You are on page 1of 789

THE PLANNING, DESIGN & CONSTRUCTION OF .

,/

Overhead

/

,_f?:w~r lines

~~~~==::j:j: f: •

if ---;;--

The planning, design and construction of overhead power lines

EDITORIAL COMMITTEE

S Bisnath AC Britten

DH Cretchley D Muftic

T Pillay

R Vajeth

February 2005

THE PLANNING, DESIGN AND CONSTRUCTION OF OVERHEAD POWER LINES

Prepared by a staff of experts from Eskom Holdings Ltd under the direction of Thavanthiran Pillay and Sanjeev Bisnath

This book is dedicated to the memory of Mike Korber, one of the authors who sadly passed away while developing the material. His contribution is greatly appreciated by the Eskom team.

While the authors and Eskom Holdings Ltd have made every effort to ensure the accuracy of this document, they cannot be held responsible for any errors or any infringment of copyright and patent rights, nor any direct or consequential loss or damage suffered by any person or organisation, however caused, which may result from the use of this document.

All rights to this document are reserved. Except where allowed by the Copyright Act, no part of this document may be reproduced, stored in a retrieval system or transmitted in any form or by any means, either electronic or mechanical, without prior written permission from Eskom Holdings Ltd.

Published by Crown Publications cc

2 Theunis Street, Bedford Gardens, Johannesburg

February 2005

CONTENTS

Chapter 1 OVERHEAD LINES

Jan Reynders, Dzevad Muftic, Antony Britten and Dave Cretchley

1.1 Introduction 2

1.2 Voltage, Impedance and Power Limit 4

1.3 The Design Task 5

1.4 The Design of Overhead Power Lines 6

1.5 References 10

1.6 Bibliography 10

Chapter 2 MODELLING

Mike Korber, Dzevad Muftic and Dipeen Dama

2.1 Introduction 12

2.2 Power Transfer Limits 13

2.3 Other Factors Limiting the Power Transfer of Overhead Power Lines 13

2.4 Inductance of Transposed and Non-transposed Three-phase

Overhead Power Lines 16

2.5 Calculation of the Inductance of Bundled Conductors 18

2.6 Line Characteristics and Parameters for Three-phase Single Circuit System 20

2.7 Overhead Power Line Modelling 25

2.8 Unbalanced Systems 32

2.9 Concluding Remarks 41

2.10 References 41

2.11 Bibliography 41

Chapter 3

POWER SYSTEM ANALYSIS AND PLANNING Riaz Vajeth and Roy Estment

3.1 Introduction 44

3.2 Identify the Need for Network Expansion 46

3.3 Formulation of Alternative Options 48

3.4 Analysis of the Options 57

3.5 Selection of the Best Option 59

3.6 Sensitivity Analyses 60

3.7 Economic Justification 61

3.8 Current and Future Challenges for the Planning Engineer 63

3.9 Concluding Remarks 64

3.10 References 65

Chapter 4

HIGH VOLTAGE DIRECT CURRENl (HVDC) TRANSMISSION Nelson Ijumba and Gary Sibilant

4.1 Introduction 68

4.2 Background 68

4.3 HVDC Transmission Lines 79

4.4 Future Developments in HVDC 81

4.5 HVDC Systems in Africa 85

4.6 Concluding Remarks 89

4.7 Acknowledgements 89

4.8 References 89

4.9 Bibliography 91

4.10 Appendices 92

Chapter 5

ENVIRONMENTAL IMPACT MANAGEMENT Dave Cretchley and Jose Clara

5.1 Introduction 96

5.2 Environmental Impacts Caused by Power Lines 97

5.3 Managing the Impacts 99

5.4 Environmental Management System 105

5.5 Detailed Environmental Impact Management in Design and Construction .. 107

5.6 Servitude Rights Acquisition (,Negotiations') 112

5.7 Overhead Line Setting-out Survey and Profiling 112

5.8 Construction Phase: EMP 112

5.9 Concluding Remarks 113

5.10 References 113

5.11 Appendices 115

Chapter 6 CORONA

Antony Britten and Arthur Burger

6.1 Introduction 122

6.2 Corona Phenomena on Overhead Power Lines 124

6.3 Factors Influencing the Generation of Corona on Conductors 133

6.4 Practical Design Procedures and Considerations 141

6.5 Design Limits for Radio Interference, Audible Noise and Corona Losses .. 148

6.6 Consideration of Conductor Surface Gradients and External Fibre Optic

Cable 151

6.7 Concluding Remarks 152

6.8 References 152

6.9 Appendices 155

Chapter 7

POWER FREQUENCY ELECTRIC AND MAGNETIC FIELDS Pieter Pretorius

7.1 Introduction 158

7.2 Typical Electric and Magnetic Field Levels 160

7.3 Overhead Power Line Electric Field Design Limit 162

7.4 Human Exposure Guidelines : 162

7.5 Application of ICNIRP Exposure Guidelines 164

7.6 Field Mitigation 165

7.7 Interference with Computer Monitors 171

7.8 Concluding Remarks 172

7.9 References 172

Chapter 8

MODELLING OF ELECTRIC FIELDS Shawn Nielsen

8.1 Introduction 176

8.2 Electric Field Theory 176

8.3 Modelling Methods 177

8.4 Step-by-step Illustration 182

8.5 Practical Examples 189

8.6 Measurement of Electric Field Intensity 196

8.7 Concluding Remarks 197

8.8 References 197

Chapter 9 LIGHTNING Arthur Burger

9.1 Introduction 200

9.2 South African Studies of the Effects of Lightning on Overhead Power Lines 200

9.3 Interaction of Lightning with Overhead Power Lines 201

9.4 Physics of Lightning 201

9.5 Direct Lightning Strikes to Overhead Power Lines 208

9.6 The Use of Overhead Ground Wires to Limit the Effects of Lightning 209

9.7 Concluding Remarks 209

9.8 References 210

Chapter 10

INTRODUCTION TO DESIGN OPTIMISA TIOI\j Dzevad Muftic, Pierre Marais and Dumsani MtoJo

10.1 Introduction 212

10.2 Electrical Design Optimisation 212

10.3 Optimisation of Line Components 213

10.4 Final Optimisation of Line 216

10.5 Reliability-based Design 217

10.6 References 221

10.7 Bibliography 221

Chapter 11

BASIC ELECTRICAL DESIGN Dzevad Muftic and Arthur Burger

11.1 Introduction 224

11.2 Number and Size of Sub-conductors 224

11.3 Phase Configuration and Spacing 225

11.4 Line (Phase Conductors and Ground Wire) Configuration 226

11.5 Required Withstand Voltages of Air Gap Insulation 227

11.6 Clearances and Line Geometry 228

11.7 Electrical Characteristics ofthe Line 232

11.8 Concluding Remarks 232

11.9 References 233

11.10 Appendices 234

Chapter 12

INSULATION CO-ORDINATION Sanjay Narain and Kristoff Sadurski

12.1 Introduction 242

12.2 Definition of Insulation Co-ordination 243

12.3 Types of Voltage Stresses 244

12.4 Insulation Strength 249

12.5 Live-line Maintenance 261

12.6 Fires Under Overhead Lines 270

12.7 Line Compaction 273

12.8 Eskom Practice 274

12.9 Concluding Remarks 274

12.10 References 274

Chapter 13 THERMAL RATING

Rob Stephen, Dzevad Muftic, Johan Swan and Arthur Burger

13.1 Introduction 278

13.2 Factors Affecting the Thermal Limit of Conductors 279

13.3 Calculation of Conductor Temperature in the Steady-state 280

13.4 Dynamic Behaviour of Conductor Temperature 288

13.5 Deterministic Thermal Rating 289

13.6 Probabilistic Conductor Rating Technique 290

13.7 Real-time Line Monitoring of Overhead Power Lines 294

13.8 Concluding Remarks 297

13.9 References 298

13.10 Appendices 300

Chapter 14 COUPLING

Antony Britten, Mike Korber and Rukesh Ramnarain

14.1 Introduction 310

14.2 The Basics of Coupling at Power Frequencies 310

14.3 SafetyIssues 319

14.4 Capacitive and Inductive Coupling and the Operation of Overhead Power

Lines .'. -: 320

14.5 Practical Considerations and Mitigation Techniques 321

14.6 Telephone Interference 321

14.7 Geomagnetically Induced Currents 322

14.8 Concluding Remarks 323

14.9 References 323

Chapter 15

CONDUCTOR OPTIMISATION

Dzevad Muftic, RiazVajeth and Dipeen Dama

15.1 Introduction 326

15.2 Overview of Conductors for Overhead Power Lines 326

15.3 Selection of Optimal Conductor 333

15.4 Concluding Remarks 351

15.5 References 351

15.6 Bibliography 352

15.7 Appendices 353

Chapter 16

GROUND WIRE OPTIMISp,TION

Dzevad Muftic, Sanjay Narain and Rukesh Ramnarain

16.1 Introduction 356

16.2 Position of Ground Wires 356

16.3 Electrogeometric Model 357

16.4 Ground Wire Selection 367

16.5 Losses in the Ground Wire 373

16.6 Optical Fibre Ground Wire (OPGW) 373

16.7 Concluding Remarks 373

16.8 References 374

Chapter 17

INSULATOR SELECTION Fabio Bologna

17.1 Introduction 376

17.2 Function and Main Parameters of Insulators for Overhead Power Lines 376

17.3 Corona Rings 383

17.4 The Influence of The Environment on the Performance of Outdoor

Insulators 383

17.5 Site Severity Monitoring Techniques 388

17.6 Flashover Caused by Pollution 388

17.7 The Influence of Leakage Current on the Flashover Process 392

17.8 Influence of Discharge Activity on Polymer Insulator Surfaces 392

17.9 Concluding Remarks 396

17.10 References 396

17.11 Appendices 398

Chapter 18 INSULATION DESIGN

Sanjay Narain and Dzevad Muftic

18.1 Introduction 410

18.2 Air gap Clearances 410

18.3 Insulator Parameters 411

18.4 Concluding Remarks 422

18.5 References 422

18.6 Appendices 424

Chapter 19

LINE HARDWARE

jacques Calitz, Bharat Haridass, Bertie jacobs, johann Retief and Pieter du Plessis

19.1 Introduction 442

19.2 Preliminary Design of Hardware 443

19.3 Detailed Design of Load Bearing Hardware 443

19.4 Compatibility of Hardware 446

19.5 joints and Dead-ends 447

19.6 Live-line Maintainability 448

19.7 Material Characteristics of Hardware 449

19.8 Material Properties 449

19.9 Mechanical Properties .450

19.10 Other Factors in Material Selection 451

19.11 Non-load Bearing Hardware for Overhead Power Lines 452

19.12 Testing and Quality Control of Hardware .452

19.13 Overhead Line Oscillation 455

19.14 Concluding Remarks 477

19.15 References 478

19.16 Bibliography 479

19.17 Appendices 481

Chapter 20

SUPPORTING STRUCTURES

jose Diez-Serrano and Pierre Marais

20.1 Introduction 484

20.2 Tower Types 485

20.3 Support Structure Design Methodology 491

20.4 Conceptual Design 492

20.5 Geometry Definition 492

20.6 Loading 495

20.7 Analysis and Design 500

20.8 Failure Sequencing and Component Design Loads 505

20.9 Final Design Documentation 506

20.10 Loading Trees 507

20.11 Tower Acceptance Tests 509

20.12 Evaluation 511

20.13 Concluding Remarks 511

20.14 References 512

20.15 Bibliography 512

Chapter 21 FOUNDA nONS

Willem Combrinck and Pierre Marais

21.1 Introduction 514

21.2 Foundation Requirements 514

21.3 Foundation Design Process 514

21.4 Soil Investigations 516

21.5 Soil Classification 516

21.6 Foundation Loading Factors 519

21.7 Foundation Types 520

21.8 Application of Various Foundation Systems 534

21.9 Structural Design of Foundation Systems 535

21.10 Concluding Remarks 539

21.11 References 539

Chapter 22 EARTHING

Arthur Burger, Rukesh Ramnarain and Luthando Peter

22.1 Introduction 542

22.2 Concluding Remarks 564

22.3 References 564

22.4 Appendices 566

Chapter 23 LAND SURVEY Norman Banks

23.1 Introduction 576

23.2 Planning 576

23.3 Implementation 577

23.4 Survey Methods 578

23.5 Concluding Remarks 585

23.6 Bibliography 586

Chapter 24 TOWER SPOTTING

Suren Natesan, Allan Muir and John Mokoena

24.1 Introduction 588

24.2 Line Optimisation Process Flow 590

24.3 Inputs and Design Criteria 592

24.4 Processing: Sag Calculations and Templating 596

24.5 Outputs 604

24.6 Concluding Remarks 606

24.7 References 606

24.8 Bibliography 607

Chapter 25

INTRODUCTION TO COMMUNICATION SYSTEMS Donald Andrews

25.1 Introduction 610

25.2 Telecommunications and Power Networks 610

25.3 Telecommunications Opportunities for Power Network Operators 612

25.4 Power Line Carriers 612

25.5 Aerial Fibre Optic Cables 613

25.6 Cellular Base Station Antennas on Structures 615

25.7. Concluding Remarks 616

25.8 Bibliography 616

Chapter 26

POWER LINE CARRIERS Andrew Pollard

26.1 Introduction 618

26.2 The Role of Power Line Carriers in Power System Operation 619

26.3 Basics of Power Line Carriers (PLCs) 621

26.4 The Overhead Line as a Telecommunications Transmission 622

26.5 Practical Issues of Overhead Power Lines 632

26.6 Other Overhead Line Criteria that Affect PLC Performance 639

26.7 Impact of Corona and Arcing Noises on PLC Performance 641

26.8 Concluding Remarks 643

26.9 References 643

Chapter 27

AERIAL FIBRE OPTIC CABLES David Smith

27.1 Introduction 646

27.2 Types of Aerial Optical Cable in General Use 647

27.3 Factors to be Considered in Selection of Cable Type Used 651

27.4 Installation Practices 661

27.5 Concluding Remarks 665 .

27.6 References 666

Chapter 28

CELLULAR BASE STATION ANTENNAS Pieter Pretorius

28.1 Introduction 670

28.2 Technical Considerations 671

28.3 Environmental Considerations 674

28.4 Concluding Remarks 675

28.5 References 675

Chapter 29 CONSTRUCTION

Pierre Marais and Barrie Badenhorst

29.1 Overhead Power Line Construction 678

29.2 Project Management 713

29.3 Concluding Remarks 720

29.4 References 720

PREFACE

The Eskom Power Series was conceived as a result of concern over the continuing loss of critical technical skills and experience.

A series of events, spanning the past few decades, has reshaped the electricity supply industry and retaining a skilled workforce remains a major priority. Each year, a growing number of technical experts leaves the industry, depriving it of a wealth of experience, knowledge and expertise.

Eskom's R&D team has created the Eskom Power Series, each book of which has been written by experienced specialists and consultants in the utility field.

Eskom's strong R&D programme, which is specifically focused on utility issues, will, by supplying the most up-to-date information available, ensure that the series is kept relevant and practical. The aim of the series is to serve as a guide and a legacy by collating Eskom's knowledge and experience. Written by Eskom staff who are familiar with local Southern African conditions, the Power Series offers a source of reliable, reputable and high level technical information and practice. We anticipate that the series will grow according to the needs of industry.

The confidence we place in our Power Series is justified by the many world-class innovative technologies developed by Eskom, which has also created a highly efficient infrastructure. For example, Eskom built the world's largest coal-fired power station, which uses a drycooling system developed by engineers in South Africa. Eskom is also developing 'pebble bed modular reactor' nuclear technology, designed to be inherently safe. The great success of our rural electrification programme, achieved in a relatively short time, places us among world leaders in this field. Eskom also pioneered and was the first to build power lines at voltages of 765 kV in high altitude conditions, a challenging feat.

In addition to many local and international awards presented to Eskom each year, we were awarded the Community Development Programme of the Year award in 2003, at the Global Energy Awards in New York City. This award, arguably one of the most prestigious in the global utility industry, affirms our place among other major utilities in the world.

The Power Series is a practical series targeted at individuals working within the utility environment. Volume One offers a useful guide to the design of overhead power lines for voltages of 132 kV and above, focusing mainly on alternating current (AC) lines. A brief chapter covers high voltage direct current (HVDC). The chapters are well balanced in their coverage of electrical, mechanical, environmental and civil engineering. The authors regard overhead power line design to be a multidisciplinary field which must not be considered in isolation. Line construction and communication over power lines are covered briefly.

Volume Two deals with the fundamentals and practice of overhead line maintenance and provides a comprehensive insight into its philosophy and practice (132 kV and above).

High voltage insulators, a major component in utility power systems, must be selected, applied and maintained with care. Volume Three, titled The Practical Guide to Outdoor High Voltage Insulators', is devoted to their application and maintenance.

The extensive information contained in the Power Series can be applied in utilities worldwide and is accessible to any interested person.

I would like, in conclusion, to acknowledge the high quality of the work that went into the development of this reference book and hope that readers will find it useful in their application.

ACKNOWLEDGEMENTS

Many people doubted the possibility of publishing such a comprehensive manual in view of the schedules of the Eskom engineers, technologists and technicians. This was indeed a challenge for our team but I must commend Sanjeev Bisnath for his dedication and drive in managing this project. Special thanks are also due to Eskom Holdings Ltd Board for the confidence they have shown us in their continued support of R&D, both in funding and resources.

Tony Britten assisted greatly in the technical development of this reference book and deserves acknowledgement for going beyond the call of duty to ensure that an international standard was upheld. The technical editing for this publication was challenging and I must thank Dave Cretchley for his invaluable contribution.

Dzevad Muftic made significant contributions in collaborating with the authors when problems were encountered and added great value in collating the text. Jan Reynders is acknowledged for his general guidance. Riaz Vajeth did a sterling job in getting the chapters developed and also assisted with the technical editing.

Thavanthiran (Logan) Pillay, PrEng Eskom Transmission Research Manager

DH Cretchley Eskom

D Muftic Eskom

EDITORIAL COMMITTEE

S Bisnath Eskom

AC Britten Eskom

T Pillay Eskom

R Vajeth Eskom

SPECIAL CONTRIBUTORS

M Cope Eskom

D Dama Eskom

K Grant

Crown Publications

G Moodley

Eskom (formerly)

T Motloung Eskom (formerly)

A Muir

Eskom (formerly)

M Mulaudzi Eskom (formerly)

V Pillay Eskom

J van den Berg Eskom

B Vorster

Crown Publications

T van den Berg Grab u Graphics

M Lloyd Eskom

P Marques Eskom

JP Reynders

University of Witwatersrand

1 I

CHAPTER 1

OVERHEAD LJ:NES

BY JAN REYNDERS, DZEVAD MUFTIC, ANTONY BRITTEN AND

DAVE CRETCHLEY

@Eskom

S Bisnath Eskom

AC Britten Eskom

EDITORIAL COMMITTEE

DH Cretchley Eskom

D Muftic Eskom

T Pillay Eskom

R Vajeth Eskom

SPECIAL CONTRIBUTORS

M Cope Eskom

D Dama Eskom

K Grant

Crown Publications

G Moodley

Eskom (formerly)

T Motloung Eskom (formerly)

A Muir

Eskom (formerly)

M Mulaudzi Eskom (formerly)

V Pillay Eskom

J van den Berg Eskom

B Vorster

Crown Publications

T van den Berg Grab u Graphics

M Lloyd Eskom

P Marques Eskom

JP Reynders

University of Witwatersrand

CHAPTER 1

OVERHEAD LJ:NES

BY JAN REYNDERS, DZEVAD MUFTIC, ANTONY BRITTEN AND

DAVE CRETCHLEY

@Eskom

SYNOPSIS

OVERHEAD LINES IN THE CONTEXT OF A POWER SYSTEM

This chapter introduces the role played by overhead power lines in the establishment and operation of a modern electric utility network. The interaction of the electrical parameters that affect the ability of the overhead power line to transmit power are described. This serves to introduce the theme of the book, namely, the design and construction of novel, compacted overhead power lines which offer reliable, cost-effective and environmentally acceptable performance.

1.1 INTRODUCTION

Society in all parts of the world is profoundly dependent on a reliable and low-cost source of electricity for the home and work. Electricity is not an end in itself, but it makes communication, mass transport and energy for the home, business and factory easy to provide. It is hard to imagine an effective hospital without electricity; business and industry would cease if all the facilities made possible by electricity were removed.

Electricity has become indispensable because people need energy that is highly adaptable and controllable. In addition, natural sources of energy are frequently in areas remote from the user and may not be in the form the user needs. Electrical energy offers efficient conversion from a variety of sources to electricity, easy transport to the user (sometimes over very long distances), and re-conversion to the form that the user needs. All of this can be achieved in a clean, controllable and efficient manner. Thus, it is not surprising that the electrical utility industry is one of the largest and most complex in the world.

Overhead power lines are the means of transporting electrical energy, and they need to fulfil their function with a high level of reliability, while having the lowest possible impact on the environment.

Overhead power lines have been in existence for more than a hundred years, as shown in Table 1.1 below [1]:

Table 1.1: History of overhead power lines.

AC/DC Length Voltage Date Location
(km) (kV)
First line DC 50 2.4 1882 Germany
First single- AC 21 4 1889 Oregon, USA
phase line
Fi rst th ree- AC 179 12 1891 Germany
phase line From these humble beginnings, there are now DC lines operating at ± 600 kY and three phase AC lines operating up to 1150 kY, with power transfer capabilities approaching 10GW.

:HAPTER'

OVERHEAD LINES IN THE CONTEXT OF A POWER SYSTEM

In Southern Africa, the highest voltage used so far is 765 kV (AC) and ± 533 kV (DC). One of the dominant characteristics is the development of long distance transmission, as shown in Figure 1.1.

Southern African Power Pool

Utility Country
SNEL DR Congo
TANESCO Tanzania
ESCOM Malawi
ZESCO Zambia
ENE Angola
Nampower Namibia
BPC Botswana
ZESA Zimbabwe
EdM Mozambique
SEB Swaziland
Eskom South Africa
LEC Lesotho Figure 1.1: The Southern African transmission grid.

The Southern African network is being developed in response to the needs of the region as a whole. Despite this level of achievement, there is continual refinement underway. Significant reductions in cost and improvements in performance have been achieved by using compact guyed structures and other innovations. In the past three decades, environmental impact has assumed considerable importance. The influence the structures and electric and magnetic fields have on the environment must conform to increasingly demanding requirements. Audible noise and radio interference are also significant issues which have to be addressed. Some of the above environmental factors can have a critical influence on the ultimate design of a line. The designer must therefore blend all the restrictions, with the requirement of producing a cost-effective and highly reliable line.

This book draws on a wealth of practical and theoretical experience on the planning, design and construction of overhead power lines. It is intended to serve as a reference source for planners and design teams, by providing a comprehensive overview of the theoretical considerations and their application to the actual design of optimised lines. It is complemented by the Eskom book entitled, The fundamentals and practice of overhead line maintenance' [2].

Before presenting an overview of the design process and the structure of the book, it is necessary to introduce some elementary concepts.

CHAPTER 1

OVERHEAD LINES INTHE CONTEXT OFA POWER SYSTEM

1.2 VOLTAGE,IMPEDANCE AND POWER LIMIT

1.2.1 EQUIVALENT CIRCUIT

An overhead power line may be represented by the simple equivalent circuit shown in Figure 1.2 [1,3]. For three-phase lines, it is assumed that balanced conditions pertain and that a single-phase circuit is an adequate representation of the line. The series resistance and inductance, as well as the shunt capacitance, can be readily calculated from the line geometry or found in tables (see Chapter 2).

v

5

R L
Is Ir
II I "NN '000' I II •
C C V
I I r
• •



Figure 1.2: Pi-equivalent circuit for an overhead power line.

The subscript's' on the voltage and current applies to the sending-end of the line and the subscript 'r' on the voltage and current applies to the receiving-end of the line.

1.2.2 POWER LIMIT

The maximum power that a line can transmit is given by equation (1.1):

P = vy,

l X

(1.1)

where X is the series reactance of the line, and is normally much larger than the series resistance. PL also refers to what is known as the steady-state stability limit of the line.

This equation describes some very important properties of the limits of performance of an overhead power line:

The power that can be transmitted is directly proportional to the product of the sending- and receiving-end voltages. Since these are usually more-or-Iess the same, the power limit is proportional to the square of the operating voltage. It is for this reason that utilities use higher voltages to transmit higher amounts of power. The increase in series reactance, for comparable line lengths, does not increase significantly with voltage, and hence, if the operating voltage of a line is doubled, the increase in the power limit is almost fourfold.

:HAPTER 1

OVERHEAD LINES INTHE CONTEXT OFA POWER SYSTEM

Obviously, increases in voltage require more insulation and wider servitudes, but the economics and environmental impact of moving to higher voltages, in comparison to having a number of parallel lines, are usually in favour of the higher voltage option. The power that can be transmitted is inversely proportional to the series reactance, and hence the length, of the line. The geometry of the conductor bundles, phase separation and length of the line dictate this reactance. Compact configurations can decrease the series reactance significantly. This becomes very important for long lines.

The series reactance is inductive and can be reduced through the use of series capacitors. This is often implemented on lines of 500 km or more in length. In this way, the transmitted power can be significantly increased, but at the risk of causing sub-synchronous resonance.

For lines shorter than about 100 krn, the above equation is still applicable. However, in such cases, the maximum power transmitted is often limited by the thermal rating of the line conductors or substation terminal equipment, and not by voltage drop and stability considerations.

1.3 THE DESIGN TASK

Bearing in mind the above considerations, line designers are usually presented with an approximate line length, an operating voltage and a power limit. It is their task to produce a line that operates at the desired voltage and is capable of delivering the power required. At the same time, it must satisfy the environmental, availability, reliability, safety and security requirements, all within a specified budget.

Environmental considerations have a major influence on the design of overhead power lines. They are generally covered by statutory requirements, utility policies and related guidelines, but designers do in any case need to be aware of the environmental impacts of overhead power lines. There are major constraints for overhead power lines in respect of electric and magnetic fields, corona (audible and radio noise), land occupation and limitation for further use, wild life protection, and visual impact. Environmental constraints on current development in Southern Africa can have a significant effect on the economics and performance of overhead power lines.

The availability of a power supply is generally measured in terms of the number of hours in a year for which the power system is able to perform its function. The availability of lines as a part of the power system is a component of total availability. For lines in particular, a distinction can be made between the electrical and mechanical (structural) components of reliability. As a measure of effectiveness, availability was considered from the early stages of the development of overhead power lines, although it was not formalised in terms of measured values. What has changed, and is still changing, is that the criteria of availability are becoming stricter and stricter. What was once considered an acceptable level of availability is not acceptable today. The gap between the ultimate hypothetical goal of 100% availability and reality is slowly being reduced. The implications for the design of overhead power lines are significant.

CHAPTER 11

OVERHEAD LINES INTHE CONTEXT OFA POWER SYSTEM

Reliability resolves itself into two important criteria, namely, availability and quality of supply. It is defined as the probability of continuous power transfer under a set of operating conditions, during a specified time of operation. Like availability, it has economic implications and for each project the question of how much the improvement in reliability will cost, and whether it is worth the cost, must be answered [4].

The stochastic (statistically random) nature of all the main influencing factors and the behaviour of the power delivery system suggest the use of probabilistic rather than deterministic methods and approaches in design. In overhead power line design, this is a relatively new approach, but it has already brought about a better understanding of power system behaviour and an improvement in reliability within acceptable cost limits.

Safety and security are two of the other main requirements for power delivery systems. For overhead power lines, these have the following meaning:

1.3.1 SAFETY

Electrical: phase-to-ground clearances and earthing (step- and touch-potentials) and their impact on the safety of the public.

Mechanical and structural: safety during construction and maintenance and in the rare occurrences of conductor or structure failure.

1.3.2 SECURITY

Security is the measure of the ability of the overhead power line to avoid or resist a collapse from component failure or from the expected loads imposed on it. The emphasis is on the behaviour of the line as an integrated structural system.

Safety and security and their improvement also have particular economic implications. Optimal design should consider this aspect in the same way as availability and reliability.

1.4 THE DESIGN OF OVERHEAD POWER LINES

The main objective of optimising the design of overhead power lines is to attain the lowest cost in construction and maintenance, together with the required availability in operation. There are many ways of achieving these objectives. The main approaches developed in Eskom will be presented in this book.

1.4.1. LINE DESIGN PROCESS

The process of line design optimisation is outlined in the sequence diagram, Figure 1.3. It reflects the way this book is structured:

It starts with Section A - PLANNING.

Within Planning, the first chapter is power line Modelling (Chapter 2). This is one of the first activities in which the expected electrical performance of the new line is clarified at a conceptual level.

::HAPTER 1

-

OVERHEAD LINES IN THE CONTEXT OF A POWER SYSTEM I

It is followed by Power system analysis and planning (Chapter 3). The principles of the techno-economic evaluation of each project, as well as the basic requirements in terms of power transfer capacity and performance, are discussed.

Section A is concluded by HVDC transmission (Chapter 4). This chapter deals with some basic principles and possible options of High Voltage Direct Current Transmission.

Section B - ENVIRONMENT.

This section starts with the discussion in broad terms of Environmental impact management (Chapter 5). All relevant issues of environmental impact, as well as relevant weather data and line route information, are considered.

More specifically, the electrical aspects of environmental impact are dealt with in Chapters 6 to 9. They cover:

Corona (Chapter 6),

Power frequency electric and magnetic fields (Chapter 7), Modelling of electric fields (Chapter 8),

Lightning (Chapter 9).

They are followed by Section C - DESIGN OPTIMISATION

This section has three sub-sections:

C.1 Electrical design optimisation C.2 Line components

C.3 Line design

Section C starts with the Introduction to design optimisation

(Chapter 10).

The first sub-section, Electrical design optimisation. leads on to in-depth discussion of issues such as:

Basic electrical design Insulation co-ordination Thermal rating

Coupling

(Chapter 11). (Chapter 12), (Chapter B). (Chapter 14).

The optimisation of line components in the next sub-section, is one of main and the most comprehensive activities in the development of overhead power lines. The main component issues are:

Conductor optimisation Ground wire optimisation Insulator selection Insulation design

Line hardware

Supporting structures Foundations

Earthing

(Chapter 15), (Chapter 16), (Chapter 17), (Chapter 18), (Chapter 19), (Chapter 20), (Chapter 21), (Chapter 22).

CHAPTER 11

OVERHEAD LINES IN THE CONTEXT OF A POWER SYSTEM

The final sub-section in design optimisation is line design. It covers Land survey (Chapter 23), and Tower spotting (Chapter 24).

Under Section D - COMMUNICATION, more detail is presented on:

Introduction to communication systems Power line carriers

Aerial fibre optic cables

Cellular base station antennas

(Chapter 25), (Chapter 26), (Chapter 27). (Chapter 28).

The final, Section D - OVERHEAD POWER LINE CONSTRUCTION includes the chapter Construction (Chapter 29), in which line construction technology is discussed in some detail, with the emphasis on design-related aspects.

This brief overview of the main activities in the design and construction of overhead power lines demonstrates that it is multidisciplinary, interactive work. It requires the participation of several different disciplines or professions. Network planners, economists, environmentalists, electrical engineers, structural designers, civil engineers, surveyors and others, are all required to work together to obtain the optimum result.

The days are past when the electrical engineer (traditionally the dominant profession in overhead power lines development) could survey the line, design the conductors, decide on the necessary insulation and design the supporting structures and foundations. The depth of development and the use of more sophisticated optimisation methods in each discipline have gone beyond the capability of an individual in any discipline. This is why multidisciplinary teamwork is essential.

:HAPTER 1

START

OVERHEAD LINES IN THE CONTEXT OF A POWER SYSTEM

POWER LINES MODELLING

BASIC TECHNICAL - - - - ~

REQUIREMENTS

PLANNING - POWER SYSTEM ANALYSIS

ENVIRONMENT - WEATHER DATA, LINE ROUTE DATA

ELECTRICAL ASPECTS OF ENVIRONMENT

,-----

-----,

LINE OPTIMISATION

ELECTRICAL DESIGN, ACt DC TRANSMISSION

TRANSFER CAPACITY

<':3 c Q) s:: o .0.. E

o u

Q) s:: :.:::

CONDUCTOR & GROUND

LINE OPTIMISATION
TOWER SPOTTING .. EARTHING DESIGN
L - - f - - - - - _j
COUPLING COMMUNICATION OVER
POWERLINES CONSTRUCTION

OPERATION, MAINTENANCE

Figure 1.3: Line design optimisation sequence diagram.

CHAPTER 1

OVERHEAD LINES IN THE CONTEXT OF A POWER SYSTEM

It emphasises another important characteristic - interactive work during each stage of the whole process of optimisation. No activity can be done in isolation. For example, conductor optimisation has to have the active participation of the structural designer, tower design requires input on the electrical requirements (clearances, insulation) etc.

To neglect or ignore this aspect, and to attempt to do the optimisation as a set of independent activities, will result in a sub-optimal design. Another feature, which has developed from Eskom's practice, is that optimised designs should be verified and improved by means of full-scale electrical and mechanical tests. These include, where practicable, tests of the lightning and switching impulse strengths, live-line insulation assessments, corona performance, mechanical strength of the tower and conductor vibration. The dielectric and corona tests, in particular, must be done at a suitable altitude, so as to replicate the conditions existing in many parts of Southern Africa.

Full-scale electrical and mechanical certification tests, and insulator pollution surveys are a significant component of overhead power lines development. The foregoing tests should be seen as indispensable in cases where the optimisation process leads to the development of novel designs in particular.

1.5 REFERENCES

[1] Glover J.D. and Sarma M.S., Power system analysis and design, 3rd ed., Brooks/Cole, 2002.

[2] Pillay T. and
maintenance
2004.
[3] Grainger J.J.
1994. Bisnath S. (eds), The fundamentals and practice of overhead line (132 kV and above), 1st ed., Johannesburg, Crown Publications,

and Stevenson W.D., Power system analysis, McGraw-Hili Inc.,

[4] Dorf R.C, The electrical engineering handbook, CRC Press, 1993.

1.6 BIBLIOGRAPHY

[1] El-Hawary M.E., Electrical power systems, IEEE Press, 1995.

[2] Faulkenberry L.M. and Coffer W., Electrical power distribution and transmission, Prentice Hall, 1996.

[3] Billinton R. and Allen R.N., Reliability evaluation of power systems, Plenum Press, 1996.

[4] Conradie S.R. and Messerschmidt L.J.M., A symphony of power - the Eskom story, Johannesburg, C van Rensburg Publishers, 2000.

[5] Kiessling F., Nefzger P., Nolasco J.F. and Kaintzyk U., Overhead power lines, Springer, 2003.

;HAPTER 1

CHAPTER 2

MODELLJ:NG

BY MIKE KORBER, DZEVAD MUFTICAND DIPEEN DAMA

@Eskom

MODELLING

SYNOPSIS

This chapter introduces electrical engineers to the most relevant parameters and techniques used in the electrical design of high voltage overhead power lines. Although the theoretical aspects are emphasised quite extensively in this chapter, when selecting the best design, application and theory need to be integrated to arrive at an optimal design solution. The chapter shows that lumped parameter representations of overline power line models need to be understood before further analysing interconnected electric power systems. Lumped parameter representations allow simpler algorithms to be developed for solving complex networks that involve overhead power lines. The approach for long lines is different since an infinite number of incremental lines, each with a differential length, must be considered. The analyses of long lines provide a more accurate solution for a line of arbitrary length I (m). The chapter ends with consideration of symmetrical components and line transpositions.

2.1 INTRODUCTION

CHAPTER. 2

The economic and operational advantages that can be realised by the availability of a high-capacity transmission network indicate the role of High Voltage (HV) and Extra High Voltage (EHV) lines in the transmission of bulk loads. In addition, the reduction of losses makes it more economical to operate at high voltages. From a systems point of view, the main interest is in the electrical performance characteristics of the overhead power line. These can be expressed in terms of the following four line parameters:

Line inductance in Him (Henrys per metre).

Line shunt capacitance (susceptance) in F/m (Farads per metre). Line resistance in Om (Ohms per metre).

Line shunt conductance in S/m (Siemens or mho (1/0hms) per metre).

If the overhead power line is asymmetrical it is generally not theoretically possible to express the parameters on a per phase basis. However, if some approximations are tolerated, it can be done as a practical convenience.

Line parameter data are necessary for detailed studies of overhead power line phenomena. The simplest constants are series impedance and shunt capacitance for the positive sequence as used in power flow studies. More detailed line parameters are sometimes needed for steady-state problems at power frequency. As an example, it may be necessary to study the current and voltage variations on a long untransposed line or the current distribution among the subconductors of a bundle. Line parameters are also necessary for interpreting transient problems such as switching studies.

Line parameters are sometimes measured after a line has been built. However, some of them are needed in the design stage so they must be computed in advance. The following sections explain the basic equations used in the calculation of line parameters. Modern computations of line parameters use computer programs, and remove the need for tedious hand calculations.

MODELLING

2.2 POWER TRANSFER LIMITS

For a normal line the series reactance is the dominating impedance element. Recalling the load capacity of a power line equation in Chapter 1, the role of the reactance X is illustrated in equation (2.1) below:

p = VSVR L X

(2.1)

where X is the series reactance of the line and is normally much larger than the series resistance. Vs and V R are the sending end and receiving end voltages respectively

To achieve high transmission capacities, designers explore possible means of reducing the magnitude of the reactance.

2.3 OTHER FACTORS LIMITING THE POWER TRANSFER OF OVERHEAD POWER LINES

In most circumstances, power flow limits are the result of constraints over electrical phase shift, voltage drop or thermal effects in lines, cables or substation equipment. In addition to electrical phase shift, voltage magnitude decreases with distance. Generally, for overhead power lines, the maximum allowable drop in voltage is limited to between 5% and 10% of the sending end voltage. The power flow (MVA or MW) that corresponds to the maximum allowable decrease in voltage magnitude is called the voltage drop limit.

To define the load transfer capability, it is very useful to consider two features characterising the 'Ioadability' of an overhead power line:

For a reactive lossless single-phase overhead power line, the surge impedance Zs is defined in equation (2.2):

(2.2)

where XL is the series reactance, Y c the shunt susceptance, L the series inductance and C the shunt capacitance

The Surge Impedance Loading (SIL) on a three-phase basis is defined in equation (2.3):

SIL=(VLL)2 MW Zs

(2.3)

where V LL (kV) is the phase-to-phase voltage

A line loaded to its SIL has no reactive power flowing into or out of it. A line loaded at SIL is known to be loaded at its 'natural loading'.

Figure 2.1 shows a single-phase overhead power line that serves as an appropriate departure point in the analysis.

CHAPTER 21

MODELLING

CHAPTER 2

Figure 2.1: Currents defined for a single-phase overhead power line.

In computing the total inductance per m of line, inductance due to magnetic flux inside (L;) and outside (Lo) the two conductors can be distinguished. The total inductance L1 of the single conductor of one-phase overhead power line is given by equation (2.4):

L = L + L == f.lrf.lo + ~ In ~ Him

1 I 0 8n 2n r

(2.4)

where:

D

=

the distance between conductors (m) conductor radius (m)

4n x10-7, permeability of free space (Him) relative permeability of material (conductor)

internal inductance of conductor (independent of wire size [1]) external inductance of conductor

r

=

=

=

=

=

Substituting the value of f.lo in equation (2.4) gives the following simplified expression (2.5) for L1:

L1 =2x10-7(~+ln~)H/m

4 r

Expressing the size of conductor in terms of the geometric mean radius (r' = 0.7788 x r; valid for solid conductor; e-1/4 = 0.7788), the inductance can be expressed in a simpler form, as shown in equation (2.6):

(2.5)

D

L1 = 2x 10In(-) Him

r'

(2.6)

To get a feel for the order of magnitude involved, the inductance of a single-phase overhead power line, with the following data, is given by equation (2.7):

r = 1.15 cm D = 4 m

u, = 1 (copper)

L = 2 x 2 x 10-7In( 4 ) = 24.4 X 10-7 HIm

(1.15 x 0.7788 I 100)

The factor 2 used in the above equation shows that the flux linkages per metre are double for two conductors. Multiplying the result of equation (2.7) by ro = 21tf = 314 (for f = 50 Hz), the corresponding reactance X = 0.767 Q/km of the line is obtained.

(2.7)

Another form for the effective inductance for a single-phase line is shown in equation (2.8):

L = L, + L2 - 2Mn HIm

(2.8)

This is in fact the effective inductance of two series-connected coils (with opposing polarities), each having a self-inductance L, and L2, characterised by a mutual inductance M'2 as given by equation (2.9):

iJo 1

M12 = -In- HIm (2.9)

21t D

where D is the separation distance (in m) between the two conductors carrying current in opposite directions

The last term in equation (2.8) is negative because of the opposite directions of the currents 11 and 12 (I = I, = -12), For these currents, the volt drop along the reactance is given by equation (2.10):

IJ.V=V,-V2=jrox1xL VIm

(2.10)

where:

= complex number operator

O) = 21tf - circular frequency

The volt drops for the two conductors are symmetrical in nature. They represent the voltage drop in each conductor. The volt drop expressed by equation (2.10) can be extended to the n-phase overhead power line. It should be noted that the single and three-phase lines are but special cases of the more general case, when the apparent self inductance of a phase i (L) is given by equation (2.8) and the mutual inductance between phases i and j, (Mij), is evaluated by equation (2.9):

IJ.V = jroxI x(L + L M) VIm

I I I IJ

(2.11 )

Figure 2.2 shows the case for an n-phase overhead power line.

Figure 2.2: The case of n-phase overhead power line.

MODELLING I

CHAPTER 21 1 S

MODELLING

It is important to make the following observations:

Equation (2.11) is applicable only in the case where IJ;=O; it cannot be applied to a line where return currents are flowing in the ground.

In adding the contribution of all mutually induced voltage drops in equation (2.11). the summation over i must be extended (except in the case when i=j, self-impedance).

2.4 INDUCTANCE OFTRANSPOSEDAND NON-TRANSPOSED THREE-PHASE OVERHEAD POWER LINES

The application of the foregoing technique is used for finding the inductance of a threephase overhead power line.

The line consists of three conductors of radius R. positioned in a horizontal plane at a distance 0 apart. as illustrated in Figure 2.3. Figure 2.3a shows a section of the untransposed line and Figure 2.3b shows the same line when transposed. When transposed. the line length is divided into three equal sections and each phase conductor occupies all three possible positions in a transposition cycle.

(a)

D

(b) :: tD ~


1/3 113 1/3

Figure 2.3: Untransposed line (a), Transposed line (b).

Using Equation (2.11) for the untransposed line. the voltage drop on phase 1 can be computed by equation (2.12):

t" V1 = J' CDI1 J:lQ_ (_1_ + In _1_) + J' CDI2 J:lQ_ In _1_ + JCDI3 J:lQ_ In _1_ Vim (2.12)

211: 4 R 211: 0 211: 20

By using the current relation given in equation (2.13):

(2.13)

and by eliminating 13 from equation (2.13). equation (2.14) is obtained by collecting terms.

(2.14)

CHAPTER 2

MODELLING

Similarly, the voltage drop for phases 2 and 3 is obtained using equations (2.15) and (2.16):

P 1 0

f,.V2 = jcoI2 _Q_(- + In-) Vim

21t 4 R

(2.15)

P 1 20 P

f,. V3 = jcoI3 _Q_ (- + In -) + jcoI2 _0 In 2 Vim

21t 4 R 21t

(2.16 )

The results contained in equations (2.14) to (2.16) can be summarised in matrix form, as shown in equation (2.17):

f,.V, 1 20 In2 0
-+In- I,
4 R
f,.V2 =jco& 0 1 0 0 (2.17)
-+In- x 12 Vim
21t 4 R
f,.V3 0 In2 1 20 13
-+In-
4 R Before commenting on the results obtained for the untransposed line in equation (2.17), the results for the transposed line can be treated in a similar manner, the only difference being that the three distinct sections of the line created by the transposition must be considered separately. The volt drop for phase one is therefore given by equation (2.18):

1 [. Po 1 20 . Po ] 1 [. Po 1 0 ]

f,.V1=- JcoI1-(-+ln-)+JcoI2-ln2 +- JcoI1-(-+ln-) +

3 21t 4 R 2n 3 2n 4 R

2[jCOI1 ft(..!. + In 20) + jc.oI2 ftln2] V 1m

3 2n 4 R 2n

(2.18)

Reducing the above expression, equation (2.19) is obtained:

. P [1 0l0]

f,.V1 = JcoI1-o - + In-- Vim

21t 4 R

(2.19)

Because of the complete symmetry between phases, the expression for phases two and three can be written directly as shown in equations (2.20) and (2.21):

. Po [1 0l0]

f,.V2 = JcoI2 - - + In-- Vim

2n 4 R

(2.20)

. Po l1 0l0]

f,.V3 = JcoI3- -+In-- Vim

2n 4 R

(2.21)

Summarising in matrix form, equation (2.22) is obtained:

CHAPTER 21

MODELLING

l1V1 _!_ + In .v4D 0 0 11
4 R
l1V2 =jeo~ 0 _!_ + In .v4D 0 x 12 VIm (2.22)
2rc 4 R
l1V3 0 0 _!_ + In .v4D 13
4 R An immediate observation is that, for the transposed line, the volt drop in each phase is proportional to the current flowing in the respective phase, but independent of the other currents. Therefore, the volt drop can be expressed by the expression as shown in equation (2.23):

l1Vj =jXxIj = jeoxLxIj

(2.23)

where X has the same value for each phase and the dimension of Om

Example 2.1

It is interesting to compare the numerical values of the reactance for the transposed and untransposed lines. Consider, for example, a line with:

R= 1.5 cm (conductor radius)

D= 6 m (phase spacing)

For these values the impedance matrices from equations (2.17) and (2.22) become:

l435~10~ 0.69 o j
3.92x10--4 o Om Untransposed
0.69 4.35x10--4
l421~10~ 0 o j
4.21 x10--4 o Om Transposed
0 4.21 x10--4 2.5 CALCULATION OFTHE INDUCTANCE OF BUNDLED CONDUCTORS

CHAPTER 1.

At extra high voltage (EHV), i.e., voltages above 220 kV, corona with its power loss and particularly its interference with communication circuits, will be excessive if the circuit has only one conductor per phase (see Chapter 6).

In the EHV range, the single conductor surface gradient is reduced considerably by having two or more conductors per phase in close proximity. The phase current will now divide between the sub-conductors. Unless the bundle is transposed, the phase current will not divide exactly between the bundle conductors, but the difference is not of practical importance.

The Geometric Mean Radius (GMR) of a symmetric phase bundle for single conductors with uniform current density (no skin effect), is given by equation (2.24):

(2.24)

where:

f.I = relative permeability of the conductor (u= 1 for aluminium and copper) r = radius of the conductor (m)

It leads to the geometric mean radius r' = 0.7788 x r, already quoted above.

For a symmetrical bundled conductor with n sub-conductors, the GMR can be approximated using equation (2.25):

GMR=~nxr'xRbn-l m

(2.25)

where Rb is the radius of the bundle and n is the number of conductors per bundle

When the bundle is described by a uniform conductor spacing S, its radius Rb can be calculated as shown in equation (2.26):

R = S m

b 2sin(nln)

(2.26)

The geometric mean distance (GMD) between the centres of the phase bundles for a three-phase single circuit is given by equation (2.27):

(2.27)

where d12, d13, and d23 are distances between phases 1, 2, 3 in m

Reduced reactance and reduced phase distance (compaction) are also important advantages of bundling. The reduction of reactance results from the increase of the equivalent radius of the bundle compared with the radius of a single conductor and reduced geometric mean distance. The method of calculating the reactance of symmetrically bundled conductors involves, as before, the calculation of the self and mutual inductances as shown in equation (2.28):

GMD

L =2X10-7XI(m)Xln(~ ) H

n nxr'xRbn-1

X =314xL Q;and

_ 1

M = 2 x 10 7 x I x In(--) H GMD

(2.28)

MODELLING

CHAPTER 2/

MODELLING

2.6 LINE CHARACTERISTICS AND PARAMETERS FOR THREE-PHASE SINGLE CIRCUIT SYSTEM

The basic overhead power line consists of phase conductors (the three-phase system is predominant), ground wire and the earth return (Figure 2.4).

V11 -------------------------V21 V 12 V 22

V 13 V2J

V1g---------------------------------------------------------1V2g

V11' V12, etc., are phase-to-ground voltages

Earth retu rn cu rrent

~--------------- !

Figure 2.4: A three-phase line with ground wire and earth return.

2.6.1 SERIES IMPEDANCE

The network equation of a three-phase line with one ground wire, as given in Figure 2.4, is given by equation (2.29):

V11 z., Z12 Z13 Zlg 11 V21
V12 Z21 Z22 Z23 Z2g 12 V22
X +
V13 Z31 Z32 Z33 Z3g 13 V23
V1g Zgl Zg2 Zg3 Zgg Ig V2g
or in the form:
[V,] = [Z] x [I] + [V2] (2.29)

(2.30)

The self-impedance of the phase conductors and ground wire (diagonal elements in the above matrix) and earth return paths, are given by equation (2.31):

Z;; = R; + 41tx10-7 xfx[jln(4h; Id;) + 2x(P + jQ)] Om

(2.31 )

wherei= 1,2,3,g

! CHAPTER 2

The mutual impedance (non-diagonal elements in the matrix above) will be given by equation (2.32):

(2.32)

where j, j = 1, 2, 3, g

D'jj = distance between conductor and image of another (m) Djj = distance between conductors i, j (m)

Rj = resistance of conductor (Om)

= frequency (Hz)

= height of conductor / ground wire above ground (m) = diameter of conductor (m)

f

h.

I

d.

I

Qp }

Carson's correction terms for the earth return effect

Equations for calculating Zjj and Zjj' originally developed by Carson, were based on the assumption of an earth of uniform conductivity and semi-infinite in extent. For the assumption of perfectly conducting ground P = Q = O. For any other assumption, P and Q can be calculated from equations (2.33) and (2.34):

n 1 k2 2 .

P = - - r;; kcos8+ -[(0.6728 + In-)cos28 + 8s1n28]

8 3,,2 16 k

(2.33)

k3 cos 38 nk 4 cos 48

+ -----

45,fi 1536

1 2 1

Q = -0.0386 +-In- +--kcos8-

2 k s-Ii

(2.34)

nk2 cos28 k3 cos38 k4 2

+ 45,fi --[(In-+1.0895)cos48+8sin48]

64 384 k

with

k = 8.565X10-4XDH

(2.35)

where:

D = 2hj (for self-impedance) or Djj (for mutual impedance) (m) 8 = angle (see Figure 2.5); for self-impedance = 0 radian

p = soil resistivity (Om)

MODELLING I

CHAPTER 21

MODELLING

2.6.2 SHUNT CAPACITANCE

CHAPTER 2

The earth affects the capacitance of an overhead power line because its presence alters the electric field of the line. If it is assumed that the earth is a perfect conductor in the form of a horizontal plane, it can be seen that the field of the charged conductors is not the same as it would be if the earth plane were not present.

The presence of the earth can be accounted for in a simple way by means of so called 'image charges'. These are of the same magnitude as the conductor charge but of opposite sign and placed 'underground' in a symmetrical position with respect to the ground plane.

In accordance with electrostatic theory, the potential difference resulting from the two charges +Q and -Q is given by equation (2.36):

(2.36)

where r 1 and r2 are the distances from a point P to the two charges that contribute to the potential at P

Line voltages measured to ground are a function of line charges. In steady-state AC conditions, the relationship is given by equation (2.37):

(2.37)

where q = charge (phasor) on the conductor per unit length and Pij (known as Maxwell Potential Coefficients, see Chapter 6) are computed from the geometry of the line, as shown in equations (2.38) and (2.39) below:

1

Pii = -2-xln(2Hi Ir;) m/Farad nEo

(2.38)

P'ij = -21 xln(D';/Dij) m/Farad nEo

(2.39)

where:

distance between conductor i and image conductor j' in (m); (see Figure 2.5) distance between conductors i and j, in (m); (see Figure 2.5)

permittivity of free space (Farads/m)

average height of conductors (m)

radius of each conductor (m)

r. =

,

Hi

..

o r

,

,

,

,

,

, , ,

• 0.,

J

D' .. = 2 H.

" I

Figure 2.5: Conductors and their symmetrical images.

For a three-phase configuration with ground wire, as per Figure 2.4, the electrostatic equation is shown in equation (2.40):

V11 P11 P12 P13 P,g Q1
V12 P21 P22 P23 P2g Q2
X
V13 P31 P32 P33 P3g Q3
V1g Pg1 Pg2 Pg3 Pgg Qg
or in simple notation:
[V]=[p]x[Q] (2.40)

(2.41 )

To determine the currents in terms of voltages, the following matrix form relations hold:

[I] = jcox [pr1 x [V] = jcox [C] x [V] = [Y] x [V]

(2.42)

2.6.3 RESISTANCE

where [C] is the capacitance matrix and [Y] is the admittance matrix

The resistance of a conductor is the most significant cause of power loss. Unless specified, the term resistance means 'effective' resistance and is given by equation (2.43):

R = Power loss in conductor Q

1 I 12

(2.43)

where the power is in Watts and the current I (A) is the rms current in amperes. The effective resistance is equal to the DC resistance only if the distribution of the current is uniform throughout the conductor cross-sectional area. For AC lines, some non-

MODELLING I

CHAPTER 21 23

MODELLING

uniformity exists which will be discussed after reviewing the fundamentals of the effective resistance.

The direct current resistance is given by equation (2.44):

R = £!

A

(2.44)

where:

p = resistivity of the conductor (.om)

I = length (m)

A = cross-sectional area of the conductor (m-)

The international standard of conductivity is that of annealed copper (lACS). Commercial hard drawn copper wire has a conductivity of 97.3% of lACS, and aluminium a conductivity of about 61% of lACS. At 20°C, the resistivity p of copper is 1.77 x 10.8 .om and for aluminium is 2.77 x 10-8 .om.

For stranded conductors, the DC resistance is slightly higher than that given by equation (2.44) because spiralling strands make them longer than the conductor itself.

The length of each wire (I;) for each layer can be calculated from equation (2.45):

JIxD. 1+( __ 1?

A

(2.45)

where:

L =
I
D. =
I
A = length of straight wire (m) mean diameter of the layer (m)

lay length; the length in which the wire makes one full turn (m)

2.6.4 THE AC RESISTANCE OF CONDUCTORS

The increased resistance is about 1% for a three-layer conductor.

2.6.4.1

I CHAPTER 1

The AC resistance of a conductor is greater than its DC resistance due to the skin effect, hysteresis, eddy currents and proximity effects.

Skin Effect

Uniform distributions of currents throughout conductors exist only for direct currents (DC). When the frequency of AC increases, the non-uniformity becomes more noticeable. This phenomenon is known as 'skin effect'. In a round conductor the current density increases from the interior towards the surface. At power frequency the skin effect is a significant factor in large conductors and higher effective resistances can occur, typically a few percent above the DC value at 20°C.

2.6.4.2

2.6.4.3

Hysteresis and Eddy Currents

The current in the aluminium strands tends to follow the helical path of the strands in each layer. If the number of such layers is even, then the axial magnetic field produced by one layer tends to be cancelled by the next. If, however, the number of aluminium strand layers is odd, the cancellation does not take place and the resulting net helical current flow tends to produce an axial magnetic field that yields magnetically-induced losses in the steel core wires. These 'core magnetisation' losses in the steel core increase the effective conductor resistance per unit length.

Another effect of the axial field and magnetic material in the core is the so-called transformer effect, which results in a non-linear distribution of current density in the aluminium layers.

Proximity Effect

Another cause of non-uniform current density is the proximity effect. When two conductors carrying alternating current are spaced relatively close to each other, their mutual inductance affects the current distribution in each wire. The non-uniform crosssectional current distribution results in a higher current density on the near side of the conductor, when the current in the conductors flows in opposite directions, and at the far side for currents flowing in the same direction. Bare overhead conductors are normally installed with such large spacings that proximity effects can be ignored.

2.6.5 GENERAL MODEL FOR CALCULATING THE AC RESISTANCE OF A CONDUCTOR

A model for the calculation of the AC resistance of a conductor is described in Appendix 13.1.

2.7 OVERHEAD POWER LINE MODELLING

The general equation relating voltage and current on an overhead power line assumes that the parameters discussed so far are uniformly distributed along the line. It should be mentioned that lumped parameters give good accuracy for short (less than 100 km) and medium length lines (less than 300 km).

The series impedance represents a short line sufficiently well. Figure 2.6 shows in the It (pi) model a representation of a single-phase equivalent of a medium length overhead power line, with half the capacitance of the phase to neutral lumped at each end.

The voltage drop along the conductors of a line is a function of the currents. For AC steady-state conditions, phasors can be used. In the general case, for phase 1 of a multiphase overhead power line, the following equation describes the voltage drop:

:~ = [ZJ[I]

(2.46)

with:

MODELLING I

CHAPTER 21

MODELLING

M [I]

[ZJ =

= vector of conductor voltages = vector of conductor currents

[RJ + jco[LJ the series impedance of the matrix

The matrix of series impedance in equation (2.46) is complex and symmetric. The diagonal elements Zii = Rii + jcoLii are the series self-impedance per unit length of the loop formed between conductor i and the ground return. The diagonal elements Zij = Rij + jcoLij are the series mutual impedance between conductors i and j. The mutual resistance in the mutual coupling, however, is unexpected. It is introduced by the presence of ground and describes the phase shift that takes place in the coupling elements. All resistance elements are calculated by means of Carson's equation, which assumes that the earth has uniform conductivity and is bounded by a flat plane of infinite extent. Only the first terms of Carson's equations are considered.

Vs

C 2

C 2

Figure 2.6: The norninal rt circuit of a medium length overhead power line.

The n model suits the description of a medium length overhead power line if the total shunt admittance, normally a pure capacitance, is divided in two equal parts placed at the sending and the receiving ends of the line. In this case, the model is called a nominal n circuit.

Short and medium length overhead power lines can be considered as special cases of long overhead power lines.

2.7.1 LONG LINES THEORY

I CHAPTER 2

Long lines are described by giving the relationship between the voltage V, current I, and their angle in terms of the distance x at a specific point.

For a long line, the nominaln model in Figure 2.6 does not represent an overhead power line accurately because it does not account for the parameters of the line being equally distributed, but lumps them in one resistance, inductance and capacitance. Figure 2.7 shows the parameters distributed along elements of the line, dx.

MODELLING I

.!S zdx I+dI ,!r
,. 1 i ,. J~
Ys VI ydx IV+dv VR
1
dx J x

Figure 2.7: Distributed parameters of a long line. Considering the differential length dx, the change in voltage and current can be expressed as follows:

dV(x) = zI(x) dx

(2.47)

dI(x) = yY(x) dx

(2.48)

where:

z = series impedance I unit length (Qm) y = shunt admittance I unit length (S/m)

Differentiating equations (2.47) and (2.48) with respect to x and substituting the second equation into the first gives the following expressions:

d2y(x) = zyV(x) dx2

(2.49)

d2I(x) = zyI(x) dx2

(2.50)

For the length s (x = s) the relationship between the voltage and current at the sending end (V) and the receiving end (Vr) of the line can be obtained (Figure 2.7).

Considering x = 0 as the boundary condition for the equations above, hyperbolic solution for equations (2.49) and (2.50) can be written [1,2]:

Vs = v, x cosh(Fr x s) + r, x H x sinh(Fr x s) (2.51)

Is = v, x Jf x sinh(Fr x s) + I, x cosh(Fr x s)

(2.52)

CHAPTER 21 .,

'10DELLING

:HAPTER 2

Four constants A, B, C and D are defined as follows:

A = cosh(,(zY x s)

B = ~ xsinh(,(zY x s)

C = H xsinh(,(zY x s)

D=A

(2.53)

(2.54)

(2.55)

(2.56)

With these four constants and the observation that AD-BC= 1, equations (2.51) and (2.52) can be written as follows:

v, = DVs -BIs

I, = -CVs + AIs

Or in matrix form:

The ABCD constants in equations (2.53) to (2.56) can be rewritten as follows:

A = cosh(yxs)

B = 'Z; x sinhrv x s]

1 C=-xsinh(yxs)

Zc

D=A

where:

Z - f! c-VY"

= surge impedance in Q

= propagation constant in per m

(2.57) (2.58) (2.59) (2.60)

(2.61 )

(2.62) (2.63)

(2.64)

(2.65)

(2.66)

(2.67)

MODELLING

It is possible, however, to And the equivalents of a long overhead power line (>200km) and to represent the line accurately, insofar as measurements at the ends of the line are concerned, by a network of lumped parameters. Assume a n: circuit similar to one in Figure 2.8:

Vs

Figure 2.8: Equivalent TC circuit of a long line.

With reference to Figure 2.8, the source voltage and current at the sending end are given by the following equations:

v, = Vr(1 + Z;Yrr.) + IrZ"

(2.68)

(2.69)

where:

Zrr = total impedance for a long line (0) Yrr = total admittance for a long line (S)

A comparison between equations (2.57) to (2.60) and (2.68) to (2.69) shows that:

A=1+ Z"Yrr. (2.70)
2
B=Z" (2.71)
2
c=y + Z"Yrr (2.72)
rr 4
D=A (2.73) CHAPTER '2

MODELLING

CHAPTER 2

From this, the equations for the main components of the n equivalent circuit of the line as a function of the constants of the long line can be written as follows:

Z1[ = B = t sinh(.[zY x I) = z; sinh(yx I)

(2.74)

2(A-1) 2 Y

Y = = -tanh(-xl)

1[ B Z 2

c

(2.75)

where:

= Length of the line (m)
z = series impedance I unit length (£1m)
y = shunt admittance I unit length (5/m)
Z1[ = total impedance of the line (£1)
Y" = total shunt admittance of the line (5) Example 2.2

Consider a 1000 km, 400 kV overhead power line. The following parameters have been calculated prior to the consideration of length:

r = 0.027 £1 Ikm Iph x = 0.33 £1 Ikm Iph

b = wC = 2.4 X 10.65 Ikm Iph

Calculate the voltage at the receiving end for an open line (zero-load line) when the generator voltage is 400 kV.

(The length of 1000 km is extreme and has been chosen merely to illustrate the concept).

Solution:

Using the length of the line (s = 1000 krn), and the parameters given above, the following can be calculated:

r' = 0.027 x 1000 = 27 £1 Iphase (ph) x' = 0.33 x 1000 = 330 £1 Iphase (ph)

b' = 2.4 X 10-6 x 1000 = 2.4 X 1O-35/phase (ph)

Considering the shunt conductance g to be negligible when compared to the susceptance b, the A, B, C and D constants can be calculated as shown below:

A = cosh(yxs)=cosh(.[zY xs)=coshU(r+ jx)xjb xs)= 0.6293+ j2.829x10-2

B = z; sinh(yxs)= ~ sinh(~(r+ jx)xjb xs)=20.284+ j288.391

C=_!_xsinh(yxs)= II xsinh(~(r+ jx)xjb xs) =- 2.392x10-s + j2.095x10-J

r; ~~

D=A=0.629+ j2.829x10·2

Therefore,

z, =B=20.284 + j288.391

Y 2x(A-1) 2x(0.629+j2.829x10-2-1) 1530 10-5 ·2572 10-3

n . X + J. x

B 20.284+j288.391

Hence, the following can be deduced:

Rn = 20.284 Q I ph Xn= 288.391 Q/ph Gn= 1.530x10-5 S/ph Bn = 2.572x10-3 S/ph

It is known that: Vs = Vr(1 + Zn Yn) + IrZn.

2

B I 0 h I d H Vs-_Vr(1+ZnYn).

ut r = since t e ine is open-circuite. ence,

2

Therefore, from the equivalent TC circuit, the receiving end voltage is :

V = 2Vs

r 2+(Zn Yn)

2x400

2 + (20.284 + j288.391)x(1.530 x10-5 + j2.572x10-3)

v, =634.346- j28.517 =634.9871-2.57° kV

As can be seen, when the load is zero or low, the receiving end voltages are much higher than the nominal voltage of the line (Ferranti effect). These overvoltages are not acceptable and it is necessary to overcome the problem by using compensating shunt reactors.

This example also shows that the lumped parameters ZIT and Yn are lower than those calculated from a simple scaling up of the distributed parameters. Hence, in the analysis of long lines, the equivalent TC model must be applied. If the distributed parameter only had been used in the above example, the calculated voltage would have been some 4% higher.

MODELLING I

CHAPTER 21 31

MODELLING

2.8 UNBALANCED SYSTEMS

CHAPTER 2

In the previous sections, the line was modelled on the premise that complete phase balance or symmetry is maintained. Under this assumption, the parameters are studied on a perphase basis. Further, it was assumed that knowledge of the voltage and current for one phase implied knowledge of the corresponding variables for the other two phases.

In a non-symmetrically loaded or faulted system neither the currents nor the voltages will possess three-phase symmetry. It is no longer possible to limit the analysis to one phase because coupling exists between the three-phases. It is necessary to treat each phase individually.

The unbalanced phasors of a three-phase system can be resolved into three balanced system of phasors called symmetrical components.

The balanced sets of components are:

1. Positive sequence components consisting of three phasors equal in magnitude, displaced from each other by 120° in phase and having a same phase sequence as the original phasors.

2. Negative sequence components consisting of three phasors equal in magnitude, displaced from each other by 120° in phase and having the phase sequence opposite to that of the original phasors.

3. Zero sequence components consisting of three phasors equal in magnitude, with zero phase displacement from each other, i.e., they are co-phasal.

v~ V~

Positive sequence components

Negative sequence Zero sequence

components components

Figure 2.9: Symmetrical components of a three-phase system.

Note: Each of the original unbalanced phasors is the sum of its symmetrical components as shown below:

v, = Va+ + Va_ + VaO Vb = Vb+ + v; + VbO v, = Ve+ + v.: + VeO

(2.76)

where Va' Vb and Ve are the original unbalanced phase voltages and the subscripts +, -, 0 denote the positive, negative, and zero sequence components respectively. The currents are resolved in the same way as shown below:

I, = Ia+ + Ia_ + laO Ib = Ib+ + Ib_ + IbO Ie = Ie+ +Ie_ +IeO

(2.77)

From the definition of the symmetrical components, the following conditions apply:

I - I -j120° - I -j2rr/3

b+ - a+e - a+e

I = I e-j120°

c- a-

(2.78)

The substitution ofthe conditions in equations (2.77) into equations (2.78) gives:

I, = Ia+ + Ia_ + laO

Ib = a2Ia+ + ala_ + laO I, = ala+ + a2Ia_ + laO

(2.79)

where:

a == ei120 = cos 1200 + jsin 1200 = - 0.5 + j 0.866

It can be shown that the following identities apply for the operator a:

a2 = - 0.5 - j 0.866 a3 = 1

1+a+a2=0

MODELLING

CHAPTER 21 33

MODELLING

CHAPTER?

The value of the current symmetrical components is extracted as per equation (2.80):

1 2

1+ = - X (I, + alb + a Ie) 3

1 2

I_ = "3 X (I, + a Ib + ale)

1

10 = - X (I, + Ib + L) 3

(2.80)

Assuming. for a three-phase line. Za = Zb = Ze = ZL' and ZN as an impedance of return path. the following relationships hold:

V+=ZLxI+ V_ = ZL xr.

Vo = (ZL + 3 X ZN) X 10

(2.81)

Sequence impedances are defined. according to the relationships in equation (2.81). as:

Z+ = ZL Z_ = ZL

z, = ZL +3xZN

positive sequence impedance negative sequence impedance zero sequence impedance

The significance of the sequence components for the analysis of faults on overhead power lines can be demonstrated by an example. Consider a single-phase-to-ground fault on a 132 kV line connected through a transformer to a generator as shown in Figure 2.10.

Example 2.3

Consider the single-phase to ground fault depicted in Figure 2.10 below. Assume that the generator neutral and transformer star-point are solidly grounded. The sequence impedances are all referred to the 132 kV side.

40 MVA 40 MVA 101132 kV

8-~

X + = X . = Xo = 44.4 Q/phase

Z+= Z.= 15 + j45 Q/phase Zo = 65 + j170 Q/phase

I fault

X + = 92.2 Q/phase X. = 76.2 Q/phase Xo = 21.1 Q/phase

Figure 2.10: Single-phase short-circuit to ground on a 132 kV line.

Using the symmetrical components method, calculate the single-phase to ground fault current Ifault on the 132 kV line.

The following sequence reactance information has been given in Figure 2.10:

Transformer X+ = 92.2 Q/phase
X = 76.2 Q /phase
Xo = 21.1 Q /phase
Generator X+ = X. = Xo = 44.4 Q /phase
Overhead power line z+ = Z_ = 15 + j45 Q /phase
Zo = 65 + j170 Q /phase Based on the above sequence reactance information, as well the information in Figure 2.10, the following symmetrical component network can be derived.

j44.4 Q Iphase

15 + j45 Q Iphase

+

L j44.4 Q /phase

15 + j45 Q Iphase

10 j21.1 Q/phase

65 + j170 Q Iphase

o

95 + j538.3 Q Iphase

Figure 2.11: Symmetrical components circuit for the network in Figure 2.10.

MODELLING I

CHAPTER 21

MODELLING

CHAPTER 2

The solution can be realised from the equivalent scheme of Figu re 2.11:

Uph

I = I = Io = ---'----

+ _ z+ + z_ + Zo

132000 = 24.231 - j137 A

Jj x (95 + j538.3)

Hence, I r; = 124.231 - j137.31 = J(24.231)2 + (137.3)2 = 139.42 A

Therefore, I Ifau'tl = 3x I r; = 3 x 139.42 = 418.26 A

Example 2.4

For long (>150 km) overhead power lines, transpositions are usually required to minimise the voltage unbalance caused on a line. This unbalance results from the electromagnetic coupling effects between the phases of long untransposed overhead power lines carrying significant power. Such lines have different mutual coupling impedances between the pairs of phase conductors and consequently the power flows through each of the three conductors of the line are different. This balancing problem can be very difficult to compensate for, because the unbalance is continually changing due to varying load conditions (i.e., magnitude and power factor).

Unbalanced voltages can cause adverse effects on equipment and the electric distribution system. The effects of voltage unbalance can be detrimental to equipment such as induction motors, power electronic converters, and variable speed drives (VSDs). Therefore, it is important to perform unbalance studies to determine the extent of voltage unbalance that would exist on a line, and from this, to determine a suitable transposition swap sequence to minimise the voltage unbalance [3].

The definition of voltage unbalance is given by the following equation:

Voltage unbalance

(2.82)

x 100%

where:

V = negative sequence voltage

neg

V = positive sequence voltage

pas

A simplified empirical method can be used to calculate unbalance without calculating the sequence components first, if it is assumed that the zero sequence components will be small. The variables described in the empirical equation below are line-to-neutral voltages at a specific point in the network where the voltage unbalance needs to be calculated [4].

71x(V _Vmid_3Vmin)

max 4 4

Voltage unbalance = ------'--------'--

»:

%

(2.83)

where:

Vave = average of the three line-to-neutral voltages

V max = maximum value of the three line-to-neutral voltages V mid = middle value of the three line-to-neutral voltages

V min = lowest value of the three line-to-neutral voltages

To analyse the above equation further, consider the scenario shown in Figure 2.12 below.

Transposition point I

Transposition point 2

8 'I' 117 km H 117km H 117 km

SOURCE
Swap sequence 3 Swap sequence 3 LOAD END Figure 2.12: Simple model of an overhead power line network.

The above diagram shows a simplified network using a balanced voltage source to supply power to a three-phase balanced load at the end of a 351 km 400 kV overhead power line. The overhead power line has the following parameters:

Tower type

Phase conductor type Sub-conductors Sub-conductor spacing Ground wire type 3-phase balanced load

400 kV guyed suspension tower.

Zebra (ACSR, stranding: 54/7, diameter: 28.62 mm). 4.

380 mm.

19/2.7 mm galvanised steel wires. 100 MW (power factor = 1).

MODELLING

CHAPTER 21

MODELLING

25.1 m

CHAPTER 2

The attachment points of the ground wire and the phases of the overhead power line are described by Figure 2.13.

Ground wire Ground wire

• i •

~T~

.......... ·e .. ·· ·· ··...?:·1} .. ·~ _ !:.J} .. ~ e .

Phase A Phase B Phase C

31.4sm

8.5 m

8.5 m

Figure 2.13: 400 kV Guyed suspension tower configuration indicating the conductor attachment heights.

DlgSilent [5] software simulation tools were used to model the full system for investigative purposes.

A Swap Sequence System (SSS) was devised to investigate all possible transposition sequences that may exist for a line. For any three-phase line there exist six possible swap sequences at a transposition point, as shown in Figure 2.14. Sequence one is considered as 'no transposition'. The left hand side of any of the sequences is considered as the input side.

><: ~



1 2 3
~ ~
><
4 5 6 Figure 2.14: Swap sequences defined for phasing and unbalance study.

MODELLING

The swapping options one to six shown above were applied to cover all combinations on the two transposition points while calculating the voltage unbalance at the load. Ideally, there should be two transposition points along a line between the sending and receiving ends. These points should be at the 1/3 and 2/3 positions on the line to balance the mutual coupling as far as possible.

Various swap sequences were investigated to determine the smallest phase unbalancing at the load. Electrostatic studies (i.e., no load) were first performed to establish the expected voltage unbalance due to the electrostatic interaction between the phases of the lines. The swap sequence used for Transposition Point 2 (TP2) was held constant while different swap sequences where inserted at Transposition Point 1 (TP1). This procedure was repeated for various fixed swap sequences held at TP2. Thirty-six different combinations were considered and the results revealed the following conclusions for the highest and lowest unbalance results respectively. These results are shown in Table 2.1 below.

Table 2.1: Results of the electrostatic studies considering transposition.

Transposition point 1 Transposition point 2 % Voltage unbalance
Swap sequence number Swap sequence number Load end
1 1 1.932
3 3 0.229 As can be seen from the above table, the unbalance at the load for an untransposed overhead power line was calculated as 1.932%, and the unbalance for a transposed line as 0.229%. Thus, by transposing at equal distances along a overhead power line and using swap sequence three at both transposition points, the unbalance seen by the load is improved significantly.

A further investigation (considering a transposed overhead power line i.e., swap sequences 3 used at both TP1 and TP2) was carried out to determine the sensitivity of the voltage unbalance (on a transposed overhead power line) to the load's specifications. The investigation entailed varying the three-phase balanced load from 100 MW to 400 MW in steps of 100 MW with a constant power factor. This study was repeated with different power factors from 1.0 to 0.90 lagging. The results, as summarised in the Table 2.2, show the effect of the line loading for different power factors on the voltage unbalance. The results indicate that the highest voltage unbalance can be expected for heavily loaded lines with low power factors.

CHAPTER 21

10DELLING

HAPTER 2

Table 2.2: Results of the study investigating the effects of load and power factor on voltage unbalance.

Voltage unbalance %
Load
PF = 1 PF = 0.95 PF = 0.9
OMW 0.229 0.229 0.229
100 MW 0.407 0.433 0.446
200MW 0.612 0.690 0.735
300MW 0.878 l078 1.226
400MW 1.270 1.835 2.431 The above results have been graphically depicted in Figure 2.15 below.

Conceptual study

3

2.5
2
<J)
u
C
«I
~ 1.5
..0
c
::J
~ 0.5

o

OMW

100MW

200 MW Load

300 MW

400 MW

--PF = 1 ---+-PF = 0.95 .- PF = 0.9

Figure 2.15: Graphical depiction of the results shown in Table 2.2.

Analysing the results in the table, and the graphs shown above, it can be seen that an increase in load or a decrease in power factor will have a negative effect on the calculated voltage unbalance values of the line. This investigation further revealed that voltage unbalance is sensitive to the magnitude of the load, and its power factor.

The findings of the unbalance studies of electromagnetic coupling between the phases of overhead power lines during transposition studies can be summarised as follows:

The swap sequence system (SSS) enables line designers to analyse all possible transposition sequences that may exist for various overhead power line configurations. Since the system also takes into account the no-transposition sequence ('SSS1'), the analysis helps to determine the number of transposition points required.

Ideally, a long three-phase line should contain two transposition points at least at the 1/3 and 2/3 positions on the line. Where more lines in parallel exist, the SSS can assist in determining whether fewer transposition points could be used to reduce the project costs associated with constructing expensive transposition towers.

Even with ideal 1/3 and 2/3 position transpositions, the voltage unbalance contributions of a line increase with increased line loading and decreased power factor.

2.9 CONCLUDING REMARKS

The basic parameters and techniques used in the electrical modelling of high voltage overhead power lines have been discussed in this chapter. It has been shown that lumped parameter representations of power line models can be used to develop simpler algorithms for solving complex interconnected electric power systems. The electromagnetic coupling effects between the phases of an overhead power line and the effects of the resulting voltage unbalance have been described in detail by means of examples.

2.10 REFERENCES

[1] Elgerd 0.1., Electric energy system theory, New York, McGraw Publishing, 1971.

[2] Gross CA., Power system analysis, New York, John Wiley and Sons, 1975.

[3] IEC/TR 1000-2-1, Electromagnetic compatibility (EMC) - Part 2, 1990.

[4] Smit I., Koch R.G. and Hennesy T.o.J., Eskom power quality reference guide, Rev. 2, April 1996.

[5] DIgSILENT, User manual for PowerFactory software, http;lIwww.digsilent.de

2.11 BIBLIOGRAPHY

[1] Gonen T., Electric power transmission engineering, New York, John Wiley and Sons, 1972.

[2] Gracey G.C, Overhead electric power lines, London, Ernest Benn Limited, 1974.

[3] Braunstein A., Power system in steady state, Tel Aviv University, 1970.

[4] Stevenson W.o., Elements of power system analysis, New York, McGraw Hill, 1982.

[5] EPRI, Transmission line reference book (345 kY and above), z- ed. revised., Palo Alto, Project UHY, 1987.

MODELLING

CHAPTER 21

"'IODELLING

:HAPTER 2

CHAPTER 3

POWER SYSTEM ANAL YSJ:S AND PLANNJ:NG

BY RIAZVAJETHAND ROY ESTMENT

@Eskom

POWER SYSTEM ANALYSIS AND PLANNING

SYNOPSIS

The transmission planning process starts with the definition of needs. It then moves through the steps of formulating and analysing alternatives. choosing the best option. performing sensitivity analyses and carrying out the economic justification. This chapter deals with each of these steps and provides an introduction to the various components that make up a typical power system.

Also covered are aspects relating to power system analysis. project costing and reliability calculations. In addition. the need for a set of planning criteria that will establish the standards for network reliability. power quality. safety and environmental considerations is highlighted. The chapter concludes with comments about the challenges facing the power system planner. including the need for maximising profits. issues around deregulation. protecting the environment. customer demands and ageing networks.

3.1 INTRODUCTION

The power system planner designs electric power networks that are of an appropriate technical standard and an adequate level of reliability. The proposed configuration must be economically viable and environmentally acceptable. The time required to execute the project is also relevant since this can affect its viability. It is therefore important to start planning studies several years before the project is required.

All these activities call for creative thinking capabilities. as there are usually several uncertainties. scenarios and options. Often with too much haste. an attractive option may be left out [1]. A carefully conceived network plan will thus yield handsome rewards for the network owner.

Power system planning covers the three main activities of a utility. namely. Generation. Transmission and Distribution. The focus in this chapter is on transmission planning. although many of the concepts are common to the other two activities. The aim of this chapter is to give an overview of the planning process and introduce important aspects relating to power system planning. It will serve as a basis for the next steps in the asset creation process. namely. the design of the overhead power line. the main emphasis of this book.

The Transmission Planning process can be divided into seven activities as shown in Figure 3.1.

:HAPTER 3

POWER SYSTEM ANALYSIS AND PLANNING

Identify the need for network
expansion.

Formulate alternative options to
satisfy this need.

'\
Analyse these options to check
their compliance with agreed
technical and environmental
standards and discard those that
-, do not comply. ~

Select the best option after
considering the Net Present
Value of each option over the
-, project life cycle.

/' '\
Perform sensitivity analyses
for each of the uncertain variables
to ensure the chosen option is
robust and flexible. ~

Justify the preferred plan from an
economic point of view, using as
a basis the investment criteria of
\. the utility. ~

Based on the above, obtain
approval of the plan and initiate
\. its execution. Figure 3.1: The planning process.

CHAPTER 31

POWER SYSTEM ANALYSIS AND PLANNING

3.2 IDENTIFYTHE NEED FOR NETWORK EXPANSION

The identification of a problem in a utility transmission system, or the need to strengthen it, is usually the result of:

Increased loads at existing supply points or as predicted by the load forecast. New loads.

Additional generating capacity.

Decommissioning or mothballing of generating plant. Need for improved reliability of supply.

Need for improved quality of supply.

Customer requirements.

Statutory considerations such as the government or electricity regulator making it essential to adhere to power quality standards.

Need to reduce transmission losses.

Need to reduce operating and maintenance costs. Refurbishment of an existing system.

Interconnection with neighbouring utilities and countries. Planning for emergencies after catastrophic failures.

The load forecast is one of the main triggers for an expansion plan. Formulating an accurate load forecast is difficult, since it involves predicting the future. Factors that influence the forecast include historical load growths, current and predicted future state of the local and international economy, the political situation and the demand for raw materials and other products.

Once the need is identified, the planner commences with power system studies. The most basic studies are those of load flow and fault level, which simulate voltage profiles and power flows in a power system under steady state or fault conditions. It is wise to verify that input data is of a sufficiently high quality to ensure that the network model represents the real condition as closely as possible.

Load flow studies are conducted to determine whether the existing or proposed network is capable of supplying the loads connected to it under steady state conditions. Important inputs are, e.g., the transmission line parameters (see Chapter 2), network topology, substation and generation station configurations. Important outputs of a load flow study are the real and reactive power flows, voltage levels, power factors and network losses. Modern computer programs can provide speedy calculation of these results, even when the system is large.

The calculation of line and overall network losses is important. The planner must work closely with the line designer to ensure that conductors are chosen to minimise network losses from an overall system point of view. This is explained in more detail in Chapter 15.

Fault level studies are conducted primarily to determine whether existing or new switchgear and other equipment are capable of interrupting the fault currents. Fault level studies are also required to assess whether a high voltage direct current (HYDC) convertor station will function properly, or if harmonics will be a problem. The planner must be aware that single-phase fault currents can sometimes be higher than three-phase fault currents, especially in the vicinity of power stations.

CHAPTER 3

POWER SYSTEM ANALYSIS AND PLANNING

In addition to load flow and fault studies, stability studies [2,3] are also performed to determine if the power system will remain stable under normal operating conditions and regain stability after being subjected to a disturbance. There are three categories of stability studies that were defined by Cigre-IEEE in 2003 [4], namely:

Rotor angle stability (both small-signal and transient). Frequency stability.

Voltage stability.

Rotor angle stability studies are conducted to determine whether the generators will remain in synchronism after a disturbance. These studies are typically performed to check for stability following the integration of a new power source or a major system reconfiguration close to one or more existing power stations. The study time frame for these simulations is normally up to 10 seconds. The two types of rotor angle stability are large-disturbance or transient stability and small-signal stability.

Large-disturbance rotor angle stability or transient stability, as it is commonly referred to, is concerned with the ability of the power system to maintain synchronism when subjected to a severe disturbance, such as a short-circuit on an overhead power line or the interconnection of two systems. The resulting system response involves large excursions of generator rotor angles. These studies are normally done in the time frame of less than one second. The results are critically affected by the size and type of machines, by the type of loads connected and by the nature of the disturbance.

Small-signal rotor angle stability is concerned with the ability of the power system to maintain synchronism under small disturbances such as a load increase or the switching of a line. In modern power systems, small-signal stability is largely a problem of insufficient damping of oscillations due to the swinging of the units at a generation station with respect to the rest of the power system, or with respect to units at other locations in the power system. These oscillations are mainly caused by two or more groups of closely coupled machines being interconnected by weak ties. Small-signal analysis can determine whether the various options such as HVDC, FACTS, SVc, Power System Stabiliser applications and AVR tuning, can eliminate the identified stability problems.

Frequency stability refers to stability issues that occur beyond the transient period, normally 0 to 30 minutes post-fault, during which time slower responding equipment or controls in the power system may be affected, such as, boiler controls at power stations, tap-changers, overload tripping and under-frequency load shedding. The associated stability problems happen occasionally and could involve uncontrolled cascading outages or an unusual combination of circumstances and events that causes part of the system to separate into electric islands. Instability that may result occurs in the form of sustained frequency swings leading to tripping of generating units or loads.

Voltage stability studies are performed mainly to determine how close to the load transfer limit or voltage collapse point the network is being operated, or to evaluate by how much the voltage will fluctuate for a small change of load. Voltage instability occurs due to the inability of the system to meet reactive power demands or when the network transfer limit is reached. Important to voltage stability studies is the modelling of loads and control devices such as transformer tap-changers.

CHAPTER 31

'OWER SYSTEM ANALYSIS AND PLANNING

Depending on circumstances, other studies, such as those listed below, may be required. Based on need, it is up to the planner to determine when such studies are required.

Voltage unbalance (negative phase sequence) studies are conducted to analyse systems with large unbalanced loads (for example, single phase loads, traction loads) or to determine the need for transpositions of overhead power lines.

Sub-synchronous resonance (SSR) studies are performed when a power station is connected to a main load centre through heavily compensated overhead power lines using series capacitors or series FACTS devices. These studies quantify the risk of an electromechanical resonance developing between the turbo-generator and the power system at sub-synchronous frequencies [5].

Voltage dip studies are performed to determine the effects of different network configurations on the magnitude of voltage dips at specific points in the network when a fault occurs. These studies are also performed to check that sensitive loads will not be subjected to unacceptable voltage dips.

Other studies may be performed on long lines to determine steady state overvoltage due to the Ferranti effect when a line is open at one end (via load flow studies), or the transient overvoltage when switching occurs (via travelling wave studies). Such studies indicate to the planner when line reactors, magnetic type voltage transformers (VTs) or switching resistors are required to limit such overvoltages. These studies are also relevant to the line designer in terms of insulation co-ordination.

Harmonic studies are performed to assess the impact of harmonic generators or amplifiers (such as ACfDC rectifiers for large loads, FACTS devices, HVDC converters or fixed series and shunt compensation devices) and to determine the size of the harmonic filters required to reduce the harmonics to acceptable levels.

Other studies, such as fast transient analysis and detailed modelling of control circuits may be required for the design of HVDC converters and FACTS devices and for the connection of large converter loads (for example, large rectifiers supplying aluminium smelters or electrolysis plants). Equipment suppliers sometimes conduct these studies to ensure compliance with the statutory quality of supply standards [6].

3.3 FORMULATION OF ALTERNATIVE OPTIONS

The planner must produce alternative configurations, which are likely to resolve the problem at hand. The quality of such configurations depends on the planner's knowledge, experience and imagination. It is crucial to establish a team with multi-disciplinary skills so that all technical, operational, cost, economic and environmental aspects are taken into account. In addition, it is wise to consult with the end customer who will benefit from the improvements and who may contribute to the solution, be it sometimes through modifications at his own facility.

In this process, it is important to consider the effect of the configurations on the environment. Most countries have introduced legislation requiring that an Environmental Impact Assessment (EIA) report be compiled by the developer, before permission to proceed with the project will be granted. While requirements, processes, and procedures vary from country to country or even regionally within a country, the EIA process typically requires public participation and acceptance before the project is allowed to proceed (see Chapter 5).

:HAPTER 3

POWER SYSTEM ANALYSIS AND PLANNING

The quality of supply specified by the customer determines the minimum performance required. For example, does the customer require a firm supply ((n-1) contingency) or would a single radial supply be sufficient? Are there any special requirements regarding voltage regulation or harmonic levels? Modern industrial and computerised processes call for increasingly improved quality of supply. Voltage dips as a result of overhead power line or other equipment faults can have major impacts on both the network owner's equipment and customer equipment [7]. This makes it important for the planner to specify required line and substation performance and reliability figures, as this will affect the design and cost of the installations.

When reinforcing an existing point of supply, the appropriate voltage is often the same as that already in use. For supplies to a new load or power source, the voltage needs careful consideration, taking into account the existing network and the required transfer capacity. Most planners are of the opinion that, when in doubt, the higher of two voltages is often the correct solution, because of the long economic life of the new infrastructure.

When several lines are already connected to an existing supply point, or the required increase in capacity is substantial, a higher voltage may be more appropriate. If environmental constraints make a new line impractical, it may be necessary to recycle an existing servitude (right-of-way) by replacing an existing line with another of higher voltage and capacity. The original circuit can be reinstated, if necessary, by building a dual voltage multi-circuit line (for example, one 400 kV circuit and one 132 kV circuit on one structure).

It is also interesting to note that multi-phase lines are currently under research with some experimental twelve-phase lines already in operation in the USA [8]. These configurations give superior electrical performance and allow more power flow for a restricted power corridor. However, multi-phase lines are generally not economically viable at present. Complex support structures and the high cost of the transformers are their main disadvantages.

Having decided on the voltage level, it is necessary to select suitable equipment such as transformers, switchgear, overhead power line designs, capacitors, reactors, and FACTS devices. It is also necessary to decide on substation arrangements. Normally, the planner will have the choice between a limited number of standard transformers and switchgear. Because of the costs of additional switchgear and protection, the better scheme is often to provide firm supplies with two transformers rather than three, unless the load is substantial (such as an aluminium smelter).

In the formulation of alternative solutions, it is important to look at all practical options and scenarios. Integrated electricity planning is a process through which the best solution to a particular electricity supply problem can be obtained, while taking into account all possible scenarios. These are typically higher or lower than expected load growths, or a higher than expected capital cost or cost of finance.

To overcome network inadequacies, two kinds of remedies can be applied, namely, demand-side and supply-side solutions.

CHAPTER 31

POWER SYSTEM ANALYSIS AND PLANNING

Supply-side options are the more traditional ways of solving capacity shortages. They are also more extensive than demand-side options. The following can be considered as possible supply-side options:

Increased overhead power line capacity. Increased transformer capacity.

VAr compensation.

Increased generation capacity. Import from other utilities.

Network reconflguration to optimise existing network capacity. Projects to improve the performance of the existing network.

A different approach (to match supply and demand) is to influence the demand instead of the supply. This can be done in a number of ways, for example:

More efficient use. Load interruption. Peak clipping.

Load shifting. Valley filling.

Strategic conservation. Tariff structures.

These options can be implemented through interruptible supply agreements, time-of-use tariffs, energy efficient appliances or building designs, or connecting new customers with load patterns that will improve the system load factor. Demand-side options may defer the need for system strengthening or even avoid it when growth is low.

Alternative solutions should therefore include demand- and supply-side options where possible. The planner must be aware that initially, most customers will believe that their load pattern cannot be modified.

The impact of long-term generation expansion plans on the transmission system also needs to be analysed. The impact of interconnections with neighbouring electric utilities (local as well as international) must be considered and analysed.

3.3.1 SELECTION OF THE SYSTEM VOLTAGE LEVEL

In most countries where a power network already exists, selecting a transmission voltage will be dictated by prevailing voltages. However, it is the planner's responsibility to ensure that the chosen voltage is optimum for the particular application.

The following factors playa leading role when selecting a voltage level:

Use of standard voltages, for example, those recommended by IEC 60038 [9]. EXisting network voltage levels.

Estimated cost of line losses over the life cycle.

Alternating current versus direct current transmission [10] (see Chapter 4). Projected normal and emergency power flow during the life cycle of typically 25 years.

:::HAPTER 3

POWER SYSTEM ANALYSIS AND PLANNING

Length of lines.

Cost of substations. Environmental impact.

From an environmental point of view, it is beneficial to consider higher voltages and HVDC, since this means less conductor mass to transmit the same amount of power. However, the high cost of convertor stations in the case of DC and the high cost of transformation in the case of AC, may dictate that a lower voltage or an AC solution is more feasible. In some cases, underground cable needs consideration in highly built-up areas. The cost of underground cable is high and overhead transmission is the preferred choice in most cases [11].

Typical Eskom practice is to consider voltages in the 11 kV, 22 kV and 33 kV range for distribution of up to 15 MVA per line circuit for relatively short distances, and slightly longer with appropriate voltage boosting technologies. 22 kV is now the preferred distribution voltage for new systems.

The use of voltages in the 66 kV, 88 kV and 132 kV range is used mainly to transfer power over longer distances. Normal power flows of up to 60 MVA over distances up to 150 km are transmitted this way. These lines are normally designed to withstand shortterm emergency loadings of up to 150 MVA. 400 kV is the preferred voltage for the Eskom transmission backbone. 275 kV was selected in the 1960s as the next level above 132 kV. It has now been superseded by 400 kV, except in cases where extensions to existing 275 kV networks are required.

Voltages in the 220 kV, 275 kV and 400 kV range are used to transmit power above 150 MVA over distances of up to 2000 km, with intermediate substations. In these cases, voltage stability must be given careful consideration (with appropriate compensation).

For transmitting power over distances longer than 2000 km, extra-high and ultra-high voltages (765 kV) and HVDC (± 533 kV) are considered.

765 kV was chosen for reinforcing the existing 400 kV grid supplying the Cape. Lower than expected growth rates during the past 15 to 20 years have, however, precluded its wider use.

To determine the optimum transmission voltage, the planner should compare various voltage levels for a given load growth scenario. The capital cost of the overhead power lines and substations, as well as the life cycle cost of the line losses and equipment maintenance, must be estimated for each voltage level. The voltage level resulting in the lowest overall total cost is then selected. If two voltages yield similar total costs, then the higher voltage is usually selected so as to allow flexibility for a higher load growth scenario.

3.3.2 COMPONENTS OF A TYPICAL POWER SYSTEM

A typical power system consists of the major equipment types described in the following pages.

CHAPTER 3)

POWER SYSTEM ANALYSIS AND PLANNING

3.3.2.1

3.3.2.2

I CHAPTER 3

Generators

These usually consist of a mix of hydro-electric, coal-, gas- and nuclear-based power stations. More recently, the demand for renewable energy sources has led to the research and development of wind turbine, small-hydro, solar, wave and tidal based power stations. In addition paper plants and sugar processing plants have material that can be burnt to generate electricity.

As electrical energy cannot be easily stored, generators match their output to the load requirement. They have automatic voltage and frequency controls and sufficient spinning reserve to meet the expected demands of the day. Pumped-storage schemes are valuable for using off-peak power to pump water to an upper reservoir for availability to generate energy during peak times. The electric model representing each of these configurations is set up by the planning engineer to predict the machine performance.

Planning the size, mix and location of new generation is beyond the scope of this book.

Transformers

Transformers are required to interconnect systems operating at different voltages. The most common applications are to step up the generator voltage to transmission voltages and to step down from transmission voltages to provide supplies to distribution networks or large end-user customers.

Selecting transformers generally involves deciding on the number of units and the voltage ratios, MVA ratings, insulation type, winding connection and vector group, impedance, type of tap-changer, and range and size of taps. Most utilities have selected a limited number of standard transformers. For each voltage combination, there is a high and a low power rating, a standard impedance value and a defined voltage range for the tapchanger. The higher MVA rating is dictated by the maximum current allowed on the low voltage side.

The number of transformers in service depends on the relative costs of transformers and switch bays, and the required degree of reliability. Where switch bays are equipped with expensive high-voltage circuit breakers, it is economic to use as few transformers as possible. However, if only one transformer is used, its outage would generally lead to the system being unable to supply any energy. Most of the time there are at least two units available, so that the loss of one will not cause a loss of supply. An installation of this type is known as a 'firm' supply. The additional cost is usually less than the value of the un-supplied energy that would be incurred by the customer as a result of the loss of one transformer. The repair time on faulty transformers is often very long. To cater for this, one or several standard transformers (of each type) are kept as strategic spares.

In emergencies, transformers can be overloaded for short periods of time. The acceptable limit depends upon the preceding load and the ambient temperature. Overloading should not occur regularly since it accelerates degradation of the insulation, reducing the life of the transformer and increasing the risk of a fault.

3.3.2.3

3.3.2.4

POWER SYSTEM ANALYSIS AND PLANNING:

Transformers are efficient and the losses are therefore low. However, with increasing pressure to minimise costs, it is advisable to quantify the cost of transformer losses over their lifetime, since manufacturers can build low-loss transformers at a certain cost premium.

SWitchgear

Switchgear comprises all devices used to connect or disconnect parts of an electricity supply system. This includes isolators or disconnecting switches of varying complexity, and various types of circuit breakers.

Switchgear must be able to withstand the currents that flow during a fault. Circuit breakers must be able to interrupt the current in a circuit on full load and are designed so that the transient overvoltages, when opening or closing the contacts, are kept to reasonable limits. This problem arises especially when switching long lines, which contain trapped charge. Shunt reactors and magnetic voltage transformers (VT's) can help to drain the trapped charge. These phenomena are studied by means of travelling wave models.

Switchgear, in the same way as transformers, has also been standardised. For the planner, it is important to consider that SWitchgear may maloperate on rare occasions and this can result in system blackouts.

Overhead Power Lines

Overhead power lines transmit bulk power over long distances. Their characteristics and design are discussed in detail in other chapters of this book.

The aspects most important to the planner are:

Electric model as discussed in Chapter 2.

The phase and bundle spacing and their impact on power transfer. The predicted line performance.

Emergency restoration times.

The servitude requirements and whether new lines will be needed in the same corridor at a future date.

The time required to obtain the servitude and to construct the line. Environmental considerations (see Chapter 5).

The line rating under normal and emergency conditions.

Capital cost estimates and the operation and maintenance costs. The cost of losses.

Whether the line can be maintained live and the duration of planned maintenance activities.

Coupling, induction and unbalance studies and the need for transpositions, especially for long lines or highly loaded shorter lines.

Telecommunication paths over overhead power lines are also relevant. The planner, telecommunications engineer and line designer should collaborate to ensure that an optimum solution is obtained, for both power line carrier and optical fibre.

This all calls for a close working relationship between the planning engineer and the line design engineer.

CHAPTER 31

POWER SYSTEM ANALYSIS AND PLANNING

3.3.2.5

Capacitors

Series and shunt capacitors are used on the transmission network to increase the power transfer capacity of a line or to maintain and regulate system voltages within specified limits.

3.3.2.5.1 Series capacitors

Series capacitor banks are more expensive per MVAr installed than shunt capacitors banks. This is because the whole installation must be insulated to withstand the line voltage. Shunt capacitors should therefore be considered first for overcoming power transfer limitations due to voltage drop.

However, for long lines where the permissible amount of shunt compensation is limited by voltage stability or harmonic resonance, both series and shunt capacitors may be justified. Before installing or modifying series capacitors, sub-synchronous resonance (SSR) studies should be conducted to verify that no power stations are exposed to SSR.

3.3.2.5.2 Shunt capacitors

Shunt capacitor banks are installed for voltage regulation or power factor correction. As a result, they slightly decrease the loading of overhead power lines and power transformers.

To reduce reactive power flows and to control voltages effectively, shunt capacitors are normally installed as close as possible to the load centres. The reactive power generated by shunt capacitors is proportional to the square of the voltage. A volt drop of 5% therefore leads to a reactive power drop of about 10%.

3.3.2.6

Reactors

Series and shunt reactors are used on the transmission network to regulate the reactive power flow and maintain system voltages. Reactors absorb extraneous reactive power.

3.3.2.6.1 Series reactors

Series reactors are used to:

Reduce excessive fault levels.

Balance load sharing between parallel circuits of unequal impedance.

Series reactors are normally built without a magnetic circuit so as to maintain their linear characteristics when the current is high. This makes them expensive when compared with iron-cored reactors.

3.3.2.6.2 Shunt reactors

Shunt reactors are used to compensate the (capacitive) reactive power (MVAr) generated by high voltage overhead power lines or underground cables. The role of the shunt reactor is to:

I CHAPTER 3

3.3.2.7

POWER SYSTEM ANALYSIS AND PLANNING I

Prevent excessive voltage rise during light-load periods (Ferranti effect).

Prevent generators from absorbing excessive reactive power (which may compromise transient stability if not corrected).

Reduce the reactive power swing when a line trips.

Reduce switching overvoltages (by draining the trapped charge).

To reduce or eliminate the risk of secondary arcs on long overhead power lines during single phase auto-reclosing (and thus increase the chances that auto-reclosing will be successful), small neutral reactors must be installed between the neutral of the (three phase) line reactor and earth.

Flexible AC Transmission Systems (FACTS) Devices

Power and voltage are controlled on a transmission system by selecting appropriate generation (scheduling), changing taps on transformers. operating phase-shifting transformers or switching reactors and capacitors in or out. Until recently, control centre operators performed these actions. The objective was to maintain desired power flows and acceptable voltages under changing system conditions.

The advent of thyristors for power control applications and the development of microprocessors has made it possible to provide automatic control of power flows and system voltages. This is quicker and more accurate than manual operation. This new technology is called FACTS and offers the following advantages:

Higher power transfers and operation closer to the stability limit.

Improved dynamic stability and damping performance (reduced power swing). More effective voltage control and compensation on long overhead power lines. Mitigation of power and voltage swings after disturbances.

Mitigation of heavy post-disturbance through-flows that might cause cascading trips, or circuit isolation by out-of-step protection.

An improvement in quality of supply.

The most common FACTS devices are Static VAr Compensators (SVCs) and HVDC systems. The technology for these devices is well proven and readily available. Other FACTS devices have also been developed, but they remain largely experimental

[12,13].

Emerging FACTS devices include:

Thyristor-controlled braking resistor (to reduce over-speed of generators). Thyristor-controlled series compensation.

Static condensor.

Unified power flow controller. Large-scale energy storage devices.

CHAPTER 31 S5

POWER SYSTEM ANALYSIS AND PLANNING

3.3.2.8

Protection

Protection schemes are designed to:

Clear the fault in the minimum time.

Isolate only the minimum number of network components. Limit damage to healthy equipment.

In some instances, maintain stability.

Protection systems often collect information concerning plant status and faults, and transmit the data to the control centre via the Supervisory Control and Data Acquisition (SCADA) system. To ensure safe and secure operation of the system, a SCADA system conveys information about system voltages, loading of lines and transformers, and status of switchgear to the control centre.

At transmission level, the SCADA system also provides a facility for the control centre to operate the switchgear by remote control. This permits fast restoration of supplies in the case of an interruption or a change in system conditions.

3.3.2.8.1 Protection: overhead power line

I CHAPTER 3

The most common primary line protection is impedance or distance protection. It is important to understand the theory of operation so that its inherent limitations can be appreciated.

By measuring voltage and current, the line impedance is continuously calculated by the impedance relay. Under healthy conditions the system impedance, which includes the load impedance, is high and predominantly resistive. In the event of a fault, the impedance as measured from the relay drops to a value equal to that of the line itself or less, depending on the position of the fault. The impedance also becomes more inductive. The relay is set to operate only if the impedance drops below pre-determined values. A time delay is associated with each value.

For overhead power lines, an under-reach philosophy is normally applied to ensure proper discrimination between the zones of surveillance of the various distance relays in the interconnected network. The relays at either end of a line are set to operate in the fast 'zone one' time only if the measured impedance is 80% or less than the line impedance. The relays at both ends will operate in the event of a fault in the central 60% of the line. If a fault occurs within 20% of either end, only the relay closer to the fault will operate. For this reason, teleprotection (typically power line carrier, fibre optic, or microwave) is installed to send a control signal to the relay at the remote end of the line to trip the local breaker. One advantage of this philosophy is that a trip signal is sent from both ends for a fault occurring in the central part of the line. This caters for faults with significant fault impedance. Distance relays typically provide two backup settings, with time delays to ensure clearance of faults if the primary protection malfunctions.

POWER SYSTEM ANALYSIS AND PLANNING

Most lines are also equipped with sensitive inverse-time earth-fault protection to cover high resistance faults. such as a midspan flashover. broken conductors falling to earth. insulator flashovers or high resistance to ground faults.

3.3.2.8.2 Protection: carrier application to line protection

This is described in Chapter 26. However. it is useful to introduce the topic here.

Power Line Carrier (PLC) uses the power conductors to transmit communication signals (protection. data and voice). Coupling capacitors are used to inject the signal onto the phase conductors. while line traps prevent the injected signal from being short-circuited to ground by the low impedance of the substation busbar at each end of the line. Carrier frequencies lie in the range SO to 500 kHz. Above 500 kl-lz, signal attenuation becomes excessive. while below 30 kHz the size and cost of the line traps and coupling capacitors become impractical.

When carrier is used as part of the line protection scheme. it is important that the signal is not corrupted by any high-energy line faults or stub effects on the line. On the other hand. faults external to the line must not cause malfunctioning of the protection system.

3.3.2.8.3 Protection: auto-redosing

Most line faults are transient in nature. It is thus normal practice to provide an automatic redosing feature on all line circuit breakers. Auto-redosing may be three-phase or single-phase. Auto-redosing can also be staged in several steps. for example. with single phase first. followed by three-phase if the fault is still present. Typical redosing delays are 300-400 ms.

3.4 ANALYSIS OFTHE OPTIONS

In theory. system planning problems usually generate a number of possible solutions that would satisfy the planning criteria. The task of the planner is to filter out and discard those options that would be impractical because of environmental. physical or time constraints. or which would be too expensive. Because they take time and effort. planning studies are performed to test only the alternative solutions that can conceivably be implemented in practice and have a reasonable chance of emerging as the best technoeconomic option.

A high-level cost estimate is often made as a preliminary filtering exercise to discard. for example. those options which require too much capital investment.

It is important to bear in mind the capabilities and limitations of the software used for planning studies. It is the responsibility of the planner to ensure that the models used and assumptions made (implicit as well as explicit) are valid for the type of problem being studied. The results of the studies must be interpreted in the light of such assumptions and limitations.

CHAPTER 31

POWER SYSTEM ANALYSIS AND PLANNING

3.4.1 PLANNING CRITERIA AND LIMITS

3.4.1.1

3.4.1.2

3.4.1.3

3.4.1.4

3.4.1.5

I CHAPTER 3

The criteria and limits against which proposed expansion or strengthening plans are evaluated include the technical and statutory requirements which must be observed. and other limits or targets which indicate that the system may be reaching a point where problems are likely to occur. If technical or statutory limits (such as those specified by the local Regulatory Authority) are exceeded. action to rectify the situation must be considered and evaluated. However. the situation may still be acceptable in some cases. and the question of whether or not immediate action is required should be decided on a case-by-case basis.

Line Loading Criteria

Normal line loading limits based on conductor temperature are calculated using conservative assumptions for the simultaneous occurrence of ambient temperature. wind velocity. solar radiation. emmissivity and solar absorption coefficients. In practice. it is possible to obtain higher line loading by applying a probabilistic approach or by adopting a real-time monitoring system. These are discussed in detail in Chapter 13.

Transformer Loading Criteria

Transformers are designed to withstand the continuous loading indicated on the nameplate. It is usual to allow an overload of up to 120% under contingency conditions. Long-term overloading of transformers will result in the reduction of their service life.

Voltage Criteria

Under normal network conditions. with all components healthy and in service. it is common practice to stipulate that the voltage is to be maintained between 0.95 and 1.05 per unit (pu). During unhealthy network conditions. it is usual to allow a minimum voltage of 0.90 pu. When switching components such as lines. capacitors or transformers. the maximum voltage change should not exceed 3% for a healthy network and 5% for a network under contingency. It is also advisable to ensure that a 5% rise in load will not cause more than a 5% change in voltage. Maximum negative sequence voltages for unbalanced loads should be kept below 1 to 1.5% of the nominal voltage.

Stability Criteria

It is usual practice to expect a healthy network to remain stable after a line or busbar fault has been cleared in normal protection times.

For voltage stability. it is usually recommended that power transfer across long overhead power lines be kept to below 95% of the voltage collapse point.

Reliability Criteria

A system cannot be made 100% reliable as planned and forced outages do occur. Multiple outages are always possible despite having. in theory. a very low probability of occurrence (i.e .. if outages are truly independent). From a purely economic point of view. optimum reliability is obtained when the total saving achieved by the utility in reducing the load by one MW is just equal to the value of this unsupplied MW to the customer involved [14].

POWER SYSTEM ANALYSIS AND PLANNING I

In practice, however, the selection of the level of reliability provided to individual customers or groups of customers is a policy decision made by each electric utility. Factors taken into account in determining this level of reliability include socio-economic conditions, the cost of financing projects, and policies pursued by the local Regulatory Authority.

For small loads it is usually not economically justified to have more than a single, unfirm, supply. For large loads, the customer will often require a firm supply, as the amount of load not supplied in an outage condition is substantial. The socio-economic consequences of a prolonged outage can be serious, not only in terms of loss of production, but also in terms of damage caused to the plant. Hence, most transmission systems are designed to withstand the loss of any single component (line, transformer, etc). This is called the (n-1) criterion.

3.5 SELECTION OF THE BEST OPTION

In order to select the best option, capital and life cycle costs, such as losses, and operation and maintenance costs, need to be estimated.

Typical capital costs for standard transformers, overhead power lines, shunt capacitors and reactors, SVCs and terminal equipment should be entered into the planner's database. These costs are used for approximate capital estimates to filter out alternatives that are too expensive. They are accurate enough for comparison purposes, as any errors are likely to affect all alternatives in the same manner. It is important to note that such standard costs may not be accurate enough for an economic justification.

Detailed cost estimates are compiled for the alternatives selected from the initial filtering process. Total annual costs (capital as well as energy losses and operation and maintenance costs) of each alternative are calculated over the economic life of the plant (typically 20-25 years) and discounted back to a present value. The net discount rate and other financial parameters to be used must be in accordance with the utility's capital investment policies.

The costs and benefits of the alternatives, which are all technically comparable (i.e., they provide network solutions for equal time durations), are then evaluated to determine which is the most attractive from an economic point of view.

Energy losses are based on the load factor and the average incremental cost of generation per MWh. These costs, together with other financial parameters should be determined annually by the electric utility. This is to ensure consistent evaluations of all capital projects throughout the utility.

A Net Present Value (NPV) analysis is carried out for each alternative [14]. The results from these calculations are very sensitive to the assumptions made, particularly those in respect of load growth and net discount rate. The NPV is the sum for each year over the life of the plant, of the discounted benefits minus costs of the alternative being considered. This can be expressed by the equation (3.1) [14]:

CHAPTER 3

POWER SYSTEM ANALYSIS AND PLANNING

n

NPV = L (B, - CJ / (1 + r)"

o

(3.1 )

where:

n =
B =
C =
r = Year 0, 1, 2, ... n-1

Benefits stream (e.g., increased revenue from sale of electricity) Costs stream (e.g., capital and operating costs)

Discount rate

Critical in the above analysis is the choice of the discount rate which relates to the cost of obtaining capital. This is closely linked to the level of risk in the project alternative and is an indication of the economic and political stability of a country.

3.6 SENSITIVITY ANALYSES

A certain amount of risk is inherent in every decision based on predictions. Because assumptions such as load growth, discount rate, the impact of demand side management (DSM) programmes, projected expenditure, and the availability and application of new technologies are uncertain, planners have sometimes to investigate several sets of assumptions, all having some probability of occurrence. The best course of action may not always be obvious, as different solutions are likely to be more suited to different scenarios.

Risk can be defined as the magnitude and probability of regret (product of two components) resulting from an incorrect course of action. With this approach, the optimum solution will be the one which results in the minimum regret for all sets of assumptions considered, i.e., the one that will have the least cost allowing for the occurrence of any of the feasible scenarios. This means that the adopted solution should be sufficiently robust and flexible to accommodate most scenarios.

For important projects, the (expected) total cost of a solution should be determined by:

Assigning probabilities to various scenarios (load growth, reliability of supply etc.). Establishing the cost of each of these scenarios.

Multiplying costs and probabilities to obtain the expected cost of the solution.

The solution with the lowest (expected) total cost assessed in this manner would be the optimum choice from a minimum regret point of view. While it may not be feasible to obtain sufficient data to conduct a complete analysis in this way (the calculations are also considerable), planners should keep the concept of 'minimum regret' in mind when recommending a techno-economic solution.

It is clear from the above that solutions with short lead times and where it is possible to stop some of the investment after the decision to go ahead, will be preferred over others.

CHAPTER 3

POWER SYSTEM ANALYSIS AND PLANNING I

3.7 ECONOMIC JUSTIFICATION

Electric utilities around the world are under increasing pressure by regulatory bodies, governments, customers, investors, and other stakeholders to prove that they are investing capital wisely. An economic justification for a project provides assurance to stakeholders that the money is spent efficiently.

Utilities have different capital investment policies and criteria, based on local circumstances. These investment criteria can, however, be grouped into four broad categories:

Positive NPV investments.

Least economic cost investments. Investments to reduce operating costs. Statutory or strategic investments.

The remainder of this section gives a general description of the four categories. Specific projects must, however, be evaluated in accordance with the utility's own investment criteria.

3.7.1 POSITIVE NPV INVESTMENTS

3.7.1.1

This method is recommended for system expansion required to supply one or more readily identifiable new loads. These new loads typically consist of large new customers applying for supply, or existing large customers applying for increased supplies. In these cases, the utility normally furnishes the customer with a quotation for the type and quality of supply specified by the load, which the customer is free to accept or reject. The quotation typically makes allowance for recovery of capital expenditure from the customer via an extension charge which is calculated in accordance with the utility'S pricing policy. The utility will normally embark on capital expenditure only after the customer has formally accepted the terms offered in the quotation.

The NPV of the investment is the result of discounted cash flow calculations over the expected number of years that the asset(s) will be in service (typically 25 years for transmission primary equipment and 15 years for secondary equipment). The following factors are considered during the analysis:

Project capital costs. Additional generation costs.

Additional operation and maintenance costs. Additional transmission losses.

Desired Return on Investment (ROI) and economic life.

Sensitivity Analysis

Experience has shown that, depending on the assumptions made regarding future input parameters, the results of discounted cash flows can vary widely. NPV sensitivity analysis is therefore essential. Sensitivity analyses involving the following parameters are usually performed:

CHAPTER 3!

POWER SYSTEM ANALYSIS AND PLANNING

Discount rate (or desired ROI). Load growth.

Capital expenditure.

Future tariff increases.

Cost of generation.

3.7.1.2

Capital Pay-Back Period

Another useful measure of investment risk exposure is the project capital expenditure pay-back period. The pay-back period measure indicates in how many years the cumulative revenue will be equal to the cumulative cost to generate the additional sales and system losses plus the capital cost of the project. In other words, how long will it take for the project to break-even and thereafter, begin generating a profit?

3.7.2 LEAST ECONOMIC COST INVESTMENT METHOD

This specific investment criterion is used to justify extensions to the existing transmission system which normally serves as the backbone system supplying many end customers. The extensions will improve the network reliability under contingency conditions, for example, the addition of a third transformer at an existing substation with two transformers or, the addition of another overhead power line as a back-up to an existing one. The least cost investment evaluation method is also suitable for justifying strategic spares, large maintenance and refurbishment expenditure. This approach takes into account the probability of network outages and the line loading, as well as the impact of lost load to the network owner and the end customers. It is usually referred to as the probabilistic approach.

This is a relatively well researched area [15], though few utilities are adopting this approach for the economic justification. The tendency is to rather use the (n-1) criterion. The probabilistic approach is used more as a tool to prioritise projects. The method is described in detail below.

This type of economic justification is based on the least-cost planning concept. This is not simply least-cost to the utility, but to the entire community of customers plus the utility. It is consistent with social welfare economic theory. The criterion minimises the total cost of electricity to the partnership consisting of the utility plus the customers who benefit from the extension of the transmission system. In other words, a new scheme or asset should be built only if the extra annual costs incurred by the utility for providing and operating the asset are less than the annual saving to the customers. This saving may arise from the improved reliability or quality of supply provided by that asset.

The extra costs incurred by the utility are the annuallevelised cost of the capital invested in the extension, plus the extra operating and maintenance costs incurred by the extension, less the savings in energy losses due to the extension.

The annual saving to the customers because of the extension includes the cost reductions resulting from the improved quality and reliability of supply. The benefits derived by the customers are difficult to assess, because they involve expected outage times and the value of unserved energy. The cost of unserved energy is not easy to estimate, because it is a function of when it occurs, its duration and its magnitude. It will also vary from customer to customer.

CHAPTER 3

POWER SYSTEM ANALYSIS AND PLANNING I

It should be borne in mind that customers may overestimate the consequences of supply interruptions if they are aware that a high estimate will increase the likelihood that the utility will carry out the extension. It is therefore advisable to check the survey results against other information sources: for example, company balance sheets and income statements, the willingness of the company to provide standby generation or Uninterruptible Power Supplies (UPS) for their critical loads, or their willingness to pay extra for improved reliability.

The costs of such projects are usually recovered through the overall tariff as a transmission charge. Ideally, utilities should promote cost reflective tariffs to minimise the amount of cross-subsidisation that can occur. However, political pressure often prevents utilities from achieving this objective.

As the load grows, the new infrastructure results in reduced transmission system losses and a reduction in unserved energy. This benefit, in monetary terms, grows over the life cycle of the infrastructure. However, the initial cost implications for the new infrastructure remain constant. The operation and maintenance costs increase annually at a rate less than the growth rate of the benefits. This implies that there is a finite point in time (break-even point) during the life cycle when the benefits will exceed the costs of the new infrastructure. After this break-even point has been reached, the infrastructure generates profit for the utility.

3.7.3 OPERATING COST REDUCTION INVESTMENTS

Proposed expenditure that is intended to reduce operating costs can be evaluated by considering the level of reduction that can be achieved by the capital investment. For example, if the maintenance cost of an old circuit breaker is higher than the annual capital plus maintenance costs of a new breaker, new ones should replace the old breakers.

3.7.4 STATUTORY OR STRATEGIC INVESTMENTS

An electric utility must sometimes invest capital to comply with legal requirements, irrespective of whether any financial benefit is derived, other than avoidance of legal costs and fines. These requirements may arise from environmental protection legislation (e.g., noise, corona, electric or magnetic fields, etc.), public or occupational health and safety regulations (e.g., substation HV yard fencing, HV yard lighting, labelling, portable earths, etc.), municipal fire regulations, or other legislation affecting the industry.

It is good practice to classify investments as statutory or strategic only if the expenditure must be made and cannot be justified by any other means.

3.8 CURRENT AND FUTURE CHALLENGES FOR THE PLANNING ENGINEER

Deregulation of the electricity industry is affecting many countries. This will pose a challenge to the Transmission Planner as the need to maximise the ROI will need to be balanced against network reliability and performance. There is an ever-increasing need to operate the network closer to its limits which is referred to as 'sweating the assets'.

CHAPTER 31

POWER SYSTEM ANALYSIS AND PLANNING

Energy trading between interconnected countries and utilities will also mean that a new and varied planning approach will need to be implemented, since these networks will have been designed with different planning criteria.

Environmental issues are affecting the viability of projects to the extent that it is becoming increasingly difficult to obtain servitudes for construction of new overhead power lines. Planners are looking at ways to re-use existing servitudes. Some consideration is being given to multi-circuit and multi-phase lines, compact and higher temperature lines, voltage and thermal uprating, development of superconductors, the development of HVDC lines and lower cost underground transmission systems [8,10,11].

Many studies are also being conducted to find ways to tap-off small amounts of power from long HVAC and HVDC lines when they pass small communities or load areas. The challenge is to do this in a reliable and cost effective manner.

Highly interconnected networks are making the analysis and control of networks extremely complex. Since local line or substation faults can trigger widespread blackouts, the challenge is to balance the benefits of interconnected networks with the risk of system blackouts.

Adverse weather is also making it difficult to design cost-effective lines. This applies to lines that are exposed to violent weather such as tornados, ice and lightning. In addition, the need to have excellent line performance means that rigorous line and substation inspections and maintenance are crucial. There are many instances of trees growing high enough to cause a line fault, so causing blackouts in an interconnected power system.

In many countries, networks are in excess of 40 years old. These networks form the backbone of many electric systems and their ability to continue to perform into the future is a challenge, especially since customers' equipment is becoming more sensitive to poor power quality.

The use of optical fibres on overhead power lines to transmit bulk digital information is now common practice and the financial benefit of this additional use should be considered in the planning stage. This also increases reliance on the fact that overhead power lines should not fail under severe conditions, as there is now more than just power that will be interrupted.

3.9 CONCLUDING REMARKS

This chapter has introduced the technical and financial concepts involved in the planning of new overhead power line or substation projects. The most important challenges facing the planning engineer have been highlighted, namely, the changes due to deregulation, sweating of the assets, ageing networks, rapid load growth, technological advances and the impact of environmental constraints.

CHAPTER 3

POWER SYSTEM ANALYSIS AND PLANNING

3.10 REFERENCES

[1] Ilbury C. and Sunter C; The mind of a fox - scenario planning in action, Cape Town, Human and Rousseau and Tafelberg Publishers, 2001.

[2] Kundur P., Power system stability and control, New York, McGraw-Hili Inc., 1994.

[3] Bayliss C; Transmission and distribution electrical engineering, z- ed., Oxford, Newnes, 2002.

[4] IEEE/Cigre joint task force on stability terms and definitions, Definition and classification of power system stability, Draft Report, Paris, 2002.

[5] Anderson P.M., Agrawal B.L. and van Ness J. E., Subsynchronous resonance in power systems, New York, Wiley - IEEE Computer Society Pr, January 1999.

[6] SABS NRS 048-2, Electricity supply - quality of supply, 1996.

[7] Vajeth R., Dama D. and Mtolo D., Cost of a network fault affecting a transmission supply point, Eskom Research Report No. RES/RRl03122561, November 2003.

[8] Grant J. et aI., Higher phase order transmission line research, Clgre Report No. 220-02, 1981.

[9] IEC 60038, IEC standard voltages,1983.

[10] Rudervall R., Charpentier J.P. and Sharma R., High Voltage Direct Current (HVDC) Transmission Systems Technology Review Paper, World Bank, 2000.

[11] Cigre, Comparison of high voltage lines and underground cables, Cigre Brochure 110, 1997.

[12] Song Y.H. and Johns AT., Flexible AC transmission systems (IEE Power Series 30), London, Inspect IEE, 1999.

[13] Hingorani N.G. and Gyugi L., Understanding FACTS: concept and technology of flexible AC transmission systems, New York, Wiley-IEEE Press, 1999.

[14] Khatib H., Economic evaluation of projects in the electricity supply industry, London, IEE,2003.

[15] Billington R. and Allan R.N., Power system reliability in perspective, London, IEEE Press, 1991.

CHAPTER)

POWER SYSTEM ANALYSIS AND PLANNING

CHAPTERJ

CHAPTER 4

HJ:GH VOLTAGE DJ:RECT C'URRENT (HVDC) TRANSM:ESSJ:ON

BY NELSON IJUMBAAND GARY SIBILANT

@Eskom

-JIG H VO LTAG E DIRECT CURRENT (HVDC) TRANSMISSIO N

SYNOPSIS

In the past 50 years, the installed capacity of HVDC schemes has increased from 20 MW to 80 GW. About 30% of this capacity has been developed in the last 10 years. Most of the HVDC schemes are for point-to-point power delivery over long distances. However, improved developments in converter technology have opened up possibilities of using HVDC systems in distribution networks. Such applications include back-to-back coupling of asynchronous networks or networks operating at different frequencies, connections to off-shore installations and interconnections across environmentally sensitive areas, such as game parks. This chapter introduces historical aspects, and current and future developments in HVDC technology. Emphasis is placed mainly on the applications of this technology and less on technical detail and the principles of operation, since these are covered in detail in other references.

4.1 INTRODUCTION

Whilst by far the majority of the world's electrical transmission is based on High Voltage Alternating Current (HVAC) technology, High Voltage Direct Current (HVDC) technology does provide a viable alternative under certain circumstances. This chapter describes the basic properties of HVDC systems, to compare the key differences between HVDC and HVAC, and shows where HVDC applications differ from those of HVAC systems.

The chapter deals broadly with the evolution of HVDC since its introduction in the 1950s, the principles of HVDC, contemporary technological trends, economic factors, and operational issues. The subject of HVDC line design is introduced. The information given here is thus an introduction to HVDC, and not a design guide.

4.2 BACKGROUND

4.2.1 HISTORICAL DEVELOPMENT

The first power transmission systems were Direct Current (DC). However, the low generated DC voltages limited the distance and capacity of such systems. Following the development of the power transformer in the late 19th century, it became possible to transmit power over longer distances using High Voltage Alternating Current (HVAC) transmission systems. Technological advances in high voltage diode valves, from mercury arc rectifiers to thyristors and Insulated Gate Bipolar Transistors (IGBTs), have made it possible to transmit DC power at high voltages and over long distances.

In 1951, a 30 MW experimental HVDC plant was completed in Russia. It consisted of one 116 km, 200 kV overhead power line from Moscow to Kasira. In 1954, the first commercial HVDC transmission system was commissioned in Sweden. The system used mercury arc valves and provided a 96 krn, 100 kV and 20 I'1W underwater link between Gotland and the Swedish mainland.

The development of thyristor valve converters led to increased application of HVDC transmission. In 1972 the first system using thyristor valves was commissioned in Canada, as the 320 MW Eel River Scheme, and connected the power systems of New Brunswick and Quebec provinces. The world's largest HVDC scheme is in Brazil where 6300 MW

:HAPTER4

HIGH VOLTAGE DIREC.T C.URRENT (HVDC)TRANSMISSION

of power is delivered from the Itaipu hydro-electric scheme over 800 km at a voltage of ±600 kV (the world's highest operating voltage) [1]. Figure 4.1 shows the development in HVDC installed capacity on an annual basis for schemes throughout the world.

Annual installed HVDe capacity

8000

7000 6000 I 5 I

~ 5000

c

.~ 4000

u -0

~

Bi 3000

" :::; ~_________...,J,JLJ Jt _ J~~~ _

Year

4.2.2 TECHNICAL AND ECONOMIC JUSTIFICATION

Figure 4.1: Increase in HYDe transmission capacity worldwide.

HVDC transmission systems have mainly been applied in the following cases:

Provision of underwater links longer than 30-50 km, where AC cable transmission would not be economical due to the requirement for reactive compensation e.g., the Gotland link in Sweden.

Provision of asynchronous links between AC systems operating at different frequencies e.g., the Itaipu scheme linking 50 Hz generators in Paraguay to the 60 Hz system in Brazil.

Transmission of large amounts of power over long distances e.g., the Cahora Bassa scheme, delivering 1800 I'1W over 1414 km from Songo, Mozambique to Apollo in South Africa.

Furthermore, the very rapid control of power flow, which is possible on HVDC systems, has a positive impact on system stability.

As a result of recent advances in converter technology, HVDC transmission schemes, (both high voltage, long distance, and lower voltage, shorter distance) are increasingly being seen as alternatives to HVAC systems. For example [2]:

CHAPTER 41

HIGHVOLTAGE DIRECT CURRENT (HVDC) TRANSMISSION

The development of Voltage Source Converters (VSC). This has resulted in more compact HVDC systems that are economical to implement at lower power levels and over shorter distances. The VSC technology also allows alternative energy sources to feed into the grid more economically (e.g., the Gotland and Tjaereborg wind projects). De-regulation of the electricity industry in which bi-directional power flows will be essential to facilitate electricity trading. Utilities will also be contractually required to provide better power control facilities.

Environmental impact is becoming a significant factor in the design and operation of transmission systems. HVDC transmission systems require less servitude (or rightof-way (ROW)) for a given level of power transfer.

HVDC transmission systems have not been widely used, for the following reasons:

Difficulties with telecommunications where local infrastructure is lacking.

At the point of coupling of an HVDC system with an AC network, a fault-Ievel-toload ratio of at least 3: 1 is required. In the case of weak AC networks, this criterion cannot always be met. This precludes the use of HVDC, when it would otherwise have been advantageous.

Technical complexity in the design and operation of the schemes. Technically difficult to tap-off power at intermediate locations. Expensive converter stations.

The cost of an HVDC transmission system depends on factors such as power to be transported, transmission medium (i.e., cable or overhead power line), environmental impact, safety and regulatory requirements. The total cost of an HVDC scheme can be divided into two main components: the converter station and the transmission system.

Erection. commissioning 8%

Freight. insurance 5%

AC filters 10%

Engineering 10%

Converter transformers 16%

••

Civil works. building 14%

Valves 20%

Figure 4.2: Distribution of converter station costs [3].

Figure 4.2 shows the cost distribution of a converter station. In a case study of four stations in 2000, costs varied from 75 to 220 $/kW, depending on power and voltage ratings [3].

Besides the capital cost of an overhead power line, additional costs vary depending on land costs (the servitude), and accessibility of terrain to be crossed. Comparison with HVAC overhead power line costs is on the basis of a double AC circuit and a bipolar

CHAPTER 4

HIGH VOLTAGE OIRE.CT CURRENT (HVOq TRANSMISS.ION

HYDC overhead power line. For transmission of 2000 MW, the breakeven distance is about 650 km (Figure 4.3). According to Woodford, the capital costs of an HYDC overhead power line can be 80%-100% of those for the same AC line and pole-to-ground voltages [2]. However, the HYDC line is capable of carrying twice the power over the same distance. The total cost of an HYDC transmission system is initially high due to the cost of the converter stations, which can be up to three times those of an AC substation (assuming the same overhead power line costs) [3].

The concept of 'critical line length' is well known. If this length is exceeded, the HYDC overhead power line is cheaper per megawatt transmitted compared to an HYAC overhead power line. For overhead power lines the critical line length is between 600 and 1000 km, and for underground cable between 40 and 80 km. In practice, the actual value of the critical distance will vary from scheme to scheme, depending on factors such as optimal voltage, cost of losses, the depreciation period and the desired rate of return on investment and cost of servitude [3]. For a given optimal voltage, the cost of losses reduces the breakeven distance because of the comparatively lower losses in HYDC systems.

As a preliminary guide for the choice between HYAC and HYDC transmission, one manufacturer applies the principle that if at least 1000 MW are to be transmitted over 1000 km, then HYDC transmission is more cost effective.

900
~ 800
Vl 700
::>
c 600
.2
S 500
(]) 400
u
.r: 300
Q_
200
100
0 AC price DC price

0000000000000 0000000000000 (V) '" Lf) -0 r-- co a-- 0 - ('l (V) '" Lf)

Distance (km)

Figure 4.3: Variation of HVDC transmission costs with distance [3].

4.2.3 SYSTEMS CONFIGURATION AND COMPONENTS

HYDC transmission systems can be subdivided into the following three categories [4]:

Point-to-point transmission: This is used to transport power from one point to another. Cables and overhead power lines are used here.

Back-to-back: This is used to connect two AC systems of different frequencies, or asynchronous systems.

Multi-terminal: This is used for power sharing and where tap-off points are required.

CHAPTER -4

-I I G H VO LTAGE DIRECT CU RRE NT (HVDC) TRANSMISSION

4.2.3.1

4.2.3.2

4.2.3.3

CH.APTER 4

Configurations

An HYDC link can be monopolar, bipolar or homo polar.

Monopolar Configurations

This is the basic configuration using one conductor. usually at negative polarity. and ground or water as the current return path (see Figure 4.4). The return can also be metallic. in the case of high values of ground resistivity. or where there are objections to the interference caused by ground or water currents.

AC system

I AC

I system

Metallic return

(Optional)

Figure 4.4: Monopolar configuration [4].

Bipolar Configuration

This configuration is shown in Figure 4.5. It comprises two conductors. positive and negative. The converters at either end of the line are equally rated and the junction point between the converters at either end is grounded. The currents in the two poles are equal so there is no ground current. The two poles can operate independently. If one pole is faulted. the other can be operated by ground return to deliver half of the rated load. Bipolar systems cause less harmonic interference than mono polar systems in adjacent telecommunication lines. A third conductor can be used in cases of high values of ground resistivity or unacceptably high levels of ground currents. The conductor requires little insulation and can serve as a ground wire for the overhead power lines.

AC system

+

AC system

Figure 4.5: Bipolar HVDe configurations [4].

4.2.3.4

HIGH VOLTAG,E DIRECT CURRENT (HVDC) TRANSMISSION

Homopolar Configuration

The homopolar configuration has two or more conductors operating at the same polarity. The negative polarity is usually preferred because it generates less radio interference and audible noise. The ground is used as the return path. If one conductor is faulted, the rest can carry the centre load (see Figure 4.6). This makes homopolar configurations advantageous over bipolar configurations, if the ground circuit is acceptable. Using ground as the return is usually not acceptable because the ground current can cause corrosion of underground pipelines and other buried metallic structures in the vicinity of the earth electrodes.

+

AC system

AC system I

I

1'- - - - - - -<41- - - - - - - -, I

4.2.4 COMPONENTS

Figure 4.6: Homopolar configuration [4].

4.2.4.1

An HVDC system consists of three main sections: AC rectification, DC transmission and inversion back to AC. Each of these sections has components for operation and control. The main components are discussed below:

Converters (Inverters)

The converters at either end of the overhead power line consist of equipment for conversion of AC to DC (in the converter side) and DC to AC (in the inverter side). Varying the firing angle of the thyristor valves controls the output voltage levels. Each single thyristor valve normally consists of a number of series connected thyristors and associated auxiliary circuits. The thyristor valves are usually arranged in a twelve-pulse group, with three quadruple valves shown in Figure 4.7.

CHAPTE - 4

-IIGHVOLTAGE DIRECT CURRENT (HVDC) TRANSMISSION

Quadrivalve (3 in a 12-pulse converter)

Valve Valve panel

(12 in a 12-pulse (2-10 panels

converter) per valve)

Saturable

DC Panel inductor

...-4----. ~-+----. """'::+-""'M

Thyristor level ( I 0-20 per panel: 10-100 per valve)

AC

Surge arrestor

Damping

res't6,tor

Power suppl'y Thyristor ! electronics ]'

r Electro

", '--_-.- __ _,L_:;'P._~i_caJ interface

Datnping '-"l-, --]f

capaclt,or '" f-'" .".1

.!:! " "t eo

:::::f .. II I.- C

~ :: Optical:: ~ '§ 9 :: fibres :: .~.~

E II 1I..r: 0

~ ::1= E

AC

DC Optical fibres (two per thyristor)

rV~I~~'b~~~

elect~ _ L

GOnverter contro~

Figure 4.7: Thyristor valve configuration [5].

There are three main types of converters, depending on the commutation process:

4.2.4.1.1 Naturally commutated converters (NCC).

These are most commonly used in HYDC systems. They consist of thyristors stacked in series to form valves, which operate at very high voltages. The thyristors conduct after being switched on, until reverse biased. The voltage level and power flow are regulated by changing the firing angle of the thyristors.

4.2.4.1.2 Capacitor com mutated converters (CCC)

These provide an improved performance on NCCs. The commutation capacitor, inserted between the converter transformer and the thyristor valve, is used to reduce commutation failure, common in NCCs.

4.2.4.1.3 Forced commutated converters (FCC)

The converters are also known as VSCs. They consist of GTOs or IGBTs, which can be turned on and off rapidly. The VSC cornmutates at high frequency using Pulse Width Modulation (PWM), which offers the option of controlling active and reactive power independently. VSC converters use fewer components and require about 20% of the area needed by conventional converter stations. The choice of application of a commutation system depends on the nature of the HVDC transmission system (Table 4.1).

CHAPTER 4

4.2.4.2

4.2.4.3

4.2.4.4

HIGH VOLTAGE DIRECT CURRENT (HVDC) TRANSMISSION

Table 4.1: Application of commutation systems [3].

Long distance Long distance Interconnections Windmill Feed of small
transmission transmission of asynchronous connection to
over land over sea networks network isolated loads
Natural
commutated X X
HYDe with
Q,1rl lines
I Natural
cornrnutated X X
HYDC with
sea cables
Capacitor
Commutated
Converters X
(Ccq in
back-to-beck
Capacitor
Cornmutated
Corwerters X X
I (CCC/With,
OH ines
Capacitor
Commutated
Converters X X
(Ccq with
sea cables
YSC
Converters in X X
back-co-back
VSC
Converters X X X X X
with land or
sea cables Thyristor Valves

The most common thyristor valve configuration is the 12-pulse group. with three quadruple valves (see Figure 4.7). This is mostly applied in bipolar schemes that have two AC and three DC connections. Each valve consists of series connected thyristors (10-20) and associated electronic circuitry. Communication to and from the thyristors is via optical fibres. The valves are air insulated and water cooled.

VSC Valves

The VSC is a two- or multi-level converter comprising phase reactors and AC filters. Each valve consists of a number of series connected IGBTs with associated electronic circuitry. The valves and control and cooling equipment are built in modules. which makes them easier to transport and install.

Converter Transformers

Converter transformers are used to step up the AC voltage that is then rectified. On the inverter side. the transformers step down the inverted DC voltage. They are usually single phase three-winding transformers and contribute to the commutation reactance.

CHAPTE

HIGH VOLTAGE DIRECT CURRENT (HVDC) TRANSMISSION

4.2.4.5

4.2.4.6

4.2.4.7

4.2.4.8

4.2.4.9

:HAPTER 4

AC Filters and Capacitor Banks

The filters are installed to limit the harmonics generated on the AC side. For a 12-pulse converter, the 11th, 13th, nrd and 25th order harmonics are generated. Capacitor banks are installed to compensate for some of the reactive power consumed by the converters. The filter banks also contribute to reactive power compensation.

In CCCs, reactive power is compensated by capacitors connected in series between the converter valves and transformers. This eliminates the need for switched compensation, thus reducing the size and capacity of the AC switchyard.

In VSCs, there is no need for converter reactive power compensation and the generated harmonics are directly related to the power frequency. This greatly reduces the number of filters required, compared to thyristor converters.

DC Filters

The harmonics generated by converters can create interference in telecommunication systems. DC filters are installed to reduce the harmonics and the level of telephone interference, known as the equivalent disturbing current. The filters are not required in back-to-back stations, or where cable transmission is used. Usually, DC filters are smaller and less expensive than AC filters. Active DC filters are used in some modern installations.

DC Reactors

These are usually included in each pole of a converter station. They assist DC filters in filtering harmonic currents and smoothing the DC output from the converter. Either air or iron cored and oil insulated reactors may be used.

Surge Arrestors

Surge arrestors are installed to provide protection against overvoltages, regardless of their source. They are connected across each valve in the converter bridge, across each converter bridge, and in the AC and DC switchyards. Metal oxide arrestors are most commonly used in modern HVDC schemes.

Earth Electrodes

An earth electrode or ground return is used in one form or another in most HVDC schemes. The earth electrode may be used continuously or under fault conditions, depending on the specific application. The aim of the earth electrode is to achieve the lowest possible earth resistance, to avoid steep electrical gradients in the soil and high ohmic losses, which would otherwise lead to problems of leakage current corrosion of nearby buried metallic objects.

HIGH VOLTAGE D.IRECT CURRENT (HVDe) TRANSMISSION

DC smoothmg

Converter reactor

ACbus~ ~--~--+-~~~~

CB

I

I

r:I .I.

Reactive power SOUI·ce

: DC line:

.I. I]

Reactive power sou rce

-~·------~-·------:----·--III

r--"----, EI ectrodes :

CB: Circuit breaker

Figure 4.8: Schematic of bipolar HVDC system showing main components [4].

4.2.5 OPERA nON AND CONTROL

Converter operation involves the regulation of thyristor firing angles to obtain power flow from AC to DC (rectification) and from DC to AC (inversion). The rectification or inversion is achieved through line or natural commutation, i.e., a thyristor, once turned on, will continue to conduct until it is reversed biased and the current reduces to zero.

The main purpose of the control process is to:

Prevent large fluctuations in DC due to variations in AC system voltage. Maintain the DC voltage near the rated value.

Maintain high power factor levels at the sending and receiving ends. Prevent commutation failure in inverters.

Since HVDC systems convey large amounts of power, precise control of current and voltage is essential. Therefore, the following quantities at each converter end have to be continuously and precisely measured: DC current, DC side voltage, delay angle (a) and extinction angle (y) [4]. There must be a reliable communication link between the converter stations for data exchange. Power line carrier (PLC) and microwave links are two of the most commonly used communications methods. Fibre optic links are increasingly being considered because they are less prone to interference. However, the availability of reliable power supplies for repeater stations is an important factor for consideration, especially if the overhead power line is through remote, inaccessible areas.

Again, since HVDC systems transmit large amounts of power and are critical for the interconnection of power systems, their reliability is essential. Data on the performance of different HVDC schemes is compiled and presented at Cigre Study Committee B4 meetings. The average figures for the unavailability of HVDC systems are summarised

CHAPTt::R <4

HIGH VOLTAGE DIRECT CURRENT (HVDC) TRANSMISSION

in Table 4.2. The forced and scheduled unavailability for the periods 2001 to 2002 are shown in Figures 4.9 and 4.10. According to the 2002 Cigre report, 89% of the forced outages were due to equipment failure on the AC side, and only 7% were attributable to major DC components failure [6]. In 2001 and 2002 about 84% of the forced outages were attributed to equipment on the AC side of the converters, compared to 11% for major DC equipment [7].

Table 4.2: Average reliability of HYDC schemes [6,7].

Period 1982-1992 1999 2000 2001 2002
Forced outages (%) 1.62 2.36 1.13 1.34 1.60
Scheduled outages (%) 5.39 4.01 5.37 5.93 4.10 Figures 4.9 and 4.10 give a comparative breakdown of the scheduled energy unavailabilty percentages for 32 HVDC schemes in 2001 and 2002 respectively (the scheme numbers refer to the same HVDC system in each case).

Scheduled energy unavailability for 200 I
70 I
60
~
>.. 50
.'=
.D
'iii 40
>
rd
C
:J 30
>..
eo
s;
OJ
c 20
L.U ~~~iJJ:J OJ] 0 o_~o II [jJJ 0 0 ~ oj] 0 0 =0
10 ~ C
o .D_u 0,
M '" r-, a- M '" ~ ~ N M '" r-, a- -
- - N N N N M
HVDe scheme number Figure 4.9: Reliability of HYDC systems in 2001 [7].

CHAPTER .4

HI.GH VOLTAGE DIRECT CURRENT (HVDC) TRANSMISSION

Scheduled energy unavailability for 2002

70
60
~ 50
.2
:0
'(;j 40
>
ro
c
::J 30
~
'-
Q)
c 20
u.J
10
0 HVDC scheme number

Figure 4.10: Reliability of HVDe Systems in 2002 [7].

4.3 HVDCTRANSMISSION LINES

4.3.1 ADVANTAGES AND DISADVANTAGES OF HVDC TRANSMISSION

The advantages and disadvantages of HVDC transmission are summarised in Table 4.3.

Table 4.3: Summary of the advantages and disadvantages of long-distance HVDe transmission.

Advantages Disadvantages
Greater power per conductor. Expensive converter stations.
Simple, less expensive line construction. Technically difficult to tap off power at
intermediate locations.
Possible monopolar operation. Constrained telecommunications
systems.
No charging current, no skin effect. Technical complexity in the design and
operation of the schemes.
Reduced corona and radio interference. High reliability in terms of redundancy
is expensive to achieve.
Allows for asynchronous operation.
Low short-circuit current.
Easy to control line power (thyristor
control). HIGH VOLTAGE DIRECT CURR.ENT (HVDC) TRANSMISSION

Table 4.3: Continued

Line carries no useless reactive power.

Does not contribute to the instability of the AC system under fault conditions.

Does not require reactive compensation. only real power is transmitted

4.3.2 DIFFERENCES BETWEEN DC AND AC TRANSMISSION

The choice between HVDC and HVAC transmission depends mainly on the performance requirements of the line. Some of the major differences between HVDC and HVAC lines in terms of overvoltages and insulation are summarised in Table 4.4.

Table 4.4: Main differences between HVDC and HVAC transmission.

Item

AC versus DC differences

Insulation

HVDC requires less insulation for the same power transmitted because the crest value for DC equals the rms value. whereas the crest value for AC is --J2times the rms value.

Overvoltages

Insulator pollution

Withstand of polluted insulators

HVDC overvoltages are more severe than AC overvoltages because of the differences in overvoltage generation mechanisms and waveshape.

More pollution on HVDC insulators than on AC. leading to higher leakage currents. Special design of insulators for HVDC lines needed. including choice of specific creepage.

Insulators withstand lower DC voltages than AC voltages for the same amount of pollution and insulator profile.

Polarity effects on noise

Weather effects on radio and audible noise

Radio noise tolerability

Radio and audible noise are produced predominantly by positive HVDC poles whereas in HVAC. all phases produce radio and audible noise.

Radio and audible noise from HVDC lines is higher in fair weather than in wet weather. in contrast to HVAC lines. In general. design limits for DC are lower than for AC.

Noise pulse repetition rate is lower for HVDC than for HVAC (low frequency noise is less likely to interfere with communication systems).

Telephone interference

Because of the inherently high levels of harmonic currents

and voltages on DC lines compared with AC lines. harmonic interference induced into adjacent telephone systems can be severe. Special filtering is required to limit the interference to tolerable levels. The interference from monopolar lines is higher than that from bipolar lines.

:HAPTER 4

HIGH VOLTAGE DIRECT CURRENT (HVDC)TRANSMISSION

Table 4.4: Continued

Corona power The ratio of foul to fair weather power loss is much lower for
loss HVDC than for HVAC. Maximum corona losses for HVDC are
normally lower than for HVAC.
Electric field at Because DC electric fields are not time-varying, they do not
ground level induce currents in the bodies of people or animals. Induced
currents caused by harmonic ripples are negligible.
Currents in the Human perception levels for DC currents are higher than for
body AC.
Air ions at Air ion concentrantions near HVDC lines can be significantly
ground level above ambient levels (could contribute to noise generation in
PLC systems and coupling into adjacent telephone lines).
Spark discharges Spark discharges can occur under HVAC lines, but seldom under
HVDC lines (space charge suppression of electric fields, which
inhibits development of leader discharges). Spark voltages are
not time varying.
Magnetic field The magnitude of the magnetic field from HVDC lines is similar
to that of the earth's magnetic field. There are no known
biological effects.
Current The current in sub-conductors of bundles is governed by
distribution on resistance in the case of DC, and predominantly by inductance
conductors in AC.
Bundle Bundles reduce the surge impedance of HVAC lines. The surge
conductors impedance plays no role in HVDC under steady-state conditions. 4.4 FUTURE DEVELOPMENTS IN HVDe

4.4.1 VOLTAGE SOURCE CONVERTERS

The key component of a converter bridge, the thyristor, continues to be developed, with its voltage and current ratings increasing. GTOs and IGBTs are used in the VSC bridge configurations. This bridge is being applied in many new developments [3,8,9]. Its special properties include the ability to control real and reactive power independently at the connection bus to the AC system. Reactive power can rapidly be controlled between capacitive or inductive modes.

Unlike the conventional line cornmutated converter, the voltage source converter, as an inverter, does not require an active AC voltage source to commutate. The VS inverter can generate an AC three-phase voltage and supply electricity to a load as the only source of power. The inverter is suitable for supplying power to weak AC systems or passive loads such as rural loads.

The operation of voltage source converters is achieved by PWM, with which it is instantly possible to create any phase angle and amplitude combination. With PWM, it is possible

CHAPTER 4



HIGHVOLTAG'E DIRECT CURRENT (HVDC) TRANSMISSION

to control active and reactive power independently on each side. With a suitable control system, PWM voltage source converters can control power to enhance and preserve AC system synchronism, and act as rapid phase angle power flow regulators with a 3600 range of control. They have fewer components than the thyristor converters and so require less space (about 20% of the traditional, thyristor-based HVDC converter stations). They can also be constructed in modular units that are easy to install. PWM converters eliminate harmonics and do not require harmonic filtering (see Figure 4.11) [3].

Figure 4.11: Voltage source converter with pulse width modulation (PWM) [10].

There is considerable flexibility in the configuration of VSC bridges. Many two-level converter bridges can be assembled with appropriate harmonic cancellation properties to generate acceptable AC system voltage waveshapes. Another option is to use multilevel converter bridges to provide harmonic cancellation. In addition, both two-level and multilevel converter bridges use pulse width modulation (PWM) to eliminate low order harmonics. With PWM, high-pass filters may still be required since PWM adds to higher order harmonics.

As the VSC bridge technology develops for higher DC voltage applications, it will be possible to eliminate converter transformers. This is already possible with the low voltage applications in use today. It is expected that present developments in power electronics will continue to provide new configurations and applications for HVDC converters.

Presently, the power transmission ranges for VSC schemes are from a few MW up to 330 MW and for DC voltages up to ±150 kV. The converters are known as HVDC Light® under Asea Brown Boveri (ABB) and HVDC Plus® under Siemens. The development of

CHAPTE:R 4

HIGH VOLTAGE DIRECT CURRENT (HVDe) TRANSMISSION

PWM based VSCs has made it possible for HVDC systems to cater for relatively small power applications (up to 200 MW) over short distances (hundreds of krn), thus opening up opportunities for implementation in distribution networks. Possible applications include coupling of AC distribution networks, interconnection of renewable energy sources such as off-shore wind farms with grid systems, and supplying off-shore installations such as oil rigs and connections across environmentally sensitive areas [3,8,10].

The first HVDC VSC converter for distribution systems was developed in Sweden in 1997 by ABB. It was commercially implemented in 1999 on the Gotland Project (50 MW, 80 kV, 70 km link between Gotland and the mainland). During the past seven years, the total installed capacity of such schemes has increased to over 800 MW (see Figure 4.12). Applications elsewhere in the world include:

The Tjaereborg Project in Denmark, for an 8 MVA, 9 kV, 4 km link between a wind farm and the main grid. The system was implemented for optimum and individual speed control of the wind turbines to maximise turbine efficiency.

The Murraylink (220 MW, 150 kV, 180 km DC cable link) and Directlink (180 MW, 80 kV, 60 km DC cable link) projects in Australia, interconnecting asynchronous networks to facilitate energy trading.

The Cross Sound Link (USA). It is the world's largest HVDC light scheme (330 MW, 150 kV, 40 krn), designed for improved power reliability and promotion of energy trading in the New York and New England electricity markets.

The Eagle Pass Scheme (36 I'1W, 16 kV) between U.S. and Mexico. It is a back-to-back configuration for energy trading and power flow control to stabilise voltage levels [10].

1997

1999

2000

2002

HVDC distribution schemes

900
800
~ 700
L 600
~
z-
'u 500
ro
a.
ro 400
u
"'0
~ 300
~
.....
Vl
c 200
H
100
0
i-
I-
1-

- .. ~. I-
I . I-
I-
t ~ .-

I-
I ----,
J I Year

Figure 4.12: Growth of voltage source converter schemes (Source: ABB, 2004).

CHAPTER 'II

HIGH VO LTA.G E DIRECT CURRENT (HVD C) TRAN SMISSION

4.4.2 INSULATED GROUND WIRES FOR MV DISTRIBUTION ON HDVC LII\JES

Various alternatives and schemes for cost effective MV distribution have been identified and researched in the past. Coupled with the need to transfer substantial amounts of power through existing corridors and infrastructure, there is growing pressure to reduce the impact on the environment.

The feasibility of operating insulated ground wires on HVAC lines for MV distribution was validated by results from experimental schemes such as the Volta River Scheme in Ghana [11]. The implementation of a similar scheme on HVDC lines would provide the means of transferring more power through existing servitudes and limiting environmental impact. Some of the expected benefits include the possibility of supplying remote villages along the route of HVDC overhead power lines with electricity [12]. This will be a costeffective way of supplying small and rural loads with power. For instance, line vandalism has been a problem in rural and developing areas simply because the communities do not draw the benefits from the overhead power lines running through their areas. However, to construct a new MV line for such small loads is difficult to justify economically, even if it is in the same servitude as an existing line. This is because the aggregate load will be too small in comparison with the distance.

Calculations and simulations indicate that there is a possibility of using the insulated ground wires of HVDC overhead power lines for MV distribution. For schemes such as Apollo-Songo and Inga-Kolwezi, where monopolar lines run over separate servitudes, one line can conceivably be equipped with ground wire to transport MV power and the other with an OPGW telecommunications cable [13].

This type of design is subject to meeting certain operational requirements, such as shielding the main conductor from direct lightning strikes. The effects of harmonic coupling must be evaluated to establish optimum harmonic filtering. Other requirements should include the implementation of a redundant system to ensure continuity of supply.

4.4.3 PLANNING TOOL FOR MVDC

Medium Voltage Direct Current (MVDC) transmission can be used to bridge greater distances with low voltage and low power [14]. Network planners need tools to compare AC and DC alternatives to find the best technical and economic solution for a specific situation. To compare the voltage regulation capabilities, line losses and transmission costs for both AC and HVDC transmission systems, a computer software program called Technical and Economical Software Analysis Tool (TESAT) has been developed using Matlab® . This tool takes the following input parameters via a graphic user interface: conductor parameters (such as the series impedance [D/km], shunt conductance [!lS/km], cost and capacity rating of the different conductors), system parameters, load parameters and financial parameters. The main outputs are three-dimensional graphs that portray the resulting voltage regulation, losses and transmission costs of the selected conductors (see Figure 4.13). This is done for both AC and DC transmission technologies [14].

CHAPTER -4

HIGH VOLTAGE DIRECT CURR.ENT (HVDC) TRANSMISSION

It has been found that voltage regulation capability favours DC transmission. The inverter at the load-end can cater for a voltage drop of more than 30% with no commutation problems. Line losses are determined from the difference between sending-end power and receiving-end power.

When comparing transmission costs for an AC or DC interconnection, both the investment cost of the line and terminals, and the cost of losses should be evaluated when selecting the most suitable scheme.

DC converter losses are also taken into consideration when determining break-even distances. The economic value is analysed through I\IPY calculations [14].

Conductor Parameters

Financial Parameters

engine vo~;,,;on

Reg~~t~n ~ Tra~sts

Graphs and quantified comparisons among conductors

Figure 4.13: Internal layout of TESAT [14].

4.5 HVDC SYSTEMS IN AFRICA

4.5.1 CAHORA BASSA - APOLLO ±533 KY HYDC SCHEME

The Cahora Bassa - Apollo scheme was built in the early 1970s to supply the growing power needs of Southern Africa. The scheme links Mozambique and South Africa via two 1414 km mono polar, ±533 kY HYDC overhead power lines, from Songo in Mozambique to Apollo in South Africa (see Table 4.5).

When the scheme was built, Mozambique and its neighbouring countries did not require the amount of energy generated by the proposed hydroelectric scheme. Although Eskom did not have a pressing need for the energy generated by Cahora Bassa at that stage, the importation of power from Mozambique provided the following benefits:

CHAPTER 41

You might also like