Ionospheric Scintillation Effects on
Global Positioning System Receivers
By
Mark Frederick Knight
Thesis submitted for the degree of
Doctor of Philosophy
Department of Electrical and Electronic Engineering
Faculty of Engineering
The University of Adelaide
Adelaide, South Australia
December 2000
ii
CONTENTS
ABSTRACT ................................................................................................................VII
DECLARATION .......................................................................................................... IX
ACKNOWLEDGMENTS .............................................................................................. XI
LIST OF FIGURES ................................................................................................... XIII
LIST OF TABLES ......................................................................................................XIX
ABBREVIATIONS .....................................................................................................XXI
LIST OF SYMBOLS ................................................................................................ XXIII
PUBLICATIONS ..................................................................................................... XXIX
1. INTRODUCTION ................................................................................................... 1
1.1.
MOTIVATION ............................................................................................................2
1.2.
THESIS OUTLINE AND CONTRIBUTIONS ......................................................................3
2. BACKGROUND ..................................................................................................... 5
2.1.
IONOSPHERIC SCINTILLATIONS ..................................................................................5
2.1.1. The ionosphere.....................................................................................................5
2.1.2. Morphology of scintillations................................................................................7
2.1.3. Statistical characteristics of scintillations ..........................................................10
2.1.4. Wide Band Scintillation Model .........................................................................15
2.1.5. Phase screen model ............................................................................................16
2.1.5.1. Deterministic phase screen .........................................................................17
2.1.5.2. Random phase screen .................................................................................18
i
2.1.6. Summary ............................................................................................................19
2.2.
GLOBAL POSITIONING SYSTEM ...............................................................................20
2.2.1. Principles of GPS positioning............................................................................20
2.2.2. GPS receiver tracking loops...............................................................................23
2.3.
A REVIEW OF SCINTILLATION EFFECTS ON GPS.......................................................25
2.3.1. Carrier tracking loops.........................................................................................26
2.3.2. Code tracking loops ...........................................................................................29
2.3.3. Codeless and Semi-Codeless receivers ..............................................................30
2.3.4. Navigation data ..................................................................................................30
2.3.5. Acquisition.........................................................................................................31
2.3.6. Optimum tracking of the carrier phase...............................................................32
2.3.7. Scintillation effects on navigation......................................................................32
2.4.
SUMMARY ..............................................................................................................33
3. CARRIER TRACKING LOOPS ............................................................................. 35
3.1.
CARRIER LOOP MODEL ............................................................................................36
3.2.
THE IMPACT OF PHASE SCINTILLATIONS ON CARRIER PHASE TRACKING LOOPS ........42
3.2.1. Phase tracking errors and thresholds..................................................................42
3.2.2. The effects of pre-detection filtering on phase errors ........................................49
3.2.3. Carrier phase range errors ..................................................................................51
3.2.4. Doppler errors ....................................................................................................54
3.2.5. Summary ............................................................................................................55
3.3.
THE IMPACT OF AMPLITUDE SCINTILLATIONS ON CARRIER PHASE TRACKING LOOPS 57
3.3.1. Background ........................................................................................................58
3.3.2. Phase errors from the linear model ....................................................................60
3.3.2.1. Amplitude scintillations only......................................................................60
3.3.2.2. Amplitude and phase scintillations.............................................................69
3.3.2.3. Amplitude scintillations and dynamics.......................................................72
3.3.2.4. Additional comments..................................................................................73
3.3.3. Phase errors from the non-linear model .............................................................74
3.3.4. The effects of pre-detection filtering on phase errors ........................................77
3.3.5. Summary ............................................................................................................79
3.4.
CARRIER LOOP TRACKING THRESHOLDS ..................................................................81
3.4.1. Optimum loop bandwidths.................................................................................85
ii
3.4.2. WBMOD predictions of T and S4......................................................................87
3.4.3. Velocity and elevation angle effects ..................................................................88
3.4.3.1. Elevation angle effects................................................................................88
3.4.3.2. Satellite and receiver velocity.....................................................................91
3.4.4. Summary............................................................................................................94
3.5.
THE IMPACT OF FADE DEPTH AND DURATION ON CYCLE SLIPS .................................96
3.5.1. 1st Order loops....................................................................................................99
3.5.1.1. Constant velocity ......................................................................................105
3.5.2. 2nd Order loops.................................................................................................106
3.5.2.1. Constant acceleration................................................................................108
3.5.3. Pre-detection filters..........................................................................................109
3.5.4. Summary..........................................................................................................110
3.6.
SCINTILLATION EFFECTS ON CARRIER PHASE DIFFERENTIAL GPS..........................111
3.7.
CARRIER FREQUENCY TRACKING LOOPS ...............................................................115
3.7.1. The impact of phase scintillations on frequency tracking loops ......................117
3.7.2. The impact of amplitude scintillations on frequency tracking loops ...............120
3.8.
CONCLUSIONS ......................................................................................................121
4. CODE TRACKING LOOPS ................................................................................. 123
4.1.
CODE LOOP MODEL ...............................................................................................123
4.2.
THE IMPACT OF PHASE SCINTILLATIONS ON CODE TRACKING LOOPS......................130
4.3.
THE IMPACT OF AMPLITUDE SCINTILLATIONS ON CODE TRACKING LOOPS..............132
4.3.1. Slow amplitude fluctuations ............................................................................138
4.4.
FREQUENCY-SELECTIVE SCINTILLATION EFFECTS..................................................143
4.5.
CONCLUSIONS ......................................................................................................152
5. CODELESS AND SEMI-CODELESS RECEIVERS ................................................ 153
5.1.
CODELESS PROCESSING TECHNIQUES ....................................................................153
5.2.
THEORETICAL ANALYSIS .......................................................................................155
5.3.
THRESHOLD CURVES.............................................................................................158
5.4.
SCINTILLATION MEASUREMENTS...........................................................................160
5.4.1. Overview of scintillation data..........................................................................160
5.4.1.1. Novatel Millennium data ......................................................................160
iii
5.4.2. A comparison of models with measurements ..................................................162
5.5.
CONCLUSIONS .......................................................................................................163
6. NAVIGATION DATA ......................................................................................... 165
6.1.
BACKGROUND ......................................................................................................165
6.2.
THE IMPACT OF PHASE SCINTILLATIONS ON NAVIGATION DATA .............................168
6.3.
THE IMPACT OF AMPLITUDE SCINTILLATIONS ON NAVIGATION DATA .....................170
6.4.
THE COMBINED EFFECT OF SCINTILLATIONS ON NAVIGATION DATA.......................175
6.5.
A NOTE ON WORD ERROR PROBABILITIES ..............................................................177
6.6.
CONCLUSIONS .......................................................................................................179
7. ACQUISITION ................................................................................................... 181
7.1.
ACQUISITION MODEL.............................................................................................181
7.2.
DETECTION AND FALSE ALARM PROBABILITIES .....................................................182
7.2.1. Phase scintillation effects.................................................................................187
7.2.2. Correlation sidelobes........................................................................................190
7.3.
ACQUISITION TIMES ..............................................................................................193
7.3.1. Acquisition search strategy ..............................................................................193
7.3.2. Mean time to acquire........................................................................................194
7.3.2.1. Amplitude correlation times .....................................................................196
7.3.2.2. Short amplitude correlation times.............................................................197
7.3.2.3. Long amplitude correlation times .............................................................200
7.3.3. Independence....................................................................................................204
7.3.4. False alarms......................................................................................................206
7.4.
CONCLUSIONS .......................................................................................................207
8. OPTIMUM TRACKING OF THE CARRIER PHASE ............................................. 209
8.1.
WIENER FILTER APPROACH ...................................................................................209
8.1.1. Causal Wiener filters........................................................................................210
8.1.2. Non-causal Wiener filters ................................................................................217
8.1.3. Doppler errors ..................................................................................................218
8.1.4. Optimum post-loop filters................................................................................222
8.2.
DIRECT DETERMINATION OF THE MMSE ..............................................................225
iv
8.2.1. Doppler errors ..................................................................................................227
8.3.
CONCLUSIONS ......................................................................................................229
9. SCINTILLATION EFFECTS ON NAVIGATION ................................................... 231
9.1.
PREDICTING THE PERFORMANCE OF A SINGLE LINK ...............................................231
9.2.
PREDICTING THE PERFORMANCE OF MULTIPLE LINKS ............................................233
9.3.
PREDICTING NAVIGATIONAL ACCURACY ...............................................................234
9.4.
PREDICTIONS BASED ON WBMOD .......................................................................238
9.5.
CONCLUSIONS ......................................................................................................245
10.
SUMMARY .................................................................................................... 247
10.1. OVERVIEW ............................................................................................................247
10.2. CONCLUSIONS ......................................................................................................249
10.3. FURTHER RESEARCH .............................................................................................252
APPENDIX A: SCINTILLATION MODEL ................................................................ 255
APPENDIX B: GPS TRACKING LOOP SIMULATORS ............................................ 265
APPENDIX C: TRACKING THRESHOLDS AND CYCLE SLIPS ................................ 269
APPENDIX D: THERMAL NOISE ERRORS ............................................................. 275
APPENDIX E: DOPPLER ERRORS ......................................................................... 285
APPENDIX F: IONOSPHERIC PIERCE POINT VELOCITY ...................................... 289
APPENDIX G: WBMOD PREDICTIONS OF FC ..................................................... 295
BIBLIOGRAPHY ...................................................................................................... 297
v
vi
Abstract
The Global Positioning System (GPS) is used extensively in both the military and civilian
communities for such diverse activities as navigation, surveying, remote sensing, asset
management and precise timing. The tremendous popularity of GPS has stemmed from
the low cost and small size of modern GPS receivers, and from the high accuracy and
reliability of the system. This second factor has also resulted in GPS being considered as a
sole means of navigation for critical safety of life applications such as precision approach
and landing for aircraft and narrow channel navigation for ships.
A number of environmental factors are known to affect the performance of GPS,
including electromagnetic interference, multipath, foliage attenuation, atmospheric
delays and ionospheric scintillations. In this thesis, the effects of ionospheric scintillations
on GPS will be examined.
Ionospheric scintillations are rapid variations in the amplitude and phase of
transionospheric radio signal resulting from density irregularities in the ionosphere.
Scintillations have the capacity to affect both the accuracy and reliability of GPS systems
by compromising the performance of the code and carrier tracking loops of a receiver. In
order to quantify this effect, a widely used stochastic model of scintillation activity is
combined with various tracking and acquisition models to produce a collection of
receiver performance measures. These include the magnitude of the code and carrier
range measurement errors, a measure of the tracking state of the carrier loop, the mean
time to acquire, and the bit error probability for the navigation data. An advantage of the
stochastic model chosen in this study is that it is linked to an existing predictive
scintillation model which is based on large amounts of scintillation data collected over the
previous 20 years or so. Consequently, by combining these models it is possible to predict
the performance of a given receiver type at any future time and location.
vii
viii
Declaration
This work contains no material which has been accepted for the award of any other
degree or diploma in any university or other tertiary institution and, to the best of my
knowledge and belief, contains no material previously published or written by another
person, except where due reference has been made in the text.
I give consent to this copy of my thesis, when deposited in the University of Adelaide
library, being available for loan and photocopying.
Signature ________________________________
Date ________________________________
ix
x
Acknowledgments
I wish to thank the following people for their assistance and support during my
candidature.
My academic supervisor Professor Doug Gray of the Electrical and Electronic Engineering
Department, the University of Adelaide, for his technical guidance, encouragement and
support over the years.
My supervisor, Dr Anthony Finn of Surveillance Systems Division (SSD), Defence Science
and Technology Organisation (DSTO) for providing me with an opportunity to pursue
my research objectives as part of my work commitments to DSTO.
The senior management of Surveillance systems Division for providing a framework in
which I could undertake further studies.
The Cooperative Research Centre for Sensor Signal and Information Processing (CSSIP)
for the use of their facilities and for the knowledge gained through the various seminars
and courses they have held.
My wife Lorraine for her patience and support over the last few years.
xi
xii
List of Figures
Figure 2.1-1: Typical daytime electron density profile for the mid-latitudes ........................ 6
Figure 2.1-2: Illustration of the equatorial Fountain Effect ...................................................... 7
Figure 2.1-3: Scintillation regions of the world .......................................................................... 8
Figure 2.1-4: Example of the scintillation indices produced by WBMOD ........................... 16
Figure 2.1-5: Modelled diffraction patterns based on a thin screen diffraction model and
an isolated Gaussian shaped irregularity ......................................................................... 17
Figure 2.1-6: Scintillation statistics for a random, Gaussian distributed density layer ...... 19
Figure 2.2–1: Architecture of a generic GPS receiver .............................................................. 23
Figure 2.2–2: Signal processing model of a generic code or carrier tracking loop .............. 24
Figure 3.1-1: Model of a generic Costas phase locked loop .................................................... 36
Figure 3.1-2: Non-linear baseband model of an I.Q Costas phase locked loop ................... 39
Figure 3.1-3: Linear baseband model of an I.Q Costas phase locked loop ........................... 39
Figure 3.1-4: Closed loop transfer function model of a Costas phase locked loop .............. 40
Figure 3.2-1: Threshold spectral strength as a function of the spectral index and the loop
noise bandwidth for second and third order loops ......................................................... 45
Figure 3.2-2: Threshold spectral strength as a function of the loop noise bandwidth for
second and third order loops .............................................................................................. 46
Figure 3.2-3: RMS phase scintillation errors as a function of spectral strength for a second
order Costas phase locked loop .......................................................................................... 47
Figure 3.2-4: Mean time between cycle slips for a second order Costas phase locked loop
in the presence of phase scintillations ............................................................................... 48
Figure 3.2-5: The effects of pre-detection filtering on the RMS phase scintillation error for
a second order Costas phase locked loop ......................................................................... 51
Figure 3.2-6: Variance of the carrier phase range error as a function of the loop noise
bandwidth for a second order Costas loop in the presence of phase scintillations .... 53
Figure 3.2-7: Variance of the Doppler error as a function of the loop noise bandwidth for a
second order Costas loop in the presence of phase scintillations ................................. 55
Figure 3.3-1: Phase error variance as a function of S 4 for a first order I.Q Costas loop with
an ideal AGC ......................................................................................................................... 64
xiii
Figure 3.3-2: Phase error variance as a function of S 4 from simulations for a first order I.Q
Costas loop with an ideal AGC ...........................................................................................64
Figure 3.3-3: Phase error variance as a function of S 4 for a first order I.Q Costas loop with
a fast AGC ..............................................................................................................................67
Figure 3.3-4: Phase error variance as a function of S 4 for a first order I.Q Costas loop with
a slow AGC ............................................................................................................................68
~
Figure 3.3-5: σ φ2εp A for a first order I.Q Costas loop with both a fast and a slow AGC ..70
()
Figure 3.3-6: σ φ2ˆ
p
as a function of S 4 for a first order I.Q Costas loop with both a fast and
a slow AGC ............................................................................................................................72
Figure 3.3-7: Non-linear model of a first order Costas loop with phase scintillations
translated back through the VCO and loop filter to the discriminator output.............75
Figure 3.3-8: σ ϑ2 and σ φ2T as a function of S 4 for a first order I.Q Costas loop with a fast
AGC ........................................................................................................................................77
Figure 3.3-9: The impact of pre-detection filtering on the power spectral density of
amplitude scintillations ........................................................................................................78
Figure 3.3-10: Scatter plot of the fade depth after pre-detection filtering versus the fade
depth before filtering from simulated scintillation data ................................................79
Figure 3.4-1: Normalised threshold amplitude as a function of the phase scintillation
spectral strength for both an ideal and a fast AGC ..........................................................83
Figure 3.4-2: The probability of losing lock for a 2nd order Costas phase locked loop as a
function of S 4 and the loop bandwidth ............................................................................84
Figure 3.4-3: Tracking threshold as a function of T, S 4 , and loop bandwidth ....................85
Figure 3.4-4: Threshold loop noise bandwidth as a function of the phase scintillation
spectral strength ....................................................................................................................86
Figure 3.4-5: T and S 4 values obtained from WBMOD for high solar activity ...................88
Figure 3.4-6: Mean and RMS of the geometry factors for T and S 4 w obtained from
WBMOD for a period of high solar activity.......................................................................89
Figure 3.4-7: Effective scan velocity as a function of elevation angle ...................................90
Figure 3.4-8: Ionospheric pierce point velocity due to satellite motion ................................92
Figure 3.4-9: Ionospheric pierce point speed as a function of elevation angle ....................93
Figure 3.5-1: Model of a phase locked loop for zero signal amplitude .................................97
Figure 3.5-2: Probability of a cycle slip as a function of the fade depth and duration for a
first order Costas loop based on theory with Bn = 15Hz ..............................................102
xiv
Figure 3.5-3: Probability of a cycle slip as a function of the fade depth and duration for a
first order Costas loop based on theory with Bn = 5Hz ................................................ 102
Figure 3.5-4: Probability of a cycle slip as a function of the fade depth and duration for a
first order Costas loop based on simulations with Bn = 15Hz ..................................... 103
Figure 3.5-5: Probability of a cycle slip as a function of the fade depth and duration for a
first order Costas loop based on simulations with Bn = 5Hz ....................................... 103
Figure 3.5-6: Probability of a cycle slip as a function of the fade depth and duration for a
first order Costas loop based on sims (Atan2(Q,I) and I.Q discriminators) .............. 104
Figure 3.5-7: Loop filter for a second order phase locked loop ........................................... 107
Figure 3.5-8: Probability of a cycle slip as a function of the fade depth and duration for a
second order Atan(Q/I) Costas loop based on simulations ......................................... 108
Figure 3.5-9: Signal amplitude after pre-detection filtering for a range of infinitely deep
fades with varying durations ............................................................................................ 110
Figure 3.6-1: σ ∆φ as a function of f o and the baseline length in the presence of phase
scintillations ........................................................................................................................ 113
Figure 3.6-2: RMS phase error as a function of S 4 and the loop bandwidth .................... 114
Figure 3.7-1: Representation of a generic frequency locked loop ........................................ 115
Figure 3.7-2: Linear equivalent model of a frequency locked loop ..................................... 116
Figure 3.7-3: Difference between the threshold spectral strength of a frequency locked
loop and the threshold spectral strength of a Costas phase locked loop ................... 119
Figure 4.1-1: Representation of a generic delay locked loop ................................................ 124
Figure 4.1-2: Delay error function of an Early-Late power discriminator ......................... 125
Figure 4.1-3: Linear model of a delay locked loop with an Early-Late power discrim .... 128
Figure 4.3-1: Delay error variance as a function of S 4 .......................................................... 140
Figure 4.4-1: Scintillation waveforms at the L1 frequency and 10 MHz above L1 under
frequency-selective scintillation conditions . .................................................................. 146
Figure 4.4-2: The impact of a single phase-changing screen with a power law in-situ
density profile on 0.0978 µs pulses that are bandlimited to ±10.23 MHz .................. 147
Figure 4.4-3: Figure 4.4-2 normalised by the peak pulse height .......................................... 147
Figure 4.4-4: The impact of a single phase-changing screen with a power law in-situ
density profile on 0.978 µs pulses that are bandlimited to ±1.023 MHz .................... 148
Figure 4.4-5: Early-Late gate error function for bandlimited 0.0978 µs pulses .................. 149
Figure 4.4-6: Figure 4.4-5 normalised by the peak discriminator error .............................. 149
Figure 4.4-7: Early-Late gate error function normalised by the peak discriminator error for
0.978 µs pulses ..................................................................................................................... 150
xv
Figure 4.4-8: Delay errors and the amplitude scintillation waveform as a function of
ground position for bandlimited 0.0978 µs pulses .........................................................151
Figure 4.4-9: Delay errors and the amplitude scintillation waveform as a function of
ground position for bandlimited 0.978 µs pulses ...........................................................151
Figure 5.2-1: The effect of a reduced amplitude threshold on the fade duration ..............158
Figure 5.3-1: Tracking thresholds for both codeless and semi-codeless receivers as a
function of T and S 4 ..........................................................................................................159
Figure 5.4-1: The percentage of time from 8:00pm to 10:00pm (local time) that the Novatel
Millennium loses lock on one or more satellites at Parepare, Indonesia ................161
Figure 5.4-2: The percentage of time that the L2 code and carrier loops lose lock as a
function of S 4 ......................................................................................................................163
Figure 6.2-1: of a bit error and a word error as a function of the phase error variance, σ φ2εp ,
for a first order loop ............................................................................................................170
Figure 6.3-1: Probability of bit and word errors as a function of S 4 for Bn = 2Hz ...........174
Figure 6.3-2: Probability of bit and word errors as a function of S 4 for Bn = 15Hz .........174
Figure 6.4-1: Probability of bit and word errors as a function of S 4 and σ φ2εp ..................176
Figure 6.5-1: Probability of a word error as a function of S 4 under the assumption that the
amplitude is independent between successive epochs .................................................178
Figure 7.1-1: A square-law acquisition detector for a GPS receiver . ...................................182
Figure 7.2-1: Pd as a function of S 4 for five values of C N o ..............................................186
Figure 7.2-2: Equivalent C N o as a function of S 4 for five values of C N o .....................187
Figure 7.2-3: Threshold T as a function of f o for a threshold variance of 1 rad2 ..............189
Figure 7.2-4: Pfa for a peak sidelobe level of -22dB as a function of S 4 ............................190
Figure 7.2-5: CDF and PDF of the sidelobe levels for the GPS Gold codes and the
corresponding Pfa ..............................................................................................................191
Figure 7.3-1: Mean acquisition time ratio as a function of S 4 ..............................................198
Figure 7.3-2: RMS acquisition time ratio as a function of S 4 ...............................................199
Figure 8.1-1: Wiener filter model of a phase locked loop .....................................................210
Figure 8.1-2: Phase error variance as a function of the spectral index for both causal and
non-causal Wiener filters ...................................................................................................218
Figure 8.1-3: Post-loop filtering schemes for loop phase estimates......................................223
Figure 8.2-1: Relationship between the two components of the mean square tracking error
as a function of the loop noise bandwidth ......................................................................226
xvi
Figure 8.2-2: Threshold values of T above which phase scintillation energy dominates
over dynamics in the choice of an optimum loop bandwidth ..................................... 228
Figure 9.3-1: Probability of a navigation or RAIM outage as a function of both the single
link probability of losing lock and the number of visible satellites ............................ 236
Figure 9.4-1: Percentage of links that may be stressed to the point of losing lock from
WBMOD. 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=135, K p =4⋅3, 10° mask
angle, C N o =41⋅5dBHz, Bn =15Hz, Loop order = 3, Coded L1 loop ......................... 239
Figure 9.4-2: Percentage of links that may be stressed to the point of losing lock from
WBMOD. 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=135, K p =4⋅3, 10° mask
angle, C N o =44dBHz, Bn =15Hz, Loop order = 3, Coded L1 loop ............................ 240
Figure 9.4-3: Percentage of links that may be stressed to the point of losing lock from
WBMOD. 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=135, K p =4⋅3, 10° mask
angle, C N o =41⋅5dBHz, Bn =2Hz, Loop order = 3, Coded L1 loop ........................... 241
Figure 9.4-4: Percentage of links that may be stressed to the point of losing lock from
WBMOD. 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=135, K p =4⋅3, 10° mask
angle, C N o =44dBHz, Bn =2Hz, Loop order = 3, Coded L1 loop .............................. 241
Figure 9.4-5: Percentage of links that may be stressed to the point of losing lock from
WBMOD. 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=135, K p =4⋅3, 0° mask
angle, C N o =41⋅5dBHz, Bn =15Hz, Loop order = 3, Coded L1 loop ......................... 242
Figure 9.4-6: Percentage of links that may be stressed to the point of losing lock from
WBMOD. 23 Oct. 2000, 12:00 noon UTC, 90th percentile, SSN=135, K p =4⋅3, 10° mask
angle, C N o =41⋅5dBHz, Bn =15Hz, Loop order = 3, Coded L1 loop ......................... 243
Figure 9.4-7: Percentage of links that may be stressed to the point of losing lock from
WBMOD. 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=100, K p =4⋅3, 10° mask
angle, C N o =41⋅5dBHz, Bn =15Hz, Loop order = 3, Coded L1 loop ......................... 243
Figure 9.4-8: Percentage of links that may be stressed to the point of losing lock from
WBMOD. 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=135, K p =4⋅3, 10° mask
angle, C N o =41⋅5dBHz, Bn =0⋅2Hz, Loop order = 3, Semi-codeless L2 loop ........... 244
Figure 9.4-9: Percentage of links that may be stressed to the point of losing lock from
WBMOD. 23 Sept. 2000, 12:00 noon UTC, 65th percentile, SSN=135, K p =4⋅3, 10° mask
angle, C N o =41⋅5dBHz, Bn =0⋅2Hz, Loop order = 3, Semi-codeless L2 loop ........... 244
Figure A-1: Geometry of the thin phase screen diffraction model ...................................... 258
xvii
Figure A-2: Sample realisation of in-situ phase, phase scintillations, and amplitude
scintillations obtained from the Fresnel-Kirchoff diffraction model ..........................262
Figure B-1: Simulink model of a Costas phase locked loop ..............................................266
Figure B-2: Simulink model of a delay locked loop ...........................................................266
Figure B-3: Simulink model of a combined Costas phase locked loop / Delay locked loop
channel .................................................................................................................................267
Figure C.1: 1σ phase error from all sources other than dynamics versus the steady state
dynamic phase error at the tracking threshold of an I.Q Costas loop ........................271
Figure G-1: f c as a function of elevation angle from WBMOD ..........................................296
xviii
List of Tables
Table 3.1-1: Costas loop discriminator functions .................................................................... 37
Table 3.1-2: Transfer functions and noise bandwidths of a phase locked loop .................. 41
Table 3.7-1: Frequency locked loop discriminator functions .............................................. 116
Table 4.1-1: Delay locked loop discriminator functions ..................................................... 124
Table 4.2-1: Conversion factors for code and carrier errors ................................................ 131
Table E.1:
Steady state tracking errors in the presence of dynamics .............................. 286
xix
xx
Abbreviations
ADO
Australian Defence Organisation
AGC
automatic gain control
AS
Anti-spoofing
BPSK
binary phase shift keying
C/A-Code
coarse/acquisition code
CDF
cumulative probability distribution function
CPDGPS
carrier phase differential GPS
CW
continuous wave
dB
decibel
DGPS
differential GPS
dip
dipole (Earth’s magnetic axis)
DLL
delay locked loop
DoD
Department of Defense
DOP
dilution of precision
DSSS
direct sequence spread spectrum
ECEF
Earth centred Earth fixed
FLL
frequency locked loop
GDOP
geometric dilution of precision
GPS
Global Positioning System
HDOP
horizontal dilution of precision
I
inphase signal
IF
intermediate frequency
IID
independent and identically distributed
INS
inertial navigation system
IPP
ionospheric pierce point
J/S
jammer to signal ratio
L1 and L2
GPS L-band carrier frequencies
NCO
numerically controlled oscillator
P-Code
precision code
P(Y)-Code
The encrypted precision code (the Y-Code)
xxi
PDF
probability density function
PDI
pre-detection integration
PDOP
position dilution of precision
PLL
phase locked loop
PPS
Precise Positioning Service
PRN
pseudorandom noise
PSD
power spectral density
PVT
position, velocity and time
Q
quadrature signal
RF
radio frequency
RMS
root mean square
SA
Selective Availability
SNR
signal to noise ratio
SPS
Standard Positioning Service
TEC
total electron content
UERE
user equivalent range error
UTC
coordinated universal time
VCO
voltage controlled oscillator
WADGPS
Wide Area Differential GPS
WBMOD
Wide Band Scintillation Model
WGS-84
World Geodetic System 1984
xxii
List of Symbols
Signal amplitude
A
~
A
Signal amplitude after the Pre-Detection Integrate & dump (PDI) filters
A
~
AN
Quiescent (unperturbed) signal amplitude
Filtered amplitude normalised by the quiescent amplitude, A
~
ATh
Amplitude threshold for carrier lock
Bn
Loop noise bandwidth
Bn
Design loop noise bandwidth
Bn
Optimum loop noise bandwidth
o
BI
Bandwidth of the PDI filters
Ck L
Height integrated irregularity strength
C No
Carrier to noise power density ratio
C No
Equivalent C N o for codeless and semi-codeless receivers
Eq
d
Navigation data
Erfc( x )
Complementary error function
F (s )
Open loop transfer function
Fo (s )
Optimum open loop transfer function
f IF , ω IF
IF carrier frequency in Hz and radians/s respectively
fn , ωn
Loop natural frequency in Hz and radians/s respectively
fn o , ωn
o
Optimum loop natural frequency in Hz and radians/s respectively
fo
Frequency corresponding to the ionospheric outer scale size
fc
Fresnel cutoff frequency of the amplitude scintillation power spectrum
f A (A )
Probability Density Function (PDF) of the amplitude, A
~
f A~ (A)
~
PDF of the amplitude after PDI filtering, A
~
f A~ (A N )
~
PDF of the normalised amplitude, AN
f I (I )
PDF of the signal intensity, I
fφ p (ϕ )
PDF of carrier phase scintillations, φ p
N
xxiii
fϑ (ϕ )
PDF of the modulo π reduced phase tracking error, ϑ
G (s )
Transfer function of the PDI filters
g
AGC gain factor
gN
Normalised AGC gain factor
H (s )
Closed loop transfer function
H o (s )
Optimum closed loop transfer function
hi
Height of the Ionospheric Pierce Point (IPP)
I
Signal intensity
IN
Normalised signal intensity
IE , IP, IL
Early, prompt and late I channel signals following the PDI filters
I n ( x)
Modified Bessel function of the first kind of order n
K X (τ )
Autocovarianec function of X
L
Loop order
lo
Ionospheric outer scale size
m
Parameter of the Nakagami-m distribution
N ( µ ,σ )
The Gaussian PDF with mean µ and variance σ 2
No
Power Spectral Density (PSD) of the thermal noise on the IF
NC
Number of cells in the uncertainty region for acquisition
n
IF thermal noise term
nd
Discriminator noise term
nd N
Discriminator noise term normalised by the AGC
n IE , n IP , n IL
Early, prompt and late I channel noise following the PDI filters
nQE , nQP , nQL Early, prompt and late Q channel noise following the PDI filters
Pd
Probability of detection for acquisition
Pd
Average probability of detection
Pfa
Probability of false alarm for acquisition
Pfa
Average probability of false alarm
Pe
Probability of a navigation data bit error
Pw
Probability of a navigation data word error
PL
Probability of losing carrier lock
PL
Average probability of losing carrier lock
PCS
Probability of a cycle slip
xxiv
p
Spectral index (slope) of the phase scintillation power spectrum
pE , pP , pL
Early, Prompt and Late PRN codes
QE , QP , QL
Early, prompt and late Q channel signals following the PDI filters
Q( x)
Gaussian probability integral
R X (τ )
Autocorrelation function of X
RE , RP , RL
Early, prompt and late code autocorrelation functions
rect( x )
Rectangle function
S4
Normalised (by the mean) standard deviation of the signal intensity
S4w
S 4 predictions obtained from the weak scintillation model
Sφ (s )
PSD of the carrier phase
Sφ d (s )
PSD of the dynamics component of the carrier phase
Sφ p (s )
PSD of the phase scintillation component of the carrier phase
S nd (s )
PSD of the discriminator noise term
Sτ (s )
PSD of the code delay process
sinc(x )
Sinc function (ie. sin (πx ) πx )
T
Period of the pre-detection integrate-and-dump filters
T
Spectral strength of phase scintillations at a frequency of 1 Hz.
T
Mean time to cycle slip
T ACQ
Mean time to acquire
Tc
PRN code chip width
TCT (x )
Correlation time of x
Td
Dwell time for acquisition
Tr
Time to re-visit a cell during acquisition
Tv
Verification time for acquisition
ve
Effective scan velocity of the propagation path through the ionosphere
vI
IPP velocity
vd
Irregularity drift velocity
vr
Component of the IPP velocity due to receiver motion
vs
Component of the IPP velocity due to satellite motion
zF
Radius of the first Fresnel zone
Γ(x )
Gamma function
Γ( x , y )
Incomplete Gamma function
xxv
γ
Slope of the delay error function (Section 4 only)
δ (t )
Dirac Delta function
ε T2
Total transient distortion produced by dynamics
η
Detection threshold for acquisition
ν
Carrier frequency (Hz)
ζ
Damping factor for a second order loop
ρe
The effective loop SNR
σ ACQ
Standard Deviation (STD) of the acquisition time
σn
STD of the thermal noise on either the I or Q channels
σ nd
STD of the discriminator noise term
σ nd
N
STD of the normalised discriminator noise term
σφ
STD of the carrier phase
σφp
STD of the carrier phase due to phase scintillations
σ φε
STD of the carrier phase tracking error
σϑ
STD of the modulo π reduced phase tracking error
σ φεp
STD of the carrier phase tracking error due to phase scintillations
σ φT
STD of the carrier phase error due to thermal noise
σ τT
STD of the delay error due to thermal noise
σ φˆ
STD of the carrier phase range error due phase scintillations
p
σ ωεp
STD of the Doppler error due phase scintillations
σ φ2ε
Th
Linear model tracking error threshold
σ φ2T
Th
σ φ2ε
o
Component of the tracking error threshold associated with thermal noise
Minimum phase error variance obtained by an optimisation process
τ
Fade duration (Section 3) or code delay (Section 4)
τd
Component of the code delay due to dynamics
τp
Component of the code delay due to phase scintillations
τo
Component of the code delay due to other sources
τε
Code loop delay tracking error
τˆ
Code loop delay estimate
φ
Carrier phase process
xxvi
φd
Component of the carrier phase process due to dynamics
φp
Component of the carrier phase process due to phase scintillations
φo
Component of the carrier phase process due to other phase noise
φˆ
Carrier loop phase estimate
φε
Carrier loop phase tracking error
φεr
Carrier phase range error
φˆp
Contribution to the carrier phase range error from phase scintillations
φεSS
Steady State carrier loop phase tracking error
φε
The maximum phase error for linear operation of the discriminator
T
ϑ
Carrier phase error reduced modulo π
Θ
Position step (radians)
Ω
Velocity step (radians/s)
Λ
Acceleration step (radians/s2)
χ n2
Chi-squared distribution with n degrees of freedom
ω εp
Contribution to the Doppler error due to phase scintillations
∆
Single difference operator for carrier phase DGPS
∇∆
Double difference operator for carrier phase DGPS
⊗
Convolution operator
[X ( s )]pos
Laplace transform of x (t ) such that x (t ) = 0 for t < 0
[X (s )]+
Factorisation of X (s ) such that all poles & zeros are in the left half s plane
[X (s )]−
Factorisation of X (s ) such that all poles & zeros are in the right half s plane
xxvii
xxviii
Publications
• Knight, M. F., Gray, D. A., “Maximum Likelihood Estimation of Ionospheric Total Electron
Content Using GPS”, ISSPA 96 - Fourth International Symposium on Signal Processing
and its Applications, Gold Coast, Australia, 1996.
• Knight, M. F., Finn, A., “The Impact of Ionospheric Scintillations on GPS Performance”,
Proceedings of ION GPS-96, The Ninth International Technical Meeting of the Satellite
Division of the Institute of Navigation, Kansas City, Missouri, 1996.
• Knight, M. F., Finn, A., “The Effects of Ionospheric Scintillations on GPS”, Proceedings of
ION GPS-98, The Eleventh International Technical Meeting of the Satellite Division of
the Institute of Navigation, Nashville, Tennessee, 1998.
• Knight, M. F., Cervera, M., Finn, A., “A Comparison of Predicted and Measured GPS
Performance in an Ionospheric Scintillation Environment”, Proceedings of ION GPS-99,
The Twelfth International Technical Meeting of the Satellite Division of the Institute of
Navigation, Nashville, Tennessee, 1999.
• Knight, M. F., Cervera, M. A., Finn, A., “A Comparison of Measured GPS Performance
with Model Based Predictions in an Equatorial Scintillation Environment”, Proceedings of
the IAIN World Congress, San Diego, California, 2000.
• Knight, M. F., Finn, A., Cervera, M. A., Thomas, R., “The Performance of GPS in the
Presence of Ionospheric Scintillations”, Proceedings of the 4th International Symposium on
Satellite Navigation Technology & Applications, Brisbane, Australia, 1999.
• Cervera, M. A., Knight, M. F., “Time Series Modelling of Intensity and Phase Scintillation
at GPS Frequencies”, Proceedings of the International Beacon Satellite Symposium 97,
Soporon, Hungary, 1997.
• Knight, M. F., Finn, A., Cervera, M. A., “Ionospheric Effects on Global Positioning System
Receivers”, DSTO Research Report, DSTO-RR-0121, 1998.
• Knight, M. F., Finn, A., “The Performance of Global Positioning System Receivers in a
Combined Navigation Warfare / Ionospheric Scintillation Environment”, DSTO Research
Report, DSTO-RR-0151, 1999.
• Knight, M. F., Finn, A., “Information Document on the Combined Impact of MSS
Interference and Ionospheric Scintillations on GPS Receiver Performance”, International
Telecommunications Union, Radiocommunication Study Groups, Doc. AUS/6, 1999.
xxix
xxx
Chapter 1
Introduction
T
he NAVSTAR Global Positioning System (GPS) is a satellite based radio navigation
system that provides accurate three dimensional position, velocity and time
information globally and continuously under all weather conditions. The GPS system can
be conveniently divided into three segments; (i) the Space Segment which consists of the
GPS satellites, (ii) the Control Segment which comprises a network of monitor stations
and ground antennas, and (iii) the User Segment which consists of the GPS receivers and
associated systems. Because of the high accuracy, low cost and portability of GPS
receivers, applications for GPS are appearing in many different areas including air, land
and sea navigation, surveying, geodesy, and military applications to name a few.
The accuracy and reliability of GPS is a function of both system and environmental
factors. System factors are associated with the three GPS segments described above and
include errors in the satellite clock and ephemeris information, hardware channel biases,
satellite geometry effects and thermal noise errors. Environmental factors are associated
with propagation phenomena and include electromagnetic interference from external
sources, ionospheric effects (including those associated with both the quiescent and the
disturbed ionosphere), tropospheric delays, obscuration and multipath. Depending on the
circumstances, the most significant environmental factor can be the disturbed ionosphere.
For GPS, the principal manifestation of a disturbed ionosphere is ionospheric
scintillations.
Ionospheric scintillations are rapid variations in the amplitude and phase of
transionospheric radio signals resulting from density irregularities in the ionosphere.
Scintillations show strong diurnal, seasonal, geographic and solar cycle dependence being
at their most severe during the evening hours, in the months of the equinox, at equatorial
latitudes and during the years of solar maximum. As we are currently at solar maximum
1
(year 2000), it is expected that both the frequency and severity of scintillation activity will
remain at a high level over the coming year or so.
Scintillations affect GPS receivers at the tracking loop level and so have the potential to
disrupt all GPS systems, including both single and dual frequency receivers and both
stand-alone and differential systems. The errors introduced into the code and carrier
tracking loops of a GPS receiver result in an increase in range measurement errors and
under extreme conditions can lead to a complete loss of signal lock. Other effects include
an increase in the probability of errors within the GPS navigation data and an increase in
the time taken to acquire the GPS signal when a receiver is first turned on. However, as
scintillations are unlikely to affect all of the satellites in a receiver’s field of view
simultaneously, their impact on navigational accuracy will be through a degradation in
the geometry of the available constellation. Consequently, the coverage of both the
satellites and the irregularities, as well as the intensity of scintillation activity will all
contribute to the accuracy of the final position solution.
1.1. Motivation
GPS will become the primary navigation system for the Australian Defence Organisation
(ADO) and will be installed on all air, sea and land based platforms, as well as forming an
integral part of the guidance mechanisms of many weapons. The positional accuracy of
GPS affords the possibility of enhancing many ADO operations, including navigation,
precision approach and landing, logistic support, the management of assets, mine
warfare, and the targeting and guidance of weapons. In addition, GPS allows combined
operations between air, sea and land based forces to be executed with flexibility and
precision through the use of a common reference grid and precise position, velocity and
time information. The cost effectiveness, accuracy, reliability and convenience of GPS will
ensure that it becomes an essential part of most military systems, replacing existing, more
costly navigation systems. GPS has also found an enormous market in the civilian
community in such diverse areas as aircraft and marine navigation, surveying, remote
sensing, geodesy, geographic information systems (GIS), and precise timing.
The Surveillance Systems Division (SSD) of the Defence Science & Technology
Organisation (DSTO) has been tasked with the job of assessing the impact of
environmental factors on the performance of GPS systems. These factors include
electromagnetic interference, multipath, foliage attenuation, atmospheric delays and
2
ionospheric scintillations. The areas that may be affected by equatorial scintillations cover
nearly 50% of the earth’s surface and include regions such as Northern Australia and
South East Asia which are of considerable operational interest to the ADO. For this
reason, SSD has an interest in quantifying the effects of scintillations on GPS performance
and of developing tools to predict the occurrence of significant scintillation events. The
results obtained from this study will also be of interest to civilian users, particularly those
attempting to achieve high levels of accuracy in equatorial regions. This is especially true
now that the United States Department of Defense has disabled Selective Availability1
which, until recently, has been the largest source of error for civilian GPS users.
1.2. Thesis outline and contributions
The principal contribution of this research is to use a widely accepted stochastic model of
scintillation activity to investigate the effects of scintillations on GPS receivers and
systems. This model has the advantage of being closely linked to the Wide Band
Scintillation Model (WBMOD2) which allows various statistical scintillation parameters to
be predicted for a given time, location and satellite-receiver geometry. The individual
contributions of this research can be summarised as follows:
Chapter 3
Expressions are derived for the variance of the carrier phase tracking error for a Costas
carrier tracking loop as a function of various amplitude and phase scintillation
parameters. These expressions are then used to determine the strength of scintillation
activity that will cause a carrier loop to lose lock. The sensitivity of the scintillation
parameters to geometrical factors such as the satellite elevation angle and satellite and
receiver motion are also discussed.
Expressions are also derived for the variance of the carrier phase range errors for a Costas
carrier tracking loop as a function of the scintillation parameters. These expressions are
then used to determine the errors that would be experienced by a carrier phase
differential GPS system over different baseline lengths.
1
Selective Availability or SA is an error introduced by the US Department of Defense to
intentionally degrade the accuracy of the civil GPS service.
2
The Wide Band Scintillation Model or WBMOD combines empirically derived models of the
global distribution and behaviour of ionospheric irregularities with a propagation model to predict
the characteristics of scintillations on a user specified transionospheric link.
3
Chapter 4
Expressions are derived for the variance of the code phase tracking errors and the variance
of the code phase range errors (pseudorange errors) for a code tracking loop as a function of
scintillation parameters.
Chapter 5
Expressions are derived for the effects of scintillations on codeless and semi-codeless
tracking loops. The predicted performance of a semi-codeless loop is then compared with
the measured performance obtained from a receiver located in the equatorial region
during a period of moderately high scintillation activity.
Chapter 6
Expressions are derived for the probability of a data bit error in the GPS navigation
message as a result of scintillations. The likely effects of these errors on the performance
of a receiver are also discussed.
Chapter 7
The effects of scintillations on acquisition are investigated for a square-law acquisition
detector, and expressions are derived for the mean time to acquire the GPS signal under
amplitude scintillation conditions.
Chapter 8
Optimum filters are derived for carrier phase tracking loops that minimise phase tracking
errors in the presence of scintillations. The effects of dynamics on the structure of these
optimum filters is also discussed.
Chapter 9
The problem of translating predictions of the impact of scintillations on individual
satellite-receiver links to predictions of navigational accuracy are discussed. The utility of
WBMOD for predicting the effects of scintillations on GPS and its limitations are also
discussed and some examples are given.
Appendix A
A technique is described for generating simulated amplitude and phase scintillation data
using a model based on a single, thin diffracting screen containing randomly distributed
ionospheric irregularities. The simulated scintillation data obtained from this model is
used throughout the thesis to validate the theoretical results.
4
Chapter 2
Background
This chapter is divided into three main sections. In Section 2.1, the morphology of
ionospheric scintillations is discussed and various scintillation models are introduced. In
Section 2.2, a brief overview of the GPS system is given and a model of GPS receiver code
and carrier tracking loops is provided. Finally, in Section 2.3 relevant literature is
reviewed and the problems to be addressed in this thesis are identified.
2.1. Ionospheric scintillations
Ionospheric scintillations are rapid variations in the amplitude and phase of
transionospheric radio signals resulting from electron density irregularities in the
ionosphere. Scintillations are therefore intimately linked to the underlying physical
processes in the ionosphere that give rise to irregularities. In this section, these processes
will be described along with the morphology of the associated scintillations. The models
of scintillation activity that will be used in subsequent chapters will then be discussed,
including; (i) a stochastic model based on data from previous solar maxima which will
form the basis of most of the analytical work that follows, (ii) a model which combines
various climatological irregularity models with a propagation model to produce
predictions of scintillation strength and occurrence, and (iii) a propagation model based
on a simple thin diffracting screen that enables simulated scintillation data to be
generated for various simulation tests.
2.1.1. The ionosphere
The ionosphere is a region of the upper atmosphere in which the density of free electrons
is large enough to have an appreciable effect on the propagation of radio waves [27], [80].
Although both the lower and upper boundaries of the ionosphere are not well defined,
for most practical purposes they can be considered to occur at roughly 50km and 1000km
5
respectively. Below this is the neutral atmosphere (the troposphere) and above is the
protonosphere that eventually merges with the interplanetary medium.
The ionosphere is formed by the ionising effects of solar X-ray and ultraviolet radiation
on neutral gasses in the upper atmosphere. As solar radiation penetrates the atmosphere,
its intensity is reduced through absorption while at the same time the density of the
atmosphere (and hence its capacity to produce ions) increases. Together, these two effects
lead to the formation of a region of maximum electron density, referred to as a Chapman
layer [27], at altitudes between about 250km and 400km. A typical daytime electron
density profile for a mid-latitude location is given in Figure 2.1-1. In this figure, it is
apparent that the ionosphere forms into several layers, the largest of these being the
F2-layer which extends from about 200 to 1000 km in height.
600
H eight (km )
500
400
F2
300
200
F1
100
E
D
0
10 9
10 10
10 11
10 12
10 13
Electron D ensity (e - /m 3 )
Figure 2.1-1: A typical daytime electron density profile for a mid-latitude location.
The peak density of the F2-layer undergoes large diurnal variations, reaching a maximum
at approximately 1400hrs local time and a minimum just before dawn. The height of the
F2-layer peak also shows a diurnal dependence, tending to fall at dawn and rise during
the evening hours. At low geomagnetic latitudes, the F2-layer height continues to rise
during the evening reaching a maximum height of approximately 500km at about 1900hrs
local time. This effect is due to an upward ExB force created by an Eastward electric field
in the E-layer that becomes enhanced soon after sunset. At these altitudes, the free ions
recombine very slowly after dusk and so the plasma density remains relatively high.
Under the influence of pressure gradients and gravity, the equatorial plasma in the
heightened F2-layer is forced downwards along the magnetic field lines, creating regions
of enhanced electron density at approximately 15 to 20 degrees either side of the
6
geomagnetic equator. These enhanced regions are referred to as the Equatorial Anomaly,
and the process by which they are created is known as the Fountain Effect (see for example
[49], [27] and the illustration in Figure 2.1-2). The Equatorial Anomaly is an important
phenomenon in the study of scintillations as it is responsible for the formation of the
plasma density irregularities that give rise to scintillations.
ExB
e-
B
e-
E
200N
Magnetic Equator
20oS
Figure 2.1-2: Illustration of the equatorial Fountain Effect which gives rise to the Equatorial
Anomaly. E and B represent the electric and magnetic field vectors respectively.
2.1.2. Morphology of scintillations
Scintillations occur predominantly in the equatorial band that extends from about 200S to
200N of the magnetic equator, and in the auroral and polar cap regions (see Figure 2.1-3).
The processes that produce scintillations in these two regions are quite different, leading
to significant differences in the characteristics of the resulting scintillations.
Auroral and polar cap scintillations are mainly the result of geomagnetic storms1 that are
associated with solar flares2 and coronal holes3. Unlike equatorial scintillations, they
show little diurnal variation in their rate of occurrence, and can last from a few hours to
many days, beginning at any time during the day [52]. Large and rapid variations in the
plasma density are often associated with auroral and polar cap scintillations [10] and can
lead to significant errors in differential GPS (DGPS) systems as well as rapid changes in
the apparent range and range rate [52] & [53]. Auroral scintillations also show a seasonal
dependence which is the reverse of that observed at low latitudes, being greatest from the
1
Large variations in the strength and direction of the Earth’s magnetic field.
2
Sudden increases in the intensity of solar electromagnetic radiation associated with sunspot activity.
3
Low density regions of the solar corona that are associated with solar winds (high energy charged particles
from the sun).
7
autumn equinox through winter to the spring equinox, and a minimum during the
summer months [2]. Indeed, the geomagnetic disturbances that excite auroral and polar
cap scintillations tend to suppress the onset of equatorial scintillations during solar
maximum periods [3], [27] & [53]. Because geomagnetic storm activity is linked to solar
activity through solar flares and coronal holes, auroral and polar cap scintillations also
show a strong dependence on the 11 year solar cycle, being most intense during solar
maximum periods, but almost non-existent during minima.
Polar Region
Mid-Latitude
Equatorial Region
Mid-Latitude
Polar Region
Figure 2.1-3: Map of the world showing the approximate locations of the polar, mid-latitude and
equatorial regions. Scintillations are mainly confined to the equatorial and polar regions.
Equatorial scintillations, on the other hand, are produced by irregularities in the F-layer
of the equatorial ionosphere following the passage of the evening terminator1 and tend to
disappear soon after midnight. In these regions, the most severe scintillations are
associated with the crests of the equatorial anomaly which are centred approximately 15°
either side of the magnetic equator [1]. As equatorial scintillations are coupled to the
anomaly, they tend to be worse during the years of solar maximum when the anomaly is
at its greatest. As we are currently at solar maximum (year 2000), it is expected that
scintillation activity will now be at its greatest, and will remain so for at least a year or so.
Equatorial scintillations also show a strong seasonal dependence, being greatest during
the months of April to August2 in the Pacific longitude sector, but a minimum during
these months in the American, African and Indian sectors. This situation is reversed
1
The terminator is the boundary that divides day from night.
2
Centred on the June Solstice.
8
during the months of September to March1 [53]. During the seasons of high scintillation
activity, the equinoctial months of March and September tend to suffer the highest levels
of activity, although this does not appear to be true at all longitudes [8].
Equatorial scintillations are mainly produced by irregularities created by instabilities in
the F-layer of the ionosphere during the evening hours. After sunset, the lower regions of
the F-layer recombine more rapidly than the upper regions, leading to an unstable
situation akin to a heavy fluid being supported on a lighter fluid2. This situation
eventually leads to the formation of bubbles of low density plasma which are forced
upwards through the denser upper regions. As the bubbles grow, steep density gradients
on the walls cause smaller irregularities to form [69]. At GPS frequencies, these smaller
irregularities, which can be of the order of the Fresnel zone radius or less (< 300m), are
responsible for scintillations [3]. The low density bubbles eventually form into
irregularity patches, or Plumes, which can reach heights of up to 1500km at the magnetic
equator. Once formed, the plumes extend along the magnetic field lines in a North-South
direction for over 2000km, leading to an accumulation of irregularities in the Northern
and Southern anomaly regions (±150 to ±200 dip latitudes3). Because of the higher
background densities in these regions, the irregularities tend to produce much stronger
scintillation effects than at the magnetic equator. Irregularity plumes typically have EastWest extents of between one and several hundred km’s and tend to move in an Easterly
direction with velocities of the order of 50 to 200 m/s [85]. Consequently, scintillations
are often experienced in patches that can last for an hour or so with periods of little or no
activity in between [3]. Eventually, in the absence of solar radiation, the irregularities
begin to fade along with the associated scintillation activity. This usually occurs around
local midnight, although at times scintillations can persist until early morning.
Scintillations can also occur during daylight hours and at mid-latitudes when Sporadic-E
is present in the E-layer. Sporadic-E are thin layers of highly dense plasma at heights of
about 100km in which large density gradients can exist. However, scintillations produced
by Sporadic-E are much less common and less predictable than those produced by the
F-layer processes described above.
1
Centred on the December Solstice.
2
This is referred to as a Rayleigh-Taylor instability.
3
“dip” refers to the Earth’s magnetic dipole or magnetic axis.
9
In the discussions that follow, only equatorial scintillations will be considered as they
affect the largest number of people and tend to be more severe than their auroral
counterparts [53]. In addition, the latitude band that is affected by equatorial scintillations
covers approximately 50% of the Earth’s surface, compared to only 7% for the auroral and
polar cap regions. However, it should be mentioned that during intense magnetic storms,
auroral disturbances can extend well into the mid-latitudes, disrupting GPS through both
scintillation activity and large density gradients. An example of this was the magnetic
storm in March 1989 during which auroral scintillation effects were felt over most of the
continental United States causing narrow bandwidth receivers to frequently lose signal
lock [52]. Such events are, however, quite uncommon.
2.1.3. Statistical characteristics of scintillations
Scintillations are produced by changes in the phase velocity of sections of a satellite signal
wavefront as it propagates through irregularities in the ionosphere. As absorption in the
ionosphere is negligible at L-band frequencies, the amplitude of the emergent wave is
unaffected by the irregularities. However, as the wave propagates towards the ground,
interference across the wavefront creates complex amplitude and phase diffraction
patterns that are a function of both the range to the irregularities and the cross-range
position. Scintillations are produced when these spatial diffraction patterns are
transformed into temporal ones, either through relative motion between the receiver and
the patterns, or by changes in the structure of the irregularities with time. Although
diffraction is the principal cause of scintillations, weak focussing and defocusing through
refraction can introduce additional amplitude and phase variations. However, for
refractive effects to be significant at L-band frequencies, the density gradients in the
ionosphere must be extremely large.
In this thesis, the effects of ionospheric scintillations are modelled as a complex
modulation of the unperturbed GPS signal. Furthermore, the phase and amplitude
components of this modulation are modelled as Wide Sense Stationary (WSS) stochastic
processes that are produced by a random distribution of irregularities of different sizes.
Consequently, they are defined in terms of their power spectral densities, probability
density functions and variances. Although a deterministic model based on a collection of
Gaussian shaped irregularities was also investigated [56], this approach was not taken
any further as the resulting waveforms were found not to represent the vast majority of
measured scintillation data.
10
The power spectral density of phase scintillations follows an inverse power law
relationship of the form [76], [35]
Sφ p ( f ) =
(
T
f o2
+f
)
2 p 2
radians2/Hz
(2.1-1)
where T is the magnitude of the power spectral density at a frequency of 1Hz (as f o << 1 ,
Sφ p ( f = 1) ≈ T ), f is the frequency of phase fluctuations,
f o is a frequency that
corresponds to the ionospheric outer scale size1, and p is termed the spectral index and
usually lies in the range 1 to 4, typically being 2.5 at equatorial latitudes [35]. The spectral
strength can be represented by the following expression [13], [76]
T ∝ G ve( p −1) λ2Ck L sec(θ )
(2.1-2)
where,
- G is a factor that depends on the direction of propagation of the radio wave and the
geometry and orientation of the irregularities,
- ve is the effective velocity of the propagation path through the contours of plasma
density,
- λ is the carrier wavelength,
- Ck L is referred to as the height integrated irregularity strength and is a measure of the
strength of the irregularity spatial power spectrum at a scale size of 1km and the
thickness of the irregularity layer, and
- θ is the off-vertical incidence angle of the propagation path at the irregularity layer.
The effective velocity, ve , is a function of the velocity of the ionospheric pierce point2
through the irregularity layer, v I , the drift velocity of the irregularities, vd (typically 50
to 200 m/s [85]), and the geometry and orientation of the irregularities. For GPS, the
pierce point velocity, v I , consists of a component due to receiver motion, v r , and a
component due to satellite motion, v s (typically 60 to 450 m/s, depending on the
elevation and azimuth angles of the satellites [9]). By changing v I through receiver
motion, the phase scintillation log spectrum will either be translated to the left or right,
depending on the direction of motion of the receiver in relation to the vector sum of vd
1
2
f o = ve lo where lo is the outer scale size (maximum irregularity size) and ve is the effective velocity.
The ionospheric pierce point (IPP) is the point at which the satellite signal ray path intersects the ionosphere
at the irregularity height. This is normally taken to be the height of the F2-layer peak which is roughly 400km.
11
and v s and the geometry of the irregularities. However, the spectral index, p, will remain
unchanged as it depends only on the spectrum of irregularity sizes in the ionosphere1.
Thus, receiver motion will shift the phase scintillation log spectrum along the frequency
axis, but will not alter its slope. This effect, in conjunction with the higher levels of carrier
loop stress experienced during receiver motion, has the potential to significantly alter a
receiver’s tolerance to scintillation activity.
The power spectral density of amplitude scintillations follows a similar power law
relationship for high fluctuation frequencies, but is heavily attenuated at low frequencies.
The cutoff frequency of the amplitude scintillation power spectrum (the Fresnel cutoff
frequency) is given by [103]
fc =
where z F =
ve
Hz
2 zF
(2.1-3)
λz1 z 2
is the Fresnel zone radius, and z1 and z 2 are the distances between
z1 + z 2
the ionospheric irregularity layer and the satellite and receiver respectively. Notice that
f c is also the frequency that corresponds to the peak of the amplitude scintillation power
spectrum [103]. Fresnel filtering occurs because amplitude scintillations are mainly
produced by diffraction effects which are only significant when the irregularity scale size
is of the order of the Fresnel zone radius. At typical ionospheric heights (~400km for the
F2-layer peak) and assuming vertical propagation, z F is of the order of 276m at the GPS
L1 frequency. For an irregularity drift velocity of 100m/s (a typical equatorial value) and
assuming v I = 0 , the Fresnel cutoff frequency is approximately 0.26Hz.
Two commonly used measures of the strength of scintillation activity are the RMS phase,
σ φ p , and the RMS intensity normalised by the mean, S 4 . The RMS phase is obtained
from the integral of the power spectral density of phase scintillations as follows
∞
σ φ p = E{φ p2 } =
∫ Sφ p ( f ).df
(2.1-4)
−∞
1
For moderate levels of scintillation activity, p is related to the slope of the one dimensional irregularity
spatial power spectrum, q , by p ≈ q + 1 . This approximation ignores the effects of diffraction on the phase.
12
where φ p is the carrier phase (assumed to be zero-mean). Consequently, σ φ p is a function
of T, p and the outer scale size parameter, f o . In practice, however, the integral in
Equation (2.1-4) is limited to some low frequency which is related to the coherent
integration time of the receiver. Indeed, as the statistics of scintillations are only
stationary for a few tens of seconds [35], this integral is unlikely to produce a realistic
result if 1 f o is much greater than a few tens of seconds.
The intensity scintillation index, S 4 , is the normalised RMS intensity and is given by
S4 =
E{I 2 }
E{I } 2
−1
(2.1-5)
where I = A 2 is the signal intensity. S 4 is also a function of T and p, but includes a
Fresnel filtering factor, F 1, and the Fresnel zone radius, z F , which together account for the
low frequency cutoff in the amplitude scintillation power spectrum. Under moderately
disturbed ionospheric conditions, S 4 can be approximated by the following expression
which is based on Rino’s weak-scatter theory [13], [76]
S 4 w 2 ∝ F z F ( p −1)λ2 C k L sec(θ )
∝
F z F ( p −1)
G v e ( p −1)
.T
(2.1-6)
The following expression can be used to derive an approximate value for S 4 under strong
scintillation conditions (assuming that refractive focusing effects do not occur) [13], [77]
S 4 = 1 − exp( − S 4 w 2 )
(2.1-7)
Both σ φ p and S 4 w show a simple dependence on the carrier frequency, ν [13], [27]. For
low to moderate levels of scintillation activity ( S 4 < 0.5), S 4 scales with the carrier
frequency as ν −( p+3)/ 4 . Under strong scintillation conditions, S 4 is approximately equal to
1 at all frequencies (unless focusing occurs which may drive S 4 slightly higher than 1).
The RMS phase, on the other hand, shows a ν −1 dependence for both weak and strong
1
The Fresnel filtering factor, F, is a function of the geometry and orientation of the irregularities, as well as
the spectral index, p.
13
scintillations, unless the scintillation activity is exceptionally strong. Consequently, the
GPS L2 frequency1 is affected slightly more by scintillation activity than the higher GPS
L1 frequency2 (by a factor of about 1.4 for S 4 and 1.3 for σ φ p ). This adds to the problem
of the inherently higher level of susceptibility of the L2 channel as a result of the lower
signal level of the L2 P-Code3 (the RMS carrier phase jitter resulting from thermal noise
on the L2 P-Code is
2 times greater than the carrier jitter on the L1 P-Code, and twice as
large as the carrier jitter on the C/A-Code for normal satellite signal levels).
Measurements of the probability density functions (PDF’s) of scintillations have shown
that phase scintillations follow a zero-mean Gaussian PDF, while amplitude scintillations
follow the Nakagami-m PDF [64]. Although other distribution functions have been
proposed for scintillations, the Gaussian / Nakagami-m distribution functions were
found to provide the best fit based on chi-square tests of observed intensity and phase
scintillation data [34], [101]. The Nakagami-m PDF is defined by the mean square
amplitude, 〈 A2 〉 , and by the m parameter which is a function of the strength of amplitude
scintillation activity, viz
f A (A) =
2m m .A 2m−1
Γ( m).〈 A2 〉 m
e − m. A
2 〈 A2 〉
, A≥0
(2.1-8)
where A is signal amplitude, Γ( ) is the Gamma function and m is a parameter that is
linked to the strength of scintillation activity by m = 1 S 42 [23]. The Nakagami-m PDF
approximates the Gaussian PDF for small values of S 4 , and becomes equal to the
Rayleigh PDF for S 4 = 1 (ie. for extremely strong scintillation activity). If the scintillation
statistics are assumed to be stationary, then by conservation of energy, and assuming no
absorption in the ionosphere, 〈 A2 〉 is independent of the strength of scintillation activity.
Phase scintillations follow a zero-mean Gaussian PDF and are therefore defined
completely by the variance, viz
fφ p (ϕ ) =
1
2π σ φ p
e
−ϕ 2 2σ 2
φp
1
The GPS L2 frequency is 1227.6 MHz.
2
The GPS L1 frequency is 1575.42 MHz.
3
The L2 P-Code is 3dB lower in power than the L1 P-Code. Refer to Section 2.2.
14
(2.1-9)
Unlike the Rayleigh or Rician fading models, the Nakagami-m model is not linked to the
phase distribution through analytical expressions (ie. the Nakagami-m PDF for amplitude
and Gaussian PDF for phase cannot be derived from an underlying signal model).
Consequently, little is known of the joint statistics of amplitude and phase scintillations,
although measurements suggest that they are negatively correlated with a correlation
coefficient of approximately –0.6 [34]. However, it is the correlation between the
amplitude and the rate of change of phase that is important in the study of tracking loop
behaviour. Again, little is known of this correlation, although it is likely that the deep
fades associated with large values of S 4 will be accompanied by rapid changes in the
carrier phase [76], [33]. This is expected to put tracking loops under more stress than
would be anticipated if amplitude and phase scintillations were considered to be
independent.
2.1.4. Wide Band Scintillation Model
The Wide Band ionospheric scintillation MODel (WBMOD [82]) is a global model of
ionospheric scintillation activity that enables users to predict the levels of scintillation
activity at a given time and location. The parameters provided by WBMOD include the
spectral index of phase, p, the spectral strength of phase, T, occurrence statistics and the
amplitude and phase scintillation indices, S 4 and σ φ p respectively. WBMOD consists of
two parts; (i) a collection of empirically derived models of the global distribution and
characteristics of ionospheric irregularities, and (ii) a power law phase screen propagation
model which allows the strength of scintillation activity to be calculated in a user defined
system. The propagation model assumes a spectral index of 2.5 at equatorial latitudes,
and calculates T from a series of eight parameters provided by the irregularity model
(based on Equation (2.1-2)). These parameters include the in-situ spectral slope, q = p − 1 ,
the height integrated irregularity strength, Ck L , the in-situ drift velocity of the
irregularities, vd , the phase screen height, three parameters describing the geometry and
orientation of the irregularities, and the outer scale size, lo . The only external inputs
required by WBMOD are the carrier frequency, the satellite and receiver locations, local
time and date, and solar and geomagnetic activity levels. In addition, the user must
decide between two types of output, namely (i) the percentage of time that a specified
level of scintillation activity is exceeded, or (ii) the level of scintillation activity associated
with a given percentile of occurrence.
15
An example of the output provided by WBMOD is given in Figure 2.1-4. The scintillation
indices σ φ p and S 4 are provided in the top panel, the spectral strength, T, is in the centre
panel and the spectral index, p, is in the lower panel. All plots are a function of latitude at
a longitude of 1200E and are at the 70th percentile (ie. the activity is expected to be
stronger than the specified level for only 30% of the time). This example represents a
period of high solar activity during the evening hours when the levels of scintillation
activity are expected to be at their greatest. The two humps at approximately 250N and
50S correspond to the crests of the Northern and Southern equatorial anomaly (at this
longitude, the magnetic equator is roughly 100 North of the geographic equator). Notice
that when amplitude scintillations, and therefore S 4 , are small, phase scintillations, and
therefore σ φ and T, are also likely to be small.
Scintillation Index
1
(a)
Solid − S4
Dashed − σφ (rads)
0.5
0
−20
−10
0
10
20
30
40
−10
0
10
20
30
40
0
10
20
Geographic Latitude (degrees)
30
40
10
(b)
−5
10
−10
10
Spectral Slope, p
Spectral Strength, T
0
−20
2.7
(c)
2.6
2.5
2.4
−20
−10
Figure 2.1-4: An example of the scintillation indices produced by WBMOD (April 10, 21:00 hrs
local time, longitude 1200E, 70th percentile, L2 frequency, R12=150, Kp=2, Phase stability 10s).
In subsequent chapters, the various statistical parameters produced by WBMOD will be
used to determine the performance of GPS receivers in a scintillation environment. By
combining these results with the occurrence statistics provided by WBMOD, predictions
can be made about the likely performance of a receiver at a given time and location.
2.1.5. Phase screen model
A simple model, based on Fresnel-Kirchoff diffraction theory, which demonstrates the
effects of a thin phase screen on a vertically propagating plane wave is given in
Appendix A. This model assumes that the plasma density irregularities are concentrated
16
within a thin layer or phase screen at a height that is typical of the F2-layer peak height
(approximately 400km). The resulting patterns of amplitude and phase variations on the
ground are then derived from the phase screen using simple diffraction theory. This
model provides an insight into the types of irregularities that are likely to produce
scintillations at L-band, as well as the characteristics of the resulting signals. It also allows
simulated scintillation data to be created for the tracking loop simulator described in
Appendix B.
2.1.5.1. Deterministic phase screen
In this section, the results obtained by modelling the irregularity layer as a series of
discrete Gaussian shaped lenses is discussed. At GPS frequencies, irregularities with scale
sizes of the order of the Fresnel zone radius1 or smaller are likely to produce the most
significant scintillation events. Larger irregularities produce very little amplitude
variation and only gradual phase variation, unless the density gradients are extremely
large. Very small irregularities (tens of metres or less) produce quite complex diffraction
patterns, but at an intensity which is too low to have a significant effect on GPS.
0
−5
5
Power (dB)
5
Power (dB)
Power (dB)
5
0
−5
0
−5
20
1
0.02
15
10
5
0
−1.5
−1
−0.5
0
Distance (km)
(a)
0.5
1
1.5
Phase (rad)
−10
Phase (rad)
−10
Phase (rad)
−10
0.5
0
−0.5
−1.5
0.01
0
−0.01
−1
−0.5
0
Distance (km)
(b)
0.5
1
1.5
−0.02
−1.5
−1
−0.5
0
0.5
1
1.5
Distance (km)
(c)
Figure 2.1-5: Modelled diffraction patterns based on a thin screen diffraction model and an
isolated Gaussian shaped irregularity (see Appendix A). Scale sizes of the irregularities are
(a) 1km, (b) 100m, and (c) 10m. The irregularities were centred at a height of 400km with a peak
density variation of 500% over the background.
In Figure 2.1-5, three examples are given of the diffraction patterns produced by an
isolated Gaussian shaped irregularity using the model described in Appendix A. From
this figure, it is apparent that the 100m irregularity, which is approximately one third the
size of the first Fresnel zone radius, produces the most significant amplitude variations
1
The Fresnel zone radius is approximately equal to 276m at the GPS L1 frequency and 312m at the GPS L2
frequency, assuming an irregularity height of 400km.
17
(centre panel). Panel (a) shows the effects of a very large irregularity (1km) and panel (c) a
very small irregularity (10m). Both of these will have much less of an impact on GPS
receivers than the 100m irregularity. In addition, the high frequency phase variations
associated with diffraction tend to be more pronounced when the irregularity size is of
the order of the Fresnel zone radius. For much larger irregularities, the phase variations
merely follows the in-situ density profile. Smaller irregularities produce prolific high
frequency phase variations but at a level that is unlikely to affect GPS. In general, because
the Fresnel zone radius at L-band frequencies is quite small, large density gradients are
required in order to produce significant scintillation effects. This tends to restrict
scintillation activity to the equatorial anomaly and polar regions where large density
gradients are known to exist.
2.1.5.2. Random phase screen
Although the effects of isolated irregularities on transionospheric signals has been
reported in the literature (eg [42]), in general irregularities occur in large numbers with a
range of sizes and densities (the spectrum of irregularity densities measured in-situ using
rockets has a power-law form). By providing a more realistic in-situ density profile for the
irregularity layer, the phase screen diffraction model will produce time series amplitude
and phase scintillation data which has more realistic statistics (Appendix A). In Figure
2.1-6, a realisation of a random density layer with a Gaussian PDF and an in-situ spectral
slope1 of 2 has been used in place of the deterministic phase screen from the previous
section. The wavenumber power spectrum of the vertically integrated density profile is
given in panel (a) along with a straight line representing a spectral slope of -2 (on a loglog scale). The low frequency cutoff at a wavenumber of approximately -31 dBmetres-1 is
produced by assuming an outer scale size, lo , of 1.3km for the irregularities. The power
spectra of the resulting phase and amplitude scintillations (panels (b) and (c) respectively)
also have a spectral slope of 2. However, the amplitude scintillations display a low
frequency cutoff, k c 2, at a wavenumber that corresponds to the Fresnel zone radius
(approximately –26 dBmetres-1 for an irregularity height of 400km). The phase
scintillation spectrum also shows evidence of Fresnel oscillations beginning at a
wavenumber of about –26 dBmetres-1. These appear as a series of nulls in the power
spectrum.
1
For the integrated density profile.
2
k c = f c ve = 1
18
2 z F from Equation (2.1-3).
The measured PDF’s of both the amplitude and phase are also consistent with the models
given in Section 2.1.3 (ie. Gaussian for phase and Nakagami-m for amplitude - panels (d)
and (e) respectively). The smooth curves overlying both of these plots are the theoretical
distribution functions obtained from the PDF expressions given in Section 2.1.3.
PSD of in−situ Phase, S4=0.4034, phasevar=0.2072
PSD of Amplitude, S4=0.4034, phasevar=0.2072
PSD of Phase, S4=0.4034, phasevar=0.2072
40
40
40
30
30
30
0
−10
Amplitude (dB)
10
Phase (dB)
Phase (dB)
20
20
10
0
−40
−40
−10
−35
−30
−25
−20
−15
−10
−5
0
−20
−45
10
0
−10
−20
−30
20
−20
−40
−35
Wavenumber (dBmetres^−1)
−30
−25
−20
−15
−30
−45
−10
−40
(a)
−35
−30
−25
−20
−15
−10
Wavenumber (dBmetres^−1)
Wavenumber (dBmetres^−1)
(b)
(c)
PDF of Phase, S4=0.4034, phasevar=0.2072
PDF of Intensity, S4=0.4034, phasevar=0.2072
0.9
1.2
0.8
1
0.7
0.8
p(Phase)
p(Intensity)
0.6
0.5
0.6
0.4
0.3
0.4
0.2
0.2
0.1
0
−1.5
−1
−0.5
0
0.5
Phase (radians)
(d)
1
1.5
0
0
0.5
1
1.5
2
2.5
Intensity
(e)
Figure 2.1-6: Scintillation statistics produced by a random, Gaussian distributed density layer
with an in-situ spectral slope for the integrated density of 2. Shown are the in-situ spectrum (a),
the phase scintillation power spectrum (b), the amplitude scintillation power spectrum (c), the
phase PDF (d) and the Intensity PDF (e).
2.1.6. Summary
In this section, the morphology and statistical characteristics of scintillations were
discussed. It was revealed that scintillations are generally restricted to specific times and
locations and that these can be predicted using models such as the Wide Band
Scintillation model. Based on studies of transionospheric scintillation data, it was decided
that scintillations can be modelled as a stochastic process in which the amplitude follows
a Nakagami-m distribution and the phase follows a zero-mean Gaussian distribution.
Furthermore, both the amplitude and phase can be assumed to have a power-law power
spectral density with a low frequency cutoff for the amplitude. A technique for
generating scintillation time series for simulation studies was also discussed. This
19
technique is based on a thin phase screen model for the irregularity layer and produces
scintillation statistics that are consistent with the stochastic model described above.
Details of the phase screen model are given in Appendix A.
2.2. Global Positioning System
GPS is a satellite based radio navigation system that provides accurate position, velocity
and time information globally and continuously under all weather conditions. Although
GPS is owned and operated by the US Department of Defense (DoD) and was developed
primarily for defence applications, it is now used widely in both the defence and civilian
communities in most countries. A convenient way of describing the GPS system is to
divide it into the following three segments:
(i) The Space Segment,
(ii) The Control Segment, and
(iii) The User Segment (the GPS receivers).
The Space Segment consists of 24 GPS satellites arranged in 6 orbital planes each of which
are inclined at 550 to the equator. The coverage provided by the GPS constellation ensures
that at least 4 satellites are visible at any time, anywhere on the Earth. As will be shown
later, this is an important requirement to ensure accurate three-dimensional positioning
by a receiver. The Control Segment consists of four monitor stations and four ground
antennas which are distributed around the Earth, and a master control station located in
Colorado Springs. The purpose of the Control Segment is to ensure that the Space
Segment is operating within specification, and to provide adjustments where necessary.
Communication between the Control Segment and the GPS satellites is via an S-band
uplink from one of the four ground station antennas. The User Segment consists of GPS
receivers, both military and civilian, and the associated infrastructure such as differential
stations.
2.2.1. Principles of GPS positioning
GPS receivers estimate three-dimensional position by solving four independent time
delay range measurement equations to four satellites in view. These equations can be
represented as follows:
ρ i = [x Si , y Si , z Si ] − [x R , y R , z R ] + c∆t
= Ri + c∆t
20
for i=1 to 4,
(2.2-1)
where [ x Si , y Si , z Si ] is the 3 dimensional position vector of satellite i with respect to the
centre of the Earth, [ x R , y R , z R ] is the position vector of the GPS receiver, c is the speed of
light, ∆t is the receiver clock offset from the satellite system time ( c∆t is usually referred
to as the receiver clock bias term), Ri is the true range to satellite i, and ρ i is referred to as
the Pseudorange to satellite i (assumed here to be error free). The four unknowns in the
range measurement equations are the x, y and z components of the GPS receiver position
vector (ie. the location of the receiver) and the receiver clock bias. The four pseudorange
measurements are found by measuring the delay in the propagation of the GPS signal
from four satellites to the receiver, and will usually be contaminated by a variety of error
sources. The four satellite position vectors [ x Si , y Si , z Si ] are obtained from the satellite
ephemeris information which is transmitted by each satellite as part of the GPS
navigation message1.
GPS pseudorange measurements are obtained by correlating the pseudorandom noise
(PRN) ranging codes transmitted by each GPS satellite with a replica code generated
within the receiver. The time delay, τ, which must be applied to the replica code in order
to achieve a correlation peak is related to the pseudorange, ρ, by ρ = cτ . A second
estimate of the satellite range can be obtained by integrating the carrier phase rate
measurements that are generated within the carrier tracking loops. However, although
the resulting carrier phase range measurements are relatively noise free compared to the
code measurements, they are subject to an unknown integer cycle ambiguity. Therefore, a
combination of these two measurements is often used in order to derive low noise,
unambiguous estimates of the satellite pseudorange.
The GPS signal is transmitted on two carrier frequencies, L1=1575.42 MHz and
L2=1227.6 MHz, each of which are bi-phase modulated by PRN ranging codes and GPS
navigation data. The PRN codes serve two purposes; (i) to create direct sequence spread
spectrum signals with good multiple access rejection and interference immunity, and (ii)
to enable the GPS receiver to measure satellite ranges by code correlation. Two PRN
codes are provided for this purpose, the Precision code or P-Code at 10.23 Mbits/s which
is modulated onto both GPS carriers, and the Coarse/Acquisition code or C/A-Code at
1
The navigation message is a 50 bits per second data stream that is modulated onto each satellite carrier and
includes information about the system time, clock correction factors, satellite health, and synchronisation
information for the military codes.
21
1.023 Mbit/s which is modulated onto the L1 carrier only. The US DoD reserves the right
to deny access to the higher accuracy available from the P-Code by encrypting it with a
second code referred to as the W-Code1. The resulting Y-Code is then only available to
authorised users who are equipped with the appropriate code decryption keys. This
process is known as Anti-spoofing (AS) and its principal function is to protect authorised
users from deceptive jamming (spoofing) by hostile forces. Prior to 1 May 2000, the US
DoD also deliberately degraded the accuracy available to unauthorised users by dithering
the satellite clocks and introducing small errors into the satellite ephemeris information.
This process was known as Selective Availability (SA) and again could only be removed
by users who had access to the code decryption keys. Although SA errors are currently
set to zero, it is nevertheless possible for the US DoD to resume SA in the future if the
need arises. In GPS terminology, the more accurate P(Y)-Code service is referred to as the
Precise Positioning Service (PPS), and the less accurate C/A-Code service is referred to as
the Standard Positioning Service (SPS).
GPS pseudorange measurements are contaminated by a number of errors, including
ephemeris errors, ionospheric delays, tropospheric delays, RF channel biases, multipath
and thermal noise. Expressions for the code and carrier phase pseudorange
measurements that include these errors are as follows:
Code:
Carrier:
ρ = R + c∆t + d TROPO + d IONO ( f ) + bS + bR + nTρ + n Mρ ,
φ = R + c∆t + d TROPO − d IONO ( f ) + bS + bR + nTφ + n Mφ + Nλ ,
(2.2-2)
(2.2-3)
where R is the true range to the satellite, c∆t is the receiver clock bias, d TROPO is the
tropospheric delay error, d IONO is the ionospheric delay error, bS and bR are the satellite
and receiver inter-channel biases (hardware biases), nT and n M are the thermal noise
and multipath errors, and Nλ is the cycle ambiguity in the carrier phase measurement.
Now that SA has been turned off, the principal source of error is likely to be the
ionospheric delay, d IONO . Under quiescent ionospheric conditions, the ionospheric delay
is proportional to the integrated electron density in the ionosphere (also called the Total
Electron Content or TEC), and inversely proportional to the square of the carrier
frequency, f (ie. d IONO = kTEC f 2 , where k is a constant [27]). As TEC is the same on
the two GPS carrier frequencies for a particular satellite to receiver link, a dual frequency
1
The chipping rate of the W-Code is 20 times less than the chipping rate of the P-Code.
22
receiver is capable of measuring, and therefore removing, ionospheric delays from
satellite range measurements. Consequently, authorised military users who are equipped
with P(Y)-Code receivers, and therefore have access to both carrier frequencies, will have
the capacity to eliminate ionospheric delays directly [81]. However, as the C/A-Code is
only modulated onto the GPS L1 carrier at present, unauthorised users are unable to
remove ionospheric delays in this way and must rely on a correction factor that is derived
from a broadcast ionospheric model [51]. In the future, modulation of the L2 carrier by
the C/A-Code and a possible second civil frequency (referred to as the L5 frequency) will
dramatically change this situation for civil users.
2.2.2. GPS receiver tracking loops
Figure 2.2–1 is an illustration of a generic GPS receiver in which the code and carrier
tracking loops are shown embedded between an RF front end stage1 and a navigation
processor. A typical GPS receiver contains many such tracking loop channels, each of
which consists of a Costas phase locked loop (PLL)2 for carrier tracking linked to a noncoherent delay locked loop (DLL) for code tracking (see for example [47] & [92]). In most
receivers, the code loop is also Doppler aided by the carrier loop to improve its
robustness to dynamics3.
Channel 1
Code
Loop
B.P.
Filter
LO
IF
τ%̂
%̂
φ
Carrier
Loop
Navigation
Processor
xˆ, yˆ, zˆ
vˆx, vˆy, vˆz
Nav. Data
Channel n
Figure 2.2–1: Architecture of a generic GPS receiver.
1
The RF front end typically consists of an antenna, a low noise preamplifier, a down-conversion mixer and an
image rejection filter.
2
As 50Hz navigation data remains on the carrier after removal of the code, the carrier loop must be capable
of tracking a suppressed carrier signal. Consequently, a Costas loop is used rather than an ordinary PLL.
3
Carrier aiding of the code loop removes virtually all of the line of sight dynamics from the code loop,
allowing the code loop bandwidth to be significantly reduced.
23
In Figure 2.2–1, the mixer at the front of the code tracking loop is driven by a replica
carrier from the carrier tracking loop and is responsible for down-converting the GPS
intermediate frequency (IF) signal to a baseband IQ pair. The mixer at the front of the
carrier tracking loop is driven by a replica code from the code tracking loop and is
responsible for removing the satellite PRN code from the IF carrier. Consequently, under
normal tracking conditions, the carrier loop receives an IF carrier which is modulated by
the navigation data, and the code loop receives a PRN code which is modulo 2 added to
the navigation data. Although other tracking techniques do exist (eg [44] & [71]), the
Costas PLL/DLL combination is by far the most common for GPS.
Both the code and carrier tracking loops can be represented by the signal processing
model illustrated in Figure 2.2–2. Here τ and φ represent the phases of the code and
carrier components of the GPS signal at the input to the tracking loop, and τ# and φ#
represent the phase estimates at the loop output. The loop contains two filters, a predetection filter G(s) which reduces the levels of thermal noise prior to the phase
discriminator, and a loop filter F(s) which controls the order and bandwidth of the
tracking loop. The phase discriminator is responsible for measuring the difference
between the input phase and the loop phase estimate and is in general a non-linear
%
device. The outputs of the tracking loop are Doppler estimates, τ%# and φ# , which are
integrated in the Navigation Processor to produce the code and carrier range estimates. In
addition, the navigation data is derived from the in-phase channel of the Costas loop at a
point immediately after the pre-detection filter G(s).
Data Detect
Signal + Noise
(φ or τ )
G(s)
Phase
Discriminator
Signal estimate
(φ# or τ# )
VCO
F(s)
φ%# or τ%#
+
Doppler Aiding
Figure 2.2–2: Signal processing model of a generic code or carrier tracking loop.
The function of the tracking loop is to provide estimates of the desired input phase
process while rejecting unwanted phase noise. For GPS, the desired phase process is the
Doppler introduced by satellite and receiver motion, while the phase noise is a
combination of thermal noise, multipath, oscillator phase noise and ionospheric
24
scintillations. Important design objectives for the tracking loops are to minimise the phase
noise on the Doppler estimates, and to minimise the tracking error between the input
phase and the estimated phase processes (ie. τ ε = τ - τ# and φε = φ − φ# ). The second
objective is associated with the ability of the loop to remain in phase lock and is probably
the most important of the two under strong scintillation conditions.
2.3. A review of scintillation effects on GPS
To date there has been little detailed work done in the area of scintillation effects on GPS.
The principal reason for this is that GPS was not declared to be fully operational until the
end of 1994 [70], which is several years after the last solar maximum. Consequently, the
effects of scintillations on a fully operational system are only now being observed. The
majority of the work described in the GPS literature has either focused on the
characteristics of the disturbed ionosphere or has dealt with the issue of GPS performance
at a very qualitative level (see for example [2], [5], [9], [11], [14], [29], [53]& [98]). Other
researchers have performed tests on GPS receivers using satellite signal simulators and
either simulated or real scintillation data [12], [24], [63] & [94]. This work has shown that
full code-correlation receivers are generally quite robust to moderate levels of scintillation
activity, but that pseudorange noise and occasional loss-of-lock can occur if the
scintillation activity becomes very strong. Nichols et al [66] correlated loss-of-lock events
for a codeless receiver1 with the scintillation statistics σ φ and S 4 for a Northern Auroral
region. Although this work has demonstrated that GPS receivers are indeed susceptible to
the effects of scintillations, again it has not been accompanied by any detailed analytical
studies. More recently, a number of researchers have begun to investigate the
performance of GPS receiver tracking loops using phase locked loop simulations and
simulated scintillation data. This has included simulation tests based on discrete
irregularity structures [56], as well as those based on stochastic models of scintillations
activity [41], [58] & [73]. Hegarty et al [41] used simulators for both the L1 code and
carrier tracking loops tracking loops to determine the tracking errors as a function of S 4
and the quiescent C N o . His results suggest that very narrow bandwidth code loops are
unlikely to be significantly affected by scintillations, but that carrier loops will suffer an
increase in measurement noise and a loss of continuous carrier lock. He also found that
codeless receivers will be significantly affected by scintillations and will even lose lock in
1
For a description of codeless receivers, see Chapter 5.
25
the presence of quite mild scintillation activity. The principal advantage of using the
simulation approaches outlined in [41], [56], [58] & [73] is that they overcome the
problems associated with attempting to mathematically analyse the inherently non-linear
code and carrier tracking loops. However, these approaches also fail to provide any deep
insight into the problem and do not reveal the links that may exist between the
scintillation statistical parameters and the receiver performance parameters.
Despite a lack of detailed research in the navigation community, a considerable amount
of work has been done in the communications field on the effects of multipath fading on
various types of communications systems, including Direct Sequence Spread Spectrum
systems (see for example [30], [46] & [48]). However, most of this work has focused on the
calculation of error rates in the transmitted message rather than the performance of the
code and carrier tracking loops. For GPS, the probability of losing lock on the GPS signal
and the magnitude of the measurement errors in the tracking loops is considered to be of
far greater importance than errors in the navigation message. This is partly because of the
large amount of diversity associated with the navigation message (diversity in both space
and time), and partly because measurement errors in the tracking loops tend to be of
more importance for a navigation system such as GPS than they are for a communications
system. In addition, the majority of the communications literature deals with the more
common Rayleigh and Rician fading models rather than the Nakagami-m model [64]
which has been found to be more suitable under amplitude scintillation conditions [7],
[34] & [101]. This is mainly because the Rayleigh and Rician fading models have
traditionally been used to describe multipath effects, but also because they contain
information about the joint statistics of amplitude and phase and so provide a more
complete description of the fading statistics (the Nakagami-m distribution for amplitude
and the Gaussian distribution for phase cannot be derived from an underlying signal
model). Nevertheless, a few researchers in the communications field have looked at the
effects of Nakagami-m fading on communications systems, and have identified
ionospheric scintillations as a possible source of Nakagami-m fading (see for example
[25], [26], [32], [61], [62], [65] & [102]). However, the work in this area has again focused
on the calculation of error rates in the transmitted message rather than the performance of
the code and carrier tracking loops. Indeed, in all of these cases, it has been assumed that
the carrier is perfectly synchronised to the receiver local oscillator (ie. it is assumed that
the carrier phase error, φε , is zero). Consequently, one of the principal objectives of the
work described in this thesis was to derive expressions for the performance of the code
and carrier tracking loops, navigation data demodulation and acquisition in terms of the
26
parameters of the scintillation model outlined in Section 2.1.3 and in [34], [35], [76], [77] &
[101].
2.3.1. Carrier tracking loops
Under normal (quiescent) conditions, the tracking loop model of Figure 2.2–2 can be
linearised by assuming that the phase tracking errors are small. This enables the nonlinear discriminator element to be ignored which significantly simplifies the analysis. All
text books on phase-locked loop theory discuss this approach and use it to obtain many of
the measures which define loop performance (see for example [15], [36] & [43]). However,
when the phase errors are large, either as a result of large amounts of direct phase noise,
electromagnetic interference or amplitude fading, the linear model approximation may be
significantly violated and the loop runs the risk of losing lock. Under these highly nonlinear conditions, the behaviour of a phase locked loop is not well understood. Indeed,
the only closed form analytical expressions that exist to define loop performance under
these conditions are the probability density function of the phase errors (also called the
“Tikhonov density function” and discussed in [97]), and the mean time to cycle slip (see
for example [43]). Although these expressions have only been derived for a standard first
order phase-locked loop that is subject to additive, white thermal noise, Lindsey and
Charles [59] have shown that they are also a reasonable approximation for higher order
loops and for non-white noise under certain conditions. In addition, Holmes [43] gives
equivalent forms for these expressions that apply to an I.Q Costas phase locked loop. In a
number of communications papers (eg [30], [46], [48]), the Tikhonov density function is
used to determine the effects of imperfect carrier synchronisation on the bit error rates in
a communications system that is subject to multipath fading. Also, Weber [100] has
looked in detail at the effect of Rayleigh, Rician and Log-normal fading on a standard
phase-locked loop, but again has only applied his results to the calculation of error rates
in the received data. However, none of these researchers have looked at the effects of
Nakagami-m fading on the performance of a phase locked loop. Also, none have looked
at the probability of losing lock on the received signal, nor at the errors introduced into
the code and carrier phase estimates. This is mainly because the code and carrier phase
estimates provide information about the range to a satellite and so are of far less
importance to a communications system than they are to a navigation system such as
GPS. In addition, none of these papers have looked at the effects of a post-detection AGC
on the susceptibility of the tracking loop to amplitude scintillations, nor the effects of
relative motion between the transmitter and the receiver. Van Dierendonck [92] discusses
the need for a post-detection AGC or a normalised discriminator such as Atan(Q/I) in
27
order to control the bandwidth of the tracking loop when the signal strength is unknown.
Without such control, the instantaneous loop bandwidth may change significantly during
periods of strong amplitude fading resulting in tracking difficulties, particularly under
dynamic conditions. The behaviour of the AGC in the presence of amplitude scintillations
will depend very much on its time constant. If the time constant of the AGC is short in
relation to the correlation time of the amplitude, or the discriminator is normalised, the
amplitude fluctuations are likely to be tracked quite accurately by the carrier loop,
provided that the amplitude fading is not too deep (if the amplitude fade is very deep, the
AGC will be limited by thermal noise). Under these conditions, the amplitude
scintillations will be decoupled from the effects of phase scintillations and will merely act
as a scaling factor for the receiver thermal noise. However, if the AGC time constant is
long, it will be unable to accurately track the amplitude fluctuations and both the loop
bandwidth and the damping factor will vary with time. Consequently, the amplitude will
be strongly coupled to the effects of phase scintillations through the loop transfer
function. Under these conditions, Weber [100] assumes that the bandwidth of the
amplitude scintillations is narrow in relation to the loop noise bandwidth and derives an
expression for the phase error variance as a function of the amplitude. The average
variance is then obtained from this conditional variance using the Nakagami-m PDF.
Consequently, the objective of this part of the research was to derive expressions for the
performance of the carrier tracking loops in terms of the scintillation parameters T, p, S 4
and σ φ from Section 2.1.3, and to investigate the impact of different AGC regimes on
carrier loop performance. The two principal performance measures to come out of this
work were the variance of the phase tracking error and the variance of the loop phase
estimate. The first of these is useful for determining the probability of losing carrier lock,
the second is important for determining errors in carrier phase DGPS as well as errors in
the estimation of velocity. Carrier loop tracking thresholds were then derived as a
function of the two principal scintillation indices, T and S 4 , and compared with
predictions obtained from the scintillation model WBMOD [82]. The dependence of these
indices on both the satellite elevation angle and the satellite and receiver velocities was
also examined.
Another significant component of this research was an investigation into the relationship
between the fade depth and duration and the probability of a cycle slip for a carrier loop
subject to a single fade with a simple rectangular profile. This problem is tackled using
28
both analytical techniques and simulations based on the carrier loop simulator from
Appendix B. Kintner et al [50] also considers this issue using measured scintillation data
and a real GPS receiver located in the equatorial region. His analysis shows that an
increase in fade duration, possibly as a result of satellite and receiver motion, can cause a
receiver channel to lose lock provided that the fade depth is near or below the tracking
threshold of the carrier loop. He also indicates that when the velocity of the ionospheric
pierce point matches the drift velocity, vd , a situation referred to as velocity resonance can
occur in which the fade duration becomes extremely long and greatly increases the
probability of losing lock.
2.3.2. Code tracking loops
A delay locked-loop operates in essentially the same way as a phase-locked loop. The two
main differences lie in the discriminator algorithm and in the need to generate a replica
code for the code correlators. As the GPS carrier wavelength is more than a thousand
times shorter than the length of a code chip, the effects of phase scintillations on the code
loop can be ignored. However, the effects of amplitude scintillations on a code loop are
essentially the same as their effects on a carrier loop and can be dealt with in much the
same way. Probably the main difference in performance between the two loops is
associated with the much narrower bandwidth of the code loop and the presence of postdetection integration within discriminator (see for example [47]). Although the principal
reason for these differences is to reduce thermal noise errors in the code loop, a second
effect will be to reduce the impact of amplitude scintillations, particularly if the fade
duration is short in relation to the loop time constant. Indeed, using simulations Hegarty
et al [41] found that the effects of amplitude scintillations on a very narrow bandwidth
(0.1 Hz) delay locked loop were negligible. However, using a real GPS receiver, Coco et al
[24] observed that strong amplitude scintillations could increase the pseudorange RMS as
well as introducing large pseudorange spikes. Therefore, the key objective of this part of
the work was to derive variance measures for the delay-locked loop which take into
account the effects of post-detection integration, and to then relate these to the accuracy
with which the code pseudorange can be measured. In order to achieve this using
analytical techniques, assumptions must be made about the bandwidth of the code loop
in relation to the correlation time of the amplitude scintillations.
A second objective of this work was to demonstrate that the distortion of the GPS ranging
codes produced by frequency selective scintillation effects is likely to be negligible at GPS
29
frequencies. The work done by Bogusch et al. [16] & [17] on the effects of ionospheric
disturbances on the performance of code correlators demonstrates that the irregularities
must be extremely dense (such as those produced by a high altitude nuclear blast) and the
code bandwidth must be relatively wide before significant code distortion can occur. At
GPS frequencies, the distortion of the GPS codes is expected to be negligible under
naturally occurring ionospheric conditions, even for the wider bandwidth P-Code.
2.3.3. Codeless and Semi-Codeless receivers
It is generally accepted that scintillations will have the most profound effect on codeless
and semi-codeless tracking loops. This is based on studies of real GPS receivers exposed
to scintillation effects (see for example [57], [66], [67] & [98]) and a number of more
theoretical studies (eg. [41] & [93]). The reasons for this are that the tracking loops of
codeless and semi-codeless receivers generally have both a significantly reduced SNR and
a very narrow loop bandwidth. As a result, receivers of this sort may frequently revert to
L1 only tracking under strong scintillation conditions. This is unlikely to have a very
significant effect on the positional accuracy of a receiver (assuming SA is the dominant
source of error), but will compromise a receiver’s ability to measure ionospheric TEC.
Consequently, it is only likely to be of importance for receivers that form part of a WAAS
network and are therefore required to monitor TEC. However, with the introduction of a
second civil frequency in the near future (referred to as the L5 frequency), it is unlikely
that this problem will persist much beyond the current solar maximum.
The objective of this part of the work, therefore, was to develop analytical expressions
that define the performance of codeless and semi-codeless receivers in terms of the
scintillation parameters given in Section 2.1.3. The result of this analysis were then
compared with measurements obtained from a semi-codeless GPS receiver co-located
with an ionospheric scintillation monitoring receiver in an area known to experience
scintillation effects.
2.3.4. Navigation data
A number of researchers in the communications field have looked at the effects of
Nakagami-m fading on the error rate in communications systems. Wojnar [102] obtained
an expression for the average bit error probability for a non-selective Nakagami-m fading
channel based on an expression for the conditional bit error probability which applies to
both coherent and non-coherent PSK and FSK. This result was based on earlier work by
30
Nesenbergs [65], Esposito [32], and Barrow [6]. Miyagaki [61] tackled the problem of nonselective Nakagami-m fading on coherent M-ary PSK, while Crepeau [25] & [26]
considered the cases of frequency hopped non-coherent binary FSK, non-coherent M-ary
FSK and differentially coherent binary PSK. In addition, Eng [31] analysed the problem of
frequency-selective Nakagami-m fading on Direct Sequence CDMA and derived
expressions for the bit error rate of a RAKE receiver. However, none of this work has
included the effects of carrier loop phase tracking errors, and in particular the impact of
direct phase noise from sources such as phase scintillations. Consequently, the objective
of this part of the work was to determine the effects of Nakagami-m amplitude fading
and carrier loop tracking errors from phase scintillations and thermal noise on the
demodulation performance of the navigation data. This was done using an approach
similar to those described in the cited papers, but with the Tikhonov density function
used to account for the effects of imperfect carrier synchronisation.
2.3.5. Acquisition
McDonough & Whalen [60] (pages 262-265) use the Rayleigh PDF for the signal
amplitude to find an average probability of detection for an envelope detector under
multipath fading conditions. The approach used here is to use the Nakagami-m PDF and
a square-law detector (see for example [43], [60], [72], [74] & [84]) that incorporates postdetection integration in order to derive a similar average probability of detection. This
requires both the amplitude and phase variations of the scintillations to be slow in
relation to the integration period of the detector. This condition that is likely to be met
under normal scintillation conditions and for typical integration periods.
The false alarm probability of an acquisition detector is generally considered to be
independent of the signal level once the design parameters of the detector are fixed (ie.
the pre and post-detection integration periods and the detection threshold). However,
correlation sidelobes produced by the GPS C/A-Codes can significantly increase the false
alarm probability if they are not accounted for when the design parameters of the
detector are chosen [92]. As amplitude scintillations lead to occasional enhancements in
the signal strength, the impact of correlation sidelobes can be even greater. This effect has
been examined using the Nakagami-m PDF to account for these enhancements, as well as
the PDF of the sidelobe levels (based on a cumulative distribution function reported by
Spilker [87]) to determine an average false alarm probability.
31
The impact of a reduced probability of detection on the mean time to acquire has also
been investigated for a simple single-dwell, serial search strategy of the sort described in
[72] & [84]. The effects of both short and long amplitude correlation times have been
addressed, although a number of simplifying approximations have been made in order to
arrive at a closed form analytical expression.
2.3.6. Optimum tracking of the carrier phase
As shown by Van Trees [95] & Brown & Hwang [19], the optimum filter for tracking a
signal that is corrupted by additive thermal noise can often be mapped into an equivalent
phase-locked loop structure, provided that the transfer functions of the two have the
same general form. This allows the optimum bandwidth, loop order and damping factor
of the carrier loop to be obtained through a procedure that is independent of phaselocked loop theory. By applying this approach to the case where a signal is corrupted by
amplitude and phase scintillations, the optimum phase locked loop structure for a
minimum probability of losing lock is obtained as a function of the scintillation
parameters from Section 2.1.3. This approach has also been extended to the case where
dynamics are present, although the resulting optimum loop transfer function cannot
always be mapped into an equivalent phase locked loop structure. In a similar way, the
optimum causal and non-causal post-loop filters for minimum carrier phase range errors
were obtained.
2.3.7. Scintillation effects on navigation
As scintillations cannot be eliminated by pre-processing prior to the tracking loops1, the
most straightforward mitigation strategy involves simply avoiding the times and
locations for which scintillation activity is most likely to be a problem. Scintillation
models such as WBMOD [82] provide predictions of the occurrence and severity of
scintillation activity and are therefore very useful for planning operations in areas that
may be affected by scintillations. Other models such as the Scintillation Network Decision
Aid or SCINDA, [21] & [38], collect scintillation data from multiple receiver sites and
process the data using models of plume formation, evolution and destruction to forecast
scintillation activity (the SCINDA model reverts to WBMOD in the absence of current
1
For electromagnetic interference, techniques such as adaptive filtering and adaptive null steering antennas
can be used to reduce the effects of interference prior to the tracking loops without the need to modify the
receiver hardware.
32
scintillation measurements). By linking the analytical performance measures described in
the previous sections with the scintillation statistics generated by WBMOD or SCINDA,
predictions can be made about the likely performance of a receiver at a given time and
location. It can be shown ([76] & [77]) that the principal amplitude and phase scintillation
parameters, S 4 and T, are directly proportional to a third parameter, referred to as the
height integrated irregularity strength or C k L , which is a measure of the strength of
scintillation activity (see Equation (2.1-2)). In WBMOD, the distribution of Log (C k L) in
equatorial regions is modelled as the sum of three Gaussians, each of which have
different means and variances (see for example [13] & [82]). Therefore, as many of the
receiver performance measures are functions of S 4 and T, they are amenable to
averaging using this distribution function. Consequently, using WBMOD the
performance of a single satellite-receiver link can be predicted for a given time and
location, either in terms of a percentile or as an average based on the distribution function
of C k L .
Ordinarily, in the absence of scintillations, the navigational accuracy of a GPS receiver is
found by estimating the single link error (referred to as the User Equivalent Range Error
or UERE) and translating this error into equivalent position and time errors using satellite
geometry factors (the Dilution Of Precision or DOP factors: see for example Kaplan [47]
pp. 261-269). The DOP factors assume that the range errors are the same on each satellite
link and are uncorrelated between the individual links. However, when the effects of
scintillations are included, WBMOD provides additional information from which the
tracking status and range errors on each satellite link can be obtained independently. The
problem then arises as to how these individual link predictions can be combined in order
to determine the likely impact on navigational accuracy. In this section, it is shown that
the joint statistics of the scintillation indices S 4 and T on the individual satellite-receiver
links are required in order to solve this problem. However, as yet there are no models or
data available from which this information can be derived. Nevertheless, in [57] data
obtained from a receiver located in an active scintillation environment was analysed to
show that the probability of losing lock simultaneously on multiple satellite links is
extremely small. Further analysis of this data is required in order to determine the
required statistics and the factors on which they depend.
33
2.4. Summary
In this chapter, an overview of the morphology and statistical characteristics of
scintillations was given. It was revealed that scintillations can be modelled as a stochastic
process in which the amplitude follows a Nakagami-m distribution and the phase follows
a zero-mean Gaussian distribution. Furthermore, both the amplitude and phase can be
assumed to have a power-law power spectral density with a low frequency cutoff for the
amplitude. The scintillation model WBMOD was also described and its potential for
predicting the statistics of various key scintillation parameters was explained. These
statistics link WBMOD to the stochastic scintillation model as they define the distribution
functions and spectral characteristics of the scintillations. A second scintillation model
based on diffraction from a thin phase changing screen was also described. The primary
purpose of this model was to generate scintillation time series for the simulation tests
carried out in subsequent chapters.
The second part of this chapter provided an overview of the GPS system, including a brief
description of the architecture of the front end of a GPS receiver and the operation of the
receiver tracking loops. This was then followed by a literature review and an overview of
each chapter in the thesis. The main point to come out of this review was that although
various researchers have conducted simulation studies and measurement campaigns
aimed at quantifying the effects of scintillations on GPS, there has been very little detailed
analytical work done in the area. Consequently, the main thrust of this thesis is to link the
various statistics associated with the stochastic scintillation model outlined above to
measures of the tracking and acquisition performance of a GPS receiver.
34
Chapter 3
Carrier tracking loops
This chapter examines the effects of scintillations on carrier tracking loops. As carrier
loops are generally the weakest link in a receiver channel, significantly more effort has
been devoted to this chapter than to the following chapter on code tracking loops. A brief
outline of each section in this chapter is given below:
•
In Section 3.1, a signal processing model of the carrier tracking loop is given. This
model is used throughout the thesis and is based on a Costas suppressed carrier tracking
loop.
•
In Section 3.2, the effects of phase scintillations on a Costas carrier loop is
investigated. Phase scintillation effects include an increase in carrier tracking errors and
an increase in carrier phase range errors. The first of these is associated with the ability of
the carrier loop to remain in phase lock, the second is associated with errors in carrier
phase differential GPS. The latter is dealt with in greater detail in Section 3.6 where the
decorrelation with distance of the phase scintillation error is examined.
•
In Section 3.3, the impact of amplitude scintillations on a Costas carrier loop is
investigated. This is done for an I.Q discriminator that is normalised by a post-detection
AGC. Because the joint statistics of amplitude and phase scintillations are unknown at
this stage, it was decided to initially treat their effects separately, and to then deal with
their combined effects by assuming independence. The potential problems associated
with this assumption are highlighted.
•
In Section 3.4, a rule of thumb tracking threshold for the Costas loop is defined in
terms of the amplitude and phase scintillation indices, T and S 4 , and the loop bandwidth.
This threshold is then compared with WBMOD predictions of T and S 4 obtained for a
stationary receiver located in a region of high scintillation activity. The effects of satellite
elevation angle and both satellite and receiver velocity on the scintillation indices, and
therefore on the probability of exceeding the tracking thresholds, are examined using
WBMOD data and the scintillation model from Section 2.1.3.
35
•
In Section 3.5, both analytical techniques and simulations are used to determine the
relationships between the depth and duration of a simple rectangular fade and the
probability of a carrier cycle slip. The intention is to show that short duration fades have
very little impact on the tracking loop performance, irrespective of the fade depth.
•
Finally, in Section 3.7 the effects of scintillations on a Frequency Locked Loop (FLL)
are examined. FLL’s are used as a primary means of carrier tracking in some receivers,
but more often as a fall-back strategy when phase locked loop operation is no longer
possible. In this section, the susceptibilities of FLL’s and PLL’s to scintillations are
compared.
3.1. Carrier loop model
Figure 3.1-1 is a representation of a generic Costas carrier phase tracking loop. A brief
description of the operation of the Costas loop is given below.
pp
Ip
1
T
IF
1
T
Sign( )
t
dτ
∫
t −T
Phase
Discrim
t
dτ
∫
t −T
Navigation Data
F(s)
%̂
φ
Qp
900
+
VCO
ω IF
Figure 3.1-1: Model of a generic Costas phase locked loop.
The Costas carrier loop generates in-phase (I) and quadrature (Q) signals by mixing the
GPS IF with I and Q reference signals produced by a voltage controlled oscillator (VCO).
The PRN ranging code is then removed from the I and Q signals by mixing with an inphase replica code, p P , that is generated within the code tracking loop (see Chapter 4).
This causes the energy of the GPS signal to be collapsed into the bandwidth of the
navigation data which is 100Hz between nulls. The resulting I and Q signals are then
36
filtered by a pair of pre-detection integrate and dump filters1 before being passed into the
Costas loop discriminator. The function of the discriminator is to measure the phase error
between the IF carrier and the VCO reference while simultaneously removing navigation
data from the carrier. A list of the most common discriminator types along with their
corresponding phase error functions is given in Table 3.1-1.
Discriminator
Discriminator Output
sign ( I ).Q
~
A sin(φ ε )
I.Q
~
0.5 A 2 sin(2φ ε )
Q I
tan( φ ε )
Atan2(Q, I ) , Atan (Q I )
φε
~
Table 3.1-1: Common Costas loop discriminator functions. A is the filtered signal amplitude and
φε
is the phase error.
Atan2( y , x )
is the four quadrant arctangent function (ie.
- π ≤ Atan2( y , x ) ≤ π ).
From the discriminator, the phase errors are passed through a loop filter, F (s ) , and then
on to the VCO. The loop filter controls the order and bandwidth of the tracking loop and
must be adjusted according to the expected dynamics and noise conditions in order to
maintain optimum tracking performance (ie. minimum phase tracking error). The filtered
phase errors force the frequency of the loop VCO to be shifted in a direction that reduces
tracking errors in subsequent phase measurements. In this way, the VCO tracks both the
frequency and the phase of the IF carrier. Estimates of the line-of-sight Doppler and the
ambiguous phase range (the integrated Doppler) are obtained directly from the filtered
phase errors by applying appropriate scaling factors.
In the analysis that follows, the effects of scintillations are modelled as a modulation of
the complex GPS signal 2. Based on this model, the IF signal can be represented by
1
The integrate and dump filters are synchronised to the navigation data and have integration periods less than
or equal to the length of a navigation data bit. Without synchronisation, the SNR of the filtered I & Q signals
would be significantly degraded by changes in the sign of the navigation data.
2
The modulation is of the form A(t )exp( − jφ p (t )) , where A(t ) and φ p (t ) are the amplitude and phase
scintillation processes respectively.
37
IF (t ) = A(t ) p (t − τ (t ) )d (t − τ (t ) )sin (ω IF t + φ (t ) ) + n (t )
(3.1-1)
where:
A(t ) is the signal amplitude,
p (t − τ (t ) ) is the satellite PRN code3,
d (t − τ ( t )) is the satellite navigation message3,
τ ( t ) is the code delay,
ω IF is the IF carrier frequency,
φ (t ) = φ d (t ) + φ p (t ) + φ o (t ) is the phase of the GPS carrier,
φd ( t ) represents the effects of satellite and platform dynamics,
φ p (t ) represents the effects of phase scintillations,
φo ( t ) represents other effects such as VCO phase noise,
n( t ) = nc ( t )cos(ω IF t ) + n s ( t )sin(ω IF t ) is stationary, zero-mean, narrowband Gaussian
thermal noise with a power spectral density of N o W/Hz within the IF band, and
n c (t ) and n s (t ) are stationary, zero-mean, Gaussian noise processes which are
independent and identically distributed (IID).
The IF signal is mixed with I and Q reference signals from the VCO and a prompt code,
p p = p (t − τˆ(t )) , from the code tracking loop to produce a pair of baseband I and Q
signals. The mixing process also generates double frequency terms centred on 2ω IF , but
these are eliminated by the pre-detection integrate and dump filters in the following
stage. If it is assumed that the pre-detection filters are synchronised to the navigation
data, and the phase error, φε , is relatively constant over the integration period, then after
filtering the I and Q signals will become4
~
I P = Ad ( t − τ )cos(φε ) + n IP ,
~
QP = Ad ( t − τ )sin (φε ) + nQP
(3.1-2)
where φε = φ (t ) − φ#(t ) is the carrier phase tracking error5, φ#( t ) is the loop's estimate of the
3
Both the PRN ranging codes and the navigation data are represented by a ±1 bit sequence.
4
The VCO signal is assumed to be of the form 2 sin(ω IF t + φˆ(t )) .
5
In this Chapter, the terms “phase error”, “carrier phase error” and “carrier phase tracking error” are used
interchangeably.
38
~ 1
carrier phase, τ = τ ( t ) , A =
T
t
∫ A(u)du
is a filtered version of the signal amplitude, T is
t −T
the integration period of the pre-detection filters, and
n IP and nQP are uncorrelated,
baseband Gaussian noise processes with σ n IP 2 = σ nQP 2 = N o T . This step assumes that
the replica prompt code is perfectly aligned with the satellite code and is therefore
removed completely from the carrier (ie τˆ = τ ). The filtered I and Q signals are then
processed in the Costas loop discriminator to produce the phase error estimates given in
Table 3.1-1. In all cases, the discriminator algorithm eliminates the navigation data from
the phase error estimates allowing the loop bandwidth to be reduced to a few Hz.
The Costas loop can be represented in an equivalent form in which the mixers and predetection filters are replaced by an adder, and the phase discriminator is replaced by the
appropriate phase error function. An example of this baseband model for the I.Q
discriminator is given in Figure 3.1-2. The 1 g factor represents gain control from an
AGC.
nd
φ
+
+
-
φˆ
φε
~
0.5A2 sin(2φε )
1/g
+
F(s)
1/s
Figure 3.1-2: Non-linear baseband model of an I.Q Costas phase locked loop.
For small phase errors, the approximation 0.5 sin (2φε ) ≈ φε can be used to produce the
linear baseband model shown in Figure 3.1-3. Similar approximations can be made for the
other three discriminator types from Table 3.1-1.
39
nd
φ
+
-
+
φε
~
A2
φˆ
1/g
+
F(s)
1/s
Figure 3.1-3: Linear baseband model of an I.Q Costas phase locked loop.
The noise term nd represents the effects of additive thermal noise translated to the
discriminator output, and includes products between the I and Q noise terms and the I
and Q signal terms which are created within the discriminator. For the I.Q discriminator,
nd is given by (from Equation (3.1-2))
[
]
~
nd = A d (t − τ ) cos(φε )nQP + sin(φε )n IP + nQP n IP
(3.1-3)
The gain control shown in these figures is typically provided by a post-detection AGC
and is necessary to ensure that tracking loops based on un-normalised discriminators
such as I.Q and sign ( I ).Q operate within their design parameters [92]. Without such
control, the bandwidth and damping factor of the loops would be strongly affected by the
~
signal amplitude, A . AGC’s of this sort can either be applied after the discriminator (as
shown in these diagrams) or at the IF stage prior to the tracking loop. In either case, their
effect on the tracking loop will be the same. For an I.Q discriminator, the AGC gain factor,
~
g, is given by I P 2 + Q P 2 ≈ A 2 . If it is assumed that the discriminator is normalised (eg.
Atan (Q I ) ), or the AGC is capable of accurately tracking the signal amplitude, the closed
loop transfer function of the Costas loop is given by (see for example [36])
H (s) =
ˆ (s )
F ( s)
Φ
=
Φ (s ) s + F ( s )
(3.1-4)
where Φ (s ) and Φ̂ (s ) are the Laplace transforms of φ (t ) and φˆ(t ) respectively. Typical
loop transfer functions and their corresponding noise equivalent bandwidth’s6 for the
three loop orders are given in Table 3.1-2 (assuming active loop filters). As shown in
6
The single-sided noise equivalent bandwidth is given by Bn =
40
1
2
∞
∫ H(f )
−∞
2
.df .
Figure 3.1-4, the closed loop transfer function can be used to simplify the linear baseband
model of the Costas loop.
~
nd A2
φ
+
H(s)
φˆ –
φε
+
+
Figure 3.1-4: Closed loop transfer function model of a phase locked loop
1st Order Loop
2nd Order Loop
3rd Order Loop
F (s )
ωn
2ζω n + ωn2 s
2ω n + 2ω n2 s + ω n3 s 2
H (s )
ωn
s + ωn
2ζω n s + ω n2
2ω n s 2 + 2ω n2 s + ω n3
s3 + 2ω n s 2 + 2ω n2 s + ωn3
Bn
ωn 4
s 2 + 2ζω n s + ω n2
ωn
1
ζ +
2
4ζ
.
ω n 12
Table 3.1-2: Open and closed loop transfer functions and single-sided noise bandwidth’s for a
phase locked loop. ω n is the loop natural frequency and ζ is the damping factor for a second order
loop.
The three loop transfer functions represent the optimum Wiener filters for tracking a
phase step (1st order), a frequency step (2nd order with ζ = 1
2 ) and a frequency ramp
(3rd order). These optimum filters were derived by Jaffe and Rechtin [45] based on a
minimisation of the mean-square loop phase error in the presence of dynamics and noise.
The derivation of an optimum Wiener filter in the presence of scintillations and dynamics
will be discussed in Chapter 8.
41
3.2. The impact of phase scintillations on carrier
phase tracking loops
In this section, the variance of the phase tracking errors and the variance of the phase
range errors will be derived for a phase locked loop in the presence of phase scintillations
and thermal noise. The variance of the phase tracking errors will be used to determine a
threshold value for the spectral strength of phase scintillations beyond which loss-of-lock
is likely to occur. The variance of the phase range errors will be used in Section 3.6 to
investigate the effects of phase scintillations on carrier phase DGPS.
3.2.1. Phase tracking errors and thresholds
The mean-square carrier phase tracking error resulting from direct phase noise and
thermal noise for the linearised carrier phase tracking loop is given by (based on Figure
3.1-4)
{ } ∫[1− H( f )
∞
E φε2 =
2
2
]
Sφ ( f ) + H ( f ) S n d′ ( f ) .df
(3.2-1)
−∞
where 1 − H ( f ) is the transfer function of the phase errors, S φ ( f ) is the power spectral
density (PSD) of the input phase process and S n′d ( f ) is the PSD of the normalised
~
thermal noise term, nd′ = nd A 2 . The input phase process can be represented by (from
Equation (3.1-1))
φ ( t ) = φ d (t ) + φ p ( t ) + φ o (t )
(3.2-2)
where φd (t ) , φ p (t ) and φo ( t ) represent the contributions from satellite and platform
dynamics, ionospheric phase scintillations and other phase noise sources respectively.
The autocorrelation function of φ (t ) is therefore
Rφφ (t1 , t2 ) = E {φ (t1 ).φ (t 2 )}
{[
][
]}
= E φ d (t1 ) + φ p (t1 ) + φo (t1 ) ∗ φ d (t2 ) + φ p (t2 ) + φo (t2 )
= Rφ d φ d (t1 , t2 ) + Rφ pφ p (τ ) + Rφ oφ o (τ ) + cross - correlation terms
42
(3.2-3)
where t1 and t 2 are two instants in time and τ = t 2 − t1 . In this analysis, it is assumed that
φ p (t ) and φo (t ) are stationary, zero-mean, Gaussian random signals, and that φ d (t ) may
be deterministic or random, but is generally not zero-mean. It is also assumed that φd ( t ) ,
φ p (t ) and φo ( t ) are independent as they are produced by entirely different physical
processes. Consequently, the cross-correlation terms in Equation (3.2-3) are all zero. The
corresponding PSD of φ (t ) is thus
Sφ ( f ) = Sφ d ( f ) + Sφ p ( f ) + Sφ o ( f )
{
where Sφd ( f ) = E Φ d ( f )
2
(3.2-4)
} is the PSD of the dynamics component (see Appendix E),
Sφ p ( f ) is the PSD of ionospheric phase scintillations (Equation (2.1-1)), and Sφo ( f ) is the
PSD of the other phase noise sources. In the analysis that follows, it is assumed that
thermal noise and ionospheric phase scintillations are the principal sources of phase
~
noise, and that amplitude scintillations are not present (ie. A = A where A is the
unperturbed signal amplitude). Under these conditions, the mean-square phase tracking
error reduces to
∫ [1 − H ( f )
∞
σ φ2ε
=
2
2
]
Sφ p ( f ) + H ( f ) S nd′ ( f ) .df
−∞
(3.2-5)
= σ φ2εp + σ φ2T
where σ φ2εp and σ φ2T are the phase scintillation and thermal noise components of the
{ }
tracking error variance, and E φε2 is equal to the phase error variance, σ φ2ε , as both
phase scintillations and thermal noise terms are zero-mean. Equation (3.2-5) can be
simplified by making the following substitutions
σ φ2T =
Bn
1
, from Equation (D-19), Appendix D
1 +
C N o 2T C N o
Sφ p ( f ) =
1− H ( f )
2
=
(f
T
2
o
+ f2
f
f
2k
)
p 2
2k
+ f n 2k
, from Equation (2.1-1), and
, from Table 3.1-2 with s = j 2πf
43
where k is the loop order (1, 2 or 3), f n = ω n 2π is the loop natural frequency in Hertz, f o
is the outer scale size frequency, and a damping factor of 1 / 2 has been assumed for
second order loops. The phase scintillation component of the tracking error variance is
∞
f 2k
T
∫ (f 2k + f n 2k ). (f 2 + f 2 )p 2 .df
−∞
o
σ φ2εp =
(3.2-6)
Unfortunately, it is difficult to obtain a closed form solution to this equation. However,
for p < 2k and f o much smaller than f n , the following approximation can be made
(f
f 2k
2k
+ f n 2k
) (f
T
.
2
o
+f
)
2 p 2
≈
(f
2k − p
Tf
2k
+ f n 2k
)
(3.2-7)
This is based on the observation that the phase error transfer function, 1 − H ( f ) , is a highpass filter, and so the low frequency components of the phase scintillation power
spectrum will have a negligible effect on the phase errors. Therefore, letting f o = 0 will
not significantly affect the phase error variance. As f o is usually very much smaller than
f n , this approximation will be accurate under the following conditions (based on p < 2k )
1st order loop: p < 2
2nd order loop: p < 4
3rd order loop: p < 6
As the carrier tracking loop in a GPS receiver is usually 3rd order (unaided) or 2nd order
(aided), and p is in the range 1 to 4 (typically 2.5 at equatorial latitudes), this
approximation is considered to be quite accurate under most circumstances. Equation
(3.2-6) then becomes (using Spiegel’s table of integrals [86], Equation 15.20)
∞
σ φ2εp
≈
Tf
2k − p
∫ (f 2k + f n 2k ).df ,
p < 2k
−∞
≈
πT
kf n
p −1
(3.2-8)
sin([p − 1]π 2k )
,
1 < p < 2k
From Appendix C, the linear model tracking threshold for an I.Q Costas loop is given by
σ φ2ε
44
Th
π
=
12
2
radians2
(3.2-9)
By equating the total phase error variance from Equation (3.2-5) with the linear model
threshold given above, the following expression can be obtained for the threshold
spectral strength
T Th
= σ φ2ε
− σ φ2T
Th
kf n
p −1
sin ([p − 1]π 2k )
π
(3.2-10)
In Figure 3.2-1, T Th is plotted as a function of the spectral index, p, and the loop noise
bandwidth, Bn , for both a second and third order carrier tracking loop under the
specified threshold condition ( C N o = 41.5dBHz7 and T=20ms are assumed). The
relationships between the loop noise bandwidth and the loop natural frequency, f n , for
the two loop orders are given in Table 3.1-2.
Second Order loop
Third Order loop
5
0
Spectral Strength, T (dB)
Spectral Strength, T (dB)
5
−5
−10
−15
−20
−25
−30
−35
0
−5
−10
−15
−20
−25
−30
−35
20
3
20
3
15
15
10
2
10
2
5
Spectral Index, p
1
5
Noise bandwidth, Bn (Hz)
Spectral Index, p
1
Noise bandwidth, Bn (Hz)
Figure 3.2-1: The threshold spectral strength, T Th , as a function of the spectral index, p, and the
loop noise bandwidth, Bn , for a second order loop (left panel) and a third order loop (right panel).
C N o = 41.5dBHz and T = 20ms.
Figure 3.2-2 shows a cross-section through each of the plots in Figure 3.2-1 for p=2.5 (a
typical equatorial value). By comparing these results with Figure 2.1-4 which was
obtained using the WBMOD scintillation model, it is clear that in the region of the
equatorial anomaly, it is possible for the spectral strength to exceed the specified tracking
threshold and cause the carrier loop to lose lock. It is also clear that the loop noise
7
Throughout this thesis, C N o = 41.5dBHz will frequently be used. This is based on a nominal GPS signal
level
of
–160dBW
[81]
and
a
system
noise
temperature,
TS ,
of
512K.
Thus
C N o = −160 − 10Log 10 ( kTS ) = 41.5 where k is Boltzman’s constant.
45
bandwidth strongly influences a receiver's tolerance to phase scintillations and that
narrow bandwidth receivers tend to be far more susceptible (this becomes more
pronounced for larger values of p). Although it is not apparent from these plots, it can be
seen from Equation (3.2-5) that when the carrier to noise density ratio is significantly
reduced, a point will be reached at which the thermal noise term, σ φ2T , dominates the
variance expression. When this occurs, wider bandwidth receivers will tend to be more
susceptible to loss-of-lock. This may occur when strong amplitude scintillations are
present or when the receiver is being affected by electromagnetic interference or
attenuation from foliage or other sources.
0
2
Spectral Strength, T (dBrad /Hz)
−5
−10
−15
−20
−25
−30
2
4
6
8
10
12
14
16
18
20
Noise bandwidth, Bn (Hz)
Figure 3.2-2: The threshold spectral strength, T Th , as a function of the loop noise bandwidth B n
for a second order loop (upper curve) and a third order loop (lower curve). p = 2.5, C N o =
41.5dBHz and T = 20ms.
As INS8 aided receivers tend to adopt a very narrow tracking loop bandwidth, this result
suggests that aided receivers will be more susceptible to phase scintillations than unaided
receivers, although their tolerance to amplitude scintillations and interference will be
better. This situation will become even worse if a receiver is in a state of open carrier loop
aiding9, as it will no longer be able to track the high power, low frequency components of
8
INS - Inertial Navigation System.
9
With open carrier loop aiding, the VCO frequency is controlled by an INS. This technique is usually used as
a weak signal hold-on strategy under conditions of strong interference.
46
the phase scintillations. However, INS aiding should also allow a receiver to recover
more quickly from a state of loss-of-lock when the scintillation activity eventually passes.
A third observation that can be made from these plots is that in general, second order
loops have a higher tolerance to phase scintillations than third order loops (by
approximately 2 to 3 dB for p=2.5). In Chapter 8, it will be shown that the optimum loop
order for minimum phase error is either first order or second order, depending on p, but
is never third order.
0.5
RMS tracking error (radians)
0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
−30
−25
−20
−15
−10
−5
0
2
Spectral Strength, T (dBradians /Hz)
Figure 3.2-3: A comparison between the RMS phase scintillation error obtained from simulations
(dotted lines) with those obtained from theory (solid line) for a second order Costas phase locked
loop with p = 2.5, f o = 0.05Hz, T = 20ms and no thermal noise. The five different lines represent
2Hz, 5Hz, 10Hz and 20Hz loop noise bandwidths respectively (upper to lower curves).
In order to verify Equation (3.2-8), a number of simulations were conducted using the
tracking loop simulator from Appendix B and simulated phase scintillation data from the
model given in Appendix A. In Figure 3.2-3, the RMS phase scintillation error, σ φεp ,
obtained from both simulations (dotted lines and circles) and theory (solid lines) is
plotted as a function of the Spectral Strength, T, for a range of loop noise bandwidths.
These results show that the linear model is relatively accurate when the phase error
variance is below the tracking loop threshold (given by the solid horizontal line above
0.25 radians). However, when T is increased to a point beyond the tracking threshold, the
simulations fail to provide a clearly defined cutoff between tracking and loss of lock.
47
Rather, the frequency of cycle slips increases until the system is effectively no longer
tracking (the sudden increases in the RMS error near the tracking threshold are the result
of uncorrected cycle slips in the data).
In Figure 3.2-4, the mean time between cycle slips obtained from simulations is plotted as
a function of T and B n for a second order loop with p = 2.5. The solid line marked with a
200 represents an average of 200 seconds between cycle slips. The other solid line
represents an average of 10 seconds between cycle slips. Also shown as a dotted line is
the threshold based on the linear loop model from Figure 3.2-2 (the upper curve). These
results suggest that the linear model threshold is quite conservative and represents a
situation in which the loop is suffering from frequent cycle slips rather than a complete
loss of lock.
Spectral strength, T (dBradians2/Hz)
0
10
−5
200
−10
−15
−20
−25
−30
2
4
6
8
10
12
14
16
18
20
Noise bandwidth, Bn (Hz)
Figure 3.2-4: Mean time between cycle slips (in seconds) for a second order Costas phase locked
loop in the presence of phase scintillations with p = 2.5, f o = 0.05Hz, C N o =41.5dBHz and
T=20ms. The dashed curve represents the theoretical threshold from Figure 3.2-2 (upper curve).
Finally, it is clear from Figure 3.2-3 that the accuracy of Equation (3.2-8) is reduced as the
loop noise bandwidth increases. This is primarily a result of pre-detection filtering and
will be discussed further in the next section.
48
3.2.2. The effects of pre-detection filtering on phase errors
In the analysis so far, it has been assumed that the pre-detection filters have a negligible
effect on the phase errors. This assumption is based on the observation that the majority
of the energy in the phase error power spectrum is within the 50Hz noise bandwidth10 of
the pre-detection filters. However, for wide bandwidth receivers it unclear whether this
approximation is valid. In the analysis that follows, the transfer function of the tracking
loop is modified to include the effects of the pre-detection filtering. The results of this
analysis are then compared with the simulations obtained from the previous section over
a range of loop bandwidths.
To account for the effects of filtering, Equation (3.1-2) can be modified as follows
1
~
I p = A d (t − τ )
T
1
~
Q p = A d (t − τ )
T
t
∫ cos(φε (u )).du + n IP ,
t −T
t
(3.2-11)
∫ sin(φε (u )).du + nQP
t −T
where I p and Q p are the prompt I and Q signals after the pre-detection filters. This result
also assumes that the amplitude is approximately constant over the T second period of
the pre-detection filters. As will be shown in Section 3.3, this is a reasonable
approximation under most circumstances. If it is assumed that φε is small (ie. sin (φε ) ≈ φε
and cos(φε ) ≈ 1 )11, then to a first approximation, I p and Q p can be simplified as follows
~
I p = A d (t − τ ) + n IP ,
t
1
~
Q p = A d (t − τ )
φε (u ).du + nQP
T
t −T
~
= A d (t − τ )[φε (t ) ⊗ g (t )] + nQP
∫
where g (t ) =
(3.2-12)
1
t-T 2
rect
is the impulse response of the pre-detection filters, and ⊗
T
T
represents the convolution integral.
The output of the I.Q discriminator is then
10
The double-sided noise bandwidth of a T second integrate and dump filter is 1/T Hz (= 50Hz for T = 20ms).
11
These approximations are based on the first terms in the Taylor series expansions of sin( ) and cos( ).
49
~
I p .Q p = A 2 [φε (t ) ⊗ g (t )]+ n d
(3.2-13)
If the transfer function of the pre-detection filters is denoted by G ( f ) , the cascade of the
pre-detection filters and the loop filter can be approximated by G( f )F ( f ) . The closed
loop transfer function is therefore
H( f ) =
G ( f ). F ( f )
j 2πf + G ( f ). F ( f )
(3.2-14)
and the transfer function of the phase errors is given by
1− H( f ) =
j 2πf
j 2πf + G ( f ). F ( f )
(3.2-15)
where G ( f ) = sinc( fT )exp(− jπfT ) (Holmes, [43] pp. 423) 12. This new transfer function can
be substituted into the expression for the phase error variance to account for the effects of
pre-detection filtering. For a second order Costas loop, this gives
∞
σ φ2ε
=
∫ 1− H (f )
−∞
∞
2
.Sφ p ( f ).df + σ φ2
T
(3.2-16)
≈ T 1− H ( f ) f
∫
2
−∞
−p
.df + σ φ2
T
where
1− H ( f ) =
2
[
]
fN 4
[
]
sinc( fT )2 1 + 2 f N 2 − 2 f N 2 sinc( fT ) cos(πfT ) + 2 f N sin(πfT ) + f N 4
(3.2-17)
and f N = f f n . In Figure 3.2-5, the RMS phase scintillation error obtained from this
modified transfer function is plotted as a function of T for a range of loop noise
bandwidths. Also shown are the simulation results from Figure 3.2-3. It is clear from this
figure that the simulations now match the theory much more closely, particularly for the
wider loop bandwidths.
12
G ( f ) consists of two factors, (i) a sinc( fT ) attenuation factor, and (ii) an exp(− jπfT ) phase shift factor
associated with the filter delay. Of the two, the phase shift factor has by far the greatest impact on φ .
ε
50
0.5
RMS tracking error (radians)
0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
−30
−25
−20
−15
−10
−5
0
Spectral Strength, T (dBradians2/Hz)
Figure 3.2-5: The effects of pre-detection filtering on the RMS phase scintillation error for a
second order Costas loop with p = 2.5, f o = 0.05Hz, T = 20ms and no thermal noise. The five
different lines represent 2Hz, 5Hz, 10Hz and 20Hz loop noise bandwidths respectively (upper to
lower curves). The dashed lines represent the results of simulations from the previous section.
3.2.3. Carrier phase range errors
Another quantity of interest in the study of GPS receiver performance is the error in the
estimate of range obtained from the carrier phase, φεr (t ) = φ d (t ) − φˆ(t ) (referred to here as
the carrier phase range error). The carrier phase range error reflects the second function of
the carrier phase tracking loop which is to provide estimates of the satellite and receiver
dynamics (represented by φd (t ) ) while rejecting unwanted phase noise from other
sources. Consequently, φεr (t ) represents the error in the loop’s estimate of φd (t ) and is of
interest to systems that make use of carrier phase range measurements such as carrier
phase DGPS, or require precise velocity information. The mean-square value of the carrier
phase range error is given by
{ } ∫ [1− H( f )
∞
E φε2r =
2
Sφd ( f ) + H ( f )
2
[S
φp (
]]
f ) + Sφo ( f ) + S nd′ ( f ) .df .
(3.2-18)
−∞
The contribution to the carrier phase range error from phase scintillations is simply the
carrier loop’s estimate of the phase scintillation process, φ p (t ) . This is given by
51
φˆ p (t ) = h(t ) ⊗ φ p (t ) , where h(t ) is the impulse response of the loop filter and ⊗ is the
convolution integral. The variance of φˆp (t ) is therefore
∞
σ φ2ˆ =
p
∫ H( f )
2
(3.2-19)
Sφ p ( f ).df .
−∞
By making the following substitutions
Sφ p ( f ) =
H( f )
2
=
(f
T
2
o
+ f
f n 2k
f
2k
+ fn
2k
)
2 p 2
, from Equation (2.1-1), and
[1 + 2(k − 1)( f
]
f n )2( k −1) , from Table 3.1-2
the variance of the carrier phase range error resulting from phase scintillations becomes
∞
σ φ2ˆ
p
=
∫ (f 2k + f n 2k )[1 + 2(k − 1)( f
−∞
f n 2k
]
f n )2( k −1) ×
(f
T
2
o
+ f
)
2 p 2
.df .
(3.2-20)
A closed form solution to this integral is again difficult to obtain. However, as the carrier
tracking loop is essentially a low pass filter for the carrier phase, the low frequency
components of the phase scintillation power spectrum will provide the greatest
contribution to σ φ2ˆ . Therefore, the outer scale size parameter, f o , will have a significant
p
effect on σ φ2ˆ
p
and must be carefully modelled in order to produce accurate results (ie. it
cannot be set to zero as before). In Figure 3.2-6, σ φ2ˆ
p
is plotted as a function of the loop
noise bandwidth for a second order Costas phase locked loop and for three values of f o
(this is based on a numerical solution to Equation (3.2-20)). It is clear from these plots that
σ φ2ˆ
p
is very sensitive to f o , but relatively insensitive to the noise bandwidth. This is
because the majority of the energy in the phase scintillation power spectrum is well below
the lowest noise bandwidth for typical values of f o and p. The slightly higher values at
very low noise bandwidth's are the result of a hump in the transfer function of the second
order loop near f = f n .
52
Second order loop
1.2
1.1
2
Variance (radians )
1
fo = 0.04 Hz
0.9
0.8
fo = 0.05 Hz
0.7
0.6
fo = 0.06 Hz
0.5
0.4
0.3
0
5
10
15
20
Noise bandwidth, Bn (Hz)
Figure 3.2-6: The variance of the carrier phase range error, σ φ2ˆ , as a function of the loop noise
p
bandwidth for a second order Costas phase locked loop in the presence of phase scintillations. p =
2.5, T = −25 dBradians2/Hz, and f o = 0.05Hz.
These results suggest that a reasonable approximation for σ φ2ˆ
p
can be obtained by
ignoring the loop transfer function. This implies that for all sensible values of the loop
noise bandwidth, the majority of the phase scintillation energy is tracked by the carrier
loop. Equation (3.2-20) can therefore be approximated by
∞
σ φ2ˆ ≈
p
∫ (f
−∞
T
2
o
+ f
)
2 p 2
.df .
(3.2-21)
From a table of integrals (eg Gradshteyn [37], Equation 3.241-4), this can be reduced to
σ φ2ˆ =
p
T π Γ(( p − 1) 2 ) f o1− p
.
Γ( p 2 )
(3.2-22)
The results obtained from this approximation are given in Figure 3.2-6 as a series of
horizontal dotted lines. It is clear from these plots that the error in this approximation is
quite small for Bn greater than a few Hertz.
The carrier phase range error is generally only of concern to users who require precise
carrier phase range measurements for carrier phase DGPS. For such applications, the
53
distance over which the phase errors become decorrelated is of interest. This is discussed
further in Section 3.6.
3.2.4. Doppler errors
Carrier Doppler measurements are used for the precise determination of velocity in GPS.
The error introduced into these measurements by phase scintillations is given by
ωεp (t ) =
dφˆp (t )
dt
(3.2-23)
radians/s
where φˆ p (t ) is the carrier loop’s estimate of the phase scintillation process. The variance
of the Doppler error is thus
∞
σ ω2 εp =
=
∫ Sωεp ( f ).df
−∞
∞
−∞
∞
=
(3.2-24)
2
∫ (2πf ) Sφˆ p ( f ).df
∫ (2πf )
2
2
H ( f ) Sφ p ( f ).df
(radians/s)2
−∞
where Sωεp ( f ) and Sφˆ ( f ) are the power spectral densities of the Doppler errors and the
p
phase estimate errors from phase scintillations respectively. It is clear from this equation
that the Doppler errors are a filtered version of the phase scintillations, where the filter
transfer function is given by j 2πfH ( f ) . As this is a high pass filter (at least for second and
third order carrier loops), the approximation Sφ p ( f ) = T f
−p
can once again be used to
obtain the following expression (based on Equation (3.2-20))
∞
σ ω2 εp
≈
2
∫ (2πf )
−∞
(f
f n 2k
2k
+ fn
2k
)[1 + 2(k − 1)( f
]
f n )2( k −1) × T f
−p
.df ,
k >1
(3.2-25)
This can be solved using a table of integrals (eg. Spiegel [86], Equation 15.20) to give the
following result
σ ω2εp ≈
54
4 Tπ 3 f n3− p
k
1
2( k − 1)
sin([3 − p ]π 2k ) + sin([p − 1]π 2k ), 1 < p < 3
(3.2-26)
In Figure 3.2-7, σ ω2εp is plotted as a function of the loop noise bandwidth, Bn , for p=2.5
and T=−25 dBradians2/Hz.
2.5
Variance (radians/s)
2
2
1.5
1
0.5
0
0
5
10
15
20
Noise bandwidth, Bn (Hz)
Figure 3.2-7: Variance of the Doppler error as a function of the loop noise bandwidth for a second
order Costas phase locked loop in the presence of phase scintillations. p = 2.5, T = −25
dBradians2/Hz.
It is clear from this figure that as the loop noise bandwidth increases, the Doppler errors
increase, unlike the phase estimate errors which remain approximately constant. This is
because a wider bandwidth receiver will allow more of the high frequency components of
the phase scintillation energy to be present on the carrier loop phase estimates (high
frequency phase fluctuations contribute more to the Doppler errors than low frequency
fluctuations). The results given in Figure 3.2-7 can be converted into equivalent velocity
errors by multiplying by the factor (c ω T )2 , where c is the speed of light and ω T is the
angular frequency of the GPS carrier (either 2π∗L1, or 2π∗L2). For the spectral strength
value specified in Figure 3.2-7, the variance of the velocity errors is only of the order of a
few tens of (cm/s)2 which is probably negligible for all but a few precision applications .
3.2.5. Summary
In this section, the effects of phase scintillations on a Costas carrier tracking loop was
examined. An expression was derived for the variance of the carrier phase tracking error
in terms of the bandwidth and order of the tracking loop and the spectral strength and
55
spectral index of phase scintillations. By assigning a tracking threshold to this expression,
the conditions under which a Costas loop would be expected to lose carrier lock were
determined. In general, it was found that the susceptibility of a carrier loop to phase
scintillations increased as the loop bandwidth decreased. Also, the effects of phase
scintillations became worse as the spectral strength and the spectral index increased (ie.
as the amount of high frequency phase scintillation energy increased).
Expressions were also obtained for the variance of the phase range errors. Essentially,
these are errors in the carrier loop’s estimate of the satellite-to-receiver range and are
mainly of interest for carrier phase differential GPS. It was found that phase range errors
are predominantly affected by the ionospheric outer scale size as well as the spectral
strength and spectral index of phase scintillations, but show very little dependence on the
loop bandwidth. The outer scale size parameter is a function of the large scale structure of
the ionosphere. As this is not accurately modelled by WBMOD, nor by any of the other
scintillation models known to the author, the evaluation of the phase range error was
considered to be beyond the scope of this thesis.
56
3.3. The impact of amplitude scintillations on carrier
phase tracking loops
In the previous section, the linear loop model was used to obtain a simple closed form
expression for the variance of the carrier phase tracking error for a Costas loop in the
presence of phase scintillations and thermal noise. This expression was then compared
with the linear model threshold derived in Appendix C to determine the strength of
phase scintillation activity required to force the Costas loop to lose lock. An assumption
inherent in this analysis was that the carrier to noise density ratio of the GPS signal was
relatively large, implying that the GPS signal was unaffected by amplitude scintillations.
In this section, the effects of amplitude scintillations on the carrier phase errors will be
examined using both the linear loop model and a non-linear approach. Variance
measures will be derived as a function of the amplitude scintillation index, S 4 , and the
loop noise bandwidth for an I.Q Costas loop. However, as these measures tend to be a
poor indicator of loss-of-lock, an alternative approach will be used in the next section to
determine suitable tracking thresholds for the Costas loop when both amplitude and
phase scintillations are present together.
A complicating factor associated with the analysis of amplitude scintillations is that, if
large enough, they have the capacity to significantly alter the transfer function of the
tracking loop. This causes the effects of amplitude and phase scintillations to be coupled
so that the two must be considered together when deriving a single variance measure.
Unfortunately, this requires a knowledge of the joint statistics of amplitude and phase
which, at this stage, is unknown for scintillations (refer to the end of Section 2.1.3). This
problem can be circumvented to some extent by assuming that the discriminator is
normalised (eg. Q I or Atan (Q I ) ), or that a post-detection AGC is present (for I.Q or
sign(I ).Q discriminators). In doing so, the effects of amplitude scintillations are
translated to the thermal noise term allowing the two effects to be dealt with separately
(phase scintillations and thermal noise are associated with entirely different physical
processes and are therefore independent). In the analysis that follows, the effects of
different AGC time constants will also be examined.
57
3.3.1. Background
As shown by Weber [100], if the bandwidth of the amplitude is relatively small compared
to the loop noise bandwidth, the PDF of the phase errors for a first order phase locked
loop is given by the Tikhonov density function. For the first order Costas phase locked
loop, this is (Equation (C-6) and [43])
fϑ (ϕ ) =
exp(ρe cos(2ϕ ))
πI o (ρe )
,
ϕ ≤
π
2
(3.3-1)
where ϑ = [φε mod π ] is the phase error reduced modulo π, ρe is the effective loop signal
to noise ratio (SNR), and I o is the modified Bessel function of the first kind of order zero.
Although an equivalent expression has not yet been found for higher order loops,
Lindsey and Charles [59] have verified experimentally that the distribution of phase
errors for a second order loop is very close to the Tikhonov density function. As shown
by Viterbi [97], Lindsey and Charles [59] and Holmes [43] (for the Costas loop), a good
approximation to the effective loop SNR is the reciprocal of the variance obtained from
the linear model, viz
ρe =
1
(3.3-2)
4σ φ2ε
where, from Equation (3.2-5)
∫ [1 − H ( f )
∞
σ φ2ε =
−∞
2
]
Sφ p ( f ) + H ( f ) S nd′ ( f ) .df
2
(3.3-3)
= σ φ2εp + σ φ2T
and σ φ2εp and σ φ2T are the contributions to the tracking error variance from phase
scintillations and thermal noise respectively (other phase noise sources have been
ignored). For reasonably slow amplitude fluctuations, the transfer function of the tracking
~
loop, H ( f ) , will be a function of the signal amplitude, A , and the post detection AGC
~
~
gain factor, g (which is also a function of A ). If A and g are assumed to vary slowly with
~
time, then over a time period, τ, for which A and g are approximately constant, the I.Q
Costas loop can be characterised by the following expression (see Figure 3.1-3)
58
ˆ (f )
Φ ετ ( f ) = Φτ ( f ) − Φ
τ
F(f )
~
= Φτ ( f ) − A 2 Φ ετ ( f ) + N dτ ( f ) .
j 2πf . g
[
]
(3.3-4)
where Φτ ( f ) , Φ̂τ ( f ) , Φ ετ ( f ) and N dτ ( f ) are the Fourier Transforms of the random
processes φ (t ) , φˆ(t ) , φε ( t ) and nd (t ) which have been truncated to zero outside of the
time interval 0 to τ seconds. Rearranging this expression gives
(
)
~
N (f )
j 2πf
A2 g F ( f )
. d~τ 2
.Φτ ( f ) −
Φ ετ ( f ) =
~2
~2
j 2πf + A g F ( f )
j 2πf + A g F ( f ) A
(
)
{
(
)
(3.3-5)
}
If it is assumed that E Φτ ( f ). N dτ ( f )* = 0 (as nd (t ) is both zero-mean and independent of
φ (t ) ), the expectation of the power spectral density of φε ( t ) is given by
2
~ 2 1
~ 2 1 N dτ ( f )
2
2
1
E Φ ετ ( f ) = 1 − H ′ f , A E Φτ ( f ) + H ′ f , A E
~
A2
τ
τ
τ
( )
( )
(3.3-6)
~
(
A g )F ( f )
is the modified loop transfer function. In the limit as
( )
~
j 2πf + (A g )F ( f )
2
~
where H ′ f , A =
2
τ → ∞ , this becomes
( )
( )
( )2 S nd~ (4 f )
~
~ 2
~
Sφε f , A = 1 − H ′ f , A Sφ ( f ) + H ′ f , A
A
(3.3-7)
Consequently, as a function of the signal amplitude, the phase error variance based on the
linear model is given by
∞
( ) ∫ Sφε ( f , A~ ).df
~
σ φ2ε A =
−∞
∞
(3.3-8)
∫ 1 − H ′( f , A) Sφ p ( f ) + H ′( f , A)
−∞
~
~
= σ φ2εp (A ) + σ φ2T (A )
=
~
2
~
2
S n d ( f )
.df
~
A4
This expression is useful for determining both an average phase error variance and a
“rule of thumb” tracking threshold and will be discussed further in the next section.
59
3.3.2. Phase errors from the linear model
3.3.2.1. Amplitude scintillations only
If, for the moment, the effects of phase scintillations are ignored, the linear model
variance becomes (from Equation (3.3-8))
( )
∞
( ) ∫ H ′( f , A~ )2 Snd~ 4 f
A
~
σ φ2T A =
.df
(3.3-9)
−∞
This represents the contribution to both the tracking error variance and the variance of
the phase range errors from amplitude scintillations and thermal noise. Equation (D-14)
from Appendix D can be used to reduce the above expression to
σ φ2T
()
2 ~
~
~ σ nd A
A = 2TBn A . ~ 4
A
~
~
= 2TBn A .σ nd′ 2 A
()
()
()
(3.3-10)
()
()
~
where Bn A is the single-sided noise equivalent bandwidth of the tracking loop as a
~
function of the amplitude, nd′ = n d A 2 , and T is the integration period of the
pre-detection filters. Again, based on Equation (D-18) from Appendix D, the
()
~
discriminator noise variance σ n d′ 2 A can be expanded to give (for an I.Q Costas loop)
()
()
N
N
~
~
σ φ2T A = 2TBn A × ~o2 1 + ~o2
TA TA
( )
~
Bn A
=
C No
1
1
~ 4
~ 2 +
2T C N o AN
AN
(3.3-11)
~
~
where C N o = A 2 2 N o , A is the nominal (unperturbed) signal amplitude, and AN = A A
is a normalised signal amplitude. The single-sided noise bandwidth is given by
()
~ 1
Bn A =
2
∞
∫ H ′( f , A) .df
~
2
−∞
(A~ ~g )F ( f ) .df
∫ j2πf + (A g )F ( f )
(A~ g )F ( f ) .df
1
= ∫
~
2
j 2πf + (A
g )F ( f )
1
=
2
∞
2
2
2
−∞
∞
2
2
N
−∞
60
N
2
N
N
(3.3-12)
where g N = g A 2 is a normalised AGC gain factor. For first and second order loops, the
noise equivalent bandwidths become (from Table 3.1-2)
1. First order:
~
~
~ AN 2 ω n AN 2
Bn , where Bn is the design loop noise bandwidth.
Bn A =
=
gN 4 gN
()
2. Second order:
~ ω
Bn A = n
2
()
~ 2
A
1
N ζ + , where ζ is the damping factor.
4ζ
g N
To be consistent with the non-linear analysis that will be presented in the next section,
()
~
only first order loops will be considered. Substituting the above expression for Bn A for
a first order loop into Equation (3.3-11) gives
()
Bn
~
σ φ2T A =
C No
1
1
+
~ 2
g N 2T C N o AN g N
(3.3-13)
For the I.Q Costas loop, the AGC gain factor will be of the form
g=
[
1 k
∑ I Pi 2 + QPi 2
k i =1
[
]
=
~
1 k ~2
Ai + 2 Ai d (t i − τ ) n IPi cos(φε ) + nQPi sin (φε ) + n IPi 2 + nQPi 2
∑
k i =1
=
1 k ~2
∑ Ai + ε g
k i =1
(
)
]
(3.3-14)
where I Pi and QPi are given by Equation (3.1-2), and ε g is the error in the AGC gain.
The k samples in Equation (3.3-14) represent the outputs from the pre-detection filters
over the previous kT seconds. Three different models are considered for the AGC. These
are:
~
1. An ideal AGC for which g = A 2 . This assumes that k = 1 and ε g = 0 .
{ }
~
2. A fast AGC for which g = A 2 + E ε g . In this case, k is assumed to be small
enough to allow the AGC to accurately track the signal amplitude, but large
enough to average the effects of thermal noise.
{ }
3. A very slow (or non-existent) AGC for which g = A 2 + E ε g .
61
From Equation (3.3-14), the expectation of the error in the AGC gain factor is given by
∑ [2 Ai d (ti − τ )(n IPi cos(φε ) + nQPi sin(φε ))+ n IPi 2 + nQPi 2 ]
1
E ε g = E
k
k
{ }
1
=
k
=
~
i =1
∑ [E{n IPi 2 }+ E {nQPi 2 }]
k
(3.3-15)
i =1
2No
T
(from Equation (D - 7), Appendix D)
Consequently, after normalising, the three AGC gain factors become
~
1. Ideal: g N = AN 2 .
2. Fast:
~
g N = AN 2 +
3. Slow:
gN = 1+
where
{ }=
E εg
A
2
1
T C No
(3.3-16)
1
T C No
1
.
T C No
We now substitute the above expressions for g N into Equation (3.3-13) to derive
expressions for the contributions to the tracking error variance from thermal noise and
amplitude scintillations.
Case 1: Ideal AGC
For the ideal AGC model, Equation (3.3-13) becomes
( )
Bn
~
σ φ2T AN =
C No
1
1
~ 2 +
~ 4
2T C N o AN
AN
(3.3-17)
and the average phase error variance is given by
∞
σ φ2T
= σ φ2T (A N ). f A~
∫
N
(A N ).dA N
0
1
Bn 1
1
E ~ 4 ,
=
E ~ 2 +
C N o AN 2T C N o AN
62
(3.3-18)
radians
2
where f A~
N
(A N ) is the PDF of the normalised signal amplitude
fluctuations, f A~
N
~
AN . For slow amplitude
(A N ) is assumed to follow the Nakagami-m distribution with
~
AN 2 = 1 .
The two expectation terms in Equation (3.3-18) can be simplified as follows
1 ∞ 1
E ~ 2 =
f A~ (A N ).dA N
AN 0 A N 2 N
∫
∞
(
)
2 m m A N 2 m −1
.
exp − mA N 2 .dA N
=
2
Γ(m )
AN
∫
1
(3.3-19)
0
∞
(
)
2m m
A N 2m − 3 exp − mA N 2 .dA N
=
Γ(m )
∫
0
From a table of integrals (eg. Spiegel [86], Equation 15.77), this becomes
1 2m m Γ(m − 1)
E ~ 2 =
,
AN Γ(m ) 2m m−1
m
=
m −1
1
, S4 < 1
=
1 − S4 2
m>1
(3.3-20)
Also,
1 ∞ 1
E ~ 4 =
f A~ (A N ).dA N
AN 0 A N 4 N
∫
=
=
m2
2
m − 3m + 2
1
,
1 − 3S 4 2 + 2 S 4 4
(3.3-21)
m>2
,
S4 < 1
2
Therefore, for an ideal AGC the phase error variance can be expressed as
σ φ2T =
Bn
C No
1
1
+
2
2T C N o 1 − 3S 4 2 + 2 S 4 4
1 − S 4
(
)
This expression is only valid for S 4 < 1
S4 > 1
(
,
)
2 . For S 4 = 1
radians 2
(3.3-22)
2 it becomes infinite and for
2 it becomes negative implying that the loop is likely to lose lock (at least for this
AGC model). In Figure 3.3-1, the phase error variance is plotted as a function of S 4 for
C N o = 44 dBHz (typical of a strong satellite signal), and C N o = 30 dBHz (a very weak
63
satellite signal). It is clear from these plots that as S 4 approaches 1
2 , the phase error
variance increases very rapidly. Indeed, the rather conservative tracking threshold of
(π 12)2
radians2 derived in Appendix C is only surpassed when S 4 is larger than about
0.7, even for the weak signal case. S 4 > 0.7 represents quite strong scintillation activity,
and would only be expected to occur in equatorial regions during the evening hours and
under solar maximum conditions. Consequently, at other times and locations, the effects
of amplitude scintillations on the carrier phase errors is likely to be negligible.
x 10
−4
0.02
8
0.018
7
0.016
6
0.014
radians2
radians
2
9
5
0.012
0.01
4
0.008
3
0.006
2
0
0.2
0.4
0.6
0.8
0.004
0
1
0.2
0.4
S4
0.6
0.8
1
S4
Figure 3.3-1: Phase error variance as a function of S 4 for a first order I.Q Costas loop with an
ideal AGC. Parameter values are T = 20ms, Bn = 5Hz, C N o = 44 dBHz (left panel), and C N o =
30 dBHz (right panel).
x 10
−4
0.02
0.018
7
0.016
6
0.014
2
8
radians
radians
2
9
5
0.012
0.01
4
0.008
3
0.006
2
0
0.2
0.4
0.6
0.8
1
0.004
0
0.2
0.4
S4
0.6
0.8
1
S4
Figure 3.3-2: Phase error variance as a function of S 4 from simulations for a first order I.Q
Costas loop with an ideal AGC (the circles denote simulation results). The unmarked curve
represents the theoretical results from Equation (3.3-22). Parameter values are T = 20ms, Bn =
5Hz, C N o = 44 dBHz (left panel), and C N o = 30 dBHz (right panel).
64
In order to verify Equation (3.3-22), a number of simulations were conducted using
simulated amplitude scintillation data and the tracking loop simulator from Appendix B
~
with an ideal AGC (ie. g = A 2 ). In Figure 3.3-2, phase error variance values obtained from
simulations are plotted against S 4 for the two values of C N o used in Figure 3.3-11. It is
clear from these plots that the simulations are in good agreement with the theory,
particularly for values of S 4 less than 1
2 . For larger values of S 4 , the occurrence of
frequent cycle slips complicates the process of estimating the variance. Nevertheless, the
simulation results do confirm that amplitude scintillations are of little concern unless S 4
exceeds 1
2.
Case 2: Fast AGC
For the fast AGC model, the phase error variance of a first order Costas phase locked loop
is given by (from Equations (3.3-13) and (3.3-16))
( )
Bn
~
σ φ2T AN =
C No
1
1
+
~ 2
~ 2 ~ 2
AN + 1 [T C N o ] 2T C N o AN AN + 1 [T C N o ]
[
]
[
]
(3.3-23)
Again, the Nakagami-m PDF can be used to find the average phase error variance as
follows
∞
σ φ2T = σ φ2T (A N ). f A~
∫
N
(A N ).dA N
(3.3-24)
0
Bn
1
1
1
E ~ 2 ~ 2
E ~ 2
,
+
C N o AN + 1 [T C N o ] 2T C N o AN AN + 1 [T C N o ]
=
[
]
[
]
radians
where
∞
1
1
E ~ 2
. f A~ (A N ).dA N = X 1 , and
=
2
AN + 1 [T C N o ] 0 A N + 1 [T C N o ] N
] ∫[
[
]
(3.3-25)
∞
1
1
E ~ 2 ~ 2
. f A~ (A N ).dA N = X 2
=
2
2
AN AN + 1 [T C N o ] 0 A N A N + 1 [T C N o ] N
[
1
] ∫
[
]
The simulation results are represented by the small circles in Figure 3.3-2.
65
If we let I N = AN2 , the first integral becomes
∞
X1 =
1
∫ [I N + 1 (T C N o )]. f I N (I N ).dI N
0
∞
m
m m I N −1 . exp(− mI N )
.dI N
=
Γ(m ) [I N + 1 (T C N o )]
(3.3-26)
∫
0
From a table of integrals (Gradshteyn [37], Equation 3.383-10), we have
∞ ν −1
∫
0
x
. exp(− µx )
.dx = β ν −1. exp(βµ ).Γ(ν ).Γ(1 − ν , βµ )
x+β
where arg(β ) < π , Re(µ ) > 0 , Re(ν ) >
(3.3-27)
∞
02
and Γ(a, b ) = exp(− t )∗ t a −1.dt is the incomplete
∫
b
gamma function. X 1 then becomes
X 1 = m m (1 [T C N o ])m−1 . exp(m [T C N o ]).Γ(1 − m, m [T C N o ])
(3.3-28)
and X 2 becomes
X2 =
mm
(1 [T C N o ])m − 2 . exp(m [T C N o ]).Γ(2 − m, m [T C N o ])
m −1
(3.3-29)
Substituting these two expressions back into Equation (3.3-24) gives
σ φ2T
Bn .m m exp(m [T C N o ])
Γ(2 − m, m [T C N o ])
=
Γ(1 − m, m [T C N o ]) +
m −1
2(m − 1)
C N o (T C N o )
(3.3-30)
Equation (3.3-30) has been used to evaluate the phase error variance as a function of S 4
for C N o = 44 dBHz and C N o = 30 dBHz . This is shown in Figure 3.3-3 along with the
results of a series of simulations based on an AGC with k=10 (the ideal AGC curves from
Figure 3.3-1 are also included for comparison). The simulations show quite good
agreement with the theory when S 4 is less than about 0.9, but tend to produce much
smaller values when S 4 is very large. This is probably the result of a failure to account for
the non-linear behaviour of the tracking loop in Equation (3.3-30). It was also observed in
simulations that for large values of S 4 , cycle slips would only occur when the amplitude
scintillation rate was significantly reduced. This is because the probability of a cycle slip
2
For both X 1 and X 2 , these conditions are met for all values of m.
66
depends not only on the fade depth, but also on the fade duration (see Section 3.5).
Consequently, when the amplitude scintillation rate is reduced, the fade durations
increase and the probability of a cycle slip increases. This suggests that loss-of-lock may
only occur when the amplitude scintillation rate is quite small, even when the strength of
scintillation activity is very large.
−4
9
x 10
0.02
0.018
8
Ideal
Ideal
0.016
0.014
2
6
radians
radians2
7
5
0.012
0.01
4
0.008
3
0.006
2
0
0.2
0.4
0.6
0.8
1
0.004
0
0.2
0.4
0.6
0.8
1
S4
S4
Figure 3.3-3: Phase error variance as a function of S 4 for a first order I.Q Costas loop with a fast
AGC ( k = 10 ). Parameter values are T = 20ms, Bn = 5Hz, C N o = 44 dBHz (left panel), and
C N o = 30 dBHz (right panel). The circles denote simulation results.
Case 3: Slow AGC
For a very slow (or non-existent) AGC, the AGC gain factor is a constant. The phase error
variance is then (from Equations (3.3-13) and (3.3-16))
( )
Bn
~
σ φ2T AN =
C No
1
1
+
~
2
[1 + 1 [T C N o ]] 2T C N o AN [1 + 1 [T C N o ]]
(3.3-31)
and the average variance is given by
∞
σ φ2T
= σ φ2T (A N ). f A~
∫
N
(A N ).dA N
0
=
Bn
C No
1
1
+
2
T
C
N
1
1
o ]] 2T C N o [1 + 1 [T C N o ]]1 − S 4
[ + [
(
,
)
radians 2
(3.3-32)
Variance values obtained from Equation (3.3-32) are plotted in Figure 3.3-4 for
C N o = 44 dBHz and C N o = 30 dBHz (the ideal AGC curves from Figure 3.3-1 have again
been included for comparison). These plots suggest that in the absence of an input phase
process, the phase error variance is only affected by amplitude scintillations when S 4 is
67
very close to 1. The simulations results, which have also been included in Figure 3.3-4,
confirm this result.
9
x 10
−4
0.02
0.018
8
Ideal
Ideal
0.016
7
2
radians
radians2
0.014
6
5
0.012
0.01
4
0.008
3
0.006
2
0
0.2
0.4
0.6
0.8
1
0.004
0
S4
Figure 3.3-4:
0.2
0.4
0.6
0.8
1
S4
Phase error variance as a function of S 4 for a first order I.Q Costas loop with a
slow AGC. Parameter values are T = 20ms, Bn = 5Hz, C N o = 44 dBHz (left panel), and C N o =
30 dBHz (right panel). The circles denote simulation results.
Although these results appear to suggest that a slow AGC is the best choice to overcome
the effects of amplitude scintillations (compare Figure 3.3-3 with Figure 3.3-4), as will be
shown in the next section, a slow AGC will also significantly increase the tracking errors
associated with phase scintillations and dynamics. Consequently, in an overall system
sense, a slow AGC may not necessarily perform any better than a fast AGC (a similar
argument can be used when comparing non-ideal AGC models with the ideal AGC
model).
~
It is clear from Equation (3.3-13) that when AN = 0 (ie. in the absence of a signal), the
phase error variance becomes infinite for all three AGC models. This occurs because a
first order tracking loop becomes an integrator for the white noise process nd (t ).ω n g
when the amplitude is zero (ie. the phase error will be given by φε (t ) =
t
∫ nd (t ).ω n
g .dt
−∞
~
(from Figure 3.1-3 with A = 0 )). Consequently, φε ( t ) will become a Random Walk process
(non-stationary) with a mean of zero and a variance that is proportional to the time t (see
for example Van Trees [96]). Therefore, on occasions when the amplitude approaches zero
(this occurs more often when S 4 is very large), the phase error variance will increase with
68
time and without bound until loss-of-lock occurs. However, if the amplitude recovers
before the loop reaches the point of losing lock, the feedback mechanism in the loop will
be restored and the phase error variance will return to much lower levels. Clearly then,
the duration of the deep amplitude fades will strongly influence both the phase error
variance and the probability of a cycle slip. Consequently, for large values of S 4 , very
slow amplitude scintillations are likely to produce a much larger average variance than
faster scintillations, even when the bandwidths of the two scintillations are much less
than the loop bandwidth. For this reason, the variance corresponding to S 4 ≈ 1 was found
to be highly dependent on the amplitude scintillation rate. The relationship between the
duration of a deep amplitude fade and the probability of a cycle slip will be examined in
more detail in Section 3.5 using a simple rectangular model for the fade profile.
3.3.2.2. Amplitude and phase scintillations
From Equation (3.3-8), the phase scintillation component of the phase error variance is
given by
∞
( ) ∫ 1 − H ′( f , A~ )2 Sφ p ( f ).df
~
σ φ2εp A =
where Sφ p ( f ) =
(f
T
2
o
+f
)
2 p 2
(3.3-33)
−∞
( )
~
(Equation (2.1-1)), and 1 − H ′ f , A =
j 2πf
~2
j 2πf + AN g N F ( f )
(
)
(Equation (3.3-5)). Consequently, the open loop transfer function of the I.Q Costas loop,
~
~
~
F ( f ) , is scaled by a factor α = AN2 g N . For an ideal AGC (ie. g N = AN2 ⇒ α = 1 ), σ φ2εp A
()
ceases to be a function of the signal amplitude and the effects of amplitude and phase
scintillations can be treated separately (ie. the results presented in Section 3.2 will apply
without modification). For a non-ideal AGC ( α ≠ 1 ), both the loop bandwidth and the
damping factor will be influenced by the amplitude. From Table 3.1-2, it is clear that for a
first order loop, both the loop natural frequency, ω n , and the loop bandwidth, Bn , will be
scaled by a factor α. For a second order loop, ω n and the damping factor, ζ, will both be
scaled by a factor
α , while the bandwidth will become equal to
ωn
2
1
αζ + 4ζ .
Consequently, if α is reduced by amplitude fading3, both the bandwidth and damping
3
~
For both of the non-ideal AGC models, a reduction in AN will cause a reduction in α.
69
factors of the two loops will be reduced. This effect will be far more pronounced for a
slow AGC for which α can become very small during periods of deep fading.
For a second order loop with a damping factor of 1
∞
2 , Equation (3.3-33) reduces to
T
( ) ∫ 2 2 f2
.
.df
2
2
2 p 2
(
)
(
)
(
)
f
f
f
f
f
f
2
α
α
−
+
+
−∞
n
n
o
~
σ φ2εp A =
4
(3.3-34)
( )
~
This is plotted in Figure 3.3-5 as a function of the fading intensity, 20 log10 AN , for both
the fast and slow AGC models. Also shown are a pair of horizontal dotted lines which
0
0
−5
−5
2
Variance (dB radians )
Variance (dB radians2)
represent the linear model tracking threshold given by Equation (3.2-9).
Loss−of−lock
−10
Tracking
−15
−20
−25
−30
−25
Loss−of−lock
−10
Tracking
−15
−20
−25
−20
−15
−10
−5
Fading intensity (dB)
0
5
10
−30
−25
−20
−15
−10
−5
0
5
10
Fading intensity (dB)
()
~
Figure 3.3-5: σ φ2εp A as a function of the fading intensity for a first order I.Q Costas loop with a
fast AGC (left panel) and a slow AGC (right panel). Parameter values are T = 20ms, Bn = 5Hz,
C N o = 44 dBHz (lower curve), C N o = 30 dBHz (upper curve), T = −25 dBradians2/Hz,
p=2.5, f o =0 Hz.
It is clear from this figure that when the AGC is unable to track the signal amplitude (the
right panel), the phase errors produced by phase scintillations become very large.
Essentially, the deep fades associated with large values of S 4 cause the instantaneous
bandwidth of the tracking loop to become narrow, resulting in large phase tracking
errors. A similar effect occurs for the fast AGC model when C N o is small (upper curve
in the left panel). Note that the expected value of Equation (3.3-34) could have been found
()
~
by averaging σ φ2εp A
∞
using the Nakagami-m PDF (ie. σ φ2εp = σ φ2εp (A ). f A~ (A ).dA ).
∫
0
70
However, as σ φ2εp provides no information about carrier phase range errors and very
little information about the probability of losing lock, this was not done.
The results given in Figure 3.3-5 are based on the assumption that the amplitude and
phase are independent of one another. In particular, they assume that the rate of change
of phase (embodied in the parameter T) is independent of the amplitude. If this
assumption is invalid, and the rate of change of phase tends to increase as the amplitude
decreases, then the actual variance will be larger than is predicted by Equation (3.3-33)
(ie. Equation (3.3-33) may take on the form
σ φ2εp
∞
( ) ∫ 1 − H ′( f , A~ )2 Sφ p ( f , A~ ).df
~
A =
where
−∞
( ) ( ) ( f o2 + f 2 )− p 2 ). An analysis of simulated scintillation data based on the
~
~
Sφ p f , A = T A
model given in Appendix A suggests that the rate of change of phase is highly correlated
with the fade depth. However, as this model is based on a greatly simplified view of the
real world, these results should be treated with some caution. As yet, there have been no
equivalent studies on real scintillation data to test the validity of this observation [33].
The phase range errors produced by phase scintillations, σ φ2ˆ (see Equation (3.2-19)), will
p
be affected in a similar way by amplitude scintillations. As a function of the signal
amplitude, the variance of the phase range errors is given by
σ φ2ˆ
p
∞
( ) ∫ H ′( f , A~ )2 Sφ p ( f ).df
~
A =
(3.3-35)
−∞
and the average variance is
∞
σ φ2ˆ = ∫ σ φ2ˆ (A N ). f A~
p
p
N
(A N ).dA N
(3.3-36)
0
where f A~
N
(A N ) is the Nakagami-m PDF. As discussed at the beginning of Section 3.2.3,
the variance of the phase range errors tends to be affected more by the outer scale size
parameter, f o , and less by the loop bandwidth. However, if deep fading results in very
narrow loop bandwidth’s, this situation may change, depending on the value of f o . For a
second order loop with a damping factor of 1
2 , Equation (3.3-35) reduces to
71
σ φ2ˆ
p
∞
( ) ∫ 2 2( f2n f )2 + f n 2 . 2 T 2 p 2 .df
α ) + 2( f n f ) ( f o + f )
−∞ ( f n − f
~
A =
2
4
(3.3-37)
The average variance for a second order loop, σ φ2ˆ , is plotted in Figure 3.3-6 for both the
p
slow and fast AGC models (based on Equation’s (3.3-36) and (3.3-37)). It is clear from this
figure that the effects of amplitude scintillations on the phase range errors resulting from
phase scintillations is quite small, even for the slow AGC model and for small values of
0.8
0.75
0.75
2
Variance (radians )
0.8
2
Variance (radians )
C No .
0.7
0.65
0.6
0.2
0.7
0.65
0.3
0.4
0.5
0.6
S4
0.7
0.8
0.9
1
0.6
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
S4
Figure 3.3-6: σ φ2ˆ as a function of S 4 for a first order I.Q Costas loop with a fast AGC (left panel)
p
and a very slow AGC (right panel). Parameter values are T = 20ms, Bn = 5Hz, C N o = 44 dBHz
(lower curves), C N o = 30 dBHz (upper curves), T = −25 dBradians2/Hz, p =2.5, f o = 0.05 Hz.
3.3.2.3. Amplitude scintillations and dynamics
The effects of a changing loop bandwidth are also observed when the loop is subject to
dynamics, particularly when the order of the dynamics4 exceeds the order of the tracking
loop resulting in errors that are not zero-mean. Again, at times when the signal amplitude
is heavily attenuated by scintillations, the instantaneous loop bandwidth will be small
and the loop will become unresponsive to dynamics. If it is assumed that the bandwidth
of the amplitude scintillations is narrow enough for the phase errors to settle at their
steady state values, then using Table E.1 (Appendix E) it is possible to determine the
steady state error as a function of the amplitude. For a second order loop in the presence
of a constant acceleration, the steady state tracking error is given by (from Table E.1)
4
1st order: Phase step, 2nd Order: Constant velocity, 3rd Order: Constant acceleration.
72
φεSS =
2π ao
∗
λ ω n2
radians
(3.3-38)
where ao is the acceleration in m/s2. Under amplitude scintillation conditions, the loop
natural frequency becomes a function of the signal amplitude. For a second order loop
this is given by (based on Equation (3.3-12))
( )
~
ω n AN = α ω n
~
A
= N
g
N
ω n
(3.3-39)
where ω n is the nominal loop natural frequency. The steady state phase error is thus
( )
2π ao
~
φεSS AN =
∗
λ ω n2
gN
~
A2
N
radians
(3.3-40)
An equivalent expression for a first order loop in the presence of a constant velocity, vo , is
( )
2π vo
~
φεSS AN =
∗
λ ωn
gN
~
A2
N
radians
(3.3-41)
Consequently, the steady state errors for the first and second order loops are scaled by the
same factor, 1 α . The biases represented by Equation’s (3.3-40) and (3.3-41) affect the
tracking thresholds for the phase error variance. From Equation (C-3) (Appendix C), a
new tracking threshold can be obtained based on the steady state error, viz
σ φ2ε
1 π
~
AN ) =
(
Th
9 4
( )
~
− φεSS AN
2
radians 2
(3.3-42)
Therefore, in the presence of dynamics, amplitude fading may result in both an increase
in σ φ2ε and a decrease in the threshold, σ φ2ε
Th
. Together, these two effects will result in
an increase in the probability of losing lock.
3.3.2.4. Additional comments
For an ideal AGC, or for a fast AGC for which C N o is large and S 4 is less than 1
2 , the
effects of amplitude scintillations are decoupled from the loop transfer function and from
the effects of phase scintillations. As the AGC normalised noise term, nd′ , is zero-mean
and uncorrelated between successive T second epochs, the phase error variance resulting
73
from thermal noise is given by (from Equation (D-14))
σ φ2T = 2TBnσ n2d′
where (from Equation (D-28))
σ n2d′ =
~2
A
1
E N2
2T C N o g N
1
1
E 2
+
2T C N o g N
~
~
If we assume that g N ≈ AN2 and AN is Nakagami-m distributed, this becomes
σ n2d′ =
1
1
E ~2
2T C N o A
N
1
1
E ~4
+
2T C N o AN
1
1
1
=
+
2
2
4
2T C N o 1 − S 4
2T C N o 1 − 3S 4 + 2 S 4
(
.
)
The phase error variance is thus
σ φ2T =
Bn
C No
1
1
+
2
2
4
2T C N o 1 − 3S 4 + 2 S 4
1 − S 4
(
)
(
)
which is the same as Equation (3.3-22). However, this result has been obtained without
assuming that the bandwidth of the amplitude scintillations is narrow in relation to the
loop noise bandwidth. It only requires that the amplitude bandwidth be narrower than
the pre-detection filter bandwidth, 1 T , that C N o is relatively large, and that S 4 is less
than 1
2 . As this result is based on Equation (D-14), it is also independent of the loop
order.
3.3.3. Phase errors from the non-linear model
In the previous section, Equation (3.3-8) was used to determine an average phase error
variance based on the linear loop model. In this section, the Tikhonov PDF for the
reduced phase error (Equation (3.3-1)) will be used to determine an equivalent non-linear
model variance.
Consider the following conditional form of the Tikhonov PDF
74
( ()
)
( ( ))
~
~ exp ρ e A cos(2ϕ )
fϑ ϕ A =
,
~
πI o ρ e A
( )
()
~
where ρ e A =
1
4σ φ2ε
()
~
A
()
~
and σ φ2ε A
ϕ ≤
π
2
(3.3-43)
is given by Equation (3.3-8) 5. The conditional
Tikhonov PDF can then be used to obtain an expression for the tracking error variance
that is equivalent to Equation (3.3-8), but based on a non-linear model of the tracking
loop, viz
π 2
( ) ∫ ϕ 2 fϑ (ϕ A~ ).dϕ
~
σ ϑ2 A =
(3.3-44)
−π 2
Although this non-linear approach provides a more accurate variance measure than the
linear model, it does have a number of limitations. These are; i) the Tikhonov PDF strictly
only applies to a first order phase locked loop (although Lindsey and Charles [59] have
verified experimentally that the PDF of a second order loop is very similar), ii) it does not
take into account the effects of satellite and receiver dynamics, and iii) the approximation
of setting ρ e equal to the reciprocal of the linear model variance (Equation (3.3-2)) is only
accurate for reasonably high SNR’s (Weber [100], Lindsey and Charles [59] and Viterbi
[97]). This last restriction implies that the effects of phase scintillations can only be
included if S 4 is assumed to be relatively small. However, as shown by Van Trees [96], if
the loop is first order and the spectral index is equal to 2, the phase noise associated with
scintillations can be considered to be equivalent to additional white thermal noise at the
input. This is illustrated in Figure 3.3-7 where the input phase process, φ p (t ) , associated
with phase scintillations has been translated back through the VCO and loop filter to
produce a term γ =
1 dφ p (t )
at the discriminator output.
ω N dt
γ
nd gain control
~
− 0 .5 A 2 sin( 2 φ ε )
+
+
+
+
-
+
ωN
1/s
5
This expression will also be used in Section 6 to determine the bit error rate in the navigation data
in the presence of scintillations.
75
Figure 3.3-7: An equivalent non-linear model of a first order Costas phase locked loop with phase
scintillations translated back through the VCO and loop filter to the discriminator output.
If it is assumed that Sφ p ( f ) = Tf −2 (ie. f o is assumed to be small and p = 2 ), the power
spectral density of γ is given by
Sγ ( f ) =
(2πf )2 S ( f )
φp
ωN
(2π )2 T
=
(3.3-45)
ωN
where (2π )2 T ω N is a constant. Consequently, for a first order loop with p = 2 , γ is
white and can be treated as if it were produced by additive white thermal noise.
Therefore, under these conditions it is quite reasonable to use the non-linear model to
describe the phase errors, even when S 4 is quite large.
An average non-linear model variance can be found by applying the Nakagami-m PDF,
f A~ (A ) , to the variance given by Equation (3.3-44). Thus
∞
σ ϑ2 = ∫ σ ϑ2 (A ). f A~ (A ).dA
(3.3-46)
0
Theoretically, all of the linear model analysis given earlier (apart from the Doppler
analysis) could be repeated using the equations given above in order to obtain equivalent
non-linear model results. If only amplitude scintillations and thermal noise are
considered, Equation (3.3-46) represents both the average tracking error variance and the
average phase range variance from the non-linear model (ie. σ ϑ2 becomes the non-linear
equivalent of σ φ2T
6).
In Figure 3.3-8, σ ϑ2 is compared with both σ φ2T and the results of
simulations for the fast AGC model (from Figure 3.3-3). It is clear from this figure that the
two models give very similar results until S 4 is quite large, at which point the non-linear
model variance begins to fall below the linear model variance.
6
This is used in Section 3.6 to account for the effects of amplitude scintillations and thermal noise
on carrier phase DGPS systems.
76
Notice that because the Tikhonov PDF approaches a uniform distribution when the linear
( )
~
model phase error variance is very large (ie. fϑ ϕ A = 1 π ,
ϕ ≤
π
), the largest possible
2
π 2
value of σ ϑ2 that can be produced by this model is
∫ϕ
2
π .dϕ ≈ 0.82 radians 2 (~7.5 cm2
−π 2
at GPS L1). Consequently, even for very low values of C N o and for S 4 =1, σ ϑ2 will never
exceed this level (indeed, loss-of-lock is likely to occur well before this point).
14
x 10
−4
0.025
Variance (radians2)
Variance (radians2)
12
10
8
6
0.02
0.015
0.01
4
2
0
0.005
0.2
0.4
0.6
0.8
1
S4
0
0.2
0.4
0.6
0.8
1
S4
Figure 3.3-8: σ ϑ2 (lower curves) and σ φ2T (upper curves) as a function of S 4 for a first order I.Q
Costas loop with a fast AGC. Parameter values are T = 20ms, Bn = 5Hz, C N o = 44 dBHz (left
panel), and C N o = 30 dBHz (right panel). The circles denote simulation results.
3.3.4. The effects of pre-detection filtering on phase errors
In Equation (3.1-2), the amplitude following the pre-detection filters is given by
~ 1
A=
T
t
∫ A(u ).du
(3.3-47)
t −T
~
where A is the unfiltered signal amplitude. In general, it has been assumed that A is
Nakagami-m distributed and approximately equal to A. This is based on the observation
that the bandwidth of the pre-detection filters, 1 T , is usually much greater than the
bandwidth of the amplitude scintillations, which is typically less than a few hertz (see
Appendix G). In this section, the validity of this assumption will be examined.
~
The power spectral density of the filtered signal amplitude, A , is given by
77
S A~ ( f ) = G ( f ) S A ( f )
2
(3.3-48)
where G ( f ) = sinc( fT )exp(− jπfT ) is the transfer function of the pre-detection filters, and
S A ( f ) is the power spectral density of the unfiltered signal amplitude. In Figure 3.3-9,
both S A ( f ) and S A~ ( f ) are given for f c =1Hz and p=2.5 ( S A ( f ) = 1 is assumed for f < f c ).
It is clear from this figure that the impact of the pre-detection filters on amplitude
scintillations is insignificant for fluctuation frequencies less than about 10Hz.
Consequently, for typical values of f c (of the order of 1Hz or less), the impact of the predetection filters on the total amplitude variance is likely to be quite small (ie. very little
scintillation energy lies above 10Hz for typical values of p and f c ). However, this may
not be true under high velocity conditions when f c may be quite large on certain
satellite-receiver links.
Sinc(fT)
1
0.5
Amplitude PSD
0
−5
0
5
10
15
20
5
10
15
20
0
−20
−40
−5
0
Frequency (dBHz)
Figure 3.3-9: The impact of a 20ms pre-detection filter on the power spectral density of amplitude
scintillations. The upper panel represents the magnitude of the filter transfer function. The lower
panel represents the power spectral densities of amplitude for p=2.5 and f c = 1 Hz (in the lower
panel, the upper and lower curves represents the unfiltered and filtered amplitudes respectively).
Unfortunately, the power spectral density does not provide a complete picture of the
effects of pre-detection filtering on the amplitude. Using simulated scintillation data
obtained from the model described in Appendix A, it appears that the very deep fades
that cause the greatest loop stress are often quite short in duration. Consequently, the predetection filters are likely to have a greater impact on deep amplitude fades than on
78
shallower fades. In Figure 3.3-10, a scatter plot of the fade depth after filtering versus the
fade depth before filtering is given for a 40s segment of simulated scintillation data for
which
p = 2.5 , and S 4 ≈ 0.75 . Also given are equivalent average plots
f c = 1Hz,
for f c = 0.5Hz, 1Hz and 2Hz based on approximately 20min of scintillation data. It is clear
from these plots that on average, the pre-detection filters significantly attenuate the very
deep fades. It is also clear that for larger values of f c (and so shorter average fade
durations), the effects of filtering become even more apparent. Indeed, it appears that the
average fade depth after filtering tends to plateau at different fade depths depending on
the value of f c . Consequently, for very large values of f c it is possible that pre-detection
filtering on its own may significantly reduce the effects of amplitude scintillations within
the tracking loop. However, as f c is usually much less than about 1Hz for a stationary
receiver (often considerably so – see Appendix G), this effect can probably be ignored for
10
10
5
5
0
0
Filtered fade (dBW)
Filtered fade (dBW)
stationary receivers.
−5
−10
−15
−5
−10
−15
2Hz
1Hz
−20
−20
−25
−40
−25
−40
0.5Hz
−30
−20
−10
0
Unfiltered fade (dBW)
10
−30
−20
−10
0
10
Unfiltered fade (dBW)
Figure 3.3-10: Scatter plot of the fade depth after filtering versus the fade depth before filtering for
a 40s segment of simulated scintillation data with f c = 1Hz , p = 2.5 , and S 4 ≈ 0.75 (left panel).
Equivalent average plots for
f c = 0.5Hz , 1Hz and 2 Hz based on approximately 20min of
simulated scintillation data (right panel). Fade depths are shown relative to a normalised quiescent
signal level of 0dBW.
3.3.5. Summary
In this section, the effects of amplitude scintillations on a Costas carrier tracking loop was
examined. In order to approach this problem using analytical techniques, it was decided
to assume that the discriminator algorithm was I.Q, normalised by a post-detection AGC.
79
As other discriminator types7 are not as amenable to direct analysis as the I.Q
discriminator, a simulation approach would be required in order to assess their
susceptibilities. This was considered to be beyond the scope of this thesis.
In Section 3.3.2, expressions were derived for the variance of the thermal noise errors for
an I.Q Costas phase locked loop in the presence of amplitude scintillations. Three
different AGC cases were considered; (i) an ideal AGC that provided a perfect estimate of
the GPS signal strength, (ii) a fast AGC with a time constant much shorter than the
duration of a typical amplitude fade, and (iii) a very slow AGC. It was shown that the
effects of amplitude scintillations on the phase error variance is negligible, unless the
amplitude scintillation index, S 4 , is very large. It was also shown that for non-ideal
AGC’s, the bandwidth of the tracking loop may fluctuate with the amplitude. If phase
scintillations and Doppler errors are also present, this could result in a significant increase
in carrier tracking errors (ie. phase scintillation errors may become larger during deep
fades as a result of a momentary reduction in the loop bandwidth). Finally, it was
observed that thermal noise errors, and therefore amplitude scintillation effects, increase
with the loop bandwidth. This is the reverse of the situation observed for phase
scintillations.
The analysis carried out in Section 3.3.2 was based on a linearised model of the Costas
carrier loop. In Section 3.3.3, the thermal noise variance was re-calculated using the
Tikhonov PDF, which is based on a non-linear model of a 1st order Costas loop. It was
found that significant variations between the linear model (based on a fast AGC) and the
non-linear model only occurred for very large values of S 4 . However, as shown in the
next section, it is highly likely that loss-of-lock or frequent cycle slips will occur under
these conditions anyway.
7
For example, the arctangent or decision-directed discriminators.
80
3.4. Carrier loop tracking thresholds
The principal objective of this section is to determine the strength of scintillation activity
required to force the Costas loop to lose lock. However, as discussed in Appendix C, it is
difficult, if not impossible, to precisely define a point at which a phase locked loop will
transition to a state of loss-of-lock1. Usually, a threshold is defined beyond which the
linear model approximations are significantly violated. Although this does not guarantee
that the loop will lose lock, it does suggest that the probability of frequent cycle slips and
perhaps loss of lock will become very high.
In Section 3.3.3, it was shown that the average variance measures for the linear and
non-linear loop models, σ ϑ2 and σ φ2ε respectively, diverged when S 4 was very large.
Although this implies that the linear model approximations are being violated, it does not
indicate how frequently this is occurring. For example, a very deep fade for a short
duration may produce the same average linear model variance as a shallower fade for a
longer duration (or a series of shallower fades for shorter durations). Consequently, a
comparison between the average phase error variance measures is not considered to be a
good measure of loss-of-lock.
In this section, the Nakagami-m PDF is used to determine the percentage of time that the
amplitude falls below the tracking threshold for the linear loop model. This is then used
as a basis for deciding whether loss-of-lock is likely to occur in the tracking loop. Inherent
in this approach is the assumption that the bandwidth of the amplitude scintillations is
narrow in relation to the carrier loop bandwidth. This ensures that fade durations below
the tracking threshold are sufficiently long to produce carrier cycle slips and loss-of-lock
(the impact of a reduced fade duration on loop behaviour is discussed in Section 3.5). The
justification for this assumption is that for a power law PSD, the majority of the
amplitude scintillation energy is near the cut-off frequency, f c , which is typically much
less than the loop bandwidth, Bn (see Appendix G). Although this condition is likely to
be met for a stationary or slowly moving receiver, it is not guaranteed under high
dynamic conditions. Under such conditions, the effects of amplitude scintillations may be
1
Loss-of-lock is defined as the point at which the VCO frequency drifts away from the IF frequency
and the phase errors (reduced modulo π) become uniformly distributed.
81
significantly suppressed on certain links, particularly for narrow bandwidth tracking
loops (again, see Section 3.5 for more details).
From Equation (3.3-8), the variance of the phase tracking error resulting from
scintillations and thermal noise is given by
()
()
()
~
~
~
σ φ2ε A = σ φ2εp A + σ φ2T A
()
~
where σ φ2εp A represents the contribution from phase scintillations (Equation (3.3-33))
()
~
and σ φ2T A represents the contribution from thermal noise (Equation (3.3-11)). By
()
~
equating σ φ2ε A
with the linear model tracking threshold, σ φ2ε
Th
(Equation (C-3),
~
Appendix C), a threshold amplitude, ATh , can be obtained below which the tracking loop
would be expected to lose lock, viz
( )
~
σ φ2ε ATh = σ φ2ε
(3.4-1)
Th
~
Unfortunately, a closed form expression for ATh is difficult to obtain for the non-ideal
()
~
AGC models. However, for the ideal AGC model, σ φ2εp A ceases to be a function of the
signal amplitude and is given by Equation (3.2-8). Consequently, the threshold variance
for thermal noise only becomes σ φ2T
Th
= σ φ2ε
Th
− σ φ2εp . By rearranging the standard
expression for the variance of the phase tracking error for an I.Q Costas PLL (Equation
(D-19), Appendix D), the signal amplitude corresponding to this new threshold can be
obtained as follows
[
1+ 1+ β
~
ATh = A
βT C N o
where β =
2σ φ2T
TBn
Th
]
(3.4-2)
, A is the nominal (unperturbed) signal amplitude, and C N o is the
nominal carrier to noise density ratio. In Figure 3.4-1, the normalised threshold
~
amplitude, ( ATh A ), is plotted as a function of the spectral strength, T, for both the ideal
AGC model (solid line) and the non-ideal AGC model (dotted line). It is clear from these
~
plots that the two models return approximately the same value of ATh A . It is also
apparent that for the specified loop bandwidth (5Hz), the carrier loop will lose lock when
82
the spectral strength exceeds about –11 dBradians2/Hz, even in the absence of amplitude
fading.
0
Normalised threshold amplitude (dB)
Normalised threshold amplitude (dB)
0
−5
−10
−15
−30
−25
−20
−15
−1
−2
−3
−4
−5
−6
−7
−8
−30
−10
−25
−20
−15
−10
Spectral Strength, T (dBradians2/Hz)
2
Spectral Strength, T (dBradians /Hz)
~
Figure 3.4-1: The normalised threshold amplitude, ATh A , as a function of the phase scintillation
spectral strength, T, for both the ideal AGC model (dotted line) and the fast AGC model (solid
line). Parameter values are T = 20ms, Bn = 5Hz, p = 2.5, f o = 0.05Hz, C N o = 44 dBHz (left
panel), and C N o = 30 dBHz (right panel).
Using the Nakagamai-m PDF, the probability that the amplitude will drop below the
~
threshold ATh (thus resulting in cycle slips or loss-of-lock) can be found as follows
~
ATh
PL =
∫ f A~ (A ).dA
0
=
2m
m
~
Γ( m).〈 A 2 〉 m
(3.4-3)
~
ATh
∫
A
~
2 m −1 −m. A2 〈 A2 〉
e
.dA
0
In Figure 3.4-2, the probability of losing carrier lock is plotted as a function of S 4 and Bn
for C N o = 41.5 dBHz under the assumption that phase scintillations are not present and
the AGC is ideal (ie. σ φ2εp = 0 is assumed). This figure clearly shows that the probability
of losing lock in the presence of amplitude scintillations increases as the loop bandwidth
increases. However, even for a very wide bandwidth receiver at S 4 =1, this probability is
still quite small for the specified nominal signal level. If the duration of the amplitude
fading is relatively short, this may only result in an occasional cycle slip rather than a
complete loss of signal lock. Indeed, for an inertially aided receiver for which the loop
bandwidth is likely to be very narrow (less than a few Hertz), the effects of amplitude
83
scintillations may be negligible (particularly since an inertial aiding unit will usually
assist the receiver to recover lock once the fading has passed).
2
P(loss of lock) (%)
1.5
1
Bn=20Hz
10Hz
0.5
5Hz
2Hz
0
0.4
0.5
0.6
0.7
0.8
0.9
1
S4
Figure 3.4-2: The probability of losing lock for a 2nd order Costas PLL as a function of S 4 and Bn .
Parameter values are C N o = 41.5 dBHz and σ φ2εp = 0.
Although, Equation (3.4-3) does not provide a clearly defined threshold for S 4 above
which loss of lock will occur, based on simulations it has been found that a threshold of
1% is quite a good choice, particularly for 2nd and 3rd order loops (for 1st order loops, the
frequency of cycle slips merely increases with S 4 with no clearly defined threshold).
Using this somewhat conservative threshold, it is still clear that amplitude scintillations
alone are unlikely to significantly affect GPS receivers, unless the activity is very severe
and the bandwidth of the tracking loop is quite wide. However, as will be shown later,
this does not necessarily apply to codeless and semi-codeless receivers for which the SNR
is significantly reduced.
In Figure 3.4-3, the tolerance of a receiver to scintillations is plotted as a function of T and
S 4 using the approach outlined above. It is clear from this plot that as the loop
bandwidth increases the tolerance to phase scintillation increases while the tolerance to
amplitude scintillation decreases. Indeed, using Wiener filter analysis it is possible to find
an optimum loop bandwidth and order that minimises the phase tracking error, and
therefore the probability of losing lock, for any combination of T and S 4 (See Chapter 8).
84
0
Spectral Strength, T (dBradians2/Hz)
Bn=20Hz
−5
−10
10Hz
5Hz
−15
2Hz
−20
−25
−30
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
S4
Figure 3.4-3: Tracking threshold as a function of T, S 4 , and Bn . Parameter values are
C N o = 41.5 dBHz and p = 2.5.
Figure 3.4-3 is based on the underlying assumption that the spectral strength of phase
scintillations, T, is independent of the amplitude, A (note that this does not imply that T
is independent of S 4 ). With reference to Section 3.3.2.2, if T can be expressed as a function
~
of A, then σ φεp will become a function of A and the derivation of ATh will become more
complicated.
3.4.1. Optimum loop bandwidths
The optimum loop bandwidth for a minimum probability of losing lock is found by
minimising Equation (3.4-3). As the amplitude PDF must always be positive, this can be
~
achieved by minimising the threshold amplitude, ATh , from Equation (3.4-2). The first
~
derivative of ATh with respect to Bn is given by
~
∂ATh
A
=
∂Bn 2 βT C N 1+ 1+β
o
[
]
[
1
1+ 1+β
−
β
2 1+β
] ∂β
∂Bn
(3.4-4)
where β is a function of Bn . For the ideal AGC model, β can be expressed in the following
form (from Equations (3.4-2) and (3.2-8))
85
2 σ φ2ε
β=
where α =
2π
π
k sin([p − 1]π 2k ) ξ
Th
− αTBn(1− p )
TBn
p −1
and ξ =
(3.4-5)
ωn
is a constant for a given loop order. The
Bn
only real, non-trivial roots to Equation (3.4-4) are found by solving
∂β
= 0 to give
∂ Bn
1
Bn opt
( p −1)
pαT
=
σ 2
φε Th
(3.4-6)
~
Equation (3.4-6) represents the bandwidth that minimises ATh , and therefore the
probability of losing lock, for a given phase scintillation spectral strength, T. In Figure
3.4-4, the optimum bandwidth is plotted as a function of T for a phase scintillation
spectral index, p, of 2.5. The corresponding phase error variance due to phase
scintillations is given by
σ φ2ε
Th
σ φ2εp
=
σ φ2ε
Th
p
which is very close to the threshold variance,
, particularly for p close to unity. This implies that the optimisation process
attempts to keep the bandwidth as small as possible (and thus the phase scintillation
~
error as large as possible) in order to minimise ATh .
Optimum Loop Bandwidth, Bn (Hz)
8
7
6
5
4
3
2
1
0
−30
−25
−20
−15
2
Spectral Strength, T (dBradians )
Figure 3.4-4: Threshold loop noise bandwidth as a function of the phase scintillation spectral
strength, T, for the three loop orders (o: 1st order, ˚: 2nd order, ∇: 3rd order loops).
86
3.4.2. WBMOD predictions of T and S4
By combining Equations (2.1-6) (the weak scatter formula for S 4 ) and (2.1-7) (the strong
scatter approximation), it is possible to obtain a relatively simple expression which links
T and S 4 through the various geometrical factors discussed in Section 2.1.3, viz
T=−
where γ ∝
F z F ( p −1)
G v e ( p −1)
(
ln 1 − S 4 2
γ
)
(3.4-7)
is a factor that depends mainly on the satellite-receiver geometry
and the orientation and elongation of the ionospheric irregularities. The logarithm of
Equation (3.4-7) is thus
1
− γ dB
TdB = 10 log10 ln
1 − S 2
4
(3.4-8)
Consequently, for a given geometry factor, γ dB , the locus of possible values of T and S 4
is a single line (at least for the scintillation model discussed in Section 2.1). From
WBMOD, it appears that for a stationary GPS receiver, γ dB is usually greater than about
20dB in equatorial regions above a 5° elevation angle. Therefore, T and S 4 values will
normally lie below the line obtained from Equation (3.4-8) by letting γ dB =20dB. This is
illustrated in Figure 3.4-5 where the line obtained from Equation (3.4-8) with γ dB =20dB is
plotted along with a series of dots representing the T and S 4 values for each visible
satellite at 64 locations near to the equatorial anomaly. Also shown in Figure 3.4-5 are the
threshold curves from Figure 3.4-3.
The T and S 4 values in this figure were obtained from WBMOD under the following
conditions, which represents strong scintillation activity:
• An area of 60 x 60 degrees centred on 100N and 1200E.
• A sunspot number of 150 (ie. high solar activity).
• A magnetic activity index, K p of 5.6.
• The September Equinox.
• 12:00 hrs GMT (~20:00 hrs local time at 1200E).
• 2 hours duration from 12:00 hours GMT.
• A 5° elevation mask angle (elevation angle cutoff).
• The 90th percentile of scintillation activity.
87
Very similar results were obtained at other times and at other equatorial locations.
Spectral Strength, T (dBradians2/Hz)
0
Bn=20Hz
−5
Bn=10Hz
−10
Bn=5Hz
−15
Bn=2Hz
−20
−25
−30
−35
−40
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
S4
Figure 3.4-5: T and S 4 values obtained from WBMOD for a period of high solar activity plotted
over the tracking thresholds from Figure 3.4-3. C N o = 41.5 dBHz is assumed.
This result suggests that a stationary receiver is probably more likely to lose lock from
amplitude scintillations than from phase scintillations, except when the bandwidth is
very narrow (ie. for medium to wide bandwidths, it is unlikely that the threshold will be
exceeded unless S 4 is very large). However, an increase in the effective scan velocity, ve ,
perhaps due to receiver motion, will reduce γ and force the locus in Figure 3.4-5 to move
upwards. As a result, phase scintillations will have more of an effect on loop
performance, as will the choice of the loop bandwidth. However, the optimum
bandwidth for minimum tracking error will still be quite different for each satellite link as
a result of the different geometries and signal levels on each link.
3.4.3. Velocity and elevation angle effects
3.4.3.1. Elevation angle effects
From the equations given in Section 2.1.3, it appears likely that the geometry factors for T
and S 4 will be larger at low elevation angles as a result of a larger ionospheric pierce
point velocity due to satellite motion, v s , and a larger Fresnel zone radius, z F ,
88
respectively. This is demonstrated in Figure 3.4-6 which was obtained using WBMOD
data under the same conditions as Figure 3.4-5.
1.5
x 10
−37
T/CkL
Mean
RMS
1
0.5
0
0
10
20
30
40
50
60
70
80
90
−35
S4w2/CkL
x 10
Mean
RMS
2
1
0
0
10
20
30
40
50
60
70
80
90
Elevation angle (degrees)
Figure 3.4-6: The mean and RMS of the geometry factors for T (upper panel) and S 4 w (lower
panel) obtained from WBMOD for a period of high solar activity.
The two geometry factors used in Figure 3.4-6 were obtained by dividing T (Equation
(2.1-2)) and S 4 obtained from the weak scintillation model (Equation (2.1-6)) by the
height integrated irregularity strength parameter, C k L , viz
T
= K1 ∗ G v e( p −1) sec(θ )
Ck L
(3.4-9)
S4w2
= K 2 ∗ F z F ( p −1) sec(θ )
Ck L
(3.4-10)
(
where, from Equation (2.1-7), S 4 w 2 = − ln 1 − S 42
)
and K1 and K2 are factors that are
independent of the geometry.
For the phase geometry factor, the principal contribution to the elevation angle
dependence originates from the effective scan velocity, ve , and from the sec(θ ) term. As
the elevation angle decreases, the ionospheric pierce point velocity due to satellite
89
motion, v s , increases which in general leads to an increase in ve (see Figure 3.4-7) 2. A
more thorough discussion of the effects of velocity on ve is given in the next section and
in Appendix F.
ve (m/s)
300
200
100
0
0
10
20
10
20
30
40
50
60
70
80
90
30
40
50
60
70
80
90
ve(p−1)
10000
5000
0
0
Elevation angle (degrees)
Figure 3.4-7: Effective scan velocity, ve , as a function of the satellite elevation angle for a
stationary receiver. The elevation angle dependence is largely due to satellite motion and the
individual curves represent different satellite trajectories. p = 2.5.
For the amplitude geometry factor, the Fresnel zone radius, z F , and the sec(θ ) terms
provide the greatest contribution to the elevation angle dependence. z F can be
approximated by z ′F sec(θ ) where z ′F = hi λ is the Fresnel zone radius of a vertically
propagating plane wave. Therefore, the majority of the elevation angle dependence for
S 4 w can be accounted for through a single sec(θ ) ( p +1) 2 term. For highly elongated
irregularities, the Fresnel filter factor, F , tends to be mainly a function of p, and therefore
shows very little dependence on the elevation angle [78].
It is also anticipated that at low elevation angles, the carrier to noise density ratio of the
GPS signal will be reduced. This is a result of a combination of additional atmospheric
absorption, a greater distance to the satellites, and satellite and receiver antenna gain
2
The actual effect on v e will depend on the vector sum of v s and the drift velocity, v d , and on the
orientation and elongation of the irregularities. v e can be quite small if the vector sum of v s and
vd is either very small or approximately aligned to the principal irregularity axis [76].
90
pattern effects. From Equation (3.4-2), it is clear that a reduction in C N o will lead to an
~
increase in the threshold amplitude, ATh , which in turn will increase the probability of
losing lock. Consequently, the threshold curves of Figure 3.4-5 will be shifted to the left,
and the loop will become even more susceptible to the effects of amplitude scintillations.
Consequently, through a combination of a reduced C N o and larger geometry factors,
satellite links that penetrate the peak of the equatorial anomaly at low elevation angles
are likely to be significantly more stressed than high elevation angle links. Notice that this
is not guaranteed to occur on all links as there are large variations in the geometry factors
between the individual links (ie. the RMS values of the geometry factors for both T and
S 4 are also large – see Figure 3.4-6).
3.4.3.2. Satellite and receiver velocity
In the presence of receiver dynamics, two effects will alter a receiver’s tolerance to
scintillations. These are:
(i) the additional stresses imposed directly upon the tracking loops by dynamics, and
(ii) the change in the scintillation rate caused by motion of the receiver through the
interference patterns.
The stresses introduced into the carrier tracking loop by dynamics may be accounted for
if the characteristics of the dynamics are known. If a steady state phase error, φεSS , is
produced by a constant dynamic process, then a modified tracking threshold of the form
2
σ φ2ε
Th 2
π φ
= − εSS radians 2 can be obtained from Equations (C-3) and (E-7) and used
3
12
in place of Equation (3.4-1). The effects of transient errors may be accounted for by adding
an extra term to the tracking error variance to account for the Total Transient Distortion, ε T2
(Equation (E-11)). This results in another modified tracking threshold of the form
2
σ φ2ε
Th 3
π
= − ε T2 . If amplitude scintillations influence the transfer function of a
12
tracking loop, both the steady state and transient errors produced by dynamics will
change. Indeed, during deep fades, the instantaneous bandwidth of the tracking loop
may be reduced, resulting in an increase in both φεSS and ε T2 during those times.
The second effect of dynamics is related more directly to scintillations and can be
accounted for by adjusting the spectral strength parameter, T, and the Fresnel cutoff
91
frequency, f c . As shown in Equation (2.1-2), T can be related to the effective scan
velocity, ve , through the expression T = Bv ep −1 , where B is a constant for a given set of
ionospheric conditions and satellite-receiver geometry. The effective scan velocity is a
function of the ionospheric drift velocity, vd , the ionospheric pierce point velocity, v I
(consisting of the satellite component, v s , and the receiver component, v r ), and the
geometry and orientation of the irregularities. At equatorial latitudes, vd is typically of
the order of 50 to 200 m/s in an Easterly direction and is accounted for in WBMOD
through a drift velocity model. For an ionospheric height of 350km, v s is typically
between about 60 and 500 m/s depending on the elevation and azimuth angles of the
GPS satellites (Appendix F). In Figure 3.4-8, v s is plotted as a function of the elevation
angle over a 24-hour period for a receiver located at 100N and 1200E. It is clear from this
figure that v s shows a strong dependence on the satellite elevation angle, being much
larger at low elevation angles. It is also apparent that despite the effects of Earth’s
rotation, the East-West component of v s is usually in an Easterly direction for all satellites
(ie. v y is usually positive). Consequently, the East-West component of v s is in the same
direction as the drift velocity, vd , which tends to reduce ve somewhat (ie. the ionospheric
pierce point scans across the plasma density contours at a reduced rate as a result of
Vs (m/s)
satellite motion).
500
Vx (m/s)
0
0
500
10
20
30
40
50
60
70
80
90
10
20
30
40
50
60
70
80
90
10
20
70
80
90
0
Vy (m/s)
−500
0
400
200
0
0
30
40
50
60
Elevation angle (degrees)
Figure 3.4-8: Ionospheric pierce point velocity due to satellite motion (upper panel), North velocity
component, v x , (middle panel) and East velocity component, v y , (lower panel) for a GPS receiver
located at 100N and 1200E. The ionospheric height is assumed to be 350km. v s2 = v x2 + v 2y .
92
The pierce point velocity due to receiver motion, v r , is a function of the receiver velocity
and the elevation and azimuth angles of the satellite (Appendix F). In Figure 3.4-9, v r is
plotted as a function of the satellite elevation angle and the direction of motion of the
receiver in relation to the satellite azimuth angle3 ( v r is plotted as a percentage of the
receiver velocity, v R ). In this figure, it is assumed that the receiver is moving in a
horizontal plane with velocity v R and at a height that is small in relation to the mean
ionospheric height (ie. level aircraft flight will meet this requirement). It is clear from this
figure that v r is approximately equal to the receiver velocity, except at moderate to low
elevation angles, and when the receiver motion has a large component that is aligned
with the satellite azimuth angle. Although WBMOD allows only one end of a link to be in
motion (usually the satellite end), it is possible to account for receiver motion by
translating vr to the satellite end and determining an equivalent satellite velocity4.
100
Vr (%)
80
60
40
20
0
300
200
100
Direction (deg)
0 0
20
40
60
80
Elevation Angle (deg)
Figure 3.4-9: Ionospheric pierce point speed as a percentage of the receiver speed and as a function
of the satellite elevation angle and the direction of motion. The ionospheric height is assumed to be
350km.
From the foregoing discussion, it is clear that the speed with which the satellite
propagation path cuts across the plasma density contours is given approximately by
3
00 represents motion of the receiver towards the satellite.
4
If v′s is the pierce point velocity due to the modified satellite motion, then v s′ = v s + v r where v s
is the unmodified satellite pierce point velocity. This approach is discussed further in Appendix F.
93
vd − (v s + v r ) , where vd , v s , and vr are two-dimensional velocity vectors in a horizontal
plane at the mean ionospheric height. For isotropic irregularities, this result is also the
effective scan velocity, ve . However, as the irregularities are in general highly
anisotropic, the degree of anisotropy and the orientation of the irregularities must all be
taken into account in order to determine ve .
In general, very high receiver velocities are likely to lead to an increase in ve , although
not on all satellite links. On average, this will lead to an increase in the spectral strength,
T, with a consequent increase in the probability of losing lock due to phase scintillations.
With reference to Figure 3.4-5, this effect will be manifested as an upward shift in the
curve that represents the likely combinations of T and S 4 . If this shift is sufficiently large
(say 10dB or so), it may cause narrow bandwidth receivers to become much more
susceptible to scintillations than wide bandwidth receivers. Although the amplitude
scintillation strength, S 4 , is not influenced by ve , the Fresnel cutoff frequency, f c ,
increases with ve (Equation (2.1-3)). Consequently, the duration of the deep fading events
that lead to loss of lock in a receiver will be reduced. This will be an advantage for narrow
bandwidth receivers for which the time constant of the tracking loops may exceed the
duration of the deep fading event (see Section 3.5). Although this effect is not accounted
for in the threshold curves of Figure 3.4-5, for very high receiver velocities, it is likely that
loss of lock will mainly occur as a result of phase scintillations (ie. the effects of amplitude
scintillations can be ignored on many of the links).
3.4.4. Summary
Expressions were derived for the probability of losing lock, PL , as a function of both the
tracking loop parameters and the scintillation statistics. It was shown that in general,
amplitude and phase scintillation activity must be at a high level before loss of carrier
lock will occur. It was also shown that as the carrier loop bandwidth increases, the
susceptibility to amplitude scintillations increases, but the susceptibility to phase
scintillations decreases. Consequently, for a given set of signal and scintillation
conditions, an optimum bandwidth exists which minimises the probability of losing lock.
Predictions of the amplitude and phase scintillation indices, S 4 and T, based on the
scintillation model WBMOD suggest that even during times of severe scintillation activity
(ie. during high solar activity, at equinox and near the equator), the majority of a
94
receiver’s channels will remain in lock. This is despite a relatively conservative threshold
being chosen for PL and a large percentile for the WBMOD predictions. However, this
result assumes that the receiver is stationary, uses full code correlation tracking loops and
is not subject to any other sources of loop stress.
The relationship between the geometry of the propagation path and the strength of
scintillation activity was investigated. The intention was to determine under what
conditions a satellite-receiver link would be subject to the greatest scintillation stresses,
and therefore when it would be most likely to lose lock. The two geometry factors
examined in this section were the elevation angle of the propagation path and its velocity
through the ionosphere. It was found that in general, for a given level of ionospheric
disturbance (embodied in the parameter C K L ; the height integrated irregularity
strength), both S 4 and T tended to be larger at low elevation angles. Consequently,
propagation paths that penetrate highly disturbed regions of the ionosphere at low
elevation angles will have the greatest probability of losing lock. It was also found that an
increase in the effective scan velocity of the propagation path through the irregularity
layer, ve , will result in an increase in the phase scintillation index, T, and therefore an
increase in the susceptibility of narrow bandwidth tracking loop to scintillations. ve is a
function of satellite motion, receiver motion, ionospheric drift and the irregularity
geometry. The effects of both satellite and receiver motion on ve were examined and a
technique for incorporating receiver motion into the WBMOD model was discussed.
Although the dependence of ve on the various geometry and velocity factors is quite
complex, it can be said that in general under very high velocity conditions, ve is likely to
increase on most satellite-receiver links which will increase the probability of losing lock,
particularly for narrow bandwidth tracking loops.
95
3.5. The impact of fade depth and duration on cycle
slips
The results given in Sections 3.3 and 3.4 do not explicitly take into account the amplitude
scintillation rate. In these sections, it was assumed that if the amplitude scintillation
bandwidth was much less than the carrier loop bandwidth5, the carrier loop would lose
lock once the amplitude had fallen below the tracking threshold, ATh . However, if the
deep fades occur with very short durations, this assumption may become invalid. This
may occur in the presence of high velocity dynamics, particularly if a receiver is aided by
an Inertial Navigation System (INS) and can therefore adopt a very narrow tracking loop
bandwidth. In this section, the relationship between fade depth and duration and the
probability of a cycle slip is investigated for a simple rectangular fade. Although the issue
of loss-of-lock is not dealt with directly in this analysis, it can be assumed that if the
probability of a cycle slip becomes very large, the carrier loop has a much greater chance
of losing lock.
From Holmes [43], the mean time to cycle slip for a 1st order I.Q Costas phase locked loop
is given by
T =
π2
ρ e I o2 (ρ e )
2 Bn
(3.5-1)
where Bn is the loop noise bandwidth, ρ e = 1 4σ φ2T is the effective loop SNR (the thermal
noise variance, σ φ2T , is obtained from the linear loop model), and I o (
)
is the modified
Bessel function of the first kind of order zero. Based on simulations ([43] page 199), it has
been found that an approximate mean time to cycle slip for a 2nd order loop can be
obtained from Equation (3.5-1) by increasing σ φ2T by 1dB. Viterbi [97] has shown that the
cycle slipping rate is the inverse of the mean time to cycle slip for a 1st order loop, and
approximately so for a 2nd order loop. As the slipping process is approximately Poisson
distributed ([43], page 95), the probability of slipping in τ seconds from a state of zero
error is given by
5
A justification for this assumption is given in Appendix G using data obtained from WBMOD for
a stationary receiver at an equatorial location.
96
PCS = 1 − exp(− τ T )
2τB n
= 1 − exp − 2
π ρ I 2 (ρ )
e
o
e
(3.5-2)
If τ is taken as the duration of the fade, and ρ e is calculated for a given fade depth and
loop bandwidth6, then the probability that one or more cycle slips will occur over the
duration of the fade will be given by this expression.
Equation (3.5-2) implies that for an infinitely deep fade (ie. ρ e = 0 ), the probability of a
cycle slip is one, irrespective of the fade duration, τ. However, Viterbi’s equation is
modelled on an ideal I.Q discriminator for which the discriminator noise term, nd ,
becomes infinite when ρ e = 0 . For a real tracking loop, the characteristics of the
discriminator will tend to override this effect and prevent PCS from becoming large when
the fade duration is very short. Indeed, for an arctangent discriminator, there is a
minimum fade duration below which the probability of a cycle slip is zero, irrespective of
the fade depth. In the analysis that follows, a correction to Equation (3.5-2) is derived for
the case of an infinitely deep fade. The probability of a cycle slip is then given as a
function of fade depth and duration for two representative loop bandwidths and
compared with the results obtained from simulations. The impact of loop order and
dynamics is also discussed.
~
For an infinitely deep fade (ie. A = 0 ), the phase locked loop behaves as if it was an open
~
loop. With reference to Figure 3.1-2, when A = 0 the phase locked loop takes on the
following form
φ
nd
F(s)
+
+
φε
–
φˆ
1/s
Figure 3.5-1: Model of a phase locked loop for zero signal amplitude.
6
Note C N o Faded = C N o Unfaded − F where F is the fade depth in decibels.
97
where nd is the discriminator noise term which is zero-mean but is generally not
Gaussian. nd can be found by applying the discriminator algorithms given in Table 3.1-1
to the prompt I and Q signals from the pre-detection filters. In the absence of a GPS
signal, the I and Q signals have the following form (from Equation (3.1-2))
I P = n IP
Q P = nQP
(3.5-3)
where n IP and nQP are zero-mean, Gaussian and IID. For the arctangent discriminators,
nd is given by
nQP
nd = Atan
n IP
or Atan2 nQP , n IP
(
)
(3.5-4)
7
From studies of narrowband Gaussian noise processes (see for example [39] pages 294
onwards), it is known that the arctangent of the ratio of two zero-mean, Gaussian, IID
random variables has a uniform PDF. Consequently, the PDF of nd is also uniform and
can be represented by
f nd (nd ) =
1
2 3σ nd
n
d
rect
2 3σ n
d
(3.5-5)
where σ n d is the standard deviation of nd and is equal to π 2 3 for the Atan
discriminator, and π
3 for the Atan2 discriminator.
For the normalised I.Q discriminator, the PDF is considerably more difficult to determine.
However, using numerical techniques it has been found that σ nd = 1 2 2 8.
7
Atan2( y , x ) is the four quadrant arctangent function.
8
This is found by evaluating
∞
∞
∫ ∫ nd
2
(
)
−∞ − ∞
bivariate Gaussian PDF and nd = n IP nQP
98
(
)
f nIP ,nQP n IP , nQP .dn IP .dnQP where f nIP ,nQP n IP , nQP is a
(n
2
IP
)
+ nQP 2 .
3.5.1. 1st Order loops
For a 1st order phase locked loop, the phase error γ seconds after the start of an infinitely
deep fade is given by
φε (to , t ) = φ (t ) − φˆ(to , t )
t
= φ (t ) − nd (u )ω n .du + φˆ(to )
to
∫
for t ≥ to
(3.5-6)
where F (s ) = ω n is the transfer function of the loop filter, to is the time at which the fade
begins, t = to + γ , and φˆ(to ) is the loop phase estimate at time to . If it is assumed that
φ (t ) − φˆ(to ) = 0 (ie. the input phase process, φ (t ) , is constant and the initial phase error,
φε (to ) , is zero), then
t
φε (to , t ) = ∫ nd (u )ω n .du
for t ≥ to
(3.5-7)
to
Therefore, φε (to , t ) is a random walk (or Brownian motion) process which begins from
zero at time t = to . As φε (to , t ) is zero-mean, its variance is given by
{
σ φ2ε (γ ) = E φε (to , t )2
}
t
t
= E nd (u )ω n .du n d (v )ω n .dv
to
to
∫
∫
t
= ωn
2
t
∫ ∫ E{nd (u )nd (v )}.du.dv
to to
t t
= ωn2
∫ ∫ σ nd
to to
= ω n σ nd 2
2
2
(3.5-8)
u−v
rect
.du.dv
T
Tγ
Consequently, for a 1st order loop the variance of the phase error increases linearly with
the fade duration, γ. If the magnitude of the phase error exceeds the threshold of the
discriminator, φε
T
9,
a cycle slip will occur when the signal level eventually returns to
normal. Consequently, even for an infinitely deep fade, the probability of a cycle slip will
9
φε
T
= π 2 for an I.Q or Atan discriminator, and π for an Atan2 discriminator.
99
be quite small if the total fade duration, τ, is very small. This is particularly true if the
loop bandwidth, which is proportional to ω n , is also very small.
As the loop is effectively open circuit during an infinitely deep fade, the concept of a cycle
slip becomes meaningless during this time. A cycle slip will only occur if the magnitude
of the phase error exceeds φε
T
when the signal level returns to normal (or at least
becomes large enough for the loop to re-lock). Nevertheless, by comparing the phase
errors for an infinitely deep fade with φε
T
, it is possible to determine an upper limit on
the probability of a cycle slip for different fade durations. This assumes that σ φ2ε (γ ) is at a
maximum value when the fade is infinitely deep.
The Random Walk described by Equation (3.5-7) is essentially a first order Markoff
process for which the PDF of φε at time t k depend only on the value of φε at time t k −1 .
The time separation between consecutive samples in this model (ie. t k − t k −1 ) is given by
the hold period, T, of the integrate and dump pre-detection filters. Consequently, the joint
[
]
PDF of the sequence of phase errors, φε k , φε k −1 , 4, φε 0 is given by
(φε k ,4,φε 0 ) = fφε 0 (φε 0 )∏ fφε
k
fφε
where f
φε i φε i −1
k ,4,φε 0
(φ
εi
i =1
i φ ε i −1
(φ
εi
φε i − 1
)
(3.5-9)
)
φε i −1 is the conditional PDF of φεi given φεi −1 and φε 0 is the phase
error immediately after the start of the fade. For the Atan and Atan2 discriminators, this
is given by 10
f
φε i φε i −1
(φ
)
ε i φε i −1 =
φε − φε
i
i −1
rect
2 3σ n ω n T
2 3σ nd ω n T
d
1
(3.5-10)
which is based on the following discrete version of Equation (3.5-7)
i
φε i = ∑ n d j ω n T
j =1
= φεi −1 + ndi ω n T
10
An equivalent expression for the normalised I.Q discriminator has not been found.
100
(3.5-11)
The probability that the magnitude of the phase error exceeds the threshold, φε
T
, at any
time during the fade is therefore
(
(
)≥ φ )
P max φε k , φε k −1 , 4, φε 0
φε T
=1−
∫
=1−
(2
φε T
4
−φε T
ε T
( )∏ fφε φε (φε i φε i −1 ).dφε k 4dφε 0
0
i
i −1
k
∫
f φ ε φε 0
−φ ε T
φε T
1
3σ n d ω n T
)k −φε∫ T
(3.5-12)
i =1
φε T
4
φε − φε
i −1
rect i
2 3σ n ω nT
i =1
d
k
∫ ∏
−φ ε T
.dφ 4dφ
ε0
εk
Although this has not been solved in closed form, it has been evaluated using Matlab
for a range of fade durations11 and found to correspond very closely with the results of
simulations.
One of the implications of this result is that for an Atan discriminator, it is impossible for
the phase error to exceed φε
T
and produce cycle slips prior to a certain time which is a
function of the loop bandwidth, irrespective of the fade depth (assuming that the link is
not subject to any dynamics and the phase error is initially zero). The reason for this is
that the discriminator noise term, nd , is limited to ± φε
maximum possible phase error at time kT is kTω n φε
T
T
(Equation (3.5-5)), and so the
radians. Consequently, a cycle slip
cannot occur before 1 ω n seconds from the beginning of the fade. Indeed, if this is less
than the pre-detection integration period, T, the pre-detection filters will play a
significant role in limiting the effects of the fade (see Section 3.5.3).
In Figure 3.5-2 and Figure 3.5-3, the probability of a cycle slip, PCS , is plotted as a function
of the fade depth and duration based on Equation (3.5-2) and the correction for an
infinitely deep fade given by Equation (3.5-12). Thus if PCS1 and PCS 2 represent the cycle
slip probabilities from Equations (3.5-2) and (3.5-12) respectively, the corrected curves are
given by min (PCS1 , PCS 2 ) .
11
The total fade duration, τ, is assumed to be kT seconds.
101
100
80
80
60
60
Pcs %
Pcs %
100
40
20
40
20
0
0
5
5
0
0
30
−5
−10
Duration (dBs)
10
30
−5
20
−10
Duration (dBs)
Depth (dB)
20
10
Depth (dB)
Figure 3.5-2: Probability of a cycle slip as a function of the fade depth and duration for a 1st order
Costas carrier tracking loop based on theory. The left panel represents an ideal I.Q discriminator
from Equation (3.5-2). The right panel incorporates the correction based on an infinitely deep fade
100
100
80
80
60
60
Pcs %
Pcs %
for an Atan discriminator. Parameter values are Bn = 15Hz, T=20ms, C N o = 40dBHz.
40
20
40
20
0
0
5
5
0
0
30
−5
−10
Duration (dBs)
20
10
30
−5
−10
Depth (dB)
Duration (dBs)
20
10
Depth (dB)
Figure 3.5-3: Probability of a cycle slip as a function of the fade depth and duration for a 1st order
Costas carrier tracking loop based on theory. The left panel represents an ideal I.Q discriminator
from Equation (3.5-2). The right panel incorporates the correction based on an infinitely deep fade
for an Atan discriminator. Parameter values are Bn = 5Hz, T=20ms, C N o = 40dBHz.
The following observations can be made from these figures:
• For long duration fades (seconds to hundreds of seconds), PCS increases sharply when
the fade depth approaches the theoretical tracking threshold given by Equation (3.4-2)
(16.2 dB for the 15Hz bandwidth tracking loop and 20.3 dB for the 5Hz bandwidth
tracking loop for C N o = 40dBHz). However, as the fade duration decreases,
increasingly larger fade depths are required for the same value of PCS .
102
• Narrow bandwidth tracking loops have a much greater resistance to fades than wide
bandwidth loops.
• The correction for an Atan discriminator based on an infinitely deep fade greatly
reduces PCS for short duration fades, particularly for the narrow bandwidth tracking
loops.
• Although Equation (3.5-2) strictly only applies to an ideal I.Q discriminator, the
correction associated with Equation (3.5-12) should provide an upper limit on PCS for
100
100
80
80
60
60
Pcs %
Pcs %
an Atan discriminator (as σ n d will be a maximum when the fade is infinitely deep).
40
20
40
20
0
0
5
5
0
0
30
−5
−10
Duration (dBs)
10
30
−5
20
−10
Duration (dBs)
Depth (dB)
20
10
Depth (dB)
Figure 3.5-4: Probability of a cycle slip as a function of the fade depth and duration for a 1st order
Costas carrier tracking loop based on simulations. The left panel represents an ideal I.Q
discriminator. The right panel represents an Atan(Q/I) discriminator. Parameter values are
100
100
80
80
60
60
Pcs %
Pcs %
Bn = 15Hz, T=20ms, C N o = 40dBHz.
40
20
40
20
0
0
5
5
0
0
30
−5
−10
Duration (dBs)
20
10
30
−5
−10
Depth (dB)
Duration (dBs)
20
10
Depth (dB)
Figure 3.5-5: Probability of a cycle slip as a function of the fade depth and duration for a 1st order
Costas carrier tracking loop based on simulations. The left panel represents an ideal I.Q
discriminator. The right panel represents an Atan(Q/I) discriminator. Parameter values are
Bn = 5Hz, T=20ms, C N o = 40dBHz.
103
In Figure 3.5-4 and Figure 3.5-5, the probability of a cycle slip is plotted as a function of
fade depth and duration for both an ideal I.Q discriminator and an Atan(Q/I)
discriminator based on simulations12. It is clear from these plots that the simulations
match the theory from Figure 3.5-2 and Figure 3.5-3 quite closely.
The following observations can be made from these figures:
• For the ideal I.Q discriminator, the theory appears to break down for very short
duration fades (ie. less than about 0.1s or so). This may be the result of a failure to
correctly account for the effects of pre-detection filtering in Equation (3.5-2). This will
be discussed further in Section 3.5.3.
• For the Atan(Q/I) discriminator, Equation (3.5-2) appears to provide a good fit for
longer duration fades, and the correction for infinitely deep fades appears to provide a
reasonably good fit for short duration fades.
Simulations were also performed for 1st order Costas loops based on both normalised I.Q
and Atan2(Q,I) discriminators. The results of these tests are given in Figure 3.5-6 for a
15Hz bandwidth tracking loop (this can be compared with the right panels of Figure 3.5-2
100
100
80
80
60
60
Pcs %
Pcs %
and Figure 3.5-4).
40
20
40
20
0
0
5
5
0
0
30
−5
−10
Duration (dBs)
20
10
30
−5
−10
Depth (dB)
Duration (dBs)
20
10
Depth (dB)
Figure 3.5-6: Probability of a cycle slip as a function of fade depth and duration for a 1st order
Costas carrier tracking loop based on simulations. The left panel represents an Atan2(Q,I)
discriminator. The right panel represents a normalised I.Q discriminator. Parameter values are
Bn = 15Hz, T=20ms, C N o = 40dBHz.
12
Several hundred simulations were performed for each combination of fade depth and duration.
104
In both cases, it is clear that the sharp upward trend in PCS occurs at a greater fade depth
than is predicted by the theory. For the Atan2 discriminator, the reason for this is that the
discriminator threshold, φε
T
, is twice as large as the corresponding threshold for an I.Q
or Atan discriminator (ie. π compared with π 2 radians). Consequently, an Atan2
discriminator is capable of tolerating twice as much thermal noise as the other two
discriminators before the non-linear region is encountered. However, for infinitely deep
fades (not shown), it was found that the tracking loop based on an Atan2 discriminator
followed the Random Walk theory very closely (ie. PCS for the Atan2 and Atan tracking
loops converged for infinitely deep fades). However, for the normalised I.Q
discriminator, a tradeoff is occurring between thermal noise errors, and thus PCS , and the
ability of the loop to track dynamics under deep fading conditions. Consequently,
although the normalised I.Q tracking loop appears to perform better at low signal levels
(ie. PCS is generally less), its ability to respond to dynamics tends to be worse. As this
analysis assumes that the link is not subject to any dynamics, this tradeoff is not apparent
in these results. In addition, because the statistics are not known for a normalised I.Q
discriminator, the Random Walk model cannot be used to define PCS under extremely
deep fades.
3.5.1.1. Constant velocity
So far, this analysis has only considered 1st order loops that are not subject to any line of
sight dynamics. One of the reasons for this is that the Fokker-Planck non-linear stochastic
differential equation which defines loop performance (see Appendix C) has only been
solved for a 1st order loop which is driven by thermal noise. When a 1st order loop is
subject to a constant velocity, the phase tracking error will no longer be zero-mean (see
Appendix E) and the probability of a cycle slip will be different. However, if the velocity,
and thus φεSS , is relatively small, the impact on PCS should not be very great. For an
infinitely deep fade, the phase error will consist of a combination of a Random Walk due
to thermal noise, and a linearly increasing phase error resulting from the dynamics (ie.
the loop will be an open circuit and will no longer be capable of tracking the input phase
process, φ (t ) , during the fade). For a constant velocity of ν radians/s, the input phase
process γ seconds after the start of an infinitely deep fade is given by (from Equation
(3.5-6))
105
t
φε (to , t ) = vt − ∫ nd (u )ω n .du + φˆ(to )
to
for t ≥ to
(3.5-13)
If it is assumed that φˆ(to ) = vto − φεSS , where φεSS is the steady state tracking error, then
t
φε (to , t ) = vγ + φεSS − ∫ n d (u )ω n .du
for t ≥ to
to
(3.5-14)
t
= φεSS +
∫ [ v − nd (u )ω n ].du
for t ≥ to
to
An equivalent discrete form of this expression is
i
[
φεi = φεSS + ∑ vT − nd j ω n T
j =1
[
]
(3.5-15)
]
= φεi −1 + vT − ndi ω n T
Consequently, for Atan and Atan2 discriminators, the PDF of φεi given φεi −1 is now
f
φε i φε i −1
(φ
)
ε i φε i −1 =
[
φε − φε + vT
i −1
rect i
2 3σ n ω n T
2 3σ nd ω n T
d
1
]
(3.5-16)
which is the rectangle function centred on φε i −1 + vT . An equivalent form of Equation
(3.5-12) can then be used to determine a new upper limit on the probability of a cycle slip.
Clearly, this new upper limit will depend heavily on the magnitude of the velocity. For
sustained higher order dynamics (eg. a constant acceleration), the 1st order loop will
quickly lose lock, even in the absence of scintillation activity.
3.5.2. 2nd Order loops
For a 2nd order Costas loop, the loop filter takes on the form illustrated in Figure 3.5-7.
Under steady state conditions, the output of the integrator in the upper path provides an
estimate of the line of sight velocity ( v̂ in radians/s) which is then combined with the
phase error estimate provided by the lower path. The resulting filtered phase error is then
passed to the VCO which is represented by the second integrator block. The velocity
estimate, v̂ , provided by the loop filter enables the second order loop to track a constant
velocity with zero steady state error (see Appendix E).
106
φε + nd
ωn 2
v̂
1/s
+
1/s
φˆ
2ζωn
Figure 3.5-7: Loop filter for a 2nd order phase locked loop. ζ is the damping factor.
In the absence of dynamics, it has been shown that the expression for PCS given in
Equation (3.5-2) is a good approximation, provided that σ φT is reduced by 1dB (Viterbi
[97]). However, under infinitely deep fading conditions, the loop filter is once again
driven only by the discriminator noise term, nd , which tends to create both a Random
Walk process (from the lower path), and an integrated Random Walk process (from the
upper path). If velocity is present, the integrated Random Walk component will begin
from an initial value of v̂ radians/s. The phase error γ seconds after the onset of an
infinitely deep fade is therefore
u
t
2
φε (to , t ) = vt −
2ζn (u )ω n + ω n n d (w ).dw + vˆ .du + vˆto
d
to
to
∫
∫
for t ≥ to
(3.5-17)
If it is assumed that vˆ = v and the initial phase error is zero, this reduces to
u
2
φε (to , t ) = 2ζnd (u )ω n + ω n n d (w ).dw.du
to
to
t
∫
∫
t
= 2ζnd (u )ω n .du + ω n 2
∫
to
for t ≥ to
(3.5-18)
t u
∫ ∫ nd (w).dw.du
for t ≥ to
to to
In discrete form, this is given by
i
i
φεi = ∑ 2ζnd j ω n T + ω n 2 ∑
j =1
j
∑ ndm T 2
j =1 m =1
i
=
∑ nd j ω nT [2ζ + ω nT (i − j + 1)]
j =1
= φεi −1 + ω n T 2ζn di +
(3.5-19)
nd j ω nT
j =1
i
∑
Equation (3.5-19) implies that for a 2nd order loop, the phase error is not a first order
Markoff process. Consequently, the result given in Equation (3.5-12) cannot be used to
107
evaluate the probability of a cycle slip. However, it is clear from the additional
[2ζ
+ ω nT (i − j + 1)] factor in Equation (3.5-19) that the phase error will grow much more
rapidly for a 2nd order loop, particularly if the fade duration is quite long (ie. if the time
index, i, is large). Nevertheless, using simulations it has been found that PCS for a 2nd
order loop is only slightly greater than PCS for a 1st order loop and tends to follow the
theory for a 1st order loop quite closely. In Figure 3.5-8, PCS is plotted as a function of fade
100
100
80
80
60
60
Pcs %
Pcs %
depth and duration for a 2nd order loop which is based on an Atan(Q/I) discriminator.
40
20
40
20
0
0
5
5
0
0
30
−5
10
Duration (dBs)
30
−5
20
−10
−10
Depth (dB)
Duration (dBs)
20
10
Depth (dB)
Figure 3.5-8: Probability of a cycle slip as a function of fade depth and duration for a 2nd order
Atan(Q/I) Costas carrier tracking loop based on simulations. The left panel represents a 15Hz
bandwidth tracking loop. The right panel represents a 5Hz bandwidth tracking loops. Parameter
values are T=20ms, C N o = 40dBHz.
3.5.2.1. Constant acceleration
In the presence of a constant acceleration, the phase error for a 2nd order loop has a nonzero mean, φεSS , which is a function of the magnitude of the acceleration and the loop
bandwidth (see Appendix E). The phase tracking error is given by
u
t
at 2
2
ˆ
−
φε (to , t ) =
2ζn (u )ω n + ω n nd (w ).dw + vˆ .du + φ (to )
d
2
to
t o
∫
∫
for t ≥ to
(3.5-20)
at 2
where φˆ(to ) = o − φεSS and vˆ = ato . The error component associated with the dynamics
2
is given by
108
t
at 2
at 2
− ato .du + o − φεSS
2
2
to
∫
(3.5-21)
aγ 2
+ φεSS
2
=
Consequently, the total phase error is given by the Random Walk of Equation (3.5-18)
added to the quadratic error term given above. For sustained higher order dynamics or
very large accelerations, the 2nd order loop would be expected to lose lock irrespective of
the presence of scintillations.
A similar approach can be applied to the analysis of 3rd order loops in the presence of
sustained velocity, acceleration and jerk.
3.5.3. Pre-detection filters
In the previous two sections it has been shown that very narrow fades have a negligible
impact on tracking loop performance, irrespective of their depth. A second factor which
comes into play for very narrow fades is the effect of the pre-detection filters. From
Equation (3.1-2), the I and Q signals at the output of the pre-detection filters is given by
~
I P = A d (t − τ )cos(φε ) + n IP ,
~
Q P = A d (t − τ )sin(φε ) + nQP
(3.5-22)
~
where A is a filtered version of the received GPS signal amplitude. For a rectangular fade
of depth 10 log10 (1 − β ) dB and duration d seconds, the unfiltered amplitude is given by
t
A(t ) = 1 − βrect
d
(3.5-23)
and the filtered amplitude is
β
~
A(t ) = 1 −
T
t
u
∫ rect d .du
(3.5-24)
t −T
~
In Figure 3.5-9, A(t ) is plotted as a function of time for a range of infinitely deep fades (ie.
β=1) with varying durations. It is clear from this figure that for fades less than T seconds,
the pre-detection filters will heavily suppress the fade. Also, as the sample time of the
109
sample and hold circuits is uncorrelated with the fade times, the filtered signal amplitude
is effectively based on a random sampling of these waveforms (at a sample rate of T s).
Normalised Signal Amplitude
1.2
1
0.8
0.6
0.4
0.2
0
−0.2
−0.08 −0.06 −0.04 −0.02
0
0.02
0.04
0.06
0.08
Time (s)
Figure 3.5-9: The filtered signal amplitude as a function of time for a range of infinitely deep fades
with varying durations (4T, 2T, T, T/2, T/4 and T/32 seconds).
3.5.4. Summary
In this section, the impact of fade depth and duration on the probability of a carrier cycle
slip was examined. It was shown that the probability of a cycle slip can become extremely
small if the fade duration is sufficiently short, irrespective of the fade depth. It was also
shown that this behaviour becomes more pronounced as the bandwidth of the tracking
loop is reduced. By assuming an infinitely deep fade and taking account of the
characteristics of the discriminator, it was possible to develop a crude correction to the
standard expression for the probability of a cycle slip which took account of this effect.
Through the use of simulations, it was shown that this correction used in conjunction
with the standard expression produced a relatively accurate measure of the probability of
a cycle slip. The effects of line-of-sight velocity, acceleration, higher loop orders and predetection filtering was also examined.
110
3.6. Scintillation effects on carrier phase differential
GPS
Carrier phase differential GPS techniques can be used to calculate the relative locations of
GPS receivers separated by hundreds of kilometres to centimetre level accuracy in real
time. As scintillations affect both the amplitude and phase of the incoming signals, it is
reasonable to assume that they will impact on the accuracy of the phase range
measurements made by carrier phase DGPS (CPDGPS) receivers. One of the most
important processes involved in CPDGPS is the resolution of the integer cycle
ambiguities in the carrier phase measurements. This involves forming a number of carrier
phase observables including the single difference observable ∆φiab (the difference in the
carrier phase measurements to satellite i from two receivers, a and b), and the double
difference observable ∇∆φijab (the difference in the single difference observables between
two satellites, i and j). Errors in these measurements will increase the time taken for a
receiver to resolve the cycle ambiguities and therefore the time required to obtain an
accurate carrier phase measurement. This is likely to be of greatest concern to systems
that attempt to resolve ambiguities on the fly from moving platforms.
In this section, the effects of amplitude and phase scintillations on the single difference
phase observable, ∆φ , will be examined. If we assume that the ionospheric irregularities
are infinitely long, field aligned, rod like structures [16], then the scintillation patterns on
the ground will show negligible variation in a North-South direction. If the East-West
velocity of the patterns is given by v p m/s (a function of ve ), then for a pair of receivers
separated by S metres in an East-West direction, the component of the single difference
phase observable that is associated with phase scintillations and thermal noise is given by
[
][
∆φ = φˆp1 (t ) + n1 (t ) − φˆp 2 (t ) + n2 (t )
]
(3.6-1)
where,
φˆp1 (t ) = h1 (t ) ⊗ φ p (t ) is the component of the carrier loop phase estimate produced by
phase scintillations at site #1 ( ⊗ denotes the convolution integral),
φˆp 2 (t ) = h2 (t ) ⊗ φ p (t − γ ) is the component of the carrier loop phase estimate produced
by phase scintillations at site #2,
111
h1 (t ) and h2 (t ) are the impulse responses of the two carrier tracking loops,
φ p (t ) is the phase scintillation time series at site #1,
γ = S v p is the time taken for the scintillation pattern to traverse the distance
between the two sites (moving from site #1 to site #2), and
n1 (t ) and n2 (t ) are associated with thermal noise and amplitude scintillations at the
two sites.
Although the amplitude scintillations may be correlated between the two sites, it is quite
straightforward to show that n1 (t ) and n2 (t ) are uncorrelated with each other and with
φˆp (t ) , irrespective of the baseline length. As ∆φ is zero-mean in the presence of phase
scintillations and thermal noise, its variance is given by
{ }
σ ∆2φ = E ∆φ 2
= σ φ2ˆ
where σ φ2ˆ
p1
and σ φ2ˆ
p2
p1
+ σ φ2ˆ
p2
{
}
− 2 E φˆp1 (t ) ∗ φˆp 2 (t ) + σ n21 + σ n22
(3.6-2)
are the variances of the phase estimates associated with phase
scintillations, and σ n21 and σ n22 are the thermal noise variances (given by σ φ2T or σ ϑ2 from
Section’s 3.3.2 or 3.3.3). If we assume that the majority of the phase scintillation energy is
within the bandwidth’s of the two carrier loops, then φˆp1 (t ) ≈ φ p (t ) and φˆp2 (t ) ≈ φ p (t − γ ) .
The variance of the single difference phase observable then becomes
[
]
σ ∆2φ = 2 σ φ2p − Rφ pφ p (γ ) + σ n21 + σ n22
∞
where σ φ2p ≈
∫
Sφ p ( f ).df =
−∞
(3.6-3)
1 p
T π Γ(( p − 1) 2 ) f o −
from Section 3.2.3, Equation (3.2-22), and
Γ( p 2 )
{
Rφ pφ p (γ ) = E φ p (t )φ p (t − γ )
∞
=
∫
}
(3.6-4)
Sφ p ( f ). cos(2πfγ ).df
−∞
is the autocorrelation function of φ p (t ) . The following simplification can be made using a
table of integrals (eg. Gradshteyn [37], Equation 3.771-2)
Rφ pφ p (γ ) =
112
2T π K q (2πf oγ ) πγ
Γ( p 2 )
fo
q
(3.6-5)
where q = ( p − 1) 2 , Γ(x ) is the Gamma function, T is the spectral strength, and K y (x ) is
the Bessel function of imaginary argument. In Figure 3.6-1, the RMS difference in carrier
phase between two GPS receivers is plotted as a function of the baseline length in an EastWest direction in the presence of phase scintillations (the five curves represent different
outer scale size parameters, f o ). These curves assume that the two receivers will be
subject to the same phase scintillation spatial patterns, but with a delay that is a function
of the baseline length, S, and the pattern velocity, v p . In this figure, the effects of thermal
noise and amplitude scintillations have been ignored as they will be independent of the
baseline length (ie. it has been assumed that σ n21 and σ n22 are both zero). This figure
shows that although the single difference error depends on the outer scale size parameter,
f o , in general the baseline length must be less than a few hundred metres in order to
significantly reduce the impact of phase scintillations.
14
fo=0.01Hz
RMS phase error (cm)
12
10
0.03Hz
8
0.05Hz
6
0.07Hz
0.09Hz
4
2
0
0
0.5
1
1.5
2
2.5
3
3.5
4
Baseline (km)
Figure 3.6-1: σ ∆φ as a function of f o and the baseline length in the presence of scintillations.
Parameter values are T = −15 dBradians2/Hz, p = 2, v p = 150m/s), and f = L1.
This analysis assumes that the irregularities are infinitely long, field-aligned, rod like
structures. Under this model, the component of σ ∆2φ which is produced by phase
scintillations is dependent only on the East-West component of the baseline length. For a
pair of receivers placed at two arbitrary locations, the variance of the single difference
observable can be obtained from the previous expression, but with γ replaced by
S sin (α ) v p , where α is the azimuth of the baseline. If it is assumed that the irregularities
are not field aligned, then the decorrelation with distance will be much less dependent on
113
the direction of the baseline and the direction of velocity of the patterns (indeed, for
vertical propagation it will be independent).
In Figure 3.6-2, the RMS phase error from thermal noise and amplitude scintillations ( σ ϑ ,
Section 3.3.3) is plotted as a function of S 4 for a range of loop bandwidths. This result is
based on the fast AGC model and the Tikhonov PDF for the modulo π reduced phase
error, ϑ. By comparing this with Figure 3.6-1, it is apparent that the contribution to the
phase estimate error from thermal noise and amplitude scintillations,
σ n21 + σ n22 , is
quite small compared to the contribution from phase scintillations, even over relatively
short baselines, quite small values of f o and large values of S 4 .
0.45
RMS phase error (cm)
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0.2
Bn=20Hz
10Hz
5Hz
2Hz
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
S4
Figure 3.6-2: The RMS phase error (modulo π) as a function of S 4 and the loop bandwidth.
Parameter values are C/N0 = 38dBHz, T = 0, and f = L1. Simulation results are also shown for
Bn =20Hz (the dotted line).
Using WBMOD, it is theoretically possible to determine σ ∆φ for each of the satellites in a
receiver’s field of view. However, the accuracy of the results will depend on the accuracy
with which the parameters T, p, f o and ve are modelled in WBMOD. Although the
models for T, p and ve are quite sophisticated, the model for f o is rather crude.
Consequently, at this stage WBMOD is not considered to be a very effective tool for
predicting the effects of scintillations on CPDGPS.
114
3.7. Carrier frequency tracking loops
Frequency locked loops (FLLs) or Automatic Frequency Control loops (AFCs) track the
frequency of the GPS carrier and are normally used during times when the carrier phase
is difficult to track or during acquisition1. Indeed, many receivers use FLL assisted PLL’s
and will automatically transition to FLL operation when phase locked loop tracking is no
longer possible [99]. The FLL discriminator estimates the carrier frequency by measuring
the change in the carrier phase over a finite interval of time, ∆t . As these frequency
estimates will in general be sensitive to changes in the sign of the navigation data, they
are usually obtained within the period of a data bit2. The general form of an FLL is very
similar to that of a PLL (compare Figure 3.7-1 with Figure 3.1-1). The principal differences
lie in the discriminator algorithm (see Table 3.7-1) and in the additional integrator prior to
the loop VCO (the loop filter, F (s ) , is identical to the phase locked loop filter from
Table 3.1-2).
pp
1
T
IF
1
T
t
∫ dτ
t −T
Ip
Frequency
Discrim
t
∫ dτ
t −T
F(s)/s
Qp
900
VCO
+
ω IF
Figure 3.7-1: Representation of a generic frequency locked loop.
For small frequency errors (ie. φε 2 − φε1 is small compared to 1 radian, where φεi = φε (ti ) ),
the output of each discriminator is proportional to dφε dt . Under this condition, the
1
However, some receivers such as the Miniature Airborne GPS Receiver or MAGR only track the
carrier frequency.
2
This is not true of decision directed and Atan2 discriminators.
115
linear equivalent circuit of an FLL has a similar form to the linear equivalent circuit of a
PLL (see Figure 3.7-2).
Discriminator
Discriminator output
β
∆t
~
A 2 sin(φε 2 − φε1 )
βsign(α )
∆t
~
A 2 sin(φε 2 − φε1 )
Atan (Q P2 I P2 ) − Atan (QP1 I P1 )
φε 2 − φε1
∆t
∆t
Atan 2( β ,α )
∆t
φ ε 2 − φε 1
∆t
∆t
∆t
~
Table 3.7-1: Typical frequency locked loop discriminators. A is the filtered signal amplitude,
α = I P1 . I P2 + Q P1 .Q P2 , β = I P1 .Q P2 − I P2 .Q P1 , ∆t = t2 − t1 ≤ 10 ms , I P1 and Q P1 are the I and Q
signals at time t1 , and I P2 and Q P2 are the I and Q signals at time t 2 .
nd
+
ν
-
νε
+
~
A2
1/g
+
F(s)
νˆ
1/s
Figure 3.7-2: Linear equivalent model of a frequency locked loop (for either of the first two
discriminator types).
The transfer functions and noise bandwidths of a frequency locked loop are the same as
those given in Section 3.1 for a phase locked loop. The mean-square frequency error is
also very similar to the mean-square carrier phase error and is given by
{ } ∫ [1− H ( f )
∞
E ν ε2 =
2
2
]
Sν ( f ) + H ( f ) S nd ( f ) .df
(3.7-1)
−∞
where f is the fluctuation frequency of the carrier frequency ν, S nd ( f ) is the power
spectral density of frequency errors associated with thermal noise, and Sν ( f ) is the
116
power spectral density of input frequency processes such as dynamics and phase
scintillations.
3.7.1. The impact of phase scintillations on frequency
tracking loops
The power spectral density of the input frequency process is given by
{
}
= E {( j 2πf ) Φ ( f ).(− j 2πf ) Φ ( f ) }
Sν ( f ) = E ν ( f ).ν ( f )*
*
(3.7-2)
= (2πf )2 .Sφ ( f )
where Sφ ( f ) is the power spectral density of the input phase process. For phase
scintillations, Sφ ( f ) = Sφ p ( f ) =
(f
T
2
o
+f2
)
p 2
. As the frequency errors resulting from phase
scintillations are zero-mean, the mean-square frequency error is
∞
σ ν2εp =
2
2
∫ 1 − H ( f ) (2πf ) Sφ p ( f ).df
−∞
∞
=
T(2πf )2
f 2k
(3.7-3)
∫ ( f 2k + f n2k ) ( f 2 + f 2 )p 2 .df
−∞
o
≈ (2π ) T
2
∞
f
2 k + 2− p
∫ ( f 2k + f n2k ).df ,
p < 2k + 2
−∞
For 3 < p < 2k + 2 this becomes
σ ν2εp ≈
(2π )3 T
2kf np−3 sin([p − 3]π
2k )
(3.7-4)
For p < 3 the integral in Equation (3.7-3) is infinite.
Equation (3.7-3) implies that there is no limit to the rate of phase fluctuation that can be
detected by the FLL discriminator. Consequently, very low level but high frequency
phase fluctuations can introduce significant frequency jitter. Under these circumstances,
the frequency error is given by
117
ν ε = ν − νˆ
dφ dφˆ
−
dt
dt
=
(3.7-5)
dφε
dt
=
where φε = φ − φˆ . However, as the incoming phase process, φ , is only sampled every T
seconds, what is actually detected by the discriminator is
νε =
φ 2 − φ1 φˆ2 − φˆ1
−
T
T
(3.7-6)
φ − φε 1
= ε2
T
where φi and φεi imply φ (ti ) and φε (ti ) respectively, and t 2 − t1 = T . Consequently, the
power spectral density of the incoming phase process will be limited to 1 2T Hz.
Although aliasing will cause the power spectral density to fold back on itself for
frequencies beyond 1 2T Hz, we can approximate the effect of sampling by limiting the
integral in Equation (3.7-3) to B 1 2T Hz. By ignoring spectral foldover in this way, the
resulting variance will be slightly less than the actual value. However, the error should
not be too large as the power spectral density falls off according to a power-law
relationship.
The variance obtained by limiting the power spectral density to 1 2T Hz is useful for
determining the tracking state of the FLL. From [47] and [99], the tracking threshold of an
FLL is defined as the point at which the 3σ frequency jitter from all sources equals π 2
radians in one T second period. Therefore, it is given by
σ ν2ε
Th
π
=
6T
2
(radians/s)2
(3.7-7)
Consequently, for the FLL to remain in lock, the following condition must be met
{ }
E ν ε2 < σ ν2ε
{
Th
E [φε 2 − φε 1 ]
118
2
}
, or
2
π
< , radians 2
6
(3.7-8)
The threshold spectral strength for phase scintillations is therefore (from Equation (3.7-3))
T Th =
σ 2
ν ε
(2π )
2
1 2T
− σ ν2T
Th
f
2k + 2 − p
p < 2k + 2
,
(3.7-9)
∫ ( f 2k + f n2k ).df
−1 2 T
where σ ν2T is the thermal noise component of the frequency error (see Appendix D,
Equation (D-21)). In Figure 3.7-3, the difference in the threshold spectral strengths
between a frequency locked loop and a phase locked loop are plotted as a function of the
spectral index and the loop noise bandwidth (Equation (3.2-10) is used for the phase
locked loop threshold). It is clear from this figure that the frequency locked loop is in
general less susceptible to phase scintillations than the phase locked loop (ie. the
difference in thresholds is positive for all values of p and Bn ). This is particularly true for
narrow bandwidth tracking loops and large values of p.
Spectral strength, T (dB)
20
15
10
5
0
20
3
15
10
2
5
Spectral Index, p
1
Noise Bandwidth, Bn (Hz)
Figure 3.7-3: The difference between the threshold spectral strength of a frequency locked loop and
the threshold spectral strength of a Costas phase locked loop. Parameter values are
C N o =41.5dBHz, T=20ms, and k=2 (ie. second order loops).
119
3.7.2. The impact of amplitude scintillations on frequency
tracking loops
From Equation (D-21) in Appendix D, the thermal noise variance in an FLL is given by
σ ν2ε =
4 F3 Bn
1
1 + C N
2
T
T C No
o
(radians/s)2
where F3 = 1 for high C N o and 2 for low C N o (near the tracking threshold). By making
use of the following relationship
C N o Th =
~2
ATh
A2
× C No
(3.7-10)
~
where C N o Th is the threshold carrier to noise density ratio, ATh is the threshold
amplitude, and A is the quiescent or unperturbed signal amplitude, the following
~
expression can be obtained for ATh
1 + 1 + 2α
~
ATh = A
αT C N o
where α =
alone, σ ν2ε
T σ ν2T
Th
2 F3 Bn
Th
, σ ν2T
Th
= σ ν2ε
Th
(3.7-11)
− σ ν2εp is the threshold variance due to thermal noise
is given by Equation (3.7-7), and σ ν2εp is the contribution to the tracking
error variance from phase scintillations (Equation (3.7-3)).
Equation (3.7-11) can be compared with the equivalent expression for a PLL which is
(from Equation (3.4-2))
1+ 1+ β
~
ATh = A
βT C N o
where β =
alone, σ φ2ε
2σ φ2T
TBn
Th
Th
, σ φ2T
Th
= σ φ2ε
Th
− σ φ2εp is the threshold variance due to thermal noise
= (π 12 )2 (Equation (3.2-9), and σ φ2εp is the contribution to the tracking error
variance from phase scintillations (Equation (3.2-8)).
120
(3.7-12)
Consequently, if T and C N o are the same for each loop, the ratio of the two threshold
amplitude values is given by
~
ATh
~
ATh
FLL
PLL
=
[1 + 1 + 2α ]β
[1 + 1 + β ]α
(3.7-13)
If the effects of phase scintillations are ignored (ie. σ ν2εp = σ φ2εp = 0 ), and F3 = 2 is assumed
(ie. near the tracking threshold), Equation (3.7-13) reduces to
2 . Consequently, the
threshold signal power is 3dB higher for an FLL assuming that T is the same for both
loops. As a result, FLL’s will be marginally more susceptible to amplitude scintillations
than PLL’s.
These results suggest that in general, FLL’s are more robust to scintillations than PLL’s.
Also, because FLL’s are much less susceptible to phase scintillations, the optimum
bandwidth of an FLL should be less than the optimum bandwidth of a PLL.
3.8. Conclusions
The analysis carried out in this chapter suggests that in general, the carrier tracking loops
of full code correlation GPS receivers are quite robust to scintillations, even when the
levels of scintillation activity are quite high. It was shown that as the carrier loop
bandwidth increases, the susceptibility to amplitude scintillations increases, while the
susceptibility to phase scintillations decreases. Consequently, an optimum bandwidth
exists for minimum probability of losing lock which depends on the relative contributions
of amplitude and phase scintillation activity, as well as the quiescent signal level and the
presence of dynamics.
For a given level of ionospheric disturbance, the geometry of the propagation path affects
the ratio of amplitude to phase scintillation activity as well as the absolute levels of
scintillation activity. Therefore, geometry will affect both the optimum bandwidth of a
tracking loop and its overall susceptibility to scintillations (for this reason, the optimum
bandwidths will be different for each channel in a receiver). It was found that
propagation paths that penetrate highly disturbed regions of the ionosphere at low
elevation angles generally experienced higher levels of amplitude and phase scintillation
121
activity. It was also found that an increase in the effective scan velocity of the propagation
path through the irregularity layer resulted in an increase in the phase scintillation
spectral strength, and therefore an increase in the susceptibility of narrow bandwidth
tracking loops to scintillations. Although the dependence of the scan velocity on the
geometry and the receiver velocity is quite complex, it can be said that in general, under
very high velocity conditions, the scan velocity is likely to increase on most propagation
paths, thus increasing the probability of losing lock.
Carrier tracking loops are generally very robust to signal fades of short duration,
particularly if the bandwidth of the tracking loop is narrow. Indeed, for fades with a
sufficiently short duration, the probability of a cycle slip can approach zero, irrespective
of the fade depth. However, the precise effect on a tracking loop will depend on the
discriminator algorithm, the quiescent signal level, and the presence of other factors such
as dynamics.
RMS carrier phase errors of several centimetres can be introduced into satellite range
measurements as a result of scintillations. Generally, these errors will become
decorrelated over distances of a few kilometres, depending upon the magnitude of the
ionospheric outer scale size parameter, f o , and the geometry of the baseline. This may
have a significant impact on carrier phase DGPS observations made in equatorial regions
during solar maximum, particularly for baselines of a kilometre or more.
Frequency locked loops are more robust to phase scintillations but slightly less robust to
amplitude scintillations than phase locked loops for the same loop bandwidth and predetection integration period. Therefore, receivers that make use of frequency locked
loops, either as a primary means of carrier tracking or as a fall-back strategy to phase
locked loops, are likely to be more tolerant to scintillations than receivers that employ
only phase locked loops.
Many of the results presented in this chapter are based on the assumption that the phase
scintillation spectral strength, T, is uncorrelated with the amplitude (ie. the rate of change
and strength of phase scintillations are uncorrelated with the amplitude). If T is
negatively correlated to the amplitude (ie. T increases when the amplitude decreases),
then the combined effects of amplitude and phase scintillations on carrier tracking loops
may be greater than is suggested by these results.
122
Chapter 4
Code tracking loops
This chapter examines the effects of scintillations on code tracking loops. In Section 4.1, a
signal processing model of a code tracking loop is given which is based on a generic noncoherent delay locked loop with a normalised Early-Late power discriminator. In Section
4.2, it is shown that the effects of phase scintillations on a code tracking loop are
negligible and can therefore be ignored. In section 4.3, the impact of amplitude
scintillations on the range measurement accuracy of a code loop is examined. It is shown
that unless the levels of amplitude scintillation activity are very large (ie. S 4 close to
unity), the additional thermal noise associated with amplitude scintillations is on average
quite low, although occasional noise spikes may occur when the amplitude undergoes
very deep fades. Finally, in Section 4.4 it is shown that because GPS signals are
narrowband, it is expected that frequency selective scintillation effects will produce
negligible distortion of the PRN codes.
4.1. Code loop model
Figure 4.1-1 is a representation of a generic, non-coherent delay locked loop (DLL). The
function of the DLL is to track the GPS PRN codes and to provide estimates of the code
delay from which pseudorange measurements can be obtained. The DLL mixes the
baseband I and Q signals from the carrier tracking loop with an early code, p E , a prompt
code, p P , and a late code, p L which are produced by a PRN code generator and a 3 bit
shift register1. The resulting early, prompt and late I and Q signals are then filtered by a
bank of pre-detection filters and passed into the DLL discriminator. The function of the
discriminator is to determine the difference in code phase between the received GPS
1
The shift register spacing is typically ½ code chips
123
signal and the replica signal represented by the prompt code, p P . A list of the most
I
Q
L P E
pL pP pE
∫ dt
IE
∫ dt
IP
∫ dt
IL
∫ dt
QE
∫ dt
QP
∫ dt
QL
Code Gen.
Discriminator
common discriminator types and their delay error functions is given in Table 4.1-1.
VCO
τ#%
F(s)
+
+
fc
Carrier
Aiding
Figure 4.1-1: Representation of a generic delay locked loop.
Discriminator algorithm
Discriminator name
Delay error function
∑ (I E − I L ).I P + ∑ (QE − QL ).QP
∑ (I E2 + QE2 )− ∑ (I L2 + QL2 )
Dot product
α (R E − RL ).RP
Early-Late power
α RE 2 − RL 2
Early-Late envelope
β (RE − RL )
∑
I E2 + Q E2 −
∑
I L2 + Q L2
(
)
Table 4.1-1: Common delay locked loop discriminators. α and β are functions of the signal
~
~
amplitude ( α = A 2 , β = A ), R (τ ) is the autocorrelation function of the PRN code,
RE = R (τ ε + Tc 2 ) , RL = R (τ ε − Tc 2 ) , RP = R (τ ε ) , τ ε = τ − τˆ is the delay error and Tc is the
code chip width.
The discriminator errors given in Table 4.1-1 are a function of the autocorrelation function
of the PRN code which is given by
1
R (τ ) =
TP
TP / 2
∫ p(t + τ ). p(t ).dt
− TP / 2
τ
, τ ≤ Tc
1−
=
Tc
0
, τ > Tc
124
(4.1-1)
where Tc is the code chip width (approximately 1µs for the C/A-Code and 100ns for the
P-Code), and TP is the code period (1ms for the C/A-Code and 7 days for the P-Code).
Although the period of the P-Code is extremely long, its autocorrelation function can be
well approximated over relatively short code segments. The delay error function of an
Early-Late power discriminator is illustrated in Figure 4.1-2.
α
Tc/2
-3Tc/2 -Tc -Tc/2
Tc 3Tc/2
0
τε
−α
Figure 4.1-2: Delay error function of an Early-Late power discriminator.
It is clear from this figure that for delay errors in the range − Tc 2 ≤ τ ε ≤ Tc 2 , the delay
error function is a straight line given by − 2ατ ε Tc . Errors larger than this will eventually
cause the code loop to lose lock (ie. the code phase estimate will drift away from the true
code phase). From the discriminator, the code error estimates are passed through a code
loop filter, F (s ) , and then on to the code VCO. The frequency of the code VCO, and
therefore the code chipping rate, is then adjusted in a direction that minimises subsequent
delay errors (this is a very similar process to the one that occurs in a phase locked loop).
In addition, a carrier aiding signal from the carrier loop is used to remove the majority of
the Doppler induced code phase error allowing the use of a much narrower loop
bandwidth and a lower loop order (typically 1st order).
An analysis of the code loop which tackles the issue of scintillation induced amplitude
variations is given below.
The early, prompt and late codes from the code generator are mixed with the baseband I
and Q signals to give (from Equation 3.1-1)
125
I E′ = A(t ) p (t − τ )p (t − τˆ + Tc 2)d (t − τ )cos(φε ) + n ′IE
′
Q E′ = A(t ) p (t − τ )p(t − τˆ + Tc 2)d (t − τ )sin(φε ) + nQE
I P′ = A(t ) p (t − τ ) p(t − τˆ)d (t − τ )cos(φε ) + n ′IP
′
Q P′ = A(t ) p (t − τ )p (t − τˆ)d (t − τ )sin(φε ) + nQP
(4.1-2)
I L′ = A(t ) p (t − τ ) p(t − τˆ − Tc 2)d (t − τ )cos(φε ) + n ′IL
′
Q L′ = A(t ) p (t − τ )p (t − τˆ − Tc 2)d (t − τ )sin(φε ) + nQL
where the superscript ‘ indicates that the signals are at a point immediately prior to the
pre-detection filters. If it is again assumed that the pre-detection filters are synchronised
to the navigation data, and that the carrier phase errors are relatively constant over the
integration period, T, then the I and Q signals will become (only I E is shown)
1
I E = d (t − τ ) cos(φε )
T
t
∫ A(u) p(u − τ ) p(u − τˆ + Tc
2).du + n IE
(4.1-3)
t −T
where σ n2IE = N o T . If the integral in Equation (4.1-3) is divided into L equal parts in
which A(u) is approximately constant, we have
I E = d (t − τ ) cos(φε )
1
T
t −( j −1)T / L
A j p (u − τ ) p(u − τˆ + Tc 2).du + n IE
j =1
t − jT / L
L
∑
∫
(4.1-4)
where A j represents the constant amplitude value at each T L second interval. By
removing A j from the integral and letting L = 20 (giving T L = 1ms which is the period
of the C/A-Code, assuming T=20 ms), the integral in Equation (4.1-4) becomes the
autocorrelation function of the code2. Thus,
1
I E = d (t − τ ) cos(φε )
L
L t −( j −1)T / L
Aj
p(u − τ ) p(u − τˆ + Tc 2).du + n IE
T t − jT / L
j =1
L
∑
1
= d (t − τ ) cos(φε )R (τ ε + Tc 2 )
L
2
∫
(4.1-5)
L
∑ A j + n IE
j =1
This is also an accurate approximation for the P-Code even though its period is much greater than 1ms.
126
~
The amplitude summation term can then be replaced with the filtered amplitude A
giving (for all six signals)
~
I E = A d (t − τ ) R (τ ε + Tc 2 )cos(φε ) + n IE
~
Q E = A d (t − τ ) R (τ ε + Tc 2 )sin(φε ) + nQE
~
I P = A d (t − τ ) R (τ ε )cos(φε ) + n IP
~
Q P = A d (t − τ ) R (τ ε )sin(φε ) + nQP
~
I L = A d (t − τ ) R (τ ε − Tc 2 )cos(φε ) + n IL
~
Q L = A d (t − τ ) R (τ ε − Tc 2 )sin(φε ) + nQL
(4.1-6)
Notice that the prompt I and Q signals contain an R(τ ε ) factor. This was ignored in the
analysis given in Chapter 3 as τ ε was assumed to be zero (ie. perfect code tracking). The
statistics of the various thermal noise terms, n IE , nQE , 4 etc, are examined in Appendix D.
This approximation assumes that the majority of the power in the amplitude scintillation
power spectrum is at frequencies below 1 kHz (ie less than L T Hz). This is considered to
be quite an accurate approximation as the low frequency cutoff in the amplitude
scintillation power spectrum is typically less than a few Hertz, even at very high platform
velocities3. Therefore, for a spectral index, p, of 2.5
4
and a low frequency cutoff, f c , of
1 Hz, the power spectrum of amplitude at 1 kHz will be approximately 75dB below the
cutoff value.
The filtered I and Q signals ( I E , Q E , I P , Q P , I L , Q L ) are processed in the code loop
discriminator to produce code delay errors of the form given in Table 4.1-1. In all cases,
the navigation data and carrier phase error terms will be eliminated by this process.
Consequently, the code loop is referred to as non-coherent (ie. it does not require the
carrier phase to remain in lock).
If it is assumed that the summations in the code loop discriminator are over m successive
periods of the integrate and dump filter (ie. Tm seconds in total), the coefficients of the
delay error function for the three un-normalised discriminators are
3
Scintillation frequencies are determined by the relative velocity between the spatial diffraction patterns and
the receiver. At high relative velocities, the spectrum is translated to higher frequencies.
4
A spectral index of 2.5 is typical of equatorial scintillation activity.
127
αi =
where
i
~
∑ Aj2
j = i − m +1
and βi =
i
~
∑ Aj
(4.1-7)
j = i − m +1
jT
~
Aj =
∫ A(u).du
are the amplitude values after filtering. This assumes that τ ε is
( j −1)T
relatively constant over the time interval Tm (a reasonable assumption if the loop is
expected to track the delay).
The code loop can be represented in an equivalent linear form in which the mixers and
pre-detection filters are replaced by an adder, and the discriminator is replaced by the
appropriate delay error function. For an Early-Late power discriminator, this
representation is given in Figure 4.1-3.
Discriminator
nd
τ
+
+
–
τˆ
τε
− 2α
TC
+
1/s
1/g
δ
F(s)
+
Carrier Aiding
Figure 4.1-3: Linear model of a delay locked loop with an Early-Late power discriminator.
In this form, the delay errors are assumed to be in the linear range, − Tc 2 ≤ τ ε ≤ Tc 2 (for
a ½ chip spacing between the Early, Prompt and Late codes). Consequently, the loop is
assumed to be in lock. In Figure 4.1-3, nd is the thermal noise translated to the
discriminator output and includes cross-terms between the input thermal noise term and
the I and Q signals. An expression for nd for the Early-Late power discriminator is given
in Section 4.3.
The delay on the received GPS PRN code, τ , can be represented in the following way
τ = T +τ d +τ p +τ o
128
(4.1-8)
where T represents the constant contributions to the code delay (including the mean
satellite to receiver range, satellite and receive clock biases, benign ionospheric delays,
benign tropospheric delays and hardware channel biases), τ d represents the effects of
satellite and platform motion, τ p represents the effects of ionospheric phase scintillations
on the code, and τ o represents other delay noise effects (eg tropospheric scintillations and
VCO oscillator jitter etc.).
For a normalised discriminator, the transfer function of the DLL is given by
H ( s) =
T# (s)
T ( s)
=
F ( s)
s + F ( s)
(4.1-9)
The transfer functions and noise bandwidth’s of the delay locked loop are the same as
those defined for the phase locked loop in Table 3.1-2, although the loop order is usually
no greater than 2 because of carrier aiding.
129
4.2. The impact of phase scintillations on code
tracking loops
The mean-square delay tracking error resulting from code delay noise and thermal noise
for the linearised code tracking loop is given by
∞
{ } = ∫ 1 − H ( f )
E τε
2
−∞
2
2
Sτ ( f ) + H ( f ) S nd ( f ). df
(4.2-1)
where Sτ ( f ) is the PSD of the input delay process and S nd ( f ) is the PSD of the thermal
noise, nd . If it is assumed that the carrier phase advance and code group delay are equal
in magnitude under scintillations conditions (as they are under quiescent ionospheric
conditions, Davies [27]), the code delay can be related to the carrier phase advance by
τ p (t ) = −φ p (t )
fP
,
π
2 fL
chips
(4.2-2)
where φ p (t ) is the component of the carrier phase associated with phase scintillations (in
radians), f L is the L-band carrier frequency (L1 = 1575.42 MHz, L2 = 1227.6 MHz) and
f P is the PRN code chipping rate (C/A-Code: 1.023 Mchips/s, P-Code: 10.23 Mchips/s).
One of the undesirable consequences of carrier aiding of the code loop is that the code
error resulting from ionospheric effects is doubled. This occurs because at L-band
frequencies, the carrier phase is advanced and the code phase is delayed by equal
amounts (Davies [27]). Consequently, in the presence of carrier aiding, the component of
the code delay associated with phase scintillations becomes
τ p (t ) = −2 ∗φ p (t )
= −ξφ p (t ),
fP
2πf L
(4.2-3)
chips
In Table 4.2-1, the scaling factor ξ is given for the four combinations of carrier frequency
and code chipping rate which may be encountered in GPS.
130
ξ
C/A-Code
P-Code
L1
2.07x10-4
2.07x10-3
L2
2.65x10-4
2.65x10-3
Table 4.2-1: Scaling factor, ξ, as a function of carrier frequency and code type.
The PSD of the ionospheric delay process is given by
{ ( ) ( )}
= E {ξ F (φ (t ))∗ F (φ (t )) }
Sτ p ( f ) = E F τ p (t ) ∗ F τ p (t ) *
*
2
p
p
= ξ 2 Sφp ( f )
where F (
)
(4.2-4)
chips 2 /Hz
denotes the Fourier Transform. Consequently, the variance of the delay error
resulting from phase scintillations (in chips2) is ξ 2 times smaller than the corresponding
phase error variance for the carrier loop (in radians2) for the same loop order and
bandwidth. Therefore, as the thermal noise errors on the code and carrier loops are of a
comparable size, it is reasonable to expect that phase scintillations will have a negligible
effect on the delay errors (ie. they will be swamped by the effects of thermal noise).
The effects of phase scintillations on the code pseudorange measurements may, however,
be significant, particularly if f o is very small. The variance of the code pseudorange error
is given by
∞
σ τ2ˆp
=
∫ H(f )
2
Sτ ( f ).df
−∞
2
= ξ σ φ2ˆ , chips 2
(4.2-5)
p
= (Tcξ ) σ φ2ˆ , m 2
2
p
Where σ φ2ˆ is the variance of the carrier phase range error due to phase scintillations, and
p
Tc is the code chip width in metres. The low frequency components of the phase
scintillation power spectrum provide the greatest contribution to σ τ2ˆp (ie. σ τ2ˆp is very
sensitive to f o ). However, as these are associated more with the background ionosphere
than with scintillations, it can be argued that phase scintillations will probably have a
negligible effect on the pseudorange.
131
4.3. The impact of amplitude scintillations on code
tracking loops
In this section, an expression is derived for the delay error variance of a delay locked loop
in the presence of amplitude scintillations and thermal noise. To simplify the analysis,
other sources of noise including phase scintillations, dynamics and oscillator phase noise
etc. have been ignored. It is also assumed that code distortion caused by frequency
selective scintillation effects is negligible at GPS frequencies because of the narrow
bandwidth of the GPS PRN codes (see Section 4.4).
The discriminator used in this analysis is an Early-Late power discriminator which is
normalised by a post-detection AGC (see Table 4.1-1). The AGC ensures that the principal
effect of amplitude scintillations is to scale the thermal noise component of the tracking
error rather than altering the loop transfer function. The output of a normalised EarlyLate power discriminator is given by (see Table 4.1-1).
δ=
1
g.k
∑ [(I Ei 2 + Q Ei 2 )− (I Li 2 + QLi 2 )]
k
(4.3-1)
i =1
where k represents the number of T second epochs over which data is averaged in the
discriminator, I E i , Q E i , I Li and Q Li represent the early and late I and Q signals from the
pre-detection filters (see Equation (4.1-6)), and g is the output of a post-detection AGC.
This expression is based on the assumption that the discriminator is operating within the
linear region (ie. − Tc 2 ≤ τ ε ≤ Tc 2 ). The AGC output can be approximated by (assuming
that τ ε is small)
1
g=
k
=
≈
∑ [I Pi 2 + QPi 2 ]
k
i =1
k
1
k
∑ [Ai 2 RP 2 + 2 Ai d (ti − τ )RP (n IPi cos(φε ) + nQPi sin(φε ))+ n IPi 2 + nQPi 2 ]
1
k
∑ Ai 2 + ε g
i =1
k
~
~
(4.3-2)
~
i =1
where RP = R (τ ε ) is the cross-correlation function of the prompt code with the satellite
code and is approximately equal to one for small values of τ ε , and ε g is the thermal noise
component of the AGC gain factor.
132
The expression for the discriminator output can be expanded to give
δ=
1
g.k
∑ [Ai 2 (RE 2 − RL 2 )+ 2 Ai d (ti − τ )(RE n Ei − RL n Li )+ n IEi 2 + nQEi 2 − n ILi 2 − nQLi 2 ]
k
~
~
(4.3-3)
i =1
where RE = R (τ ε + Tc 2 ) and RL = R (τ ε − Tc 2 ) are the cross-correlation functions of the
early and late codes with the satellite code (assuming a ½ chip correlator spacing), τ ε is
the true delay error, and n Ei = n IEi cos(φε ) + nQEi sin(φε ) and n Li = n ILi cos(φε ) + nQLi sin(φε )
are zero-mean, white, Gaussian random variables. The first term in the discriminator
expression (Equation (4.3-3)) is proportional to the delay error and is given by
1
g.k
∑ Ai 2 (RE 2 − RL 2 )
(
) ∑ Ag
k
~
i =1
1
= RE 2 − R L 2
k
2
≈ RE − R L
2
k
i =1
~2
i
(4.3-4)
for ε g ≈ 0
= γτ ε
where γ is the slope of the discriminator delay error function. For delay errors in the
region − Tc 2 ≤ τ ε ≤ Tc 2 , the discriminator slope is
γ =−
2
Tc
(4.3-5)
where Tc is the code chip width. The remaining terms in Equation (4.3-3) represent the
effects of thermal noise and amplitude scintillations and are given by
ndN =
nd
1
=
g
g .k
∑ [2 Ai d (ti − τ )(RE n Ei − RL n Li )+ n IEi 2 + nQEi 2 − n ILi 2 − nQLi 2 ]
k
~
(4.3-6)
i =1
where nd is the equivalent thermal noise term for an un-normalised discriminator. As
n IEi , nQEi , n ILi , and nQLi are zero-mean random variables which are independent5 and
~
identically distributed (IID) and independent of Ai and g 6, the thermal noise term, ndN ,
must also be zero-mean. The variance of ndN is therefore given by
5
For a ½ chip correlator spacing, n IEi , nQEi , n ILi , and nQLi are all independent (see Appendix D).
6
g is a function of Ai , n IPi and nQPi which are all independent of n IEi , n QEi , n ILi , and nQLi for a ½
~
chip correlator spacing.
133
{
σ n2dN = E ndN 2
}
1
= E
2
(g .k )
∑∑ 2 Ai d (ti − τ )(RE n Ei − RL n Li )∗ 2 A j d (t j − τ )(RE n E j − RL n L j )+
k
i =1
~
~
j =1
∑∑ 2 Ai d (ti − τ )(RE n Ei − RL nLi )∗ (n IE j 2 + nQE j 2 − n IL j 2 − nQL j 2 )+
k
2
k
k
~
(4.3-7)
i =1 j =1
∑∑ (n
k
k
i =1 j =1
2
IEi
+ nQEi 2 − n ILi 2 − nQLi 2 ∗ n IE j 2 + nQE j 2 − n IL j 2 − nQL j 2
)(
)
Again, as the four thermal noise terms are zero-mean and IID, the expectation of the
second term in the variance expression is zero, and both the first and third terms are zero
when i ≠ j . Squaring also eliminates the navigation data from the first term (ie.
d (ti − τ )2 = 1 ). The variance expression therefore reduces to
(
) (
{(
)}
)
k
1
~2
2
2
2
2
2 2
4 A
R
n
R
n
n
n
n
n
σ n2dN = E
−
+
+
−
−
i
E
E
L
L
IE
QE
IL
QL
∑
i
i
i
i
i
i
2
(g.k ) i =1
~
1
2
1 k A 2
= 2 ∑ 4 E i2 E RE n Ei − RL n Li 2 + E 2 E n IEi 2 + nQEi 2 − n ILi 2 − nQLi 2
k i =1 g
g
~
1
2
1 k A 2
= 2 ∑ 4 E i2 RE 2 E n Ei 2 + RL 2 E n Li 2 + E 2 4 E nαβi 4 − 4 E nαβi 2
k i =1 g
g
(
{ }
(
{ })
)
(4.3-8)
{ } [ { }]
where α represents either I or Q, and β represents either E or L. From Equation (D-5) in
Appendix D, the variance of nαβi is given by N o T where T is the pre-detection
integration period. Also, as nαβi is a real, zero-mean, Gaussian random variable,
}] = 3(N T ) . In addition, the amplitude sequence is assumed to be
~
~
~
stationary and so E { A }= E { A }= E { A }. Therefore, the variance reduces further to
{
} [{
E nαβi 4 = 3 E nαβi 2
2
2
o
2
2
i
σ n2dN =
2
j
~2
A
1
N
4N o 2
2
R E + R L E 2 + 2 o E 2
k .T
T g
g
(
)
(4.3-9)
For a small delay error, the early and late autocorrelation functions, RE and RL , are
approximately equal to ½ giving
σ n2dN ≈
134
2No
k .T
~2
A
N 1
E
2 + 4 o E 2
T g
g
(4.3-10)
~
The signal amplitude A and the AGC gain factor, g, can be normalised by dividing by the
nominal (unperturbed) signal amplitude, A, as follows
~
A
~
AN = ,
A
gN =
g
A2
(4.3-11)
~
~
By substituting A = A ∗ AN , g = A 2 ∗ g N and C N o = A 2 2 N o into the variance expression,
the following result is obtained
σ n2dN ≈
~ 2
A
1
1
2
E 2
E N 2 +
k .T C N o g N T C N o g N
(4.3-12)
The discriminator noise, ndN (t ) , consists of a sequence of random variables that are
maintained at constant for kT seconds, but are uncorrelated between successive kT second
epochs. Therefore, based on the analysis given in Appendix D for the Costas loop, the
power spectral density of ndN is given by
S ndN ( f ) = σ n2dN kT sinc 2 ( fkT )
(4.3-13)
and the variance of the delay error resulting from thermal noise becomes
σ τ2T =
=
1
γ2
∞
∫ H(f )
2
S n dN ( f ).df
−∞
∞
σ n2dN kT
γ
2
(4.3-14)
∫ H(f )
2
sinc 2 ( fkT ).df
−∞
As the bandwidth of the closed loop transfer function, H ( f ) , must be smaller than the
bandwidth of the sinc( fkT ) function7, the sinc 2 ( fkT ) term can be approximated by one,
giving
σ τ2T ≈
=
7
2σ n2dN kT 1 ∞
2
H ( f ) .df
2
γ
2 − ∞
∫
2σ n2dN
γ
(4.3-15)
kTBn
2
In order to correctly track the desired signal (in this case the code delay process), the bandwidth of the pre-
detection filters must be greater than the design bandwidth of the tracking loop.
135
where Bn =
1
2
∞
∫ H(f )
2
.df is the loop noise bandwidth and γ = − 2 Tc is the slope of the
−∞
delay error function. Combining Equations (4.3-12) and (4.3-14) gives
~
1
Bn Tc 2 AN 2
2
E 2
E 2 +
2 C N o g N T C N o g N
~
2
1
Bn AN
2
E 2
=
E 2 +
2 C N o g N T C N o g N
σ τ2T =
m2
(4.3-16)
chips
2
~
~
If it is assumed that there are no amplitude scintillations (ie. AN 2 = 1 and g N ≈ AN 2 = 1 ),
the delay error variance expression reduces to the standard form for a delay locked loop
(see Equation (D-20), Appendix D), viz
σ τ2T =
Bn
2C No
2
1 + T C N
o
chips 2
(4.3-17)
Discussion
As the code loop is usually aided by Doppler estimates from the carrier loop, its noise
bandwidth can be made quite narrow in order to minimise the effects of thermal noise.
Indeed, if external Doppler aiding is provided by an inertial measurement unit, the code
loop bandwidth can be as small as 0.1Hz [47]. Under these conditions, the bandwidth of
the discriminator errors can be reduced significantly by increasing the size of k (ie.
summing more terms in the discriminator). This will also reduce the effects of amplitude
scintillations by an amount which depends on both the sample rate of the discriminator
( 1 kT ), and the bandwidth of the scintillations. To quantify this effect, it is first necessary
to show that the signal intensity under amplitude scintillation conditions is chi-squared
2
where m = 1 S 42 ).
distributed with 2m degrees of freedom (ie. χ 2m
The PDF of a chi-squared random variable, X, with 2m degrees of freedom is given by [74]
(x 2 )m−1 exp(− x 2 )
, x>0
f X (x ) =
2Γ(m )
0,
x≤0
If we let x = 2m I I
(4.3-18)
~
where I = A 2 is the signal intensity at the output of the pre-
detection filters, and I = A 2 is the average value of I, the PDF of I can be found from
136
f I (I ) = f X (x (I ))
where x (I ) =
dx
dI
(4.3-19)
dx 2m
2mI
=
. This leads to the following expression for the PDF of the
and
I
dI
I
intensity
m m .I m-1
exp(− m.I I ),
f I (I ) = Γ(m ). I m
0,
I>0
(4.3-20)
I≤0
which is the Nakagami-m PDF for intensity (this can be obtained from Equation (2.1-8)
through a simple change of variables). If k successive values of the random variable X are
summed, the number of degrees of freedom of the resulting random variable will be
increased. The amount by which the number of degrees of freedom increases will depend
on the summation period, kT , and the correlation time, TCT , of X. In Haykin, page 246
[39], the correlation time of a zero-mean, wide-sense stationary random process, Z (t ) , is
the time taken for the autocorrelation function, RZ (τ ) , to reduce to a small fraction of
RZ (0 ) (say 1%). For a random process, X (t ) , that is not zero-mean, the correlation time
can be defined as the time taken for the autocovariance function, K X (τ ) , to reduce to a
small fraction of K X (0) . Therefore, the correlation time, TCT , can be defined by
K X (TCT ) = ε
(4.3-21)
where K X (τ ) = R X (τ ) − [E {X (t )} ]2 is the autocovariance function of X (t ) , and ε is a very
small number. Consequently, TCT defines the separation required between successive
samples of X (t ) in order for those samples to be uncorrelated (ie. to have a zero
correlation coefficient). If we make the following definition
1,
α = T TCT ,
1 k ,
TCT ≤ T
T < TCT < kT
TCT ≥ kT
(4.3-22)
k
then a random variable given by Y = α
∑ Xi
is chi-squared distributed with 2mαk
i =1
effective degrees of freedom (ie. Y ~ χ 22mαk ). Y can be related to the normalised AGC gain
factor, g N , in the following way
137
k
Y =α
∑Xi
i =1
1 k
I Ni
= 2mαk
k
i =1
≈ 2mαk ∗ g N
∑
where I Ni = I i
I
(4.3-23)
is the normalised signal intensity. This expression only applies for
moderately deep fades where the effects of thermal noise on the AGC gain factor can be
ignored (ie. ε g ≈ 0 ). If we let m ′ = mαk , then Y = 2m ′ ∗ g N where Y ~ χ 22m ′ . Therefore, g N
is also approximately Nakagami-m distributed but with m = m ′ . The increase in the
number of degrees of freedom (and the consequent decrease in the apparent strength of
scintillation activity) is unlikely to be very significant unless k is quite large and/or TCT is
quite small. Such a situation is most likely to occur when a GPS receiver is aided by an
inertial measurement unit and can therefore adopt a very narrow loop bandwidth (ie. k is
large), and when the GPS ray path is moving rapidly in relation to the irregularity layer
(ie. TCT is small). Both of these conditions may be encountered when a GPS receiver is
operating within a jet aircraft.
4.3.1. Slow amplitude fluctuations
In Chapter 3, the impact of deep, slow fades on the transfer function of a 1st order Costas
carrier tracking loop was accounted for by expressing the loop noise bandwidth as a
function of the amplitude. This allowed the tracking error variance to be expressed as a
function of amplitude, and from there an average variance could be obtained using the
Nakagami-m PDF. An equivalent approach can also be used to analyse the DLL, although
the results will only apply for very slow amplitude fluctuations because of the narrow
bandwidth of most DLL’s.
Using the approach outlined in Section 3.3 for the Costas carrier loop (see Equations
(3.3-4) to (3.3-7)), the following expression can be obtained for the power spectral density
of the delay errors as a function of the signal amplitude (assuming that the amplitude
remains approximately constant for a time period which greatly exceeds the time constant
of the tracking loop)
138
( )
( )2 S n2d ~( f4 )
( )
~
~ 2
~
Sτε f , A = 1 − H ′ f , A Sτ ( f ) + H ′ f , A
γ A
(4.3-24)
The component of the delay error variance associated with thermal noise and amplitude
scintillations is given by
σ τ2T
()
2 S nd ( f )
(
)
.df
~
∫
A4
−∞
2 ~
~
2kTBn (A ) σ n d (A )
=
⋅
1
~
A = 2
γ
∞
~
H′ f ,A
(4.3-25)
~
A4
γ2
()
~
where γ = − 2 Tc . σ n2d A represents the variance of the discriminator noise as a function
~
of the amplitude (ie. conditioned on the random variable A ) for an un-normalised
~
discriminator. An expression for σ n2d A can be obtained from Equation’s (4.3-6) to
()
(4.3-10) by letting g=1. This gives
()
~ 2No
σ n2d A ≈
kT
~2 4No
A + T
(4.3-26)
From Equation (3.3-12), the loop noise bandwidth of a 1st order loop is given by
~
~ A2
Bn A = N
g
N
()
Bn
(4.3-27)
Inserting Equations (4.3-26) and (4.3-27) into (4.3-25) results in
()
~
σ τ2T A =
Bn
2C No
1
2
+
~2
g N T C N o . g N AN
2
chips
(4.3-28)
The expected value of the delay error variance is therefore
σ τ2T =
Bn
C
2 No
1
1
2
E
E
+
~2
g N T C N o g N AN
(4.3-29)
The differences between Equations (4.3-16) and (4.3-29) will only become significant for
large values of S 4 when the probability of a deep fade becomes sufficiently large. The
simplifications applied to the fast AGC model (model #2) for the I.Q Costas loop can be
applied to Equation (4.3-29) to obtain the following result (see Equation’s (3.3-23) to
(3.3-30))
139
σ τ2T =
Bn .m m exp(m T C N o )
2Γ(2 − m, m T C N o )
Γ(1 − m, m T C N o ) +
m −1
(m − 1)
2C N o (T C N o )
(4.3-30)
where Γ(a, b ) is the Incomplete Gamma function (the integral from b to infinity), and
~
g N = AN2 +
1
has been assumed (from Equation (3.3-16)). Equations (4.3-16) and
T C No
(4.3-30) are compared in Figure 4.3-1 over a range of typical code loop bandwidths for the
C/A-Code and for C N o = 38 dBHz. Also included as a dotted line are the results of
simulations for the case Bn = 2 Hz. This result assumes that the dynamics driving the
input delay process is being tracked by the DLL with zero steady state error.
RMS error (m)
15
10
5
Bn=2Hz
1Hz
0.5Hz
0.2Hz
0
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
S4
Figure 4.3-1: The delay error variance as a function of S 4 from Equation (4.3-16) (lower curves)
and (4.3-30) (upper curves). The small circles and the dotted line represent the results of
simulations for Bn = 2 Hz . Parameter values are T=20ms and C N o = 38dBHz .
It is clear from this figure that the two models do not diverge significantly until the
scintillation activity is moderately strong ( S 4 > 0.7 or so), and that the RMS pseudorange
error resulting from scintillations does not become excessive until the scintillation activity
is very strong ( S 4 > 0.9 say). However, this level of activity is likely to be close to, if not
beyond, the tracking threshold of the carrier loop. Moreover, the effect is small when
compared with the RMS error from the background ionosphere which tends to dominate
the pseudorange error budget for a stand-alone SPS receiver. Nevertheless, for DGPS or
140
WADGPS users, the increase in thermal noise errors from amplitude scintillations may
become a significant factor under strong scintillation conditions.
Nominal figures for the RMS thermal noise errors for a standard DLL are 1.5m for the
C/A-Code, and 20cm for the L1 P(Y)-Code ([47] Chapter 7). However, modern receivers
frequently achieve a significant reduction in these levels by smoothing the code
measurements with carrier phase measurements obtained from the carrier loop. In the
absence of significant multipath, this approach typically reduces RMS thermal noise
errors to a few tenths of a metre. This approach will also reduce thermal noise errors in
the presence of amplitude scintillations, provided that the carrier loop remain in lock and
does not suffer from an excessive number of cycle slips. Although this condition is likely
to be met under moderately strong scintillation conditions, as explained below, it may not
hold under very strong scintillation conditions.
The RMS errors plotted in Figure 4.3-1 represent an ensemble average based on a
distribution of GPS signal levels given by the Nakagami-m PDF. As it is assumed that
scintillations are ergodic, these measures also represent the statistics of an individual
realisation of the delay error process. In practice, a time sequence of delay errors obtained
under amplitude scintillation conditions will contain a series of noise spikes, separated by
several seconds or more, which represent momentary increases in the DLL tracking error
resulting from deep fading events. These noise spikes, which become larger and more
frequent as the strength of scintillation activity increases, tend to provide the largest
contribution to the RMS error. Under strong scintillation conditions, it is likely that the
carrier loop will either lose lock or suffer frequent cycle slip during these events which
may preclude carrier smoothing during those times. Consequently, the impact of
amplitude scintillations on the carrier smoothed code observable may be quite
pronounced when the scintillation activity is strong.
The DLL analysis presented in this chapter parallels the linear model analysis for the PLL
in that it is based on the premise that the delay errors are unbounded. However, from
Figure 4.1-2 it is clear that for delay errors beyond ±TC 2 (for an E-L spacing of 1 chip),
the discriminator delay error function will begin to return to zero. Indeed, when the delay
errors exceed ±3TC 2 , the delay error function will become equal to zero and will remain
at that level until the delay error approaches the period of the code sequence (1ms for the
C/A-Code). This behaviour implies that if the magnitude of the delay error exceeds
TC 2 , the feedback mechanism in the DLL will force the delay error beyond 3TC 2 , at
141
which point the loop will lose lock. However, during these times, the reported delay error
will be at a maximum when the true delay error is at ±TC 2 , and will then become
smaller as the true delay error increases beyond this point. Although this non-linear
discriminator model differs significantly from the simple linear model used to obtain
Equations (4.3-16) and (4.3-30), the differences are only significant when the delay errors
become very large. From the linear model, it is clear that this will occur when the
amplitude is heavily attenuated during deep fading events. Indeed, even if the PLL was
to remain in lock during these events, the DLL may lose lock, and be unable to regain lock
until the fade has passed and the loop has reverted to re-acquisition mode (the DLL
cannot cycle slip and then regain lock in the same way as the PLL). In the case of the
Costas PLL, the non-linear behaviour of the discriminator was accounted for by making
use of the Tikhonov PDF for the modulo π reduced phase error. For the DLL, an
equivalent non-linear PDF does not exist for the delay error. A crude approximation that
has been applied to this problem has been to formulate the RMS error as a function of the
amplitude using Equation (4.3-28), and to then restrict the errors to some upper limit
based on a nominal tracking threshold for the DLL ( d 3 code chips, where d is the
correlator spacing [47]). As before, the RMS error is then found by averaging over all
possible amplitude values using the Nakagami-m PDF. This procedure results in an RMS
error curve which lies slightly above the curve obtained from Equation (4.3-16). This
suggest that for moderately strong scintillation activity (ie. S 4 > 0.7 or so), the depth and
frequency of the deep fading events which force delay errors beyond the tracking
threshold is sufficiently large to significantly affect the average RMS delay error.
Consequently, for large values of S 4 , the analytical expressions for the RMS delay error
must be treated with some caution.
Nevertheless, based on simulations for which the cutoff frequency of the amplitude
scintillation power spectrum, f c , is much less than Bn , the RMS delay error appears to
follow the upper curve given by Equation (4.3-30). For larger values of f c , the RMS errors
become smaller, particularly when S 4 is large.
142
4.4. Frequency-selective scintillation effects
In the analysis so far, it has been assumed that all frequency components in the GPS code
spectrum undergo the same amplitude and phase variations at the receiver. This
situation, which is also referred to as frequency-flat or frequency-nonselective fading (see for
example [72], [75] or [88]), may produce fluctuations in the amplitude and delay of a code
sequence, but will not result in any code pulse distortion. Consequently, tracking errors
and pseudorange noise will be produced mainly by additive thermal noise, assuming that
phase scintillations have a negligible effect.
However, if the bandwidth of the code is sufficiently large and/or the scintillation
activity is sufficiently strong, variations will exist in the amplitude and phase of the
scintillation waveforms across the code spectrum (ie. the frequency response of the
propagation path will no longer be flat). If these variations are large enough, they will
produce distortions in the code sequence which may affect the performance of the code
loop discriminator. When a channel’s frequency response exhibits statistical decorrelation
across its bandwidth, it is said to be frequency-selective. The most common source of
frequency-selective fading in a wireless communications system is multipath. In a
multipath environment, the multipath reflectors introduce different delays and reflection
coefficients relative to the direct path, and so the electrical lengths of each path will be a
function of the frequency. Consequently, each frequency component in the signal
spectrum will be subject to slightly different amplitude and phase variations at the
receiver. From the model given in Appendix A, it is clear that this is very similar to
scintillation effects, where the received signal is a composite of multiple rays scattered
from different points on a phase screen, each of which are subject to different phase
advances.
The parameter that characterises the degree of frequency selectivity of a propagation
channel is the coherence or correlation bandwidth, Bcoh . The coherence bandwidth
determines the frequency separation for which the fading statistics of two frequency
components are essentially uncorrelated. Therefore, if the bandwidth of the transmitted
code sequence is much greater than Bcoh , significant frequency-selective fading may occur
leading to code pulse distortion. The coherence bandwidth is related to the RMS code
delay jitter, σ τ , by the following relationship (see for example [17], [55], [83] & [88])
143
Bcoh =
Consequently, if B code > B coh or σ τ >
1
2πσ τ
Hz
(4.4-1)
1
, where Bcode is the two-sided bandwidth of
2πBcode
the code sequence, code pulse distortion may occur. For the GPS C/A-Code and
P(Y)-Code, this becomes σ τ > 23m ( Bcode = 2.046 MHz) and σ τ > 2 ⋅ 3m ( Bcode = 20.46
MHz), respectively.
Unfortunately, there are few measurements of either the delay jitter or the coherence
bandwidth for transionospheric radio channels at GPS frequencies. However, in [79] Rino
derives an expression for the single-point, two-frequency coherence function, R (δf ; f ) , of
a transionospheric radio channel that is subject to scintillations. R (δf ; f ) is a measure of
the correlation between the time varying transfer functions of the propagation channel at
two frequencies, f 1 and f 2 8, and can be used to determine the coherence bandwidth, viz
R (B coh ; f ) = µ
(4.4-2)
where B coh is defined as the value of δf for which R (δf ; f ) drops to some pre-defined
value µ (see for example Steele, [88]). In [79], it is shown that for highly anisotropic (rodlike) irregularities, such as might be encountered at equatorial latitudes, R (δf ; f ) can be
expressed as a function g (⋅) of the form
δf
R (δf ; f ) = g H ×
f
( p −1) 2
(4.4-3)
where H is a measure of the strength of scintillation activity (proportional to S 42 under
weak scatter conditions, [79]), and p is the spectral index (the slope of the phase
scintillation power spectrum). Therefore, for a given level of scintillation activity, H, the
two-frequency coherence function and the channel coherence bandwidth will depend
only on the ratio δf f . Consequently, the following relationship will hold
B coh1 B coh 2
=
f1
f2
8
δf = f 1 − f 2 and f = ( f 1 + f 2 ) 2 .
144
(4.4-4)
where B coh1 is the coherence bandwidth at frequency f 1 for a channel with a scintillation
strength of H at f 1 , and B coh 2 is the coherence bandwidth at f 2 for a channel with the
same scintillation strength H at f 2 .
In [55], Knepp determines the channel coherence bandwidth at VHF frequencies (centred
on 155.5 MHz) by isolating the component of the code delay spread due to scintillations
and applying this to Equation (4.4-1). At these frequencies, it was found that the channel
coherence bandwidth could be as low as 0.5 MHz under very disturbed scintillation
conditions. Indeed, the measured S 4 values on a two-way propagation path were in
excess of 2.25, implying a one-way S 4 slightly greater than one. This indicates strong
Rayleigh fading with some focusing effects to drive S 4 above one. By applying these
results to Equation (4.4-4), it can be shown that under very intense scintillation
conditions, the channel coherence bandwidths at the two GPS carrier frequencies will be
B cohL1 = f L1 ×
B cohL 2
0.5
= 5.1 MHz
155.5
(4.4-5)
0.5
= f L2 ×
= 4.0 MHz
155.5
Therefore, under very strong scintillation conditions, the P(Y)-Code may suffer from
frequency-selective scintillation effects, and thus code pulse distortion, whereas the C/ACode is unlikely to be affected. In addition, the L2 channel is likely to be affected more by
frequency-selective scintillation effects than the L1 channel, partly because the coherence
bandwidth at L2 is less for a given level of scintillation activity (as shown above), and
partly because the scintillation activity is stronger at L2 for a given set of irregularity
conditions.
In order to test whether frequency-selective scintillation effects will cause significant code
pulse distortion, the phase screen model from Appendix A was used to generate
amplitude and phase scintillation waveforms at a number of frequencies across the code
spectrum (note that in all cases, the in-situ carrier phase perturbations were Gaussian
distributed with a power-law power spectral density). Examples of these waveforms at
the GPS L1 frequency and 10 MHz above L1 are given in Figure 4.4-1. It is clear that
although the general shape of the waveforms are quite similar, the detailed structures are
very different between these two frequencies. In particular, it appears that the differences
are at their greatest when the amplitude fading is at a maximum.
145
Signal phase (radians)
4
0
2
L1
L1
Attenuation of signal power (dB)
10
−10
−20
−30
0
−2
0.5
1
1.5
2
2.5
3
3.5
−4
0
4
0.5
1
1.5
0.5
1
1.5
2
2.5
3
3.5
4
2
2.5
3
3.5
4
4
L1 + 10 MHz
L1 + 10 MHz
10
0
−10
−20
−30
0
0
0.5
1
1.5
2
2.5
3
3.5
2
0
−2
−4
0
4
Position (km)
Position (km)
Figure 4.4-1: Amplitude and phase scintillation waveforms (left and right respectively) at the GPS
L1 frequency (upper two panels) and 10 MHz above GPS L1 (lower two panels). The scintillation
index S 4 is approximately equal to 0⋅8.
The amplitude and phase scintillation waveforms, A(x, f ) and φ p (x, f ) , were then
converted to a complex modulation for each frequency component, f, in the GPS code
spectrum, viz
(
)
M (x, f ) = A(x, f )exp jφ p (x, f )
(4.4-6)
where x represents a position in the scintillation pattern (the East-West ground position
in these examples). If P ( f ) is the Fourier transform of an ideal pulse with a pulse-width
equal to the GPS code chip-width, Tc , then the distorted code pulse, p ′(x, t ) , can be found
by taking the inverse Fourier transform of the product of M (x, f ) and P ( f ) as follows
p ′(x, t ) =
∞
∫ M (x, f )P( f )exp( j2πft ).df
(4.4-7)
−∞
Examples of the effects of frequency selective scintillations on a single code pulse are
given in Figure 4.4-2 to Figure 4.4-4. In these figures, it is assumed that the code pulse
spectrum is limited to the first nulls of the
p ′(x, t ) =
sinc( fTc )
power spectrum (ie.
1 Tc
∫ M (x, f )P( f )exp( j2πft ).df
, where 1 Tc = 1.023 MHz for the C/A-Code and
−1 Tc
10.23 MHz for the P(Y)-Code). Figure 4.4-2 shows the full effect of scintillations on a
P(Y)-Code pulse as a function of time and ground position, while Figures 4.4-3 and 4.4-4
represent the distorted waveforms that have been normalised by their respective peak
146
pulse values. Consequently, the last two figures provide a more accurate picture of the
true distortion to the code pulse shape.
Figure 4.4-2: The impact of a single phase-changing screen with a power law in-situ density
profile on 0.0978 µs pulses that are bandlimited to ±10.23 MHz (representative of a single P(Y)Code chip).
Figure 4.4-3: Figure 4.4-2 normalised by the peak pulse height.
147
Figure 4.4-4: The impact of a single phase-changing screen with a power law in-situ density
profile on 0.978 µs pulses that are bandlimited to ±1.023 MHz (representative of a single C/ACode chip).In this example, the reconstructed pulses are normalised by their peak values.
By applying these pulses to a simple Early-Late gate error function, the error in the code
loop discriminator can be found. The Early-Late gate error function is given by the
following expression
ε (x, τ ε ) =
∞
∞
1
p ′(x, t ) p (t + τ ε + Tc 2 ).dt − p ′(x, t )p (t + τ ε − Tc 2 ).dt
Tc
−∞
−∞
∫
∫
(4.4-8)
where τ ε is the code tracking error, p (t + τ ε + Tc 2 ) is a locally generated (ideal) early
pulse, and p (t + τ ε − Tc 2 ) is a locally generated late pulse. This is equivalent to a
normalised version of the Early-Late envelope discriminator from Table 4.1-1 (ie.
ε (x,τ ε ) = R E (x ) − R L (x ) ). In Figures 4.4-5 to 4.4-7, the Early-Late gate error function is
plotted for the distorted code pulses from the previous figures. Figures 4.4-6 and 4.4-7
represent the error function normalised by their respective peak values to isolate the
effects of code distortion from frequency-flat amplitude fading.
148
Figure 4.4-5: Early-Late gate error function for the bandlimited 0.0978 µs pulses.
Figure 4.4-6: Figure 4.4-5 normalised by the peak discriminator error.
149
Figure 4.4-7: Early-Late gate error function normalised by the peak discriminator error for the
0.978 µs pulses.
From the results presented in these figures, it is clear that the effects of
frequency-selective fading are much greater for the wider bandwidth P(Y)-Code than for
the C/A-Code, as would be expected. However, it is also apparent that even under
relatively strong scintillation conditions (ie. S 4 ≈ 0.8), the distortions to the P(Y)-Code are
only significant for relatively short periods of time. As will be shown next, these times are
usually associated with deep amplitude fading.
In Figures 4.4-8 and 4.4-9 (left panels), the code delay errors are plotted as a function of
the ground position, x, for τ ε = 0 . As τ ε = 0 represents a situation in which the replica
code is correctly aligned to the received code, the errors in these figures are associated
entirely with distortions to the code pulse shape. It is clear from these figures that the
discriminator errors associated with the P(Y)-Code (in chips) are much greater than those
associated with the C/A-Code. It is also clear from the scatter plots on the right that
significant tracking errors are usually associated with the time periods during which the
amplitude is deeply faded. Consequently, when the scattered rays come close to complete
cancellation on the ground (ie. during deep fading), the amplitude and phase response of
the channel attains its greatest sensitivity to frequency. However, as the carrier tracking
loops of a GPS receiver are likely to suffer from cycle slips and tracking problems during
150
these times, it is questionable whether these errors will be of practical importance in a real
GPS receiver.
P−Code tracking error. RMS = 0.0312 chips
P−Code tracking error v Fade depth
Error (chips)
0.2
0.15
0.1
0
0
1
2
3
4
5
6
7
8
9
10
Signal power (db)
10
|Error| (chips)
0.1
−0.1
0.05
0
−10
−20
−30
0
1
2
3
4
5
6
7
8
9
0
10
Ground position (km)
−5
0
5
10
15
20
Fade depth (dB)
Figure 4.4-8: The delay errors (upper left panel) and amplitude scintillation waveform (lower left
panel) as a function of ground position for the bandlimited 0.0978 µs pulses. The scatter plot on
the right compares the delay errors to the fade depth.
CA−Code tracking error. RMS = 0.00865 chips
CA−Code tracking error v Fade depth
0.05
0.045
0.05
0.04
0
−0.05
0
0.035
1
2
3
4
5
6
7
8
9
10
Signal power (db)
10
0
|Error| (chips)
Error (chips)
0.1
0.03
0.025
0.02
0.015
−10
0.01
−20
0.005
−30
0
1
2
3
4
5
6
7
8
9
10
Ground position (km)
0
−5
0
5
10
15
20
Fade depth (dB)
Figure 4.4-9: The delay errors (upper left panel) and amplitude scintillation waveform (lower left
panel) as a function of ground position for the bandlimited 0.978 µs pulses. The scatter plot on the
right compares the delay errors to the fade depth.
The procedures outlined above are very similar to those followed by Bogusch [16], [17]
and Knepp [54] using their more sophisticated multiple phase screen model. However,
their results suggest that the coherence bandwidth at L-band frequencies (2 GHz) could
be as low as 0.25 MHz under very strong scintillation conditions [17]. Using Equation
(4.4-4), this implies that BcohL1 ≈ 0.2 MHz , which is roughly 25 times smaller than the
151
value derived from the measurements in [55]. Although this result implies that code
distortion could be quite severe under strong scintillation conditions (ie. B cohL1 << B code ),
it was not clear in [17] whether significant distortion was only observed during periods of
deep amplitude fading.
4.5. Conclusions
In this chapter, the effects of scintillations on code tracking loops was examined. It was
found that phase scintillations have a negligible effect on code loops, and that the
additional RMS thermal noise error associated with amplitude scintillations is only small,
unless S 4 is close to unity. Nevertheless, under strong amplitude scintillation conditions,
it is likely that error spikes will exist in the code pseudorange measurements during times
when the amplitude is deeply faded. However, as the carrier loop is likely to be stressed
to the point of losing lock during these times, this effect may not be regarded as
important for most GPS users. It was also found that because of the very narrow
bandwidth of GPS signals, frequency selective scintillation effects are unlikely to produce
significant code distortion under naturally occurring ionospheric conditions, except
possibly for the P(Y)-Code during times when the amplitude is deeply faded.
152
Chapter 5
Codeless and semi-codeless receivers
In this Chapter, the effects of scintillations on codeless and semi-codeless receivers1 is
examined. Because codeless tracking loops have a much lower signal to noise ratio than
full code correlation tracking loops, their susceptibility to scintillations is expected to be
far greater. In Section 5.1, the various codeless tracking techniques employed in nonmilitary, dual frequency receivers are outlined. In Section 5.2, a theoretical analysis is
given of the effects of amplitude and phase scintillations on codeless tracking loops.
Essentially, the reduced signal to noise ratio of codeless tracking loops increases their
susceptibility to amplitude scintillations, while their much narrower loop bandwidths
increases their susceptibility to phase scintillations. Finally, in Section 5.3 the theoretical
performance measures of semi-codeless tracking loops are compared with measurements
taken from a semi-codeless receiver located in a region of high scintillation activity.
5.1. Codeless processing techniques
Codeless and Semi-codeless receivers obtain L2 code and carrier phase measurements
without requiring access to the military Y-Code. These receivers are frequently used in
WAAS and LAAS2 systems and are therefore important for civilian applications such as
air traffic control. The two most commonly used codeless techniques are squaring and
cross-correlation [91]. Squaring removes both the navigation data and the Y-Code from the
L2 signal and produces a carrier at twice the L2 frequency. However, this technique does
not provide code phase information, and with such a high carrier frequency the process
of resolving carrier cycle ambiguities can be difficult and time consuming. Crosscorrelation is an improvement on the squaring technique that produces a carrier at the
1
To improve the readability of this Chapter, the term “codeless” will frequently imply both
codeless and semi-codeless tracking loops.
2
WAAS: Wide Area Augmentation Systems, and LAAS: Local Area Augmentation Systems.
153
difference in frequency between L1 and L2 as well as a measure of the code delay
difference between the two frequencies. The much lower frequency allows carrier cycle
ambiguities to be resolved more rapidly, while the code delay difference provides an
unambiguous measure of the ionospheric delay on each carrier.
Both techniques can be enhanced by employing semi-codeless processing prior to the
codeless tracking loops. In a semi-codeless receiver, the P-Code (which is known a priori)
is removed from the GPS signal to produce an L2 carrier which is modulated by the
encryption code (also called the W-Code) and the navigation data. As the bandwidth of
the encryption code is 20 times less than the P-Code bandwidth, the resulting signal can
be filtered to reduce the noise power by up to 13dB [4]. Consequently, a semi-codeless
receiver will have a 13dB advantage in signal to noise ratio (SNR) over a purely codeless
receiver. Both techniques, however, suffer a considerable reduction in SNR over full code
correlation P(Y)-Code tracking.
Another semi-codeless technique that is frequently used in Novatel™ receivers involves
tracking the P-Code directly using a standard tracking loop, but with a pre-detection
bandwidth that is equal to the W-Code bandwidth (500 kHz) 3. Consequently, the WCode is treated in essentially the same way as the navigation data in a full code
correlation tracking loop. An advantage of this technique is that a true estimate of the L2
P-Code pseudorange is produced, although the associated thermal noise errors are at a
greatly elevated level.
In order to cope with the reduced SNR, codeless receivers employ very narrow tracking
loop bandwidths, typically much less than 1 Hz. To reduce the consequent dynamic
stresses associated with satellite and receiver motion, codeless carrier tracking loops are
aided by carrier phase error estimates provided by the more robust L1 C/A-Code carrier
tracking loops. These estimates tend to reduce, but not eliminate, the phase scintillation
errors on the codeless loops.
3
The pre-detection integration period, T, is therefore 2 × 10 −6 s.
154
5.2. Theoretical analysis
Based partly on the above discussion, the various aspects of codeless tracking loops
which determine their susceptibility to ionospheric scintillations can be summarised as
follows:
1. The L2 signal power is approximately 6dB below the L1 C/A-Code signal power.
2. The SNR of codeless tracking loops is significantly less than the SNR of full code
correlation tracking loops.
3. Both amplitude and phase scintillations are slightly stronger at the lower L2
frequency as a result of the inverse frequency scaling of scintillations.
4. L1 carrier aiding of the L2 codeless loop reduces the effects of phase scintillations.
5. Carrier aiding virtually eliminates dynamic stresses on the codeless loops allowing
the tracking loop bandwidth of the carrier loop to be significantly reduced.
The first two factors can be accounted for by determining an equivalent C N o for the
codeless tracking loops. This can then be used to determine a tracking threshold under
amplitude scintillation conditions based on the variance expression for a standard phase
locked loop.
The degradation in C N o for codeless and semi-codeless tracking loops is given by [91]
7.8 + 10 log10 ( B I ) − C N o
C
dB
A
10 x 10
→
6 BI
C No C
(5.2-1)
A
where B I is the pre-detection bandwidth (10MHz for a codeless loop, 500kHz for a semicodeless loop), and C N o C
A
is the carrier to noise density ratio of the C/A-Code. The
equivalent C N o is thus [91]
C No
Eq
(C N
=
)
2
o C A
6BI
(5.2-2)
As both squaring and cross-correlation eliminates the navigation data from the GPS
signal, a standard (non-Costas) phase locked loop can be used to track the resulting
carrier. For a standard phase locked loop, the variance of the tracking error is given by the
inverse of the loop SNR [73], viz
155
σ φ2ε = (Loop SNR )−1
=
=
Bn
C No
6 Bn B I
(C N
(5.2-3)
Eq
)
2
o C A
where Bn is the tracking loop bandwidth. If a variance threshold is chosen for σ φε ,
threshold values of the signal amplitude can be found for both codeless and semicodeless receivers as a function of Bn , B I , and the nominal C N o
~
ATh = A 4
(C N
6 Bn B I
o C A ∗σ φT Th
where A is the nominal signal amplitude and σ φT
C A
, viz
)
Th
2
(5.2-4)
is the threshold RMS error once the
effects of phase scintillations have been removed (see Section 3.4). As before, these
threshold values can be used to determine the probability of losing lock as a function of
S 4 (see Equation (3.4-3)).
The third factor (ie. the difference in the strength of scintillations at the L2 frequency) is
automatically accounted for by using the L2 frequency as an input to the WBMOD
scintillation model. In Section 2.1.3, it was stated that S 4 and σ φ p scale in the following
way with the carrier frequency, ν
S 4 ∝ ν −( p +3) 4 , for low to moderate levels of scintillation activity
≈1
σ φp ∝ν
, for strong scintillations
(5.2-5)
−1
From Equations (2.1-1) and (2.1-4) it is also apparent that for a constant outer scale size
parameter, f o , T also scales with the carrier frequency as T ∝ ν −2 (ie. σ φ2p ∝ T ).
Consequently, at the L2 frequency, S 4 is between 1 and 1.4 times larger than the
corresponding L1 value, while T is approximately 1.65 times larger.
The fourth factor (ie. carrier aiding of the codeless loops) is accounted for by assuming
that the majority of the phase scintillation energy is associated with refraction effects
156
which show a simple ν −1 dependence with frequency. Consequently, scintillation
induced phase variations on the L2 carrier, φ p L 2 , can be related to the corresponding
variations on the L1 carrier, φ pL1 , through 4
φ pL 2 =
ν L1
φp
ν L 2 L1
(5.2-6)
where ν L1 = 1575.42 MHz and ν L 2 = 1227.6 MHz . Aiding of the L2 carrier loop is
primarily intended to eliminate Doppler errors associated with satellite and receiver
motion. The Doppler correction term is given by
ν L2
φd
ν L1 L1
(5.2-7)
where φ d L1 is the phase error associated with Doppler on the L1 . The ν L2 ν L1 factor in
this expression is also applied to the phase scintillation errors on L1, φ pL1 , before they are
removed from the L2 carrier loop. The residual phase scintillation error is thus [91]
φ p′ L 2 = φ pL 2 −
ν L2
φp
ν L1 L1
ν 2
= φ pL 2 1 − L2
ν L1
= 0.393 φ pL 2
(5.2-8)
Consequently, the phase scintillation spectral strength, T, obtained from models such as
WBMOD at the L2 frequency must be scaled by a factor of 0.3932 (-8.1dB) to account for
the effects of carrier aiding.
The last factor (ie. a very narrow loop bandwidth) provides a considerable amount of
resistance to the effects of amplitude scintillations. By implementing a narrow loop
bandwidth, the codeless receiver reduces thermal noise errors on the phase estimates
which in turn increases the SNR of the tracking loop. This helps to overcome some of the
4
It is worth noting that if diffraction effects were to predominate, this simple relationship would
not necessarily apply. However, as the bandwidths of the codeless loops are very narrow, it is
reasonable to assume that a large proportion of the energy in the carrier loop phase tracking errors
is below the Fresnel cutoff frequency and so can be attributed mainly to the effects of refraction.
157
effects associated with the lower signal level at the L2 frequency and the reduced SNR of
the codeless tracking loops. A second effect of a narrow loop bandwidth is that the
duration of the deep fading events which lead to loss-of-lock may now be much less than
the time constant of the tracking loop. For loop bandwidths as low as 0.1Hz, this could
greatly improve a loop’s tolerance to amplitude scintillations, particularly if the
amplitude scintillation rate is increased by receiver motion. However, this effect will be
offset somewhat by the greatly reduced amplitude threshold of codeless tracking loops.
Figure 5.2-1 is an illustration of the impact of a reduced amplitude threshold on the fade
duration using a short segment of simulated amplitude scintillation data obtained from
the model in Appendix A. For deep fading events, the signal amplitude will be below the
semi-codeless threshold for much longer periods of time, effectively resulting in a longer
duration fade.
Semi-codeless
Coded Loop
Figure 5.2-1: Illustration of the effect of a reduced amplitude threshold on the fade duration.
5.3. Threshold curves
In Figure 5.3-1, threshold curves are plotted for both codeless and semi-codeless receivers
(both techniques) based on the following assumptions:
•
The loop bandwidths are 0.1Hz.
•
C N o for the C/A-Code tracking loop is 44dBHz (at 38dBHz, the codeless loop is
already very close to losing lock for a loop bandwidth of 0.1Hz). Therefore, C N o
Eq
is
10.2dBHz for the codeless loop, and 23.2dBHz for the semi-codeless loop.
•
Carrier aiding of the codeless loops reduces T by 8.1dB.
•
The amplitude scintillation bandwidth is narrow compared to the loop bandwidth.
This assumption is necessary in the absence of a suitable approach to the problem of
very narrow loop bandwidths and may result in an overestimation of the
susceptibility of tracking loops to amplitude scintillations.
The threshold curve for the semi-codeless-2 technique (see caption of Figure 5.3-1) is
simply found using the approach outlined in Section 3.4 with T = 2 × 10 −6 s and the
158
Spectral strength, T, reduced by 8.1dB. Also shown as a dotted line is the curve obtained
from Equation (3.4-8) with γ dB = 20dB (ie. the approximate upper limit on [T, S 4 ] values
for a stationary receiver obtained from WBMOD).
−15
Spectral Strength, T (dB)
−20
−25
Semi−codeless−2
−30
−35
Codeless
Semi−codeless−1
−40
−45
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
S4
Figure 5.3-1: Tracking thresholds for both codeless and semi-codeless receivers as a function of the
Spectral Strength, T, and the amplitude scintillation index, S 4 . A loop bandwidth of 0.1Hz has
been assumed. Semi-codeless-1 refers to the process of removing the P-Code from the L2 carrier
prior to codeless processing. Semi-codeless-2 refers to the technique of adopting a much wider predetection bandwidth in a Costas loop to accommodate the W-Code.
It is clear from this figure that because the various threshold curves lie well below the
dotted line obtained from WBMOD, the probability of losing lock is likely to be
reasonably high on links that are affected by scintillations. By comparing this with Figure
3.3-3, it is also clear that codeless and semi-codeless tracking loops are considerably more
susceptible to scintillations than full code correlation tracking loops. The very narrow
loop bandwidth increases their susceptibility to phase scintillations, despite a reduction
in phase scintillation energy through aiding from the C/A-Code carrier loop. Similarly,
the reduced SNR greatly increases their susceptibility to amplitude scintillations, despite
a very narrow loop bandwidth.
5.4. Scintillation measurements
The objective of this section is to check the validity of the semi-codeless performance
models derived in the previous sections using data obtained from co-located Novatel
159
Millennium and PAQ12 Ionospheric Scintillation Monitoring (ISM) receivers. The
Millennium uses full code correlation for the L1 C/A-Code and a semi-codeless
tracking technique for the L2 Y-Code. The ISM receivers are high data rate L1 SPS
receivers which are designed specifically to measure the scintillation indices S 4 and T
[92]. Although data from the peak of the current solar cycle has not yet been processed,
ISM data from the September 1998 and March 1999 equinoxes shows sufficient evidence
of scintillation activity for an analysis to be performed.
5.4.1. Overview of scintillation data
The data used in this study was obtained from co-located Millennium and ISM
receivers deployed at Parepare in Indonesia. Parepare is located at 4ºS, 119.6ºE, which
places it beneath the crest of the southern anomaly in a region of potentially strong
scintillation activity. The 12 month smoothed mean monthly sunspot number (SSN)5 for
the March 1999 equinox was approximately 83.8, which represents a moderate level of
solar activity. At the time of writing (November 2000), the measured SSN’s for the
September 1999, March 2000 and September 2000 equinoxes were 102.3, 119.8 and 128.7,
respectively. In addition, it is predicted that the peak of the current solar cycle will occur
in December 2000, and that it will take until late 2003 before the solar activity drops
below the level measured in September 1998 (69.5). Consequently, over the coming few
years it is expected that scintillation activity and it impact on GPS will continue, although
mainly during the equinoctial months.
5.4.1.1. Novatel Millennium data
The Novatel Millennium provides an indication of the tracking state of the L1 and L2
code and carrier tracking loops as part of its standard output6. A second set of lock
indicators obtained by comparing the reported code and carrier pseudorange
measurements with the range calculated from the satellite ephemeris information was
found to give virtually the same results. Consequently, for convenience it was decided
that the tracking state indicator would be used in all subsequent analysis as an indicator
of a loss of valid pseudorange data. In addition, a 10º elevation angle mask was chosen to
5
Sunspot numbers were obtained from the Australian Ionospheric Prediction Service (IPS) web site
at “http://www.ips.gov.au”.
6
Tracking state information is contained within the Novatel format RGEA/B data blocks [68].
160
avoid multipath effects and the possibility of satellite obscuration on low elevation angle
links.
L1 Code
L1 Code
2
1
0.5
0
1
0
L1 Carrier
L1 Carrier
2
1
0.5
0
1
0
L2 Code
L2 Code
20
10
0
20
10
0
L2 Carrier
L2 Carrier
20
10
0
0
5
10
15
20
Day into September 1998
25
30
20
10
0
0
5
10
15
20
25
30
Day into March 1999
Figure 5.4-1: The percentage of time from 8:00pm to 10:00pm (local time) that the Novatel
Millennium loses lock on one or more satellites at Parepare, Indonesia (left panel: September
1998 equinox, right panel: March 1999 equinox). The white bars represent the loss of one satellite.
The solid bars represent the simultaneous loss of two or more satellites.
In Figure 5.4-1, the percentage of time between 8:00pm and 10:00pm (local time) that the
Millennium loses lock on one or more satellites is plotted as a function of the day for
both the September 1998 and March 1999 equinoxes. It is clear from this figure that the
carrier loops lose lock slightly more often than the code loops (by a factor of about 1.5).
This supports the assertion that in the absence of external Doppler aiding, the code loop
loses lock soon after the carrier loop. It is also apparent that the semi-codeless L2 loops
are considerably more susceptible to the effects of scintillations than the L1 loops. Indeed,
on average the L2 carrier loop loses lock on one or more satellites about 30 times more
often than the L1 carrier loop, although this factor varied significantly from day to day. It
is also apparent that for the moderate levels of solar activity represented by these two
months, the impact on navigation is relatively minor. On the worst day (16th March 99),
the L1 code loop lost lock on 1 satellite for approximately 1.7% of the two hour period of
interest (roughly 2 minutes in total). For a mask angle of 10º, the minimum number of
available satellites during this period was 8. Consequently, even during these two
161
minutes, the loss of one satellite would not prevent navigation or even seriously degrade
the satellite geometry. However, it is clear that the effect of scintillations on the semicodeless L2 channels was significantly greater. On the 16th March 1999 (a bad day), the L2
carrier loop lost lock on one or more satellites for approximately 27% of the two hour
period of interest (a total time of 32 minutes). In addition, two and three satellites were
lost simultaneously for about 3% and 0.03% of the time, respectively.
5.4.2. A comparison of models with measurements
The PAQ12 ISM receiver provides measurements of the four scintillation parameters,
S 4 , σ φ p , T and p once every minute. By feeding these parameters into the single link
performance models given in Section 5.2, it is possible to predict the probability of an
outage on the L2 semi-codeless tracking loops. Unfortunately, because the Millennium
and ISM receivers are co-located and have approximately the same L1 carrier loop
bandwidths, the L1 tracking loops of the Millennium would be expected to lose lock at
about the same time as the ISM tracking loops (ie. a deep fade will affect both receivers
simultaneously). Consequently, it is not possible to check the validity of the tracking
thresholds for a full code correlation tracking loop using this approach. Indeed, during
the September 1998 equinox, it was found that on average ISM data was missing from
97% of L1 code and 95% of L1 carrier epochs for which the Millennium had lost lock.
However, as the L2 loops lose lock at much lower levels of scintillation activity, only 54%
of L2 code and 52% of L2 carrier loss-of-lock epochs had missing ISM data.
Unfortunately, because the S 4 measurements vary so much from epoch to the next, it is
not possible to interpolate between valid measurements to eliminate this problem.
In Figure 5.4-2, the percentage of time that the Millennium loses lock is plotted as a
function of S 4 using data from both the September and March equinoxes (from 6:00pm to
local midnight). Data for which the Spectral Strength, T, is greater than -30dBradians2 has
been ignored to ensure that amplitude scintillations have the predominant effect.
Included in the lower panel of this figure is the probability of losing lock on the carrier
loop based on the theoretical analysis presented in Section 5.2 and using the ISM
measurements of S 4 (the upper and lower error bars represent ±1σ values of the
measured C N o from the Millennium). If the tracking loops re-acquire the signal
shortly after the amplitude rises above the tracking threshold, then the measured and
theoretical probability of an outage should be comparable. From Figure 5.4-2, this appears
162
to be true until S 4 becomes large, at which point the measurement curve begins to flatten
off. This flattening is believed to be the result of deep fades which result in the
simultaneous loss of both Millennium L1 and ISM data. Although, these results indicate
that the expression for the probability of losing lock, PL , is relatively accurate, they also
suggest that the 1% threshold discussed in Section 3.4 may be relatively conservative.
6
15
5
L2 Carrier %
L2 Code %
4
3
2
10
5
1
0
0
0.2
0.4
0.6
0.8
1
S4
0
0
0.2
0.4
0.6
0.8
1
S4
Figure 5.4-2: The percentage of time that the L2 code and carrier loops lose lock as a function of
S 4 . Data for which the phase scintillation Spectral Strength, T, is greater than -30dBradians2 has
been ignored. The error bars represent theory based on quiescent L2 C N o values between
34dBHz (upper bars) and 44dBHz (lower bars).
An equivalent analysis of the relationship between loss-of-lock and the spectral strength
parameters, T and σ φ p , has not been given because of a lack of confidence in the integrity
of the ISM measurements of T and σ φ p .
5.5. Conclusions
Codeless tracking loops are far more susceptible to the effects of scintillations than full
code correlation tracking loops. However, the different tracking techniques employed in
codeless receivers do have significantly different susceptibilities, with semi-codeless
techniques being generally more robust than purely codeless techniques. The poor
performance of codeless tracking loops under scintillation conditions may result in a
degradation in the accuracy of systems such as the Wide Area Augmentation System
(WAAS) which rely on dual frequency SPS receivers for the measurement of ionospheric
delays. However, by the time WAAS and equivalent systems become operational, it is
163
expected that solar activity will have declined to the point where scintillations are no
longer regarded as a significant threat. Also, with the imminent introduction of a second
C/A-Code signal on the L2 frequency, and with approval being given for a future L5 civil
signal, the effects of scintillations on codeless receivers is unlikely to be an issue for future
solar maxima.
Measurements of the strength of scintillation activity and loss-of-lock taken during times
of strong scintillation activity are generally in quite good agreement with the theory.
They also show that because scintillations are very patchy, the chances of losing lock on
several channels simultaneously are very small, even under strong scintillation
conditions. This issue is discussed further in Chapter 9.
164
Chapter 6
Navigation data
In this chapter, the effects of scintillations on the process of demodulating the navigation
data is examined. In Section 6.1, background theory is given which enables the bit error
rates and word error rates to be calculated under quiescent signal conditions. In Sections
6.2 and 6.3, scintillation effects on the navigation data are determined by treating
amplitude and phase scintillation effects separately. It is shown that even under very
intense scintillation conditions, word error rates increase to only a few percent with
amplitude scintillations providing the greatest contribution. The results of these two
sections are then used to determine the combined effects of amplitude and phase
scintillations on the navigation data. Finally, in Section 6.5 the effects of a slowly varying
amplitude waveform on the word error rate is examined.
6.1. Background
The GPS navigation message is broadcast by each satellite and contains information about
the satellite ephemerides, clock and ionospheric correction factors, timing information
and constellation status. The navigation message consists of 25 data frames, each
containing 5 subframes and each subframe containing 10 words of 30 bits. Therefore, at a
data rate of 50 bits/s, the complete navigation message takes 12.5 minutes to be
downloaded by a receiver. The first three subframes of each frame contain the same clock
and ephemeris information which is considered critical to the operation of a receiver.
Consequently, this information is made available to a receiver at a rate of once every 30
seconds.
In a GPS receiver, the navigation data is extracted from the in-phase channel of the carrier
tracking loop at a point immediately after the pre-detection filters (see Figure 3.1-1). In
the analysis that follows, it is assumed that the carrier tracking loop remains locked to the
165
GPS carrier and synchronised to the navigation data. Under these assumptions, the inphase signal is given by (from Equation (3.1-2))
~
I P = A d (t − τ ) cos(φε ) + n IP
(6.1-1)
~
where A is the signal amplitude after the pre-detection filters, d (t − τ ) is the navigation
data, φε is the carrier phase error and n IP is distributed N (0,σ n ) where σ n = N o T .
Two new random variables, X 0 and X1 , can be introduced to represent I P during the
transmission of a binary 0 data bit ( d (t − τ ) = −1 : Hypothesis H 0 ) and a binary 1 data bit
( d (t − τ ) = 1 : Hypothesis H 1 ), viz
~
I P = X 0 = − A cos(φε ) + n IP
~
I P = X 1 = A cos(φε ) + n IP
H0 :
H1 :
(6.1-2)
~
Under quiescent conditions (ie. no scintillations), A is approximately constant and φε is
small compared to 1 radian. Consequently, X 0 and X 1 can be approximated by
X 0 = −A + n IP
(6.1-3)
X 1 = A + n IP
~
where A = A is a constant. The probability density functions of X 0 and X1 are therefore
f X 0 (x0 ) = N (− A,σ n )
f X1 (x1 ) = N (A,σ n )
(6.1-4)
and the probability of a bit error is given by (see for example Haykin [39])
∞
0
0
−∞
Pe = p (H 0 )∫ f X 0 (x 0 ).dx0 + p (H 1 ) ∫ f X1 (x1 ).dx1
∞
=
∫ f X 0 (x0 ).dx0
(6.1-5)
0
=
A
1
Erfc
2σ
2
n
where Erfc( ) is the Complementary Error Function1, and p (H 0 ) and p( H1 ) are the
1
Ercf (x ) =
166
2
π
∞
2
∫ exp(− y ).dy
x
probabilities of binary 0 and 1 data bits respectively2. This can also be expressed in terms
of the carrier to noise density ratio, C N o = A 2 (2 N o ) , and the energy per bit,
Eb = TA 2 2 , as follows
(
1
Erfc T . C
2
Eb
1
= Erfc
N
2
o
Pe =
No
)
(6.1-6)
Under normal tracking conditions, T = 20 ms and C N o ≈ 41.5 dBHz resulting in a bit
error probability of 3 ⋅ 4 × 10 −125 (this assumes a nominal satellite signal power level of
-160 dBW at the ground and a noise temperature of 530 K).
The probability of a word error, assuming no error correction, is given by (Hegarty [40])
Pw = 1 − (1 − Pe )m
≈ mPe
(6.1-7)
for Pe << 1
where m is the number of bits per word (30 for the GPS navigation message). This results
in a word error rate of approximately 10 −123 under the signal conditions outlined above.
Similarly, the probability of a word error in any of the first three subframes (ie. the critical
navigation data) is given by 1 − (1 − Pw )30 ≈ 3.1 × 10 −122 . These results demonstrate that
under quiescent signal conditions, the bit error rates and word error rates are negligible
for GPS.
In the following sections, the effects of scintillations on the navigation data will be
examined by treating the GPS receiver as a BPSK communications system that is subject
to non-dispersive fading (ie. frequency and time-flat fading). The principal difference in
this analysis over most other analyses is that the signal intensity is assumed to follow the
Nakagami-m distribution, and the resulting bit error rates are linked to the scintillation
parameters discussed earlier.
2
For binary data, p (H 0 ) + p (H 1 ) = 1 .
167
6.2. The impact of phase scintillations on navigation
data
In this section, the effects of phase scintillations on the bit error probability will be
examined under the assumption that amplitude scintillations are absent (ie. it will be
assumed that the amplitude is constant).
When the satellite signal is modulated by phase scintillations, the phase error, φε ,
becomes a random variable. Therefore, the random variables X 0 and X1 given in
Equation (6.1-2) become functions of the two random variables n IP and φε as follows
X 0 = −A cos(φε ) + n IP
X 1 = A cos(φε ) + n IP
(6.2-1)
where the amplitude A is assumed to be constant (ie. for the moment, amplitude
scintillations have been ignored). If it is assumed that the linearising approximations
made in the analysis of the tracking loops are not significantly violated, the variance of φε
can be obtained from the linear model transfer function of the tracking loop as follows
(from Equation (3.2-5))
{ }
E φε 2 = σ φ2ε =
∞
f 2k
∫ (f 2k + f n 2k )⋅ (f
−∞
T
2
o
+ f2
)
p
.df +
2
Bn
C No
1
1 + 2T C N
o
(6.2-2)
= σ φ2εp + σ φ2T
where σ φ2εp and σ φ2T
are contributions to the phase error variance from phase
scintillations and thermal noise respectively. This assumes that the pre-detection filters
have a negligible effect on the phase errors produced by phase scintillations. As the
integration period of the pre-detection filters coincides with the duration of a navigation
data bit, this also implies that the phase errors will remain approximately constant over
each data bit. This is an important assumption on which the following analysis is based.
As X 0 is now a function of the phase error, φε , the new PDF of X 0 is given by
f X 0 (x 0 ) =
∞
∫ f X 0 φε (x0 ϕ ). fφε (ϕ ).dϕ
−∞
168
(6.2-3)
where f X 0 φ (x0 ϕ ) is the conditional PDF of X 0 given φε , and fφε (ϕ ) is the PDF of
ε
φε . The probability of a bit error is thus (from Equation (6.1-5))
∞
Pe =
∫ f X 0 (x0 ).dx0
0
∞
=
∞
∫ ∫ f X 0 φε (x0 ϕ ). fφε (ϕ ).dϕ .dx0
0 −∞
∞
(6.2-4)
∞
= fφε (ϕ ) f X 0 φε (x0 ϕ ).dx0 .dϕ
0
−∞
∫
∫
∞
=
∫ fφε (ϕ )Pe (ρ b ,ϕ ).dϕ
−∞
where ρb = T . C N o , and Pe (ρb , ϕ ) is the conditional bit error probability given ρb and ϕ .
As f X 0 φ (x0 ϕ ) is distributed N (− A cos(ϕ ),σ n ) , the conditional bit error probability is
ε
given by
Pe (ρb , ϕ ) =
=
Acos(ϕ )
1
Erfc
2
2σ n
(
1
Erfc cos(ϕ ) T . C N o
2
(6.2-5)
)
If it is assumed that the phase errors follow the Tikhonov density function for a Costas
loop (Equation (C-6)), the probability of a bit error is given by
π 2
Pe =
∫
−π 2
exp(ρ e cos(2ϕ )) 1
. Erfc cos(ϕ ) T . C N o .dϕ
2
πI o (ρ e )
(
)
(6.2-6)
where ρe = 1 4σ φ2ε and σφ2ε is the phase error variance from the linear model (Equation
(6.2-2)). Although this result strictly only applies to a system that is based on a first order
Costas loop, it is also likely to be quite accurate for higher loop orders, particularly when
the spectral index, p, is close to 2. The reasons for this were discussed earlier in Chapter 3.
In Figure 6.2-1, the bit error probability and word error probability are plotted as a
function of σ φ2εp for a first order Costas loop with T = 20ms, C N o = 41.5dBHz and
Bn =2Hz. The maximum phase error variance is set to (π 12) radians 2 which is the rule
2
of thumb tracking threshold for linear operation (Equation (C-3), Appendix C). Notice
169
that σ φ2εp can be related to the scintillation parameters T, p and f o for a specific loop
order and bandwidth through Equation (6.2-2).
−40
−30
Word error probability (dB)
Bit error probability (dB)
−50
−60
−70
−80
−40
−50
−60
−70
−90
0.02
0.03
0.04
0.05
0.06
0.07
2
Phase error variance (radians )
−80
0.02
0.03
0.04
0.05
0.06
0.07
2
Phase error variance (radians )
Figure 6.2-1: Probability of a bit error (left panel) and a word error (right panel) as a function of
the phase error variance, σ φ2εp , for a first order loop with T=20ms, C N o = 41.5dBHz and
Bn =2Hz.
The results presented in Figure 6.2-1 suggest that phase scintillations have a relatively
minor effect on the process of demodulating the navigation data. Even when the phase
error variance is near the carrier loop tracking threshold, the probability of a word error is
only 0.1% (-30dB) and the probability of a word error in any of the first three subframes is
approximately 3% ( 30 Pw ). As will be shown in the next section, these are much smaller
than the corresponding probabilities under amplitude scintillation conditions with S 4 = 1 .
6.3. The impact of amplitude scintillations on
navigation data
In this section, the effects of amplitude scintillations on the bite error probability will be
examined under the assumption that phase scintillations are absent (ie. φεp = φ p = 0 ).
Under quiescent signal conditions, φε
is extremely small and the simplifying
approximation cos(φε ) ≈ 1 can be made. However, when amplitude scintillations are
present, occasional deep fading can result in a significant increase in the thermal noise
contributions to φε , particularly when the tracking loop bandwidth is wide.
170
~
Consequently, both the amplitude, A , and the phase error, φε , will become random
variables. X 0 and X1 can therefore be represented by
~
X 0 = − A cos(φε ) + n IP
~
X1 = A cos(φε ) + n IP
(6.3-1)
~
where A is Nakagami-m distributed (Equation (2.1-8)), n IP is distributed N (0,σ n ) and
φε is assumed to follow the Tikhonov PDF. In this analysis, it is assumed that the effects
~
of the pre-detection filters on the amplitude are negligible (ie. A ≈ A , where A is the
signal amplitude prior to filtering). For amplitude scintillations, the Tikhonov PDF is
considered to be a reasonable choice as the phase errors are driven entirely by white,
Gaussian thermal noise. The effective SNR for the Tikhonov PDF is given by ρe = 1 4σ φ2ε ,
~
where σφ2ε is a function of A and is given by (from Equation (3.3-9))
( ) ∫ H ′( f , A~ )2 Snd~ 4( f ).df
A
()
~
~
σ φ2ε A = σ φ2T A =
∞
(6.3-2)
−∞
For a first order loop, this reduces to (from Equation (3.3-13))
()
()
Bn
~
~
σ φ2ε A = σ φ2T A =
C No
1
1
+
~
2
g N 2T C N o AN g N
(6.3-3)
~
~
where AN = A A is the normalised signal amplitude, and g N = g A 2 is the normalised
AGC gain factor. These equations assume that the bandwidth of the amplitude
scintillations is small compared with the nominal loop noise bandwidth Bn . The PDF of
X 0 is given by
f X 0 (x 0 ) =
π 2
∞
∫ ∫ f X 0 φε , A~ (x0 ϕ , A ). fφε A~ (ϕ A ). f A~ (A )dϕ .dA
(6.3-4)
0 −π 2
where
fX
~
0 φε , A
(x0 ϕ , A ) = N (− A cos(ϕ ),σ n )
(6.3-5)
~
is the conditional PDF of X 0 given φε and A . Also
fφ
~
ε A
(ϕ A ) = exp(ρ e cos(2ϕ )) ,
πI 0 (ρ e )
ϕ ≤
π
2
(6.3-6)
171
~
is the conditional PDF of φε given A with ρe =
f A~ (A ) =
1
~ , and
4σ φ2ε A
()
mA 2
2m m A 2m−1
− ~ ,
exp
~
〈 A2 〉
Γ(m ).〈 A 2 〉 m
A≥0
(6.3-7)
is the Nakagami-m PDF for amplitude. The probability of a bit error is then (from
Equation (6.1-5))
∞
Pe =
∫ f X 0 (x0 ).dx0
0
∞ ∞
=
=
π 2
∫ ∫ ∫ f X 0 φε , A~ (x0 ϕ , A ). fφε A~ (ϕ A ). f A~ (A )dϕ .dA.dx0
0
0 −π 2
∞
π 2
∫ ∫
0 −π 2
fφ
ε
(6.3-8)
∞
~ (ϕ A ). f ~ (A ) f
~ (x 0 ϕ , A ).dx 0 dϕ .dA
A
A
X φ ,A
0 0 ε
∫
By substituting the appropriate PDF expressions from Equations (6.3-5) to (6.3-7) into
Equation (6.3-8), the probability of a bit error becomes
π 2
∞
Pe =
∫ ∫
0 −π
A cos(ϕ )
mA 2 1
exp(ρ e cos(2ϕ )) 2m m A 2m−1
.d .dA
− ~ . Erfc
.
exp
~
2
2
m
〈A 〉 2
2σ ϕ
I 0 (ρ e )
π
(
).
m
A
Γ
〈
〉
n
2
(6.3-9)
Although a closed form solution to this integral is difficult to obtain, a numerical solution
can be produced for a given nominal (undisturbed) carrier to noise density ratio and
noise bandwidth. To obtain a numerical solution, it is convenient to replace the dummy
variable A with a variable ρo which represents the instantaneous carrier to noise density
ratio of the satellite signal. By making the following substitutions
A = 2N o ρo
σ n2 = No T
~
〈 A2 〉 = 2 C N o ∗ N o
(6.3-10)
where C N o is the nominal carrier to noise density ratio, the following expression can be
obtained for the probability of a bit error
∞
Pe =
π 2
∫ ∫
0 −π
172
mρ o
exp(ρ e cos(2ϕ )) m m ρ o m −1
.
exp −
m
πI 0 (ρ e )
Γ(m ). C N o
C No
2
(
)
1
. Erfc Tρ o cos(ϕ ) .dϕ .dρ o
2
(6.3-11)
In Figure 6.3-1, Pe is plotted as a function of the intensity scintillation index S 4 ( = 1 m2 )
for a first order loop with T=20ms, C N o = 41.5 dBHz and Bn =2Hz (upper of the two
~
curves). In this case, the AGC is assumed to be ideal (ie. g N = AN 2 ), and the effective loop
SNR is given by (from Equation (6.3-3))
ρe =
=
=
1
4σ φ2ε
(A~ )
C No
(6.3-12)
1
1
4 Bn ~ 2 +
~ 4
2T C N o AN
AN
Tρ o2
2 Bn [2Tρ o + 1]
If it is assumed that the phase errors are negligible (ie. cos(φε ) ≈ 1 ), the PDF of X 0 reduces
to
f X 0 (x 0 ) =
∞
∫ f X 0 A~ (x0 A ). f A~ (A ).dA
(6.3-13)
0
The probability of a bit error is then
∞
Pe =
∫ f X 0 (x0 ).dx0
0
∞ ∞
=
∫ ∫ f X 0 A~ (x0 A ). f A~ (A ).dA.dx0 .
0 0
∞
=
m m ρ o m −1
∫ Γ(m). C N o m exp − C Noo . 2 Erfc(
mρ 1
(6.3-14)
)
Tρ o .dρ o
0
Wojnar [102], demonstrated that this integral could be expressed in terms of an
incomplete Beta function ratio as follows
1 β m (m+α ) (m,1 2 )
2
β (m,1 2 )
1
= I m (m+α ) (m,1 2 )
2
Pe =
(6.3-15)
where α = T . C N o , β (a , b) is the Beta function, βn ( a , b) is the Incomplete Beta function,
and I n (a , b) = βn (a , b) β (a , b) is the Incomplete Beta function ratio (see for example
Gradshteyn [37]). In Figure 6.3-1, the lower curves represent Pe obtained from Equation
173
(6.3-15) under the assumption that the phase errors are negligible. These curves show that
for a narrow bandwidth receiver (in this case 2Hz), the direct effect of amplitude
scintillations on the probability of a bit error is far more significant than the effect of an
increase in the level of thermal noise in the feedback path. For wider noise bandwidth’s,
the contributions to Pe from phase errors in the feedback path becomes more significant
but are still relatively small (see Figure 6.3-2).
−10
−30
Word error probability (dB)
Bit error probability (dB)
−20
−40
−50
−60
−70
−80
−30
−40
−50
−60
−70
−90
−80
−100
0.4
0.5
0.6
0.7
0.8
0.9
1
0.4
0.5
0.6
S4
0.7
0.8
0.9
1
S4
Figure 6.3-1: Probability of a bit error (left panel) and a word error (right panel) as a function of
S 4 for T = 20ms, C N o = 41.5dBHz and Bn = 2Hz. The upper curves represent a situation in
which the phase errors have been included (Equation (6.3-11)). The lower curves represent a
situation in which they have been ignored (Equation (6.3-15)).
Bit error probability (dB)
−30
−40
−50
−60
−70
−80
−90
−100
0.4
0.5
0.6
0.7
0.8
0.9
1
S4
Figure 6.3-2: Probability of a bit error as a function of S 4 for T=20ms, C N o = 41.5dBHz and
Bn = 15Hz. The upper curve represent a situation in which the phase errors have been included.
The lower curve represent a situation in which they have been ignored.
174
So far, only the effects of an ideal AGC have been considered. In order to account for the
effects of a non-ideal AGC (either fast or very slow), the linear model variance expression
(Equation (6.3-3)) must be modified to include a non-ideal AGC gain factor, g N .
However, as shown in Section 3.3, when the loop is only subject to amplitude
scintillations and thermal noise, a non-ideal AGC will result in smaller phase tracking
errors. Therefore, the phase errors will have even less of an effect on the bit error rates
than has already been discussed. However, this is not necessarily true when the loop is
also subject to phase scintillations as its ability to track phase variations will be impaired.
6.4. The combined effect of scintillations on
navigation data
When, as is normally the case, amplitude and phase scintillations are present together,
their impact on the navigation data can be found using Equation (6.3-11), but with σφ2ε
based on Equation (3.3-8) rather than Equation (6.3-2). Again, this assumes that the
Tikhonov PDF is a valid choice for the phase error density function (ie. it assumes that
p ≈ 2 ). From Equation (6.3-11), the bit error probability is given by
∞
π 2
∫ ∫
Pe =
0 −π
where
mρ o
exp(ρ e cos(2ϕ )) m m ρ o m −1
.
exp −
m
πI 0 (ρ e )
Γ(m ). C N o
C No
2
ρe =
1
4σ φ2ε
σ φ2ε
and
()
~
A
(
)
1
. Erfc Tρ o cos(ϕ ) .dϕ .dρ o
2
( )
∞
( ) ∫ 1 − H ′( f , A~ )2 Sφ p ( f ) + H ′( f , A~ )2 Snd~ 4 f
A
~
A =
−∞
.df
(6.4-1)
from
Equation (3.3-8). For a first order I.Q Costas loop, this becomes
∞
() ∫
~
σ φ2ε A =
−∞
f2
~
f 2 + f n AN 2 g N
(
T
.
2
) (f
2
o
+ f2
)
p
.df +
2
Bn
C No
1
1
+
~ 2
g
N 2T C N o AN g N
For an ideal AGC (or a fast AGC with a large C N o and S 4 ≤ 1
σ φ2ε
∞
() ∫
~
A =
−∞
f2
2
f + fn
2
.
(f
T
2
o
+f
)
2 p 2
.df +
(6.4-2)
2 ), this reduces to
Bn
1
1 +
ρ o 2Tρ o
(6.4-3)
()
~
= σ φ2εp + σ φ2T A
175
In Figure 6.4-1, the probability of a bit error is plotted as a function of S 4 and σ φ2εp for a
first order loop with an ideal AGC, T = 20ms, C N o = 41.5dBHz and Bn = 2Hz. Note that
σ φ2εp can be represented in terms of the phase scintillation spectral parameters T, p and
Word error probability (dB)
Bit error probability (dB)
f o for a given loop order and bandwidth.
−30
−40
−50
−60
−70
0.06
1
0.04
0.8
Phase error var. (rad )
−20
−30
−40
−50
0.06
1
0.04
0.8
0.6
0.02
2
−10
0.2
0.6
0.02
0.4
0
2
S4
Phase error var. (rad )
0.4
0
0.2
S4
Figure 6.4-1: Probability of a bit error (left panel) and a word error (right panel) as a function of
S 4 and σ φ2εp for T = 20ms, C N o = 41.5dBHz and Bn = 2Hz.
It is clear from Figure 6.4-1 that even under very strong amplitude and phase scintillation
conditions (ie. S 4 ≈ 1 and σ φεp ≈ σ φε
Th
), the probability of a word error in the navigation
data is only a few percent. Because of the high level of redundancy in the navigation data
(both within the navigation message and between satellites), it is unlikely that this level
of impairment will have much of an impact on a tracking GPS receiver. Indeed, the
complete loss of the navigation data will only affect GPS operation if the outage is long
enough for the ephemeris data to be significantly in error. As the ephemeris data can be
regarded as being accurate for many tens of minutes or more, short losses of a few
seconds to minutes would be inconsequential to a receiver (under sever scintillation
conditions, loss of code and carrier lock would be of more importance). However,
navigation data errors may affect the process of downloading almanac data from a
satellite during acquisition, thus extending a receiver’s time to first fix.
176
6.5. A note on word error probabilities
An implicit assumption in the analysis so far has been that the amplitude scintillation
waveform remains approximately constant for the duration of a navigation data bit (ie.
for 20 ms). Based on a knowledge of the typical fluctuation rates of amplitude
scintillations, it is expected that this assumption will be valid under most scintillation
conditions.
Word error probabilities were calculated by assuming independence between the bit error
probabilities of successive data bits. Consequently, from Equation (6.1-7) the word error
probability is given by
Pw = 1 − (1 − Pe )30
(6.5-1)
where Pe is the average bit error probability. As the bit error probability is conditioned
on the amplitude, this assumption implies that the amplitude is independent between
consecutive data bits. As explained later (see Section 7.3.3 on acquisition) this in turn
implies that the amplitude scintillation waveform must be fluctuating at a very rapid rate.
However, depending on the cutoff frequency of the amplitude scintillation power
spectrum3, f c , it is known that the amplitude waveform may vary quite slowly in
relation to the navigation data. For a slowly varying waveform, there will be fewer,
longer duration deep fades over a given interval of time. Therefore, the occasions during
which the bit error probabilities are at an elevated level are likely to be clustered and
associated with these longer duration deep fades. As the loss of only one data bit is
required for the loss of a word, it is expected that this condition will reduce Pw
somewhat.
To test this hypothesis, it is assumed that the amplitude remains approximately constant
for the 0.6 seconds duration of a word, but may vary between consecutive words (in fact,
this result will be the same if the amplitude is assumed to remain constant for longer
periods of time). The average word error probability is then
3
f c is a function of the satellite-receiver geometry and the ionospheric drift velocity (see Appendix
G).
177
∞
∫[
]
Pw = 1 − [1 − Pe (A )]30 f A~ (A ).dA
(6.5-2)
0
where Pe (A ) =
∞
∫ f X 0 A~ (x0 A ).dx0 = 2 Erfc
1
0
is a conditional bit error probability.
2σ n
A
Replacing A with ρ o gives
∞
[
(
Pw = 1 − 1 − 0.5Erfc Tρ o
∫
0
)]
30
m m ρ o m−1
Γ( m). C N m
o
mρ o
exp −
C No
.dρ o
(6.5-3)
In Figure 6.5-1, Pw is plotted as a function of S 4 for T=20ms and C N o = 41.5dBHz using
both Equations (6.5-1) and (6.5-3). It is clear that by assuming a slowly varying amplitude
waveform, the word error probability is reduced by as much as 5dB for values of S 4 near
to one. In practice, it is anticipated that the actual values of Pw will lie somewhere
between these two curves.
−10
Word error probability (dB)
−20
−30
−40
−50
−60
−70
−80
0.4
0.5
0.6
0.7
0.8
0.9
1
S4
Figure 6.5-1: Probability of a word error as a function of S 4 for T=20ms and C N o = 41.5dBHz.
The upper curve represent a situation in which the amplitude is assumed to be independent
between successive T second epochs. The lower curve represent a situation in which the amplitude
is assumed to be constant during each word.
Although this approach could be extended to include the effects of amplitude
scintillations on a subframe, the much greater length of a subframe (10 words = 6s) means
that the assumption of a constant amplitude is probably no longer valid.
178
6.6. Conclusions
In general, phase scintillations have much less of an effect on the navigation data than
amplitude scintillations. When phase scintillations are at the tracking threshold of the
carrier loop, the word error probability is less than 0.1%, compared to approximately 4%
when S 4 = 1 (assuming a nominal, quiescent GPS signal level and rapid amplitude
fluctuations). However, for a very slowly varying amplitude waveform, the word error
probability can fall to only 1% for S 4 = 1 under the same quiescent signal conditions.
Consequently, even under conditions for which the carrier loop is likely to lose lock, the
word error probability will only be a few percent, assuming that amplitude and phase
scintillations are uncorrelated.
It is considered unlikely that the small error rates associated with scintillations will cause
much of a problem for GPS for the following reasons:
• There is significant redundancy within the navigation message, particularly with
regard to the ephemeris data and clock correction factors which are repeated once
every frame (ie. every 30 seconds).
• There is also significant redundancy between the navigation messages transmitted by
different satellites. Each satellite transmits the same almanac data which contains
health, ionospheric correction factors and low precision orbital information etc. for all
of the satellites in the constellation.
• As much of the navigation data consists of slowly varying correction factors, loss of
the navigation data will only cause a gradual degradation in navigational accuracy.
Nevertheless, during the acquisition process, navigation data errors may have a more
significant effect, particularly is the receiver is tracking only one satellite and is
attempting to download almanac data following a cold start.
179
180
Chapter 7
Acquisition
In this chapter, the effects of scintillations on the acquisition performance of a GPS
receiver is examined. In Section 7.2, the effects of scintillations on the probability of
detection and the probability of false alarm are investigated for a full code period, squarelaw, Neyman-Pearson type detector. It is shown that both amplitude and phase
scintillations have a negligible effect on the probability of false alarm, but that amplitude
scintillations can significantly reduce the probability of detection. The effect of this
reduced probability of detection on the mean time to acquire the GPS signal is then
examined in Section 7.3 for a single dwell, serial search strategy for which there is
assumed to be no a priori information about the code phase. This is then extended to a
situation in which the correlation time of the amplitude scintillations is much longer than
the time required to execute one pass of the search domain. The results show that
amplitude scintillations increase the mean time to acquire, and that the effect is more
pronounced for longer amplitude correlation times (ie. for slower scintillations).
7.1. Acquisition model
Acquisition is the process of synchronising a local reference signal to the received GPS
signal prior to closure of the code and carrier tracking loops. The process involves a two
dimensional search for the GPS signal in both Doppler frequency and code phase.
As shown in Figure 7.1-1, the acquisition detector is essentially a square-law detector for
the GPS signal where the test statistic is given by
Z=
1
k
k
∑ I i2 + Qi2
(7.1-1)
i =1
The output of the detector, Z, is compared with a threshold, η, to determine whether a
satellite signal is present and whether it is correctly aligned with the reference signal. If
the threshold is exceeded, it is assumed that both the code delay and carrier frequency of
181
the reference signal are sufficiently close to those of the satellite signal for tracking to
begin.
p( t − tˆ)
t
1 dτ
T ∫
T
t−T
IF
t
1 dτ
T ∫
Ii2
+
T
t−T
1 k
k i=1
∑
Z
Qi2
900
cos(ω̂IFt)
Figure 7.1-1: A square-law acquisition detector for a GPS receiver.
Usually, when the reference is incorrectly aligned, it is assumed that the signal produced
at the output of the code and carrier mixers behaves like zero-mean, white Gaussian
noise. However, because correlation sidelobes1 may be present, particularly during C/ACode acquisition, this model is strictly not correct and a more accurate approach is to
assume that a signal of much lower strength is present under these conditions [92]. In the
analysis that follows, the white Gaussian noise model will be used predominantly, and
the error in this model will be discussed in Section 7.2.2.
7.2. Detection and false alarm probabilities
For an IF signal of the form A(t ) p(t − τ )d (t − τ ) cos(ω IF t + φ (t )) + n(t ) and a reference signal
(
)
of the form 2 p (t − τˆ )cos ωˆ IF t + φˆ , the I and Q signals immediately after the pre-detection
filters are given by [20]
sin(ω ε T 2 )
~
R (τ ε )cos(ω ε (t − T 2 ) + φε (t )) + n I (t )
I = A(t )d (t − τ )
ωε T 2
sin(ω ε T 2 )
~
R (τ ε )sin(ω ε (t − T 2 ) + φε (t )) + nQ (t )
Q = A(t )d (t − τ )
ωε T 2
(7.2-1)
where ωε = ω IF − ω# IF is the error in the frequency estimate, τ ε = τ − τ# is the error in the
1
Correlation sidelobes can result from correlation between the reference signal and another satellite signal, or
between an incorrectly aligned reference signal and the desired satellite signal.
182
code delay estimate, φε (t ) = φ (t ) − φˆ
is the phase error, and
R(τ ) is the code
autocorrelation function (Equation (4.1-1)). These two expressions assume that the
~
amplitude, A(t ) , and the phase, φ (t ) , of the GPS signal do not change appreciably over
the T second integration period of the filters, and that the navigation data does not
change sign. As T is typically of the order of a few ms, these conditions are usually met,
even in the presence of scintillations.
For a typical GPS receiver, the separation between bins in the Code/Doppler search
domain is of the order of ½ a chip and 3 ( 4T ) Hz respectively [20]. T can be made quite
small for strong satellite signals (eg. 1ms for a 750 Hz bin spacing), but must be large for
weak signals (eg. 10ms for a 75 Hz bin spacing). Increasing T to account for weak signals
will result in longer search times, even if the dwell time in each bin, kT, is kept constant,
as the number of Doppler bins required for a given frequency uncertainty will increase.
As the signal strength is rarely known a priori, assumptions must be made about the
receiver antenna gain pattern and the satellite signal power etc. in order to obtain a good
estimate of the signal strength. However, such assumptions are unlikely to take into
account effects such as obscuration and attenuation by nearby obstacles, nor the effects of
scintillations.
In the analysis that follows, it is assumed that if the correct bin is selected in the
Code/Doppler search domain, τ ε and ωε will both be zero. If the bins are separated by
τ ∆ chips and ω ∆ radians/s, the maximum error in this assumption will be
ξ=
sin (ω ∆ T 4 )
R (τ ∆ 2 )
ω ∆T 4
(7.2-2)
To account for this error, the GPS signal power can be multiplied by a correction factor,
ξ 2 , prior to calculating the probability of detection. If, ω ∆ = 2π ∗ 3 4T radians/s and
τ ∆ = 1 2 a chip, this correction factor will be -4.6dB.
By assuming ω ε = τ ε = 0 , the I and Q samples produced at the output of the pre-detection
filters for a correctly aligned reference signal can be represented by
~
I i = Ai cos(φεi ) + n Ii
~
Qi = Ai sin(φεi ) + nQi
(7.2-3)
183
The navigation data, d (t − τ ) , has been ignored in these expressions as it will be
eliminated by squaring in the subsequent stage. When the reference signal is incorrectly
aligned, it is assumed that I i = n Ii and Qi = nQi .
The probability density function (PDF) of the test statistic, Z, in the presence of noise
only, or in the presence of an incorrectly aligned signal (Hypothesis H 0 ) is given by the
central chi-squared distribution with 2k degrees of freedom [84]
k
k z k −1
zk
f Z H 0 (z ) =
. exp −
,
2 Γ(k )
2σ 2
n
2σ n
z≥0
(7.2-4)
where, σ n 2 = N o T is the thermal noise variance on either the I or Q channels and k is the
order of the post-detection integrator. The PDF of the test statistic in the presence of a
correctly aligned satellite signal (Hypothesis H1 ) is given by the non-central chi-squared
distribution with 2k degrees of freedom [84]
f Z H1 (z ) =
k
2σ n 2
z
~2
A
(k −1) 2
~
kA
z
I k −1
σ 2
n
(
~2
exp − k z + A
2σ n 2
),
z≥0
(7.2-5)
~
where A = σ n 2T C N o is the signal amplitude immediately after the pre-detection
filters, and I k−1 (
)
is the modified Bessel function of the first kind of order k-1. If a
detection threshold of η is chosen, the probability of a correct detection is given by
∞
Pd =
∫ f Z H1 (z ).dz
η
∞
=
k
z
∫ 2σ n 2 A~ 2
(k −1) 2
η
~
kA
z
I k −1
σ 2
n
(
~2
exp − k z + A
2σ n 2
)
.dz
(7.2-6)
and the probability of a false detection (false alarm) is given by [84]
∞
Pfa =
∫ f Z H0 (z ).dz
η
∞
k
k z k −1
− zk .dz
.
exp
=
2
2σ 2
n
2σ n Γ(k )
η
∫
= exp(− η ′).
184
k −1
η′ j
∑ j!
j =0
(7.2-7)
where η ′ = kη 2σ n 2 . As the random variables, I i2 + Qi2 , generated at the output of the
square law detector are independent, the Central Limit theorem can be invoked for large
values of k to allow Z to be approximated by a Gaussian distribution. The corresponding
probabilities of detection and false alarm are then [84]
β −γ k
Pd = Q
1 + 2γ
(7.2-8)
Pfa = Q (β )
(7.2-9)
η
~
where Q is the Gaussian probability integral, β = k
− 1 and γ = A 2 2σ n 2 is the
2
2σ
n
signal to noise ratio.
Equations (7.2-8) and (7.2-9) (or (7.2-6) and (7.2-7)) can be used to select the design
parameters k and η once the required values of Pd and Pfa have been chosen for a given
application. σ n 2 can also be adjusted through the pre-detection integration period, T, but
is restricted somewhat by the presence of navigation data and by uncertainties in the
carrier Doppler.
From these equations it is clear that Pfa is not a function of the GPS signal characteristics,
and so will not be affected by scintillations. However, Pd is a function of the GPS signal
~
amplitude, A , and so will be directly affected by amplitude scintillations. In the presence
of amplitude scintillations, the PDF of the test statistic Z under Hypothesis H 1 becomes
~
conditional on the amplitude, A , and can be represented by f Z H
~
1, A
(z A ) . The marginal
PDF of Z H1 is thus
f Z H1 (z ) =
∞
∫ f Z H1, A~ (z A ). f A~ (A ).dA
(7.2-10)
0
where f A~ (A ) is the Nakagami-m distribution for amplitude (Equation (2.1-8)). The
average probability of detection is therefore
185
∞
Pd =
∫ ∫
η
∞
=
∞
f Z H , A~ (z A ). f A~ (A ).dA.dz
1
0
∞
f
~ (z A ).dz f ~ (A ).dA
η Z H1, A
A
∫ ∫
0
(7.2-11)
∞
=
∫ Pd (A ). f A~ (A ).dA
0
where Pd (A ) is the probability of detection as a function of the signal amplitude and can
be obtained from either Equation (7.2-6) or (7.2-8). In Figure 7.2-1, Pd is plotted as a
m ) for five values of C N o and for Pfa = 0.01% (using Equations
function of S 4 ( = 1
(7.2-8) and (7.2-9)). It is clear from this figure that the probability of detection decreases as
S 4 increases, and that the effect is more pronounced for smaller values of C N o .
Consequently, satellite links that penetrate the peak of the ionospheric anomaly at low
elevation angles are likely to have the poorest acquisition performance ( C N o is likely to
be lower and S 4 larger under these conditions).
1
0.95
40
38
0.9
36
0.85
0.8
34
0.75
0.7
32dBHz
0.65
0.6
0.55
0
0.2
0.4
0.6
0.8
1
S4
Figure 7.2-1: Pd as a function of S 4 for five values of C N o (32 to 40dBHz in 2dBHz steps).
T=1ms, k=20, and Pfa =0.01%.
By inverting Equation (7.2-8) and solving for the signal to noise ratio, an equivalent C N o
for quiescent ionospheric conditions can be obtained that will produce the same
186
probability of detection as Equation (7.2-11). If α is defined as the inverse of Equation
(7.2-8) and we let Pd = Pd , we have
α = Q −1 (Pd )
=
β −γ ′ k
1 + 2γ ′
(7.2-12)
where γ ′ is the equivalent signal to noise ratio and β is constant for a given Pfa . Solving
for γ ′ gives
γ ′ = α 2 + β k + α α 2 + 2 β k + k k
(7.2-13)
from which the equivalent C N o can be obtained using C N o′ = γ ′ T . In Figure 7.2-2,
C N o′ is plotted as a function of S 4 using the same five values of C N o that were used in
Figure 7.2-1. It is clear from this figure that C N o′ decreases as S 4 increases and that the
effect is more pronounced for large values of C N o .
40
40
Equivalent C/No (dBHz)
38
38
36
36
34
32
34
32
30
0
0.2
0.4
0.6
0.8
1
S4
Figure 7.2-2: Equivalent C N o as a function of S 4 for five values of C N o (32 to 40dBHz in
2dBHz steps). T=1ms, k=20, and Pfa =0.01%.
187
7.2.1. Phase scintillation effects
If τ ε and ωε are assumed to be zero when the correct Code/Doppler bin is selected, the
test statistic, Z, can be represented by Equation (7.1-1) with I i and Qi given by Equation
(7.2-3). By making the following substitutions
n Ii = nci cos(φεi ) + n si sin(φεi )
(7.2-14)
nQi = nci sin (φεi ) − n si cos(φεi )
where nci and nsi are uncorrelated, zero-mean, baseband Gaussian noise processes with
variances of σ n 2 = N o T , the test statistic becomes
Z=
1
k
k
∑ (Ai + nci )
~
2
+ n si2
(7.2-15)
i =1
In this form, it is clear that the carrier phase (and therefore phase scintillations) do not
affect Z and so will not influence the detection process for acquisition, provided that the
phase does not vary significantly over the integration period of the pre-detection filters. A
measure of the phase variation over the integration period of the filters is the expectation
of the phase variance over that period. This is given by
σφp
T
1
where φ p =
T
t
∫ φ p (u ).du
2
1 t
2
φ p (u ) − φ p .du
= E
T t −T
∫[
]
(7.2-16)
is the average value of φ p (t ) over a T second period (which is
t −T
also the output of a T second integrate and dump filter which operates directly on the
phase). Equation (7.2-16) can be simplified as follows
σφp
T
2
=
1
T
2
2
∫ [E {φ p (u ) }− 2 E{φ p (u )φ p }+ E {φ p }].du
t
t −T
1 t
φ p (u ).du + E φ p 2
= Rφ p (0 ) − 2 E φ p
T t −T
∫
{ }
= Rφ p (0 ) − E φ p 2
188
{ }
(7.2-17)
where Rφ p (0 ) =
∞
∫ Sφ p ( f ).df
is the power in the phase scintillations prior to filtering,
−∞
∞
{ } ∫ G ( f ) .S
2
2
E φp =
φp
( f ).df
is the power in the phase scintillations after a T second
−∞
integrate
and
dump
filter
which
operates
directly
on
the
phase,
and
G ( f ) = sinc( fT )exp(− jπfT ) is the transfer function of such a filter. Thus,
∫ [1 − G ( f )
∞
σφ p
2
T
=
−∞
2
].S
φp
( f ).df
(7.2-18)
∞
=
∫ [1 − sinc( fT ) ]. (f
T
2
−∞
2
o
+ f2
)
p 2
.df
This can be rearranged to obtain a threshold, T, below which the approximation can be
considered to be valid, viz
T Th =
where σ φ p
T
2
σφp
2
T
Th
∞
1
1 − sinc( fT )2 .
− ∞
fo2 + f 2
∫[
]
(
)
p
df
.
2
(7.2-19)
is a threshold variance.
Th
19.46
19.44
T (dBW/Hz)
19.42
19.4
19.38
19.36
19.34
19.32
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0.08
0.09
0.1
fo (Hz)
Figure 7.2-3: Threshold T as a function of f o for a threshold variance of 1 rad2, T=10 ms and
p=2.5.
189
In Figure 7.2-3, the threshold spectral strength, T, is plotted as a function of f o for a
threshold variance of 1 rad2, T=10 ms (a typical upper limit) and p=2.5. It is clear from this
figure that the spectral strength must be enormous in order for phase scintillations to
cause a significant deviation in the carrier phase over a typical filter integration period.
By comparison, a typical large value for the spectral strength parameter in equatorial
regions during high solar activity is about -20 dBW/Hz (ie. about 40dB below the values
given in Figure 7.2-3). Consequently, for the acquisition model described earlier, the
effects of phase scintillations can be ignored.
7.2.2. Correlation sidelobes
The impact of correlation sidelobes on the probability of false alarm can be found by
obtaining a new Pfa based on the probability of detection equation (Equation (7.2-8)), but
using a signal level that is significantly less than the nominal GPS signal level. For the
GPS Gold codes (the C/A-Code), the largest sidelobe is approximately 22dB below the
correlation peak, although this can vary by several decibels depending on the Doppler
offset [87] & [47]. Consequently, the worst case Pfa will be associated with a very strong
satellite signal that is producing a sidelobe at the maximum level.
0.1
Pfa (%)
0.08
0.06
40
0.04
38
0.02
0
0
36
34
32
0.2
0.4
0.6
0.8
1
S4
Figure 7.2-4: Pfa for a peak sidelobe level of -22dB as a function of S 4 for five values of C N o
(32dBHz to 40dBHz in 2dBHz steps). T=1ms, k=20, and the design Pfa =0.01%.
190
In Figure 7.2-4, the average false alarm probability, Pfa , obtained from Equation (7.2-8)
using the maximum sidelobe level is plotted as a function of S 4 for the same five values
of C N o that were used in Figure 7.2-1. The threshold, η, has once again been chosen for
a design Pfa of 0.01% based on the assumption that the input to the detector is white
Gaussian noise when the desired satellite signal is incorrectly correlated. It is clear from
these plots that P fa increases as S 4 increases, and that the effect becomes more
pronounced as the GPS signal level increases. This is because enhancements in the
sidelobe levels caused by amplitude scintillations only become a problem when the GPS
signal level is relatively large. For small signals, the sidelobe energy remains below the
noise floor at the output of the pre-detection filters, even when the signal level has been
significantly enhanced.
The results given in Figure 7.2-4 are based on the maximum sidelobe level which will
only occur infrequently. In [87], the cumulative probability distribution function of the
sidelobe levels is given for the GPS Gold codes for a range of Doppler shifts from 0 to
± 5kHz. This distribution function is obtained by averaging the results for all 1023 Gold
codes in the GPS family, for all possible code time offsets, and for all possible code pairs.
By differentiating this function, the PDF of the sidelobe levels, f S (s ) , can be found (see
1
0.1
0.5
0.08
0
0
1
2
3
4
5
6
−3
x 10
pdf
1000
500
0
0
Pfa (%)
cdf
Figure 7.2-5, left panel for the case where a Doppler shift of a few kHz is assumed).
0.06
0.04
0.02
1
2
3
4
Sidelobe Level, S
5
6
−3
0
0
x 10
0.2
0.4
0.6
0.8
1
S4
Figure 7.2-5: CDF and PDF of the sidelobe levels for the GPS Gold codes (left panel) and the
corresponding Pfa (right panel). T=1ms, k=20, and the design Pfa =0.01%. A Doppler shift
between the satellite and local codes of a few kHz is assumed.
191
The average probability of false alarm for a single interfering signal can then be obtained
by averaging the probability of detection expression over both the amplitude (to account
for scintillations) and the sidelobe levels as follows
∞∞
Pfa =
∫ ∫ Pd (s ∗ A). f A~ (A). f S (s ).dA.ds
(7.2-20)
00
where f A~ (A) is the Nakagami-m PDF for a nominal GPS signal, and s is the sidelobe level
(s and A are independent random variables). When Equation (7.2-20) was evaluated for
the five values of C N o used in Figure 7.2-4, the effects of sidelobes on Pfa were found to
be negligible, even at high S 4 (see Figure 7.2-5, right panel). However, as the sensitivity
to sidelobes and scintillations will depend very much on the design parameters Pfa , k
and T, this result really only applies to this particular example. In addition, these results
assume that only one interfering signal is present. In reality, there may be up to 12
satellites visible (possibly more in the future) which will increase the probability that a
strong sidelobe will be present.
Van Dierendonck [92] suggests that the detection threshold, η, should be adjusted to
account for the worst case sidelobe level while maintaining Pfa at a desired level. This
can be achieved by solving the probability of detection equations ((7.2-8) or (7.2-6)) for η
with Pd = P fa
design
and using a signal level that corresponds to the maximum sidelobe
level. If this is done, the tracking threshold, η, will be substantially larger and the effects
of sidelobes and scintillations will probably be negligible.
Note that because of the extremely low sidelobe levels for the P(Y)-Code, the effect
outlined above will not be apparent during direct P(Y)-Code acquisition (at present,
direct P(Y)-Code acquisition is only available to the US military).
192
7.3. Acquisition times
In the previous section, the probability of detection and the probability of false alarm
were calculated for a single cell in the code phase / carrier Doppler uncertainty region.
The detector type used in this analysis was a fixed integration time, square-law detector
based on a Neyman-Pearson detection strategy. In this section, the time required to
identify the correct cell will be examined for a single dwell1, serial search strategy for
which there is assumed to be no prior information about the code phase. It will also be
assumed that the code Doppler is zero, and that the search is only conducted in the code
domain (ie. the frequency of the replica carrier is assumed to be approximately correct).
7.3.1. Acquisition search strategy
In the absence of a priori code phase information, a serial search begins at the start of the
uncertainty region and progresses through each cell in sequence until a successful
detection is made. If a detection is not made before the end of the region is reached, the
search returns to the beginning and is repeated. If a false alarm occurs, time will be
expended in order to verify that the detection is incorrect before the search can continue.
This is usually referred to as verification time and may include the time required to re-run
the detector on a particular cell, or the time associated with a failed attempt to revert to
code tracking mode. The total time required to make a correct detection is therefore a
function of the time spent in each cell (the dwell time, Td = kT ), the number of cells in the
uncertainty region, N C , the probability of detection, Pd , the probability of false alarm,
P fa , and the verification time, Tv = K v Td where K v is a factor greater than one.
In the absence of scintillations, the mean time to acquire, T ACQ , and the RMS acquisition
time, σ ACQ , for a uniform serial search are given by [84]
2 − Pd
T ACQ = N C Td K v Pfa + 1
2 Pd
(
1
)
(7.3-1)
Single dwell: The detector’s decision is based on a single, fixed time observation of the received signal plus
noise. The alternative is multiple dwell in which multiple observations are used to verify the first observations
193
σ ACQ = N C Td (K v Pfa + 1)
1
1
1
+ 2−
12 Pd Pd
(7.3-2)
In Section 7.2, it was shown that when scintillations are present, the probability of
~
detection becomes a function of the signal amplitude, A (if sidelobes are ignored, the
probability of false alarm will be unaffected by scintillations). The average mean time to
acquire and the RMS acquisition time must then be found by taking an ensemble average
of these two parameters over all possible realisations of the amplitude. In order to do this,
assumptions must be made about the statistics of the amplitude over the time period
required for acquisition. In particular, the joint statistics of the amplitude at time periods
separated by N C Td seconds must be determined (ie. the time between successive re-visits
to the correct code phase cell, assuming that no false alarms have occurred).
In the following sections, it will be assumed that the amplitude is approximately constant
during the relatively short dwell time, Td , in each cell, but that significant variations may
occur between the start and end of the acquisition process. As before, it will be assumed
that these variations are described by the Nakagami-m distribution.
7.3.2. Mean time to acquire
This analysis closely follows that given by Peterson & Ziemer [72], but has been modified
to account for variations in the signal amplitude between successive re-visits to the
correct code phase cell. In order to keep the time between re-visits, Tr , constant, it will be
assumed that false alarms do not occur, and so Tr = N C Td . The justification for this
assumption is that most acquisition systems are designed to have a very small P fa such
that K v P fa << 1 . Therefore, from Equations (7.3-1) and (7.3-2) it is clear that to a first
approximation, the mean and RMS acquisition times are largely unaffected by false
alarms. As P fa is also not affected by scintillations (if sidelobes are ignored), it is expected
that this approximation will also hold under scintillation conditions. The error in this
approximation will be discussed further in Section 7.3.4.
If the nth cell is assumed to be the correct code phase cell, and there are j missed
detections, then the total acquisition time is given by
194
T ACQ (n, j ) = nTd + jN C Td
(7.3-3)
where the first term represents the time required to detect the correct cell on the final
(successful) pass, and the second term represents the time expended in the j unsuccessful
passes. The probability of this event occurring, assuming that there is no prior
information about the location of n, is given by 2
(
j
∏[
)
1
~
~
~
Pr n, j A sj =
Pd ( As )
1 − Pd ( As −i )
NC
i =1
[
]
(7.3-4)
]
~
~ ~
~
~
where A sj = As , As −1 , As −2 , 4 , As − j is a vector that represents the amplitude of the GPS
signal at the times during which the acquisition detector is testing the correct code phase
~
~
cell (ie. As −i = A(t s −i ) where t s −i is the time corresponding to the ( j − i + 1)th pass through
~
the correct cell, and t s −i − t s −i −1 = Tr ). As a function of A sj , the mean time to acquire is
therefore
{
( )
~
~
T ACQ A sj = E n, j T ACQ (n, j ) A sj
}
NC ∞
∑∑ T ACQ (n, j )Pr (n, j
=
~
A sj
n =1 j =0
(7.3-5)
)
~
By taking the expectation over all possible realisations of A sj , this becomes
{
( )}
~
T ACQ = E A~ T ACQ A sj
sj
=
NC ∞
∑∑ E A~ sj {
(
~
T ACQ (n, j )Pr n, j A sj
n =1 j =0
(7.3-6)
)}
Substituting Equations (7.3-3) and (7.3-4) into the above expression gives
T ACQ =
Td
NC
j
∞
~
~
1 − Pd ( As −i ) +
n E A~ Pd ( As )
sj
n =1
i =1
j =0
NC
∑ ∑
]
(7.3-7)
j
~
~
jE A~ Pd ( As )
1 − Pd ( As −i )
sj
j =0
i =1
∞
NC
∏[
∑
∏[
]
0
2
By definition,
∏ f (i ) = 1 .
i =1
195
In order to obtain a closed form solution to this expression, it is necessary to simplify the
following two terms
Term1=
∞
j =0
(7.3-8)
j
~
~
1 − Pd ( As −i )
Pd ( As )
i =1
(7.3-9)
∞
Term2=
j
∑ E A~sj Pd ( As )∏ [1 − Pd ( As−i )]
∑
j =0
jE A~
sj
~
~
i =1
∏[
]
These terms can only be simplified once the joint statistics of the amplitude scintillations
have been established. This will be discussed next.
7.3.2.1. Amplitude correlation times
The correlation time, TCT , of amplitude scintillations is a function of both the Fresnel
zone radius, z F = hi λ 3, and the relative velocity between the satellite ray path and the
irregularity structure (which in turn is a function of the irregularity drift velocity, the
receiver velocity, satellite motion and satellite/receiver geometry). TCT is typically of the
order of a few seconds, but may extend to a few tens of seconds if the ionospheric pierce
point tracks the irregularity drift [50]. For a receiver that is attempting to acquire the GPS
signal with no prior information about the code phase (ie. a cold start), the number of
cells, N C , is typically of the order of 2046 (ie. 2 x the number of code chips in the C/ACode assuming a cell spacing of ½ a code chip). For a dwell time of 20ms, this equates to a
re-visit time, Tr = N C Td , of 41 seconds. Consequently, for a cold start it is likely that the
correlation time of the amplitude will be much less than the time to re-visit the correct
cell. This situation is discussed in Section 7.3.2.2. However, if a receiver is re-acquiring
following a relatively short period of loss of lock (ie. a warm start), the correlation time
may be greater than the re-visit time. The re-acquisition times of modern GPS receivers
following a period of signal loss of greater than 1 minute are typically less than a few
seconds (GPS World, January 2000, pages 34–54). Consequently, for such receivers, Tr
would be expected to be only a few seconds which may be less than TCT . This situation is
discussed in Section 7.3.2.3.
3
hi is the height of the ionospheric pierce point as discussed in Section 2.1.
196
7.3.2.2. Short amplitude correlation times
If it is assumed that the amplitude correlation time is much less than the re-visit time (ie.
during a cold start or in the presence of rapid amplitude scintillations), it is likely that the
~
individual amplitude values in the A sj vector will be independent (see Section 7.3.3 for a
~
justification of this assumption). Consequently, the joint PDF of A sj becomes the product
of the individual marginal PDFs as follows
f A~
sj
(A~ s , A~ s−1 , A~ s−2 , 4 , A~ s− j ) = f A~s (A~ s ) f A~s−1 (A~ s−1 ) f A~s−2 (A~ s−2 )4 f A~s− j (A~ s− j )
(7.3-10)
where each marginal PDF follows the Nakagami-m distribution. The expectation
expressions within Equations (7.3-8) and (7.3-9) can then be simplifies as follows
j
~
~
E A~ Pd ( As )
1 − Pd ( As −i )
sj
i =1
∏[
]
∞
∞
0
0
i =1
∞
∞
j
0
0
~
= 5 Pd ( A s )
∫ ∫
∏ [1 − Pd (A s−i )]f A~sj (A s , 4, A s− j ).dA s 4dA s− j
~
= 5 Pd ( A s )
∫ ∫
j
~
~
~
~
~
∏ [1 − Pd (A s−i )]f A~s (A s )4 f A~s− j (A s− j )dA s 4dA s− j
~
~
~
~
(7.3-11)
~
i =1
}∏ [1 − E A~s−i {Pd (A~ s−i )}]
{
s
j
~
= E A~ Pd ( A s )
= Pd (1 − Pd )
i =1
j
This result assumes that the amplitude scintillations are a stationary random process, and
~
~
so Pd = E A~ Pd ( A i ) = E A~ Pd ( A k ) . Consequently, Equations (7.3-8) and (7.3-9) reduce to
i
{
}
k
{
}
∞
Term1=
∑ Pd (1 − Pd )
j
j =0
∞
Term2=
∑ jPd (1 − Pd )
j
j =0
(7.3-12)
(7.3-13)
Substituting these two terms back into Equation (7.3-7) results in the following expression
T ACQ =
Td
NC
∞
j
n Pd (1 − Pd ) + N C
n =1
j =0
NC
∑ ∑
j
jPd (1 − Pd )
j =0
∞
∑
(7.3-14)
197
This expression has the same form as that derived by Peterson & Ziemer [72] (ignoring
false alarms), except that the standard probability of detection has been replaced by Pd .
As shown by Peterson & Ziemer, Equation (7.3-14) can be reduced to
T ACQ =
2 − Pd
Td
+ N C Td
2
2 Pd
(7.3-15)
2 − Pd
≈ N C Td
for large values of N C and Pd
2 Pd
Indeed, as this result assumes that Tr is large enough for the amplitude to be
independent between successive re-visits, the additional verification time associated with
~
false alarms will not alter the statistics of A sj . Consequently, the effects of false alarms
can readily be incorporated to obtain (see Peterson & Ziemer for a justification of this
step)
2 − Pd
T ACQ = N C Td K v Pfa + 1
2 Pd
(
)
(7.3-16)
The percentage increase in the mean acquisition time as a result of scintillations is then
found by dividing Equation (7.3-16) by (7.3-1) as obtain
T ACQ
T ACQ
scint
no scint
=
Pd (2 − Pd )
Pd (2 − Pd )
(7.3-17)
where Pd and Pd are the probabilities of detection in the presence of scintillations and in
the absence of scintillations respectively.
2
32dBHz
1.5
34
36
38
1
0
0.2
0.4
0.6
S4
198
0.8
40
1
Figure 7.3-1: Mean acquisition time ratio as a function of S 4 for C N o =32 to 40 dBHz. T=1ms,
k=20, P fa =0.01%.
In Figure 7.3-1, Equation (7.3-17) is plotted as a function of S 4 for five values of C N o .
The worst case increase in the mean acquisition time is by a factor of approximately two,
although this corresponds to quite a low quiescent C N o . However, at normal signal
levels (around 40dBHz) the increase in the acquisition time is relatively small, even under
strong scintillation conditions.
A similar argument can be used to demonstrate that the effects of scintillations on the
RMS acquisition time can be found by substituting Pd in place of Pd in Equation (7.3-2).
Again, this only applies to the situation in which TCT is much less than Tr . The
corresponding RMS acquisition time ratio is therefore
σ ACQ
σ ACQ
scint
no scint
=
Pd 2 − 12 Pd + 12
Pd
Pd
(7.3-18)
Pd 2 − 12 Pd + 12
In Figure 7.3-2, the RMS acquisition time ratio is plotted as a function of S 4 for the same
four values of C N o that were used in Figure 7.3-1. From these two figures, it is clear that
both the mean and the spread of the acquisition times increases.
3.5
3
2.5
2
32dBHz
34
36
1.5
1
0
38
0.2
0.4
0.6
0.8
40
1
S4
Figure 7.3-2: RMS acquisition time ratio as a function of S 4 for C N o =32 to 40 dBHz. T=1ms,
k=20, P fa =0.01%.
199
7.3.2.3. Long amplitude correlation times
If the correlation time of the amplitude increases, or conversely the re-visit time
~
decreases, it may no longer be valid to assume that the components of A sj are
~
independent. Under these conditions, it is not valid to represent the joint PDF of A sj as
the product of the individual marginal PDF’s, as in Equation (7.3-10). The difficulty that
arrises under these circumstances is that because the joint PDF is unknown, the two
expectation terms in Equations (7.3-8) and (7.3-9) cannot be solved. Therefore, an
expression for T ACQ cannot be found. Indeed, if the joint PDF were known, it is likely that
the complexity involved in attempting to solve Equation (7.3-7) would be so high that a
closed form solution would be very difficult to find.
The approach used here to achieve an increase in the correlation time of the amplitude
while still allowing T ACQ to be solved in closed form is to simply repeat amplitude values
within the original sequence. For example, if each amplitude value is repeated once only,
the original sequence becomes
[ (
[(
)
~a
~ ~
~
A
~
′ sj = As , As −1 = As −2 ,
′
A sj = ~ b
~
~
~
A ′ sj = As = As −1 , As −1 ,
)
(
~
~
As −2 , As −3 =
~
~
As −2 = As −3 ,
(
)
)
~
As − 4 ,
~
As −3 ,
]
~
As −4 , 4 ,
with probability ½
~
~
As −4 = As −5 , 4 , with probability ½
(
) ]
(7.3-19)
where now only every second amplitude value is independent (ie.
~ ~
~
~
p A i , A i −n = p A i p A i −n for n > 1 ). Based on this model, the joint PDF’s of the
(
) ( ) (
)
amplitude sequence become
(
)
(
) [ ( ) (A~s − A~s−1 )]
1 ~
~
~ ~
p As , As −1 = p As −1 p As + δ
2
1 ~
~ ~
~
j = 2 : p A s , A s −1 , A s −2 = p As −2 p
2
~ ~
~
~
j = 3 : p As , As −1 , As −2 , As −3
j = 1:
(
(
)
) [ (A~s )δ (A~s−1 − A~s−2 )+ p(A~s−1 )δ (A~s − A~s−1 )]
(7.3-20)
)
(
1 ~
p As −3
2
etc. for j > 3
=
(
) [p(A~s )δ (A~s−1 − A~s−2 )p(A~s−2 )+ p(A~s−1 )δ (A~s − A~s−1 )δ (A~s−2 − A~s−3 )]
( )
( )
~
~
where the notation p Ai has been used in place of f A~ A i to improve clarity, and δ (
i
) is
the Dirac Delta function [39]. The joint PDF can be generalised for arbitrary values of j to
give
200
(
~
~
p As , 4 ,As − j
=
) j even
j2 ~
1 ~
~
~
p As − j
p As −2 q+1 δ As −2 q+ 2 − As −2 q +1 +
2
q=1
(
(
)∏ (
~
~
p As , 4 ,As − j
)(
j 2
q =1
) ∏ p(A~s−2q+2 )δ (A~s−2q+1 − A~s−2q )
(7.3-21)
) j odd
( j −1) 2
~ ( j −1) 2 ~
1 ~
~
~
~ ~
~
~
~
p As −2 q δ As −2 q+1 − As −2 q + δ As − As −1
p As −2 q+1 δ As −2 q − As −2 q−1
= p As − j p As
2
q =1
q =1
(
) ( )∏ (
)(
) (
)∏ (
)(
( )
~
~
The sequence A ′ sj is wide sense stationary and the marginal density functions, p Ai , are
~
again described by the Nakagami-m PDF. The autocorrelation function of A ′ sj is given by
{
~~
R (nTr ) = E Ai Ai −n
∞∞
=
∫∫
}
(
(7.3-22)
)
~~
~ ~
~ ~
Ai Ai −n p Ai , Ai −n .dAi .dAi −n
00
where n is an integer. As it has been assumed that amplitude values separated by more
than one sample are independent, the autocorrelation function reduces to
{ }
~
R (0) = E A 2 = A2
{ }2 = (A )2 ,
~
R (nTr ) = E A
R (Tr ) =
∞∞
n >1
∫ ∫ Ai Ai−1 p(Ai , Ai−1 ).dAi .dAi−1
~~
~ ~
~
~
(7.3-23)
00
∞∞
=
∫ ∫ Ai Ai−1 2 p(Ai−1 )[p(Ai )+ δ (Ai − Ai−1 )].dAi .dAi−1
~~
1
~
~
~
~
~
~
00
=
1 2
2
A + (A )
2
where R (0) ≥ R (Tr ) ≥ R (nTr ) as A 2 ≥ (A ) . Consequently, the correlation time of the
2
amplitude sequence is now between Tr and 2Tr seconds (originally, it was less than Tr
seconds).
~
The joint PDF of A′ sj (Equation (7.3-21)) can be used to obtain closed form expressions for
Equations (7.3-8) and (7.3-9), which in turn can be used to derive a simplified expression
for T ACQ . Expanding Equation (7.3-8) gives
201
)
term1 = E A~′
s∞
where E A~′
s∞
{Pd (A~s )[1 + Q (A~s−1 )+ Q (A~s−1 )Q (A~s−2 )+ Q (A~s−1 )Q (A~s−2 )Q (A~s−3 )+ 5] }
( ) [
(7.3-24)
( )]
~
~
= E A~′ for j → ∞ , and Q Ai = 1 − Pd Ai . By taking account of the two
sj
~
separate amplitude sequences in A ′ sj (see Equation (7.3-19)), this can be reduced to
term1 =
( )
(
) (
1
~
~
~
E ~ a Pd As 1 + Q As −2 + Q As −2
′
A
2
s∞
( )
)2 + Q (A~s−2 )2 Q (A~s−4 )+ Q (A~s−2 )2 Q (A~s−4 )2 + 5 +
( ) ( )(
) ( )(
1
~
~
~
~
~
~
E ~ b Pd As −1 1 + Q As −1 + Q As −1 Q As −3 + Q As −1 Q As −3
2 A′s∞
)2 + Q (A~s−1 )Q (A~s−3 )2 Q (A~s−5 )+ 5
(7.3-25)
for j → ∞ . Using properties of the expectation such
where E ~ a = E ~ a and E ~ b = E ~ b
A′s∞
A′s∞
A′sj
{ ( )(
~
~
as E ~ a Pd As Q As −2
A′s∞
term1 =
=
=
1
Pd 1 + Q
2
[
A′sj
) }= E A~s {Pd (A~s )}E A~s−2 {Qd (A~s−2 )}, Equation (7.3-25) becomes
]1 + Q 2 + Q 2
1
Pd + Pd + Pd Q 1 + Q
2
[
[
+ 5 etc + Pd + Pd Q 1 + Q
2
][
∞
] ∑Q 2
2
+ 5 etc
j
j= 0
1
Pd + 2 Pd − Pd2 [2 − Pd ]
2
]1 + Q 2 + Q 2
j
Q2
j=0
∞
∑
(7.3-26)
where
Q = 1 − Pd
{[
( )] }
{ ( )[ ( )]}
~ 2
Q 2 = E 1 − Pd Ai
= 1 − 2 Pd + Pd2
~
~
Pd Q = E Pd Ai 1 − Pd Ai = Pd − Pd2
However, as
∞
∑Q 2
j
=
j= 0
0 ≤ Q2 =
∞
∑ 1 − 2Pd − Pd2
j
, and
j= 0
∞
∞
0
0
~ 2 ~ ~
~ ~
∫ [1 − Pd (Ai ) ] p(Ai ).dAi ≤ ∫ p(Ai ).dAi = 1
we can say that
∞
∑
j=0
202
j
Q2 =
1
2 Pd − Pd2
(7.3-27)
Therefore, Equation (7.3-26) reduces to
1
1
term1 = Pd + 2 Pd − Pd2 [2 − Pd ]
2
2 P − P 2
d
d
=1
(7.3-28)
This result is the same as that obtained for both the constant amplitude and the
independent amplitude sequence. In a similar way, Equation (7.3-9) can be expanded
~
using the A ′ sj model from Equation (7.3-19) to give
term2 =
( ) (
)
(
1
~
~
~
E ~ a Pd As Q As −2 + 2Q As −2
′
A
2
s∞
( ) (
)
)2 + 3Q (A~s−2 )2 Q (A~s−4 )+ 4Q (A~s−2 )2 Q (A~s−4 )2 + 5 +
( )(
)
( )(
1
~
~
~
~
~
~
E ~ b Pd As −1 Q As −1 + 2Q As −1 Q As −3 + 3Q As −1 Q As −3
′
A
2
s∞
)2 + 4Q (A~s−1 )Q (A~s−3 )2 Q (A~s−5 )+ 5
(7.3-29)
By once again evaluating the two expectation terms, this expression reduces to
term2 =
=
=
=
2
1
2
2
2
2
2
Pd Q + 2Q + 3Q Q + 4Q + 5 etc + Pd Q 1 + 2Q + 3Q + 4Q Q + 5 etc
2
1
Pd
2
Q
∞
∑ (1 + 2 j )Q 2
j
+
j= 0
1
Pd Q + Pd Q (2Q + 1)
2
[
j
2 j Q 2 + Pd Q Q
j= 0
∞
∑
∞
]∑ Q 2 j
[
∑ (2 + 2 j )Q 2
j=0
+ 2 (Q + 1) Pd + Pd Q
j= 0
1
4 Pd − 3Pd 2 − 3Pd2 + 2 Pd Pd2
2
∞
∞
∑
Q2
j
j=0
∞
j= 0
j
+
∞
∑ (1 + 2 j )Q 2
j=0
j
]∑ j Q 2 j
+ 2 4 Pd − 2 Pd 2 − 2 Pd2 + Pd Pd2
∞
∑
j= 0
j
j Q2
(7.3-30)
Also,
∞
∑ jQ 2
j
=
j= 0
∞
∑ j 1 − 2 Pd − Pd2
j
j= 0
=
1 − 2 Pd + Pd2
2 P − P 2
d
d
(7.3-31)
2
By substituting Equations (7.3-27) and (7.3-31) into Equations (7.3-30), it is possible to
show after some manipulation that term 2 reduces to
203
term2 =
2
1 [Pd − 2]
− 1
2 2P − P 2
d
d
(7.3-32)
~
Consequently, although the amplitude sequence A ′ sj is somewhat artificial, it does allow
Equations (7.3-8) and (7.3-9) to be reduced to simple closed form expressions while still
obeying the requirements that the marginal density functions are Nakagami-m
distributed and the correlations time is greater than that of an independent amplitude
sequence.
By substituting these reduced expressions for terms 1 and 2 back into Equation (7.3-7), the
mean acquisition time becomes
2
1 [P − 2]
n + N C d
− 1
2 2P − P 2
n =1
d
d
NC
T ACQ =
Td
NC
∑
=
Td
NC
N (N + 1) N 2
C C
+ C
2
2
=
Td N C Td [Pd − 2]
+
2
2 2 Pd − Pd2
[P − 2]2
d
− 1
2 Pd − Pd2
(7.3-33)
2
It is relatively straightforward to prove that this new mean acquisition time is greater
than or equal to the mean acquisition time obtained by assuming an independent
amplitude sequence (ie. T ACQ from Equation (7.3-33) ≥ T ACQ from Equation (7.3-15)). This
implies that as the rate of the amplitude scintillations decreases, the mean time to acquire
increases, at least for the amplitude model used in this analysis. However, this effect is
only likely to be important when the time taken to re-visit the correct cell, Tr , is relatively
small. This will generally be the case when the receiver is re-acquiring following a short
outage and has a good knowledge of the code phase and carrier frequency. When Tr is
large, such as during a cold start, the probability that a typical amplitude scintillation
sequence will be correlated between successive re-visits is likely to be quite small.
204
7.3.3. Independence
The analysis given in the previous section was based on the assumption that if the
amplitude scintillation sequence was sampled at a sufficiently low rate, successive
samples would be independent. This section provides a justification for this assumption.
As discussed in Appendix A, the majority of the amplitude scintillation energy is created
by irregularities of the order of the first Fresnel zone radius. For GPS, this is
approximately 300m or so depending of the height and elevation angle of the
irregularities. Irregularities much larger or much smaller than this produce negligible
diffraction energy, although large, extremely dense irregularities may produce refractive
fading effects. However, generally it can be assumed that amplitude scintillation patterns
are produced by the composite effect of numerous irregularities of the order of the first
Fresnel zone radius. The amplitude diffraction pattern produced by an isolated
irregularity has the following general form
zF
2
where z F is the first Fresnel zone radius. It is clear from this diagram that the pattern
decays very rapidly beyond z F 2 , and that the spacing between the peaks is independent
of the irregularity size (as z F is not a function of irregularity size). Consequently, if the
amplitude scintillation waveform produced by an isolated irregularity is sampled with a
spacing in excess of z F , it is likely that only one sample will contain a significant amount
of energy from the scintillation pattern. Adjacent irregularities will produce similar
patterns, and superposition can be used to determine the combined effects of these
irregularities (ie. a composite pattern can be obtained). As the sizes and locations of the
irregularities within a larger plume structure can be assumed to be random, it is
reasonable to assume that samples of the pattern taken more than z F m apart will be
approximately independent. If it is then assumed that the irregularities are part of a frozen
flow, and that motion of the frozen flow causes the pattern to move past the receiver with
a relative velocity of ve m/s, then samples taken more than z F v e seconds apart can be
205
assumed to be independent. For example, for v e = 100 m/s, z F = 275 m, and τ d = 20 ms
(typical values), then samples taken more than 2.75s apart ( N C > 125 cells) will be
approximately independent.
7.3.4. False alarms
If the amplitude values are assumed to be independent between successive re-visits (ie.
the amplitude correlation time is relatively short), T ACQ can be calculated as the sum of
two parts; one associated with detecting the correct cell assuming no false alarms, the
other associated with the false alarm verification time (see for example [84]).
Consequently, under these conditions T ACQ becomes (based on Equations (7.3-1) and
(7.3-15))
2 − Pd
T ACQ = N C Td K v Pfa + 1
2 Pd
(
)
(7.3-34)
However, if the amplitude correlation time is long enough for the amplitude sequence to
be correlated between re-visits, the additional verification time associated with false
alarms may shift the sequence more towards an independent sequence by increasing the
average re-visit time. This will tend to reduce T ACQ somewhat. In this section, it will be
demonstrated that the probability that the amplitude sequence statistics are significantly
affected by false alarms is extremely small, even for quite large verification times. This
assumes that scintillations do not greatly affect the false alarm rate, even when sidelobe
enhancements are considered (this is justified in Section 7.2.2).
If we let:
K v = false alarm verification time factor,
Tv = K v Td = false alarm verification time,
q = the number of false alarms between successive re-visits (ie. in N C cells),
N fa = the number of cells in which false alarms can occur,
then,
N C Td is the time taken for one unsuccessful pass of the uncertainty region with
no false alarms,
N fa = N C − 1 as one cell out of N C in the uncertainty region is the correct cell,
T p = qK v Td is the false alarm penalty time for one pass,
206
Tr = N C Td + T p is the total re-visit time,
N − 1
P q N fa = C P fa q 1 − Pfa
q
(
)
(
)NC −1−q ,
N C −1
E {q N fa }= ∑ q ∗ P ( q N fa )
,
q =0
= (N C − 1)Pfa
{
}
E T p N fa = (N C − 1)Pfa ∗ K v Td
Consequently, the average false alarm penalty time for one pass of the uncertainty region
is T p = (N C − 1)Pfa ∗ K v Td seconds. As a proportion of the total time, this is
(N C − 1)Pfa ∗ K v Td
N C Td
=
(N C − 1)K v Pfa
NC
(7.3-35)
≈ K v Pfa
Thus, if K v Pfa << 1 , false alarms will have a negligible impact on both the average re-visit
time and the statistics of the amplitude sequence. For example, if Pfa = 0.01% (a typical
level), K v << 10,000 is required. In general, this condition is expected to be met, although
the actual value of K v will depend on the verification methodology. For example, rechecking the cell once will result in K v = 1 , whereas attempting to establish carrier lock
immediately will probably result in K v >> 1 (Campanile [20] suggests that K v >> 10 ).
If K v Pfa << 1 , false alarms can be ignored and Equation (7.3-33) can be used to give the
mean acquisition time for the amplitude model described by Equation (7.3-19). However,
if K v Pfa << 1 is violated, the additional verification time will tend to increase T ACQ
somewhat, while the increase in the degree of independence of the amplitude sequence
will tend to reduce T ACQ slightly. As the two effects are linked, it is difficult to account for
them correctly.
7.4. Conclusions
Scintillations increase acquisition times by reducing the probability of detection while
searching the code phase / carrier Doppler uncertainty region. For satellite signals which
have a relatively low signal to noise ratio, the mean time to acquire may increase by a
207
factor of two or more, and the RMS acquisition time by a factor of three, depending on the
characteristics of the detector and the precise value of the carrier to noise density ratio,
C N o . Amplitude scintillations have by far the greatest impact on detection probabilities.
For example, the probability of detection can drop from almost one to approximately 0.8
at S 4 = 1 for a C N o of 36 dBHz (this represents an equivalent C N o of only 31 dBHz
under quiescent signal conditions). It was also found that for a slowly varying amplitude
waveform (ie. slow in relation to the time required to search the uncertainty region), the
increase in the mean time to acquire may be larger than for a rapidly varying amplitude
waveform.
Phase scintillations, on the other hand, were found to have virtually no effect on detection
probabilities for the square-law detector studied in this chapter. Similarly, correlation
sidelobes produced by competing satellite signals or by an incorrectly aligned signal were
found to have a negligible impact on false alarm probabilities, even under severe
scintillation conditions.
This chapter examined the effects of scintillations on the time to locate the correct code
phase and carrier Doppler for a square-law acquisition detector. However, the problems
that may be encountered when attempting to transition to a state of code and carrier
tracking under scintillation conditions have not been addressed. From earlier chapters, it
seems likely that under very intense scintillation conditions, it may not be possible to
achieve code and carrier lock, even though the correct code phase and carrier Doppler
may be known. Consequently, the acquisition time may effectively be extended to include
the time required for the scintillation patch to pass, or at least for the level of scintillation
activity to drop to the point where carrier tracking can begin.
208
Chapter 8
Optimum tracking of the carrier phase
The objective of this Chapter is to determine an optimum phase locked loop for the GPS
carrier that provides minimum phase tracking error under the specified ionospheric
scintillation and dynamic conditions. In Section 8.1, this is achieved by finding an
optimum filter using Wiener filter theory, and then mapping this filter into the structure
of an equivalent phase locked loop. Although in practice it is unlikely that GPS receivers
will be designed with scintillations in mind, this exercise nevertheless gives some insight
into the benefits that may arise from adopting an optimum loop configuration, and the
sensitivity of this optimum to receiver dynamics (normally, the characteristics of a
tracking loop are based solely on the dynamics and nominal signal to noise ratios). In
Section 8.2, the optimum bandwidth for minimum mean square tracking error is
determined directly for each loop order. Although this approach is not as generic as the
Wiener filter approach (ie. it assumes that the filter is a phase locked loop), it nevertheless
allows the optimum to be determined for arbitrary values of the spectral index, p.
8.1. Wiener filter approach
If the GPS signal is modelled as the sum of a phase process, φ (t ) , and thermal noise, w(t ) ,
which are assumed to be jointly stationary, the Wiener filter is the filter that minimises
[
]
2
the mean-square carrier phase tracking error, E φ ( t ) − φ#( t ) . Figure 8.1-1 is an
illustration of the relationship between the Wiener filter (represented by the dotted box)
and the equivalent phase locked loop filter, F (s) .
209
φ (t) + w(t)
+
φˆ(t)
F(s) s
+
-
Figure 8.1-1: Wiener filter model of a phase locked loop.
8.1.1. Causal Wiener filters
As the carrier tracking loop is causal (ie. future phase measurements are unavailable), the
Wiener filter must also be causal. The causal Wiener filter for the tracking loop illustrated
in Figure 8.1-1 is given by (see for example Brown & Hwang [19], Van Trees [95])
H o (s ) =
1
Sφ + w,φ (s )
−
Sφ + w (s )
pos
[Sφ +w (s )]+ [
]
(8.1-1)
where Sφ + w (s ) is the power spectral density of the combined signal φ (t ) + w(t ) , Sφ + w,φ (s )
is the cross-spectral density of φ (t ) + w(t ) with φ (t ) , + and - denotes all poles and zeros in
the left and right half of the complex s plane respectively (ie. X (s ) = X + (s ) X − (s ) where
X + (s ) = X − (s )* ), and
[ ]pos
denotes the transform of a positive-time function f (t ) such
that f (t ) = 0 for t < 0 . If it is assumed that φ (t ) and w(t ) are uncorrelated, then
Sφ + w (s ) = Sφ (s ) + S w (s ) and Sφ + w,φ (s ) = Sφ (s ) . The Wiener filter then becomes
H o (s ) =
1
Sφ (s )
−
Sφ (s ) + S w (s )
pos
[Sφ (s ) + S w (s )]+ [
]
(8.1-2)
When w(t ) is white with a power spectral density of βN o (where β is a constant), the
Wiener filter reduces to (based on the result by Yovits and Jackson [104])
H o (s ) = 1 −
βN o
[Sφ (s ) + βN o ]+
The transfer function of the optimum loop filter is then
210
(8.1-3)
H o (s )
Fo (s ) = s
1 − H o (s )
[
Sφ (s ) + βN o
= s
βN o
]+ − 1
(8.1-4)
Initially, it is assumed that the input phase process is produced entirely by phase
(
scintillations with a power spectral density of Sφ p ( f ) = T. f o 2 + f 2
)
−p 2
(ie. dynamics and
other sources of phase noise are ignored for the moment). This is representative of a
situation in which the receiver is either stationary or INS aided and phase scintillations
are the dominant source of phase noise. It is also assumed that the thermal noise term,
w(t ) , is zero-mean and white, but with a power spectral density that is scaled by the
effects of amplitude scintillations. A justification for this assumption is given below:
If it is assumed that the filter is a phase locked loop with an imperfect AGC, then the
thermal noise at the loop input can be represented by
w(t ) =
n (t )
g (t )
(8.1-5)
where n (t ) is zero-mean, white Gaussian thermal noise at the loop input (ie. on the IF
signal from Equation (3.1-1)), and g (t ) is the AGC gain factor from Section 3.2 which has
~
been translated back to the IF (ie. g (t ) = A(t )2 + 2 N o T > 0 for an imperfect AGC – see
Equation (3.3-16)). If n (t ) is assumed to be independent of g (t ) 1, the mean value of w(t ) is
given by
n (t )
E {w(t )}= E
g (t )
1
= E {n (t )}E
g (t )
(8.1-6)
= 0.
Therefore, w(t ) is zero-mean.
1
~
This implies that the amplitude fades are not too deep (ie. A(t ) does not become too small), or the AGC
time constant is relatively large.
211
The autocorrelation function of w(t ) is
Rw (τ ) = E {w(t ) w(t + τ )}
n (t ) n (t + τ )
= E
g (t ) g (t + τ )
1
= E {n (t ) n (t + τ )}E
g (t ) g (t + τ )
(8.1-7)
1
= N oδ (τ )E
.
g (t )
= N oδ (τ )β
{
}
where N o = E n (t )2 is the power spectral density of the thermal noise on the IF, δ (τ ) is
the Dirac Delta function and β is a constant. The power spectral density of w(t ) is thus
S w ( f ) = F {N oδ (τ )β }
(8.1-8)
= βN o
where F { } denotes the Fourier Transform. This result demonstrates that w(t ) is white,
~
although it is not necessarily Gaussian. For an ideal AGC (ie. g = A(t )2 ), the power
spectral density of w(t ) reduces to
1
Sw ( f ) = No E ~ 2
A(t )
1
=
(
2 C N o 1 − S4 2
(8.1-9)
).
In Section 3.2, it was shown that this result is only likely to be accurate for S 4 less than
about 1
2 . As n (t ) is assumed to be independent of g (t ) and φ p (t ) , it is straightforward
to show that w(t ) is uncorrelated with φ p (t ) , viz
n (t )
E φ p (t ) w(t ) = E φ p (t )
g (t )
{
}
φ p (t )
= E {n (t )}E
g (t )
= 0.
212
(8.1-10)
However, as A(t ) and φ p (t ) are produced by the same ionospheric processes, it cannot
also be said that w(t ) and φ p (t ) are independent. Nevertheless, these results are sufficient
to make use of Equation (8.1-3) to find an optimum Wiener filter.
The power spectral density of w(t ) and φ p (t ) can be represented in terms of the complex
frequency variable s = j 2πf as follows
S w (s ) = βN o
Sφ p (s ) =
(ω
(2π )p T
2
o
−s
(8.1-11)
)
2 p 2
where ω o = 2πf o is the outer scale size angular frequency. The causal Wiener filter is then
H o (s ) = 1 −
[S
φp
= 1−
βN o
(s ) + βN o ]+
(
(2π ) p T βN + ω 2 − s 2
o
o
p 2
ωo2 − s 2
(
= 1−
(8.1-12)
1
)
)
p 2 +
1
X (s )
The denominator of X (s ) can be separated into two factors which represent repeated
poles at s = ±ω o . These are (ω o − s)
p 2
and (ω o + s ) p 2 . However, the numerator is more
difficult to factorise. The zeros can be found by solving
(2π )p T
(
βN o + ω o 2 − s 2
)
p 2
=0
to give
T
s = ± ω o 2 − (2π )2
βN o
2 p
. exp(2πj (1 + 2n ) p )
(8.1-13)
where n is an integer. Consequently, depending on the value of p, there are potentially an
infinite number of zeros and so no unique solution to the factorisation problem. When m
zeros are present, the Wiener filter can be represented by
213
H o (s ) = 1 −
1
m
m
(zi + s )
(zi − s )
i =1
⋅ i =1
(ω + s ) p 2 (ω − s ) p 2
o
o
∏
= 1−
∏
+
(8.1-14)
(ω o + s )p 2
m
∏ (zi + s )
i =1
where z i are the zeros which are given by Equation (8.1-13). In order to proceed, two
integer values are chosen for the spectral index, p. These are p=2 and p=4.
Case 1: p = 2
For p = 2, the Wiener filter simplifies considerably as shown below
H o (s ) = 1 −
=
(ω o + s )
(z + s )
z − ωo
z+s
(8.1-15)
where z = ω o 2 + (2π )2 T βN o . The loop filter then becomes
s
Fo (s ) =
(z − ω o )
s + ωo
(8.1-16)
Consequently, in the limit as ω o approaches zero (ie. for an infinitely large ionospheric
outer scale size), the Wiener filter approaches a first order loop with a loop natural
frequency of
ωn o =
lim
F (s )
ωo → 0 o
T
.
= 2π
βN o
(8.1-17)
As the spectral index, p, at equatorial latitudes is typically equal to 2.5 [82], a first order
loop gives a good approximation to the optimum tracking loop at these latitudes. The
corresponding phase error variance is given by (from Equation (3.2-1))
214
∫ [1 − H o ( f )
∞
2
σ φε =
2
]
Sφ p ( f ) + H o ( f ) S w ( f ) .df
2
−∞
(8.1-18)
A simplified version of this expression for the case where w(t ) is white is (Van Trees [95])
n
σφε 2 = βN o ∑ ( z i − pi )
(8.1-19)
i =1
where pi and z i are the poles and zeros of Sφ p (s ) + βN o . For p=2, this becomes
σφε 2 = βN o ( z − ω o )
(8.1-20)
Therefore, in the limit as ω o approaches zero, the phase error variance becomes
lim
σ 2 = βN o ω n
ω o → o φε
o
(8.1-21)
= 2π TβN o
Consequently, the variance increases equally with both the phase scintillation energy, T,
and the amplitude scintillation energy, β . However, as would be expected, the optimum
loop bandwidth increases with the phase scintillation energy, but decreases with the
amplitude scintillation energy (the optimum loop bandwidth is proportional to ω n o ).
Case 2: p=4
For p=4, the Wiener filter is given by
(ω o + s )2
(z1 + s )(z 2 + s )
z z − ω o 2 + s[z1 + z 2 − 2ω o ]
= 1 2
z1 z 2 + s[z1 + z 2 ] + s 2
H o (s ) = 1 −
(8.1-22)
where z1 and z 2 are the zeros which are given by (from Equation (8.1-13))
z1 , z 2 = ω o 2 B j (2π )2
T
βN o
(8.1-23)
215
If
we
let
(2π )2
θ = atan
ω 2
o
ω o 2 B j (2π )2
T
βN o
T
= k exp(B jθ ) ,
βN o
k = ω o 4 + (2π )4
where
T
βN o
and
, the zeros become
z1 , z 2 = k exp(B j θ 2)
[
]
= k cos(θ 2) B j sin (θ 2)
[
= k 2 1 + cos(θ ) B j 1 − cos(θ )
= k + ωo 2 B j k − ωo 2
]
(8.1-24)
2
By substituting Equation (8.1-24) into (8.1-22), the Wiener filter becomes
(
)
k − ω o 2 + s 2 k + ω o 2 − 2ω o
H o (s ) =
k + s 2 k + ωo 2 + s 2
(
)
(8.1-25)
In the limit as ω o approaches zero, this reduces to
H o (s ) =
where ω n o = 2π 4
2ω n s + ω n 2
(8.1-26)
s 2 + 2ω n s + ω n 2
T
. In this form, the Wiener filter represents an active second order
βN o
loop with a damping factor of ζ = 1
2 and a loop natural frequency of ω n (see Table 3-
2). As 4 is the upper limit for the spectral index parameter [27], this result suggests that a
third order loop will not provide an optimum solution unless dynamics are also present.
The corresponding phase error variance is given by (from Equation (8.1-19))
[
]
σ φε 2 = βN o ( z1 − ω o ) + ( z 2 − ω o )
(
)
= βN o 2 k + ω o 2 − 2ω o
(8.1-27)
In the limit as ω o approaches zero, this simplifies to
lim
ωo → 0
σ φε 2 = 2 βN o ω n
o
= 2π 4 4T(βN o )3
216
(8.1-28)
Consequently, for p=4 the variance is far more sensitive to amplitude scintillation energy,
β , than to phase scintillation energy, T.
8.1.2. Non-causal Wiener filters
Although the non-causal Wiener filter is not a practical filter structure for a phase locked
loop, it does provide a lower bound on the phase tracking error which cannot be
surpassed by any filter type (Van Trees [95]). As φ p (t ) and w(t ) are uncorrelated, the
non-causal Wiener filter for arbitrary values of p is given by
H o (s ) =
=
Sφ p (s )
Sφ p (s ) + S w (s )
(8.1-29)
(2π )p T
(2π )p T + βN o (ω o 2 − s 2 )
p 2
and the corresponding phase error variance is
∫ [1 − H o ( f )
∞
2
σ φε =
]
Sφ p ( f ) + H o ( f ) S w ( f ) .df
2
−∞
Sφ p ( f )S w ( f )
∞
=
2
∫ Sφ p ( f ) + S w ( f ).df
(8.1-30)
−∞
∞
=
TβN
∫ T + βN (f 2 +o f 2 )p 2 .df
−∞
o
o
In the limit as f o approaches zero, this expression reduces to
lim
fo → 0
σ φε
2
2πβN o T
=
p sin(π p ) βN o
1 p
,
p >1
(8.1-31)
Although a closed form expression for the transfer function of an optimum causal filter
cannot be obtained for arbitrary values of p, it is possible to determine the variance for
arbitrary p without the need to factor the input spectrum. This given by (Van Trees [95])
217
Sφ p ( f )
ln 1 +
.df
βN o
−∞
∞
2
σ φε = βN o
∫
T
ln 1 +
βN f 2 + f 2
−∞
o o
(8.1-32)
∞
= βN o
∫
(
)
p
.df
2
2
Phase error variance (rad )
0.01
0.008
0.006
0.004
0.002
0
1
1.5
2
2.5
3
3.5
4
Spectral Index p
Figure 8.1-2: Phase error variance as a function of the spectral index p for T = − 25 dBW / Hz ,
S 4 = 0 , C N o = 415
. dBHz and f o ≈ 0 . The lower line represents the non-causal Wiener filter,
the upper line represents the causal Wiener filter. The two circles correspond to the variance
values obtained from Equations (8.1-21) and (8.1-28) for p=2 and p=4 respectively.
A comparison between these two variance measures is given in Figure 8.1-2 for
T= − 25 dBW / Hz , S 4 = 0 (ie. β = 1 A 2 ), C N o =415
. dBHz and f o ≈ 0 . It is clear from these
plots that the errors associated with the causal Wiener filter are always larger than those
associated with the non-causal filter (as would be expected).
8.1.3. Doppler errors
The optimum filters obtained in the previous sections were based on the assumption that
the tracking loops are only subject to scintillations and thermal noise. However, the
tracking loops of a real GPS receiver will also encounter Doppler errors resulting from
relative motion between the satellite and the receiver. An optimum filter which takes
account of Doppler errors can be obtained by adding a Doppler term to the power
spectral density of the input phase process, Sφ (s ) .
218
In the approach taken by Jaffe and Rechtin [45], optimum loop filters were obtained for a
phase locked loop that was subject to thermal noise and dynamics consisting of a step in
position, velocity and acceleration. This resulted in 1st, 2nd, and 3rd order tracking loops
respectively. An equivalent approach can be taken here by adding the power spectral
densities of a step in position, velocity and acceleration to the phase scintillation terms.
From Appendix E, the power spectral density of the dynamics, Sφ d (s ) , is given by
{
Sφd (s ) = E Φ d (s ) × Φ d (s )∗
=−
=+
=−
Θ2
,
position step
,
velocity step
,
acceleration step
s2
Ω2
s
4
Λ2
s6
}
(8.1-33)
where Θ, Ω and Λ are the magnitudes of the dynamic processes in radians, radians/s and
radians/s2 respectively (note that these can be related to the quantities given in Appendix
E through Θ = 2π ro λ , Ω = 2πvo λ , and Λ = 2π a o λ , where λ is the carrier wavelength).
Equation (8.1-33) can be generalised as follows
Sφd (s ) = (− 1)n
Γ2
(8.1-34)
s 2n
where Γ is either Θ, Ω or Λ, and n is the order of the dynamics (1, 2 or 3 for position,
velocity and acceleration respectively). The total power spectral density of the input
phase process is therefore
Sφ (s ) = Sφ p (s ) + Sφd (s )
=
(ω
(2π )p T
2
o
− s2
+ (− 1)n
)
p 2
Γ2
(8.1-35)
s 2n
Assuming ω o = 0 and letting s = j 2πf gives
Sφ ( f ) =
T
f
p
+
Γ2
(2πf )2n
(8.1-36)
The corresponding closed loop transfer function is therefore (from Equation (8.1-3))
219
Ho ( f ) = 1 −
βN o
(8.1-37)
[Sφ ( f ) + βN o ]+
In the analysis that follows, four cases are considered corresponding to p=2 and p=4, and
steps in both position and velocity (other values of p result in filters that do not map into
phase locked loop structures).
Case 1: Position step (n=1), p=2
For a position step with p=2, Sφ ( f ) becomes
Sφ ( f ) =
T
f2
+
Θ2
(8.1-38)
(2πf )2
Consequently, the optimum tracking loop is 1st order with a loop natural frequency of
ωn
o
= 2π
T + (Θ 2π )2
βN o
(8.1-39)
Note that this reverts to Equation (8.1-17) when dynamics are absent (ie. Θ=0).
Case 2: Velocity step (n=2), p=2
For a velocity step with p=2, Sφ ( f ) becomes
Sφ ( f ) =
T
f2
+
Ω2
(8.1-40)
(2πf )4
Consequently,
[Sφ ( f ) + βN o ]
+
T
Ω2
= 2 +
+ βN o
4
(2πf )
f
+
f 4 + Tf 2 βN o + Ω 2
=
f 4 βN o
[(2π ) βN ]
4
+
o
( j 2πf )2 + ( j 2πf )A + B (− j 2πf )2 + (− j 2πf )A + B
=
2
2
(
)
(
)
C
j
f
C
j
f
2
2
π
π
−
=
220
( j 2πf )2 + ( j 2πf )A + B
C ( j 2πf )2
(8.1-41)
+
where B =
Ω
, A = D + 2B , C =
βN o
(2π )2 T . The optimum closed loop filter
1
, D=
βN o
βN o
is therefore
Ho ( f ) = 1 −
βN o
( j 2πf ) + ( j 2πf )A + B
C ( j 2πf )2
2
=
( j 2πf )A + B
( j 2πf )2 + ( j 2πf )A + B
=
( j 2πf )2ζω n + ω n2
( j 2πf )2 + ( j 2πf )2ζω n + ω n2
(8.1-42)
which is the transfer function of a 2nd order loop with a loop natural frequency, ω n , and
damping factor, ζ, given by
ωn
o
ζ =
= B = 2π
1
A
=
2 B
2
4
Ω 2 (2π )4
βN o
(2π )
2
T
2Ω βN o
(8.1-43)
+1
Consequently, the loop natural frequency (and thus bandwidth) depends only on the
magnitude of the dynamics and the amplitude scintillations. However, the damping
factor is greater than the normal critical damping factor of 1
2 by a factor which
depends on all three effects (ie. dynamics and both the amplitude and phase scintillation
characteristics).
Case 3: Position step (n=1), p=4
For a position step with p=4, Sφ ( f ) becomes
Sφ ( f ) =
T
f4
+
Θ2
(2πf )2
(8.1-44)
Consequently, by inspection from Case 2 it is clear that the optimum tracking loop is 2nd
order with
221
ω n o = 2π
ζ =
1
2
4
T
βN o
Θ2
+1
8π 2 TβN o
(8.1-45)
Consequently, the loop natural frequency depends only on the magnitude of the
scintillation activity, and the damping factor depends on all three effects.
Case 4: Velocity step (n=2), p=4
For a velocity step with p=4, Sφ ( f ) becomes
Sφ ( f ) =
T
f4
+
Ω2
(2πf )4
(8.1-46)
And the optimum tracking loop is a 2nd order loop with
ωn
o
ζ =
= 2π
4
T + Ω 2 (2π )4
βN o
1
(8.1-47)
2
Consequently, phase scintillations and dynamics affect the loop bandwidth equally.
Although this analysis has not been continued for higher order dynamics, it appears that
if the order of the dynamics is expected to be large (ie. 2n >> p ), the order of the optimum
tracking loop will be determined solely by the dynamics. Indeed, at equatorial latitudes
where p ≈ 2 , it appears that the strength of phase scintillations will only affect the
damping factor, ζ, for a velocity step. The loop order will be determined by the dynamics,
and the loop bandwidth will be a function of both the dynamics and the strength of
amplitude scintillations.
8.1.4. Optimum post-loop filters
The optimum loop filters discussed so far have been designed to minimise the phase
errors in the carrier tracking loop. To reduce the effects of scintillations on the phase
estimates without compromising this first design objective, a second filter can be placed
in cascade with the tracking loop. This can be done in one of two ways as shown in
Figure 8.1-3.
222
+
φ (t) + w(t)
-
G2(s)
φˆ(t)
G1(s)
φˆ(t)
F(s)
+
1s
Figure 8.1-3: Post-loop filtering schemes to reduce scintillation phase noise in the loop phase
estimates.
The transfer functions of the resulting cascaded systems are
Case 1:
K1 (s ) = s.H o (s ).G1 (s ) , and
(8.1-48)
Case 2:
K 2 (s ) = [1 − H o (s )].G2 (s ) .
(8.1-49)
The causal Wiener filter for either system is given by (from Equation (8.1-2))
K oC (s ) =
Sφd (s )
Sφ (s ) + βN o
1
[Sφ (s ) + βN o ]+ [
−
pos
]
(8.1-50)
where Sφd (s ) is the power spectral density of the dynamics (ie. the desired signal for a
GPS receiver) and Sφ (s ) = Sφ d (s ) + Sφ p (s ) + Sφ o (s ) is the power spectral density of the
input phase process. Notice that Equation (8.1-3) cannot be used in this case as the noise
is no longer white (ie. it is of the form w(t ) + φ p (t ) + φo (t ) ). The Wiener filter represented by
Equation (8.1-50) is designed to minimise the errors from all noise sources while
providing a best estimate of the dynamics φd (t ) . By substituting Equation (8.1-50) into
Equations (8.1-48) and (8.1-49), the following optimum post-loop filters can be obtained
for a causal system
Case 1:
Case 2:
G1o (s ) =
[
1
s Sφ (s ) + βN o
G2 o (s ) =
]
+
− βN o
Sφd (s )
Sφ (s ) + βN o
][
Sφd (s )
βN o Sφ (s ) + βN o
1
[
.
−
pos
]
]
−
, and
pos
(8.1-51)
(8.1-52)
223
As the cascaded system does not necessarily need to be causal, equivalent non-causal
filters can also be found. The non-causal Wiener filter is given by
K oNC (s ) =
Sφ d (s )
(8.1-53)
Sφ (s ) + βN o
The non-causal filter applies to systems that are not required to provide range and
velocity estimates in real time and so may have access to future phase estimates.
Consequently, filter of this sort may be approximated arbitrarily closely by introducing a
processing delay. Notice that although the cascade represented by K oNC (s ) may be noncausal, the loop filter, H o (s ) , must always be causal. The two non-causal post-loop filters
are given by
G1o (s ) =
Case 1:
Sφ d (s )
[
G2 o (s ) =
Case 2:
[
s Sφ (s ) + βN o − βN o Sφ (s ) + βN o
[
Sφd (s )
βN o Sφ (s ) + βN o
]−
]− ]
, and
.
(8.1-54)
(8.1-55)
The error associated with the estimate of the Doppler process, φ d (t ) , is given by
φε d (t ) = φˆ(t ) − φ d (t )
(8.1-56)
The corresponding mean square error is therefore
{
E φε d (t )
where Sφε
d
(s ) = 1 − K o (s ) 2 Sφ d (s ) +
2
j∞
}
1
=
2πj
∫ Sφε d (s ).ds
(8.1-57)
− j∞
[
]
K o (s ) Sφ p (s ) + βN o is the power spectral density of
2
the Doppler estimate error. Using Equation (8.1-57), the mean-square error can be found
for the optimum loop filter, H o (s ) , the optimum causal cascaded filter, K oC (s ) , and the
optimum non-causal cascaded filter, K oNC (s ) . Although a comparison of these errors has
not been carrier out here, it is expected that the optimum non-causal cascaded filter will
produce the minimum error as it has access to both past and future information.
224
8.2. Direct determination of the MMSE
The principal advantage of the Wiener filter approach is that it allows the optimum filter
for minimum mean square error (MMSE) to be found without regard for the filter
structure. However, a drawback with this approach is that it results in a filter which does
not readily map into a phase locked loop unless the spectral index, p, is either 2 or 4. By
minimising the mean square error for each loop order directly, it is possible to determine
an optimum loop bandwidth and MMSE for all values of p. The optimum loop order can
then found by comparing the MMSE’s for each of the three loop orders and selecting the
minimum. In this section, an expression will be derived for the optimum loop bandwidth
for MMSE for all three loop orders in the presence of scintillations and thermal noise. It
will also be demonstrated that dynamics may strongly influence the choice of an
optimum loop bandwidth and order, and will in many cases take precedence over
scintillation effects.
The variance of the phase tracking error for a phase locked loop in the presence of
scintillations is given by (from Equation (3.2-5))
σ φ2 = σ φ2 + σ φ2
ε
εp
(8.2-1)
T
where (based on Equation (8.1-8))
σ φ2T = 2 Bn S w ( f )
(8.2-2)
= 2 Bn βN o
and (from Equation (3.2-8))
σ φ2εp =
kf np−1 sin
πT
,
([p − 1]π 2k )
1 < p < 2k
(8.2-3)
and (from Table 3-2)
πf n 2 ,
B n = 3πf n 2 2 ,
5πf 3 ,
n
= αf n
1st Order
2 nd Order
3 rd Order
(8.2-4)
Therefore, the derivative of σ φ2 with respect to the loop natural frequency, f n , is given by
ε
225
∂σ φ2ε
∂f n
Letting
∂σ φ2
ε
∂f n
=
2k )
+ 2αβN o
( p − 1)πT
=
2αβN o k sin([p − 1]π 2k )
=
=
o
(1 − p )πT
sin([p − 1]π
(8.2-5)
= 0 gives
fn o
Bn
kf np
ωn
1 p
(8.2-6)
o
2π
Bn o
α
is the optimum loop bandwidth for MMSE for all three loop orders. As shown in
Figure 8.2-1, Equation (8.2-5) is guaranteed to return the bandwidth for MMSE as the
variance of the tracking error consists of a monotonically increasing component due to
thermal noise added to a monotonically decreasing component due to phase
scintillations. It can also be shown that for a 1st order loop and p=2, or a 2nd order loop and
p=4, ω n o reduces to the two results given in Section 8.1 for the Wiener filter solution.
M.S.E.
Phase Scintillations
Amplitude Scintillations
& Thermal Noise
Bn
Figure 8.2-1: Illustration of the relationship between the two components of the mean square
tracking error as a function of the loop noise bandwidth, B n . The dotted line represents the sum,
σ φ2ε .
226
The corresponding MMSE can be found by substituting Equation (8.2-6) back into
Equation (8.2-1) to give
σ φ2ε
o
2αβN o p
( p − 1)πT
=
( p − 1) 2αβN o k sin([p − 1]π 2k )
1 p
(8.2-7)
8.2.1. Doppler errors
The effects of dynamics on the optimum loop bandwidth can be treated in one of two
ways; (i) assume that the dynamics are constant and the loop is in steady state, or (ii)
assume that the dynamics are introduced suddenly and produces a transient error in the
tracking loop. In the first case, the dynamics will either produce a constant error if the
order of the dynamics is equal to the order of the tracking loop, or zero error if the order
of the dynamics is less than the order of the tracking loop (see Appendix E, Section E.1).
In the second case, the transient error can be accounted for by introducing a term referred
to as the Total Transient Distortion, ε T2 . This is based on the approach taken by Jaffee and
Rechtin [45] that was discussed in Section 8.1.3.
Using Jaffee and Rechtin’s approach, the optimisation problem becomes one of
minimising the following variance expression
σ φ2ε = σ φ2εp + σ φ2T + ε T2
(8.2-8)
From Appendix E, the Total Transient Distortion is given by
∞
ε T2
∫ 1− H(f )
=
2
Sφ d ( f ).df
(8.2-9)
−∞
where Sφd ( f ) =
Γ2
(2πf )
2n
∞
ε T2 =
∫
−∞
=
from Section 8.1.3. ε T2 can be expanded to give
f 2k
f 2 k + f n 2k
×
Γ2
Γ2
(2πf )2n
.df
2k (2πf n )2 n −1 sin([2n − 1]π 2k )
(8.2-10)
,
0.5 < n < k + 0.5
227
The condition 0.5 < n < k + 0.5 implies that the loop order must be greater than or equal to
the order of the dynamics in order for ε T2 to be finite (eg. either a position or velocity step
for a 2nd order loop, but not an acceleration). In principle, the optimum loop bandwidth
for MMSE can once again be found by minimising Equation (8.2-8) with Equation (8.2-10)
used in place of ε T2 . However, in practice a simple analytical expression cannot readily be
obtained for all values of n, k and p as the derivative of Equation (8.2-8) is a polynomial
with a non-integer order. What can be done is to solve the polynomial using numerical
techniques for a specific set of conditions, or obtain an analytical solution for integer
values of p. Alternatively, it is possible to determine a value for the spectral strength, T,
above which the phase scintillation component dominates over the dynamics component.
This can be found by equating Equations (8.2-3) and (8.2-10) to give
T=
Γ 2 sin([p − 1]π 2k ) f n p−2n
(2π )2n sin([2n − 1]π
2k )
,
0.5 < n < k + 0.5 and 1 < p < 2k
(8.2-11)
In Figure 8.2-2, T (in decibels) is plotted as a function of Γ for a 2nd order loop that is
subject to both a position step (Γ=Θ) and a velocity step (Γ=Ω). In addition, two loop
bandwidths are considered that represent typical upper and lower values for a phase
0
−20
−40
0
−20
−40
0.02 0.04 0.06 0.08
Position Step (m)
0
0
0.5
1
Velocity Step (m/s)
0
2
T dBrad /Hz
T dBrad2/Hz
0
2
T dBrad /Hz
T dBrad2/Hz
locked loop.
−20
−40
−20
−40
0.02 0.04 0.06 0.08
Position Step (m)
0.2 0.4 0.6 0.8 1
Velocity Step (m/s)
Figure 8.2-2: Threshold values of T above which phase scintillation energy will dominate over
dynamics in the selection of an optimum loop bandwidth. The upper two panels represents a loop
bandwidth of 2Hz. The lower two panels represent a loop bandwidth of 15Hz. p=2.5.
228
Also note that because of the ambiguity in a sinusoid, it is only necessary to look at
instantaneous phase steps of less than ½ a cycle (~0.095m at the GPS L1 frequency). For
this reason, Θ is often modelled as a uniform random variable on (− π , π ) such that
Sφd ( f ) =
{ }=
E Θ2
π2 3
(2πf )2 (2πf )2
. The corresponding value of T can then be found from the two
left panels of Figure 8.2-2 by assuming a position step of π
3 radians or 0.055m at GPS
L1. Notice that this implies that T must be greater than about –10dBradians2/Hz in the
presence of a random phase step before it will begin to dominate over dynamics in the
selection of an optimum loop bandwidth.
8.3. Conclusions
Using Wiener filter theory, it can be shown that the optimum causal filter for minimum
carrier phase tracking error has the same structure as a first order phase locked loop
when the spectral index, p, is equal to two, and a second order loop when p = 4.
Therefore, as p is usually close to 2.5 at equatorial latitudes, it seems likely that a first
order loop will be the best choice for carrier tracking in an equatorial scintillation
environment. However, this result is predicated on the assumption that amplitude
scintillations can be treated as a scaling factor for the thermal noise power spectral
density, and that other direct phase processes such as dynamics are absent.
For p = 2 (close to the typical equatorial value) and for a step in position, the optimum
phase locked loop structure is a first order loop with a bandwidth that depends on both
the strength of scintillation activity (ie. T and S 4 ), and the magnitude of the phase step.
However, for p = 2 and a step in velocity, the optimum loop order is determined by the
dynamics (ie. it is second order) and the bandwidth is a function of the magnitude of the
velocity step and the strength of amplitude scintillations only. Although this analysis has
not been carried out for dynamics with higher orders, it appears that if the order of the
dynamics is large (ie. 2n >> p ), the order of the optimum tracking loop will be decided by
the dynamics.
Based on a direct determination of the optimum bandwidth for a second order phase
locked loop, it appears that the strength of phase scintillations must be extremely large
before it begins to dominate over dynamics, even when the magnitude of the dynamics is
quite modest. Consequently, in practice the choice of loop order and bandwidth should
229
be based on the dynamics, the quiescent GPS signal level and perhaps the anticipated
strength of amplitude scintillation activity. The level of phase scintillation activity is only
likely to be important when dynamics are greatly minimised through the use of inertial
aiding, or in the case of codeless and semi-codeless receivers, through L1 aiding of the L2
carrier loop.
230
Chapter 9
Scintillation effects on navigation
In this thesis, a number of receiver performance measures have been derived by
combining a stochastic ionospheric scintillation model with various receiver tracking loop
models. By linking these measures with a climatological scintillation model such as the
Wide Band Scintillation Model, WBMOD, the performance of individual satellite-receiver
links can be predicted for a given time and location. This is discussed further in Section
9.1. However, because WBMOD does not account for large scale structures such as
equatorial plumes which affect the distribution of irregularities in the sky, it is unable to
model the spatial and temporal patchiness of scintillations, nor the night-to-night
variability that is frequently observed on scintillating links. Consequently, it is not well
suited to predicting the performance of multiple satellite links simultaneously, and so
cannot easily be used to assess the impact of scintillations on navigational accuracy. This
is discussed in more detail in Sections 9.2 and 9.3. Nevertheless, in Section 9.4 it is shown
that WBMOD can be used to determine the number of links that may be stressed to the
point of losing lock for a given time, location and percentile. Although this does not
indicate the likelihood of simultaneous losses of lock, it does illustrate when and where
significant scintillation events are likely to occur for a given receiver type.
9.1. Predicting the performance of a single link
For any given satellite-receiver geometry, time and date, uncertainty will exist about the
size and density of the irregularities along a specified propagation path. In WBMOD, this
uncertainty is accounted for by providing a probability density function (PDF) for the
height integrated irregularity strength parameter1, Ck L . Based on Equations (3.4-9) and
(3.4-10), it is clear that T and S 4 can be related to Ck L through the expressions
1
Ck L is the height-integrated strength of the irregularity spatial power spectrum at a scale size of
1km (see Equation (2.1-2) and [82] and [76]).
231
T = α Ck L , and
(9.1-1)
S 42 = 1 − exp(− β Ck L )
where α and β are based on a combination of deterministic geometrical factors and
accurately modelled random parameters (see for example [82] and [76]). T and S 4 can
(
)
also be related to each other through the expression T = − ln 1 − S 42 γ where γ = β α
(Equation (3.4-7)). Therefore, for a given set of conditions (ie. time, date and satellitereceiver geometry), WBMOD’s internal models provide information from which the
parameters α, β , and γ and the distribution functions of T and S 4 , f T (T ) and f S 4 (S 4 ) ,
can be deduced. These functions can then be used to determine average values for many
of the receiver performance measures derived in earlier chapters. These include the
variance of the code and carrier phase range errors, the probability of losing lock, PL , the
probability of a navigation data bit error, Pe , and the probability of detection for
acquisition, Pd . Also, because T can be expressed as a function of S 4 and the geometry
factor γ, it is only necessary to determine the PDF of S 4 (or T) in order to find average
values for the specified performance measures (ie. the joint PDF of T and S 4 can be
(
)
expressed as f T, S 4 (T, S 4 ) = δ (T − T ′) f S 4 (S 4 ) where T′ = − ln 1 − S 42 γ ).
In WBMOD, the PDF of log(C k L ) in equatorial regions is modelled as the sum of two
Gaussians [82]. In principle, this allows Equation (9.1-1) to be used to determine the PDF
of S 4 for a given value of β . In practice, however, WBMOD does not provide this
information as part of its standard output. Nevertheless, it is possible to deduce the PDF
of S 4 directly by differentiating the cumulative distribution function of S 4 which can be
obtained from WBMOD’s predictions of S 4 over a range of different percentiles.
If the PDF of S 4 is known for a particular link, the average probability of losing lock can
be found as follows
∞ 1
PL =
∫ ∫ PL (T, S4 ) f T, S 4 (T, S 4 ).dS4 .dT
0 0
1
= PL (T ′, S 4 ) f S 4 (S 4 ).dS 4
∫
(9.1-2)
0
where PL (T, S 4 ) is the probability of losing lock as a function of T and S 4 (from Equation
232
(
)
(3.4-3) with m = 1 S 42 ), T′ = − ln 1 − S 42 γ , f S4 (S 4 ) is the PDF of S 4 , T ≥ 0 and 0 ≤ S 4 ≤ 1 .
In a similar way, the average values of Pe and Pd can be found for each visible satellite
link (ie. from Equation (6.3-11) and (6.3-15) we have Pe (S 4 ) , and from Equation (7.2-11)
we have Pd (S 4 ) ).
9.2. Predicting the performance of multiple links
In order to determine the impact of scintillations on navigational accuracy, it is necessary
to find the probability of losing lock on multiple satellite links simultaneously. The
average probability of losing lock on n satellite links simultaneously is given by
1
1 n
Pn = 4
PLk Tk′ , S 4k ∗ f S4 ,4S4 S 41 ,4 S 4n .dS 41 4dS 4n
n
1
k =1
0
0
∫ ∫∏ (
(
)
(
)
(9.2-1)
)
where PLk T′k , S 4k is the probability of losing lock on link k as a function of the S 4 index
on that link, and f S4
1 ,4S4n
(S41 ,4 S 4n ) is the joint PDF of
S 4 on the n links. Equation
(9.2-1) implies that for a given set of S 4 values, the individual probabilities of losing lock
are independent of one another. In other words, the probability of losing lock on n links
simultaneously is simply the product of the probabilities of losing lock on each link. This
is based on the observation that although the strength of scintillation activity may be
correlated between the links (perhaps as a result of a large plume structure that is
penetrated by several links simultaneously), the individual scintillation patterns, and in
particular the deep fades that give rise to loss-of-lock, are likely to be independent. The
justification for this assumption is that scintillation patterns are produced by small scale
irregularities of the order of the Fresnel zone radius or smaller (< 300m or so), and so it is
unlikely that two propagation paths will penetrate the same group of irregularities at the
same time. This assumption may, however, break down if the ionospheric pierce points of
the two propagation paths happen to be extremely close.
Although WBMOD does not provide information about the joint PDF of S 4 on multiple
satellite links, if it is assumed that the links are independent (ie. the irregularity regions
are assumed to be highly “patchy”), then the average probability of losing lock on n links
simultaneously is simply the product of the average probability of losing lock on each
link, viz
233
n
Pn =
∏ PLk
(9.2-2)
k =1
n
(S 41 ,4 S4n ) = ∏ f S4k (S 4k ) has been assumed. Under these conditions, the
1 ,4S4n
where f S4
k =1
probability of simultaneously losing lock on n links is expected to be very small, given
that the individual probabilities are also likely to be quite small. Measurements of loss-oflock taken from a Novatel Millennium receiver during the September 1998 and March
1999 equinoxes tends to support this view. In Figure 5.4-1 from Chapter 5, the percentage
of time between 8:00pm and 10:00pm that the Novatel Millennium loses lock on both
the L1 and semi-codeless L2 channels is plotted as a function of day. It is clear from these
plots that the percentage of time that one link was lost (represented by the white sections
of the bars) was always much greater than the percentage of time that two or more links
were lost simultaneously (represented by the solid sections). Indeed, on days during
which significant scintillation activity occurred, the average ratio of the percentage of
time that two or more links were lost simultaneously compared to only one link was 4.7%
for the semi-codeless L2 carrier loop. Again, this supports the view that the simultaneous
loss of multiple links becomes much less common as the number of links, n, increases.
9.3. Predicting navigational accuracy
Equation (9.2-1) gives the average probability of losing lock on n satellite-receiver links
based on the joint PDF of S 4 on those links. However, on its own this provides no
information about the navigational accuracy, nor the probability of a complete navigation
outage. In this section, an approach is outlined which addresses these problems by
assuming that the joint statistics of scintillation on multiple satellite-receiver links are
known.
Consider a situation in which any n of m visible links have lost lock, and let i index the
different ways in which this can occur. The probability of any one of these is denoted as
Pni and is given by (based on Equation (9.2-1))
1
1
n
Pni = 4
PLk Tk′ , S 4k
k =1
0
0
∫ ∫∏ (
234
) ∏ [1 − PLk (Tk′ , S 4k )] × f S41 ,4S4m (S41 ,4 S 4m ).dS41 4dS4m
m
k = n +1
(9.3-1)
If it is assumed that a navigation outage occurs when less than four satellites are tracking,
and a RAIM2 failure occurs when less than six are tracking, the probabilities of these two
events are given by
mj
P (Navigation outage) =
P(m− j )i
j =0 i =1
3
∑∑
mj
P (RAIM outage) =
P(m− j )i
j =0 i =1
5
(9.3-2)
∑∑
m
m!
where =
represents the number of different satellite-receiver combinations
j
(
m
j )! j !
−
for which j of the m visible satellites are tracking, and P(m− j ) are the probabilities
i
associated with each possible satellite-receiver combination (ie. the ith satellite
combination for which m-j satellite links have lost lock).
Unfortunately, Equation (9.3-2) cannot be evaluated as the joint PDF of S 4 is unknown.
However, if it is assumed that the probability of losing lock on each satellite link is the
same (and given by PL ), then
Pn = PL n (1 − PL )
m−n
(9.3-3)
where Pn = Pni as all combinations associated with the loss of n satellites now have the
same probability. In practice, this is an extremely unlikely situation given the
inhomogeneous nature of the ionosphere and the vastly different satellite-receiver
geometries on each link. Nevertheless, this assumption allows Equation (9.3-2) to be
simplified and evaluated for a given value of PL . It also illustrates the sensitivity of
Equation (9.3-2) to the number of satellites, m, and the single link probability of losing
lock, PL . Under this assumption, Equation (9.3-2) simplifies to
2
Receiver Autonomous Integrity Monitoring or RAIM is a technique whereby six or more satellite
pseudorange measurements are cross-checked to determine their integrity. As RAIM is performed
within a receiver, it eliminates the need for external integrity information. Although five satellites
are required for fault detection, an additional satellite is needed for both the detection and
exclusion of faults [47], page 313.
235
m (m− j )
j
(
1 − PL )
PL
j
j =0
3
P (Navigation outage) = ∑
(9.3-4)
m (m− j )
j
P (RAIM outage) =
(
1 − PL )
PL
j
j = 0
5
∑
In Figure 9.3-1, the probability of a navigation outage and a RAIM outage are given for
different values of PL and m based on Equation (9.3-4). Although these results must be
treated with caution, they do show that a small increase in m, perhaps as a result of
improved satellite visibility or the use of a supplementary navigation system such as
GLONASS, will greatly reduce the risk of navigation or RAIM outages. They also shows
that RAIM is far more vulnerable to failure than a loss of navigation.
0
0
−10
dB
dB
−10
m=6
m=6
−20
−20
−30
−30
0
0
−10
dB
dB
−10
m=8
m=8
−20
−20
−30
−30
0
0
−10
dB
dB
−10
m=10
m=10
−20
−20
−30
−30
0
0
−10
dB
dB
−10
m=12
−30
0
m=12
−20
−20
0.1
0.2
0.3
0.4
0.5
PL
0.6
0.7
0.8
0.9
1
−30
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
P
L
Figure 9.3-1: The probability of a navigation outage (left panel) and a RAIM outage (right panel)
as a function of both the single link probability of losing lock, PL , and the number of visible
satellites, m.
236
Although a navigation outage is very unlikely, it is still possible for scintillations to
reduce navigational accuracy by degrading the satellite-receiver geometry. For example,
if a receiver is located slightly to the North of the anomaly peak, it may lose several
satellites towards the South in the direction of the anomaly. Consequently, the satellites
available to the receiver will be skewed towards the North which will probably degrade
the horizontal accuracy of the receiver in a North-South direction.
In order to assign probabilities to different levels of navigational accuracy, it is necessary
to determine a DOP3 value for each satellite-receiver combination. If a probability is
assigned to each satellite combination, Pni , and a DOP value is assigned, DOPni , the
probability of the DOP exceeding some acceptable level, β , can be represented by
P (DOP > β ) =
∑ Pni (DOPni )
(9.3-5)
DOPn > β
i
However, as this expression cannot be evaluated without a knowledge of the joint
statistics of S 4 and T, it will not be considered any further.
3
The Dilution of Precision or DOP is a factor which converts the average pseudorange error into
equivalent navigation and time errors by taking into account the geometry of the satellite
constellation. DOP factors exist for horizontal position error, vertical position error, 3 dimensional
position error, time error, and position and time error.
237
9.4. Predictions based on WBMOD
Although WBMOD does not provide information about the PDF of S 4 and T, or the joint
statistics of scintillations on multiple satellite links, it does allows the scintillation stresses
on individual links to be predicted for a given time, location and percentile. This enables
areas to be identified within a broader region of interest in which significant scintillation
effects may occur for a given receiver type. It also allows factors such as the sunspot
number (SSN), the magnetic activity index ( K p ), the carrier loop bandwidth, the
quiescent signal level, and the elevation mask angle1 etc. to be varied and the resulting
impact on carrier loop performance to be examined. In Figure 9.4-1, the percentage of
links above an elevation angle of 10° that may be stressed to the point of losing lock is
plotted over an area that encompasses both the northern and southern anomalies in the
South East Asia / Australian region. These plots were obtained by passing WBMOD’s
predictions of the scintillation indices, S 4 and T, into the carrier loop model from Section
3.4. For each satellite-receiver link, the tracking state was then found by comparing the
predicted probability of losing lock, PL (obtained from S 4 and T), with a 1% threshold. In
this particular plot, the following parameter values were chosen:
Date:
23/09/2000 (the Sept. equinox nearest to the current solar maximum)
Time:
12:00 UTC (approximately 8:00pm local time at a longitude of 120°E)
Duration:
~ 20 minutes
Percentile:
90%
SSN:
135 (based on predictions for September 2000 obtained from the IPS2)
Kp :
4.3 (a moderate level of geomagnetic activity)
Mask angle: 10°
C No :
41.5 dBHz
Bn :
15 Hz
It is important to realise that this plot does not indicate the likelihood of simultaneous
losses of lock, merely the number of links that may be stressed to the point of losing lock
for a given percentile. Nevertheless, it could be said that if many links are stressed, and
1
The elevation mask angle is the satellite elevation angle below which receiver measurements are
ignored. It is primarily intended to reduce contamination of the navigation solution by multipath
and thermal noise.
2
IPS: the Australian Ionospheric Prediction Service.
238
the chosen percentile is relatively small, the probability of simultaneous losses of lock
should become much larger.
Figure 9.4-1: The percentage of links that may be stressed to the point of losing lock from
WBMOD. Parameter values are: 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=135,
K p =4⋅3, 10° elevation angle mask, C N o =41⋅5dBHz, Bn =15Hz, Loop order = 3, Coded L1 loop.
It is clear from Figure 9.4-1 that scintillations are only likely to be a problem between
6:00pm and 10:00pm local time3 and within the region of the anomaly. It is also apparent
that directly beneath the anomaly crest at approximately 120°E (~ 8:00pm local time), the
percentage of affected links increases to about 70%. However, care must be taken in
drawing too many conclusions from this result as it only applies to WBMOD predictions
at the 90th percentile and is based on the rather conservative threshold of 1% for PL .
Additional simulations (not shown) revealed that if the percentile was reduced to 65%,
the percentage of links affected by scintillations dropped to zero at all times and
locations. Therefore, based on WBMOD, it can be said that under the specified conditions,
the probability of losing lock on any satellite-receiver link in the region would be less
than 35%.
In Figure 9.4-2, the quiescent carrier to noise density ratio was raised to 44dBHz with all
other parameters left unchanged. By comparing this with Figure 9.4-1, it is clear that the
3
Because of Earth rotation, 90°E and 150°E represent 6:00pm and 10:00pm local time respectively
for a simulation time of 12:00 noon UTC.
239
signal to noise ratio has a significant influence on the tolerance of wide bandwidth
receivers to scintillations. From Figure 3.4-3, it is apparent that this is because wide
bandwidth receivers have a greater susceptibility to amplitude scintillations which are
strongly influenced by the quiescent signal level (essentially, an increase in C N o will
reduce the probability of the amplitude falling below the tracking threshold).
Figure 9.4-2: The percentage of links that may be stressed to the point of losing lock from
WBMOD. Parameter values are: 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=135,
K p =4⋅3, 10° elevation angle mask, C N o =44dBHz, Bn =15Hz, Loop order = 3, Coded L1 loop.
In Figure 9.4-3 and Figure 9.4-4, the predictions are repeated for a 2Hz bandwidth
receiver for both C N o = 41.5dBHz and 44dBHz respectively. It is clear from these figures
that variations in the carrier to noise density ratio have much less of an effect on narrow
bandwidth receivers than on wide bandwidth receivers. This suggests that the majority of
the predicted outages in these figures are due to phase scintillations which are not
influenced by the quiescent signal level. Again, this conclusion is consistent with
Figure 3.4-3 for a narrow bandwidth receiver.
In the four figures presented so far, the elevation mask angle was set to 10°. In Figure
9.4-5, the effect of reducing the mask angle to 0° is examined. By comparing Figure 9.4-5
with Figure 9.4-1, it is apparent that a reduced mask angle increases the extent of the
region affected by scintillations, but also reduces the impact of scintillations directly
beneath the anomaly peak. This second effect is a result of an increase in the number of
satellite-receiver links that are directed away from the anomaly peak when the additional
240
low elevation angle links are included (ie. directly beneath the anomaly peak, the low
elevation angle links all point away from the peak and will therefore be largely
unaffected by scintillations).
Figure 9.4-3: The percentage of links that may be stressed to the point of losing lock from
WBMOD. Parameter values are: 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=135,
K p =4⋅3, 10° elevation angle mask, C N o =41⋅5dBHz, Bn =2Hz, Loop order = 3, Coded L1 loop.
Figure 9.4-4: The percentage of links that may be stressed to the point of losing lock from
WBMOD. Parameter values are: 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=135,
K p =4⋅3, 10° elevation angle mask, C N o =44dBHz, Bn =2Hz, Loop order = 3, Coded L1 loop.
241
Figure 9.4-5: The percentage of links that may be stressed to the point of losing lock from
WBMOD. Parameter values are: 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=135,
K p =4⋅3, 0° elevation angle mask, C N o =41⋅5dBHz, Bn =15Hz, Loop order = 3, Coded L1 loop.
So far, all simulations have been conducted for the equinox of the 23rd September 2000.
Simulations conducted under the conditions outlined in Figure 9.4-1 and Figure 9.4-3, but
for days that were two months removed from the September and March equinoxes
revealed no evidence of scintillation effects at all, even at a reduced signal level
(30 dBHz). Also, simulations conducted on days that were one month removed from the
equinoxes showed that scintillation effects were significantly reduced. An example of this
is given in Figure 9.4-6 which is based on the 23rd October 2000 (all other parameters are
the same as Figure 9.4-1). Therefore, according to WBMOD, it appears that receivers are
unlikely to suffer any effects from scintillations beyond about one month from the
equinoxes, even near solar maximum. Obviously, more simulations would need to be
conducted in order to rigorously test this observation.
In Figure 9.4-7, the impact of reducing the sunspot number to 100 is given. Based on
current predictions, this level of solar activity is unlikely to be reached until about
October 2002. When the sunspot number was reduced to 70 (approximately August 2003),
the observed effects were negligible, and when reduced further to 50 (approximately
March 2004), no effects were observed at all under the specified conditions.
242
Figure 9.4-6: The percentage of links that may be stressed to the point of losing lock from
WBMOD. Parameter values are: 23 Oct. 2000, 12:00 noon UTC, 90th percentile, SSN=135,
K p =4⋅3, 10° elevation angle mask, C N o =41⋅5dBHz, Bn =15Hz, Loop order = 3, Coded L1 loop.
Figure 9.4-7: The percentage of links that may be stressed to the point of losing lock from
WBMOD. Parameter values are: 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=100,
K p =4⋅3, 10° elevation angle mask, C N o =41⋅5dBHz, Bn =15Hz, Loop order = 3, Coded L1 loop.
243
Figure 9.4-8: The percentage of links that may be stressed to the point of losing lock from
WBMOD. Parameter values are: 23 Sept. 2000, 12:00 noon UTC, 90th percentile, SSN=135,
K p =4⋅3, 10° elevation mask angle, C N o =41⋅5dBHz, Bn =0⋅2Hz, Loop order = 3, Semi-codeless
L2 tracking loop.
Figure 9.4-9: The percentage of links that may be stressed to the point of losing lock from
WBMOD. Parameter values are: 23 Sept. 2000, 12:00 noon UTC, 65th percentile, SSN=135,
K p =4⋅3, 10° elevation angle mask, C N o =41⋅5dBHz, Bn =0⋅2Hz, Loop order = 3, Semi-codeless
L2 tracking loop.
244
In Figure 9.4-8, the effects of scintillations on a semi-codeless tracking loop of the sort
discussed in Chapter 5 are given. By comparing this with Figure 9.4-1, it is clear that the
predicted effects are much more severe than for a full correlation L1 tracking loop.
Indeed, even at the 65th percentile, significant effects were observed on the semi-codeless
tracking loops when virtually no effects were observed on a full correlation tracking loop
(see Figure 9.4-9). Consequently, it would be expected that the probability of
simultaneously losing two or more satellites would be much greater for the semi-codeless
channels (the measurements reported in Section 5.3.1.1 for lower levels of solar activity
tend to support this view).
9.5. Conclusions
WBMOD provides information from which the PDF’s of the two principal scintillation
indices, S 4 and T, can be deduced. Using these functions, it is possible to determine the
average values of various receiver performance measures such as the variance of the code
and carrier phase range errors, the probability of losing lock, the probability of a
navigation data bit error, and the probability of detection for acquisition.
However, WBMOD is not well suited to predicting the performance of multiple channels
simultaneously, and so cannot be used to determine the overall impact of scintillations on
navigational accuracy. This is because it does not account for the large scale structures
such as equatorial plumes which tend to restrict scintillation activity to certain parts of
the sky. The statistics that are required from a scintillation model in order to determine
navigational accuracy are the joint PDF’s of S 4 and T on each of the propagation paths.
Although this information is not available at present, it is clear that if scintillations are
assumed to be independent between the individual propagation paths, the probability of
losing lock simultaneously on multiple channels would be expected to be quite small,
given that the individual probabilities are also quite small. Consequently, the probability
of a navigation outage or a loss or RAIM is likely to be very small, even under conditions
for which all links are affected by scintillations (ie. when the irregularities cover the entire
sky).
Predictions of the percentage of links that may be stressed to the point of losing lock
based on WBMOD clearly show that scintillation effects are mainly restricted to:
245
-
Solar maximum (or at least high sunspot numbers).
-
The equinox.
-
Approximately 6:00pm to 10:00pm local time.
-
The Northern and Southern anomalies.
However, even under these conditions, the probability of losing lock simultaneously on
multiple links would be very small.
246
Chapter 10
Summary
In Section 10.1, a brief overview of the thesis is given and the main results from each
chapter are summarised. In Section 10.2, conclusions are drawn about the overall
performance of GPS in a scintillation environment by drawing together the results from
all previous chapters. Finally, in Section 10.3, future research directions are examined
including areas in which the scintillation model can be improved, and the potential
impact of new developments in GPS on the susceptibility of GPS to scintillations.
10.1.
Overview
In this thesis, a stochastic model of scintillation activity was combined with various
receiver tracking and acquisition models to determine the likely impact of scintillations
on GPS. A summary of the key points to come out of this work is given below:
•
In general, the carrier tracking loops of full code correlation GPS receivers are quite
robust to scintillations, even under very strong scintillation conditions.
•
An optimum bandwidth exists for minimum probability of losing lock which depends
on the relative contributions of amplitude and phase scintillation activity, as well as the
quiescent signal level and the presence of dynamics.
•
For a given level of ionospheric disturbance, the geometry of the propagation path
affects the ratio of amplitude to phase scintillation activity as well as the absolute levels of
scintillation activity. Therefore, geometry will affect both the optimum bandwidth of a
tracking loop and its overall susceptibility to scintillations.
•
Carrier tracking loops are generally very robust to signal fades of short duration,
particularly if the bandwidth of the tracking loop is narrow.
•
RMS carrier phase errors of several centimetres can be introduced into satellite range
measurements by scintillations. These may have a significant impact on carrier phase
DGPS observations made in equatorial regions during solar maximum, particularly for
baselines of a kilometre or more.
247
•
Frequency locked loops are more robust to phase scintillations but slightly less robust
to amplitude scintillations than phase locked loops for the same loop bandwidth and predetection integration period. Therefore, receivers that make use of frequency locked
loops, either as a primary means of carrier tracking or as a fall-back strategy to phase
locked loops, are likely to be more tolerant to scintillations than receivers that employ
only phase locked loops.
•
Phase scintillations have a negligible effect on code tracking loops and the additional
thermal noise associated with amplitude scintillations is only small, unless S 4 is close to
unity. Nevertheless, under strong amplitude scintillation conditions, it is likely that error
spikes will exist in the code pseudorange measurements during times when the
amplitude is deeply faded.
•
Codeless tracking loops are far more susceptible to the effects of scintillations than full
code correlation tracking loops. The poor performance of codeless tracking loops may
result in a degradation in the accuracy of systems such as WAAS which rely on dual
frequency SPS receivers for the measurement of ionospheric delays.
•
Even under conditions for which the carrier loop is likely to lose lock, the probability
of a word error in the navigation data will only be a few percent. Therefore, because of
the high levels of redundancy that exist in the navigation data, it is expected that
scintillations will have negligible effect on a receivers ability to acquire the navigation
information.
•
Scintillations increase acquisition times by reducing the probability of detection. For
satellite signals which have a relatively low signal to noise ratio, the mean time to acquire
may increase by a factor of two or more, and the RMS acquisition time by a factor of
three, depending on the characteristics of the detector and the signal to noise ratio.
•
Wiener filter theory was used to determine the optimum structure of a phase locked
loop that is subject to both scintillations and dynamics. It was found that in the absence of
line of sight dynamics, the optimum loop order was determined by the slope of the phase
scintillation power spectrum. However, in the presence of dynamics, the order of the
dynamics would usually determine the optimum loop order. It was also found that the
magnitude of the dynamics and the strength of amplitude scintillation activity would
determine the optimum loop bandwidth, unless phase scintillation activity was very
strong.
•
WBMOD provides information from which the probability of occurrence of
scintillations at different levels can be determined for individual satellite-receiver
propagation paths. This information can be used to determine average values for many of
the receiver performance measures derived in earlier chapters. However, WBMOD is not
248
well suited to predicting the performance of multiple receiver channels simultaneously,
and so cannot be used to determine the overall impact of scintillations on navigational
accuracy.
•
Although information about the joint statistics of scintillations on multiple
propagation paths is not available, it is clear that if scintillations are assumed to be
independent between the individual paths, the probability of losing lock simultaneously
on multiple channels would be very small, given that the individual probabilities are also
quite small.
10.2.
Conclusions
The analysis carried out in this thesis suggests that equatorial scintillations will have a
relatively minor effect on the navigation performance of stand-alone GPS receivers. This
is partly because the patchy nature of scintillations introduces a degree of independence
between the individual satellite-receiver links, and partly because the coverage of
satellites in the equatorial region is generally very good. Consequently, the probability of
simultaneously losing lock on enough satellites to significantly degrade the satellite
geometry is relatively small. Also, on the satellite links that remain in lock, the additional
noise introduced into the code tracking loops by amplitude scintillations is unlikely to
contribute significantly to the overall pseudorange error. Data obtained from GPS
receivers deployed throughout South East Asia during the past three equinoxes strongly
supports this view.
However, if the visible constellation is reduced, either through an increase in the
elevation mask angle or obscuration from nearby obstacles, fewer satellite losses would
be required before navigational performance was significantly degraded. In addition,
Receiver Autonomous Integrity Monitoring or RAIM requires at least six visible satellites
in order to be effective. Consequently, under conditions of limited sky coverage, strong
scintillation activity may degrade the performance of RAIM.
The combination of conditions under which a receiver is likely to experience the greatest
stresses from equatorial scintillations are as follows:
• High solar activity.
• Low geomagnetic activity.
• During the months of the equinoxes (March/April and September/October).
• For several hours following local sunset.
249
• Within bands approximately 15° wide and centred on the crests of the northern and
southern anomalies.
However, simulations based on WBMOD reveal that scintillation effects may extend well
beyond the nominal northern and southern boundaries of the anomaly. This is caused by
a few satellite-receiver links penetrating the peak of the anomaly at low elevation angles,
even when the receiver is well removed from the region of the anomaly. Consequently,
areas of Northern Australia that are generally considered to be south of the southern
anomaly may still be affected by scintillations on certain low elevation angle links. In
addition, plume structures are known to reach enormous heights near the equator and
may extend the influence of scintillations well beyond that predicted by WBMOD (the
WBMOD model assumes that all irregularities are concentrated at height that is typical of
the F2-layer peak height).
The analysis carried out in Chapters 3 and 9 suggests that narrow bandwidth receivers,
such as those used in tightly coupled GPS-INS systems, are more susceptible to the effects
of phase scintillations than wide bandwidth receivers. Consequently, narrow bandwidth
receivers are more sensitive to factors that influence the phase scintillation rate such as
receiver dynamics. Wide bandwidth receivers, on the other hand, are more susceptible to
the effects of amplitude scintillations and thermal noise, and so are affected more by
factors that influence the signal to noise ratio such as the antenna gain pattern and
electromagnetic interference (EMI).
Most military and some civilian aircraft are likely to be fitted with tightly coupled
GPS-INS systems to improve their immunity to EMI and high dynamics. Although the
GPS receivers in such systems will adopt a narrow bandwidth and so will become more
susceptible to phase scintillations, the INS units will be unaffected by scintillations and so
will continue to provide a navigation solution during any scintillation induced GPS
outages. This will also help the GPS receiver to recover when the scintillation activity has
eventually passed. It is also likely that the high velocities of jet aircraft will allow the
satellite-receiver links to pass through the scintillation patches much more rapidly. Again,
the reduced dwell time within the patches will help to mitigate their effects on GPS.
250
It should also be mentioned that under high velocity conditions, the amplitude
scintillation rate will increase on many satellite links causing the assumption Bn > f c 1 to
be violated (ie. the duration of deep fades will be reduced on many links). This is likely to
reduce the impact of amplitude scintillations somewhat, particularly for narrow
bandwidth receivers for which Bn is already quite small. However, as phase scintillations
are likely to be the principal factor causing loss of lock at high receiver velocities, this
effect may be negligible in terms of the number of satellites lost.
Although the higher chipping rate of the P-Code offers P(Y)-Code receivers greater
resistance to interference, it does not afford them any protection against scintillations.
Indeed, the lower power levels of the P(Y)-Code tends to make P(Y)-Code receivers
slightly more susceptible to the effects of amplitude scintillations. Added to this is the
higher level of scintillation activity on the L2 frequency as a result of the inverse
frequency scaling of scintillations. However, military P(Y)-Code receivers are likely to be
designed and constructed much better than civilian receivers, which may give them
superior performance under conditions of reduced signal level (ie. in the presence of
amplitude scintillations).
Codeless and semi-codeless tracking loops are considerably more susceptible to the
effects of scintillations than full correlation tracking loops. The very narrow bandwidths
of codeless tracking loops increases their susceptibility to phase scintillations, despite a
reduction in phase scintillation energy through carrier aiding from the L1 C/A-Code
carrier loops. Similarly, the reduced signal to noise ratio of codeless channels greatly
increases their susceptibility to amplitude scintillations, despite their very narrow loop
bandwidths. Therefore, it seems likely that systems which rely on codeless and semicodeless receivers such as the Wide Area Augmentation System (WAAS) may suffer very
adverse effects under strong scintillation conditions. However, by the time these systems
are actually operational, the expectation is that the second civil signal on the L2 frequency
will be available which will mitigate the need for codeless and semi-codeless processing.
RMS carrier phase errors of several centimetres can be introduced onto individual
satellite-receiver links by phase scintillations. Generally, these errors will become
decorrelated over distances of a few km, depending upon the ionospheric outer scale size
1
Where Bn is the loop bandwidth and f c is the cutoff frequency of the amplitude scintillation
power spectrum.
251
parameter, f o , and the direction of the baseline. This may have a significant impact on
carrier phase DGPS measurements taken in equatorial regions during solar maximum.
Unfortunately, the modelling of f o in WBMOD is very primitive at this stage.
Under very intense scintillation conditions, the mean time to acquire the GPS signal may
increase by a factor of two or more, although only when the carrier to noise density ratio
of the GPS signal is at a reduced level. However, under these conditions it is uncertain
whether a channel will have the ability to transition to a stable tracking state anyway.
Nevertheless, as only a few of the satellites in a receiver's field of view are likely to be
subject to such high levels of activity, even during solar maximum, this effect is not
expected to be overly important.
The impact of scintillations on the navigation data also appears to be negligible. It is
likely that a tracking loop will lose lock or become unreliable well before navigation data
demodulation errors become significant. Indeed, much of the navigation data is repeated
on each satellite-receiver link and the update rate required from such information is
relatively low anyway.
The performance measures discussed in this thesis have been based on a number of
assumptions and approximations, some of which have already been discussed. However,
other sources of loop stress such as oscillator phase noise, multipath, foliage attenuation,
obscuration, antenna gain pattern variations, the elevation angle dependence of C N o ,
and EMI etc. have largely been ignored. Each of these effects is likely to reduce a receivers
tolerance to the effects of scintillations.
10.3.
Further research
Research into the effects of ionospheric scintillations on GPS is by no means complete.
Possible areas for further research include:
1. Determining the joint statistics of scintillations on multiple satellite links. Data
obtained from scintillation monitoring receivers located in equatorial regions could be
used to determine the correlation between scintillation activity on different satellitereceiver links. Factors that may influence this correlation include the separation
between individual ionospheric pierce points (IPPs), the location and local solar time
252
at the IPPs, and other factors that affect the overall levels of scintillation activity such
as the sunspot number and the magnetic activity index.
2. The correlation between signal amplitude and the strength of phase scintillations. If
the amplitude is negatively correlated to the rate of change of the carrier phase (ie. the
rate of change of carrier phase increases when the amplitude decreases), then the
combined effects of amplitude and phase scintillations on GPS may be more severe
than is predicted in this thesis.
3. The impact of highly dense ionospheric structures, or plasma lenses, on the
performance of GPS receivers. Structures of this sort produce refractive focusing and
de-focusing effects which can alter the statistics of scintillations and drive S4 values
well above one. Predicting the occurrence of lenses and developing models of the
resulting signal statistics at L-band frequencies are two areas that could be looked at
further. In [56], the effects of a collection of discrete, Gaussian shaped ionospheric
lenses on the performance of carrier tracking loops was investigated using the
diffraction model from Appendix A and the tracking loop simulator from Appendix
B. However, an analytical approach to this problem has yet to be developed. Also, it
has not been established whether naturally occurring ionospheric lenses will have a
sufficiently short focal length at L-band frequencies to cause significant scintillation
effects for GPS.
4. The development of a model to account for the non-stationary nature of
scintillations. In this thesis, it has been assumed that scintillations can be described
by wide-sense stationary random processes. However, scintillations tend to occur in
patches, the duration of which will depend on the dimensions of the irregularity
patch and the speed with which the satellite-receiver ray path scans through the
patches. Also, statistics such as S4 and T may change with time as the ray path scans
through a particular patch. Models that account for the resulting non-stationarity will
not only help with the analysis of single link performance, but will also assist in the
development of models of the joint statistics of scintillations on multiple satellite
links.
5. The validation and extension of models such as WBMOD and SCINDA for the South
East Asia / Australian region. The WBMOD and SCINDA models are based largely
on data obtained from the American longitude sector. Consequently, it is necessary to
determine whether these models also apply to the South East Asia / Australian region
by analysing scintillation data obtained from this region. DSTO in conjunction with
253
the AFRL2, LAPAN3, and DSTC4 have deployed a network of ionospheric scintillation
monitoring receivers in Indonesia, Malaysia and Papua & New Guinea to measure
scintillations and to compare their occurrence statistics with models such as WBMOD
(see for example [57], [22]). The data obtained from these sites is also being used to
investigate the probability of scintillation activity occurring simultaneously on
multiple satellite links.
6. The validation of analytical results against hardware simulations. For the past few
years, various groups within the United States5 have been investigating the effects of
simulated scintillation data on real GPS receivers by appropriately modulating the
signals produced by satellite signal simulators. However, these tests have yet to
include the effects of high receiver velocities, inertial aiding, and EM interference on
receivers that are subject to scintillations.
7. Accounting for new developments in GPS, including:
• Alternative tracking and acquisition architectures.
• The use of beamforming antennas. Antennas that are capable of steering beams
towards the GPS satellites will gain a significant advantage under amplitude
scintillation conditions as a result of an increase in the signal to noise ratio. Antennas
of this sort are already in existence (eg. the Navsys Corporation High gain Advanced
GPS Receiver or HAGR™ [18]).
• The introduction of a second civil signal at 1176.45 MHz (referred to as the L5
signal), the new military code or M-Code signals, the addition of a C/A-Code signal
at the L2 frequency, and a general increase in satellite signal power levels following
the launch of the modified block IIR and new generation block IIF satellites. Under
the current launch schedule, it is anticipated that the constellation will consist
primarily of block IIF and the proposed new block III satellites by the time the next
solar maximum occurs in 2011.
• Increased satellite coverage through the use of supplementary satellite navigation
systems such as the Russian GLONASS system, the European Geostationary
Navigation Overlay System (EGNOS), and the proposed European Galileo system.
2
AFRL: The Air Force Research Laboratory, USA.
3
LAPAN: The Ionospheric Research and Development Centre, Indonesia.
4
DSTC: The Defence Science and Technology Centre, Malaysia.
5
The GPS & Navigation Systems division of the SPAWAR Systems Center in San Diego, and the
Air Force Research Laboratory and Wright Patterson Laboratory, AFRL and WPL, in Cambridge
Massachusetts and Dayton Ohio respectively.
254
Appendix A
Scintillation model
The scintillation model used in this study is essentially that of Titheridge [89] and Davies
[28] and is based on the solution of the Fresnel-Kirchoff integrals for an assembly of field
aligned irregularities. The irregularities are assumed to produce variations in TEC in
directions normal to the earth's magnetic field lines, but no variations along the field
lines. For convenience, the phase perturbations are assumed to be concentrated within a
thin phase screen located at a typical F2 layer peak height. In reality, such perturbations
would result from the cumulative effect of numerous small irregularities located along
the ray path.
A.1 Deterministic phase screen
The deterministic model, which is essentially that described by Titheridge and Davies,
assumes that the irregularities are discrete rod like lenses which are aligned with the
Earth’s magnetic field and produce Gaussian phase perturbations normal to the field. The
phase perturbation produced by k such lenses is given by
− (x − x0i )2
Φ 0i . exp
li 2
i =1
k
Φ( x ) =
∑
(A-1)
where x is the horizontal position in a direction normal to the Earth’s magnetic field lines
(ie. in an East-West direction), and Φ 0i , li , and x0i are the peak phase variations, scale
sizes, and centres of the irregularities respectively. The peak phase variations, Φ 0i , are
related to the peak TEC variations, ∆TEC0i , through the expression (Davies [27])
Φ 0i = 40.3∗2π
∆TEC0i
cf
(radians)
(A-2)
where f is the GPS carrier frequency and c is the speed of light. If it is assumed that the
variations in plasma density are Gaussian in both the x and y (vertical) directions, ∆TEC0i
can be found as a function of the scale sizes and peak density variations. The variation in
255
plasma density over the background is given by1
(x − x )2 ( y − y )2
0
0
∆N ( x, y ) = ∆N 0 exp −
−
2
2
lx
ly
(A-3)
where ∆N 0 is the peak density variation. As TEC is the vertical integral of the electron
density,
∞
∆TEC ( x ) =
∫ ∆N ( x, y ).dy
−∞
(x − x0 )
= ∆N 0l y π exp −
lx2
2
(A-4)
The peak TEC variation is therefore
∆TEC0 = ∆N 0l y π
(A-5)
∆N 0 can then be defined as a fraction, p, of the background electron density, N b , as
follows
∆N 0 = pN b ,
p ≥ −1
(A-6)
An additional constraint which can be applied in order to establish a relationship
between the scale sizes, l x and l y , and ∆N 0 is the maximum permissible electron density
gradient, N gm . From Equation (A-3), the maximum density gradients in both the x and y
directions are
∂∆N
∂x
∂∆N
∂y
=
∆N 0 2 e
,
lx
y=y0
(A-7)
=
∆N 0 2 e
,
ly
x=x0
(A-8)
max
max
The maximum peak density variation which will ensure that N gm is not exceeded is then
1
This assumes that the density profile is symmetrical in both the x (horizontal) and y (vertical)
planes.
256
∆N 0 = min(l x , l y ). N gm e 2
(A-9)
By combining Equations (A-1), (A-2), (A-5), (A-6) and (A-8), an irregularity layer can be
defined once values for N b and N gm are assigned. If it is assumed that the irregularities
are located at the F2 layer peak, N b will be equal to the peak density of the F2 layer,
NmF2.
The peak phase variations can either be positive, which corresponds to an enhancement
in TEC (a defocusing type irregularity), or negative which corresponds to a depletion in
TEC (a focusing type irregularity). Both types of irregularity have been shown to exist in
the ionosphere [90]. The radio wave is also assumed to experience no attenuation as it
passes through the phase screen. Consequently, any amplitude fluctuations are caused
entirely by the effects of interference across the wavefront as it propagate towards the
ground.
At the ground, and relative to the undisturbed wave, the in-phase, I, and quadrature, Q,
components of a vertically propagating plane wave2 can be found by solving the FresnelKirchoff integrals (see for example [89] & [28]). This results in the following two terms
I = 1 − 2 sin(Po + Φ 2 )sin(Φ 2 )dx / rλ
∫
(A-10)
ir
Q = 2 cos(Po + Φ 2 )sin(Φ 2 )dx / rλ
∫
(A-11)
ir
where Φ = Φ (x ) , Po = − π 4 − 2π (r − h ) λ , h is the screen height and r joins the elements dx
on the phase screen to a point on the ground (see Figure A-1). As these two integrals
contain a sin (Φ 2 ) term which approaches zero when the emergent wave is unperturbed,
the integrals need only be calculated over the region of the irregularities in the plane of
the phase screen. This region is denoted as ir.
2
In this analysis it is assumed that the GPS satellites are at sufficiently high altitudes to make the
simplifying assumption that the all incident waves are plane waves.
257
Incident wave
Irregularity layer
dx
Emergent wave
h~400km
r
Figure A-1: Geometry of the thin phase screen diffraction model.
The resulting phase and amplitude variations are obtained from the I and Q components
as follows
Phase = Atan (Q I )
(A-12)
Amplitude = Q 2 + I 2
For irregularities much larger than the first Fresnel zone radius, z F , diffraction effects are
minimal and a geometric optics solution can be employed. As a result, phase variations
on the ground will closely resemble those in the ionosphere and amplitude variations will
be negligible. On the other hand, for scale sizes of the order of the Fresnel zone radius or
smaller, or for very large plasma density gradients, rapid variations in both amplitude
and phase will occur. Irregularities smaller than z F produce diffraction effects, whereas
those containing large density gradients produce significant refraction and hence
interference effects. Both cause rapid variations in the amplitude and phase of the GPS
signals on the ground. For a Gaussian shaped irregularity, the threshold conditions for
diffraction and interference are (from [89])
Diffraction:
Interference:
l < zF =
l<
λz1 z 2
≈ λz 2
z1 + z 2
(A-13)
Φ 0h
37 f
(A-14)
where Φ 0 is the peak phase variation, z1 and z2 are the distances between the
ionospheric irregularity layer and the satellite and receiver respectively. Typical upper
limits on the scale sizes of irregularities which are likely to produce these effects are
258
Diffraction:
300m (L2),
275m (L1)
Interference:
220m (L2),
190m (L1)
where it has been assumed that hi = 400km and ∆TEC0 = 1 TECu (approximately 1% of a
typical equatorial TEC value during solar maximum). This analysis also assumes that the
irregularities are directly overhead and that propagation is vertical. For E-region
irregularities (h ~ 100 km), the equivalent dimensions are approximately one half.
It is expected that irregularities of the order of the Fresnel Zone radius or slightly smaller
will produce the most significant scintillation effects. Larger irregularities are unlikely to
produce fully developed diffraction effects and would require very large peak densities in
order to produce significant interference effects. Irregularities much smaller than the
Fresnel zone radius will produce prolific diffraction effects, but with very small peak
phase variations. Examples of the effects of different scale sizes on the amplitude and
phase diffraction patterns is given in Section 2.1.5.
The phase and amplitude fluctuations derived from this model are a function of position
in an East-West direction. Because the irregularities are assumed to be field aligned,
fluctuations do not exist in a North-South direction. Consequently, the temporal
variations in phase and amplitude experienced by a GPS receiver will depend on the
East-West component of the irregularity velocity, the GPS platform velocity and the
satellite velocity. These velocity components are encapsulated in the effective scan
velocity, ve , which was discussed in Section 2.1.3.
A.2 Random phase screen
A more realistic model for the phase screen based on in-situ measurements3 of electron
density fluctuations is the power law phase screen model. This model assumes that the
phase perturbations on the emergent wave are random with a Gaussian distribution and
a power law wavenumber power spectrum which is given by
(
S Φ ( k ) = A. k o 2 + k 2
)
−p 2
(A-15)
where k is the wavenumber, A determines the strength of scintillation activity,
ko = 2π lo m -1 is an outer scale wavenumber, lo is the outer scale size, and p is the
3
In-situ measurements of electron density are made with probes flown on rockets and satellites.
259
spectral index. The in-situ wavenumber power spectrum of phase (or equivalently of
TEC) has the same spectral index as the power spectrum of scintillations measured on the
ground ([16] and [49]). Consequently, p is in the range 1 to 4 and is typically 2.5 at
equatorial latitudes.
Realisations of a random phase screen Φ (x ) with the desired spectral characteristics can
be obtained from an Nth order Autoregressive (AR) process of the form
N
Φ ( nX ) = w(nX ) +
∑ hi Φ([n − i]X )
(A-16)
i =1
where n is an integer, X represents the spacing between successive points on the phase
screen, w(nX ) is a white Gaussian noise process and hi are the coefficients of the AR
model. Equation (A-16) can be rearranged as follows
N
Φ ( nX ) −
∑ hi Φ([n − i]X ) = w(nX )
(A-17)
i =1
Multiplying (A-17) through by Φ (nX ) and taking the expectation gives
N
r0 −
∑ hi ri = σ w 2
(A-18)
i =1
where r j = RΦ ( jX ) is the Autocorrelation function of phase at a separation of jX m and
σ w 2 is the variance of the white noise process. In a similar way, multiplying through by
Φ ([n − 1]X ) and taking the expectation gives
N
r1 −
∑ hi ri −1 = 0
(A-19)
i =1
By repeating this process N+1 times, the following matrix expression can be obtained
r0
r
1
6
rN
r1 5 rN 1
1
0
6 − h1
r0
= σ w2
6
8 6 6
5 5 r0 − h N
0
which can also be expressed as
260
(A-20)
P.h = σ w 2δ
(A-21)
The coefficients of the AR model are therefore
h = σ w 2 . P −1.δ
(A-22)
Consequently, in order to determine h , the autocorrelation function of phase RΦ (x ) must
be found. As the wavenumber power spectrum of phase is an even function, RΦ (x ) is
given by
∞
∫ SΦ (k ).e
RΦ ( x ) =
jkx
.dk
−∞
(A-23)
∞
∫
= 2 S Φ ( k ). cos(kx ).dk
0
Substituting (A-15) into (A-23) gives
∞
RΦ ( x ) = 2 A ∫
0
cos( kx )
(k
2
o
+ k2
)
p 2
. dk
(A-24)
This is a general expression for the autocorrelation function of phase for a power law
wavenumber power spectrum. Unfortunately, it is difficult to solve in closed form for
arbitrary p. Consequently, numerical techniques must be used in order to determine
RΦ ( x ) and P -1 for any p.
However, for the special case of p=2, the following simple solution for RΦ (x ) can be
obtained from a table of integrals
∞
RΦ ( x ) = 2 A
cos(kx )
∫ ko2 + k 2 .dk
0
(A-25)
Aπ − k o x
e
=
ko
Giving
1
Aπ e − k o X
P=
ko 6
− Nk X
e o
e −k o X
1
5
5 e − Nk o X
6
8
6
5
1
(A-26)
261
Substituting (A-26) into (A-22) and solving for the AR coefficients gives h1 = e − k o X and
h2 5 hN = 0 . Therefore, for p=2 a first order AR process will generate suitable realisations,
viz
Φ ( nX ) = w( nX ) + Φ ([n − 1]X ).e − ko X
(A-27)
An example of simulated scintillation data obtained from the Fresnel-Kirchoff diffraction
model using the AR process from Equation (A-27) is given in Figure A-2. Parameter
values for the AR process are lo = 1km (outer scale size), X = 1m (step size), and
hi = 400km (ionospheric height). In addition, it was assumed that the velocity of the
propagation path normal to the receiver was 150m/s and that the L1 frequency was being
used. Consequently, for this example, the AR process is of the form
Φ ( nX ) = w( nX ) + 0.994Φ ([n − 1]X )
(A-28)
In−situ phase (rad)
where w(nX ) is assumed to have a variance of 0.0025 radians2.
2
1
0
−1
−2
0
5
10
Phase (rad)
2
1
0
−1
−2
0
10
20
30
40
50
60
Time (s)
70
80
90
100
10
20
30
40
50
60
Time (s)
70
80
90
100
5
Power (dB)
15
Positions (km)
0
−5
−10
0
Figure A-2: A sample realisation of the In-situ phase (upper panel), phase scintillations (middle
panel), and amplitude scintillations (lower panel) obtained from the Fresnel-Kirchoff diffraction
model.
262
Notice that the phase process at the receiver (middle panel) bears only a vague
resemblance to the in-situ phase process (upper panel). For much larger in-situ phase
gradients, the two phase processes would become even more different. The RMS phase
and S 4 values for this particular example are 0.41 radians and 0.39 respectively. These
were obtained by averaging numerous different realisations based on Equation (A-28)
and the phase screen model.
263
264
Appendix B
GPS tracking loop simulators
In this thesis, software simulations of GPS receiver code and carrier tracking loops are
used to validate analytical results and to determine where the analytical results begin to
fail. The advantage of simulation over a purely analytical approach is that it allows the
loop non-linearities to be taken into account without the need to introduce any
approximations. These non-linearities include the discriminator, the I and Q channel
mixers and the code correlators. Although the majority of the work described in this
thesis is based on stand-alone code and carrier tracking loops, the behaviour of a
combined code and carrier tracking loop channel can be investigated relatively easily
through the simulation approach described here.
The tracking loop simulators are based on Simulink for Matlab and are driven by
amplitude and phase scintillation data produced by the model described in Appendix A.
Simulink diagrams of a stand-alone Costas carrier tracking loop and a delay locked
loop are given in Figures B-1 and B-2. An equivalent diagram of a combined tracking loop
is given in Figure B-3. In all cases, the simulated scintillation data is stored in amplitude
and phase lookup tables and is extracted at a rate which depends on the value of ve
chosen for the simulation.
For the Costas loop simulator, it is assumed that the prompt code estimate is perfectly
aligned with the satellite code (ie. τ ε = 0 ) and is therefore removed completely from the
GPS signal. It is also assumed that the pre-detection filters are correctly synchronised to
the navigation data. Similarly, for the delay locked loop it is assumed that the carrier
phase error is negligible (ie. φε = 0 ) and so the Q channel consists only of thermal noise
(ie. Ap (t )sin(φε ) + nQ ≈ nQ ). In the combined simulator, the code and carrier phase errors
affect both tracking loops. In addition, because of the dispersive nature of scintillations,
the effects of phase scintillations on the code delay is assumed to be equal in magnitude
but opposite in sign to the effect on the carrier phase.
265
Amplitude
Phase
nQP
Ionospheric
Scintillations
f(u)
f(u)
QP filter
sin( )
Mux
Discriminator
nIP
Dynamics
f(u)
f(u)
f(u)
IP filter
cos( )
Cs 2+Ds+1
3rd
Es 2
As+1
Bs
2nd
Wn 1st
Order
1/s
VCO
Figure B-1: Simulink model of a Costas phase locked loop.
nIE
f(u)
IE filter
Ionospheric
Scintillations
nIP
f(u)
IP filter
nIL
f(u)
IL filter
Mux
f(u)
Discriminator
f(u)
R(t)
Early
R(t)
Prompt
R(t)
Late
nQE
QE filter
nQP
QP filter
nQL
QL filter
f(u)
Dynamics
f(u)
1/s
NCO
Figure B-2: Simulink model of a delay locked loop.
266
DLL
filter
KCode scale
Dynamics
Phase
K-
Amplitude
Carrier scale
Ionospheric
Scintillations
Receiver
Channel
Simulator
_______________________________________________________________________________
f(u)
nIE
f(u)
cos( )
IE filter
nIP
f(u)
IP filter
2
3
Carrier
Phase
nIL
Amp
f(u)
IL filter
nQE
f(u)
f(u)
QE filter
sin( )
nQP
f(u)
Mux
f(u)
Code
Discriminator
f(u)
Carrier
Discriminator
QP filter
nQL
DLL
filter
f(u)
QL filter
R(t)
Early
R(t)
Prompt
PLL
filter
R(t)
Late
1.0
1
1/s
Code
Phase
NCO
Carrier
Aiding
1/s
VCO
Figure B-3: Simulink model of a combined Costas phase locked loop / Delay locked loop channel.
The upper panel represents the environment model. The lower panel represents the tracking loops.
In all simulations, the thermal noise blocks are based on Gaussian distributed random
number generators which are fed with different seeds. This helps to ensure that the
individual noise sequences are uncorrelated, or at least approximately so. For the delay
locked loop, this implies that the separation between the Early, Prompt and Late signals
must be ½ chips or greater (see Appendix D). The code spacing is set within Early,
Prompt and Late autocorrelation functions which replicate the behaviour of the three
code correlators in a delay locked loop. In addition, all simulators provide the capacity to
vary the loop order, the loop bandwidth and the discriminator type.
267
268
Appendix C
Tracking thresholds and cycle slips
Tracking thresholds define the point at which a tracking loop transitions from a state of
stable tracking to one of loss-of-lock. When a tracking loop loses lock, the VCO frequency
drifts away from the signal frequency and the loop phase estimates become meaningless.
This is qualitatively different from a condition of frequent cycle slips during which the
loop continues to track the phase and frequency of the signal between successive slips. In
practice, however, it is difficult to distinguish between the two states, particularly when
the cycle slipping rate is very high. Consequently, it is often useful to define tracking
thresholds in terms of acceptable tracking performance. This may take the form of a
maximum acceptable probability of a cycle slip based on the non-linear analysis
described in Section C.2, or a maximum acceptable phase error variance based on the
linear model analysis described in Section C.1. The second of these threshold measures is
based on the assumption that a loop will lose lock when the magnitude of the phase
tracking error exceeds a certain boundary. This is generally set at a fairly conservative
level under the assumption that the probability of losing lock increases sharply when the
linear model approximations are significantly violated.
Because tracking loops are highly non-linear in the threshold region, Monte Carlo
simulation techniques are usually required in order to establish the true tracking
performance. The code and carrier tracking loop simulator described in Appendix B takes
account of the non-linearities in a tracking loop and provides an indication of where lossof-lock is likely to occur for various noise and dynamic conditions.
For an unaided GPS receiver, loss of carrier lock is usually followed soon after by a loss of
code lock. Consequently, the tracking threshold of such a receiver is typically set by the
carrier tracking loop. However, when a receiver is aided with Doppler measurements
from an Inertial Navigation System (INS), the tracking threshold will be determined by
the code loop. This is because Doppler aiding allows the VCO to continue tracking the
frequency of the desired signal 4, even when the carrier loop has been forced to lose lock.
4
For a prolonged outage, changes in the carrier frequency as a result of satellite motion will
eventually cause the VCO frequency to drift away from the carrier frequency.
269
However, in the analysis that follows, the receiver is assumed to be unaided and the loop
threshold is assumed to be determine by the carrier loop.
C.1 Linear analysis
The tracking loop threshold derived from the linear loop model is based on the
assumption that loss of lock occurs when the linearising approximations are significantly
violated. If the carrier phase errors are modelled as zero-mean, Gaussian5 random
variables, the probability of a phase error exceeding some threshold Φ T is given by
p (φε > Φ T ) = 1 −
ΦT
∫
−ΦT
(ϕ − φ )2
1
εSS
.dϕ
exp −
2
2π σ φε
2
σ φε
(C-1)
where φεSS is the steady state phase error resulting from relative motion between the
satellite and the receiver (see Appendix E). This can also be expressed in terms of the
complementary error function as follows
p (φε > Φ T ) =
Φ −φ
1
εSS
Erfc T
2
2σ φε
+ Erfc Φ T + φεSS
2σ φ
ε
(C-2)
A widely used rule of thumb tracking threshold for a Costas loop is that the 3-sigma
phase jitter from all sources other than dynamic stress errors must be less than 45° for the
loop to remain in lock (Ward [47], page 157). This corresponds to an error in the linear
loop model of approximately 36% for an I.Q Costas loop (ie. the phase error estimate from
the discriminator is in error by a factor of 1 − [0.5 sin (2Φ T ) Φ T ] ≈ 0.36 ). Thus for φεSS = 0 ,
the tracking threshold can be represented by 3σ ϕε = Φ T = 45o . From Equation (C-2), this
corresponds to a probability of approximately 0.27% of the phase error exceeding Φ T ,
which is the probability of a Gaussian random variable lying more than 3 standard
deviations from the mean. In Figure C.1, the RMS phase error from all sources other than
dynamic stress errors, σϕε , is plotted as a function of the steady state phase tracking
(
error, φεSS , for Φ T = 45o and p φε > Φ T ) = 0 ⋅ 27% . This curve represents the locus of
points for which the loop is on the verge of losing lock under the specified threshold
conditions (from Equation (C-2)). Notice that for a zero steady state error, the threshold
5
This is based on the assumption that the thermal noise and ionospheric phase scintillations are
both zero-mean and Gaussian distributed.
270
RMS jitter is approximately 150 ( π 12 radians) which is consistent with the initial
assumption. Also, notice that the straight section of Figure C.1 implies that the following
approximate relationship holds between the RMS phase error and the steady state
dynamic phase error for the loop to remain in lock
3σφε + φεSS ≤ 450
(C-3)
16
RMS phase error (degrees)
14
12
10
8
6
4
2
0
0
10
20
30
40
50
Steady state dynamic phase error (degrees)
Figure C.1: 1σ phase error from all sources other than dynamics versus the steady state dynamic
phase error at the tracking threshold of an I.Q Costas loop.
C.2 Non-linear analysis
In this section, various results from the non-linear analysis of tracking loops are presented
without proof. Closed form expressions are given for the PDF of phase errors, the mean
time to cycle slip and the probability of a cycle slip for a Costas loop. Although these
expressions have only been obtained for a 1st order tracking loop under fairly restrictive
signal conditions, various researchers have shown that the performance of higher order
loops can be closely approximated by these expressions with only minor adjustments to
the SNR . The principal drawback in the use of results from the non-linear model is that
they assume that the system is driven by additive white Gaussian thermal noise only.
However, in the presence of scintillations, coloured phase noise will also be present.
Nevertheless, Van Trees ([96], pages 55-56) demonstrated that by a simple translation of
the phase noise process back through the loop integrator, phase noise with a power
271
spectral density of the form Tf −2 could be considered to behave like additive white
Gaussian thermal noise at the input (also see Section 3.3.3).
The non-linear analysis is based on the solution of the non-linear stochastic differential
equation which defines the operation of a standard phase locked loop. This is given by
φ = φε +
F ( s)
[A sin(φε ) + nd ]
s
(C-4)
The origin of this equation can be understood by referring to Figure 3.1-2 with the non~
linear element 0.5 A 2 sin(2φε ) replaced with Asin(φε ) for a standard phase locked loop. By
assuming a first order loop with the input phase process, φ, constant (ie. the system is
driven by additive white thermal noise only), the following PDF for the phase errors
reduced modulo 2π can be obtained (for proof, see Holmes [43], pages 114-118)
fϑ (ϕ ) =
where I 0 (
)
exp(ρ cos(ϕ ))
,
2πI 0 (ρ )
ϕ ≤π
(C-5)
is the modified Bessel function of the first kind of order zero, ρ is the loop
SNR which is given by ρ = 1 σ φ2ε = A2 (2 N o Bn ) for a standard phase locked loop, and ϑ
is the phase error reduced modulo 2π (ie. ϑ = φε mod 2π ). This PDF is usually referred to
as the “Tikhonov density function”. For a first order Costas loop with an I.Q
discriminator, the Tikhonov density function is given by (Holmes [43], page 274)
fϑ (ϕ ) =
exp(ρ e cos(2ϕ ))
,
πI 0 (ρ e )
ϕ ≤
π
2
(C-6)
where ρe is the effective loop SNR which is given by ρe = 1 4σφ2ε , and ϑ is the phase
error reduced modulo π (ie. ϑ = φε mod π ). For an I.Q loop in the presence of white
Gaussian thermal noise only, the variance of the phase errors is given by the following
expression (Appendix D)
σ φ2T =
Bn
C N0
1
1 + 2T C N
0
(C-7)
The variance of the phase errors reduced modulo π for a first order I.Q Costas loop can be
obtained from the Tikhonov PDF as follows (assuming that the errors are zero-mean)
272
π 2
2
{ }= ∫ ϕ
π
σ ϑ =E ϑ
2
-
2
fϑ (ϕ ).dϕ
(C-8)
2
where σ ϑ 2 ≈ σ φ2ε when the phase errors are small (ie. when they rarely exceed ± π 2
radians).
For a standard phase locked loop in the presence of white Gaussian thermal noise, the
mean time to cycle slip is given by (Homes [43], page 95, Gardner [36], page 38)
T =
π 2 ρI o 2 (ρ )
2 Bn
(C-9)
where Bn is the single sided noise bandwidth of the tracking loop. For a standard phase
locked loop, a cycle slip is defined as an increase in the magnitude of the phase error by
2π radians. This usually leads to a jump in the loop’s estimate of carrier phase by an
integer number of carrier cycles.
A more generalised expression for the mean time to cycle slip which includes the steadystate phase tracking error, φεSS , is [59]
∞
π tanh (πφεSS ρ ) 2
( − 1) n I n 2 ( ρ )
T =
I o ( ρ ) + 2∑
2
2 BnφεSS
n =1 1 + ( n ρφεSS )
[
]
(C-10)
where I n ( ) is the modified Bessel function of the first kind of order n.
For an I.Q Costas loop, the mean time to cycle slip includes the effective loop SNR and is
given by (Holmes [43], page 200)
T =
π 2 ρe I o 2 (ρe )
2 Bn
(C-11)
where for a Costas loop, a cycle slip is defined as an increase in the magnitude of the
phase error by π radians (leading to an integer number of half cycle jumps in the carrier
phase). The corresponding expression for the mean time to cycle slip in the presence of a
steady-state phase tracking error, φεSS , is
273
∞
π tanh (2πφεSS ρe ) 2
( − 1) n I n 2 ( ρe )
(
)
T =
I
2
+
ρ
∑
o e
2
4 BnφεSS
n =1 1 + (n 2 ρeφεSS )
[
]
(C-12)
Equations (C-11) and (C-12) can be obtained from the standard phase locked loop
expressions by doubling both σφε and φεSS . This simple relationship holds because the
phase error characteristic of an I.Q Costas loop discriminator is sin (2φε ) 2 , while the
corresponding characteristic for a standard phase locked loop is sin (φε ) . Consequently,
the carrier phase tracking errors of an I.Q Costas loop need only be half as large as those
of a standard phase locked loop to have the same impact in terms of cycle slips.
By applying the linear model tracking threshold for σφε (Equation (C-3)) to Equation (C11), the mean time to cycle slip becomes T = 1260 Bn seconds. This demonstrates that the
tracking threshold obtained from the linear model analysis is quite conservative. For
example, for a noise bandwidth of 10Hz, the mean time to cycle slip is more than 2
minutes.
If it is assumed that the slipping process is approximately Poisson distributed, then the
probability of a slip within t seconds from a condition of zero error is (Holmes [43], page
95)
t
ps = 1 − exp −
T
274
(C-13)
Appendix D
Thermal noise errors
In this Appendix, the statistics of thermal noise errors on the code and carrier tracking
loops are examined. A method is outlined for obtaining the standard expression for the
variance of the thermal noise errors on a Costas carrier loop. This is then extended to
include the effects of amplitude scintillations on an I.Q Costas loop that is normalised by
an AGC.
D.1 Thermal noise prior to the discriminator
From Equation (3.1-1), the IF signal at the input of the receiver tracking loops is given by
IF ( t ) = A( t ) p(t − τ ( t ))d (t − τ ( t ))sin (ω IF t + φ ( t )) + n( t )
where n( t ) = nc ( t )cos(ω IF t ) + n s ( t )sin (ω IF t ) is a narrowband representation of thermal
noise at the IF stage. n(t ) is assumed to be a wideband, stationary, zero-mean Gaussian
random process with a power spectral density of N o W Hz within the IF band.
Similarly, the two baseband noise processes, nc ( t ) and ns ( t ) , are wideband, stationary,
zero-mean, Gaussian noise processes with power spectral densities of N o W Hz at
baseband. After mixing with the VCO reference signals, the noise is separated into I and
Q components as follows (note that the double frequency terms have been ignored as
they will be eliminated by filtering in the pre-detection filters).
n I = nc (t ) sin(φˆ(t )) + n s (t ) cos(φˆ(t ))
nQ = nc (t ) cos(φˆ(t )) − n s (t ) sin(φˆ(t ))
(D-1)
where n I and nQ are again uncorrelated noise processes with the same statistics as nc ( t )
and ns ( t ) . The I and Q signals are mixed with Early, Prompt and Late replica codes from
the code generator and filtered by pre-detection integrate and dump filters to produce the
three I and Q pairs given by Equation (4.1-6). These are then converted into phase
tracking errors by the code and carrier tracking loop discriminators. The noise on the six I
and Q signals after filtering can be represented by the following vector
275
n = [n IE , nQE , n IP , nQP , n IL , nQL ]
(D-2)
The covariance between any two of these six signals is given by
σ αX σ βY = E {(nαX − E{nαX }) (n βY − E {n βY })}
(D-3)
where α and β represents either I or Q, and X and Y represents either E, P or L. As all of
these noise terms are zero-mean, the covariance simplifies to
σ αX σ βY = E {nαX n βY }
1 t
1 t
nα (u ) p(u − x ).du
n β ( v ) p (v − y ).dv
= E
T t −T
T t −T
∫
=
t
1
T2
∫
(D-4)
t
∫ ∫ E{nα (u)n β (v) p(u − x) p(v − y )}.du.dν
t −T t −T
where nα and n β are given by Equation (D-1), p (u − x ) and p (v − y ) represent the PRN
codes with delays of u − x and v − y seconds respectively, and x and y represent the
delays associated with either the Early, Prompt or Late codes. As nα and n β are
uncorrelated and independent of the PRN codes, the covariance expression can be
simplified as follows
σ αX σ βY =
t
1
T2
t
∫ ∫ E{nα (u)n β (v )}E{p(u − x) p(v − y )}.du.dv
t −T t −T
1
= T2
0 ,
1
= T 2
0 ,
t
2
∫ E{nα (u) }E{p(u − x ) p(u − y )}du ,
t −T
α=β
α≠β
(D-5)
t
∫ N o R( x − y )du ,
t −T
α=β
α≠β
No
R( x − y ) ,
= T
0 ,
α=β
α≠β
where E {p(u − x ) p (u − y )}= R ( x − y ) is the autocorrelation function of the PRN code
{
}
(given by Equation (4.1-1)), and E nα (u ) 2 = Rn (0) = N o is the power spectral density of
nα (or n β ). This result also relies on n I and nQ being sufficiently wideband for
276
E {nα (u )nα (v )}= 0 for u ≠ v .
Based on Equation (D-5), the covariance matrix of the six noise terms for an Early-Late
spacing of 2d code chips6 is given by
0
0
0
R (d )
R ( 2d )
1
0
1
0
0
R (d )
R ( 2d )
0
1
0
0
R (d )
N o R (d )
T
E n n =
0
1
0
R (d )
R (d )
T 0
R( 2d )
0
0
1
0
R (d )
0
0
1
R( 2d )
R (d )
0
{ }
(D-6)
For an Early-Late spacing of 1 code chip (d = ½ ; typical of most GPS receivers), the
covariance matrix becomes
0 0.5 0
0
0
1
0
1
0 0.5 0
0
N 0.5 0
1
0 0.5 0
E nT n = o
1
0 0.5
T 0 0.5 0
0 0.5 0
1
0
0
0
0
0 0.5 0
1
{ }
(D-7)
Consequently, when the Early-Late spacing is greater than or equal to 1 code chip, the
noise on all Early and Late signals is uncorrelated (ie. R (2d ) = 0 ). However, this is not the
case for narrow correlator spacing receivers such as the Novatel GPSCard™ for which
d < ½ chips.
D.2 Thermal noise errors in the absence of scintillations
For a phase locked loop, the mean-square phase tracking error resulting from thermal
noise is given by
{ }
E φε2 = σ φ2T
∞
=
∫
2
(D-8)
H ( f ) S nd ( f ).df
−∞
6
The spacing between the Early and Prompt codes and between the Prompt and Late codes is
assumed to be d chips.
277
where H ( f ) is the closed loop transfer function (Table 3.1-2), and S nd ( f ) is the power
spectral density of the discriminator noise term (labelled nd in Figures 3.1-2 and 4.1-3). In
the analysis that follows, it is assumed that amplitude scintillations are absent and that
nd is a zero-mean random variable. Also, the sample-and-hold circuits which form part
of the pre-detection filters will maintain nd at a constant level for a period of T seconds.
However, between successive T second epochs, the values of nd will be uncorrelated. As
will be shown in Equation (D-24), this is a result of n IP and nQP being uncorrelated
between epochs.
The discriminator noise can be viewed as the output of a sample-and-hold circuit fed by a
white noise sequence, w(t ) . Therefore, we can express nd (t ) in the following way (from
Haykin [39], Section 7.3)
nd (t ) = wδ (t ) ⊗ g h (t )
where
wδ (t ) =
(D-9)
∞
∑ w(kT )δ (t − kT )
is an instantaneously sampled version of
w(t ) ,
k = −∞
t −T 2
g h (t ) = rect
represents the “hold” function of the sample-and-hold, and ⊗
T
represents the convolution integral. The corresponding power spectral density of nd (t ) is
{
S nd ( f ) = E F {wδ (t ) ⊗ g h (t )}
= S wδ ( f ) Gh ( f )
2
}
(D-10)
2
where S wδ ( f ) is the power spectral density of wδ (t ) , and Gh ( f ) = T sinc( fT )exp(− jπfT )
is the Fourier transform of g h (t ) . As w(t ) is a white noise sequence, wδ (t ) is also white.
Therefore, S wδ ( f ) = N wδo which is a constant. The variance of nd (t ) is given by
∞
σ n2d =
∫ S nd ( f ).df
−∞
∞
=
∫ S wδ ( f ) Gh ( f )
2
.df
−∞
∞
=
∫ N wδo T
−∞
= N wδo T
278
2
sinc 2 ( fT ).df
(D-11)
Consequently, S wδ ( f ) = σ n2d T and Equation (D-10) becomes
S nd ( f ) = σ n2d T sinc 2 ( fT )
(D-12)
Therefore, from Equation (D-8), the thermal noise variance of the phase errors is given by
∞
σ φ2T
=
∫ H(f )
2
σ n2d T sinc 2 ( fT ).df
−∞
(D-13)
1 ∞
2
H ( f ) sinc 2 ( fT ).df
= 2σ n2d T
2 − ∞
∫
The closed loop transfer function, H ( f ) , is a low-pass filter with a bandwidth much
smaller than the bandwidth of the sinc( fT ) function7. Consequently, sinc 2 ( fT ) can be
approximated by one, giving
1 ∞
2
σ φ2T ≈ 2σ n2d T ∫ H ( f ) .df
2 −∞
= 2TBnσ n2d ,
(D-14)
radians 2
∞
where Bn =
2
1
∫ H ( f ) df is the single-sided noise equivalent bandwidth of the tracking
2 −∞
loop. Noise equivalent bandwidth’s for the three loop orders are given in Table 3.1-2 as a
function of the loop natural frequency, ω n .
Equation (D-14) is independent of the algorithm chosen for the Costas loop discriminator.
In order to proceed, however, it is necessary to specify a discriminator algorithm so that
σ n2d can be found. For the I.Q Costas loop, the discriminator noise term nd is given by
(from Equation (3.1-3))
[
]
~
nd = A d (t − τ ) cos(φε )nQP + sin(φε )n IP + nQP n IP
7
(D-15)
The single-sided noise equivalent bandwidth of H ( f ) is normally less than 20Hz, whereas the
equivalent bandwidth for the sinc( fT ) function is 1 2T Hz = 25Hz for T=20ms.
279
~
If it is assumed that the signal amplitude is constant (ie. A = A where A is a constant),
and the discriminator is normalised by an ideal post-detection AGC (ie. g = A 2 ), the noise
term will become
nd′ =
nd
A
2
=
[
]
1
1
d (t − τ ) cos(φε )nQP + sin(φε )n IP + 2 nQP n IP
A
A
(D-16)
As nQP and n IP are uncorrelated and zero-mean, the variance of nd′ is given by
{ }
σ n2d′ = E nd′ 2
=
1
A2
{ }
cos 2 (φε )E nQP 2 +
1
A2
{ }
sin 2 (φε )E n IP 2 +
1
A4
{ }{ }
(D-17)
E nQP 2 E n IP 2
{ } { }
Letting E nQP 2 = E n IP 2 = N o T (from the diagonal elements of Equation (D-6)) gives
σ n2d′ =
=
No
N
1 + o2
2
TA TA
(D-18)
1
1
1+
2T C N o 2T C N o
where C N o = A 2 2 N o is the nominal carrier to noise power density ratio of the GPS
signal8. Consequently, the phase error variance of a normalised I.Q Costas loop is given
by (from Equations (D-14) and (D-18))
σ φ2T = 2TBnσ n2d′
=
Bn
C No
1
1 + 2T C N ,
o
(D-19)
radians 2
Equivalent expressions for the thermal noise variance of both the Delay Locked Loop
(DLL) and the Frequency Locked Loop (FLL) are given below (see for example Kaplan
[47]).
DLL:
8
σ τ2T =
4 F1d 2 Bn
C No
4 F2 d
2(1 − d ) + T C N
o
chips 2
(D-20)
If it is assumed that the nominal satellite signal power at the ground is -160dBW [81] and the
noise temperature is 530K, the nominal carrier to noise density ratio is C N o = 41 ⋅ 5 dBHz .
280
σ ν2T =
FLL:
1
1 + T C N
π T C No
o
F3 Bn
2
2
Hz 2
(D-21)
where F1 is the discriminator correlator factor (1 for time shared tau-dithered early/late
correlators, 1/2 for dedicated early/late correlators), F2 is the discriminator type factor (1
for early/late discriminators, 1/2 for dot product discriminators), F3 is 1 at high C N o
and 2 at low C N o and d is the correlator spacing (in chips). Although the discriminator
algorithms are quite different for the DLL and FLL, it is clear that the variance
expressions have the same general form as Equation (D-19).
D.3 Thermal noise errors in the presence of amplitude scintillations
In the presence of an AGC, the discriminator noise term is given by (based on Eqn (3.1-3))
nd′ =
where g =
[
[
]
nd 1 ~
= A d (t − τ ) cos(φε )nQP + sin(φε )n IP + nQP n IP
g
g
]
(D-22)
1 k 2
∑ I i + Qi 2 is the AGC gain factor. If nQP and n IP are assumed to be
k i=1
~
independent of A and g (ie. independent of amplitude scintillations), and both nQP and
n IP are uncorrelated and zero-mean, the noise term nd′ will also be zero-mean. This is
demonstrated below
If we let n x = cos(φε )nQP + sin(φε )n IP and n y = nQP n IP , it can easily be shown that
{ }
{
}
E {n x } = 0 , E n y = 0 and E n x n y = 0 . Therefore
1 ~
E {n d′ } = E Ad ( t − τ )n x + n y
g
~
Ad ( t − τ )
1
= E
E {n x } + E E n y
g
g
= 0.
[
]
{ }
(D-23)
and so nd′ is zero-mean. It is can also be shown that under the assumptions outlined
above, nd′ is uncorrelated between successive T second epochs. The autocorrelation
function of nd′ for a lag of T seconds is given by
281
Rnd′ (T ) = E {n d′ (t ) n d′ (t + T )}
[
]
[
]
1 ~
1
~
A(t )d (t − τ )n x (t ) + n y (t ) ∗
A(t + T )d (t + T − τ )n x (t + T ) + n y (t + T )
= E
g (t + T )
g (t )
~
~
A(t )d (t − τ ) A(t + T )d (t + T − τ )
= E
E {n x (t )n x (t + T )}+
g (t )g (t + T )
~
A(t )d (t − τ )
E
E n x (t )n y (t + T ) +
g (t )g (t + T )
~
A(t + T )d (t + T − τ )
E
E n y (t )n x (t + T ) +
g (t )g (t + T )
{
}
{
}
1
E
E n y (t )n y (t + T )
g (t )g (t + T )
{
}
(D-24)
{
}
As E n x n y = 0 , the second and third terms in this expression are zero. Also, as n x and
{
}
n y are uncorrelated between epochs, E {n x (t )n x (t + T )}= 0 and E n y (t )n y (t + T ) = 0 .
Consequently, Rnd′ (T ) = 0 and so nd′ is also uncorrelated between successive epochs (note
that Rnd′ (τ ) = 0 for τ > T also). The variance of nd′ is given by
{
σ n2d′ = E nd′ (t )2
}
~2
A
1
= E 2 E n x 2 + E 2 E n y 2
g
g
{ }
{ }
~2
A
1
= E 2 E nQP 2 cos 2 (φε ) + n IP 2 sin 2 (φε ) + 2nQP n IP cos(φε )sin(φε ) + E 2 E nQP 2 n IP 2
g
g
{
}
{
}
~2
A
1
= E 2 E nQP 2 cos 2 (φε ) + E n IP 2 sin 2 (φε ) + E 2 E nQP 2 E n IP 2
g
g
[{ }
{ }
]
{ }{ }
(D-25)
{ } { }
Letting E nQP 2 = E n IP 2 = N o T (from the diagonal elements of Equation (D-7)), the
variance expression reduces to
σ n2d′ =
No
T
~2
A
N 1
E 2 + o E 2
g T g
(D-26)
~
The signal amplitude A can be normalised by dividing by the nominal (unperturbed)
signal amplitude A as follows
282
~
~
A
AN =
A
{ } { }A
~
~
where E AN 2 = E A 2
2
(D-27)
= 1 under amplitude scintillation conditions. By substituting
~
~
A = A∗ AN and C N o = A 2 2 N o into the variance expression, the following result is
obtained
σ n2d′ ≈
~ 2
A
1
1
1
E 2
E N 2 +
2T C N o g N 2T C N o g N
(D-28)
where g N = g A 2 is a normalised AGC gain factor, and C N o is the carrier power to
noise density ratio of the GPS signal.
Equation (D-28) represents the noise variance at the output of a normalised I.Q
discriminator. The only assumptions made in this analysis are that the amplitude and
AGC gain factor are independent of the thermal noise terms, nQP and n IP . However, no
assumptions have been made about the bandwidth of the amplitude scintillations.
D.3 A note on sample-and-hold circuits
The outputs of the pre-detection integrate-and-dump filters will generally be held
constant for T seconds by a zero-order sample-and-hold circuit. The purpose of the
sample-and-hold is to maintain the I and Q signals at a constant level so that φˆ(t ) is fixed
for the subsequent T second integrate-and-dump period. Using arguments similar to
those given in Equations (D-9) to (D-12), it is relatively straightforward to show that the
sample-and-hold does not change the variances or power spectral densities of the six
noise terms from Equation (D-2), although it may alter their appearance in the time
domain (ie. they will become stepped rather than continuous). Therefore, if nαX (t )
represents one of the noise terms from Equation (D-2), S nαX ( f ) = N o sinc( fT )2 and
σ n2αX = N o T , irrespective of whether a sample-and-hold circuit is present after the
integrate-and-dump filters.
283
284
Appendix E
Doppler errors
The effects of satellite and receiver motion on the phase of the GPS signal can be
represented by the following Doppler expressions
2π
λ
1
1 3
2
ro + v o t + 2 a o t + 6 jo t +4
Carrier:
φd ( t ) = u( t )
Code:
1
1
1
τ d ( t ) = u( t ) ro + v o t + a o t 2 + jo t 3 +4
c
2
6
radians
seconds
(E-1)
(E-2)
where u(t ) is the unit step function, ro , vo , ao and jo are constants which define the
range, velocity, acceleration and jerk components of the relative motion between the
satellite and receiver, λ is the carrier wavelength and c is the speed of light. These
equations assume that the dynamics are applied at time t=0, and that prior to t=0 there is
no relative motion between the satellite and receiver. The Laplace transforms of φd ( t )
and τ d ( t ) are given by the following generalised expression
v
a
j
r
θd ( s ) = k o + 2o + 3o + 4o +4
s
s
s s
(E-3)
where θd ( s ) represents either φd ( s ) or τ d ( s ) , and k is a constant which equals 2π λ for
the carrier tracking loop and 1 c for the code tracking loop. The phase errors at the
output of the tracking loop are given by
θε ( s ) = [1 − H ( s )]θd ( s ) =
sθ d ( s )
s + F ( s)
(E-4)
where F (s) is the loop filter and H (s) is the closed loop transfer function (from Table 4.22). The phase error as a function of time, θε ( t ) , can be obtained from Equation (E-4) by
taking the inverse Laplace transform.
285
E.1 Steady state errors
The steady state tracking error is given by the Final Value Theorem of Laplace transform
theory as follows
θεSS =
lim
lim
sθ ( s ) .
θε ( t ) =
t→∞
s→0 ε
(E-5)
where θε ( t ) is the inverse Laplace transform of θε ( s ) . For the generalised Doppler
process of Equation (E-3), the steady state tracking error becomes
θεSS = k ∗
lim
s→0
ro s
vo
ao
jo
+
+
+ 2
+4 .
s + F ( s ) s + F ( s ) s( s + F ( s )) s ( s + F ( s ))
(E-6)
Equation (E-6) results in the following steady state phase errors as a function of the loop
order and the specified component Doppler processes.
jo t 3
6
u (t )vo t
1st Order
2nd Order
3rd Order
4th Order
vo
ωn
∞
∞
1st Order
0
2nd Order
0
3rd Order
aot 2
2
u (t )ro
0
k
0
0
u (t )
k
u (t )
∞
ao
ωn
0
2
k
jo
ωn3
Table E.1: Steady state tracking errors as a function of the loop order and loop natural frequency
ω n for the specified Doppler process.
It is clear from Table E.1 that if the loop order is less than the order of the Doppler process
minus one, the loop will lose lock. However, when the loop order is greater than or equal
to the order of the Doppler process, the steady state tracking error will be zero (assuming
that the loop filter is active). When the loop order is equal to the order of the Doppler
process minus one, the steady state phase error is given by (based on Table E.1)
286
d k θ d (t )
dt k
ωnk
θ εSS =
(E-7)
where k is the loop order. This expression implies that the steady state error can be
reduced by increasing the loop bandwidth (ie. increasing ω n ). For the frequency locked
loop, the steady state error is given by the first derivative of the phase error as follows
f εSS =
dθ εSS (t )
=
dt
d k +1θ d (t )
dt k +1
ωnk
radians/s
(E-8)
Consequently, an FLL will in general be more robust to Doppler effects than a PLL of the
same order.
E.2 Transient errors
For a typical GPS receiver closed loop transfer function, the transient errors will
overshoot the steady state error by only a small amount ([92], page 390). Nevertheless, the
contribution to the loop phase error from transient dynamic effects can be accounted for
by including a Total Transient Distortion term in the phase error variance expression [45].
The Total Transient Distortion is given by
∞
εT 2 = ∫ θε ( t ) 2 . dt
(E-9)
0
If we replace θd ( s ) with a truncated version of the series given in Equation (E-3), and we
assume that θε ( t ) is bounded (ie. The loop remains in lock), then θε ( t ) is a deterministic
power signal and its power spectral density is defined by
{
Sθε ( f ) = E θε ( f )
2
{
}
= E [1 − H ( f )]θd ( f )
2
{
2
= 1 − H ( f ) E θd ( f )
}
2
}
(E-10)
2
= 1 − H ( f ) Sθd ( f )
The Total Transient Distortion then becomes
287
∞
2
εT =
∫ Sθε ( f ). df
−∞
∞
=
∫ 1− H( f )
−∞
2
(E-11)
Sθd ( f ). df
Equation (E-11) assumes that the steady state tracking error is zero (ie the order of the
loop is greater than the order of the Doppler process). If this assumption is violated, the
Total Transient Distortion will become infinite.
288
Appendix F
Ionospheric pierce point velocity
Satellite motion
The steps required in order to determine the Ionospheric Pierce Point (IPP) velocity due
to satellite motion, v s , are outlined below. It is assumed that suitable algorithms are
available to convert between the geodetic coordinate system (latitude, longitude and
height) and the Cartesian Earth Centred Earth Fixed (ECEF) coordinate system.
Step 1
Using the satellite Almanac (or Ephemeris) parameters, calculate the satellite ECEF
coordinates, rS = ( x S , y S , z S ) . The appropriate equations can be found in Appendix II of
ICD-GPS-200 [81].
Step 2
Convert the satellite ECEF coordinates to local level coordinates as follows:
The ECEF line of sight coordinates of the satellite from the receiver are
rLS = ( x LS , y LS , z LS ) = rS − rR
where rR = ( x R , y R , z R ) are the ECEF coordinates of the receiver. Convert the receiver
ECEF coordinates to latitude and longitude coordinates, (φ R , λR ) , using an appropriate
transform. Convert the line of sight satellite coordinates to local level coordinates using
the following matrix
cos(λR )
0
− sin(λR )
R = − sin(φ R )cos(λR ) − sin(φ R )sin(λR ) cos(φ R )
cos(φ R )cos(λR ) cos(φ R )sin(λR ) sin(φ R )
The local level coordinates are then
rL ′ = R ∗ rLS ′
289
where rL = ( x L , y L , z L ) are in the East, North and Up directions respectively, and the
x L y L plane is tangent to the Earth at the receiver location.
Step 3
Determine the elevation and azimuth angles of the satellite using local level coordinates.
Only those satellites for which the elevation angles are above some low elevation angle
mask are considered further (a typical mask angle is 50)
zL
Elevation, e = arctan
2
2
( x + y )
L
L
Azimuth, a = arctan x L
yL
Step 4
Calculate the earth centred angle, Ψ (the angle between a line joining the centre of the
Earth and the receiver, and a line joining the centre of the Earth and the IPP).
cos(e) Re
Ψ = 90 − e − arcsin
Re + hI
where Re = 6378 km is the Earth’s radius, and hI = 350 km is the mean ionospheric height.
Step 5
Calculate the IPP in geodetic coordinates. The IPP is the point below which the line of
sight vector penetrates the mean ionospheric height.
φ I = φ R + Ψ cos(a )
λI = λR +
Ψ sin(a )
cos(φ I )
where (φ I , λ I ) are the IPP latitude and longitude respectively.
Step 6
Determine the velocity of the IPP in ECEF coordinates as follows:
290
Convert the IPP geodetic coordinates (φ I , λ I , h I ) to ECEF coordinates, rI = ( x I , y I , z I ) ,
using an appropriate transform. Perform this step at two time instants separated by a
small time increment, τ, to obtain the two coordinates, rI 1 and rI 2 . The ECEF velocity of
the IPP is then approximately
v I = ( v xI , v yI , v zI ) =
rI 2 − rI 1
τ
(An alternative approach is to solve the expression Re + h I = rR + β rLS for the scalar
factor β (0 < β < 1) at two different satellite locations separated in time by τ seconds. The
IPP in ECEF coordinates is then rI = rR + β rLS and the IPP velocity, v I , is once again
found using the equation given above. This approach by-passes steps 2 to 5. However, if
this approach is used, it is still necessary to find the local level vector, rL , in order to
obtain the satellite elevation angle for masking purposes.)
Step 7
Translate the IPP ECEF velocity into a local level velocity as follows:
Form a translation/rotation matrix based on the first IPP coordinate rI 1 (or (φ I 1 , λ I 1 , h I 1 )
in geodetic coordinates)
cos(λ I 1 )
0
− sin(λI 1 )
R1 = − sin(φ I 1 )cos(λ I 1 ) − sin(φ I 1 )sin(λ I 1 ) cos(φ I 1 )
cos(φ I 1 )cos(λI 1 ) cos(φ I 1 )sin(λ I 1 ) sin(φ I 1 )
Convert the velocity vector to a local level vector as follows
v LI ′ = R1 ∗ v I ′
where v LI = ( v LxI , v LyI , v LzI ) are the velocity components in the East, North and Up
directions respectively. The IPP velocity due to satellite motion is then v s = v LI and the
speed of the IPP is simply v s = v LI m/s.
291
Receiver motion
The steps required to determine the IPP velocity due to receiver motion, v r , are outlined
below.
Step 1
Determine the ECEF coordinates of two receiver locations separated by τ seconds in time.
The receiver must be characterised in terms of its location rR (in ECEF coordinates), and
its velocity in local level coordinates, v R = ( v xR , v yR , v zR ) (East, North and Up directions
respectively). Following τ seconds of motion, the location of the receiver in local level
coordinates (with respect to rR ) is given by v R ∗ τ . Consequently, the ECEF line of sight
vector from rR to the new receiver location is
rLS ′ = R −1 ∗ v R ′ ∗τ
where R −1 is the inverse of the transformation matrix given earlier. Therefore, the two
receiver locations in ECEF coordinates will become
rR1 = rR
rR 2 = rR + rLS
Step 2
If it is assumed that the satellite position, rS , remains fixed over the τ second period of
interest, the procedures outlined earlier for satellite motion can be repeated in order to
derive the two IPP coordinates corresponding to the two receiver locations (ie. repeat
step’s 2 to 6 for satellite motion). As before, these two coordinates can then be combined
in order to obtain ECEF, and finally local level IPP velocity measures (ie. v r ).
Combined motion
It is quite straightforward to combine the effects of satellite and receiver motion by
calculating two values of rS ( rS1 and rS 2 ) and two values of rR ( rR1 and rR 2 ) which
represent the two time instants of interest. Steps 2 to 7 of the analysis of satellite motion
can then be used to determine v I based on combined motion.
292
However, to a first approximation, the IPP velocity can simply be found by summing the
two velocity vectors obtained by treating the satellite and receiver motion separately.
Equivalent satellite velocity
For GPS calculations, the WBMOD model is called using the FILE option. This requires
that the receiver be stationary at a known location, and that the satellite position and
velocity vectors be specified in a file that is passed to WBMOD. Unfortunately, WBMOD
does not include an option in which both the satellite and receiver are capable of moving
independently of one another. To account for this, it is possible to determine an
equivalent satellite velocity vector that includes the effects of both satellite and receiver
motion. The receiver is then considered to be stationary and the equivalent velocity vector
is passed to WBMOD within the satellite file. The equivalent satellite velocity vector is
calculated as follows:
The ECEF satellite vector can be described in terms of the ECEF receiver and IPP vectors
as follows
rS = rR +
1
(rI − rR )
β
where β is the scalar factor described earlier. If rR is considered to be stationary, and β is
assumed to be constant, the time derivative of the above expression is given by
∂rR 1 ∂rI ∂rR
−
+
∂t
∂t β ∂t
v
= I
β
v SE =
where v I is the IPP ECEF velocity which includes the effects of both satellite and receiver
motion, and v SE is the equivalent satellite velocity. The value of v SE given by this
expression will ensure that WBMOD creates an IPP velocity vector which accounts for
both satellite and receiver motion, and that β is a constant over some small time interval,
∂t . The v SE vector must then be transformed into a local level velocity vector for
WBMOD by multiplying by the following matrix
293
cos(λS )
0
− sin(λS )
RS = − sin(φ S )cos(λS ) − sin(φ S )sin(λS ) cos(φ S )
cos(φ S )cos(λS ) cos(φ S )sin(λS ) sin(φ S )
where φ S , λS are the satellite latitude and longitude respectively.
294
Appendix G
WBMOD predictions of fc
Many of the results derived in this thesis are based on the assumption that the bandwidth
of the amplitude scintillations is much less that the bandwidth of the tracking loops. This
assumption greatly simplifies the problem of analysing tracking errors and allows
analytical expressions to be obtained for measures such as phase error variance and the
probability of losing lock. The cutoff frequency of the amplitude scintillation power
spectrum, f c , is an important indicator of the validity of this assumption as the majority
of the amplitude scintillation energy is expected to be concentrated near to f c (above f c ,
the PSD of amplitude scintillations falls off according to a power law expression of the
form k A f − p where k A is a constant). From Equation (2.1-3), f c is given by
fc =
ve
2 zF
Hz
(G-1)
where ve is the effective velocity, z F ≈ λz is the Fresnel zone radius, λ is the carrier
wavelength, and z is the distance to the irregularity layer. For a given satellite-receiver
geometry, WBMOD provides predictions of ve based on internal models of the
ionospheric drift velocity, vd . In addition, simple geometry can be used to determine z as
a function of the elevation angle, e, for an assumed ionospheric height of hi , viz
z = re sin(e )2 + (1 + hi re )2 − 1 − sin(e )
(G-2)
where re is the radius of the Earth. In Figure G-1, WBMOD predictions of ve under the
same conditions as those used in Figure 3.4-5 were used to obtain f c as a function of the
elevation angle for hi =350km. Each point in the upper panel of this figure represents one
propagation path at one instant in time between 12:00 noon and 14:00 hours GMT. It is
clear from this figure that on average, f c tends to be slightly larger at low elevation
angles. This implies that the increase in z F at low elevation angles tends to be
outweighed by a larger increase in ve on some of the low angle links. It is also clear that
f c is generally less than about 0.3Hz (often considerably less), which is well below the
295
bandwidth of a typical carrier loop. However, it is possible that in the presence of high
platform velocities, f c may increase significantly on some links. If the carrier loop
bandwidth is also very narrow, perhaps due to INS aiding, the assumptions made about
the amplitude bandwidth may be violated, particularly for the narrower bandwidth code
loops.
fc (Hz)
0.4
0.3
0.2
0.1
0
0
10
20
10
20
30
40
50
60
70
80
90
30
40
50
60
70
80
90
Mean (Hz)
0.2
0.15
0.1
0.05
0
0
Elevation angle (degrees)
Figure G-1: f c as a function of the elevation angle from WBMOD.
296
Bibliography
[1]
Aarons, J. A., “Equatorial Scintillations: A Review”, IEEE Transactions on Antennas
and Propagation, vol. AP-25, No. 5, pp. 729-736, 1977.
[2]
Aarons, J. A., Basu, S., “Ionospheric Amplitude and Phase Fluctuations at the GPS
Frequencies”, Proceedings of ION GPS-94, Seventh International Technical Meeting
of the Satellite Division of the Institute of Navigation, pp. 1569-1578, 1994.
[3]
Aarons, J. A., et al., “Gigahertz scintillations associated with equatorial patches”, Radio
Science, vol. 18, No. 3, pp. 421-434, 1983.
[4]
Ashjaee, J., Lorenz, R., “Precision GPS Surveying After Y-Code”, Proceedings of ION
GPS-92, Fifth International Technical Meeting of the Satellite Division of the
Institute of Navigation, pp. 657-659, 1992.
[5]
Bandyopadhayay, T., Guha, A., DasGupta, A., Banerjee, P., Bose, A., “Degradation of
navigational accuracy with global positioning system during periods of scintillation at
equatorial latitudes”, Electronics Letters, vol. 33, No. 12, pp. 110-111, 1997.
[6]
Barrow, B. B., “Error probabilities for data transmission over fading radio paths”, SHAPE
Air Defence Technical Centre, Den Haag, Report TM-26, 1962.
[7]
Basu, S., et at., “250MHz/GHz scintillation parameters in the equatorial, polar, and
auroral environments”, IEEE Journal on Selected Areas in Communications, vol.
SAC-5, No. 2, pp. 102-115, 1987.
[8]
Basu, S., Basu, S., “Equatorial scintillations - a review”, Journal of Atmospheric and
Terrestrial Physics, vol. 43, No. 5/6, pp. 473-489, 1981.
[9]
Bishop, G. J., Basu, S., “Impacts of Ionospheric Fading on GPS Navigation Integrity”,
Proceedings of ION GPS-94, Seventh International Technical Meeting of the
Satellite Division of the Institute of Navigation, pp. 577-585, 1994.
[10]
Bishop, G. J., Bullet, T. W., Holland, E. A., “GPS Measurements of L-Band Scintillation
and TEC in the Northern Polar Cap Ionosphere at Solar Maximum”, Proceedings of the
International Beacon Satellite Symposium, 1994.
[11]
Bishop, G. J., Mazella, A. J., Holland, E. A., Rao, S., “An Overview of Ionospheric
Effects and Mitigation in RF Communication, Navigation and Surveillance”, Ionospheric
Effects Symposium IES-96, 1996.
[12]
Bishop, G. J., et al., “Test Bed for Evaluation of GPS Receivers’ Performance in
Ionospheric Scintillation - A Progress Report”, Proceedings of the ION 1998 National
Technical Meeting, pp. 637-648, 1998.
297
[13]
Bishop, G. J., et al., “GPS Applications to Global Ionospheric Monitoring: Requirements
for a Ground-Based System”, Proceedings of ION GPS-92, Fifth International
Technical Meeting of the Satellite Division of the Institute of Navigation,
pp. 393-353, 1992.
[14]
Bishop, G. J., et al., “Upcoming Ionospheric Impacts on GPS at Solar Maximum, what do
we know / what do we need ?”, Proceedings of ION GPS-96, Ninth International
Technical Meeting of the Satellite Division of the Institute of Navigation,
pp. 595-604, 1996.
[15]
Blanchard, A., “Phase-Locked Loops, Application to Coherent Receiver Design”, John
Wiley & sons, 1979.
[16]
Bogusch, R. L., Guigliano, F. W., Knepp, D. L., Michelet, A. H., “Frequency Selective
Propagation Effects on Spread-Spectrum Receiver Tracking”, Proceedings of the IEEE,
vol. 69, No. 7, pp. 787-796, 1981.
[17]
Bogusch, R. L., Guigliano, F. W., Knepp, D. L., “Frequency-Selective Scintillation
Effects and Decision Feedback Equalization in High Data-Rate Satellite Links”,
Proceedings of the IEEE, vol. 71, No. 6, pp. 754-767, 1983.
[18]
Brown, A., Holm, E., Groves, K., “GPS Ionospheric Scintillation Measurements using a
Beam Steering Antenna Array for improved Signal/Noise”, Proceedings of the 54th
Annual Meeting of the Institute of Navigation, 1998.
[19]
Brown, R. G., Hwang, P. Y. C., “Introduction to random signals and applied Kalman
filtering - Second Edition”, John Wiley & sons, New York, 1992.
[20]
Campanile, J., et al., “GPS Acquisition Performance in the Presence of Jamming”,
Proceedings of ION GPS-92, Fifth International Technical Meeting of the Satellite
Division of the Institute of Navigation, pp. 1569-1578, 1992.
[21]
Caton, R. G., McNeil, W. J., Groves, K. M., Basu, S., Sultan, P., “Real-Times UHF and
L-Band Scintillation Measurements with the Scintillation Network Decision Aid
(SCINDA)”,Proceedings of the Ionospheric Effects Symposium - IES 99, pp. 1A6-1
to 1A6-8, 1999.
[22]
Cervera, M. A., et al., “Validation of WBMOD in the South-East Asian Region”, paper
submitted to Radio Science journal in June 2000.
[23]
Chytil, B., “The distribution of amplitude scintillation and the conversion of scintillation
indices”, Journal of Atmospheric and Terrestrial Physics, vol. 29, pp. 1175-1777,
1967.
[24]
Coco, D. S., Coker, C., Valladares, C. E., Bishop, G. J., Mazzella, A. J., Fremouw, E.
J., Howell, D., “High Speed GPS Scintillation Measurements”, Proceedings of the
Ionospheric Effects Symposium - IES 99, pp. 2A6-1 to 2A6-6, 1999.
298
[25]
Crepeau, P. J., “Uncoded and coded performance of MFSK and DPSK in Nakagami fading
channels”, IEEE Transactions on Communications, vol. COM-40, No. 3, pp. 487-493,
1992.
[26]
Crepeau, P. J., “Performance of FH/BFSK with Generalised Fading in Worst Case
Partial-Band
Gaussian
Interference”,
IEEE
Journal
in
Selected
Areas
of
Communications, vol. SAC-8, No. 5, pp. 884-886, 1990.
[27]
Davies, K., “Ionospheric Radio”, vol. 31, IEE Electromagnetic Waves Series, Peter
Peregrinus Ltd., 1990.
[28]
Davies, K., Whitehead, J. D., “A radio lens in the ionosphere”, Journal of Atmospheric
and Terrestrial Physics, vol. 39, pp. 383-387, 1977.
[29]
Doherty, P. H., Klobuchar, J. A., “Expected Ionospheric Scintillation Effects During the
Next Solar Maximum – Observations From Previous Solar Cycles”, Proceedings of the
Ionospheric Effects Symposium - IES 99, 1999.
[30]
Eng, T., Milstein, L. B., “Partially Coherent DS-SS Performance in Frequency Selective
Multipath Fading”, IEEE Transactions on Communications, vol. 45, No. 1,
pp. 110-118, 1997.
[31]
Eng, T., Milstein, L. B., “Coherent DS-CDMA Performance in Nakagami Multipath
Fading”, IEEE Transactions on Communications, vol. 43, pp. 1134-1143, 1995.
[32]
Esposito, R., “Error probabilities for the Nakagami channel”, IEEE Transactions on
Information Theory, vol. IT-13, pp. 145-148, 1967.
[33]
Fremouw, E. J., “DSTO Workshop on Ionospheric and Atmospheric effects on Spaced
Based Assets”, private correspondence, DSTO, Jan 1999.
[34]
Fremouw, E. J., Livingston, R. C., Miller, D. A., “On the statistics of scintillating
signals”, Journal of Atmospheric and Terrestrial Physics, vol. 42, pp. 717-731, 1980.
[35]
Fremouw, E. J., et al., “Early results from the DNA Wideband satellite experiment Complex-signal scintillation”, Radio Science, vol. 13, No. 1, pp. 167-187, 1978.
[36]
Gardner, F. M., “Phaselock Techniques, 2nd Ed.”, John Wiley & sons, 1979.
[37]
Gradshteyn, I. S., Ryzhik, I. M., “Table of Integrals, Series, and Products: Fourth Ed.”,
Academic Press, New York, 1965.
[38]
Groves, K. M., et al., “Equatorial scintillation and system support”, Radio Science, vol.
32, No. 5, pp. 2047-2064, 1997.
[39]
Haykin, S., “Communication Systems ”, Second Edition, John Wiley and sons Inc.,
1983.
[40]
Hegarty, J. C., “Analytical Derivation of Maximum Tolerable In-Band Interference Levels
for Aviation Applications of GNSS”, NAVIGATION, Journal of the Institute of
Navigation, vol. 44, No. 1, pp. 25-34, 1997.
299
[41]
Hegarty, C., Bakry El-Arini, M., Kim, T., Ericson, S., “Scintillation Modelling for
GPS/WAAS Receivers”, Proceedings of the Ionospheric Effects Symposium - IES 99,
pp. 3A6-1 to 3A6-8, 1999.
[42]
Heron, M., L., "Diffraction from discrete and homogeneously structured ionospheric
irregularities", Radio Science, vol. 14, No. 1, pp. 97-102, 1979.
[43]
Holmes, J. K., “Coherent Spread Spectrum Systems”, Krieger Publishing Company,
1990.
[44]
Hurd, W. J., Statman, J. I., Vilnrotter, V. A., “High Dynamic GPS Receiver Using
Maximum Likelihood Estimation and Frequency Tracking”, IEEE Transactions on
Aerospace and Electronic Systems, vol. AES-23, No. 4, pp. 425-437, 1997.
[45]
Jaffe, R., Rechtin, E., “Design and Performance of Phase-Lock Loops Capable of
Near-Optimum Performance Over a Wide Range of Input Signal and Noise Levels”, IRE
Transactions, IT-1, pp. 66-76, 1955.
[46]
Kam, P. Y., Teo, S. K., Some, Y. K, Tjhung, T.T., “Approximate results for the bit error
probability of binary phase shift keying with noisy phase reference”, IEEE Transactions
on Communications, vol. 41, No. 7, pp. 1020-1022, 1993.
[47]
Kaplan,
E.
D,
“Understanding
GPS:
Principles
and
Applications”,
Mobile
Communications Series, Artech House, 1996.
[48]
Kaplan, G., Ram, U., “Bounds on performance for the noisy reference PSK channel”,
IEEE Transactions on Communications, vol. 38, No. 10, pp. 1699-1707, 1990.
[49]
Kelly, M. C., “The Earth’s Ionosphere; Plasma Physics and Electrodynamics”, vol. 43,
International Geophysics Series, Academic Press Inc., 1989.
[50]
Kintner, P. M., Kil, H., “Fading time scales associated with GPS signals and potential
consequences”, Proceedings of the Ionospheric Effects Symposium - IES 99,
pp. 5A3-1 to 5A3-7, 1999.
[51]
Klobuchar, J. A., “Ionospheric Time-Delay Algorithm for Single-Frequency GPS Users”,
IEEE Transactions on Aerospace and Electronic Systems, vol. AES-23, No. 3,
pp. 325-331, 1987.
[52]
Klobuchar, J. A., “Ionospheric Effects on GPS”, GPS World, pp. 48-51, April 1991.
[53]
Klobuchar, J. A., “Global Positioning System: Theory and Applications, Volume I”,
Progress in Astronautics and Aeronautics, vol. 163, Chapter 12, pp. 485-515, 1996.
[54]
Knepp, D. L., “Multiple Phase-Screen Calculation of the Temporal Behaviour of
Stochastic Waves”, Proceedings of the IEEE, vol. 71, No. 6, pp. 722-737, 1983.
[55]
Knepp, D. L., Houpis, H. L. F., “Altair VHF/UHF Observations of Multipath and
Backscatter Enhancement”, IEEE Transactions on antennas and propagation, vol. 39,
No. 4, pp. 528-534, 1991.
300
[56]
Knight, M. F., Finn, A., “The Impact of Ionospheric Scintillations on GPS Performance”,
Proceedings of ION GPS-96, The Ninth International Technical Meeting of the
Satellite Division of the Institute of Navigation, pp. 555-564, 1996.
[57]
Knight, M. F., Cervera, M., Finn, A., “A Comparison of Predicted and Measured GPS
Performance in an Ionospheric Scintillation Environment”, Proceedings of ION GPS-99,
The Twelfth International Technical Meeting of the Satellite Division of the
Institute of Navigation, pp. 1437-1450, 1999.
[58]
Kumar, R., Munjal, P., “Simulation Studies for Ionospheric Scintillation”, Proceedings
of the 54th Annual Meeting of the Institute of Navigation, 1998.
[59]
Lindsey, W. C., Charles, F. J., “A Model Distribution for the Phase Error in
Second-Order
Phase-Locked
Loops”,
IEEE
Transactions
on
Communication
Technology, vol. COM-14, pp. 662-664, 1966.
[60]
McDonough, R. N., Whalen, A. D., “Detection of Signals in Noise, Second Ed.”,
Academic Press, 1995.
[61]
Miyagaki, Y., Morinaga, N., Namekawa, T., “Error Probability Characteristics for
CPSK Signal Through m-Distributed Fading Channel”, IEEE Transactions on
Communications, vol. COM-26, No. 1, pp. 88-100, 1978.
[62]
Morris, J. M., “Burst Error Statistics of Simulated Viterbi Decoded BPSK on Fading and
Scintillating Channels”, IEEE Transactions on communications, vol. 40, No. 1,
pp. 34-41, 1992.
[63]
Morrissey, T. N., et al, “GPS Receiver Performance Characterization Under Simulated
Ionospheric Scintillation Environments”, Proceedings of the IAIN World Congress,
2000.
[64]
Nakagami, M., “The m-distribution - A general formula of intensity distribution of rapid
fading”, in Statistical Methods in Radio Wave Propagation, W.C. Hoffman Ed.,
Symposium Publications Division, Pergamon Press, 1960.
[65]
Nesenbergs, M., “Binary error probability due to an adaptable fading model”, IEEE
Transactions on Communications, vol. COM-12, pp. 64-73, 1964.
[66]
Nichols, J., Hansen, A., Walter, T., Enge, P., “High Latitude Measurements of
Ionospheric Scintillation Using the NSTB”, Proceedings of the ION 1999 National
Technical Meeting, 1999.
[67]
Nichols, J., Hansen, A., Walter, T., Enge, P., “Observations of Equatorial Scintillations
Using GPS Receivers”, Proceedings of ION GPS-99, The Twelfth International
Technical Meeting of the Satellite Division of the Institute of Navigation,
pp. 1451-1460, 1999.
301
[68]
NovAtel Communications Ltd., “GPSCard Command Description Manual Revision
2.0, Software Version 3.3”, 1995.
[69]
Ossakow, S. L., Chaturvedi, P. K., “Morphological studies of rising equatorial spread F
bubbles”, Journal of Geophysical Research, vol. 83, 1978.
[70]
Parkinson, B. W., “Global Positioning System: Theory and Applications, Volume I”,
Progress in Astronautics and Aeronautics, vol. 163, Chapter 1, pp. 3-28, 1996.
[71]
Peterson, B., Hartnett, R. J., Fiedler, R., Nebrich, A., “Frequency Domain Techniques
for Fast GPS Acquisition and Interference Detection/Rejection”, NAVIGATION, Journal
of the Institute of Navigation, vol. 43, No. 3, pp. 237-255, 1996.
[72]
Peterson, R. L., Ziemer, R. E., Borth, D. E., “Introduction to Spread Spectrum
Communications”, Prentice-Hall, New Jersey, 1995.
[73]
Pullen, S., et al., “A Preliminary Study of the Effect of Ionospheric Scintillation on WAAS
User Availability in Equatorial Regions”, Proceedings of ION GPS-98, Eleventh
International Technical Meeting of the Satellite Division of the Institute of
Navigation, pp. 687-699, 1998.
[74]
Proakis, J. G., “Digital Communications, 2nd Edition”, Mc Graw Hill, New York, 1989.
[75]
Proakis, J. G., Salehi, M., “Communication Systems Engineering”, Prentice Hall, 1994.
[76]
Rino, C. L., “A power law phase screen model for ionospheric scintillation: 1. Weak
scatter”, Radio Science, vol. 14, No. 6, pp. 1135-1145, 1979a.
[77]
Rino, C. L., “A power law phase screen model for ionospheric scintillation: 2. Strong
scatter”, Radio Science, vol. 14, No. 6, pp. 1147-1155, 1979b.
[78]
Rino, C.L., “On the application of phase screen models to the interpretation of ionospheric
scintillation data”, Radio Science, vol. 17, No. 4, pp. 855-867, 1982.
[79]
Rino, C. L., Gonzalez, V. H., Hessing, A. R., “Coherence bandwidth loss in
transionospheric radio propagation”, Radio Science, vol. 16, No. 2, pp. 245-255, 1981.
[80]
Rishbeth, H., “Basic physics of the ionosphere: a tutorial review”, Journal of the
Institution of Electronic and Radio Engineers, vol. 58, No. 6, pp. S207-S223, 1988.
[81]
Rockwell International Corporation, Satellite Systems Division, “ICD-GPS-200,
NAVSTAR GPS Space Segment / Navigation User Interfaces”, 1987.
[82]
Secan, J. A., Bussey, R. M., Fremouw, E. J., Basu, S., “An improved model of equatorial
scintillation”, Radio Science, vol. 30, No. 3, pp. 607-617, 1995.
[83]
Shanmugan, K. S., “Modelling and Simulation of Wireless and Satellite Communication
Systems”, Course presented at the Institute for Telecommunications Research, the
University of South Australia, April 2000.
[84]
Simon, M. K., Omura, J. K., Scholtz, R. A., Levitt, B. K., “Spread Spectrum
Communications Handbook, Revised Edition”, McGraw-Hill, 1994.
302
[85]
Spatz, D. E., Franke, S. J., Yeh, K. C., “Analysis and interpretation of spaced receiver
scintillation data recorded at an equatorial station”, Radio Science, vol. 23, No. 3,
pp. 347-361, 1988.
[86]
Spiegel, M. R., “Mathematical Handbook”, Mc Graw-Hill, New York, 1968.
[87]
Spilker, J. J., “Signal Structure and Performance Characteristics”, Global Positioning
System papers published in Navigation, The Institute of Navigation, pp. 29-54,
1980.
[88]
Steele, R., “mobile Radio Communications”, IEEE Press N.Y. and Pentech Press
London.
[89]
Titheridge, J. E., “The diffraction of satellite signals by isolated ionospheric irregularities”,
Journal of Atmospheric and Terrestrial Physics, vol. 33, pp. 47-69, 1971.
[90]
Tsunoda, R. T., Livingston, R. C., McClure, J. P., Hanson, W. B., “Equatorial Plasma
Bubbles: Vertically Elongated Wedges From the Bottomside F Layer”, Journal of
Geophysical Research, vol. 87, No. A11, pp. 9171-9180, 1982.
[91]
Van Dierendonck, A. J., “Understanding GPS Receiver Terminology: A Tutorial”, GPS
World, pp. 34-44, January 1995.
[92]
Van Dierendonck, A. J., “Global Positioning System: Theory and Applications, Volume
I”, Progress in Astronautics and Aeronautics, vol. 163, Chapter 8, pp. 329-407, 1996.
[93]
Van Dierendonck, A. J., Hua, Q., Fenton, P., Klobuchar, J. A., “Commercial
Ionospheric Scintillation Monitoring Receiver Development and Test Results”,
Proceedings of the 52nd Annual Meeting, The Institute of Navigation, pp. 573-582,
1996.
[94]
Van Dierendonck, A. J., Reddan, P., Nicholson, M. J., “GPS Receiver Performance
Characterization under Simulated Ionospheric Scintillation Environments”, Proceedings
of the Ionospheric Effects Symposium - IES 99, pp. 3A7-1 to 3A7-8, 1999.
[95]
Van Trees, H. L., “Detection, Estimation , and Modulation Theory, Part I”, John Wiley
& sons, New York, 1968.
[96]
Van Trees, H. L., “Detection, Estimation , and Modulation Theory, Part II”, John Wiley
& sons, New York,
[97]
Viterbi, A. J., “Principles of Coherent Communications”, McGraw-Hill, New York,
1966.
[98]
Wanninger, L., “Effects of the Equatorial Ionosphere on GPS”, GPS World, pp. 48-54,
1993.
[99]
Ward, P. W., “Performance Comparisons Between FLL, PLL and a Novel FLL-AssistedPLL Carrier Tracking Loop Under RF Interference Conditions”, Proceedings of ION
303
GPS-98, The Eleventh International Technical Meeting of the Satellite Division of
the Institute of Navigation, pp. 783-795, 1998.
[100] Weber, W. J., “Performance of phase-locked loops in the presence of fading
communications channels”, IEEE Transactions on Communications, vol. COM-24,
No. 5, pp. 487-499, 1976.
[101] Whitney, H. E., Aarons, J. A., Allen, R. S., Seemann, D. R., “Estimation of the
cumulative amplitude probability distribution function of ionospheric scintillations”,
Radio Science, vol. 7, No. 12, pp. 1095-1104, 1972.
[102] Wojnar, A. H., “Unknown bounds on performance in Nakagami channels”, IEEE
Transactions on communications, vol. COM-34, No. 1, pp. 22-24, 1986.
[103] Yeh, K. C., Liu, C. H., “Radio wave scintillation in the ionosphere”, Proceedings of the
IEEE, vol. 70, pp. 324-360, 1982.
[104] Yovits, M. C., Jackson, J. L., “Linear Filter Optimization with Game-Theoretic
Considerations”, IRE National Convention Record, part 4, pp. 193-199, 1955.
304