated within

a GPS receiver because the corrections can be broadcasted using the same band (L-Band) as

GPS signals. At the same time, the bandwidth is not so serious a problem in satellite-based

broadcasting when compared with Internet-based distributing.

Still, compared with the Internet, satellite-based broadcasting also has several disadvantages.

Firstly, special receivers are required. It takes time and effort to develop a special receiver to

receive and decode corrections broadcasted by satellites. Internet, on the other hand, is a

public infrastructure, for which users have many accessible devices. This is especially

relevant for meteorological and timing applications, in which processing is conducted indoor

where Internet connections are normally available. Secondly, the cost of satellite-based

broadcasting is much more expensive than that of Internet-based distributing. Thirdly, it

takes a longer time to uplink the correction messages to satellite and downlink to the users.

For example, as tested by JPL, it takes 1.5 s to uplink the global correction messages to the

geostationary satellite and broadcast to the users (Muellerschoen et al., 2001). Similar

Page 159: 7 Tesis PPP Kongzhe Chen

135

transmission time has also been demonstrated by Lahaye et al. (1998). A much shorter

transmission time is required for Internet distributing. Figure 4.19 shows the transmission

time over the Internet, which was obtained by sending 106-byte length packets from a

computer at the University of Calgary to a computer at the University of York and back to

the sender. In this one-day’s test, the packets were sent out at a rate of 1 Hz using UDP

transport protocol. The transmission time was calculated by subtracting the time of sending

from the time of returning. From the figure, we can see, most of time, the packet was

returned within 60 ms after over 7000km traveling. Less than 1% packets were returned over

100 ms. Fourthly, in satellite-based broadcasting, users can rarely reply to the server. In this

case, the two-way communication and extensible data transmission rate of Internet-based

distributing are attractive to some special users. Finally, the satellite-based broadcasting

approach cannot provide global coverage. Geostationary satellites can only provide

corrections to users at latitudes from 76°N to 76°S because of the limitation of their orbits

(Armatys et al., 2003). For example, currently, real-time GPS•C corrections are broadcast via

the MSAT geostationary satellites (Kassam, 2003). But they cannot provide stable and

reliable communication links to users in all parts of Canada. Since geostationary satellites

such as MSATs are seen at low elevation in high latitudes, they are vulnerable to signal

blockage.

Considering the above, the satellite-based broadcasting and Internet-based distributing

methods are complementary to each other. Users will benefit from the coverage and

reliability if the orbit and clock products are broadcast through both approaches. Therefore,

Page 160: 7 Tesis PPP Kongzhe Chen

136

some organizations plan to adopt both of them to distribute their corrections, such as JPL and

NRCan. As for users, they can choose either approach based on the reliability, timeliness,

convenience, and available devices.

Figure 4.19 Transmission Time over the Internet

4.6.2 UDP/IP Multicast

In this research, JPL real-time orbit and clock products are received over the Internet. Several

techniques can be used for distributing products efficiently over the Internet.

Page 161: 7 Tesis PPP Kongzhe Chen

137

Multicast

Unlike satellite-based broadcasting, Internet-based distributing will encounter the problem of

bandwidth exhaustion. There are two distinct methods with which packets can be sent over

the Internet: unicast and multicast. Unicast is point-to-point communication whereas

multicast is point-to-multipoint or multipoint-to-multipoint communication. Unicasting

creates a stream for each receiver. This can be bandwidth intensive when multiple streams

are requested and cause quite a burden on the internal network. In the multicast case, the

server puts together a packet with a group address and gives it to the network. The routers

recognize the addresses included in the group address and subsequently duplicate the packet

for each receiver. Multicast technology eases the burden on the broadcast server as well as

the internal network. Multicast technique is also suitable for returning GPS data from real-

time tracking network to processing centers, because real-time stations can return real-time

data stream to multiple processing centers. For example, currently, JPL shares real-time data

stream from some stations with NRCan.

UDP Transport Protocol

There are two types of transport-layer protocols widely used in the Internet: Transmission

Control Protocol (TCP) and User Datagram Protocol (UDP) (Fapojuwo, 2003). TCP is a

connection-oriented protocol, meaning that TCP will set up, maintain, and tear down a

connection. The sender will wait for a reply from the receiver, and resend the packets if the

receiver has not received the packet correctly. TCP is designed for reliable transmission, but

Page 162: 7 Tesis PPP Kongzhe Chen

138

not for timely transmission. On the other hand, UDP is based on connectionless technology,

which means that the datagram is sent without first setting up a connection. The UDP is

designed to transfer packets in a timely manner, and does not re-send the lost packets. UDP is

also a much simpler protocol than TCP and has fewer overheads. Another advantage of UDP

is that it is much easier to implement multicast using UDP than TCP. Because UDP is a one-

to many and many-to-one protocol, while TCP is a one-to-one protocol. The drawback of

UDP is that it does not guarantee that the packet will arrive at its final destination.

When timeliness is a more important factor than reliability, UDP is usually used. Otherwise,

TCP is used. Therefore, UDP is more suitable for real-time services than TCP, such as real-

time distribution of GPS observations and products. Real-time orbit and clock products

update at a high rate, 1~2 s for clock corrections, 20~50 s for orbit corrections. Low packet

loss rate is compensated by the high update rate. Currently, GPS data stream and GPS•C

corrections are transmitted over NRCan's network using UDP multicast (ICD-GPS•C, 2001).

UDP is also adopted by JPL and GFZ for data transmission. For example, IGDG products are

distributed using UDP transport protocol to users. UDP transport protocol is also

recommended by IGS Real-Time Working Group for real-time distribution of data and

products (Muellerschoen and Caissy, 2004).

When directly receiving orbit and clock products from servers using UDP transport protocol,

such as the servers of GPS•C and IGDG, users do not know whether the missing packets are

unsent or lost. Therefore, here, the test done to show the transmission time in Section 4.6.1 is

used to show the packet loss rate of distributing products using UDP transport protocol. As

Page 163: 7 Tesis PPP Kongzhe Chen

139

shown in Figure 4.20, the packet loss rate is zero most of time. The packet loss rate was

calculated on a one-minute basis. Only in two minutes did the packet loss rate reach 5%,

which means 3 out of 60 packets were lost. Only 69 packets were lost in the entire test and

the total packet loss rate is 69/86400=0.08%. Less than one packet was lost out of 1000

packets on average.

Figure 4.20 Packet Loss Rate Using UDP Transport Protocol

4.7 Real-Time Correction Formats

The formats of the real-time orbit and clock products will determine the update interval and

bandwidth required to distribute the products. Several industry standard formats have been

Page 164: 7 Tesis PPP Kongzhe Chen

140

implemented for differential services, such as RTCM-104 and RTCA-159. They are widely

used in real-time kinematic or wide area differential GPS systems. However, they are not

suitable to encode the high precision orbit and clock products for global coverage. Two

formats have been proposed by NRCan and JPL, which are currently used to carry their real-

time orbit and clock products. These formats will be discussed in the following.

4.7.1 RTCM-104

RTCM-104 was first recommended by the Radio Technical Commission for Maritime

Services Special Committee 104 as standards for DGPS correction messages in November

1983. Types of messages have been added ever since then to the subsequent versions

(Muellerschoen and Caissy, 2004).

RTCM-104 has been widely accepted as a standard for local area DGPS services. Many GPS

receivers are now integrated with RTCM capability. But the scalar corrections in the RTCM

message are only applicable to measurements whose errors are correlated with those in the

measurements used to calculate the corrections. The spatial decorrelation of these scalar

corrections in this format makes them unsuitable for carrying the globally effective

corrections.

Page 165: 7 Tesis PPP Kongzhe Chen

141

4.7.2 RTCA-159

RTCA-159 is a standard format developed by the Radio Technical Commission for Aviation

Special Committee 159 for aircraft navigation using GPS. It was first proposed as the

standard to serve the Wide Area Augmentation System (WAAS) (Muellerschoen and Caissy,

2004). Subsequently, it was also adopted by some other wide area differential systems. In

this standard, corrections are encoded in a vector format including separate corrections for

satellite clocks, orbits and ionospheric effects. In this point, it can be chosen to carry real-

time corrections which are globally effective. RTCA-159 has also been widely used in GPS

community as a standard and is integrated into many GPS receivers. But the resolutions of

this format are limited. For example, the resolution for clock corrections is 0.125 m, and the

same resolution is applied to orbit corrections in each positioning component (ICD-GPS•C,

2001). Considering that the corrections of this format are originally proposed for code

processing, the resolutions were adequate. On the other hand, some corrections discussed in

this chapter are obtained by processing using both code and phase measurements, and the

accuracy of the corrections has been improved to decimetre level. A modified RTCA-159

format has been proposed by NRCan for the high precision corrections.

4.7.3 GPS••••C Format

GPS•C uses a subset of the RTCA-159 message types and has implemented several non-

standard RTCA-159 messages. Therefore, GPS•C products are encoded in a modified

Page 166: 7 Tesis PPP Kongzhe Chen

142

RTCA-159 (MRTCA) format. For example, messages of types 32, 33, 34, 35 and 45, which

were reconstructed from the standard RTCA-159 message 2 and 25, respectively, carry fast

and slow corrections with a higher resolution. In these types of messages, the resolution of

the corrections was increased to 4 mm for orbits and clocks. The bandwidth required to

distribute GPS•C messages is 106 byte/s for update intervals of 2 s and 20 s for clocks and

orbits respectively (ICD-GPS•C, 2001).

4.7.4 IGDG Format

GPS•C adopted a modified RTCA-159 format, which keeps some aspects of the original

RTCA-159. For example, the active PRN list (PRN mask) is encoded in a separate message

(Message Type 1). The full time of week and GPS week number are put in another separate

message (Message Type 12). The preambles of the messages carrying orbit and clock

corrections only provide the GPS time modulo 256 s (ICD-GPS•C, 2001). Message Type 1

and 12 are not sent out as frequently as the orbit and clock correction messages. If one of

them is lost, which is likely to happen when distributed using UDP, the integrity of the

correction data will be broken and a long sequence of corrections will be useless. The

separate message format also adds difficulties for decoding. For example, to decode the clock

corrections of a single satellite, at least information in four messages should be used,

including messages for PRN mask, full time tag, slow clock corrections, and the fast clock

corrections.

Page 167: 7 Tesis PPP Kongzhe Chen

143

A simplified format was proposed by JPL and used to encode the IGDG orbit and clock

products. The format has been described in details by Armatys (2002). The IGDG 44-byte

messages are generated at 1 Hz. There are no separate PRN mask and time tag messages. The

44-byte message contains orbit corrections and metre-level clock corrections for one satellite,

and centimetre-level clock corrections for all satellites. Therefore, it is very convenient to

decode the corrections, and one message loss only affects the availability of one satellite. The

bandwidth required to distribute IGDG messages is only 44 byte/s with update interval of 1 s

for clock corrections. The resolutions of IGDG clock corrections is 1/128 m=7.8125 mm,

which may be not high enough for corrections with improved accuracy, as discussed in

section 4.5.2.

4.7.5 Application of Real-Time Orbit and Clock Products

Currently, many wireless technologies are Internet accessible, such as GSM (Global System

for Mobile Communication) and CDPD (Cellular Digital Packet Data). Furthermore, the

third generation wireless network’s devices can provide millions of bits per second in terms

of bandwidth (Fapojuwo, 2003). Therefore, users can adapt wireless devices, which are very

important in kinematic application, to receive orbit and clock products distributed over the

Internet. For example, even CDPD, an older first-generation technology with a bandwidth of

14.4 kbps (Fapojuwo, 2003), is enough to receive GPS•C and IGDG orbit and clock

products.

Page 168: 7 Tesis PPP Kongzhe Chen

144

The real-time orbit and clock corrections are based on the broadcast ephemeris, which

normally update every two hours. If corrections based on one set of broadcast ephemeris are

applied to another set of broadcast ephemeris, the positioning accuracy will be degraded. The

Issue Of Data Ephemeris (IODE), which exists in broadcast ephemeris with a numerical

range of 0 to 255, is also introduced into the real-time corrections. The IODE items in the

corrections indicate which set of broadcast ephemeris has been used in generating the

corrections. Normally, corrections based on old ephemeris will continue for several minutes

for user accumulating new ephemeris between the ephemeris updates.

Page 169: 7 Tesis PPP Kongzhe Chen

145

CHAPTER 5

PRECISE POINT POSITIONING USING

SINGLE-FREQUENCY GPS DATA

This chapter investigates precise point positioning method using only single-frequency GPS

data. A new ionospheric estimation model has been proposed which can estimate vertical

TEC accurate to 1~2 TECU and has shown great promise for real-time single-frequency

point positioning. The chapter begins with a brief description of single-frequency point

positioning methods, followed by studying several models used for ionospheric effects

mitigation. The new ionospheric estimation model is then proposed which estimates

horizontal gradients of ionospheric delay using a single GPS receiver. The chapter concludes

with some numerical results to investigate the accuracy of zenith ionospheric delay

estimation and positioning using the new model. Some positioning results using other

ionospheric mitigation models are also presented as a comparison.

5.1 Single-Frequency Point Positioning

Most GPS receivers in use today are single-frequency receivers. Point positioning using

single-frequency observations and precise GPS orbit and clock products has been

Page 170: 7 Tesis PPP Kongzhe Chen

146

investigated by researchers for a long time (Heroux and Kouba, 1995). However, the precise

GPS orbit and clock products, which have been proven to be significant for point positioning

using dual-frequency measurements, can only offer limited improvement for point

positioning using single-frequency data.

Compared with point positioning using broadcast orbits and clocks, the accuracy of dual-

frequency point positioning has improved from tens of metres with SA on or a couple of

metres with SA off to sub-decimetre by applying precise GPS orbit and clock products. On

the other hand, precise GPS orbits and clocks can just improve the accuracy of single-

frequency point positioning to a couple of metres from about 10 m when broadcast orbits and

clocks are used after SA was turned off (Kouba, 2003). Ionospheric effects are the main

factors for single-frequency users to further improve positioning accuracy using un-

differenced data. Lachapelle et al. (1994b; 1996) presented kinematic point positioning

accuracy of a couple of metres using shipborne and airborne datasets with precise orbit and

clock products, and found ionospheric effects were the dominant error sources. Therefore,

the key issue for single-frequency point positioning is how to mitigate ionospheric effects.

In the past decade, ionospheric products, new mapping functions and analysis models have

been developed to improve the accuracy of single-frequency point positioning. IGS has been

providing the total electron content of ionosphere on a global scale since 1998 (Feltens and

Schaer, 1998). Using the products, Ovstedal (2002) has demonstrated an accuracy of 1~2

metres with IGS precise orbits and clocks. But only code measurements have been used. In

this research, improved accuracy will be presented by using both code and phase

Page 171: 7 Tesis PPP Kongzhe Chen

147

observations. Ionosphere-free combination can also be formed on single-frequency GPS

observations, known as GRAPHIC (Group And Phase Ionospheric Correction) (Yunck,

1996a). Montenbruck (2003) has obtained 1.5 m 3D positioning accuracy in LEO (Low Earth

Orbit) satellite orbit determination using this model. In some research, ionospheric delay

parameters have also been estimated as unknowns along with coordinates. Beran et al. (2003)

have presented a couple of metres accuracy using single-frequency observations from a static

geodetic receiver with one or two (a bias and a drift) zenith ionospheric parameters

estimated.

In these researches, because the ionospheric effects, which became the dominant error

sources after precise orbit and clock products were used, could not be mitigated effectively

using un-differenced single-frequency data to fully exploit the precise carrier phase

observables, only accuracies of one metre to a couple of metres were obtained. Therefore, the

PPP methodology, which is promised to provide centimetre level accuracy, is only applicable

to dual-frequency GPS users at present.

Consequently, single-frequency point positioning has not been considered to be able to

provide high precision positioning results at accuracy such as decimetre level. To achieve

decimetre or even centimetre level accuracy using single-frequency observations, differential

methods were widely used, such as Differential GPS (DGPS) and Real-Time Kinematic

(RTK) systems. Differential methods, however, need the supports of base stations with

precisely known coordinates, which not only increase the operational cost of equipment and

labour but it is also not easy to establish in remote survey areas.

Page 172: 7 Tesis PPP Kongzhe Chen

148

In this research, precise point positioning method using single-frequency GPS data will be

investigated. It aims to derive decimetre to sub-metre accurate position solutions by

processing un-differenced single-frequency code and carrier phase observations with precise

GPS orbit and clock products. To mitigate ionospheric effects, a new ionospheric estimation

model has been proposed and will be described. The new model, which estimates horizontal

gradients along with the zenith delay of ionosphere, has demonstrated its potential to provide

position solutions at decimetre level accuracy using single-frequency data. On the other hand,

several models, such as the Klobuchar model and the Global Ionospheric Model (GIM), will

also be examined by processing both code and carrier phase measurements to investigate the

obtainable accuracy of these models.

5.2 Ionospheric Models

In single-frequency data processing, the observation equation models can be described as

follows:

)()(

)(

112/1

1

PdmDCBTc

dddddtdtcP

PPPgd

reliontroporbs

rsr

ερ

++−+

++++−+= (5.1)

)C(dm)DCBDCBT(c

dddd)dtdt(cC

CC/PP/Pgd

reliontroporbs

rsr

111121

1

ερ

++−−+

++++−+= (5.2)

Page 173: 7 Tesis PPP Kongzhe Chen

149

)(w m)N)t()t((

dddd)dtdt(csr

reliontroporbs

rsr

111101011

1

Φ++++−+

+−++−+=Φ

εδδφφλ

ρ (5.3)

where,

1P is the P-Code pseudorange measurement on 1L (m);

1C is the C/A-Code pseudorange measurement (m);

1Φ is the carrier phase measurement on 1L (m);

srρ is the true geometric range (m);

c is the speed of light (m/s);

sdt is the satellite clock error (s);

rdt is the receiver clock error (s);

orbd is the satellite orbit error (m);

tropd is the tropospheric delay (m);

iond is the ionospheric delay on 1L (m);

gdT is the group delay differential of satellite (s);

21 P/PDCB is the differential code bias between 1P and 2P of receiver (s);

11 C/PDCB is the differential code bias between 1P and 1C (s);

reld is the relativistic effects (m);

1wδ is the phase windup on 1L (m);

1λ is the wavelength on 1L (m/cycle);

Page 174: 7 Tesis PPP Kongzhe Chen

150

1N is the integer phase ambiguity on 1L (cycle);

)t(r01φ is the initial phase offset of the receiver on 1L (cycle);

)t(s01φ is the initial phase offset of the satellite on 1L (cycle);

1Pdm is the P-code multipath on 1L (m);

1Cdm is the C/A code multipath (m);

1mδ is the carrier phase multipath on 1L (m).

In both GPS receivers and satellites, GPS signals are delayed by hardware. The delays in 1P

and 2P pseudorange measurements are different. For satellites the differences is normally

called group delay differential gdT . For receivers, the term Differential Code Bias ( 21 P/PDCB )

is often used. With double differencing between satellites and receivers, these biases are

removed completely. As discussed in Chapter 3, users do not need to consider these terms

when applying the traditional dual-frequency ionosphere-free combinations, regardless if

they use broadcast ephemeris or IGS precise orbits and clocks. However, for single-

frequency point positioning, they have to be taken into account (Kouba, 2003). Users can

obtain the gdT information from the broadcast navigation messages (Wilson et al., 1999),

while the 21 P/PDCB can be absorbed into receiver clock offset estimation in single-frequency

point positioning. There are also biases between 1P and 1C measurements, namely 11 C/PDCB ,

which are satellites related (Jefferson et al. 2001). As the clock information in broadcast

ephemeris or in IGS precise clock products is fully consistent with P1/P2 code

measurements, for users who use 1C measurements, 11 C/PDCB should also be applied. The

Page 175: 7 Tesis PPP Kongzhe Chen

151

values of 11 C/PDCB for all GPS satellites can be downloaded from CODE DCB website

(2004).

Satellite orbit error orbd and clock error sdt can be reduced to centimetre level by precise

orbits and clocks. Zenith tropospheric delay can be modeled to decimetre or even centimetre

level accuracy using simple pressure, humidity and temperature models or meteorological

measurements. It can also be estimated, but the estimation of additional parameters may

degrade the solution if errors in observables have not been mitigated effectively. Relativistic

effects reld and phase windup 1wδ can be modeled to centimetre level accuracy. The initial

phase offsets of both satellite and receiver are absorbed by the float ambiguity estimation.

The ionospheric delay iond , which can be almost removed using dual-frequency

observations, can be mitigated by models discussed in the following when only single-

frequency observations are available. As discussed in Chapter 3, antenna phase center offset

and variations and site displacement effects such as solid earth tides and ocean loading

should be taken into account in precise point positioning using single-frequency data.

Several methods have been developed to deal with ionospheric effects in single-frequency

point positioning, including:

I. Using broadcast Klobuchar model (Klobuchar, 1996);

II. Using Global Ionospheric Model (GIM) provided by IGS and other organizations

(Schaer et al., 1998);

Page 176: 7 Tesis PPP Kongzhe Chen

152

III. Using single-frequency ionosphere-free combination (Yunck, 1996a; Montenbruck,

2003);

IV. Estimating ionospheric delay parameters (Beran et al., 2003).

The first and likely the most popular method, is using the Klobuchar model with broadcast

ionospheric coefficients (Klobuchar, 1996). This method can be implemented in real-time or

post-mission using broadcast ephemeris or precise orbit and clock products. The drawback to

the Klobuchar model with broadcast ionospheric coefficients is that it can only mitigate

50~60% of total ionospheric errors (Klobuchar, 1996). Even using precise orbit and clock

products, the end results can only be accurate to a couple of metres (Ovstedal, 2002). Post-fit

ionospheric coefficients have been developed that can help improve the performance of the

Klobuchar model. Since the middle of July 2000, CODE has been providing Klobuchar-style

ionospheric coefficients on a regular basis that best fit its IONEX data. The post-fit

coefficients have demonstrated much better performance than the coefficients originally

broadcast by the GPS for the single-frequency user (Ovstedal, 2002). Currently, CODE post-

fit coefficients have several days’ latency, so they can only be used in post-mission. CODE is

also computing predicted Klobuchar-style coefficients, but the improvement is not as

significant as the post-fit ones.

The second method is to use the Global Ionospheric Model (GIM) provided by IGS and other

organizations (Schaer et al., 1998). Since 1998, IGS has provided ionospheric TEC grid

parameters with a latency of about 11 days. Currently, the IGS ionospheric products can

provide an accuracy of 2 TECU (1 TECU corresponds to 0.163 m range error in 1P ) at grid

Page 177: 7 Tesis PPP Kongzhe Chen

153

points (Ovstedal, 2002). But the accuracy does degrade for interpolated points as their

temporal resolution is 2 hours and spatial resolution is 5 degrees (longitude) × 2.5 degrees

(latitude) (IGS website, 2004). As investigated by Ovstedal (2002), this method could

provide better results than Klobuchar model using the same GPS dataset and ephemeris. But

as only code measurements were processed, the achievable accuracy is limited at a couple of

meters. The use of the model is also limited by the 11 days’ latency of the IGS ionospheric

products despite that the precise GPS orbits and clocks have been available with much

shorter latency or even in real-time as discussed in Chapter 4. In the following, this model

will also be used along with both code and phase measurements to further exploit the

potential of the IGS ionospheric products.

The third method is based on the use of a single-frequency ionosphere-free combination

(Montenbruck, 2003). The single-frequency ionosphere-free combination, which averages

the code and carrier-phase measurements on the same frequency, has been introduced as

GRAPHIC (Group And Phase Ionospheric Correction) and can be expressed as follows

(Yunck, 1996a):

222

222

1111101011

21111

)()P(mdm)N)t()t((

)DCBT(cwddd)dtdt(c

P

Psr

P/Pgdreltroporb

sr

Φ++

++

+−+

−+++++−+=

Φ+

εεδφφλ

δρ (5.4)

Though the first-order ionospheric error can be completely removed with the combination,

the phase ambiguity is introduced and the noise of this combination is dominated by the code

noise (about half of the magnitude of the code noise). It is also impossible to solve coordinate

Page 178: 7 Tesis PPP Kongzhe Chen

154

parameters and ambiguities using observations from only a single epoch. A global

adjustment using cumulative measurements has to be carried out, and a long convergence

time is required for the float ambiguities, i.e. 2~4 hours (Heroux et al., 2004). This method is

suitable for post-mission with long data tracking sessions. Montenbruck has demonstrated

1.5 m 3D positioning accuracy in determining orbits of LEO satellites. Because this research

is focused on real-time applications, this model will not be used.

The last method is to estimate zenith ionospheric delay using code and/or phase observations

(Beran et al., 2003). Mapping functions are used to map the zenith ionospheric delay to slant

delays. Using post-mission precise GPS orbits and clocks, they obtained an accuracy of a

couple of metres with metre level biases using single-frequency observations from a static

geodetic receiver with one or two (a bias and a drift) estimated ionospheric parameters. This

method is also applicable in real-time navigation using real-time precise GPS orbit and clock

products. However, zenith ionospheric delays for different satellites at ionospheric pierce

points vary significantly, it is not adequate to model ionospheric delays for all satellites using

only one zenith delay and mapping functions. The limitations of using one zenith delay and

mapping functions have been investigated in Klobuchar et al. (1993) who described that

applying mapping function to regions with large horizontal electron density gradients would

lead to errors of several TECU. Therefore, overall this model does not show any

improvement over the simple Klobuchar model since it could only achieve comparable

accuracy by using IGS precise orbit and clock products and both code and phase

Page 179: 7 Tesis PPP Kongzhe Chen

155

measurements (Beran et al., 2003). In the next section, estimating ionospheric horizontal

gradients using single-frequency observables from a single GPS receiver will be discussed.

Given in Table 5.1 are the vertical TEC (VTEC) of each satellite at the ionospheric pierce

point at a specific epoch obtained via interpolation from IGS final ionospheric products to the

S1 pillar on the roof of Engineering Building at the University of Calgary. IGS Final

ionospheric products were interpolated to the ionospheric pierce point of each satellite using

the recommended interpolation procedure (Schaer et al., 1998). The epoch is from December

3, 2003, which is an ionospheric quiet day with Ap index of 4. From the table, we can see

that even with a quiet ionospheric condition and at mid-latitude station, the differences in the

VTECs at the ionospheric pierce points of different satellites can be up to 11 TECU.

Table 5.1 VTEC of Each Satellite at the Ionospheric Pierce Point

PRN 1 2 3 8 10 13 27 28 29 30

Azimuth (deg) 153 61 60 286 290 166 34 229 318 111

Elevation (deg) 11 19 32 54 25 39 87 23 12 45

VTEC at ionospheric pierce point (TECU) 25.5 22.9 22.6 20.8 18.3 23.3 22.0 21.1 14.3 23.4

5.3 Precise Point Positioning with Ionospheric Delay Estimated

Beran et al. (2003) could only obtain positioning accuracy of a couple of metres by

estimating zenith ionospheric delay parameters along with coordinates. They have not

considered the horizontal gradients of electron density, which may account for the metre

Page 180: 7 Tesis PPP Kongzhe Chen

156

level biases. Giffard (1999; 2000) has also tried to estimate vertical TEC using observables

from a single GPS receiver. Without estimating the horizontal gradients, he only obtained an

accuracy of 3 to 10 TECU of vertical TEC estimates when compared with IGS products.

Horizontal electron density gradients have long been investigated by researchers. As

described in Leitinger (1993), the most typical gradients are:

the general equatorward increase of total electron content (TEC) in mid-latitudes

during daytime;

the west to east increase of TEC in the morning for all seasons;

and the east to west increase in the afternoon in winter.

Though some models, such as the Klobuchar model, have considered these gradients, the

actual gradients can differ considerably from the average values because of the day-to-day

variability of the ionization (Leitinger, 1993). Normally, people use TECU/km to denote

electron content changes versus distance or TECU/deg to denote electron content changes

versus latitude or longitude. Vo and Foster (2001) have shown TEC gradients are correlated

with the background TEC. High gradients values occurred in the sunlit sector with TEC

gradients up to 10 TECU/deg found in the post-noon ionosphere. Hernández-Pajares et al.

(1998) suggested a 2 TECU/deg gradient with low solar activity for tomographic modeling.

Horizontal electron density gradients have been described as a common phenomenon in

middle-latitude region (Gail et. al., 1993), but they have also been investigated in other

regions (Huang, 1997; Ohta and Hayakawa, 2000). Schaer et al. (1999b) and Bock et al.

Page 181: 7 Tesis PPP Kongzhe Chen

157

(2000) attempted to introduce ionospheric gradient parameters in GPS network processing,

and they found these parameters might absorb part of the satellite- and epoch-specific biases.

Dai et al. (2001) have also made similar attempts to estimate ionospheric gradient parameters

for ambiguity resolution in the hopes that the ionospheric gradient parameters could absorb a

significant amount of the spatially correlated ionospheric biases. No work has been reported

on estimating ionospheric gradient parameters using un-differenced single-frequency GPS

data.

5.3.1 Estimating Ionospheric Horizontal Gradients with Un-differenced GPS

Measurements

IGS Final ionospheric TEC grids, which are accurate up to 2 TECU or even better at grid

points (Ovstedal, 2002), can be used to demonstrate the gradients numerically. The IGS Final

ionospheric TEC grids for December 31st, 2003 were obtained for this purpose. The Ap index

of that day is 19, which means a typical ionospheric condition. A mid-latitude IGS station,

AMC2 (38.8° N, 104.5° W), was selected to test the ionospheric gradients. The ionospheric

TEC grids were interpolated to the ionospheric pierce points for satellites observed at AMC2

with different azimuth and elevation angles.

Figures 5.1 to 5.4 show VTEC at ionospheric pierce points for satellites observed by AMC2

at different azimuth and elevation angles. The VTEC shown in these figures is consistent

with the three types of gradients discussed above. In Figure 5.1, VTEC was interpolated to

Page 182: 7 Tesis PPP Kongzhe Chen

158

the ionospheric pierce points for satellites in the north (0° azimuth) and south (180° azimuth)

with elevation angles from 0° to 90° at 16:00 local time. It presents the equatorward increase

of TEC in the mid-latitudes during daytime. In Figure 5.2, VTEC was interpolated to the

ionospheric pierce points for satellites in the east (90° azimuth) and west (270° azimuth) with

elevation angles from 0° to 90° at 6:00 local time. It highlights the west to east increase of

TEC in the morning. In Figure 5.3, VTEC was interpolated to the ionospheric pierce points

for satellites in the east (90° azimuth) and west (270° azimuth) with elevation angles from 0°

to 90° at 18:00 local time. It illustrates the east to west increase of TEC in the afternoon. In

Figure 5.4, VTEC was interpolated to the ionospheric pierce points for satellites at an

elevation angle of 30° with azimuth angles from 0° to 360° at different local time. It shows

the general TEC changes against the azimuth angles at different local time periods.

If no horizontal gradients existed, the vertical TEC should be the same at the ionospheric

pierce points for satellites with different elevation angles and different azimuth angles. The

tangent-like curves in Figures 5.1, 5.2 and 5.3 and the cosine-like curves in Figure 5.4

indicate that ionospheric horizontal gradients are present at different times and directions.

Page 183: 7 Tesis PPP Kongzhe Chen

159

Figure 5.1 The Equatorward Increase of TEC at Local Time 16:00

Figure 5.2 The West to East Increase of TEC at Local Time 6:00

Page 184: 7 Tesis PPP Kongzhe Chen

160

Figure 5.3 The East to West Increase of TEC at Local Time 18:00

Figure 5.4 The VTEC at Ionospheric Pierce Point for Satellites at 30°°°° Elevation Angle against

Azimuth at Different Local Time

Page 185: 7 Tesis PPP Kongzhe Chen

161

The cosine-like behaviour of the VTEC gradients in Figure 5.4 and the tangent-like

behaviour of the VTEC gradients from Figures 5.1 to 5.3 suggest that mapping functions can

be derived by tilting the zenith.

Following Klobuchar (1996) and Davis et al. (1993), the ionospheric group delay in the

measurement to a satellite whose signal penetrates the ionosphere at altitude z and

horizontal position vector x

as measured from the site, can be expressed as:

( ) ( )

( ) ( )

( ) ( )α

α

,eded

dlxzNf

.dlzN

f.

dllNf

.,ed

g

g

ion

+=

⋅+=

=

0

202

2

340340

340

(5.5)

where e is the elevation angle, α is the azimuth angle, 0N is the electron density with

absence of gradients and gN

is the horizontal gradient of electron density at x

=0, and

( ) ( ) ( )eFddlzNf

.ed z 0020

340 == (5.6)

( ) ( ) ⋅= dlxzNf

.,ed gg

2

340α (5.7)

( )ed 0 , zd and ( )eF0 are the slant group delay, zenith delay and mapping function with the

absence of horizontal gradients of electron density, when the group delay is independent of

azimuth. ( )α,ed g is the delay caused by the horizontal gradients of electron density.

Page 186: 7 Tesis PPP Kongzhe Chen

162

( )α,ex

, ( )zN g

, and dl can be expressed as (Davis et al., 1993):

( ) ( )ααα sinecosnecotz,ex +≈ (5.8)

( ) ( ) ( )ezNnzNzN eng +=

(5.9)

( )dzeFdl 0≈ (5.10)

where n and e are unit vectors in north and east directions, nN and eN are the horizontal

gradients of electron density in north and east directions, respectively.

Therefore, the gradient delay can be calculated as:

( ) ( )

( ) ( )

( ) [ ]αα

αα

α

sinGcosGecoteF

dzzzNf

.sindzzzN

f.

cos

ecoteF,ed

en

en

g

+=

0

22

0

340340 (5.11)

where nG and eG are the horizontal gradients in north and east directions respectively, and

they can be expressed as:

( )

( )

=

=

dzzzNf

.G

dzzzNf

.G

ee

nn

2

2

340

340

(5.12)

Therefore, the total ionospheric group delay for satellite with elevation e and azimuth α is:

Page 187: 7 Tesis PPP Kongzhe Chen

163

( ) ( ) ( ) [ ]ααα sinGcosGecoteFeFd,ed enzion ++= 00 (5.13)

If a single layer model mapping function is used, ( )eF0 can be expressed as (Manucci et al.,

1993):

( )( )

20

1

1

+−

=

ecosHR

ReF

E

E

(5.14)

where ER is the mean earth radius (6371 km) and H is the ionospheric shell height, which

can be roughly selected between 300 km and 500 km (Schaer, 1999a). For example, the value

of 450 km has been used by IGS ionospheric products. Additional details about mapping

functions will be discussed in Section 5.3.2.

The new model, as shown in Equation 5.13, can be used to estimate zenith ionospheric delay

along with the horizontal gradients using single-frequency code and carrier phase

measurements. It is denoted as “ionospheric estimation model” in the thesis. In Section 5.3.3

and Section 6.5, the performance of this model will be compared with the Klobuchar model

and the Global Ionospheric Model.

5.3.2 Ionospheric Mapping Functions

Except for the ionosphere-free model, all the previously discussed models need to use

ionospheric mapping functions. Several mapping functions are used with slight differences at

Page 188: 7 Tesis PPP Kongzhe Chen

164

low elevation. These mapping functions include the broadcast model, Single Layer Model

(SLM) and Modified Single Layer Model (MSLM) mapping functions.

The broadcast model mapping function is given as (Klobuchar, 1987):

−⋅+=90

9621

e)e(F (5.15)

where e is the elevation angle.

The single-layer model (SLM) mapping function was given in Equation 5.14. The SLM

assumes that ionospheric electron density is concentrated on a thin shell of height H above

the mean Earth radius ER (Manucci et al., 1993). In practice, the shell height can be selected

from 300 to 500 km. 350, 400 and 450 km are most commonly used, and are denoted as

SLM350, SLM400 and SLM450 in this thesis.

The modified single-layer model (MSLM) mapping function can be expressed as (Schaer,

1999a):

( )( )( )

2

901

1

+−

=

esinHR

ReF

E

E β (5.16)

where β is a correction factor which is close to unity. In this research, the coefficients,

which best fit the JPL extended slab model mapping function, are adopted with H = 506.7

Page 189: 7 Tesis PPP Kongzhe Chen

165

km and β = 0.9782 (CODE Ionosphere Map website, 2004). It is denoted as MSLM506 in

the thesis.

The mapping functions discussed above differ primarily at low elevation angles, e.g., lower

than 15°. At high elevation angles, they are very close. This behaviour can be seen in Figure

5.5 which shows the mapping functions. To highlight the differences more clearly, using the

IGS selected SLM450 as a reference, the values from which other mapping functions differ

are plotted in Figure 5.6 using a large scale. At elevation greater than 15°, the differences of

all mapping functions to SLM450 are less than 0.2. Even at elevation 5°, the differences are

still less than 0.3, which is less than 10% of the value of the mapping functions at this

elevation. It is interesting to see that at all elevations, MSLM506 is smaller than SLM450,

SLM350 and SLM400 are bigger than SLM450. On the other hand, the broadcast model

mapping function is smaller than SLM450 at elevations less than 55°, but bigger than

SLM450 at elevation greater than 55°. Because of the strong correlation existing between the

mapping functions and the zenith ionospheric delay estimates, these mapping functions may

provide similar position solutions but different zenith ionospheric delay estimates. The

numerical results will further confirm this.

Page 190: 7 Tesis PPP Kongzhe Chen

166

Figure 5.5 Mapping Functions

Figure 5.6 Mapping Function Difference

Page 191: 7 Tesis PPP Kongzhe Chen

167

In this research, the broadcast model mapping function will be used along with Klobuchar

model. The Global Ionospheric Model will adopt SLM450 to keep consistent with IGS

ionospheric products. The same mapping function will also be used in the proposed

ionospheric estimation model, because the use of SLM450 provides better position solutions

than other mapping functions as tested in Section 5.3.3.

5.3.3 Results of Positioning and Ionospheric Delay Estimation

To test the performance of the ionospheric estimation model in positioning and ionospheric

delay estimation, a dataset collected in GPS week 1251 from IGS station AMC2 (38.8° N,

104.5° W, 1911.4 m) was processed. The necessary GPS observation files and IGS Final

orbit and clock products were obtained from the IGS website. A dual-frequency receiver was

used at AMC2 station but only single-frequency observations, 1P and 1Φ , were processed.

The P3-RT software package, which will be discussed in Chapter 6, was used to conduct the

processing. The ionospheric activity was typical throughout the selected week. The Ap

indices are shown in Table 5.2.

The index Kp and its related indices ap and Ap have been widely used to measure the world

wide geomagnetic activity. The three-hour-range Kp is obtained as the mean value from the

standardized K index (Ks) of 13 magnetic observatories. The three-hour-range ap index is

derived from the Kp index by a linear scale. The Kp index can be one of the 28 values ranged

from 0o to 9o. The corresponding ap index is ranged from 0 to 400. To measure the daily

Page 192: 7 Tesis PPP Kongzhe Chen

168

world wide geomagnetic activity, the daily Ap index is usually used, which is the average of

the 8 three-hour-range ap indices in one day. In this thesis, the Ap index is obtained from the

International Service of Geomagnetic Indices (ISGI) website (2004).

Table 5.2 Ap Indices of GPS Week 1251

Day 28 29 30 31 1 2 3

Ap index 13 5 6 19 24 15 22

The zenith ionospheric delay, zd , was modeled as a random walk process with a spectral

density of h/m2 . The gradient parameters, nG and eG , were also modeled as random

walk processes with a spectral density of h/dm2 . An elevation cut-off angle of 7° was

used. The receiver’s clock was modeled as a white noise process. The float ambiguity

parameters were estimated as constants. Station coordinates were estimated as random walk

processes with a spectral density of h/km60 . The sample interval of GPS measurements is

30 s.

The positioning errors are shown in Figure 5.7 while Figures 5.8, 5.9 and 5.10 show the

estimated zenith ionospheric delay and horizontal gradients in the east and north respectively.

As shown in Table 5.3, about 30 cm accuracy in the horizontal components and 50 cm

accuracy in the vertical component were obtained at this mid-latitude station for one week. In

Figure 5.7, the correlation between ionospheric conditions and positioning errors was also

indicated. The Ap indices are bigger in December 31st, January 1st and January 3rd than in

Page 193: 7 Tesis PPP Kongzhe Chen

169

other days. As a result, the positioning errors are bigger in these days as compared to other

days.

Figures 5.8 to 5.10 have clearly shown the daily ionospheric activity at AMC2. The half

cosine patterns in Figure 5.8 show the diurnal variations of TEC, with some disturbances.

Figures 5.9 and 5.10 show decimetre level gradients in the east and north directions, and they

also indicate three typical ionospheric gradients found at mid-latitude stations. Figure 5.9

indicates that TEC increases from west to east in the morning and from east to west in the

afternoon. A positive gradient in the morning (about 9:00 local time) and a negative gradient

in the afternoon (about 18:00 local time) were obtained in the east direction daily. The

negative gradients in the north direction shown in Figure 5.10 indicate TEC increases

towards equator at mid-latitude regions.

The zenith ionospheric delay estimates were also transferred to VTEC and compared with

IGS Final ionospheric products. Estimates obtained at every 2 hours were used for the

comparison as that matched the temporal resolution of IGS ionospheric TEC grids. The first

estimate at GPS time (1251, 0) was also omitted because of the un-converged ambiguity. The

comparison was shown in Figure 5.11. The unit is in 0.1 TECU to be consistent with IGS

products. The RMS difference is about 1.2 TECU with a bias of about -0.6 TECU.

Page 194: 7 Tesis PPP Kongzhe Chen

170

Figure 5.7 Positioning Errors Using Ionospheric Estimation Model for GPS Week 1251

Table 5.3 Positioning Accuracy Using Ionospheric Estimation Model with SLM450

RMS (m) BIAS (m) STD (m) Latitude 0.260 -0.097 0.241

Longitude 0.248 -0.084 0.234

Height 0.486 -0.050 0.483

Page 195: 7 Tesis PPP Kongzhe Chen

171

Figure 5.8 Zenith Ionospheric Delay Estimates for GPS Week 1251

Figure 5.9 Ionospheric Gradients in the East Direction

Page 196: 7 Tesis PPP Kongzhe Chen

172

Figure 5.10 Ionospheric Gradients in the North Direction

Figure 5.11 Vertical TEC Comparison

To compare the performance of different mapping functions for positioning and ionospheric

delay estimation, the same dataset was also processed using the broadcast model mapping

Page 197: 7 Tesis PPP Kongzhe Chen

173

function, SLM350, SLM400 and MSLM506. The positioning accuracy statistics are shown in

Table 5.4. The VTEC estimation statistics, as compared with IGS ionospheric products, are

presented in Table 5.5.

Table 5.4 Positioning Accuracy Using Ionospheric Estimation Model with Different Mapping Functions (Unit: m)

Broadcast SLM350 SLM400 SLM450 MSLM507

RMS BIAS RMS BIAS RMS BIAS RMS BIAS RMS BIAS

Latitude 0.264 -0.095 0.265 -0.092 0.260 -0.094 0.260 -0.097 0.268 -0.103

Longitude 0.254 -0.097 0.260 -0.098 0.253 -0.091 0.248 -0.084 0.243 -0.067

Height 0.490 -0.066 0.504 -0.128 0.489 -0.086 0.486 -0.050 0.502 0.009

Table 5.5 Statistics of VTEC Estimation with Different Mapping Functions (Unit: 0.1 TECU)

Broadcast SLM350 SLM400 SLM450 MSLM507

RMS 20.60 21.69 17.53 12.28 11.25

BIAS -18.19 -19.39 -14.37 -5.89 -0.05

As shown in Tables 5.4 and 5.5, different mapping functions provide similar positioning

accuracy with small differences in the height component, even though their VTEC estimates

vary significantly. The ionospheric estimation model provides the best positioning results

when using SLM450, but its VTEC estimates are not the best among these mapping

functions when compared with the IGS ionospheric products. On the other hand, the VTEC

estimates using MSLM507 are most consistent with the IGS ionospheric products, which use

the SLM450 mapping function. Using the same mapping function, SLM450, a -0.6 TECU

bias between the ionospheric estimation model and IGS ionospheric products exists. The

accuracy of IGS ionospheric grids, which were selected as references in this research, is

Page 198: 7 Tesis PPP Kongzhe Chen

174

about 2 TECU at the grid points at a sample interval of 2-hour. The accuracy degrades after

interpolation to the ionospheric pierce point at a selected epoch. The biases may also be due

to the errors in the IGS ionospheric grids. The accuracy of vertical TEC estimation will be

further investigated if more accurate ionospheric products are available to be used as

references.

The broadcast model, SLM300 and SLM400 seem to have underestimated the vertical TEC

with negative biases of 1 to 2 TECU when compared to the IGS ionospheric products. The

mapping function comparison in Figure 5.6 can be used to explain the biases. As shown in

Figure 5.6, SLM300 and SLM400 are bigger than SLM450, and the broadcast model is also

bigger than SLM450 at elevations less than 55°. When mapped to the slant TEC, the

underestimated vertical TEC is counteracted by the overestimated mapping functions. As a

result, these mapping functions provided similar slant ionospheric group delays or phase

advances for observations, and offered similar positioning results.

To compare the positioning accuracy with other models, the dataset was also processed using

the Klobuchar model and GIM. An elevation cut-off angle of 10° was used instead of the 7°

used for the ionospheric estimation model to estimate the horizontal gradients. The settings

for coordinate and receiver clock parameter estimation are the same as used in the

ionospheric estimation model. The zenith tropospheric delay, if estimated, is estimated as a

random walk process with a spectral density of h/cm1 . The float ambiguity parameters, if

estimated, were estimated as constants. The results are shown in Tables 5.6 and 5.7.

Page 199: 7 Tesis PPP Kongzhe Chen

175

When using the Klobuchar model, two processing strategies have been tested. The first

processing strategy uses only code observables, while the second one uses both code and

phase observables. As shown in Table 5.6, the phase observables can slightly improve the

positioning accuracy for this mid-latitude station. The improvements are mainly in the

vertical component. As will be seen in Chapter 6, the phase observables would not

necessarily improve the position solutions all the time when using Klobuchar model.

Another processing strategy has been tested using GIM in addition to the code processing,

and the processing using both code and carrier phase. Using both code and phase

observables, it estimates zenith tropospheric delay (ZTD) instead of modeling it as done in

the first two processing strategies. As shown in Table 5.7, the phase observables can improve

the positioning accuracy but estimating zenith tropospheric delay does not improve the

accuracy. When processing using the Klobuchar model, because of the big ionospheric

residuals, ZTD estimation is not recommended. Because there are 7 parameters (3

coordinates, 1 receiver clock offset and 3 ionospheric parameters) to be estimated in the

ionospheric estimation model in addition to the float ambiguity parameters, estimating ZTD

will degrade the solution if errors in measurements have not been mitigated effectively. With

this in mind, the zenith tropospheric delay will not be estimated in the following processing,

it will be modeled using Saastamoinen model and meteorological settings.

Page 200: 7 Tesis PPP Kongzhe Chen

176

Table 5.6 Positioning Accuracy Using Klobuchar Model

Code (m) Code & Phase (m) Latitude 0.872 0.613

Longitude 0.304 0.978

Height 1.743 1.056

Table 5.7 Positioning Accuracy Using GIM

Code (m) Code & Phase (m) Code & Phase with ZTD estimated (m)

Latitude 0.501 0.319 0.319

Longitude 0.221 0.352 0.354

Height 0.756 0.417 0.436

From Tables 5.3, 5.5, 5.6 and 5.7, several conclusions can be made. VTEC (or zenith

ionospheric delay) can be estimated at 1~2 TECU accuracy using un-differenced single-

frequency GPS observations at mid-latitude region, which has great promise for single-

frequency point positioning. Decimetre level accuracy is obtainable using single-frequency

measurements. Phase observables can improve positioning accuracy at mid-latitude stations

using the GIM, although estimating ZTD cannot improve the positioning accuracy. The

improvement of using phase observables to the Klobuchar model is not obvious. The

ionospheric estimation model and GIM perform better than the Klobuchar model. The

ionospheric estimation model can provide comparable accuracy to GIM, which uses the IGS

ionospheric products with a latency of about 11 days. Because several types of real-time

precise GPS orbit and clock products have become available with comparable accuracy to the

IGS post-mission orbits and clocks, the accuracy obtained using the ionospheric estimation

Page 201: 7 Tesis PPP Kongzhe Chen

177

model and Klobuchar model is obtainable in real-time. In Section 6.5, single-frequency

precise point positioning results will be presented using JPL real-time GPS products.

Page 202: 7 Tesis PPP Kongzhe Chen

178

CHAPTER 6

NUMERICAL RESULTS AND ANALYSIS

This chapter will present numerical results for precise point positioning, receiver clock offset

estimation and atmospheric sensing using real-time orbit and clock products. Section 6.2 will

show position determination results using dual-frequency measurements. Datasets were

collected from receivers at a fixed site as well as on a vehicle and aircraft. These datasets are

processed to show the performance of positioning under different dynamic environments.

The results of receiver clock offset estimation are shown in Section 6.3. Section 6.4 shows

zenith tropospheric delay estimation and water vapour sensing results. The results of precise

point positioning using single-frequency measurements are presented in Section 6.5. Some of

the processing was conducted in real-time using real-time orbit and clock products received

over the Internet or a serial port, while other processing was done in a simulated real-time

mode with products generated for real-time applications. All processing tasks were

accomplished by the P3-RT software package, which will be described in the next section.

Page 203: 7 Tesis PPP Kongzhe Chen

179

6.1 Software Development and Parameter Modeling

A software package, P3-RT, has been developed to perform all processing tasks required in

this thesis. Details of the software and the modeling for parameters are described in the

following.

6.1.1 P3-RT Software Package

Interface

P3-RT was developed using C++ on the Microsoft Windows Operating System and features

a user-friendly interface. Figure 6.1 shows the interface of P3-RT for setting up new

processing tasks.

In the Settings dialog, user can set basic settings such as elevation cut-off angle and

measurement noise. If a standardized antenna name is selected, antenna phase center offset

and variations would be applied in the software according to the parameters in the

standardized IGS antenna table.

Initial coordinates can also be set for the processing of static positioning, timing or

atmospheric sensing. In static positioning, if the precisely known coordinates are given, the

coordinates can be used to calculate accuracy statistics. In timing and atmospheric sensing,

precisely determined coordinates can be used to fix the position parameters, which thus can

be removed from the unknowns in post-mission PPP processing. In this research however,

Page 204: 7 Tesis PPP Kongzhe Chen

180

coordinates are still estimated in timing and atmospheric sensing data processing in real-time

or simulated real-time mode.

Estimating ZWD and modeling ZHD are recommended in this thesis when dual-frequency

data is used. Meteorological settings can be used to calculate an apriori ZWD. User can also

choose to model both ZHD and ZWD based on precise meteorological measurements which

can be input from a source file. But as discussed in Section 3.2, the spatial and temporal

variability of water vapour makes modeling ZWD difficult and cannot be modeled any better

than 1~2 cm even when using precise meteorological measurements.

Finally, users can also exclude “unhealthy” satellites from the processing manually in post

processing.

Page 205: 7 Tesis PPP Kongzhe Chen

181

Figure 6.1 P3-RT Interface – Setup

Figures 6.2 and 6.3 show the interface of P3-RT during kinematic and static processing,

respectively.

Page 206: 7 Tesis PPP Kongzhe Chen

182

Figure 6.2 P3-RT Interface – Kinematic Processing

Page 207: 7 Tesis PPP Kongzhe Chen

183

Figure 6.3 P3-RT Interface – Static Processing

Figures 6.4 and 6.5 show the interface of P3-RT for results analysis. Figure 6.4 displays the

trajectory in horizontal and vertical components separately for kinematic positioning. If

ground-truth coordinates, such as position solutions from double differencing techniques, are

provided, the position errors can also be displayed as shown in Figure 6.5.

Page 208: 7 Tesis PPP Kongzhe Chen

184

Figure 6.4 P3-RT Interface – Trajectory

Figure 6.5 P3-RT Interface – Position Errors

Page 209: 7 Tesis PPP Kongzhe Chen

185

Real-Time and Post-Mission Capability

The primary purpose of P3-RT is for real-time processing. Currently, it can also perform

post-mission processing using precise products (SP3 format orbits and RINEX format

clocks) from IGS or other agencies, or simulated real-time processing using products

generated for real-time applications. All of these modes have been used in this research.

In real-time mode, GPS raw data can be accessed through a computer’s serial port or over the

Internet. Acquiring GPS raw data over the Internet is useful for atmospheric sensing, in

which data from a GPS network is normally processed. Currently, the data stream from

several types of receivers can be processed directly in P3-RT. These include receivers

manufactured by Javad, Leica, Trimble, etc. More data stream formats will be supported in

the future. Precise GPS orbit and clock products can also be received through the serial ports

or over the Internet. Two formats of real-time orbit and clock products are currently

supported. These formats are the GPS•C format and the IGDG format discussed in Chapter

4. Figure 6.6 shows the real-time PPP settings.

Page 210: 7 Tesis PPP Kongzhe Chen

186

Figure 6.6 Real-Time PPP

In post-mission mode, GPS raw data can be input in the standard RINEX format. More

precise GPS products are also supported in post-mission. Orbits in the SP3 format and clocks

in RINEX format from IGS or other agencies at any sample intervals are supported. Products

with a low sample rate for orbits and clocks would be automatically interpolated to the time

tag of GPS measurements. Though designed for real-time applications, the high sample rate

orbit and clock products in the GPS•C and IGDG formats can also be used in post

processing. Figure 6.7 shows the PPP settings for post-mission.

In simulated real-time mode, only previously logged real-time orbit and clock products in the

GPS•C and IGDG formats are supported. The software intentionally delays the logged real-

Page 211: 7 Tesis PPP Kongzhe Chen

187

time orbit and clock products for a period. This delay is used to simulate the real-time latency

of the products that would be found in a normal real-time process. The simulated real-time

processing has also been used by other researchers when a pure real-time processing is

impractical (Muellerschoen et al., 2001). In this research, the test for receiver clock offset

estimation was conducted in this simulated real-time mode because the GPS raw data of

station AMC2 was not available in real-time.

Figure 6.7 Post-Mission PPP

Page 212: 7 Tesis PPP Kongzhe Chen

188

Positioning, Timing and Atmospheric Sensing with P3-RT

P3-RT is a realization of PPP methodology. It can be used in positioning, timing and

atmospheric sensing applications in real-time and post-mission.

For positioning applications, P3-RT can offer position solutions using dual-frequency or

single-frequency measurements. As will be demonstrated later in this chapter, it can provide

sub-centimetre level static position solutions or sub-decimetre level kinematic position

solutions in real-time or post-mission using dual-frequency receivers. The models discussed

in Chapter 5 have been implemented into P3-RT for single-frequency data processing. Using

single-frequency measurements, sub-metre level kinematic positioning accuracy has also

been obtained.

Sub-nanosecond level receiver clock offset estimates that will be presented in Section 6.3

show the promise of P3-RT in timing applications.

Sub-centimetre level ZTD estimates will be shown in Section 6.4. Provided precise pressure

and temperature measurements are available in real-time, P3-RT is capable of outputting sub-

millimetre level PWV estimates in real-time.

6.1.2 Modeling for Parameters

In P3-RT, invariable parameters, such as the receiver coordinates in static mode and the

ambiguity parameters, are estimated as constants. The receiver clock offset is modeled as a

Page 213: 7 Tesis PPP Kongzhe Chen

189

white noise process. Other variable parameters, such as the coordinates in kinematic mode,

the tropospheric parameters and/or the ionospheric parameters, can be modeled as random

walk processes or first-order Gauss-Markov processes. Typical modelling settings for the

processing in this research are discussed in the following.

In this research, the receiver coordinate parameters are estimated as constants in the static

mode. For kinematic positioning, they are modeled as random walk processes with spectral

densities that will be defined according to the actual dynamic environments. For example, for

land vehicles, spectral densities of h/km60 and h/km6 can be used for horizontal and

vertical components respectively. While for airborne datasets, spectral densities of

h/km100 ~ h/km1000 and h/km20 can be used for horizontal and vertical

components respectively. For shipborne datasets, spectral densities of h/km60 and

h/m60 can be used for horizontal and vertical components respectively. When static

datasets are processed in an epoch-by-epoch mode, a spectral density of h/km60 is used

for both horizontal and vertical components.

The tropospheric zenith wet delay, if estimated, is modeled as a random walk with a spectral

density of h/mm6 to h/cm2 depending the dynamic environments and meteorological

conditions. The tropospheric horizontal gradient parameters, if estimated, are modeled as

random walk processes with a spectral density of h/mm.30 .

Page 214: 7 Tesis PPP Kongzhe Chen

190

In single-frequency point positioning using the ionospheric estimation model, the zenith

ionospheric delay is modeled as a random walk with a spectral density of h/m2 . The

ionospheric horizontal gradient parameters are modeled as random walk processes with a

spectral density of h/dm2 .

The receiver clock is modeled as a white noise process. The float ambiguity parameters are

estimated as constants.

The apriori uncertainties for parameters would be set based on the initial values of the

parameters. For example, the uncertainties for coordinate parameters can be safely set as 100

km if only approximate coordinates are provided.

The initial values of the float ambiguities are normally calculated from the code

measurements. An uncertainty of 10 m is used in this research.

The initial value of the tropospheric zenith wet delay is calculated from meteorological

settings. Because the zenith wet delay is normally at the decimetre level, plus the residual of

the zenith hydrostatic delay, which should be less than one decimetre using a pressure model

that will be discussed in Section 6.4, the uncertainty of the zenith wet delay can be set as

0.1~1 m. For the tropospheric horizontal gradients, which are normally at the centimetre

level, an uncertainty of 0.01 m is used.

In single-frequency point positioning using the ionospheric estimation model, the initial

value of the zenith ionospheric delay is calculated from the broadcast Klobuchar model.

Page 215: 7 Tesis PPP Kongzhe Chen

191

Considering that the model performance and typical zenith ionospheric delay, the uncertainty

of the zenith ionospheric delay can be set as 3 m. For the ionospheric horizontal gradients,

which are normally at the decimetre level, an uncertainty of 0.1 m is used.

Normally, a large uncertainty should used for receiver clock offset, i.e., 3×105 m in range or

1 ms in transmission time. This is due to the fact that most receivers try to synchronize their

internal clocks with GPS time within 1 ms.

6.2 PPP Using Dual-Frequency Measurements

In this section, several datasets collected using dual-frequency GPS receivers are processed

using real-time orbit and clock products, including JPL IGDG products, NRCan GPS•C code

solution products and GPS•C phase solution products. Some datasets are processed in

simulated real-time or post-mission mode because GPS measurements or orbit and clock

products are not available in real-time.

6.2.1 Static PPP Using IGDG Products

A real-time static test was carried out on December 3rd, 2003. One Javad Legacy dual-

frequency receiver was set up on the S2 pillar on the roof of the Engineering Building at the

University of Calgary. JPL IGDG orbit and clock products were received over the Internet.

The sample interval was set to 10 s. The coordinates of S2, which were determined in ITRF-

Page 216: 7 Tesis PPP Kongzhe Chen

192

1993 with centimetre level accuracy, were transformed to ITRF-2000 using coefficients

provided by McCarthy and Petit (2004) and used as true coordinates. The positioning errors

are shown in Figure 6.8. The accuracy statistics of the position solutions after convergence is

presented in Table 6.1.

It took approximately 20 minutes for the positioning errors to converge to the decimetre

level. But it took more than 1 hour to converge to the centimetre level. The long convergence

time may be caused by multipath effects on the roof, which have been investigated in Ray

(2000). The antenna used in this test is a JPSLEGANT antenna with a small ground plane,

which can partially mitigate the multipath effects. After convergence, the RMS of the

positioning errors is about 2 cm in each positioning component. The centimetre level biases

would be partially caused by the reference coordinates. The coordinates of S2 were

determined several years ago with an accuracy of centimetre level.

Figure 6.8 Real-Time Static Positioning Using IGDG Products

Page 217: 7 Tesis PPP Kongzhe Chen

193

To further test the obtainable accuracy of PPP in static mode using JPL IGDG products,

accurate coordinates for the test stations in ITRF-2000 are required to serve as the ground-

truth. The coordinates of IGS stations are estimated on a daily basis in the ITRF-2000 frame

currently. IGS stations are also normally set up in multipath-friendly environment with a

Choke Ring antenna. Raw data and coordinates of these stations can be downloaded from

IGS website. In this test, GPS data acquired for the day of August 4, 2004 at IGS station

ALGO was processed in a simulated real-time mode using the JPL real-time orbit and clock

products acquired from JPL server. The positioning results are shown in Figure 6.9 and the

accuracy statistics for results after convergence are given in Table 6.1. It can be seen that the

coordinate estimates could converge to the centimetre level within 30 minutes. This

convergence time is much shorter than the convergence time with data affected by multipath.

After the convergence, all position coordinate components are accurate at the sub-centimetre

level. The results in Table 6.1 indicate that PPP is capable of providing real-time sub-

centimetre level accuracy for static control survey.

Figure 6.9 Static Positioning Using IGS Dataset

Page 218: 7 Tesis PPP Kongzhe Chen

194

Table 6.1 Accuracy Statistics of Static Positioning Results (Unit: cm)

Real-time dataset IGS dataset

Latitude 2.3 0.9

Longitude 1.4 1.0 RMS

Height 2.6 0.7

Latitude -1.7 0.8

Longitude -1.2 0.3 BIAS

Height 2.4 0.0

Latitude 1.6 0.3

Longitude 0.8 0.9 STD

Height 0.8 0.7

6.2.2 Kinematic PPP Using IGDG Products

Two kinematic datasets were processed using JPL IGDG orbit and clock products. The first

dataset was collected on a vehicle and processed in real-time. Another dataset was collected

by an aircraft and processed in a simulated real-time mode.

Land Vehicle Kinematic Positioning

A kinematic positioning test using a land vehicle was conducted on September 30th, 2003.

The vehicle was driven along the highway at a speed of 80 km/h near Springbank, Alberta. In

order to establish a reference trajectory for the vehicle, a reference receiver was set up at one

control point of the Springbank Baseline Network so double difference data processing could

be performed to establish a reference for accuracy assessment. Both the control point and

vehicle used Javad Legacy dual-frequency receivers with the JPSLEGANT antennas. A

Page 219: 7 Tesis PPP Kongzhe Chen

195

CDPD radio was used to receive JPL IGDG real-time precise orbit and clock products via the

Internet. The sample rate of the two GPS receivers was set to 1 Hz. The PPP solutions are

obtained using P3-RT software while the double difference solutions are obtained using the

commercial software package GrafNav from the Waypoint Consulting Inc. With a relatively

short baseline length (7 km on average and maximum of 12 km), the ambiguity-fixed

position results from the GrafNav can serve as the ground-truth to assess the positioning

accuracy of PPP solutions.

The positioning differences between PPP and double difference solutions are shown in

Figure 6.10, the trajectory of vehicle is presented in Figure 6.11, and the positioning accuracy

statistics after convergence is given in Table 6.2. They indicate that sub-decimetre accurate

positioning results have been obtained in real-time using the precise point positioning method

with a convergence time of about half an hour.

Figure 6.10 Positioning Errors with Vehicle Dataset

Page 220: 7 Tesis PPP Kongzhe Chen

196

Figure 6.11 Vehicle Trajectory on September 30th, 2003

Airborne Kinematic Positioning

The airborne dataset, which was provided by the Mosaic Mapping System Inc. (2004), was

collected on August 28th, 2004 at 40 km north of Halifax, Nova Scotia. A NovAtel Black

Diamond GPS receiver and Model 512 antenna were set up on a helicopter. The sample rate

of the two GPS receivers was 1 Hz. The helicopter was flying at an altitude of approximately

250 m above the ground at 50 knots. The distance between the rover and base is less than 10

km. During the test, a NovAtel DL-4 receiver and an antenna with a ground plane were used

as the base station. The double-difference ambiguity-fixed position solutions, provided by

Mosaic Mapping System Inc. (2004), have been used as the ground-truth.

The positioning errors, trajectory and accuracy statistics after position convergence are

shown in Figure 6.12, Figure 6.13, and Table 6.2.

Page 221: 7 Tesis PPP Kongzhe Chen

197

Figure 6.12 Positioning with Aircraft Dataset

Figure 6.13 Aircraft Trajectory on August 28th, 2004

Page 222: 7 Tesis PPP Kongzhe Chen

198

Table 6.2 Accuracy Statistics of Kinematic Positioning Results (Unit: cm)

Vehicle dataset Aircraft dataset

Latitude 7.8 2.8

Longitude 5.9 6.8 RMS

Height 7.9 4.9

Latitude 3.6 -0.2

Longitude -3.5 -1.5 BIAS

Height -0.7 -1.5

Latitude 6.9 2.8

Longitude 4.8 6.7 STD

Height 7.9 4.6

As shown in Figures 6.10 and 6.12, about 20 to 30 minutes was taken for the positioning

errors converge to the decimetre level in kinematic mode. After convergence, sub-decimetre

level accuracy was obtained in each positioning component. These tests confirmed that PPP

is a very promising technique for high precision real-time kinematic positioning.

The results obtained from processing the aircraft dataset are even better than those obtained

from the vehicle dataset. It took only 20 minutes for the positioning errors to converge to the

decimetre level. Satellite geometry, observation environment, and improvement in the IGDG

orbit and clock products are possible reasons for the better results.

Shown in Figures 6.14 and 6.15 are the satellite geometry for the aircraft and vehicle

datasets. It is obvious that the observation environment in the air is much better than that on a

vehicle. Though at least 5 satellites were tracked when driving along the high way, many

loss-of-locks were found in the vehicle dataset because signals were blocked by hills or

Page 223: 7 Tesis PPP Kongzhe Chen

199

buildings. The loss-of-locks definitely degraded the results because it takes time for the float

ambiguities of the new tracked satellites to converge.

The aircraft dataset was collected about 1 year later than the real-time vehicle testing. More

tracking stations have been added to the JPL IGDG real-time tracking network during this

period and new processing strategies were adopted by JPL in April 2004 (Muellerschoen,

2004). All these led to improvements of JPL IGDG products.

Figure 6.14 Satellite Geometry for Vehicle Dataset on September 30th, 2003

Page 224: 7 Tesis PPP Kongzhe Chen

200

Figure 6.15 Satellite Geometry for Aircraft Dataset on August 28th, 2004

6.2.3 PPP Using GPS••••C Products

In this section, the two types of NRCan precise orbit and clock products were tested in real-

time and post-mission for precise point positioning.

Positioning with GPS••••C Code Solution Products

GPS•C code solution products are broadcast over satellite and the Internet in real-time

(Kassam, 2003; Chen et al., 2002). Using orbit and clock products received over the Internet,

half-metre accuracy has been obtained using data from IGS stations. The accuracy was

degraded to the metre level when processing a dataset collected on the roof of Engineering

Building because of multipath effects in code measurements (Chen et al., 2002).

Page 225: 7 Tesis PPP Kongzhe Chen

201

In this test, a CPGPS radio was used to receive GPS•C code solution orbit and clock

products. The real-time test was carried out on December 2nd, 2003. The settings are the same

as the real-time static test on December 3rd, 2003 that has been described in Section 6.2.1.

The phase smoothed, ionosphere-free code observables were processed in real-time in an

epoch-by-epoch mode. The coordinates from PPP processing, which were based on NAD83,

were transformed into ITRF93 using coefficients provided by Kouba (2002) and compared

with the true coordinates. The positioning errors and satellite geometry are shown in Figures

6.16 and 6.17. The accuracy statistics are presented in Table 6.3.

The positioning errors are correlated with the satellite geometry. Sub-metre level accuracy

was achieved in real-time. The positioning errors were dominated by the measurement noises

and multipath effects in the ionosphere-free code combinations, and errors in the GPS•C

code solution orbit and clock products.

Figure 6.16 Kinematic Positioning Using GPC••••C Code Solution Products

Page 226: 7 Tesis PPP Kongzhe Chen

202

Figure 6.17 Satellite Geometry of S2 on December 2nd, 2003

Positioning with GPS••••C Phase Solution Products

GPS•C phase solution orbit and clock products are still at the testing stage. Their latency is

still several hours, and they are not continuously available because of occasionally missing

real-time raw data from several tracking stations (Collins, 2004). For this test, several days

worth of orbit and clock products were provided by NRCan. One dataset collected on S1

pillar on the roof of Engineering Building on June 10th, 2004 using a Javad receiver was

processed using the phase solution products with a sample interval of 10 s. Dual-frequency

code and phase measurements were used for position determination in an epoch-by-epoch

mode. The positioning errors are presented in Figure 6.18 and Table 6.4.

Page 227: 7 Tesis PPP Kongzhe Chen

203

The positioning accuracy and convergence time are comparable with those obtained using the

kinematic dataset with the IGDG orbit and clock products. But the GPS•C products are still

not as consistent as the IGDG products. The large errors during the time period from GPS

time 403200 s to 417600 s were due to the degraded orbit and clock products (Collins, 2004).

Currently only 20 stations are used by NRCan to estimate the phase solution orbit and clock

products as compared to over 60 stations used by JPL for IGDG products computation.

Figure 6.18 Kinematic Positioning Using GPC••••C Phase Solution Products

Page 228: 7 Tesis PPP Kongzhe Chen

204

Table 6.3 Accuracy Statistics of Positioning Using GPS••••C Products (Unit: m)

Code Solution

(December 2, 2003) Phase Solution (June 10, 2004)

Latitude 0.426 0.048

Longitude 0.232 0.065 RMS

Height 0.686 0.065

Latitude 0.122 0.010

Longitude -0.070 0.007 BIAS

Height -0.117 0.011

Latitude 0.408 0.047

Longitude 0.221 0.065 STD

Height 0.676 0.064

6.2.4 Summary

From the results presented above, the following conclusions can be made.

Sub-centimetre to sub-metre level accuracy was obtained by processing dual-frequency

measurements. But the accuracy is related to the performance of precise orbit and clock

products. Using the IGDG real-time orbit and clock products, sub-centimetre level static and

sub-decimetre kinematic accuracy is obtainable in real-time. GPS•C code solution products,

which are only suitable for code processing, can provide sub-metre level kinematic accuracy

in real-time. GPS•C phase solution products, which are not yet available in real-time, have

the potential to offer comparable accuracy to that of the IGDG products but they are still not

as consistent. The tests using datasets collected on the roof of the Engineering Building have

indicated the qualities of these products. Though the datasets were collected at different

Page 229: 7 Tesis PPP Kongzhe Chen

205

periods and/or different sites, the datasets are one full day in length. Therefore, the satellite

geometry should be similar for these tests on the roof. The observation conditions on the roof

are similar. The ionospheric effects might be different significantly for different seasons, but

dual-frequency ionosphere-free combinations were used in these test.

When processing phase measurements, the convergence time is affected by a variety of

factors, including satellite geometry, multipath effects, performance of precise orbit and

clock products, etc. As demonstrated by the discussed tests, the typical convergence time is

about 20~30 minutes for the positioning errors to converge to the decimetre level. In static

mode, 30 minutes to 1 hour is required for the positioning results to converge to centimetre

level. To shorten the convergence time, fixing or pseudo-fixing the ambiguity may be

required in real-time positioning (Gao and Shen, 2002). In post-mission processing,

backward processing can be used to get consistent accuracy for the entire data period (Gao et

al., 2005).

6.3 Receiver Clock Offset Estimation Using PPP Methodology

In addition to position determination, PPP can also output receiver clock offset solutions

which have the potential to support precise timing applications. Unlike the positioning

accuracy assessment, in which the ground-truth can be established using either double

differencing solutions or comparing with coordinates from precisely surveyed stations, the

receiver clock offset estimates are time reference dependent and normally no ground-truth is

Page 230: 7 Tesis PPP Kongzhe Chen

206

available for the clock offset value of a specific receiver. To estimate precise orbit and clock

products, usually one receiver clock, equipped with a hydrogen maser external frequency, is

fixed and used as a time reference. For example, the JPL IGDG orbit and clock products are

generated using the clock of IGS station AMC2 as the reference clock (Muellerschoen,

2003). One can assess the accuracy of receiver clock offset estimation from PPP by

processing the GPS data from the AMC2 station equipped with the reference clock. The

resultant receiver clock estimates from PPP solutions for station AMC2 should theoretically

equal zero using the precise orbit and clock products that are referenced to the clock, and the

variations in the solutions should then directly reflect the quality of the clock solutions using

PPP method. As described in Section 2.3, PPP using the IGDG products also provides a new

method to recover UTC(USNO) with an accuracy of a few nanoseconds in real-time using a

single GPS receiver.

6.3.1 Receiver Clock Offset Estimation Using IGDG Products

In this test, the receiver clock offset was estimated as a white noise process using GPS data

from ACM2 station acquired from June 12th, 2004 to June 14th, 2004. The IGDG orbit and

clock products were used to do processing in a simulated real-time mode. The station

coordinates were estimated as constants. ZWD was estimated along with horizontal gradients

as random walk processes, while ZHD was modeled using Saastamoinen model. Figures

6.19, 6.20 and 6.21 present the receiver clock offset estimates for ACM2 station. The clock

offset estimates are also considered as errors in clock offset estimation as the theoretical

Page 231: 7 Tesis PPP Kongzhe Chen

207

value of the clock offset is zero. For the purpose of correlation analysis, which will be

discussed in Section 6.3.2, ZTD estimates were also displayed in the figures. Table 6.4

provides the statistics of the estimation accuracy.

It took about half an hour for the float ambiguities to converge for processing the data each

day. The receiver clock offset estimates remain less than 0.2 ns after the ambiguity

convergence. The RMSs of receiver clock offset estimates after the ambiguity convergence

in the three test days are 0.077 ns, 0.106 ns and 0.091 ns.

The accuracy of about 0.1 ns indicates that PPP is capable of providing real-time receiver

clock offset estimates at sub-nanosecond accuracy, making it a promising tool for time

transfer.

Page 232: 7 Tesis PPP Kongzhe Chen

208

Figure 6.19 Receiver Clock Offset and ZTD Estimates on June 12th, 2004

Page 233: 7 Tesis PPP Kongzhe Chen

209

Figure 6.20 Receiver Clock Offset and ZTD Estimates on June 13th, 2004

Page 234: 7 Tesis PPP Kongzhe Chen

210

Figure 6.21 Receiver Clock Offset and ZTD Estimates on June 14th, 2004

Table 6.4 Receiver Clock Offset Estimation Accuracy

Day RMS (ns) BIAS (ns) STD (ns)

June 12th, 2004 0.077 0.018 0.075

June 13th, 2004 0.106 -0.003 0.106

June 14th, 2004 0.091 0.037 0.083

Page 235: 7 Tesis PPP Kongzhe Chen

211

6.3.2 Analysis of Receiver Clock Offset Estimation

In the test above, the float ambiguities and the station coordinates were estimated as

constants. On the other hand, ZHD was modeled and ZWD was estimated along with its

horizontal gradients. The ZTD presented in Figures 6.19, 6.20 and 6.21 is the sum of the

modeled ZHD and the estimated ZWD. A negative correlation between errors in ZWD

estimates and errors in clock offset estimates has been investigated by various researchers.

Hackman and Levine (2003) illustrated the following relationship between them:

wetclk d.d 43−= (6.1)

where clkd and wetd are the errors in clock offset estimate and ZWD estimate, respectively.

The errors in receiver clock offset estimates and the errors in ZWD estimates in the test

above should follow the similar relationship. However, in the test, no precise ZWD

measurements were available for stations AMC2 to serve as references. Therefore, the errors

in ZWD estimates could not be calculated to show the correlation between errors in clock

offset estimates and ZWD estimates. IGS tropospheric products, which has been claimed to

be accurate up to 4 mm for IGS stations (IGS website, 2003), only provide ZTD for IGS

stations. The sample interval of the products is 2 hours, which is too long to demonstrate the

correlations. Instead, in Figures 6.19, 6.20 and 6.21, the ZTD estimates were presented. The

negative correlation between the ZTD and the receiver clock offset estimates is still obvious,

especially before the ambiguity convergence.

Page 236: 7 Tesis PPP Kongzhe Chen

212

Due to the correlation between the ZWD and the receiver clock offset estimation, the error

sources affecting the ZWD estimation will also corrupt the clock solution. The error sources

include errors in orbits, clocks, multipath and the higher-order ionospheric effects. They will

be analyzed in Section 6.4 and they would introduce about triple errors in the receiver clock

offset estimates than in the ZWD estimates.

Still, in order to use the estimates for timing, all instrumental biases should be calibrated to

relate the internal clock to the external hardware clock driving the receiver (Petit et al.,

2001). Special cables that are less temperature sensitive may also be required (Larson et al.,

2000).

6.4 Atmospheric Sensing Using PPP Technique

Atmospheric sensing with GPS, as discussed in Section 2.4, is equivalent to ZWD or ZTD

estimation, given that precise pressure measurements accurate up to the sub-mbar level are

available. To assess the accuracy of atmospheric sensing with PPP method, accurate values

of ZTD, ZWD or PWV are required for using as the ground-truth. IGS has provided high

accuracy ZTD combinations for IGS stations since 1998 (Gendt, 1998). The current accuracy

of this combination is given to be at the 4 mm level (IGS website, 2004), which is

sufficiently good enough for the products to be used as the ground-truth.

Page 237: 7 Tesis PPP Kongzhe Chen

213

Radiometers, which can be calibrated to provide ZWD at about 1.8 mm accuracy or PWV at

0.3 mm accuracy (Rocken et al., 1993), are usually used to provide reference measurements

to assess the accuracy of ZWD or PWV estimates from GPS.

6.4.1 Comparison with IGS Final Tropospheric Products

In this test, three types of precise orbit and clock products with different accuracies and

latencies were used to assess the obtainable accuracy of ZTD estimates. They are the IGS

Final, JPL NRT and JPL IGDG products. JPL IGDG orbit and clock products were tested

with simulated real-time processing, while the other two were used in post-mission tests.

Raw data from GPS week 1251 for IGS station AMC2 and the IGS Final products were

downloaded from the IGS website. The JPL NRT products for the same period were

downloaded from the JPL website. IGDG real-time orbit and clock products were received

over the Internet. P3-RT was used to process the GPS raw data with different precise

products. The real-time IGDG products were used in a simulated real-time processing.

Though the IGS Final products are not available in real-time, the products were used to

process the data from AMC2 for comparison. Only observations at 5-minute interval from

AMC2 were processed to avoid clock interpolation. The sample interval of IGS Final

tropospheric delay combinations is 2 hours, so only the ZTD estimates at those epochs were

used for comparison. There were a total of 84 such epochs and the results are shown in

Figure 6.22. A summary of the results obtained using different products is provided in Table

6.5.

Page 238: 7 Tesis PPP Kongzhe Chen

214

Figure 6.22 ZTD Estimates Compared with IGS Tropospheric Products

Table 6.5 ZTD Estimation Statistics

Products RMS (mm) BIAS (mm) STD (mm)

IGS Final 4.1 -2.0 3.6

JPL IGDG 5.2 -2.2 4.7

JPL NRT 5.9 0.1 5.9

The statistics shows that PPP method has the potential to provide ZTD estimates at an

accuracy of about 5~6 mm in real-time. As shown in Figure 6.22, almost all the differences

when using IGS Final products are less than 10 mm while the differences can be more than

15 mm when using JPL IGDG products or NRT products. The processing results using the

IGS Final products provided the best results, but the IGS Final products are not available in

real-time, as it has a latency of 13 days. The -2 mm biases when using IGS Final and JPL

IGDG may be caused by the use of different processing strategies in the ZTD estimation. The

ZTDs of IGS stations are estimated as a total delay using a single hydrostatic mapping

Page 239: 7 Tesis PPP Kongzhe Chen

215

function in some analysis centers (Heroux, 2003). In this test, the ZHD was modeled and

mapped with the hydrostatic mapping function and pressure model, and ZWD, including the

un-modeled ZHD, is estimated with the wet mapping function. The output ZTD is the sum of

the modeled ZHD and the estimated ZWD. The effects of different mapping functions for the

zenith delay estimation will be investigated in Section 6.4.2. The spikes in the results of JPL

IGDG products and NRT products may be caused by the errors in orbit and clock products.

The statistics in Table 6.5 show that the quality of the satellite orbit and clock products

directly affects the obtainable accuracy of ZTD estimates.

Because ZHD can be determined to better than 1 mm given accurate surface pressure (Bevis

et al., 1992), it is possible for PPP method to produce real-time ZWD at an accuracy of 6~7

mm if precise pressure measurements can be obtained in real-time. The vertical integrated

water vapour overlaying a receiver, in terms of Precipitable Water Vapour (PWV), can be

related to the ZWD at the receiver using Equation 2.6. The dimensionless constant of the

transfer factor Π is approximately equal to 0.15. Therefore, PPP has the potential to

determine real-time PWV to an accuracy of 1 mm to satisfy the required accuracy for GPS

meteorological applications (Gutman and Benjamin 2001).

6.4.2 Comparison with Radiometer Measurements

From July to September in 2004, a real-time water vapour sensing test was conducted in the

University of Calgary. One week’s results, from September 2 to 8 will be discussed in this

section.

Page 240: 7 Tesis PPP Kongzhe Chen

216

A Radiometrics 1100 WVR, which has been set up on the roof of the Engineering

Building, was used to derive the “true” PWV values in this test to assess the PWV estimates

derived from GPS observations using PPP methodology. The radiometer was set up to make

direct measurements of line-of-sight slant water vapour to all GPS satellites during the test

period. The WVR tracks each satellite for approximately 40 s. Consequently, it takes about 6

minutes to track all satellites in view in a given cycle (Gao et al., 2004). In order to obtain

optimal accuracy for the hydrostatic delay values, precise pressure sensors are required at the

observation site. A Paroscientific MET3A sensor was set up beside the radiometer. The

MET3A collected meteorological data continuously during the test period – measuring

pressure, temperature and humidity at 30 s interval. The pressure measurements were

interpolated to the sample interval (10 s) of GPS measurements to calculate the zenith

hydrostatic delay. The accuracy of the pressure observations is better than 0.1 mbar for this

instrument (Nicholson, 2004), so that the corresponding errors in the hydrostatic delay values

are considered negligible. The temperature measurements of MET3A, which were used to

calculate the transfer factor Π , are accurate up to 0.1 degree (Nicholson, 2004).

A Javad Legacy dual-frequency receiver was used during the test period to output GPS

measurements at an interval of 10 s. A Javad JPSLEGANT antenna was set up on S1 pillar

(51°04’ N, 114°07’ W, 1116.82 m) on the roof beside the radiometer (about 1.5 m away).

JPSLEGANT is an antenna with a flat ground plane so it can partially mitigate the multipath

effects. The MET3A is located about 2 m above the GPS antenna, which was set up at almost

the same height as the radiometer. Assuming a pressure scale height of 8 km, a 2 m height

Page 241: 7 Tesis PPP Kongzhe Chen

217

difference will lead to about 0.5 mm ZHD difference. The pressure measurements of the

MET3A were corrected to the height of GPS antenna in post-mission to isolate the ZHD. The

setup of the radiometer, JPSLEGANT antenna and MET3A sensor is shown in Figure 6.23.

Figure 6.23 Radiometer, GPS Antenna and MET3A Instruments

In some researches, ZWD and ZHD were estimated as a total ZTD using a single mapping

function, which can be either the hydrostatic mapping function or wet mapping function

(Duan et al., 1996). This method is practical when no gradient components were estimated

and the elevation cut-off angle is as big as 15°, since the difference between the hydrostatic

and wet mapping functions at 15° is ~0.03 (Niell, 1996). In this research, GPS data at low

elevation angles was included to separate gradient components from the azimuthally

homogeneous components. The difference between the hydrostatic and wet mapping

Radiometer

GPS Antenna

MET3A

Page 242: 7 Tesis PPP Kongzhe Chen

218

functions can be up to 0.26 at 7°, which is the elevation cut-off angle used in this research.

Estimating only a total delay using a single mapping function will therefore lead to a bias in

zenith tropospheric delay estimates. Figure 6.24 shows the Niell Mapping Functions for the

testing site during the testing period, where the hydrostatic and wet mapping functions are

denoted as NMFh and NMFw, respectively. The difference between the hydrostatic and wet

mapping functions is shown in Figure 6.25.

Figure 6.24 Niell Mapping Functions

Page 243: 7 Tesis PPP Kongzhe Chen

219

Figure 6.25 Difference between the Wet and Hydrostatic Mapping Functions

Therefore, in this test, ZHD was modeled and mapped using the hydrostatic mapping

function, while ZWD was estimated using the wet mapping function. However, precise

pressure measurements were only available with a latency of several hours. So in real-time

processing a pressure model, the model based on Equation 6.2, has been used to calculate

pressure values. The un-modeled ZHD was absorbed to ZWD estimates. The ZWD plus the

un-modeled ZHD was then estimated using wet mapping functions. The modeled ZHD and

estimated ZWD (including the un-modeled ZHD) were output as a total ZTD from the real-

time processing. The precise pressure measurements were then used to isolate the hydrostatic

delays from the ZTD in post-mission before they were transferred to PWV and compared

with the measurements from WVR.

Page 244: 7 Tesis PPP Kongzhe Chen

220

During the test, the RMS difference between the calculated pressure value and the precise

pressure measurements from MET3A is just 5.92 mbar, which corresponds to about 13 mm

ZHD. Estimating such a small amount of un-modeled ZHD with the wet mapping function

did not greatly affect the total ZTD. The contribution of the small un-modeled ZHD to the

errors in the PWV estimates will be discussed in Section 6.4.3. The difference between the

calculated pressure value and the precise pressure measurements from MET3A is shown in

Figure 6.26, where the pressure value is calculated using the following formula by assuming

a pressure scale height of 8 km.

( )80000

/hePP −×= (6.2)

where 0P is the pressure at sea level and 1013.25 mbar was selected in this research, h is the

site height in metre and 1118.62 m, which is the height of MET3A, was used to calculate the

pressure values in Figure 6.26 and compare with MET3A measurements, and 1116.82 m,

which is the height of the GPS antenna, was used in the real-time processing.

Page 245: 7 Tesis PPP Kongzhe Chen

221

Figure 6.26 Difference between the Calculated and Measured Pressures

During September 2-8, 2004, JPL IGDG real-time precise orbit and clock products were

acquired over the Internet from a JPL server at a rate of 1 Hz. Real-time GPS observations

were output at interval of 10 s from the Javad receiver. The PPP numerical computation was

conducted using the software package P3-RT installed in a computer at the office ENF405

with an Internet connection and a serial port connected to the GPS receiver. ZTDs were

output in real-time.

After removed the zenith hydrostatic delays, which were calculated using the Saastamoinen

model and precise pressure measurements from MET3A that have been converted to the

height of radiometer and GPS antenna, the remaining ZWD were converted to PWV and

Page 246: 7 Tesis PPP Kongzhe Chen

222

compared with water vapour measurements from the WVR. Equation 2.8 and 2.10 given by

Bevis et al. (1994) were used to calculate the transfer factor Π . The surface temperatures

required in these equations were obtained from the MET3A measurements.

The slant water vapour measurements to each satellite from the WVR were mapped to zenith

water vapour using the wet Niell Mapping Function. The zenith water vapour measurements

for each cycle of observations were averaged and then compared with the average value of

GPS-derived zenith water vapour estimates over the same time period, about 6 minutes per

cycle. The averaged PWV measurements from the radiometer, the averaged GPS-derived

PWV, and the differences between them are shown in Figures 6.27 to 6.33. To demonstrate

the correlation between days, the differences between GPS-derived and radiometer-measured

PWV are also presented in Figure 6.34, from top to bottom in the order of September 2, 3, 4,

5, 6, 7, and 8. The results are given in local time series and are offset 5 mm between days.

For each day, the first one-hour ZWD estimates were used only for the ambiguity

convergence. Because of an Internet connection problem during the test period, a 4-hour

period of orbit and clock products were lost on September 2, 2004. The statistics of the

results are shown in Table 6.6.

It can be observed that the PWV difference between the WVR measurements and GPS

estimates is about 1 mm, with very small (less than 0.3 mm) random biases over the seven

days. The results indicate the potential to determine PWV to an accuracy of 1 mm in real-

time using precise orbit and clock products and PPP methodology.

Page 247: 7 Tesis PPP Kongzhe Chen

223

Figure 6.27 PWV from GPS and WVR on September 2nd, 2004

Figure 6.28 PWV from GPS and WVR on September 3rd, 2004

Figure 6.29 PWV from GPS and WVR on September 4th, 2004

Page 248: 7 Tesis PPP Kongzhe Chen

224

Figure 6.30 PWV from GPS and WVR on September 5th, 2004

Figure 6.31 PWV from GPS and WVR on September 6th, 2004

Figure 6.32 PWV from GPS and WVR on September 7th, 2004

Page 249: 7 Tesis PPP Kongzhe Chen

225

Figure 6.33 PWV from GPS and WVR on September 8th, 2004

Figure 6.34 PWV Comparison between GPS and WVR from September 2nd to 8th, 2004

Page 250: 7 Tesis PPP Kongzhe Chen

226

Table 6.6 Statistics of PWV Comparison

Date RMS (mm) BIAS (mm) STD (mm)

Sept. 2nd, 2004 0.93 0.04 0.92

Sept. 3rd, 2004 0.99 0.08 0.99

Sept. 4th, 2004 1.09 0.30 1.05

Sept. 5th, 2004 0.77 -0.28 0.72

Sept. 6th, 2004 0.65 0.01 0.65

Sept. 7th, 2004 0.64 0.12 0.62

Sept. 8th, 2004 0.94 -0.21 0.91

To demonstrate the importance of mapping functions in zenith tropospheric delay estimation,

the GPS data was also processed in post-mission with different estimation strategies. The

first method is to model ZHD with a hydrostatic mapping function and the precise pressure

measurements from MET3A, and to estimate ZWD with a wet mapping function. Only ZWD

was output and converted to PWV and compared with the radiometer measurements. Two

other methods are to estimate ZTD with a single mapping function, wet or hydrostatic

mapping function. Still, ZHD calculated from the precise pressure measurements was

removed from the estimated ZTD. The remaining ZWD was thus converted to PWV. Table

6.7 shows the accuracy statistics of different methods when compared with radiometer

measurements.

Page 251: 7 Tesis PPP Kongzhe Chen

227

Table 6.7 Accuracy Statistics of Different Strategies (Unit: mm)

Estimate ZWD Estimate ZTD with NMFh Estimate ZTD with NMFw Day

RMS BIAS STD RMS BIAS STD RMS BIAS STD

2 0.90 0.04 0.90 1.10 0.64 0.89 14.90 -14.76 2.10

3 1.02 0.13 1.01 1.36 0.80 1.10 14.37 -14.23 1.97

4 1.02 0.16 1.01 1.37 0.94 0.99 14.88 -14.79 1.71

5 0.70 -0.11 0.69 0.74 0.18 0.71 14.31 -14.23 1.51

6 0.70 0.22 0.66 1.02 0.69 0.76 13.65 -13.58 1.42

7 0.63 0.20 0.60 0.96 0.67 0.69 13.16 -13.05 1.69

8 0.87 -0.22 0.84 1.04 0.38 0.96 14.61 -14.49 1.83

As shown in Table 6.7, estimating ZWD with a wet mapping function and modeling ZHD

with a hydrostatic mapping function and pressure measurements provides the best results.

Though they are slightly worse than the real-time results in September 6 and 8, the results

overall are better than the real-time ones. Estimating ZTD with a hydrostatic mapping

function also provides millimetre level accuracy, but the biases are positive in all days,

compared with the random biases when estimating only ZWD in real-time or post-mission

processing. The biases are strongly correlated with the PWV values. If we use the results

from the first method as the reference, the biases will be about 0.5~0.6 mm in September 2,

3, 4, 7 and 8, and about 0.3~0.4 mm in September 5, 6. The PWVs range between 10 and 18

mm in September 2, 3, 4, 7 and 8, and between 7 and 13 mm in September 5 and 6.

Therefore, mapping ZWD with a hydrostatic mapping function produces a bias of about 3%

to 4% of the ZWD value in magnitude. The bias is consistent with the difference between the

hydrostatic and wet mapping functions at the elevation cut-off angle, which is 7° in this test.

The difference between the hydrostatic and wet mapping functions is about 0.26 at the

Page 252: 7 Tesis PPP Kongzhe Chen

228

elevation angle of 7°, which is about 3.3% of the wet mapping function at this elevation

(~7.92).

On the other hand, estimating ZTD with a wet mapping function would not provide useful

information to PWV estimation because the ZTD is dominated by ZHD. The bias introduced

by mapping ZHD with a wet mapping function is also about 3% to 4% of the ZHD value in

magnitude (about 2 m ZHD during the test). The difference between the hydrostatic and wet

mapping functions is about 3.5% of the hydrostatic mapping function at the elevation angle

of 7° (~7.65).

Therefore, the relative bias introduced by mapping ZWD using a hydrostatic mapping

function or mapping ZHD using a wet mapping function is determined by the relative

difference between the mapping functions at the elevation cut-off angle.

Based on the analysis above, the best way for PPP meteorology is to model and map ZHD

with precise pressure measurements and the hydrostatic mapping function, and estimating

ZWD with the wet mapping function. If pressure measurements are not available during the

data processing, a pressure model can be used to calculate an approximate pressure value as

was adopted by the real-time processing in this research. If even a pressure model is not

practicable, ZTD should be estimated with a hydrostatic mapping function. Estimating ZTD

with a wet mapping function should not be used in any case.

Page 253: 7 Tesis PPP Kongzhe Chen

229

6.4.3 Analysis of Real-Time Water Vapour Sensing Results

The 1 mm level RMS of GPS-WVR precipitable water vapour difference can be mainly

attributed to errors in GPS zenith wet delay estimates, the transfer factor Π (as calculated

from surface temperature measurements) and the WVR retrieval coefficients and/or

calibration errors. Details of these potential error sources are described in the following.

Errors in GPS Orbits and Clocks

Errors in the orbit and clock products will directly affect the zenith wet delay estimation. As

shown in Section 6.4.1, the difference between ZTD estimates using the IGS Final products

and the IGDG products can be up to more than 1 mm (4.1 for the IGS Final and 5.2 for the

IGDG). If one assumes that the errors in orbits and clocks are not correlated with other

errors, the ZTD errors introduced by the IGDG orbit and clock products would be about 3

mm, which corresponds to 0.45 mm PWV.

The contribution of errors in orbits and clocks can also be analyzed in another way. As

investigated in Section 6.2, IGDG real-time orbit and clock products can provide sub-

centimetre level positioning accuracy in static mode. The zenith tropospheric error caused by

the errors in orbits and clocks should be less than 3.3 mm considering that the positioning

errors in height component are typically 3 times as large as the zenith tropospheric errors

(Rocken et al., 1993). Therefore, their contribution to PWV errors should be less than 0.5

mm.

Page 254: 7 Tesis PPP Kongzhe Chen

230

In summary, the contribution of errors in orbits and clocks should be in the range of 0.4 to

0.5 mm. They are the main error sources in the PWV estimation.

Errors in Zenith Hydrostatic Delay

In this research, the pressure measurements from the MET3A barometer were used along

with Saastamoinen model to calculate and remove the ZHD from the ZTD output from GPS

data processing. Since the accuracy of the pressure observations is better than 0.1 mbar in

this research, even after interpolation, the corresponding errors in hydrostatic delay values

therefore are considered negligible (Gao et al., 2004). In real-time processing, a pressure

model was used to calculate pressure value, which led to approximately 13 mm ZHD being

combined into the ZWD and mapped with a wet mapping function. As discussed above,

when an elevation cut-off angle of 7° is used, mapping ZWD using a hydrostatic mapping

function or mapping ZHD with a wet mapping will introduce an error of about 3.5% of the

mapped parameter value in magnitude. In this case, the error would be approximate 0.46 mm

in ZWD, equivalent to 0.07 mm in PWV. The total contribution to PWV errors should be less

than 0.1 mm.

Multipath Effects

The test site was located at an environment with significant multipath effects as previously

discussed. Even though an antenna with a small ground plane was used, the residual

multipath effects seem evident. Similar diurnal variations in PWV errors can be found in

Page 255: 7 Tesis PPP Kongzhe Chen

231

Figure 6.34, especially during 12:00 to 14:00 local time. The big negative differences (with

respect to WVR truth) during this period could be caused by multipath effects.

Higher-order Ionospheric Effects

In this research, dual-frequency observations were used to form ionosphere-free

combinations which have neglected the higher-order ionospheric effects. The residual

second-order ionospheric effect is about 0.11 mm/TECU for code combinations and half of

this effect for phase combination (Bassiri and Hajj, 1993). During the test period, the vertical

TEC was less than 30 TECU, ranging between 5 to 15 TECU most of times. Considering that

ionospheric mapping function is smaller than tropospheric mapping function as shown in

Figure 5.5 and Figure 6.24, the higher-order ionospheric effect on the PWV estimation

should be at a 0.1 mm level. In post-processing, corrections for the second-order ionospheric

effects have been calculated using the equations discussed in Section 3.2. But they have not

shown any improvement over the ionosphere-free combinations, because the IGDG orbit and

clock products are estimated using ionosphere-free combinations without considering the

higher-order ionospheric effects (Muellerschoen, 2003).

Mapping Functions

Both wet and hydrostatic NMFs were not 100% perfect at very low elevations. When

compared with the radiosonde profiles, NMFh and NMFw have shown biases of –0.0011 and

–0.0179 respectively at 5° elevation (Niell, 1996). Considering big mapping functions at this

elevation (NMFh=10.1428, NMFw=10.7441), the relative errors in the mapping functions

Page 256: 7 Tesis PPP Kongzhe Chen

232

are approximately 0.01% and 0.17%. The errors are even smaller at 7° elevation. Therefore,

the contribution should be less than 0.1 mm for an elevation cut-off angle of 7°.

Errors in Transfer Factor Π

The relative error in Π is approximately equal to the relative error in mT (Bevis et al., 1994).

In this research, mT is calculated from surface temperature using the linear relation estimated

by Bevis et al. (1994) with an RMS relative error of less than 2%. This regression was

determined by investigating data in a 2-year period from 13 stations in the United States,

from Fairbanks, Alaska, to West Palm Beach, Florida. The linear relation was considered

suitable for the test site in Calgary, with a similar accuracy (Gao et al., 2004). From the

figures, it is observed that the PWV in Calgary during the test period is about 15 mm.

Therefore, the uncertainty in mT contributes to the PWV difference should be less than 0.3

mm.

WVR Retrieval Coefficients and Calibration

Biases may be introduced to the PWV measurements if the radiometer is not calibrated

carefully. Rocken et al. (1993) claimed that a radiometer could be calibrated to provide ZWD

at 1.8 mm accuracy or PWV at 0.3 mm accuracy. Because the retrieval coefficients for the

radiometer used in this research were derived from radiosonde data local to Boulder,

Colorado, an uncertainty of about 1 mm level for the radiometer PWV observations can be

assumed (Gao et al., 2004).

Page 257: 7 Tesis PPP Kongzhe Chen

233

6.5 PPP Using Single-Frequency Measurements

In this section, precise point positioning results using single-frequency measurements are

presented. All data was processed using the IGDG orbit and clock products either in real-

time or simulated real-time mode. As discussed in Chapter 5, the key issue for precise point

positioning using single-frequency measurements is how to mitigate the ionospheric effects.

Therefore the performance of single-frequency precise point positioning is correlated with

the ionospheric conditions, which is also regionally dependent (Skone, 1999; Komjathy,

1997). Data collected at different ionospheric conditions and from different ionospheric

regions was therefore processed in this research. Different ionospheric mitigation models

were used to investigate the performances of these models at different ionospheric

conditions.

6.5.1 Positioning at Mid-Latitude Stations

A real-time static test was carried on December 3rd, 2003. One Javad Legacy dual-frequency

receiver was set up on the S1 pillar on the roof of the Engineering Building at the University

of Calgary. Only measurements on 1L were used in processing using the ionospheric

estimation model proposed in Chapter 5. The IGDG orbit and clock products were received

over the Internet. The sample interval was set to 10 s. The coordinates of S1, which were

determined in ITRF-1993, were transferred to ITRF-2000 using coefficients provided by

McCarthy and Petit (2004) and used as the true coordinates. The positioning errors and

Page 258: 7 Tesis PPP Kongzhe Chen

234

zenith ionospheric delay estimates are shown in Figure 6.35 and 6.36. The accuracy statistics

is presented in Table 6.8.

Figure 6.35 Positioning Errors Using Ionospheric Estimation Model on December 3rd, 2003

Figure 6.36 Zenith Ionospheric Delay Estimates on December 3rd, 2003

Page 259: 7 Tesis PPP Kongzhe Chen

235

The date for the real-time test was selected randomly, but it turned out to be an ionospheric

quiet period. The Ap index is 4 on December 3rd, 2003. The half cosine shape of the zenith

ionospheric delay also indicates a quiet ionospheric condition. At the beginning of the

processing, an apriori value, which was calculated using the initial coordinates and the

Klobuchar model, was set as the initial zenith ionospheric delay. As shown in Figure 6.36,

the initial zenith ionospheric delay did not deviate from the true value noticeably. About 30-

minute was required for the zenith ionospheric delay and ambiguities to converge. A

positioning accuracy of about half metre in 3D was obtained in the real-time test for that day.

To test the performance of the ionospheric estimation model under ionospheric disturbed

conditions, a dataset of 24 hours was processed in a simulated real-time mode using the

saved IGDG orbit and clock products. The Ap index on July 27th, 2004 is 186, indicating a

day with an extremely disturbed ionosphere. The GPS data was collected on the S1 pillar

with a sample interval of 10 s to match the first test. The positioning errors, zenith

ionospheric delay estimates, and satellite geometry are shown in Figures 6.37, 6.38 and 6.39.

The accuracy statistics is presented in Table 6.8.

Page 260: 7 Tesis PPP Kongzhe Chen

236

Figure 6.37 Positioning Errors Using Ionospheric Estimation Model on July 27th, 2004

Figure 6.38 Zenith Ionospheric Delay Estimates on July 27th, 2004

Page 261: 7 Tesis PPP Kongzhe Chen

237

Figure 6.39 Satellite Geometry of S1 on July 27th, 2004

Table 6.8 Accuracy Statistics of Single-Frequency Point Positioning at S1 (Unit: m)

December 3rd, 2003 July 27th, 2004

Latitude 0.277 0.490

Longitude 0.223 0.434 RMS

Height 0.371 0.691

Latitude 0.027 -0.228

Longitude -0.180 -0.017 BIAS

Height 0.085 -0.070

Latitude 0.268 0.433

Longitude 0.195 0.434 STD

Height 0.361 0.688

As shown in Figure 6.39, many satellites lost lock during testing because of the disturbed

ionosphere, which could even lead to missing solutions in some epochs when the number of

satellites is less than 7. The positioning errors are not as smooth as those in the ionospheric

Page 262: 7 Tesis PPP Kongzhe Chen

238

quiet day. The estimates for the zenith ionospheric delay also do not have a half cosine-like

trend and are very noisy. Sub-metre level accuracy was obtained for each positioning

component. The accuracy is degraded by a factor of 2 when compared with the results

obtained on December 3rd, 2003 with an ionospheric quiet condition.

Therefore, in the middle latitude regions, sub-metre level positioning accuracy is obtainable

in real-time, using the ionospheric estimation model and real-time precise GPS orbit and

clock products.

6.5.2 Positioning Using Data from Different Ionospheric Regions

In this section, datasets collected at different ionospheric conditions from three stations

located at different ionospheric regions were processed using different models, including the

ionospheric estimation model, the Klobuchar model, and the Global Ionospheric Model

(GIM). The purpose of the tests is to compare the performance of these models in different

regions and under different ionospheric conditions. All datasets were processed in a

simulated real-time mode using the IGDG orbit and clock products acquired from the real-

time water vapour sensing test in August 2004. The IGDG orbit and clock products were

available continuously in real-time during that month except for several hours which were

lost because of an internal network problem on August 25th. The data that has been processed

includes five days with the most disturbed ionospheric conditions, five days with the quietest

ionospheric conditions, and one day with typical ionospheric conditions in August 2004. The

approximate coordinates of the three stations and the Ap indices of these days are shown in

Page 263: 7 Tesis PPP Kongzhe Chen

239

Table 6.9 and 6.10. The three stations are GLPS located at the equatorial region, the S1

located at the mid-latitude region, and FAIR located at the high-latitude region. The data

interval is 30 s for GLPS and FAIR and 10 s for S1. For the purpose of consistency, all

datasets were processed independently epoch by epoch at an interval of 30 s. August 19th was

selected instead August 25th as an ionospheric quiet day because part of the orbit and clock

products was lost in August 25th. August 14th was selected as a day with typical ionospheric

conditions. The ionospheric products required for processing using GIM were downloaded

from IGS website.

Table 6.9 Station Coordinates

GLPS S1 FAIR

Latitude -00° 44’ 35” 51° 04’ 46” 64° 58’ 41”

Longitude -90° 18’ 13” -114° 07’ 58” -147° 29’ 57” Height 1.8 m 1116.8 m 319.0 m

Table 6.10 Ap Indices in August 2004

Date 4 8 24 3 19 14 20 10 21 31 30

Ap index 3 3 3 3 4 7 15 16 16 28 42

Given in Tables 6.11, 6.12 and 6.13 are the accuracy statistics (RMS) using the ionospheric

estimation model for stations GLPS, S1 and FAIR in these 11 days. The tables are sorted in

ascending Ap order. From the tables, we can see the ionospheric estimation model can

provide metre level accuracy at all stations and for all testing days. In ionospheric quiet days

at mid- and high- latitude stations, half metre level accuracy was obtained.

Page 264: 7 Tesis PPP Kongzhe Chen

240

The performance of the ionospheric estimation model is strongly correlated with ionospheric

conditions. At each station, the accuracy is higher in ionospheric quiet days than in disturbed

days. Accuracy is best at mid-latitude, worst at equatorial regions and high-latitude is in

between, which is consistent with the ionospheric conditions in these regions. In the

equatorial region, the values of the peak electron density are the highest among the three

regions while the mid-latitude ionosphere is the least variable and undisturbed (Komjathy,

1997). The ionospheric activity is more complicated in high-latitude region than mid-latitude

region due to factors such as auroral activity and ionospheric trough (Skone, 1999;

Komjathy, 1997).

Table 6.11 Accuracy Statistics of Ionospheric Estimation Model for GLPS (Unit: m)

Day Ap Latitude Longitude Height

4 3 0.167 0.396 0.451

8 3 0.270 0.286 0.532

24 3 0.289 0.472 0.740

3 3 0.111 0.359 0.485

19 4 0.189 0.695 0.740

14 7 0.421 0.820 0.734

20 15 0.190 0.352 0.532

10 16 0.262 0.348 0.497

21 16 0.282 0.853 1.006

31 28 0.286 0.556 0.735

30 42 0.369 0.772 0.917

Page 265: 7 Tesis PPP Kongzhe Chen

241

Table 6.12 Accuracy Statistics of Ionospheric Estimation Model for S1 (Unit: m)

Day Ap Latitude Longitude Height

4 3 0.259 0.295 0.364

8 3 0.237 0.281 0.406

24 3 0.281 0.247 0.342

3 3 0.314 0.322 0.398

19 4 0.319 0.276 0.407

14 7 0.306 0.308 0.392

20 15 0.325 0.260 0.621

10 16 0.363 0.293 0.520

21 16 0.260 0.399 0.434

31 28 0.205 0.253 0.547

30 42 0.449 0.270 0.636

Table 6.13 Accuracy Statistics of Ionospheric Estimation Model for FAIR (Unit: m)

Day Ap Latitude Longitude Height

4 3 0.176 0.191 0.415

8 3 0.231 0.207 0.686

24 3 0.233 0.222 0.605

3 3 0.231 0.258 0.456

19 4 0.246 0.302 0.631

14 7 0.388 0.410 0.702

20 15 0.457 0.563 0.741

10 16 0.483 0.339 0.628

21 16 0.279 0.240 0.749

31 28 0.300 0.248 0.652

30 42 0.494 0.319 1.057

Shown in Tables 6.14 to 6.16 are the accuracy statistics (RMS) for the three stations using

the Klobuchar model over the 11 days. The results from the processing using only code

Page 266: 7 Tesis PPP Kongzhe Chen

242

observations and the results from the processing using both code and phase observations are

shown in order to demonstrate the improvement using phase observations.

Compared to the ionospheric estimation model, the accuracy from the Klobuchar model is

much worse; only 1 to 3 metres accuracy was obtained. But the Klobuchar model is not as

sensitive to ionospheric conditions as the ionospheric estimation model. Unlike the

ionospheric estimation model, the Klobuchar model provides the best results at the high-

latitude station FAIR, though it still provides better results at the mid-latitude station S1 than

the equatorial station GLPS. This may be due to the fact that the Klobuchar model just uses 8

coefficients to fit the ionospheric activity on a global scale. The coefficients perform better at

some regions, which may not be characterized with quieter ionospheric conditions, than other

areas.

Processing using both code and carrier phase observations does not guarantee better results

than the code only processing when using the Klobuchar model. This is especially true at the

equatorial station GLPS where the accuracy even degraded when processing using code and

phase observations as shown in Table 6.14. The Klobuchar model is not accurate enough to

exploit the higher accuracy of phase observations at equatorial stations. As shown in Tables

6.15 and 6.16, only slight improvements are found at mid- and high- latitude stations for

phase processing solutions.

Page 267: 7 Tesis PPP Kongzhe Chen

243

Table 6.14 Accuracy Statistics of Klobuchar Model for GLPS (Unit: m)

Code Code & Phase Day Ap

Latitude Longitude Height Latitude Longitude Height

4 3 0.511 0.653 1.566 0.477 1.551 1.927

8 3 0.607 0.373 1.257 0.701 0.894 2.836

24 3 1.317 0.500 1.496 1.483 1.084 1.818

3 3 0.514 0.436 1.587 0.757 0.712 1.154

19 4 1.172 0.613 1.749 0.906 1.519 2.069

14 7 1.171 0.761 1.223 1.058 1.186 2.254

20 15 0.834 0.606 1.669 0.756 1.093 1.890

10 16 0.689 0.328 1.215 0.989 1.183 2.338

21 16 1.212 0.627 1.184 1.022 1.141 1.772

31 28 0.882 0.775 2.578 1.158 0.980 2.187

30 42 0.869 0.541 2.936 0.745 1.290 3.225

Table 6.15 Accuracy Statistics of Klobuchar Model for S1 (Unit: m)

Code Code & Phase Day Ap

Latitude Longitude Height Latitude Longitude Height

4 3 0.465 0.288 1.257 0.703 0.516 1.095

8 3 0.480 0.308 1.311 0.764 0.761 1.609

24 3 0.530 0.322 2.066 0.523 0.534 1.352

3 3 0.460 0.291 0.976 0.335 0.477 0.669

19 4 0.518 0.333 2.367 0.527 0.588 2.097

14 7 0.440 0.326 1.923 0.536 0.627 0.874

20 15 0.488 0.330 2.603 0.734 0.470 1.287

10 16 0.693 0.353 1.380 0.412 0.572 0.838

21 16 0.584 0.323 2.361 0.607 0.574 1.071

31 28 0.578 0.348 1.180 0.422 0.411 1.009

30 42 0.536 0.368 1.069 0.498 0.400 0.927

Page 268: 7 Tesis PPP Kongzhe Chen

244

Table 6.16 Accuracy Statistics of Klobuchar Model for FAIR (Unit: m)

Code Code & Phase Day Ap

Latitude Longitude Height Latitude Longitude Height

4 3 0.409 0.238 0.844 0.253 0.187 0.865

8 3 0.693 0.452 0.818 0.388 0.236 0.885

24 3 0.788 0.478 1.653 0.499 0.465 0.987

3 3 0.358 0.298 0.850 0.357 0.290 0.820

19 4 0.891 0.552 1.803 0.462 0.362 1.147

14 7 1.131 0.659 1.432 0.647 0.362 1.313

20 15 0.790 0.499 2.047 0.482 0.280 0.768

10 16 0.643 0.558 0.927 0.432 0.367 0.740

21 16 0.841 0.538 1.549 0.388 0.295 1.074

31 28 0.317 0.256 0.929 0.358 0.307 0.824

30 42 0.530 0.445 0.891 0.652 0.548 1.335

Given in Tables 6.17 to 6.19 are the accuracy statistics (RMS) for the three stations using the

GIM over the 11 days. The results from the processing using only code observations and the

results from the processing using both code and phase observations are presented.

The GIM provides much better results at all stations than the Klobuchar model. It can even

provide slightly better results than the ionospheric estimation model at the high latitude

station, though the latter is much better at the equatorial station. The accuracy of the GIM is

metre or even sub-metre level at mid- and high-latitude stations but it is about 1.5 metres at

equatorial stations.

Like the ionospheric estimation model, the GIM provides the best results at mid-latitude

stations and performs the worst at equatorial stations. This is consistent with the ionospheric

Page 269: 7 Tesis PPP Kongzhe Chen

245

conditions and the performance of the IGS ionospheric products in these regions. The IGS

tracking network, which is used to create the ionospheric products, is unevenly distributed

(Fedrizzi et al., 2002). It is much denser in the mid-latitude region than in the equatorial

region (Komjathy, 1997). Also, the resolution of the IGS final ionospheric TEC grids, which

is 5 deg (longitude) x 2.5 deg (latitude) at a 2-hour’ interval, is not high enough to recover

TEC at any given location and time in the equatorial region. This is because the vertical TEC

can change up to 20-TECU within one hour or several degrees in this region.

Unlike the Klobuchar mode, the carrier phase observations can almost always improve the

ionosphere estimation accuracy when using the GIM because the IGS ionospheric products

can model ionospheric delay much better than the broadcast Klobuchar coefficients.

Table 6.17 Accuracy Statistics of GIM for GLPS (Unit: m)

Code Code & Phase Day Ap

Latitude Longitude Height Latitude Longitude Height

4 3 0.338 0.486 1.578 0.364 0.309 0.980

8 3 0.425 0.318 1.920 0.432 0.790 1.026

24 3 0.643 0.406 1.771 0.636 0.644 1.001

3 3 0.348 0.378 1.670 0.260 0.323 1.341

19 4 0.420 0.483 1.907 0.594 0.645 1.456

14 7 0.578 0.640 1.978 0.598 0.960 1.406

20 15 0.496 0.471 2.162 0.360 0.336 1.185

10 16 0.376 0.352 1.999 0.431 0.517 1.035

21 16 0.586 0.540 1.974 0.363 0.522 1.349

31 28 0.655 0.612 1.649 0.663 0.625 1.252

30 42 0.515 0.361 1.584 0.346 0.603 0.979

Page 270: 7 Tesis PPP Kongzhe Chen

246

Table 6.18 Accuracy Statistics of GIM for S1 (Unit: m)

Code Code & Phase Day Ap

Latitude Longitude Height Latitude Longitude Height

4 3 0.374 0.221 0.615 0.188 0.215 0.355

8 3 0.391 0.215 0.573 0.245 0.219 0.474

24 3 0.418 0.217 0.631 0.238 0.281 0.494

3 3 0.410 0.219 0.565 0.243 0.290 0.270

19 4 0.363 0.231 0.629 0.256 0.263 0.466

14 7 0.387 0.233 0.613 0.279 0.363 0.403

20 15 0.375 0.230 0.599 0.324 0.244 0.540

10 16 0.413 0.226 0.638 0.241 0.228 0.442

21 16 0.462 0.235 0.564 0.469 0.289 0.654

31 28 0.391 0.300 0.730 0.322 0.236 0.671

30 42 0.406 0.260 0.804 0.308 0.199 0.523

Table 6.19 Accuracy Statistics of GIM for FAIR (Unit: m)

Code Code & Phase Day Ap

Latitude Longitude Height Latitude Longitude Height

4 3 0.249 0.146 0.517 0.145 0.165 0.371

8 3 0.277 0.255 0.570 0.199 0.205 0.621

24 3 0.278 0.205 0.620 0.248 0.228 0.551

3 3 0.281 0.198 0.573 0.200 0.171 0.407

19 4 0.264 0.214 0.580 0.186 0.211 0.505

14 7 0.352 0.316 0.691 0.333 0.278 0.824

20 15 0.377 0.221 0.679 0.355 0.218 0.574

10 16 0.285 0.281 0.711 0.338 0.278 0.658

21 16 0.283 0.246 0.701 0.234 0.186 0.724

31 28 0.325 0.222 0.617 0.285 0.230 0.656

30 42 0.437 0.312 0.921 0.481 0.327 0.880

Page 271: 7 Tesis PPP Kongzhe Chen

247

6.5.3 Positioning Using Kinematic Datasets

To test the performance of these ionospheric models in pure kinematic positioning, the

airborne dataset, used for dual-frequency positioning in Section 6.2, was processed with the

IGDG orbit and clock products in a simulated real-time mode again. This time however, only

measurements on 1L were used. The double differencing ambiguity-fixed position solutions

served as the ground-truth. The Ap index of that day is 7.

Figures 6.40 to 6.44 show the positioning errors in each positioning component using the

three models. The accuracy statistics for all processing is shown in Table 6.20.

Figures 6.40, 6.42 and 6.44 present results from the processing using both code and phase

observations, while Figures 6.41 and 6.43 show the much noisier results from the code only

processing. It took approximately 10 to 20 minutes for the positioning errors to converge to a

sub-metre level with phase processing. When using the ionospheric estimation model and

GIM, the positioning errors remained within the sub-metre level after ambiguity

convergence. However the ambiguities may diverge when using the Klobuchar model. This

may be caused by the change over time of large ionosphere residuals, which are evident by

the varying bias in the height component in Figure 6.41 when processing using only code

measurements. When processing phase measurements using the Klobuchar model, the float

ambiguities can absorb part of the ionospheric residuals for short periods, but the varying

residuals would make the ambiguities diverge, leading to worse solutions. This can explain

Page 272: 7 Tesis PPP Kongzhe Chen

248

why processing phase measurements using the Klobuchar model does not assure better

results than processing code measurements, as indicated in Section 6.5.2.

As shown in Table 6.20, promising results were obtained using different models. About 20

cm accuracy was obtained in each positioning component when using the ionospheric

estimation model and GIM by phase processing. The accuracy is though much worse when

using the Klobuchar model, but a metre level accuracy was still obtained through phase

processing. When processing using only code measurements, the GIM and Klobuchar model

can only provide accuracies of about 0.7 m and 2.5 m respectively. The accuracy obtained

with the kinematic dataset is even better than the accuracy obtained with static datasets. This

is because the accuracy of single-frequency precise point positioning is correlated with not

only the ionospheric conditions but also the satellite geometry. The performance of these

models in kinematic positioning confirms that sub-meter level accuracy is obtainable using

single-frequency measurements.

Figure 6.40 Positioning Using Ionospheric Estimation Model

Page 273: 7 Tesis PPP Kongzhe Chen

249

Figure 6.41 Positioning Using Klobuchar Model with Code Measurements

Figure 6.42 Positioning Using Klobuchar Model with Code and Phase Measurements

Page 274: 7 Tesis PPP Kongzhe Chen

250

Figure 6.43 Positioning Using GIM with Code Measurements

Figure 6.44 Positioning Using GIM with Code and Phase Measurements

Page 275: 7 Tesis PPP Kongzhe Chen

251

Table 6.20 Single-Frequency Point Positioning with Airborne Dataset (Unit: m)

Klobuchar GIM Ionospheric

Estimation Code Code & Phase Code Code & Phase

Latitude 0.154 0.419 0.418 0.482 0.131

Longitude 0.271 0.322 0.347 0.222 0.243

Height 0.217 2.317 0.840 0.448 0.262

6.5.4 Summary

The following conclusions can be drawn from the analysis of the single-frequency precise

point positioning results:

The performance of all ionosphere models is correlated with the ionospheric conditions. For

each model, the positioning accuracy is higher on ionospheric quiet days than on disturbed

days, and performance is better at mid- and high-latitude stations than at equatorial stations.

At mid- and high-latitude stations, almost all models can provide the metre level accuracy on

ionospheric quiet days. At equatorial stations, even the best model can only provide about

one metre level accuracy during ionospheric disturbed periods.

Based on the same dataset, the ionospheric estimation model and GIM offer better

performance than the Klobuchar model. The ionospheric estimation model and GIM provide

comparable accuracy at mid-latitude stations. But the GIM is slightly more accurate at high-

latitude stations, while the ionospheric estimation model is much better at equatorial stations.

Page 276: 7 Tesis PPP Kongzhe Chen

252

Carrier phase observations can always improve positioning accuracy when using the GIM

and precise GPS orbit and clock products, but they can only slightly improve the positioning

accuracy at mid- and high-latitude stations when using the Klobuchar model.

The demonstrated positioning accuracy using the Klobuchar model and the ionospheric

estimation model is obtainable in real-time using IGDG real-time orbit and clock products.

The same positioning accuracy is only obtainable with latency of about 11 days when using

the GIM and IGS ionospheric products.

Page 277: 7 Tesis PPP Kongzhe Chen

253

CHAPTER 7

CONCLUSIONS AND RECOMMENDATIONS

Real-time position determination, receiver clock offset estimation and atmospheric sensing

using PPP methodology have been investigated in this thesis. From the research results

presented in this thesis, some conclusions and recommendations are summarized in the

following.

7.1 Conclusions

A number of error sources, which can be removed completely or mitigated partially by

double differencing techniques, should be taken into account in PPP to exploit this novel

technique for different applications.

Several types of real-time precise GPS orbit and clock products are now available from

different agencies, such as JPL and NRCan. Internet multicast and satellite broadcast are two

complementary approaches for real-time distribution of precise GPS orbit and clock

products. Users can choose either approach based on the reliability, timeliness, convenience,

and available devices that best suit their field of application.

Page 278: 7 Tesis PPP Kongzhe Chen

254

JPL IGDG real-time orbit and clock products can be received over the Internet with a latency

of about 4 s, which only adds an extrapolation error of less than 1 cm in real-time processing.

A real-time software package, P3-RT, has been developed for real-time precise point

positioning, timing and atmospheric sensing. The software can output position solutions,

receiver clock offset and ZTD estimates within 50 ms provided the precise GPS orbit and

clock products are accessible in real-time. The computational efficiency and implementation

flexibility of PPP make it a promising tool for real-time positioning, timing and atmospheric

sensing.

Sub-centimetre level static and sub-decimetre level kinematic positioning accuracy is

obtainable in real-time using dual-frequency code and phase measurements, and real-time

precise GPS orbit and clock products.

Among several types of real-time orbit and clock products tested in this research, IGDG

products from JPL provide the most consistent and accurate results in real-time; GPS•C

phase solution products, which are still at the development stage and have a latency of

several hours, provide comparable accuracy but with less consistency as the IGDG products.

On the other hand, GPS•C code solution products can provide sub-metre level positioning

accuracy in real-time with dual-frequency code measurements.

The convergence time of the float ambiguity in un-differenced phase measurement is affected

by a variety of factors, including satellite geometry, multipath effects, performance of precise

orbit and clock products, etc. Typical convergence time is about 20~30 minutes for the

Page 279: 7 Tesis PPP Kongzhe Chen

255

positioning errors to converge to the decimetre level. In a static mode, a longer time,

generally in the range of 30 minutes to 1 hour, is required for the positioning results to

converge to the centimetre level.

Using real-time precise GPS orbit and clock products and PPP methodology, precipitable

water vapour overlying the GPS receiver can be estimated in real-time with an accuracy of

about 1 mm, if precise pressure and temperature measurements are available in real-time.

Issues related to real-time atmospheric sensing have been investigated. Besides troposphere

horizontal gradients, elevation cut-off angle, antenna phase center offset and variations,

mapping functions also play an important role in the ZWD or PWV estimation. The best

strategy is to model and map the ZHD with precise meteorological measurements and

hydrostatic mapping function and estimate the ZWD with a wet mapping function. If precise

meteorological measurements are not available in real-time, a pressure model can be used

instead. Estimating the total ZTD using a single mapping function will lead to big bias in the

ZWD or PWV estimates, which is related to the relative difference between the hydrostatic

and wet mapping functions at the elevation cut-off angle.

Errors in real-time orbit and clock products contribute most to the errors in real-time

atmospheric sensing using PPP methodology. Other main error contributions come from the

multipath, the higher-order ionospheric effects and the transfer factor. The multipath effects

can be reduced by site and antenna selection. A more accurate transfer factor can be obtained

by an analysis on the local data. However, the higher-order ionospheric effects cannot be

Page 280: 7 Tesis PPP Kongzhe Chen

256

mitigated by users unless they can be taken into account in the real-time orbit and clock

products.

Receiver clock offset can be estimated in real-time to an accuracy of about 100-picosecond

with respect to the reference clock of the precise GPS orbit and clock products, therefore

holding great promise for real-time time transfer applications.

JPL IGDG real-time products provide a new method to recover UTC(USNO) at an accuracy

of a few nanoseconds in real-time using a single GPS receiver. Timing using PPP

methodology and IGDG products, which keeps the flexibility of the one-way time transfer,

can offer much better performance for recovering UTC(USNO) in real-time than the

traditional one-way time transfer method. The capability of this method to recover

UTC(USNO) is comparable to the common-view method which requires that one clock be

precisely linked to UTC(USNO).

Ionosphere horizontal gradients have been investigated at a mid-latitude station. An

ionospheric estimation model has been proposed to estimate the ionosphere horizontal

gradients along with the zenith delay. The new model can provide VTEC estimates with an

accuracy of 2 TECU when compared with the IGS Final ionospheric products.

The performance of single-frequency precise point positioning is correlated with the

ionospheric activity. Several models have been investigated which can mitigate ionospheric

effects at different levels. The ionospheric estimation model can offer sub-metre level

position solutions with single-frequency data in real-time. The performance of the

Page 281: 7 Tesis PPP Kongzhe Chen

257

ionospheric estimation model is comparable to the global ionospheric model using the IGS

ionospheric products which have a latency of 11 days.

The real-time tests which have been conducted with the P3-RT software for positioning using

single or dual-frequency GPS measurements and for atmospheric sensing have indicated the

potential of PPP for real-time positioning and meteorology applications.

7.2 Recommendations

Real-time precise GPS orbit and clock products are essential for the implementation of PPP

processing for real-time positioning, timing and meteorological applications. Currently, only

several types of precise GPS products can be provided to users in real-time. This is because a

real-time global tracking network is required to estimate high precision GPS orbits and

clocks. The denser the network, the more consistent and accurate the products can be.

Therefore, more agencies should be involved to provide real-time GPS data to processing

centers.

The 20 to 30 minutes convergence time of the float ambiguities will affect PPP for real-time

applications. Progresses have been made by researchers to improve the ambiguity

convergence. A convergence time of several minutes would be highly desired.

Higher-order ionospheric effects, which can be up to several centimetres in magnitude at

zenith during times of high TEC, are one of the major error sources for meteorological

parameter determination using GPS. Although algorithms have been developed to mitigate

Page 282: 7 Tesis PPP Kongzhe Chen

258

the higher-order ionospheric effects to the millimetre level, they would not benefit PPP users

unless the higher-order ionospheric effects have been taken into account in the data

processing for the orbit and clock products. Some agencies have plans to mitigate the higher-

order ionospheric effects in their products. For example, the IGS has considered mitigating

the second-order ionospheric effects during data processing (Dow, 2004).

The IGS ionospheric products, which are currently accurate up to 2-TECU at grid points, can

be used to produce sub-metre level positioning solutions for single-frequency GPS users.

However, the accuracy degrades under ionospheric disturbed conditions and in equatorial

regions. In addition, the products are only available with a delay of 11 days. As precise GPS

orbit and clock products are expected to be widely available in real-time in the near future,

precise ionospheric products with higher resolutions and a shorter latency will be in high

demand by single-frequency users.

The absolute zenith and slant tropospheric delays estimated from PPP data processing can be

used for troposphere tomographic modeling. Also the absolute zenith and slant ionospheric

delays obtained from PPP using un-differenced data can be used for ionosphere tomographic

modeling.

Real-time PPP processing using precise GPS orbit and clock products, precise pressure and

temperature measurements for real-time precipitable water vapour estimation is significant

and should be investigated in the future.

Page 283: 7 Tesis PPP Kongzhe Chen

259

REFERENCES

Alber, C.R., Rocken, W.C. and Braun, J. (2000), Inverting GPS double differences to obtain GPS single path phase delays, Geophysical Research Letters, Vol. 27, pp. 2661-2664.

Allan, D. and Weiss, M. M. (1980), Accurate time and frequency transfer during common-view of a GPS satellite, Proceedings of 34th Annual Frequency Control Symposium, pp. 334-336.

Allan, D.W., Ashby, N. and Hodge C.C. (1997), The Science of Timekeeping, Hewlett Packard Application Note 1289.

Anderle, R.J. (1976), Point Positioning Concept Using Precise Ephemeris, International Geodetic Symposium on Satellite Doppler Positioning, Las Cruces, N. M.

Armatys, M. (2002), Real-Time Global Differential GPS Correction Message User Guide.

Armatys, M., Muellerschoen, R., Bar-Sever, Y. and Meyer, R. (2003), Demonstration of Decimetre-level Real-time Positioning of an Airborne Platform, Proceedings of ION National Technical Meeting 2003, Anaheim, California, January 22-24, 2003.

Ashby, N. and Spilker JR., J.J. (1996), Introduction to Relativistic Effects. In Parkinson & Spilker, Jr. (Eds.), Global Positioning System: Theory and Applications Volume I. Progress in Astronautics and Aeronautics, Volume 163, American Institute of Aeronautics and Astronautics, Inc.

Bar-Sever, Y.E., and Kroger, P.M. (1996), Strategies for GPS-based estimates of troposphere delay, Proceedings of ION GPS -96, Kansas, Mo., 17–20 Sept. 1996.

Bar-Sever, Y.E., Kroger, P.M. and Borjesson, J.A. (1998), Estimating horizontal gradients of tropospheric path delay with a single GPS receiver. Journal of Geophysical Research, Vol. 103, No. B3, pp. 5019-5035, 1998.

Bar-Sever, Y.E. And J. Dow (2002), Position Paper for the Real Time Applications and Products Session. IGS Towards Real-Time Network, Data, Analysis Center 2002 Workshop, Ottawa, Canada, April 8-11, 2002.

Page 284: 7 Tesis PPP Kongzhe Chen

260

Bar-Sever, Y.E. (2004), IGS Tropospheric Products and Services at a Crossroad, IGS Analysis Center Workshop 2004, March 1-5, 2004, Berne, Switzerland.

Bassiri, S., and Hajj, G. A. (1993), Higher-order ionospheric effects on the global positioning systems observables and means of modeling them, Manuscripta Geodetica, vol. 18, pp. 280–289.

Blewitt, G. (1989), Carrier phase ambiguity resolution for the Global Positioning System applied to geodetic baselines up to 2000 km, Journal of Geophysical Research, 94(B8), pp. 10,187-10,203.

Beran, T., Kim, D. and Langley, R.B. (2003). High-precision single-frequency GPS point positioning, Proceedings of ION GPS 2003, Portland, Oregon, September 9-12.

Beutler, G., J. Kouba, and T. Springer (1995), Combining the Orbits of the IGS Analysis Centers, Bull. Geod. 69, pp. 200-222.

Bevis M., Businger, S., Herring, T.A., Rocken, C., Anthes, R.A. and Ware, R.H. (1992), GPS Meteorology: Remote Sensing of Atmospheric Water Vapour Using the Global Positioning System, Journal of Geophysical Research, Vol. 97, No. D14, pp. 15,787-15,801, October 1992.

Bevis, M., Businger, S., Chiswell, S., Herring, T.A., Anthes, R.A., Rocken, C. and Ware, R.H. (1994), GPS Meteorology: Mapping zenith wet delays onto precipitable water, Journal of Applied Meteorology, Vol. 33, pp. 379-386.

Bisnath, S., Wells, D. and Dodd, D. (2003), Evaluation of Commercial Carrier Phase-Based WADGPS Services for Marine Applications, Proceedings of ION GPS 2003, 9-12 September, Portland, OR.

Bock, H., Beutler, G., Schaer, S., Springer, T. A. and Rothacher, M. (2000), Processing aspects related to permanent GPS arrays, Earth Planets Space, Vol. 52 No. 10, pp. 657-662.

Braasch, M.S. (1996), Multipath Effects. In Global Positioning System: Theory and Applications Volume I, Parkinson, B.W., Spilker, J.J., Eds. Washington: The American Institute of Aeronautics and Astronautics, Inc., 1996, pp. 547-568.

Brown, R. G. and Hwang, P. Y. C. (1992), Introduction to Random Signals and Applied Kalman Filtering, Second Edition, John Wiley & Sons, Inc.

Page 285: 7 Tesis PPP Kongzhe Chen

261

Businger, S., S. R. Chiswell, M. Bevis, J. Duan, R. A. Anthes, C. Rocken, R. H. Ware, M. Exner, T. VanHove, and F. S. Solheim (1996), The promise of GPS in atmospheric monitoring. Bull. Am. Meteorol. Soc., 77, 5-18.

Caissy, M., Héroux, P, Lahaye, F., MacLeod, K., Popelar, J., Blore, J., Decker, D. and Fong, R. (1996), Real-Time GPS Correction Service of the Canadian Active Control System, Proceedings of ION GPS-96, September, 1996.

CDGPS Receiver User's Guide (2003), Province of British Columbia Ministry of Sustainable Resource Management Base Mapping and Geomatic Services.

Chen, K., Gao, Y. and Shen, X. (2002), An Analysis of Single Point Positioning with Real-Time Internet-based Precise GPS Data. Proceedings of 2002 International Symposium on GPS/GNSS, November 6-8, 2002, Wuhan, China.

Chen, K. (2004), Real-Time Precise Point Positioning and Its Potential Applications. Proceedings of ION GNSS 2004, Long Beach, California, September 21-24, 2004.

Chen, G. and Herring, T.A. (1997), Effects of atmospheric azimuthal asymmetry on the analysis of space geodetic data, Journal of Geophysical Research, 102, No. B9, 20, 489–20,502, 1997.

Clynch, J.R. (2000), GPS Marine Position Improvement in the Post SA Era. Proceeding ION GPS 2000, 19-22 September, Salt Lake City, UT.

CODE DCB website (2004), CODE'S 30-DAY GPS P1-C1 DCB SOLUTION. Available: http://www.aiub.unibe.ch/ionosphere/p1c1.dcb.

CODE Ionosphere Map website (2004), Global Ionosphere Maps Produced by CODE. Available: http://www.aiub.unibe.ch/ionosphere.

Collins, P., Lahaye, F., Kouba, J. and Héroux, P. (2001), Real-Time WADGPS Corrections from Undifferenced Carrier Phase, Proceedings of ION-NTM-2001, January 22-24, Long Beach, California.

Collins, P., Mireault, Y. and Héroux, P. (2002), Strategies for Estimating Tropospheric Delays with GPS, Position Location and Navigation Symposium 2002, Palm Springs, 15-18 April 2002.

Collins, P. (2004), personal communication.

Page 286: 7 Tesis PPP Kongzhe Chen

262

Dai, L., Wang, J., Rizos, C. and Han, S. (2001), Real-time carrier phase ambiguity resolution for GPS/GLONASS reference station networks. Int. Symp. on Kinematic Systems in Geodesy, Geomatics & Navigation (KIS2001), Banff, Canada, 5-8 June, pp. 475-481.

Dach, R., Schildknecht, Th., Springer, T., Dudle, G. and Prost, L. (2002), Continuous time transfer using GPS carrier phase, IEEE Trans. Ultrason., Ferroelect., Freq. Control, Vol. 49, pp.1480–1490.

Davis, J. L., Herring, T. A., Shapiro, I. I., Rogers, A. E. E. and Elgered, G. (1985), Geodesy by Radio Interferometry: Effects of Atmospheric Modelling Errors on Estimates of Baseline Length, Radio Science, 20, No. 6, pp. 1593-1607.

Davis, J. L., Elgered, G., Niell, A. E. and Kuehn, C. E. (1993), Ground-based measurement of gradients in the “wet” radio refractivity of air, Radio Sci., Vol. 28, No. 6, pp. 1003-1018.

Defraigne, P. and Petit, G. (2003), Time transfer to TAI using geodetic receivers, Metrologia, 40,184-188.

Dodson A. H., Chen, W, Penna, N.T. and Baker, H.C. (2001). GPS Estimation of Atmospheric Water Vapour from a Moving Platform. Journal of Atmospheric and Solar-Terrestrial Physics, Vol. 63, pp.1331-1341.

Douša, J. (2001), The Impact of Ultra-Rapid Orbits on Precipitable Water Vapour Estimation using a Ground GPS Network, Phys. and Chem. of the Earth, Part A, 26/6-8, pp. 393-398, 2001.

Dow, J.M. (2004), [IGSMAIL-4963]: Recommendations from IGS 10th Anniversary Symposium and Workshop. Available: http://igscb.jpl.nasa.gov/mail/igsmail/2004/msg00187.html.

Duan J., Bevis, M., Fang, P., Bock, Y., Chiswell, S., Businger, S., Rocken, C., Solheim, F., van Hove, T., Ware, R.H., Mc-Clusky, S., Herring, T.A. and King, R.W. (1996), GPS Meteorology: Direct Estimation of the Absolute Value of Precipitable Water, J. of Appl. Meteorol., 35, 830-838, 1996.

Fang, P., Bevis, M., Bock, Y., Gutman, S. and Wolfe, D. (1998), GPS meteorology: Reducing systematic errors in geodetic estimates for zenith delay, Geophysical Research Letters, 25, 3583-3586, 1998.

Page 287: 7 Tesis PPP Kongzhe Chen

263

Fang, P., Gendt, G., Springer, T. and Mannucci, T. (2001). IGS near real-time products and their applications, GPS Solutions, 4(4), 2-8.

Fapojuwo, A. (2003), personal communication.

Fedrizzi, M., R. Langley, M. Santos, A. Komjathy, E. de Paula, and I. Kantor (2002), Mapping the Low-latitude Ionosphere with GPS. GPS World, Vol. 13, No. 2, pp. 41-47.

Feltens, J., and S. Schaer (1998), IGS Products for the Ionosphere, in Proceedings of the 1998 IGS Analysis Centers Workshop, ESOC, Darmstadt, Germany, February 9-11, 1998.

Francisco, S.G. (1996). GPS Operational Control Segment. In Global Positioning System: Theory and Applications Volume I, Parkinson, B.W., Spilker, J.J., Eds. Washington: The American Institute of Aeronautics and Astronautics, Inc., 1996, pp. 435-466.

Gabor, M.J. and Nerem, R.S. (2002). Satellite-Satellite Single-Difference Phase Bias Calibration As Applied to Ambiguity Resolution. Navigation, Journal of the Institute of Navigation, Vol. 49, No. 4, pp. 223-242.

Gail, W.B., Prag, A.B., Coco, D.S. and Coker, C. (1993). A Statistical Characterization of Local Mid-latitude Total Electron Content. Journal of Geophysical Research, Vol. 98, No. A9, pp. 15,717-15,727.

Gao, Y., F. Lahaye, P. Heroux, X. Liao, N. Beck and M. Olynik (2001). Modelling and Estimation of C1 - P1 Bias in GPS Receivers, Journal of Geodesy, Vol. 74, No.9., pp. 621-626.

Gao, Y. and Shen, X. (2002). A New Method for Carrier Phase Based Precise Point Positioning, Navigation, Journal of the Institute of Navigation, Vol. 49, No. 2.

Gao, Y., Skone, S., Chen, K., Hoyle, V. and Muellerschoen, R. (2004), Real-Time Sensing Atmospheric Water Vapour Using GPS Precise Orbit/Clock Products, Proceedings of ION GNSS 2004, Long Beach, California, September 21-24, 2004.

Gao, Y., Wojciechowski, A. and Chen, K. (2005), Airborne Kinematic Positioning Using Precise Point Positioning Methodology, Geomatica (accepted).

Page 288: 7 Tesis PPP Kongzhe Chen

264

Ge, M., Calais, E., and Haase, J. (2000), Reducing satellite orbit error effects in near real-time GPS zenith tropospheric delay estimation for meteorology, Geophysical Research Letters, 27, 2000.

Gelb, A. (1974), Applied Optimal Estimation, MIT Press, Cambridge, MA.

Gendt, G. (1998). IGS Combination of Tropospheric Estimates – Experience from Pilot Experiment, Proceedings of 1998 IGS Analysis Center Workshop, J.M. Dow, J. Kouba and T. Springer, Eds. IGS Central Bureau, Jet Propulsion Laboratory, Pasadena, CA, pp. 205-216.

Gendt, G.; Reigber, Ch.; Dick, G. (2001), Near real-time water vapour estimation in a German GPS network : first results from the ground program of the HGF GASP Project (COST-716 Workshop, Oslo, 10-12 July 2000), Physics and Chemistry of the Earth (A), 26, 6-8, 413-416.

Gendt, G., Dick, G., Reigber, C., Tomassini, C., Liu, Y. (2003), Demonstration of NRT GPS Water Vapour Monitoring for Numerical Weather Prediction in Germany, International Workshop on GPS Meteorology, Tsukuba, Japan, January 14-17, 2003.

Giffard, T (1999), Estimation of GPS Ionospheric Delay Using Ll Code and Carrier Phase Observables, 31st Annual Precise Time and Time Interval (PTTI) Meeting. December 7-9, 1999, Dana Point, California.

Giffard, T (2000), Recovering UTC (USN0,MC) with Increased Accuracy Using a Fixed, Ll-CA Code, GPS Receiver, 32nd Annual Precise Time and Time Interval (PTTI) Meeting. November 28-30, 2000, Reston, Virginia.

Gifford, A., Pace, S. and McNeff, J. (2000), One-Way GPS Time Transfer 2000, 33rd Annual Precise Time and Time Interval (PTTI) Meeting. November 28-30, 2000, Reston, Virginia.

Goad, C.C. (1998). Single-site GPS models. In GPS for Geodesy, Teunissen P.J.G and Kleusberg A. (Eds), Springer, 1998. pp. 437-456.

Gutman, S.I. and Benjamin, S.G. (2001), The Role of Ground-Based GPS Meteorological Observations in Numerical Weather Prediction, GPS Solutions, Volume 4, No. 4, pp. 16-24.

Page 289: 7 Tesis PPP Kongzhe Chen

265

Hackman, C and Levine, J. (2003), New Frequency Comparisons Using GPS Carrier-Phase Time Transfer, 2003 IEEE International Frequency Control Symposium & PDA Exhibition Jointly with the 17th European Frequency and Time Forum, May 5-8, 2003, Tampa, Florida, USA.

Hajj, G. A., Kursinski E. R., Romans L. J., Bertiger W. I. and Leroy S. S. (2002), A Technical Description of Atmospheric Sounding by GPS Occultations, J. of Atmosphere and Solar-Terrestrial Physics, 64, 451-469.

Haase, J., Ge, M., Vedel, H. and Calais, E. (2003), Accuracy and variability of GPS Tropospheric Delay Measurements of Water Vapour in the Western Mediterranean. J. of Appl. Meteorol., Vol. 42, No. 11. November 2003, pp. 1547-1568.

Han, S.C., Kwon, J.H. and Jekeli, C. (2001) Accurate absolute GPS positioning through satellite clock error estimation, Journal of Geodesy, Vol. 77, pp. 33-43.

Hatch, R. (1982). The Synergism of GPS Code and Carrier Measurements. Proceedings of Third International Geodetic Symposium on Satellite Doppler Positioning, Washington, D.C., pp. 1213–1232.

Heflin, M.B. (2000), [IGSMAIL-3082]: Near Real Time Products from JPL. Available: http://igscb.jpl.nasa.gov/mail/igsmail/2000/msg00429.html.

Heflin, M.B. (2004), personal communication.

Hernández-Pajares M., Juan, J.M. and Sanz, J. (1999), New approaches in global ionospheric determination using ground GPS data, Journal of Atmospheric and Solar Terrestrial Physics. Vol. 61, pp. 1237-1247.

Heroux, P., Caissy, M. and Gallace, J. (1993), Canadian active control system data acquisition and validation. Proceedings of the 1993 IGS Workshop. University of Berne, pp. 49-58.

Heroux, P. and Kouba, J. (1995). GPS precise point positioning with a difference, Geomatics '95, Ottawa, Canada.

Heroux, P. (2003), personal communication.

Heroux, P., Gao, Y., Kouba, J., Lahaye, F., Mireault, Y., Collins, P., Macleod, K., Tetreault, P. and Chen, K. (2004). Products and Applications for Precise Point Positioning -

Page 290: 7 Tesis PPP Kongzhe Chen

266

Moving Towards Real-Time. Proceedings of ION GNSS 2004, Long Beach, CA, September 21-24, 2004.

Hofmann-Wellenhof, B., Lichtenegger, H. and Collins, J. (2000). GPS Theory and Practice. Springer-Verlag, Wien New York, fifth, revised edition.

Huang, Y.N. (1997), Spatial correlation of the ionospheric total electron content at the equatorial anomaly crest. Journal of Geophysical Research. vol. 89, pp. 9823-9827

Hutsell, S., Forsyth, M. and McFarland, C. (2002), One-Way GPS Time Transfer: 2002 Performance, 34th Annual Precise Time and Time Interval (PTTI) Meeting. December 2-5, 2002, Reston, Virginia.

ICD-GPS-200-C. (2000). ARINC Research Corporation, CA, USA.

ICD-GPS•C (2001), NRCan, GPS·C Interface Control Document, Active Control System Section Geodetic Survey Division, Geomatics Canada, Natural Resources Canada.

IGS ACC website (2004), IGS Analysis Center Coordinator (ACC) at GFZ Potsdam. Available: http://www.gfz-potsdam.de/pb1/igsacc/index_igsacc.html.

IGS RTWG website (2004), Real Time Working Group. Available: http://igscb.jpl.nasa.gov/projects/rtwg/.

IGS website (2004), IGS Products. Available: http://igscb.jpl.nasa.gov/components/prods.html.

ISGI website (2004). Available: http://www.cetp.ipsl.fr/~isgi/homepag1.htm.

Jefferson, D., Heflin, M.B. and Muellerschoen, R.J. (2001), Examining the C1-P1 pseudorange bias. GPS Solutions, Vol. 4, No. 4, pp. 25–30.

JPL IGDG website (2004), IGDG (Internet-based Global Differential GPS). Available: http://gipsy.jpl.nasa.gov/igdg/index.html.

Kalman, R. E. (1960), A New Approach to Linear Filtering and Prediction Problems, Transaction of the ASME, Journal of Basic Engineering, pp. 35-45.

Page 291: 7 Tesis PPP Kongzhe Chen

267

Kassam, A. (2003), CDGPS Service Announcement. Available: http://www.cdgps.com/e/news.htm.

Kee C., Wide area Differential GPS. In Global Positioning System: Theory and Applications Volume II, Parkinson, B.W., Spilker, J.J., Eds. Washington: The American Institute of Aeronautics and Astronautics, Inc., 1996, pp. 81-114.

Klepczynski, W.J. (1996). GPS for Precise Time and Time Interval Measurement. In Global Positioning System: Theory and Applications Volume I, Parkinson, B.W., Spilker, J.J., Eds. Washington: The American Institute of Aeronautics and Astronautics, Inc., 1996, pp. 435-466.

Klobuchar, J.A. (1987), Ionospheric Time-Delay Algorithm for Single-Frequency GPS Users, IEEE Transaction on Aerospace and Electronic System, Vol. 23, pp. 325-331.

Klobuchar, J.A., Basu, S and Doherty, P. (1993), Potential Limitations in Making Absolute Ionospheric Measurements Using Dual Frequency Radio Waves From GPS Satellite, Proceedings of Ionospheric Effects Symposium, IES 93, pp. 187-194.

Klobuchar, J.A. (1996), Ionospheric effects on GPS, In Parkinson & Spilker, Jr. (Eds.), Global Positioning System: Theory and Applications Volume I. Progress in Astronautics and Aeronautics, Volume 163, American Institute of Aeronautics and Astronautics, Inc.

Komjathy, A. (1997), Global Ionospheric Total Electron Content Mapping Using the Global Positioning System, Ph.D. dissertation, Department of Geodesy and Geomatics Engineering Technical Report No. 188, University of New Brunswick, Fredericton, New Brunswick, Canada, 1997.

Kouba, J., and Héroux, P. (2001a). GPS Precise Point Positioning Using IGS Orbit Products, GPS Solutions, Vol.5, No.2, pp. 12-28.

Kouba, J. and Springer, T. (2001b), New IGS Station and Satellite Clock Combination, GPS Solutions, Vol. 4, No. 4, pp. 31–36.

Kouba, J., (2002), The GPS Toolbox ITRF Transformations, GPS Solutions, Vol. 5, No. 3, pp.88-90.

Kouba, J (2003), A Guide to Using International GPS Service (IGS) Products. IGS Central Bureau, Pasadena, CA: Jet Propulsion Laboratory.

Page 292: 7 Tesis PPP Kongzhe Chen

268

Kruse L.P., Sierk, B., Springer, T. and Cocard, M. (1999), GPS Meteorology: Impact of predicted orbits on precipitable water estimates, Geophysical Research Letters, Vol. 14, pp. 2045-2048, 1999.

Kuo, Y.H., Guo, Y.R. and Westwater, E. (1993), Assimilation of Precipitable Water Measurements into a Mesoscale Numerical Model, Mon. Wea. Rev., Vol. 121, pp. 1215-1238, 1993.

Lachapelle, G., Klukas, R., Qiu, W. and Melgard, T.E. (1994a) Single Point Satellite Navigation Accuracy - What The Future May Bring, Proceedings of the IEEE Position, Location and Navigation Symposium, Las Vegas, Nevada, April 11-15, 1994, pp. 16-22.

Lachapelle, G., R. Klukas, D. Roberts, W. Qiu and C. McMillan (1994b), One-Metre Level Kinematic Point Positioning Using Precise Orbits and Satellite Clock Corrections, Proceedings of ION GPS-94, September 20-23, 1994, Salt Palace Convention Center - Salt Lake City, UT.

Lachapelle, G., Cannon, M.E., Qiu, W. and Varner, C. (1996), Precise Aircraft Single Point Positioning Using Post-Mission Orbits and Satellite Clock Corrections, Journal of Geodesy, Vol. 70, pp. 562-571.

Lahaye, F., M. Caissy, J. Popelar, and R.J. Douglas (1998), Real-time GPS monitoring of atomic frequency standards in the Canadian Active Control System (CACS), Proc. 30th Precise Time and Time Interval Meeting, 187-199, 1998.

Landis, G. and White, J. (2002), Limitation of GPS Receiver Calibrations, 34th Annual Precise Time and Time Interval (PTTI) Meeting. December 2-5, 2002, Reston, Virginia.

Langley, R. B. (1997), The GPS Error Budget. GPS World, Vol. 8, No. 3, pp. 51-56.

Langley, R.B. (1998a). Propagation of GPS Signals. In GPS for Geodesy, Teunissen P.J.G and Kleusberg A. (Eds), Springer, 1998. pp. 111-149.

Langley, R.B. (1998b). GPS Receivers and Observables. In GPS for Geodesy, Teunissen P.J.G and Kleusberg A. (Eds), Springer, 1998. pp. 151-185.

Larson, K. and Levine, J. (1999), Carrier-phase time transfer, IEEE Trans. Ultrason., Ferroelect., Freq. Control. Vol. 46, No. 4, pp. 1001–1012, July 1999.

Page 293: 7 Tesis PPP Kongzhe Chen

269

Larson, K., Levine, J., Nelson, L. and Parker, T. (2000), Assessment of GPS carrier-phase stability for time-transfer applications, IEEE Trans. Ultrason., Ferroelect., Freq. Control. Vol. 47, pp. 484–494.

Leitinger, R. (1993), The effect of horizontal gradients of ionization on position determination and the availability of relevant data. In: Environmental Effects on Spacecraft Positioning and Trajectories, Proceedings of the Twentieth General Assembly of the International Union of Geodesy and Geophysics, Ed. A.V. Jones, Vienna. International Union of Geodesy and Geophysics and the American Geophysical Union, Washington, D.C., Geophysical Monograph 73, IUGG Volume 13, 39-46.

Leick, A. (2004), GPS Satellite Surveying, John Wiley & Sons, Inc., 3rd Edition.

Lewandowski W., Tisserand L. (2004), Determination of the differential time corrections for GPS time equipment located at the OP, PTB, AOS, KRISS, CRL, NIST, USNO and APL, Report BIPM-2004/06, pp 29.

Lombardi, M.A., Nelson, L.M., Novick, A.N. and Zhang, V.S. (2001), Time and Frequency Measurements Using the Global Positioning System, Cal. Lab. Int. J. Metrology, pp. 26-33, July-September 2001.

MacLeod, K. (2004), [IGS-RTWG-11]: Updated List of RTIGS Stations. Available: http://igscb.jpl.nasa.gov/mail/igs-rtwg/2004/msg00005.html.

Mader, G.L. (1999), GPS Antenna Calibration at the National Geodetic Survey, GPS Solutions, Vol. 3, No. 1, pp. 50-58.

Malys, S., Larezos, M., Gottschalk, S., Mobbs, S., Winn, B., Feess, W., Menn, M., Swift, E., Merrigan, M. and Mathon, W. (1997), The GPS Accuracy Improvement Initiative, Proceedings of ION GPS-97, September 1997.

Mannucci, A.J., B.D., Wilson and C.D., Edwards, A new method for monitoring the earth's ionospheric total electron content using the GPS global network, Proceedings of GPS-93, 1323-1332, September 1993.

Matsakis, D., Senior, K. and Breakiron, L. (1999), Analysis noise, short-baseline time transfer, and a long-baseline GPS carrier-phase frequency scale, 31st Annual Precise Time and Time Interval (PTTI) Meeting. December 7-9, 1999, Dana Point, California.

Page 294: 7 Tesis PPP Kongzhe Chen

270

McCarthy, D.D. and Petit, G. (2004), IERS Conventions 2003, IERS Technical Note 32.

Mendes, V.B. and Langley, R.B. (2000a). An analysis of high-accuracy tropospheric delay mapping functions, Physics and Chemistry of the Earth, Vol. 25, No. 12, pp. 809-812.

Mendes, V.B., G. Prates, L. Santos, and R.B. Langley (2000b), An Evaluation of the Accuracy of Models for the Determination of Mean Weighted Temperature of the Atmosphere. Proceedings of ION NTM 2000, January 26-28, 2000, Anaheim, CA. USA.

Montenbruck, O. (2003), Kinematic GPS positioning of LEO satellites using ionosphere-free single frequency measurements, Aerospace Science and Technology, Vol. 7, No. 5, pp. 396-405.

Muellerschoen, R.J., Bertiger, W.I., Lough, M., Stowers, D. and Dong, D. (2000), An Internet-Based Global Differential GPS System, Initial Results. Proceedings of ION National Technical Meeting 2000, Anaheim, CA, January 2000.

Muellerschoen, R.J., Reichert, A., Kuang, D., Heflin, M., Bertiger, W. and Bar-Sever, Y.E. (2001), Orbit Determination with NASA’s High Accuracy Real-Time Global Differential GPS System, Proceedings of ION GPS-2001, Salt Lake City, UT, September 2001.

Muellerschoen, R.J. (2003), personal communication.

Muellerschoen, R.J. and Caissy, M. (2004), Real-Time Data Flow and Product Generation for GNSS. IGS Analysis Center Workshop 2004, March 1-5, 2004, Berne, Switzerland.

Neilan R.E., Angelyn M., Springer, T., Kouba, J., Ray, J. and Reigber, C. (2000). International GPS Service 2000: Life without SA. Proceeding ION GPS 2000, 19-22 September, Salt Lake City, UT.

Nicholson, N., V. Hoyle, S. Skone, M.E. Cannon and G. Lachapelle (2003), 4-D Troposphere Modeling Using a Regional GPS Network in Southern Alberta, Proceedings of ION GPS 2003, 9-12 September, Portland, OR.

Nicholson, N.A. (2004), personal communication.

Niell, A.E. (1996), Global mapping functions for the atmosphere delay at radio wavelengths. Journal of Geophysical Research, Vol. 101, No. B2, pp. 3227-3246.

Page 295: 7 Tesis PPP Kongzhe Chen

271

Niell, A.E., A.J. Coster, F.S. Solheim, V.B. Mendes, P.C. Toor, R.B. Langley and C.A. Upham (2001), Comparison of measurements of atmospheric wet delay by radiosonde, water vapour radiometer, GPS and VLBI, Journal of Atmospheric and Oceanic Technology, Vol. 18, pp. 830-850, 2001.

Ohta, K. and M. Hayakawa (2000), Three-dimensional ray-tracing for very low latitude whistlers, taking into account the latitudinal and longitudinal gradients of ionosphere, Journal of Geophysical Research, Vol. 105, no. A8, pp. 18,895-18,900.

Ovstedal, O. (2002). Absolute Positioning with Single Frequency GPS Receivers, GPS Solutions, Vol. 5, No. 4, pp. 33-44.

Pacione, R., Vespe, F., Faccia, R. and Colucci, G. (2002), The Italian Near-Real Time GPS Fiducial Network for Meteorological Applications. IGS Towards Real-Time Network, Data, Analysis Center 2002 Workshop, Ottawa, Canada, April 8-11, 2002.

Parkinson, B.W. (1996), Introduction and Heritage of NAVSTAR, the Global Positioning System. In Global Positioning System: Theory and Applications Volume I, Parkinson, B.W., Spilker, J.J., Eds. Washington: The American Institute of Aeronautics and Astronautics, Inc., 1996, pp. 3-28.

Petit, G., Jiang, Z., White, J., Beard, R., and Powers, E. (2001), Absolute calibration of an Ashtech Z12-T GPS receiver. GPS Solutions, Vol. 4, No. 4, pp. 41–46.

Powers, Ed. (2002), [IGSMAIL-3992]: AMC2 station modifications. Available: http://igscb.jpl.nasa.gov/mail/igsmail/2002/msg00330.html.

Ray, J. and Petit, G. (1999a), IGS/BIPM pilot project to study time and frequency comparisons using GPS phase and code measurements, the 14th Meeting of the Consultative Committee for Time and Frequency, BIPM, Sevres, France. April 20-22, 1999.

Ray, J. (1999b), [IGSMAIL-2320]: Handling mixed receiver types. Available: http://igscb.jpl.nasa.gov/mail/igsmail/1999/msg00564.html.

Ray, J., Arias, F., Petit, G., Springer, T., Schildknecht, Th., Clarke, J. and Johansson, J. (2001), Progress in carrier phase time transfer, GPS Solutions, Vol. 4, No. 4, pp. 47–54.

Ray, J. and Senior, K. (2003), IGS/BIPM pilot project: GPS carrier phase for time/frequency transfer and timescale formation, Metrologia, 40(3), 2003, S270-S288.

Page 296: 7 Tesis PPP Kongzhe Chen

272

Ray, J. K. (2000), Mitigation of GPS Code and Carrier Phase Multipath Effects Using a Multi-Antenna System. Department of Geomatics Engineering, Ph.D. Thesis, The University of Calgary. UCGE Report No. 20136.

Reigber, Ch., Gendt, G., Dick, G. and Tomassini, M. (2002), Near-real-time water vapour monitoring for weather forecasts. GPS World, Vol 13, 2002. pp. 18-27.

Rivers, M. and Osborne, S. (1999), 1999 GPS Time Transfer Performance, 31st Annual Precise Time and Time Interval (PTTI) Meeting. December 7-9, 1999, Dana Point, California.

Rocken, C., Ware, R.H., van Hove, T., Solheim, F., Alber, C., Johnson, J., Bevis, M. and Businger, S. (1993), Sensing atmospheric water vapour with the Global Positioning System, Geophysical Research Letters, 20(23), 2631-2634, 1993.

Rocken, C., van Hove, T., Johnson, J., Solheim, F., Ware, R., Bevis, M., Chiswell, S. and Businger, S. (1995), GPS/STORM - GPS sensing of atmospheric water vapour for meteorology, J. Atm. Ocean. Tech., 12, 468-478.

Rocken, C., van Hove, T. and Ware, R. (1997), Near real-time GPS sensing of atmospheric water vapour, Geophysical Research Letters, Vol. 24, No 24, pp 3221-3224, 1997.

Rocken, C., Sokolovskiy, S., Johnson, J. and Hunt, D. (2001), Improved mapping of tropospheric delays, J. Atm. Ocean. Tech., Vol. 18, pp. 1205-1213, 2001.

Rocken, C., Braun, J., van Hove, T., Johnson, J. and Kuo, Y.H. (2003), Developments in ground-based GPS meteorology. International Workshop on GPS Meteorology, Tsukuba, Japan, January 14-17, 2003.

Rothacher, M., T.A. Springer, S. Schaer, G. Beutler (1997), Processing strategies for regional GPS Networks, IAG General Assembly, Rio, Brazil, September 2-9, 1997.

Rothacher, M and G. Mader (2002a), Receiver and Satellite Antenna Phase Center Offsets and Variations, Proceedings of IGS Network, Data and Analysis Center Workshop 2002, Ottawa, Canada, April 8-11.

Rothacher, M. and Beutler, G. (2002b). Advanced Aspects of Satellite Positioning, Lecture Note, University of Calgary, Aug. 2002.

Page 297: 7 Tesis PPP Kongzhe Chen

273

Saastamoinen, J. (1972), Atmospheric Correction for the Troposphere and Stratosphere in Radio Ranging of Satellites, Geophysical Monograph 15, Henriksen (Eds), pp. 247-251.

Schaer, S., W. Gurtner, and J. Feltens (1998), IONEX: The IONosphere Map EXchange Format Version 1, February 25, 1998, in Proceedings of the 1998 IGS Analysis Centers Workshop, ESOC, Darmstadt, Germany, February 9-11, 1998.

Schaer, S. (1999a). Mapping and predicting the Earth's ionosphere using the Global Positioning System. Geodaetisch-geophysikalische Arbeiten in der Schweiz, 59.

Schaer, S., Beutler, G., Rothacher, M., Brockmann, E., Wiget, A. And Wild, U. (1999b), The Impact of the Atmosphere and Other Systematic Errors on Permanent GPS Networks, Proceedings of the IUGG 99 General Assembly, Birmingham, UK, July 1999.

Schaer, S. (2000), [IGSMAIL-2827]: Monitoring (P1-C1) code biases. Available: http://igscb.jpl.nasa.gov/mail/igsmail/2000/msg00166.html.

Scherneck, H-G, (2003), personal communication.

Schildknecht, T., Beutler, G., Gurtner, W. and Rothacher, M. (1990), Towards sub-nanosecond GPS Time Transfer using Geodetic Processing Technique, Proceedings of the 4th EFTF, pp. 335-346.

Schildknecht, T. and Springer, T. (1998), High Precision Time and Frequency Transfer Using GPS Phase Measurements, 32nd Annual Precise Time and Time Interval (PTTI) Meeting. November 28-30, 2000, Reston, Virginia.

Schildknecht, Th. and Dudle, G. (2000), Time and frequency transfer: high precision using GPS phase measurements, GPS World, Vol. 11, No. 2, pp. 48–52.

Schmid R. and Rothacher, M. (2002), Estimation of Elevation-Dependent GPS Satellite Antenna Phase Center Variations, IGS Towards Real-Time Network, Data, Analysis Center 2002 Workshop, April 8-11, 2002, Ottawa, Canada.

Schuler, T., A. Posfay, G. W. Hein, and R. Biberger (2001), A Global Analysis of the Mean Atmospheric Temperature for GPS Water Vapour Estimation, Proceedings of ION-GPS 2001, Salt Lake City, Utah, Sept. 11-14, 2001.

Page 298: 7 Tesis PPP Kongzhe Chen

274

Senior, K., Powers, E. and Matsakis, D. (1999), Attenuating day-boundary discontinuities in GPS carrier-phase time transfer, 31st Annual Precise Time and Time Interval (PTTI) Meeting. December 7-9, 1999, Dana Point, California.

Skone, S. (1999), Wide Area Ionosphere Grid Modelling in the Auroral Region, Ph.D. Dissertation, University of Calgary, Calgary, Canada.

Skone, S. and S. Shrestha (2003), 4-D modeling of water vapour using a regional GPS network, Proceedings of the ION National Technical Meeting, Anaheim, CA, Jan., 2003.

Skone, S. (2005), personal communication.

Spilker, JR., J.J. (1996a), Signal Structure and Theoretical Performance. In Global Positioning System: Theory and Applications Volume I, Parkinson, B.W., Spilker, J.J., Eds. Washington: The American Institute of Aeronautics and Astronautics, Inc., 1996, pp. 57-119.

Spilker, JR., J.J. (1996b), GPS Navigation Data. In Global Positioning System: Theory and Applications Volume I, Parkinson, B.W., Spilker, J.J., Eds. Washington: The American Institute of Aeronautics and Astronautics, Inc., 1996, pp. 121-176.

Spilker, JR., J.J. (1996c), Tropospheric Effects on GPS. In Global Positioning System: Theory and Applications Volume I, Parkinson, B.W., Spilker, J.J., Eds. Washington: The American Institute of Aeronautics and Astronautics, Inc., 1996, pp. 517-546.

Sun, H. P., Ducarme, B., and Dehant, V. (1995), Effect of the Atmospheric Pressure on Surface Displacements, Journal of Geodesy, Vol. 70, pp.131-139.

Talbot, N. (1988), Optimal weighting of GPS carrier phase observation based on the signal-to-noise ratio, Proceedings of The International Symposia on Global Positioning Systems, October 17-19, 1988, Queensland, Australia.

Thayer, G. D. (1974), An Improved Equation for the Radio Refractive Index of Air. Radio Science, 9(10), pp. 803-807.

Tiberius, C.C.J.M., Jonkman, N. and Kenselaar, F. (1999), The Stochastics of GPS Observables. GPS World, 10(2), pp. 49-54.

Page 299: 7 Tesis PPP Kongzhe Chen

275

Tetewsky, A.K. and Mullen, F.E. (1996), Carrier Phase Wrap-up Induced by Rotating GPS Antennas, GPS World, Vol. 8, Num. 2.

Teunissen P.J.G and Kleusberg, A. (1998). GPS Observation Equations and Positioning Concepts. In GPS for Geodesy, Teunissen P.J.G and Kleusberg A. (Eds), Springer, 1998. pp. 187-230.

Tregoning P., Boers, R. and O’Brien, D. (1998), Accuracy of absolute precipitable water vapour estimates from GPS observations, Journal of Geophysical Research, 103, 28,701-28,710.

vanDam, T. M. and Wahr, J. M. (1987), Displacements of the Earth's Surface due to Atmospheric Loading: Effects on Gravity and Baseline Measurements, Journal of Geophysical Research, 92, pp. 1281-1286.

vanDam, T. M., Blewitt, G., and Heffin, M. B. (1994), Atmospheric Pressure Loading Effects on Global Positioning System Coordinate Determinations, Journal of Geophysical Research, 99, pp. 23,939-23,950.

vanDam, T. M. (2003), personal communication.

Vo, H. B. and Foster, J. C. (2001), A quantitative study of ionospheric density gradients at midlatitudes, Journal of Geophysical Research, Vol. 106, No. A10, pp. 21,555-21,564

Wang, C. and P. Dare (2004), Precipitable Water Vapour Monitoring Using the Global Positioning System in Atlantic Canada. Proceedings of ION GNSS 2004, Long Beach, California, September 21-24, 2004.

Ware, R., Braun, J., Gutman, S., Ha, S.Y., Hunt, D., Kuo, Y.H., Rocken, C., Sleziak, M., Van Hove, T., Weber, J., Xie, Y., Anthes, R. and MacDonald, A. (2004), Real-Time Water Vapour Sensing with SuomiNet - Today and Tomorrow, Bulletin of the American Meteorological Society (in review), 2004.

Wilson, B.D., Yinger, C.H., Fess, W.A., & Shank, K. (1999). New and improved—The satellite broadcast interfrequency biases. GPS World, 10 (9), 56-66.

Witchayangkoon, B. and P.C.L. Segantine. (1999). Testing JPL's PPP Service. GPS Solutions, Vol. 3, No. 1, pp. 73-76.

Page 300: 7 Tesis PPP Kongzhe Chen

276

Witchayangkoon, B. (2000), Elements Of GPS Precise Point Positioning, Ph.D. Thesis, The University of Maine.

Wu, J.T., Wu, S,C., Hajj, G.A., Bertiger, W.I. and Lichten, S.M. (1993). Effects of antenna orientation on GPS carrier phase, Man. Geodetica, Vol. 18, pp. 91-98.

Xu, G. (2003), GPS • Theory, Algorithms and Applications, Springer Heidelberg, 340 pages, in English.

Yuan L., Anthes, R.A., Ware, R.H., Rocken, C., Bonner, W., Bevis, M. and Businger, S. (1993), Sensing Climate Change Using the Global Positioning System, Journal of Geophysical Research, Vol. 98, No. D8, pp 14,925-14,937, 1993.

Yunck, T.P. (1996a), Orbit Determination. In Parkinson & Spilker, Jr. (Eds.), Global Positioning System: Theory and Applications Volume II. Progress in Astronautics and Aeronautics, Volume 163, American Institute of Aeronautics and Astronautics, Inc.

Yunck, T.P., Y. E. Bar-Sever, W. I. Bertiger, B. A. Iijima, S. M. Lichten, U. J. Lindqwister, A. J. Mannucci, R. J. Muellerschoen, T. N. MUSNOn, L. Romans, and S. C. Wu (1996b), A Prototype WADGPS System for Real Time Sub-Metre Positioning Worldwide, Proceedings of ION GPS 96, Kansas City, Kansas, September, 1996.

Zhu, S. Y., Massmann F.H., Yu Y., Reigber C. (2003), Satellite antenna phase center offsets and scale errors in GPS solutions, Journal of Geodesy, 76 (11-12), 668-672.

Zumberge, J.F., Neilan, R.E. and Mueller, I.I. (1995), Densification of the IGS Global Network, Proceedings of the IGS Workshop on Densiflcation of the IERS Terrestrial Reference Frame through Regional GPS Networks, May, 1995, J. F. Zumberge and R. Liu, eds.

Zumberge, J.F., Heflin, M.B., Jefferson, D.C., Watkins, M.M. and Webb, F.H. (1997), Precise point positioning for the efficient and robust analysis of GPS data from large networks. Journal of Geophysical Research, Vol. 102, 5005-5017.

Zumberge, J.F., M.M. Watkins and F.H. Webb (1998), Characteristics and applications of precise GPS clock solutions every 30 seconds, Navigation, 44(4), 449-456.

Zunberge, J. (1999). Automated GPS Data Analysis Service, GPS Solutions, Vol. 2, No. 3, pp. 76-78.

Page 301: 7 Tesis PPP Kongzhe Chen

277

Zumberge, J. and Gendt, G. (2000). The demise of selective availability and implication for the International GPS Service, position paper presented at the IGS Network Workshop 2000, held in Oslo, Norway, July, Phys. Chem. Earth, Vol. 26, No. 6-8.