154
UCGE Reports Number 20231 Department of Geomatics Engineering Hardware Simulator Characterization of Assisted GPS (URL: http://www.geomatics.ucalgary.ca/links/GradTheses.html) by Milidu Dharshaka Karunanayake November 2005

UCGE Reports Number 20231

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: UCGE Reports Number 20231

UCGE Reports

Number 20231

Department of Geomatics Engineering

Hardware Simulator Characterization of Assisted GPS

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

by

Milidu Dharshaka Karunanayake

November 2005

Page 2: UCGE Reports Number 20231

UNIVERSITY OF CALGARY

Hardware Simulator Characterization of Assisted GPS

by

Milidu Dharshaka Karunanayake

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF SCIENCE

DEPARTMENT OF GEOMATICS ENGINEERING

CALGARY, ALBERTA

November, 2005

© Milidu Dharshaka Karunanayake 2005

Page 3: UCGE Reports Number 20231

iii

Abstract The Federal Communications Commission (FCC) E-911 mandate, Location-Based

Services, as well as personal and vehicular navigation applications are driving the need

for navigation capability in degraded signal environments such as in urban areas and

indoors. Since the position accuracy yielded by GPS methods is better than other

positioning technologies, most wireless carriers are looking at Assisted GPS (AGPS) as

the solution to meet the FCC criteria.

In this thesis, the performance of AGPS is analyzed using a hardware simulator

and compared to High Sensitivity GPS (HSGPS) and conventional GPS receivers. It is

found that an AGPS receiver provides greater acquisition sensitivity and similar tracking

performance as an HSGPS receiver. The time-to-first fix (TTFF) is considerably lower

due to the assistance data provided to the AGPS receiver. The effect of aiding

information is also investigated. Results indicate that for weak signals, timing accuracy

has a significant effect on TTFF, with more accurate timing leading to lower TTFFs. In

terms of initial position, as the user to reference distance increases, the TTFF also

increases. No clear trend was observed in the position domain. In terms of Radio

Frequency Interference (RFI), the AGPS receiver was able to tolerate 5 to 10 dB more

interference power than an HS receiver and 10 to 15 dB more than a conventional

receiver in acquisition mode. Both AGPS and HS have similar performance while

tracking. In all RFI cases, all the receivers were able to tolerate more interference while

tracking than in acquisition mode.

Page 4: UCGE Reports Number 20231

iv

Acknowledgements The completion of this thesis would not be possible without the help of many people and

I would like to acknowledge them at this time. I would like to thank …

• My parents, without whom it would not be possible to accomplish what I have so

far in my life. Thank you for being my greatest cheerleaders and when necessary,

for being my critics too. A special thank you to my brother Charaka.

• My supervisor, Dr. Elizabeth Cannon; she gave me an opportunity to apply my

knowledge in Electrical Engineering in a different field and consequently,

renewed my enthusiasm, my lifelong thirst to learn. Thank you for your financial

support, encouragement, patience and mentoring and for allowing me to be

myself throughout my studies.

• My co-supervisor Dr. Gérard Lachapelle. Thank you for your financial assistance,

positive feedback and your great sense humour, which provided many laughs.

After all, laughter is the best medicine.

• SiRF Technology Inc. for proving us with the AGPS receiver. A special thank

you to Geoff Cox for answering my many questions. Lionel Garin and Greg

Turetzky at SiRF are also acknowledged.

• Sanjeet Singh for his friendship; we have travelled the same road (at least in terms

of academics) since 1997 and it has been quite an adventure.

• Tao Hu for being the best officemate that a guy like me, who loves to talk, can

ask for. Xie xie!

• Sameet Deshpande for your friendship and expertise.

• My colleagues in Geomatics Engineering:

Haitao Zhang and Ping Lian for exposing me to the Chinese culture

PLAN group members that have worked with me since the start of my

studies especially Salman Syed, Rob Watson, Mark Petovello, Olivier

Julien, Nyunook Kim, Scott Crawford, Diep Dao, Anastasia Salycheva,

Page 5: UCGE Reports Number 20231

v

Chaminda Basnayake, Bo Zheng, Seema Phalke and Glenn MacGougan.

Thank you all for making the PLAN group a fun place to work.

Victoria Hoyle, Natalya Nicholson and Ruben Yousef

• God for all that he has given me. “Every good and perfect gift is from above”

James 1:17.

Page 6: UCGE Reports Number 20231

vi

Dedication To my parents, for all their support, guidance and love

Across 2. Frequency shifts 4. Data 7. AGPS receiver 8. LBS device 12. Thesis topic 16. MS 18. Satellite orbit

Down 1. Author 3. Satellite 5. Cell 6. First fix time 8. Almighty 9. User 10. Direct signal 11. Positioning system 13. Unaided 14. Technology to transmit codes 15. Interference 17. Reference receiver

Answers: Go to http://www.geocities.com/dhardevil/puzzle

Page 7: UCGE Reports Number 20231

vii

Table of Contents Abstract ............................................................................................................ iii Acknowledgements........................................................................................... iv

Dedication .......................................................................................................... vi Table of Contents ............................................................................................. vii List of Tables ...................................................................................................... x

List of Figures................................................................................................... xii List of Abbreviations........................................................................................ xv

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

1.1 Relevant Research............................................................................................... 4

1.2 Research Objectives............................................................................................ 5

1.3 Thesis Outline ..................................................................................................... 8

CHAPTER 2: GPS, HSGPS and AGPS .............................................................. 9

2.1 GPS Overview .................................................................................................... 9

2.2 GPS Signal Structure ........................................................................................ 10

2.2.1 Spread Spectrum Basics...................................................................... 11

2.2.2 Code Division Multiple Access ........................................................... 13

2.2.3 L1 and L2 Signals ............................................................................... 13

2.2.4 Auto Correlation ................................................................................. 15

2.2.5 Cross Correlation ............................................................................... 17

2.3 GPS Observations and Error Sources ............................................................... 18

2.3.1 Pseudorange ....................................................................................... 18

2.3.2 Carrier Phase...................................................................................... 20

2.3.3 Doppler Measurement ........................................................................ 21

2.3.4 Error Sources...................................................................................... 22

2.4 Receiver Architecture ....................................................................................... 26

2.5 High Sensitivity GPS ........................................................................................ 28

Page 8: UCGE Reports Number 20231

viii

2.6 Assisted GPS..................................................................................................... 33

2.6.1 Aiding Parameters .............................................................................. 36

2.6.2 Timing Information ............................................................................. 38

2.6.3 AGPS Field Tests ................................................................................ 41

CHAPTER 3: Acquisition and Tracking Tests ..................................... 43

3.1 Hardware GPS RF Simulation .......................................................................... 43

3.2 Test Setup.......................................................................................................... 45

3.3 Acquisition........................................................................................................ 49

3.3.1 All Satellites with the Same Power ..................................................... 49

3.3.2 One Satellite with a Strong Signal ...................................................... 53

3.4 Tracking ............................................................................................................ 56

3.5 Chapter Summary ............................................................................................. 62

CHAPTER 4: Assistance Data ......................................................................... 63

4.1 Timing Accuracy .............................................................................................. 63

4.2 Initial Position................................................................................................... 70

4.3 Position Uncertainty.......................................................................................... 77

4.4 Chapter Summary ............................................................................................. 82

CHAPTER 5: RF Interference on AGPS................................................. 84

5.1 Sources of RF Interference ............................................................................... 84

5.2 Interference Effects........................................................................................... 89

5.3 Interference Tests.............................................................................................. 92

5.3.1 CW Interference .................................................................................. 93

5.3.2 AM Interference .................................................................................. 98

5.3.3 FM Interference ................................................................................ 102

5.4 Chapter Summary ........................................................................................... 107

Page 9: UCGE Reports Number 20231

ix

CHAPTER 6: User Dynamics on AGPS................................................. 108

6.1 Effect of Velocity............................................................................................ 108

6.1.1 Acquisition ........................................................................................ 108

6.1.2 Tracking ............................................................................................ 114

6.2 Effect of Acceleration ..................................................................................... 116

6.2.1 Acquisition ........................................................................................ 116

6.2.2 Tracking ............................................................................................ 119

6.3 Chapter Summary ........................................................................................... 121

CHAPTER 7: Conclusions and Recommendations .......................... 123

7.1 Conclusions..................................................................................................... 123

7.2 Recommendations for Future Work................................................................ 125

REFERENCES................................................................................................ 128

Appendix A: Network-based Positioning Technologies .............. 136

Page 10: UCGE Reports Number 20231

x

List of Tables

Table 2.1: Cross correlation probability of C/A code [from Kaplan, 1996]..................... 18

Table 2.2: GPS Error Sources [from Lachapelle, 2002]................................................... 24

Table 2.3: Processing Gain Example ................................................................................ 30

Table 3.1: Acquisition Performance (All Same Power) ................................................... 50

Table 3.2: Acquisition Performance (One Strong Signal) ................................................ 53

Table 3.3: Tracking Performance ..................................................................................... 56

Table 3.4: Average Number of Satellites Tracked............................................................ 59

Table 3.5: 2D and 3D RMS Errors ................................................................................... 60

Table 4.1: 2D and 3D RMS Errors for Different Timing Accuracies .............................. 68

Table 4.2: Average Number of Satellites for Different Timing Accuracies..................... 68

Table 4.3: 2D RMS Errors for Different Initial Position Offsets ..................................... 74

Table 4.4: 2D RMS Errors for Different User to Reference Distances with a Fixed

Horizontal Uncertainty.................................................................................... 77

Table 4.5: 2D RMS Errors for Different Horizontal Uncertainties (11 km User-to-

Reference Distance) ........................................................................................ 81

Table 4.6: 2D RMS Errors for Different Horizontal Uncertainties (30 km User-to-

Reference Distance) ........................................................................................ 82

Table 5.1: Types of RFI and possible sources [Kaplan, 1996]......................................... 85

Table 5.2: Mobile Frequencies and Power Levels [from Paddan et al., 2003]................. 86

Table 5.3: Acquisition and Tracking Threshold under CW Interference ......................... 94

Table 5.4: Acquisition and Tracking Threshold under AM Interference ......................... 98

Page 11: UCGE Reports Number 20231

xi

Table 5.5: Acquisition and Tracking Threshold under FM Interference ........................ 103

Table 6.1: 2D RMS Errors for Different Accelerations (Acquisition Test) ................... 119

Table 6.2: 2D RMS Errors for Different Accelerations (Tracking Test)........................ 121

Page 12: UCGE Reports Number 20231

xii

List of Figures

Figure 1.1: Gizmondo GPS Mobile Device [www.gizmondo.com]................................... 3

Figure 2.1: GPS Signal Spectrum [from Deshpande, 2004]............................................. 11

Figure 2.2: GPS Satellite Transmitter Unit [from Spilker and Parkinson, 1996] ............ 15

Figure 2.3: GPS Receiver Architecture ............................................................................ 26

Figure 2.4: AGPS System................................................................................................. 34

Figure 3.1: Spirent GSS6560 Hardware Simulator........................................................... 44

Figure 3.2: Schematic of Simulator Test Setup ................................................................ 46

Figure 3.3: Test Setup Components.................................................................................. 47

Figure 3.4: TTFF for AGPS for 125 µs Timing (All Satellites Same Power) .................. 52

Figure 3.5: TTFF for AGPS 125 µs Timing Accuracy (One Strong Signal).................... 55

Figure 3.6: Average C/N0 (for all satellites) during Tracking Test .................................. 58

Figure 3.7: Number of Satellites Tracked during Tracking Test ...................................... 59

Figure 3.8: Tracking Test – Position Errors...................................................................... 60

Figure 4.1: Test Setup for Timing and Assistance Data ................................................... 64

Figure 4.2: Normalized TTFF for Different Timing Accuracies...................................... 65

Figure 4.3: Normalized TTFF for 125 µs Timing Accuracy ............................................ 65

Figure 4.4: Position Errors for Different Timing Accuracies at -138 dBm...................... 67

Figure 4.5: Position Errors for Different Timing Accuracies at -140 dBm...................... 67

Figure 4.6: Scenario Representation for a User-to-Reference Distance of 22 km............ 72

Figure 4.7: Average TTFF for Different User-to-Reference Distances............................ 73

Figure 4.8: Position Errors for 2 km and 55 km Initial Position Offsets .......................... 73

Page 13: UCGE Reports Number 20231

xiii

Figure 4.9: TTFF Comparison of Initial Position Offsets for Two Scenarios .................. 76

Figure 4.10: Scenario Representation (11 km User-to-Reference Distance).................... 79

Figure 4.11: TTFF Performance for Different Horizontal Uncertainties ......................... 80

Figure 4.12: Position Errors for 5 km and 50 km Horizontal Uncertainties (11 km User-

to-Reference Distance)................................................................................. 81

Figure 5.1: Interference Test Setup................................................................................... 93

Figure 5.2: 2D Error vs. Relative CW Interference Power (Acquisition) ........................ 96

Figure 5.3: 2D Error vs. Relative CW Interference Power (Tracking)............................. 96

Figure 5.4: 2D RMS Error for each CW Interference Power Interval.............................. 97

Figure 5.5: 2D Error vs. Relative AM Interference Power (Acquisition) ...................... 100

Figure 5.6: 2D Error vs. Relative AM Interference Power (Tracking)........................... 100

Figure 5.7: 2D RMS Error for each AM Interference Power Interval............................ 101

Figure 5.8: Average C/N0 vs. the Relative FM Interference Power (Acquisition) ......... 104

Figure 5.9: Average C/N0 vs. the Relative FM Interference Power (Tracking) ............. 104

Figure 5.10: 2D Error vs. Relative FM Interference Power (Acquisition) ..................... 105

Figure 5.11: 2D Error vs. Relative FM Interference Power (Tracking) ......................... 105

Figure 5.12: 2D RMS Error for each FM Interference Power Interval .......................... 106

Figure 6.1: Dynamics Test Setup.................................................................................... 109

Figure 6.2: Dynamics Test Trajectory ............................................................................ 110

Figure 6.3: Average TTFFs for Different Velocities ...................................................... 111

Figure 6.4: 2D Position Error for AGPS (Acquisition) .................................................. 112

Figure 6.5: 2D Position Error for HSGPS (Acquisition) ................................................ 112

Page 14: UCGE Reports Number 20231

xiv

Figure 6.6: AGPS Position Errors for 72 km/h, 180 km/h and 360 km/h....................... 113

Figure 6.7: AGPS Velocity Errors for Acquisition Test................................................. 114

Figure 6.8: 2D Position Error for AGPS and HSGPS (Tracking) .................................. 115

Figure 6.9: AGPS Velocity Errors for Tracking Test ..................................................... 116

Figure 6.10: Average TTFFs for Different Accelerations .............................................. 117

Figure 6.11: AGPS Position Errors for 0.5 m/s2 and 4 m/s2 (Acquisition) ..................... 118

Figure 6.12: AGPS Position Errors for 0.5 m/s2 and 4 m/s2 (Tracking) ......................... 120

Page 15: UCGE Reports Number 20231

xv

List of Abbreviations

Abbreviations and Acronyms

ADC Analog-to-Digital Converter AGC Automatic Gain Controller AGPS Assisted GPS A-GNSS Assisted GNSS ALI Automatic Location Identification AM Amplitude Modulation AOA Angle of Arrival BOC Binary Offset Carrier BPSK Binary Phase Shift Keying C/A Coarse-Acquisition C/N0 Carrier-to-Noise CDGPS Canada-wide Differential GPS CDMA Code Division Multiple Access CW Continuous Wave DGPS Differential GPS DLL Delay Lock Loop DOP Dilution of Precision DS Direct Sequence DSP Digital Signal Processor E-911 Enhanced 911 EOTD Enhanced Observed Time Difference FAA Federal Aviation Administration FCC Federal Communications Commission FM Frequency Modulation GNSS Global Navigation Satellite System GPS Global Positioning System GSM Global System for Mobile Communications HDOP Horizontal Dilution of Precision HSGPS High Sensitivity GPS IF Intermediate Frequency LBS Location-Based Services LMU Location Measurement Unit LNA Low Noise Amplifier

Page 16: UCGE Reports Number 20231

xvi

LOS Line-of-Sight MS Mobile Station NDU Navigation Data Unit P-code Precise Code PLL Phase Lock Loop PN Pseudo Noise PPS Precise Positioning Service PRN Pseudo Random Noise PVT Position, Velocity & Time RFI Radio Frequency Interference RMS Root Mean Square SNR Signal-to-Noise Ratio SPS Standard Positioning Service SV Satellite Vehicle TCXO Temperature Compensated Crystal Oscillator TOA Time of Arrival TTB Time Transfer Board TTFF Time-To-First Fix UERE User equivalent Range Error UHF Ultra High Frequency VDOP Vertical Dilution of Precision VHF Very High Frequency WAAS Wide Area Augmentation System WLAN Wireless Location Area Network

Page 17: UCGE Reports Number 20231

1

CHAPTER 1: INTRODUCTION

The Global Positioning System (GPS) has become a critical part of the navigation

infrastructure not only within the United States but also in other nations around the

world. With the introduction of the Enhanced-911 (E-911) mandate from the Federal

Communications Commission (FCC) in the United States, it has now become necessary

to provide positions in various environments including urban canyons and indoors. These

environments are referred to as weak/degraded signal environments since the GPS line-

of-sight (LOS) signal can be attenuated by as much as 20-25 dB or more [MacGougan,

2003]. This mandate requires all cell phone service providers to be able to determine the

positions of users by December 31, 2005. The mandate specifies that a mobile 911 caller

location must be established for 67 percent of calls to within 50 metres, and 95 percent of

calls to within 150 metres for wireless handset-based solutions. The requirements are less

stringent for network-based positioning, with 67 percent of calls to within 100 metres,

and 95 percent of calls to within 300 metres [FCC, 2003]. The FCC equivalent body in

Canada, the Canadian Radio-television and Telecommunications Commission (CTRC)

has not yet mandated E-911 Phase II requirements; hence the Canadian carriers are not

pressured to rollout automatic location identification (ALI) technology across their

networks [Mindbranch, 2005].

Wireless carriers have two basic options for meeting the FCC mandate: handset-

based solutions such as Assisted GPS (AGPS) or network-based positioning

technologies. In AGPS, “aiding” parameters are provided by the network to a GPS

Page 18: UCGE Reports Number 20231

2

receiver in a cellular handset to assist in GPS signal acquisition. Aiding improves the

sensitivity of the GPS receiver, allowing it to operate in weak signal environments such

as urban canyons and indoors since conventional GPS receivers (without aiding or high

sensitivity methods) are intended to be used in open sky areas. With network-based

positioning technologies such as Time of Arrival/Time Difference of Arrival

(TOA/TDOA) and Angle of Arrival (AOA), the mobile network in conjunction with

network-based position determination equipment is used to position the mobile device

[Klukas, 1997]. These technologies, which are discussed in Appendix A, require wireless

telecom carriers to make modifications to existing cell sites and the position accuracy

yielded from these techniques is generally lower than GPS-based solutions [LaMance et

al., 2002].

The convergence of wireless technology and GPS has led to the development of a

new set of applications to serve the location-based needs of users. These applications are

known as Location-Based Services (LBS). The intent of LBS is to use accurate real-time

user position information to connect users to nearby points of interest (such as retail

businesses, public facilities or travel destinations), to advise them of current conditions

(such as traffic and weather), or to provide routing and tracking services [Liu, 2000].

The impact of GPS has even entered mobile gaming. GPS-enabled video games

played on mobile devices such as the Gizmondo and location-enabled mobile phones are

adding another dimension to gaming. Gizmondo, which is shown in Figure 1.1, integrates

a GPS receiver (SiRFstarIIe/LP 12-channel GPS chipset), an antenna and SiRFXTrac

high sensitivity tracking software to deliver surreal gaming experiences, and a host of

Page 19: UCGE Reports Number 20231

3

LBS such as “Where Am I,” “Find the Nearest,” and “Tracking.” [Whitford, 2005;

Gizmondo, 2005]

Figure 1.1: Gizmondo GPS Mobile Device [www.gizmondo.com]

For LBS, instead of using GPS, other positioning technologies can also be used

but most LBS applications require outdoor position accuracies of 50 metres or better.

Typical GPS positioning accuracy is 5-10 metres outdoors which is considerably better

than positioning accuracies provided by other positioning technologies. The E-911

mandate, LBS, as well as personal and vehicular navigation applications, are driving the

need for navigation capability in degraded signal environments such as in urban areas and

indoors. Since the position accuracy yielded by GPS methods is better than other

positioning technologies, most wireless carriers are looking at AGPS as the solution to

meet the FCC criteria.

Page 20: UCGE Reports Number 20231

4

1.1 Relevant Research

The GPS L1 carrier is modulated with the Coarse-Acquisition (C/A) code for civilian use.

The signal containing this code, which repeats every millisecond, can be integrated for

extended periods in order to obtain a higher signal-to-noise ratio (SNR) [Peterson et al.,

1995]. Conventional GPS receivers typically use integration times which are less than the

nominal maximum 20 ms coherent interval and are limited in terms of their operational

environments to where there are strong signals (-130 dBm). In order to extend the

capabilities of GPS into many indoor environments where a receiver has to acquire

signals with power levels of -150 dBm or lower, High Sensitivity GPS (HSGPS)

receivers were designed. These receivers employ longer dwell times which can be used to

acquire GPS signals at very low power levels [Chansarkar and Garin, 2000]. In general,

the coherent integration period is limited to 20 ms due to the navigation bits as well as

residual frequency errors during the coherent integration period. Residual frequency

errors are caused by satellite motion, receiver clock instability and user motion-induced

Doppler effects [ibid]. External timing information can be used to extend coherent

integration for more than 20 ms [Krasner et al., 2002; Shewfelt et al., 2001].

The ability to acquire and track weak GPS signals depends on the capabilities of

the receiver to maximize the coherent integration interval prior to non-coherent

accumulation while minimizing residual frequency errors during coherent integration.

The non-coherent integration period, which is the squared output of the coherent interval,

can be much longer compared to coherent integration. However, this procedure results in

a squaring loss [Ray, 2003].

Page 21: UCGE Reports Number 20231

5

There are a number of factors affecting HSGPS performance that have to be taken

into consideration in the design process of the receivers. The thermal noise should be

minimized to maintain tracking and avoid carrier tracking error. In addition, residual

frequency errors should be reduced, which can be accomplished by using a more stable

oscillator [MacGougan, 2003].

The total dwell-time of HSGPS receivers can be up to hundreds of milliseconds

while for conventional GPS it is less than the 20 ms coherent integration interval

maximum. In general, high sensitivity methods can be implemented in either aided or

unaided modes. In unaided mode, the high sensitivity receiver lacks the ability of the

aided receiver to acquire weak signals if it has no a priori knowledge about GPS time and

the current position. However, if an HSGPS receiver is initialized with assistance data by

first acquiring and tracking four or more GPS satellites with strong signals, it has the

same functional capability as an AGPS receiver as long as it can maintain timing,

approximate position, and satellite ephemeris. In aided mode (AGPS), initialization in a

strong signal environment is unnecessary as the network provides the assistance data.

1.2 Research Objectives

When performing GPS testing, two options are available: field testing and simulator

testing. The advantage of field testing is that the results obtained indicate the performance

in real-world conditions. However, the disadvantages are that field testing is more time

consuming and often more expensive than simulator testing. Furthermore, repeatability is

almost impossible in the field. In recent years, advances in simulation technology have

Page 22: UCGE Reports Number 20231

6

contributed to the development of state-of-the-art hardware GPS RF signal simulators.

The simulators provide a controllable environment and results can be verified with

multiple tests that assess repeatability. In this thesis, the latter option is employed. All of

the testing was conducted using the Spirent GSS6560 hardware simulator.

Over the years, extensive testing has been conducted in HSGPS. High-sensitivity

receiver development since 2000 has yielded receivers capable of tracking signal levels

well below nominal levels in real-time [Chansarkar and Garin, 2000]. MacGougan et al.

[2002] demonstrated the operation of the SiRFstarII HSGPS receiver indoors, most

notably demonstrating an ability to track multiple satellites to levels at least 9 to 10 dB

below that which standard receivers can track, and 25 dB below nominal levels. In a

garage environment, the availability of a solution with the SiRF receiver was greater than

99%, with horizontal position-domain errors near 10 m RMS observed. Further similar

tests by Lachapelle et al. [2004] in a North American residence show similar solution

availability of greater than 98% with position errors of approximately 17 m RMS

horizontally. While results as discussed above are promising, these tests focused on

tracking in an indoor environment. In all of the tests, outdoor acquisition was required in

order to initialize the receivers before indoor operation. Thus, in recent years, more focus

has been on testing GPS augmentations such as AGPS and hybrid systems that provide

better performance.

In terms of AGPS testing, field tests have been conducted by SiRF Technology

Inc. [Garin et al., 1999; Garin et al, 2002], Global Locate Inc. [van Diggelen, 2001b] and

QUALCOMM [Biacs et al, 2002]. Syrjärinne [2001] discusses AGPS in detail but from a

Page 23: UCGE Reports Number 20231

7

theoretical perspective and his major contributions are mostly related to GPS time

recovery and decision-based integrity monitoring. However, little research has been

conducted to investigate the effect of aiding parameters on AGPS acquisition and

tracking performance which is the major thrust of this research. The objectives of this

thesis are as follows:

• Assess the fundamental signal acquisition and tracking capability of an AGPS

receiver under weak signal conditions

• Investigate the effect of timing accuracy and approximate position as well as

other aiding parameters on AGPS acquisition and tracking performance

• Investigate the acquisition and tracking performance of an AGPS receiver in

the presence of Radio Frequency interference (RFI), and

• Investigate the effects of user dynamics on AGPS acquisition and tracking

performance

SiRF Technologies Inc. has developed an AGPS receiver (SiRFLocTM) capable of

making measurements in GPS signal degraded environments. All of the testing done in

this thesis will use this receiver. The SiRFLoc receiver will be discussed more thoroughly

in Chapter 3.

Results from simulation tests conducted using a hardware simulator will be

quantified in terms of Time-To-First-Fix (TTFF) as well as position accuracy.

Page 24: UCGE Reports Number 20231

8

1.3 Thesis Outline

This thesis is organized in the following manner: Chapter 1 has introduced the need for

AGPS as well as the research objectives. Chapter 2 gives an overview of GPS, HSGPS

and AGPS. The GPS receiver architecture is also presented in this chapter. In Chapter 3,

the acquisition and tracking simulation tests performed to obtain the sensitivity of an

AGPS receiver are presented.

In Chapter 4, the effects of various aiding parameters on AGPS acquisition are

investigated. Parameters such as timing uncertainty, initial user position and position

uncertainty are examined. Chapter 5 investigates the effect of various RFI sources on

AGPS acquisition and tracking. In this chapter, Continuous Wave (CW), Amplitude

Modulation (AM) and Frequency Modulation (FM) in-band interference will be studied.

AGPS performance under user dynamics is presented in Chapter 6. Finally, in

Chapter 7, conclusions obtained from research as well as recommendations for future

work are provided.

Page 25: UCGE Reports Number 20231

9

CHAPTER 2: GPS, HSGPS and AGPS

This chapter gives a brief overview of the GPS system including the signal structure and

various measurements. It also discusses the theory behind High Sensitivity GPS and

Assisted GPS.

2.1 GPS Overview

The GPS is a satellite-based radio navigation system capable of providing position in

most places and environments in the world. This system was developed by the US

Department of Defense to support military forces by providing world-wide, real-time

position and timing information [Parkinson et al., 1995]. Even though it was originally

developed for the military, GPS is widely used in civilian applications [Spilker and

Parkinson, 1996]. Presently, the system currently consists of 27 (nominally 24) satellites

which provide continuous information for the user to compute three dimensional position

and velocity as well as time (PVT). The satellites orbit approximately 20,000 km above

the Earth’s surface (26,000 km from the Earth’s centre) and have an orbital period of

11 hours 58 minutes [ICD, 2003]. The principle behind GPS is one-way TOA ranging

whereby the user determines the TOA of the signals transmitted by the GPS satellites.

These ranges are used to compute the user navigation solution. A 3D position

computation requires range information from at least three satellites. However, since the

user GPS receiver clock is not synchronized with the satellite clocks, an additional

satellite measurement is required to solve for the receiver clock offset [Kaplan, 1996].

Page 26: UCGE Reports Number 20231

10

GPS operates on two signal frequencies, L1 (1575.42 MHz) and L2

(1227.60 MHz), using code division multiple access (CDMA) technology to transmit the

ranging codes [ICD, 2003]. GPS provides different accuracy levels for civilian and

military users. Civilian users have access to the C/A-code which provides the Standard

Positioning Service (SPS) while military users use a Precise (P)-code to get the Precise

Positioning Service (PPS). The P-code is encrypted and hence not available for civilian

users. The SPS provides an accuracy of 36 m (2D RMS 95%) in the horizontal plane and

77 m (95%) in the vertical direction [Stenbit, 2001]. However, recent field tests have

shown 2D accuracies of 2 m (1-σ RMS) with dual frequency single point positioning

[e.g. Cannon et. al, 2004].

2.2 GPS Signal Structure

The current GPS signal structure was specifically developed for positioning military

personnel and hence required good resistance to jamming signals [Parkinson et al., 1995].

The spread spectrum concept was used to transmit ranging codes to provide the desired

anti-jamming performance. A pseudo random noise (PRN) sequence with a high chipping

rate was used to transmit the navigation information onto the GPS frequencies [ICD,

2003]. Spread spectrum signals have power levels below the noise level and can be

recovered only with an appropriate spreading code. The two spreading codes used in the

current GPS signals are the C/A-code and the P-code. These spreading codes were

selected from a family of Gold codes [Kaplan, 1996]. As previously mentioned, each

satellite transmits the signal on two frequencies (L1 and L2) with the P-code present on

Page 27: UCGE Reports Number 20231

11

both the frequencies. The C/A-code is transmitted only on L1. The CDMA technique of

transmitting different spreading codes for each satellite on the same frequency is used to

distinguish the signals from the different satellites [ICD, 2003]. The current GPS signal

structure is shown in Figure 2.1. The concept of spread spectrum and CDMA are

discussed briefly in the next two sections.

Figure 2.1: GPS Signal Spectrum [from Deshpande, 2004]

2.2.1 Spread Spectrum Basics

The spread spectrum concept consists of transmitting information over a large bandwidth

and using a PRN sequence to spread the information [Peterson et al., 1995]. The amount

of bandwidth required for transmission is determined by the PRN sequence bandwidth. It

should be noted that all modulation techniques which use a bandwidth wider than

required for transmission are not spread spectrum techniques. The spread spectrum

technique is useful for long distance communication with less interference problems

[Kaplan, 1996]. During the recovery of the spread spectrum signal, any interference

signal is spread thereby reducing its power level below the noise. Spread spectrum solves

two important communication problems, namely pulse jamming and low probability of

Page 28: UCGE Reports Number 20231

12

detection [Peterson et al., 1995]. The pulse jammer power level is reduced during signal

recovery in the spread spectrum method. The spread spectrum can be recovered only

when the PRN signal used for spreading is known [ibid]. This reduces the chance of

signal detection by other users in the same frequency band.

For GPS signals, direct sequence (DS) spread spectrum is used. It consists of

modulating the information signal using a spreading carrier signal [ibid]. A binary phase

shift keying (BPSK) signal is used to spread (modulate) the navigation data signal. The

BPSK signal is a square wave (±1) and the phase of the modulated signal changes by 180

degrees with a change in the sign of the signal. Consider a data modulated carrier signal,

S(t):

))(cos()( ttAtS Φ+= ω (2.1)

where

A is the amplitude of the carrier signal (Volts),

ω is the carrier frequency (Hz), and

Φ is the data modulation signal.

BPSK spreading is performed by multiplying the S(t) by a function c(t), which

represents the spreading waveform and the resulting signal, ST(t):

))(cos()()( tttActTS Φ+= ω (2.2)

This spread spectrum signal is then transmitted and is received by the receiver

after a delay of T. To recover the signal, the receiver must replicate the spreading signal

Page 29: UCGE Reports Number 20231

13

used at the transmitter and match the phase of the spreading signal. The received signal,

SR(t), is given as follows:

))(cos()()( φω +−Φ+−= TttTtActRS (2.3)

where φ is the random phase error (radians).

The spreading signal, c(t), has values of ±1, which when multiplied with the

received signal c(t-T) will have a value of one when the phase of the replica signal

matches the incoming signal. This allows for the recovery of the information in Equation

2.3 except for some random phase error [Tsui, 2000]. A concept similar to the one

described above is used in GPS for transmission and recovery of information.

2.2.2 Code Division Multiple Access

A CDMA signal is a spread spectrum signal with all the signals using the same centre

frequency. The spreading codes used are a set of orthogonal or near-orthogonal codes

[Kaplan, 1996]. An orthogonal code has zero correlation with the other codes used in the

system. The codes do not have zero cross-correlation due to side lobes of the codes and

hence there is a possibility of a cross-correlation peak, resulting from correlation between

the same or different codes, being higher than the autocorrelation peak when the desired

signal is weak. Auto-correlation and cross-correlation will be discussed in Sections 2.2.4

and 2.2.5.

2.2.3 L1 and L2 Signals

As previously stated, GPS satellites transmit on two frequencies on the L-band frequency

called L1 and L2. The two carrier frequencies are modulated by the spread spectrum

codes with a unique PRN associated with each satellite. The signals are further modulated

Page 30: UCGE Reports Number 20231

14

by a 50 Hz navigation data message [ICD, 2003]. The C/A and P-codes are in quadrant

phase with each other on the L1 frequency. The C/A-code is 1023 bits long and is

available to civilian users. The P-code is a complex binary sequence, approximately

266.4 days long and is allocated such that each satellite transmits a one week portion of

the entire sequence. Since the P-code is reserved for military applications, it is encrypted

using a Y-code. This encrypted code is transmitted instead of the P-code on both

frequencies [ibid].

A GPS satellite uses a 10.23 MHz reference clock to generate both the L1 and L2

frequencies. The reference clock is usually a cesium standard and generates a clock

frequency slightly lower than 10.23 MHz to account for relativistic effects [Spilker and

Parkinson, 1996]. The GPS signal broadcast on the L1 and L2 frequencies have the

following signal structure [Kaplan, 1996]:

)t1f2sin()t(N)t(C/A1A)t1f2cos()t(N)t(P1A)t(1L π+π= (2.4)

)t2f2cos()t(N)t(P2A)t(2L π= (2.5)

where

A1 is the L1 signal amplitude,

A2 is the L2 signal amplitude,

P(t) is the P-code,

C/A(t) is the C/A-code,

N(t) is the navigation data,

cos(2π f1t), sin(2π f1t), cos(2π f2t) are the unmodulated L1 and L2 signals,

and

Page 31: UCGE Reports Number 20231

15

L1(t) and L2(t) are the modulated L1 and L2 signals.

Figure 2.2 shows a block diagram of the GPS satellite transmitter unit. The

navigation data unit (NDU) generates the cosine and sine of the carrier signal which are

modulated by a 50 Hz navigation data signal. This modulated signal is then spread using

the C/A-code and the P(Y)-code [Kaplan, 1996]. The NDU block performs the function

of modulating the signal, and the synthesizer is used to manipulate the signals according

to the bandwidth specifications of the signal. For the L1 signal, the combiner combines

the C/A-code and the P(Y)-code signals onto one signal. Both the L1 and the L2 signals

are transmitted to the Earth using an L-band antenna.

Figure 2.2: GPS Satellite Transmitter Unit [from Spilker and Parkinson, 1996]

2.2.4 Auto Correlation

The autocorrelation characteristics of GPS PRN codes are fundamental to the signal

acquisition and demodulation processes in a GPS receiver [Tsui, 2000]. The correlation

of a code with itself is called autocorrelation, while the correlation between two codes is

called cross-correlation, which will be discussed in the next section. The autocorrelation

function involves replicating the code and then shifting its phase while multiplying it

with the original function. When the phases of the two signals match, the maximum

correlation is obtained. The autocorrelation function for a Pseudo Noise (PN) sequence,

Navigation Data Unit

(NDU)

L-band synthesizer

Combiner L-band

antenna

Page 32: UCGE Reports Number 20231

16

PN(t), whose amplitude is ±A, chipping period is Tc and period is NTc is given by

Equation 2.6 [Macabiau et al., 2001].

∫ +=cT

c

dttPNtPNT

R0

)()(1)( ττ (2.6)

A PN sequence of length N = 2n-1, where n is the number of shift register stages

used to generate the sequence is called a maximum length sequence [Kaplan, 1996]. The

autocorrelation function yields –A2/N outside the correlation interval because the number

of negative values (-1) is always one greater than number of positive values (+1) in a

maximum length PN sequence [Peterson et al., 1995]. An autocorrelation function for a

maximum length PN sequence is the infinite series of triangular functions with period

NTc. The negative correlation amplitude (–A2/N) is obtained when the phase shift,τ, is

greater than ±Tc, (or multiples of ±Tc(N±1)) and represents a constant term in the series

[Macabiau et al., 2001].

GPS PRN codes have periodic correlation triangles and a peak spectrum that has

similar characteristics to the maximum length PN sequences [Kaplan, 1996]. However

the GPS codes are not maximum length PN sequences. A simple 10-bit linear code

generator can generate 1023 sequences but all the autocorrelation functions have

considerable power in the side lobes which affects the signal detection at low signal

strengths. This problem was overcome by combining sequences from two 10-bit shift

registers (G1 and G2) to generate the C/A-code [Spilker and Parkinson, 1996]. The

combination of two sequences from the C/A-code generator yields 1023 possible

combinations. The correlation properties of these sequences were studied and 32 codes

Page 33: UCGE Reports Number 20231

17

with the best cross-correlation properties were selected for the GPS satellites [Kaplan,

1996].

The autocorrelation function of the GPS C/A-code has the same period and shape

in the correlation domain as the maximum length PN sequences. However, there are

small correlation values in the interval between the maximum correlation intervals. These

small fluctuations in the autocorrelation function of the C/A-code result in the deviation

of the line spectrum from the sin(x)/x envelope [Spilker and Parkinson, 1996]. The

1 KHz line spectrum spacing is the same for all the C/A-codes and the 10-bit maximum

length sequence code. The ratio of power in each of the C/A-code line spectrum to the

total power can fluctuate by nearly 8 dB with respect to the -30 dB levels that would be

obtained if every line contained the same power [Kaplan, 1996]. The autocorrelation

function of the P-code has similar characteristics to the C/A-code.

2.2.5 Cross Correlation

A GPS receiver generates a replica of the GPS PRN code and shifts its phase to align

with the PRN code for each satellite. The PRN codes for different satellites should have

poor cross-correlation properties among them to allow acquisition of the correct PRN

signal. The GPS C/A-code length is 1023 chips which causes the cross-correlation

properties to be poor for some codes. The C/A-code autocorrelation peaks are higher than

cross-correlation peaks by just 21-24 dB, which can cause false acquisition [Kaplan,

1996]. Table 2.1 lists the C/A-code cross correlation power probabilities.

Page 34: UCGE Reports Number 20231

18

Table 2.1: Cross correlation probability of C/A code [from Kaplan, 1996]

The P-code is not a maximum length sequence but since its period is very long, its

autocorrelation and cross-correlation properties are almost ideal. The cross-correlation

peak between the P-codes is 127 dB lower than the autocorrelation peak, which is much

better compared to 24 dB difference for the C/A-codes [Kaplan, 1996]. The P-code is not

discussed in detail in this thesis since only the C/A-code is used for the research.

2.3 GPS Observations and Error Sources

Three different types of measurement information can be extracted from a GPS satellite

signal, namely a pseudorange measurement, a carrier phase measurement, and the

Doppler measurement.

2.3.1 Pseudorange

A GPS pseudorange measurement is the apparent distance between receiver and satellite

obtained as a difference between transmission and reception time [Leick, 1995]. The term

pseudo comes from the fact that the measured range has an unknown clock bias which

has to be estimated [Misra and Enge, 2001]. GPS measurements suffer from various

errors arising out of clock and other propagation errors as follows:

Cumulative Probability

of Occurrence

Cross correlation for any

two codes (dB)

0.23 -23.9

0.50 -24.2

0.99 -60.2

Page 35: UCGE Reports Number 20231

19

pionotroporb tdTtdtctdtddttp ερ +−++++= ))()(()()()()( (2.7)

where

)(tp is the pseudorange measurement at time t (m),

)(tρ is the true range between satellite and receiver at time t,

)(tdorb is the orbital error,

)(tdtrop is the tropospheric error at time t,

)(tdiono is the ionospheric error at time t,

c is the speed of light (≈ 2.99 x 108 m/s),

)(tdt is the satellite clock error at time t,

)(tdT is the receiver clock error at time t, and

pε is the combined error due to multipath and receiver noise.

Orbital error, satellite clock error, and atmospheric delay are common to standard

and HSGPS measurements. These effects are spatially correlated and can be reduced by

differencing pseudorange measurements with a receiver at a known location (i.e.

differential GPS (DGPS)) or by analytical modeling often based on parameters included

in the broadcast navigation message. The receiver clock error is included as an unknown

parameter in single point and DGPS methods. Noise on the pseudorange measurement

depends on the received signal strength and the correlation method used by the receiver.

Multipath is the result of reflected signals interfering with the direct LOS signal and is a

dominant source of error in GPS methods that utilize the pseudorange measurement.

When no LOS signal is available, measurements are made on multipath signals only.

Page 36: UCGE Reports Number 20231

20

DGPS is normally implemented by differencing the ranges to common satellites from two

receivers. If the coordinates of one station are known, an accurate position of the second

station (rover) can be determined [Misra and Enge, 2001]. DGPS methods does not

reduce multipath, and the noise of a differenced observation is larger than that of an

individual measurement by a factor of √2 [Hofmann-Wellenhof et al., 1994].

2.3.2 Carrier Phase

A carrier phase measurement is a range measurement computed from the GPS carrier

signal information. The total number of the carrier cycles from the GPS satellites to the

user are measured and converted into a range measurement using the carrier wavelength

[Kaplan, 1996]. The receiver cannot determine the number of integer cycles before the

signal is acquired. This is referred to as the integer cycle ambiguity. This ambiguity must

be resolved before the carrier phase measurement can be used for position computation. It

can be represented as follows [Wells et al., 1986]:

( ) θελρλϕφ ++−+−++=−= NtionodttropdtdTtdtcorbdttt )()()()()()()( (2.8)

where

φ (t) is the carrier phase measurement at time t (m),

λ is the carrier wavelength (m/cycle),

ϕ(t) is the carrier phase measurement (cycles),

N is the integer carrier phase ambiguity (cycles), and

εθ is the carrier multipath and measurement noise (m).

Page 37: UCGE Reports Number 20231

21

The definitions of all other symbols in the above equation are the same as in

Equation 2.7. The carrier phase measurement with the ambiguity resolved to the correct

integer provides a very accurate range measurement and is used to provide

centimetre-level position accuracies.

2.3.3 Doppler Measurement

The Doppler effect is the change in reception frequency due to the relative motion of the

transmitter and receiver and is a direct measure of the rate of change of range between the

two points [Ashjaee, 1985]. Thus the Doppler measurement can be used to calculate the

velocity between the transmitter and the receiver. In GPS, Doppler is a measure of the

instantaneous phase rate of a tracked satellite’s signal [Ward, 1996]. Thus, the velocity of

the user with respect to GPS satellites can be determined. The Doppler measurement does

not only include effects due to motion but also the receiver clock drift [Lipp and Gu,

1994]. Doppler observation equation is given as follows:

pionotroporb tTdttdctdtdtdtt ερφ &&&&&&&& +−++++= ))()(()()()()()( (2.9)

where

)(tφ& is the Doppler measurement at time t,

)(tρ& is true geometric range rate between satellite and receiver at time t,

)(tdorb& is the orbital drift error,

)(tdtrop& is the tropospheric delay drift error at time t,

)(tdtrop& is the ionospheric delay drift error at time t,

Page 38: UCGE Reports Number 20231

22

)(ttd& is the satellite clock drift at time t,

)(tTd & is the receiver clock drift at time t, and

pε& is the combined drift error due to multipath and receiver noise.

The effects of troposphere, ionosphere, orbital error and satellite clock drift can

partly be compensated by differencing or by the parameters in the navigation message

[MacGougan, 2003]. The effect of multipath on the Doppler measurement is fairly small

as it is derived from the phase range measurements (which are less effected by

multipath). However the effect of the receiver clock is variable and depends on the

quality of the oscillator used in the GPS receiver. This error is fairly large for low cost

GPS receivers and is commonly estimated as an unknown parameter. Consequently, a

minimum of four Doppler measurements are needed to estimate the user 3D velocity and

receiver clock drift.

2.3.4 Error Sources

GPS measurements have various errors including satellite clock errors, orbital errors,

atmospheric errors, receiver clock error, multipath and interference [Wells et al., 1986].

The satellite clock error is the drift in the satellite clock with respect to the GPS time

reference. The GPS master control station synchronizes the satellite clock with the GPS

clock during the upload of the navigation information, and this offset is transmitted in the

navigation message. The satellite orbital error is the difference between the satellite’s

position using the ephemeris and the actual values [ICD, 2003].

Page 39: UCGE Reports Number 20231

23

When the GPS signal travels through the troposphere, its path will bend slightly

due to the refractivity of the troposphere [Kaplan, 1996]. The change of the refractivity

from free space to the troposphere causes the the GPS signal to slow down which results

in a delay of the GPS signal. This tropospheric delay is a function of the temperature,

pressure, and relative humidity [Spilker and Parkinson, 1996]. Hopfield [1969] and

Saastamoinen [1972] have developed different tropospheric delay models which can

reduce the tropospheric error by about 90%.

The ionosphere is the layer of the atmosphere that extends from about 60 km to

over 1000 km of height above the Earth’s surface. It is a significant source of range and

range-rate errors for GPS users requiring high-accuracy measurements. The ionospheric

variation is generally large compared to the troposphere and is more difficult to model.

The Doppler induced by ionospheric changes is a function of the total electron content

(TEC). A TEC unit represents 1016 electrons per square metre of ionospheric cross-

section [Spilker and Parkinson, 1996]. Ionospheric errors can be eliminated using dual

frequency measurements from GPS. The single frequency ionospheric Klobuchar model

described in ICD [2003] can reduce the ionospheric error by up to 50%. Ionospheric

errors can be further reduced using better ionospheric estimation models. For example,

Wide Area DGPS systems such as Wide Area Augmentation System (WAAS) and

Canada-wide Differential GPS (CDGPS) can be used to provide ionospheric corrections

[Cannon et al., 2004].

Page 40: UCGE Reports Number 20231

24

The user clock is often inaccurate and not synchronized with the GPS clock,

which results in the user clock error. The approximate magnitudes of different errors are

listed in Table 2.2.

All errors except multipath and noise can be reduced using techniques such as

single-differencing, double-differencing and DGPS corrections. As previously

mentioned, multipath is the error caused by the reflected GPS signals entering the

receiver front-end and mixing with the direct signal. Its effect will be more pronounced

for static receivers close to large reflectors. It is specific to a receiver/antenna and

depends on the surrounding environment [Braasch and Van Grass, 1991].

Table 2.2: GPS Error Sources [from Lachapelle, 2002]

GPS Error Source Error magnitude (1 σ)

(m)

Satellite clock and orbital errors 2.3

Ionosphere on L1 7.0

Troposphere 0.2

Code multipath* 0.01-10

Code noise 0.6

Carrier multipath 50x10-3

Carrier noise 0.2-2x10-3

* Outdoor only; may be much larger indoors

Another major source for degradation of the GPS accuracy and reliability is RFI.

Since there are other sources of errors which further degrade GPS accuracy, this makes

RFI mitigation even more difficult. Adding to the problem, GPS satellites and users are

Page 41: UCGE Reports Number 20231

25

mobile which makes it difficult to integrate the signals over long periods of time to

average out the effects of noise. Satellite and user motion introduce Doppler effects, slow

power fluctuations (due to changes in the effective antenna gain and path loss) and fast

power changes (due to multipath fading, blockage and shadowing) [Heppe and Ward,

2003]. A Doppler fluctuation makes it difficult to distinguish between user motion and

receiver clock drift. Power fluctuations make it difficult to determine the thresholds for

acquisition and tracking while atmospheric errors introduce range and range-rate errors.

Two major concerns for providing users with a reliable GPS solution are RFI and

jamming. Unintentional interference can be caused by RF transmitters, harmonics of

ground transmitters, radar signals and accidental transmission of signals in the wrong

frequency band [Spilker and Parkinson, 1996]. These signals, or the harmonics of the

signals, near the GPS frequencies (L1 and L2) are potential sources of interference.

Interference can also be caused by ionospheric scintillation and evil waveforms

transmitted by the GPS satellites themselves [Jakab, 2001; Geyer and Frazier, 1999].

Pulsed interference can result from radar signals in nearby frequency bands which are not

properly filtered [Littlepage, 1999].

Continuous Wave (CW) interference can be either a pure tone or a narrow band

modulated signal such as AM or FM [Macabiau et al., 2001]. It adds to the signal

spectrum and can affect the carrier tracking. Higher order harmonic emissions from AM

and FM radio broadcast transmitters fall close to the GPS L1 frequency and cause

interference. More details about these types of RFI and their affects on GPS signal

acquisition and tracking will be discussed later in Chapter 5.

Page 42: UCGE Reports Number 20231

26

2.4 Receiver Architecture

A conventional GPS receiver consists of three blocks which process the incoming GPS

signal in three different frequency ranges. The RF section operates on the incoming GPS

signals at the GHz frequency range, the signal processing section operates on the signal at

the MHz/KHz frequency range and the data processing section operates at the Hz

frequency range. A conventional GPS Receiver block diagram is shown in Figure 2.3.

The RF section is responsible for receiving the GPS signal from the antenna and

down converting it to an intermediate frequency (IF). The down conversion process can

be performed in a single stage or in multiple stages [Kaplan, 1996]. Each stage consists

of a local oscillator, mixer and band pass filter to eliminate the undesired mixer product.

The RF section amplifies the signal and also determines its precorrelation bandwidth. The

IF signal is sampled at a desired sampling rate using an automatic gain controller (AGC)

and an analog-to-digital converter (ADC) [Tsui, 2000].

Figure 2.3: GPS Receiver Architecture

Page 43: UCGE Reports Number 20231

27

The signal processor acquires and tracks the signals and determines the navigation

data bit value. Acquisition involves performing a two-dimensional search in the code and

Doppler range. It involves a carrier wipe-off wherein the carrier from the incoming GPS

signal is removed and a code wipe-off wherein the PRN code from the incoming GPS

signal is removed. Once the carrier is wiped off, the residual frequency component is the

Doppler. The acquisition process must replicate both the carrier and code of the satellite

in order to acquire it. To acquire the signal, correlation is done over a period called the

predetection integration period, which is chosen depending on the acquisition scheme,

TTFF requirement, data bit prediction and Doppler frequency [Tsui, 2000]. When the

replica signal correctly matches the code and Doppler of the received signal, a GPS

signal peak is obtained. This peak is easily distinguishable from other peaks at the

nominal power (-130 dBm).

Once the signal is acquired, the tracking loops are used to keep lock on the signal

and to detect the navigation data bit transitions. A phase lock loop (PLL) and a

frequency lock loop (FLL) are used to track the carrier signal whereas a delay lock loop

(DLL) is used to track the code phase [Spilker and Parkinson, 1996]. This section

generates the pseudorange and the Doppler measurements, computes the Carrier-to-Noise

(C/N0) ratio of the signal to determine signal quality and determines the thresholds for

acquisition and tracking processes. It also extracts the raw navigation data from the data

bits collected.

Position, velocity and time are computed by the navigation processor using the

raw pseudorange, Doppler and navigation bit stream provided by the signal processor.

Page 44: UCGE Reports Number 20231

28

The navigation processor must decode the 50 Hz navigation data and compute the GPS

satellite positions. With that information, it can then estimate position and time by least-

squares adjustment of the pseudoranges. With accurate position information, Doppler

measurements can be used to estimate velocity precisely. The algorithms and techniques

used in navigation processing vary with each receiver implementation and depend on the

purpose of the receiver. Beyond the typical least-squares approach, some receivers

implement heavy filtering and employ other error detection and mitigation techniques.

Current GPS receivers combine the receiver blocks to reduce cost and size and to

have a greater level of integration. Advances in GPS receiver technology have made it

possible to have a 12-channel receiver with the capability of computing the navigation

information at a 10 Hz rate, being smaller in size than a credit card, and at an affordable

price [Ray, 2003].

Conventional GPS receivers typically use integration times less than the nominal

maximum 20 ms coherent interval and are limited in terms of their operational

environments to where there are strong signals (-130 dBm). In order to extend the

capabilities of GPS into many indoor environments where a receiver has to acquire and

track signals with power levels of -150 dBm or lower, High Sensitivity GPS receivers

were designed.

2.5 High Sensitivity GPS

The navigation data bit duration puts a limit on the coherent integration period. This limit

puts a constraint on the processing signal gain in the acquisition process which

Page 45: UCGE Reports Number 20231

29

determines the GPS signal level that can be acquired [Ward, 1996]. To acquire weak

signals, the predetection integration time has to be extended beyond 20 ms. This can be

achieved by performing coherent integration for 20 ms and non-coherent integration for

the desired duration [Choi et al., 2002]. Non-coherent integration squares and sums the

signal across the coherent integration periods. This allows for a coherent integration time

to be less than 20 ms and a predetection integration time beyond 20 ms. The total gain

using coherent and non-coherent accumulation is given by:

SQlossMTB preGtot −+= )log(10)*log(10 (2.10)

where

Gtot is total processing gain (dB) with respect to pre-correlation SNR

Bpre is pre-detection bandwidth (Hz)

T is total coherent integration time (ms)

M is number of non-coherent accumulations of the coherent output,

and

SQloss is squaring loss due to non-coherent accumulation (dB).

The limitations of coherent correlation accumulation are data bit transitions and

residual frequency errors. Predicting the data bit transitions and limiting residual

frequency errors during coherent correlation is necessary to obtain the optimal gain prior

to non-coherent accumulation. This is because reduction of the ensuing squaring loss is

paramount to successful non-coherent accumulation.

Page 46: UCGE Reports Number 20231

30

The processing gain example for a high sensitivity receiver is given in Table 2.3.

For a high-sensitivity GPS receiver, the desired acquisition sensitivity is -150 dBm or

lower.

Table 2.3: Processing Gain Example

For a -150 dBm GPS signal with an IF bandwidth of 2.046 MHz, the IF SNR is

about -40 dB. After accounting for implementation losses from 2-bit quantization and

Signal strength at Antenna -150.0 dBm

IF Bandwidth 2.046 MHz Teff 363.0 Kelvin

Noise power -109.2 dBm 10 log (k*Teff * BW) where k is Boltzmann’s constant k = 1.3806503 × 10-23 m2 kg s-2 K-1

IF SNR -40.1 dB Signal Strength (dBm) - Noise Power (dBm)

Coherent addition Pre-detection

bandwidth (Bpre) 2.046 MHz

Coherent interval (T) 20.0 Ms Length of coherent integration Coherent gain (dB) 46.1 dB 10 * log ( Bpre * T) = 10 log (2046 * 20)

Implementation losses 2.0 dB 1.2 dB for 2-bit quantization, 0.5 dB due to filter and 0.3 dB for other losses

Actual coherent gain 44.1 dB Perfect coherent gain - implementation losses

SNR after coherent addition 4.0 dB IF SNR + actual coherent gain

Non-coherent (NC) addition Squaring loss 3.8 dB From Squaring Loss

Total integration period 1,000 milliseconds Total integration time = Number of NC sums (M) * coherent interval

Number of NC sums, M 50.0 coherent

intervals integration period/length of coherent interval

Non coherent gain (dB) 17.0 dB = 10*log10(M)

Expected Output SNR (dB) 17.2 dB coherent SNR + non coherent gain -

squaring loss

Page 47: UCGE Reports Number 20231

31

filter loss, 20 ms of coherent integration will increase the SNR to about 4 dB. To detect a

signal, one requires a SNR of at least 14 dB, so non-coherent accumulation is needed. In

this case, the total integration period is 1 second, which means 50 non-coherent

accumulations are performed to provide an extra non-coherent gain of about 17 dB. The

final expected SNR is about 17 dB, which is higher than the required 14 dB to detect a

signal.

As stated previously, non-coherent integration requires the coherent integration

values and consequently, the noise increases. This is called the squaring loss. The

squaring loss depends upon the SNR after coherent integration and before non-coherent

integration [Ray, 2003]. This loss can be reduced by multiplying the adjacent coherent

integration samples over the desired period [Chansarkar and Garin, 2000]. Lin et al.

[2002] proposed an incoherent integration scheme to reduce the squaring loss present in

non-coherent integration. In this scheme, the absolute amplitudes of the coherent

integrations are summed up instead of squaring before summation which reduces

squaring loss. Multiple thresholds for detection with different coefficients based on the

false detection probability were chosen to compensate for the power loss during the

correlation due to the Doppler frequency mismatch and the code phase transition.

The ability to acquire and track weak GPS signals depends on the capabilities of

the receiver to maximize the coherent integration interval prior to non-coherent

accumulation while minimizing residual frequency errors during coherent integration. In

addition, the design of the receiver must also minimize the impact of thermal noise to

maintain signal tracking. The ability to predict the sign of the bits and the timing of the

Page 48: UCGE Reports Number 20231

32

navigation message signal modulation directly affects the ability to perform long

coherent integration. Thermal noise also induces frequency error jitter depending on the

carrier tracking loop bandwidth. Thermal noise can often be a dominant source of carrier

tracking error, especially for weak GPS signal tracking [MacGougan, 2003].

Residual frequency errors during coherent integration cause a reduction in the

coherent signal gain and higher squaring loss for non-coherent accumulation. Residual

frequency error sources include oscillator instability and user motion-induced Doppler

effects. An error in the receiver clock frequency manifests as a measured Doppler effect

as well as any phase noise in the clock signal makes the Doppler appear to change

rapidly. The receiver clock stability depends on the external oscillator parameters and the

frequency synthesizer used within the Digitizer [Watson, 2005]. A high stability clock

provides less jitter and variations in the Doppler and code measurements, thus enabling

edge-to-edge correlation. A stable clock also enables longer pre and post detection

integration to be performed on the signal without loss of signal due to drift [Sudhir et al.,

2001].

In general, high sensitivity methods can be implemented in either aided or

unaided modes. In unaided mode, the high sensitivity receiver lacks the ability of the

aided receiver to acquire weak signals if it has no a priori knowledge about GPS time and

the current position. However, if an HSGPS receiver is initialized with assistance data by

first acquiring and tracking four or more GPS satellites with strong signals, it has the

same functional capability as an AGPS receiver as long as it can maintain timing,

approximate position, and satellite ephemeris. In aided mode (AGPS), initialization (the

Page 49: UCGE Reports Number 20231

33

need to acquire and track four satellites in a strong signal environment) is unnecessary as

the network provides the assistance data.

2.6 Assisted GPS

The idea of providing assistance to a GPS receiver is not recent. Taylor and Sonnet

[1981] proposed transmitting an acquisition-aiding signal generated by an earth-based

control station to user terminals via a geostationary satellite to simplify user equipment.

The aiding signal identifies satellites in view having best geometry and includes Doppler

prediction data as well as GPS satellite coordinates and identification data associated with

user terminals within an area being served by the control station. As a result, the concept

of aiding has been applied in many variations for almost quarter of a century [van

Diggelen, 2001b].

In present systems, AGPS requires a server with a reference GPS receiver that has

clear LOS views of available satellites. The server collects satellite almanac, ephemeris,

approximate user position (which is defined by the initial reference position and its

associated uncertainty) and timing assistance data from the reference receiver, computes

the assistance information and sends the assistance information to the rover receiver. The

receiver, which typically resides in the mobile cellular handset, uses this information to

speed up the acquisition process (see Figure 2.4).

AGPS methods can be divided into two categories, depending on where the user

GPS position calculation is performed. If the position is calculated at the user, called the

Mobile Station (MS), the method is called MS-based GPS. Alternatively, if the network

Page 50: UCGE Reports Number 20231

34

calculates the position, it is called MS-assisted GPS [Syrjärinne, 2001a]. MS-based GPS

requires assistance data from the network prior to position computation in the handset.

MS-assisted GPS is a distributed system where the function of the handset is to compute

the pseudorange or coarse position and send this information to the network (server),

which then computes the MS position. MS-assisted GPS has the significant downside on

being totally dependent on network coverage and channel capacity [Syjarinne, 2001b]. If

no assistance is available from the network, AGPS architecture could allow the GPS

receiver to work in standalone mode. However, in this case, the acquisition sensitivity of

the receiver would be lower.

Figure 2.4: AGPS System

Page 51: UCGE Reports Number 20231

35

In principle, the assisting network does not necessarily have to be a cellular

network. It is equally possible to transmit the latest navigation data and time assistance,

for example, through a Blue Tooth-connectivity layer or a Wireless Local Area Network

(WLAN). However, currently, the assistance usually comes from a cellular network. For

example, Global System for Mobile Communications (GSM) standards already support

the delivery of assistance and the major cellular manufacturers are committed to make the

necessary modifications for AGPS [Syrjärinne, 2001a].

Both the CDMA and the GSM communities have developed standards for

Network Plane AGPS messaging. The standards are outlined in TIA/EIA/IS-801-1 for

CDMA networks while the GSM standards are outlined in 3GPP2 C.S0022-0-1 and

3GPP TS 25.331. The minimum operational performance for AGPS handsets are

presented in TIA 916 for CDMA and 3GPP2 C.P9004-0 and 3GPP TS 25.171 V6.0.0 for

GSM networks [Bryant, 2004]. There is considerable similarity between the assistance

fields included in the two protocols. In addition, for both networks, the minimum

performance standards are tested using statistical testing for five tests.

The five tests include sensitivity, nominal accuracy, dynamic range, multipath

scenario and moving scenario with a periodic update. The nominal accuracy tests are for

static accuracy under typical signal strength conditions rather than weak signal conditions

and with no multipath present. The performance in the presence of multipath is tested

separately as are the performances under weak signal conditions and under typical land-

based dynamic conditions [ibid].

Page 52: UCGE Reports Number 20231

36

2.6.1 Aiding Parameters

The network can provide various types of assistance data including ephemeris, almanac

and approximate user position. The ephemeris is valid for up to four hours and can take

up to 30 seconds to download a satellite’s ephemeris in a conventional GPS receiver. By

providing the ephemeris from an external source, an AGPS-capable mobile handset can

focus on acquisition and position computation earlier, which leads to a much lower

TTFF.

The almanac can be used to predict the approximate location of a satellite and is

typically used when the ephemeris is not provided. The Almanac coefficients remain

reasonably accurate for months. There are only 10 coefficients per satellite and each of

these consists of fewer bits than the corresponding Ephemeris coefficients. Each satellite

transmits the Almanac coefficients for the entire constellation. The clock corrections are

also included. However, the Almanac coefficients are intended only to provide coarse

satellite location suitable for determining which satellites are visible and the approximate

signal Doppler frequency offsets [Bryant et al., 2001].

Another parameter the network can provide to an AGPS receiver is the position of

the reference station receiver. It is assumed that the rover/user receiver is within 100 km

of the reference station and therefore, both receivers have roughly the same visible

satellites. From this information, an AGPS receiver (i.e. a mobile cellular phone user) can

determine the satellites in view [Garin et al., 1999].

To achieve rapid positioning, the range of frequency uncertainty in acquiring the

satellites at the client receiver must be reduced as much as possible in order to reduce the

Page 53: UCGE Reports Number 20231

37

search time. For this reason, it is common practice in AGPS systems for the server to

transmit Doppler information to the client receiver. Even with this information, the

frequency uncertainty of the receiver local oscillator still remains as an obstacle to rapid

acquisition. Today’s technology can produce oscillators which have a frequency

uncertainty on the order of ±1 part per million (ppm) at a cost low enough to permit

incorporation into a consumer product such as a cell phone [Weill et al., 2004]. But even

with such an oscillator, 1 ppm translates into about ±1575 Hz of frequency uncertainty at

the GPS L1 frequency. Assuming that the coherent integration time during satellite

searching is 20 milliseconds (the length of a navigation message data bit), the frequency

bins in the search would have a 50 Hz spacing. This means, a total of 2 × 1575/50 = 63

frequency bins might have to be searched to find the first satellite. Once the first satellite

is acquired, the local oscillator offset can be determined, and the frequency uncertainty in

searching for the remaining satellites can thereby be reduced [ibid].

The received GPS signals are shifted in frequency due to the relative receiver-

satellite motion which is the so-called Doppler frequency shift. The receiver must find the

frequency of the signal before it can lock onto it. Knowledge of the satellite position and

velocity data and the approximate receiver position reduces the number of frequency bins

to be searched because the receiver directly computes the Doppler frequency shift instead

of searching over the entire possible frequency range. By giving the receiver ephemeris,

almanac and approximate user position, and controlling the local oscillator drift by

synchronizing with a network, the search space is reduced corresponding to the

Page 54: UCGE Reports Number 20231

38

calculated satellite Doppler frequency. Reducing the number of frequency bins which

must be searched to acquire the signal reduces the TTFF.

2.6.2 Timing Information

Prediction of the data bits for enabling coherent integration up to 20 ms or for data wipe-

off (coherent integration longer than 20 ms by cancelling the sign of the incoming bits)

requires precise timing [MacGougan, 2003]. If the GPS receiver knows the absolute time,

it can also find out the exact period of a certain satellite's C/A code. Since the GPS

receiver knows what kind of codes the satellites are sending, it can begin to search for the

right period of the code. On its own, the absolute time does not help the GPS receiver

significantly. A rough estimate of the GPS receiver's position and the position of the

satellites are also needed. An accurate enough position for the MS can be achieved using

GSM Location Services such as Enhanced Observed Time Difference (EOTD). MS

position and satellite positions are needed to remove the transfer delays from the satellite

to the MS, which are also used in solving the period of the satellite code [Kinnari, 2001].

As mentioned previously, a C/A code search has two variables; code phase and

Doppler frequency. Without any assistance data, the Doppler frequency uncertainty

sequence is about 12 kHz. This uncertainty depends on three factors: satellite Doppler

error contribution (±4.5 KHz), local oscillator drift which introduces about ±1.5 KHz

assuming 1 ppm oscillator and user Doppler uncertainty, which is ±300 Hz [Kubrak et

al., 2004]. The C/A code length is 1023 chips and the code phase is usually searched in

0.5 chip increments. The combination of one code bin and one Doppler bin is referred to

Page 55: UCGE Reports Number 20231

39

as a cell. In acquisition, all these cells have to be swept through a few times to remove the

possible error due to noise peaks. With assistance data, the number of possible cells can

be decreased. The satellite's Doppler uncertainty is received in the assistance data and,

thus, the searched Doppler frequencies can be reduced [Kinnari, 2001].

With an accurate reference position and accurate time, the searched code phase

can also be decreased to a few chips. If reference position is known with the uncertainty

of a few hundred metres and the time uncertainty is less than 10 µs, the amount of

searched chips can be decreased to 10 chips. This means that only half of the frequencies

and about one hundredth of the chips have to be searched. With this assistance

information, the total number of the searched cells can be decreased significantly. When

the GPS receiver has synchronized to the first satellite, the clock error of the assisted time

can be solved. The synchronization of the other satellites can be done even more

accurately because the exact time is now known. The GPS receiver knows the exact C/A

code phase where the synchronization should be found [Syrjärinne, 2001b].

Depending on the accuracy of the assisted time, the width of the searched C/A

code window in acquisition can be chosen. The aim of the C/A code search is to get the

GPS receiver synchronized to the C/A codes sent by the satellites. With accurate assisted

time, the searched C/A code window can be narrowed. If the assisted time is not accurate

enough, it has no use in acquisition and the whole C/A code has to be searched. Because

of the correlator implementation used in the GPS receiver, there are a few reasonable

widths of the searched C/A code window, which define a few boundary values for the

accuracy of the assisted time.

Page 56: UCGE Reports Number 20231

40

The optimum target of the time accuracy is that the assisted time error is less than

10 µs. In this case, the search for the synchronization of the incoming code can be started

from the assumed code phase. Now there is no need to do any sweep over the whole code

sequence and no satellite acquisition is needed. The value, 10 µs, for the time accuracy is

due to the width of the correlator. The width of the correlator used is ten chips and, thus,

ten chips can be correlated at the same time. The frequency of the C/A code is

1.023 MHz, which means that the period of one chip is approximately 1 µs and the period

of ten chips is approximately 10 µs. If the error is less than that, the searched C/A code

phase is still in the correlator and it can be found at once. The reference position received

from the LMU must also be quite accurate. The reference position error may not be more

than a few hundred metres [ibid].

In a non-synchronized network, such as GSM, the handset needs to treat the

supplied code phase measurements as relative code phase offsets. In order to do this, the

handset locks on to the first satellite using its Doppler information and a potentially large

search in the time (code phase) domain. Once it locks onto the first satellite, it calculates

the difference between the code phase measurement for this satellite and that supplied in

the assistance data. This offset is then applied to all of the other code phase estimates in

order to determine where to search for those satellites. In a non-synchronized network, it

will take a longer time to lock on to the first satellite but once it gets that one, the narrow

code phase search using the assistance data can be applied to the other satellites [Harper

et al, 2004].

Page 57: UCGE Reports Number 20231

41

2.6.3 AGPS Field Tests

As mentioned in Chapter 1, AGPS field tests have been performed by many GPS

companies. SnapTrackTM, now a QualComm Company, performed tests in various

environments to test the SnapTrack server-aided GPS architecture and Digital Signal

Processing (DSP) software-based receiver solution. The position accuracy ranged from

four metres for the outdoor open site test to 84 m for a 50-story glass/steel building. The

receiver provided position results from 89% (for the glass/steel building test) to 100% of

the time for the outdoor test as well as inside a sport utility vehicle [Moeglein and

Krasner, 1998].

Field tests have also been carried out by SiRF Technology Inc. using SiRFLocTM

client, which is a multimode GPS receiver able to work with assistance data or in

standalone mode. The test was carried in several locations: in a parking lot, in a narrow

walkway between tall buildings, a shopping mall with a glass roof, a two-story office

building and a restroom inside a two-story office building. These tests were carried out in

MS-based GPS mode and the RMS error was around 100 m with the exception of the

narrow walkway where it was 183 m [Garin et al., 1999]. Tests performed by Garin et al.

[2002] found that for weak signals with a C/N0 of 23 dB-Hz, the TTFF was

approximately 40 seconds while the TTFF was about five seconds for open sky

conditions.

Global Locate Inc. also carried out AGPS field tests using the GL-16000TM in

downtown San Francisco. These tests were done inside a shirt pocket, inside a steel truck

going at 112 km/h, a parking lot and a shopping mall. On the bottom floor of the parking

Page 58: UCGE Reports Number 20231

42

lot (four floors from the top), the receiver was able to acquire signals between -150 dBm

to -155 dBm [van Diggelen, 2001b].

Page 59: UCGE Reports Number 20231

43

CHAPTER 3: Acquisition and Tracking Tests

This chapter discusses acquisition and tracking tests performed to obtain the sensitivities

of AGPS, HSGPS and conventional GPS receivers. All of the tests were performed using

a hardware GPS simulator, which is described in the following section.

3.1 Hardware GPS RF Simulation

When testing, it is valuable to have conditions whereby the environment is controllable.

This is very difficult, if not impossible, with field testing. Other advantages of using

simulators are that it is easier to isolate the variables of interest, and results can be

verified with multiple tests that assess repeatability. In recent years, advances in

simulation technology have contributed to the development of state-of-the-art hardware

GPS RF signal simulators. The system at the University of Calgary is the GSS6560,

which is comprised of a control computer and two synchronous 12-channel L1 C/A code

hardware signal simulation units [Boulton, 2002]. These units, shown in Figure 3.1, will

be referred to as simulator vehicles throughout this thesis. The simulator allows real-time

control of the signal level (±20 dB with respect to -130 dBm) for each satellite

corresponding to one channel of signal [ibid]. Some of the simulator capabilities are as

follows:

• Control of the signal power for each channel

• Satellite constellation definition and modeling

• Atmospheric effects modeling (ionospheric/tropospheric)

• Vehicle motion modeling for aircraft, cars, and spacecraft

Page 60: UCGE Reports Number 20231

44

• Simulated vehicle trajectories

• Multipath simulation

• Antenna gain pattern manipulation

• Terrain obscuration modelling

Figure 3.1: Spirent GSS6560 Hardware Simulator

The GSS6560 can be combined with the GSS4766 Interference Simulation to test

satellite navigation equipment in the presence of intentional or unintentional RF

interference. The GSS4766 is capable of a large power and frequency range, multiple

operating modes and multiple channel configurations. It has the following characteristics:

• Broad range of signal options to simulate many types of sources

• CW, AM, FM, and variable bandwidth noise (optional) signals available.

• Pulsed mode available for CW and Noise.

Page 61: UCGE Reports Number 20231

45

• Large power range

• Multi-channel configurations available

• Interactive and modelled modes of operation

The GSS4766 Interference System was used to conduct the RFI tests, which are

described in Chapter 5.

3.2 Test Setup

The tests performed in this section use the SiRFLocTM AGPS evaluation kit. However,

for comparison purposes, two other receivers are also used. All three receivers used are

described below:

• SiRFLocTM – This is a 12-channel L1 C/A-code AGPS receiver. Aiding

information is provided via another receiver, referred to as the Time Transfer

Board (TTB). The TTB acts as the reference receiver and can provide assistance

data including timing and reference position and associated uncertainties with the

reference initial position. The TTB employs a temperature compensated crystal

oscillator (TCXO) to provide a frequency accuracy of ±0.5 ppm. For more details

about the SiRFLoc system architecture, refer to Garin et al. [1999].

• SiRFXTracTM – This is an HSGPS receiver which is similar to the SiRFstarII but

with improvements made in acquisition sensitivity. For the XTrac, more non-

coherent integration is used in acquisition. In tracking mode, 10 ms integration is

performed with the StarII while 20 ms of integration is performed with the XTrac

[Cox, 2005]

Page 62: UCGE Reports Number 20231

46

• NovAtel OEM4 – This is a 24-channel L1/L2 dual frequency geodetic receiver

which is optimized for high accuracy performance under normal operating

conditions.

The expected acquisition sensitivities of the receivers are as follows: -150 dBm

for the AGPS and -142 dBm for the XTrac [SiRF, 2005]. The OEM4 is not expected to

acquire much below nominal signal strengths and the expected tracking threshold for this

receiver is -139 dBm [MacGougan et al., 2002]. The AGPS and HSGPS receivers are

expected to track down to at least -156 dBm, which was the value reported for the

SiRFstarII [MacGougan, 2003].

The test setup used for acquisition and tracking tests conducted is shown in

Figures 3.2 and 3.3.

Figure 3.2: Schematic of Simulator Test Setup

Page 63: UCGE Reports Number 20231

47

Figure 3.3: Test Setup Components

The setup for both the acquisition and tracking tests is similar. The only

difference is that for the acquisition tests, all the receivers are set to a cold start mode at

each power level. In cold start mode, the receiver has no acquisition aiding information

available, meaning it has no information about the current time, the orbits of the satellites

or its current position; therefore, it has to perform a full search to acquire available

satellites.

However, for tracking, all receivers are initialized at -130 dBm and then the

power level of each satellite is lowered until no position fix is obtained. No cold starts are

performed when the power level is lowered.

Page 64: UCGE Reports Number 20231

48

The Low Noise Amplifier (LNA) provides a 30 dB gain while also setting the

effective noise floor for the receivers under test. The noise figure of the LNA indicates

the amount of noise power the LNA will contribute to the total receiver noise. The more

noise the LNA contributes, the higher the noise floor and less sensitive the receiver. The

noise figure is usually expressed in decibels (dB), and is with respect to the thermal noise

power at the system impedance, at a standard noise temperature (usually 20o C, 293 K)

over the bandwidth of interest. It is determined by measuring the ratio, usually expressed

in dB, of the thermal noise power at the output, to that at the input, and then subtracting

from that result, the gain, in dB, of the system. Typical noise figures range from 0.5 dB

for very low noise devices, to 4 to 8 dB.

The 5 dB attenuator provides the LNA with burnout protection. A 10 dB

attenuator was inserted prior to the LNA in order to provide signals as low as -160 dBm.

All hardware simulation tests are designed with 10 to 12 satellites in simulation with no

orbital, atmospheric, multipath or any other errors. Only the effect of noise on receiver

performance will be seen. This is important because it represents only errors due to weak

signal conditions which is what is been investigated. However, this means the errors will

be less than what is expected of a real-life scenario since multipath will produce

significant errors indoors and in an urban canyon environment.

It should be noted that although the AGPS receiver is cold started for each

acquisition test trial, the TTB provides assistance data to the AGPS receiver at the

beginning of each trial, after the cold start has been performed. The cold start is done so

that the procedure is consistent for all of the receivers tested.

Page 65: UCGE Reports Number 20231

49

3.3 Acquisition

Acquisition tests were performed under two scenarios, (1) all satellites having the same

power which is decreased incrementally, and (2) one satellite having a strong signal while

the signal strengths for the remaining satellites are decreased incrementally. The two

scenarios were tested for two reasons. In some weak signal environments, it is possible

that one strong signal may be present. For example, in an indoor environment, there may

be a window which permits at least one strong signal. When a GPS receiver has

synchronized with the first satellite, the clock error of the first satellite can be solved.

With this information, synchronization with other satellites can be done even more

accurately since the GPS receiver knows the exact C/A code phase where the

synchronization should be found, assuming the position is known within a certain

accuracy [Kinnari, 2001]. A more accurate frequency search window can be defined from

the first satellite as it will assist in narrowing down the clock bias. The AGPS receiver

under test has implemented these techniques. Therefore, it is necessary to test the

acquisition sensitivity under both scenarios.

3.3.1 All Satellites with the Same Power

For this test, the power level of each satellite in simulator vehicle 1, which is connected

to the rover receiver (AGPS receiver), was decreased by 1 dB, starting from -130 dBm

and continuing until the receiver could no longer obtain a position fix. Simulator vehicle

2 was connected to the TTB, which is the reference receiver, and the power level of this

vehicle was kept at -130 dBm throughout this test. This is because in reality the reference

Page 66: UCGE Reports Number 20231

50

receiver will reside at the base station and is assumed to operate under good signal

conditions.

At each power level, at least twenty trials were conducted from a cold start. A

trial is considered to be an acquisition from a cold start followed by five position fixes.

To obtain a position fix, the receiver is required to acquire and track at least four

satellites. The receivers were given five minutes to acquire the signal, meaning a position

fix must be obtained within 5 minutes in order for a trial to be successful. The elevation

mask was set to 5° for all the receivers. A timing accuracy of 125 µs was provided to the

SiRFLoc AGPS receiver by the TTB. In real-life AGPS implementations, the timing

assistance can vary from a few microseconds to a few seconds [Bryant, 2004]. The

horizontal uncertainty was set to 2 km while the vertical uncertainty was 50 m.

The acquisition performance threshold when all the signals have the same power

is shown in Table 3.1. The SiRFLoc AGPS receiver was able to acquire and obtain a

position fix when the power level was above -145 dBm while the HSGPS SiRFXTrac

receiver was able to acquire above -140 dBm. The OEM4 was only able to acquire until a

power level of -133 dBm.

Table 3.1: Acquisition Performance (All Same Power)

Acquisition Threshold (dBm) Receiver

Min Max Mean SiRFLoc (AGPS) -145 -145 -145 SiRFXTrac (HSGPS) -140 -139 -140 NovAtel OEM4 -133 -132 -133

Page 67: UCGE Reports Number 20231

51

In this scenario, the AGPS receiver is 5 dB better than an HSGPS receiver and

12 dB better than a conventional receiver. The results are as expected; the OEM4 is not

expected to acquire weak signals while the acquisition sensitivity of the XTrac is close to

the specifications.

The average normalized TTFF for the AGPS receiver when all of the satellites

had the same power is shown in Figure 3.4. In the context of this figure, normalized

means that all TTFF values were divided the maximum TTFF value of the acquisition

tests, which was yielded when the power level of all the satellites was -145 dBm. From

the figure, it can be seen that the TTFF remains relatively the same until -136 dBm. For

power levels below -142 dBm, the TTFF increases significantly. This may be due to the

concept of “acquisition maps” that is employed in the AGPS receiver. Different signal

strength ranges have different coherent and mostly, non-coherent integration times as

well as different acquisition strategies may be used depending on the signal strength

[Cox, 2005]. One can assume that the -142 dBm to -145 dBm may fall under the weakest

signal strength range for the AGPS receiver and the combination of integration times and

acquisition strategy may increase the TTFF significantly.

Page 68: UCGE Reports Number 20231

52

Figure 3.4: TTFF for AGPS for 125 µs Timing (All Satellites Same Power)

At the bottom of Figure 3.4, the percentage of position fixes is shown. From the

nominal power level up to -143 dBm, all trials resulted in a position fix; the success was

100%. However, at -145 dBm, which is the acquisition threshold of the AGPS receiver

Page 69: UCGE Reports Number 20231

53

with all satellites having the same power, the percentage of position fixes was

considerably lower, approximately 35%. At -146 dBm, the failure rate was 100%.

3.3.2 One Satellite with a Strong Signal

This test is similar to the above test with one minor modification. One of the satellites in

vehicle 1 was always kept at -130 dBm while the remaining satellites were decreased by

1 dB. In this scenario, the highest elevation satellite was chosen as the strong signal

satellite. Once again, the signals for the TTB were kept at -130 dBm. At each power

level, at least twenty trials were conducted from a cold start. Table 3.2 shows the

acquisition performance with one strong signal.

Table 3.2: Acquisition Performance (One Strong Signal)

Receiver Acquisition Threshold (dBm)

SiRFLoc (AGPS) -153

SiRFXTrac (HSGPS) -140

NovAtel OEM4 -133

With this particular test, the AGPS receiver was able to acquire and obtain a

position fix down to -153 dBm while no change in performance was noticed with any of

the other receivers. As mentioned previously, this improvement is achieved based on a

new signal processing architecture implemented in the SiRFLoc product. The acquisition

sensitivity under the one strong signal scenario is similar to the one found in the

specifications, which specifies the acquisition sensitivity of the SiRFLoc receiver as

Page 70: UCGE Reports Number 20231

54

-152 dBm. Overall, a 13 dB improvement was noticed with an AGPS receiver compared

to an HSGPS receiver such as the SiRFXTrac and a 20 dB improvement over a

conventional receiver in acquisition sensitivity.

The average normalized TTFF for the AGPS receiver with one strong signal is

shown in Figure 3.5. Here, the normalization was once again performed by dividing all

TTFF values by the TTFF when the power level of all the satellites was -145 dBm (from

the previous acquisition scenario since that yielded the maximum TTFF value). From the

figure, it can be seen that the TTFF remains relatively the same until -148 dBm. For

power levels below -151 dBm, the TTFF increases significantly. It can be concluded that

there is a significant benefit from the acquisition of the first satellite in an AGPS

implementation.

At the bottom of Figure 3.5, the percentage of position fixes is shown. A fix is

considered to be a position solution with at least four satellites. For power levels between

-130 dBm and -151 dBm, the AGPS receiver is able to provide a position fix 100% of the

time. However, at -153 dBm, which is the acquisition threshold of the AGPS receiver

with one strong signal, a position fix was obtained less than 10% of the time. Since the

acquisition sensitivity of the AGPS receiver is specified as -152 dBm by SiRF

Technology Inc., this result is not surprising. At -154 dBm, no position fixes were

obtained.

Page 71: UCGE Reports Number 20231

55

Figure 3.5: TTFF for AGPS 125 µs Timing Accuracy (One Strong Signal)

Page 72: UCGE Reports Number 20231

56

3.4 Tracking

A tracking test was performed to investigate the tracking threshold of three different

receiver technologies: AGPS, HSGPS and a conventional GPS receiver. At the start of

the tracking test, the receivers were given -130 dBm signals for 20 minutes. This was

done to make sure that all of the receivers were able to acquire and track the signals.

Then the power level of the satellites in simulator vehicle 1 was decreased by 1 dB every

one minute. No cold starts were performed when the power level was lowered. The

signals from simulator vehicle 1 were directed to all three receivers via a signal splitter

while the simulator vehicle 2 was connected to the TTB and was kept at -130 dBm for the

entire test. The TTB provided 125 µs timing to the AGPS receiver.

The tracking thresholds for the three receivers under test are shown in Table 3.3.

As can be seen, the tracking performance of the AGPS receiver was similar to that of the

HSGPS receiver. As expected, the OEM4 did not perform well under signal degradation

and was only able to track signals as low as -140 dBm, 15 dB less than all other receivers.

Table 3.3: Tracking Performance

Receiver Tracking Threshold (dBm)

SiRFLoc (AGPS) -155 SiRFXTrac (HSGPS) -155

NovAtel OEM4 -140

C/N0 is the best measurable value of the signal quality present at the input to a

GPS receiver. The C/N0 is an instantaneous measure of the ratio of the carrier power

present to noise power density measured per Hertz of bandwidth. With a minimum

Page 73: UCGE Reports Number 20231

57

guaranteed LOS GPS signal power at -130 dBm, the nominal C/N0 level is 44 dB-Hz

[Lachapelle, 2003]. Theoretically, C/N0 is not dependent on the receiver used; however,

each receiver must compute its own C/N0 value based on the measured signal. This

explains why the C/N0 values are different for each receiver (see Figure 3.6).

The raw pseudorange data was extracted at 1 Hz from each receiver and

processed using the University of Calgary’s C3NAVG2TM software [Petovello et al.,

2000]. C3NAVG2 uses a least-squares algorithm to estimate epoch-to-epoch positions,

which is suited to analyze the impact of errors on positions and velocities since no

filtering is performed. The processed data was used to obtain the position errors for all

receivers under test. It should be noted that all position results presented in this chapter

have been post-processed using the C3NAVG2 software.

Differential positioning methods were not employed; only single point positioning

was used. It should be noted that the position results will be optimistic since only noise

was simulated. No additional errors such as orbital errors, atmospheric effects and

multipath were added to the simulation.

The tracking results, in terms of average C/N0 for all satellites tracked, and the

associated simulator channel power, are shown in Figure 3.6 while Figure 3.7 gives the

number of satellites tracked. The tracking test position errors are shown in Figure 3.8.

Page 74: UCGE Reports Number 20231

58

Figure 3.6: Average C/N0 (for all satellites) during Tracking Test

Table 3.4 shows the results for the average number of satellites tracked for four

power intervals. Similarly, the 2D and 3D errors for all receivers for the same four

intervals are shown in Table 3.5. The four intervals were chosen to be representative of

different power levels. The first interval, -130 dBm, is nominal GPS signal power. The

second interval, from -130 dBm to -140 dBm represents the interval where even the

conventional receiver was able to track. The third interval, from -140 dBm to -150 dBm

is indicative of a weak signal environment. The fourth interval, which is below

Page 75: UCGE Reports Number 20231

59

-150 dBm, represents the most difficult conditions for a receiver since the signals are at

least attenuated by 20 dB.

Table 3.4: Average Number of Satellites Tracked

No. of Satellites Power (dBm)

AGPS XTrac OEM4

-130 7.8 7.9 8.7

-130 to -140 7.9 7.9 8.7

-140 to -150 6.1 5.9 N/A

-150 to -155 4.9 4.8 N/A

Figure 3.7: Number of Satellites Tracked during Tracking Test

Page 76: UCGE Reports Number 20231

60

Table 3.5: 2D and 3D RMS Errors

AGPS XTrac OEM4

RMS Error (m) Power (dBm)

2D 3D 2D 3D 2D 3D

-130 1.1 1.6 1.1 1.7 0.3 0.5 -130 to -140 1.3 1.9 1.3 2.1 0.5 0.7 -140 to -150 5.3 8.0 6.0 8.8 N/A N/A -150 to -155 37.7 53.6 24.7 31.6 N/A N/A

Figure 3.8: Tracking Test – Position Errors

Page 77: UCGE Reports Number 20231

61

It is clear from Table 3.5 that the OEM4 provides very accurate position

solutions. This is expected as it is optimized for high accuracy performance. Its 2D and

3D errors are significantly lower than any of the other two receivers during the -130 to

-140 dBm interval. However in weak signal environments (below -140 dBm), it is not

able to track at all which shows the trade-off between accuracy and tracking capability.

The OEM4 also had much better availability than the AGPS and HSGPS receivers. The

number of satellites tracked by AGPS and HSGPS receivers was very similar for all four

power intervals. As previously stated, no additional errors were simulated; the position

results are all due to noise. Therefore, the position results are very optimistic relative to

those expected under field conditions.

The 2D and 3D errors for the HSGPS and AGPS receivers are similar until a

power level of -150 dBm. In fact, as can be seen from Table 3.5, the AGPS and the

HSGPS receivers had very similar availability and position errors throughout the tracking

test. It can be concluded that aiding provides “coarse” estimates intended to assist

acquisition and this “coarse” assistance is not useful in improving tracking performance

because when the receiver is in tracking mode, it provides a much more precise GPS time

and location. Another reason for the similar performance between the AGPS and

HSGPS while tracking is due to the fact that the SiRFLoc (AGPS) and the SIRFXTrac

(HSGPS) receivers from SIRF Technology Inc. have very similar architectures in terms

of tracking; both receivers are based on the SiRFstarII architecture [Turetzky et al, 1999].

The main difference is that the AGPS receiver is capable of accepting assistance data so

one would expect improvements only is acquisition.

Page 78: UCGE Reports Number 20231

62

3.5 Chapter Summary

In this chapter, acquisition and tracking tests were performed to analyze the performance

of the AGPS receiver. Overall, a 13 dB improvement was observed with an AGPS

receiver compared to an HSGPS receiver such as the SiRFXTrac and a 20 dB

improvement over a conventional receiver in acquisition sensitivity. No improvement

was noted in tracking performance between the AGPS and HSGPS receivers. This is

expected since aiding provides “coarse” estimates intended to assist acquisition and this

“coarse” assistance is not useful in improving tracking performance. In addition, both

receivers are based on the SiRFstarII architecture.

Page 79: UCGE Reports Number 20231

63

CHAPTER 4: Assistance Data

In this chapter, the effect of timing accuracy, initial position and the position uncertainty

on AGPS receiver performance are investigated.

4.1 Timing Accuracy

As previously stated, the accuracy of the code phase prediction is directly proportional to

the accuracy of timing assistance. In order to investigate the effect of timing accuracy on

AGPS acquisition performance, a test was designed with 10 satellites in simulation with

no orbital, atmospheric or any other errors. Simulator vehicle 1 was connected to the

AGPS receiver and the power level of each satellite was decreased by 2 dB, starting from

-130 dBm. At each power level, at least 30 trials were conducted. A trial is considered

to be an acquisition from a cold start mode, followed by five position fixes. The receiver

is then cold started to start a new trial.

The elevation mask of the AGPS receiver was set to 7.5°, which is the default

value. Data on the AGPS receiver was collected at least 90 seconds after the TTB was

initialized to ensure that the reference receiver (TTB) was able to get a position fix with

at least seven satellites. The TTB, which was connected to simulator vehicle 2, was given

“strong” signals, -130 dBm. The test setup for all of the tests performed in this chapter is

shown in Figure 4.1. The precise timing accuracy provided by the TTB to the AGPS

receiver was set to one of three levels: 125 µs, 250 µs and 500 µs.

Page 80: UCGE Reports Number 20231

64

Figure 4.1: Test Setup for Timing and Assistance Data

The effect of timing assistance on the TTFF is shown in Figure 4.2. All TTFF

values have been normalized, meaning all values were divided by the maximum TTFF

value of all the trials in this test. Figure 4.3 shows the TTFF performance for 125 µs for

signal strengths from -130 dBm to -145 dBm.

From Figure 4.2, it can be concluded that up to -136 dBm, it is difficult to

distinguish the effect of different timing accuracies on AGPS performance. However, as

the signals get weaker, a trend can be seen. It is clear that a more accurate timing

accuracy leads to lower TTFFs. For example, at -142 dBm with 125 µs timing accuracy, a

30% improvement in TTFF can be expected over 250 µs and about a 45% improvement

over 500 µs.

Page 81: UCGE Reports Number 20231

65

Figure 4.2: Normalized TTFF for Different Timing Accuracies

Figure 4.3: Normalized TTFF for 125 µs Timing Accuracy

Page 82: UCGE Reports Number 20231

66

From Figure 4.3, which shows the TTFF performance with 125 µs timing

accuracy, it can be seen that the TTFFs increase significantly for signals lower

than -142 dBm. Once again, this may be due to the “acquisition maps” mentioned in

Section 3.3.1.

The position accuracy results at the three different timing accuracies are shown in

Figures 4.4 and 4.5 for -138 dBm and -140 dBm signal power. The 2D and 3D RMS

statistics are listed in Table 4.1. Table 4.2 presents the number of satellites acquired for

each of the scenarios listed in Table 4.1.

The 2D and 3D RMS statistics, which is used throughout the thesis, are computed

using the RMS values for latitude, longitude and height. The ‘2D RMS’ error is

calculated by squaring the RMS values for latitude and longitude and then taking the

square root. For the computation of the ‘3D RMS’ error, in addition to the RMS of

latitude and longitude, the RMS value of the height component is also used. These

statistics are not to be confused with Distance Root-Mean-Square (DRMS) error. This is

a measurement used to describe the accuracy of a fix. It is twice the square root of the

sum of the squares of all radial errors surrounding a true point divided by the total

number of measurements.

Page 83: UCGE Reports Number 20231

67

Figure 4.4: Position Errors for Different Timing Accuracies at -138 dBm

Figure 4.5: Position Errors for Different Timing Accuracies at -140 dBm

Page 84: UCGE Reports Number 20231

68

Table 4.1: 2D and 3D RMS Errors for Different Timing Accuracies

Power: -138 dBm 2D RMS Error (m) 3D RMS Error (m)

All

Fixes No First

Fix First Fix

Only All

Fixes No First

Fix First Fix

Only 125 µs 16.1 7.8 32.7 27.4 13.0 56.0 250 µs 14.0 10.6 23.8 20.6 17.2 31.5 500 µs 10.5 8.4 16.7 18.0 14.7 28.2 Power: -140 dBm 2D RMS Error (m) 3D RMS Error (m)

All

Fixes No First

Fix First Fix

Only All

Fixes No First

Fix First Fix

Only 125 µs 35.2 30.4 58.2 46.7 39.1 82.2 250 µs 24.2 16.7 52.1 32.3 24.7 63.6 500 µs 18.6 16.9 27.4 31.7 27.4 51.9

Table 4.2: Average Number of Satellites for Different Timing Accuracies

Power: -138 dBm No. of Satellites All Fixes No First Fix First Fix Only

125 µs 8.1 8.5 6.5 250 µs 7.3 7.5 6.5 500 µs 7.4 7.7 5.9

Power: -140 dBm No. of Satellites All Fixes No First Fix First Fix Only

125 µs 7.3 7.6 5.1 250 µs 7.2 7.5 7.0 500 µs 6.4 6.6 5.0

In Table 4.1, the first column includes all five position fixes from every trial

performed for that scenario. The column labelled ‘No First Fix’ means that for each trial,

the first position fix of that trial is excluded from the results presented in this column;

only four position fixes per trial are used to create the statistics for that column. Finally,

Page 85: UCGE Reports Number 20231

69

the column labelled ‘First Fix Only’ includes only the first position fix of each of the

trials.

The table is organized in this manner because the first position fix of each trial

seem to affect the results significantly and this can be seen clearly from the table. For

example, at -138 dBm with 125 µs timing, the 2D RMS error is 16.1 m. Without

including the first position fix of each trial, the 2D RMS error reduces significantly, down

to a value of 7.8 m. The 2D RMS error for only first fix is considerably higher, about

33 m.

At -138 dBm, the 2D RMS errors for 125 µs, 250 µs and 500 µs timing accuracies

are 16.1 m, 14.0 m and 10.5 m respectively while at -140 dBm, the 2D RMS errors are

35.2 m, 24.2 m and 18.6 m. These are the results with all of the position fixes included.

Similarly, the 3D RMS errors for the three different timing accuracies tested are 27.4 m,

20.6 m and 18.0 m at -138 dBm and at -142 dBm, the RMS errors are 46.7 m, 32.3 m and

31.7 m. There is a clear trend, the more accurate timing accuracies results in larger 2D

and 3D RMS errors. It is clear from Table 4.2 that the larger errors are not due to a lower

number of acquired satellites for the position fix. In fact, at both power levels, the

average number of satellites for all fixes was higher at a 125 µs timing accuracy than at a

250 or 500 µs time accuracy.

The statistics for the first fix (shown in columns labelled ‘First Fix Only’) clearly

show that for more accurate timing accuracies, the 2D and 3D RMS errors are much

larger. For example, at -138 dBm with 125 µs timing accuracy, the 2D RMS error for the

first fix is 32.7 m while the errors are 23.8 m and 16.7 m for 250 µs and 500 µs timing

Page 86: UCGE Reports Number 20231

70

accuracies respectively. A similar trend is seen for the timing accuracy tests conducted

with -140 dBm.

Without the first position fix (which is presented in the columns labelled ‘No First

Fix’), the trend discussed above does not hold. For example, at -138 dBm, the 125 µs

timing accuracy provided a lower 2D RMS error than both 250 µs and 500 µs timing

accuracy.

It can be concluded that there is a trade-off between the TTFF and position

accuracy. The more precise timing accuracies result in lower TTFFs but larger

positioning errors. These larger positioning errors are mainly due to the poor accuracy of

the first position fix.

4.2 Initial Position

In AGPS, it is assumed that the user receiver is close to the reference station and

consequently, an initial position for the user receiver can be approximated to be the same

location as the reference station. However, with the initial position, an uncertainty also

needs to be defined.

In the AGPS implementation under test, the uncertainty in the initial position is

taken into account with the use of two parameters: the Estimated Horizontal Error and the

Estimated Vertical Error. For example, the estimated horizontal error defines the radius

of uncertainty where the user might be located compared to the actual position of the

reference station. The receiver can use the initial position and uncertainties to determine

the satellites in view, reduce its search space and initialize the navigation solution

Page 87: UCGE Reports Number 20231

71

[Bryant, 2004]. In this section, the effect of the initial position will be investigated. The

effect of the position uncertainty will be investigated in Section 4.3.

A test was carried out to investigate the effect of the initial position on AGPS

acquisition performance. Simulator vehicle 1 was set at a fixed location with a power

level of -140 dBm. The location of simulator vehicle 2 (which is connected to the

reference receiver that supplies the initial position to the user AGPS receiver) was varied

such that the user-to-reference receiver distance was set to one of six values: 2 km, 10

km, 22 km, 30 km, 39 km or 55 km. The power level of vehicle 2 was set to -130 dBm.

The horizontal uncertainty was set to the same value as the user-to-reference

distance tested (i.e. if the reference was 22 km away from the user receiver, the horizontal

uncertainty was also set to 22 km; see Figure 4.6). The vertical uncertainty was set to

50 m for all cases. At least 30 trials were conducted for each of the distances. Here, a trial

is considered to be an acquisition from a cold start mode, followed by thirty position

fixes. The elevation mask of the AGPS receiver was set to 5° and the timing accuracy

provided by the TTB was 125 µs.

This scenario is intended to describe a real-life situation whereby the centre of the

cell sector can be taken as the initial position and the radius of the cell as the horizontal

uncertainty since the user can be anywhere within the specified radius of the cell sector.

Therefore, in urban areas, where there are many cell sectors, the horizontal uncertainty

would be smaller than in rural areas.

Page 88: UCGE Reports Number 20231

72

Figure 4.6: Scenario Representation for a User-to-Reference Distance of 22 km

The average TTFF for each initial position offset (i.e. user-to-reference distance)

tested is shown in Figure 4.7. The C3NAVG2 position domain results for two different

initial position offsets (2 km and 55 km) are shown in Figure 4.8 while the 2D RMS error

statistics for the tests are shown in Table 4.3.

Clearly, as the user-to-reference distance increases, the TTFF also increases. If the

user-to-reference distance was 55 km instead of 2 km, the average TTFF will be more

than twice compared to the 2 km situation. This result is significant, as it clearly indicates

that the AGPS TTFF performance will be better in urban areas than in rural areas,

assuming the cell coverage is better in an urban centre. This is clearly an advantageous

situation since most cellular phone users conduct their daily activities in an urban area.

One could also say that the rural users are less likely to need AGPS since they are more

likely to have a clear view of sky.

Page 89: UCGE Reports Number 20231

73

Figure 4.7: Average TTFF for Different User-to-Reference Distances

Figure 4.8: Position Errors for 2 km and 55 km Initial Position Offsets

Page 90: UCGE Reports Number 20231

74

Table 4.3: 2D RMS Errors for Different Initial Position Offsets

User-to-Reference Distance (km) 2D RMS Error (m)

2 19.2

11 24.1

22 32.0

30 24.3

39 18.5

55 36.2

No clear trend was observed in the position domain, with the position accuracy

between 18 – 36 m. The 2 km and 39 km position offset produced the best positioning

accuracy, approximately 19 m. The largest position offset, 55 km, produced the highest

errors, about 36 m. One might say that the position accuracy may also be better in an

urban area as compared to a rural area. However, keep in mind, that an urban area

presents many challenges including reduced satellite availability and multipath, which

will undoubtedly affect the position accuracy in a real-life scenario. The effect of

multipath on AGPS is not studied in this thesis.

A second simulation was created to evaluate the effect of the initial position offset

(i.e. user to reference distance) with a fixed horizontal uncertainty. This can be

considered to be a case where poor position assistance is provided to the AGPS receiver

Page 91: UCGE Reports Number 20231

75

from the reference receiver since the user to reference distance is much greater than the

horizontal uncertainty specified.

To investigate this scenario, the horizontal uncertainty was fixed to 2 km while

the user-to-reference distance was set to one of six values: 0 km, 11 km, 22 km, 30 km,

39 km and 55 km. As in the previous case, the power level of the vehicle connected to the

AGPS receiver was set to -140 dBm. At least 30 trials were conducted with thirty

position fixes in each trial. Once again, the elevation mask of the AGPS receiver was set

to 5° and the timing accuracy provided by the TTB was 125 µs.

Figure 4.9 shows the average normalized TTFF for the case where the horizontal

uncertainty was fixed to 2 km as well as the previous case where the horizontal

uncertainty matched the user-to-reference distance. Table 4.4 presents the 2D RMS errors

for the fixed horizontal uncertainty.

From Figure 4.9, it can be seen that up to a 22 km user-to-reference distance, the

TTFF for both the matching uncertainty and a 2 km fixed horizontal uncertainty is the

same. This result is surprising since the user was 22 km away from the reference receiver

yet the assistance provided indicated that it was at most 2 km away. If the AGPS receiver

was to accept the position assistance as correct, degradation in TTFF should have been

noticed and this was not the case up to 22 km. This may indicate that the AGPS receiver

under test have some protection built in for the case of poor position assistance. After all,

in real life, most cell sectors are less than 30 km, so it is not surprising that the AGPS

receiver may have algorithms to detect poor position assistance up to 30 km.

Page 92: UCGE Reports Number 20231

76

Figure 4.9: TTFF Comparison of Initial Position Offsets for Two Scenarios

However, as expected, when the user-to-reference is increased beyond 22 km, a

significant increase in TTFF was noticed when the horizontal uncertainty was fixed to

2 km. The AGPS receiver is still able to recover from the poor position assistance

provided but it takes significantly longer to acquire the signals. This is expected since the

fixed uncertainty of 2 km is an erroneous assistance.

Page 93: UCGE Reports Number 20231

77

Table 4.4: 2D RMS Errors for Different User to Reference Distances with a Fixed

Horizontal Uncertainty

User-to-Reference Distance (km) 2D RMS Error (m)

0 13.9

11 21.3

22 24.8

30 19.8

39 24.4

55 26.6

No clear trend can be seen with the positioning results. Most of the user-to-

reference distances tested resulted in 2D RMS errors around 20 – 25 m. It appears that

with a fixed horizontal uncertainty (2 km in this test), the user-to-reference distance has

no effect on the position accuracy.

4.3 Position Uncertainty

As mentioned in the previous section, the horizontal uncertainty defines the radius of

uncertainty where the user might be located compared to the actual position of the

reference station. This information, along with the initial position, can be used to reduce

the search space.

Page 94: UCGE Reports Number 20231

78

A test was carried out to investigate the effect of the varying horizontal

uncertainty on AGPS performance with a fixed user-to-reference offset. Only the effect

of the horizontal uncertainty was examined in this thesis.

Simulator vehicle 1 was placed at a fixed location. Similarly, simulator vehicle 2

(which is connected to the reference receiver) was placed at approximately 11 km away

from the user receiver (vehicle 1). This was purposely done to isolate the effect that an

exact initial position may have on receiver position calculation.

The power level of vehicle 1 was set to -140 dBm while the TTB was given

nominal signals (-130 dBm). The estimated horizontal error parameter (uncertainty) was

set to one of five values: 5 km, 10 km, 20 km, 50 km or 100 km. The estimated vertical

error was kept at 50 m for all cases. At least 30 trials were conducted at each horizontal

uncertainty (see Figure 4.10 for a representation of the scenario).

Another similar test was also performed, with a few modifications. The main

modification was that the initial user-to-reference distance was set to 30 km instead of the

11 km in the above test. In this test, the estimated horizontal error parameter was set to

following values: 5 km, 20 km, 30 km, 50 km and 100 km. Once again, the elevation

mask of the AGPS receiver was set to 5° and the timing accuracy provided by the TTB

was 125 µs.

Page 95: UCGE Reports Number 20231

79

Figure 4.10: Scenario Representation (11 km User-to-Reference Distance)

The average normalized TTFF for each horizontal uncertainty for both the 11 km

and 30 km user-to-reference offsets is shown in Figure 4.11. From the results, it can be

concluded that as the horizontal uncertainty remains within the user-to-reference

distance, there is no observable difference in TTFF. For example, for the 11 km user-to-

reference distance, all horizontal uncertainties below 11 km had similar TTFFs. The same

result was observed for the 30 km user-to-reference distance where uncertainties of 5 km,

20 km and 30 km produced similar TTFFs. However, when the horizontal uncertainty is

larger than the user-to-reference distance, the TTFF increases.

Page 96: UCGE Reports Number 20231

80

Figure 4.11: TTFF Performance for Different Horizontal Uncertainties

The position domain results for two different horizontal uncertainties (5 km and

50 km) for the 11 km user-to-reference distance are shown in Figure 4.12. The 2D RMS

error for the 11 km user to reference offset is shown in Table 4.5 while Table 4.6 shows

the 2D RMS error for the 30km user to reference offset. The position domain results were

processed using the C3NAVG software.

Page 97: UCGE Reports Number 20231

81

Figure 4.12: Position Errors for 5 km and 50 km Horizontal Uncertainties (11 km

User-to-Reference Distance)

Table 4.5: 2D RMS Errors for Different Horizontal Uncertainties (11 km User-to-

Reference Distance)

Horizontal Uncertainty (km) 2D RMS Error (m)

5 21.9

10 25.5

20 19.3

50 20.5

100 20.5

Page 98: UCGE Reports Number 20231

82

Table 4.6: 2D RMS Errors for Different Horizontal Uncertainties (30 km User-to-

Reference Distance)

Horizontal Uncertainty (km) 2D RMS Error (m)

5 26.5

20 33.1

30 19.7

50 26.9

100 35.9

No clear trend can be seen with the positioning results. In fact, all horizontal

uncertainties had similar results for the 11 km initial offset, with the 2D error around 20 –

25 m for all five uncertainties tested. For the 30 km initial offset, the 2D RMS error

varies from 20 m (for 30 km horizontal uncertainty) to 36 m for 100 km horizontal

uncertainty. It can be concluded that there is an effect on the TTFF domain but no impact

in the position domain when the horizontal uncertainty is varied.

4.4 Chapter Summary

In this chapter, the effect of aiding information was investigated. It was found that for

weak GPS signals (-140 dBm or lower), timing accuracy had a significant effect on

TTFF, with more accurate timing leading to lower TTFF values. However, there was a

trade-off between TTFF and position accuracy with more accurate timing leading to

larger positioning errors due to the inaccuracy of the first fix. Previous field tests

Page 99: UCGE Reports Number 20231

83

performed with the SiRFLoc receiver showed that the first position fix always had a

larger error than the subsequent fixes [Garin et al., 2002].

In terms of initial position, as the user to reference distance increases, the TTFF

also increases. No clear trend was observed with the position accuracy. In terms of

position uncertainty, it was found that if the horizontal uncertainty was within the user-to-

reference distance, the TTFF remains unchanged. For a horizontal uncertainty larger than

the user-to-reference distance, the TTFF increases. No clear trend was observed in the

position domain. The AGPS receiver under test is able to recover from poor/erroneous

position assistance but as expected, it takes significantly longer to acquire the signals.

Page 100: UCGE Reports Number 20231

84

CHAPTER 5: RF Interference on AGPS

In this chapter, the effect of various RFI on AGPS acquisition and tracking will be

investigated. Continuous Wave, Amplitude Modulation and Frequency Modulation

in-band interference will be studied.

5.1 Sources of RF Interference

As mentioned in Section 2.3.4, RFI is a major source for degradation of the GPS

accuracy and reliability. Since there are other sources of errors which further degrade

GPS accuracy, this makes RFI mitigation more difficult. Satellite and user motion

introduce Doppler effects, slow power fluctuations (due to changes in the effective

antenna gain and path loss) and fast power changes (due to multipath fading, blockage

and shadowing) [Heppe and Ward, 2003]. Doppler fluctuations make it difficult to

distinguish between user motion and receiver clock drift. Power fluctuations make it

difficult to determine the thresholds for acquisition and tracking while atmospheric errors

introduce range and range-rate errors.

The signals, or the harmonics of the signals, near the GPS frequencies (L1 and

L2), are potential sources of interference. Interference can also be caused by ionospheric

scintillation and evil waveforms transmitted by the GPS satellites themselves [Geyer and

Frazier, 1999]. Unintentional interference can be caused by RF transmitters, harmonics of

ground transmitters, radar signals and accidental transmission of signals in the wrong

frequency band [Spilker and Parkinson, 1996]. Pulsed interference can result from radar

Page 101: UCGE Reports Number 20231

85

signals in nearby frequency bands which are not properly filtered [Littlepage, 1999].

Table 5.1 summarizes various types of RFI.

Table 5.1: Types of RFI and possible sources [Kaplan, 1996]

Type Typical source

Wideband-Gaussian Intentional noise jammers

Wideband phase/frequency modulation Television transmitter’s harmonics or near-band microwave link transmitters

Wideband-spread spectrum Intentional spread spectrum jammers or near-field of pseudolites

Wideband pulse Radar transmitters

Narrowband phase/frequency modulation AM stations transmitter’s harmonics

Narrowband swept continuous wave Intentional CW jammers or FM stations transmitter’s harmonics

Narrowband continuous wave Intentional CW jammers or near-band unmodulated transmitter’s carriers

CW interference can be either a pure tone or a narrow band modulated signal such

as AM or FM [Macabiau et al., 2001]. It adds to the signal spectrum and can affect the

carrier tracking. A carrier tracking loop will lock onto the interference frequency for a

pure tone signal generating erroneous carrier phase and Doppler measurements (provided

the CW power level is considerably high). Broadband noise increases the amount of noise

in the GPS spectrum without distorting the signal spectrum [Heppe and Ward, 2003].

Swept CW interference is more damaging than CW interference because it can cover

multiple Doppler frequencies and affect more than one receiver channel at the same time.

Pulse interference can cause malfunctioning of the AGC which affects the tracking loops

[Hegarty et al., 2000].

Page 102: UCGE Reports Number 20231

86

Mobile phones use FM signals for communication and the incorporation of GPS

into a cellular handset means that a jammer will be operating nearby at the cellular

frequency. For example, GSM phones used in Europe work either on the 900 MHz or

1800 MHz frequency bands while North American GSM phones primarily use the 1900

MHz band. CDMA technology is the basis for Interim Standard 95 (IS-95) and operates

in both the 800-MHz and 1900-MHz frequency bands in the US. The major US carriers

using CDMA are Air Touch, Bell Atlantic/Nynex, GTE and Primeco. In 1994, the FCC

announced it was allocating spectrum specifically for Personal Communication Services

(PCS) technologies at the 1900 MHz band [About.com, 2005]. A summary of handset

frequencies and power levels that will manifest themselves as out-of-band GPS jammers

is provided in Table 5.2.

Table 5.2: Mobile Frequencies and Power Levels [from Paddan et al., 2003]

Cellular Standard Transmit Freq (MHz) Max. Handset Output Power

GSM 880-915 and 1710-1785 +33 dBm

IS-95 824-849 +23 dBm PCS 1850-1910 +24 dBm

A high jammer power level can cause the generation of unwanted mixing

products (spurs) if the level exceeds the linear range of the circuit blocks. An out-of-band

jammer mixes with spectral components to create spurs in the same frequency band as the

desired signal. If the power levels are high enough, the resulting spurs may possibly

exceed the linear range of the circuit, resulting in the receiver's inability to retain GPS

Page 103: UCGE Reports Number 20231

87

signal lock. There is a possibility of these signals interfering with GPS signals and

causing problems for acquisition. Therefore, care must be taken to isolate the mobile

signals from GPS to avoid interference and jamming of the GPS signals [Deshpande,

2004].

AGPS-capable cell phones will be subjected to RFI from other sources which can

cause degradation of GPS accuracy and reliability. Higher order harmonics of AM and

FM radiobroadcast transmitters emissions fall close to the GPS L1 frequency (1575.42

MHz), which can potentially cause interference. With an AM broadcast, the harmonic

order is very high (985) and the likelihood of RFI is minimal. However, for an FM

broadcast, the harmonic order is lower (15 to 18) and the maximum effective isotropic

radiated power (EIRP) is higher (50 to 60 dBW). Analog TV broadcast maximum EIRP

limits are higher than for FM, while harmonic orders are lower (2 to 9 for RFI signals

within 2 MHz of GPS L1) and will cause more interference [Erlandson and Frazier,

2002].

Buck and Sellick [1997] analyzed the effects of the harmonics of the TV signals

interfering with GPS frequencies and they were found to be in the L1 signal spectrum

causing a non-linear effect. The strongest suspected interference signal was at

525.25 MHz which is the video carrier of a local UHF TV station (Channel 23). Thus the

1575.75 MHz signal was the third harmonic of the local station video carrier. The GPS

L1 frequency divided by three is 525.14 MHz and the transmitted TV signal's lower side

band suppression was at 524.50 MHz thereby allowing full power at this frequency. This

jump in power will produce a high level of interference resulting in a reduced SNR.

Page 104: UCGE Reports Number 20231

88

Filters were found to be effective in eliminating these interference signals by having high

attenuation for the undesired signals. The TV and Air Traffic Control (ATC) frequencies

have high transmitter powers and their harmonics fall in the GPS L1 frequency band. The

best protection for a GPS receiver is to use RF filtering to exclude the unwanted

interference. Spurious transmissions from RF transmitters in the GPS frequency band

should be measured to allow its suppression [Johannessen et al., 1990].

Previously, research has been done on the effect of interference on GPS

performance. Betz [2000] developed expressions to describe the effect of narrowband

interference on code tracking accuracy and C/N0 ratio, which shows that interference at a

frequency mid-way between the carrier and the first null has the greatest overall effect.

The expressions depend on the early-late spacing, the integration time, the unjammed

carrier-to-noise density ratio, and the tracking loop’s equivalent rectangular bandwidth.

Deshpande [2004] analyzed the interference effects on the acquisition process. RF

interference distorts the autocorrelation peak and leads to a false acquisition. However,

the power required to prevent or jam the acquisition process largely depends on the types

of interference. A relative CWI power of 15 dB is needed to jam this process while a

relative FM power of 35 dB is needed.

Other research was performed by Johnston [1999], Burns et al. [2002], and

Deshpande and Cannon [2004]. Most of that research was performed using a software

receiver and focused on acquisition. However, little research has been conducted to

investigate the effect of interference on AGPS receiver acquisition and tracking

performance.

Page 105: UCGE Reports Number 20231

89

5.2 Interference Effects

RFI has the same effect on GPS acquisition or tracking as signal blockage, foliage

attenuation, ionospheric scintillation and multipath, which is to reduce the C/N0 for all

the GPS signals. A jammer reduces the SNR of the GPS signals affecting acquisition and

tracking of the signals in the GPS receiver. Spoofing is another form of interference

which transmits a stronger version of the GPS signal to capture the receiver loops and

fool the receiver [Heppe and Ward, 2003]. Pseudolites operating at close range to a

receiver can jam the GPS receiver. The primary aspect of the GPS architecture that

makes it vulnerable is the low power of the signal which is actually below the noise floor

until it is de-spread with an appropriate PRN code. The RFI effect depends on the details

of the receiver design, especially the front-end bandwidth and early-late spacing in the

discriminator [Macabiau et al., 2001]. It has a different effect on the code tracking

accuracy than it does on some other aspects of the GPS receiver [Geyer and Frazier,

1999]. Several types of perturbations like thermal noise, atmospheric disturbances,

multipath and interference can affect the GPS signal. Geyer and Frazier [1999] conducted

tests on a C/A-code receiver for the Federal Aviation Administration (FAA) to determine

the vulnerability of the GPS receivers to RFI. This allowed the FAA to establish

interference standards for GPS receivers used in civil aviation. These tests were focused

on the C/A-code receiver’s tracking degradation and loss of lock under different

interference conditions. The GPS signal was found to be vulnerable to very high

frequency (VHF) transmissions and CW interference.

Page 106: UCGE Reports Number 20231

90

RFI detection should be given high priority because it provides an instantaneous

warning of the potential loss of GPS integrity. It can be detected using a jamming-to-

noise (J/N) power ratio meter [Kaplan, 1996]. The J/N meter is implemented in the AGC

of the GPS receiver front-end. This meter keeps a check on the thermal noise level and

any signal different from it, is detected as the presence of an interference signal.

The C/N0 for a satellite vehicle (SV) signal without interference is termed as

unjammed C/N0 [ibid]. The difference between the unjammed C/N0 and the acquisition or

the tracking threshold gives an indication of the possible interference tolerance and is

termed as effective C/N0. The unjammed C/N0 and the effective C/N0 are used to

compute the maximum jammer-to-signal (J/S) level at the receiver input from which the

RFI power can be determined. The unjammed C/N0 depends upon the GPS receiver

parameters and is computed as follows [Kaplan, 1996]:

Hz)-(dB L - Nf - (kTo) 10log- Ga Sr / 0 +=NC (5.1)

where

Sr is the received GPS signal power (dBW),

Ga is the antenna gain towards the SV (dBic),

10log(kTo) is the thermal noise density (dB-Hz) ≅ -204 dBW-Hz,

k is the Boltzmann’s constant (watt-sec/K) = 1.30 x 10-23,

To is the thermal noise reference temperature (K) = 290 K,

Nf is the noise figure of the receiver (dB), and

L is the implementation loss plus ADC loss (dB).

Page 107: UCGE Reports Number 20231

91

Signal information is lost during conversion of the signal from analog to digital

by the ADC which is referred to as the ADC loss. The level to which the unjammed C/N0

is reduced by the RFI is called the equivalent C/N0 power density ratio. The equivalent

C/N0 power density ratio is related to unjammed C/N0 and J/S as given by the following

equation [Kaplan, 1996]:

Ratio)Power ( -1(J/S)/QR)) -1((C/No) eq]/[ 0 +=NC (5.2)

where

C/N0 is the unjammed carrier-to-noise power in a 1 Hz bandwidth

expressed as a ratio,

J/S is the jammer-to-signal power expressed as a ratio,

R is the GPS PRN code chipping rate (chips/sec), which is 1.023x106

chips for the C/A code and 10.23x106 chips for the P code, and

Q is the spread spectrum processing gain adjustment factor, and is 1

for narrow band jammer, 1.5 for wide spread spectrum jammer and

2 for wideband Gaussian noise jammer.

Equation (5.2) can be expressed in terms of dB-Hz, which is shown in Equation

(5.3). This equation can be rearranged to obtain J/S [Kaplan, 1996]:

Hz)(dB]/10/)/(1010/)/(10[log10]/[ 0 −+−−= QRSJNoCeqNC (5.3)

(dB) ]10 - (10 [QR log 10 J/S ) -(C/No)/10)/10-([C/No]eq= (5.4)

Page 108: UCGE Reports Number 20231

92

For a C/A-code receiver, with a received signal strength, Sr = -159.6 dBW and

assuming the antenna has unity gain toward the SV (Ga = 0), a noise figure of 4 dB and

an implementation loss of 2 dB, then the unjammed C/N0 is 38.4 dB-Hz. For Q=2, and

assuming an equivalent C/N0 threshold of 28 dB-Hz, the J/S = 34.7 dB [Kaplan, 1996].

This tolerance looks good in terms of dB but when converted to the actual signal power,

it is just 3 pW. The RF transmitter transmits signals with high power levels (in terms of

Watts) and hence the harmonics of these signals can have power levels greater than 3

pW. This will result in jamming of the GPS receiver and hence RFI detection and

mitigation is important in a GPS receiver.

A number of techniques have been designed to increase the robustness of a GPS

receiver to RFI signals [Littlepage, 1999]. RFI can be mitigated at various stages of the

GPS receiver from the instant of receiving the GPS signals by the antenna to the position

computation stage. RFI signals will have full effect when the interference signal is

unobstructed and the antenna provides adequate gain to the signal.

5.3 Interference Tests

In the following sections, the effect of CW, AM and FM interference on AGPS

acquisition and tracking performance is investigated. The Agilent signal generator

(E4431B) was used to generate the various interference signals. The GPS signals

generated using GSS6560 and the interference signals are combined using an interference

combiner GSS4766. The interference combiner introduces a loss while combining the

input signals. This loss was measured and found to be around 8-10 dB for each channel.

Page 109: UCGE Reports Number 20231

93

This was taken into account to obtain the desired signal strength at the output of the

interference combiner.

5.3.1 CW Interference

Continuous wave interference was tested in an in-band range of the GPS L1 frequency.

Figure 5.1 shows the test setup used to conduct interference tests for various types of

RFI. All simulation tests in the interference section were designed with 10 to 12 satellites

in simulation with no orbital, atmospheric or any other errors. Only the effect of noise on

receiver performance under various types of RFI will be seen.

Figure 5.1: Interference Test Setup

A timing accuracy of 125 µs was provided to the AGPS receiver. The horizontal

uncertainty was set to 30 km while the vertical uncertainty was set to 50 m. The values

used for timing accuracy and the position uncertainties represent typical values used in an

AGPS implementation. Data on the AGPS receiver was collected at least 90 seconds after

Page 110: UCGE Reports Number 20231

94

the TTB was initialized to ensure that the TTB was able to obtain a position fix with at

least seven satellites. The TTB was given nominal signals, -130 dBm.

The GPS L1 frequency was used as the carrier frequency for the CW signals

analyzed. The interference power levels were varied from 0 to +50 dB relative to the GPS

signal power, which was set to -130 dBm. It should be noted that only the three SiRF

receivers were used for the acquisition tests while the OEM4 was used only for the

tracking tests. The elevation mask was set to 5°.

Table 5.3 shows the amount of CW interference power (relative to -130 dBm) that

each of the receivers is able to tolerate in acquisition and tracking. In acquisition, the

AGPS was able to tolerate 30 dB relative CW interference power, meaning it was able to

tolerate up to -100 dBm of CW interference power. Both the HSGPS and standard

receivers were able to withstand -110 dBm of CW interference power. AGPS was able to

tolerate 10 dB more CW interference than HSGPS and standard receivers. Better

performance from AGPS is expected since the acquisition sensitivity of the AGPS

receiver is higher than the HSGPS receiver.

Table 5.3: Acquisition and Tracking Threshold under CW Interference

Receiver Acquisition (dB) Tracking (dB)

AGPS 30 40

HSGPS 20 40

Standard 20 25

OEM4 N/A 40

Page 111: UCGE Reports Number 20231

95

In tracking, the AGPS and HSGPS receivers were able to track up to 40 dB

relative CW interference power. It is not surprising that the AGPS performance in

tracking is similar to the HSGPS since aiding provides “coarse” estimates intended to

assist acquisition. This “coarse” assistance is not useful in improving tracking

performance because when the receiver is in tracking mode, it has a much more precise

GPS time and location.

The Standard SiRF receiver was only able to track with 25 dB relative CW

interference power. However, the conventional geodetic OEM4 receiver was able to track

up to 40 dB, the same level as the AGPS and HSGPS receivers. In all cases, the tracking

threshold was at least 5 dB higher than the acquisition threshold. In the case of the

HSGPS receiver under CW interference, a 20 dB improvement in tracking over

acquisition was seen.

The 2D errors as a function of relative CW interference power for acquisition and

tracking tests are shown in Figure 5.2 and Figure 5.3. The 2D RMS error for different

interference power levels during acquisition and tracking is shown in Figure 5.4.

It should be noted that the position results were obtained by C3NAVG2TM post-

processing with a Horizontal DOP limit of 5 and Vertical DOP limit of 5. Dilution of

Precision (DOP) is a dimensionless number that accounts for the contribution of relative

satellite geometry to errors in the position determination. DOP has a multiplicative effect

on the user equivalent range error (UERE). All of the position results in the interference

section are obtained using the above criteria.

Page 112: UCGE Reports Number 20231

96

Figure 5.2: 2D Error vs. Relative CW Interference Power (Acquisition)

Figure 5.3: 2D Error vs. Relative CW Interference Power (Tracking)

Page 113: UCGE Reports Number 20231

97

Figure 5.4: 2D RMS Error for each CW Interference Power Interval

During acquisition, all of the receivers were able to give a position within less

than 4 m (RMS) until 20 dB of interference power. The AGPS, which had the most

tolerance under CW interference, was able to acquire up to 30 dB but the position

accuracy degrades at the upper limits to approximately 8 m. This shows a trade-off

between interference tolerance and position accuracy. In addition, the acquisition success

at 30 dB for the AGPS receiver was approximately 25%.

In tracking mode, the OEM4 receiver provided the best position accuracy. This is

consistent with previous results; research performed by Kim [2005] has also found that

the OEM4 provided better position results than the SiRFXTrac receiver. Once again, at

the upper limits of each receiver, severe performance degradation was seen. In the case of

the AGPS receiver, the 2D RMS error was greater than 30 m. At 40 dB, more than 25%

of the errors had a 2D RMS error of over 30 m.

Page 114: UCGE Reports Number 20231

98

5.3.2 AM Interference

An AM signal is a CW signal whose amplitude varies as a function of the modulating

signal. AM is used widely for radio communications. The AM signal was tested with the

carrier frequency at the GPS L1 frequency and the modulating signal was set to 10 Hz

while the modulation depth of the AM signal was set to 50%. The modulation depth

determines the amount of modulation present in the signal with higher modulation depths

resulting in more noise [Deshpande, 2004].

The interference power levels were varied from 0 to +50 dB relative to the GPS

signal power, which was set to -130 dBm. Table 5.4 shows the acquisition and tracking

thresholds of each of the receivers used.

Table 5.4: Acquisition and Tracking Threshold under AM Interference

Receiver Acquisition (dB) Tracking (dB)

AGPS 30 40

HSGPS 25 40

Standard 20 25

OEM4 N/A 40 In acquisition, the AGPS was able to tolerate 30 dB relative AM interference

power, meaning it was able to tolerate up to -100 dBm of AM interference power. The

HSGPS receiver was able to withstand -105 dBm of AM interference power while the

standard receiver was able to acquire up to -110 dBm. The AGPS receiver was able to

tolerate 5 dB more AM interference than an HSGPS receiver and 10 dB more than a

conventional receiver. As in the case with CW interference, three of the receivers

Page 115: UCGE Reports Number 20231

99

(AGPS, HSGPS and geodetic conventional receivers) are able to track up to 40 dB

relative AM interference power.

The SiRF Standard tracking performance was comparably lower; it was only able

to tolerate 25 dB of AM interference. The better performance from the OEM4 compared

to the SiRF Standard, even though they are both conventional GPS receivers with similar

tracking performance in normal GPS conditions, is due to the fact that the OEM4 uses

narrow chip spacing and its tracking loops have a lower bandwidth. In all cases, the

tracking threshold was higher than acquisition threshold.

The 2D errors as a function of relative AM interference power while acquiring

and tracking are shown in Figure 5.5 and Figure 5.6. The 2D RMS error for different

interference power levels during acquisition and tracking is shown in Figure 5.7.

From Figure 5.6, one can clearly see that at the upper limits of each receiver, the

position accuracy degrades significantly. For the SiRF Standard receiver, at its tracking

threshold of 25 dB of AM interference power, the 2D errors exceed 50 m.

Page 116: UCGE Reports Number 20231

100

Figure 5.5: 2D Error vs. Relative AM Interference Power (Acquisition)

Figure 5.6: 2D Error vs. Relative AM Interference Power (Tracking)

Page 117: UCGE Reports Number 20231

101

Figure 5.7: 2D RMS Error for each AM Interference Power Interval

During acquisition under AM interference, all of the receivers were able to give a

position within less than 5 metres (RMS) up to 20 dB of AM interference power. As in

the case of CW interference, the AGPS receiver was able to acquire up to 30 dB relative

interference power but the position accuracy degraded at the upper limits, close to 25 m

of 2D RMS error. At 30 dB for the AGPS receiver, there is a large error of over 180 m

which resulted in a biased 2D RMS error. The HDOP at this epoch was 1.85 and four

satellites were used to compute the position. Removing that one large error, the 2D RMS

error becomes 8.1 m, which is similar to the acquisition results with CW interference.

In the case of tracking with the AGPS receiver, the 2D RMS error at 35 dB

appears to be greater than that of 40 dB. This is due to four epochs that affected the

results. When these four epochs are removed from the solution, the 2D RMS error for

Page 118: UCGE Reports Number 20231

102

AGPS at 35 dB is about 2.75 m, which is much lower than the 2D RMS error at 40 dB.

Significant position accuracy degradation is noticed with the SiRF Standard receiver at

25 dB. There were many large errors; more than 15% of the errors at 25 dB were over 30

m.

5.3.3 FM Interference

An FM signal is a continuous wave signal whose frequency varies as a function of the

modulating signal. FM signals are used for radio broadcasts in the 88-108 MHz

frequency range, audio in television and for cellular transmission at various frequencies

listed in Table 5.2. The signal level of these FM signals is very high and the high order

harmonics of FM signals in the GPS frequency band will have considerable power as

compared to GPS signal levels. This will result in interference from the FM signal that

needs to be mitigated.

The GPS L1 frequency was used as the carrier frequency for the FM signals

analyzed. The modulating frequency was kept at 10 Hz while the frequency deviation

was set to 1 MHz. The modulation frequency in the FM signal decides the rate at which

the frequency deviates from the centre frequency. A smaller modulation frequency has

less frequency variation and the interference signal appears like a CW frequency

[Deshpande, 2004]. The interference power levels were varied from 0 to +50 dB relative

to the GPS signal power, which was set to -130 dBm.

The acquisition and tracking threshold of each of the receivers used is shown in

Table 5.5. In acquisition, the AGPS was able to tolerate 10 dB more FM interference than

an HSGPS receiver and 15 dB more than a conventional receiver. In tracking, AGPS and

Page 119: UCGE Reports Number 20231

103

HSGPS receivers were able to tolerate up to a 40 dB relative AM interference power, the

same level as for CW and AM interference. The Standard receiver was only able to

tolerate up to 30 dB relative power while the OEM4 was able to tolerate up to 35 dB of

relative FM interference power. Once again, in all cases, the tracking threshold was

higher than the acquisition threshold.

Table 5.5: Acquisition and Tracking Threshold under FM Interference

Receiver Acquisition (dB) Tracking (dB)

AGPS 35 40

HSGPS 25 40

Standard 20 30

OEM4 N/A 35 As mentioned previously, C/N0 is the best measurable value of the signal quality

present at the input to a GPS receiver. It is an instantaneous measure of the ratio of the

carrier power present to noise power density measured per Hertz of bandwidth. Figure

5.8 and Figure 5.9 show the average C/N0 for all satellites and the relative FM

interference power during acquisition and tracking tests. In acquisition, all of the

receivers are able to tolerate up to 20 dB of relative interference power before the C/N0

decreased.

The 2D error as a function of relative FM interference power is shown in Figure

5.10 and Figure 5.11. The 2D RMS error for different interference power levels during

acquisition and tracking is shown in Figure 5.12.

Page 120: UCGE Reports Number 20231

104

Figure 5.8: Average C/N0 vs. the Relative FM Interference Power (Acquisition)

Figure 5.9: Average C/N0 vs. the Relative FM Interference Power (Tracking)

Page 121: UCGE Reports Number 20231

105

Figure 5.10: 2D Error vs. Relative FM Interference Power (Acquisition)

Figure 5.11: 2D Error vs. Relative FM Interference Power (Tracking)

Page 122: UCGE Reports Number 20231

106

Figure 5.12: 2D RMS Error for each FM Interference Power Interval

During acquisition under FM interference, all of the receivers were able to give a

position within less than 5 m (RMS) up to 20 dB of FM interference power. Once again,

the position accuracy degraded at the upper limits, similar to the results seen with CW

and AM interference. For example, the AGPS receiver was able to provide position

within less than 9 metres at 35 dB relative FM interference power.

In tracking mode, the position accuracy with FM interference was much better

than under CW or AM interference. All of the receivers provided position within an

accuracy of less than 5 m (RMS), even at the upper limits. The OEM4 receiver provided

the best position accuracy in tracking under FM interference.

Page 123: UCGE Reports Number 20231

107

5.4 Chapter Summary

In this chapter, the effect of CW, AM and FM in-band interference (close to the GPS L1

frequency) on receiver acquisition and tracking performance was analyzed. It was found

that the AGPS receiver was able to tolerate 5 to 10 dB more interference power than an

HS receiver and 10 to 15 dB more than a conventional receiver in acquisition mode. Both

AGPS and HS had similar performance while tracking. In all RFI cases, all the receivers

were able to tolerate more interference while tracking than in acquisition. In tracking, the

OEM4 receiver provided the best positioning accuracy under all RFI cases.

Page 124: UCGE Reports Number 20231

108

CHAPTER 6: User Dynamics on AGPS

In this chapter, the effect of user dynamics on AGPS acquisition and tracking will be

investigated. User dynamics increase the frequency uncertainty by about 1.46 Hz per

kilometre per hour [van Diggelen, 2001a]. As a result, user movement increases the

Doppler search range (by as much as ±300 Hz for high dynamic situations) in the

acquisition process and consequently, the TTFF should also increase. They also limit the

duration of the pre-detection integration time since the Doppler varies quickly for high

user dynamics. Aiding data provided by the cellular network server helps to reduce the

search space and may improve the acquisition performance under dynamic conditions of

an AGPS receiver compared to other GPS technologies such as HSGPS receivers.

6.1 Effect of Velocity

6.1.1 Acquisition

An acquisition test was performed to investigate the effect of different velocities

on receiver acquisition performance. For this test, the power level of each satellite in

simulator vehicles 1 and 2 was kept at -130 dBm. As shown in Figure 6.1 which

illustrates the test setup, simulator vehicle 2 was connected to the TTB (which is the

reference receiver that provides assistance to the AGPS receiver) and this vehicle was

stationary throughout the test.

Page 125: UCGE Reports Number 20231

109

Figure 6.1: Dynamics Test Setup

A 20-km trajectory as shown in Figure 6.2 was created. Initially, the vehicle

accelerated to a prescribed velocity over a 15 second interval and then maintained this

velocity. At this velocity, the simulated car followed the 20-km trajectory for the

remainder of the test. The acquisition test was started after the prescribed velocity was

reached. The velocity of vehicle 1 was set to one of five levels: 36 km/h, 72 km/h,

108 km/h, 180 km/h and 360 km/h. These dynamic scenarios represent different

velocities that may be experienced while driving. The 360 km/h is used to test the

sensitivity of the receiver as this would require a really fast automobile. The

specifications state that the AGPS receiver is able to operate up to 300 km/h while the

limit for the SiRFXTrac is listed as approximately 1850 km/h (515 m/s). It is worth

noting that the specifications are quite different for the two receivers despite being

designed from the SiRFstarII architecture.

Page 126: UCGE Reports Number 20231

110

Figure 6.2: Dynamics Test Trajectory

For the AGPS receiver, the precise timing accuracy provided by the TTB was set

to 125 µs. The horizontal uncertainty was set to 30 km. This uncertainty is suitable since

the entire trajectory is only 20 km. The vertical uncertainty was only set to 50 m for all of

the velocities tested since the height was not varied during the test (i.e. the test trajectory

was purely horizontal). The position of the simulated automobile was taken as the truth

trajectory.

At each velocity, at least 30 trials were conducted. For the dynamic test, a trial is

considered to be an acquisition from a cold start followed by 60 position fixes to give

sufficient samples for analysis. The receiver is then cold started to start a new trial. It is

assumed that with a cold start, ephemeris, almanac and an initial position is not known

prior to acquiring the GPS signal. The raw pseudorange data was extracted at 1 Hz from

each receiver and post-processed using the C3NAVG2TM software.

Figure 6.3 shows the TTFF values for different velocities for both the AGPS and

HSGPS receivers. All TTFF values have been normalized, meaning all values were

divided by the maximum TTFF value of all the trials in this particular test. The TTFF is

at least four times lower with the AGPS receiver over the HSGPS receiver. With the

Page 127: UCGE Reports Number 20231

111

HSGPS receiver, an increase in TTFF is seen as the velocity increases but the TTFF for

the AGPS remains constant. As expected, the aiding data provided to the AGPS receiver

helps to reduce the search space in acquisition and consequently, a TTFF improvement is

seen with AGPS over the HSGPS. The 2D position errors as a function of velocity for

both receivers are shown in Figure 6.4 and Figure 6.5.

Figure 6.3: Average TTFFs for Different Velocities

From Figure 6.4 and Figure 6.5, it is clear that as the velocity increases, the

position accuracy decreases. The AGPS and HSGPS receivers have similar performance.

In Figure 6.6, the AGPS position errors (latitude, longitude and height) for different

velocities are shown. From this figure, it is clear that the longitude errors are larger for

360 km/h compared to 180 km/h test. The height errors remain relatively the same since

the test trajectory was purely horizontal.

Page 128: UCGE Reports Number 20231

112

Figure 6.4: 2D Position Error for AGPS (Acquisition)

Figure 6.5: 2D Position Error for HSGPS (Acquisition)

Page 129: UCGE Reports Number 20231

113

Figure 6.6: AGPS Position Errors for 72 km/h, 180 km/h and 360 km/h

The velocity errors as a function of velocity for the AGPS receiver are shown in

Figure 6.7. It is clear that as the velocity increases, the velocity errors also increase.

Page 130: UCGE Reports Number 20231

114

Figure 6.7: AGPS Velocity Errors for Acquisition Test

6.1.2 Tracking

The above results indicate the effect of user dynamics on AGPS acquisition performance.

In order to determine the effect on tracking performance, another test was conducted. The

setup and procedure was similar to the one described above with the only exception being

that the receiver was not cold started between trials. Figure 6.8 shows the 2D position

error for AGPS and HSGPS receivers while tracking under different velocities.

Page 131: UCGE Reports Number 20231

115

Figure 6.8: 2D Position Error for AGPS and HSGPS (Tracking)

It is clear that as the velocity increases, the position accuracy decreases. This

trend is observed clearly with the AGPS receiver. However, with the HS receiver, the 2D

error results for 100 km/h do not seem to fit the general trend. This is due to a higher

latitude RMS for this case compared to all other tests. If 95% of the best data are used,

the increasing trend holds for all values. Overall, one can conclude that the 2D error

performance is similar for both AGPS and HSGPS receivers.

Figure 6.9 shows the velocity errors for the AGPS receiver while tracking under

different velocities. It is clear that as the velocity increases, the velocity errors also

Page 132: UCGE Reports Number 20231

116

increase. During tracking, the velocity errors appear to be less than the errors seen during

the acquisition test. For example, at 180 km/h, the velocity errors are 12 km/h during

acquisition while they are approximately 7.5 km/h while tracking.

Figure 6.9: AGPS Velocity Errors for Tracking Test

6.2 Effect of Acceleration

6.2.1 Acquisition

Another test was conducted to investigate the effect of acceleration on AGPS acquisition

performance. The test setup was similar to the one used in the velocity section. Once

again, the power level of each satellite in simulator vehicles 1 and 2 was kept

Page 133: UCGE Reports Number 20231

117

at -130 dBm. The acceleration of vehicle 1 was set to one of four levels: 0.5 m/s2, 1 m/s2,

2 m/s2, and 4 m/s2. These values were chosen to represent moderate accelerations that

may be experienced while driving a vehicle. For example, the acceleration of a vehicle

that goes from 0 to 100 km/h in 8.0 seconds is approximately 3 m/s2.

Figure 6.10 shows the average TTFF under different accelerations. The position

errors for the AGPS receiver for two accelerations (0.5 m/s2 and 4 m/s2) are shown in

Figure 6.11 while 2D RMS errors are shown in Table 6.1.

Figure 6.10: Average TTFFs for Different Accelerations

As in the case with varying velocity, the TTFF remains the same for the AGPS

receiver under different accelerations. For the HSGPS receiver, an increase in TTFF is

Page 134: UCGE Reports Number 20231

118

seen as the acceleration increases. It should be noted that one large TTFF (more than

twice the average) biased the 1.0 m/s2 HSGPS result. As expected, the aiding data

provided to the AGPS receiver helps to reduce the search space in acquisition and

consequently, a TTFF improvement is seen with AGPS over the HSGPS.

Figure 6.11: AGPS Position Errors for 0.5 m/s2 and 4 m/s2 (Acquisition)

From Figure 6.11, it can be seen that for the AGPS receiver, the errors increase as

the acceleration increases. Position errors larger than 50 m are indicated on the 50 m line,

Page 135: UCGE Reports Number 20231

119

at the top of the figure with red, green and blue dots. Clearly, there are many more errors

over 50 m for the 4 m/s2 case as compared to the 0.5 m/s2 case.

Table 6.1: 2D RMS Errors for Different Accelerations (Acquisition Test)

AGPS HSGPS Acceleration (m/s2)

2D RMS Error (m) 2D RMS Error (m) 0.5 16.1 19.1 1.0 26.9 34.0 2.0 75.3 74.6 4.0 359.3 606.1

As acceleration increases, the 2D RMS errors increase for both AGPS and

HSGPS receivers. In the case where the acceleration is 4.0 m/s2, the 2D RMS error for

both receivers is very large. This is due to the fact that at this acceleration, there were

many errors that were over 2000 m that biased the results. When large errors occurred,

only four satellites were used to compute the position and in most cases, two or three

satellites were rejected by the post-processing software due to range residual checking. If

95% of the best data is chosen for the 4.0 m/s2 case, the resulting 2D RMS error is 59.6 m

for AGPS and 70.8 m for HSGPS. For most of the accelerations tested, the position

accuracies for both AGPS and HSGPS receivers are similar.

6.2.2 Tracking

A test was conducted to investigate the effect of acceleration on AGPS tracking

performance. The test setup was similar to the one used in the acquisition section except

no cold starts were performed during the trial. Once again, the acceleration of vehicle 1

was set to one of four levels: 0.5 m/s2, 1 m/s2, 2 m/s2, and 4 m/s2.

Page 136: UCGE Reports Number 20231

120

The position errors for the AGPS receiver for two accelerations (0.5 m/s2 and

4 m/s2) are shown in Figure 6.12. The 2D RMS errors for both AGPS and HSGPS

receivers are shown in Table 6.2.

Figure 6.12: AGPS Position Errors for 0.5 m/s2 and 4 m/s2 (Tracking)

Page 137: UCGE Reports Number 20231

121

Table 6.2: 2D RMS Errors for Different Accelerations (Tracking Test)

AGPS HSGPS Acceleration (m/s2)

2D RMS Error (m) 2D RMS Error (m) 0.5 14.8 35.0 1.0 23.1 25.2 2.0 82.9 149.2 4.0 151.6 190.7

From Figure 6.12, it can be seen that for the AGPS receiver, the errors increase as

the acceleration increases. This was also the conclusion that was reached for the velocity

scenario in Section 6.1.2. There are many more errors over 50 m for the 4 m/s2 case as

compared to the 0.5 m/s2 case.

A trend similar to the performance of the AGPS under constant acceleration can

be seen for the HSGPS receiver as well. However, for the HSGPS receiver, a larger 2D

RMS error is seen for the 0.5 m/s2 as compared to 1.0 m/s2. If 95% of the best data is

chosen for the HSGPS receiver, the resulting 2D RMS error is 2.3 m for 0.5 m/s2 and

7.9 m for 1.0 m/s2.

6.3 Chapter Summary

In this chapter, the effect of user dynamics on AGPS acquisition and tracking was

investigated. It was found that for an AGPS receiver, increasing velocity had no effect on

TTFF performance but the position accuracy decreased. With an HSGPS, the TTFF

increased as the velocity increased and the position accuracy decreased.

Page 138: UCGE Reports Number 20231

122

When tested with different accelerations, the TTFF remained the same for an

AGPS receiver while it increased for an HSGPS receiver. In terms of position, the

accuracy of the position decreased for both AGPS and HSGPS receivers with increasing

acceleration, whether in acquisition or tracking mode.

Page 139: UCGE Reports Number 20231

123

CHAPTER 7: Conclusions and Recommendations

The FCC E-911 mandate, Location-Based Services, as well as personal and vehicular

navigation applications are driving the need for navigation capability in degraded signal

environments such as in urban areas and indoors. Since the position accuracy yielded by

GPS methods is better than other positioning technologies, most wireless carriers are

looking at AGPS as the solution to meet the FCC criteria.

In this thesis, a GPS RF signal simulator was used to assess the signal acquisition

and tracking capability of a representative AGPS receiver. Extensive testing was

performed to analyze the effect of assistance data on AGPS acquisition and tracking.

Furthermore, performance of AGPS under various types of RFI and user dynamics was

investigated.

7.1 Conclusions

The following conclusions can be drawn from the work presented in this thesis:

• The AGPS receiver provided a 13 dB improvement in acquisition sensitivity

compared to an HSGPS receiver such as the SiRFXTrac and a 20 dB

improvement over a conventional receiver. No improvement was noticed in

tracking performance between the AGPS and HSGPS receivers. It was concluded

that aiding provides “coarse” estimates intended to assist acquisition and this

“coarse” assistance is not useful in improving tracking performance.

• In terms of assistance data, many interesting results were observed with the AGPS

receiver. For weak signals, timing accuracy had a significant effect on TTFF, with

Page 140: UCGE Reports Number 20231

124

more accurate timing leading to lower TTFF values. However, there was a trade-

off between TTFF and position accuracy with more accurate timing leading to

larger positioning errors due to the inaccuracy of the first fix.

• With the initial position, as the user to reference distance increases, the TTFF also

increases. In terms on position uncertainty, it was found that if the horizontal

uncertainty was within the user to reference distance, the TTFF remains

unchanged. For a horizontal uncertainty higher than the user to reference distance,

the TTFF increases. No clear trend was observed in the position domain.

• As for position uncertainty, it was found that as the horizontal uncertainty remains

within the user to reference offset, there is no observable difference in TTFF.

Once again, no clear trend was observed in position accuracy.

• It was found that the AGPS receiver was able to tolerate 5 to 10 dB more

interference power than an HS receiver and 10 to 15 dB more than a conventional

receiver in acquisition mode. Both AGPS and HS had similar performance while

tracking. In all RFI cases, all the receivers tested (SiRFLoc, SiRFXTrac, SiRF

Standard and the OEM4) were able to tolerate more interference while tracking

than in acquisition mode.

• For an AGPS receiver, increasing velocity had no effect on TTFF performance

but the position accuracy decreased. The TTFF is at least four times lower with

the AGPS receiver over the HSGPS receiver. With an HSGPS, the TTFF

increased as the velocity increased and the position accuracy decreased. The

Page 141: UCGE Reports Number 20231

125

velocity errors increased with increasing velocity during both acquisition and

tracking.

• When tested with different accelerations, the TTFF remained the same for an

AGPS receiver while it increased for an HSGPS receiver. In terms of position, the

accuracy of the position decreased for both AGPS and HSGPS receivers with

increasing acceleration, whether in acquisition or tracking mode.

7.2 Recommendations for Future Work

The following recommendations are made to improve the work presented in this thesis:

• All of the tests were conducted without the addition of many errors including

multipath, which is a major cause of error in real-life situations especially for high

sensitivity and AGPS receivers. Testing needs to be conducted to assess the

impact of multipath on AGPS acquisition and tracking performance.

• The advantage of hardware simulator testing is that it easier to isolate the

variables of interest and perform multiple, repeatable tests. However, ultimately,

the users of AGPS will be using their cellular phones in the outside world.

Therefore, field tests should be performed to assist in the prediction of real-life

AGPS performance from the simulation tests. Some field test results investigating

the effect of aiding parameters on AGPS have been published in Singh et al.

[2005]. The field results are similar to the results found in this thesis. But more

field tests can be performed in terms of interference and with user dynamics.

Page 142: UCGE Reports Number 20231

126

• When performing testing, the environment was not taken into consideration. A

few tests should be performed in a variety of environments. Indoor signal

replication including the effects of multipath has been performed by Hu et al.

[2005]. This work involved collecting field data from an HSGPS receiver and

reproduced the results using a hardware simulator. Similar tests can be performed

with an AGPS receiver.

• The characterization of an AGPS system was done using only one AGPS receiver.

Conducting the tests performed in this thesis with another AGPS receiver may

lead to different results. However, the conclusions drawn from the research will

remain the same.

• The effect of vertical uncertainty needs to be investigated.

With the arrival of Galileo, the concept of Assisted GPS is moving

towards Assisted GNSS (A-GNSS). The Galileo system proposes several free signals,

E5a (1176 MHz) and E5b (1207 MHz) as well as a narrower signal based on BOC (1,1)

modulation in L1 with the same central frequency as GPS L1 [Monnerat et al., 2004].

The Galileo L1 signal offers many advantages particular for LBS applications. Firstly, the

signal power is higher for Galileo, about 5 dB above GPS L1. Secondly, the signal design

calls for a dataless channel which removes the coherent integration limitation of 20 ms

[ibid].

The first modernized GPS satellite, IIR-M, was launched in late September 2005.

It has the capacity to implement the new military signals as well as the second civilian

signal, L2C. One of the main advantages of the L2C is that it has 45 dB cross-correlation

Page 143: UCGE Reports Number 20231

127

protection compared to 21 dB for L1 C/A-code. Better cross-correlation properties help

L2 receivers reject narrowband interference signals [Fontana et al., 2001]. The most

inherent advantage of the modernized signals and Galileo is the increase in satellite

availability which will improve poor availability currently experienced in urban canyon

environments.

Other developments in indoor positioning include techniques that do not use GPS

methods. Chun et al. [2005] discusses the potential of using wireless local area networks

(WLANs) to achieve indoor positioning by using reference points to estimate errors

between access points and the user. Also, there are some new techniques related to using

digital television (DTV) signals in obtaining user location information. The Rosum TV-

GPS combines broadcast TV signals with GPS to provide reliable indoor/outdoor

coverage [Rabinowitz and Spilker, 2003]. Therefore, one can safely say that

improvements in outdoor/indoor positioning will continue, with or without the use of

GPS. In the near future, however, GPS is here to stay and will be augmented using

different methods including the above mentioned WLANs and television signals.

Page 144: UCGE Reports Number 20231

128

REFERENCES

Ashjaee, J. (1985), GPS Doppler Processing for Precise Positioning in Dynamic Applications, IEEE OCEANS, November 15-19, Volume 17 , pp. 195-203.

Betz, J.W (2000), Effect of Narrowband Interference on GPS Code Tracking Accuracy, Proceedings of ION NTM 2000, Anaheim, CA, January 26-28, pp. 16-27.

Biacs, B., G. Marshall, M. Moeglein and W. Riley (2002), The Qualcomm/SnapTrack Wireless-Assisted GPS Hybrid Positioning System and Results from Initial Commercial Deployments, Proceedings of ION GPS 2002, Portland, OR, September 24-27, pp. 378-384.

Boulton, P. (2002), GSS6550 Multi-Channel GPS/SBAS Simulator Product Specification, Issue 1.00, May 21.

Braasch, M.S. and F. Van Graas (1991), Guidance Accuracy Considerations for Real time GPS Interferometry, Proceedings of ION GPS 1991, Albuquerque, NM, September 9-13, pp. 373-386.

Bryant, R. (2004), Lessons Learnt in Assisted GPS, Proceedings of the 2004 International Symposium on GNSS/GPS, Sydney, Australia, December 6-8.

Bryant, R., S. Dougan and E. Glennon (2001), GPS Receiver Algorithms & System for Weak Signal Operation, Proceedings of ION GPS 2001, Salt Lake City, UT, September 11-14, pp. 1500-1510.

Buck, T. and G. Sellick (1997), GPS RF Interference via a TV Video Signal, Proceedings of ION GPS 1997, Kansas City, MO, September 16-19, pp. 1497-1502. Burns, J., C. Cutright and M. Braasch (2002), Investigation of GPS Software Radio Performance in Combating Narrow Band Interference, Proceedings of ION AM 2002, Albuquerque, NM, June 24-26, pp. 523-530.

Page 145: UCGE Reports Number 20231

129

Cannon, M.E., S. Skone, M.D. Karunanayake and A. Kassam (2004), Performance Analysis of Real-Time Canada-Wide DGPS Service (CDGPS), Geomatica, Vol. 58, No. 2, pp. 95-105.

Chansarkar, M. and L. Garin (2000), Acquisition of GPS Signals at Very Low Signal to Noise Ratio, Proceedings of ION NTM 2000, Anaheim, CA, January 26-28, pp. 731-737.

Choi, I.H., S. H. Park, D. J. Cho, S. J. Yun, Y. B. Kim and S. J. Lee (2002), A Novel Weak Signal Acquisition Scheme for Assisted GPS, Proceedings of ION GPS 2002, Portland, OR, September 24-27, pp. 177-183.

Chun, J.W., J. Chey, J.G. Lee, G. Jee and W. Kim (2005), WLAN Differential Indoor Positioning Using Reference Points, Proceedings of ION NTM 2005, San Diego, CA, January 24-26, pp. 1113-1117.

Cox, G. (2005), Personal Communication, SiRF Technology Inc., San Jose, CA.

Deshpande, S. (2004), Study of Interference Effects on GPS Acquisition, M.Sc. Thesis, UCGE Report No. 20199, Department of Geomatics Engineering, University of Calgary.

http://www.geomatics.ucalgary.ca/links/GradTheses.html

Deshpande, S. and M.E. Cannon (2004), Interference Effects on the GPS Signal Acquisition, Proceedings of ION NTM 2004, San Diego, CA, January 26-28, pp. 1026-1036.

Erlandson, R. and R. Frazier (2002), An Updated Assessment of the GNSS L1 Radio Frequency Interference Environment, Proceedings of ION GPS 2002, Portland, OR, September 24-27, pp. 591-599.

FCC Report (2003), Report and Order and Second Further Notice of Proposed Rule Making, Report No. FCC 03-290, Washington, DC.

http://hraunfoss.fcc.gov/edocs_public/attachmatch/FCC-03-290A1.pdf

Fontana, R., W. Cheung and T. Stansell (2001), The Modernized L2 Civil Signal, GPS World, September 2001.

Page 146: UCGE Reports Number 20231

130

Garin, L. J., M. Chansarkar, S. Miocinovic, C. Norman and Hilgenberg (1999), Wireless Assisted GPS-SiRF Architecture and Field Test Results, Proceedings of ION GPS 1999, Nashville, TN, September 14-17, pp. 489-497.

Garin, L. J., M. S. Phatak and H. Gehue (2002), Indoor GPS Test Methodology and Indoor Performance Tests, Proceedings of ION GPS 2002, Portland, OR, September 24-27, pp. 192-199.

Geyer, M. and R. Frazier (1999), FAA GPS RFI Mitigation Program, Proceedings of ION GPS 1999, Nashville, TN, September 14-17, pp. 107-113.

GPS World., http://www.gpsworld.com/ (Accessed November 12, 2005).

Grewal, M.S., R. Lawrence, A.P. Andrews (2001), Global Positioning Systems, Inertial Navigation, and Integration, John Wiley and Sons Publications, New York, NY, USA.

Harper, N., P. Nicholson, P. Mumford and E. Poon (2004), Process for Improving GPS Acquisition Assistance Data and Server-side Location Determination for Cellular Networks, Proceedings of the 2004 International Symposium on GNSS/GPS, Sydney, Australia, December 6-8

Heppe, S. and P. Ward (2003), RFI & Jamming and its Effects on GPS Receivers, Based on Communication Theory, Navtech Seminars Course 452, Dahlgren, VA, May 5-6.

Hofmann-Wellenhof, B., H. Lichtenegger and J. Collins (1994), Global Positioning System Theory and Practice, Third Revised Edition, Springer-Verlag, New York, NY, USA.

Hopfield, H.S. (1969) Two-Quartic Tropospheric Refractivity Profile for Correcting Satellite Data, Journal of Geophysics Res., 74(18), 4487-4499.

Hu, T., G. Lachapelle and R. Klukas (2005), Indoor GPS Signal Replication Using a Hardware Simulator, Proceedings of ION GNSS 2005, Long Beach, CA, September 13-16, in press.

ICD-GPS-200 (2003), Interface Control Document, Navstar GPS Space Segment and Navigation User Interface, ARINC Research Corporation, El Segundo, CA, January 14.

Page 147: UCGE Reports Number 20231

131

Jakab, A. J. (2001), Quality Monitoring of GPS Signals, M.Sc. Thesis, UCGE Report No. 20149, Department of Geomatics Engineering, University of Calgary.

http://www.geomatics.ucalgary.ca/links/GradTheses.html

Johannessen, R., S. Gale and M. Asbury (1990), Potential interference sources to GPS and solutions appropriate for applications to civil aviation, IEEE AES Magazine, vol. 5, Issue 1, January 1990, pp. 3-9. Johnston, H. (1999), A Comparison of CW and Swept CW Effects on a C/A Code GPS Receiver, Proceedings of ION GPS 1997, Kansas City, MO, September 16-19, pp. 149-158.

Kaplan, E.D. (1996), Understanding GPS: Principles and Applications, Artech House Inc., Norwood, MA.

Kim, N. (2005), Interference Effects on GPS Receivers in Weak Signal Environments, M.Sc. Thesis, Department of Geomatics Engineering, University of Calgary, in progress.

Kinnari, T. (2001), Accurate Time Transfer in Assisted GPS, M.Sc. Thesis, Tempere University of Technology, Finland.

Klukas, R. W. (1997), A Superresolution Based Cellular Positioning System Using GPS Time Synchronization, PhD Thesis, UCGE Report No. 20114, Department of Geomatics Engineering, University of Calgary.

http://www.geomatics.ucalgary.ca/links/GradTheses.html

Krasner, N.F, G. Marshall and W. Riley (2002), GPS Receiver and Methods for Processing GPS Signals, SnapTrack, United States Patent# 5663734.

Kubrak, D., C. Macabiau and M. Monnerat (2005), Analysis of a software-based A-GPS acquisition performance using statistical processes, Proceedings of ION NTM 2005, San Diego, CA, January 24-26, pp. 1082-1092.

Lachapelle, G. (2002), Navstar GPS: Theory and Applications, ENGO 625, University of Calgary, Calgary, AB.

Page 148: UCGE Reports Number 20231

132

Lachapelle, G., H. Kuusniemi, D. T. H. Dao and M.E. Cannon (2003), HSGPS Signal Analysis and Performance under Various Indoor Conditions, Proceedings of ION GPS 2003, Portland, OR, September 24-27, pp. 1171-1184.

LaMance, J., J. DeSalas and J. Jarvinen (2002), Assisted GPS: A Low-Infrastructure Approach, GPS World, May 2002.

Langley, R.B. (1993), The GPS Observables, GPS World, Volume 4, No. 4, pp. 52-59.

Leick, A. (1995), GPS Satellite Surveying, Second Edition, John Wiley & Sons Publication, New York, NY, USA.

Lin, D. M., J. B.Y. Tsui, L. Lee, Y. T. Liou and J. Morton (2002), Sensitivity Limit of A Stand-Alone GPS Receiver and An Acquisition Method, Proceedings of ION GPS 2002, Portland, OR, September 24-27, pp. 2515-2523.

Lipp, A. and X. Gu (1994), Cycle-Slip Detection and Repair in Integrated Navigation Systems, Proceedings of IEEE PLAN Symposium, 11-15 April, pp. 681 - 688.

Littlepage, R. (1999), The Impact of Interference on Civil GPS, Proceedings of ION AM 1999, Cambridge, MA, June 28-30, pp. 821-828.

Liu, Z. (2000), A Java-Based Wireless Framework for Location-Based Services Applications, M.Sc. Thesis, UCGE Report 20161, University of Calgary, Calgary, AB, Canada.

Macabiau, C., O. Julien and E. Chatre (2001), Use of Multicorrelator Techniques for Interference Detection, Proceedings of ION NTM 2001, Long Beach, CA, January 22-24, pp. 353-363.

MacGougan, G. D. (2003), High Sensitivity GPS Performance Analysis in Degraded Signal Environments, M.Sc. Thesis, UCGE Report No. 20176, Department of Geomatics Engineering, University of Calgary.

http://www.geomatics.ucalgary.ca/links/GradTheses.html

Page 149: UCGE Reports Number 20231

133

MacGougan, G. D., G. Lachapelle, R. Klukas, K. Siu, L. Garin, J. Shewfelt and G. Cox (2002), Degraded GPS Signal Measurements With A Stand-Alone High Sensitivity Receiver, Proceedings of ION NTM 2002, San Diego, CA, January 28-30, pp. 191 - 204.

Mindbranch (2005), The Canadian Market for Commercial Location Based Services, Yankee Group.

http://www.mindbranch.com/listing/product/R388-0465.html (Accessed Sept. 13, 2005)

Misra, P. and P. Enge (2001), Global Positioning System Signals, Measurements, and Performance, Ganga-Jamuna Press, Lincoln, MA, USA.

Moeglein, M. and N. Krasner (1998), An Introduction to SnapTrackTM Server-Aided Technology, Proceedings of ION GPS 1998, Nashville, TN, September 15-18, pp. 333-342.

Monnerat, M., R. Couty, N. Vincent, O. Huez and E. Chatre (2005), The Assisted GNSS, Technology and Applications, Proceedings of ION GNSS 2004, Long Beach, CA, September 21-24, pp. 2479 – 2488.

Paddan, P., P. Naish and M. Phocas (2003), GPS Radio IP Design for Cellular Applications, GPS World, February 2003, pp. 30-45.

Petovello, M., M.E. Cannon and G. Lachapelle (2000), C3NAVG2TM Operating Manual, Department of Geomatics Engineering, University of Calgary, May.

Parkinson, B.W., T. Stansell, R. Beard and K. Gromov (1995), A History of Satellite Navigation, Journal of Institute of Navigation, Vol. 42, Special Issue 1, pp. 109-164.

Peterson, R., R. Ziemer and D. Borth (1995), Introduction to Spread Spectrum Communications, Prentice Hall Inc., New Jersey, NY.

Rabinowitz, M. and J.J. Spilker (2003), A New Positioning System Using Television Synchronization Signals http://www.rosum.com/RosumCorpWhitePaper_A_New_Positioning_System_Using_Television_Synchronization_Signals.pdf

Page 150: UCGE Reports Number 20231

134

Ray, J. (2003), Advanced GPS Receiver Technology, ENGO 699.73 Course Notes, University of Calgary, Calgary, AB.

Saasamoinen, J. (1971) Atmospheric Correction for the Troposphere and Stratosphere in Radio Ranging of Satellite, International Symposium on the Use of Artificial Satellite, Henriksen (ed.), 3rd Washington, 247-251.

Singh, S., M. E. Cannon, R. Klukas and G. Cox (2005), Field Test Assessment of Assisted GPS and High Sensitivity GPS Receivers under Weak/Degraded Signal Conditions, Proceedings of ION GNSS 2005, Long Beach, CA, September 13-16, in press.

SiRF Technology Inc., http://www.sirf.com/ (Accessed April 12, 2005).

Shewfelt, J, R. Nishikawa, C. Norman and G.F Cox (2001), Enhanced Sensitivity for Acquisition in Weak Signal Environments Through The Use of Extended Dwell Times, Proceedings of ION GPS 2001, Salt Lake, UT, September 11-14, pp. 155-162.

Spilker, J.J. Jr. and B.W. Parkinson (1996), Overview of GPS Operation and Design, Global Positioning System: Theory and Applications, Vol. I, American Institute of Aeronautics and Astronautics Inc., Washington, DC.

Stenbit, J. (2001), Global Positioning System Standard Positioning Service Performance Standard, Command, Control, Communications, and Intelligence, Department of Defense, Washingtion, DC.

Sudhir, N.S, C. Vimala and J.K. Ray (2001), Receiver Sensitivity Analysis and Results, Proceedings of ION GPS 2001, Salt Lake City, UT, September 11-14, pp. 1420-1426.

Syrjärinne, J. (2001a), Studies of Modern Techniques for Personal Positioning, PhD Thesis, Tempere University of Technology, Finland.

Syrjärinne, J. (2001b), Wireless-Assisted GPS Keeping Time with Mobiles, GPS World, January, pp. 22-31.

Taylor, R.E. and J.W. Sennott (1981), Navigation System and Method, NASA, United States Patent# 4445118.

Page 151: UCGE Reports Number 20231

135

Tsui, Y. (2000), Fundamentals of Global Positioning System Receivers: A Software Approach, John Wiley & Sons Inc., New York, NY.

Turetzky, G., L. Peng, R. Tso and J. Knight (1999), Catching the Wave: SiRF’s New Architecture Tackles GPS for Consumer Products, Proceedings of ION GPS 1999, Nashville, September 14-17, pp. 771-775.

van Diggelen, F. (2001a), Course 240B: Indoor GPS II, ION GPS-2001 Tutorial, Salt Lake, UT, September 11.

van Diggelen, F. (2001b), Global Locate Indoor GPS Chipset & Services, Proceedings of ION GPS 2001, Salt Lake, UT, September 11-14, pp. 1515-1521.

Ward, P. (1996), GPS Receiver Search Techniques, IEEE Proceedings of Position Location and Navigation Symposium, Atlanta, GA, April 22-26, pp. 604-611.

Watson, J. R. (2005), High-Sensitivity GPS L1 Signal Analysis for Indoor Channel Modelling, M.Sc. Thesis, UCGE Report No. 20215, Department of Geomatics Engineering, University of Calgary.

http://www.geomatics.ucalgary.ca/links/GradTheses.html

Weill, L.R., N. Kishimoto, S. Hirata and K. Chin (2004), The Next Generation of a Super Sensitive GPS System, Proceedings of ION GNSS 2004, Long Beach, CA, September 21-24, pp. 1924-1935. Wells, D.E., N. Beck, D. Delikaraoglou, A. Kleusberg, E.J. Krakiwsky, G. Lachapelle, R.B. Langley, M. Nakiboglou, K.P. Schwarz, J.M. Tranquilla, and P. Vanicek (1986) Guide to GPS Positioning, Canadian GPS Associates, Fredericton. Whitford, M. (2005), Got Game? GPS World, March 2005, pp 30-45.

Page 152: UCGE Reports Number 20231

136

Appendix A: Network-based Positioning Technologies

The FCC E-911 mandate and location-based services have been the major driving forces

for cellular network operators to locate mobile users. There are many ways to locate a

mobile terminal. These techniques can be categorized as network-based techniques,

handset-based techniques or hybrid techniques.

The network-based techniques carry out the measurements on the network side

and consequently, modifications are necessary. The techniques belonging to this category

include Cell-ID, Angle of Arrival (AOA), Timing Advance (TA) and Time of Arrival

(TOA); these methods are discussed here.

Cell-ID

The simplest method for locating the mobile terminal is the Cell-ID solution. When a

mobile phone is switched on, it makes contact with a Base Transceiver Station (BTS) in

its vicinity; therefore, all mobile phones are automatically located by the network with an

accuracy of the radius of the connected cell. Depending on the transmission power of the

BTS, the cell radius varies from a few hundreds of metres in urban areas to as large as 35

kilometres in rural areas for the GSM system [Silventoinen and Rantalainen, 1995]

Angle of Arrival

The Angle of Arrival (AOA) technique is based on the signal directions measured at

multiple BTSs. In order to determine the signal direction, an antenna array is needed at

Page 153: UCGE Reports Number 20231

137

the BTS. Therefore, major modifications at the BTS are required. However, smart BTS

antennas will be used in the future 3G system, AOA measurements may be available

without additional antenna array [Chen, 2002].

.

Timing Advance

The Timing Advance (TA) technique is based on the existing TA parameter in the GSM

network. The TA parameter, which is determined by the network, is a round trip

propagation delay of the signal between the mobile terminal and the serving BTS. The

serving BTS measures the round-trip delay of the access bursts carried in the Slow

Associated Control Channel (SACCH). The error of the distance derived from the TA

parameter is about ± 550 metres in the GSM system. Three TA parameters are needed in

order to determine a 2D-position. Therefore, the network has to perform two or more

positioning handovers in order to obtain the additional two or more TA measurements

from the neighbour BTSs. If the mobile phone is in idle mode, the network needs to

establish a call to the mobile terminal in order to determine the TA parameters [Chen,

2002].

.

Time of Arrival

Time of Arrival (TOA) positioning method is based on measuring the time of arrivals

from known uplink bursts (from MS to different BTSs). At least three measurements are

needed for positioning. TOA system needs a synchronous network system, which means

that TOA cannot be used in GSM without modifications. The TOA method used in GSM

is called Time Difference of Arrival (TDOA) [Kinnari, 2001].

Page 154: UCGE Reports Number 20231

138

The values needed in TDOA method are TOA measurements from a Location

Measurement Unit (LMU), real time difference (RTD) measurement, also from LMU,

and the coordinates of the BTSs. RTD values are needed to remove the time difference

between asynchronous BTSs in a manner similar to the way it is done in Enhanced

Observed Time Difference (EOTD). As a result, three TDOA measurements give two

hyperbolas, whose intersection is the position of MS.

When TOA measurement is needed, MS is forced to do an asynchronous

handover. In an asynchronous handover, MS sends 70 access bursts, which are used as

known bursts in TOA measurement, to its serving BTS and neighbour BTSs. The TOA

value is measured using all 70 access bursts to improve accuracy. The accuracy can also

be improved by removing multipath signals, which can be done, for example, using

LMU's multipath rejection techniques, antenna diversity and frequency hopping. The

accuracy of the TDOA positioning is close to accuracy in EOTD measurement,

approximately 50-100 m.

When an application wants to know the position of the MS, it sends a positioning

request to Serving Mobile Location Centre (SMLC). Depending on the accuracy level,

SMLC decides how many measurements are needed. SMLC receives the needed

information for positioning such as TOA and RTD measurements from LMUs and the

coordinates of the BTSs. With this information, SMLC calculates the position of MS and

sends it to the application [Syrjärinne, 2001a].