Download as pdf or txt
Download as pdf or txt
You are on page 1of 353

COMPUTER MODELLING OF

ELECTRICAL POWER SYSTEMS


Second Edition

J Arrillaga and N R Watson


University of Canterbury, Christchurch, New Zealand

JOHN WILEY & SONS, LTD


Chichester New York Weinheim Brisbane Singapore Toronto
Copyright Q 2001 by John Wiley & Sons, Ltd
Baffins Lane, Chichester,
West Sussex, PO19 IUD, England
National 01243 179111
International ($44) 1243 119777
e-mail (for orders and customer service enquiries): cs-books@wiley.co.uk
Visit our Home Page on http:Nwww.wiley.co.uk or http://www.wiley.com
All Rights Reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, scanning or
otherwise, except under the terms of the Copyright Designs and Patents Act 1988 or under the terms of a
licence issued by the Copyright Licensing Agency, 90 Tottenham Court Road, London, W l P 9HE, UK,
without the permission in writing of the Publisher, with the exception of any material supplied
specifically for the purpose of being entered and executed on a computer system, for exclusive use by the
purchaser of the publication.
Neither the authors nor John Wiley & Sons Ltd accept any responsibility or liability for loss or damage
occasioned to any person or property through using the material, instructions, methods or ideas contained
herein, or acting or refraining from acting as a result of such case. The authors and Publisher expressly
disclaim all implied warranties, including merchantability of fitness for any particular purpose,
Designations used by companies to distinguish their products are often claimed as trademarks. In all
instances where John Wiley & Sons is aware of a claim, the product names appear in initial capital or
capital letters. Readers, however, should contact the appropriate companies for more complete
information regarding trademarks and registration.
Other Wiley Editorial Ofices
John Wiley & Sons, Inc., 605 Third Avenue,
New York, NY 10158-0012, USA
Wiley-VCH Verlag GmbH
Pappelallee 3, D-69469 Weinheim, Germany
Jacaranda Wiley Ltd, 33 Park Road, Milton,
Queensland 4064, Australia
John Wiley & Sons (Canada) Ltd, 22 Worcester Road
Rexdale, Ontario, M9W IL1, Canada
John Wiley & Sons (Asia) R e Ltd, 2 Clementi Loop #02-01,
Jin Xing Distripark, Singapore 129809

British Library Cataloguing in Publicadon Data


A catalogue record for this book is available from the British Library
ISBN 0 471 81249 0
Typeset in 10/12pt Times by Laser Words, Chennai, India
PREFACE

This book describes the use of component models and efficient computational tech-
niques in the development of computer programs representing the steady and dynamic
states of electrical power systems.
The content, although directed to practising engineers, should also be appropriate
for advanced power system courses at final year degree and masters levels.
Some basic knowledge of power system theory, matrix analysis and numerical tech-
niques is presumed, and specific references are given to help the uninitiated to pick up
the relevant material.
An introductory chapter describes the main computational and transmission system
developments justifying the purpose of the book. This is followed by two chapters
describing the modelling of power transmission systems, one on conventional plant
components and the other on FACTS devices and HVDC links.
A general-purpose single-phase a.c. load-flow program is described in Chapter 4 with
particular reference to the Newton Fast-Decoupled algorithm. The load-flow subject
is extended in Chapter 5 with the description of algorithms for the incorporation of
FACTS and HVDC transmission in conventional programs.
The remaining chapters consider the power system in the dynamic state. Chapter 6
covers the subject of electromagnetic transients with reference to the EMTP method,
and provides examples of its application to fast transients and general power system
disturbances. Attention is also given to the representation of power electronic compo-
nents.
Chapter 7 describes electromechanical models of a.c. power system plant and their
use in multi-machine transient stability studies.
Finally, a combination of the electromechanical and electromagnetic models
described in Chapters 6 and 7 is presented in Chapter 8 for the assessment of Transient
Stability in systems containing HVDC links and/or FACTS devices.
The authors should like to acknowledge the considerable help received from
their earlier and present colleagues, both at UMIST (Manchester) and University
of Canterbury (New Zealand). In particular they wish to single out E Acha,
G Anderson, C P Arnold, G Bathurst, P S Bodger, A Brameller, H W Dommel,
B J Harker, M D Heffernan, B C Smith, B Stott, K S Turner and A R Wood. They
also wish to thank Mrs G M Arrillaga for her active participation in the preparation of
the manuscript.

J Arrillaga
N R Watson
INDEX

Index Terms Links

ABCD parameters 29
a.c.-d.c. converters 62 192
a.c.-d.c. load flow
sequential 147 148 153
unified 142 151
Acha’s formulation 33
admittance matrix 5
compound admittances 13
formation rules 17
geometrical 34
nodal 83
primitive 6 31 40
system 51
advanced series compensator 54
aluminium conductor steel reinforced
conductors 35
automatic voltage regulators 237 261
axes 231

backward Euler method 254 255


benchmark model 359
Bergeron line model 162 167 211 212 213
Bessel functions 34
branch switching 250

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

bus conditions 86

cage factor 286


capacitance 164
Carson’s technique 32 166
chatter 188
CIGRE HVDC benchmark model 359
commutation overlap angle 64
commutation reactance 64 68
compensation method 181
complex penetration 35 36
compound admittances 13
computer systems
graphical interface 2 203 206
languages 2
memory 2
simulation engine 2
software 2
conduction overlap angle 64
conductor impedance 34
constant current control 71
constant power control 73
converters
a.c.-d.c. 62 192
model 137
multiple 149
quasi steady-state simulation 325
d.c. link 330
network representation 334
rectifier loads 325

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

converters (Cont.)
in transient stability program 339
terminal voltage control 147
three-phase model 74
coupled three-phase lines 24
current-source representation 181
curve fitting 175 310
effectiveness 313

damping coefficient 232


d.c. convergence tolerance 151
d.c. link 330
control 69
constant current 71
constant power 73
extinction angle 71
power-frequency 74
network representation 339
d.c. per unit system 138
d.c. power modulation 332
decoupled Newton load flow 98
decoupled three-phase algorithm 117
delta-delta transformer 9
difference equation, exponential form 190 191
direct axis 231
distorted waveforms, data extraction from 310
Dommel’s method (numerical integrator
substitution) 161 162
double-cage rotor model 286
Dubanton’s formulae 33
dynamic ordering 96

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

earth impedance 31
earth return impedance 38
electromagnetic transients (EMTP) 1 161
and d.c., see EMTDC
Dommel’s’ method (numerical integrator
substitution) 161 162
nodal equations 179
PSCADEMTDC program 202
fast transient example 211 214 216
frequency-dependent transmission line 211
GTO test system 207 209
STATCOM 215 217
structure 202
subsynchronous resonance 218
test cases 207 209
version 3: 204 206 208
real time digital simulation 219
root matching 190
series impedance approximation 36
state variables 221
choice 224
formulation 221
procedure 222
subsystems 183
switching discontinuities 186
chatter 188
firing errors 186
synchronous machines 195
transformers 199

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

electromagnetic transients (EMTP) (Cont.)


transmission lines and cables 166
Bergeron line model 162 167 211 212 213
frequency-dependent 173 211
multi-conductor 170
PI section model 166
electromechanical transients 161
EMTDC 202 203 205
TS data transfer 306
TS/EMTDC interface 300
equivalent circuits component 300
interface variables derivation 304
location 315
method 313
equivalent circuits converter 137
induction motors
double cage 287
single cage 285
Norton capacitor 165
inductor 163
synchronous machines 247
rectifier load 326
single conductor line 174
subsystems 185
synchronous machines 197 247
Thevenin
fault 303
power flow 303
prefault 348
thyristor and snubber 193
thyristor controlled series compensator 55

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

equivalent circuits converter (Cont.)


transformer 8
star-delta 45
two-winding 57
TS/EMTDC interface 300
two-coupled windings 200
unified power flow control 62
equivalent-pi 7 22 245
extinction angle control 71

FACTS devices 3 53
in load flow solutions 129
phase shifter 58 130
static compensator (STATCOM) 60
static tap changing 56 130
static VAR compensator 59
thyristor controlled series compensator 54 131
unified power flow controller 61 132
fast decoupled load flow 100 123
fast transient analysis 211 214 216
faults 249 345
unbalanced 293
flexible a.c. transmission systems see FACTS
devices
Fortescue technique 11
frame of reference
synchronous machines 231
three-phase system analysis 11
frequency-dependent model, transmission
lines 173 211

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

geometrical admittance 34
geometrical impedance 33
geometrical mean radius 33
governor power flow 110
Graetz bridge 62
GTO test system 207 209

high voltage direct current transmission


(HVDC) 3 62
a.c.-d.c. converter 62
CIGRE benchmark model 359
commutation reactance 64 68
d.c. link control 69
constant current 71
constant power 73
extinction angle 71
power-frequency 74
inversion 67
in load flow solutions 135
converter model 137
converter terminal voltage control 147
d.c. convergence tolerance 151
multiple/multiterminal systems 149
numerical example 155
solution techniques 142
test system and results 151
rectification 63
three-phase model 74

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

hybrid program 298


structure 317
test system 322
hydro turbines 241

impedance
conductor 34
earth 31
earth return 38
geometrical 33
sea return 38
series 18 31 36
inductance 163
induction machines 284
contactors 291
electrical equations 285
large slip 286
equivalent circuits
double cage 287
single cage 285
mechanical equations 284
network representation 288
in transient stability program 288
injected nodal current 345
integration 1 252 351
accuracy 351
backward Euler 254 255
predictor-corrector methods 354
Runge-Kutta 252 356
Runge-Kutta-Gill 252 254

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

integration (Cont.)
stability 352
stiffness 230 353
trapezoidal 1 253
problems 255
programming 256
saturating AVR exciter 261
synchronous machine 258
synchronous machine controller limits 259
inversion 67 141

Jacobian approximations 119


Jacobian matrix 89 90 92 93 95
98 99 117 142 143

lead-lag circuits 242


linear transformation 5
load 243
network representation 249
load flow 1 81
a.c.-d.c.
sequential 147 148 153
unified 142 151
algorithm 87 88
analytical definition 86
convergence criteria and tests 104
decoupled Newton 98
FACTS device incorporation 129

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

load flow (Cont.)


phase shifting 130
static tap changing 130
thyristor controlled series capacitance 131
unified power flow controller 132
fast decoupled 100 123
HVDC transmission incorporation 135
converter model 137
converter terminal voltage control 147
d.c. convergence tolerance 151
multiple/multiterminal systems 149
numerical example and results 155
solution techniques 142
test system and results 151
ideal solution properties 81
Newton-Raphson method 1 82 87
characteristics 97
convergence aids 96
nodal method 82
numerical example 105 106
one voltage known 85
problems faced by 81
stability assessment 105 110
three-phase 110
computer program 123
decoupled three-phase algorithm 117
equation derivation 115
fast decoupled algorithm 123
generator model 123
notation 111
variables specified 115

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

load flow (Cont.)


voltage regulator 113
Y matrix methods 82 84
Z matrix methods 82
long-distance lines 3 22
lumped resistance 169
lumped series impedance 31

machine switching 251


mesh matrix method 245
metal oxide surge arresters 211 214 216
method of companion circuits 162
modal analysis 23
multi-conductor lines 170

negative sequence system 293


Newton-Raphson method 1 82 87
characteristics 97
convergence aids 96
nodal analysis
advantages 5
load flow 82
stability 246
nodal conductance approach 162
nodal equations 179
modification for switching 180
non-linearities 181 184
compensation method 181

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

non-linearities (Cont.)
current-source representation 181
piecewise linear representation 183 184
non-voltage controlled bus 87
Norton equivalent
capacitor 165
inductor 163
synchronous machines 247
numerical integrator substitution 161 162

on-load tap changing 56


ordering, optimal 95
overhead lines 31

Pade approximation 356


Park’s transformation 196
phase-shifting 10 58
in load flow solutions 130
phase-vector type phase-locked oscillator 194
PI section model 166
piecewise linear representation 183 184
Pollaczek’s equations 166
polynomial fitting 175
post-disturbance power flows 105 110
Potier reactance 276
power-frequency control 74
power systems computed aided design
(PSCAD) 203
see also PSCAD/EMTDC program

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

predictor -corrector methods 354


pre-ordering 95
primitive admittance matrix 6 31 40
propagation constant 176 177
propagation function 176
PSCAD/EMTDC program 202
fast transient example 211 214 216
frequency-dependent transmission line 211
GTO test system 207 209
STATCOM 215 217
structure 202
subsynchronous resonance 218
test cases 207 209
version 3 204 206 208

quadrature axis 231


quasi steady-state converter simulation 325
d.c. link 330
network representation 334
rectifier loads 325
in transient stability program 339

rational polynomial fitting 175


real time digital simulation 219
rectification in HVDC transmission 63
rectifier loads 325
dynamic 327
equivalent circuit 326
network representation 334

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

rectifier loads (Cont.)


operating mode identification 328
static 326
relays 289
distance 291
overcurrent
directional 291
instantaneous 289
inverse definite minimum time lag 289
in transient stability program 292
undervoltage 291
resistance 163
RLC branch, Norton equivalent 165
r.m.s. power 306
root matching 190
Runge-Kutta-Gill method 252 254
Runge-Kutta method 252 356

salient machine saturation 273


saturating AVR exciter 261
saturation 199 201 268 271
classical model 271
curve representation 275
effects 276
network representation 277
salient machine 273
simple representation 275
in transient stability program 278
sea return impedance 38
sensitivity analysis 110

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

sequential a.c.-d.c. load flow 147 148 153


series capacitor bank 28
series impedance 18 31 36
short circuit analysis 345
fault calculations 348
system equations 346
shunt admittance 20 31
shunt capacitance 121
shunt capacitor bank 27
single-phase modelling 7
skin effect 34
correction factors 39 40
slack bus 87
software development 2
sparsity programming 94
speed governors 237 239 261
stability 229
advanced component models 268
equations 230
frames of reference 231
hybrid program 298
structure 317
test system 322
induction machines 284
integration 252
load flow for 105 110
loads 243
long-term 229
power electronic control 297
program structure 262 263
relays 289

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

single-phase modelling (Cont.)


system representation 245
faults 249
load 249
mesh matrix 245
nodal matrix 246
switching 250
synchronous machines 246
test system 359
transient 229
transmission network 245
turbines 279
unbalanced faults 293
star-delta transformer 9 45 202
star-star transformer 42
STATCOM (static compensator) 60 215 217
state variables 221
choice 224
formulation 221
procedure 222
static phase shifter 58
static tap changing 56
in load flow 130
static VAR compensator 59 339
stiffness 230 353
submarine cables 36
subsynchronous resonance 218
subsystems 18 183
surge arresters 211 214 216
swing bus 87
switch representations 186

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

single-phase modelling (Cont.)


switching
branch 250
discontinuities 186
chatter 188
firing errors 186
machine 251
synchronous machines
controllers 237
automatic voltage regulators 237 261
lead-lag circuits 242
limits 259
speed governors 237 239
turbines 241
electrical equations 232
electromagnetic transients 195
equivalent circuits 197 247
frame of reference 231
machine models 236
mechanical equations 231
network representation 246
saturation 199 268 271
classical model 271
curve representation 275
effects 276
network representation 277
salient machine 273
simple representation 275
in transient stability program 278
steady state equations 233
subtransient equations 235

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

synchronous machines
controllers (Cont.)
three-phase load flow 111
time constants 230
transient equations 234
trapezoidal integration 258

tap changing 56 130


Taylor’s theorem 88
thermal turbines 241
Thevenin equivalent circuit
fault 303
power flow 303
prefault 348
Thevenin resistance 182
three-phase converter model 74
three-phase load flow 110
computer program 123
decoupled three-phase algorithm 117
equation derivation 115
fast decoupled algorithm 123
generator model 123
notation 111
variables specified 115
voltage regulator 113
three-phase system analysis 11
compound admittances 13
frame of reference 11

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

three-phase transformer models 39


connections 42
independent phase tap control 47
primitive admittance 40
sequence components 48
three-phase transmission lines 18
equivalent-pi 22
mutually coupled 24
series elements 28
series impedance 18
shunt admittance 20
shunt elements 27
terminal connections 26
thyristor controlled phase angle 58
thyristor controlled series capacitor 54
in load flow solutions 131
time constants 230
transformers 199
delta-delta 9
equivalent-pi model 8
off-nominal tap setting 8
single-phase models 8
star-delta 9 45 202
Star-Star 42
three-phase models 39
connections 42
independent phase tap control 47
primitive admittance 40
sequence components 48
Wye Delta
connection models 43
sequence components models 48

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

transient stability 229


hybrid program 298
structure 317
test system 322
program structure 262 263
transmission lines 166
Bergeron model 162 167 211 212 213
frequency-dependent 173 211
long-distance 3 22
multi-conductor 170
overhead 31
PI section model 166
sectionalization 28
single-phase 7
three-phase 18
equivalent-pi 22
mutually coupled 24
series elements 28
series impedance 18
shunt admittance 20
shunt elements 27
terminal connections 26
underground/submarine 36 166
transmission systems 5
admittance matrix 51
nodal analysis, advantages 5
overhead lines 31
single-phase models 7
subsystems 18
three-phase system analysis 11

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

transmission systems (Cont.)


compound admittances 13
frame of reference 11
underground/sumarine cables 36
trapezoidal integration 1 253
problems 255
programming 256
saturating AVR exciter 261
synchronous machine 258
controller limits 259
trapezoidal rule 162
triangular factorization 95
TSE hybrid 298
structure 317
test system 322
TS/EMTDC interface 300
equivalent circuits component 300
interface variables derivation 304
location 315
method 313
turbines 241 279

underground cables 36 166


unified a.c.-d.c. load flow 142 151
unified power flow controller 61
in load flow solutions 132

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

valve group model 192 193


firing control 194
VAR compensator, static 59 339
voltage
controlled bus 87
converter terminal 147
low value 244
regulators, automatic 237 261

waveform distortion, data extraction from 310


Wye Delta transformers
connection models 43
sequence components models 48

Y matrix methods 82 84

Z matrix methods 82
zero sequence system 294

This page has been reformatted by Knovel to provide easier navigation.


CONTENTS

Preface xi

1 Introduction
1.1 General Background
1.2 The New Computer Environment
1.3 Transmission System Developments
1.4 Theoretical Models and Computer Programs

2 Transmission Systems 5
2.1 Introduction 5
2.2 Linear Transformation Techniques 5
2.3 Basic Single-phase Modelling 7
2.3.1 Transmission lines 1
2.3.2 Transformer on nominal ratio 8
2.3.3 Off-nominal transformer tap representation 9
2.3.4 Phase-shifting representation 10
2.4 Three-phase System Analysis 11
2.4.1 Discussion of the frame of reference 11
2.4.2 The use of compound admittances 13
2.4.3 Rules for forming the admittance matrix of simple networks 17
2.4.4 Network subdivision 18
2.5 Three-phase Models of Transmission Lines 18
2.5.1 Series impedance 18
2.5.2 Shunt admittance 20
2.5.3 Equivalent IT model 22
2.5.4 Mutually coupled three-phase lines 24
2.5.5 Consideration of terminal connections 26
2.5.6 Shunt elements 27
2.5.7 Series elements 28
2.5.8 Line sectionalization 28
2.6 Evaluation of Overhead Line Parameters 31
2.6.1 Earth impedance matrix [Z,] 31
2.6.2 Geometrical impedance matrix [Z,] and admittance matrix [Yg] 33
2.6.3 Conductor impedance matrix [Zc] 34
2.6.4 Series impedance approximation for electromagnetic transients 36
vi CONTENTS

2.7 Underground and Submarine Cables 36


2.8 Three-phase Models of Transformers 39
2.8.1 Primitive admittance model of three-phase transformers 40
2.8.2 Models for common transformer connections 42
2.8.3 Three-phase transformer models with independent phase tap control 47
2.8.4 Sequence components modelling of three-phase transformers 48
2.9 Formation of the System Admittance Matrix 51
2.10 References 51

3 FACTS and HVDC Transmission 53


3.1 Introduction 53
3.2 Flexible a.c. Transmission Systems 53
3.2.1 Thyristor controlled series compensator (TCSC) 54
3.2.2 Static on-load tap changing 56
3.2.3 Static phase shifter 58
3.2.4 Static VAR compensator 59
3.2.5 The static compensator (STATCOM) 60
3.2.6 Unified power flow controller (UPFC) 61
3.3 High Voltage Direct Current Transmission 62
3.3.1 The a.c.-d.c. converter 62
3.3.2 Commutation reactance 68
3.3.3 d.c. link control 69
3.3.4 Three-phase model 74
3.4 References 79

4 Load Flow 81
4.1 Introduction 81
4.2 Basic Nodal Method 82
4.3 Conditioning of Y Matrix 84
4.4 The Case Where One Voltage is Known 85
4.5 Analytical Definition of the Problem 86
4.6 Newton-Raphson Method of Solving Load Flows 87
4.6.1 Equations relating to power system load flow 89
4.7 Techniques Which Make the Newton-Raphson Method
Competitive in Load Flow 94
4.7.1 Sparsity programming 94
4.7.2 Triangular factorization 95
4.7.3 Optimal ordering 95
4.7.5 Aids to convergence 96
4.8 Characteristics of the Newton-Raphson Load Flow 97
4.9 Decoupled Newton Load Flow 98
4.10 Fast Decoupled Load Flow 100
4.11 Convergence Criteria and Tests 104
4.12 Numerical Example 105
4.13 Load How for Stability Assessment 105
4.13.1 Post-disturbance power flows 105
CONTENTS vii

4.13.2 Modelling techniques 110


4.13.3 Sensitivity analysis 110
4.14 Three-phase Load Flow 110
4.14.1 Notation 111
4.14.2 Synchronous machine modelling 111
4.14.3 Specified variables 115
4.14.4 Derivation of equations 115
4.14.5 Decoupled three-phase algorithm 1I7
4.14.6 Structure of the computer program 123
4.15 References 127

5 Load Flow under Power Electronic Control 129


5.1 Introduction 129
5.2 Incorporation of FACTS Devices 129
5.2.1 Static tap changing 130
5.2.2 Phase-shifting (PS) 130
5.2.3 Thyristor controlled series capacitance (TCSC) 131
5.2.4 Unified power flow controller (UPFC) 132
5.3 Incorporation of HVDC Transmission 135
5.3.1 Converter model 137
5.3.2 Solution techniques 142
5.3.3 Control of converter ax. terminal voltage 147
5.3.4 Extension to multiple and/or multiterminal d.c. systems 149
5.3.5 d.c. convergence tolerance 151
5.3.6 Test system and results 151
5.3.7 Numerical example 155
5.4 References 158

6 Electromagnetic Transients 161


6.1 Introduction 161
6.2 Background and Definitions 162
6.3 Numerical Integrator Substitution 162
6.3.1 Resistance 163
6.3.2 Inductance 163
6.3.3 Capacitance 164
6.4 Transmission Lines and Cables 166
6.4.1 Bergeron line model 167
6.4.2 Multi-conductor transmission lines 170
6.4.3 Frequency-dependent model 173
6.5 Formulation and Solution of the System Nodal Equations 179
6.5.1 Modification for switching and varying parameters 180
6.5.2 Non-linear or time varying parameters 181
6.6 Use of Subsystems 183
6.7 Switching Discontinuities 186
6.7.1 Voltage and current chatter due to discontinuities 188
6.8 Root-matching Technique 190
6.8.1 Exponential form of difference equation 190
viii CONTENTS

6.8.2 Root-matching implementation 191


6.8.3 Numerical illustration 191
6.9 a.c./d.c. Converters 192
6.10 Synchronous Machine Model 195
6.1 1 Transformer Model 199
6.12 The PSCADIEMTDC Program 202
6.12.1 Structure of the program 202
6.12.2 PSCADEMTDC Version 3 204
6.12.3 PSCADEMTDC test cases 207
6.13 Real Time Digital Simulation 219
6.14 State Variable Analysis 22 1
6.14.1 State variable formulation 22 1
6.14.2 Solution procedure 222
6.14.3 Choice of state variables 224
6.15 References 225

7 System Stability 229


7.1 Introduction 229
7.1.1 The form of the equations 230
7.1.2 Frames of reference 23 1
7.2 Synchronous Machines-Basic Models 23 1
7.2.1 Mechanical equations 23 1
7.2.2 Electrical equations 232
7.3 Synchronous Machine Automatic Controllers 237
7.3.1 Automatic voltage regulators 237
7.3.2 Speed governors 239
7.3.3 Hydro and thermal turbines 24 1
7.3.4 Modelling lead-lag circuits 242
7.4 Loads 243
7.4. I Low-voltage problems 244
7.5 The Transmission Network 245
7.6 Overall System Representation 245
7.6.1 Mesh matrix method 245
7.6.2 Nodal matrix method 246
7.6.3 Synchronous machine representation in the network 246
7.6.4 Load representation in the network 249
7.6.5 System faults and switching 249
7.7 Integration 252
7.7.1 hoblems with the trapezoidal method 255
7.7.2 hogramming the trapezoidal method 256
7.7.3 Application of the trapezoidal method 258
7.8 Structure of a Transient Stability Program 263
7.8.1 Overall structure 263
7.8.2 Structure of machine and network iterative solution 264
7.9 Advanced Component Models 268
7.9.1 Synchronous machine saturation 268
7.9.2 Detailed turbine model 279
CONTENTS ix

7.9.3 Induction machines 284


7.9.4 Relays 289
7.9.5 Unbalanced faults 293
7.10 References 295

8 System Stability under Power Electronic Control 297


8.1 Introduction 291
8.2 Description of the Algorithm 298
8.2.1 Data flow 299
8.2.2 Modifications required to the component programs 300
8.3 TSEMTDC Interface 300
8.3.1 Equivalent circuit components 300
8.3.2 Interface variables derivation 304
8.4 EMTDC to TS Data Transfer 306
8.5 Data Extraction from Distorted Waveforms 310
8.5.1 CFA effectiveness 313
8.6 Interface Method 313
8.7 Interface Location 315
8.8 Structure of the Hybrid Program 317
8.9 Test System and Results 322
8.9.1 Response of the individual programs 322
8.9.2 TSE hybrid response 323
8.10 Quasi Steady-state Converter Simulation 325
8.10.1 Rectifier loads 325
8.10.2 d.c. link 330
8.10.3 Representation of converters in the network 334
8.10.4 Inclusion of converters in the transient stability program 339
8.1 1 Static VAR Compensation Systems 339
8.1 1.1 Representation of SVS in the overall system 342
8.12 References 343

Appendix I Fault Level Derivation 345


1.1 Short Circuit Analysis 345
I. 1.1 System equations 346
I. 1.2 Fault calculations 348

Appendix I1 Numerical Integration Methods 351


11.1 Introduction 35 1
11.2 Properties of the Integration Methods 35 1
11.2.1 Accuracy 35 1
11.2.2 Stability 352
11.2.3 Stiffness 353
11.3 Predictor-Corrector Methods 354
11.4 Runge-Kutta Methods 356
11.5 References 357
X CONTENTS

Appendix I11 Test System used in the Stability Examples 359


111.1 Reference 362

Index 363
I
INTRODUCTION

1.1 General Background

The first edition of this book described the development of FORTRAN based power
system programs for implementation in the mainframe computer technology of the
1970s. Since then, the ubiquitous personal computer (PC) has become the workhorse
of most power system engineers and the new edition takes this fact into account.
Nevertheless, the basic algorithms have remained largely unchanged; what has altered
in the past two decades is the general acceptance of power electronic devices for the
control of power flow and its related stability improvement. Both high voltage d.c. and
FACTS technologies are now an integral part of modern power transmission systems.
Their incorporation in power system analysis is not straightforward, however, and this
new edition describes suitable models of these devices.
The primary subject of computer modelling is the load flow problem, which finds
application in all phases of power system analysis. Due to space limitations, only the
solution of the basic load flow equations is normally considered. It is acknowledged,
however, that the load flow problem is not restricted to the solution of the basic
continuously differentiable equations. There is probably not a single routine program
in use anywhere that does not model other features. Such features often have more
influence on convergence than the performance of the basic algorithm.
The most successful contribution to the load flow problem still is the application
of Newton-Raphson and derived algorithms. These were finally established with the
development of programming techniques for the efficient handling of large matrices
and, in particular, the sparsity-oriented ordered-elimination methods. The Newton algo-
rithm was first enhanced by taking advantage of the decoupling characteristics of power
flow and, finally, by the use of reasonable approximations directed towards the use of
constant Jacobian matrices.
In dynamic studies, a significant modelling tool has been the application of implicit
integration techniques which allow the differential equations to be algebraized and then
incorporated with the network algebraic equations to be solved simultaneously. The use
of implicit trapezoidal integration has proved to be very stable, permitting step lengths
greater than the smallest time constant of the system. This technique allows detailed
representation of synchronous machines with their voltage regulators and governors,
induction motors and non-linear loads, such as FACTS and HVDC converters.
Trapezoidal integration is also an integral part of the Electromagnetic Transient
Programs (EMTP). With the availability of cheap computer power the EMTP has
2 1 INTRODUCTION

become universally accepted for the solution of all sorts of transient problems and
constitutes the main addition to the present edition.
In particular, electromagnetic transient programs are now used extensively for the
analysis of system disturbances and in power system protection design. The availability
of fast EMTP solutions, and particularly the possibility of Real Time Digital Simu-
lators (RTDS) to interface with real protection hardware, has diminished the use of
conventional fault simulation, based on quasi-steady-state linear system behaviour. The
increasing presence of power electronic devices, for which conventional fault simula-
tion is totally inadequate, has added to the demise of the earlier programs. However,
the Fault Level (MVAf) at points of common coupling and the Short Circuit Ratio
(SCR) at nodes with converter equipment are still derived from Fault Simulation. Thus
in the new edition a concise version of the earlier Faulted System Studies chapter has
been included as an appendix.

1.2 The New Computer Environment


Earlier implementations of power system programs were severely restricted by the
lack of flexibility of mainframe computers as well as limitations in graphical support,
memory and storage space. Now the evolution of computer technology has removed
most of these limitations, and made the PC a universal platform for power system
simulation.
Abundance of cheap memory and new operating systems permit the computers to
accommodate large executable code with extensive data storage. In addition, the utilisa-
tion of virtual memory technology (paging to hard disks) has eliminated the traditional
limitation of insufficient memory space for executable binary and simulation data. The
dynamic memory allocation facility has also enabled optimised usage of the memory
space. Application code can be made relocatable and dynamically loaded into the
memory and called from the main program only when it is needed. Therefore, memory
space is released immediately when it is no longer required.
Traditionally, power system programs have been designed using a single program-
ming language, Fortran being the dominant language. However, there is little graphic
support available through Fortran and, hence, many power system programs lack
a graphical user interface. Therefore, a text-based database file is commonly used,
allowing users to construct the database by typing in the power system data using an
ordinary text editor. Manual database editing has the advantage that data errors are
often immediately obvious.
Recent advancements in software development tools have made it possible for the
program binary generated by different compilers on different programming languages
to be able to interact with each other. This enables programmers to exploit the strengths
of different languages in the development of single computer software. The software
can be separated into a graphical user interface and a simulation engine. The graphical
user interface and the simulation engine have different development tools requirements
and, hence, different languages are chosen to construct them. Fortran is retained as the
language of the simulation engine, due mainly to its implicit handling of complex-
number arithmetic. Normally, the simulation engine is also programmed using basic
Fortran commands (i.e. avoiding extended features of some Fortran compilers) so that
it is transportable across different operating systems/platforms.
1.4 THEORETICAL MODELS AND COMPUTER PROGRAMS 3

1.3 Transmission System Developments


Large increases in transmission distances and voltage level, as well as national and
international interconnections, have resulted in a more sophisticated means of active
and reactive power control.
HVDC links and FACTS devices have now been generally accepted to achieve
more flexible power transmission and their incorporation in conventional power system
programs presents a challenge to power system modelling. Moreover, considering the
large power ratings of such devices, their presence exercises considerable influence on
the rest of the system and must be accurately represented.
Whenever possible, any equivalent models used to simulate the power electronic
components should involve traditional power system concepts for easy incorporation
within existing power system programs. However, the number of degrees of freedom of
d.c. power transmission and FACTS is higher and any attempt to model the behaviour
in the more restricted ax. framework will have limited applications.
Long-distance transmission presents voltage and power balancing problems which
can only be accurately assessed by using correspondingly accurate models of power
transmission plant in the phase frame of reference. With reference to power system
disturbances, the behaviour of FACTS and HVDC transmission cannot be accurately
modelled by conventional fault study or stability programs. Disturbances are best
simulated with electromagnetic transient programs. Thus the modelling of linear and
non-linear components, as well as the frequency dependence of transmission lines for
use in the electromagnetic transient programs have received considerable attention in
recent times and are given extensive coverage in this edition of the book.

1.4 Theoretical Models and Computer Programs


The transition from power system analysis to efficient computer programs is a very
laborious exercise. Present commercial programs are the result of considerable skills
and many engineering years. Thus, newcomers to the power system area usually acquire
existing commercial packages rather than building their own. However, they need to
understand their capabilities and limitations to be able to incorporate new components
and more advanced hardware technologies.
Instruction manuals provide full information on the practical structure of the
programs but lack the technical background necessary for the user to perform the
inevitable modifications required in the long run.
It is almost expected that a specialist book of this type should provide a comprehen-
sive survey and comparison of the various conventional alternatives. Such an approach,
although academically satisfying, would detract from the main object of the book and
would occupy invaluable space. Instead, up-to-date modelling techniques generally
recognized as efficient are described from theoretical and practical considerations.
2
TRANSMISSION SYSTEMS

2.1 Introduction
The conventional power transmission system is a complex network of passive compo-
nents, mainly transmission lines and transformers, and its behaviour is commonly
assessed using equivalent circuits consisting of inductance, capacitance and resistance.
This chapter deals with the derivation of these equivalent circuits and with the
formation of the system admittance matrix relating the current and voltage at every
node of the transmission system.
Among the many alternative ways of describing transmission systems to comply
with Kirchhoff s laws, two methods, mesh and nodal analysis, are normally used. The
latter has been found to be particularly suitable for digital computer work, and is almost
exclusively used for routine network calculations.
The nodal approach has the following advantages:
The numbering of nodes, performed directly from a system diagram, is very simple.
0 Data preparation is easy.
The number of variables and equations is usually less than with the mesh method
for power networks.
0 Network crossover branches present no difficulty,
0 Parallel branches do not increase the number of variables or equations.
0 Node voltages are available directly from the solution, and branch currents are
easily calculated.
0 Off-nominal transformer taps can easily be represented.

2.2 Linear Transformation Techniques


Linear transformation techniques are used to enable the admittance matrix of any
network to be found in a systematic manner. Consider, for the purposes of illustration,
the network drawn in Figure 2.1.
Five steps are necessary to form the network admittance matrix by linear transfor-
mation, i.e.
(i) Label the nodes in the original network.
6 2 TRANSMISSION SYSTEMS

Figure 2.1 Actual connected network

(ii) Number, in any order, the branches and branch admittances.


7
(iii) Form the primitive network admittance matrix by inspection.
This matrix relates the nodal injected currents to the node voltages of the primitive
network. The primitive network is also drawn by inspection of the actual network. It
consists of the unconnected branches of the original network with a current equal to the
original branch current injected into the corresponding node of the primitive network.
The voltages across the primitive network branches then equal those across the same
branch in the actual network.
The primitive network for Figure 2.1 is shown in Figure 2.2.
The primitive admittance matrix relationship is:

[YPRIM 1
Off-diagonal terms are present where mutual coupling between branches is present.

(iv) Form the connection matrix [C].

Figure 2.2 Primitive or unconnected network


2.3 BASIC SINGLE-PHASE MODELLING 7

This relates the nodal voltages of the actual network to the nodal voltages of the
primitive network. By inspection of Figure 2.1,

or in matrix form

(v) The actual network admittance matrix which relates the nodal currents to the
voltages by

(2.4)

can now be derived from

which is a straightforward matrix multiplication.

2.3 Basic Single-phase Modelling


Under perfectly balanced conditions, transmission plant can be represented by single-
phase models, the most extensively used being the equivalent-rc circuit.

2.3.1 Transmission lines


In the case of a transmission line, the total resistance and inductive reactance of the
line is included in the series arm of the n-equivalent and the total capacitance to neutral
is divided between its shunt arms.
8 2 TRANSMISSION SYSTEMS

qrF-T[ Y&

Figure 2.3 Transformer equivalent circuit

2.3.2 Transformer on nominal ratio


The equivalent-lr model of a transformer is illustrated in Figure 2.3, where yoc is the
reciprocal of zoc (magnetizing impedance) and ysc is the reciprocal of zsc (leakage
impedance). zsc and zoc are obtained from the standard short-circuit and open-circuit
tests.
This yields the following matrix equation:

I Is I I -Ysc + YW/2 I Ysc I I vs 1


where
ysc is the short-circuit or leakage admittance,
yW is the open-circuit or magnetizing admittance.

The use of a three-terminal network is restricted to the single-phase representation


and cannot be used as a building block for modelling three-phase transformer banks.
The magnetizing admittances are usually removed from the transformer model and
added later as small, shunt-connected admittances at the transformer terminals. In the
per unit system, the model of the single-phase transformer can then be reduced to a
lumped leakage admittance between the primary and secondary busbars.

2.3.3 Off-nominal transformer tap representation


A transformer with turns ratio a interconnecting two nodes i, k can be represented
by an ideal transformer in series with the nominal transformer leakage admittance as
shown in Figure 2.4(a).
If the transformer is on nominal tap (a = l), the nodal equations for the network
branch in the per unit system are:
Iik = yikvi - YikVk (2.7)
Iki = Y i k vk - Y i k vi (2.8)
In this case, l i k = -Iki.
For an off-nominal tap setting and letting the voltage on the k side of the ideal
transformer be V , , we can write
2.3 BASIC SINGLE-PHASE MODELLING 9

Figure 2.4 Transformer with off-nominal tap setting

Iki = yik(Vk - V t ) , (2.10)

lik = --.Iki (2.11)


a
Eliminating V, between Equations (2.9) and (2. lo), we obtain
Yik
Iki = yikvk - -vi,
a
(2.12)

(2.13)

A simple equivalent n circuit can be deduced from Equations (2.12) and (2.13),
the elements of which can be incorporated into the admittance matrix. This circuit is
illustrated in Figure 2.4(b).
The equivalent circuit of Figure 2.4(b) has to be used with care in banks containing
delta-connected windings. In a star-delta bank of single-phase transformer units, for
example, with nominal turns ratio, a value of 1.0 per unit voltage on each leg of the
star winding produces under balanced conditions 1.732 per unit voltage on each leg
of the delta winding (rated line to neutral voltage as base). The structure of the bank
requires in the per unit representation an effective tapping at J?; nominal turns ratio
on the delta side, i.e. a = 1.732.
For a delta-delta or star-delta transformer with taps on the star winding, the equiv-
alent circuit of Figure 2.4(b) would have to be modified to allow for effective taps
to be represented on each side. The equivalent-circuit model of the single-phase unit
can be derived by considering a delta-delta transformer as comprising a delta-star
transformer connected in series (back-to-back) via a zero-impedance link to a star
delta transformer, i.e. star windings in series. Both neutrals are solidly earthed. The
leakage impedance of each transformer would be half the impedance of the equiva-
lent delta-delta transformer. An equivalent per unit representation of this coupling is
shown in Figure 2.5. Solving this circuit for terminal currents:
I’ (V’ - V ” ) y
I,=-=
ff ff

(2.14)

I’ = ?VP
-I, = - Y Y
- -vs, (2.15)
B B2
10 2 TRANSMISSION SYSTEMS

or in matrix form:

(2.16)

These admittance parameters form the primitive network for the coupling between
a primary and secondary coil.

2.3.4 Phase-shifting representation


To cope with phase-shifting, the transformer of Figure 2.5 has to be provided with
a complex turns ratio. Moreover, the invariance of the product VI* across the ideal
transformer requires a distinction to be made between the turns ratios for current and
voltage, i.e.
v I* = -v’I‘*,
PP
or
V p= ( a + jb)V’ = aV’,

I’ I‘
-
a - j b - --
IP = -- a*‘
Thus, the circuit of Figure 2.5 has two different turns ratios, i.e.
aU = a! +j b for the voltages,
and
ai = (II - j b for the currents.
Solving the modified circuit for terminal currents:
I’ (V’ - V”)y
I,=-=
ffi ffi

- (Vplffu - VS/P>Y = -vp


- Y - -v,,
Y (2.17)
Qi Uuai QiS

rp 1 I‘ V 1:0

I
Figure 2.5 Basic equivalent circuit in p.u. for coupling between primary and secondary coils
with both primary and secondary off-nominal tap ratios of ct and /?
2.4 THREE-PHASE SYSTEM ANALYSIS 11

1' Y Y
-1, = - = -vp - -vs. (2.18)
B %B B2

.I;
Thus, the general single-phase admittance of a transformer including phase
shifting is:

[Yl = (2.19)
--
a"B P2
It should be noted that, although an equivalent lattice network similar to that in
Figure 2.5 could be constructed, it is no longer a bilinear network as can be seen from
the asymmetry of y in Equation (2.19). The equivalent circuit of a single-phase phase-
shifting transformer is thus of limited value and the transformer is best represented
analytically by its admittance matrix.

2.4 Three-phase System Analysis


2.4.1 Discussion of the frame of reference
Sequence components have long been used to enable convenient examination of the
balanced power system under both balanced and unbalanced loading conditions.
The symmetrical component transformation is a general mathematical technique
developed by Fortescue whereby any 'system of n vectors or quantities may be
resolved, when n is prime, into n different symmetrical n phase systems'[l]. Any set
of three-phase voltages or currents may therefore be transformed into three symmet-
rical systems of three vectors each. This in itself would not commend the method, and
the assumptions that lead to the simplifying nature of symmetrical components must
be examined carefully.
Consider, as an example, the series admittance of a three-phase transmission line,
shown in Figure 2.6, i.e. three mutually coupled coils. The admittance matrix relates
the illustrated currents and voltages by

where
(2.21)

and

(2.22)
12 2 TRANSMISSION SYSTEMS

l a

Figure 2.6 Admittance representation of a three-phase series element (a) series admittance
element; (b) admittance matrix representation

By the use of the symmetrical components transformation the three coils of


Figure 2.6 can be replaced by three uncoupled coils. This enables each coil to be treated
separately with a great simplification of the mathematics involved in the analysis.
The transformed quantities (indicated by subscripts 012 for the zero, positive and
negative sequences respectively) are related to the phase quantities by
[ V O I Z=
I [ T ~ I -[ '~ a ~ x I t (2.23)
[10121 = [Tsl-'[labc] (2.24)
'
= [Tsl- [yabcl [Ts1 [ VOI21 v (2.25)
where [Ts] is the transformation matrix.
The transformed voltages and currents are thus related by the transformed admittance
matrix,
[ y o 1 2 1 = i ~ s 1 - I[ y a b c l [ ~ s l . (2.26)
Assuming that the element is balanced, we have
Yaa = Ybb = Ycc.
Yab = Ybc = Yea,
= Ycb = Yac,
Yba (2.27)
and a set of invariant matrices [TI exists. Transformation (2.26) will then yield a
diagonal matrix [ y 0 1 2 ] .
In this case, the mutually coupled three-phase system has been replaced by three
uncoupled symmetrical systems. In addition, if the generation and loading may be
2.4 THREE-PHASE SYSTEM ANALYSIS 13

assumed balanced, then only one system, the positive sequence system, has any current
flow and the other two sequences may be ignored. This is essentially the situation with
the single-phase load flow.
If the original phase admittance matrix [yabc] is in its natural unbalanced state,
then the transformed admittance matrix [yo121 is full. Therefore, current flow of one
sequence will give rise to voltages of all sequences, i.e. the equivalent circuits for the
sequence networks are mutually coupled. In this case, the problem of analysis is no
simpler in sequence components than in the original phase components and symmetrical
components should not be used.
From the above considerations, it is clear that the asymmetry inherent in all power
systems cannot be studied with any simplification by using the symmetrical component
frame of reference. Data in the symmetrical component frame should only be used when
the network element is balanced, for example synchronous generators.
In general, however, such an assumption is not valid. Unsymmetrical interphase
coupling exists in transmission lines and to a lesser extent in transformers, and this
results in coupling between the sequence networks. Furthermore, the phase shift intro-
duced by transformer connections is difficult to represent in sequence component
models.
With the use of phase co-ordinates the following advantages become apparent:

(i) Any system element maintains its identity.


(ii) Features such as asymmetric impedances, mutual couplings between phases
and between different system elements, and line transpositions are all readily
considered.
(iii) Transformer phase shifts present no problem.

2.4.2 The use of compound admittances


When analysing three-phase networks, where the three nodes at a busbar are always
associated together in their interconnections, the graphical representation of the network
is greatly simplified by means of compound admittances, a concept which is based on
the use of matrix quantities to represent the admittances of the network.
The laws and equations of ordinary networks are all valid for compound networks
by simply replacing single quantities by appropriate matrices [2].
Consider six mutually coupled single admittances, the primitive network of which
is illustrated in Figure 2.7.

Figure 2.7 Primitive network of six coupled admittances


14 2 TRANSMISSION SYSTEMS

The primitive admittance matrix relates the nodal injected currents to the branch
voltages as follows:

(2.28)

6x1
Partitioning Equation (2.28) into 3 x 3 matrices and 3 x 1 vectors, the equation
becomes:

where
--- (2.29)

(2.30)

Graphically, we represent this partitioning as grouping the six coils into two
compound coils (a) and (b), each composed of three individual admittances. This is
illustrated in Figure 2.8.
On examination of [Y&] and [ Y b a ] , it can be seen that
[Ybal = EY,bIT*
if, and only if, Yi& = y&i for i = 1 to 3 and k = 4 to 6. That is if, and only if, the
couplings between the two groups of admittances are bilateral.
In this case, Equation (2.29) may be written

(2.31)
2.4 THREE-PHASE SYSTEM ANALYSIS 15

Figure 2.8 Two coupled compound admittances

Figure 2.9 Sample network represented by single admittances

[YAl

Figure 2.10 Sample network represented by compound admittances

The primitive network for any number of compound admittances is formed in exactly
the same manner as for single admittances, except in that all quantities are matrices of
the same order as the compound admittances.
The actual admittance matrix of any network composed of the compound admittances
can be formed by the usual method of linear transformation; the elements of the
connection matrix are now n x n identity matrices where n is the dimension of the
compound admittances.
16 2 TRANSMISSION SYSTEMS

Figure 2.11 Primitive networks and corresponding admittance matrices. (a) Primitive network
using single admittances; (b) Primitive admittance matrix; (c) Primitive network using compound
admittances; (d) Primitive admittance matrix

If the connection matrix of any network can be partitioned into identity elements of
equal dimensions greater than one, the use of compound admittances is advantageous.
As an example, consider the network shown in Figures 2.9 and 2.10, which repre-
sents a simple line section. The admittance matrix will be derived using single and
compound admittances to show the simple correspondence. The primitive networks
and associated admittance matrices are drawn in Figure 2.1 1. The connection matrices
for the single and compound networks are illustrated by Equations (2.32) and (2.33),
2.4 THREE-PHASE SYSTEM ANALYSIS 17

respectively.

(2.32)

(2.33)

Th exact equivalence, with appropriate matrix partitionin is clear.


The network admittance matrix is given by the linear tran,xmation equation,

This matrix multiplication can be executed using the full matrices or in partitioned
form. The result in partitioned form is

I I I

2.4.3 Rules for forming the admittance matrix of simple


networks
The method of linear transformation may be used to obtain the admittance matrix of
any network. For the special case of networks where there is no mutual coupling,
simple rules may be used to form the admittance matrix by inspection. These rules,
which apply to compound networks with no mutual coupling between the compound
admittances, may be stated as follows:

(a) Any diagonal term is the sum of the individual branch admittances connected to
the node corresponding to that term.
(b) Any off-diagonal term is the negated sum of the branch admittances which are
connected between the two corresponding nodes.
18 2 TRANSMISSION SYSTEMS

2.4.4 Network subdivision


To enable the transmission system to be modelled in a systematic, logical and conve-
nient manner, the system must be subdivided into more manageable units. These units,
called subsystems, are defined as follows: A subsystem is the unit into which any
part of the system may be divided such that no subsystem has any mutual couplings
between its constituent branches and those of the rest of the system. This definition
ensures that the subsystems may be combined in an extremely straightforward manner.
The system is first subdivided into the most convenient subsystems consistent with
the definition above.
The most convenient unit for a subsystem is a single network element. In previous
sections, the nodal admittance matfix representation of all common elements has been
derived.
The subsystem unit is retained for input data organization. The data for any subsystem
is input as a complete unit, the subsystem admittance matrix is formulated and stored
and then all subsystems are combined to form the total system admittance matrix.

2.5 Three-phase Models of Transmission Lines


Transmission line parameters are calculated from the line geometrical characteristics.
The calculated parameters are expressed as a series impedance and shunt admittance
per unit length of line. The effects of ground currents and earth wires are included in
the calculation of these parameters [3,43.

2.5.1 Series impedance


A three-phase transmission line with a ground wire is illustrated in Figure 2.12(a). The
following equations can be written for phase a
Va - Vh = la(& +j w L ) +I b ( j w L b ) +I c ( j o L c ) (2.34)
+ jwLgIg- j U L m I n + Vn,
Vn = In(&+ j w & ) - I, j w L 0 - I b j W L t b - I, j w k c - Igj w k B , (2.35)
and substituting
In =Ia+Ib+Ic+Ig, (2.36)
Va - V i = la(& + j w L ) + Ib jwLb + Ic jwLC
+ j W L g l g - j W L n ( & + Ib + IC + I g ) + v n . (2.37)
Regrouping and substituting for V,, i.e.
A V , = Va - V:

(2.38)
2.5 THREE-PHASE MODELS OF TRANSMISSION LINES 19

Figure 2.12 (a) Three-phase transmission series impedance equivalent; (b) Three-phase trans-
mission shunt impedance equivalent

and writing similar equations for the other phases, the following matrix equation results:

(2.41)

Since we are interested only in the performance of the phase conductors, it is


more convenient to use a three-conductor equivalent for the transmission line. This
20 2 TRANSMISSION SYSTEMS

is achieved by writing matrix Equation (2.41) in partitioned form as follows:

(2.42)

(2.43)
(2.44)
From Equations (2.42) and (2.43), and assuming that the ground wire is at zero
potential:
A vabc = Zabclabc 3 (2.45)

(2.46)

2.5.2 Shunt admittance


With reference to Figure 2.12(b), the potentials of the line conductors are related to
the conductor charges by the matrix Equation [3]:

(2.47)

Similar considerations as for the series impedance matrix lead to


Vabc = PLbcQabc 9 (2.48)
where Pbbc is a 3 x 3 matrix which includes the effects of the ground wire. The
capacitance matrix of the transmission line of Figure 2.12 is given by

(2.49)

The series impedance and shunt admittance lumped-n model representation of the
three-phase line is shown in Figure 2.13(a) and its matrix equivalent is illustrated
2.5 THREE-PHASE MODELS OF TRANSMISSION LINES 21

Figure 2.13 Lumped-7r model of a short-three phase line series impedance. (a) Full circuit
representation; (b) Matrix equivalent; (c) Using three-phase compound admittances

in Figure 2.13(b). These two matrices can be represented by compound admittances,


(Figure 2.13(c)), as described earlier.
Following the rules developed for the formation of the admittance matrix using the
compound concept, the nodal injected currents of Figure 2.13(c) can be related to the
nodal voltages by the equation:

I, I

(2.50)
1 I

6x1 6x6 6x1


22 2 TRANSMISSION SYSTEMS

This forms the element admittance matrix representation for the short line between
busbars i and k in terms of 3 x 3 matrix quantities.
This representation may not be accurate enough for electrically long lines. The
physical length at which a line is no longer electrically short depends on the wavelength;
therefore, if harmonic frequencies are being considered, this physical length may be
quite small. Using transmission line and wave propagation theory, more exact models
may be derived [5,6].However, for normal mains frequency analysis, it is considered
sufficient to model a long line as a series of two or three nominal-n sections.

2.5.3 Equivalent a model


For long lines a number of nominal n models are connected in series to improve
the accuracy of voltages and currents, which are affected by standing wave effects.
For example, a three-section IT model provides an accuracy to 1.2% for a quarter
wavelength line (a quarter wavelength corresponds with 1500 and 1250 km at 50 and
60 Hz, respectively).
As the frequency increases, the number of nominal IT sections to maintain a particular
accuracy increases proportionally, e.g. a 300 km line requires 30 nominal IT sections
to maintain the 1.2% accuracy for the 50th harmonic. However, near resonance the
accuracy departs significantly from an acceptable value.
The computational effort can be greatly reduced and the accuracy improved with
the use of an equivalent r model derived from the solution of the second order linear
differential equations describing wave propagation along transmission lines [ 5 ] .
The solution of the wave equations at a distance x from the sending end of the
line is:
V(X) = exp(-p)vi + exp(yx>vr, (2.5 1)
I(X) = (Z’)-’y[eXp(-yx)Vi - exp(yx)vr~, (2.52)
where y = = (Y + j/3 is the propagation constant, Z’ = r + j 2 n f L is the series
+
impedance per unit length, Y’ = g j 2 n f C is the shunt admittance per unit length,
and Vi and Vr the forward and reverse travelling voltages, respectively.
Depending on the problem in hand, e.g. if the evaluation of terminal quantities
only is required, it is more convenient to formulate a solution using two-port matrix
equations. This leads to the equivalent n model, shown in Figure 2.14, where
Z = Z, sinh(yf), (2.53)

Y*=Y2=-
1 cosh(yf) - 1
Z,, sinh(yf)
1
= - tanh
Z,
):( , (2.54)

and
z,= (2.55)

is the characteristic impedance of the line.


Due to the standing wave effect of voltages and currents on transmission lines, the
maximum values of these are likely to occur at points other than at the receiving end or
sending end busbars. These local maxima could result in insulation damage, overheating
2.5 THREE-PHASE MODELS OF TRANSMISSION LINES 23

Figure 2.14 The equivalent j7 model of a long transmission line

or electromagnetic interference. It is thus important to calculate the maximum values


of currents and voltages along a line and the points at which these occur.
In the case of multiconductor transmission lines, the nominal n series impedance
and shunt admittance matrices per unit distance, [Z'] and [Y']respectively, are square
and their size is fixed by the number of mutually coupled conductors. The derivation of
the equivalent 77 model for harmonic penetration studies from the nominal n matrices
is similar to that of the single phase lines, except that it involves the evaluation of
hyperbolic functions of the propagation constant which is now a matrix:
[Yl = ([z~l[yw2. (2.56)
There is no direct way of calculating sinh or tanh of a matrix, thus a method using
eigenvalues and eigenvectors, called modal analysis, is employed [6] that leads to
the following expressions for the series and shunt components of the equivalent n
circuit [7]:
[ZIEPM = l[Z'I[Ml - [
sinh(y l )
yl ][MI-', (2.57)

where 1 is the transmission line length, [&pM is the equivalent IT series impedance
matrix, [MI is the matrix of normalized eigenvectors,
sinh( y11)
0 ...
O 1

and y, is the j t h eigenvalue for j / 3 mutually coupled circuits. Similarly

(2.59)

where [ Y ] E P Mis the equivalent n shunt admittance matrix.


Computer derivation of the correction factors for conversion from the nominal IT to
the equivalent n model, and their incorporation into the series impedance and shunt
24 2 TRANSMISSION SYSTEMS

series impedance and


shunt admittance matrices

Calculate the Calculate


diagonal correction
eigenvalues and
Formmatrix matrices of factors and
eigenvectors solution for hyperbolic apply to give
product [ Y'J[Z']
and form (MI acceptable
eigenvalue
calculate functions

Figure 2.15 Structure diagram for calculation of the equivalent n model

admittance matrices, is carried out as indicated in the structure diagram of Figure 2.15.
The LR2 algorithm of Wilkinson and Reinsch [8] is used for accurate calculations in
the derivation of the eigenvalues and eigenvectors.

2.5.4 Mutually coupled three-phase lines

When two or more transmission lines occupy the same right of way for a considerable
length, the electrostatic and electromagnetic coupling between those lines must be
taken into account.
Consider the simplest case of two mutually coupled three-phase lines. The two
coupled lines are considered to form one subsystem composed of four system busbars.
The coupled lines are illustrated in Figure 2.16, where each element is a 3 x 3
compound admittance and all voltages and currents are 3 x 1 vectors.

Figure 2.16 Two coupled three-phase lines


2.5 THREE-PHASE MODELS OF TRANSMISSION LINES 25

The coupled series elements represent the electromagnetic coupling while the coupled
shunt elements represent the capacitive or electrostatic coupling. These coupling param-
eters are lumped in a similar way to the standard line parameters.
With the admittances labelled as in Figure 2.16, and applying the rules of linear
transformation for compound networks, the admittance matrix for the subsystem is

q.
defined as follows:

IA

IB

1,
ID VfJ
12 x 1 12 x 12 12x 1
(2.60)
It is assumed here that the mutual coupling is bilateral. Therefore Y21 = YT12, etc.
The subsystem may be redrawn as in Figure 2.17. The pairs of coupled 3 x 3
compound admittances are now represented as a 6 x 6 compound admittance. The
matrix representation is also shown. Following this representation and the labelling of

FYs1]
Figure 2.17 A 6 x 6 compound admittance representation of two coupled three-phase lines:
(a) 6 x 6 matrix representation; (b) 6 x 6 compound admittance representation
26 2 TRANSMISSION SYSTEMS

the admittance blocks in the figure, the admittance matrix may be written in terms of
the 6 x 6 compound coils as

(2.61)

This is clearly identical to Equation (2.60) with the appropriate matrix partitioning.
The representation of Figure 2.17 is more concise and the formation of
Equation (2.61) from this representation is straightforward, being exactly similar to
that which results from the use of 3 x 3 compound admittances for the normal single
three-phase line.
The data that must be available to enable coupled lines to be treated in a similar
manner to single lines are the series impedance and shunt admittance matrices. These
matrices are of order 3 x 3 for a single line, 6 x 6 for two coupled lines, 9 x 9 for
three and 12 x 12 for four coupled lines.
Once the matrices [Z,] and [ Y,] are available, the admittance matrix for the subsystem
is formed by application of Equation (2.61)
When all the busbars of the coupled lines are distinct, the subsystem may be
combined directly into the system admittance matrix. However, if the busbars are not
distinct then the admittance matrix as derived from Equation (2.61) must be modified.
This is considered in the following section.

2.5.5 Consideration of terminal connections


The admittance matrix as derived above must be reduced if there are different elements
in the subsystem connected to the same busbar. As an example, consider two parallel
transmission lines as illustrated in Figure 2.18.
The admittance matrix derived previously related the currents and voltages at the
four busbars A l , A2, B1 and B2. This relationship is given by

(2.62)

~ vB2

The nodal injected current at busbar A, I,, is given by

(2.63)

(2.64)
2.5 THREE-PHASE MODELS OF TRANSMISSION LINES 27

Busbar @ & /BI Busbar @

-+-
A2

Figure 2.18 Mutually coupled parallel transmission lines

Also, from inspection of Figure 2.18


VA = VAI = VA2, VB = VBl = vB2, (2.65)
The required matrix equation relates the nodal injected currents, 1, and IB, to the
voltages at these busbars. This is readily derived from Equation (2.62) and the condi-
tions specified above. This is simply a matter of adding appropriate rows and columns,
and yields

(2.66)

This matrix [ Y A B ] is the required nodal admittance matrix for the subsystem.
It should be noted that the matrix in Equation (2.62) must be retained as it is required
in the calculation of the individual line power flows.

2.5.6 Shunt elements


Shunt reactors and capacitors are used in a power system for reactive power control.
The data for these elements are usually given in terms of their rated MVA and rated
kV; the equivalent phase admittance in p.u. is calculated from these data.
Consider, as an example, a three-phase capacitor bank shown in Figure 2.19. A
similar triple representation to that for a line section is illustrated. The final two forms
are the most compact and will be used exclusively from this point on.

I. I

Figure 2.19 Representation of a shunt capacitor bank


28 2 TRANSMISSION SYSTEMS

Figure 2.20 Graphic representation of series capacitor bank between nodes i and k

The admittance matrix for shunt elements is usually diagonal as there is normally
no coupling between the components of each phase. This matrix is then incorporated
directly into the system admittance matrix, contributing only to the self-admittance of
the particular bus.

2.5.7 Series elements


Any element connected directly between two buses may be considered a series element.
Series elements are often taken as being a section in a line sectionalization which is
described later in the chapter.
A typical example is the series capacitor bank which is usually taken as uncou-
pled, i.e. the admittance matrix is diagonal. This can be represented graphically as in
Figure 2.20.
The admittance matrix for the subsystem can be written by inspection as:

(2.67)

2.5.8 Line sectionalization


A line may be divided into sections to account for features such as the following:
0 Transposition of line conductors.
0 Change of type of supporting towers.
0 Variation of soil permitivity.
0 Improvement of line representation (series of two or more equivalent-lr networks).
0 Series capacitors for line compensation.
0 Lumping of series elements not central to a particular study.
An example of a line divided into a number of sections is shown in Figure 2.21.
2.5 THREE-PHASE MODELS OF TRANSMISSION LINES 29

The network of Figure 2.21 is considered to form a single subsystem. The resultant
admittance matrix between bus A and bus B may be derived by finding, for each section,
the ABCD or transmission parameters, then combining these by matrix multiplications
to give the resultant transmission parameters. These are then converted to the required
admittance parameters.
This procedure involves an extension of the usual two-port network theory to multi-
two-port networks. Currents and voltages are new matrix quantities and are defined in
Figure 2.22. The ABCD matrix parameters are also shown.

1 I 1 5 1 1 7
I
f
I I I/
I
-
I
I
I
I
c-
I.
I/ 41
I\

41

Transposition

Figure 2.22 Two-port network transmission parameters. (a) Normal two-port network;
(b) Transmission parameters; (c) Multi-two-port network; (d) Matrix transmission parameters
30 2 TRANSMISSION SYSTEMS

I ,Bus B
220 kV
220166 kV

Bus A
s 220166 kV (i)6 6
Bus ckV
Bus B

Bus A
220 kV

i I BUS C
Section ! Section i Section ,
No 1 No 2 No 3
(ii)

Figure 2.23 Sample system to illustrate line sectionalization. (a) System single line diagram;
(b) system redrawn to illustrate line sectionalization

The dimensions of the parameter matrices correspond to those of the section being
considered, i.e. 3, 6, 9, or 12 for 1, 2, 3 or 4 mutually coupled three-phase elements,
respectively. All sections must contain the same number of mutually coupled three-
phase elements, ensuring that all the parameter matrices are of the same order and
that the matrix multiplications are executable. To illustrate this feature, consider the
example of Figure 2.23.

Features of interest
(a) As a matter of programming convenience, an ideal transformer is created and
included in Section 1.
(b) The dotted coupling represents coupling which is zero. It is included to ensure
correct dimensionality of all matrices.
(c) In the p.u. system, the mutual coupling between the 220 kV and 66 kV lines is
expressed to a voltage base given by the geometric mean of the base line-neutral
voltages of the two parallel circuits.

In Table 2.1, [u] is the unit matrix, [O] is a matrix of zeros, and all other matrices
have been defined in their respective sections.
Once the resultant ABCD parameters have been found, the equivalent nodal admit-
tance matrix for the subsystem can be calculated from the following equation.

[YI = (2.68)
2.6 EVALUATION OF OVERHEAD LINE PARAMETERS 31

Table 2.1 ABCD parameter matrices for the common section types

-[YsPl- [ yss 1
Transformer

Shunt element

Series element
WI
Note: All the above matrices have dimensions corresponding to the number
of coupled three-phase elements in the section.

2.6 Evaluation of Overhead Line Parameters


The lumped series impedance matrix [Z] of a transmission line consists of three compo-
nents, while the shunt admittance matrix [ Y ] contains one.
[ZI = [GI + [&I + [&I, (2.69)
[YI = [YgI, (2.70)
where
[ZC] is the internal impedance of the conductors (Sl.km-'),
[Z,] is the impedance due to the physical geometry (shown in Figure 2.24) of the
conductor's arrangement (Q.km-'),
[&I is the earth return path impedance (R.km-'), and
[Y,] is the admittance due to the physical geometry of the conductor (SZ-'.krn-l).
In multiconductor transmission all primitive matrices (the admittance matrices of
the unconnected branches of the original network components) are symmetric and,
therefore, the functions that define the elements need only be evaluated for elements
on or above the leading diagonal.

2.6.1 Earth impedance matrix [&I


The impedance due to the earth path varies with frequency in a non-linear fashion.
The solution of this problem, under idealized conditions, has been given in the form
of either an infinite integral or an infinite series [ 9 ] .
32 2 TRANSMISSION SYSTEMS

i
vi- Y j

Ground

Yi+ Y j

- Image of
conductors
i'

Figure 2.24 Conductor and its image

As the need arises to calculate ground impedances for a wide spectrum of frequen-
cies, the tendency is to select simple formulations aiming at a reduction in computing
time, while maintaining a reasonable level of accuracy.
Consequently, what was originally a heuristic approach [ l o ] is becoming the more
favoured alternative, particularly at high frequencies.
Based on Carson's work, the ground impedance can be concisely expressed as
te = 1000J(r, B)(Q.km-'), (2.71)
where
ze E [&I,
~ ( r0), = -w-,
WVa
7r
6) + j Q ( r , e)),

eij = arctan (-)xi -+


Yi
xj

Yj
for i # j ,

ei j = 0 for i = j ,
w = 27rf (rades-I),
f = frequency (Hz),
Yi = height of conductor i(m),
2.6 EVALUATION OF OVERHEAD LINE PARAMETERS 33

xi - x, = horizontal distance between conductors i and j (m),


Pa = permeability of free space = 7 n x lO-'(H.m-'),
P = earth resistivity (S2.m).
Carson's solution to Equation (2.71) is defined by eight different infinite series which
converge quickly for problems related to transmission line parameter calculation, but
the number of required computations increases with frequency and separation of the
conductors.
More recent literature has described closed form formulations for the numerical
evaluation of line-ground loops, based on the concept of a mirroring surface beneath
the earth at a certain depth. The most popular complex penetration model which has
had more appeal is that of C. Dubanton [ 111, due to its simplicity and high degree of
accuracy for the whole frequency span for which Carson's equations are valid.
Dubanton's formulae for the evaluation of the self- and mutual impedances of
conductors i and j are

(2.72)

(2.73)

where p = 1/4= is the complex depth below the earth at which the mirroring
surface is located.
An alternative and very simple formulation has been recently proposed by Acha [ 123,
which for the purpose of harmonic penetration yields accurate solutions when compared
with those obtained using Carson's equations.

2.6.2 Geometrical impedance matrix [&I and admittance


matrix [Y,]
If the conductors and the earth are assumed to be equipotential surfaces, the geometrical
impedance can be formulated in terms of potential coefficients theory.
The self-potential coefficient +ij for the ith conductor and the mutual potential
coefficient +i, between the ith and jth conductors are defined as follows,
+ii = In (2Yi/ri)7 (2.74)
+ij = ln (Qj/(xi - xi)), (2.75)
where ri is the radius of the ith conductor (m) while the other variables are as defined
earlier.
For bundle conductors, ri is replaced by the Geometrical Mean Radius (GMRi),
given by

where
GMRi = (a~ ,

R ~ ~ =~ radius
d l ~of bundle,
n = number of conductors in bundle.
34 2 TRANSMISSION SYSTEMS

The use of GMR ignores proximity effects, hence it is valid only if the subconductor
is much smaller than the spacing between phases of the line.
Potential coefficients depend entirely on the physical arrangement of the conductors
and need only be evaluated once.
For practical purposes the air is assumed to have zero conductance and

[z,] = jwK'[@I Chkm-', (2.76)

where [3r3 is a matrix of potential coefficients and K' = 2 x


The lumped shunt admittance parameters [ Y ] are completely defined by the inverse
relation of the potential coefficients matrix, i.e.

where = permittivity of free space = 8.857 x 10-12(F.rn-').


As [Z,] and [Y,] are linear functions of frequency, they need only be evaluated once
and scaled for other frequencies.

2.6.3 Conductor impedance matrix [Q]


This term accounts for the internal impedance of the conductors. Both resistance
and inductance have a non-linear frequency dependence. Current tends to flow on
the surface of the conductor, this skin effect increases with frequency and needs to
be computed at each frequency. An accurate result for a homogeneous non-ferrous
conductor of annular cross-section involves the evaluation of long equations based on
the solution of Bessel functions,

(2.78)

where
xe =jdFiGCZ9
xi =jJFiGE,
re = external radius of the conductor (rn),
ri= internal radius of the conductor (m),
J O = Bessel function of the first kind and zero order,
J b = derivative of the Bessel function of the first kind and zero order,
N o = Bessel function of the second kind and zero order,
Nb = derivative of the Bessel function of the second kind and zero order,
a, = conductivity of the conductor material at the average conductor temperature.
The Bessel functions and their derivatives are solved, within a specified accuracy, by
means of their associated infinite series. Convergence problems are frequently encoun-
tered at high frequencies and low ratios of conductor thickness to external radius, i.e.
(re - ri)/ret necessitating the use of asymptotic expansions.
2.6 EVALUATION OF OVERHEAD LINE PARAMETERS 35

A new closed form solution has been proposed based on the concept of complex
penetration [lo]; unfortunately errors of up to 6.6% occur in the region of low order
harmonic frequencies.
To overcome the difficulties of slow convergence of the Bessel function approach
and the inaccuracy of the complex penetration method at relatively low frequency, an
alternative approach based upon curve fitting to the Bessel function formula has been
proposed by Acha [12].
Lewis and Tuttle [I31 presented a practical method for calculating the skin effect
resistance ratio by approximating ACSR (aluminium conductor steel reinforced)
conductors to uniform tubes having the same inside and outside diameters as the
aluminium conductors, see Figure 2.25(a). Figure 2.25(b) illustrates the skin effect
ratio for different models and various tube ratios for ACSR conductors. Skin effect

Steel strand core

Aluminium strands

5.0
tlr;= 1.0

''vO 50 100 150 200 250 300


(b) f(fIR,,,)

Figure 2.25 ACSR hollow conductor: (a) conductor geometry; (b) skin effect resistance for
different models
36 2 TRANSMISSION SYSTEMS

modelling is important for long lines. Although the series resistance of a transmission
line is typically a small component of the series impedance, it dominates its value at
resonances.

2.6.4 Series impedance approximation for electromagnetic


transients
Based on Semiyen's complex depth of penetration concept, PSCADEMTDC uses the
following expressions for the self and mutual series impedance parameters of the line
(on the assumption that D, << Dij in Figure 2.24):

jowo
zi;= -
27r
(In -
)i:(
+-0.3565 + p p coth-l(0.777&M)
7r@ 27rR~
S2.m-', (2.80)

where
Rc and l$. are the conductive and external conductor radii,

De =/" jmpo '


pc = conductor resistivity (S2.m),
p g = ground resistivity (S2.m).

2.7 Underground and Submarine Cables [14]


A unified solution similar to that of overhead transmission is difficult for underground
cables because of the great variety in their construction and layouts.
The cross-section of a cable, although extremely complex, can be simplified to that
of Figure 2.26 and its series per unit length harmonic impedance is calculated by the
following set of loop equations.

where
[ dVl/h
- dV2/&] =
dV3/h
[z: 0
2;2
Z;,
z;*
z:~]
z;3
[li] , (2.81)

2; 1 is the sum of the following three component impedances,


Z&re-outside is the internal impedance of the core with the return path
outside the core,
Z&re.insu,ationis the impedance of the insulation surrounding the core,
Z:heath-inside is the internal impedance of the sheath with the return path
inside the sheath.
2.7 UNDERGROUND AND SUBMARINE CABLES [ 141 37

LOOP 2

Loop 3

External

Figure 2.26 Cable cross-section

Similarly
;'2 = iheath-outside -k Z:heath/armour-insulation + Zknour-inside 9 (2.82)
and
;'3 = Zkour-outside + ':rmour/earth-insulation + Z:arth-inside' (2.83)
The coupling impedances Z',2 = Z;, and Z;, = Z;, are negative because of opposing
current directions (1, in negative direction in loop 1, and I 3 in negative direction in
loop 2), i.e.
';2 = = -Z:heath-mutua13 (2.84)
z;, = z;, = -z'armour-mutual 9
(2.85)
where
Zihea[h-mutual is the mutual impedance (per unit length) of the tubular sheath
between inside loop 1 and the outside loop 2.
is the mutual impedance (per unit length) of the tubular armour;
Z~rmour-mutual
between the inside loop 2 and the outside loop 3.
Finally, Z;, = Z;, = 0 because loop 1 and loop 3 have no common branch.
The impedances of the insulation are given by

(2.86)

where
p is the permeability of insulation in H.rn-',
routside is the outside radius of insulation,
rinside is the inside radius of insulation.
If there is no insulation between the armour and earth, then Zfnsulation
= 0.
38 2 TRANSMISSION SYSTEMS

The internal impedances and the mutual impedance of a tubular conductor are a
function of frequency, and can be derived from Bessel and Kelvin functions.

(2.87~)

mr = 4- 1
1 -s2’
(2.88)

mq = J.- S2
(2.89)
1-9’
with
81r x f pr
K = (2.90)
%C

s=
9
-, (2.91)
r
q = inside radius,
r = outside radius,
R&, = d.c. resistance in !3km-l.
The only remaining term is Z&rth-inside in Equation (2.83) which is the earth return
impedance for underground cables, or the sea return impedance for submarine cables.
The earth return impedance can be calculated approximately with Equation (2.87a) by
letting the outside radius go to infinity. This approach, also used by Bianchi and Luoni
[lS] to find the sea return impedance is quite acceptable considering the fact that sea
resistivity and other input parameters are not known accurately.
Equation (2.81) is not in a form compatible with the solution used for overhead
conductors, where the voltages with respect to local ground and the actual currents in
the conductors are used as variables. Equation (2.81) can easily be brought into such
a form by introducing the appropriate terminal conditions, namely with

VI = vcore - Vsheath, 11 = Icore,

and
2.8 THREE-PHASE MODELS OF TRANSFORMERS 39

Equation (2.81) can be rewritten as

- [ dvcore/&
dVsheath/&
dVarmour/k
][ =
Zkc
zgc zgs
Zks

Zkc ZL Zia
ZLa
Zia] [ Icore

Isheath]

Iarmour
9
(2.92)

where
Z,: = Z’,I + 2Z;, + Z,; + 2Zi3 + Z i 3 ,
ZLS = z:, = zl,, + z;,+ 2z;, + z;y
ZLa = zl,=,z;, = zgs= z;, + Zi3,
z:, = Z;, + 2Zi3 + Zi3,
z’
aa
= z‘33’
Because a good approximation for many cables having bonding between the sheath
and the armour, and the armour earthed to the sea, is Vsheath = VarmOur = 0, the system
can be reduced to
-dVcore/& = Zlcore, (2.93)
where Z is a reduction of the impedance matrix of Equation (2.92).
Similarly, for each cable the per unit length harmonic admittance is:
jwC‘, 0 0
(2.94)

where C[ = 2m0sr/ln(r/q). Therefore, when converted to core, sheath and armour


quantities,
Y; -Y\ 0

[
- dIsheath/&
d ~ ~
=
[ -Y\
~ 0 ~ -Y;u
where Yi = j d i . If, as before, Vsheafh = Va,o,r
Y’+ Y;
~Y ; + Y~ ; ]
-Y;

= 0, Equation (2.95) reduces to


-dIcoreIb = Y ; Vcore (2.96)
Therefore, for frequencies of interest, the cable per unit length harmonic impedance,
Z’, and admittance, Y’, are calculated with both the zero and positive sequence values
being equal to the Z in Equation (2.93), and the Y’ in Equation (2.96), respectively.
In the absence of rigorous computer models, such as described above, power compa-
nies often use approximations to the skin effect by means of correction factors. Typical
corrections used by the NGC (UK) and EDF (France) are given in Table 2.2.

2.8 Three-phase Models of Transformers


The inherent assumption that the transformer is a balanced three-phase device is justi-
fied in the majority of practical situations, and traditionally, three-phase transformers
are represented by their equivalent sequence networks.
40 2 TRANSMISSION SYSTEMS

Table 2.2 Corrections for skin effect in cables


Company Voltage (kV) Harmonic order Resistance
NGC 400, 275 (Based on 2.5 h 21.5 +
0.74R1(0.267 1 . 0 7 3 A )
s q h . conductor at 5
in. spacing between
centres)
132 h 2 2.35 R I(0.187 + 0.532Ji;)
EDF 400, 225 h 2 2 +
0.74R1(0.267 1.073Jj;)
150, 90 h 1 2 +
Ri (0.187 0 . 5 3 2 A )

More recently, however, methods have been developed [3,4] to enable all three-phase
transformer connections to be accurately modelled in phase co-ordinates. In phase co-
ordinates, no assumptions are necessary although physically justifiable assumptions
are still used in order to simplify the model. The primitive admittance matrix, used
as a basis for the phase co-ordinate transformer model is derived from the primi-
tive or unconnected network for the transformer windings and the method of linear
transformation enables the admittance matrix of the actual connected network to be
found.

2.8.1 Primitive admittance model of three-phase transformers

Many three-phase transformers are wound on a common core and all windings are,
therefore, coupled to all other windings. Therefore, in general, a basic two-winding
three-phase transformer has a primitive or unconnected network consisting of six
coupled coils. If a tertiary winding is also present the primitive network consists of
nine coupled coils. The basic two-winding transformer shown in Figure 2.27 is now
considered, the addition of further windings being a simple but cumbersome extension
of the method.
The primitive network, Figure 2.28, can be represented by the primitive admittance
matrix which has the following general form:

(2.97)

v6

The elements of matrix [Y]can be measured directly, i.e. by energizing coil i and
short-circuiting all other coils, column i of [ Y l can be calculated from yki = &pi.
2.8 THREE-PHASEMODELS OF TRANSFORMERS 41

Figure 2.27 Diagrammatic representation of two-winding transformer

Figure 2.28 Primitive network of two-winding transformer. Six coupled coil primitive network.
(Note the dotted coupling represents parasitic coupling between phases)

Considering the reciprocal nature of the mutual couplings in Equation (2.97), 21


short circuit measurements would be necessary to complete the admittance matrix.
Such a detailed representation is seldom required.
By assuming that the flux paths are symmetrically distributed between all windings,
Equation (2.97) may be simplified to Equation (2.98).

H t I I
I

(2.98)
-Ym Y; Y; Y, YE YE
Y; -Ym I
Y; I
Y; I
Ys I
Y;
H k I

I I I I

where
y,l,, is the mutual admittance between primary coils;
y$ is the mutual admittance between primary and secondary coils
on different cores;
y; is the mutual admittance between secondary coils.
42 2 TRANSMISSION SYSTEMS

t
where ypi = y/$ , y Sl. = y/# and Mii = y/ffipi
fori=1,2or3andj=4,5or6
Figure 2.29 Primitive network

For three separate single-phase units, all the primed values are effectively zero. In
three-phase units, the primed values, representing parasitic interphase coupling, do have
a noticeable effect. This effect can be interpreted through the symmetrical component
equivalent circuits.
If the values in Equation (2.98) are available, then this representation of the primitive
network should be used. If interphase coupling can be ignored, the coupling between
a primary and a secondary coil is modelled as for the single-phase unit, giving rise to
the primitive network of Figure 2.29.
The new admittance matrix equation is

(2.99)

2.8.2 Models for common transformer connections

The network admittance matrix for any two-winding three-phase transformer can now
be formed by the method of linear transformation.
As a simple example, consider the formation of the admittance matrix for a star-star
connection with both neutrals solidly earthed in the absence of interphase mutuals. This
example is chosen as it is the simplest computationally.
The connection matrix is derived from consideration of the actual connected network.
For the star-star transformer, illustrated in Figure 2.30, the connection matrix [C]
relating the branch voltages (i.e. voltages of the primitive network) to the node voltages
2.8 THREE-PHASE MODELS OF TRANSFORMERS 43

l-a "a

l-C "b

Figure 2.30 Network connection diagram for three-phase star-star transformer

(i.e. voltages of the actual network) is a 6 x 6 identity matrix, i.e.

(2. loo)

The nodal admittance matrix [Y]NODE is given by:

[YINODE = [Cl" YlPRIM[Cl. (2.101)


Substituting for [C] yields:
[YINODE = [YIPRIM, (2.102)
Let us now consider the Wye G-Delta connection, illustrated in Figure 2.31. The
following connection can be written by inspection between the primitive branch volt-
ages and the node voltages:

(2.103)
44 2 TRANSMISSION SYSTEMS

y; I I v:

Figure 2.31 Network connection diagram of Wye G-Delta transformer

or
(2.104)

we can also write


(2.105)

and using [YIPRIM from Equation (2.98)

(2.106)
Moreover, if the primitive admittances are expressed in per unit, with both the
primary and secondary voltages being one per unit, the Wye-Delta transformer model
must include an effective turns ratio of 4.The upper right and lower left quadrants
of matrix (2.106) must be divided by and the lower right quadrant by 3.
In the particular case of three-single phase transformer units connected in Wye G-
Delta all the y' and y" terms will disappear. Ignoring off-nominal taps (but keeping
in mind the effective f i turns ratio in per unit) the nodal admittance matrix equation
2.8 THREE-PHASE MODELS OF TRANSFORMERS 45

relating the nodal currents to the nodal voltages is:

~,

v,"
(2.107)
where Y is the transformer leakage admittance in p.u. An equivalent circuit can be
drawn, corresponding to this admittance model of the transformer, as illustrated in
Figure 2.32.
The large shunt admittances to earth from the nodes of the star connection are
apparent in the equivalent circuit. These shunts are typically around 10 p.u. (for a 10%
leakage reactance transformer).
The models for the other common connections can be derived following a similar
procedure.
In general, any two-winding three-phase transformer may be represented using two
coupled compound coils. The network and admittance matrix for this representation is
illustrated in Figure 2.33.
It should be noted that
V s p l = [YpslT*

as the coupling between the two compound coils is bilateral.


Often, because more detailed information is not required, the parameters of all three
phases are assumed balanced. In this case, the common three-phase connections are
found to be modelled by three basic submatrices.

....
y//3
I J
Primary Secondary

Figure 2.32 Equivalent circuit for star-delta transformer


46 2 TRANSMISSION SYSTEMS

Figure 2.33 Two-winding three-phase transformer as two coupled compound coils

Table 2.3 Characteristic submatrices used in forming the transformer


admittance matrices
Transformer Self- Mutual
connection admittance admittance
Bus P Bus S YPP ys, Yps, ysp
Wye G Wye G YI YI -YI
Wye G WYe YI yII/3 -yll/3
Wye G Delta Yl YII Ylll
WYe WYe YII/3 yII/3 -y11/3
WYe Delta YIIl3 YII YIII
Delta Delta Yll YI I -Y11

The submatrices, [ Ypp] [ Y,,] etc., are given in Table 2.3 for the common connections,
where
I 2Yt I -Yt I -Yt I
Y,= 9 Ylll =

Finally, these submatrices must be modified to account for off-nominal tap ratio as
follows:
(i) Divide the self-admittance of the primary by a*.
(ii) Divide the self-admittance of the secondary by /J2.
(iii) Divide the mutual admittance matrices by a/J.
It should be noted that in the p.u. system, a delta winding has an off-nominal tap of A.
For transformers with ungrounded Wye connections, or with neutral connected
through an impedance, an extra coil is added to the primitive network for each
unearthed neutral and the primitive admittance matrix increases in dimension. By noting
that the injected current in the neutral is zero, these extra terms can be eliminated from
the connected network admittance matrix [ 161.
Once the admittance matrix has been formed for a particular connection it represents
a simple subsystem composed of the two busbars interconnected by the transformer.
2.8 THREE-PHASE MODELS OF TRANSFORMERS 47

2.8.3 Three-phase transformer models with independent phase


tap control

Disregarding interphase mutual couplings, the per unit primitive admittance matrix in
terms of the transformer leakage admittance (yti) is

where a l , a2 and a3 are the off-nominal taps on windings 1, 2 and 3, respectively. In


addition, any windings connected in delta will, because of the per unit system, have
an effective tap of a.
The nodal admittance matrix for the transformer windings is:

where [C] is the connection (windings to nodes) matrix.


As an example, [Ynode] for a star-delta transformer with earthed neutral is as follows:

[Yncdel =
48 2 TRANSMISSION SYSTEMS

2.8.4 Sequence components modelling of three-phase


transformers
In most cases, lack of data will prevent the use of the general model based on the prim-
itive admittance matrix and will justify the conventional approach in terms of symmet-
rical components. Let us now derive the general sequence components equivalent
circuits and the assumptions introduced in order to arrive at the conventional models.
With reference to the Wye G-Delta common-core transformer of Figure 2.3 1, repre-
sented by Equation (2.106), and partitioning this matrix to separate self and mutual
elements, the following transformations apply:

where

and a = eJ2nf3.
Therefore

YOPI2 = (2.108)
0

Secondaly side:
The delta connection on the secondary side introduces an effective f i turns ratio and
the sequence components admittance matrix is

(2.109)
2.8 THREE-PHASE MODELS OF TRANSFORMERS 49

Mutual terms:
The mutual admittance submatrix of Equation (2.106), modified for effective turns
ratio. is transformed as follows:

0 0 0
- 0 -(ym + y3L-30" 0 . (2.110)

Recombining the sequence components submatrices yields:

~
I;
1;

2.35 and 2.36, respectively.


1.v;
V;

(2.1 11)
Equation (2.1 11) can be represented by the three sequence networks of Figures 2.34,

In general, therefore, the three sequence impedances are different on a common-core


transformer.
The complexity of these equivalent models is normally eliminated by the following
simplifications:
0 The 30" phase shifts of Wye-Delta connections are ignored.

Figure 2.34 Zero-sequence node admittance model for a common-core grounded Wye-Delta
transformer (3) (01982 IEEE)
50 2 TRANSMISSION SYSTEMS

(Y, + YJ /30" Delta


1

~Y,-Y,',-~Y,+YJ /30" (Y, - YJ - (Y, + YJ /30"

Figure 2.35 Positive-sequence node admittance model for a common-core grounded


Wye-Delta transformer (3) (01982 IEEE)

Figure 2.36 Negative-sequence node admittance model for a common-core grounde


Wye-Delta transformer (3) (01982 IEEE)

Table 2.4 Typical symmetrical-component models for the six


most common connections of three-phase transformers (3) (0 1982

-
IEEE)

Bus P Bus Q PosSeq

-
P 'SC 0
WyeG WyeG

-
W y e ~ wye P&Q

-
Wye G Delta -p 'SC Q

-
Wye Wye P
&
Q

Wye Delta

Delta Delta
2.10 REFERENCES 51

The interphase mutuals admittances are assumed equal, i.e. yh = yg = y:. These
are all zero with uncoupled single-phase units.
The differences (yp - ym) and (ys - ym) are very small and are, therefore, ignored.
With these simplifications, Table 2.4 illustrates the sequence impedance models of
three-phase transformers in conventional steady-state balanced transmission system
studies.

2.9 Formation of the System Admittance Matrix


It has been shown that the element (and subsystem) admittance matrices can be manipu-
lated efficiently if the three nodes at the busbar are associated together. This association
proves equally helpful when forming the admittance matrix for the total system.
The subsystem, as defined in section 2.4, may have common busbars with other
subsystems, but may not have mutual couplings terms to the branches of other subsys-
tems. Therefore, the subsystem admittance matrices can be combined to form the
overall system admittance matrix as follows:

0 The self-admittance of any busbar is the sum of all the individual self-admittance
matrices at that busbar.
The mutual admittance between any two busbars is the sum of the individual mutual
admittance matrices from all the subsystems containing those two nodes.

2.10 References

1. Clarke, E, (1943). Circuit Analysis of a.c. Power Systems, Vol. 1, John Wiley and Sons,
New York.
2. &on, G, (republished in 1965). Tensor Analysis of Networks, MacDonald, London.
3. Chen, M S and Dillon, W E, ( 1 974). Power-system modelling, Proceedings of the IEEE,
62, (7), pp. 901.
4. Laughton, M A, ( I 968). Analysis of unbalanced polyphase networks by the method of phase
co-ordinates, Part I. System representation in phase frame of reference, Proceedings of the
IEE, 115, (8), pp. 1163- I 172.
5 . Kimbark, E W, (1950). Electrical Transmission of Power and Signals, John Wiley and Sons,
New York.
6. Wedepohl, L M and Wasley, R G, (1966). Wave propagation in multiconductor overhead
lines, Proceedings of the IEE, 113, (4), pp. 627-632.
7. Bowman, K J and McNamee, J M, (1964). Development of equivalent pi and T matrix
circuits for long untransposed transmission lines, IEEE Transactions on Power Apparatus
and Systems, PAS-84, pp. 625-632.
8. Wilkinson, J H and Reinsch, (197 1). Handbook for Automatic Computations, Vol. If, Linear
Algebra, Springer-Verlag, Berlin.
9. Carson, J R, (1926). Wave propagation in overhead wires with ground return, Bell Systems
Technical Journal, 5 , pp. 539-556.
10. Deri, A, Tevan, G, Semlyen, A and Castanheira, A, (1981). The complex ground return
plane, a simplified model for homogeneous and multi-layer earth return, IEEE Transactions
on Power Apparatus and Systems, PAS-100, pp. 3686-3693.
52 2 TRANSMISSION SYSTEMS

11. Semlyen, A and Deri, A, (1985). Time domain modelling of frequency dependent three-
phase transmission line impedance, IEEE Transactions on Power Apparatus and Systems,
PAS-104, pp. 1549-1555.
12. Acha, E, (1988). Modelling of power system transformers in the complex conjugate
harmonic space, Ph.D. thesis, University of Canterbury, New Zealand.
13. Lewis, V A and Tuttle, P D, (1958). The resistance and reactance of aluminium
conductors steel-reinforced, IEEE Transactions on Power Apparatus and Systems, PAS-77,
pp. 1189-1215.
14. Dommel, H W, (1978). Line constants of overhead lines and underground cables, Course
E.E. 553 notes, University of British Columbia.
15. Bianchi, G and Luoni, G, (1976). Induced currents and losses in single-core submarine
cables, IEEE Transactions on Power Apparatus and Systems, PAS-95, pp. 49-58.
16. Dillon, W E and Chen, M S. (1972). Transformer modelling in unbalanced three-phase
networks, IEEE Summer Power Meeting, Vancouver.
3
FACTS AND HVDC
TRANSMISSION

3.1 Introduction
As well as the passive components described in the previous chapter, modern transmis-
sion systems contain a growing number of power electronic devices with the aim of
enhancing the power transfer controllability. These are discussed in this chapter under
the two categories of FACTS (Flexible a.c. Transmission Systems) and HVDC (High
Voltage Direct Current Transmission).
Power electronic devices are not amenable to easily manageable equivalent circuits,
as their constituent elements vary with the operating condition in a non-linear fashion.
Therefore, the steady state models of the power electronic devices are described in this
chapter in the form of mathematical relationships instead of circuit equivalents.

3.2 Flexible a.c. Transmission Systems [l]


In the conventional free flow mode of operation of a.c. transmission networks, the
power flow on individual transmission circuits is determined by the characteristics
of the transmission network itself. Moreover, for stable operation sufficient transmis-
sion margin must be available at all times to accommodate the almost instantaneous
redistribution of power flow that results from a power system disturbance.
Instead of this free flow mode of operation, the power flow through one or more
transmission lines can be controlled in a predetermined manner, through the appli-
cation of power electronic controlled devices at strategic locations. The use of such
devices reduces the need for network reinforcements by improving the utilization and
performance of existing facilities.
The main control actions in a power system, such as changing transformer taps,
switching current or governing turbine steam pressure, are currently achieved through
the use of mechanical devices, which necessarily impose a limit on the speed at which
control action can be made. FACTS devices, based on solid-state control, are capable
of control actions at far higher speed. The three parameters that control transmission
line power flow are line impedanceand the magnitude and phase of line end voltages.
Conventional control of these parameters, although adequate during steady-state and
slowly changing load conditions, cannot, in general, be achieved quickly enough to
54 3 FACTS AND HVDC TRANSMISSION

Sending end Receiving end

Figure 3.1 Simplified transmission system

Figure 3.2 Effect of different power flow controllers

handle dynamic system conditions. It is shown below that the use of FACTS technology
can change this situation.
The real and reactive power transferred via the simplified transmission line of
Figure 3.1 are determined by the following relationships:
p = - ' s V R sin 8, (3.1)
XL
VS
Q = -[v, case - VR] (3.2)
XL
Figure 3.2, graph (l), shows the variation of real power transmission with voltage
phase angle difference (6) for the case when (Vsl = l V ~ = l V . The effect of the
different power flow controllers is also shown in Figure 3.2, graph (2) is achieved
by the insertion of series capacitance, graph (3) by the presence of a constant voltage
source in the middle of the line in the form of a static VAR compensator and graph
(4)by the use of a phase shifter.
The characteristics and modelling of FACTS devices used to implement the above
power flow controls are discussed next.

3.2.1 Thyristor controlled series compensator (TCSC)


The TCSC unit shown in Figure 3.3 constitutes the basic module of an Advanced Series
Compensator (ASC). By exercising phase-angle control of the switching instants the
ASC alters the transmission line impedance in a controllable manner [2,3].
3.2 FLEXIBLE A.C. TRANSMISSION SYSTEMS [I] 55

: Figure 3.3 Thyristor controlled series capacitor (TCSC)


Transmission line

Figure 3.4 Equivalent circuit of the basic TCSC module

A number of these modules in series permits, within specified limits [4] the following:

0 Flexible, continuous control of the effective line impedance, and thus of the compen-
sation level.
0 Direct dynamic control of power flow within the network.
0 Damping of local and inter-area oscillations, as well as subsynchronous resonances.
Figure 3.5 shows the ASC voltage V C (diagram (a)), the thyristor pair asymmetrical
current pulses i~ (diagram (b)), the capacitor current ic (diagram (c)) and the line
current i~ (diagram (d)), which is assumed sinusoidal. The chosen time reference is
the positive-going zero crossing of the TCSC voltage.
The waveforms of Figure 3.5 are derived by applying the Laplace Transform to the
circuit of Figure 3.4. The resulting steady state current [4] has the following expression:
cos B cos fi&t
iT(t) = A C O S U -~A (3.3)
cos ka
where

A=--- @;
@; - w2
1
0;= -
LC
k =q / w

and the firing angle a = R - a.


56 3 FACTS AND HVDC TRANSMISSION

4-
f Line current

Figure 3.5 ASC voltage and current waveforms

I
I
I
I
2 Inductive I
$ I
Q3, 1

I Capacitive
1
-E I
I
1
I
I 1 I I

Figure 3.6 Effective ASC circuit impedance

The ASC circuit effective reactance at fundamental frequency (XI) is derived from
Fourier analysis of the waveforms and has the following expression:
1 A 4A
XI=--- -[2a + sin 2a] + cos 2a[k tan k a - tan a]. (3.4)
WC R6X noc(2k - 1)
Equation (3.4) has poles at:
ka = (2n - l)n/2 for n = 1 , 2 , 3 , . , .
A plot of the effective impedance is shown in Figure 3.6.

3.2.2 Static on-load tap changing


The voltage magnitude difference between the ‘sending’ and ‘receiving’ ends is
primarily responsible for the transfer of reactive power according to Equation (3.2)
3.2 FLEXIBLE A.C. TRANSMISSION SYSTEMS [ l ] 57

Figure 3.7 Equivalent circuit of the two-winding transformer

when 8,the phase angle difference between the sending and receiving end voltages, is
kept small for stability reasons.
On-Load Transformer tap Changing (OLTC) is widely used to satisfy Equation (3.2)
while maintaining a constant voltage on the receiving end transformer secondary side.
Figure 3.7 shows the equivalent circuit of a two-winding transformer with complex
taps on both primary and secondary windings [ 5 ] .
The primary and secondary windings are represented by ideal transformers in series
with their respective impedances Z, and Z,. They both include complex tap ratios,
which are different for the voltage and the current, i.e.

TVLVV= Ti" for the primary

and
= V;
UvLvu for the secondary.

To account for transformer core loss, the equivalent circuit of Figure 3.7 includes a
magnetizing admittance branch Yo.
The following expressions can be written for the circuit of Figure 3.7:

(3.5)

or
(3.8)

where
58 3 FACTS AND HVDC TRANSMISSION

Putting Equations ( 3 . 3 , (3.6) and (3.7) into matrix form:

(3.10)

Matrix Equation (3.10) represents the transformer shown in Figure 3.7. However, a
reduced equivalent matrix can be written using only the external nodes p and s, i.e.

3.2.3 Static phase shifter


As shown in Figure 3.2 (graph 4), the power flow can be kept at its peak value over
a range of phase angles. This is achieved by means of a quadrature voltage injection,
V,, which produces the required phase angle shift in the voltages to keep the power
flow at a level which is independent of phase angle 6.
No action is taken in the region of 0 < 8 < 90", where the power follows the same
trace as case 1. After 90", V,, is altered so the phase angle between the sending and
receiving end voltages remains at 90". When the phase is greater than 90" + a , the
power flow is decreased sinusoidally towards zero as the angle increases. The voltage
phase angle can be adjusted with a Thyristor Controlled Phase Angle (TCPA) device
or phase shifter, as shown in Figure 3.8.

I I
- Transmission Line

Figure 3.8 Thyristor Controlled Phase Angle (TCPA)


3.2 FLEXIBLE A.C. TRANSMISSION SYSTEMS [ l ] 59

The general model described in the previous section can be adapted for phase-shifting
[ 5 ] , if the following substitutions are made in Equation (3.1 1):
Tv = cos@,, + jsin@rv,
(3.12)
Uv = cos Cpuv + j sin @.,,
Then the transfer admittance matrix equation becomes:

(3.13)

3.2.4 Static VAR compensator [6]


The OLTC does not provide reactive power and, therefore, the receiving voltage level
can only be maintained if there exists a 'flexible' and rapidly controllable source of
reactive power, either locally or somewhere at the sending end. Power electronics can
provide such sources in the form of static VAR Compensators (SVC).
An SVC (shown in Figure 3.9) is a parallel combination of capacitors and inductors,
the latter under phase angle control (TCR). This combination provides a fast variable
source (or sink) of reactive power. TCRs were designed to operate with thyristor
switches and have already been used extensively worldwide in the past two decades.
The operation of the TCR is illustrated by the equivalent circuit and waveforms
shown in Figure 3.10.
The following two equations represent the positive and negative half-cycles of the
reactor current.
(3.14)
WL

(3.15)

I Transmission line
I

I
T
Figure 3.9 Static VAR Compensator (SVC) using (TCR)
60 3 FACTS AND HVDC TRANSMISSION

+I

Figure 3.10 Thyristor controlled reactor

and the fundamental component of this current is

(3.16)

3.2.5 The static compensator (STATCOM) [7]


Besides their relatively high cost, SVCs produce harmonic currents that require some
filtering. A more flexible alternative recently proposed is the static equivalent to
the synchronous compensator, hence termed the ‘STATCOM’, based on the prin-
ciple of a voltage-sourced converter which uses gate turn-off devices and, therefore,
does not need the expensive passive reactive components. The basic components of
a STATCOM are illustrated in Figure 3.11. For an ideal lossless unit, the (three-
phase) inverted voltage (Vinv) is in phase with the ax. source voltage ( V s ) . Then
(from Equation 3.2) when V , > Vinv, the unit absorbs VARs whereas for V , < Vinv it
provides VARs.

Figure 3.11 Static compensator (STATCOM)


3.2 FLEXIBLE A.C. TRANSMISSION SYSTEMS [ I ] 61

While primarily designed for reactive power, the STATCOM is potentially capable of
storing energy (e.g. in the form of superconducting coils) for use to improve transient
stability.

3.2.6 Unified power flow controller (UPFC)


The UPFC is a device capable of controlling the active and reactive power, as well as
the voltage magnitude. Within specified operating limits, the UPFC can control the three
variables simultaneously. Thus, the UPFC can operate as a shunt VAR Compensator,
as a thyristor controlled series compensator, or as a phase-shifter controller.
The main components of the UPFC, shown in Figure 3.12, are the shunt and series
converters. The output voltage of the latter is added in series to the a.c. terminal voltage
VO via a coupling transformer. Therefore, the injected voltage, V c ~acts , as a series
voltage source and alters the sending-end voltage at node rn. The active and reactive
power exchanged between the series converter and the a.c. system is given by the
product of the transmission line current, I,, and the series voltage source, V c ~ .
The active power provided by the series converter is derived from the a.c. power
system via the shunt converter and a d.c. link. The shunt converter can generate
or absorb controllable reactive power in both operating modes, i.e. rectification and
inversion. The reactive power can be independently controlled to maintain the shunt
converter terminal a.c. voltage magnitude at a specified value.
The steady-state model of the UPFC [8] is straightforward, as shown by the equiv-
alent circuit in Figure 3.13.
It consists of two ideal voltage sources representing the fundamental component of
the voltage waveforms at the a s . terminals of the shunt and series converters.
These are:
V,,R = V,~(cos + j sin OUR) (3.17)
~ V,R(COS 6cR + j sin & R )
V c= (3.18)
V,,R and V c ~have specified upper and lower limits, whereas their corresponding
phase angles 6,,R and &R can be placed anywhere between 0 and 2n radians.

__* 0
lk Im
_3 *---.

Figure 3.12 UPFC schematic diagram 181 (0 1997 IEE)


62 3 FACTS AND HVDC TRANSMISSION

a--n
f

Figure 3.13 Equivalent circuit of the UPFC [8] 0 1998 IEE)

ca

Figure 3.14 Basic three-phase rectifier bridge

3.3 High Voltage Direct Current Transmission


3.3.1 The ax.-d.c. converter
The basic converter configuration is the Graetz bridge shown in Figure 3.14.
For large power ratings, static converter units generally consist of a number of series
and/or parallel connected bridges, some or all bridges being phase-shifted relative
to the others. With these configurations, 12-pulse and higher pulse numbers can be
achieved to reduce the distortion of the supply current with limited or no filtering. A
multiple bridge rectifier can, therefore, be modelled as a single equivalent bridge with
a sinusoidal supply voltage at the terminals.
The following basic assumptions are normally made in the development of the
steady-state model [9,10]:

(i) The forward voltage drop in a conducting valve is neglected so that the valve
may be considered as an ideal switch. This is justified by the fact that the voltage
drop is very small in comparison with the normal operating voltage. It is, further,
quite independent of the current and should, therefore, play an insignificant part
in the commutation process since all valves commutating on the same side of the
3.3 HIGH VOLTAGE DIRECT CURRENT TRANSMISSION 63

bridge suffer similar drops. Such a voltage drop is taken into account by adding
it to that of the d.c. line resistance. The transformer windings resistance is also
ignored in the development of the equations, though it should also be included
to calculate the power loss.
(ii) The converter transformer leakage reactances, as viewed from the secondary
terminals, are identical for the three phases, and variations of leakage reactance
caused by on-load tap-changing are ignored.
(iii) The direct current ripple is ignored, i.e. sufficient smoothing inductance is
assumed on the d.c. side.
Rectification Rectification in HVDC transmission is normally achieved via full-
bridge configurations and steady state operation is optimized by means of transformer
on-load tap-changing.
Referring to the voltage waveforms in Figure 3.15 and using as time reference the
instant when the phase-to-neutral voltage in phase b is a maximum, the commutating
voltage of valve 3 can be expressed as:

eb sin ut + - ,
- e, = Z/Z;;vterrn ( ;> (3.19)

where a is the off-nominal tap-change position of the converter transformer. The shaded
area in Figure 3.15(b) indicates the potential difference between the common cathode

wt=O

Figure 3.15 Diode rectifier waveforms. (a) Alternating current in phase b; (b) Common anode
(ca) and cathode (cc) voltage waveform; ( c ) Rectified voltage
64 3 FACTS AND HVDC TRANSMISSION

(cc) and common anode (ca) bridge poles for the case of uncontrolled rectification.
The maximum average rectified voltage is therefore:

However, uncontrolled rectification is rarely used in large power conversion.


Controlled rectification is achieved by phase-shifting the valve conducting periods
with respect to their corresponding phase voltage waveforms.
With delay angle control, the average rectified voltage (shown in Figure 3.16) is
thus

(3.21)

In practice, the voltage waveform is that of Figure 3.17, where a voltage area (SA) is
lost due to the reactance (X,)of the a.c. system (as seen from the converter), referred
to as commutation reactance. The energy stored in this reactance has to be transferred
from the outgoing to the incoming phase, and this process results in a commutation
or conduction overlap angle (u). Referring to Figure 3.17 and ignoring the effect of
resistance in the commutation circuit, area SA can be determined as follows:

X, di,
eb -e, = 2-- (3.22)
o dt'

Figure. 3.16 Thyristor controlled waveform. (a) Alternating current in phase 'b'; (b) Rectified
d.c. voltage waveforms
3.3 HIGH VOLTAGE DIRECT CURRENT TRANSMISSION 65

Figure 3.17 Effect of commutation reactance, (a) Alternating current; (b) d.c. voltage wave-
forms

where e,, e b are the instantaneous voltages of phases a and b, respectively, and i, is
the incoming valve (commutating) current.

(3.23)

Finally, by combining Equations (3.20),(3.21)and (3.23),the following a.c.-d.c.


voltage relationship is obtained.

(3.24)

It must be emphasized that the commutating voltage (Vterm) is the a.c. voltage at
the closest point to the converter bridge where sinusoidal waveforms can be assumed.
The commutation reactance (X,) is the reactance between the point at which Vterm
exists and the bridge. Where filters are installed, the filter busbar voltage can be used
as Ve,. In the absence of filters, V,,,, must be established at some remote point and
X , must be modified to include the system impedance from the remote point to the
converter.
With perfect filtering, only the fundamental component of the current waveform will
appear in the a.c. system. This component is obtained from the Fourier analysis of the
current waveform in Figure 3.17,and requires information of i, and u.
66 3 FACTS AND HVDC TRANSMISSION

Taking as a reference the instant when the line voltage (eb - e,) is zero, Equation
(3.22) can be written as
X , dic
&avtermsin or = 2- -, (3.25)
w dt
and integrating with respect to (or):

/% sin wtd(wt) = X , / di,, (3.26)

or
--1 uVtermcos wt + K = X&. (3.27)
Jz
From the initial condition: i, = 0 at wt = a,the following expressions for K and i,
are obtained
1
K = --aVterm cos 01, (3.28)
A
avterm
1, = -[cos a - cos w t ] . (3.29)
Ax,
From the final condition: i, = t d at wt =a + u, the following expressions for I d
and u are obtained.

(3.30)

(3.31)

Equation (3.29) provides the time-varying commutating current and Equation (3.3 1)
the limits for the Fourier analysis.
Fourier analysis of the a.c. current waveform, including the effect of commutation
(shown in Figure 3.17) leads to the following relationship between the r.m.s. of the
fundamental component and the direct current:

(3.32)

where

k='
+
[cos 2a - cos ~ ( C Y u)I2 + [2u + sin 2a - sin 2(a + u)I2 (3.33)
4[cos - coS(a + u)]
for values of u not exceeding 60".
The values of k are very close to unity under normal operating conditions, i.e.,
when the voltage and currents are close to their nominal values and the a.c. voltage
waveforms are symmetrical and undistorted.
Taking into account the transformer tap position, the current on the primary side
becomes c

46
I, = k-atd.
TT (3.34)
3.3 HIGH VOLTAGE DIRECT CURRENT TRANSMISSION 67

When using per unit values based on a common power and voltage base on both
sides of the converter, the direct current base has to be f i times larger than the
alternating current base and Equation (3.34)becomes

(3.35)
Using the fundamental components of voltage and current and assuming perfect
filtering at the converter terminals, the power factor angle at the converter terminals
is 4 (the displacement between fundamental voltage and current waveforms) and the
following applies:
P = f i v t e r m l p cos 4 = V d l d , (3.36)
or
(3.37)
and
(3.38)
Inversion Owing to the unidirectional nature of current flow through the converter
valves, power reversal (i.e. power flow from the d.c. to the a.c. side) requires direct
voltage polarity reversal. This is achieved by delay angle control which, in the absence
of commutation overlap, produces rectification between 0" < a < 90" and inversion
between 90" -= a < 180". In the presence of overlap, the value of a at which inversion
begins is always less than 90". Moreover, unlike with rectification, full inversion (i.e.
a = 180") cannot be achieved in practice. This is due to the existence of a certain
deionization angle y at the end of the conducting period, before the voltage across the
commutating valve reverses, i.e.
+ u 5 180" - yo,
If the above condition is not met (yo being the minimum required extinction angle),
a commutation failure occurs; this event would upset the normal conducting sequence
and preclude the use of the steady-state model derived in this chapter.
The inverter voltage, although of opposite polarity with respect to the rectifier, is
usually expressed as positive when considered in isolation.
Typical inverter voltage and current waveforms are illustrated in Figure 3.18. By
similarity with the waveforms of Figure 3.17,the following expression can be written
for the inverter voltage in terms of the extinction angle:

(3.39)
which is the same as Equation (3.24)substituting y for a,
It should by now be obvious that inverter operation requires the existence of three
conditions, as follows:
(i) An active a.c. system which provides the commutating voltages.
(ii) A d.c. power supply of opposite polarity to provide continuity for the unidirec-
tional current flow (i.e. from anode to cathode through the switching devices).
(iii) Fully controlled rectification to provide firing delays beyond 90".
68 3 FACTS AND HVDC TRANSMISSION

Figure 3.18 Inverter waveforms. (a) Alternating current; (b) d.c. voltage waveforms

Qi I Qr

Figure 3.19 P and Q vector diagram

When these three conditions are met, a negative voltage of a magnitude given
by Equation (3.39) is impressed across the converter bridge and power (-VdId) is
inverted. Note that the power factor angle (4) is now larger than go", i.e.
P = JSv,,,r, cos rp = -&VtermIp cos(n - 4) (3.40)
Q = d3Vte,.,,Ip sin 4 = 3 V t e m I p sin(n - 4). (3.41)
Equations (3.40) and (3.41) indicate that the inversion process still requires reactive
power supply from the a.c. side, The vector diagram of Figure 3.19 illustrates the sign
of P and Q for rectification and inversion.

3.3.2 Commutation reactance


Figure 3.20 shows the general case of n bridges connected in parallel on the a.c. side.
In the absence of filters, the pure sinusoidal voltages exist only behind the system
source impedance ( X s s ) and the commutation reactance ( X , j ) for the jth bridge is thus
Xcj = Xss + Xtj. (3.42)
3.3 HIGH VOLTAGE DIRECT CURRENT TRANSMISSION 69

"dl

"dfl

Figure 3.20 n bridges connected in series on the d.c. side and in parallel on the a s . side

However, if the bridges are under the same controller or under identical controllers,
then it is preferable to create a single equivalent bridge. The commutation reactance
of such an equivalent bridge depends upon the d.c. connections and also the phase
shifting between bridges.
If there are k bridges with the same phase shift, then they will commutate at the
same time and the equivalent reactance must reflect this. For a series connection of
bridges, the commutation reactance of the equivalent bridge is:
Xcre,ies= kxss +Xtj, (3.43)
where j represents any of the n bridges. For bridges connected in parallel on the d.c.
side, the equivalent bridge commutation reactance is:

(3.44)
It should be noted that with perfect filtering or when several bridges are used with
different transformer phase shifts, the voltage on the a.c. side of the converter trans-
formers may be assumed to be sinusoidal and hence X,, has no influence on the
commutation.

3.3.3 d.c. link control


The invention of the high voltage mercury valve half a century ago paved the way for
the development of the HVDC transmission technology. By 1954, the first commer-
cial d.c. link came successfully into operation and was soon followed by many other
schemes, orders of magnitude larger.
Each terminal of the HVDC link consists of at least two bridges (like the one shown
in Figure 3.14) in a 12-pulse configuration using solid state valves formed by many
thyristor levels in series.
HVDC transmission systems can be configured in different ways to take into account
cost, flexibility and operational requirements. Three basic configurations are illustrated
in Figure 3.21 in order of increasing complexity.
The simplest is the back-to-back interconnection (Figure 3.2 1(a)) in which the two
converters are on the same site, as there is no transmission line. They can be designed
more economically (a 15% to 20% saving in converter plant is quoted) than those of
70 3 FACTS AND HVDC TRANSMISSION

Figure 3.21 Basic configurations

long distance schemes, and are designed for relatively low voltages (50 to 150 kV),
although still use the highest current rating of the single thyristors. The two units are
identical and each can be used on the rectification or inversion modes as ordered by
the system control.
In the monopolar link (Figure 3.21(b)), the two converter stations are joined by a
single conductor line, and earth (or the sea) is used as the return conductor, requiring
two electrodes capable of carrying the full current.
Next, and the most common configuration, is the bipolar link (Figure 3.21(c)) which
consists of two monopolar systems combined, one at positive and one at negative
polarity with respect to ground (middle). Each monopolar side can operate on its own
with ground return; but if the two poles have equal current, they cancel each other’s
ground current to practically zero. In such cases, the ground path is used for limited
periods in an emergency, when one pole is temporarily out of service.
The sending and receiving ends of the two-terminal d.c. transmission link can be
modelled as single equivalent bridges with terminal voltages v d , and v d , , respectively,
as shown in Figure 3.22. The direct current is thus given by:

(3.45)

where R d is the resistance of the link and includes the loop transmission resistance (if
any), the resistance of the smoothing reactors and the converter valves.
The prime considerations in the operation of a d.c. transmission system are to mini-
mize the need for reactive power at the terminals and to reduce system losses. These
objectives require maintaining the highest possible transmission voltage and this is
3.3 HIGH VOLTAGE DIRECT CURRENT TRANSMISSION 71

Convertor I Convertor I1

Figure 3.22 Two-terminal d.c. link

achieved by minimizing the inverter end extinction angle, i.e. operating the inverter
on constant Extinction Angle (EA) control while controlling the direct current at the
rectifier end by means of temporary delay angle backed by transformer tap-change
control.
EA control applied to the inverter automatically varies the firing angle of advance
to maintain the EA y at a constant value. Deionization imposes a definite minimum
limit on y , and the EA control usually maintains it at this limit.
Constant Current (CC) control applied to the rectifier regulates the firing angle a!
to maintain a pre-specified link current I:’, within the range of a!. If the value of a
required to maintain ISp falls below its minimum limit, current control is transferred
to the inverter, i.e. a is fixed on its minimum limit, and the inverter firing angle is
advanced to control the current.
The converter-transformer tap-change is a composite part of this control. The
rectifier transformer attempts to maintain a! within its permitted range. The inverter
transformer attempts to regulate the d.c. voltage at some point along the line to a spec-
ified level. For minimum loss and minimum-reactive-power absorption, this voltage is
required to be as high as possible, and the firing angle of the rectifier should be as low
as possible.
Figure 3.23 shows the d.c. voltage/current characteristics at the rectifier and inverter
ends (the latter have been drawn with reverse polarity in order to illustrate the oper-
ating point). The current controller gains are very large and for all practical purposes
the slopes of the constant current characteristics can be ignored. Consequently, the
operating current is equal to the relevant current setting, i.e. I d f r and Id,, for rectifier
and inverter constant current control, respectively.
The direction of power flow is determined by the current settings, the rectifier end
always having the larger setting. The difference between the settings is the current
margin Id,,, and is given by
Id,,, = Id,, - I d , , > 0. (3.46)
Many d.c. transmission schemes are bi-directional, i.e. each converter operates some-
times as a rectifier and sometimes as an inverter. Moreover, during d.c. line faults, both
converters are forced into the inverter mode in order to de-energize the line faster. In
such cases, each converter is provided with a combined characteristic as shown in
Figure 3.24, which includes natural rectification, CC control and constant EA control.
With the characteristics shown by solid lines (i.e. operating at point A ) , power is
transmitted from Converter I to Converter 11. Both stations are given the same current
command but the current margin setting is subtracted at the inverter end. When power
reversal is to be implemented, the current settings are reversed and the broken line
characteristics apply. This results in operating point B with direct voltage reversed and
no change in direct current.
72 3 FACTS AND HVDC TRANSMISSION

"d

lnvertor c.c., ,Rectifier C.C.

1% k
Figure 3.23 Normal control characteristics

'd

+ Conv. I

Figure 3.24 Control characteristics and power flow reversal


3.3 HIGH VOLTAGE DIRECT CURRENT TRANSMISSION 73

Alternative forms of control A commonly used operating mode is constant power


(CP) control. As with constant current control, either converter can control power.
The power setting at the rectifier terminal P d , , must be larger than that at the inverter
terminal Pd,, by a suitable power margin P d , , , , that is:
Pd,,, = Pd,, - Pd,, '0. (3.47)
The CP controller adjusts the CC control setting 1;' to maintain a specified power
flow P i p through the link, which is usually more practical than CC control from a
system-operation point of view. The voltage/current loci now become non-linear, as
shown in Figure 3.25.
Several limits are added to the CP characteristics, as shown in Figure 3.26. These are:
0 A maximum current limit with the purpose of preventing thermal damage to the
converter valves; normally between 1 and 1.2 times the nominal current.

vdt

Figure 3.25 Constant power characteristics

Figure 3.26 Voltage and current limits


74 3 FACTS AND HVDC TRANSMISSION

A minimum current limit (about 10% of the nominal value) in order to avoid
possible current discontinuities which can cause overvoltages.
Voltage dependent current limit (line OA in the figure) in order to reduce the power
loss and reactive power demand.
In cases where the power rating of the d.c. link is comparable with the rating of
either the sending or receiving a s . systems interconnected by the link, the frequency
of the smaller a.c. system is often controlled, to a large extent, by the d.c. link. With
power-frequency (PF) control, if the frequency goes out of pre-specified limits, the
output power is made proportional to the deviation of frequency from its nominal
value. Frequency control is analogous to the CC described earlier, i.e. the converter
with lower voltage determines the direct voltage of the line and the one with higher
voltage determines the frequency. Again, current limits have to be imposed which
override the frequency error signal.
CPEA and CCEA controls were evolved principally for bulk point-to-point power
transmission over long distances or submarine crossings and are still the main control
modes in present use.
Multi-terminal d.c. schemes have also being considered, based on the basic controls
already described. Two alternative are possible, i.e. constant voltage parallel [ 1 I ] and
constant current series [ 121 schemes.

3.3.4 Three-phase model [(13)]


Terminology and waveforms The converter model for unbalanced analysis
is considerably more complex than that developed for balanced operation. The
additional complexity arises from the need to include the effect of the three-phase
converter-transformer connection and of the converter firing control strategies. Under
unbalanced conditions, the converter-transformer modifies the source voltages applied
to the converter and also affects the phase distribution of current and power. Each
bridge operates with a different degree of unbalance, due to the influence of the
converter-transformer connections and must be modelled independently.
The converter, whether rectifying or inverting, is represented by the circuit of
Figure 3.27.
As for symmetrical operation, in fundamental frequency studies the converter is
assumed to be connecting a system with perfect filtering on the a.c. side and perfect
current smoothing on the d.c. side. By using one of the converter angles (e.g. O;.,,, in
Figure 3.27) as a reference, the mathematical coupling between the a.c. system and
converter equations is weakened and the rate of convergence of the power flow solution
improved. Computational simplicity is achieved by using common power and voltage
bases on both sides of the converter. The phase-neutral voltage is used as the base
parameter and, therefore
MVAb,,, = Base power per phase
Vbase = Phase-neutral voltage base.
The current base on the a.c. and d.c. sides are also equal. Therefore, the p.u. system
does not change the form of any of the converter equations.
3.3 HIGH VOLTAGE DIRECT CURRENT TRANSMISSION 75

G I I I I

Primary Secondary

Figure 3.27 Basic converter unit

Figure 3.28 Star-star transformer connection

The phase-to-phase source voltages referred to the transformer secondary are found
by a consideration of the transformer connection and off-nominal turns ratio. For
example, consider the star-star transformer of Figure 3.28.
The phase-to-phase source voltages referred to the secondary are:

(3.48)

(3.50)
which in terms of real and imaginary parts yield six equations.
Variables and equations With reference to Figures 3.27 and 3.29 and
Equations (3.48)-(3.50), the converter model uses 26 variables, i.e.
E l , E 2 , E 3 , 4 1 , 4 2 7 4 3 3 1 ~1 2, , 1 3 , m i 30 2 , ~ 3 ,

U12r U13r u 2 3 ~ cl, c 2 , C3r G I , a 2 9 013, al, a 2 , a 3 ,


Vdr Id.
76 3 FACTS AND HVDC TRANSMISSION

Since the terminal voltage is assumed to be undistorted apart from negative and zero
sequence at fundamental frequency, every third delay angle and commutation duration
is the same, i.e. 4 4 = ( Y I , ,u4 = p l , etc. and consequently only three commutation
durations and three delay angles need be considered.
Therefore, 26 equations have to be found involving these variables.
Starting from the d.c. side, the average d.c. voltage, found by integration of the
waveforms in Figure 3.29(b), may be expressed in the form:

.Jz
Vd = -{U21[cos(cI
n
+ 41 - c3 + n)- cos(C2 + 4 2 - c3 + n)]
+ U13[COS(C2 + 4 2 - CI - cos(C3 + - Cl)]43

+ U23[COS(C3 + 4 3 - C2) - cos(C1 + 4 1 + n - C2)]


- Id(Xc1 + x c 2 + Xc3) (3.51)

where XCi is the commutation reactance for phase i.


Another equation is derived from the d.c. system topology relating the d.c. voltages
and currents, i.e.
f W d , Id) = 0.

Phase 1

Figure 3.29 Unbalanced converter voltage and current waveform. (a) Phase voltages, (b) d.c.
voltage waveform, (c) assumed current waveshape for Phase 1 (actual waveform is indicated by
dotted line)
3.3 HIGH VOLTAGE DIRECT CURRENT TRANSMISSION 77

For example, a monopolar d.c. link with two bridges at each end provides the
following equation:
Vd, Vd2 vd, + +
v d 4 - I d R d = 0. + (3.52)
The three-phase converter transformer is represented by its nodal admittance
model, i.e.

where p and s indicate the primary and secondary sides of the transformer on the
assumption of a lossless transformer (i.e. Y,, = jb,,, etc.).
Turning to the a.c. side, on the assumption of a lossless transformer, the currents at
the converter side busbar are expressed as follows:
3
Ii exp(jwi>= - x(jbi:Ek exp(j@k) j b i ; ~ f , ,exp[j(ef,, + - e&,,,)i. (3.53)
k= I

By subtracting O:em in the above equation, the terminal busbar angles are related to
the converter angle reference.
Separating this equation into real and imaginary components, the following six equa-
tions result:
3
11 c o s w = z[b:aEk sin @k + b:,kVfemsin(&,, - e:,,)], (3S4a)
k= I
3
1 2 cos 0 2 = x[b:aEk sin @k + b:,k~:,,, sin(ekrm- e:,,)~, (3.54b)
k= I
3
1 3 cos 0 3 = x[b::Ek sin @k + b;ivfermsin(&,,, - e:,,)], (3.54c)
k-I
3
11 sin 0 1 = - x [ b , , E~ cos @k
Ik
+ b:,kVfermcOs(e,k, - e;e,)~, (3.54d)
k=l
3
12 sin w2 = - x[b::& COSok + bti~f,, cOs(ekrm- e;,,,,,)], (3.54e)
k= I

k= 1

Three equations are derived from approximate expressions for the fundamental r.m.s.
components of the line current waveforms as shown in Figure 3.29(c), i.e.

(3.55a)
78 3 FACTS AND HVDC TRANSMISSION

4 Id
12 = - -sin(T2/2), (3.55b)
n f i

(3 S5C)

where T is the assumed conducting period of each phase.


Six further equations can be derived from the transformer connection, interrelating
the secondary (converter side) currents and the primary and secondary phase voltages,
i.e. from the components of the transformer’s nodal admittance matrix equation.
As an example, the nodal admittance matrix (in per unit) relating these variables for
the star-delta connected transformer (assuming the nominal tap position a = 1 is:

-YM YIA $Y -gY


1 1
-gY

Y I A -YlA 1
-gY
2
?Y
I
-jY . . (3.56)

y / A -y/a +Y - g Iy $Y

The sum of the real powers on the three phases of the transformer secondary may
be equated to the total d.c. power, i.e.

i= I

To obtain a reference which may be applied to all transformer secondary windings,


an artificial reference node is created corresponding to the position of the zero-sequence
secondary voltage. This choice of reference results in the following two equations:
3
C E; cos 4; = 0, (3.58)
i= 1
3
C E ; sin 4; = 0. (3.59)
i= 1

The nodal admittance matrix for the star-connected transformer secondary is now
formed for an unearthed star winding. The restriction on the zero-sequence current
flowing on the secondary is, therefore, implicitly included in the transformer model
for both star and delta connections.
A total of 20 equations have been selected so far. The remaining six are obtained
from the control strategies. For three-phase inverter operation it is necessary to retain
the variable 01 in the formulation, as it is required in the specification of the symmetrical
3.4 REFERENCES 79

firing controller. Therefore, the restriction upon the extinction advance angle y, requires
the implicit calculation of the commutation angle for each phase.
Using the specification for y , as defined in Figure 3.29, the following expressions
apply:

(3.60a)

(3.60b)

(3.60~)

Two further equations result from assuming that the off-nominal tap position is the
same in the three phases, i.e.
a1 = a2, (3.61a)
a2 = a3, (3.61b)
and a final equation relates to the constant current or constant power controller, i.e.
Id = Is', (3.62a)

(3.62b)
Summarizing, the formulation of the three-phase converter model involves the
following 26 equations:
(3.51) .. . 1
(3.52) ... 1
(3.54) ... 6
(3.55) ... 3
(3.56) ... 6
(3.57) ... 1
(3.58) . .. 1
(3.59) . .. 1
(3.60) . .. 3
(3.61) . .. 2
(3.62) ... 1

3.4 References
1. Hingorani, N G, (1993). Flexible a.c. transmission systems, IEEE Spectrum, 40-45.
2. Larsen, E V, Clark, K, Miske, S S and Urbanek, J, (1994). Characteristics and rating
considerations of thyristor controlled series compensation, IEEE Transactions on Power
Delivery, 9, (2), 992-1000.
3. Larsen, E V, Bowler, C, Damsky, B and Nilsson, S,(1992). Benefits of thyristor controlled
series compensation, International Conference on Large High Voltage Electric Systems
(CIGRE), PAPER 14/37/38-04.
80 3 FACTS AND HVDC TRANSMISSION

4. Christ], N, Hedin, R, Sadek, K, Lutzelberger, P, Krause, P E, McKenna, S M, Montoya


A H and Torgenson, D, (1992). Advanced series compensation with thyristor-controlled
impedance, CIGRE paper I4/37/38-05, Paris.
5 . Fuertes-Esquivel, C R and Acha, E, (1996). Newton-Raphson algorithm for the reliable
solution of large power networks embedded FACTS devides, Proceeding of the IEE, Gener.
Transm. & Distrib., 143, (S), 447-454.
6. Miller, T J E, (1982). Reactive Power Control in Electric Systems, John Wiley & Sons, New
York.
7. Gyugyi, L, (1994). Dynamic compensation of ax. transmission lines by solid state synchro-
nous voltage sources, IEEE Transactions on Power Delivery, PD-9, (2). 904-91 1.
8. Fuertes-Esquivel, C R and Acha, E, (1998). The unified power flow controller: A critical
comparison of Newton-Raphson UPFC algorithms in power flow studies, Proceedings IEE
Generation, Transmission and Distribution, 144, (9,437-444.
9. Adamson, C and Hingorani, N G, (1960). HVDC Power Transmission, Garraway.
10. Arrillaga, J, (1999). HVDC Transmission, IEE Publications, London.
11. Lamm, U, Uhlmann, E and Danfors, P, (1963). Some aspects of tapping of HVDC trans-
mission systems, Direct Current, 8, 124- 129.
12. Adamson, C and Arrillaga, J, (1968). Behaviour of multiterminal a.c.-d.c. interconnections
with series-connected stations, Proceedings IEE, 115, (1 l), 1685- 1692.
13. Harker, B J and Arrillaga, J, (1979). Three-phase a.c.1d.c. load flow’, Proceedings ofthe
IEE, 126, (12), 1275-1281.
LOAD FLOW

4.1 Introduction

Under normal conditions, electrical transmission systems operate in their steady-state


mode and the basic calculation required to determine the characteristics of this state is
termed load flow (or power flow).
The object of load flow calculations is to determine the steady-state operating char-
acteristics of the power generatiodtransmission system for a given set of busbar loads.
Active power generation is normally specified according to economic-dispatching prac-
tice and the generator voltage magnitude is normally maintained at a specified level by
the automatic voltage regulator acting on the machine excitation. Loads are normally
specified by their constant active and reactive power requirement, assumed unaffected
by the small variations of voltage and frequency expected during normal steady-state
operation.
The solution is expected to provide information of voltage magnitudes and angles,
active and reactive power flows in the individual transmission units, losses and the
reactive power generated or absorbed at voltage-controlled buses.
The load flow problem is formulated in its basic analytical form in this chapter
with the network represented by linear, bilateral and balanced lumped parameters.
However, the power and voltage constraints make the problem non-linear and the
numerical solution must, therefore, be iterative in nature.
The current problems faced in the development of load flow are: an ever-increasing
size of systems to be solved, on-line applications for automatic control, and system
optimization. Hundreds of contributions have been offered in the literature to overcome
these problems [l].
Five main properties are required of a load flow solution method.

(i) High computational speed. This is especially important when dealing with large
systems, real time applications (on-line), multiple case load flow such as in
system security assessment, and also in interactive applications.
(ii) Low computer storage. This is important for large systems and in the use of
computers with small core storage availability, e.g. mini-computers for on-line
application.
(iii) Reliability of solution. It is necessary that a solution be obtained for ill-
conditioned problems, in outage studies and for real time applications.
82 4 LOADFLOW

(iv) Versatility. An ability on the part of load flow to handle conventional and special
features (e.g. the adjustment of tap ratios on transformers; different representa-
tions of power system apparatus), and its suitability for incorporation into more
complicated processes.
(v) Simplicity. The ease of coding a computer program of the load flow algorithm.

The type of solution required from a load flow also determines the method used:
accurate or approximate
unadjusted or adjusted
off-line or on-line
single case or multiple cases.
The first column are requirements needed for considering optimal load flow and
stability studies, and the second column those needed for assessing security of a system.
Obviously, solutions may have a mixture of the properties from either column.
The first practical digital solution methods for load flow were the Y matrix-iterative
methods [2]. These were suitable because of the low storage requirements, but had the
disadvantage of converging slowly or not at all. Z matrix methods [3] were developed
which overcame the reliability problem but a sacrifice was made of storage and speed
with large systems.
The Newton-Raphson method [4,5] was developed at this time and was found
to have very strong convergence. It was not, however, made competitive until spar-
sity programming and optimally ordered Gaussian elimination [6-81 were introduced,
which reduced both storage and solution time.
Non-linear programming and hybrid methods have also been developed but these
have created only academic interest and have not been accepted by industrial users
of load flow. The Newton-Raphson method and techniques, derived from this algo-
rithm, satisfy the requirements of solution type and programming properties better than
previously used techniques, and have now replaced them.

4.2 Basic Nodal Method


In the nodal method, as applied to power system networks, the variables are
the complex node (busbar) voltages and currents, for which some reference must
be designated. In fact, two different references are normally chosen: for voltage
magnitudes the reference is ground, and for voltage angles the reference is chosen
as one of the busbar voltage angles, which is fixed at the value zero (usually). A nodal
current is the nett current entering (injected into) the network at a given node, from a
source and/or load external to the network. From this definition, a current entering the
network (from a source) is positive in sign, while a current leaving the network (to a
load) is negative, and the nett nodal injected current is the algebraic sum of these. One
may also speak in the same way of nodal injected powers S = P jQ. +
Figure 4.1 gives a simple network showing the nodal currents, voltages and powers.
In the nodal method, it is convenient to use branch admittances rather than
impedances. Denoting the voltages of nodes k and i as Ek and Ei,respectively, and the
4.2 BASIC NODAL METHOD 83

Figure 4.1 Simple network showing nodal quantities

admittance of the branch between them as yk;, then the current flowing in this branch
from node k to node i is given by:
Iki = Yki(Ek - E i ) . (4.1)
Let the nodes in the network be numbered 0, 1, . . . , n , where 0 designates the refer-
ence node (ground). By Kirchhoff s current law, the injected current I k must be equal
to the sum of the currents leaving node k, hence:
n n

i d i=O

Since Eo = 0, and if the system is linear:


n n

If this equation is written for all the nodes except the reference, i.e. for all busbars
in the case of a power system network, then a complete set of equations defining the
network is obtained in matrix form as:

YI1 y12 YIn


Y21 y22 Y2n
(4.4)

yn1 Yn2 Yn,,

where
n
yk; = self-admittance of node k ,
ykk = i=O#k
Y k ; = - y k ; = mutual admittance between nodes k and i .

In shorthand matrix notation, Equation (4.4) is simply


I = YE. (4.5)
84 4 LOADFLOW

or in summation notation we have


n

(4.6)

The nodal admittance matrix in Equations (4.4) or (4.5) has a well-defined structure,
which makes it easy to construct automatically. Its properties are as follows:

(i) Square of order n x n.


(ii) Symmetrical, since yki = yik.
(iii) Complex.
(iv) Each off-diagonal element yki is the negative of the branch admittance between
nodes k and i , and is frequently of value zero.
(v) Each diagonal element ykk is the sum of the admittance of the branches, which
terminate on node k, including branches to ground.
(vi) Because in all but the smallest practical networks very few non-zero mutual
admittances exist, matrix Y is highly sparse.

4.3 Conditioning of Y Matrix


The set of equations I = YE may or may not have a solution. If not, a simple physical
explanation exists, concerning the formulation of the network problem. Any numerical
attempt to solve such equations is found to break down at some stage of the process
(what happens in practice is usually that a finite number is divided by zero).
The commonest case of this is illustrated in the example of Figure 4.2.
The nodal equations are constructed in the usual way as:

(4.7)

y12

L I
Figure 4.2 Example of singular network
4.4 THE CASE WHERE ONE VOLTAGE IS KNOWN 85

Suppose that the injected currents are known, and nodal voltages are unknown. In this
case, no solution for the latter is possible. The Y matrix is described as being singular,
i.e. it has no inverse, and this is easily detected in this example by noting that the
sum of the elements in each row and column is zero, which is a sufficient condition
for singularity, mathematically speaking. Hence, if it is not possible to express the
voltages in the form E = Y - I I , it is clearly impossible to solve Equation (4.7) by any
method, whether involving inversion of Y or otherwise.
The reason for this is obvious: we are attempting to solve a network whose reference
node is disconnected from the remainder, i.e. there is no effective reference node, and
an infinite number of voltage solutions will satisfy the given injected current values.
When, however, a shunt admittance from at least one of the busbars in the network
of Figure 4.2 is present, the problem of insolubility immediately vanishes in theory,
but not necessarily in practice. Practical computation cannot be performed with abso-
lute accuracy, and during a sequence of arithmetic operations, rounding errors due to
working with a finite number of decimal places accumulate. If the problem is well
conditioned and the numerical solution technique is suitable, these errors remain small
and do not mask the eventual results. If the problem is ill-conditioned, and this usually
depends upon the properties of the system being analysed, any computational errors
introduced are likely to become large with respect to the true solution.
It is easy to see intuitively that, if a network having zero shunt admittances cannot
be solved even when working with absolute computational accuracy, then a network
having very small shunt admittances may well present difficulties when working with
limited computational accuracy. This reasoning provides a key to the practical problems
of network, i.e. Y matrix, conditioning. A network with shunt admittances, which are
small with respect to the other branch admittances, is likely to be ill-conditioned, and
the conditioning tends to improve with the size of the shunt admittances, i.e. with the
electrical connection between the network busbars and the reference node.

4.4 The Case Where One Voltage is Known


In load flow studies, it usually happens that one of the voltages in the network is
specified, and instead the current at that busbar is unknown. This immediately alleviates
the problem of needing at least one good connection with ground, because the fixed
busbar voltage can be interpreted as an infinitely strong ground tie. If it is represented
as a voltage source with a series impedance of zero value, and then converted to the
Norton equivalent, the fictitious shunt admittance is infinite, as is the injected current.
This approach is not computationally feasible, however.
The usual way to deal with a voltage that is fixed is to eliminate it as a variable
from the nodal equations. Purely for the sake of analytical convenience, let this busbar
be numbered 1 in an n-busbar network. The nodal equations are then

(4.8)
86 4 LOADFLOW

The terms in El on the right-hand side of Equations (4.8) are known quantities, and
as such are transferred to the left-hand side.
+
I I - Y I I E I= Y12E2 . . . YlnEn,
12 - Y ~ I E=I Y22E2 + . YznEn,
(4.9)

p,m In - YnlE1 = Yn2E2 + + . YnnEn.


The first row of this set may now be eliminated, leaving (n - 1) equations in (n - 1)
unknowns, E2 . . . E n .In matrix form, this becomes:

or
In - YnIEI
(4.10)

I = YE. (4.1 1)
The new matrix Y is obtained from the full admittance matrix Y merely by removing
the row and column corresponding to the fixed-voltage busbar, both in the present case
where it is numbered 1 or in general.
In summation notation, the new equations are
n
Ik - YklEl = YkiEi for k = 2 . . . n , (4.12)
i=2
which is an (n - 1) set in (n - 1) unknowns. The equations are then solved by any
of the available techniques for the unknown voltages. It is noted that the problem
of singularity when there are no ground ties disappears if one row and column are
removed from the original Y matrix.
Eliminating the unknown current I1 and the equation in which it appears is the
simplest way of dealing with the problem, and reduces the order of the equations by
one. I I is evaluated after the solution from the first equation in (4.8).

4.5 Analytical Definition of the Problem


The complete definition of power flow requires knowledge of four variables at each
bus k in the system:
PI, = real or active power,
Q k = reactive or quadrature power,
vk = voltage magnitude,
0, = voltage phase angle.
Only two are known apriori to solve the problem, and the aim of the load flow is to
solve the remaining two variables at a bus.
We define three different bus conditions based on the steady-state assumptions of
constant system frequency and constant voltages, where these are controlled.
4.6 NEWTON-RAPHSON METHOD OF SOLVING LOAD FLOWS 87

(i) Voltage controlled bus. The total injected active power Pk is specified, and the
voltage magnitude Vk is maintained at a specified value by reactive power
injection. This type of bus generally corresponds to a generator where Pk is
fixed by turbine governor setting and Vk is fixed by automatic voltage regula-
tors acting on the machine excitation; or a bus where the voltage is fixed by
supplying reactive power from static shunt capacitors or rotating synchronous
compensators, e.g. at substations.
(ii) Non-voltage controlled bus. The total injected power, PkQk, is specified at this
bus. In the physical power system this corresponds to a load centre such as a
city or an industry, where the consumer demands the power requirements. Both
Pk and Q k are assumed to be unaffected by small variations in bus voltage.
(iii) Slack (swing) bus. This bus arises because the system losses are not known
precisely in advance of the load flow calculation. Therefore, the total injected
power cannot be specified at every single bus. It is usual to choose one of
the available voltage controlled buses as slack, and to regard its active power
as unknown. The slack bus voltage is usually assigned as the system phase
reference, and its complex voltage
E, = v,Le,
is, therefore, specified. The analogy in a practical power system is the generating
station which has the responsibility of system frequency control.
Load flow solves a set of simultaneous non-linear algebraic power equations for the
two unknown variables at each node in a system. A second set of variable equations,
which are linear, are derived from the first set, and an iteration method is applied to
this second set.
The basic algorithm which load flow programs use is depicted in Figure 4.3. System
data, such as busbar power conditions, network connections and impedance, are read
in and the admittance matrix formed. Initial voltages are specified to all buses; for base
+
case load flows, P, Q buses are set to 1 j 0 while P, V busbars are set to V + j 0 .
The iteration cycle is terminated when the busbar voltages and angles are such that
the specified conditions of load and generation are satisfied. This condition is accepted
when power mismatches for all buses are less than a small tolerance, q l , or voltage
increments less than q 2 . Typical figures for ql and q 2 are 0.01 p.u. and 0.001 P.u.,
respectively. The sum of the square of the absolute values of power mismatches is a
further criterion sometimes used.
When a solution has been reached, complete terminal conditions for all buses are
computed. Line power flows and losses and system totals can then be calculated.

4.6 Newton-Raphson Method of Solving Load Flows


The generalized Newton-Raphson method is an iterative algorithm for solving a set
of simultaneous non-linear equations in an equal number of unknowns.
fk(x,) =0 for k = 1 + N
(4.13)
m=l+N
88 4 LOADFLOW

Read system data and


the specified loads and
generation. Also read the
voltage specifications at

w
Form system admittance
matrix from s stem data

and angles at all system

angles in order to Iteration


satisfy the specified cycle
conditions of load and

conditions

generation and all line

Figure 4.3 Flow diagram of basic load flow algorithm

At each iteration of the Newton-Raphson method, the non-linear problem is approx-


imated by the linear matrix equation. The linearizing approximation can best be visu-
alized in the case of a single-variable problem.
In Figure 4.4, xp is an approximation to the solution, with error AxP at iteration p ,
Then
+
f ( x P A x p ) = 0. (4.14)
This equation can be expanded by Taylor’s theorem,
+
f ( x P A x p )= 0
(AxP)*
=f(XP) + A x ” f ’ ( x P )+ - 2! f ” ( X P ) + . . . . (4.15)
If the initial estimates of the variable xp is near the solution value, AxP will be
relatively small and all terms of higher powers can be neglected. Hence,
f ( x p )+ A x p f ’ ( x P )= 0 (4.16)
or
(4.17)
4.6 NEWTON-RAPHSON METHOD OF SOLVING LOAD FLOWS 89

L Tangent to
f(x)
I
I
I

i
I
b X

i
x ~ +I l /\
I ~oiution

Figure 4.4 Single-variable linear approximately

The new value of the variable is then obtained from


xP+' = X" + Axp. (4.18)
Equation (4.16) may be rewritten as
f (x') = - J A x P . (4.19)
The method is readily extended to the set of N equations in N unknowns. J becomes
the square Jacobian matrix of first order partial differentials of the functions f k ( x , ) .
Elements of [ J ] are defined by
Jkm = afk
- 9 (4.20)
axm
and represent the slopes of the tangent hyperplanes which approximate the functions
f k ( X m ) at each iteration point.
The Newton-Raphson algorithm will converge quadratically if the functions have
continuous first derivates in the neighbourhood of the solution, the Jacobian matrix
is non-singular, and the initial approximations of x are close to the actual solutions.
However, the method is sensitive to the behaviours of the functions f k ( x m ) and hence
to their formulation. The more linear they are, the more rapidly and reliably Newton's
method converges. Non-smoothness, i.e. humps, in any one of the functions in the
region of interest, can cause convergence delays, total failure or misdirection to a
non-useful solution.

4.6.1 Equations relating to power system load flow


The non-linear network governing equations are
Ik =y h E m for all k , (4.21)
mck
where I k is the current injected into bus k . The power at the k bus is then given by
sk = Pk d- jQk = EkI;

(4.22)
90 4 LOADFLOW

Mathematically speaking, the complex load flow equations are non-analytic, and
cannot be differentiated in complex form. In order to apply Newton's method, the
problem is separated into real equations and variables. Polar or rectangular co-ordinates
may be used for the bus voltages. Hence, we obtain two equations
Pk = P ( V , 8 ) or P ( e , f ) ,
and
Qk = Q(v,
8 ) or Q<elf).
In polar co-ordinates, the real and imaginary parts of Equation (4.22) become
pk = VkVm(GkmCOS8km -k Bkm (4.23)
mek

Qk = c
mck
Vk v m (Gkm sin e k m - Bkm COS e k m 1, (4.24)

where 8km = 8, - 8,.


Linear relationships are obtained for small variations in the variables 8 and V by
forming the total differentials, the resulting equations being:
For a PQ busbar:

(4.25)

(4.26)

For a PV busbar:
Only Equation (4.25) is used, since Qk is not specified.
For a slack busbar: No equations.
The voltage magnitudes appearing in Equations (4.25) and (4.26) for PV and slack
busbars are not variables, but are fixed at their specified values. Similarly, 8 at the
slack busbar is fixed.
The complete set of defining equations is made up of two for each PQ busbar and
one for each PV busbar. The number of variables is, therefore, equal to the number of
equations. Algorithm (4.19) then becomes:
P mismatches 8 corrections
for all P Q and APP-' for all PQ and

1 1
PV busbars PV busbars

Q mismatches
for all PQ
busbars
{ AQp-'
V corrections
for all PQ
busbars
Jacobian matrix
(4.27)
4.6 NEWTON-RAPHSON METHOD OF SOLVING LOAD FLOWS 91

and for m = k:

In practice, some programs express these coefficients using voltages in rectangular


+
form, i.e. e i j f i. This only affects the speed of calculation of the mismatches and
the matrix elements, by eliminating the time-consuming trigonometrical functions.
In rectangular co-ordinates, the complex power equations are given as

Pk -I-j Q k = E k c
mek
YimE>

= (ek +j f k ) c
mck
(Gkm - jBkrn)(ern -jfm)?
and these are divided into real and imaginary parts

Pk = ek x ( G k m e m - B k m f m ) -k f k C ( G k m f m -I- B k m e m ) ,
mek nick

Qk =f kc ( G k m e m - B k m f n i ) - e k C ( G k m fm + B k n i e m ) .
niek mek

At a voltage controlled bus, the voltage magnitude is fixed but not the phase angle.
Hence, both e k and f k vary at each iteration. It is necessary to provide another equation

V: = e: +f i ,
to be solved with the real power equation for these buses.
92 4 LOADFLOW

Linear relationships are obtained for small variations in e and f , by forming the
following total differentials:

Apk =
mek
apk
-Aem
aem
+ c
mek
apk
-A
afm
f m

mek met

for all buses except the slack bus;

AQk = c
mek
%,em
aem
+ c
mek
*Afm
afm

mek mek

for all non-voltage controlled buses;

for voltage controlled buses.

pi
The Jacobian matrix has the form

(4.28)
AV2 EE F F

For voltage controlled buses, V is specified but not the real and imaginary compo-
nents of voltage, e and f. Approximations can be made, for example, by ignoring the
off-diagonal elements in the Jacobian matrix, as the diagonal elements are the largest.
4.6 NEWTON-RAPHSON METHOD OF SOLVING LOAD FLOWS 93

Alternatively, for the calculation of the elements, the voltages can be considered as
E = 1 + j 0 . The off-diagonal elements then become constant.
The polar co-ordinate representation appears to have computational advantages over
rectangular co-ordinates. Real power mismatch equations are present for all buses
except the slack bus, while reactive power mismatch equations are needed for non-
voltage controlled buses only.
The Jacobian matrix has the sparsity of the admittance matrix [ Y ] and has positional
but not numerical symmetry. To gain in computation, the form of [A& AV/VIT is
normally used for the variable voltage vector. Both increments are dimensionless and
the Jacobian coefficients are now symmetric in structure though not in value. The values
of [ J ] are all functions of the voltage variables V and 8 and must be recalculated each
iteration.
As an example, the Jacobian matrix equation for the four-busbar system of Figure 4.5
is given as Equation (4.29).

I API I I A01 I
I AP3 I I 0 I H33 I H34 I 0 I N34 I I A03 I
I AP4 I=

The differences in bus powers are obtained from


(4.30)
(4.31)

A further improvement is to replace the reactive power residual AQ in the Jaco-


bian matrix equations by AQ/V. The performance of the Newton-Raphson method is
closely associated with the degree of problem non-linearity; the best left-hand defining
functions are the most linear ones. If the system power Equation (4.31) is divided
throughout by vk, only one term Qip/Vk on the right-hand side of this equation is
non-linear in vk. For practical values of elp
and Vk, this non-linear term is numerically
relatively small. Hence, it is preferable to use AQ/V instead of AQ in the Jacobian
matrix equation.
Dividing A P by V is also helpful but is less effective since the real power component
of the problem is not strongly coupled with voltage magnitudes. A further alternative

Figure 4.5 Sample system


94 4 LOADFLOW

INPUT DATA
Read all input data, form
system admittance matrix,
initialize voltages and

for voltages and angle

t
Update voltages and angles
I
i

Figure 4.6 Flow diagram of the basic Newton-Raphson load flow algorithm

is to formulate current residuals at a bus. While computationally simple, this method


shows poor convergence in the same way as Y matrix iterative methods.
A flow diagram of the basic Newton-Raphson algorithm is given in Figure 4.6.

4.7 Techniques Which Make the Newton-Raphson


Method Competitive in Load Flow
The efficient solution of Equation (4.27) at each iteration is crucial to the success of
the Newton-Raphson method. If conventional matrix techniques were to be used, the
storage ( a n 2 )and computing time (an3)would be prohibitive for large systems.
For most power system networks, the admittance matrix is relatively sparse, and in
the Newton-Raphson method of load flow the Jacobian matrix has this same sparsity.
The techniques which have been used to make the Newton-Raphson competitive
with other load flow methods involve the solution of the Jacobian matrix equation and
the preservation of the sparsity of the matrix by ordered triangular factorization.

4.7.1 Sparsity programming


In conventional matrix programming, double subscript arrays are used for the location
of elements. With sparsity programming [6],only the non-zero elements are stored, in
one or more vectors, plus integer vectors of identification.
4.7 TECHNIQUES WHICH MAKE THE NEWTON-RAPHSON METHOD 95

For the admittance matrix of order n , the conventional storage requirements are n 2
+
words, but by sparsity programming 6b 3n words are required, where b is the number
of branches in the system. Typically, b = l S n , and the total storage is 12n words. For
a large system (say 500 buses), the ratio of storage requirements of conventional and
sparse techniques is about 40:1.

4.7.2 Triangular factorization


To solve the Jacobian matrix Equation (4.27), represented here as
[AS1 = [Jl[AEl,
for increments in voltage, the direct method is to find the inverse of [J] and solve for
[AE] from
[AE] = [J]-'[AS]. (4.32)
In power systems [ J ] is usually sparse but [J]-' is a full matrix.
The method of triangular factorization solves for the vector [AE] by eliminating [J]
to an upper triangular matrix with a leading diagonal and then back-substituting for
[AE], i.e. eliminate to
[AS'] = [U][AE]
and back-substitute
[V]-'[AS'] = [AE].
The triangulation of the Jacobian is best done by rows. Those rows below the one
being operated on need not be entered until required. This means that the maximum
storage is that of the resultant upper triangle and diagonal. The lower triangle can then
be used to record operations.
The number of multiplications and additions to triangulate a full matrix is i N 3 ,
compared with N 3 to find the inverse. With sparsity programming, the number of
operations varies as a factor of N . If rows are normalized N further operations are saved.

4.7.3 Optimal ordering


In power system load flow, the Jacobian matrix is usually diagonally dominant which
implies small round-off errors in computation. When a sparse matrix is triangulated,
non-zero terms are added in the upper triangle. The number added is affected by the
order of the row eliminations, and total computation time increases with more terms.
The pivot element is selected to minimize the accumulation of non-zero terms, and
hence conserve sparsity, rather than minimizing round-off error. The diagonals are used
as pivots.
Optimal ordering of row eliminations to conserve sparsity is a practical impossibility
due to the complexity of programming and time involved. However, semi-optimal
schemes are used and these can be divided into two sections.
(a) Pre-ordering [7] Nodes are renumbered before triangulation. No complicated
programming or storage is required to keep track of row and column interchanges.
96 4 LOADFLOW

(i) Nodes are numbered in sequence of increasing number of connected lines.


(ii) Diagonal banding-non-zero elements are arranged about either the major
or minor diagonals of the matrix.
(b) Dynamic ordering [8] Ordering is effected at each row during the elimination.
(i) At each step in the elimination, the next row to be operating on is that with
the fewest non-zero terms.
(ii) At each step in the elimination, the next row to be operated on is that which
introduces the fewest new non-zero terms, one step ahead.
(iii) At each step in the elimination, the next row to be operated on is that which
introduces the fewest new non-zero terms, two steps ahead. This may be
extended to the fully optimal case of looking at the effect in the final step.
(iv) With cluster ordering, the network is subdivided into groups which are then
optimally ordered. This is most efficient if the groups have a minimum of
physical intertia. The matrix is then anchor banded.
The best method arises from a trade-off between a processing sequence which
requires the least number of operations, and time and memory requirements.
The dynamic ordering scheme of choosing the next row to be eliminated as that
with the fewest non-zero terms, appears to be better than all other schemes in spar-
sity conservation, number of arithmetic operations required, ordering times and total
solution time.
However, there are conditions under which other ordering would be preferable, e.g.
with system changes affecting only a few rows, these rows should be numbered last;
when the sub-networks have relatively few interconnections, it is better to use cluster
ordering.

4.7.4 Aids to convergence


The Newton-Raphson method can diverge very rapidly or converge to the wrong
solution if the equations are not well behaved or if the starting voltages are badly
chosen. Such problems can often be overcome by a variety of techniques. The simplest
device is to impose a limit on the size of each Ad and A V correction at each iteration.
Figure 4.7 illustrates a case which would diverge without this device.
Another more complicated method is to calculate good starting values for the 0’s
and V’s, which also reduces the number of iterations required.
In power system load flow, setting voltage controlled buses to V j 0 and +
+
non-voltage controlled buses to 1 j 0 may give a poor starting point for the
Newton-Raphson method.
If previously stored solutions for a network are available these should be used. One
or two iterations of a Y matrix iterative method [2] can be applied before commencing
the Newton method. This shows fast initial convergence unless the problem is ill-
conditioned, in which case divergence occurs.
A more reliable method is the use of one iteration of a d.c. load flow (i.e. neglecting
losses and reactive power conditions) to provide estimates of voltage angles; followed
4.8 CHARACTERISTICS OF THE NEWTON-RAPHSON LOAD FLOW 97

Normal Ax'

Required solution

Figure 4.7 Example of diverging solution

by one iteration of a similar type of direct solution to obtain voltage magnitudes. The
total computing time for both sets of equations is about 50% of one Newton-Raphson
iteration and the extra storage required is only in the programming statements. The
resulting combined algorithm is faster and more reliable than the formal Newton
method and can be used to monitor diverging or difficult cases, before commencing
the Newton-Raphson algorithm.

4.8 Characteristics of the Newton-Raphson Load Flow


With sparse programming techniques and optimally ordered triangular factorization, the
Newton method for solving load flow has become faster than other methods for large
systems. The number of iterations is virtually independent of system size (from a flat
voltage start and with no automatic adjustments) due to the quadratic characteristic of
convergence. Most systems are solved in 2 + 5 iterations with no acceleration factors
being necessary.
With good programming, the time per iteration rises nearly linearly with the number
of system buses N , so that the overall solution time varies as N . One Newton itera-
tion is equivalent to about seven Gauss-Seidel iterations. For a 500-bus system, the
conventional Gauss-Seidel method takes about 500 iterations and the speed advantage
of the Newton method is then 15:l. Storage requirements of the Newton method are
greater, however, but increase linearly with system size. It is, therefore, attractive for
large systems.
The Newton method is very reliable in system solving, given good starting approxi-
mations. Heavily loaded systems with phase shifts up to 90" can be solved. The method
is not troubled by ill-conditioned systems and the location of slack bus is not critical.
Due to the quadratic convergence of bus voltages, high accuracy (near exact solution)
is obtained in only a few iterations. This is important for the use of load flow in short-
circuit and stability studies. The method is readily extended to include tap-changing
transformers, variable constraints on bus voltages, and reactive and optimal power
scheduling. Network modifications are easily made.
98 4 LOADFLOW

The success of the Newton method is critical on the formulation of the problem-
defining equations. Power mismatch representation is better than the current mismatch
versions. To help negotiate non-linearities in the defining functions, limits can be
imposed on the permissible size of voltage corrections at each iteration. These should
not be too small, however, as they may slow down the convergence for well-behaved
systems.
The coefficients of the Jacobian matrix are not constant, they are functions of the
voltage variables V and 8, and hence vary with each iteration. However, after a few
iterations, as V and 8 tend to their final values, the coefficients will tend to constant
values.
One modification to the Newton algorithm is to calculate the Jacobian for the first
two or three iterations only and then to use the final one for all the following iterations.
Alternatively, the Jacobian can be updated every two or more iterations. Neither of
these modifications greatly affects the convergence of the algorithm but much time is
saved (but not storage).

4.9 Decoupled Newton Load Flow


An inherent characteristic of any practical electric power transmission system operating
in the steady state condition is the strong interdependence between active powers and
bus voltage angles, and between reactive powers and voltage magnitudes. Correspond-
ingly, the coupling between these ‘P- 8’ and ‘Q - V’ components of the problem
is relatively weak. Many algorithms have been proposed which adopt this decoupling
principle [9- 1 1 1 .
The voltage vectors method uses a series approximation for the sine terms which
appear in the system defining equations, to calculate the Jacobian elements and arrive
at two decoupled equations:
(4.33)
(4.34)
where for the reference node 60 = 0 and Vk = VO.The values of i?k and Qk represent
real and reactive power quantities, respectively and [TI and [ U ] are given by

(4.35)

(4.36)
mwk

(4.37)

(4.38)

where Z h and X k m are the branch impedance and reactance, respectively. [U]is
constant valued and needs to be triangulated once only for a solution. [TI is recalculated
and triangulated each iteration.
4.9 DECOUPLED NEWTON LOAD FLOW 99

The two Equations (4.33) and (4.34) are solved alternately until a solution is obtained.
These equations can be solved using Newton’s method, by expressing the Jacobian
eauations as

(4.39)

or
(4.40)
(4.41)

where

and T and U are, therefore defined in Equations (4.35-4.38).


The most successful decoupled load flow is that based on the Jacobian matrix equa-
tion of the formal Newton method, i.e.

(4.42)

If the sub-matrices N and J are neglected, since they represented the weak coupling
between ‘P - 8’ and ‘Q - V’, the following decoupled equations result
(4.43)
(4.44)

It has been found that the latter equation is relatively unstable at some distance
from the exact solution due to the non-linear defining functions. An improvement in
convergence is obtained by replacing this with the polar current mismatch formulation
[71
[AI] = [D][AV]. (4.45)

Alternatively, the right-hand side of both Equations (4.43) and (4.44) is divided by
voltage magnitude V.
(4.46)
(4.47)

The equations are solved successively using the most up-to-date values of V and 8
available. [A] and [C] are sparse, non-symmetric in value and are both functions of V
and 8.
They must be calculated and triangulated each iteration.
100 4 LOADFLOW

Further approximations that can be made are to assume that Ek = 1.0 p.u. for all
buses, and G h << B h in calculating the Jacobian elements. The off-diagonal terms
then become symmetric about the leading diagonal.
The decoupled Newton method compares very favourably with the formal Newton
method. While reliability is just as high for ill-conditioned problems, the decoupled
method is simple and computationally efficient. Storage of the Jacobian and matrix
triangulation is saved by a factor of 4, or an overall saving of 30-40% on the formal
Newton load flow. Computation time per iteration is also less than for the Newton method.
However, the convergence characteristics of the decoupled method are linear, the
quadratic characteristics of the formal Newton being sacrificed. Thus, for high accura-
cies, more iterations are required. This is offset for practical accuracies by the fast initial
convergence of the method. Typically, voltage magnitudes converge to within 0.3%
of the final solution on the first iteration and may be used as a check for instability.
Phase angles converge more slowly than voltage magnitudes but the overall solution is
reached in 2-5 iterations. Adjusted solutions (the inclusion of transformer taps, phase
shifters, inter-area power transfers, Q and V limits) take many more iterations.
The Newton meihods can be expressed as follows

(4.48)

where
E = 1 for the full Newton-Raphson method,

E = 0 for the decoupled Newton algorithm.


A Taylor series expansion of the Jacobian about E = 0 results in a first-order approx-
imation of the Newton-Raphson method whereas the decoupled method is a zero
order approximation. The method has quadratic convergence properties because of the
coupling but retains storage requirements similar to those of the decoupled method.

4.10 Fast Decoupled Load Flow


By further simplifications and assumptions, based on the physical properties of a prac-
tical system, the Jacobians of the decoupled Newton load flow can be made constant
in value. This means that they need to be triangulated only once per solution or for a
particular network.
For ease of reference, the real and reactive power equations at a node k are repro-
duced here.
Pk = vk V m ( G k , COS 6km + B k m Sin O h ) , (4.49)
mek

(4.50)
4.10 FAST DECOUPLED LOAD FLOW 101

A decoupled method which directly relates powers and voltages is derived using the
series approximations for the trigonometric terms in Equations (4.49)and (4.50).

sin6=8--
e3
6'

The equations, over all buses, can be expressed in their simplified matrix form:
[A"] = [PI, (4.51)
CCl[Vl= [el, (4.52)
where P and Q are terms of real and reactive power, respectively, and

muJk

c k m = -Bkm, m # k,
tkm is the tap ratio if a transformer is in the line.
A modification suggested is to replace Equation (4.51)by
[Al[Gl= PI,
where

mok

Hence, [A]becomes constant valued,


A similar direct method is obtained from the ucoupled vc..age vectors method
(Equations (4.33)and (4.34)).If V m r vk are put as 1.0 p.u. for the calculation of
matrix [TI,then [TI becomes constant and need be triangulated once only. This same
simplification can be used in the decoupled voltage vectors and Newton's method of
Equations (4.40)and (4.41).
Fast decoupled load flow algorithms [ 1 11 are also derived from the Jacobian matrix
equations of Newton's method (Equations (4.42)),and the decoupled version (Equa-
tions (4.43)and (4.44)).
If we make the assumptions
(i) Vk, Vm = 1.0 P.u.,
(ii) G h << Bkm,and hence can be ignored (for most transmission line reactancehesis-
tance ratios, X / R >> 1).
102 4 LOADFLOW

(iii) cOS(8k - Qm)=l.O


sin(& - 8,)=0.0
since angle differences across transmission lines are small under normal loading
conditions
this leads to the decoupled equations
[ A P ] = [ B ] [ A e ] of order ( N - l), (4.53)
[ A Q ] = [ B ] [ A V ] of order ( N - M), (4.54)
where the elements of [El are
i?km = -Bkm for m # k,

and Bkm are the imaginary parts of the admittance matrix. To simplify still further, line
resistances may be neglected in the calculation of elements of [El.
An improvement over Equations (4.53)and (4.54)is based on the decoupled Equa-
tions (4.46)and (4.47)which have less non-linear defining functions. Applying the
same assumptions listed previously, we obtain the equations
(4.55)
(4.56)
Further refinements to this method are:
(a) Omit from the Jacobian in Equation (4.55)the representation of those network
elements that predominantly affect the reactive power flow, e.g. shunt reactances
and off-nominal in-phase transformer taps. Neglect also the series resistances of
lines.
(b) Omit from the Jacobian of Equation (4.56)the angle shifting effects of phase
shifters.
The resulting fast decoupled power flow equations are then
[ A P l V l = [B’l[AQl, (4.57)
[AQlV] = [B”][AV]. (4.58)
where
1
B;, = -- for m # k,
x km

mwk
4.10 FAST DECOUPLED LOAD FLOW 103

4 Data Input

1 No No

Solve 4.573nd
update (0)

I I

Solve 4.58 and update (8)


I

output
results

Figure 4.8 Flow diagram of the fast-decoupled load flow

The equations are solved alternatively using the most recent values of V and 6 available
as shown in Figure 4.8 [ l 11.
-
The matrices B’ and B” are real and are of order (N 1) and (N - M),respectively,
where N is the number of busbars and M is he number of P V busbars. B” is symmetric
in value and so is B’ if phase shifters are ignored; it is found that the performance of
the algorithm is not adversely affected. The elements of the matrices are constant and
need to be evaluated and triangulated only once for a network.
Convergence is geometric, 2-5 iterations are required for practical accuracies, and
more reliable than the formal Newton method. This is because the elements of B’
and B” are fixed approximations to the tangents of the defining functions A P / V and
AQ/V, and are not susceptible to any ‘humps’ in the defining functions,
104 4 LOADFLOW

If A P / V and A Q / V are computed efficiently, then the speed for iterations of the
fast decoupled method is about five times that of the formal Newton-Raphson or about
f that of the Gauss-Seidel method, Storage requirements are about 60% of the formal
Newton, but slightly more than the decoupled Newton method.
Changes in system configurations are easily effected, and while adjusted solutions take
many more iterations, these are short in time and the overall solution time is still low.
The fast decoupled Newton load flow can be used in optimization studies for a
network and is particularly useful for accurate information of both real and reactive
power for multiple load flow studies, as in contingency evaluation for system security
assessment.

4.11 Convergence Criteria and Tests [13]


The problem arises in the load flow solution of deciding when the process has converged
with sufficient accuracy. In the general field of numerical analysis, the accuracy of
solution of any set of equations F ( X ) = 0 is tested by computing the ‘residual’ vector
F ( X p ) . The elements of this vector should all be suitably small for adequate accuracy
but how small is, to a large extent, a matter of experience of the requirements of the
particular problem.
The normal criterion for convergence in load flow is that the busbar power-
mismatches should be small, i.e. AQi and/or APi, depending upon the type of busbar
i, and can take different forms, e.g.
(APiI 5 cI for all PQ and PV busbars, (4.59)
1AQil 5 c2 for all PQ busbars,
where c1 and c2 are small empirical constants, and C I = c2 usually. The value of
c used in practice varies from system to system and from problem to problem. In a
large system, c = 1 MWMVAR typically gives reasonable accuracy for most purposes.
Higher accuracy, say c = 0.1 MWMVAR may be needed for special studies, such as
load flows preceding transient stability calculations. In smaller systems, or systems
at light load, the value of c may be reduced. For approximate load flows, c may be
increased but with some danger of obtaining a meaningless solution if it becomes
too large. Faced with this uncertainty, there is thus a tendency to use smaller values
of c than are strictly necessary. The criterion (Equation (4.59)) is probably the most
common in use. A popular variant on it is
(4.60)
i k
and other similar expressions are also being used.
In the Newton-Raphson algorithms, the calculation and testing of the mismatches
at each iteration are part of the algorithm.
The set of equations defining the load flow problem has multiple solutions, only one
of which corresponds to the physical mode of operation of the system. It is extremely
rare for there to be more than one solution in the neighbourhood of the initial estimates
+
for the busbar voltages (( 1 j 0 ) P.u., in the absence of anything better), and apart from
the possibility of data errors, a sensible-looking mathematically-converged solution
is normally accepted as being the correct one. However, infrequent cases of very
4.13 LOAD FLOW FOR STABILITY ASSESSMENT 105

ill-conditioned networks and systems operating close to their stability limits arise where
two or more mathematically-converged solutions of feasible appearance can be obtained
by different choices of starting voltages, or by different load flow algorithms.
A load flow problem whose data correspond to a physically unstable system operating
condition (often due to data errors, or in the investigation of unusual operating modes,
or in system planning studies) usually diverges. However, the more powerful solution
methods, and in particular Newton-Raphson, will sometimes produce a converged
solution, and it is not always easy in such cases to recognize that the solution is a
physically-unstable operating condition. Certain simple checks, e.g. on the transmission
angle and the voltage drop across each line, can be included in the program to monitor
the solution automatically .
Finally, a practical load flow program should include some automatic test to discon-
tinue the solution if it is diverging, to avoid unnecessary waste of computation, and to
avoid overflow in the computer. A suitable test is to check at each iteration whether
any voltage magnitude is outside the arbitrary range 0.5- 1.5 P.u., since it is highly
unlikely in any practical power system that a meaningful voltage solution lies outside
this range.

4.12 Numerical Example


A reduced version of the New Zealand system is used here to test the symmetrical
load flow algorithm.
The test network, illustrated in Figure 4.9, involves the main generating and loading
points of the New Zealand South Island, with the HVDC converter represented as a
load, i.e. by specified P and Q.
The computer print-out (pages 107- 109) illustrates the numerical input and output
information for the specified conditions.

4.13 Load Flow for Stability Assessment


It will be seen in Chapter 7 that load flow is an integral part of rotor angle stability
and voltage stability assessment. A partial loss of generation or transmission in a
region of heavy load also causes immediate voltage sags. The initial voltage behaviour
(transient voltage stability) is assessed by dynamic simulation using the models and
methodology described in Chapter 7. After 10 to 30 s following the disturbance, the
system will normally return to a quiescent state, the synchronizing swings having been
damped out. From then on, voltage stability can be approximately assessed solely by
means of power flow snapshots in time.
However, the power flow method and the representation of the components involved
need special consideration.

4.13.1 Post-disturbance power flows [14]


Regarding the applicability of the solution methods, decoupling techniques and fast
decoupling in particular, must be used with caution in stability studies because some of
the approximations regarding the phase angle will not hold. Decoupling can also cause
convergence difficulties in networks with high R/X ratios. This is important for voltage
106 4 LOADFLOW

lslington 220

Bromley 220 --
Bus voltage
0.9940, -44.73

'wizel---220

Benmore--Ol&
-
Benmore -220

-
Aviemore -220 , Line MVA
97.39. -41.1 1

Line MVA
Bus type: SLK

Livingstn 220

721.7,242.4
\ invercarg 220

Tiwai ----220

Manapouri 220

Figure 4.9 Reduced primary a.c. system for the South Island of New Zealand
4.13 LOAD FLOW FOR STABILITY ASSESSMENT 107

WAD FLOW PROCLAN

DEPARTNENT OF ELECTRICAL k ELECTRONIC ENCIfEERING. UNIVELSITY OF C ANTEhBuLY. NEV ZEALAQ


SYSTEI NO. 2 23 N A l 90

THE SLACK BUS IS 6


IAXINUN hUNBER OF ITERATIONS 10 NUNBEl OF BUSES 17
POYE1 TULELANCE .00100 NUNBEI OF LINES 20
PI INTO^ INDICATOR 000000000 NO OF TUNSFOHEPS 6

BUS D A T A

MAD GEMXATION IININUl NMIMUN sm


BUS NAME TYPE VOLTS NW NVM NW IVAI NVAI IVAI SUSCEPTANCE

1 INVEICARG22O 0 0.0000 200.00 51.00 0.00 0.00 0.00 0.00 0.000


2 IOXBURCK-220 0 O.oo00 150.00 60.00 0.00 0.00 0.00 0.00 0.000
3 MANAPOUR1220 0 O.OOO0 0.00 0.00 0.00 0.00 0.00 0.00 0.000
4 NANAPOURIO14 1 1.0450 0.00 0.00 690.00 0.00 0.00 0.00 0.000
5 TIWAI-220 0 0,M)OO 420.00 185.00 0.00 0.00 0.00 0.00 0.000

6 lOXBuLCKd11 1 1.0500 0.00 0.00 0.00 0.00 0.00 0.00 0.000


7 BEHIOLE-220 0 0.0000 500.00 200.00 0.00 0.00 0.00 0.00 0.000
8 BMIOIJ%OIG 1 1.0600 0.00 0.00 0.00 0.00 0.00 0.00 0.000
9 AVIENORS220 0 0.0000 0.00 0.00 0.00 0.00 0.00 0.00 0.000
10 AVIENOLeOll 1 1.0450 0.00 0.00 200.00 0.00 0.00 0.00 0.000

11 OIIAWYSTEN I 1.0500 0.00 0.00 350.00 0.00 0.00 0.00 0.000


I2 LIVINCSTN220 0 0.0000 150.00 60.00 0.00 0.00 0.00 0.00 0.000
13 ISLIt(cTON220 1 1.oooO 500.00 300.00 0.00 0.00 0.00 0.00 0.000
14 BWNLEY-220 0 0.0000 100.00 60.00 0.00 0.00 0.00 0.00 0.000
15 TEKAPo---OlI 0 1.0500 0.00 0.00 150.00 0.00 0.00 0.00 0.000
16 TEXAPD-220 0 0,0000 0.00 0.00 0.00 0.00 0 .OO 0.00 0.000
17 W I Z E G Z 2 0 0 O.OOO0 0.00 0.00 0.00 0.00 0.00 0.00 0.000

LINE D A T A

BUS NAME BUS NAlE RESISTANE REACTANCE SUSCEPTANCE

1 ISVERCARC220 3 NANAPOUR1220 0.01300 0.09000 0.25000


1 IoWERCARC220 3 IAX4POU11220 0.01300 0.09000 0.25000
3 llANAPOUII220 5 TIYAI-220 0.01Ooo 0.10000 0.29000
3 NA,YAPOUR1220 5 TIVAI-220 0.0 1000 0.10000 0.29000
1 INVEICARG220 5 TIWAI-220 0.00200 0.01000 0.04000

1 INVERCARC220 5 TIWAI-220 0.00200 0.01000 0.04000


1 ItiVERCAIC220 2 ROXBIRCH-220 0 I 01000 0.11000 0.17000
2 ROXBURGH-220 17 TYIZEd220 0.01600 0.14000 0.24000
2 ROXBURCH-220 17 TWIZEG220 0.01600 0.14000 0.24000
2 ROXBURCH-220 12 LIVINGSTN220 0.03000 0.12000 0.18000

7 BEMORE-220 17 'TYI2EG220 0.00400 0.03000 0.07000


12 LIVlfCSTN220 9 AVIENORE-220 0.00700 0.03000 0.05000
9 AVIEIORE-220 7 BENNORE-220 0.00400 0.05000 0.02000
9 AVIEIOOE-220 7 BENNOW-220 0.00400 0.05000 0.02000
12 LIVIbCSTN220 13 ISLINGTUN220 0.03000 0.18000 0.35000
108 4 LOADFLOW

17 TVIZEG220 16 TEKAPL220 0.00200 0.01000 0.02000


16 TEKAPO-220 13 ISLINCTON220 0.02000 0.13000 0.35000
17 TVIZEb220 14 BROILEY-220 0.02000 0.14000 0.45000
14 BROIILEY-220 13 ISLINCTON220 0.00200 0.01000 0.05000
17 T V I Z E b 2 2 0 13 ISLINCT0#220 0.02000 0,14000 0.45000

T R A N S F O R I E B D A T A
BUS NAME BUS NAME RESISTANCE REACTANCE TAP CODE

3 IAIYAPOURI220 4 IANAPOURIOl4 0.00060 0.01600 1.000 0


2 ROXBUllCH-220 6 ROXBUXCH-OII 0.00200 0.04000 1.000 0
17 T V I Z E d 2 2 0 11 OHAO-SYSTEI 0.00400 0.03200 1.000 0
9 AVIEIORb220 10 AVIEIORF.-Oll 0.00150 0.04500 1.000 0
7 BUMORE-220 8 BENI011E--016 0.00120 0.03200 1.000 0
16 TUAPO-220 15 TEKAPO---Oll 0.00300 0.05600 1.000 0

SOLUTION CONYELCED IN 5 P 4 AND 5 Q-V ITEXATIONS


LOAD CE8F.LATION AC WSSES IISIATCn SHVMS
IV IVAL IV IVAL nv nvu IV WAR IVAL

2020.00 916.00 2113.71 1420.67 93.92 504.69 -0.21 -0.02 0.00

POVER TUNSFUS

BUS DATA
CENXLATION LOAD SHlMI
BUS NAIE YOLTS ANGLE IV IVAL IV IVAL #VAL BUS NAP4 IV WAR

1 INVElCALG220 0.936 -12.26 0.00 0.00 200.00 Sl.00 0.00


3 IANAPOUL1220 -174.88 40.45
3 lANAPouII220 -174.88 40.45
5 TIVAI-220 49.34 39 .'6S
5 TIVAI-220 49.34 39.65
2 ROIBULQI-220 51.09 -49.40

2 lOXBULCIc220 0.982 -16.02 0.00 0.00 150.00 60.00 0.00


IISI A M -0.014
-
4.004

1 LmELCALG220 40.59 39.24


17 fyIZEb220 184.16 -25.51
I7 m z e c - 2 2 0 184.16 -25.51
12 LIvIMsIw220 245.35 -17.18
.6 M I B U L ~ 1 1 -7 13. I4 41.03
nxsuTcn 0.076 4.009
3 IANAPOUU220 1.002 -2.84 0.00 0.00 0.00 0.00 0.00
1 IWC11C220 179.54 49.24
1 INyEICIIC220 179.54 49.24
5 TIVAI-220 163.87 54.14
5 TIVAI-'220 163.87 54.14
4 IANAPOULIOI4 486.91 -206.77

4 IANAPOUU014 2.045 5.12 690.00 288.73 0.00 0.00 0.00


IISIATCH 0.088 4.003
-
3 IANAPOUII220 689.98 288.73
1IslAm 0.020 0.000
5 TIVAI-020 0.931 -12.53 0.00 0.00 420.00 185.00 0.00
3 IANAPOvII220 -160.72 49.83
3 IANAPOULI220 -160.72 49.83
1 I)NuCUC220 -49.24 42.66
1 MVELCARG220 49.24 42.66
IISIATCH -0.067 -0.016
4.13 LOAD FLOW FOR STABILITY ASSESSMENT 109

6 lOlBuLCw-011 1 .ow 0.000 723.71 212.37 0.00 0.00 0.00


2 IOXBuLCll-220 723.71 242.37
IISIATCU 0.MM o.oO0
7 BMIOLC220 0.993 -36.85 0.00 0.00 500.00 200.00 0.00
WIZEL--220
17 -323.19 6.63
AVXEIOIE-220 -88.64
9 1.28
AVIWOIE-220 -88.64
9 1.28
8
BENMOW16 0.53 -209.19
IISIATCU -0.061 4.005
8 B W I O W 1 6 1.080 -37.00 0.00 223.40 0.00 0.00 0.00
7 BAIOLC220 0.01 223.40

9 AVIEIOLC220 0.996 -34.28 0.00 0.00 0.00 0.00 0.00


IISIAKU -0.006
-
0.000

12 LIVIRSTN220
21.37 92.02
7 88.96
BENNO&-220 0.73
BWIOlE-220
7 88.96 0.73
AVIEIOLeOll -199.26
10 -93.49
IISIATCY -0.023 -o).ooo
10 AVIEIOLCOII 1.045 -29.41 200.00 115.46 0.00 0.00 0.00
9 AVIEIOIE-220 199.99 115.46
IISIATCII 0.007 0.000
11 OlUUSVSTEI 1.050 -25.43 350.00 113.38 0.00 0.00 0.00 ~

I7 l V I Z E d 2 2 0 350.00 113.38
IISIATCH -0.004 o.oO0
12 LIVlNCSl"220 0.966 -34.27 0.00 0.00 150.00 60.00 0.00 ~~

2 MSBuLGll-220 -226.60 75.10


9 AVIUOLCz20 -20.71 -94.00
13 ISLINCToN220 97.39 -41.11
IISIATCU -0.082 0.006
1 .ooo 4 5 . 1 5 0.00 437.32 500.00 300.00 0.00 ~

12 LIVIffiSTN220 -94.14 26.75


16 TEIAP0--120 -176.86 26.13
14 BMILEY-220 -61.68 67.19
17 W I Z E d 2 2 0 -167.23 17.24
IISIAtcII -0.085 0 .ow
14 BMILEY-220 0.994 44.73 0.00 0.00 100.00 60.00 0.00
17 TYIZEd220 -161.84 11.30
13 ISLIwc10N220 61.85 -71.30
IISIATCM -0.004 0.002
15 TEyAP(Ml1 I . 008 -26.72 150.00 0.00 0.00 0.00 0.00
16 "EXAPO-220 149.98 0.00
IISIATCU 0.017 -0.000
16 TUAPO-220 1.007 -31.47 0.00 0.00 0.00 0.00 0.00
17 W I Z E d 2 2 0 4 4 . 1 7 5.86
13 XSLIwc10NZZO 183.50 -18.25
15 T E I A P W I I -149.32 12.39

17 T V I Z E b 2 2 0 I .007 -31.27 0.00 0.00 0.00 0.00 0.00


IISIAtcll -0.007
- 0.001

2 MIBUICH-220 -I 78.50 51.27


2 MIBUIGII-220 -178.50 51.27
7 BWIOLE-220 327.44 18.21
16 TEKAPO-220 34.19 -7.77
14 UMILEY-220 167.31 -17.69
13 I S L I ~ N 2 2 0 173.14 -21.21
11 MIAU-SVSTEB -345.09 -74.09
IISIATCY -0.052 0.010

IWE IAXIlUI IlSIATQ IS 0.0881 YVA ON BVS 3 (IANAwuLI220)


IWE SLACK BUS CEWEUTION IS 723.709 IV 242.372 IVAI
110 4 LOADFLOW

stability assessment, where the distribution circuits are normally modelled in detail.
However, voltage stability related decoupled formulations have been proposed [ 15,
161 which can overcome these problems.

4.13.2 Modelling techniques


In the time frame between 10 and 30 s following the disturbance, the prime mover
controls respond to the speed variation and the frequency partially recovers, with all
the interconnected machine governors sharing the variation of power. This situation
has been labelled ‘governor power flow’ and represents the timeframe of longer-
term voltage stability.
After 2 or 3 min, transformers OLTC restore the voltage to sensitive loads.
The OLTC action results in transmission current increases which cause generator
terminal voltage sags and forces over-excitation; AVRs may then reach their limits.
In this timeframe, the voltage sensitive areas of the network must be modelled
in detail, including the action of transformers OLTC and of switchable capacitor
banks.
Nodal specification of load and generation should be separated; also, reactive power
compensation must be modelled separately from the load reactive power.
The generator representation (in the voltage sensitive areas) must include the reac-
tive power limits. Instead of the conventional PQ specification, in this region it is
better to model the generator by a voltage source behind a saturated synchronous
reactance. Also, as the terminal voltage drops, the armature current overload protec-
tion must be included.
As far as loads are concerned, the traditional fixed PQ specification is not always
appropriate for voltage stability analysis, where more detailed representation of the
distribution system is required.
Reference [17] provides useful information on the procedure for the power flow
simulation and on the use and interpretation of most power flow V - Q and P - V
curves for the assessment of long-term voltage stability.

4.13.3 Sensitivity analysis [17]


Once power flow convergence has been reached, the Jacobian matrix provides a
linearized model of the system around the operating point and the sensitivity of the
voltages to incremental changes in real and reactive power can be easily calculated.
This information is also useful to determine the type of control and the size and loca-
tion of reactive power compensation equipment. However, it must be kept in mind that
the linearized model is only valid for small power changes.

4.14 Three-phase Load Flow


For most purposes in the steady-state analysis of power systems, the system unbalance
can be ignored and the single-phase analysis described in previous sections is adequate.
However, in practice it is uneconomical to balance the load completely or to achieve
4.14 THREE-PHASE LOAD FLOW 111

perfectly balanced transmission system impedance, as a result of untransposed high


voltage lines and lines sharing the same right-of-way for considerable lengths.
A realistic assessment of the unbalanced operation of an interconnected system,
including the influence of any significant load unbalance, requires the use of three-
phase load flow algorithms [18-201. The object of the three-phase load flow is to
find the state of the three-phase power system under the specified conditions of load,
generation and system configuration.
The basic three-phase models of system plant and the rules for their combination
into overall network admittance matrices, discussed in Chapter 2, can be used as the
framework for the three-phase load flow. This is illustrated in Figure 4.10 for the
network shown in the lower part of the diagram in Figure 4.9.
The storage and computational requirements of a three-phase load flow program are
much greater than those of the corresponding single-phase case. The need for efficient
algorithms is, therefore, significant even though in contrast to single-phase analysis, the
three-phase load flow is likely to remain a planning, rather than an operational, exercise.

4.14.1 Notation
A clear and unambiguous identification of the three-phase vector and matrix elements
requires a suitable symbolic notation using superscripts and subscripts.
The a s . system is considered to have a total of n busbars, where:
n =nb+ng,
nb is the number of actual system busbars;
n g is the number of synchronous machines.

Subscripts i, j , etc. refer to system busbars:


i = 1, nb identifies all actual system busbars, i.e. all load busbars plus all gener-
ator terminal busbars;
i = nb + ng identifies the internal busbar at the slack machine;
+ +
i = nb 1, nb ng - 1 identifies all generator internal busbars with the excep-
tion of the slack machine.
The following subscripts are also used for clarity:
reg refers to a voltage regulator;
int refers to an internal busbar at a generator;
gen refers to a generator.
Superscripts p , m identify the three phases at a particular busbar.

4.14.2 Synchronous machine modelling


Synchronous machines are designed for maximum symmetry of the phase windings and
are, therefore, adequately modelled by their sequence impedances. Such impedances
112 4 LOADFLOW

Figure 4.10 Test system 3 x 3 matrix representation

contain all the information that is required to analyse the steady-state unbalanced
behaviour of the synchronous machine.
The representation of the generator in phase components may be derived from the
sequence impedance matrix (Zg)012 as follows:
4.14 THREE-PHASE LOAD FLOW 113

where
[T,] = [ I1 a 2
1 a
“1.
a2
(4.63)

and a is the complex operator ej2x/3.The phase component impedance matrix is thus:

[Zglabc = I zo + U2ZI + a Z 2 I zo + ZI + z 2 I zo + UZI + a 2 Z 2 1.

(4.64)
The phase component model of the generator is illustrated in Figure 4.11(a). The
machine excitation acts symmetrically on the three phases and the voltages at the
internal or excitation busbar form a balanced three-phase set, i.e.
E i = E i = E; (4.65)
and
(4.66)

For three-phase load flow, the voltage regulator must be accurately modelled as it
influences the machine operation under unbalanced conditions. The voltage regulator
monitors the terminal voltages of the machine and controls the excitation voltage
according to some predetermined function of the terminal voltages. Often, the positive
sequence is extracted from the three-phase voltage measurement using a sequence filter.
Before proceeding further, it is instructive to consider the generator modelling from
a symmetrical component frame of reference. The sequence network model of the
generator is illustrated in Figure 4.1 l(b). As the machine excitation acts symmetrically
on the three phases, positive sequence voltages only are present at the internal busbar.
The influence of the generator upon the unbalanced system is known if the voltages
at the terminal busbar are known. In terms of sequence voltages, the positive sequence
voltage may be obtained from the excitation and the positive sequence voltage drop
caused by the flow of positive sequence currents through the positive sequence reac-
tance. The negative and zero sequence voltages are derived from the flow of their
respective currents through their respective impedances. It is important to note that the
negative and zero sequence voltages are not influenced by the excitation or positive
sequence impedance.
There are infinite combinations of machine excitation and machine positive sequence
reactance which will satisfy the conditions at the machine terminals and give the correct
positive sequence voltage. Whenever the machine excitation must be known (as in fault
studies) the actual positive sequence impedance must be used. For load flow, however,
the excitation is not of any particular interest and the positive sequence impedance
may be arbitrarily assigned to any value. The positive sequence impedance is usually
set to zero for these studies.
Thus, the practice with regard to three-phase load flow in phase co-ordinates is to
set the positive sequence reactance to a small value in order to reduce the excitation
114 4 LOADFLOW

~ ~ +ve
V term, -
sequence 1

V term,
-vesequence

zero sequence

Figure 4.11 Synchronous machine models. (a) Phase component representation. (b) Symmet-
rical component representation

voltage to the same order as the usual system voltages with a corresponding reduction
in the angle between the internal busbar and the terminal busbar. Both these features
are important when a fast decoupled algorithm is used.
Therefore, in forming the phase component generator model using Equation (4.64),
an arbitrary value may be used for 21but the actual values are used for 20 and 22.
4.14 THREE-PHASE LOAD FLOW 115

There is no loss of relevant information as the influence of the generator upon the
unbalanced system is accurately modelled.
The nodal admittance matrix, relating the injected currents at the generator busbars
to their nodal voltages, is given by the inverse of the series impedance matrix derived
from Equation (4.64).

4.14.3 Specified variables


The following variables form a minimum and sufficient set to define the three-phase
system under steady-state operations:
0 The slack generator internal busbar voltage magnitude Vint ,where j = nb + ng.
(The angle Elint is taken as a reference.)
0 ,
The internal busbar voltage magnitude Vint and angles 8ini j at all other generators,
+
i.e. j = 1, nb + 1, nb ng - 1.
0 The three voltage magnitudes ( V r ) and angles (0:) at every generator terminal
busbar and every load busbar in the system, i.e. i = 1, nb and p = 1, 3.
Only two variables are associated with each generator internal busbar as the three-
phase voltages are balanced and there is no need for retaining the redundant voltages
and angles as variables. However, these variables are retained to facilitate the calcula-
tion of the real and reactive power mismatches. The equations necessary to solve the
above set of variables are derived from the specified operating conditions, i.e.
0 The individual phase real and reactive power loading at every system busbar.
0 The voltage regulator specification for every synchronous machine.
0 The total real power generation of each synchronous machine, with the exception
of the slack machine.
The usual load flow specification of a slack machine, i.e. fixed voltage in phase and
magnitude, is applicable to the three-phase load flow.

4.14.4 Derivation of equations


The three-phase system behaviour is described by the equation

where the system admittance matrix [ Y ] , as developed in Chapter 2, represents each


phase independently and models all inductive and capacitive mutual couplings between
phases and between circuits. The mathematical statement of the specified conditions is
derived in terms of the system admittance matrix

[YI = [GI + M I
as follows:
116 4 LOADFLOW

(i) For each of the three phases (p) at every load and generator terminal busbar (i),
APP = (Pf)’p - pf
n 3
= (pp)$P - vf 1v;[Gip,” cos 0:’” + sin O i m ] (4.68)
k = l m=l

and

k=l m=l

(ii) For every generator j ,

where k is the bus number of the jth generator’s terminal busbar.


(iii) For every generator j , with the exception of the slack machine, i.e. j # n b + ng.
3 n 3

p= 1 k=l m=l

(4.71)
where, although the summation fork is over all system busbars, the mutual terms
Gjk and Bjk are non-zero only when k is the terminal busbar of the jth generator.

It should be noted that the real power specified for the generator is the total real power
at the internal or excitation busbar whereas in actual practice the specified quantity is
the power leaving the terminal busbar. This, in effect, means that the generator’s real
power loss is ignored.
The generator losses have no significant influence on the system operation and may
be calculated from the sequence impedances at the end of the load flow solution, when
all generator sequence currents have been found. Any other method would require the
real power mismatch to be written at busbars remote from the variable in question,
i.e. the angle at the internal busbar. In addition, inspection of Equations (4.68) and
(4.71) will show that the equations are identical except for the summation over the
three phases at the generator internal busbar.
That is, the sum of the powers leaving the generator may be calculated in exactly
the same way and by the same subroutines as the power mismatches at other system
busbars. This is possible because the generator internal busbar is not connected to any
other element in the system. Inspection of the Jacobian submatrices derived later will
show that this feature is retained throughout the study. In terms of programming, the
generators present no additional complexity.
Equations (4.68) to (4.7 1) form the mathematical formulation of the three-phase load
flow as a set of independent algebraic equations in terms of the system variables.
4.14 THREE-PHASE LOAD FLOW 117

The solution to the load flow problem is the set of variables which, upon substitution,
make the left-hand side mismatches in Equations (4.68) to (4.7 1) equal to zero.

4.14.5 Decoupled three-phase algorithm [21]


The standard Newton-Raphson algorithm may be used to solve Equations (4.68) to
(4.71). This involves an iterative solution of the matrix equation:
A E l M

C G K P AV/V
D H L R A Vint/ Vint
for the right-hand side vector of variable updates. The right-hand side matrix in Equa-
tion (4.72) is the Jacobian matrix of first order partial derivatives.
Following decoupled single-phase load flow practice, the effects of he on reactive
power flows and A V on real power flows are ignored. Equation (4.72) may, therefore,
be simplified by assigning
[I] = [MI= [ J ] = [ N ] = 0
and
[ C ] = [GI = 0.
In addition, the voltage regulator specification is assumed to be in terms of the
terminal voltage magnitudes only and, therefore,
[D]= [HI = 0.
Equation (4.72) may then be written in decoupled form as:

(4.73)

for 1, k = 1, nb and j , 1 = 1, ng - 1 (i.e. excluding the slack generator)


[ AQP ] =[K P] [ AVzI/Vr ] (4.74)
AVreg j L R AVint //Vint /

for 1, k = 1, nb and j , I = 1, ng (i.e. excluding the slack generator).


To enable further development of the algorithm, it is necessary to consider the
Jacobian submatrices in more detail.
In deriving these Jacobians from Equations (4.68) to (4.71), it must be remembered
that
v; = v; = v; = Vin, /,
2n 2n
e; = e; - - = e; + - = eintI ,
3 3
when 1 refers to a generator internal busbar.
118 4 LOADFLOW

The coefficients of matrix Equation (4.73) are:


[Aim] = [aAPp/%?]
or
~i~
= VPV;:[G~P," sin eim- ~i~
cos eim],
except for
AGm = -B$m(Vr)2 - QY,
[BZI = [aAPgen j /qI
3
= C Vint V; [~,41"sin:0; - B;: cos e;],
p= 1

[E$I = [aPp/%int 11
3
= C Vint 1 V;[G,~" sin e;m - B [ cos
~ e,';"],
m= I

[Fjrl = [apgen ,/%in, 11.

where [Fj]]= 0 for all j # 1 because the jth generator has no connection with the Ith
generator's internal busbar, and
3 3 3
[Fill = C(-BLp(Vint i)2 - QP) + C C(Vint i)2[Gr - BLmcos8Lm].
p= I m=l p=l
m#P

The coefficients of matrix Equation (4.74) are


-[Kim] = Vr[aAQP/aVr],
where
K i m = VrVf[G,"," sin8im - Bipk" cosBim],
except
KGm = -Bi!m(VY)2 + QF,
-[L'i",] = Vr[aAVreg j / a V r ] .
Let [LT] = Vr[L$]' where k is the terminal busbar of the jth generator and L> = 0
!J
otherwise.
-[Pi1 = Vint /[aAQf/vint /I
3
= vint C V ~ I G , 'sin; ~ e,';m- B [ cos
~ e;"],
m= I

-[Rjil = [aAvreg jlavint 11


= 0 for all j , I as the voltage regulator specification does not
explicitly include the variables Vint.
4.14 THREE-PHASE LOAD FLOW 119

Although the above expressions appear complex, their meaning and derivation are
similar to those of the usual single phase Jacobian elements.
Jacobian approximations Approximations similar to those applied to the single-
phase load flow are applicable to the Jacobian elements as follows:

(i) At all nodes (i.e. all phases of all busbars)


Qr << Bzm(Vr)2.
(ii) Between connected nodes of the same phase,
cosOzm= 1, i.e. qmis small,
and
GZmsin qm<< BT.
(iii) Moreover, the phase-angle unbalance at any busbar will be small and hence an
additional approximation applies to the three-phase system, i.e.
el: = f120" for p # m.
(iv) Finally, as a result of (ii) and (iii), the angle between different phases of connected
busbars will be approximately 120", i.e.
eim = f120" for p # m
or
~ 0 ~ =
8 -0.5
; ~
and
= f0.866.
These values are modified for the f 3 0 " phase-shift inherent in the star-delta
connection of three-phase transformers.
The final approximation (iv), necessary if the Jacobians are to be kept constant, is the
least valid, as the cosine and sine values change rapidly with small angle variations
around 120". This accounts for the slower convergence of the phase unbalance at
busbars as compared with that of the voltage magnitudes and angles.
It should be emphasized that these approximations apply to the Jacobian elements
only, i.e. they do not prejudice the accuracy of the solution nor do they restrict the
type of problem which may be attempted.
Applying approximations (i) to (iv) to the Jacobians and substituting into Equations
(4.73) and (4.74) yields
120 4 LOADFLOW

and

[2:j1 =
where

with
egm = 0,
eZm = 0,
6;"'= f120" for p # m.
All terms in the matrix [MI are constant, being derived solely from the system
admittance matrix. Matrix [MIis the same as matrix [-B] except for the off-diagonal
terms which connect nodes of different phases. These are modified by allowing for the
nominal 120"angle and also including the GZmsin f9*; terms.
The similarity in structure of all Jacobian submatrices reduces the programming
complexity normally found in three-phase load flows. This uniformity has been
achieved primarily by the method used to implement the three-phase generator
constraints.
The above derivation closely parallels the single-phase fast decoupled algorithm but
the added complexity of the notation obscures this feature. At the present stage, the
Jacobian elements in Equations (4.75) and (4.76) are identical except for those terms
which involve the additional features of the generator modelling.
These functions are more linear in terms of the voltage magnitude [TI than are
the functions [AP] and [ A a ] . In the Newton-Raphson and related constant Jacobian
methods, the reliability and speed of convergence improve with the linearity of the
defining functions. With this aim, Equations (4.75) and (4.76) are modified as follows:

0 The left-hand side defining functions are redefined as [APp/VPl, [APgen j/Vint j]
and [AQp/Vf].
0 In Equation (4.73, the remaining right-hand side V terms are set to 1 p a .
0 In Equation (4.76), the remaining right-hand side V terms are cancelled by the
corresponding terms in the right-hand side vector.
4.14 THREE-PHASE LOAD FLOW 121

Recalling that [L;]’ = [aAV,, , / a V r ] , as V,, is normally a simple linear function


of the terminal voltages, [L’] will be a constant matrix.
Therefore, the Jacobian matrices [B’] and [B”] in Equations (4.77) and (4.78) have
been approximated to constants.
Zero diagonal elements in Equation (4.78) may result from the ordering of the
equations and variables. This feature causes no problems if these diagonals are not
used as pivots until the rest of the matrix has been factorized (by which time, fill-in
terms will have appeared on the diagonal). This causes a minor loss of efficiency as it
inhibits optimal ordering for the complete matrix. Although this could be avoided by
reordering the equations, the extra program complexity is not justified.
Based on the reasoning of Stott and Alsac [ 111, which proved successful in the
single-phase load flow, the [B’] matrix in Equation (4.77) can be further modified by
omitting the representation of those elements that predominantly affect MVAR flows.
The capacitance matrix and its physical significance is illustrated in Figure 4.12 for
a single three-phase line. With n capacitively-coupled parallel lines, the matrix will be
3n x 3n.
In single-phase load flows, the shunt capacitance is the positive sequence capacitance
which is determined from both the phase-to-phase and the phase-to-earth capacitances
of the line. It, therefore, appears that the entire shunt capacitance matrix predominantly
affects MVAR flows only. Thus, following single-phase fast decoupling practice, the
representation of the entire shunt capacitance matrix is omitted in the formulation
of [B’].This increases dramatically the rate of real power convergence. However, as
compared with the balanced load flow, this convergence rate deteriorates requiring, on
average, twice as many iterations.
With capacitively-coupled three-phase lines, the interline capacitance influences the
positive sequence shunt capacitance. However, as the values of interline capacitances
are small in comparison with the self capacitance of the phases, their inclusion makes no

Cac

Figure 4.12 Shunt capacitance matrices


122 4 LOADFLOW

noticeable difference. The effective tap of f i introduced by the star-delta transformer


connection is interpreted as a nominal tap and is, therefore, included when forming the
[B'] matrix.
A further difficulty arises from the modelling of the star-g/delta transformer connec-
tion. The equivalent circuit, illustrated in section 2.8 shows that large shunt admittances
are effectively introduced into the system. When these are excluded from [B'], as for a
normal shunt element, divergence results. The entire transformer model must, therefore,
be included in both [B'] and [B"].
With the modifications described above, the two final algorithmic equations may be
concisely written, i.e.

(4.79)

(4.80)

The constant Jacobians [BL] and [B;] correspond to fixeb approximated tangent
slopes to the multidimensional surfaces defined by the left-hand side defining functions.
Equations (4.79) and (4.80) are then solved according to the iteration sequence
illustrated in Figure 4.13.

&I KQ=1
KP=

t No

Figure 4.13 Iteration sequence for three-phase ax. load flow


4.14 THREE-PHASE LOAD FLOW 123

Generator models and the fast decoupled algorithm The derivation of the fast
decoupled algorithm involves the use of several assumptions to enable the Jacobian
matrices to be approximated to constant. The same assumptions have been applied to
the excitation busbars associated with the generator model as are applied to the usual
system busbars. The validity of the assumptions regarding voltage magnitudes and
the angles between connected busbars depends upon the machine loading and positive
sequence reactance. As discussed in section 4.14.2, this reactance may be set to any
value without altering the load flow solution and a value may, therefore, be selected
to give the best algorithmic performance.
When the actual value of positive sequence reactance is used, the angle across
the generator and the magnitude of the excitation voltage both become compara-
tively large under full load operation. Angles in excess of 45" and excitation volt-
ages greater than 2.0 p.u. are not uncommon. Despite this considerable divergence
from assumed conditions, convergence is surprisingly good. Convergence difficulties
may occur at the slack generator and then only when it is modelled with a high
synchronous reactance (1.5 p.u. on machine rating) and with greater than 70% full
load power.
All other system generators, where the real power is specified, converge reliably but
somewhat slowly under similar conditions.
This deterioration in convergence rate and the limitation on the slack generator
loading may be avoided by setting the generator positive sequence reactance to an
artificially low value (say 0.01 p.u. on machine rating), a procedure which does not
involve any loss of relevant system information,

4.14.6 Structure of the computer program


The main components of the computer program are illustrated in Figure 4.14. The
approximate number of FORTRAN statements for each block is indicated in paren-
thesis. The main features of each block are described in the following sections.

Data input The data input routine implements the system modelling techniques
described in Chapter 2 and forms the system admittance model from the raw data for
each system component. Examples of representative data are given in Table 4.1 with
reference to the test system of Figure 4.10.
The structure and content of the constant Jacobians B' and B" are based upon the
system admittance matrix and are thus formed simultaneously with this matrix.
Both the system admittance matrix and the Jacobian matrices are stored and processed
using sparsity techniques which are structured in 3 x 3 matrix blocks to take full
advantage of the inherent block structure of the three-phase system matrices.

Factorization of constant Jacobians The heart of the load flow program is the
repeat solutions of Equations (4.79) and (4.80) as illustrated in Figure 4.13. These
equations are solved using sparsity techniques and near optimal ordering, as discussed
in section 4.7, or like those embodied in Zollenkopf's bifactorization [22]. The constant
Jacobians are factorized before the iteration sequence is initiated. The solution of each
equation within the iterative procedure is relatively fast, consisting only of the forward
and back substitution processes.
124 4 LOADFLOW

Data input
Form and store system
admittance matrix, [B']
and 18'7 Jacobian
matrices (1645)

Factorize [a]
and [S"]
(310 )

system voltages (200)

procedure (see Fig. 4.13)


(450)

voltages, line power flows


and total system losses
(215)

Figure 4.14 Program structure

Starting values Starting values are assigned as follows:


0 The non-voltage controlled busbars are assigned 1 p.u. on all phases.
0 At generator terminal busbars all voltages are assigned values according to the
voltage regulator specifications.
0 +
All system busbar angles are assigned 0, - 120", 120" for the three phases, respec-
tively.
0 The generator internal voltages and angles are calculated from the specified real
power and, in the absence of better estimates, by assuming zero reactive power.
For the slack machine, the real power is estimated as the difference between total
load and total generation plus a small percentage (say, 8%) of the total load to
allow for losses.
For cases where convergence is excessively slow or difficult, it is recommended
to use the results of a single-phase load flow to establish the starting values. The
values will, under normal steady state unbalance, provide excellent estimates for all
4.14 THREE-PHASELOAD FLOW 125

Table 4.1 Representative data


Table of generator data
Generator Zero Impedance Pos. Impedance Neg. Impedance P Voltage
No.Name RO xo R1 XI R2 x2 (P.u.) regulator
Vphasc,

1 MAN014 0.0 0.080 0.0 0.010 0.0 0.021 15.OOO 1.045


2 ROXOll 0.0 0.150 0.0 0.010 0.0 0.091 SLACK 1.050

Transformer Data

Busbar names Leakage Tap ratio


reactance primary
primary secondary
MAN220 MAN014 0.0006 + jO.0164 0.045
ROX220 ROXO 1 1 0.0020 + j0.038 0.022

Line Data INV 220-ROX 220

+ j0.045
a 0.006 0.002 + j0.015 0.001 + j0.017
[Z,] = b 0.002 + jO.015 0.006 + j0.050 0.002 + j0.017
c I 0.001 + j0.017 1 0.002 + j0.017 I 0.007 + j0.047 I
a 0.0 + j0.35 0.0 - j.06 0.0 - jO.04

[Y,] = b 0.0 - j0.06 0.0 + j0.352 0.0 - j0.06


I
c 0.0 - jO.04 I 0.0 - j0.06 I O.O+ j0.34 I
Double Line: INV 220-TIW 220
Series Impedance

Line 1 Line 2
a b C a b C

Line 1 b

Line 2 b

C
126 4 LOADFLOW

Table 4.1 (continued)


Shunt Admittance

a +j0.045
Line 1 b -j0.008 +j0.040
c I -jO.ooS I -jO.Oll I Sj0.035 I I I 1
a -j0.007 -j0.003 -j0.003 +j0.044
Line 2 b -j0.003 -j0.005 -j0.002 -jO.01 fj0.040
-j0.002 -j0.002 -j0.004 -jO.Ol -jO.Oll +j0.036

Table 4.2 Busbar results


No. Busbar Phase A Phase B Phase C Generation
name Total
Volt Ang Volt Ang Volt Ang
1 INV220 1.0173 21.36 1.0509 -98.16 1.0351 139.44 0.0 0.0
2 ROX220 1.0319 23.30 1.0730 -96.18 1.0449 141.76 0.0 0.0
3 MAN220 1.0693 25.34 1.0816 -95.21 1.0641 144.34 0.0 0.0
4 MAN014 1.0450 -0.79 1.0545 -120.64 1.0522 118.84 0.0 0.0
5 TIW220 1.0137 21.08 1.0434 -98.61 1.0316 138.98 0.0 0.0
6 ROXOll 1.0500 -1.79 1.0653 -120.57 1.0771 118.12 0.0 0.0
7 MAN.GN 1.0669 1.69 1.0669 -118.31 1.0669 121.69 500.000 185.804
8 ROX.GN 1.0738 0.0 1.0738 -120.00 1.0738 120.00 281.277 108.106

Table 4.3 Selected print-out power flows


Sending end Busbar Receiving end Sending end Receiving end
Busbar
MW MVA R MW MVAR
No. Name No. Name
4 MAN 7 MAN.GN -163.583 -62.676 164.077 71.179
- 160.184 -47.925 159.968 55.047
- 176.232 -50.0050 175.955 59.577
3 MAN220 5 TIW220 34.710 10.919 -34.135 -19.871
33.977 8.91 1 -32.640 -19.255
38.172 6.260 -37.730 - 15.979

voltages and angles including generator internal conditions which are calculated from
the single-phase real and reactive power conditions.
Moreover, as a three-phase iteration is more costly than a single-phase iteration, this
practice can be generally recommended to provide more efficient overall convergence
and to enable the more obvious data errors to be detected at an early stage.
4.15 REFERENCES 127

For the purpose of investigating the load flow performance, flat voltage and angle
values are used in the examples that follow.

Iterative solution The iterative solution process (Figure 4.13) yields the values of
the system voltages which satisfy the specified system conditions of load, generation
and system configuration.

Output results The three-phase busbar voltages, the line power flows and the total
system losses are calculated and printed out. An example is given in Tables 4.2 and 4.3.

4.15 References

1. Stott, B, (1974). Review of load-flow calculation methods, Proceedings of the IEEE, 62,
(7), 916-929.
2. Ward, J B and Hale, H W, (1956). Digital computer solution of power-flow problems,
Transactions of AIEE, PAS-75, 398-404.
3. Brown, H E, Carter, G K, Happ, H H and Person, C E, (1963). ‘Power-flow solution by
impedance matrix iterative method’, IEEE Transactions of PAS-82, 1- 10.
4. Van Ness, J E and Griffin, J H, (1961). ‘Elimination methods for load-flow studies’, Trans-
actions of AIEE, PAS-80, 299-304.
5 . Tinney, W F and Hart, C E, (1967). ‘Power flow solution by Newton’s method’, IEEE
Transactions on Power Apparatus and Systems, PAS-86, (1 l), 1449- 1460.
6. Ogbuobiri, E C, Tinney, W F and Walker, J W, (1970). ‘Sparsity-directed decomposition
for Gaussian elimination on matrices’, IEEE Transactions on Power Apparatus and Systems,
PAS-89, (I), 141-150
7. Stott, B and Hobson, E, (197 1). ‘Solution of large power-system networks by ordered elim-
ination: A comparison of ordering schemes’, Proceedings ofthe IEE, 118, (1). 125- 134.
8. Tinney, W F and Walker, J W, (1967). ‘Direct solutions of sparse network equations by
optimally ordered triangular factorization’, Proceedings of the IEEE, 55, (1 I), 1801- 1809.
9. Despotovic, S T, (1974). ‘A new decoupled load-flow method’, IEEE Transactions on Power
Apparatus and Systems, PAS-93, (3), 884-891.
10. Stott, B, (1972). ‘Decoupled Newton load flow’, IEEE Transactions on Power Apparatus
and Systems, PAS-91, 1955- 1959.
11. Stott, B and Alsac, 0, (1974). ‘Fast decoupled load flow’, IEEE Transactions on Power
Apparatus and Systems, PAS-93, (3), 859-869.
12. Medanic, J and Avramovic, B, (1975). ‘Solution of load-flow problems in power systems
by &-couplingmethod’, Proceedings ofthe IEE, 122, (8), 801 -805.
13. Stott, B, (1972). Power-system load flow, (M.Sc. lecture notes), University of Manchester
Institute of Science and Technology.
14. Kundur, P, (1993). Power System Stability and Control, McGraw-Hill, New York.
15. Carpentier, J L, (1986). CRIC, a new active-reactive decoupling process in load-flows,
optimal power flows and system control, Proceedings of the IFAC Conference on Power
System and Plant Control, Beijing, pp. 59-64.
16. Van Cutsem, T, (1991). ‘A method to compute reactive power margins with respect to
voltage collapse’, IEEE Transactions on Power Systems, PWRS-6, (2), 145- 156.
17. Taylor, C W, (1994). Power System Voltage Stability, McGraw-Hill International Edition,
New York.
18. El-Abiad, A H and Tarisi, D C, (1967). Load-flow solution of untransposed EHV networks,
PICA, Pittsburgh, Pa., 377-384.
128 4 LOADFLOW

19. Wasley, R G and Slash, M A, (1974). ‘Newton-Raphson algorithm for three-phase load
flow’, Proceedings of the IEE, 121, (3,630.
20. Birt, K A, Graffy, J J, McDonald, J D and El-Abiad, A H, (1976). ‘Three-phase load-flow
program’, IEEE Transactions on Power Apparatus and Systems, PAS-95,59.
21. Arrillaga, J and Harker, B J, (1978). ‘Fast decoupled three-phase load flow’, Proceedings
of rhe IEE, 125, ( 8 ) , 734-740.
22. Zollenkopf, K, (1970). Bifactorization-basic computational algorithm and programming
techniques, Conference on Large Sets of Sparse Linear Equations, Oxford, pp. 75-96.
5
LOAD FLOW UNDER POWER
ELECTRONIC CONTROL

5.1 Introduction
The relatively recent availability of high voltage power electronic switching has led
to the development of HVDC transmission and FACTS technologies. Although the
original purpose of the former was a return to d.c. transmission for large distances,
most of the recently commissioned schemes are actually back-to-back; their purpose
is the interconnection of asynchronous power systems.
The FACTS technology, on the other hand, has developed to enhance the control-
lability of the synchronous transmission system, thus permitting its operation closer
to the stability limits. The basically different reasons for the existences of HVDC and
FACTS have important consequences in the load flow solutions.
This chapter describes the incorporation of these two power electronic technolo-
gies in conventional load flow solutions whenever possible, and describes alternative
methods of solution when such incorporation makes the load flow convergence difficult.

5.2 Incorporation of FACTS Devices


Chapter 3 has described the functions and steady-state models of the main FACTS
devices already in use or under serious consideration by electricity suppliers. The need
to work plant and transmission systems harder in the new deregulated environment
is likely to increase substantially the use of FACTS controllers. Accordingly, their
behaviour must be adequately represented in power system modelling.
As the variety of power electronic controllers in the network interact with each
other, the reliability of the solution takes precedence over the computational burden.
Moreover, the conditions leading to the acceptability of decoupled load flows do not
necessarily apply when the network operates under power electronics control. There-
fore, iterative methods with strong convergence characteristics are to be preferred.
Similarly, the use of a sequential solution between the a.c. network and each of the
FACTS devices present does not yield quadratic convergence because only the voltage
phasors are used as state variables, while the conditions of the controllable devices
are only updated at the end of each Newton iteration. This often leads to diverging
solutions.
Thus, the original full Newton-Raphson simultaneous technique, described in
Chapter 4, is used here to incorporate FACTS devices in the load flow solution.
130 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

Some FACTS devices present no special problem, as their steady-state behaviour


is adequately represented by the standard load flow specifications. This is the case
of static VAR compensation, whether of the TCR or STATCON types, which can be
specified by a zero active power and constant voltage magnitude, like the conventional
synchronous condensers. In general, however, the incorporation of FACTS devices
requires special consideration and is described in this section.

5.2.1 Static tap changing


The following transfer admittance matrix of a two-winding transformer with complex
taps on the primary and secondary windings has been developed in section 3.2.2 [l].

When applied to in-phase tap changing, the following substitutions must be made
in Equation (5.1):
T , = T: = T,LO",

and
u, = u; = UVLO".
For a tap changer connected between terminals k and m, to control the voltage
magnitude at bus k, the extra linearized load flow equations required at nodes k and
m are:

Finally, at each iteration, the tap controller is updated by the following equation

(5.3)

5.2.2 Phase-shifting (PS)


A transformer admittance matrix for phase-shift use has been described in section 3.2.3.
Its incorporation in Newton's load flow solution requires an enlarged Jacobian, because
the active power flowing from node k to node m, Pkm, is not used as a control variable
in the conventional formulation of the Newton-Raphson solution. Therefore, either d,,
or &, must be added to the state variables.
5.2 INCORPORATION OF FACTS DEVICES 131

The modified matrix equation is [ 11:

and

is a diagonal matrix of order equal to the number of phase-shifters in the network.


Apt, = APf;’ - pf;l is the PS real power flow mismatch vector.

A4 = - c$i is the vector of incremental changes in phase angle.


Unless the initial values of the complex tap positions are known, the phase-shifter
can be initialized with zero degree angles.

5.2.3 Thyristor controlled series capacitance (TCSC)


The transfer admittance matrix equation of the TCSC connected between nodes k and
m, shown in Figure 5.1, is

(5.6)

(5.7)

and the expression for X I in terms of the firing angle has been derived in section 3.2.1.

v.. I II I v-

Figure 5.1 TCSC module


132 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

The power equations at node k are:


Pk = - v k V m B ( I ) sin(8k - Om), (5.9)
Qk = -viB(~)
-k V k V m B ( I ) c o s ( 8 k - o m ) . (5.10)

The power mismatches required to incorporate a TCSC in the load flow solution are
contained in the following matrix equation (written for a specified power flow from k

- Apk

APm

A Qk (5.11)
AQm

-Apkm

(5.12)

(5.13)

(5.14)

The presence of series compensators can cause ill-conditioning of the Jacobian matrix
if zero voltage angle initialization is adopted. To avoid ill-conditioning, it is recom-
mended to treat the TCSC as a fixed reactance until a specified voltage angle difference
appears across the reactance. From then on, the control equations of the series compen-
sator are included in the iterative process.

5.2.4 Unified power flow controller (UPFC) [3,4]


A UPFC, connected between nodes m and k, can be designed to control the active and
reactive power flows between rn and k, as well as the voltage at node k. In its simplest
form, a loss free UPFC can be specified in the load flow solution as a P m k r Qmk (at
node m) and a P m k . l v k l generator (at node k). No special modification is thus needed
in the standard load flow solution with this model.
Upon convergence, a sequential assessment of the UPFC parameters needs to be
made to ensure that the control limits are not exceeded. This method, however, is
not applicable to cases when the UPFC is only used to control one or two of the
variables. A more flexible and reliable method is next derived with reference to the
5.2 INCORPORATION OF FACTS DEVICES 133

converter I1 12 converter
++

t t
V ~ R0vR
t t
vcR ecR

Figure 5.2 Unified power flow controller [4] (0 1998 IEE)

The power conditions at the terminals of the series converter are:


134 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

where
Ykk = Gkk + jBkk = zii + z&!,
Y m m = Gmm + jBmm = ZFi,
Ykm= Ykm = Gkm+ jBkm = -z,-d,
Y v =~G,R + jBvR = -Z;R.I
Ideally, the UPFC neither absorbs active power from nor injects it into the a.c. system
and the d.c. link voltage, Vdc, remains constant. However, to keep v d c constant, the
shunt converter must supply an amount of active power ( P V ~equal ) to the d.c. power
( v d c l 2 ) , which is equal to the power supplied by the series converter ( P c ~ )i.e.
,

PvR + PcR = 0. (5.23)


As stated in the Introduction, the state variables representing the UPFC must be
added to the network nodal voltage magnitudes and angles to permit a unified solution
by the Newton-Raphson method. In this way, the UPFC state variables are adjusted
automatically to satisfy specified power flows and voltage magnitudes.
When the UPFC controls the voltage magnitude at node k (to which the shunt
converter is connected), as well as the active power transfer from m to k and the
reactive power injection at node m (which is specified as a P, Q bus), the following
matrix equations applies:

(5.24)
where APbb represents the mismatch introduced by the presence of Equation (5.23).
If voltage control at node k is not required, the third column of the Jacobian in
Equation (5.24) is replaced by partial derivatives of the nodal and UPFC mismatch
powers with respect to the nodal voltage magnitude vk. Correspondingly, in the vari-
ables vector the shunt source voltage magnitude increment, A V V ~ / V , is, ~replaced
, by
the nodal voltage magnitude increment at node k, AVk/Vk. In this case, V,R is kept
constant at a value within the specified limits.
5.3 INCORPORATION OF HVDC TRANSMISSION 135

If a limit violation takes place in one of the voltage magnitudes of the UPFC sources,
the voltage magnitude is fixed at that limit and the regulated variable is freed.
In line with the basic load flow solution, in the absence of controlled buses or
branches, 1 p.u. voltage magnitude for all P Q buses and 0 voltage angle for all buses,
provide a suitable starting condition. However, if controllable devices are included in
the analysis, good initial estimates can be obtained by assuming a loss free UPFC and
zero voltage angles in Equations (5.15) to (5.18).
From Equations (5.17) and (5.18) for specified P m , Qm:
(5.25)

(5.26)

where

C1 = Qmref if vO, = v:.


Also, in a lossfree UPFC, the active power absorbed by the shunt converter ( P v ~ )
must be equal to the power ( P c ~delivered
) by the series converter, i.e.
PvR = -pcR (5.27)
combining (5.27) with (5.19) and (5.21), after some reduction leads to:

(5.28)

When the shunt converter operates as a voltage controller, the voltage magnitude of
the shunt source is initialized at the required voltage level and then updated at each
iteration. If the shunt converter is not used as a voltage controller, the voltage magnitude
of the shunt source is kept at a fixed value within its limits during the iterative process.

5.3 Incorporation of HVDC Transmission


The operating state of a combined a.c.-HVDC power system is defined by the vector
[V,8, KIT,
where
-
V is a vector of the voltage magnitudes at all a.c. system busbars;
3 is a vector of the angles at all a.c. system busbars (except the reference bus
which is assigned 8 = 0);
i is a vector of d.c. variables.

e
Chapter 4 has described the use of 7 and as a.c. system variables and the selection
of d.c. variables X is discussed in section 5.3.1.
The development of a Newton-Raphson based algorithm requires the formulation
of n independent equations in terms of the n variables.
136 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

The equations which relate to the a.c. system variables are derived from the specified
a.c. system operating conditions. The only modification required to the usual real and
reactive power mismatches occurs for those equations which relate to the converter
terminal busbars. These equations become:
Pf& - Pterm(ac) - Pterm(dc> = 0 (5.29)
QZm - Qterm(aC) - Qterm(dc) = 0 (5.30)
where
Pterm(ac)is the injected power at the terminal busbar as a function of the a.c.
system variables;
Pterm(dc) is the injected power at the terminal busbar as a function of the d.c.
system variables;
P;:,,, is the usual a.c. system load at the busbar;
and similarly for Qterm(dC) and Qterm(aC).
The injected powers Q,rm(dc) and P,rm(dC) are functions of the converter a.c.
terminal busbar voltage and of the d.c. system variables, i.e.
Pterm(dC) = f ( v t e r m 9 3 (5.31)
Qterm(dc) = f ( v t e r m , 9 (5.32)
The equations derived from the specified a.c. system conditions may, therefore, be
summarized as:

(5.33)

ABterm ( 7 9 8, z)
AG(v,8) =0, (5.35)
ADterm (7,8, @

- -
Wte,,X) -
5.3 INCORPORATION OF HVDC TRANSMISSION 137

53.1 Converter model


The selection of variables 7 and formulation of the equations require several basic
assumptions which are generally accepted in the analysis of steady state d.c. converter
operation. These are:

(i) The three a.c. voltages at the terminal busbar are balanced and sinusoidal.
(ii) The converter operation is perfectly balanced.
(iii) The direct current and voltage are smooth.
(iv) The converter transformer is lossless and the magnetizing admittance is ignored.

Converter variables Under balanced conditions, similar converter bridges, attached


to the same ax. terminal busbar, will operate identically regardless of the transformer
connection. They may, therefore, be replaced by an equivalent single bridge for the
purpose of single-phase load flow analysis. With reference to Figure 5.3, the set of
variables illustrated, representing fundamental frequency or d.c. quantities, permits a
full description of the converter system operation.
An equivalent circuit for the converter is shown in Figure 5.4, which includes the
modification explained in section 5.3 as regards the position of angle reference.
The variables, defined with reference to Figure 5.4, are:
V,,,,L4 converter terminal busbar nodal voltage (phase angle referred to
converter reference);
EL II/ fundamental frequency component of the voltage waveform at the
converter transformer secondary;

Figure 5.3 Basic d.c. converter (angles refer to a.c. system reference)

Figure 5.4 Single-phase equivalent circuit for basic converter (angles refer to d.c. reference)
138 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

I,, I , fundamental frequency component of the current waveshape on the


primary and secondary of the converter transformer, respectively;
0 firing delay angle;
a transformer off-nominal tap ratio;
Vd average d.c. voltage;
Id converter direct current.
These ten variables, nine associated with the converter, plus the a.c. terminal voltage
magnitude Vterm,form a possible choice of X for the formulation of Equations (5.31),
(5.32) and (5.34).
The minimum number of variables required to define the operation of the system is
the number of independent variables. Any other system variable or parameter (e.g. P d c
and Qdc) may be written in terms of these variables.
Two independent variables are sufficient to model a d.c. converter, operating under
balanced conditions, from a known terminal voltage source. However, the control
requirements of HVDC converters are such that a range of variables, or functions
of them (e.g. constant power), are the specified conditions. If the minimum number
of variables is used, then the control specifications must be translated into equations
in terms of these two variables. These equations will often contain complex non-
linearities, and present difficulties in their derivation and program implementation. In
addition, the expressions used for Pdc and Q d c in Equations (5.29) and (5.30) may be
rather complex and this will make the programming of a unified solution more difficult.
For these reasons, a non-minimal set of variables is recommend, i.e. all variables
which are influenced by control action are retained in the model. This is in contrast
to a.c. load flows where, due to the restricted nature of control specifications, the
minimum set is normally used.
The following set of variables permits simple relationships for all the normal control
strategies:
[a] = [ V d , I d , a, cos(Y,#IT.
Variable 4 is included to ensure a simple expression for Qdc. While this is important
in the formulation of the unified solution, variable 4 may be omitted with the sequential
solution as it is not involved in the formulation of any control specification; cosa is
used as a variable rather than (Y in order to linearize the equations and thus improve
convergence.

d.c. per unit system To avoid per unit to actual value translations and to enable the
use of comparable convergence tolerances for both a s . and d.c. system mismatches, a
per unit system is also used for the d.c. quantities.
Computational simplicity is achieved by using common power and voltage base
parameters on both sides of the converter, i.e. the a.c. and d.c. sides. Consequently,
in order to preserve consistency of power in per unit, the direct current base, obtained
from ( M V A B ) / V B , has to be 4 times larger than the alternating current base [ 5 ] .
This has the effect of changing the coefficients involved in the a.c.-d.c. current
relationships. For a perfectly smooth direct current and neglecting the commutation
overlap, the r.m.s. fundamental components of the phase current is related to Id by the
5.3 INCORPORATION OF HVDC TRANSMISSION 139

approximation,
(5.36)

Translating Equation (5.36)to per unit yields


4
ls(p.u,) = -&Id(p.u.),
n
and if commutation overlap is taken into account, as described in Chapter 3, this
equation becomes:
(5.37)
where k is very close to unity. In load flow studies, Equation (5.37) can be made
sufficiently accurate in most cases by letting:
k = 0.995.
Derivation of equations The following relationships are derived for the variables
defined in Figure 5.4. The equations are in per unit.

(i) The fundamental current magnitude on the converter side is related to the direct
current by the equation
(5.38)

(ii) The fundamental current magnitudes on both sides of the lossless transformer
are related by the off-nominal tap, i.e.
I , = a I,. (5.39)
(iii) The d.c. voltage may be expressed in terms of the a.c. source commutating
voltage referred to the transformer secondary, i.e.

(5.40)
The converter a.c. source commutating voltage is the busbar voltage on the
system side of the converter transformer, Vterm.
(iv) The d.c. current and voltage are related by the d.c. system configuration,
f ( v d , Id) = 0, (5.41)
e.g. for a simple rectifier supplying a passive load,
v d - I d & = 0.
(v) The assumptions listed at the beginning of this section prevent any real power
of harmonic frequencies at the primary and secondary busbars. Therefore, the
real power equation relates the d.c. power to the transformer secondary power
in terms of fundamental components only, i.e.
VdId = Els COS II/. (5.42)
140 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

(vi) As the transformer is lossless, the primary real power may also be equated to
the d.c. power, i.e.
VdId = VtermIp cos 4. (5.43)
(vii) The fundamental component of current flow across the converter transformer can
be expressed as
I , = B, sin II/ - BtaVtem sin& (5.44)
where j B , is the transformer leakage susceptance.
So far, a total of seven equations have been derived and no other independent
equation may be written relating the total set of nine converter variables.
+
Variables I , , I,, E and can be eliminated as they play no part in defining control
specifications.
Thus, Equations (5.38), (5.39), (5.42) and (5.43)can be combined into
Vd - klaVterm COS 4 = 0, (5.45)
where kl = k ( 3 a / ~ r ) .
The final two independent equations required are derived from the specified control
mode.
The d.c. model may thus be summarized as follows:
B(z,Vtermlk = 0, (5.46)
where
R(1) = Vd - klavte, cos 4,

R(2) = Vd - klavtem C O S a + -J3rI d x , ,


N 3 )=f W d ?I d ) ,

R(4) = control equation,


R ( 5 ) = control equation,
and
= [vd,I d r a , cosa, $IT.
V,,, can either be a specified quantity or an ax. system variable. The equations for
Pdc and Qdc may now be written as:

Qterm(dc1 = V t e d p sin 4
= Vte,kla I d sin4 (5.47)
and

(5.48)
or
(5.49)
5.3 INCORPORATION OF HVDC TRANSMISSION 141

Incorporation of control equations Each additional converter in the d.c. system


contributes two independent variables to the system and thus two further constraint
equations must be derived from the control strategy of the system to define the operating
state. For example, a classical two-terminal d.c. link has two converters and, therefore,
requires four control equations. The four equations must be written in terms of the ten
d.c. variables (five for each converter).
Any function of the ten d.c. system variables is a valid (mathematically) control
equation so long as each equation is independent of all other equations. I n practice,
there are restrictions limiting the number of alternatives. Some control strategies refer
to the characteristics of power transmission (e.g. constant power or constant current),
other introduce constraints such as minimum delay or extinction angles.
Examples of valid control specifications are:
(i) Specified converter transformer tap,
a - -
- 0.
(ii) Specified d.c. voltage
Vd - vsp= 0.
(iii) Specified d.c. current
Id - Isp = 0.
(iv) Specified minimum firing angle
cos - cos a m i n = 0.
(v) Specified d.c. power transmission
Vdld - P:: = 0.
These control equations are simple and are easily incorporated into the solution
algorithm. In addition to the usual control modes, non-standard modes, such as spec-
ified a.c. terminal voltage, may also be included as converter control equations (see
section 5.3.3).
During the iterative solution procedure, the uncontrolled converter variables may go
outside pre-specified limits. When this occurs, the offending variable is usually held
to its limit value and an appropriate control variable is freed.
Inverter operation All the equations presented so far are equally applicable to
inverter operation. However, during inversion, it is the extinction advance angle ( y )
which is the subject of control action and not the firing angle a. For convenience,
therefore, equation R(2) of (5.46) may be written as

(5.50)
This equation is valid for rectification or inversion. Under inversion, v d , as calculated
by Equation (5.50), will be negative.
To specify operation with constant extinction angle, the following equation is used:
cos(7r - y ) - cos(7r - y ” ) = 0
where ysP is usually y minimum for minimum reactive power consumption of the
inverter.
142 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

5.3.2 Solution techniques


The converter model developed in section 5.3.1 is to be incorporated with a fast decou-
pled a s . load flow algorithm with the minimum possible modification of the latter in
order to retain its computational advantages.
The solution is first discussed with reference to a single converter connected to an
a.c. busbar. The extension to multiple or multiterminal d.c. systems is relatively trivial
and is discussed in section 5.3.4.
Unified solution [6,7] The unified method gives recognition to the interdependence
of a.c. and d.c. system equations and simultaneously solves the complete system.
Referring to Equation (5.33,the standard Newton-Raphson algorithm involves repeat
solutions of the matrix equation:

(5.51)

where J is the matrix of first order partial derivatives.

APterm = PSL - Pterm(ac) - Ptem(dc), (5.52)


AQterm = QZm - Qterm(aC) - Qterm(dc), (5.53)

and
Pterm(dc) = f(Vterm, (5.54)
Qterm (dc 1 = f ( Vterm 9 (5.55)
Applying the a.c. fast decoupled assumptions to all Jacobian elements related to the
a.c. system equations, yields:

(5.56)

where all matrix elements are zero unless otherwise indicated. The matrices [B’] and
[B”] are the usual single-phase fast decoupled Jacobians and are constant in value. The
other matrices indicated vary at each iteration in the solution process.
A modification is required for the element indicated as Byi in Equation (5.56). This
element is a function of the system variables and, therefore, varies at each iteration.
5.3 INCORPORATION OF HVDC TRANSMISSION 143

The use of an independent angle reference for the d.c. equations results in
aPterm(dc)/Wem = 0,
i.e. the diagonal Jacobian element for the real power mismatch at the converter terminal
busbar depends on the a.c. equations only and is, therefore, the usual fast decoupled
B’ element.
In addition,
aR/aOter, = 0,
which will help the subsequent decoupling of the equation.
In order to maintain the block successive iteration sequence of the usual fast decou-
pled a.c. load flow, it is necessary to decouple Equation (5.56). Therefore, the Jacobian
submatrices must be examined in more detail. The Jacobian submatrices are:
1
D D = -a APterml avterm
Vterm

( apterm (dc)/aVterm1-
1 1
- -(aptem(ac) /avteml+ -
Vterm Vterm
Following decoupled load flow practice
1
DD = 0 + -( a p t e m (dc)/aVterm 1 9
Vtem
and since
144 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

In the above formulation, the d.c. variables X are coupled to both the real and reactive
power ax. mismatches. However, Equation (5.56) may be separated to enable a block
successive iteration scheme to be used.
The d.c. mismatches and variables can be appended to the two fast decoupled a.c.
equations, in which case the following two equations result:

, (5.57)

(5.58)
AZ

The iteration scheme illustrated in Figure 5.5 is referred to as -PDC, QDC- and the
significance of the mnemonic should be obvious.
The algorithm may be further simplified by recognizing the following physical char-
acteristics of the a.c. and d.c. systems:

0 The coupling between d.c. variables and the ax. terminal voltage is strong.
0 There is no coupling between d.c. mismatches and a.c. system angles.
0 Under all practical control strategies, the d.c. power is well constrained and this
implies that the changes in d.c. variables j z do not greatly affect the real power
mismatches at the terminals. This coupling, embodied in matrix AA' of Equa-
tion (5.57) can, therefore, be justifiably removed.

These features justify the removal of the d.c. equations from Equation (5.57) to
yield a -P, QDC-block successive iteration scheme, represented by the following two
equations:
[ABIV] = [B'J[AS], (5.59)

Programming considerations for the unified algorithms In order to retain the


efficiency of the fast decoupled load flow, the B' and B" matrices must be factorized
only once before the iterative process begins.
The Jacobian elements related to the d.c. variables are non-constant and must be
re-evaluated at each iteration. It is, therefore, necessary to separate the constant and
non-constant parts of the equations for the solution routine.
5.3 INCORPORATION OF HVDC TRANSMISSION 145

(dc) and d.c. residuals ( R )


Calculate Pterm
1
Yes
- 1

b
I Form d.c. iacobian matrix 1 lP=11

Forward reduction of vector AFI V

Solve reduced equation for Pt,,, and A P

Back substitute for A 3

I
b m

I
I
UDdate Hand 8 1I
t
1 P=O ] I

+=
I Calculate A d(a.c. system only)
1
1
Calculate Qterm (d.c.) and d.c. residuals (A)

Forward reduction of vector AGI V


I

Solve reduced equation for V,,,, and AX


+ 1 No

I I
a I

t i
Figure 5.5 Flow chart for unified a.c.-d.c. load flow
146 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

Initially, the a.c. fast decoupled equations are formed with the d.c. link ignored
(except for the minor addition of the filter reactance at the appropriate a.c. busbar).
The reactive power mismatch equation for the ax. system is:

(5.61)

where
AQ:,,, = Q:, - Qterm(aC)is the mismatch calculated in the absence of the
d.c. converter,
and
B” is the usual constant a.c. fast decoupled Jacobian.
After triangulation down to, but excluding, the busbars to which d.c. converters are

=m[
attached, Equation (5.61) becomes

[ ’ (5*62)
--
where (AQ/V)” and (AQterm/Vterm)” signify that the left-hand side vector has been
processed and matrix B”’ is the new matrix B” after triangulation.
This triangulation (performed before the iterative process) may be achieved simply
by inhibiting the terminal busbars being used as pivots during the optimal ordering
process.
The processing of AQ indicated in the equation is actually performed by the standard
forward reduction process used at each iteration.
The d.c. converter equations may then be combined with Equation (5.62) as follows:

(5.63)
where

The unprocessed section, i.e.

[ (W”+ K
AQ;iy)] -
BB”
[A:;]
(5.64)

may then be solved by any method suitable for non-symmetric matrices.


The values of AX and AVtem are obtained from this equation and AVtermis then
used in a back substitution process for the remaining A T to be completed, i.e. Equation
(5.62) is solved for AV.
5.3 INCORPORATION OF HVDC TRANSMISSION 147

The most efficient technique for solving Equation (5.64) depends on the number
of converters. For six converters or more, the use of sparsity storage and solution
techniques is justified, otherwise all elements should be stored. The method suggested
here is a modified form of Gaussian elimination where all elements are stored but only
non-zero elements processed.

Sequential method [5,8] The sequential method results from a further simplification
of the unified method, i.e. the ax. system equations are solved with the d.c. system
modelled simply as a real and reactive power injection at the appropriate terminal
busbar. For a d.c. solution, the a.c. system is modelled simply as a constant voltage at
the converter ax. terminal busbar.
The following three equations are solved iteratively to convergence:

[ A F / v ] = [B’][A8], (5.65)
[AQi/v] = [B”][Av], (5.66)
[R] = [A][AT]. (5.67)

This iteration sequence, referred to as P , Q,DC,is illustrated in the flow chart of


Figure 5.6 and may be summarized as follows:

(i) CalculateAF/v, solve Equation (5.65) and update 8.


(ii) Calculate AD/v, solve Equation (5.66) and update 7.
(iii) Calculate d.c. residuals, R, solve Equation (5.67) and update E.
(iv) Return to (i).
With the sequential method, the d.c. equations need not be solved for the entire
iterative process. Once the d.c. residuals have converged, the d.c. system may be
modelled simply as fixed real and reactive power injections at the appropriate converter
terminal busbar. The d.c. residuals must still be checked after each a.c. iteration to
ensure that the d.c. system remains converged.
However, in order to establish a direct comparison between the unified and sequential
algorithms, the d.c. equations continue to be solved until both a s . and d.c. systems
have converged. This ensures that the sequential technique is an exact parallel of the
corresponding unified algorithm.
Alternatively, the d.c. equations can be solved after each real power as well as after
each reactive power iteration and the resulting sequence is referred to as P , DC,Q,
DC. As in the previous method, the d.c. equations are solved until all mismatches are
within tolerance.

5.3.3 Control of converter ax. terminal voltage


A converter terminal voltage may be specified in two ways:

(a) By local reactive power injection at the terminal. In this case, no reactive power
mismatch equation is necessary for that busbar and the relevant variable (i.e.
148 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

(all a.c. busbars)


-.
i Calculate A g(total svstem) and d.c. residuals Ti i
Yes

. i
Solve equation 5.65 and update (8)
t

1
* Converged
) and d.c. residuals R

Solve equation 5.i6 and update (8)

I
I Fl
1

Calculate d.c. mismatches

t
I Form d.c. jacobian matrix I
I
t
I Solve eauation 5.67 and udate K 1

LTJ P=Q=O

Figure 5.6 Flow chart for sequential a.c.-d.c. load flow

AV,,,) is effectively removed from the problem formulation. This is the situation
where the converter terminal busbar is a P-V busbar.
(b) The terminal voltage may be specified as a d.c. system constraint. That is, the d.c.
converter must absorb the correct amount of reactive power so that the terminal
voltage is maintained constant.
5.3 INCORPORATION OF HVDC TRANSMISSION 149

With the unified method, the equation


SP
Vterm - Vterm -
-0 (5.68)
is written as one of the two control equations. This would lead to a zero row in Equation
(5.67) and, therefore, during the solution of Equation (5.67) some other variable (e.g.
tap ratio) must be specified instead. Although d.c. convergence is marginally slower
for the PDC, QDC iteration, the d.c. system is overconverged in this iteration scheme
and the overall convergence rate is practically unaffected.
With the sequential method, Equation (5.68) cannot be written. The terminal busbar
is specified as a P-V busbar and the control equation
Q
m
r:; (dc) - Qterm (dc) = 0
is used, where Q&, (dc) is taken as the reactive power required to maintain the
voltage constant. The specified reactive power thus varies at each iteration and this
discontinuity slows the overall convergence.

5.3.4 Extension to multiple and/or muititerminal d.c. systems


The basic algorithm has been developed in previous sections for a single d.c. converter.
Each additional converter adds a further five d.c. variables and a corresponding set of
five equations. The number of a.c. system Jacobian elements which become modified
in the unified solutions is equal to the number of converters.
As an example, consider the system shown in Figure 5.7. The system represents the
North and South Islands of the New Zealand system, 220 kV a.c. system. Converters 1
and 2 form the original 600 MW, 500 kV d.c. link between the two islands. Converter 3
represents a 420 MW aluminium smelter. A further three-terminal d.c. interconnection
has been added (Converters 4,5 and 6 ) to illustrate the flexibility of the algorithm.

'd 2 'd 6
+ t

Figure 5.7 Multiterminal d.c. system


150 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

Normally, Converter 4 will operate in the rectifier mode with Converters 5 and 6 in
the inversion mode.
The reactive power d.c. Jacobian for the unified method has the structure, shown in
Figure 5.8, where BSOD is the part of B" which becomes modified. Only the diagonal
elements become modified by the presence of the converters.
Off diagonal elements will be present in BboD if there is any a.c. connection between
converter terminal busbars. All off diagonal elements of BB" and AA" are zero.
In addition, matrix A is block diagonal in 5 x 5 blocks with the exception of the
d.c. interconnection equations.
Equation R ( 3 ) of (5.46) in each set of d.c. equations is derived from the d.c. inter-
connection. For the six-converter system, shown in Figure 5.7, the following equations
are applicable.
vdi + v d 2 - Idl(Rd1 +
R d 2 ) = 0,

vd3 - = 0,
Id1 - Id2 = 0,
vd4 + v d 6 - l d 4 R d 4 - l d 6 R d 6 = 0,
vd5 - vd6 - I d s R d ~+ l d 6 R d 6 = 0 ,
I d 4 - I d s - I d 6 = 0.
This example indicates the ease of extension of the multiple converter case.

Avterrn 6

A
(30 x 30)

Figure 5.8 Matrix equation for a multi-terminal configuration


5.3 INCORPORATION OF HVDC TRANSMISSION 151

5.3.5 d.c. convergence tolerance

The d.c. p.u. system is based upon the same power base as the a.c. system and on
the nominal open circuit a.c. voltage at the converter transformer secondary. The p.u.
tolerances for d.c. powers, voltages and currents are, therefore, comparable with those
adopted in the a.c. system.
In general, the control equations are of the form

where X may be the tap or cosine of the firing angle, i.e. they are linear and are
thus solved in one d.c. iteration. The question of an appropriate tolerance for these
mismatches is, therefore, irrelevant.
An acceptable tolerance for the d.c. residuals which is compatible with the a.c.
system tolerance is typically 0.001 p.u. on a 100 MVA base, i.e. the same as that
normally adopted for the ax. system.

5.3.6 Test system and results


The A.E.P standard 14-bus test system is used to show the convergence properties
of the ax.-d.c. algorithms, with the a.c. transmission line between busbars 5 and 4
replaced by a HVDC link. As these two buses are not voltage controlled, the interaction
between the a.c. and d.c. systems will, therefore, be considerable.
Various control strategies have been applied to the link and the convergence results
for the various algorithms are given in Table 5.1. The number of iterations ( i , j ) should
be interpreted as follows:
0 i is the number of reactive power voltage updates required.
0 j is the number of real power angle updates.

Although the number of d.c. iterations varies for the different sequences, this is of
secondary importance and may, if required, be assessed in each case from the number
of a.c. iterations. In this respect, a unified QDC iteration is equivalent to a Q iteration
and a DC iteration executed separately.
The d.c. link data and specified controls for Case 1 are given in Table 5.2 and the
corresponding d.c. link operation is illustrated in Figure 5.9. The specified conditions
for all cases are derived from the results of Case 1. Under those conditions, the a.c.
system in isolation (with each converter terminal modelled as an equivalent a.c. load)
requires (4,3) iterations. The d.c. system in isolation (operating from fixed terminal
voltages) requires two iterations under all control strategies.

Unified cases The results in Table 5.1 show that the unified methods provide fast
and reliable convergence in all cases.
For the unified methods 1 and 2, the number of iterations did not exceed the number
required for the a.c. system alone.
152 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

Table 5.1 Convergence results


Case specification Number of iterations to convergence (0.1 M W M V A R )
Specified d.c. Unified methods Sequential methods
link constraints (5 variables)
(5 variables) (4 variables)
m-rectifier and
n-inverter end PDC.QDC2 P, QDC P.Q,DC2P.DC,Q,DC 1P,Q,DC2P,DC.Q.DC

1 UrnPdrnYnVdn 4,3 4,3 4,3 4.3 44 43


2 UmPdnian Vdn 4,3 43 494 5s 4,4 Failed
3 UmPdmanVdn 4,3 4.3 4,4 5,s 44 Failed
4 UmPdniYn vdri 4,3 4,3 4,4 44 4,4 44
5 UmPdrnYnan 4,3 43 494 44 4,4 494
6 amPdmUmYn 43 4,3 43 4,3 44 43
7 d Vdn
~ m I Yn 4,3 4,3 493 43 44 43
8 amVdmYnPdn 4,3 4,3 4-4 4,4 494 4-4
Case 1 with initial
condition errors
9 50 per cent error 4.3 4,3 44 4,3 4,4 43
10 80 per cent error 5,4' 6,5" 1,6* 5,4" 44 4-3
Where 'indicates a false solution.

Table 5.2 Characteristics of d.c. link


~ ~ ~~

Converter 1 Converter 2
a.c . busbar Bus 5 Bus 4
d.c. voltage base 1 0 0 kV 1 0 0 kV
Transformer reactance 0.126 0.0728
Commutation reactance 0.126 0.0728
Filter admittance Bi 0.478 0.629
d.c. link resistance 0.334a
Control parameters for Case 1
d.c. link power 5 8 . 6 MW -
Rectifier firing angle (deg) 7 -
Inverter extinction angle (deg) 10
Inverter d.c. voltage -1 2 8 . 8 7 kV
"Filters are connected to ax. terminal busbar.
Note: All reactances are in p,u. on a 100 MVA base.

Bus 5 Bus 4
V = 1.032
a=2.8% +Id =454.2 V = 1.061

I R = 0.334

0
-t P = 58.60
0 = 18.79
1
dvd

a = 7.0
= 129.022 Vd=-128.87$.

u = 17.32
y = 10.0
u = 10.33
P = -58.31
Q = 16.78
All angles are in degrees. D.C. voltages and current are in kV and Amp respectively.
D.C. resistance is in ohms. A.C. powers (P,Q)are in MW and MVARs.
-
Figure 5.9 d.c. link operation for Case 1
5.3 INCORPORATION OF HVDC TRANSMISSION 153

Sequential cases The sequential method ( P , Q, DC) produces fast and reliable
convergence although the reactive power convergence is slower than for the a.c. system
alone.
With the removal of the variable 4, &,(dc) converges faster but the convergence
pattern is more oscillatory and an overall deterioration of a.c. voltage convergence
results.
With the second sequential method (P, DC,Q, DC),convergence is good in all cases
except 2 and 3, i.e. the cases where the transformer tap and d.c. voltage are specified
at the inverter end. However, this set of specifications is not likely to occur in practice.

Initial conditions for d.c. system Initial values for the d.c. variables E are assigned
from estimates for the d.c. power and d.c. voltage and assuming a power factor of 0.9
at the converter terminal busbar. The terminal busbar voltage is set at 1.0 p.u. unless
it is a voltage controlled busbar.
This procedure gives adequate initial conditions in all practical cases as good esti-
mates of Pter,(dc) and V d are normally obtainable.
With starting values for d.c. real and reactive powers within 35096, which are
available in all practical situations, all algorithms converged rapidly and reliably (see
Case 9).
Effect of a.c. system strength In order to investigate the performance of the algo-
rithms with a weak a.c. system, the test system described earlier is modified by the
addition of two a.c. lines, as shown in Figure 5.10.
The reactive power compensation of the filters was adjusted to give similar d.c.
operating conditions as previously.
The number of iterations to convergence for the most promising algorithms are
shown in Table 5.3 for the control specifications corresponding to Cases 1 to 4 in the
previous results.
The different nature of the sequential and unified algorithms is clearly demonstrated.
The effect of the type of converter control is also shown. For Case 11, both the d.c.
real power and the d.c. reactive powers are well constrained by the converter control
strategy. Convergence is rapid and reliable for all methods.
In all other cases, where the control angle at one or both converters is free, an
oscillatory relationship between converter a.c. terminal voltage and the reactive power
of the converter is possible. This leads to poor performance of the sequential algorithms.
To illustrate the nature of the iteration, the convergence pattern of the converter
reactive power demand and the ax. system terminal voltage of the rectifier is plotted
in Figure 5.1 1.

Bus 5 Bus 4
ix
I
rl
0 Filters

s. Filters

Figure 5.10 d.c. link operating from weak a.c. system


154 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

Table 5.3 Numbers of iterations for a weak ax. system


Case specification xi = 0.3 X; = 0.4
rn -rectifier
n -inverter Unified Sequential Unified Sequential
P , QDC P , Q , DC P , QDC P , Q , DC
(i) (ii) (i) (ii)

(i) Using the five variable formulation; (ii) using the four variable formulation.

Q term
(MVAR) 1 $el?
-
28 -

26,-

24 - 0.96 1\
22 -
20 -
18 - o-.--
nnl Y I I I I I L
0 1 2 3 4 5 6 7

Q term A
(MVAR)
28 - 1.o
26 - 0.98

24 - 0.96
22 - 0.94
20 -

18 - 0.90
0.92 0 1 1 2 3 4

Figure 5.11 Convergence pattern for a.c.-d.c. load flow with weak ax. system. (a) Sequential
methods ( P , Q, DC five variable); (b) Unified method ( P , QDC)

A measure of the strength of a system in a load flow sense is the short circuit to
converter power ratio (SCR)calculated with all machine reactances set to zero. This
short circuit ratio is invariably much higher than the usual value.
In practice, converter operation has been considered down to an SCR of 3. A survey
of existing schemes shows that, almost invariably, with systems of very low SCR,some
5.3 INCORPORATION OF HVDC TRANSMISSION 155

form of voltage control, often synchronous condensers, is an integral part of the converter
installation. These schemes are, therefore, often very strong in a load flow sense.
It may, therefore, be concluded that the sequential integration should converge in all
practical situations although the convergence may become slow if the system is weak
in a load flow sense.
Discussion of convergence properties The overall convergence rate of the
ax.-d.c. algorithms depends on the successful interaction of the two distinct parts.
The a.c. system equations are solved using the well-behaved constant tangent fast
decoupled algorithm, whereas the d.c. system equations are solved using the more
powerful, but somewhat more erratic, full Newton-Raphson approach.
The powerful convergence of the Newton-Raphson process for the d.c. equations
can cause overall convergence difficulties. If the first d.c. iteration occurs before the
reactive power voltage update, then the d.c. variables are converged to be compat-
ible with the incorrect terminal voltage. This introduces an unnecessary discontinuity
which may lead to convergence difficulties in the sequential method. In the unified
approach, the powerful convergence of the d.c. equations is dampened by the reflec-
tion of the a.c. mismatches onto the changes in d.c. variables. This gives faster and
better behaved convergence. The solution time of the d.c. equations is normally small
compared with the solution time of the a.c. equations. The relative efficiencies of the
alternative algorithms may, therefore, be assessed by comparing the total numbers of
voltage and angle updates.
In general, those schemes which acknowledge the fact that the d.c. variables are
strongly related to the terminal voltage give the fastest and most reliable performance.
The unified methods are the more reliable and the P,QDC solution is the most efficient.
Of the sequential methods, the (P,Q,D C ) solution is only marginally inferior to the
unified method.
When the busbar to which the converters are attached is voltage controlled (as is
often the case) the two approaches become virtually identical as the interaction between
ax. and d.c. systems is much smaller.
The only computational difference between a unified and a comparable sequential
iteration is that the d.c. Jacobian equations for the unified method (Equation (5.64))
is slightly larger. The difference is one additional row and column for each converter
present. In terms of computational cost per iteration, the corresponding unified and
sequential algorithms are virtually identical.

(i) In cases where the a.c. system is strong, both the unified and sequential algorithms
may be programmed to give fast and reliable convergence.
(ii) If the a.c. system is weak, the sequential algorithm is susceptible to convergence
problems. Thus, in general, the unified method id recommended due to its greater
reliability.

5.3.7 Numerical example


The complete New Zealand primary transmission system was used as a basis for a
planning study which included an extra multiterminal HVDC scheme, i.e. involving
six converter stations, as illustrated in Figure 5.7.
156 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

Representative input and output information obtained from the computer is given in
the following pages.

DEPARTYENT OF ELECTRICAL k ELECTRONIC ENCINEER.ING. UNIVERSITY OF CANTERBURY, NEV ZEALAND

SYSTEM NO. 3 23 YAR 90

XAIIYUN NUYBU OF ITEUTIONS 10


POVEL TOLELANCE 0.00100
PRINT WT INDICAML 000000000 NUIBEL OF BUSES 114
SYSTEI YVA BASE 100.00 NUIBEL OF LINES 206
D.C. L W T INDICATOR 6 NUIBER OF TUNSFUPlERS 19
NUIBER OF A.C. SYSTENS 2
SLACK BUSBALS 80 218

B U S D A T A

LOAD CENEUTION IINIYU IAXIXUY SHUNT


BUS NAYE TYPE VOLTS IV IVB nv EVIL IVAL IVAL SUSCEPTANCE

104 AVIEIOIE-220 1 1.0520 0.00 0.00 220.00 -34.40 -500.00 500.00 0.OOO
108 BDINOILE-220 1 1.0520 97.20 0.00 540.00 46.60 -500.00 500.00 0.000
118 BUY-220 0 1.0030 329.60 95.80 0.00 0.00 0.00 0.00 O.Oo0
127 CWIl-220 0 1.0520 0.00 0.00 0.00 0.00 0.00 0.00 0.000
128 CUJMZ-220 0 1.0520 0.00 0.00 0.00 0.00 0.00 0.00 0.000

129 CLUTAA-220 1 1.0300 0.00 0.00 600.00 0.00 0.00 0.00 0.000
138 CMLDINE220 0 1.0210 0.00 0.00 0.00 0.00 0.00 0.00 0.000
143 HyBS---220 0 1.0270 95.30 80.40 0.00 0.00 0.00 0.00 0.OOO

LINE D A T A

BUS NAME BUS NAIE LESISTANQ LEACTANCE SUSCEPTANCE

104 AVIEIOIE-220 108 BENYORE-220 0.00330 0.01630 0.02298


104 AVIEIOIE-220 108 BENIOLE-220 0.00330 0.01530 0.02298
104 AVIEMOIE-220 268 VAITAKI-220 0.00150 0.00730 0.01052
108 BENOS220 255 TVIZEL-220 0.00370 0.02610 0.06954
118 BUY-220 167 ISLINGTON220 0.00210 0.01651 0.05285

118 BUY-220 181 LANILM2-220 0.00110 0.00861 0.02751


127 CUJMl-220 218 1OXBuICH-220 0.00770 0.04450 0.07251
127 CUJI1-220 255 TVIZEd-220 0.00820 0.09260 0.16746
128 cu)I2-220 218 ROXBUILCH-220 0.00770 0.04450 0.07251
128 CUlU2-220 255 TVIZEG-220 0.00820 0.09260 0.16746
5.3 INCORPORATION OF HVDC TRANSMISSION 157

T l A N S F O I I E I D A T A

BUS NAME BUS NAME LESISTANCE IEACTMCE TAP CODE

6 BWTHOlPEl10 7 BWTHWE220 0.00400 0.09560 1.OW 0


6 BWrmOlF'ElIO 7 BIRITHOlPE220 0.00400 0.09560 1.OW 0
6 BIRITHOlPEl10 7 BWTHOlPM20 0.00170 0.04SW 1.ooO 0
10 EDCECOIBEIlO II EDGUXnBEZZO 0.00400 0.09560 1.OOo 0
10 EDCECOMBEllO II EDCECOMBE220 0.00400 0.09560 1.ooO 0
21 HAWAIIIS-I10 22 HAWANIS-220 0.00170 0.05140 1.040 0
21 HAYWAIDS-110 22 HAWANIS-220 0.00170 0.05140 1.000 0
2 1 HAWALDS-I10 22 HAWANIS-220 0.00410 0.10120 1.000 0
?I HIWhM5-110 22 HAWAMS-220 0.004 I0 0.10120 1.OOo 0
23 HFNDEISON110 24 HENDEISONZ20 0.00090 0.01840 1.053 0
39 IAlSDEN-IIO 40 IAlSDW-220 o.oo0oo 0.05500 1.OOO 0
39 MAUDEN-IIO 40 IAISDM-220 0.00000 0.05500 1.OW 0
48 HEVPLYI7HlIO 49 NEWLTITY220 0.00080 0.02480 1.OW 0
54 OTAHIJW-I10 55 OTAWUlN-220 0.00700 0.04100 1.000 0
54 OTAHIJW-I10 55 OTAHWU-220 0 00700
I 0.04100 l.OW 0
54 MAHIJW-I10 55 OTARllWU-220 0.00160 0.04550 1.000 0
58 PMIOSE-IIO 59 PEIROsc220 0.000w 0.02750 1.OW 0
62 STLATSOLDIIO 63 STUtfUlD220 0.00200 0.05260 1.ooO 0
66 TUUTENGAIIO 67 TAIUTI)IGAZ20 0.00080 0.02530 1.OOO 0

WUhTIUNS
. l ~ l D l .III2-1D3 .III3-ID4 .IDI-IDS .IIIS-IW .LD6-1D7 .III7-ID8 .IIIB-IW .NI9-IDlO .u)lO=O
I r ~ ~ ~ + Y M + V D 5 r V W + Y D 7 * V D ~ W 9 + Y DIIII-ID2

,
I I 0 0 0 0 0 0 0 0 25.5800 o.oo00 0 . m o.oo00 o.oo00 o.oo00 o.oo00 o.oo00 o.Ooo0 o.oo00
# 1 0 0 0 0
8 @ 0 1 0 1 0 0 0 0
0 0 0 0.0000 o.Ooo0 0.0019 o.oo00 o.oo00 o.oo00 o.oo00 o.oo00 o.oo00 o.oo00
o.oo00 o.oo00 o.ooO0 1o.oooo 0.M)Oo 2o.oooo o.oo00 o.oo00 o.oo00 o.oo00
8 @ 0 0 I -1 0 0 0 0 0.0000 0.OOOO 0.0000 O.oO00 3.oo00-20.oooO 0.0000 O.oo00 O.oo00 O.Oo0o

~,~~kID4+IDS+IDb+ID7+ID8~ID9+IDlO:O
1-1 0 0 0 0 0 0 0 0

~)~~IDkiD4+ID5rIW+I7+lD8+ID9+IDlO~O
) ( O 1 - 1 4 0 0 0 0

pC CONVEITEI WlBEl 1 INPUT DATA PC CONVEXTU WMBEl 6 lNPWf

CONVElTEl ATTACHED TO BUS NUIBEI I08 co1(yElTEl ATTACHED To BUS N U I B U 7

NOMINAL DC VOLTAGE l10.00oOo NOIINAL DC VOLTAGE 90 * 00000


I A X I I U I DC VOLTAGE 150.00000 IAIINI DC VOLTAGE 140.00000
MINIIUI DC VOLTAGE 0.00000 MINIllll DC VOLTAGE 0.00000
IAIIIUI DC CUuEm O.oo00 I A I I N I DC CUUENI 0.0000

COMIUTATION UACTAWCE (P.U.) 0.08970 COIIVIATIMI LEICIANCE (P.U.) 0.07000


N N S F W E l LEIcIANlZ (P.U.) 0.08970 hlNslolIll UACTANQ (?.U.) 0,07000
FIIINC MCLE:MINIIW (Dffi) 10.00000 FILING AffiU:MINIIUI (DED) 8.o0000
MAIIIW (OK) 110.00000 IAIIIUI (DED) 150.00000
RUlSMUEl TAP:IINIIUI (P.C.)

-
0.00000 T U N S M U U T Y : I I N I I U I (P.C.)

-
0.00000
MAIIIUI (P.C.) 0.00000 IAXIIUI (P.C.) 0.00000
INQEIFN 0.00000 INCWEW o.Ooo00
FILTEL UACTANWCE (P.U.) I .m F I L T U IEACTANCE ( P . U . ) 0.70000
WUIBEI OF BUDGES I N SEIIES 4 WIBU OF WIDGES IN SUlEs 2

ComElTot P O ~ I FACIOI CONVUTU POW FACT01


Dc LINT VOLTAGE (KV) DC Llm VOLTACE (KV) -220.00000
CONVEllUl C O W L ANGLE 10.00000 COwVUTEl COwTlOL AWCLE 8.00000
NNSFOUEI TAP T U N S M U U TAP
CowVEIlEl Dc POVEI ( I V ) 500.00000 CONVUIU. DC POVU (IV)
T E U I N A L IEACIIVE P
O W (IVAI) TEUINAL IEACTIYE POYE1 (IVAI)
AC TEUINAL VOLTACE ( K V ) AC m l I N A L VOLTAGE (KV)
W N V U I U DC CUuEIlf (KA) CONVUTEI DC CVuEnr (KA)
158 5 LOAD FLOW UNDER POWER ELECTRONIC CONTROL

SOLUTION CONYEICED IN 7 P-D ANTI 6V-Q ITERATIONS

WAD CE)IEUTION AC W S S S IISIATCH SWVMS


IV IVAl W IVAL IV IVN IV #VAL IVM

4496.80 1518.60 5226.58 791.80 194.91 -306.06 534.87 -182.89 37.85

@'ELATING STATE OF COIwELTEl Q WICM IS AlTACWED M BUS 7 (BUNTHOlPE220)

CONVELlEL IS OPEUTING IN THE MVELTION NODE


TKE CUKfLOL ANGLE IS TKE EXTINCTION ADVANCE ANGLE
DC p m SUPPLIEDm THE AC SYSTEI = IV
CDNVEIIEI AC VOLTACE TUNSMUE1 TAP CONTML ANGLE COIImATION ANGLE DC C W T DC VOLTAGE
(Y-VOLTS) (PEL CENT) (DW (DGS) (K-ANPS) (I(-VOLTS)

87.94 -8.26 8.00 22.45 I .406 -220.00

W L TUNSFEQ
LIM TUIINAL POVEL = -309.23 IV 207.61 IVAL
no1 TUNSFDUEL M COmFITEI = -309.23 IV 125.89 IVAt
REACTIVE P D V U OF FILTElS = 142.86 IVAL

BUS DATA
CENEMTION MAD SHUNT
BUS NAlE VOLTS ANGLE IV IVAl IV IVAl IVAL BUS NAIE W IVN

104 AVIEIOU-220 1.052 4.78 220.00 -33.87 0.00 0.00 0.00


108 ~ ~ ~ 1 0 ~ 41.89
~ 2 2 -10.17
0
108 BENIOLC220 41.89 -10.17
268 VAITAKI-220 136.22 -13.53
IISIATCH o.oO0 o.Oo0
108 BMIOLC-220 1.052 4.43 540.00 -88.04 97.20 0.00 0.00
104 AVIEIDlE-220 41.83 7.88
104 AVIEIOU-220 -41.83 7.88
255 W 1 2 E 6 2 2 0 26.47 6.87
IISIAtCW 5OO.oO0 -110.672
118 BUY-220 0.968 -12.95 0.00 0.00 329.60 95.80 0.00
167 ISLINCMN220 -120.18 -76.16
181 LAhDf02-220 -209.41 -19.64
IISIATCU 4.019 4.002

5.4 References
1. Fuertes-Esquivel, C R and Acha, E, (1996). Newton-Raphson algorithm for the reliable solu-
tion of large power networks with embedded FACTS devices, Proceedings of the IEE on
Generation, Transmission and Distribution, 143 (3,pp. 447-453.
2. Fuertes-Esquivel, C R, Acha, E and Ambriz-Perez, H, (1998). A thyristor controlled series
compensator model for the power flow solution of practical power networks, IEEE Trunsuc-
tions Winter Power Meeting, Paper 98WM 162.
3. Gyugyi, L, (1992). A unified power flow control concept for flexible a.c. transmission
systems, Proceedings of the IEE, 139, (4), pp. 323-333.
4. Fuertes-Esquivel, C R and Acha, E, (1998). The unified power flow controller: A critical
comparison of Newton-Raphson UPFC algorithms in power flow studies, Proceedings of the
IEE on Generation, Transmission and Distribution, 144 (9,pp. 437-444.
5. Sato, H and Amllaga, J, (1969). Improved load-flow techniques for integrated a.c.-d.c.
systems, Proc. IEE, 116. (4), pp. 525-532.
5.4 REFERENCES 159

6. Braunayal, D A, Kraft, L A and Whysong, J L, (1976). Inclusion of the converter and trans-
mission equations directly in a Newton power flow, IEEE Transactions on Power Apparatus
and Systems, PAS-75, (1). pp. 76-88.
7. Arrillaga, J, Harker, B J and Turner, K S, (1980). Clarifying an ambiguity in recent a.c.-d.c.
load-flow formulations, Proceedings of the IEE, 127, Pt.C(5), pp. 324-325.
8. Reeve, J, Fahmy, G and Stott, B, (1976). Versatile load-flow method for multi-terminal
HVDC systems, IEEE PES Summer Meeting, Paper F76-354- 1. Portland.
ELECTROMAGNETIC
TRANSIENTS

6.1 Introduction
An electrical power system is subjected to many types of disturbances, all of which
result in a transient. Accurate dynamic simulation is, therefore, essential in power
system design to minimize the disruption and possible damage of equipment due to
the overvoltages and overcurrents caused by the disturbance. Unlike electronic systems,
the size and cost of power system components makes it unrealistic to ‘bread-board’
alternative proposals. Each proposal must be checked thoroughly to ensure satisfactory
operation and optimize component parameters and controller settings, before alterations
are made. This may be done at the planning stage, where many scenarios for improving
the transient performance must be investigated, or during operation to diagnose the
cause of the disturbance.
The term ‘electromagnetic transient ’ refers to transients that involve the interaction
between the energy stored in the magnetic fields of the inductances and electric field of
the capacitances in the system. It does not include the slower ‘electromechanical tran-
sient ’ response, which involve the interaction between the mechanical energy stored in
the rotating machines and the electrical energy stored in the electrical network; these
are covered in Chapters 7 and 8.
However, no component model is appropriate for all types of transient analysis and
must be tailored to the scope of the study. The main criterion for the selection of
model components is the time range of the study. For instance, when analysing fast
transients, such as lightning phenomena, stray capacitance and inductances must be
represented and the solution step size needs to be at least h t h of the smallest time
constant introduced by the stray parameters. For system disturbances, such as short-
circuits, the dynamics of power electronic controllers, rather than stray components,
will play the major role and the solution step size can be considerably larger.
Regardless of the type of system equivalent and size of the integration steps, the
EMTP concept proposed by Dommel [ 1-4) is now universally accepted for the simula-
tion of complex power systems containing non-linearities, power electronic components
and their controllers.
6 ELECTROMAGNETIC TRANSIENTS

6.2 Background and Definitions


Most of the programs developed for power system transient simulation are based on
Bergeron’s method This method uses linear relationships (characteristics) between
current and voltage, which are invariant from the point of view of an observer travelling
with the wave. However, the time intervals or discrete steps required by the digital
solution generate truncation errors which often leads to numerical instability. The use
of the trapezoidal rule for the integration of the ordinary differential equations has
improved the situation considerably.
Dommel’s EMTP method combines the method of characteristics and the trapezoidal
rule into a generalized algorithm, which permits the accurate simulation of transients
in networks involving distributed as well as lumped parameters.
The method, generally referred to as Numerical Integration Substitution, makes use
of difference equations for the digital simulation of the power system ‘continuous’
dynamic behaviour.
It is often referred to by alternative names. One of them, proposed by G.T. Heydt [ 6 ] ,
is the Method Of Companion Circuits, as the difference equation can be viewed as a
Norton equivalent (or companion circuit) for each element in the circuit. Another is the
Nodal Conductance Approach (NDA), to emphasize the use of the nodal formulation,
each of the network components being represented by its companion circuit.

6.3 Numerical Integrator Substitution


As the name implies, Numerical Integrator Substitution involves substituting a numer-
ical integration formula into the differential equation and rearranging it to the appro-
priate form.
Dommel’s method involves substituting the trapezoidal integrator into the differen-
tial equation. While other integrators could have been used, the trapezoidal solution
was preferred due to its simplicity as well as being A-stable and reasonably accurate
in most circumstances. However, since it is based on a truncated Taylor series, numer-
ical oscillations can still occur under certain conditions due to the neglected terms.
Substituting an integrator is equivalent to substituting the appropriate finite differ-
ence approximation into the differential equations. Another important contribution of
Dommel’s work is the replacement of inductors and capacitors by a resistor and current
source in parallel (the latter representing previous history terms) and their integration
into a nodal conductance matrix and injected currents vector, respectively, to find a
solution for the complete system.
These two components are derived using the trapezoidal rule as shown in Figure 6.1.

1-At

Figure 6.1 Trapezoidal rule


6.3 NUMERICAL INTEGRATOR SUBSTITUTION 163

'-i"
+ +
-
"r
Figure Resistance

6.3.1 Resistance
This case, shown in Figure 6.2, is straightforward, i.e.
vk(t) - = RikSrn(t)

6.3.2 Inductance
The differential equation for the inductor shown in Figure 6.3 is:
dikm
VL = vk - Vjn = (6.3)
dt
which must be integrated from a known state at - At to the unknown one at t , i.e.

and, applying the trapezoidal rule, Equation (6.4) can be replaced by:
At
i k n d t ) = ikm(r-Ar) -k -((vk
2L
- + - vrn)(r-Af))
At At
= ikm(1-A') + -(vk(f-Af)
2L
- %(f-Af))
2L
+ - Vm(r))* (6.5)
This equation can be rewritten in the following form:

ik,n(l) = Ihistory(r-At) + R e1f f - urn([)),


which is recognized as a Norton equivalent, as shown in Figure 6.4.

Figure 6.3 Inductor


164 6 ELECTROMAGNETIC TRANSIENTS

Figure 6.4 Norton equivalent of inductor

The term l/R,ff = G,ff is the instantaneous term that relates the present time voltage
to a present time current contribution, and involves a pure resistance. The term
Ihistory(t-At) is the history term as this current source value is a function of quantities
at previous time steps where
At
Ihistory(f-Af) = ikm(t-Af) + -(vk(t-Af)
2L
- vm(f-At))
and
2L
Reff = -.
At
Transforming to the z-domain gives:

6.3.3 Capacitance
The differential equation for the capacitor, shown in Figure 6.5, is:
d(vk
ikm = c- =c (6.10)
dt dt
Rearranging it as an integral:

and applying trapezoidal integration:


6.3 NUMERICAL INTEGRATOR SUBSTITUTION 165

Figure 6.5 Capacitor

Hence the current in the capacitor is given by:

(6.13)

which, again, is interpreted as the Norton equivalent shown in Figure 6.6.


Transforming to the z-domain gives:

(6.16)

Combinations of components can be replaced by a single Norton equivalent, thereby


reducing the number of nodes and, hence, the computation at each time step. For
example, each tuned RLC branch forming part of a harmonic filter bank can be

Figure 6.6 Norton equivalent of capacitor


166 6 ELECTROMAGNETIC TRANSIENTS

Figure 6.7 Reduction of RLC branch to a Norton equivalent

represented by a Norton equivalent, as shown in Figure 6.7. Equally, the complete


filter bank can be represented by a single Norton equivalent that combines the equiv-
alents of the individual filter branches. The reduction process involves a series of
Thevenin to Norton transformations, the internal nodes being reduced by Gaussian
elimination; the result is a greatly simplified nodal admittance matrix.

6.4 Transmission Lines and Cables


The models used for overhead transmission lines and cables are identical except for the
derivation of their electrical parameters (described in Chapter 2). The work of Carson
[7] forms the basis of overhead transmission line parameter calculations, where either
numerical integration of Carson's integral equation, the use of Carson's series or a
complex depth approximation are used. Underground cable parameters are calculated
using Pollaczek's equations [8].
The three transmission line models commonly used in electromagnetic transient
programs are section, Bergeron and frequency-dependent line; the latter two being
classed as travelling wave models.
The section model is used for short lines, where the travel time is less than the
time step. This corresponds to approximately 15 km for a 50 ps time step. Typically
6.4 TRANSMISSION LINES AND CABLES 167

this occurs in distribution rather than transmission systems. This model is unsuit-
able for longer lines as the number of nominal sections required for an adequate
representation makes the solution very inefficient.
With the losses ignored, Bergeron’s method provides a simple and elegant travelling
wave model, which is then completed by the addition of series resistance to represent
the losses. In its original implementation, EMTP used the concept of natural modes
to represent multi-conductor lines and the main effort went into the development of
frequency-dependent modal transformations and fitting techniques to try and improve
accuracy and thus eliminate prospective instabilities. However, Gustavsen and Semlyen
[9] have indicated that, although the phase domain problem is inherently stable, the
associated modal domain may be inherently unstable; hence, regardless of the fitting
accuracy, the modal line model will be unstable. This has spurned research effort into
modelling lines directly in the phase-domain despite the complication of representing
couplings between phases [lo- 151.

6.4.1 Bergeron line model

Consider a lossless distributed parameter line as depicted in Figure 6.8, with (induc-
tance) and C’ (capacitance) per unit length. The wave propagation equations for this
line are:

(6.17a)

(6.17b)

The general solutions of Equations (6.17a) and (6.17b) are:

=fl(x- +f2(x + (6.18a)


t ) = Zfl - - + st), (6.18b)

where - and f 2 ( x + are arbitrary functions of - and + sr) to


be determined from problem boundary and initial conditions, 1 - represents a
wave travelling at velocity s in a forward direction and + a wave travelling
in a backward direction.

X= x=d

Figure 6.8 Propagation of a wave on a transmission line


168 6 ELECTROMAGNETIC TRANSIENTS

Z, the surge (or characteristic) impedance and s, the propagation velocity for a
lossless line are given by:

1
s=-

Multiplying Equation (6.18a) by ZC and adding it to, and subtracting it from, Equa-
tion (6.18b) gives the required branch equations, i.e.
+Zci(x, = 2Zcfl - (6.19a)
v(x, - Z&, = -2Z~f2(x + (6.19b)
Note that v(x, + Z c i ( x , is constant when - is constant. This can be inter-
preted by becoming a fictitious observer travelling along the line as shown in Figure 6.8
with the wave. Then (x - and v(x, +
Z c i ( x , will appear constant all along the
line. If the travel time to get from terminal to terminal m of a line of length d is

then the expression +


Z c i ( x , t ) seen by the observer when leaving terminal at
time - must be the same when the observer arrives at terminal at time t , i.e.
- + - = +
Rearranging the latter leads to the simple two-port equation for i.e.

(6.20)

where the current source from past history terms is:


1
- = - - - (6.21)
ZC
Similarly for the other end
1
ikm =
ZC
+I k ( t - (6.22)

where
I
- = - - - 5). (6.23)
ZC
The expressions - = constant and + = constant are called the character-
istic equations of the differential equations.
Figure 6.9 depicts the resulting two-port model. In this model, there is no direct
connection between the two terminals and the conditions at one end are seen indirectly,
and with time delay (travelling time), at the other through the current sources. The
past history terms are stored in a ring buffer and hence the maximum travelling time
6.4 TRANSMISSION LINES AND CABLES 169

Figure 6.9 Equivalent two-port network for lossless line

that can be represented is the time step times the number of locations in the buffer.
Since the time delay is not usually a multiple of the time step the past history terms
either side of actual travelling time are extracted and interpolated to give the correct
travelling time.
The distributed series resistance of the line is approximated by treating the line as
lossless and adding lumped resistances at both ends. Although lumped resistances can
be inserted in many places along the line by dividing the total length into many line
sections, this makes little difference and, hence, two sections are normally used, as
depicted in Figure 6.10. This lumped resistance model gives reasonable answers only
if R/4 << ZC. By assigning half of the mid-point resistance to each line section the
representation for a section modelling half the line is depicted in Figure 6.11, where:

and
-1
- s / 2 ) = Zc + Rf4 - r/2) - (EF :\:) - 5/21. (6.25)

By cascading two half line sections and eliminating the mid-point variables, as only
the terminals are of interest, the model depicted in Figure 6.12 is obtained. This is in
the same form as the earlier models except that the current source representing history
terms is more complicated as it contains conditions from both ends on the line at

Figure 6.10 Line with lumped losses


170 6 ELECTROMAGNETIC TRANSIENTS

Figure 6.11 Half line section

Figure 6.12 Bergeron transmission line model

(t - t/2). For example, the expression for the current source at end is:

The EMTDC line model separates the propagation into low and high frequency paths
so that the line can have a higher attenuation to higher frequencies.

6.4.2 Multi-conductor transmission lines


Equations (6.17a) and (6.17b) are also valid for multi-conductor lines if the scalar
voltages and currents are replaced by vectors and using inductance and capacitance
matrices. Thus, in the frequency domain these equations are:

(6.27a)
6.4 TRANSMISSION LINES AND CABLES 171

By differentiating a second time one vector either voltage or current may be elimi-
nated giving:

Traditionally the complication of having off-diagonal elements in the matrices of


Equation (6.28) and (6.29) is overcome by using natural modes. Eigenvalue analysis is
applied to produce diagonal matrices thereby transforming from coupled equations in
the phase domain to decoupled equations in the modal domain. Each equation in the
modal domain is then solved as for a single phase line by using modal travelling time
and modal surge impedance. The transformation matrices between phase and modal
quantities are different for voltage and current, i.e.
(6.30a)
(6.30b)

Substituting Equation (6.30a) in (6.28) gives:

(6.31)

Hence

[%] = = [A][vmodel. (6.32)

To derive the matrix IT,] that diagonalizes [ZLhase][ Ybhase], its eigenvalues and
eigenvectors must be found. However, eigenvectors are not unique, as when multi-
plied by a nonzero complex constant they still are valid eigenvectors; therefore, some
normalization is desirable to allow from different programs to be compared.
PSCADRMTDC uses the root squaring technique developed by Wedepohl for eigen-
value analysis [ 161. To generate frequency-dependent line models the eigenvectors
must be consistent from one frequency to the next, such that the eigenvectors form
a continuous function of frequency, so that curve fitting can be applied. A Newton
Raphson algorithm has been developed for this purpose [ 171.
On completion of the eigenvalue analysis, the following relationships are used:

[Zmodel = [Tvl-'[Zphasel[Til,
[yrncdel = [Til-'[Yphasel[Twl~
172 6 ELECTROMAGNETIC TRANSIENTS

As [z;ha,.[Ybha,] is different from [YbhaSe][Zbhase],the eigenvectors are also


different, even though their eigenvalues are identical. However, [Ti] = ( and,
therefore, only one of these matrices needs to be calculated.
Looking at mode i, i.e. the ith equation of (6.32), gives:

(6.33)

and its general solution at point x in the line is:


(6.34)

where
=
VF = forward travelling wave,
VB = backward travelling wave.

Equation (6.34) contains two arbitrary integration constants. An n-conductor line


has natural modes and thus requires 2n arbitrary constants. This is consistent with
the existence of 2n boundary conditions, one for each end of every conductor.
In matrix form, Equation (6.34) becomes:
Ymde(x) = [e-YX~
* Y:,de(k) + * Y:de(m) (6.35)

Reintroducing phase quantities by using Equation (6.32) gives:

~ ( x= +
) [ e - r x l ~ F [erx] - yB, (6.36)

is the propagation matrix, and

Similarly the solution of (6.29) for current gives:

) re- rx 1~F - [erx IL = Y ~ ( [ ~ - -


~ ( x= ~ [erxIvB),
~ I v ~ (6.37)

where
IF= forward travelling wave,
IB = backward travelling wave.

Thus, the voltage and current at the end of the line are:
V(k)= (v” + vB,, (6.38)
= + lB)= Yc(VF - B
1, (6.39)
6.4 TRANSMISSION LINES AND CABLES 173

and, similarly, at the end of the line:


- +
~ ( m =) [ e - r r ~ ~ F[erX]~’ (6.40)
=Y ~ ( [ ~ - ~-
’ I[er’1yB).
v~ (6.41)
Hence, the forward and backward travelling voltage waves at terminal are:
VF = + (6.42)
- = - ZcI(k)/2. (6.43)
And since
+
[ Y ~ I V&~ = =2[e-”]~~
the forward and backward travelling current waves at terminal are:
(6.44)

(6.45)

6.4.3 Frequency-dependent model


As the line parameters are functions of frequency, the relevant equations are first
expressed in the frequency domain and extensive use made of curve fitting to incorpo-
rate the frequency-dependent parameters [ 18-20]. Analysis proceeds by first consid-
ering a frequency-domain solution for a single conductor line of length
V k ( 0 ) = c ~ s h [ y ( ~ ) l ] V m=( ~ ) sinh[y(w)l]i,(w), (6.46)

(6.47)

where

is the propagation constant, and

is the characteristic impedance.


Y’(w) = G’ +
is the per unit length shunt admittance obtained from conductor geometry, and
= R’ +
is the per unit length series impedance, also obtained from conductor geometry. Let
= vk(w) +
= Vk(w) - zC(w)ik(w),
Fm(w) = Vm(w> +
(6.48)
174 6 ELECTROMAGNETIC TRANSIENTS

The functions F and B correspond to forward and backward travelling waves, respec-
tively in the frequency domain. From Equations (6.46), (6.47) and (6.48):
=A I
Bm(w) = A I ( W ) F k ( W ) , (6.49)
where
Al(w) = = cosh(y(w)l) - sinh(y(w)l).
Equations (6.49) yield the equivalent circuit of Figure 6.13 with the source terms
Bk and B , related to electrical quantities at the other end of the line.
To convert the frequency domain circuit of Figure 6.13 to the time domain, it is only
necessary to express the source terms Bk and in the time domain. This requires a
convolution integral in place of the multiplication in Equations (6.49):

Bk(t) = L'A(u)F,,(t - u)du

B,,(t) = l A ( u ) F k ( r - (6.50)

The lower integral limit of is set equal to the shortest possible !ransmission delay
of the line. Evaluating the convolution integrals (Equation (6.50)) at every time step
is very slow, and a recursive method is used instead [21]. The recursive convolution
method represents A ( u ) in Equation (6.50) by a sum of exponentials in and by
a low order polynomial. For example, if

/=I
and
fm(t - = + au2 +
then
Bk(t) = xe-"iu(f(t) + au2 + ,L?u)du (6.5 1 )

= - At) + + - At) + V/fm(t - 2At)], (6.52)

where and ,L? are obtained by equating the interpolating quadratic to - At),
- 2 A t ) and A, are constant coefficients obtained by integrating Equation (6.51)

Figure Single conductor line equivalent


6.4 TRANSMISSION LINES AND CABLES 175

from - At to t . The convolution integral is, therefore, replaced by a past-history term


B k ( t - A t ) , and a linear combination of past-history terms for electrical conditions at
the other end of the line.
Curve fitting for Zc and A @ ) Both Zc and A are readily defined in the frequency
domain. Representation in the time domain proceeds by first finding a rational poly-
nomial in s that matches Z or A along the axis. In EMTDC, the fitting takes place
at frequency points evenly distributed on a log scale between specified lower and
upper frequencies. The choice of lower frequency affects the shunt conductance at d.c.,
giving it a maximum value. Specifying a very low start frequency affects the accuracy
and efficiency of the fit at other frequencies.
Rational polynomial fitting is an area of active research, with many applications in
power system analysis. The method employed in the EMTDC transmission line and
cables program directly places poles and zeros in the s plane to obtain a match with
the interpolated function. The orders of the numerator and denominator polynomials
are incremented until a good match is obtained (up to a specified limit). Finding the
rational polynomial directly in terms of poles and zeros,
(6.53)
simplifies the implementation of the function in a form suitable for simulation. A partial
fraction expansion yields:
= k, +-kl
S+Pl
+-k2
S+P2
k3
+-+...
S+P3
(6.54)
which can be readily represented by an network in the case of z , or transforming
to the time domain for A @ ) ,
= kle-”’‘ + k2e-P?‘ + . . .
k, = 0 since in this case the denominator is of higher order than the numerator.
Implementation of the transmission line model is illustrated in Figure 6.14. The trans-
mission line and CABLES program calculates the RC networks, exponential sum
,
I
I
I
Coupled multi-conductor line I
I

/ Decoupled multi-mode line


I

/
I

RC network ----.
Modal
transform
-
Modal
transform

-.

Figure 6.14 Implementation of the frequency-dependent transmission-line model in EMTDC


176 6 ELECTROMAGNETIC TRANSIENTS

approximation to the propagation constant, and additionally an exponential sum for


every term in the modal transform matrices. This is because the modal transform
matrices are frequency dependent but can be treated in the same way as the propagation
constant using recursive convolution. In order to obtain continuity in calculated eigen-
values and eigenvectors at the 100 sample points, the solution at frequency N is used
as the starting point for a Newton Raphson iterative refinement at frequency N 1. +
Numerical illustration The line propagation constant, + is a function
of frequency, while the propagation function (e-(cr(w)+'p(w))') is a function of frequency
and line length
The variation of the amplitude term e-n(w)' of a single phase line is shown in
Figure 6.15(a) for a line length = 100 km. These results illustrate the line low pass
characteristics. As the line length is in the exponent, the attenuation of travelling waves
increases with it.
The variation with frequency of the phase angle of the propagation function, e-jp(w)',
is shown in Figure 6.15(b). A negative phase represents a phase lag in the travelling
waveform and its counterpart in the time domain is a time delay. The phase angle is a
continuous function, becoming more negative with frequency but, for display purposes,
it is constrained to be within the range -180" to 180". It is a difficult function to fit
and requires a high order rational function to achieve sufficient accuracy. However,
multiplication by e-Jsr, where represents the time for a wave to travel from one end
of the line to the other (in this case 0.33597 ms), results in the smooth function shown
in Figure 6.15(b). With this procedure, referred to as backwinding, the attenuation
(complex function) is easily fitted with a low order rational function. To obtain the
correct response, the model must counter the phase advance applied in the frequency
domain fitting. This is performed in the time domain implementation by incorporating

Magnitude IAttenuationl =

Phase - (Degrees) Angle (Attenuation)=e-i@(w)'

-1

With backwinding
-3OOL I
1 2 3 4 5
(b) Log ( 2 4

Figure 6.15 Propagation function


6.4 TRANSMISSION LINES AND CABLES 177

Magnitude Propagation constant

(a)
0.4 ' 3 3.5 4 4.5
I
5

-
Phase (Degrees)

1 1.5 2 4 4.5 5
(b)

Figure 6.16 Propagation constant

a time delay T. With this purpose, a buffer of past voltages and currents at each end
of the line is maintained and the values delayed by are used. Because in general is
not an integer multiple of the time step, interpolation between the values in the buffer
is required to get the correct time delay.
Figure 6.16 shows the match obtained when applying least squares fitting of a
rational function (numerator order 2, denominator order 3). The number of poles is
normally one more than the zeros for the attenuation function magnitude which must
go to zero as frequency approaches infinity.
Although the fitting is good, close inspection shows a slight error at fundamental
frequency. Any slight discrepancy at fundamental frequency shows up as a steady-state
error, which is undesirable. This occurs because the least squares fitting tends to smear
the error across the frequency range. To control this problem, a weighting factor can
be applied to specified frequency ranges (such as around fundamental frequency) when
applying the fitting procedure. When the fitting has been completed any slight error
still remaining is removed by multiplying the rational function by a constant to give
the correct value at low frequency. This sets the d.c. gain (i.e. its value when s is set
to zero) of the fitted rational function. The value of k controls the d.c. gain of this
rational function and is calculated from the d.c. resistance and the d.c. gain of the surge
impedance, thereby ensuring that the correct d.c. resistance is exhibited by the model.
At frequencies when the value of drops below a specified threshold, the trav-
elling waves will have been attenuated by the time they reach the end of the line;
therefore the associated terms in the fitted are neglected. Also frequencies well
above the Nyquist rate (e.g. 10 x Nyquist frequency for selected time-step) will not
influence the simulation results and the corresponding terms are removed.
178 ELECTROMAGNETICTRANSIENTS

Some fitting techniques force the poles and zeros to be real and stable (in the left hand
half of the s-plane) while others allow complex poles and use other methods to ensure
stable fits (such as vector-fitting). A common approach is to assume a minimum-phase
function and use real half-plane poles. Fitting can be performed either in s-domain or
z-domain, each alternative having advantages and disadvantages.
The same fitting algorithm can be used for fitting the characteristic impedance (or
admittance if using the Norton form). The number of poles and zeros is the same in
both cases. Hence the partial expansion of the fitted rational function is:

This can be implemented by using a series of RC parallel blocks (the Foster I


realization), which gives = ko, R; = k i / p i . and C; = I / k ; . Either the trapezoidal
rule can be applied to the RC network or better still recursive convolution.
The shunt conductance is not normally calculated. At low frequencies the
surge impedance becomes larger as the frequency approaches zero, i.e.

This trend can be seen in Figure 6.17 which shows the characteristic (or surge)
impedance calculated by the transmission line parameter program down to Hz.

Magnitude IZsurgel
.

Start frequency
1

-
Phase (Degrees) Angle (Zsurge)

(b) Log(2 n f )

Figure 6.17 Characteristic impedance


180 6 ELECTROMAGNETIC TRANSIENTS

Using triangular factorization (arrow l), Equation (6.57) reduces to:


(1)

(6.58)
Forward reduction (arrow 2), followed by back substitution (arrows 3) is then used
to get V,(t), i.e.
(3) (2)

(6.59)
where

Once V u ( t ) has been found, the history terms for the next time step; are calculated.

6.5.1 Modification for switching and varying parameters


To represent switching operations, or time varying parameters, matrices [Guu] and
[ G u K ]need to be altered and retriangularized. The solution is simplified by placing
the nodes with switches last, such that the initial triangular factorization is reduced to
the nodes without switches, i.e.

Switching components

(6.60)
6.5 FORMULATION AND SOLUTION OF THE SYSTEM NODAL EQUATIONS 181

followed by complete triangulation (arrow 2)

(6.61)
and then forward reduction (arrow 3) and back substitution (arrow 4). i.e.
(4) (3)

T (6.62)
Transmission lines do not introduce off-diagonal elements. This produces a block
diagonal structure for [Guu]. Each block represents a subsystem that can be solved
for independently, as any influence from the rest of the system is implemented in the
form of history terms. This allows parallel computation and is used extensively in
real time digital simulators. PSCADEMTDC performs the triangular factorization on
a subsystem basis rather than on the entire matrix.

6.5.2 Non-linear or time varying parameters


Typical non-linearities requiring representation are the saturated inductances of
transformers and reactors and the resistances of surge arresters. Non-linear effects
in synchronous machines are incorporated directly in the machine equations. As the
number of non-linear elements are limited, it is more efficient to modify the linear
solution than using a non-linear solution method for the entire network. The three
alternative approaches [22] used for this purpose are:
Current-source representation (with one time step delay).
Compensation method.
Piecewise linear representation.
Current-source representation This method uses a current source to model the
current drawn by the non-linear component, its value calculated from information
at previous time-steps. Therefore it lacks an instantaneous term and appears as an
‘open circuir’ to the voltages at the present time-step. Thus, to avoid possible related
instabilities, a correction source is used in the form of a large ‘fictitious’ Norton
resistance. However there is a one time-step delay in the correction source.
Compensation method If there is only one non-linear branch a compensation
method can be applied, whereby the non-linear branch is excluded from the network
and replaced by a current source. The total network solution v(t) is then equal to the
182 6 ELECTROMAGNETIC TRANSIENTS

value vo(t) found with the non-linear branch omitted, plus the contribution produced
by the non-linear branch, i.e.
v(l) = vO(t) - RTheveninikm(t)t (6.63)
where
Rmevenin= Thevenin resistance of the network without non-linear branch
connected between nodes and
vg(t) = open circuit voltage of network, i.e. voltage between nodes and
without non-linear branch connected.
The Thevenin resistance of the linear network is calculated by taking the difference
between the th and th columns of matrix [Guul-' . This is achieved by solving
[ G u ~ ] v ( t )= 1; with 1; set to zero except for the and elements, which are -1
and 1, respectively. It is equivalent to finding the terminal voltage when connecting a
current source (of magnitude 1) between nodes and The Thevenin resistance is pre-
computed only once, before entering the time step loop, and only needs re-computing
whenever switches open or close.
Two scalar equations are then solved simultaneously, as shown in the diagram of
Figure 6.18, i.e.
Vkrn(t) = VkmO(t) - RTheveninikmT (6.64)
Vkrn(l)= f ( i k m r dikmldt, * * *),
If Equation (6.65) is given as an analytic expression, a Newton Raphson solution
is used. When Equation (6.65) is defined point-by-point as a piecewise linear curve, a
search procedure is used instead.
Even though there may be more that one non-linear branch, using the subsystem
concept described in Section 6.6, the compensation approach can still be used, as long
as there is only one non-linear branch per subsystem.
If the non-linear branch is defined by Vkm = f ( i k m ) or Vkm = R(t)iknl the solution is
straightforward. For a non-linear inductor the flux (A = f ( i k m ) ) is the integral of the
voltage with time, i.e.

= A(l - A t ) + (6.66)

Figure 6.18 Pictorial view of simultaneous solution of two equations


6.6 USE OF SUBSYSTEMS 183

Figure 6.19 Artificial negative damping

Equation (6.66) is solved using the trapezoidal rule, giving:


Af
A(f> = -v(f> AHistory(f -
2
where
At
= A(f - A l ) + -v(f
2
- Af). (6.67)

However, numerical problems can occur with non-linear elements when Af is too
large because the regions of the non-linear characteristic between samples will be
missed. This, as shown in Figure 6.19, may produce fictitious negative damping or
hysteresis.

Piecewise linear representation The piecewise linear inductor characteristic, as


depicted in Figure 6.20, can be represented as a linear inductor in parallel with a current
source representing saturation. This model is very accurate when the compensation is
small.

6.6 Use of Subsystems


The presence of transmission lines and cables in the system being simulated introduces
decoupling into the conductance matrix. This is because the transmission line model
injects current at one terminal as a function of the voltage at the other at previous time
steps. In the present time step, there is no dependence on electrical conditions at distant
terminals of the line, frequently leading to a block diagonal conductance matrix, i.e.

Y = (6.68)
184 6 ELECTROMAGNETIC TRANSIENTS

'Compensation

-j icompensation
Linear inductor

Figure 6.20 Piecewise linear inductor

(b)

Figure 6.21 Separation of two coupled subsystems by means of linearized equivalent sources

Each decoupled block in this matrix is a subsystem, and can be solved at each time
step independently of all other subsystems.
Figure 6.21(a) illustrates coupled systems that are to be separated into subsystems.
Each subsystem in Figure 6.21(b) is represented in the other by a linear equivalent.
The Norton equivalent is constructed using information from the previous time step,
looking into subsystem (2) from bus The shunt connected at is considered
USE SUBSYSTEMS 185

part of (1). The Norton admittance is

The Norton current is

The Thevenin impedance is

and the voltage source is

If Y A is a capacitor bank, Z is a series inductor, and Y B is small, then


Y N% YA and Y N =IBA(t - A?) (the inductor current),
ZTH % and VTH = v A ( t - At) (the capacitor voltage).
When simulating HVDC systems, it can frequently be arranged that the subsystems
containing each end of the link are small, so that only a small conductance matrix need
be refactored after every switching. Even if the link is not terminated at transmission
lines or cables, a subsystem boundary can still be created by introducing a one-time
step delay at the commutating bus.
A d.c. link subdivided into subsystems is illustrated in Figure 6.22.
The method of interfacing subsystems by controlled sources is also used to interface
subsystems with component models solved by another algorithm, e.g. components
using Numerical Integration Substitution on a state variable formulation. Synchronous
machine and SVC models are framed in state variables in PSCADEMTDC and appear

y
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I

Figure 6.22 A typical subsystem equivalent for a d.c. link


186 6 ELECTROMAGNETIC TRANSIENTS

to their parent subsystems as controlled sources. As for interfacing subsystems, best


results are obtained if the voltage and current at the point of connection are stabilized,
and if each componendmodel is represented in the other as a linearized equivalent
around the solution at the previous time step. In the case of synchronous machines, a
suitable linearizing equivalent is the subtransient reactance, which should be connected
in shunt with the machine current injection.

6.7 Switching Discontinuities


The basic EMTP-type algorithm requires modification in order to model accurately and
efficiently the switching actions associated with HVDC, thyristors, FACTS devices,
or any other piecewise linear circuit. The simplest approach is to simulate normally
until a switching is detected and then update the system topology andor conductance
matrix.
A switch can be represented, as shown in Figure 6.23, by either an ordoff resistance
or an off resistance only. The former method cannot represent an ideal switch, since Ron
must be large enough not to de-condition the system conductance matrix. In practice,
this represents well real switching devices, which are themselves not ideal. The second
method avoids the use of small resistances, but requires a more severe change in
system topology since there is one fewer node in the conduction state than in the
off-state.
For either representation, the system conductance matrix must be reformed and
factorized after each change in conduction state. This considerably increases the compu-
tational requirements of the simulation in proportion to the number of switching actions
(recall that the efficiency of the EMTP technique lies in the fact that the conduc-
tance matrix is held constant to avoid refactorization). Nevertheless, for HVDC and
most FACTS applications, the switching rate is only several kHz, so that the overall
simulation is still fast.
The efficiency and elegance of the EMTP method relies on the step length being
constant. This, however, causes firing errors when modelling switching elements as
the switching instants will not normally coincide with the time steps, as shown in

I I
Figure Switch representation
187

Figure 6.24 Apparent reverse current in a diode due to placement of simulation time steps

Figure 6.24 for the case of a diode-ending conduction. At (A) the diode should be
switched off, but the negative current is only detected at (B) and the conductance
matrix is reformed at (C).
An effective solution of this problem is to apply a linear interpolation at (B) to
find all the nodal voltages at a point very close to (A). An exact solution for the zero
crossing is a non-linear problem; however, given the close proximity of points t and
+
t At, a linear interpolation between them introduces no significant error [22,23].
In the above example, the diode model must include logic to detect the switching
event and then estimate the instant t + of its occurrence between t and t + At. For
a diode, a linear interpolation on the forward current yields:
Atif(t)
(6.73)
- if(f + At)
The nodal voltages at t + are then approximately:
~ (+t At) 2 ~ ( t+) -[[v(t + At) - ~ ( t ) ] , (6.74)
At
and similarly for branch currents in the system. If simulation proceeds normally from
t + T with the new conductance matrix, subsequent solution points will be shifted by
For convenience, an additional interpolation can be made after the first time step
with the new conductance matrix to bring the simulation back onto the original time
sequence, yielding the sequence of steps illustrated in Figure 6.25. Another option is

>
New conductance matrix

Figure 6.25 A double interpolation scheme to find the switching instant and re-synchronise
time steps
188 6 ELECTROMAGNETIC TRANSIENTS

to accept unequally spaced points in the solution (which complicates Fourier anal-
ysis of the resulting waveforms), or to fit a cubic spline to the solution and re-index
to any desired time base. In the PSCADEMTDC package, the interpolation scheme
of Figure 6.25 is used but with two additional interpolations introduced to eliminate
voltage and current chatter, to be discussed in the next section.

6.7.1 Voltage and current chatter due to discontinuities


The trapezoidal integration has the effect of increasing the apparent impedance of
inductors at high frequencies approaching the Nyquist rate. A corollary is that large
voltages will be developed across inductors if high frequency currents are injected into
them. This effect is manifested most notably with respect to switching actions or point
discontinuities, which necessarily contain high frequency components [25-271.
Figure 6.26 shows an inductor in series with a diode ending conduction. Assuming
that the inductor current falls steadily to zero at T, the voltage will be constant at V L .
At the diode switches off and the inductor voltage will fall instantaneously to zero,
since dil/dr = 0. The inductor voltage is shown in Figure 6.27(a).

-
f

Figure 6.26 ~ - circuit M ..ich will display voltage chatter after the diode switches off

- , \

' -I

Figure 6.27 (a) Correct solution for voltage across the inductor. Chatter voltage at the
diode anode
6.7 SWITCHING DISCONTINUITIES 189

The trapezoidal integration equation for the inductor is:


2L
VL(t)=
At
- iL(r - A t ) ) - V L ( t- Ar). (6.75)

From (t + A t ) onwards
iL(t) = iL(t - At) =

so
V L ( t )= -VL(t - A t ) . (6.76)

The inductor voltage consequently oscillates between ~ V instead


L of falling to zero
(Figure 6.27(b)). Note, however, that the average inductor voltage is correct at zero,
and that the chatter does not grow larger.
PSCADEMTDC uses a double half-time step interpolation method to remove chatter
which relies only on trapezoidal integration [28]. With reference to Figure 6.28, the
following steps are involved:

trapezoidal integration,
interpolation back to discontinuity, between (1) and (2),
trapezoidal integration with new conductance matrix,
interpolation between (3) and (4) of half time step to remove
chatter,
trapezoidal integration to get past t At, +
interpolation between (6) and to re-synchronise time steps,
normal integration resumes.

Figure 6.28 Chatter removal


190 ELECTROMAGNETIC TRANSIENTS

The chatter problem discussed above is associated with the numerical error of the
trapezoidal (or any other integration) rule, i.e. it is inherent in the numerical integrator
substitution method. With the use of root matching techniques instead of numerical
integrator substitution [29], a chatter removal process is not required. Root matching is
always numerically stable and more efficient numerically than trapezoidal integration.
Root matching is only formulated for the branches containing at least two or more
elements (i.e. RL, RC.. RLC, LC..) and these can be intermixed in the same solution
with branches solved with other integration techniques.

6.8 Root-matching Technique


Since digital simulation requires the transformation from the s- to the z- plane, the
poles and zeros of the continuous process must be correctly represented in the digital
solution. However, when using the numerical integrator substitution method to derive a
difference equation, the poles and zeros are not inspected. They will, therefore, match
poorly those of the continuous system that is being simulated.
Ensuring the correct match for the poles and zeros is the basic purpose of the root-
matching technique [29,30].The difference equations generated by this method involve
exponential functions, as the transform equation 2-’ = e-SA‘is used rather than some
approximation to it. This concept is explained next.

6.8.1 Exponential form of difference equation


The application of Dommel’s method to a series RL branch produces the following
difference equation for the branch:
AtR At
(6.77)

Careful inspection of this equation shows that the first term is a first order approxi-
mation of e-’ where = AtR/L. The second term is a first order approximation of
(1 - e-X)/2.
This suggests that using the exponential expressions in the difference equation will
eliminate the truncation error and give accurate and stable simulations regardless of
the time step.
Equation (6.77) is thus expressed as:

(6.78)

Although the exponential form can be deduced from the difference equation devel-
oped by numerical integrator substitution, such an approach is unsuitable for most
transfer functions or electrical circuits, due to the difficulty in identifying the form
of the exponential that has been truncated. The root-matching technique provides a
rigorous method.
Previous Page

6.10 SYNCHRONOUS MACHINE MODEL [32] 195

A_ _ _ _ _ _ _ _ _ _Nor@ -operating point


______

Rectifier characteristic
at amin

amin
limit
1-
I
I- Current margin
I I
>
Figure 6.35 Classic V - I converter characteristic

simple control blocks or from specific HVDC control blocks. The d.c.-link controls
provided are a gamma or extinction angle and current control with voltage-dependent
current limits.
Power control must be implemented from general-purpose control blocks. The
general extinction angle and current controllers provided with PSCAD readily enable
the implementation of the classic V-I characteristic for a d.c. link, illustrated in
Figure 6.35.
General control modelling is made possible by the provision of a large number of
control building blocks including integrators with limits, real pole, PI control, second-
order complex pole, differential pole, derivative block, delay, limit, timer and ramp.
The control blocks are interfaced to the electrical circuit by a variety of metering
components and controlled sources.
A comprehensive report on the control arrangements, strategies and parameters used
in existing schemes has been prepared by CIGRE WG 14-02 [31].All these facilities
can easily be represented in the electromagnetic transient programs discussed in this
chapter.

6.10 Synchronous Machine Model [32]


The modelling of synchronous machines is based on Park’s transformation from phase
to dq0 quantities. With reference to Figure 6.36, the formulation of the machine
dynamics in phase components is as follows:

where
Va - iaRa
V , - icR,
(,b-ibRb) =: (I), (6.79)

(6.80)
196 6 ELECTROMAGNETICTRANSIENTS

Figure 6.36 Electrical machine windings

The inductances vary with e(t),the angle between field winding f and winding a
at time t , i.e.
Laa = La + Lm CoS(O),
Lbb = La + Lm C O S ( ~-(2~~ / 3 ) ) , (6.81)
Lc, = La + Lm CoS(2(6 - 4n/3)),
Lab = Lba = -M, - Lm COS(2(8 - ~ / 6 ) ) ,
Lbc = Lcb = -Ms - Lm COS(2(8 - n/2)), (6.82)
L,, = L,, = -M, - L, Cos(2(e + n/2)),
Laf = L f a = Mf cos(e),
Lbf = Lfb = Mf cost8 - 2 ~ / 3 ) , (6.83)
LCf = L f c = Mf - 4q3).
In dq0 axes and compact notation, the machine fluxes can be expressed as
(6.84)
where Park’s transformation is:

I I
2 2
6.10 SYNCHRONOUS MACHINE MODEL [32] 197

Expanding Equation (6.84) leads to

($) =
La
0
+ Lmd
O
:][3 [
h
+ (6.85)

where

and, similarly, for the field circuit:


@; = [ L m d l i d + [ L m d + L f I i ; .
Figure 6.37 depicts the equivalent circuit based on these equations.
The procedure to interface the machine to the rest of the electrical network is as
follows:
(1) Assume va, Vb, v, and vf from previous time step and calculate Vd, vq, vo using
transformation matrix [T(O)].
(2) Choose @ d , qq,@) (and @O if zero sequence is to be considered) as state variables

(3) These are integrated by numerical integration.

(6.88)

Lf

Figure 6.37 Equivalent circuit for synchronous machine equations


198 6 ELECTROMAGNETIC TRANSIENTS

All quantities (e.g. v d , v;, etc) on the right-hand side of Equation (6.88) are known
from the previous time step. The currents i d and i, are linear combinations of the
+
estate variables and therefore have a state variable formulation = [A]x [B]u(if
w is assumed constant).

(4) After solving for @ d , Ic/; and @, id, i; and i, (and io) can be calculated. These
are transformed back to obtain io, i h and i,. Also, the equation d8/dt = w is
used to update 8 in the transformation. In electromagnetic transient studies, the
machines mechanical behaviour is usually ignored and the integration for B is not
needed.
This solution process is depicted in Figure 6.38.
Additional rotor windings need to be represented due to the presence of damper
windings. In this case, more than two circuits exist on the d-axis and more than one
on the q-axis. For the case of a single damper winding on each of the d- and q- axes,
Equation (6.86) becomes:

where the subscript dd denotes the damper winding on the d-axis.


The extended version of Equations (6.88) being:
6.11 TRANSFORMER MODEL 199

The damper windings are short-circuited and, hence, their voltage is zero. Saturation
can be introduced in a variety of ways, a popular method being to make Lmd and Lmq
+ +
functions of the magnetising current (i.e. i d i d d if in the d-axis). For round rotor
machines, it is only necessary to saturate L m d and L,, but for start-up transients the
saturation of L, should also be included.

6.11 Transformer Model


Transformers are represented as mutually coupled windings. For a two winding trans-
former, illustrated in Figure 6.39, the voltage across the windings can be expressed as:

(6.90)

where
,511 and L22 are the self inductance of winding 1 and 2, respectively,
,512 and L21 are the mutual inductance between the windings.

To solve for the winding currents the inductance matrix has to be inverted, i.e.

(6.91)

The mutual coupling is bilateral, i.e. L I Z= L21 and the coupling coefficient between
the two coils is:
K12 = -L12
&G2'
(6.92)

Figure 6.39 Two coiled coils


200 6 ELECTROMAGNETIC TRANSIENTS

I i2

ideal
Transformer
“1
a.v2

0
I I

Figure 6.40 Equivalent circuit for two coupled windings

Rewriting Equation (6.90) using the turns ratio a = q / v 2 gives:

which can be represented by the equivalent circuit shown in Figure 6.40, where

This approach can be extended to n-coupled windings, the only difference being the
size of the inductance matrix.
However, transformer data are usually not available in this format. Instead, either
results from short-circuit and open-circuit tests are available or the magnetizing current
and leakage reactance are given based on machine rating.
A short-circuit test on winding 2, with the resistance neglected, gives:

11 = Vl/O(LI + L2).
An open circuit test with winding 2 open-circuited gives:

and, similarly, with winding 1 open-circuited gives:

From these test results, the values of L I I ,L22 and L12 are determined. These calcu-
lations are often internally performed by the program and the user only needs to enter
the leakage and magnetising reactances.
The values of the inductance matrix must be specified very accurately to avoid ill-
conditioning when subtracting two numbers of very similar magnitude. This is more
likely to happen for small magnetizing currents. For this reason the ideal transformer,
6.1 1 TRANSFORMER MODEL 201

L, L2
11
‘2
0 r
L w 0

Ideal
Vl Transformer v2
a:1

v Figure 6.41 Ideal transformer equivalent

Figure 6.42 Saturation modelling for mutually coupled windings

with the magnetizing term ignored, is often used. This is shown by the equivalent
circuit shown in Figure 6.41 and represented by the equation:

Figure 6.42 displays the modelling of saturation in mutually coupled windings.


Current source representation is used rather than varying the inductance as the latter
would require retriangulation of the matrix every time the inductance changed. When
the simulation is starting up, in order to reach the steady-state condition faster, it
is desirable to inhibit saturation. This is achieved by setting a limit value for the
flux (result of integration of voltage); the limit is removed just before applying the
disturbance to allow the flux to go into the saturation region.
Another refinement is to impose a decay time on the inrush currents, which occur
on energization or fault recovery. The additional circuit required for this purpose is
shown in Figure. 6.43.
The discretization of the transformer model, using the trapezoidal rule, does not
provide complete isolation between its terminals for d.c. That is, if a d.c. source is
applied to winding 1, a small amount of d.c. current will flow in winding 2, which
in practice cannot occur. A three-phase bank is modelled by the correct connection
of three two-coupled winding. For example, the star-delta connection is shown in
Figure 6.44.
202 6 ELECTROMAGNETIC TRANSIENTS

Integration

A IS
4

@S

Figure 6.43 Saturation with inrush decay

-LA

LB

LLC

Figure 6.44 Star delta transformer

6.12 The PSCADEMTDC Program


6.12.1 Structure of the program
PSCADEMTDC consists of a suite of programs which enable the efficient transient
simulation of a wide variety of power system networks.
EMTDC (electromagnetic transient and d.c.) [33,34] is an implementation of the
EMTP-type method, initially designed for the solution of ac-dc power systems.
It includes models of various physical components contained in functional modules.
These are among others, transformers, transmission lines, switches, surge arresters,
control systems and electric machinery. It is designed to solve any network which
consists of interconnections of resistance, capacitance, inductance, single and multi-
phase coupled circuits, distributed transmission lines and various other elements.
6.12 THE PSCAD/EMTDC PROGRAM 203

Models of transmission lines and transformers have improved substantially


throughout the years. Accurate models were developed to model the distortion of
waveforms in transmission lines and saturation in transformers. The program also
contains a model for a built-in synchronous machine with its mechanical system.
PSCAD (Power Systems Computer Aided Design) is a graphical Unix-based user
interface for EMTDC which can also run on PCs. PSCAD consists of software enabling
the user to enter a circuit graphically, create new custom-components, solve transmis-
sion line and cable parameters, interact with an EMTDC simulation while in progress,
and to process the results of a simulation [35].
The six programs constituting PSCAD are interfaced by a large number of datafiles
which are managed by a program called ‘Filemanager’. This program also provides an
environment within which to call the other five programs and to perform ‘housekeeping’
tasks associated with the Unix system. The starting point for any study with EMTDC
is to create a graphical ‘sketch’ of the circuit to be solved using the Draft program.
Draft provides the user with a canvas area and a selection of component libraries. A
library is a set of component icons any of which can be dragged to the canvas area and
connected to other components by buswork icons. Associated with each component
icon is a form into which component parameters can be entered. The user can create
component icons, the forms to go with them, and FORTRAN code to describe how
the component acts dynamically in a circuit. Qpical components are multi-winding
transformers, six-pulse groups, control blocks, filters, synchronous machines, circuit
breakers, timing logic, etc.
The output from Draft is a set of files that are used by EMTDC. EMTDC is called
from the PSCAD ‘Runtime’ program, which permits interaction with the simulation
while it is in progress. Runtime enables the User to create buttons, slides, dials and
plots connected to variables used as input or output to the simulation. At the end of
simulation, Runtime copies the time evolution of specified variables into data files.
The complete state of the system at the end of simulation can also be copied into a
‘Snapshot’ file, which can then be used as the starting point for future simulations.
The output data files from EMTDC can be plotted and manipulated by the plotting
programs ‘Uniplot’ or ‘Multiplot’. The output files can also be processed by other
packages, such as Matlab or user-written programs, if desired.
All the intermediate files associated with the PSCAD suite are in text format and
can be usefully inspected and edited. As well as compiling a circuit schematic to input
files required by EMTDC, Draft also saves a text file description of the schematic,
which can be readily distributed to other PSCAD Users. A simplified description of
the PSCADEMTDC suite is illustrated in Figure 6.45. Not shown are many batch
files, operating system interface files, setup files, etc.
EMTDC consists of a main program primarily responsible for finding the
network solution at every time step, input and output, and supporting user-defined
component models. The user must supply two FORTRAN source-code subroutines
to EMTDC- ‘dsdynf and ‘ds0ut.f ’. Usually these subroutines are automatically
generated by Draft but can be completely written or edited by hand. At the start
of simulation these subroutines are compiled and linked with the main EMTDC
object code.
Dsdyn is called each time step before the network is solved and provides an oppor-
tunity for user-defined models to access node voltages, branch currents or internal
204 6 ELECTROMAGNETIC TRANSIENTS

Tline/cable

User interaction

MTDC executabl

Figure 6.45 The PSCADEMTDC suite

variables. The versatility of this approach to user-defined component models means


that EMTDC has enjoyed wide success as a research tool. A flow chart for the EMTDC
program, illustrated in Figure 6.46,indicates that the DSOUT subroutine is called after
the network solution. The purpose of the subroutine is to process variables prior to
being written to an output file. Again, the user has responsibility for supplying this
FORTRAN code, usually automatically from Draft.
The external multiple run loop in Figure 6.46 permits automatic optimisation of
system parameters for some specified goal, or the determination of the effect of vari-
ation in system parameters.

6.12.2 PSCADEMTDC Version 3


Version 3 is a recent Windows version of the original PSCAD/EMTDC program, which
was only available for UNIX systems. The main aim of this development was to make
the simulation package available on the PC platform under the Windows operating
system. As well as implementing many algorithmic improvements, this version took a
totally different approach to the interface. As shown in Figure 6.47,instead of using
separate programs, each with its own interface, the functions of Filemanager, Draft
and Runtime were all put into one. The Project Tree window shows the projects and
libraries loaded. Double clicking on the selected item in the Project Tree gives another
window with the drawing canvas, as shown in Figure 6.48.The simulation control is on
the Toolbar and plots are placed on the drafting canvas alongside the drawing (which
formally was the function of Runtime). Messages with the stage of the simulation and
any errors appear on the Message Tree.
A hierarchical approach has been adopted were there is a main page and sub-pages
below. For example Figure 6.48 shows the main page for a fast-transient. Clicking on
6.12 THE PSCADEMTDC PROGRAM 205

2 Start

Multiple run loop Control of multiple run optimisation

Start EMTDC from data file


or snapshot

Main time loop Time =Time + delt


I
4
Solve for history terms

1
I Calculate history-term current
injectionsfor all network components

I Call DSDYN

1
-
User-definedmaster dynamic file

I Interpolation
J Interpolationalgorithm, switching
procedures and chatter removal

1
Call DSOUT User-definedoutput-definitionfile

1
Write a snapshot if required Store data for further processing

1
Runtime communication Bidirectionalsocket communication
to GUI

1
I Write output files
I
Generation of output files for future
plotting

3
L
Is run finished?

Any multiple runs?

1 ComDleted

Figure 6.46 EMTDC flow chart


206 6 ELECTROMAGNETIC TRANSIENTS

Figure 6.47 PSCADEMTDC version 3 graphical interface

Figure 6.48 Example of Main Page


6.12 THE PSCADEMTDC PROGRAM 207

Figure 6.49 Example of sub-page

the box representing the surge arrester opens the sub-page with the details of the surge
arrester, as shown in Figure 6.49.
Figure 6.50 shows the integration of Runtime and Draft by allowing plots to be
placed alongside the circuit being simulated. An alternative approach is to place all
the plots and controls on a sub-page, as shown in Figure 6.51.
Finally, the new version of PSCADEMTDC takes advantage of advances made in
computer languages for use in dynamic memory allocation.

6.12.3 PSCAD/EMTDC test cases


This section contains a selection of test cases of different time frames. Although simple,
these circuits are chosen to illustrate the main problems involved in digital transient
simulation and demonstrate the broad capability of the ElectromagneticTransient Simu-
lation programs.

(i) GTO test system A circuit particularly difficult to solve is that of Figure 6.52,
where the diode must turn ON simultaneously as the GTO is switched OFF [36].In
the absence of interpolation and chatter removal Figure 6.53 shows the large numerical
oscillations that occur and mask the behaviour of the circuit. A closer inspection of
the turn-OFF, shown in Figure 6.54,reveals that when the GTO turns OFF,although
the diode becomes forward biased, it cannot turn ON until the next time point in the
simulation. This effectively forces the inductor current to zero, i.e. there is no current
to be transferred to the diode. After switching ON, the diode current becomes negative;
208 6 ELECTROMAGNETIC TRANSIENTS

Figure 6.50 Example of integration of functions

Figure 6.51 Graphs and controls on sub-page


6.12 THE PSCADEMTDC PROGRAM 209

GTO

Diode

Figure 6.52 GTO test system

Figure 6.53 GTO test circuit no chatter removal nor interpolation

at the next simulation point the negative current is detected and the diode is switched
OFF again. Then the voltage across the diode causes it to turn ON, but goes negative
thus causing the current to be negative too
With the addition of chatter removal the simulation results, shown in Figure 6.55,
are slightly more realistic because the diode takes over the current immediately when
the GTO turns OFF.The negative spike in the load voltage is not shown completely
in the Figure 6.55, as it reaches a value of -890 V and would, therefore, completely
mask the underlying waveform if the y-axis range were altered to accommodate this.
210 6 ELECTROMAGNETIC TRANSIENTS

Figure 6.54 GTO test circuit with no chatter removal nor interpolation

Figure 6.55 With chatter removal


6.12 THE PSCADEMTDC PROGRAM 211

Figure 6.56 GTO test circuit simulation with interpolation and chatter removal

Finally, with interpolation and chatter removal the ideal waveforms shown in
Figure 6.56 are obtained.

(ii) Frequency-dependent transmission line The system of Figure 6.57, which is


part of the Bonneville Power Administration (BPA) network, is chosen to show the
need for accurate frequency-dependent modelling of transmission lines. The simulation
results as well as field test measurements are published in a paper by Meyer and
Dommel [37]. A fault is applied to Phase C (Fault resistance = 2.0 52) at t = 0.0143
seconds and removed 0.06 s later; a ground resistivity of 100 S2.m is used for the
transmission line parameter calculations.
Figures 6.58 and 6.60 show the receiving end voltage and current, using Bergeron’s
line model. The results of Figures 6.59 and 6.61, displaying the receiving end voltage
and current when a frequency-dependent transmission line model is used, are more
realistic as they show closer agreement to the field test waveforms.

(iii) Fast transient example [38] The main area of fast transient analysis is insu-
lation co-ordination. A typical example is the calculation of the rating of metal oxide
surge arresters (MOVs) to withstand lightning strikes. Figure 6.62 shows an MOV
surge arrester representation suitable for fast transients with time to crest in the range
0.5 to 40 ps. The RlL1 filter represents the effective delay of the impact of the
transient on the ZnO material. This filter has a very small impedance for slow wave-
fronts, such as switching surges. The surge arrester is normally modelled by a single
current-voltage characteristic in switching studies; however, for fast transients it must
212 6 ELECTROMAGNETIC TRANSIENTS

John Day Lower monumental


Phase A
Phase B
Phase C

Voltage: 500 kV rms L-L


Length : 222.07 km
Owned by Bonneville Power Authority
John Day Dam lo Lower Monumental
Conductor: ACSR 84/19, 1,780 MCM, Chukar,
Phase ac source,
.Diameter = 4.069 cm
.Bundle size = 2 conductors, separation = 45.72 cm
Frequency = 60 Hz, Voltage
set at 482.25 kV rms L-L,
.Dc resistance at 25 degrees Celsius = 0.0317 ohm/km/conductoi
Impedance = 15 Ohms
Phase spacing and configuration: Delta
' Height of outer conductors at tower = 23.31 metre
. Height of top conductor at tower = 31.7 metre
. Phase spacing of outer conductors = 12.2 metre
. Sag = 10 metre
Shield Wires: No. = 2
. Radius = 0.635 cm (0.00635 metre)
' Dc Resistance at 25 degrees Celsius = 1.49 ohm/km
' Height above lowest conductor at tower = 13.87 metre
. Spacing between shield wires = 7.87 metre
. Sag = 5 metre

Figure 6.57 Transmission line

Phase A

0 0.02 0.04 0.06 0.08 0.1 0.12

Phase A
I I I I I

0 0.02 0.04 0.06 0.08 0.1 0.12

Phase A
I

I I I I

0 0.02 0.04 0.06 0.08 0.1 0.12


Time (Seconds)

Figure 6.58 Line voltage using Bergeron line model


6.12 THE PSCADEMTDC PROGRAM 213

Phase A

0 0.02 0.04 0.06 0.08 0.1 0.12

Phase A
I I 1

0 0.02 0.04 0.06 0.08 0.1 0.12

Phase A

Time (Seconds)

Figure 6.59 Line voltage using frequency-dependent line model

Phase A
- 1
d
E
g O
5 -1
2

1
3
v

E O
p!
L

5 -1
0 0.02 0.04 0.06 0.08 0.1 0.12

Phase C

0 0.02 0.04 0.06 0.08 0.1 0.12


Time (Seconds)

Figure 6.60 Line current using Bergeron line model


214 6 ELECTROMAGNETIC TRANSIENTS

Phase A

0 0.02 0.04 0.06 0.08 0.1 0.12

Phase 6

0 0.02 0.04 0.06 0.08 0.1 0.12

Phase C

0 0.02 0.04 0.06 0.08 0.1 0.12


Time (Seconds)

Figure 6.61 Line current using frequency-dependent line model

0
LO

Figure 6.62 Metal oxide surge arrester model for fast transient simulations

be split into two, i.e. A0 and A l , such that for switching studies the parallel combina-
tion of A0 and A1 gives the correct current-voltage characteristic. As the impedance
of the intervening filter is negligible at these rates of change, represents the stray
inductance in and around the arrester due to leads and C the terminal to terminal
capacitance of the arrester.
To illustrate the use of this model in a fast transient study, consider a distribution
transformer subjected to a steep wave-front surge, as depicted in Figure 6.63. The
simulation step-length is 1 ns and the simulation duration 10 ps. Because the time
period of interest is less than the travel time of the transmission line the line is repre-
sented by its surge impedance. As the LV winding and neutral are strongly coupled,
being twisted together, they are combined in this simple model. Figure 6.64 shows
the current-voltage characteristics of this surge arrester model. The current surge is
6.12 THE PSCADEMTDC PROGRAM 215

Figure 6.63 Distribution transformer

0.5 '
10-8 10-6 1 o-4 1o-2 1 oo
I
1 o2
Current (Amps)

Figure 6.64 Voltage-current relationship for the two non-linear sections of the surge arrester
model

represented by the function i(t) = 50(exp -2.00 x 10-4t) - exp(-1.666 x lo-%) and
is displayed in Figure 6.65.
Finally, transmission line models are used to represent the service cable and ground.
The PSCADEMTDC simulation results are displayed in Figure 6.66.

(iv) Statcom The test circuit of Figure 6.67 is used to demonstrate the response of
a simple STATCOM to a disturbance. The STATCOM is voltage controlled with a
simple PI controller. Figures 6.68 and 6.69 display the transient response following a
three-phase to ground fault and a fault impedance of 75 52 is used to represent a fault
in the system remote from the STATCOM installation.
The terminal voltage drop caused by the fault is translated by the PI controller into
a new alpha order which increases the reactive power delivered by the STATCOM
(Figure 6.68).
216 6 ELECTROMAGNETIC TRANSIENTS

50 I 1

45

40

-
v)

?
35
g 30
E
25

10

0' I
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (Seconds) x 10-~

Figure 6.65 Surge wave-front

1 .. -8 .- - % . - _ I . . . I - . - I_ . . '_ . . ! . . .#. .

011 012 0:3 0:4 0:5 0:6 0:7 018 0:9 1

-2
h

1001
-5
1
_ _ . _ . _ I . _ . _ _ _ _ _ _ - _ _ . . . ~

1 - 0 0d0 4
Y

2000 - - . ..- . . ~. - _ I _ . . . . - . - . . - . . - . . . .
0
2
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (Seconds) x 10-~

Figure 6.66 Simulation results


6.12 THE PSCADEMTDC PROGRAM 217

-
- Example case:
A simple &pulse STATCOM

75ohms with X I R
ratio equal to 1 .
Occurs a! I .5 sec

Figure 6.67 STATCOM test case

Real Power

zE 5 0 0 1
$ 0
g
a
-50
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

Reactive Power
2 400

-200 ~ . _ _ ~ . . _ _ . . . . . . _ . _ . _ _ _ . _ _ _ _
ti
g
n o
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

(Y Order

. _ . _ _ _ . _ _-_ _ _ _ _ . _ _ _ _

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6


Time (Seconds)

Figure 6.68 Power and firing order for STATCOM


218 6 ELECTROMAGNETIC TRANSIENTS

-
Y2
a
1.5

1
-B
0)

2
Y

0-L . _
-B&0

-> 801
DC Voltage
I

20
0
’ 0.2 0.4 0.6 0.8 1 1.2 1.4
I
1.6
Time (Seconds)

Figure 6.69 Voltage for STATCOM test system

There is an inevitable overshoot and damped oscillation before reaching the new
steady-state level. The alpha order is initially increased and then settles back at the
original steady-state value, after the oscillations disappear. The new steady-state d.c.
busbar voltage, shown in Figure 6.69, is higher, giving larger reactive power for a
given alpha order. The real power undergoes a transient, as energy is required to raise
the capacitor voltage, but under steady-state conditions the real average power is zero
(ignoring losses). The proportional and integral gain were not optimized in this study
and hence they can be modified, and the system re-simulated, to obtain a better transient
response. Important factors are the time required to reach the new steady state and the
size of the overshoot while settling to the new steady state.

(v) Sub-synchronous resonance Sub-synchronous resonance occurs when the


mechanical resonant frequency of the shaft and generator masses coincides with the
fundamental frequency minus the electrical resonant frequency of the capacitor in
series with the total system impedance. This causes mechanical oscillations, which can
damage the turbine shaft.
The IEEE First Benchmark case used for Sub-Synchronous Resonance Studies
[39] is depicted in Figure 6.70. The simulation results for this case, displayed in
Figures 6.71 and 6.72, show that, even though the oscillations in the electrical torque
are dying away, the oscillations in the torque between machine and generator exciter
are increasing. The resonance can, in this case, result in catastrophic failure of the
turbine shaft.
6.I3 REAL TIME DIGITAL SIMULATION [40-43 J 219

Generator with
multi-mass
modelled
(>
compensation
(T)
Equivalent for the rest
of the AC system

m
Transmission
Line I-[@ Fault at 1.5

-
L
-
Figure 6.70 Sub-synchronous resonance test system

Machine Current A Out


51 I

-Y
1.4 1.5 1.6 1.7 1.8 1.9 2

Machine Voltage Phase-A


I I I 1 I I

S 1.4
’ 1.5 1.6 1.7 1.8 1.9 2

Capacitor Voltage Phase-A

- 1.4 1.5 1.6 1.7 1.8 1.9 2


Time (Seconds)

Figure 6.71 Current and voltage waveforms

6.13 Real Time Digital Simulation [40-431


The Real Time Digital Simulator (RTDS) is essentially a parallel processor implemen-
tation of the EMTDC program (although quite different in detail). Through careful
coding, the EMTDC algorithm and component models have been distributed over
many processors running in parallel, so that the simulation can proceed in real time for
time steps typically in the 50-75 ps range. The graphical user interface to RTDS is the
PSCAD suite, but using RTDS component libraries. As in EMTDC, the User can create
new control components, in this case written in a C-type language. The size of system
able to be simulated with the RTDS depends upon the size of the particular RTDS. The
user must also pay more regard to the use of subsystems because of inter-processor
220 6 ELECTROMAGNETICTRANSIENTS

Machine Generator Exciter Torque

-2 '
1.4 1.5 1.6 1.7 1.8 1.9
I
2

LPA to LPB Torque

-2 '
1.4 1.5 1.6 1.7 1.8 1.9
I
2

-Electrical Torque
1 I I I I I

1.4 1.5 1.6 1.7 1.8 1.9 2


Time (Seconds)

Figure 6.72 Torque in generator

communication restraints, and non-linear scaling constraints. The RTDS also provides
many analogue inputloutput ports, for direct connection via power amplifiers to control,
protection and instrumentation circuitry.
Physically, the RTDS is organized as one or more cubicles, each containing several
racks. Each rack consists of 18 Tandem Processor Cards (TPS), one Workstation Inter-
face Card (WIC) and one Inter-rack Communication Card (IRC). Each TPC contains
two NEC 77240 DSPS. A new card, the 3PC, based on the SHAC chip (Super Harvard
Architecture) is also available with a limited number of components running on it. This
card uses C language and can be plugged into an existing rack. It is approximately six
times more powerful than the TPC card and can thus run more nodes per rack.
The user interacts with the RTDS primarily by means of PSCAD software running
on a Unix based workstation, which is connected to the RTDS by an ethernet LAN.
The WIC on each rack supports communication for all TPCs on that rack with the
workstation. This includes the downloading of software to each TPC prior to a run,
instructions from the user to modify a simulation parameter during a run, and transfer
of data produced by the simulation back to the workstation for monitoring.
The IRC permits communication between racks by high speed coaxial connection.
Up to four other racks can be connected to a given rack by means of the IRC. Each
processor on a TPC is capable of simulating a single electrical node, a component
model, or part of a component model. Since the computational effort associated with
solving the nodal voltages in a subsystem at each time step scales with the square of
the number of nodes, the largest possible subsystem is fixed at that which can fit on a
single rack, i.e. 2 x 18 = 36 nodes. In practice, the largest subsystem is smaller than
this since additional processors are required to model sources and components. The
RTDS therefore makes extensive use of subsysteming for large networks. Generally,
6.14 STATE VARIABLE ANALYSIS 221

this is facilitated by the presence of transmission lines and cables and, in the case of
HVDC systems, the presence of voltage stable nodes on the a x . side due to filters, and
a current stable branch on the d.c. side due to the smoothing inductor. In many cases,
nodes with only RLC branches can be effectively eliminated if it is not necessary to
monitor them. Such is the case with several types of tuned filters which require no
extra processors, apart from the one at the node to which they are connected.
Component models in the RTDS mirror those in EMTDC but are coded in assembly
language for additional speed and efficiency. For HVDC modelling, the converter trans-
former and six pulse group are lumped into a single component. Firing pulses may also
be generated externally by real circuitry if desired. Each six-pulse group and associated
transformer require two processors, at least one additional processor is required for the
link controller. Thyristors are modelled as on/off resistances, but will reconduct if a
forward voltage is applied too soon after the end of conduction. The bridge model,
by assumption of the presence of a smoothing inductor on the d.c. side, is configured
so that the a.c. and d.c. sides may be in different subsystems. A separate 12-pulse
unipolar back-to-back link model is provided for cases with no smoothing inductor.
The converter transformer model includes saturation, hysteresis and tap changing. Only
wye/delta and wye/wye connections are supported. Converter control is either by a
generic HVDC control or by construction from a library of control components.
A useful feature of real time simulation is that many runs can be completed in a short
time, thereby covering a variety of operating conditions, contingencies and parameter
values. To facilitate this process, the RTDS provides a script language which can be
used to describe a sequence of simulation commands, output processing, and circuit
modifications which would normally be carried out by the user via PSCAD.

6.14 State Variable Analysis


State variables are those that completely define the state of the system. State variable
analysis was the dominant technique for transient simulation prior to the appearance
of Dommel’s numerical integration substitution. Due to their suitability to model non-
linearities, stable variables are still used for the modelling of some components in the
Electromagnetic Transient Programs. In particular they are suited to the modelling of
frequent switching, as they require no overhead to change the step length. Their main
disadvantages are solution speed and code complexity; also the modelling of compo-
nents with distributed parameters becomes more difficult in the state variable analysis.

6.14.1 State variable formulation


A dynamic system of order n can be represented by a system of n first-order linear
differential equations, i.e.:
il(t)= + + bllul(t)+ b l 2 ~ 2 ( t+) * . . b l , u , ( t ) ,
~ I I X I ( +~ U) I ~ X : ! ( ~ ) * . . a l , ~ , ( t )

i2(t) +
= a21X1( t ) + a22~2(t) * ‘ + b 2 1 ~(1t ) + b22~2(t)+ b 2 n 1 ~ m ( t ) ,
a2n~n(t) * * *
222 6 ELECTROMAGNETIC TRANSIENTS

Or in matrix form (ignoring the parameter t for simplicity):

'i.
Lm

In compact matrix notation Equations (6.94) and (6.96), which make up the state
variable formulation of a problem, are:
(6.96)

X = [A]x + [B]u, (6.97)


y = [CIX + [Dlu. (6.98)
This is the normal or standard form of the state variable equation; the first equation
is the state equation and the second the output equation. If there is no direct connection
between the inputs and output vector then [D]is zero. These equations can be solved
by transform methods, convolution integral, or numerically in an iterative procedure.
The form of this equation is not unique and depends on the choice of state variables.

6.14.2 Solution procedure


Figure 6.73 illustrates the standard state variable computer implementation. At each
step the state variable equations are solved using an iterative procedure, the most
common being the predictor-corrector formulation. An implicit integration method,
such as the trapezoidal rule is used to calculate the state variables at time t ; however
this method requires the value of the state variable derivatives at time t . Although
initially its value is equated to that of the previous time-step, once an estimate of the
6.14 STATE VARIABLE ANALYSIS 223

Initial Estimate of Slate


Variab!e Derivatives

Integrate

4
Use Slate Equation lo obtain
State Variable Derivatives

-
A, = [ A 1 5 +[mu,

Yes
Calculate Dependent Variables

No lf
= IClZf +IDl_u,

Increase time-step

Advance Time

1
Figure 6.73 State variable solution technique

state variables has been obtained then the state equation is used to update the estimate
of the state variable derivatives.
If convergence is not reached then the step size is halved and the procedure restarted.
The trapezoidal rule is A-stable, so that given a small enough step size it is guaranteed
to converge; however, the step size may get too small for practical simulation. Once
convergence is reached, the dependent variables are calculated from Equation (6.98).
The elements of matrices [A], [B], [ C ] and [D]are dependent on the R, L and C
values of the network, but not on the step length. This allows the step length to be
varied so as to coincide with the switching instants, thereby eliminating the problem
of numerical oscillations due to switching errors. Moreover, an optimizing algorithm
can be used to modify the nominal step length. The number of iterations is examined
and the step size increased or decreased by 10% based on whether the number of
224 6 ELECTROMAGNETIC TRANSIENTS

iterations is too small or too large. An alternative widely used approach is to apply
an integration formula directly to the state and output equations. This is just another
form of numerical integrator substitution and leads to the same difference equation as
the numerical integrator substitution method. In other words, regardless of the way the
differential equations are rearranged and manipulated, the end result is only a function
of whether a numerical integration formula is substituted in or an iterative solution
procedure is adopted. The term state variable analysis refers to the latter approach.

6.14.3 Choice of state variables


Although inductor current and capacitor voltage are the state variables chosen in most
textbooks it is better to use flux linkage (Q) of the inductor and capacitor charge (Q).
Regardless of the numerical integration algorithm used in solving the differential equa-
tions, this will result in a better solution caused by error propagation due to the presence
of local truncation errors. Moreover non-linearities in q-v or Q-1 characteristics can
be modelled more easily [44].
The state variable solution requires the number of state variables to be equal to the
number of independent energy storage elements. When this does not occur then any
initialization error will remain throughout the simulation. This in turn can alter the
simulation dramatically, as point-on-wave switching devices will switch at different
times. It is therefore important to recognize when capacitor charge (or voltage) or
inductor flux (or current) in the network is dependent or independent. This dependence
occurs when there is a cutset of inductors and current sources or a loop of capacitors
and voltage sources [45,46]. A cutset is a set of branches such that their removal results
in two disconnected sub-graphs, and following removal of the branches forming the
cutset, restoration of any one branch will create a connected graph again. If the current
in one inductor is a linear combination of the currents in the other inductors and current
sources then the inductor is dependent. This is not always obvious, due to the presence
of a large intervening network; this problem is illustrated in Figure 6.74, where it is not
immediately apparent that inductors 3, 4 3 and 6 with the current source form a cutset.

Figure 6.74 Dependent inductor


6.15 REFERENCES 225

6.15 References

1. Dommel, H W, (1969). Digital computer solution of electromagnetic transients in single-


and multiphase networks, IEEE Transactions On Power Apparatus and Systems, PAS-88 No
(4), pp. 388-399.
2. Dommel, H W and Meyer, W S, (1974). Computation of electromagnetic transients,
Proceeding of the IEEE, 62, (7) pp. 983-993.
3. IEEE PES Working Group 15.08.09, (1999). Modeling and Analysis of System Transients,
IEEE PES Special Publication TP133-0.
4. Martinez-Velasco (Ed.) (1 999). Computer Analysis of Electric Power System Transients,
Selected Readings IEEE, New York.
5. Bergeron, L, (1949). Du Coup de Belier en Hydraulique au Coup de Foudre en Electricite
Dunod, Paris (Translated version: Wiley, New York, 1961).
6. Heydt, G T, (1991). Electric Power Quality, Stars in a Circle Publications West Lafayette,
USA.
7. Carson, J R, (1926). Wave propagation in overhead wires with ground return, Bell Systems
Technical Journal, No. 5, pp. 539-554.
8. Manitoba HVDC Research Centre, EMTDC User’s Manual.
9. Gustavsen, B and Semlyen, A, (1997). Simulation of transmission line transients using
vector fitting and modal decomposition, Paper 97WM-347.
10. Angelidis, G and Semlyen, A, (1995). Direct phase-domain calculation of transmission line
transients using two-sided recursions, IEEE Transactions on Power Delivery, 10 (2), pp.
94 1-947.
11. Noda, T, Nagaoka, N and Ametani, A, (1996). Phase domain modeling of frequency-
dependent transmission lines by means of an ARMA model, IEEE Transactions on Power
Delivery, 11 (1). pp. 401-411.
12. Noda, T, Nagaoka, N and Ametani, A, (1996). Further improvements to a phase-domain
ARMA line model in terms of convolution, steady-state initialization, and stability, Paper
96 SM 457-2 PWRD.
13. Nguyen, H V, Dommel, H W and Marti, J R, (1996). Direct phase-modelling of frequency-
dependent overhead transmission lines, Paper SM 458-0 PWRD 96.
14. Gustavsen, B, Sletbak, J and Henriksen, T, (1995). Calculation of electromagnetic transients
in transmission cables and lines taking frequency dependent effects accurately into account,
IEEE Transactions on Power Delivery, 10 (2), pp. 1076-1084.
15. Marcano, F J and Marti, J R, (1997). Idempotent line model: case studies, Proceedings
of International Conference on Power System Transients, IPST’97, Seattle, June 22-26,
pp. 67-72.
16. Wedepohl, L M, (1989). Theory of natural modes in multi-conductor transmission lines,
Course Note.
17. Wedepohl, L M, Hguyen, H V and Irwin, G D, (1996). Frequency-dependent transforma-
tion matrices for untransposed transmission lines using Newton-Raphson method, IEEE
Transactions on Power Systems, 11 (3), pp 1538-1545.
18. Marti, L, (1982). Accurate modelling of frequency-dependent transmission lines in elec-
tromagnetic transient simulations, IEEE Transactions on Power Apparatus and Systems,
PAS-101 (l), pp. 147-157.
19. Marti. L, (1983). Low-order approximation of transmission line parameters for frequency-
dependent models, IEEE Transactions on Power Apparatus and Systems, 102 (1 1) pp.
3582-3589.
20. Semlyen, A, (1981). Contributions to the theory of calculation of electromagnetic transients
on transmission lines with frequency dependent parameters, IEEE Transactions on Power
Apparatus and Systems, PAS-100 No (2), pp. 848-856.
226 6 ELECTROMAGNETIC TRANSIENTS

21. Semlyen, A and Dabuleanu, A, (1975). Fast and accurate switching calculations on trans-
mission lines with ground return using recursive Convolutions, IEEE Transactions on Power
Apparatus and Systems, 94 NO (2) pp. 56 1-57 1.
22. Kuffel, P, Kent, K and Irwin, G D, (1995). Implementation and effectiveness of linear
interpolation within digital simulation, Proceedings of International Conference on Power
System Transients, IPST’95, Lisbon, Portugal, 3-5 September pp. 499-504.
23. Gole, A M, Fernando, I T, Irwin, G D, and Nayak, 0 B, (1997). Modeling of power elec-
tronic apparatus: additional interpolation issues, Proceedings of International conference on
power system transients, IPST’97, Seattle, WA, 22-26 June, pp. 23-28.
24. Dommel, H W, Bhattacharya, S, Brandwajn, V, Lauw, H K and Marti, L, (1986). EMTP
Theory Book, Vols 1 and 2, Prepared for Bonneville Power Administration.
25. Alvarado, F L, Lasseter, R H and Sanchez, J J, (1983). Testing of trapezoidal integration
with damping for the solution of power transient problems, IEEE Transactions on Power
Apparatus and Systems, 102 (12), pp. 3783-3790.
26. Lin, J and Marti, J R, (1990). Implementation of the CDA procedure in EMTP, IEEE Trans-
actions on Power Systems, 5 (2) pp. 394-402.
27. Marti, J R and Lin, J, (1989). Suppression of numerical oscillations in the EMTP, IEEE
Transactions on Power Systems, 4 (2), pp. 739-747.
28. Woodford, D and Irwin, G D, (2000). Personal Private communication.
29. Watson, N R and Irwin, G D, (1999). Accurate and stable electromagnetic transient simula-
tion using root-matching techniques, International Journal of Electrical Power and Energy
Systems, 21, (3) pp. 225-234.
30. Watson, N R, and Irwin, G D, (1998). Electromagnetic transient simulation of power
systems using root-matching techniques, Proceedings of the IEE, Part C, 145, (3,
pp. 481-486.
31. CIGRE WG14-02, (1994). A summary of the report survey of control and control perfor-
mance in HVDC schemes, Electra, No. 155, pp. 65-88.
32. Gole A, Lecture Notes on Electromagnetic Transients.
33. Woodford, D A, Ino, T, Mathur, R M, Gole, A and Menzies, R W, (1985). Validation of
digital simulation of HVDC transmission by field tests, IEE Con5 Publication on AC and
DC Power Transmission, No. 255, pp. 377-38 1.
34. Woodford, D A, Gole, A and Menzies, R W, (1983). Digital simulation of DC links
and AC machines, IEEE Transactions on Power Apparatus and Systems, PAS-102, (6),
pp. 1616-1623.
35. Manitoba HVDC Research Centre, (1994). PSCADEMTDC power systems simulation soft-
ware tutorial manual.
36. Woodford, D and Irwin, G D, (2000). Benchmark system, personal Communication.
37. Meyer, W S and Dommel, H W, (1974). Numerical modeling of frequency-dependent
transmission line parameters in an electromagnetic transient program, IEEE Transactions
on Power Apparatus and Systems, PAS-93 ( 9 , pp. 1401- 1409.
38. Manitoba HVDC Research Centre, (2000). PSCADEMTDC Version 3 Program Example.
39. IEEE, (1997). IEEE first benchmark case for sub-synchronous resonance studies, IEEE
Transactions on Power Apparatus and Systems, PAS-96 Vol ( 3 , pp. 1565- 1572.
40. Kuffel, R, Giesbrecht, Maguire, T, Wierckx, R P and McLaren, P, (1995) . RTDS -A full
digital power system simulator operating in real time, ICDS-95, USA, April p. 19.
41. Kuffel, R, McLaren, P, Yalla, M and Wang, X, (1995). Testing of the Beckwith electric M-
0430 multifunction protection relay using a real-time digital simulator (RTDS), ICDS-95,
USA, p. 49.
42. Duchen, H, Lagerkvist, Kuffel. R and Wierckx, R P, (1995). HVDC simulation and control
system testing using a real-time digital simulator (RTDS), ICDS-95, USA, p. 213.
6.15 REFERENCES 227

43. Kuffel, R, Wierckx, R P, Forsyth, P, Duchen, H, Lagerkvist and Wang, X, (1995).


Expanding an analogue HVDC simulator’s modelling using a real-time digital simulator
(RTDS), ICDS-95, USA, p. 199.
44. Chua, L 0, and Lin, P M, (1975). Computer Aided Analysis of Electronic Circuits: Algo-
rithms and Computational Techniques, Prentice-Hall, Englewood Chiffs, NJ.
45. Rajagopalan, V, (1987). Computer-Aided Analysis of Power Electronic Systems, Marcel
Dekker, New York.
46. Rohrer, R A, (1970). Circuit Theory: An Introduction to the State Variable Approach,
McGraw-Hill New York.
Previous Page

7.8 STRUCTURE OF A TRANSIENT STABILITY PROGRAM 263

7.8 Structure of a Transient Stability Program


7.8.1 Overall structure
An overview of the structure of a transient stability program is given in Figure 7.19.
Only the main parts of the program have been included and, as can be seen, the
same system may have several case studies performed on it by repeatedly specifying
switching data, When no further switching data is available, control returns to the start
to see if another system is to be studied.
With care, the program can be divided into packages of subroutines each concerned
with only one aspect of the system [13]. This permits the removal of component models
when not required and the easy addition of new models whenever necessary. Thus, for
example, the subroutines associated with the synchronous machine, the AVRs, speed
governors, etc, can be segregated from the network. Figure 7.20 shows a more detailed
block diagram of the overall structure where this segregation is indicated. The diagram
is subdivided into the five sections indicated in Figure 7.19.
While the block diagrams are intended to be self-evident several logic codes need
to be explained. These are:
KASE-This is the case study number for a particular system. It is initially set
to zero and incremented by 1 at the end of the initialization and at the end of
each case study.
KBIFAl -The sparse factored inverse of the nodal network matrix is obtained
using three bifactorization subroutines. The first and second subroutines are integer
routines, which determine busbar ordering, and non-zero element location. The
code KBIFAl is set to unity if it is necessary to enter these two subroutines,
otherwise it is set to zero.
KBIFA3 -The elements of the sparse-factored inverse are evaluated in the third
bifactorization subroutine. The code KBIFA3 is set to unity if it is necessary
to enter this subroutine, otherwise it is set to zero. When KBIFA3 is unity, it
indicates that a network discontinuity has occurred and, hence, it is also used for
this purpose.
TIME -The integration time.
H-The integration step length. Like KBIFA3, it is also used to indicate a discon-
tinuity when it is set to zero.
PRINTTIME-The integration time at which the next print out of results is
required.
MAXTIME -The predefined maximum integration time for the case study.
ITMAX -Maximum number of iterations per step since last printout of results.
Note that many data error checks are required in a program of this type but they have
been omitted from the block diagram for clarity.
264 7 SYSTEM STABILITY

7.8.2 Structure of machine and network iterative solution


The structure of this part of the program requires further description. Two forms of
solution are possible depending on whether an integration step is being evaluated or the
non-integrable variables are being recalculated after a discontinuity. A block diagram
is given in Figure 7.21.
The additional logic codes used in this part of the program are:
ERROR- The maximum difference between any integrable variable from one
iteration to another.
ITR -Number of iterations required for solution.
IHALF -Number of immediate step halving required for the solution
TOLERANCE -Specified maximum value of ERROR for convergence.
If convergence has not been achieved after a specified number of iterations, the
case study is terminated. This is done by setting the integration time equal to the
maximum integration time. The latest results are thus printed out and a new case study
is attempted.

Network calculations Synchronous AVR calculations Speed governor


machine calculations
calculations
.I .. ..

KBlFAl = 0

data input data input

(a) Section 1

Figure 7.20 Structure of transient stability program. (a) Section 1. (b) Section 2. (c) Section 3.
(d) Section 4. (e) Section 5
7.8 STRUCTURE OF A TRANSIENT STABILITY PROGRAM 265

order and nonzero


element location
~ for bifactorization

I
KBIFA3 = 1
-
I
KASE>i

at TIME = O?

initial conditions
conditions

S initial, ,
conditions

Perform

- numerical part
,of bifactorization ,

9-1 S
E
, =

(b) Section 2

Figure 7.20 (Continued)


266 7 SYSTEM STABILITY

B T

switching ops.
if branch change
set KBIF3 = 1

rder and non zer


lement location
or bifactorization

No

numerical part

+ I

F
l KBIFA3 = I
I
1
t-
(c) Section 3

Figure 7.20 (Continued)


7.8 STRUCTURE OF A TRANSIENT STABILITY PROGRAM 261

J
For further
Solve for details of
machines this part
Solve for network

o? BIFA3 = 1

+I
TIME =TIME + H

Print-out Print-out
AVR results
busbar and sync. machinl
branch results results

+
if required
-m
speed gov.
results if

ITMAX13

Set flag for step

(d) Section 4

Figure 7.20 (Continued)


268 7 SYSTEM STABILITY

P I h

t
'End of case'
I

olKase =l?

(steady-state)

Reset initial
(steady-state)
conditions
(e) Section 5

Figure 7.20 (Continued)

7.9 Advanced Component Models


7.9.1 Synchronous machine saturation
The relationship between mutual flux and the exciting MMF within a machine is not
linear and some means of representing this non-linearity is necessary if the results
obtained from a stability study are to be accurate.
In most multimachine stability programs, each machine is represented by a voltage
behind an impedance, the latter consisting of armature resistance plus either transient
or subtransient reactance. Also, the voltage magnitude may be fixed or time varying,
depending on the complexity of the model.
7.9 ADVANCED COMPONENT MODELS 269

Saturation may thus be taken into account by modifying the value of the reactance
used in representing the machines. However, as explained for the model developed in
section 7.6, it is more convenient to fix the reactance and adjust the voltage accordingly.
Saturation is a part of synchronous machine modelling where there is still uncertainty
as to the best method of simulation. The degree of saturation is not the same throughout
the machine because the flux varies by the amount of leakage flux. Also, the saturations
in the direct and quadrature axes are different, although this difference is small in the
case of a cylindrical rotor.
Various methods have been adopted to account for saturation which differ not only
in the model modification technique but also in the representation of the saturation
characteristic of the machine.

Y
Start solution

Store sufficient data ame for each AV


in case H is reduced
immediately
I

Calculate constants ame for each AV


C and M for
trapezoidal method
.I
:
Same for each
speed pov.

nonintegrable
--------

variables using
algebraic form of

ITR = 0

1
(a) Section 1

Figure 7.21 Structure of machine and network iterative solution. (a) Section 1. (b) Section 2.
(c) Section 3
270 7 SYSTEM STABILITY

4 1 ITR=ITR+l
I

I Same for eachi


nonlinear injected into network
for each sync.
machine

voltages from
bifactored

machine
non-integrable
ame for each AV
-------
- -------
Evaluate integrable
variables using
algebraic form of
trapezoidal method speed gov.
also determine
ERROR

ro No
Set flag for step

Exit
solution 7
W
(b) Section 2

Figure 7.21 (Continued)


7.9 ADVANCED COMPONENT MODELS 27 1

Re-evaluate
conditions
at beginning of

Lr''Not coverging'

Lr'
0
Time = maxtime

Exit

(c) Section 3

Figure 7.21 (Continued)

Classical saturation model Classical theory [ 141 for a cylindrical rotor machine
assumes that the saturation is due to the total MMF produced in the iron and is the
same in each axis.
It is necessary to make further assumptions in order to simplify the model as follows:

(1) The magnetic reluctances in both axes are equal. Thus, the synchronous reactances
are equal, i.e. Xd = X,.
(2) Saturation does not distort the sinusoidal variations assumed for rotor and stator
inductances.
(3) Because load-test data is not usually available, saturation is determined using the
open-circuit saturation curve.
272 7 SYSTEM STABILITY

(4) Potier reactance X, may be used in calculating saturation.


( 5 ) The total iron MMF (Fe) may be determined from
-
Fe = S r f i (7.100)
where S is the saturation factor, defined as:
iron MMF
S=l+ (7.101)
air gap MMF '
( 6 ) With reference to Figure 7.22, the saturation factor S may be determined from:

s = -AC (7.102)
AB '
Figure 7.23 shows a typical voltage and MMF diagram for a round rotor synchronous
machine. The Potier voltage Ep, the voltage behind the Potier reactance, may be deter-
mined from the terminal voltage V t and the terminal current I. The MMF required to
produce this voltage is found from the open circuit saturation curve of Figure 7.22.
Armature reactance F is found using Assumption (4) from which the field MMF (Fe)
is calculated. The voltage equivalent to Ff referred to the stator is E f. It is readily
apparent that rotating the MMF diagram through 90" gives:
Ff c( E i (in the steady state),
(7.103)

A
V

Field current (MMF)

Figure 7.22 Open circuit saturation characteristic of a synchronous machine


7.9 ADVANCED COMPONENT MODELS 273

Figure 7.23 Vector diagram of cylindrical rotor synchronous machine saturation

In Figure 7.23, the reactance x d s is the saturated value of x d . This produces an


internal machine voltage E, which lies on the quadrature axis. As I ( x d - X p ) is parallel
with I. X d s , then
-EP
- - I(Xds - x p > = -Eq (7.104)
S Ep I(Xd - X p ) Ei ’
and from this
Ei = S Eq 9 (7.105)
and
xds = xd -S xLJ +xp. (7.106)

All machine reactances subject to saturation are similarly modified.

Salient machine saturation In the case of a salient synchronous machine, it may be


assumed that the direct and quadrature axis armature reaction MMFs (Fd and F,, respec-
tively) are proportional to the reactive voltage drops I d X a d and I , . Xu,, respectively.
Assumptions 3 and 4 for the classical model also apply.
There are many different methods of accounting for the saturation effect. The
methods considered here assume that saturation in the d-axis is due, at least in part, to
the component of flux in the d-axis. The first method ignores saturation in the q-axis,
the second method accounts for quadrature axis saturation by the component of flux in
the q-axis, the third methods considers that the total flux contributes to the saturation
in both axes.
In the first method, the density of the flux due to the quadrature axis armature
reactance MMF (F,) is considered sufficiently small that saturation effects on voltages
are thus neglected on the direct axis. The other component of the armature reaction
MMF ( F d ) adds directly to the field MMF (F’ to produce a main flux which, in
turn, produces a quadrature axis voltage subject to saturation. The saturation level is
determined by the quadrature component of Potier voltage (Ep,).
274 7 SYSTEM STABILITY

The second method [15] allows for saturation in both the direct and quadrature axis
components of the Potier voltage. It is assumed that the reluctances of the d-axis and
q-axis paths differ only because of the different air gaps in each axis. The d-axis
component of Potier voltage (Epd) is thus modified by the ratio X q / & before the
q-axis saturation factor is determined. Provided that it is assumed that the vector sum
of the two saturated main flux components (Fe, and Fed) is in phase with the MMF
proportional to Potier voltage, then the saturated d and q axis synchronous reactances
(Xd.7 and Xqs) are:

(7.107)

(7.108)

A third method 1141 distinguishes between the saturation in the rotor and stator,
and saturation factors based on Ep and Ep, are obtained. This method is difficult to
implement because it is necessary to ensure that saturation is not applied twice to any
part of the machine. That is, the saturation in the field poles must be isolated from that
of the armature, giving two saturation curves. The two saturation factors for this case
may be defined as:
iron MMF in the stator
sd9 = 1 + total air gap MMF '
(7.109)

iron MMF in the rotor


sd=1+ (7.110)
direct axis air gap MMF'
where s d g acts equally on both d- and q-axes and Sd acts on the direct axis only.
Figures 7.24, 7.25 and 7.26 demonstrate the differences between the three methods
of saturation representation.

Ff
(aEi)

Figure 7.24 Salient pole synchronous machine with direct axis saturation only
7.9 ADVANCED COMPONENT MODELS 275

El

Figure 7.25 Salient pole synchronous machine with direct and quadrature axis saturation

Ei

Figure 7.26 Salient pole synchronous machine with separate stator and rotor saturation

Simple saturation representation An even simpler method of including the effect


of saturation is to calculate the saturation initially (by some means) after which it
is either held constant, or varied according to the slope of the saturation curve at the
initial point. This method is suitable for small perturbation studies where Potier voltage
and machine angle do not vary greatly.
Saturation curve representation The open-circuit saturation curve must be stored
within the computer so that a new saturation factor can be determined at every stage
of the study.
276 7 SYSTEM STABILITY

The most accurate method of storing this curve is to fit a polynomial of the form:
~f=
CO + C I V + c2v2+ c3v3... + CnV”, (7.1 1 1 )
+
by taking n 1 points on the curve. Normally, n would be 5, 7 or 9. This is a clumsy
method of both entering the data and storing it. In multimachine transient stability
studies, where the machines are represented at best by subtransient parameters and an
approximation to Potier reactance is made, nothing is achieved by such an elaborate
method of representing the saturation curve.
The problem can be simplified by assuming that most of the coefficients of the poly-
nomial are zero. A sufficiently good approximation is achieved with the Equation [ 161:
If= v + CnV”, (7.112)
where n is normally either 7 or 9. Only one point is needed to specify the curve and
if I f is always specified at a predetermined voltage, the data entries required per curve
are reduced to one, from which C, may be readily determined.
Potier reactance The Potier reactance of a machine is rarely quoted, although the
open-circuit saturation curve is normally available. In order to model the saturation
effects, it is thus necessary to estimate this reactance.
From a knowledge of the leakage reactance XI, Beckwith [17] calculated that:
X p = X1+ 0.63(X&- X I ) , (7.1 13)
and if XI is not available, then
X p = O.SX&. (7.1 14)
Equation (7.114) may be modified to account for the type of synchronous machine,
[ 181 i.e.
X p = 0.9X& (7.115)
for a salient-pole machine, as most of the saturation occurs in the poles, and
X p = 0.7X& (7.116)
for a round rotor machine, as most of the saturation occurs in the rotor teeth and the
Potier and leakage reactances have similar values.

The effect of saturation on the synchronous machine model Saturation effec-


tively modifies the ordinary differential equations describing the behaviour of the
voltages used to model the synchronous machine. Equations (7.13, (7.16), (7.19) and
(7.20) become respectively:
P E=
~ (Ef + W d - x&Yd- SdEb)/T&o, (7.117)
pE& = (-(Xq - X b Y q - SqE&)/T&o, (7.1 18)
P E= ~ (SdEb + (Xd - Xi)Id - SdE:)/T&, (7.1 19)
p E : = (S,E& - ( X b - X:)Zq - S,E:)/T:o, (7.120)
where Sd and S, are the direct and quadrature axis saturation factors.
7.9 ADVANCED COMPONENT MODELS 277

Where subtransients are considered then Equations (7.17) and (7.18) are replaced by:

Representation of saturated synchronous machines in the network Repre-


sentation of a salient but unsaturated synchronous machine in the network has been
discussed in section 7.6. When saturation occurs, a double adjustment must be made
at each step in the solution process [lo].
With the notation developed in section 7.6, the fixed admittance (Yo),which is
included with the network, is made up from unsaturated and non-salient values of
reactance, whereas the correct admittance (Y,,)is made up from saturated and salient
values. Using Figure 7.14, as before, the adjusting current to account for this change
in admittance is
Iadjs = (y,, - To)(E”- v). (7.123)

That is:

(7.124)

The current T:iis is similar to Tadj developed in section 7.6:

The current injected into the network is given by Equation (7.52) and as the terms
in the brackets contain no saliency then in the real and imaginary axis of the network:
278 7 SYSTEM STABILITY

where contains saturated but non-salient reactance terms and ?$! contains salient
and saturated reactance terms.

Note that the third part of Equation (7.126) is TO^ and not 70,". This part of the
injected current is merely the current flowing through 70and could be eliminated if 70
were not included in the network. The conditioning of the network would be reduced,
however, and in certain systems this could lead to numerical problems.
Inclusion of synchronous machine saturation in the transient stability program
Only two sub-routines need modification to allow saturation effects in synchronous
machines to be modelled. In both cases, an iterative solution is necessary for each
saturating machine, although in most instances the number of iterations is small.
Saturation is a function of the voltage behind armature resistance and Potier reac-
tance. Assuming the second method of salient machine saturation is being used, then
from Equation (7.112):

(7.129)

where
X
c,, = JCnd (7.130)
Xd

and
(7.131)

and I d and I , are given by Equations (7.121) and (7.122).


A Jacobi iterative technique is quite adequate to establish the initial conditions of the
synchronous machine and this can be incorporated in the relevant sub-routine shown
in section 2 of Figure 7.20.
During the time solution, however, saturation can vary over a large range of values
and a Newton form of iteration is an advantage especially if large integration steps are
used.
Redefining Equation (7.129) as:

(7.132)
7.9 ADVANCED COMPONENT MODELS 279

the elements of a 2 x 2 Jacobian matrix can be found. However, elements afl/aS,


and af2/aSd are small with respect to the other two elements and, if Ru is considered
to be zero, then the four elements reduce to:

This decouples the Newton method and each saturation factor may be solved inde-
pendently I1 31

(7.134)

Despite the advantages of a Newton form of solution, it can be found to be diver-


gent if too great an integration step length is used. Analysis of the functions f l and
f 2 show that they have discontinuities, when Xs = ( S d - 1)Xp and Xi = (S, - 1)Xp
respectively, although otherwise are almost linear. It is, therefore, necessary to monitor
this iterative procedure and to modify the step length, if necessary.
The evaluation of s d and S, should be performed twice during each iteration.
Considering Figure 7.21, this is during the calculation of the injected currents into
the network and the calculation of the non-integrable variables. Provided the discon-
tinuity is not encountered, convergence is achieved in one or two iterations at each
re-evaluation, especially if the saturation factors are extrapolated at the beginning of
each step.

7.9.2 Detailed turbine model


More detailed turbine models than that described earlier in the chapter are often required
for the following reasons:

(1) A longer-term transient stability study or a dynamic stability study is to be made.


(2) The turbine is a two-shafted cross-compound machine which has a separate gener-
ator on each shaft.

(3) Generator overspeed is such that an interceptor valve may operate during the
study.

A generalized model to accommodate the different types of compound turbine has


been developed by the IEEE [7]. As with the generalized AVR model, by setting certain
gains to either zero or unity and time constants to either infinity (very large) or zero,
the model can be reduced to any desired form. An interceptor valve can easily be
incorporated, as shown in Figure 7.27.
280 7 SYSTEM STABILITY

Figure 7.27 Generalized detailed turbine model including H.P.and interceptor valves

Some normal compound turbine configurations are shown in Figure 7.28, and
Table 7.5 gives typical values for these configurations using the generalized mode.
A hydroturbine can also be represented and the values given in Table 7.5 are justified
by the method of representing a lead-lag circuit, described in section 7.3, with the
time constant T5 set at i T w in the case of the simplest model.
The full set of equations for the detailed turbine model is:
pG4 = (Pgv- G4)/T4, (7.135)
pG5 = (G4 - G5)/T5, (7.136)
Piv = G5 . Pvi, (7.137)
pG6 = (Piv - G6)/T6, (7.138)
pG7 = (G6 - G7)/T7, (7.139)
P,I = K 1 . G4 + K 3 . Piv + K5 G6 + K 7 . G7,
3 (7.140)
Pm2 = K 2 . G 4 + K 4 . Piv + K 6 *G6 + K8 G7. (7.141)
Also note that:
8
C K n = 1 (7.142)
n=l

and that, in the initial steady state, the interceptor valve, if present, will be fully open
(Pvi = 1) in which case:

G4 = G5 = P i v = G6 = G7 = P m l + Pm2. (7.143)
The speed governor controlling the interceptor valve is similar to that controlling
the HP turbine except that it is set to operate at some overspeed value of slip (k,) and
not about synchronous speed. Equation (7.33) can be modified in this case to:
7.9 ADVANCED COMPONENT MODELS 281

(a)
-d-h--
Valve
Position
1
Control
valves,
steam
chest 1
Shaft

To condenser

Shaft
Position- steam

Control
valve
Position
valves,
steam
--- Shaft
chest

Valve -.Control
valves,
Position - steam
chest

-i-wF+-----
(d)

Reheater
Control

m-----
HP,IP Shaft
Valve
Position steam
chest
Crossover

LP Shaft

To condenser

Control
Valve valves,
Position + steam
chest

(f)

Figure 7.28 Common steam turbine configurations [7] (01982 IEEE) (a) Non-reheat;
(b) Tandem compound, single reheat; (c) Tandem compound, double reheat; (d) Cross
compound, single reheat; (e) Cross compound, single reheat; (f) Cross compound, double reheat
282 7 SYSTEM STABILITY

+-

pm2

Figure 7.28 (continued) Approximate linear models [7]. (OIEEE 1982)


Table 7.5 Parameters used in generalized detailed turbine model [7] (@1982 IEEE)
Turbine system figure Time constants Fractions
with typical values (seconds) with typical values (PA.) 4
b
T4 T5 T6 n K1 K2 K3 K4 K5 K6 K7 K8 ~

U
Non-reheat 10.7a TCH - - - 1 0 0 0 0 0 0 0 s
0.2- 0.5 z
Tandem-compound 10.7b TCH TRH Tco - FHP 0 Ftp 0 FLP 0 0 0 U
8
single-reheat 0.1-0.4 4- 1 I 0.3-0.5 0.3 0.4 0.3
Tandem-compound 10.7~ TCH TRH~ T R H ~ Tco FVHP 0 FHP 0 FIP 0 FLP 0
double-reheat 0.1-0.4 4-11 4-11 0.3-0.5 0.22 0.22 0.3 0.26 P
Cross-compound 10.7d TCH TRH T, - FnP 0 0 FIP ~FLP~
I FLP 0 0 8
single-reheat 0.1-0.4 4- 1 1 0.3-0.5 0.3 0.3 0.2 0.2 i5
Cross-compound 10.7e TCH TRH T, - FHP 0 F~P 0 0 FLP 0 0 s
single-reheat 0.1-0.4 4- 1 1 0.3-0.5 0.25 0.25 0.5
Cross-compound 10.7f TCH TRH, TRH~ Tco FVHP 0 0 FHP
I
IFIP fFip ~FLP~FLP
8
double-reheat 0.1-0.4 4-11 4-11 0.3-0.5 0.22 0.22 0.14 0.14 0.14 0.14
Hydro 9.9a 0 fTw - - -2 0 3 0 0 0 0 0 t;
284 7 SYSTEM STABILITY

7.9.3 Induction machines


An approach similar to that used to construct the synchronous machine models is
required if induction machines are to be explicitly modelled [ 15,191. However, speed
cannot be assumed to vary only slightly and this basic difference requires that the
equations describing the behaviour of induction machines be somewhat different from
those developed for a synchronous machine.

Mechanical equations It is necessary to express the equation of motion of an induc-


tion machine in terms of torque and not power. Also symmetry of the rotor makes its
angular position unimportant, and slip (S) usually replaces angular velocity ( w ) as the
variable, where:
s = (wg - w)/wo. (7.145)
Assuming negligible windage and friction losses and smooth mechanical shaft power,
the equation of motion is:
pS = (Tm - Te)/2Hm, (7.146)
where H m is the inertia constant measured in kWsfkVA established at synchronous
speed. The mechanical torque ( T m ) and electrical torque ( T e ) are assumed to be posi-
tive when the machine is motoring.
The mechanical torque T m will normally vary with speed, the relationship depending
on the type of load.
A commonly used characteristic is:

T m 0: {(speed)k),

where k = 1 for fan-type loads and k = 2 for centrifugal pumps. A more elaborate
torquehpeed characteristic can be used for a composite load, i.e.:

Tm 0: (a + b(speed) + c(speed)2), (7.147)


which can include the effect of striction when start-up is being considered.
In terms of slip, the torque is thus:

Tm =A + BS + CS’, (7.148)
where

A + b + c),
0: { a

B a ( b + 2c),
c 0: c.
The values of A , B and C are determined from the initial (steady state) loading of the
motor and, hence, its initial value of slip.
The electrical torque T e is related to the air gap electrical power by the electrical
frequency which is assumed coilstant and hence:

T e = Re@ i y ) / 2 nf 0 . (7.149)
7.9 ADVANCED COMPONENT MODELS 285

Figure 7.29 Steady state equivalent circuit of a single cage induction motor

Electrical equations A simplified equivalent circuit for a single-cage induction


motor is shown in Figure 7.29, with R1 and X I referring to the stator and R2 and
X 2 referring to the rotor resistance and reactance, respectively. In a similar manner to
the transient model of a synchronous machine, an induction motor may be modelled
by a Thevenin equivalent circuit of a voltage E’ behind the stator resistance R I and a
transient reactance X ‘ . The transient reactance is the apparent reactance when the rotor
is locked stationary and the slip (S) is unity and is given by:
X2Xm
X’=X1+ (7.150)
X2 Xm’ +
where X m is the magnetizing reactance of the machine. The rate of change of transient
voltage is given by:
pE’ = -j27rf sE’- (E’ - j(x0- x’>Il) / T & (7.151 )
where the rotor open-circuit time constant Tb is:
X 2 +X m
(7.152)
T b = 217f OR2
and the open-circuit reactance X O is:
XO=Xl+Xm. (7.153)
The reactances are unaffected by rotor position and the model is described in the
real and imaginary components used for the network, i.e. in the synchronously rotating
frame of reference. Thus, for a full description of the model, the following equations
are used:
V r - E: = R1 * 11, - X’ Zl,, (7.154)
V,,-E~=R1~II~+X’.Ilr, (7.155)
PEL = 2~rfOSEh- (E: + (XO - X’)Ilm)/T’O, (7.156)
PEL = - 2 f~OSE: - (EL - (XO - X’)Zlr)/T’O. (7.157)
A transient stability program incorporating an induction motor model uses the tran-
sient and open-circuit parameters but it is often convenient to allow the stator, rotor
and magnetizing parameters to be specified and let the program derive the former
parameters.
286 7 SYSTEM STABILITY

For completeness, the electrical torque may now be written as:


T e = (ErIlr +ELll,)/uo. (7.158)

Electrical equations when the slip is large Single-cage induction motors have
low-starting torques and it is often difficult to bring them to speed without either
reducing the load or inserting external resistance in the rotor circuit. As a result of the
low-starting torque, when the slip exceeds the point of maximum torque, the single-
cage model is often insufficiently accurate. These problems are overcome by the use
of a double-cage or deep-bar rotor model.
Cage factor
When a torque slip characteristic of the motor is available, then a simple solution is to
modify the torque-slip characteristic of the single-cage motor model. Double-cage or
deep-bar rotors have a resistance and reactance which varies with slip. A cage factor
K g can be included which allows for the variations of rotor resistance.
R2 = R2(0)(1 + Kg * S), (7.159)
where R2(0) is the rotor resistance at zero slip.
It is usually convenient to make the cage factor larger than that necessary to describe
the change in rotor resistance. In this way, the torque-slip characteristics of the model
can be made similar to that of the motor without the need to vary the rotor resistance
with slip. The result of varying the rotor resistance is to modify the open-circuit
transient time constant only, and this can be done quite simply at each integration step
of the simulation.
Rotor reactance does not vary with slip as greatly as rotor resistance, provided
saturation effects are ignored, and its effect on the open-circuit transient time constant is
thus small. Transient reactance (X’) varies with rotor reactance. However, this variation
on the term (XO- X’) in Equations (7.156) and (7.157) is insignificant. Thus, the only
major effect of varying rotor reactance is in Equations (7.154) and ( 7 . 1 5 9 , which
requires a technique similar to that adopted in the synchronous machine model to
account for saturation and saliency. However, the gains obtained in using two-cage
factors are insignificant and a single cage factor varying rotor resistance is usually
adopted.
Double-cage Rotor Model
An alternative to the cage factor is the use of a better rotor model, though this is often
restricted by the unavailability of suitable data.
Induction motor loads having double-cage or deep-bar rotors can be represented in a
similar manner to a single cage motor [20,21].It is assumed that the end-ring resistance
and that part of the leakage flux which links the two secondary windings, but not the
primary, are neglected. The steady state equivalent circuit, shown in Figure 7.30, can
thus be obtained where R3 and X 3 are the resistance and reactance of the additional
rotor winding. A circuit similar to that of Figure 7.29 can be obtained by substituting
the two parallel rotor circuit branches by a single series circuit, where:

R2(S) =
+ +
R2.R3(R2 R3) S2(R2.X3* R3.X2’) + (7.160)
(R2 + R3)2 + S 2 ( X 2 + X 3 )
288 7 SYSTEM STABILITY

and at standstill are:

R2(1) =
+
X22 * R3 X32 * R2 + R2 R3(R2 0 + R3) (7.171)
+ + +
(X2 X3)2 (R2 R3)2
X2 ‘ X3(X2 + X 3 ) + R22 X3 + R32 X2
X2(1) = (7.172)
( X 2 + X3)* + (R2 + R3)2
This set of non-linear equations may be solved using Newtonian techniques but by
substituting:
R2 ‘ R2(0)
R3 = (7.173)
R2 - R2(0)
and
x2 .xx
x3 = (7.174)
x 2 -xx’
where
X2( 1) - (R2(1 ) - R2(0))2
xx= (7.175)
(X2(0)- X2(1)>
the number of variables reduces to two and a simple iterative procedure yields a result
in only a few iterations [22].A reasonable starting value is X2 % ~ X derived
O from
assuming R2 i R 3 and X2 % i X 3 .

Representation of induction machines in the network This is quite simple


compared with the representation of synchronous machines, as neither saliency nor
saturation are normally considered in the induction machine models. They may, there-
fore, be considered as injected currents in parallel with fixed admittances.
Modifying Equations (7.162) and (7.163) gives a machine current of
71 = F(V - E”), (7.176)
or

(7.177)

The injected current into the network which includes is thus:

(7.178)
Ri + XI1* ~1 E; ’

where the minus sign confirms the induction machine as being assumed to be motoring.
Inclusion of induction machine in the transient stability program This is rela-
tively straightforward using the same format as developed for synchronous machines.
Most induction machines are equipped with contactors which respond to terminal condi-
tions such as undervoltage and it is sometimes necessary to model this equipment. The
characteristics and logic associated with contactors are included in section 7.9.4.
7.9 ADVANCED COMPONENT MODELS 289

7.9.4 Relays
Relay characteristics may be applied to a transient stability program and the effect
of relay operation automatically included in system studies. This permits checking
of relay settings and gives more realistic information as to system behaviour after a
disturbance, assuming 100% reliability of protective equipment. Reconstruction of the
events after fault occurrences may also be carried out.
Unit protection only responds to faults within a well-defined section of a power
system and, as the faults are pre-specified, the operation of unit protection schemes
can equally be specified in the switching data input. Thus, only non-unit protection
needs to be modelled and of these overcurrent, undervoltage and distance schemes are
the most common.
Instantaneous overcurrent relays Instantaneous, or fixed time-delay, overcurrent
relays are readily modelled. The operating point of the relay should be specified in
terms of p.u. primary current, thus avoiding the need to model the current transformer
specifically. However, the location of the current transformer must be specified, e.g. at
busbar A on branch to B, so that the correct signals are used by the relay model. The
only other piece of information required is delaytime (&I) between the relay operation
time and the circuit-breaker arc extinction time (fcb).
Initially, the circuit-breaker operating time (tcb) is set to some large value as it must
be assumed that the steady-state current is less than the relay setting. At the end of each
time step (e.g. at time t ) the current at the current transformer location is evaluated
and, if it exceeds the relay setting, the effective circuit-breaker time is set to:
tcb = t $. fdel. (7.179)
The integration then proceeds until the time-step nearest to fcb, when the circuit-
breaker opening is simulated by reducing the relevant branch admittance to a very
small value. Alternatively, the integration step length can be adjusted to open the
circuit at time fcb.
During the period between relay operation and fcb, the simulation of relay drop-off
may be desired. In this case, if current falls below a pre-specified percentage of relay
setting current, then tcb is reset to a large value.
Inverse definite minimum time lag overcurrent relays The inverse time charac-
teristics of induction disc and similar relays may easily be included in an overcurrent
relay. This may be accomplished by defining several points on the characteristic and
interpolating, but curve fitting is better if a simple function can be found.
For example, an overcurrent relay conforming to British Standard B.S. 142 would
appear to be accurately modelled by defining seven points on the curve as shown
in Figure 7.31(a). However, when plotted on a log-log graph, as in Figure 7.31(b),
the errors are more obvious and can exceed the accuracy limits laid down in the
Standard, if care is not taken. However, acceptable accuracy can be obtained by using
the approximation:
top = 3.O/log I for 1.1 5 1 5 20 (7.180)
and
lop= c a for I < 1.1,
where top is the operating time of the relay for a current of I .
290 7 SYSTEM STABILITY

25

20

4 8 12 16 20
(4 Setting multiple

100

30

3 10 30 100
(b) Setting multiple

Figure 7.31 Inverse definite minimum time lag overcurrent relay (from BS142 (1966)).
(a) Linear scale; (b) Logarithmic scale

Plug bridge setting (Spb) and time multiplier setting (Stm), both measured in per unit,
can be incorporated into the relay characteristic and the relay induction disc travel (0,)
at time t due to a current I, may be determined from the previous travel at time
t - h bv:

provided that I t 2 1.1Spb.


7.9 ADVANCED COMPONENT MODELS 291

For currents less than the definite minimum value, the relay may be assumed to reset
by spring action. Assuming that resetting from full travel takes 2s then:

(7.182)

when
I t < l.lS,b.

Initially, travel is set to zero and relay operation is assumed when D equals or
exceeds 1 p.u. If necessary, the relay operating time may be determined by linear
interpolation backwards over the last time interval.
Dt - 1.0
top = t - h, (7.183)
Dt - Dr-h
when
Dt 2 1.0
and
Dt-h < 1.0,

and from this, the circuit-breaker operating time is given by:

rcb = top -k tdel. (7.184)


Many static relays have been designed which conform to mechanical characteristics
but they have also permitted different and more suitable characteristics to be developed.
These may be modelled in a similar manner.
Undervoltage relays Apart from the fact that the relay operating current is propor-
tional to primary voltage and not primary current, these relays should be modelled in
the same manner as instantaneous or fixed time-delay overcurrent relays.

Induction machine contactors The transient analysis of industrial power systems


usually requires that many induction machines are modelled. During a disturbance,
the voltage levels throughout the system will fluctuate and may result in the machine
being disconnected from the system, albeit temporarily. Other machines may be on
automatic stand-by to maintain essential services.
It is, therefore, necessary to include models of induction machine contactors in
a transient stability program. Undervoltage protection is usually associated with the
contactors and can be modelled in the normal manner.
Directional overcurrent relay A directional overcurrent relay requires a voltage
signal as well as current. The relay may operate only when the phase difference of the
two signals is within prescribed limits and all other constraints are satisfied.
Distance relays In practice, both busbar voltage and branch current signals are
required, from which an apparent impedance Z,L6, of the system at the relaying point
can be calculated. This is then compared with the relay characteristic to determine
operation.
292 7 SYSTEM STABILITY

Figure 7.32 Three-zone distance relay characteristic

A typical three-zone distance protection relay is shown in Figure 7.32. Assuming


circular characteristics, then the settings of the relay may be identified by forward reach
Zrf L6,f measured in impedance (complex) coupled with backward reach Rb expressed
+
as a per unit of forward reach. From this information, the centre ( p j q ) and radius
(a) of each of the three circles in the impedance plane can be established:

(7.185)
q = izrf(l- Rb)sinO,f.
In the example in Figure 7.32, R b for zones 1 and 2 is zero.
The equation of the boundary of an operating zone is:
( z c o s e - p)’ + (ZsinB - ql2 - a’ = 0, (7.186)
and hence operation is defined when:

p 6, + q sin 6,) + ( p 2+ q2 - a2>I0,


- 2 ~ , ( cos (7.187)
Tomato, lens, quadrilateral or other complex characteristics may be constructed by
combining several simple characteristics of this type.
Each zone has a fixed time-delay associated with it so that the timing for circuit-
breaker action is the same as that described previously.

Incorporating relays in the transient stability program Non-unit protection


equipment usually only trips the local circuit-breaker. Therefore, it is necessary to
create dummy busbars so that a faulted branch can be switched out correctly. This
can be done automatically during the data input stage, as described in section 7.6, for
faults located on branches. Thus, a faulted branch may have several dummy busbars
7.9 ADVANCED COMPONENT MODELS 293

associated with it and care should be taken to ensure all are adequately identified.
Protected branches which are not directly faulted need not be modelled as accurately
and the whole branch may be removed if the circuit-breaker at either end is opened.
Relay characteristics should be checked at the end of each time solution and recon-
vergence after a discontinuity. It is not necessary to perform the check at each iteration,
however. This reduces the computational effort associated with relays and permits more
complex relay characteristics to be modelled at critical points in the system.
Induction machine switching should not be simulated by creating dummy branches
which can be removed from the network whenever necessary. While this is a feasible
solution, it is extremely wasteful of computational storage and effort. A more satisfac-
tory method is to identify the state of the machine, i.e. either switched in or out, by a
simple flag and when switched out to solve for the machine with zero stator current
and likewise remove its injected current from the network.
The network, however, usually includes a shunt admittance representing the machine
in the initial steady state. This problem may be overcome by injecting another current
to compensate for this admittance whenever a machine is switched out. Alternatively,
a machine liable to switching need not have its equivalent shunt admittance included
in the network at any time during the study. This simplifies periods when the machine
is switched out, but requires a different injected current to the usual when switched in.
A minor problem occurs when induction machines, which are initially switched out
of service, are included in the input data. An estimate of the full load active power of
the machine must be specified so that the load characteristics of the machine can be
adequately defined. Also, induction motors on stand-by for automatic start-up must be
modelled accurately if sensible run-up simulation is to be achieved.

7.9.5 Unbalanced faults


The models developed so far for transient stability analysis have assumed balanced
three-phase operation even during the fault period. Although three-phase faults are the
most onerous, there are occasions when unsymmetrical fault conditions need to be
analysed. It is possible to develop three-phase models of all power system equipment
but the development effort plus the extra computational costs restrict this type of
program to very simple systems. Unbalanced fault studies are relatively rare and the
unbalance only occurs for a short period of study; thus the need for a three-phase
model is limited and makes full-scale development unattractive.
A more practical approach is the use of symmetrical components. The negative-and
zero-sequence component system models can be added to the existing single-phase
(positive-sequence) model without major disruption and can be easily removed when
not required.

Negative sequence system Of the two additional symmetrical component systems,


the negative sequence is the easier. It is very similar to the positive-sequence system.
The negative-sequence impedances of the components of the transmission network
and static loads are usually the same as for the positive-sequence impedances and,
hence, no additional storage is required. Phase displacement in transformer banks is
of the opposite sign to that for the positive sequence. While phase displacement can
be ignored during balanced operation, it must be established if phase quantities are
294 I SYSTEM STABILITY

to be calculated during unbalanced operation. A simple clock notation with each hour
representing 30" shift is suitable for this purpose.
The negative-sequence impedance of synchronous machines is different from the
positive-sequence impedance. The flux produced by negative-sequence armature current
rotates in the opposite direction of the rotor, unlike that produced by positive-sequence
current, which is stationary with respect to the rotor. The flux path oscillates rapidly
between the positive-sequence d and q axis subtransient flux paths and the negative-
sequence reactance X2 may conveniently be defined as:
x2 = (x; + Xb')/2. (7.188)
This reactance is the same as the reactance which represents the machine in the
positive-sequence network. The negative-sequence resistance is given by(4):
(7.189)
where Rr is the rotor resistance. While R2 and Ra will differ, the overall difference
between the negative-sequence impedance (22)and positive-sequence impedance repre-
senting the machine is so small as to be neglected in most cases. Further, rotating
machinery do not generate negative-sequence e.m.f.s and, hence, there is no negative-
sequence Norton injected current.
Thus, ignoring d.c. equipment, the overall negative-sequence network is identical to
the positive-sequence network with all injected currents set to zero.
Negative-sequence currents have a braking effect on the dynamic behaviour of
rotating machinery. For a synchronous machine, where torque and power may be
assumed to be equivalent, the mechanical braking power (Pb) is [4]:
Pb = I:(R2 - Ra), (7.190)
which may be added directly into the mechanical equation of motion given by Equa-
tion (7.6).
A similar expression can be found for negative sequence-braking torque in an induc-
tion machine where the speed of the rotor and the negative-sequence currents are taken
into account.
Zero sequence system The zero-sequence system differs greatly from the other two
sequence systems.
The zero-sequence impedance of transmission lines is higher and, for a transformer,
its value and location depends on the phase connection and neutral arrangements.
Figure 7.33 shows the zero-sequence models for various typical transformer connec-
tions. By replacing the open circuit of transformer types 2, 3 and 4 with a very low
admittance, the topology of the zero-sequence network can be made the same as for
the other sequence networks.
The zero-sequence impedance of rotating machinery must be specified in the data
input so that its inverse can be included in the zero-sequence system admittance matrix.
As with the negative-sequence system model, there are no zero-sequence e.m.f.s gener-
ated and, hence, there are no Norton injected currents into this system.

Inclusion of negative and zero sequence systems for unsymmetrical faults The
major effect of unsymmetrical faults is to increase the apparent fault impedance. On
fault application, the negative- and zero-sequence impedances of the system at the
7.10 REFERENCES 295

Transformer type Various winding configurations

Type 1

p ~ z o - O ADP O

I I p P O P Q

Figure 7.33 Modelling of zero-sequence equivalent networks of transformers

point of fault are calculated. These are simply the inverse of the self-admittances at
the point of fault. Depending on the type of fault, the fault impedance is modified to
include the negative- and zero-sequence impedance. The fault impedance then remains
constant until changed by either branch switching or fault removal.
If negative sequence-braking effects are to be included, it is necessary to evaluate
the negative-sequence current in the relevant machines at each iteration. This is done
by injecting the negative-sequence current, determined at the point of fault, into the
negative sequence system admittance matrix [ 7 2 1.
(7.191)
where [ffz] is a zero vector except at the point of fault. The vector [v2]contains the
negative-sequence voltages at all busbars from which the machine negative-sequence
currents are readily obtained.
If phase information is required, then the zero-sequence voltages at all busbars also
need to be determined, depending on the type of fault. This is done in an identical
manner to that used for the negative-sequence system.

7.10 References
I . Elgerd, 0 I, (1971). Electrical Energy Systems Theory: An Introduction, McGraw-Hill, New
York.
2. Weedy, B M, (1979). Electric Power Systems, John Wiley and Sons, London.
3. Concordia, C, (1951). Synchronous Machines. Theory and Performance, John Wiley and
Sons, New York.
296 7 SYSTEM STABILITY

4. Kimbark, E W, (1956). Power System Stability: Synchronous Machines. Vol. 3, John Wiley
and Sons, New York.
5. Dandeno, P L, et al., (1973). Effects of synchronous machine modeling in large-scale system
studies. IEEE Transactions on Power Apparatus and Systems, PAS-92(2), pp. 574-582.
6. IEEE Committee Report, (1968). Computer representation of exciter systems, IEEE Trans-
actions on Power Apparatus and Systems, PAS-87(6), pp. 1460- 1464.
7. IEEE Committee Report, (1973). Dynamic models for steam and hydro turbines in
power-system studies, IEEE Transactions on Power Apparatus and Systems, PAS-92(6),
pp. 1904-1915.
8. Dandeno, P L and Kundur, P, (1973). A noniterative transient stability program including
the effects of variable load-voltage characteristics, IEEE Transactions on Power Apparatus
and Systems, PAS-92(5), pp. 1478- 1484.
9. Berg, G L, (1973). Power system load representation, Proceedings of the IEE, 120(3),
pp. 344-348.
10. Dommell, H W and Sato, N, (1972). Fast transient stability solutions, IEEE Transactions
on Power Apparatus and Systems, PAS-91(4), pp. 1643- 1650.
1 1, Brameller, A, et al, (1969). Practical Diakoptics for Electrical Networks, Chapman and Hall,
London.
12. Lapidus, L and Seinfeld, J H, (1971). Numerical Solution of Ordinary Differential Equa-
tions, Academic Press, New York.
13. Arnold, C P, (1976). Solutions of the multi-machine power-system stability problem, Ph.D.
thesis, Victoria University of Manchester, UK.
14. Crary, S B, (1945). Power System Stability-Steady-State Stability, Vol. 1, John Wiley and
Sons, New York.
15. Olive, D W, (1966). New techniques for the calculation of dynamic response, IEEE Trans-
actions on Power Apparatus and Systems, PAS-85(7), pp. 767-777.
16. Hammons, T J and Winning, D J, (1971). Comparisons of synchronous-machine models in
the study of the transient behaviour of electrical power systems, Proceedings of the IEE,
118( lo), pp. 1442- 1458.
17. Beckwith, S, (1937). Approximating Potier reactance, AIEE Transactions, p. 813.
18. Knable, A H, (1956). Electrical Power Systems Engineering-Problems and Solutions,
McGraw-Hill, New York.
19. Brereton, D S Lewis, D G and Young, C C, (1957). Representation of induction motor loads
during power-system stability studies, AIEE Transactions on Power Apparatus and Systems,
PAS-76, pp. 45 1-461.
20. Jordan, H E, (1979). Synthesis of double-cage induction motor design, AIEE Transactions
on Power Apparatus and Systems, PAS-78, pp. 691-695.
21. Arnold, C P and Pacheco, E J P, (1979). Modelling induction motor start-up in a multi-
machine transient stability program, IEEE PES Summer Meeting, Vancouver.
22. Pacheco, E J P, (1975). Induction motor starting in an electrical power-system transient-
stability programme, M.Sc. dissertation, Victoria University of Manchester, UK.
SYSTEM STABILITY UNDER
POWER ELECTRONIC
CONTROL

8.1 Introduction
The presence of high voltage power electronic equipment in the transmission system
has important consequences on the stability of the power system. However, the conven-
tional multi-machine stability assessment described in the last chapter is inadequate to
represent the transient behaviour of some power electronic devices. This problem is
particularly important in the case of HVDC converters due to the occurrence of inverter
commutation failures that often follow a large disturbance. Only the use of electro-
magnetic transient programs, described in Chapter 6, can provide accurate information
of the power electronic response during and immediately following clearance of the
disturbance. Although machine rotor dynamics, as discussed in Chapter 7, could be
included in the electromagnetic transient program of Chapter 6 , this is not a practical
proposition in the presence of power electronic switching, considering the different
time constants involved. For instance, electromagnetic-transient simulations require
steps of (typically) 50 ps, whereas the stability programs use steps at least 200 times
larger.
To reduce the computational requirements, some programs [ 11 contain two separate
modes. An instantaneous mode is used to model components in three-phase detail with
small time steps in a similar way to the EMTPEMTDC programs [2]. The alternative
is a stability mode based on r.m.s. quantities at fundamental frequency only, with
increased time-step lengths. The user can switch between the two modes as required
while running but, in either mode, the entire system must be modelled in the same
way. Thus, when using the instantaneous mode, a system of any substantial size would
still be very computationally intensive.
A more efficient solution is described here [3,4] which takes advantage of the
computationally inexpensive dynamic representation of the a.c. system in the stability
program, and the accurate transient simulation of the power electronic devices.
The slow dynamics of the a.c. system are adequately modelled by the stability
program while the fast dynamic responses of the HVDC and FACTS devices are accu-
rately represented by electromagnetic transient simulation. This hybrid combination is
essential to predict the first few swings in stability studies.
298 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

Finally, quasi steady-state models of the power electronic devices are also developed
in this chapter for use in longer stability studies as well as system response to small
perturbations.

8.2 Description of the Algorithm


Transient stability (TS) studies are normally carried out on the assumptions of balanced
and sinusoidal waveforms and such assumptions need to be extended to the voltage
and current waveforms at the terminals of the power electronic devices. As explained
in the introduction, part of the study requires detailed transient simulation, the results
of which cannot be directly incorporated in a TS study.
It is therefore necessary to extract fundamental frequency quantities for the TS study
from the voltage and current waveforms produced by the electromagnetic transient
program (which in this book is the PSCADEMTDC version).
The hybrid concept, referred to as TSE and illustrated in Figure 8.1, uses electrome-
chanical simulation as the steering program, and the electromagnetic transient program
is called as a subroutine. The interfacing code is written in separate routines to mini-
mize the number of modifications and thus make it easily applicable to any stability
and dynamic-simulation programs.
Initially, the TSE hybrid reads in the data files and runs the entire network in
the stability program, until electromechanical steady state equilibrium is reached.
The steady-state representation of the converter, described in Chapter 3, is perfectly
adequate as no fault or disturbance has yet been applied. Before placing the network
disturbance, the TS network is split into two independent and isolated systems, system
1 and system 2, as shown in Figure 8.2.
System 1 contains the a s . part of the system modelled by the stability program TS,
and system 2 the part of the system modelled in detail by EMTDC.
The snapshot data file is now used to initialize the EMTDC program used, instead
of the TS representation of system 2. The two programs are then interfaced and the
network disturbance can be applied. The system 2 representation in TS is isolated but
kept up to date during the interfacing at each TS time step to allow tracking between
programs. The a.c. network of system 1 modelled in the stability program also supplies
interface data to the system 2 network in TS, as shown in the lower part of Figure 8.2.
During the disturbance, the quasi steady-state representation of system 2 in TS and
the EMTDC representation of system 2 are tracked. When both of these system 2

Conventional Stability
Analysis

7- transient simulation
of power electronic

Figure 8.1 The hybrid concept


8.2 DESCRIPTION OF THE ALGORITHM 299

Stability program ENTDC program

Figure 8.2 Interfacing procedure

detailed

sequence Stability
data for Loadflow input TSlEMTDC Outflow
whole hybrid

c-y
Interface

Figure 8.3 Data flow

models produce the same results within a predefined tolerance and over a set period,
the complete system can then be reconnected and used by TS,and the EMTDC repre-
sentation terminated. This allows better computational efficiency, particularly for long
simulation runs.

8.2.1 Data flow


Data for the EMTDC solution entered in the programme database via the PSCAD
graphics. Equivalent circuits are used at each interface point to represent the rest of
the system not included in the detailed model. This system is then run until steady
300 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

state is reached and a ‘snapshot’ is taken. This snapshot holds all the relevant data for
the various system components and is used as the starting point when interfacing the
detailed model with the stability programme.
The stability program is initialized conventionally through power-flow results via a
data file. An interface data file is also read by the TSE hybrid that provides the number
and location of interface buses, analysis options, and timing information. The data flow
diagram is shown in Figure 8.3.

8.2.2 Modifications required to the component programs


The main modification of the TS program is a call of the interfacing routine in the main
loop, as shown in Figure 8.4. The complete network is split into system 1 and system
2 at the interface points, but this task is performed using separate code. The only other
direct modification inside TS is the inclusion of the interface current injections at each
TS network solution.
To enable EMTDC to be called as a subroutine from TS requires a few changes
to its structure. This is helped by splitting EMTDC into three distinct segments, i.e.
an initializing segment, the main time loop, and a termination segment. In this way,
TS can call the main time loop for discrete periods as required when interfacing. The
EMTDC options, which are normally available at the beginning of a simulation run,
are moved to the interface data file and read from there. The equivalent-circuit source
values, which TS updates periodically, are located in the user-accessible DSDYN file
of the EMTDC (described in Chapter 6).

8.3 TS/EMTDC Interface


The hybrid algorithm involves regular exchanges of information between the transient
stability and electromagnetic transient programs.
The information to be transferred from one program to the other must be sufficient
to determine the power flow in or out of the interface. As shown in Figure 8.5, possible
parameters to be used are the real power, P, the reactive power, Q, the voltage, V, and
the current, I, at the interface. Phase-angle information is also required to be able to
use separate phase frames of reference.
Appropriate equivalent circuits are needed to represent the network modelled in the
stability program when running the EMTDC and vice versa. These are illustrated in
Figure 8.6, where E, and represent the equivalent circuit of system 1 and 7, and
z2 the equivalent circuit of system 2.
8.3.1 Equivalent circuit components
In Figure 8.6, f, and z2 represent the detailed part of the system modelled by EMTDC
to be used by the TS program since TS is only using the fundamental active power
flowing in or out of the interface, the equivalent impedance 2 2 can be selected arbi-
trarily and the current source 7, can be varied to provide the correct power flow.
Moreover, to avoid numerical instability, a constant value of z2, estimated from the
initial power flow results, is used for the duration of the simulation.
8.3 TS5MTDC INTERFACE 301

-
Yes

No
Pass information
to detailed model

i Solve EMTDC

Extract information
from detailed solution
and include in stability
programme

1
1 I
Solve stability equations

c=
T = T + step length

Output results

Is T = end time?

Figure 8.4 Modified TS steering routine

The EMTDC programme represents system 1 by a Thevenin equivalent (El and zl)
as shown in Figure 8.6. The simplest ZI is an R-L series impedance, representing the
fundamental frequency equivalent of system 1. This impedance can be derived from
the results of a power flow and a fault analysis at the interface bus.
The power flow 'provides an initial current through the interface bus and the initial
interface bus. A fault analysis is also used to determine the fault current through the
302 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

f
Figure 8.5 Hybrid interface

I
+

System 1 I,
I
System 2

Figure 8.6 Equivalent circuits at the interface

interface for a short circuit fault to ground. The equivalent Thevenin source El and
Thevenin impedance ZI values shown in Figure 8.7, are obtained as follows:
Using the power flow solution:
El = inZl + V ,
and using the fault circuit: -
El = I F Z ~ .
Combining Equations (8.1) and (8.2):

and can then be found from either Equation (8.1) or Equation (8.2).
During a transient, the impedance of the synchronous machines in system 1 will
change. The net effect on the fundamental power in or out of the equivalent circuit,
however, can be represented by varying the source and keeping ZI constant.
8.3 TS/EMTDC INTERFACE 303

-
(a) Loadflow circuit

- -
(b) Fault circuit

Figure 8.7 Derivation of Thevenin equivalent circuit (a) Power-flow circuit. (b) Fault circuit

However, the waveforms obtained from EMTDC involve frequencies other than the
fundamental and, therefore, a frequency-dependent equivalent circuit (as described in
section 6.4.3)is used to represent the a.c. system.
Regarding the equivalent circuits sources, information from the EMTDC model
representing system 2 (in Figure 8.6) is used to modify the source of the equivalent
circuit of system 2 in the stability programme. Similarly, data from TS are used to
modify the source of the equivalent circuit of system 1 in EMTDC. These equiva-
lent sources are normally updated at each TS step length (refer to section 8.5). From
Figure 8.6,if both ZI and 2 2 are known, additional information is still necessary to
determine update values for the sources 7, and El. The selection and processing of
interface parameters to derive this information is discussed in section 8.4.
Moreover, the stability programme requires only positive sequence data so data from
the three a.c. phases at the interface(s) is analysed and converted to positive sequence,
i.e. for the positive sequence voltage:
-
v,, = p,+ ZVb + Z*V,), (8.4)
where
-
V,, = positive sequence voltage,
- - -
V,, V b , V, = phase voltage,
Z = 120"forward rotation vector(i.e.a = 1L 120").
304 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

Similarly, the positive sequence voltage from TS is converted to three phase data as
follows:
- -
Va = Vps, (8.5)

Some transient stability programmes have the additional capability of modelling


negative and zero sequence networks. If this option is utilized, sequence data from
a stability programme can then be converted to three phase through the following
formulae:
- -
Va = Vps + Vns + Vzs, (8.8)
- -
v b = a 2 v p s + z v n s + v,,, (8.9)
-
v, = zvps+ ans
+ V,,, (8.10)
where
-
Vns = negative sequence voltage,
-
V, = zero sequence voltage.
Similarly, any unbalance in system 2 can be accommodated in the transient stability
programme.

8.3.2 Interface variables derivation


As already explained, in Figure 8.6, El and ZI represent the equivalent circuit of system
1 modelled in EMTDC, while 2 2 and 7, represent the equivalent circuit of system 2
v
modelled in the stability program. is the interface voltage and 71 the current through
the interface which is assumed in the direction shown.
The magnitudes of the interface voltage and current, along with the phase angle
between them, are derived from the detailed EMTDC simulation. This information is
used to update the equivalent circuit source (TC) of system 2 in TS, i.e.:
From Figure 8.6:
-
El = TiZ1 +V , (8.11)
-
v = i2Z2, (8.12)
-
i 2 = i, + 7,. (8.13)
From Equations (8.12) and (8.13):
v = -z1z2
- -
+ icZ2. (8.14)
From Equation (8.1 1):
-
E~ = IIzIL(eII + ezl)+ w e V
=I121cos(6Il + & l ) + j l l Z l sin(& + & I ) +Vcos6v+ jVsin6v (8.15)
8.3 TSEMTDC INTERFACE 305

and
611 = ev - 9, (8.16)
where rb is the displacement angle between the voltage and the current.
Thus, Equation (8.15) can be written as
-
EI = IlZl cos(8v + B ) + jIlZ1 sin(& + B ) + VcosBv + j v s i n b
=IIZl(cos8vcosB-sinevsinB)+ vcosev (8.17)
I Ov cos /? + cosev sin /?)+ V sin ev],
+ ~ [ I I Z(sin
where
B = ezI- 9.
If
Ei = E l , + j E l i ,
then equating the real terms:
E l , = (7,Zl cosB+ V)cos6v + (-1121sinB)sin&, (8.18)
where zlis known and constant throughout the simulation.
From the EMTDC results, the values of V , I and 4 are also known and, hence, so
is j3. El can be determined in the TS phase reference frame from knowledge of zI
and the previous values of interface current and voltage from TS, through the use of
Equation (8.1 1).
From Equation (8.18), making
A = IlZi C O S ~ + V, (8.19)
B = - I I Z ~sin B, (8.20)
and remembering that

A cos ev + B sin ev = J F T 3 c o s ( e v *), (8.21)


where

the voltage angle Bv in the TS phase reference frame can be calculated, i.e.:

(+= tan-' )I:[ . (8.23)

The equivalent current source 7, can be calculated by rearranging Equation (8.14):


- v
I, = -L(ev - eZ2) - Lerl, (8.24)
22
where 811 is obtained from Equation (8.16).
306 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

In a similar way, data from the transient stability programme simulation can be
used to calculate a new Thevenin source voltage magnitude for the equivalent circuit
of system 1 in the EMTDC programme. Knowing the voltage and current magnitude
at the TS programme interface and the phase difference between them, by a similar
analysis the voltage angle in the EMTDC phase reference frame is:

where I,, is the real part of 7, and


V
c = -cosez2 -I, COS~, (8.26)
22
V
D = - cosO~2- I ' sin@, (8.27)
22
4 = ev -ell, (8.28)

+ = tan-' I$[ . (8.29)

Knowing the EMTDC voltage angle 8v allows calculation of the EMTDC current
angle from Equation (8.28). The magnitude value of El can then be derived from
Equation (8.1 1).

8.4 EMTDC to TS Data Transfer


TS assessment is based purely on sinusoidal waveforms, but during faults the actual
waveforms are very distorted. However, there is no consensus on a universal definition
for power when waveforms are non-sinusoidal [ 5 ] .
In the original edition of the book, the total r.m.s. real power was used as the
interfacing variable to transfer information. The total r.m.s. power is made up of
fundamental power (Ff)and harmonic power (&), i.e.:
prms = pf + ph. (8.30)
The r.m.s. power can be extracted directly from the waveforms and, therefore,
Fourier transform or curve fitting methods are not necessary, i.e.:

(8.31)

This method greatly reduces the computing time necessary for the data extraction
from EMTDC.
The choice of r.m.s. power was made on the basis that harmonic power flow will
result only from in-phase components of harmonic voltage and current. The assumption
was made that if a system contains only a low resistive component, then the harmonic
power flow is not significant [6].
This, however, is not valid for every situation and, particularly, at the inverter end of
an HVDC link, the effect of the resistive component of the network is not insignificant.
8.4 EMTDC TO TS DATA TRANSFER 307

Certain harmonic frequencies in a network may also be parallel resonant or close


to parallel resonance and exhibit more resistance than reactance. The presence of a
transient may excite the resonant frequency and greatly affect the results. It is important,
therefore, to model the fundamental power flow accurately.
Another factor to consider is that the direction of fundamental power may not neces-
sarily be the same as the direction of harmonic power. With an HVDC link, while
fundamental power is drawn into a rectifier, much of the harmonic power flow will be
in the reverse direction. The r.m.s. real power measured is the difference between these
two powers and is, therefore, not entirely representative of the fundamental power load
of the HVDC link. Conversely, at the inverter end the r.m.s. real power will include
harmonic power flow into the a.c. system. This may exaggerate the amount of true
fundamental power from the inverter.
As an illustration of the problem, Figure 8.8 shows an EMTDC simulated result of
a single phase fault at the inverter end of the CIGRE HVDC benchmark model [7]
The instantaneous power is shown, along with the total r.m.s. power over discrete one
cycle intervals. For hybrid interfacing purposes, data is transferred at discrete intervals
equal to the stability programme time-step and these are significantly larger than the
interval between the discrete output points constituting the instantaneous power. The
r.m.s. power is then taken over a fundamental period to represent its use in the hybrid
model.
Figure 8.9 shows an analysis of the single-phase fault comparing fundamental
frequency power with the total r.m.s. power. The fundamental frequency power
was derived using the curve fitting method described in section 8.5 to extract both
fundamental voltage and current. The comparison shows that, particularly during the

Instantaneous power -Rms power over T


lo00

800

600

400
F
3 200

[ 0
1
a -200
-400

-600

-800

-1000 , I I I

.O
308 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

1000
900 -
-
800

- 700 -
$ 600-
Y

zi 500-
-m 400-
$ 300-
200 -
100 -
0-
-100 -
-200 I 1

0.00 0.04 0.08 0.12 0.16 0.20 0.24 0.28 0.32 0.36 0.40

Figure 8.9 Fundamental power versus total r.m.s. power

fault time, there exists a significant amount of Ph or harmonic r.m.s. power in the total
r.m.s. power. When the three phase network is unbalanced the fundamental frequency
real power consists of positive, negative and zero sequence components, i.e.:
Pf = Pfps + Pfns + Pfzs, (8.32)
and the negative and zero sequence powers cause additional power loss in the network.
Figure 8.10 shows the sequence components of the fundamental r.m.s. power of
Figure 8.9.
Fundamental frequency negative sequence currents, in the presence of damper wind-
ings, can produce a braking torque which will retard the rotor [8]. Damper windings,
however, also serve to lower the negative sequence impedance of a machine which in
turn reduces the negative sequence voltage [9]. Which of these two opposing effects is
dominant depends on the resistance of the damper windings. High resistance windings
cause the braking torque to be the significant effect. The braking power of the negative
sequence current is:
(8.33)

where
I, = negative sequence rotor current,
R, = rotor resistance,
I,, = negative sequence stator (or armature) current,
8.4 EMTDC TO TS DATA TRANSFER 309

1000 -.
900 -
800 -
F
z 700
600 -
-
' 500 - -
8I 400 -
a 300 -
200 - -
100 - --d
r----

0 - ------ ---- I
..,...r
- - -
-
- - - -

-
-100
-200 - I

Figure 8.10 Sequence components of fundamental frequency r.m.s. power

rn = negative sequence resistance of the machine,


rp = positive sequence resistance of the machine.

The negative sequence resistance can be approximated from the rotor and the arma-
ture resistance, i.e.:
rn % ( ~ s +i ~ r ) ,
(8.35)
where
R, = armature resistance.

Retardation of the rotor can also be caused by d.c. components in the armature
windings. Three phase faults at or near machine terminals can cause d.c. components
of short circuit armature current which can have a definite braking effect on the machine
[9].The braking power in this case is:
(8.36)
(8.37)
where

I , = the effective very low frequency value of the armature current,


idc = instantaneous d.c. component of armature current.

The total r.m.s. power then is not always equivalent toeither the fundamental frequency
power or the fundamental frequency positive sequence power. A comparison of these
310 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

1000
900 -
800 -
-
- 700
600-
v
5 500-
3
-gm 400-
2 300-
200 - Ic..*

100 -
0-
-100- .--
_I.
7-
..

d
-200 I I t I I

0.00 0.04 0.08 0.12 0.16 0.20 0.24 0.28 0.32 0.36 0. 0

Figure 8.11 Comparison of total r.m.s. power, fundamental frequency power and fundamental
frequency positive sequence power

three powers is shown in Figure 8.11. The difference between total r.m.s. power and the
positive sequence power can be seen to be highly significant during the fault.
The most appropriate power to transfer from EMTDC to TS is then the funda-
mental frequency positive sequence power. This, however, requires knowledge of both
fundamental frequency positive sequence voltage and fundamental frequency positive
sequence current. These two variables contain all the relevant information and, hence,
the use of any other power variable to transfer information becomes unnecessary.

8.5 Data Extraction from Distorted Waveforms


At each step of the transient stability program power transfer information needs to be
derived from the distorted current and voltage waveforms at the points of interface.
This can be achieved using the FFT which provides accurate information for the whole
frequency spectrum. However, only the fundamental frequency is used in the stability
program and a simpler curve fitting approach is described next that provides sufficient
accuracy.
A curve fitting algorithm (CFA) can be used to extract the fundamental frequency
data, based on a least squared error technique. It can be described as follows: Assume
a sinewave signal with a frequency of o radians per second and a phase shift of $
relative to some arbitrary time To.
y ( t ) = A sin(@?- $), (8.38)
where +=d o .
8.5 DATA EXTRACTION FROM DISTORTED WAVEFORMS 311

This can be rewritten as:


y ( t ) = A sin(wt) cos(wT0) - A cos(ot) sin(wT0). (8.39)
Letting C1 = A cos(wTo) and C2 = A sin(wT0) and if sin(wt) and cos(ot) are repre-
sented by functions F I ( t ) and F2(t), respectively, then:
y(t> = CI FI ( t ) + C2F2(t). (8.40)
F l ( t ) and F 2 ( t ) are known if the fundamental frequency w is known. However, the
amplitude and phase of this frequency generally need to be found, so the equation has
to be solved for C I and C2. If the signal y ( t ) is distorted, then its deviation from a
sinusoid can be described by an error function E :
~ ( t=) y ( t ) +E . (8.41)
For a least squares method of curve fitting, {he size of the error function is measured
by the sum of the individual residual squared values such that:

i= 1
+
where x; = x(t0 i A t ) and yi = y(t0 + iAr).
From Equation (8.40):
n
E= C ( x ; - ~ I ~ l ( t-i ~) 2 ~ 2 ( t i ) ) ~ , (8.43)
i= I

where the residual value r at each discrete step is defined as:

In matrix form:

(8.45)

or

The error component can be described in terms of the residual matrix as follows:
E = [rlT[r]
312 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

This error then needs to be minimized.

(8.48)
If [A] = [FIT[F] and [B] = [FIT[X] then :
[CI = [ A I - m (8.49)

and hence

Elements of matrix [A] can then be derived as shown:

n-1

(8.50)

Similarly:

and

(8.5 1)
(8.52)

From these matrix element equations, CIand C2 can be calculated recursively using
sequential data.
8.6 INTERFACE METHOD 313

2’oo I

-5.00 I 1 I I I 1 I 1 I 1

0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20
Time (s)

Figure 8.12 Simulated fault waveform

8.5.1 CFA effectiveness


The accuracy of the CFA algorithm has been compared with the results of an FFT
for the voltage waveform of Figure 8.12. This waveform, derived with the EMTDC
program, corresponds to a single-phase fault on the rectifier side of the CIGRE bench-
mark model [7].
Several windows were used to capture the fundamental frequency information. Of
these, a discrete rectangular window of full period duration provides the best match
when data are only required at intervals of one cycle. However, if data are required at
more frequent intervals, a staggered (or overlapping) window producing results twice
per cycle is a better choice.

8.6 Interface Method


The data from each program must be interchanged at appropriate points during the
hybrid simulation run. The timing of this data interchange between the TS and EMTDC
programs is important, particularly around discontinuities caused by fault application
and removal.
The electromechanical a.c. system solution exhibits a relatively slow dynamic
response to any disturbances and the accuracy of a previous time-step information
passed from TS to EMTDC is, therefore, an adequate assumption. On the other hand,
the HVDC converter exhibits fast dynamic behaviour. Present time-step information
is, thus, more appropriate to pass from EMTDC to TS when solving the TS network
equation. Therefore, of the different possible interfacing alternatives, the method
314 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

illustrated in Figure 8.13 is selected for the information exchange. It shows the use
of the previous time-step data when passing information from TS to EMTDC but
present time-step data when passing information from EMTDC to TS. When the two
programs are run concurrently, TS passes its information to EMTDC. EMTDC then
runs to t + A t , and the information gathered over this present time-step given back to
+
TS at time t . TS now runs to t Ar, using equivalent circuit information derived for
that particular time step.
However, to cater for TS step lengths below the fundamental period, the above
method is modified as shown in Figure 8.14; it relates to the case when the step length
is one half of the fundamental period.
Following the sequential numbering on Figure 8.14, at a particular point in time,
the EMTDC and TS programs are concurrent and the TS information from system 1
is passed to update the system 1 equivalent in EMTDC. This is shown by the arrow
marked 1. EMTDC is then called for a length of half a fundamental period (arrow
2) and the curve fitted results over the last full fundamental period processed and
passed back to update the system 2 equivalent in TS (arrow 3). The information over
this period is passed back to TS at the mid-point of the EMTDC analysis window
which is half a period behind the current EMTDC time. TS is then run to catch up

I I

Figure 8.13 Interfacing methods

Stabiliity

--
step length

I I
4
I
I
8
I
I
etc.
I
1-1 I I
I TS

-- 2 6 4 c III I
Dynamic program
step length

Figure 8.14 Normal interaction protocol


8.7 INTERFACE LOCATION 315

Disturance Stabiliity
applied step length

I
I
l
I
i
I
4
l
I
t
I
*
I
I
I
- /-I
I
8 etc.
I I I
TS

v p * t
++ II
II
r Dynamic program
step length

Figure 8.15 Interaction protocol around a disturbance

to EMTDC (arrow 4), and the new information over this simulation run used to again
update the system 1 equivalent in EMTDC (arrow 5 ) . This protocol continues until
any discontinuity in the network occurs.
When a network change such as a fault application or removal occurs, the interaction
protocol is modified to that shown in Figure 8.15. The curve fitting analysis process is
also modified to avoid applying an analysis window over any point of discontinuity.
The sequential numbering in Figure 8.15 explains the flow of events. At the fault
time, the interface variables are passed from TS to the system 1 equivalent in EMTDC
in the usual manner, as shown by the arrow marked 1. Neither system 1 nor system
2 has yet been solved with the network change. The fault is now applied in EMTDC
which is then run for a full fundamental period length past the fault application (arrow
2) and the information obtained over this period passed back to TS (arrow 3). The fault
is now also applied to the TS program which is then solved for a period until it has
again reached EMTDC’s position in time (arrow 4). The normal interaction protocol
is then followed until any other discontinuity is reached.
A full period analysis after the fault is applied is necessary to accurately extract
the fundamental frequency component of the interface variables. The mechanically
controlled nature of the a.c. system implies a dynamically slow response to any distur-
bance and so, for this reason, it is considered acceptable to run EMTDC for a full
period without updating the system 1 equivalent circuit during this time.

8.7 Interface Location


The original intention of the initial hybrid algorithm [6] was to model the a.c. and d.c.
solutions separately. The point of interface location was consequently the converter
bus terminal. The detailed d.c. link model included all equipment connected to the
converter bus, such as the a.c. filters, and every other a.c. component was modelled
within the stability analysis. A fundamental frequency Thevenin equivalent was used
to represent the stability program in the detailed solution and vice versa.
316 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

An alternative approach has been proposed [ 101 where the interface location is
extended out from the converter bus into the a.c. system. This approach maintains that,
particularly for weak a.c. systems, a fundamental frequency equivalent representing the
a s . system is not adequate at the converter terminals. In this case, the extent of the a.c.
system to be included in the d.c. system depends on phase imbalance and waveform
distortion.
Although the above concept has some advantages, it also suffers from many disad-
vantages. The concept is proposed, in particular, for weak a.c. systems. A weak a.c.
system, however, is likely to have any major generation capability far removed from the
converter terminal bus as local generation serves to enhance system strength. However,
when the generation is far away from the converter, then the distance required for an
interface location to achieve considerably less phase imbalance and waveform distortion
is also likely to be significant.
The primary advantage of a hybrid solution is in accurately providing the d.c.
dynamic response to a transient stability program, and in efficiently representing the
dynamic response of a considerably sized a.c. system to the d.c. solution. Extending the
interface some distance into the a.c. system, where the effects of a system disturbance
are almost negligible, diminishes the hybrid advantage. If a sizeable portion of the a.c.
system requires modelling in detail before an interface to a transient stability program
can occur, then one might question the use of a hybrid solution at all and instead use
a more conventional approach of a detailed solution with a.c. equivalent circuits at the
system cut-off points.
Another significant disadvantage in an extended interface is that ax. systems may
well be heavily interconnected. The further into the system that an interface is
moved, the greater the number of interface locations required. The hybrid interfacing
complexity is thus increased and the computational efficiency of the hybrid solution
decreased. The requirement for a detailed representation of a significant portion of
the a.c. system serves to decrease this efficiency, as does the increased amount of
processing required for variable extraction at each interface location.
The advantages of using the converter bus are:
0 The detailed system is kept to a minimum.
0 Interfacing complexity is low.
0 Computational expense is minimized.
0 Converter terminal equipment, such as filters, synchronous condensers SVCs, and
can still be modelled in detail.
The major drawback of the detailed solution is in not seeing a true picture of the
ax. system, since the equivalent circuit is fundamental frequency based. Waveform
distortion and imbalance also make it difficult to extract the fundamental frequency
information necessary to transfer to the stability program.
The problem of waveform distortion for transfer of data from EMTDC to TS is
dependent on the accuracy of the technique for extraction of interfacing variable infor-
mation. If fundamental frequency quantities can be accurately measured under distorted
conditions, then the problem is solved. Section 8.5 has described an efficient curve
fitting algorithm to extract the required information from distorted waveforms. It has
8.8 STRUCTURE OF THE HYBRID PROGRAM 317

Rectifier 7 1 Inverter ESCR=2.0 h+-@

Figure 8.16 Test system

been shown that, using the technique described, fundamental frequency quantities can
be accurately measured.
Moreover, a simple fundamental frequency equivalent circuit is insufficient to repre-
sent the correct impedance of the a.c. system at each frequency. Instead, this can be
achieved by using a fully frequency-dependent equivalent circuit of the a.c. system
[ l 11 at the converter terminal instead of just a fundamental frequency equivalent. A
frequency-dependent equivalent avoids the need for modelling any significant portion
of the ax. system in detail yet still provides an accurate picture of the system impedance
across its frequency spectra.
To show the effect of a frequency-dependent equivalent, the test system of
Figure 8.16 is used with an inverter effective short circuit ratio of 2. The d.c. link
is represented by the CIGRE benchmark model [7].
A three-phase solid short circuit fault is applied at the inverter terminal. Three cases
are considered, the first being the entire system represented by the detailed solution.
The second and third cases relate to the hybrid solution interfaced at the converter
bus, one case with a fundamental frequency Thevenin representation of the stability
program in the detailed solution, and the other with a frequency-dependent equivalent.
The results of the inverter terminal voltages for the three cases are plotted in
Figure 8.17, and they show that the benchmark EMTDC case and the frequency-
dependent equivalent case are identical. The fundamental frequency equivalent case
(Figure 8.17(b)) shows more distortion and prolonged effects from the disturbance
than the benchmark EMTDC case.
These results show the inadequacy of a fundamental frequency equivalent at the
converter terminal, whereas the use of a frequency-dependent equivalent is perfectly
adequate.

8.8 Structure of the Hybrid Program


The hybrid algorithm interface is shown in Figure 8.18. It contains the five possible
states summarized in Table 8.1. The default state is 0 for the first time that the algorithm
is called. This occurs once TS has reached steady state electromechanical equilibrium
and is ready for interfacing, prior to a system disturbance.
Under the 0 state, EMTDC is called for a full fundamental period ahead of TS and
its interfacing variables extracted and converted to positive sequence. The state is now
set to -1.
Under the - 1 state, TS is run for one half of a fundamental period and the variables
at its interface location are compared with those of EMTDC. This is simply to check
and ensure that the data in both files is correctly set up. The TS network is then split
318 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

-1.4 I i i i i i i i i

(a) EMTDC

1.4 , i

1.2
1 .o
0.8

-& 0.6
0.4
0.2
g) 0.0
-0.2
>o -0.4
-0.6
-0.8
-1 .o
-1.2
-1.4 I 1 , , 1 1 1 1 1

(b) TSE Hybrid-Fundamental frequency equivalent

1.4 , f

0.96 1.00 1.02 1.04 1.06 1.08 1.10 1.12 1.14 1.16 1.18
Time (s)
-
(c) TSE Hybrid Frequency dependent equivalent

Figure 8.17 Inverter ax. voltage


8.8 STRUCTURE OF THE HYBRID PROGRAM 319

Table 8.1 Interfacing states


State Event
-1 Isolate TS network ready to interface with EMTDC
0 Initial time through interfacing routine
1 Normal interaction between TS and EMTDC
2 System network change signalled-run EMTDC for a full period
past the fault and return information to TS
3 TS catching up with EMTDC after a system network change

No
1
Yes 1 1
Compare variables I
s = Ts+stepl
I
I

No 1
1 and2

Set up equivalent
circuits

Yes State=l
Ts=O

Call EMTDC for a

Convert EMTDC olp

Ts=Ts+steol

A 0

Figure 8.18 TSE hybrid algorithm. [step 11 TS program step length. [Ts] TS step length
counter. [Tend] Minimum finish time for EMTDC. [K] Time counter for EMTDC and TS
comparison of variables. [Tc] Time over which comparison must be within tolerance
320 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

Yes I

eriod of TI2

variables to TS sys. 1

State=3
Ts=O
Ts=Ts+stepl
~~~~~~~~ ~

Extract TS system
2 variables
lr
Compare variables of
TS sys.2 and EMTDC

+=+&

I- K=K+stepl
No
i

Terminate EMTDC
*
Reconnect TS
systems 1 and 2
I I

C D

Figure 8.18 (Continued)


8.8 STRUCTURE OF THE HYBRID PROGRAM 321

IC
stop

1 f
Pass TS,System 1 1 f
variables to EMTDC
% Ts=Ts+step 1

Ts=O

Figure 8.18 (Continued)

into systems 1 and 2, and Norton equivalent circuits set up in each system. The state
is now set to 1, representing normal interfacing conditions, and TS is solved until its
time is concurrent with EMTDC.
Under the normal interfacing of state 1, variable information is given to EMTDC
from TS to update its equivalent circuit. The equivalent circuit of system 1 in the TS
model of system 2 is updated in the same way. EMTDC is then called for one half a
period and the information measured over the last fundamental period of its simulation
. . . . . . , -
used to update the equivalent circuit or system L in 15.
a . rn"

In the interface data file, a time can be specified as a minimum termination time
for EMTDC, after any network disturbance is cleared. Once this time is reached in
the program, the interface variables of both system 2 models are compared at each
TS time-step. If they correlate within a predefined tolerance over a set period, then
EMTDC can be terminated and the TS representation of system 2 once again takes
river nurino
" 1 1 1 . ..." nerinrl
b the
Y"1
1.. y, .
..,.. thpv
-.."J rirp hpino
I.w" -..--.-..-,if_- at anv,time
-... rhprkpd I---f --.-.-
the
_-.-vsriahleq
.
-_---_-I -_- outside
are -
the specified tolerance, the checking period time is restarted.
322 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

When a network admittance change is about to occur, such as a fault being applied
or removed, the interfacing routine is entered with a state value of 2. At this point
in time, TS is concurrent with EMTDC. The TS interface variables are passed to
EMTDC, which is then run for a full period from the network discontinuity time. It
is not accurate to apply the analysis window over the network discontinuity, so the
staggered window process must be restarted. The variables from EMTDC are passed
back to TS at the time of the disturbance and the state is changed to 3.
State 3 allows TS to catch up to EMTDC before the normal interfacing procedure
of state 1 recommences.

8.9 Test System and Results


The d.c. link of the test system is based on the CIGRE benchmark model [7] described
in Appendix 111. However, the rectifier a.c. system is modified as shown in Figure 8.19
to be representative of the New Zealand central South Island primary transmission
system. The parameters for this test system are also given in Appendix 111. A three-
phase short circuit is applied to the rectifier terminals of the d.c. link (at time 1.7 s) and
cleared three cycles later. For the EMTDC studies, the a.c. system at the faulted (recti-
fier) end is replaced by a frequency-dependent equivalent circuit, derived as explained
in section 6.4.3, which produced the harmonic impedances, shown in Figure 8.20.

8.9.1 Response of the individual programs


The differences between the TS and EMTDC programs when used independently from
each other are illustrated in Figure 8.21
Following fault clearance the rectifier is unblocked at t = 1.8s and a few ms later
the inverter takes over conduction from its by-pass valves. After an initial current
inrush into the d.c. line capacitance, the power setting is ramped up. At t = 1.97 s the
inverter is restated and the link current drops, a reduction which is more pronounced
and stays longer in the EMTDC solution.

Clyde Twizel

Livingston Aviemore

Figure 8.19 Test system on the a.c. side of the rectifier station
8.9 TEST SYSTEM AND RESULTS 323

2 3000
-3 2000-
r
o 2500- t
-6
.-
1500-
ii

m
E 1000-
8 500 -
2
0

1.20
7 1.00
n
0.80-
C
0.60-
3
0.40-
0
0 0.20-
0.00
.....
I I I I I I I I
1.50 1.65 1.80 1.95 2.10 2.25 2.40 2.55 2.70 2.85 3.00
Time (s)

Figure 8.21 d.c. current response to a three-phase short circuit at the rectifier end

Following the power ramp, the a.c. current returns to its nominal value and after a
small oscillatory settling of the controllers so do the d.c. voltage and current.

8.9.2 TSE hybrid response


The results of the EMTDC and TS simulations show substantial differences and thus
their independent use is inadequate to model the transient stability of the a.c. system.
Figure 8.22(a) displays the rectifier terminal a.c. voltage levels calculated with the
TS and TSE alternatives. The main difference between them is a transient overshoot
in the TSE solution immediately after fault clearance.
The rectifier d.c. currents, displayed for the three solutions in Figure 8.22(b), show
a very similar variation for the TSE and EMTDC solutions, except for the region
between I = 2.03 s and t = 2.14 s but the difference with the TS only solution is very
pronounced.
Figure 8.23 compares the fundamental positive sequence real and reactive powers
across the converter interface for the TS and TSE solutions.
The main differences in real power occur during the link power ramp. The difference
is almost a direct relation to the d.c. current difference between TS and TSE shown
in Figure 8.22(b). The oscillation in d.c. voltage and current as the rectifier terminal
is de-blocked is also evident.
324 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

................ TS only T S E (TS variable)


1.40 I 1
1.20
1 .oo
0.80
0.60
0.40
0.20
0.00

1.20
1.20 -- _I

z 1.00 -
g 0.80-
6 0.60-
0 0.40 -
0.20 -
0.00 I I I I I I I I
1.50 1.65 1.80 1.95 2.10 2.25 2.40 2.55 2.70 2.85 3.00
(b) Time (s)

Figure 8.22 Hybrid response comparisons. (a) Rectifier end a x . voltage. (b) Rectifier end d.c.
current

- TSE TS
1200 , I
1000

58 600
400

0
I I I I I I I I I

d
h
200

0
-200-
-- .......... ...............
~

z. -400-
p8 -600-

.-rg -800-
5 -1Ooo-
I
a -1200 -
-1400 I I I I I I I I I

Figure 8.23 Real and reactive power across interface


8.10 QUASI STEADY-STATE CONVERTER SIMULATION 325

- -
5
50
30 --
- - - - - --
I ,
I ~.‘,~’,-I~,.R-,,-~--.
1,
I -- -_-.- - - - - - -
s -
-a lo-
pi F,
.,i \...!. !..,/’*\,.,..-.
0,
5 -10-
........... ......................................................................
L..’..’....
a
c -30-
E
0
-50

Figure 8.24 Rotor angle swings using TSE

As for the reactive power Q, prior to the fault, a small amount is flowing into
the system due to a surplus MVAr’s at the converter terminal. The fault reduces this
power flow to zero. When the fault is removed and the a.c. voltage overshoots in TSE,
the reactive MVAr’s also overshoot in TSE and since the d.c. link is shut down, a
considerable amount of reactive power flows into the system.
Finally, Figure 8.24 shows the machine angle swings with respect to the Clyde
generator (see test system of Figure 8.19). These indicate that the system is transiently
stable.

8.10 Quasi Steady-state Converter Simulation


When analysing small perturbations and dynamic stability, converter equipment is
expected to operate in a controlled manner almost instantaneously when compared
with the relatively slow a.c. system dynamics. In these cases, it is acceptable to use a
modified steady-state (or quasi steady-state) model, the modifications being due to the
different constraints imposed by the load flow and stability studies. Such a model is
also suitable for representing large rectifier loads during a.c. system disturbances, with
further modifications necessary to represent abnormal rectifier operating modes.
Further to the basic assumptions listed in Chapter 3, the following need to be made
here:
0 The implementation of delay angle control is instantaneous.
0 The transformer tap position remains unchanged throughout the stability study
unless otherwise specified.
0 The direct current is smooth, though its actual value may change during the study.

8.10.1 Rectifier loads


Large rectifier loads generally consist of a number of series andor parallel connected
bridges, each bridge being phase shifted relative to the others. With these configu-
rations, high-pulse numbers can be achieved resulting in minimal distortion of the
326 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

Figure 8.25 Rectifier load equivalent circuit

Reducing aV,,,,cosa

Constant current
control characteristic

;
.
'd* 'd

Figure 8.26 Simple rectifier control characteristic

supply voltage without filtering. Rectifier loads can, therefore, be modelled as a single
equivalent bridge with a sinusoidal supply voltage at the terminals but without repre-
sentation of passive filters. This model is shown in Figure 8.25.
Rectifier loads can utilize a number of control methods which can be modelled using
a controlled rectifier with suitable limits imposed on the delay angle (a)[12].
Static loads Operating under constant current control, the d.c. equations are:
(8.53)

(8.54)

where A is the constant current controller gain and I d s is the normal d.c. current setting
as shown in Figure 8.26.
Constant current cannot be maintained during a large disturbance as a limit of delay
angle will be reached. In this event, the rectifier control specification will become one
8.10 QUASI STEADY-STATE CONVERTER SIMULATION 327

of constant delay angle and Equation (8.54) becomes:

(8.55)

Protection limits and disturbance severity determine the rectifier operating charac-
teristics during the disturbance. Shutdown occurs if I d reaches a set minimum or zero
and the voltage Vload will cause shutdown before the a.c. terminal voltage reaches zero.
The action of the rectifier load system is thus described by Equations (3.24), (3.30),
(3.34), (3.36), (3.37), (8.53) and either (8.54) or (8.55).
Dynamic loads The basic rectifier load model assumes that current on the d.c. side
of the bridge can change instantaneously. For some types of rectifier loads this may
be a valid assumption, but the d.c. load may well have an overall time constant which
is significant with respect to the fault clearing time. In order to examine realistically
the effects which rectifiers have on the transient stability of the system, this time
constant must be taken into account. This requires a more complex model to account
for extended overlap angles, when low commutating voltages are associated with large
d.c. currents.
When the delay angle (a)reaches a limiting value, the dynamic response of the d.c.
current is given by
V d = IdRd + +
Vload Ldpld, (8.56)

where 4 represents the equivalent inductance in the load circuit. Substituting for Vd
using Equation (3.24) gives

where T d c = L d / R d .
The controller time constant may also be large enough to be considered. However,
in transient stability studies where large disturbances are usually being investigated,
faults close to the rectifier load force the delay angle (a)to minimum very quickly.
Provided the rectifier load continues to operate, the delay angle will remain at its
minimum setting throughout the fault period and well into the post-fault period until
the terminal voltage recovers. The controller will, therefore, not exert any significant
control over the d.c. load current. Ignoring the controller time constant can, therefore,
be justified in most studies.

Abnormal modes of converter operation The slow response of the d.c. current
when a large disturbance has been applied to the ax. system can cause the rectifier to
operate in an abnormal mode.
After a fault application near the rectifier, the near normal value of d.c. current
( I d ) needs to be commutated by a reduced a.c. voltage. This causes the commutation
angle ( b )to increase and it is possible for it to exceed 60". This mode of operation is
beyond the validity of the equations and to model the dynamic load effects accurately
it is necessary to extend the model.
328 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

The full range of rectifier operation can be classified into four modes [13]:
Mode 1 -Normal operation. Only two valves in the bridge are involved in simulta-
neous commutation at any one time. This mode extends up to a commutation angle
of 60".
Mode 2 -Enforced delay. Although a commutation angle greater than 60" is desired,
the forward voltage across the incoming thyristor is negative until either the previous
commutation is complete or the firing angle exceeds 30". In this mode, p remains
at 60" and a ranges up to 30".
Mode 3 -Abnormal operation. In this mode, periods of three-phase short circuit and
d.c. short circuit exist when two commutations overlap. During this period, there
is a controlled safe short circuit which is cleared when one of the commutations is
complete. During the short-circuit periods, four valves are conducting. Commutation
cannot commence until 30" after the voltage crossover.
Mode 4 -Continuous three-phase and d.c. short circuit caused by two commutations
taking place continuously. In this mode, the commutation angle is 120" and the a.c.
and d.c. current paths are independent.

The waveforms for these modes are shown in Figure 8.27, and Table 8.2 summarizes
the conditions for the different modes of operation. Equations (3.24) and (3.30) do not
apply for a rectifier operating in Mode 3 and they must be replaced by:

(8.58)

and
a
Id = -Vterm(c0sa'- cos y'). (8.59)
45xc
Fourier analysis of the waveform leads to the relationship between ax. and d.c.
current given by Equation (3.32), where the factor k is now:

(8.60)

where
a' = CY - 30"
and
y' = y + 30". (8.61)
A graph showing the value of k for various delay angles (a)and commutation angles
( p ) is shown in Figure 8.28.

Identification of operating mode The mode in which the rectifier is operating can
be determined simply by use of a current factor K I . The current factor is defined as:

(8.62)

Substitution in this, using the relevant equations, yields limits for the modes.
8.10 QUASI STEADY-STATE CONVERTER SIMULATION 329

6 2 4 6

2 4 6

Figure 8.27 Rectifier voltage waveforms showing different modes of operation. (a) Mode 1,
g < 60"; (b) Mode 2, p = 60" with enforced delay al; (c) Mode 3, p z 60" with short-circuit
period a2

Table 8.2 Rectifier modes of operation


Mode Firing angle Overlap angle
1 0" 5 a 5 90" 0" 5 /A 5 60"
2 0" 5 a 5 30" 60"
3 30" 5 (Y 5 90" 60" 5 p < 120"
4 30" 5 a 5 90" 120"
330 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

?
I 1 I 1

1.2 -

Figure 8.28 Variation of k in expression I , = k ( 3 , h / n ) I d

Mode 1:
K I 5 COS(~O"- a ) (8.63)
and
K I 5 2 cos a for rectifier operation.
Mode 2.
(8.64)
L

Mode 3:
2
K I < - when a 5 30°,
43
2
K I < -cos(ct - 30") when ct > 30". (8.65)
43
Mode 4:
2
K I = - when a 5 30°,
43
2
K I = -cos(a - 30") when a > 30" (8.66)
43
This can be demonstrated in the curve of converter operation shown in Figure 8.29.
It can thus be seen that the mode of operation can be established prior to solving
for the rectifier load equations at every step in the solution.

8.10.2 d.c. link


Provided that it can be safely assumed that a d.c. link is operating in the quasi steady-
state (QSS) Mode 1, the equations developed for converters in Chapter 3 can be used.
That is, the converters are considered to be controllable and fast acting so that the
8.10 QUASI STEADY-STATE CONVERTER SIMULATION 331

Mode 4

Delay angle, a

Figure 8.29 Converter operation [ 141

normal steady-state type of model can be used at each step in the transient stability
study.
The initial steady-state operating conditions of the d.c. link will have been deter-
mined by a load flow and, in this, the control type, setting and margin will have been
established.
During the solution process at each iteration, the control mode must be established.
This can be done by assuming Mode 1 (i.e. with the rectifier on C.C. control) and,
by combining Equations (3.39), (3.45), (8.53) and (8.54), a d.c. current can be deter-
mined as:
Idsr - [(3fi/n)aiVtermi cos Yicl/Ar
I d mode I = (8.67)
+
1 (Rd - (3/n)xci)/Ar
Assuming this current to be valid, then d.c. voltages at each end of the link can be
calculated using Equations (3.24) and (3.39).
The d.c. link is operating in mode 2 (i.e. with the inverter on C.C. control) i f
Vdr mode I - Vdi mode I =< 0. (8.68)
The d.c. current for mode 2 operation is given by

For constant power control, under control mode 1, the d.c. current may be determined
from the quadratic equation:

where Pds is the setting at the electrical mid-point of the d.c. system, i.e.
pds = (Pdsr + pdsi)/2* (8.7 1)
332 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

Table 8.3 Current setting for constant power control


from quadratic equation

Id I Id2

Within Outside Id I
Outside Within Id2
Within Within Greater of I d , and Ir12
Greater Greater I d max
Greater Less I d max
Less Greater I d max
Less Less 0
Within = within the range / d m i n to / d m a x ;
Outside = outside the range l d m i n to l d m a x ;
Greater c Greater than I d m a x ;
Less = Less than I d m i " .

The correct value for Idm&I can then be found from Table 8.3. Control mode 2
is determined using Equation (8.68) and in this case the following quadratic equation
must be solved:
k r z : mode 2 - k v l d mode 2 - P d marg - p d s = 0, (8.72)
where

(8.73)

(8.74)

If the link is operating under constant power control but with a current margin then
for control mode 2:

It is possible for the d.c. link to be operating in control mode 2 despite satisfying the
inequality of Equation (8.68). This occurs when the solution indicates that the rectifier
firing angle (a,)is less than the minimum value ( a r m i n ) . In this case, the delay angle
should be set to its minimum and a solution in mode 2 is obtained.
It is also possible that when the link is operating close to the changeover between
modes, convergence problems will occur in which the control mode changes at each
iteration. This can easily be overcome by retaining mode 2 operation whenever detected
for the remaining iterations in that particular time step.
d.c. power modulation It has been shown in the previous section that under the
constant power control mode, the d.c. link is not responsive to a.c. system terminal
conditions, i.e. the d.c. power transfer can be controlled disregarding the actual a.c.
voltage angles. Since, generally, the stability limit of an ax. line is lower than its
thermal limit, the former can be increased in systems involving d.c. links, by proper
utilization of the fast converter controllability.
The d.c. power can be modulated in response to a.c. system variables to increase
system damping. Optimum performance can be achieved by controlling the d.c. system
8.10 QUASI STEADY-STATE CONVERTER SIMULATION 333

so as to maximize the responses of the a.c. system and d.c. line simultaneously
following the variation of terminal conditions.
The dynamic performance under d.c. power modulation is best modelled in three
separate levels [14]. These levels, illustrated in Figure 8.30, are (a) the a.c. system
controller, (b) the d.c. system controller and (c) the ax.-d.c. network.

(a) The ax. system controller uses ax. andor d.c. system information to derive the
current and voltage modulation signals. A block diagram of the controller and
a.c.-d.c. signal conditioner is shown in Figure 8.31.
(b) The d.c. system controller receives the modulation signals AI and A E and the
steady-state specifications for power PO current I 0 and voltage Eo. Figure 8.32(a)
illustrates the power controller model, which develops the scheduled current
setting; it is also shown that the current order undergoes a gradual increase during
restart, after a temporary blocking of the d.c. link.
The rectifier current controller, Figure 8.32(b), includes signal limits and rate
limits, transducer time constant, bandpass filtering and a voltage dependent current
order limit (VDCOL).
The inverter current controller, Figure 8.32(c), includes similar components plus a
communications delay and the system margin current (Z,,,). Finally, the d.c. voltage
controller, including voltage restart dynamics, is illustrated in Figure 8.32(d).
(c) The d.c. current I d and voltage E d derived in the d.c. system controller constitute
the input signals for the a.c.-d.c. network model which involves the steady state

ffY
DC control modes Ed 'd 'd, ' ' d i
I_ AClDC interface Qr Ql

(a) (b) (4
Figure 8.30 a.c.-d.c. dynamic control structure: (i) a.c. system controller, (ii) d.c. system
controller, (iii) ax.-d.c. network

Figure 8.31 a.c. system controller


334 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

N(P)
3~- '07

D(P)

Figure 8.32 d.c. system controller. (a) Power controller; (b) Rectifier current controller;
(c) Inverter current controller; (d) d.c. voltage controller

solution of the d.c. system (neglecting the d.c. line dynamics which are included
in the d.c. system controller). Here the actual a.c. and d.c. system quantities are
calculated, i.e. control angles, d.c. current, voltage, active and reactive power.
The converter a.c. system constraints are the open circuit secondary voltages Ear
and Eai.

8.10.3 Representation of converters in the network


Rectifiers The static load rectifier model can be included in the overall solution
of the transient stability program in a similar manner to the basic loads described in
Chapter 7.
8.10 QUASI STEADY-STATE CONVERTER SIMULATION 335

From the initial load flow, nominal bus shunt admittance (yo) can be calculated
for the rectifier. This is included directly into the network admittance matrix [Y].The
current injected into the network in the initial steady state is, therefore, zero. In general:

where
?;*
- 30
(8.77)
Yo=Ivrermo12
and
(8.78)

The static load rectifier model does not depart greatly from an impedance charac-
teristic and is well behaved for low terminal voltages, the injected current tending to
zero as the voltage approaches zero. Figure 8.33 compares the current due to a rectifier
with that due to a constant impedance load. As the injected current is never large, the
iterative solution for all ax. conditions is stable.
When the rectifier model is modified to account for the dynamic behaviour of the
d.c. load, its characteristic departs widely from that of an impedance. Immediately after
a fault application, the voltage drops to a low value but the injected current magni-
tude does not change significantly. Similarly, on fault clearing, the voltage recovers
instantaneously to some higher value while the current remains low.

Ip t

I
0.5 1.o
Vterm
linj 4
~F.L. -

Figure 8.33 Difference between impedance and static load rectifier characteristic (for
vload # 0)
336 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

When the load characteristic differs greatly from that of an impedance, a sequential
solution technique can exhibit convergence problems, especially when the voltage is
low. With small terminal voltages, the a.c. current magnitude of the rectifier load is
related to the d.c. current but the current phase is greatly affected by the terminal
voltage. Small voltage changes in the complex plane can result in large variations of
the voltage and current phase angles.
To avoid the convergence problems of the sequential solution, an alternative algo-
rithm has been developed [ 151. This combines the rectifier and network solutions into
a unified process. However, it does not affect the sequential solution of the other
components of the power system with the network.
The basis of this approach is to reduce the a.c. network, excluding the rectifier, to an
equivalent Thevenin source voltage and impedance as viewed from the primary side
of the rectifier transformer terminals. This equivalent of the system, along with the
rectifier, can be described by a set of non-linear simultaneous equations which can be
solved by a standard Newton-Raphson algorithm. The solution of the reduced system
yields the fundamental a.c. current at the rectifier terminals.
To obtain the network equivalent impedance, it is only necessary to inject 1 p.u.
current into the network at the rectifier terminals while all other nodal injected currents
are zero. With an injected current vector of this form, a solution of the nodal network
Equation (7.48) gives the driving point and transfer impedances in the resulting voltage
vector:
[Z] = [V']= [T]-l[7;,1, (8.79)

(8.80)

The equivalent circuit shown in Figure 8.34 can now be applied to find the rectifier
current (7,) by using the Newton-Raphson technique.
The effect of the rectifier on the rest of the system can be determined by superpo-
sition:
+
[VI= [TI [ZII,, (8.81)
where
[TI= [r]-"l;j] (8.82)

and [7Ynj]are the injected currents due to all other generation and loads in the system.
If the network remains constant, vector [z] is also constant and thus only needs
re-evaluation on the occurrence of a discontinuity.
Thus, the advantages of the unified and sequential methods are combined. That is,
good convergence for a difficult element in the system is achieved while the program-
ming for the rest of the system remains simple and storage requirements are kept low.
8.10 QUASI STEADY-STATE CONVERTER SIMULATION 337

'I
Rd

vd=0

Dynamic
load element

2 Vload

Figure 8.34 Equivalent system for Newton-Raphson solution

The equivalent system of Figure 8.34 contains seven variables ( V , e r m , I , , 8, @, a,


Vd and I d ) . With these variables, four independent equations can be formed. They are
Equation (3.24) and:
Vl!@- V t e r m L 8 - Z,hl!(IpL$' = 0, (8.83)
Vdld - &aVtermlp cOS(6 - 9) = 0. (8.84)
Equation (8.83) is complex and represents two equations. Substituting for V d and
I , using Equations (3.34) and (8.53) reduces the number of variables to five. A fifth
equation is necessary and with constant current control, i.e. with the delay angle (a)
within its limits, this can be written as
Id - I d s p = 0. (8.85)
Equation (3.24), suitably reorganized, and Equations (8.83)-(8.85)represent [ F ( X ) ]=
0 of the Newton-Raphson process and
[XIT = [Er, 8, a,I d ] . $9 (8.86)
When the delay angle reaches a specified lower limit (omin),the control specification,
given by Equation (8.85) changes to
a - amin = 0. (8.87)
Equation (3.24) is no longer valid. The d.c. current ( I d ) is now governed by the
differential Equation (8.57). If the trapezoidal method is being used, this equation
can be transformed into an algebraic form similar to that described in Chapter 7 .
Equation (3.24) is replaced by:
Id = kaE, cosa - kb = 0. (8.88)
The variables ka and kb contain information from the beginning of the integration
step only and are thus constant during the iterative procedure:
ka = h / ( 2 + kc h), (8.89)
338 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

where
(8.91)

t represents the time at the beginning of the integration step and h is the step length.
Commutation angle p is not explicitly included in the formulation, and since these
equations are for normal operation, the value of k in Equation (3.35) is close to unity
and may be considered constant at each step without loss of accuracy. On convergence,
p may be calculated and a new k evaluated suitable for the next step.
In Mode 3 operation, the value of k becomes more significant and for this reason
the number of variables is increased to six to include the commutation angle p. The
equations [ F ( X ) ] = 0 for the Newton-Raphson method in this case are:

VLp - VtemL8 - (8.92)

Jz
-uVterm
A
cOs(8 - @ ) f ( p )- -aVterm cosa!' + -3Inxd c = 0, (8.93)
7I 7I
Id - ka * aVtermcos CY'- kb = 0, (8.94)

cos(a!
AXc
+ p + 30) - cos(Y' + -Idavterm = 0, (8.95)

a! - CY,ln = 0. (8.96)
Although k can be calculated explicitly, a linearized form of Equation (8.60) obtained
for 01 = 30" can be used to simplify the expression. In the range 60" < 1 < 120", the
value of k can be obtained from:
= 1.01 - 0.0573 1,
f(1) (8.97)
where p is measured in radians.
In Mode 4, the a s . and d.c. systems are both short circuited at the rectifier and
operate independently. In this case, the system equivalent of Figure 8.34 reduces to
that shown in Figure 8.35. The network equivalent can be solved directly and the d.c.
current is obtained from the algebraic form of the differential Equation (8.57).

Dynamic
element

Figure 8.35 Rectifier load equivalent in Mode 4 operation


8.11 STATIC VAR COMPENSATION SYSTEMS 339

d.c. links The problems associated with dynamic rectifier loads do not occur when
the d.c. link is represented by a quasi steady-state model. Each converter behaves
in a manner similar to that of a converter for a static rectifier load. A nominal bus
shunt admittance (Lo) is calculated from the initial load flow for both the rectifier and
inverter ends and injected currents are used at each step in the solution to account for
the change from steady state calculated from Equation (8.76). Note that the steady-
state shunt admittance at the inverter (YOoi)will have a negative conductance value as
power is being supplied to the network. This is not so for a synchronous or induction
generator as the shunt admittance serves a different purpose in these cases.

8.10.4 Inclusion of converters in the transient stability


program
A flow diagram of the unified algorithm is given in Figure 8.36. It is important to note
that the hyperplanes of the functions used in the Newtonian iterative solution process
are not linear and good initial estimates are essential at every step in the procedure. A
common problem in converter modelling is that the solution converges to the unrealistic
result of converter reactive power generation. It is, therefore, necessary to check against
this condition at every iteration. With integration step lengths of up to 25 mS, however,
convergence is rapid.

8.11 Static VAR Compensation Systems


Static VAR compensation systems (SVS) of large ratings are now in common use to
control the voltage at points remote from power generation. They are also installed to
assist system stability and, therefore, to consider them as a fixed shunt element can
produce erroneous information in a transient stability study.
A composite static VAR compensator, shown in Figure 8.37, has been suggested by
a CIGRE Working Group [16]. The model is not overly complex as this would make
data difficult to obtain and would be incompatible with the overall philosophy of a
multi-machine transient stability program. The SVS representation can be simplified
to any desired degree, however, by a suitable choice of data.
The basic control circuit consists of two lead-lag and one lag transfer function
connected serially. The differential equations describing the action of the control circuit
with reference to Figure 8.37 are:

Although electronically produced, the dead band may be considered as a physical


linkage problem, as shown in Figure 8.38(a). In this example, the input ( x ) and output
( y ) move vertically. The diagram shows the initial steady-state condition in which x
and y are equal. The input ( x ) may move in either direction by an amount Db/2 before
y moves. Beyond this amount of travel, y follows x, lagging by &/2 as depicted in
Figure 8.38(b). The effect of a dead band can be ignored by setting Db to zero.
340 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

TS programme Rectifier subroutines

Calculate constants for


differential equation

( Chec for switching } I .a-


1
I I No Obtain new 1
a 1
Extract th_eveninvoltag-
Thevenin imDendancd

es from [ V ]

II

,
L

t Calculate new initial


vuT
Calculate K, if mAAn lconditons and mode 1
I operation I

‘3
I v
I Calculate mismatches
I
I Convergence Yes
I
I
I
I
I I
given mode of operation

(-%
I
i lter +
I 11ter+ll

I
I
I Calculate rectifier variables

I t
I
II - lter > 15
+
Yes
)
I
I Error messages I A.C. current on [ V ]
I
!
#
.
(Convergence?) I I
I
Figure 8.36 Unified algorithm flow diagram

Stepped output permits the modelling of SVS when discrete capacitor (or inductor)
blocks are switched in or out of the circuit. It is usual to assume that all blocks are of
equal size. During the study, the SVS operates on the step nearest to the control setting.
Iterative chattering can occur if the control system output (B3) is on the boundary
between two steps. The simplest remedy is to prevent a step change until B3 has
moved at least 0.55BSepfrom the mean setting of the step.
8.1 1 STATIC VAR COMPENSATION SYSTEMS 341

Initial
controlling

Y
vohage ( p u ~ "9v-t
Initial
susceptanm (pu)

Dead band Control system Susceptance


limits and steps

Figure 8.37 Composite static VAR compensation system (SVS) model

(a) Movement
1 x-r+z4
Movement
input x
\n

Figure 8.38 Dead band analogy and effect. (a) Physical analogy of dead band; (b) The effect
of a dead band on output

The initial MVAR loading of the SVS should be included in the busbar loading
schedule data input. However, it is possible for an SVS to contain both controllable
and uncontrollable sections (e.g. variable reactor in parallel with fixed capacitors or
vice versa). It is the total MVAR loading of the SVS which is, therefore, included in
the busbar loading. Only the controllable part should be specified in the SVS model
input and this is removed from the busbar loading leaving an uncontrollable MVAR
load which is converted into a fixed susceptance associated with the network.
Note that busbar load is assumed positive when flowing out of the network. The
sign is, therefore, opposite to that for the SVS loading.
In order to clarify this, consider an overall SVS operating in the steady state, as
shown in Figure 8.39(a). The busbar loading in this case must be specified as -50
MVAR and it may be varied between -10 MVAR and -80 MVAR provided the
voltage remains constant.
The SVS may be specified in a variety of ways, some more obvious than others,
the response of the system being identical. Three possible specifications are given in
Table 8.4. In the first example, the network static load will be +20 MVAR, while in
the second case the static load will be zero. The third example may be represented by
an overall SVS, as shown in Figure 8.39(b).
342 8 SYSTEM STABILITY UNDER POWER ELECTRONIC CONTROL

(limits
70MVAR0 and

70 MVAR)
12z 80 MVAR

Figure 8.39 Example of an overall SVS controllable and uncontrollable sections. (a) Example
of overall SVS using controllable capacitors; (b) Alternative to overall SVS in (a) using a
controllable reactor

Table 8.4 Examples of MVAR loading specification for SVS shown in


Figure 8.39(a)
Example Initial loading Maximum limit Minimum limit
1 70 100 30
2 50 80 10
3 -30 0 -70

It is convenient, when specifying the initial steady-state operation, to use MVAR.


However, this is a function of the voltage and, hence, all MVAR settings must be
converted to their equivalent per unit susceptance values prior to the start of the stability
study.

8.11.1 Representation of SVS in the overall system


The initial MVAR loading of the SVS is converted into a shunt susceptance (Bo) and
added to the total susceptance at the SVS terminal busbar. During the system study, the
deviation from a fixed susceptance device is calculated ( B 4 ) and a current equivalent
to this deviation is injected into the network.
A reduction in controlling voltage VSV will cause the desired susceptance B4 to
increase. That is, the capacitance of the SVS will rise and the MVAR output will
increase.
The injected current (Ti,,,) into the network is given by:

(8.101)
8.12 REFERENCES 343

where
Y =0 +jB4.
Although not necessary for the solution process, the MVA output from the SVS into
the system is given by:

(8.102)

8.12 References
1, Kulicke, B. Netomac digital program for simulating electromechanical and electromagnetic
transient phenomena in a x . systems, (198 1). Siemens Aktienngesel-lschaft, E15/1722- 101.
2. Woodford, D A, (1985). Validation of digital simulation of d.c. links, IEEE Transactions on
Power Apparatus and Systems, PAS-104 (9), pp. 2588-2595.
3. Anderson, G W J, Arnold, C P, Watson, N R and Arrillaga, J, (1995). A new hybrid
a.c. -d.c. transient stability program, International Conference on Power Systems Transients
(IPST), pp. 535-540.
4. Anderson, G W J, (1995). Hybrid simulation of a.c.-d.c. power systems, Ph.D. thesis,
University of Canterbury, New Zealand.
5 . Eguiluz, L I and Arrillaga, J, (1995). Comparison of power definitions in the presence of
waveform distortion, International Journal of Electrical Engineering Education, 32 (2),
pp. 141-153.
6. Heffernan, M D, Turner, K S, Arrillaga, J and Arnold, C P, (1981). Computation of
a.c. -d.c. system disturbances -Parts I, I1 and 111, Transactions on Power Apparatus and
Systems, PAS-100 (1 l), pp. 4341 -4363.
7. Szechtman, M, Weiss, T and Thio, C V, (1991), First benchmark model for HVDC control
studies, ELECTRA (133, pp. 55-75.
8. Clarke, E, (1950). Circuit Analysis of a.c. Power Systems- Vol. 11, John Wiley and Sons,
New York.
9. Kimbark, E W, (1968). Power System Stability- Vol. IIIs Synchronous Machines, Dover
Publications, New York.
10. Reeve, J and Adapa, R, (1988). A new approach to dynamic analysis of a.c. networks
incorporating detailed modelling of d.c. systems -Parts I and 11, IEEE Transactions, PD-3
(4), pp. 2005-2019.
1 I . Watson, N R, (1987). Frequency-dependent a.c. system equivalents for harmonic studies
and transient converter simulation, Ph.D. thesis, University of Canterbury, New Zealand.
12. Arnold, C P, Turner, K S and Arrillaga, J, (1980). Modelling rectifier loads for a multi-
machine transient-stability programme, IEEE Transactions on Power Apparatus and Systems,
PAS-99 (l), pp. 78-85.
13. Giesner, D B and Arrillaga, J, (1970). Operating modes of the three-phase bridge converter,
International Journal of Electrical Engineering Education, 8, pp. 373-388.
14. IEEE Working Group on Dynamic Performance and Modeling of d.c. Systems, (1980).
Hierarchical structure.
15. Turner, K S , (1980). Transient stability analysis of integrated a.c. and d.c. power systems,
Ph.D. thesis, University of Canterbury, New Zealand.
16. CIGRE Working Group 31-01, (1977). Modelling of static shunt VAR systems for system
analysis, ELECTRA (Sl), pp. 45-74.
Appendix I
FAULT LEVEL DERIVATION

1.1 Short Circuit Analysis


The main object of short circuit analysis is to calculate fault currents and voltages
for the determination of circuit-breaker capacity. This information also permits the
derivation of the MVA fault level, and thus the system strength, at any point of the
network.
A preliminary stage to the analysis is the collection of appropriate data specifying
the system to be analysed in terms of prefault voltage, loading and generating condi-
tions. Such data are then processed to form a nodal equivalent network constituted by
admittances and injected currents.
The equivalent circuits of loads, lines and transformers, discussed in Chapter 2, are
directly applicable here. The generators can be represented by a constant voltage EM
behind an approximate machine admittance y M , the value of which depends on the time
of the calculation from the instant of fault inception. This is illustrated in Figure I. l(a).
When analysing the first two or three cycles following the fault, the subtransient
admittance of the machine is normally used, whilst for longer times, it is more appro-
priate to use the transient admittance. The machine model, illustrated in Figure I. l(a), is
then converted to a nodal equivalent by means of Norton's theorem, which changes the
voltage source into a current source injected at the bus j, as shown in Figure I.l(b).
This is most effective, as otherwise a further node at j' is necessary to define the
machine admittance y'.
The injected nodal current is given by

Ij = yMEM
J j ' (1.1)

where
rM

so that
Ij =yrvj +I?. (1.3)

17 is the current required at the voltage V, to produce the machine power Py + jQy,
so
=
(IY)*v~ PY + j@'. (1.4)
346 APPENDIX I

(i1
I

Figure 1.1

Thus, from the load flow data of P M ,QM and VM, we may calculate the injected nodal
current Z j as
Ij = yj
M
vj +
PY - j@'
Vf

1.1.1 System equations


Let us take as a reference the small system of Figure 1.2. Each element is converted to
its nodal equivalent. These are connected together, as shown in Figure 1.3, and finally
simplified to the equivalent circuit of Figure 1.4.
The following equations may then be written for the network of Figure 1.4:

II = Y l l V l + Yl2(VI - v2>, (1.6)


12 = Y12(v2 - v l ) + Y22v2 + Y23(v2 - v 3 ) f y 2 4 ( v 2 - v 4 > , (1.7)
I 3 = Y 2 3 W 3 - V 2 ) + Y33V3 + Y 3 4 W 3 - V4)> (1.8)
1 4 = Y 2 4 ( v 4 - v 2 > + Y 3 4 ( v 4 - v 3 ) + Y 4 4 v 4 + Y45(V4 - V S ) , (1.9)
1 5 = Y 4 5 W 5 - V 4 ) + YSSV59 (1.10)

Figure 1.2 Example of small power system


1.1 SHORT CIRCUIT ANALYSIS 347

-
T - ,
II II
T. I.: zr
.~ -I

I- ff
.~
..A.

vvvv

-I -
I":
Y23 13
Y12

I
Y22 - Y33

Figure 1.4 Final equivalent

-yI1 y21 y3l y4I y51- 'v1'


y12 y22 y32 y42 y52 v2
= y13 y23 y33 y43 y53 v3 7 (I. 11)
y14 y24 y34 y44 y54 v4
-yl5 y25 y35 y45 y55- - v 5 -
348 APPENDIX I

- 4

Node impendance
matrix
A
k
- c -
Vi
vk

Figure 1.5 Thevenin equivalent of prefault system

where
~ i=i C yi, ~ i ,= -yij i # j.
j

Equation (I. 11) is usually written as


“1 = [YI * [VI (1.12)
where [I] and [V] are the current and voltage vectors and [Y] is the nodal admittance
matrix of the system of Figure 1.2.
The nodal admittance equation is inefficient, as it requires a complete iterative solu-
tion for each fault type and location. Instead, equation (1.12) can be written as
[V] = [Y]--’ [I]
= [Z] * [Z]. (1.13)
This equation uses the bus nodal impedance matrix [Z] and permits using the Thevenin
equivalent circuit, as illustrated in Figure 1.5, which provides a direct solution of the
fault conditions at any node. However, the use of conventional matrix inversion tech-
niques results in an impedance matrix with non-zero terms in every position Z i j .
The sparsity of the [Y] matrix may be retained by using an efficient inversion
technique and the nodal impedance matrix can then be calculated directly from the
factorised admittance matrix.

1.1.2 Fault calculations


From the initial machine data, the values of [I] are first calculated from equation (1.5)
using one per unit voltages. These may now be used to obtain a better estimate of [V],
the prefault voltage at every node from equation (1.13). If the initial data are supplied
from a load flow, this calculation will not make any difference.
The program now has sufficient information to calculate the voltages and currents
during a fault.
From Figure 1.6, the voltage at the fault bus k is:
vf = ZfIf, (I. 14)
I. 1 SHORT CIRCUIT ANALYSIS 349

Figure 1.6

[/j
V n
-
-zll
z21

zkl

-znl
212
222
.
zk2
.
Zn2
.
*
.
*
Zlk
Z2k
.
zkk
.
Znk
*

*
.

*
.
ZIn'
Z2n
.
zkn
.
znn-
*
-11'
12

Ik

-In-
*
(I. 15)

This equation describes the voltage at bus k prior to the fault. During a fault, a large
fault current I f flows out of bus k. Including this current in equation (1.16) and using
equation (1.14) gives:

vf =Z f I f =ZklIl + ' * * + Z k k l k + + Zk-In - - Z k k I f .


' * (I. 17)
or
ZfIf =v k -ZkkIf. (I. 18)
and so the fault current is given directly by:

(1.19)

Multiplying both sides by 3vk gives the MVA fault level, i.e.

(1.20)

or
(1.21)
Appendix 11
NUMERICAL INTEGRATION
METHODS

11.1 Introduction
Basic to the computer modelling of power system transients is the numerical integration
of the set of differential equations involved. Many books have been written on the
numerical solution of ordinary differential equations, but this appendix is restricted to
the techniques in common use for the dynamic simulation of power system behaviour.
It is, therefore, appropriate to start by identifying and defining the properties required
from the numerical integration method in the context of power system analysis.

11.2 Properties of the Integration Methods


11.2.1 Accuracy
This property is limited by two main causes, i.e. round-off and truncation errors. Round-
off error occurs while performing arithmetic operations and is due to the inability of the
computer to represent numbers exactly. A word length of 48 bits is normally sufficient
for scientific work and is certainly acceptable for transient stability analysis. When the
stability studies are carried out on computers with a 32-bit word length, it is necessary
to use double precision on certain areas of the storage to maintain adequate accuracy.
The difference between the true and calculated results is mainly determined by the
truncation error, which is due to the numerical method chosen. The true solution at
any one point can be expressed as a Taylor series based on some initial point and
by substituting these into the formulae, the order of accuracy can be found from the
lowest power of step length (h), which has a non-zero coefficient. In general terms,
the truncation error T ( h ) of a method using a step length h is given by:
T ( h ) = O(hP+’), (11.1 )
where superscript p represents the order of accuracy of the method.
The true solution y ( t n ) at tn is thus:
y ( t n ) = yn + O W + ’) + En, (11.2)
where y,, is the value of y calculated by the method after n steps, and E, represents
other possible errors.
352 APPENDIX I1

11.2.2 Stability
Two types of instability occur in the solution of ordinary differential equations, i.e.
inherent and induced irtstability.
Inherent instability occurs when, during a numerical step by step solution, errors
generated by any means (truncation or round-off) are magnified until the true solution
is swamped. Fortunately, transient stability studies are formulated in such a manner
than inherent instability is not a problem.
Induced instability is related to the method used in the numerical solution of the
ordinary differential equation. The numerical method gives a sequence of approxima-
tions to the true solution and the stability of the method is basically a measure of the
difference between the approximate and true solutions as the number of steps becomes
large.
Consider the ordinary differential equation:

PY = AY? (11.3)
with the initial conditions y ( 0 ) = yo which has the solution:

y ( t ) = yoel'. (11.4)
Note that A is the eigenvalue [l] of the single variable system given by the ordinary
differential equation (11.3). This may be solved by a finite difference equation of the
general multi-step form:
k k

i=O i=O

where ai and are constants.


Letting

(11.6)

and
k

i=O

and constraining the difference scheme to be stable when A = 0, then the remaining part
of equation (11.5) is linear and the solutions are given by the roots zi (for i = 1,2, . . . , k)
of m(z) = 0. If the roots are all different, then

and the true solution, in this case (A = 0), is given by

where superscript p is the order of accuracy.


11.2 PROPERTIES OF THE INTEGRATION METHODS 353

The principal root z l , in this case, is unity and instability occurs when J Z j I 2 1 (for
i = 2,3, . . . , k , i
# 1) and the true solution will eventually be swamped by this root
as n increases.
If a method satisfies the above criteria, then it is said to be stable but the degree of
stability requires further consideration.
Weak stability occurs where a method can be defined by the above as being stable,
but because of the nature of the differential equation, the derivative part of equation
(11.5) gives one or more roots which are greater than or equal to unity. It has been
shown by Dalquist [2] that a stable method which has the maximum order of accuracy
is always weakly stable. The maximum order or accuracy of a method is either k 1 +
or k + 2 depending on whether k is odd or even, respectively.
Partial stability occurs when the step length (h) is critical to the solution and is
particularly relevant when considering Runge-Kutta methods. In general, the roots zi
of equation (11.7) are dependent on the product hh and also on equations (11.6). The
stability boundary is the value of hA for which IziI = 1, and any method which has
this boundary is termed conditionally stable.
A method with an infinite stability boundary is known as A-stable (unconditionally
stable). A linear multi-step method is A-stable if all solutions of equation (11.5) tend to
zero as n + CG when the method is applied with fixed h > 0 to equation (11.3) where
h is a complex constant with ReA < 0.
Dalquist has demonstrated that, for a multi-step method to be A-stable, the order of
accuracy cannot exceed p = 2, and hence the maximum k is unity, i.e. a single-step
method. Backward Euler and the trapezoidal method are A-stable, single-step methods.
Other methods not based upon the multi-step principle may be A-stable and also have
high orders of accuracy. In this category are implicit Runge-Kutta methods in which
p < 2r, where r is the number of stages in the method.
A further definition of stability has been introduced recently [3], i.e. Z-stability which
is the multi-variable version of A-stability. The two are equivalent when the method
is linear but may not be equivalent otherwise. Backward Euler and the trapezoidal
method are C-stable single-step methods.
The study of scalar ordinary differential equations of the form of equation (11.3) is
sufficient for the assessment of stability in coupled equations, provided that h are the
eigenvalues of the ordinary differential equations. Unfortunately, not all the equations
used in transient stability analysis are of this type.

11.2.3 Stiffness
A system of ordinary differential equations in which the ratio of the largest to the
smallest eigenvalue is very much greater than unity is usually referred to as being stiff.
Only during the initial period of the solution of the problem are the largest negative
eigenvalues significant, yet they must be accounted for during the whole solution.
For methods which are conditionally stable, a very small step length must be chosen
to maintain stability. This makes the method very expensive in computing time.
The advantage of C-stability thus becomes apparent for this type of problem, as the
step length need not be adjusted to accommodate the smallest eigenvalues.
In an electrical power system, the differential equations which describe its behaviour
in a transient state have greatly varying eigenvalues. The largest negative eigenvalues
354 APPENDIX I1

are related to the network and the machine stators but these are ignored by establishing
algebraic equations to replace the differential equations. The associated variables are
then permitted to vary instantaneously.
However, the time constants of the remaining ordinary differential equations are still
sufficiently varied to give a large range of eigenvalues. It is, therefore, important that if
the fastest remaining transients are to be considered and not ignored, as so often done
in the past, a method must be adopted which keeps the computation to a minimum.

11.3 Predictor-Corrector Methods


These methods for the solution of the differential equation

p Y = F ( Y , X ) with Y ( 0 ) = Y O (11.9)
and X ( 0 ) = X O

have all developed from the general k-step finite difference equation:
k k

(11. lo)
i=O i=O
Basically, the methods consist of a pair of equations, one being explicit (PO = 0) to
give a prediction of the solution at tn+l and the other being implicit (PO# 0), which
corrects the predicted value. There are a great variety of methods available, the choice
being made by the requirements of the solution. It is usual for simplicity to maintain
a constant step length with these methods if k > 2.
Each application of a corrector method improves the accuracy of the method by one
order, up to a maximum given by the order of accuracy of the corrector. Therefore,
if the corrector is not to be iterated, it is usual to use a predictor with an order of
accuracy one less than that of the corrector. The predictor is thus not essential, as the
value at the previous step may be used as a first crude estimate, but the number of
iterations of the corrector may be large.
While, for accuracy, there is a fixed number of relevant iterations, it is desirable
for stability purposes to iterate to some predetermined level of convergence. The char-
acteristic root (z1) of a predictor or corrector when applied to the single variable
problem
p y = k y with y ( 0 ) = yo (11.1 1 )

may be found from

c(ai-
k

i=O
hAp"z'k-'' = 0. (11.12)

When applying a corrector to the problem defined by equation (11.11) and rearranging
equation (11.10) to give

(11.13)
11.3 PREDICTOR-CORRECTOR METHODS 355

the solution to the problem becomes direct. The predictor is now not necessary as
the solution only requires information of y at the previous steps, i s . at yn=i+l for
i = 1 , 2, . . . , k.
Where the problem contains two variables, one non-integrable, such that:
(11.14)

(11.15)

then
Yn+I = Cn+I + mn+lXn+I? (11.16)

and
(11.18)

Although cn+l and m,+l are constant at a particular step, the solution is iterative
using equations (11.15) and (11.16). Strictly, in this simple case, xn+l in equation (II.16)
could be removed using equation (11.15), but in the general multi-variable case this is
not so.
The convergence of this method is now a function of the non-linearity of the system.
Provided that the step length is sufficiently small, a simple Jacobi form of iteration gives
convergence in only a few iterations. It is equally possible to form a Jacobian matrix
and obtain a solution by a Newton iterative process, although the storage necessary
is much larger and, as before, the step length must be sufficiently small to ensure
convergence.
For a multi-variable system, equation (11.9) is coupled with

0 = G ( Y ,X), (11.19)
and the solution of the integrable variables is given by the matrix equation

Yn+l = Cn+l + Mn+I V n + 1 I Xn+lll. (11.20)


The elements of the vector Cn+l are given in equation (11.17) and the elements of
the sparse Mn+l matrix are given in equation (11.18).
The iterative solution may be started at any point in the loop, if Jacobi iteration is
used. Because the number of algebraic variables (X)associated with equations (11.9)
or (11.19) are small, it is most advantageous to extrapolate these algebraic variables
and commence with a solution using equation (11.19).
The disadvantage of any multi-step method (k > 2) is that it is not self-starting. After
a discontinuity k - 1 steps must be performed by some other self-starting method.
Unfortunately, it is the period immediately after a step which is most critical as the
largest negative eigenvalues are significant. As k - 1 is usually small, it is not essential
to use an A-stable starting method. Accuracy over this period is of more importance.
356 APPENDIX I1

11.4 Runge-Kutta Methods


Runge-Kutta methods are able to achieve high accuracy while remaining single-step
methods. This is obtained by making further evaluation of the functions within the
step. The general form of the equation is:

c w i = 1. (11.23)
i= I

Being single-step, these methods are self-starting and the step length need not be
constant. If j is restricted so that j < i, then the method is explicit and cI must be
zero. When j is permitted to exceed i , then the method is implicit and an iterative
solution is necessary.
Also of interest are the forms developed by Merson and Scraton. These are fourth-
order methods ( p = 4) but use five stages (v = 5 ) . The extra degree of freedom
obtained is used to give an estimate of the local truncation error at that step. This
can be used to control the step length automatically.
Although they are accurate, the explicit Runge-Kutta methods are not A-stable.
Stability is achieved by ensuring that the step length does not become large compared
with any time constant. For a pth order explicit method, the characteristic root is:

(11.24)

where the second summation term exists only when v > p and where a; are constant
and dependent on the method.
For some implicit methods, the characteristic root is equivalent to a Pade approxi-
mant to ehA.
The Pade approximant of a function f ( r ) is given by

(11.25)

and if
m
(11.26)
j=O
11.5 REFERENCES 357

If the approximant is to have accuracy of order M + N, and if f(0) = P M N ( ~ ( O ) ) ,


then
N M

j=O j=O j=O j=M+N+I

It has been demonstrated that for approximants of Ah where M = N, M = N 1 +


+
and M = N 2, the modulus is less than unity and thus a method with a characteristic
root equivalent to these approximants is A-stable as well as having an order of M N. +
11.5 References
1. Lapidus, L and Seinfeld, J H (197 1). Numerical Solution of Ordinary Dt#erential Equations,
Academic Press, New York.
2. Dalquist, G (1963). Stability questions for some numerical methods for ordinary differential
equations, Proceedings of Symposia in Applied Mathematics.
3. Zakian, V (1975). Properties of I M N and J M N approximants and applications to numerical
inversion of Laplace transforms and initial value problems, Journal of Mathematical Analysis
andApplications, 50 (l), pp. 191-222.
Appendix 111
TEST SYSTEM USED IN THE
STABILITY EXAMPLES

The d.c. link components are taken from the CIGRE HVDC benchmark model [ 11 and
are shown in Figure 111.1 and Tables 111.1 and 111.2.
The inverter a x . system is as per the benchmark model, and the rectifier a.c. system
is modified. The rectifier a.c. system is shown in Figure 111.2 and is representative
of the central South Island power system of New Zealand. The parameters for this
test system are described below. Rectifier a.c. system per unit values are based on
100 MVA and 345 kV .

1.088 0.151 3.737

.04 167.2 13.23 0.0606

83.32 37.03

Figure 111.1 CIGRE HVDC benchmark model

Table 111.1 CIGRE HVDC benchmark model parameters


Parameter Rectifier Inverter
~ ~~ ~

ax. system voltage (line to line) 345 kV 230 kV


ax. system impedance magnitude 119.03 R 52.9 R
converter transformer tap (prim.side) 1.01 0.989
equivalent commutation reactance 27 R 27 R
d.c. voltage 505 kV 495 kV
d.c. current 2 kA 2 kA
firing angle 15" 15"
d.c. power 1010 MW 990 MW
360 APPENDIX 111

Table 111.2 Parameters for the CIGRE benchmark rectifier


power base 603.73 MVA
primary voltage base 345 kV
secondary voltage base 213.4557 kV
nominal d.c. current 2000 A
nominal firing angle 15"
d.c. voltage source 4.179 p.u.
transformer leakage reactance 0.18 p a .
transformer series resistance 0.01 p.u.
thyristor forward voltage drop 8.11E-6 P.U.
thyristor on resistance 0.001325 p.u.
d.c. current transducer time constant 0.001 ssrad-'
PI controller proportional gain 1.0989 rad/A(p.u.)
PI controller time constant 0.0091 wad-'

Clyde Twizel

Livingston Aviemore

Figure 111.2 Test system

Table 111.3 Test system a.c-bus


parameters
Bus IVI (P.U.1 ec')
1 1.001 19.377
2 0.983 3.635
3 0.992 0.0
4 1 .00J -7.049
~ ~ ~~

Table 111.4 Test system ax.-line parameters


Line R(p.u.1 X(P.U.)
1-2 0.00452 0.02832
2-3 0.00334 0.01154
2-4 o.Ooo10 0.04256
361

Table 111.5 Test system steady state


d.c. conditions
Rectifier Inverter
Vdr 500 kV 490 kV
Idc 2 kA 2kA
a! 17.65" 15"
pdc lo00 MW 980 MW
QdC 583 MVAr 545 MVAr

Table 111.6 Rectifier a.c. system bus


parameters

Aviemore 1.015 1.338


Benmore 1.Ooo 0.000
Clyde 1.043 6.479
Roxburgh 1.042 6.338
Livingston 1.024 2.415
Twi zel 1.015 1.564

Table 111.7 Rectifier ax.-system line parameters


Bus (from) Bus (to) Circuit R(p.u.) X(P.U.1 B(P.U.1
Aviemore Benmore 0.00151 0.01034 0.17038
Aviemore Livingston 0.00304 0.01459 0.12685
Benmore Twizel 0.00151 0.0 1034 0.17038
Clyde Roxburgh A 0.002 16 0.01297 0.11566
Clyde Roxburgh B 0.00216 0.01297 0.11566
Clyde Twizel A 0.00405 0.04792 0.47925
Clyde Twizel B 0.00405 0.04792 0.47925
Livingston Roxburgh 0.01088 0.05209 0.45406

Table 111.8 Rectifier a.c.-system load


parameters
Bus P (MW) Q (MVAr)
Roxburgh 150.0 75.0
Twizel 100.0 15.0

Table 111.9 Rectifier a.c.-system generator parameters


~ ~ ~~ ~ ~ ~~~~~~~

Bus H (MWs/MVA) X&(p.u.) xd(p.u.) R,(p.u.) ?'&,(s) Tbo(S) x:(p.U.) T:,(s)


Aviemore 6.273 0.108 0.299 0.0024 1.47 0.43 0.066 0.053
Benmore 12.337 0.069 0.259 0.0011 8.7 0.43 0.044 0.080
Clyde 14.68 0.117 0.181 O.OO0 6.1 99.990.117 0.0
Roxburgh 6.42 0.103 0.54 0.002 7.16 0.36 0.062 0.055
362 APPENDIX I11

Table 111.10 Rectifier a.c.-system AVR parameters (a block diagram of this AVR is shown in
Figure 7.6)
Bus Regulator gain (P.u.) Regulator TC (s) Feedback gain (PA.) Feedback TC (s)
Aviemore 400.0 0.02 0.03 1.o
Benmore 50.0 0.2 0.04 0.4
Clyde 50.0 0.2 0.04 0.4
Roxburgh 50.0 0.2 0.04 0.4

Bus Exciter gain (P.u.) Exciter TC (s) Regulator max. Regulator min.
limit (P.u.) limit (P.u.)
Aviemore 1.o 0.01 1.6 -1.0
Benmore -0.05 0.5 2.0 -1.0
Clyde -0.05 0.5 2.0 -1.0
Roxburgh -0.05 0.5 2.0 - 1.0

Bus Exciter max. limit (P.u.) Exciter min. limit (P.u.)


Aviemore 1.6 -1.3
Benmore 3.0 -1.0
Clyde 3.0 -1.0
Roxburgh 3.0 -1.0

Table 111.11 Rectifier ac system speed governor parameters (a block diagram of this governor
is shown in Figure 7.8)
Bus Regulation (P.u.) Governor 7'1 Governor lead Tz Governor T3
Aviemore 5.0 16.0 2.4 0.92
Benmore 5.0 25.0 2.8 0.50
Clyde 5.O 20.0 4.0 0.50
Roxburgh 5.0 12.0 3.0 0.50

Bus Water TC (s) Turbine TC (s)


~-
Aviemore 0.3 0.15
Benmore 0.43 0.215
Clyde 0.3 0.15
Roxburgh 0.35 0.175

111.1 Reference

1. Szechtman, M, Weiss, T and Thio, C V (1991). First benchmark model for HVDC control
studies, Electra, 135, 55-75.

You might also like