You are on page 1of 296

IET PowEr and EnErgy sErIES, VOLUME 29

High Voltage Direct


Current Transmission

2nd Edition
Jos Arrillaga

The Institution of Engineering and Technology

Published by The Institution of Engineering and Technology, London, United Kingdom


First edition 1998 The Institution of Electrical Engineers
Reprint with new cover 2008 The Institution of Engineering and Technology
First published 1998
Reprinted with new cover 2008
This publication is copyright under the Berne Convention and the Universal Copyright
Convention. All rights reserved. Apart from any fair dealing for the purposes of research
or private study, or criticism or review, as permitted under the Copyright, Designs and
Patents Act, 1988, this publication may be reproduced, stored or transmitted, in any
form or by any means, only with the prior permission in writing of the publishers, or in
the case of reprographic reproduction in accordance with the terms of licences issued
by the Copyright Licensing Agency. Inquiries concerning reproduction outside those
terms should be sent to the publishers at the undermentioned address:
The Institution of Engineering and Technology
Michael Faraday House
Six Hills Way, Stevenage
Herts, SG1 2AY, United Kingdom
www.theiet.org
While the author and the publishers believe that the information and guidance given
in this work are correct, all parties must rely upon their own skill and judgement when
making use of them. Neither the author nor the publishers assume any liability to
anyone for any loss or damage caused by any error or omission in the work, whether
such error or omission is the result of negligence or any other cause. Any and all such
liability is disclaimed.
The moral rights of the author to be identified as author of this work have been
asserted by him in accordance with the Copyright, Designs and Patents Act 1988.

British Library Cataloguing in Publication Data


A CIP catalogue record for this book is available from the British Library
ISBN (10 digit) 0 85296 941 4
ISBN (13 digit) 978-0-85296-941-0

Printed in the UK by Short Run Press Ltd, Exeter


Reprinted in the UK by Lightning Source UK Ltd, Milton Keynes

Preface

The high voltage mercury-arc valve and its application to the development
of an HVDC transmission technology have been described in earlier books
by Adamson and Hingorani, Kimbark and Uhlmann. In common with
these texts the first edition of this book, published in 1983, described the
basic principles of static power conversion and their application to power
transmission by high-voltage direct current.
By then, however, in parallel with the development of microelectronic
technology there had been an equally impressive, although less publicised,
macroelectronic development in the power field sharing the same basic
ingredients, i.e. switching and silicon. The main exponent of macroelectronic technology must surely be the solid-state HVDC valve. By the time the
first edition of this book was being prepared, thyristors had already
displaced mercury-arc valves in new HVDC schemes, and the book
reflected the change. Although the basic principles of operation remain
the same, the past 15 years have seen a worldwide acceptance of HVDC
and particularly the installation of a large number of back-to-back interconnections. There have also been substantial improvements in the ratings
and reliability of thyristor valves and the appearance of more controllable
solid-state devices; the latter have encouraged a new technology called
FACTS (flexible AC transmission systems) which is proving to be very
competitive with HVDC for some specific applications. However, thyristor
technology has not remained at a standstill and a variety of new concepts
and techniques have been appearing with the aim of reducing the cost of
HVDC and extending its area of application.
This expanded edition of the book includes the main technical advances
of the past 15 years and describes the new concepts which, no doubt, will
help to make HVDC even more competitive in the new millennium.
Again, I would like to acknowledge the valuable help received early
on from all the experts mentioned in the first edition and extend my
gratitude to my present colleagues C.P. Arnold, P.S. Bodger, S. Chen,

xii Preface

W. Enright, B.C. Smith, N.R. Watson and A.R. Wood for their support
and dedication to the HVDC cause. I acknowledge the continued encouragement and financial assistance received from TransPower NZ Ltd for
our research into HVDC transmission.
It would be difficult to properly acknowledge all the sources of information used in the preparation of this book; I must, however, single out the
vast amount of work carried out by CIGRE study committee 14 on HVDC
transmission from which I have derived inspiration, the practical information and photographs obtained from industry, especially GEC-Alsthom
and ABB, and the close collaboration that I have had over the years with
the Manitoba HVDC Research Centre.

Contents
Preface
1 Introduction
1.1 Historical background
1.2 The mercury-arc valve
1.3 The silicon controlled rectifier (thyristor)
1.4 Future switching trends
1.5 The HVDC claims
1.6 The advent of a FACTS technology
1.7 References
2

Static power conversion


2.1 Introduction
2.2 Basic conversion principle
2.3 Selection of converter configuration
2.4 The ideal commutation process
2.4.1 Effect of gate control
2.4.2 Valve current and voltage waveforms
2.5 The real commutation process
2.5.1 Commutating voltage
2.5.2 Commutation reactance
2.5.3 Analysis of the commutation circuit
2.6 Rectifier operation
2.6.1 Mean direct voltage
2.6.2 AC current
2.7 Inverter operation
2.8 Power factor and reactive power
2.9 Maximum available power
2.10 Characteristic converter harmonics
2.11 Noncharacteristic harmonics
2.11.1 Harmonic crossmodulation

xi
1
1
3
4
7
8
8
9
10
10
10
13
13
14
17
18
18
19
23
24
26
27
27
28
32
33
39
42

vi Contents

2.12 Harmonic transfer generalisation


2.13 Quantified effects of system asymmetries
2.14 References

50
52
55

Harmonic elimination
3.1 Introduction
3.2 Pulse number increase
3.3 Design of AC filters
3.3.1 Design criteria
3.3.2 Design factors
3.3.3 Network impedance
3.3.4 Circuit modelling
3.3.5 Tuned filters
3.3.6 Self-tuned filters
3.3.7 High-pass filters
3.3.8 Example of recent filter arrangement
3.3.9 Type C damped filters
3.3.10 Simplified filtering for 12-pulse converters
3.4 DC-side filters
3.5 Active; filters
3.5.1 AC-side active cancellation
3.5.2 DC-side active cancellation
3.6 References

56
56
56
57
57
58
62
70
70
72
73
74
74
76
77
80
80
81
82

HVDC system development


4.1 Basic DC system configurations
4.2 Mercury-arc schemes
4.3 Evolution of the modern solid-state HVDC scheme
4.3.1 Frequency conversion
4.3.2 Asynchronous back-to-back interconnections
4.4 Operation reliability
4.5 References

84
84
86
88
93
93
97
97

Control of HVDC converters and systems


A - CONVERTER CONTROL
5.1 Basic philosophy
5.2 Individual phase control
5.3 Equidistant firing control
5.3.1 Constant-current loop
5.3.2 Inverter extinction-angle control
5.3.3 Transition from extinction-angle to current control
5.3.4 Other equidistant firing-control schemes
5.3.5 Application to 12-pulse converter groups
5.3.6 Comparative merits

100
100
100
101
103
104
105
106
106
108
108

Contents vii

B - DC SYSTEM CONTROL
5.4 Basic philosophy
5.5 Characteristics and direction of DC-power flow
5.5.1 Tap-changer control
5.5.2 Reversal of power flow
5.5.3 Modifications to the basic characteristics
5.5.4 Operational nonminimum margin angle
5.5.5 Power-flow control
5.5.6 Frequency control
5.5.7 Power/frequency control
5.6 Different control levels
5.6.1 Overall control co-ordination
5.6.2 Hierarchical power control at the New Zealand link
5.7 Telecommunication requirements
5.8 References

111
111
112
115
116
117
118
119
120
121
121
123
124
126
128

Interaction between AC and DC systems


6.1 Introduction
6.2 System strength definition
6.2.1 AC-system Thevenin equivalent
6.3 Voltage interaction
6.3.1 Dynamic voltage regulation
6.4 Dynamic stabilisation of AC systems
6.4.1 Large-signal modulation
6.4.2 Controlled damping of DC-interconnected systems
6.4.3 Damping of su bsynchronous resonances
6.4.4 Active and reactive-power co-ordination
6.4.5 Transient stabilisation of AC systems
6.5 AC-DC frequency interactions
6.6 Harmonic instabilities
6.6.1 Instability caused by individual firing control
6.6.2 Composite resonances
6.6.3 Transformer-core-related harmonic instability
6.7 AC-DC interaction following disturbances
6.7.1 AC-side fault recovery
6.7.2 DC-side fault recovery
6.8 References

129
129
130
130
132
133
136
138
138
140
141
142
143
144
145
148
150
155
155
156
157

Main design considerations


7.1 Introduction
7.2 Mercury-arc circuit components
7.2.1 Valve group
7.2.2 Converter station
7.2.3 Mercury-arc converter layout
7.3 Thyristor valves

159
159
160
160
160
162
162

viii

Contents

1A
7.5
7.6
7.7
7.8
7.9

7.10
7.11
7.12

7.13

7.3.1 Electrical considerations


7.3.2 Mechanical considerations
7.3.3 Valve-cooling system
7.3.4 Valve-control circuitry
7.3.5 Valve tests
7.3.6 Valve-hall arrangement
Station layout
Relative costs of converter components
Converter transformers
Smoothing reactors
Overhead lines
Cable transmission
Earth electrodes
Design of back-to-back thyristor converter systems
HVDC system upgrade
7.12.1 The converter stations
7.12.2 DC transmission line
7.12.3 Submarine power cables
References

162
165
167
168
171
174

174

176
178
181
182
184
190
192
194
195
197
198
198

Fault development and protection


8.1 Introduction
8.2 Converter disturbances
8.2.1 Misfire and firethrough
8.2.2 Commutation failure
8.2.3 Backfire
8.2.4 Internal short circuit
8.2.5 Bypass action
8.2.6 Bypass action in thyristor bridges
8.3 Simulation of practical disturbances
8.4 AC-system faults
8.4.1 Three-phase faults
8.4.2 Unsymmetrical faults
8.5 DC-line fault development
8.5.1 Fault detection
8.5.2 Fault clearing and recovery
8.5.3 Overall dynamic response
8.6 Overcurrent protection
8.6.1 Valve-group protection
8.6.2 DC-line protection
8.6.3 Filter protection
8.7 References

226

Transient overvoltages and insulation co-ordination


9.1 Introduction

228
228

200
200
200
201
201

207
207
208
209
210
211
214
214
216
218
219
219
221
224
225
226

Contents ix
9.2
9.3
9.4
9.5
9.6
9.7

9.8

9.9
9.10

Overvoltages excited by disturbances on the* DC side


Harmonic overvoltages excited by AC disturbances
Overvoltages owing to converter disturbances
Fast transients generated on the DC system
9.5.1 Lightning surges
9.5.2 Switching-type surges
Surges generated on the AC system
Fast transient phenomena associated with the converter
plant
9.7.1 Mercury-arc converters
9.7.2 Thyristor converters
Insulation co-ordination
9.8.1 System design
9.8.2 Surge arresters
9.8.3 Application of surge arresters
Considerations on cable overvoltage protection
References

229
231
232
233
233
235
235
238
238
240
243
243
244
245
250
251

10 DC versus AC transmission
10.1 General considerations
10.2 Power-carrying capability of AC and DC lines
10.3 A comparison of AC and DC transmission characteristics
10.4 Other considerations
10.5 Infeeds at lower voltage levels
10.6 Examples of the application of the break-even distance
10.7 Environmental effects
10.7.1 Electric field
10.7.2 Radiated interference
10.7.3 Acoustic noise
10.7.4 Visual impact and space requirements
10.8 Existing AC transmission facilities converted for use
with DC
10.9 Very long-distance transmission
10.10 References

253
253
255
258
259
263
264
266
266
267
269
270

11 New concepts in HVDC converters and systems


11.1 Introduction
11.2 Advanced devices
11.2.1 Thyristor development
11.2.2 Gate turn-off semiconductors (GTO)
11.3 New concepts for thyristor converters
11.3.1 Capacitor-commutated converter
11.3.2 Continuously-tuned AC filters
11.3.3 Outdoor valves
11.4 Compact converter stations

278
278
278
278
279
279
280
282
283
285

270
273
276

x Contents
11.5 GTO-based voltage-source converters
11.5.1 A GTO back-to-back HVDC link
11.5.2 HVDC light
11.6 DC cable developments
11.7 Direct connection of generators to HVDC converters
11.8 Small HVDC tappings
11.9 References
Index

286
287
289
289
290
292
294
296

Chapter 1

Introduction

1.1 Historical background


As early as 1881 Marcel Deprez, inspired by experiments with arc lights
across a DC generator, published the first theoretical examination of
HVDC power transmission. He soon put theory into practice and by 1882
he transmitted 1.5 kW at 2 kV over a distance of 35 miles.
The following decade witnessed the rising of alternating currents on
account of the availability of transformers and the development of induction motors. This prompted the following warning by Thomas Edison in
1887: "Take warning! Alternating currents are dangerous, they are fit
only for the electric chair ... ."
From 1889, R. Thury continued the work of Deprez by using DC
generators in series to attain high transmission voltages. Among his many
European installations, the best example of a DC transmission technology
was that from Moutiers to Lyon with a final capacity of 20 MW at 125 kV
over a distance of 230 km. This scheme operated at constant current and
was used as a reinforcement of an existing AC system. It was probably the
first recognition of AC-DC co-existence, as Thury himself put it: "The
two systems shake hands fraternally in order to give each other help and
assistance . . . "
With the advent of the steam turbine as a prime mover for the generation of power, the limitations of the Thury system were accentuated.
There had been no special problems with the low speed water turbines
driving the generators, but now the use of the steam turbine for DC
generation depended on the availability of high-speed reduction gearing.
However, by this time, some other interesting developments had taken
place.1
The constraints affecting the economic design of power generation and
transmission plant are very different. Thus, the use of a transmission

High voltage direct current transmission

system inflexibly tied to the requirements imposed by the generators will,


in general, not produce an economical power system.
AC transmission over long distances, especially via underground cable,
requires frequent shunt compensation and causes stability problems. AC
interconnections increase the fault level of the overall system. DC transmission is free from these problems and has lower losses and design costs.
These advantages were realised early and the idea of generating AC
power, converting it into DC for transmission and converting it back into
AC, was taken seriously. However, the use of static AC-DC and DC-AC
power conversion is expensive and in general the comparison of alternatives is not straightforward.
Among the many steps that can be identified with the development of
modern HVDC transmission technology, the following are worth mentioning:

a first attempt to combine the advantages of HVAC turbogeneration


and HVDC transmission was made in the 1920s by Calverley and Highfield with the 'transverter'. The idea was based on a number of transformers commutated by brushgear rotating synchronously at high
speed;
Hewitt's mercury-vapour rectifier, which appeared in 1901, and the
introduction of grid control in 1928, provided the basis for controlled
rectification and inversion;
prior to 1940, experiments were carried out in America with thyratrons
and in Europe with mercury-pool devices;
countries with long transmission distances like America, the Soviet
Union and Sweden, showed great interest in HVDC developments; in
the Soviet Union an experimental single-anode valve was constructed
during the Second World War and intensive research was carried out
in Sweden from 1940 by the Allmanna Svenska Elektriska Aktiebologet
(ASEA);
in the case of Germany, the Secretariat for Aviation encouraged the
development of HVDC technology during the War believing that
underground transmission was less vulnerable to air raids; an experimental transmission system of 15 MW at 100 kV was built between the
Charlotenburg and Moabit districts of Berlin, which was intended to be
a prototype for a 60 MW, 400 kV system of about 110 km, part of
which was built by the end of the War;
after 1945, the interrupted work on valves in the USSR was resumed as
part of a wider programme of HVDC developments; in 19502 an
underground system was brought into operation between Moscow and
Kashira, transmitting 30 MW at 200 kV over a distance of 112 km;
however, it is Dr Uno Lamm of ASEA(Sweden)3'4 who should be credited with having pioneered the development of modern HVDC transmission technology, as explained in the next section.

Introduction

1.2 The mercury-arc valve


Although the economic advantage of DC-power transmission was understood from the early days of electrical technology, its practical application
had to wait for the development of a suitably-rated electronic valve.
Among the various switching principles used in the early days of the
power-electronics industry, mercury-arc rectification was found to be the
most suitable for handling large currents. Multiphase mercury-pool
cathode valves provided with a control electrode (or grid) have been extensively used for over 50 years in industrial and railway applications.
The most successful development of the mercury valve for high-voltage
applications was carried out in Sweden, where by 1939 Dr Uno Lamm
from ASEA had invented a system of grading electrodes with a singlephase valve construction, which provided the basis for larger peak-inverse
withstand voltages.
The basic components and main auxiliaries of the ASEA high voltage
valve assembly are illustrated in Figure 1.1. In common with earlier technology, this valve includes a mercury-pool cathode with a cathode spot,
maintained by means of an auxiliary arc, which causes a continuous emission of electrons. It is the anode design that differs greatly from the
Anode assembly
Anode porcelain
Voltage divider

Equipment for
temperature
control of anodes
Control pulse input
Grid bias device
Excitation and
ignition set
Tank

Mercury diffusion
pump

Excitation anode
Ignition

Pre-vacuum tank

Cathode (mercury
pool)
Chassis
Water
Outlet
Inlet

Figure 1.1

High-voltage mercury-arc valve assembly

High voltage direct current transmission

earlier medium-voltage valve, its main aim being the elimination of the
reverse emission of electrons which causes reversal of conduction or arc
back.
With the graded electrodes it is possible to achieve a more uniform
distribution of the reverse voltage in the vicinity of the anode. This
reduces the energy of the charge carriers striking the anode material, and
with it the likelihood of arc backs. The grading electrodes are connected
to an external capacitive-resistive voltage divider, which together with the
interelectrode capacitances limits the voltage difference between them to
about 5 kV. The anode porcelain forms a vacuum-tight envelope that
functions as a supporting insulator for the different electrodes in the
anode assembly. Depending on the rated current, up to six parallel anodes
are placed on top of the stainless steel tank.
The quality of the porcelain used for the external cylinder is essential to
the viability of the HVDC valve. Under the influence of the direct-voltage
component across the valve, some ion migration occurs which causes ion
depletion at one end and ion increase at the other. This effect produces
conductivity variation and thus causes uneven voltage distribution. In later
designs the use of very high-resistance porcelain has reduced this problem
dramatically.
Another important problem was the deposition of material throughout
the valve which results from charge carriers striking the walls during
firings and blockings. This effect appears to limit the maximum direct
voltage achieved with mercury-arc bridges to about 150 kV and necessitates considerable maintenance.

1.3 The silicon controlled rectifier (thyristor)5


The appearance of the thyristor, or silicon controlled rectifier (SCR), in
the late 1950s had a dramatic effect on static-converter technology. A thyristor is basically a combination of two transistors, as illustrated in Figure

Figure 1.2

IK

Basic structure of a thyristor with anode current iA, gate current


and cathode current iK

Introduction

1.2. The collector of a pnp transistor structure forms the base of the npn
transistor structure, or vice versa. The thyristor function is based on the
regenerative action of the two coupled transistors and is modelled as a
four-layer, three-terminal device.
The complete voltage-current characteristic of a thyristor unit is illustrated in Figure 1.3. A small current injection through the gate terminal
makes a forward-biased thyristor switch from a very high to a very low
impedance stage, thus approximating the characteristic of the ideal switch,
with practically unlimited amplification factor.
Two-terminal breakover (in the absence of gate injection) can also take
place, either by sufficient forward (anode-cathode) bias (V) or excessive
rate of change of voltage dV/dt. Once the thyristor is turned on, it can
only be turned off, or blocked, by reducing the main circuit current below
a very low critical value, called the holding current (//>).
Under reverse bias, there is a critical breakdown level (VBD)> below
which the thyristor behaves like a pn junction diode, i.e. with only a low
leakage current.
The thyristor can be destroyed by excessive reverse voltage (VBD),
extended overcurrents and excessive rate of change of current (di/dt).
Therefore, when connected in series, the individual devices have to be
properly protected against overvoltages, overcurrent, di/dt and dv/dt.
For the use of thyristors in HVDC transmission to be economical, it was
necessary to improve their ratings. Although a typical thyristor in the late
1960s had a peak blocking voltage of approximately 1600 V and could
carry about 1000 A, modern devices permit in excess of 4000 A DC in a
six-pulse bridge and a 10 kV blocking voltage is not far away. Higher
voltages and currents are feasible but normally at the expense of severe
derating on other important parameters. The main limitation of the
forward region
+ on anode

reverse region
- on anode

Figure 1.3

Thyristor IV characteristic

Vs, /s: switching-point


Vh /h: holding-point

6 High voltage direct current transmission

Figure 1.4 Thyri-stmin a presspack housing


higher voltage thyristor is a greater requirement for forward recovery
time, which results in larger nominal extinction angles ( y ) during inverter
operation.
Thus, from a few kilowatts in its origin, the thyristor has quickly grown
into the megawatt range and expanded from power-utilisation to powertransmission fields at an unprecedented rate. Figure 1.4 shows a modern
thyristor encapsulated in a presspack housing and Figure 1.5 illustrates a
typical interdigitated gate geometry in a high-power thyristor; the black
areas indicate silicon (SiO2) and the white areas the A1 metallisation.
Both electrically-triggered ( E n ) and light-triggered (LTT) thyristors
are now available for use in HVDC transmission. Although these share a
common structure, the LTT is designed with a light-sensitive gate which
turns the device on vk a light source.

Figure I . 5 Typical interdigitated gate geometry in a high-powerthyristor

Introduction

Standard wafer diameters of 100 and 150 mm are currently used by the
ETT and LTT devices, giving repetitive peak off-state voltages of 7500
and 8800 V and average on-state currents of 2100 and 3500 A respectively.
Although the ratings of the individual devices are still increasing, there
is no need for dramatic growth in this respect because of the progress
made in thyristor-valve architectures; the individual devices can now be
connected in series and in multibridge configurations to process practically
unlimited power.

1.4 Future switching trends


Silicon is expected to remain the material for high-power semiconductor
devices for the next decade at least. In that period the thyristor is likely to
face strong competition for some HVDC applications from two alternative
switches or variations of them, i.e. the gate turn-off thyristor (GTO) and
the insulated gate bipolar transistor (IGBT).
The GTO has the same structure as the thyristor but permits turn off by
the application of a high negative current to the gate. Symmetrical and
reverse-conducting devices are available, although the latter only for lowpower ratings.
The symmetric reverse-blocking type suffers from slow switching speeds
and high switching losses. The anode-shorted type has limited reverseblocking capability and requires the protection of an antiparallel diode
connected across it, but has tolerable switching loss and adequately fast
switching. It is available in double-side cooled capsules with silicon
diameters up to about 100 mm, which have a peak forward blocking
voltage (FDRM) up to about 6 kV, a controllable (i.e. extinguishable)
current of up to 4 kA and /x(av) of about 1.5 kA. Higher ratings based on
150 mm silicon are expected to become available soon.
To achieve satisfactory turn off, a low-inductance diode-coupled
snubber capacitor must be employed, both to limit spike-voltage energy
dissipation and to prevent dv/dt retriggering following turn off. For use in
HVDC converters, the turn off gain of the GTO needs to be greatly
increased and the switching losses and snubber requirements must be
decreased to provide a similar flexibility with respect to series connection
of devices as is presently available with thyristors.
Another fast-developing technology is the metal-oxide semiconductor
(MOS) which is a development of the MOS field-effect transistor
(MOSFET), with the same structure. The main type for possible HVDC
application is the insulated-gate bipolar transistor (IGBT). As a MOS-gated
device, the IGBT requires only a voltage source of typically +10 to 15 kV
to give full turn-on and turn-off control, although the drive circuit still
needs enough power to charge and discharge the gate capacitance rapidly.
The main topology for high-voltage applications consists of an IGBT
module with an associated free-wheeling antiparallel diode. The ratings of

High voltage direct current transmission

these IGTBs have today reached 1600 V/1200 A, and 3300 V/1200 A
have already been announced; they also have inherent short-duration
high-current capability. At present most IGBTs are manufactured in
modular form; a module includes a number of individual chips connected
in parallel. However, presspack designs have already been developed
which permit the stacking of devices in series for high-voltage applications.
Because it has high switching speed and low switching loss, and can also be
switched snubberless in a suitably low-inductance circuit, the IGBT can be
operated at high pulse-width modulation (PWM) frequency, typically
several kHz for large units to minimise harmonics and equipment size.

1.5 The HVDC claims


After four decades of reliable HVDC operation throughout the world, it is
justifiable to reiterate with confidence the claims more cautiously made in
earlier references to the subject.6"8
The main advantages generally claimed in favour of the DC alternative
are:

DC transmission results in lower losses and costs than equivalent AC


lines, but the terminal costs and losses are higher;
AC transmission via cable is impractical over long distances; such a
restriction does not exist with DC;
DC constitutes an asynchronous interconnection and does not raise the
fault level appreciably;
the power flow in a DC scheme can easily be controlled at high speed,
and thus with appropriate controls, a DC link can be used to improve
AC-system stability;
DC stations, with or without transmission distance, can be justified for
the interconnection of AC systems of different frequencies or different
control philosophies.

This book attempts to justify the above claims with reference to the
present state of the art and particularly to thyristor technology.

1.6 The advent of a FACTS technology9


The attractive market created by the invention of the graded-electrode
high-voltage mercury valve forced an immediate response to develop a
solid-state alternative in the late 1960s. The following chapters will show
that the progress made in the solid-state valve in the past 25 years has
exceeded expectations.
Once the parent technology (HVDC) reached maturity, power-electronics related investment started to look for new applications. FACTS (flex-

Introduction

ible AC transmission systems) was then born,8 not necessarily as a direct


competitor but in order to provide solutions to specific problems at a
lower cost.
The present boundaries between HVDC and FACTS relate to the types
of solid-state device (which in the HVDC case are restricted to the silicon
controlled rectifier) and to the power rating of the schemes. However, as
the rating and acceptability of alternative solid-state devices improve, the
boundaries will gradually become blurred. HVDC will be tempted to use
the new devices and FACTS will try to influence power controllability
more directly, e.g. by developing the four-quadrant asynchronous interconnector. But that is precisely what a back-to-back HVDC link will do!
A merging of technologies is thus inevitable and the present terminology will have to give way to the wider concept of high voltage power electronics or HVPE.

1.7 References
1 'The history of high voltage direct current power transmission', Direct Current
Part I, December 1961, p.260; Part II, March 1962, p.60; Part III, September
1962, p.228; Part IV, January 1963, p.2; Part V, April 1963, p.89
2 PIMENOV, V.P.: 'The work of the Direct Current Institute (Leningrad)', Direct
Current, 1957, 3, (6), pp. 185-91
3 LAMM, U.: 'Mercury-arc valves for high voltage DC transmission', Proc. IEE,
1964, III, (10), pp. 1747-753
4 BERNERYD, S., and FUNKE, B.: 'Design of high voltage mercury-arc valves'.
IEE conference on High voltage DC transmission, Publication 22, Manchester,
UK, 1966
5 LIPS, H.P.: 'Semiconductor power devices for use in HVDC and FACTS
controllers'. International colloquium on HVDC and FACTS, Johannesburg,
South Africa, 1997, paper 6.8
6 ADAMSON, C, and HINGORANI, N.G.: 'High voltage direct current power
transmission' (Garraway Ltd, London, 1960)
7 KIMBARK, E.W.: 'Direct current transmission' (Wiley Interscience, New York,
USA, 1971)
8 UHLMAN, E.: 'Power transmission by direct current' (Springer-Verlag, Berlin/
Heidelberg, Germany, 1975)
9 HINGORANI, N.G.: 'High power electronics and flexible AC transmission
systems', IEEE Power Eng. Rev., July 1988

Chapter 2

Static power conversion

2.1 Introduction
The static conversion of power from AC to DC and from DC to AC constitutes the central process of HVDC transmission.
It is therefore important to begin the subject with a clear understanding
of the conversion principles, and of the steady-state relationships, which
exist between the various parameters involved in the process of static
power conversion.
This Chapter describes the requirements of stable converter operation,
the effect of controlled rectification and the commutation phenomena.
Detailed consideration is given to the voltage and current waveforms, and
to the reactive-power demand and harmonic problems attached to converter operation.

2.2 Basic conversion principle


The first consideration to be made in the process of static power conversion is how to achieve instantaneous matching of the AC and DC voltage
levels, given the limited number of phases and switching devices which are
economically feasible.1 With reference to Figure 2.1a, in the absence of
energy storing elements on either side and in the presence of a constant
DC voltage, the time variation of the AC voltage waveform and any
voltage-supply deviations from the nominal level, will cause theoretically
infinite current-level transients.
For practical operation, enough series impedance must therefore be
included to absorb the continuous voltage mismatch between the two sides.
If such an impedance is exclusively located on the AC side (Figure 2.16),
the switching devices transfer the instantaneous direct voltage to the AC
system according to transformer connection and ratio; thus, the circuit

Static power conversion 11

configuration is basically a voltage converter, with the possibility of altering the DC current by thyristor control
If a large smoothing reactor is placed on the DC side (Figure 2.1c), only
pulses of constant direct current flow through the switching devices into
the transformer secondary windings. These current pulses are then transferred to the primary side according to transformer connection and ratio;
thus the result is basically a current converter, with the possibility of
adjusting the direct voltage by thyristor control.
The use of voltage conversion was rejected in mercury-arc converters
owing to the impossibility of recovering from arc-back disturbances. Even
with thyristor schemes, rapid changes in the supply voltage can only be
accommodated within narrow limits and require the use of large series
impedances, which would be uneconomical in terms of reactive-power
(a)

(b)

(0

01
Figure 2.1

AC/DC voltage matching


a Unmatched circuit
b Circuit for voltage conversion
c Circuit for current conversion

VcO

12

High voltage direct current transmission


(a)

(cO

e'

'

s3 i 5

i
(e)

41 i 3 i
4
(0

V6

V2

(f)

-3

J-5
\

- -5

^2

B1

85

V3

85

D !

\ I

R1

Cathode
potential

F !

Neutral

\
Anode
potential
VB

82

f?4

VB

82

5&6 6&1 1&2 2&3 3&4 4&5 5&6 6&1 1&2 2&3
(h)

Figure 2.2

Bridge-conducting sequence and DC-voltage waveforms

Static power conversion 13

compensation. Therefore, the current-conversion principle is generally


accepted as the basis of HVDC converter design.

2.3 Selection of converter configuration


The three-phase bridge, shown in Figure 2.2, is the only configuration used in HVDC transmission. As compared with other possible alternatives, such as the three-phase double star or six-phase diametrical
connections, the bridge configuration provides better utilisation of the
converter transformer and a lower peak-inverse voltage across the converter valves.2
As the Figure shows, two valves are connected to each phase terminal,
one with the anode connected to it (shown on the upper side of the
bridge) and the other with the cathode connected to it (shown on the
lower side of the bridge). The need for two valves conducting in series is
not a drawback in high-voltage applications, particularly with solid-state
converters, because of the need for many series-connected units to withstand the voltage levels used.

2.4 The ideal commutation process2


To understand the operation of a three-phase bridge rectifier let us first
consider the idealised case of a converter bridge connected to an infinitely
strong power system (i.e. of zero source impedance). Under this condition,
the transfer of current (commutation) between valves on the same side of
the bridge takes place instantaneously. The switching sequence and the
rectified voltage waveform are illustrated in Figure 2.2 for the case of an
uncontrolled bridge rectifier (i.e. on diode operation); valves 1, 3, 5 at the
top and 4, 6, 2 at the bottom are connected to phases red, yellow and blue,
respectively.
With reference to Figures 2.2a and g, and starting at instant A, phases R
and Y are involved through conducting valves 1 and 6. This operating
state continues up to point B, after which valve 2 becomes forward biased,
since its anode, directly connected to that of valve 6, is positive with
respect to its cathode (connected to phase blue); therefore at point B the
current commutates naturally from valve 6 to valve 2 (Figure 2.2b).
A similar argument applies at point C, with reference to valves 1 and 3
on the upper half of the bridge. The anode of valve 3 (connected to phase
Y) begins to be positive with respect to its cathode (connected to phase R
through the conducting valve 1) and a commutation takes place from valve
1 to valve 3 (Figure 2.2c). This is followed by commutation from valve 2 to
valve 4 at point Z), valve 3 to valve 5 at point E, valve 4 to valve 6 at point

14

High voltage direct current transmission

F, and valve 5 to valve 1 at point G. This completes the switching-cycle


sequence.
The output waveform in Figure 2.2g shows the voltage variation of the
positive (common cathode) and negative (common anode) poles with
respect to the transformer neutral. Figure 2.2A shows the output voltage,
i.e. the voltage of the positive pole with respect to the negative pole. It is
seen that the output voltage has a ripple, or harmonic frequency, of six
times the main frequency.
Each valve carries the full value of direct current for one third of the
cycle, and there are always two valves conducting in series.

2.4.1 Effect of gate control


By delaying the firing instants of the valves with respect to the voltage
crossings, the commencement of the natural commutations described in
Section 2.4 can be delayed by a definite time interval and the effect of this
action on the direct-voltage waveforms is illustrated in Figures 2.3a and b.
It is noticeable that the voltage area, and therefore the mean direct
voltage, is reduced in proportion to the magnitude of the delay.
For delay angles above 60 some negative voltage periods begin to
appear. If the bridge output were connected to a pure resistance, the
bridge unidirectional current-conduction property would prevent reverse
current flow during these negative voltage periods, and the operation
would then be intermittent. However, the provision of a large smoothing
reactor maintains positive current flow during the negative periods, and
energy is transferred from the reactor magnetic field to the AC system.
The voltage waveforms for a delay of 90, illustrated in Figures 2.4a and
b, show equal positive and negative voltage regions (indicated by horizontal
and vertical shaded areas, respectively); the mean direct voltage is therefore zero with a 90 delay.
Beyond 90 the mean voltage is negative and bridge operation can only
be maintained in the presence of a DC power supply. This supply overcomes the negative voltage and forces the current to conduct in the same
direction (i.e. from anode to cathode), in opposition to the induced e.m.f.
in the converter transformer. This indicates that power is being supplied
to the AC system, i.e. the converter is inverting. Figures 2.5a and b illustrate the (ideal) limiting case of full inversion which would require a delay
of 180.
Three conditions are thus required to permit power inversion:
(a) an active AC-voltage source which provides the commutating voltage
waveforms;
(b) provision of firing-angle control to delay the commutations beyond a =
90;
(c) a DC power supply.

Static power conversion 15

(a)

MR)
Cathode busbar

Neutral

- Anode busbar

Reference
anode busbar

Figure 2.3

Effect offiring delay on voltage waveforms


a Common-anode and common-cathode voltages
b Direct voltage
c Voltage across valve 1

16

High voltage direct current transmission

MR)

3<Y)

S(6J

(a)

(c)

Figure 2.4

Voltage waveforms with 90 firing delay


a Common-anode and common-cathode voltages
b Direct voltage
c Voltage across valve 1

Static power conversion 17


6(V)

Figure 2.5

2(8)

4(f?)

Voltage waveforms on full inversion


a Common-anode and common-cathode voltages
b Direct voltage
c Voltage across valve 1

2.4.2 Valve current and voltage waveforms


In the presence of a large smoothing reactor on the DC side, the voltage
waveform of Figure 2.36 will produce a constant direct current, the level
of which depends on the mean voltages at both ends of the link and the
link resistance. For the idealised commutating conditions described in
Section 2.4, the valve current will be a rectangular pulse lasting 120, its

18

High voltage direct current transmission

relative position with reference to the corresponding voltage waveform


being determined by the firing-delay angle a.
During the conducting period, and disregarding the small internal
voltage drop, the voltage across the valve is zero. Outside the conducting
period, the voltage across the valve consists of varying portions of two
phase-to-phase voltage waveforms; these are the phase to which the valve
is directly connected and the phase of the conducting valve on the same
side of the bridge. Examples of valve-voltage waveforms for different
delay angles are shown in Figures 2.3c, 2.4c and 2.5c.

2.5 The real commutation process


In practice, the zero-impedance supply required to produce the voltage
and current waveforms described in Section 2.4 does not exist. Even if the
AC system impedance were negligible, there is considerable transformer
leakage reactance between the converter and the AC system. In theory, the
converter transformer is not essential to the process of static power conversion. However, there are practical reasons for using converter transformers, like the possibility of phase shifting multiple bridges and the
availability of on-load tap changing, which will become apparent when
discussing harmonics and reactive-power compensation later on. The main
effect of AC system reactance is to reduce the rate of change of current or,
in other words, to lengthen the commutating time.
During the commutation, the magnetic energy stored in the reactance of
the previously-conducting phase has to be transferred to the reactance of
the incoming phase. That energy only depends on the direct-current level
and the inductance per phase. The speed of the commutation process, on
the other hand, and therefore the rate of change of current, are also
affected by two other parameters, i.e. the supply voltage and the firingdelay angle.
It is thus essential, before analysing the commutation process, to define
clearly what is meant by commutating voltage and commutation reactance.

2.5.1 Commutating voltage


The commutating voltage can be defined as the voltage appearing on the
DC line during the periods when no commutation is taking place. In this
operating region only direct current flows through the AC system impedance, and the voltage waveform is therefore sinusoidal.
Considering the AC-current waveform distortion produced by the
converter, it will be necessary to go back to the system source voltage to
find an undistorted supply to the converter. In practice, however, phase
shifting and filtering are provided with every HVDC converter station and
the voltage waveform at the filter busbar is reasonably sinusoidal (under

Static power conversion 19

steady-state and normal operating conditions!). Such voltage can, therefore, be used as the commutating voltage.

2.5.2 Commutation reactance3


The commutation reactance can be defined as the reactance between the
commutating voltage, as defined in Section 2.5.1, and the converter valves.
Figure 2.6 shows the general case of n bridges connected in parallel on the
AC side. In the absence of filters, pure sinusoidal voltages only exist
behind the system source impedance (Xj and in such a case the commutation reactance (XCJ) for the jth bridge is given by
(2.1)

Xcj=Xss+Xtj

If the bridges are under the same controller, or under identical controllers,
it is preferable to create a single equivalent bridge. The commutation reactance of such an equivalent bridge depends upon the DC connections and
on the phase shift between the bridges.
Bridges with the same phase shift will commutate simultaneously and
the equivalent reactance must reflect this. For a series connection of k
bridges the commutation reactance of the equivalent bridge is
Xc(series) = kXss

xtj

(2.2)

where j represents any of the n bridges.


If the bridges are connected in parallel on the DC side the equivalent
bridge commutation reactance is
Xc(parallel) = Xss + - Xtj
k

Figure 2.6

(2.3)

n bridges connected in series on the DC side and in parallel on the


AC side

20

High voltage direct current transmission

With perfect filtering, or with a combination of filters and transformer


phase shift (refer to Section 2.5.1), the voltage on the AC side of the
converter transformers may be assumed to be sinusoidal and hence Xss has
no influence on the commutation. It must be noted that HVDC schemes
are normally designed for 12-pulse operation and that filters are always
provided (i.e. the system impedance can be ignored).
However, the presence of local plant components at the converter terminals, such as synchronous compensators, will affect the commutation reactance. Since the commutation produces a phase-to-phase short circuit
during a small fraction of a cycle, the synchronous machines must be
represented by their subtransient reactances.
By way of example, let us consider the two ends of the original New
Zealand HVDC link (with reference to Figures 2.7 and 2.8).
At the receiving (Haywards) end (Figure 2.7) the effect of the subtransient reactance of the synchronous compensators on the tertiaries of the
converter transformers must be taken into account. The approximate
equivalent circuit is illustrated in Figure 2.7& and the commutation reactance is

xp + xt+x"d
where Xv is the transformer secondary leakage reactance, Xp is the transformer primary leakage reactance, Xt is the transformer tertiary leakage reactance and X"(i is the subtransient reactance (direct axis) of the synchronous
condenser unit.
At the sending (Benmore) end (Figure 2.8) the subtransient reactance of
the generators is combined in parallel with the secondary reactance of the
interconnecting transformer. The primary reactance is beyond the filters
and can thus be neglected. An approximate equivalent circuit is illustrated
in Figure 2.86. Although there are two converter groups commutating on
this reactance, the commutations are not simultaneous owing to the 30
phase-shift of their respective transformers. Thus, the effective commutation reactance per group is

X"d + nXs

(2.5)

where X is the two-winding transformer leakage reactance, Xs is the interconnecting transformer secondary leakage reactance (note filters
connected to tertiary winding), X"a is the generator subtransient reactance
and n is the number of generators connected. Eqn. 2.5 is only valid for
commutation angles not exceeding 30; however, this covers most operating conditions.

Static power conversion 21


(a)
Haywards
a.c.
110 kV

sc

To North
Island
system

SC

Figure 2. 7 Receiving end of the New Zealand DC link


a Simplified diagram
b Equivalent circuit to calculate the commutation reactance

22

High voltage direct current transmission

To South island
system

o-

Figure 2.8

-CO-fw

Sending end of the New Zealand DC link


a Simplified diagram
b Equivalent circuit to calculate the commutation reactance

Static power conversion 23

2.5.3 Analysis of the commutation circuit


Let us consider the commutation process between valves 1 and 3 of a
converter bridge, connected to a system with a source voltage, vc, a commutation reactance per phase, Xo and negligible source resistance.
With reference to Figure 2.9, commutation from valve 1 to valve 3 can
start (by the firing of 3) any time after the upper voltage crossing between
t;CR and t;CY (and must be completed before the lower crossing of these
two voltages).
Since VQY > <^CR> a commutating current ic{= %) builds up at the expense
of i\ so that at all times
h +h = U

(a)
3

'3

VCY

*CB

(c)

Figure 2.9

The commutation process


a Equivalent circuit of the commutation from valve 1 to valve 3
b Voltage waveforms showing early (rectification) and late (inversion) commutations
c The commutating currents

24

High voltage direct current transmission

As the rates of change of % and -i\ are equal (provided that the commutation reactances are balanced), the voltage drops across ^ c R and XCY are the
same and thus, during the overlap period, the direct voltage vd is the mean
value of t^GY a n d t>cR.
From the circuit of Figure 2.9a and assuming XCR = XQY = XQ w e c a n
write
^CY - ^CR = 2(Xc/co)d(ic)/dt

(2.6)

Taking as a reference the voltage crossing between phases R and Y


c

sin cot

where Vc is the phase-to-phase r.m.s. voltage.


Eqn. 2.6 can also be written as
Vc sin (wt) d(cDt) = Xcdic

(2.7)

and integrating from a)t= a


- i r \^VC sin (wt) dicot) = Xc fQd(ic)

(2.8)

The instantaneous expression for the commutating current is thus


ic =

yc
/2X

[cos a - cos (cot)]

(2.9)

and substituting the final condition, i.e. ic = Id at cot = oc + u yields


Id = -p^- [cos a - cos (a + u)]

(2.10)

2.6 Rectifier operation


Typical voltage and current waveforms of a bridge operating as a rectifier
with the commutation effect included are shown in Figure 2.10, where P
indicates a firing instant (e.g. Pi is the firing instant of valve 1), *S indicates
the end of a commutation (e.g. at S5 valve 5 stops conducting) and C is a
voltage crossing (e.g. C\ indicates the positive crossing between phases
blue and red).
Figure 2.10a illustrates the positive (determined by the conduction of
valves 1, 3, 5) and negative (determined by the conduction of valves 2, 4,
6) potentials with respect to the transformer neutral. Figure 2.106 shows
the direct voltage output waveform.
The potential across valve 1, also shown in Figure 2.106, depends on the
conducting valves. When valve 1 completes the commutation to valve 3 (at
SI) the voltage across will follow the red-yellow potential difference until
P4. Between P4 and S2 the commutation from valve 2 to valve 4 (see

Static power conversion 25

Figure 2.10

Typical six-pulse rectifier operation


a Positive and negative direct voltages with respect to the transformer neutral
b Direct bridge voltage Vfj and voltage across valve 1
c,d Valve currents i\ to i$
e AC line current of phase R

26

High voltage direct current transmission

Figure 2.10a) reduces the negative potential of phase red and causes the
first voltage dent. The firing of valve 5 (at Ph) increases the potential of
the common cathode to the average of phases yellow and blue; this causes
a second commutation dent, at the end of which (at S3) the common
cathode follows the potential of phase blue (owing to the conduction of
valve 5). Finally, the commutation from valve 4 to valve 6 (between P6 and
S4) increases the negative potential of valve 1 anode and produces another
voltage dent.
Figures 2.10c and d illustrate the individual valves (1 and 4) and Figure
2.10*? the phase (red) currents, respectively.
A number of reasonable approximations have to be made to simplify the
derivation of the steady-state equations that follow. These are:

converter valves are treated as ideal switches; when calculating the


power loss, the valve resistance can be added to that of the DC transmission line;
AC systems consist of perfectly balanced and sinusoidal e.m.f.s, the
commutation reactances are equal in each phase and their resistive
components are ignored; the main effects of nonideal supply waveforms are discussed in Section 2.9.2;
direct current is constant and ripple free, i.e. the presence of a very
large smoothing reactor is assumed; the effect of nonideal DC current
waveforms is discussed in Section 2.9.2;
only two or three valves conduct simultaneously, i.e. two simultaneous
commutations are not considered; the low AC voltage and/or high DC
current required to cause simultaneous commutations are prevented in
the steady state; during disturbances, on the other hand, the converter
behaviour can only be predicted by dynamic analysis.8

2.6.1 Mean direct voltage


The following expression can easily be derived for the average output
voltage with reference to the waveforms of Figure 2.10
Vrf= (1/2)7,0 [cos a + cos (a + u)]

(2.11)

where Vc0 is the maximum average DC voltage (i.e. at no load and without
firing delay); for the three-phase bridge configuration Vc0 = (3y/2/n)Vc,
and Vc is the phase-to-phase r.m.s. commutating voltage referred to the
secondary or valve side of the converter transformer.
Eqn. 2.11 specifies the DC voltage in terms of Va a and u. However, the
value of the commutation angle is not normally available and a more
useful expression for the DC voltage, as a function of the DC current, can
be derived from eqns. 2.10 and 2.11, i.e.
SX

Vd = V c 0 cos a - / r f
n

(2.12)

Static power conversion 27

2.6.2 AC current
The r.m.s. magnitude of a rectangular current waveform (neglecting the
commutation overlap) is often used to define the converter transformer
MVA, i.e.
U = y/{ (1/TT) ln^\ I2dd(ojt)} - V2VV3

(2.13)

Since harmonic filters are normally provided at the converter terminals,


the current flowing in the AC system contains only fundamental component frequency, and its r.m.s. magnitude (obtained from the Fourier analysis described in Section 2.9) is
(2.14)

Ii = Idy/f>/n

If the effect of commutation reactance is taken into account, the current


waveform for a star-star-transformer connection is as shown in Figure
2A0e. Using eqns. 2.9 and 2.10 the currents of the incoming and outgoing
valve during the commutation are defined by eqns. 2.15 and 2.16, respectively
i

7 rf (cosa-cos<^)
cos a - cos (a + u)

i=Id-Id

for a < cot < a + u

cos a - cos (cot - 2TT/3)

cos a - cos (a + u)

(2.15)

for
a+

In between commutations the current is

< mt < a +
3

+u

(2.16)

i= Id fora + u < cot < + a

(2.17)

The fundamental component of the current waveform defined by eqns.


2.15, 2.16 and 2.17 is
I=^-Id J{[cos 2oc - cos 2(oc + u)f + [2w + sin 2oe nsin 2(oc + w)]2}/{4[cos a - cos (oc + u)]}

(2.18)

2.7 Inverter operation


The conditions for inverter operation have been described in Section 2.4.1
with reference to an ideal system without commutation reactance. In prac-

28

High voltage direct current transmission

tice, full inversion cannot be achieved and the delay angle must be less
than 180.
With reference to Figures 2.11a and e, a commutation from valve 1 to
valve 3 (at P5) is only possible as long as phase Y is positive with respect to
phase R. Furthermore, the commutation must not only be completed
before C6, but some extinction angle y\ (> y0) must be left for valve 1,
which has just stopped conducting, to re-establish its blocking ability. This
puts a limit to the maximum angle of firing a = % - (u + y0) for successful
inverter operation. If this limit were exceeded, valve 1 would pick up the
current again, causing a commutation failure.
Moreover, there is a fundamental difference between rectifier and inverter operations which prevents an optimal firing condition in the latter case.
Although the rectifier delay angle, a, can be chosen accurately to satisfy a
particular control constraint, the same is not possible with respect to angle y
because of the uncertainty of the overlap angle, u. Events taking place after
the instant of firing are beyond predictability and, therefore, the minimum
extinction angle, yo, must contain a margin of safety to cope with reasonable uncertainties (values between 15 and 20 are typically used).
The analysis of inverter operation is not different from that of rectification, carried out in Section 2.6, and will not be repeated here. However,
for convenience, the inverter equations are often expressed in terms of the
angle of advance p (= n - a) or the extinction angle y (= p - u).
Thus, omitting the negative sign of the inverter DC voltage, the following expressions apply
^

(2.19)

or
(2.20)
or
Vd = Yfl (cos p + cos y)

(2.21)

The expression for the direct current is


Id = ~ [cos y - cos p]
2X

(2.22)

2.8 Power factor and reactive power4


Owing to the firing delay and commutation angles, the converter current
in each phase always lags its voltage (refer to Figure 2.1(k). The rectifier
therefore absorbs lagging current (consumes VARs).

Static power conversion 29

(e)

Figure 2.11

Typical six-pulse inverter operation


a Positive and negative direct voltage with respect to the transformer
neutral
b Voltage across valve 1, and direct bridge voltage Vd
c, d Valve currents i\ to i
e AC line current of phase R

30

High voltage direct current transmission

In the presence of perfect filters, no distorting current flows beyond the


filtering point, and the power factor can be approximated by the displacement factor (cos </>), where (f) is the phase difference between the fundamental-frequency voltage and current components.
Under these idealised conditions, with losses neglected, the active fundamental AC power (P) is the same as the DC power, i.e.
P= J$VcIcos 4> = VdId

(2.23)

cos (j> = VJJiy/S VJ)

(2.24)

and
Substituting Vd and Id from eqns. 2.11 and 2.14 in eqn. 2.24 the following
approximate expression results
cos (p = V2[cos a + cos (a + u)]

(2.25)

The reactive power is often expressed in terms of the active power, i.e.
<2=P.tan</>
where tan (f) (derived from eqns. 2.18 and 2.24) is
sin (2a + 2u) - sin 2a - 2u
t a n <p =

(2.26)

h,

Kl.ll)

cos 2a - cos (2a + 2u)


Similarly to eqn. 2.25, the following approximate expression can be written
for the power factor of the inverter
cos 4> = V2[cos y + cos jB]

(2.28)

Referring to the AC-voltage and valve-current waveforms in Figures 2.11a


and e it is clear that the current supplied by the inverter to the AC system
lags the positive half of the corresponding phase-voltage waveform by
more than 90, or leads the negative half of the same voltage by less than
90. It can either be said that the inverter absorbs lagging current or
provides leading current, both concepts indicating that the inverter, like
the rectifier, acts as a sink of reactive power. This point is made clearer in
the vector diagram of Figure 2.12.
Eqns. 2.23, 2.25 and 2.26 show that the active and reactive powers of a
controlled rectifier vary with the cosine and sine of the control angle,
respectively. Thus, when operating on constant current, the reactive-power
demand at low powers (<^> 90) can be very high.
However, such an operating condition is prevented in HVDC converters
by the addition of onload transformer tap changers, which try to reduce
the steady-state control angle (or the extinction angle) to the minimum
specified. Under such controlled conditions, Figure 2.13 shows a typical
variation of the reactive power demand against the active power of an

Static power conversion 31

Vc

- Vc

Figure 2.12

Vector diagrams of current and power


Suffix R for rectification
I for inversion

CMp.u.)

Ptp.u.)

Figure 2.13

Variation of reactive power with active power

32

High voltage direct current transmission

HVDC converter; the reactive-power demand is shown to be approximately


60% of the power transmitted at full load.

2.9 Maximum available power 5


For a given AC-DC system configuration and operating condition there is
a unique Pj/Id characteristic, such as that illustrated in Figure 2.14. The
characteristic is derived from the steady-state equations assuming that the
DC current adjusts itself immediately to the change in current order, while
the AC system e.m.f., the inverter minimum y and the tap-changer position remain unchanged.
When considering the inverter-power capability, it is assumed that the
rectifier provides no limitation to the supply of DC current at rated DC
voltage. If the inverter is operated throughout at minimum constant y, the
resulting characteristic will represent the maximum obtainable power for
the system parameters being considered. This curve is termed the
maximum power curve (MPC). Any power can be obtained below the MPC
by increasing a and y, but power can be obtained only if one or more of
the system parameters are changed, e.g. reduced system impedance,
increased system e.m.f., using larger capacitor banks etc.
A similar MPC curve can be obtained for the rectifier at minimum
constant a. The MPC exhibits a maximum available power (MAP) point, as
can be seen in Figure 2.14. An increase of current beyond the MAP
reduces the DC voltage to a greater extent than the corresponding DCcurrent increase. This could be counteracted by changing the AC-system
conditions, e.g. by controlling the AC terminal voltage. It should be noted
that dPc/dld is positive for operation at DC currents smaller than /MAP*
whereas dPydId is negative at DC currents larger than /MAP-

MAP
MPC for y
constant

Figure 2.14

DC power-DC current curve for y minimum

Static power conversion 33

2.10 Characteristic converter harmonics4


The term harmonics is used to define the sinusoidal components of a
repetitive waveform and these consist exclusively of frequencies which are
exact multiples (harmonic orders) of the basic repetition frequency (i.e.
the fundamental). The full set of harmonics forms a Fourier series which
completely represents the original waveform.
The original waveform can thus be described by its time-domain data
(i.e. at any given instant in time the amplitude of the waveform is
displayed) or by its frequency-domain data (i.e. by the magnitudes, and
often phase, of its Fourier components).
The general trigonometric form of the Fourier series is
F((ot) = + Z {\ cos (ncot) + Bn sin (ncot)}
2

(2.29)

n=l

where co is the basic repetition frequency in radians per second, and


1

71 "u

An = \J nF((ot) cos (ncot) d(cot)


n
Bn = 1 fj2n F(cot) sin (ncot) d(cot)
n

(2.30)
(2.31)
(2.32)

where a is the angle, A / 2 is the average value of the function F(cot) and A^
and Bn are rectangular components of the nth harmonic. The corresponding vector is
An-jBn=CnZcl>n

(2.33)

where Cn = \M 2 n + B2n = crest valve and 4>n = tan'


HVDC converters generate harmonic voltages and currents on the DC
and AC sides, respectively. It is convenient to separate the converter
harmonics into two groups, termed characteristic and noncharacteristic.
The orders of the characteristic harmonics are related to the pulse
number of the converter configuration, defined as the number of nonsimultaneous commutations per cycle of the fundamental frequency.
A converter of pulse number p ideally generates only characteristic
voltage harmonics of orders pk on the DC side, and current harmonics of
orders pk + 1 on the AC side (where k is any integer).
The derivation of the characteristic harmonics is based on the following
assumptions:
(a) The supply voltages are displaced exactly by one third of a cycle in
time from each other and consist only of fundamental frequency.

34

High voltage direct current transmission

(b) T h e direct current is perfectly constant (i.e. has no frequency components). This can only be achieved if the DC smoothing reactor has infinite inductance.
(c) The valves begin conducting at equal time intervals.
(d) The commutation impedances are the same in the three phases (i.e. all
the overlap angles are the same).
Direct voltage harmonics: Using as a reference the single three-phase bridge
configuration (i.e. p = 6), the order of harmonics is n = 6k The repetition
interval (see Figure 2.106) is TT/3 and it contains three different functions
which, using as a time reference the voltage crossings, are expressed as
follows
cot + \ forO < mt < a
6J
I

1 -

(2.34)
JE

in cot = Vc. cos cot

for a < cot < a + u

(2.35)

f o r a + u < cot < ^ (2.36)


6J
3
and using the Fourier equations, the r.m.s. magnitudes of the harmonic
voltages are obtained from the equation
ccos\cot-\

cos

( + ! ) u + ( ,+ I ).2, o cos 2* ,( - I ^u
)!!

u | cos
- 2(n - \){n + 1) cos | (n + 1)

u
qcos(2oe
+ ^)f

1/2

(2.37)

Figures 2.15 and 2.16 give the sixth and 12th harmonics 2 as a percentage
of Vc0 = S(y/2) Vc/n. These curves and equations show some interesting
facts. First for a = 0 and u = 0, eqn. 2.37 reduces to
Vno = j2Vc0/(n2-l)

(2.38)

or

i^

(2.39)

giving 4.04, 0.99 and 0.44 per cent for the sixth, 12th and 18th harmonics,
respectively. Generally, as a increases, harmonics also increase, and for a =
(TT/2) and u = 0

(2.40)

Static power conversion 35

10

20

30

40

Angle of overlap u

Figure 2.15

Variation of sixth-harmonic voltage in relation to angle of delay and


overlap

which produces n times the harmonic content corresponding to a = 0. This


means that the higher harmonics increase faster with a. Eqn. 2.40 is of
some importance as it represents the maximum proportion of harmonics
in the system, particularly when it is considered that at a = 90, u is likely
to be very small.
If the converter involves two bridges, one with a star-star-connected
transformer and the other with a delta-star or star-delta transformer, their
voltages will be 30 out of phase and so the harmonics will accordingly be
out of phase. Since 30 of main frequency corresponds to a half cycle of
sixth harmonic, this harmonic will be in phase opposition in the two
bridges. Similarly for the 12th harmonic, 30 corresponds to one cycle,
giving harmonics in phase; for the 18th harmonic, 30 corresponds to one
and a half cycles, giving harmonics in opposition and so on.

36

High voltage direct current transmission

10

20

30

Angle of overlap u

Figure 2.16

Variation of 12th-harmonic voltage in relation to angle of delay and


overlap

AC current harmonics: In the absence of commutation reactance, the current


waveform for a star-star-connected converter transformer, shown in
Figure 2.17a, can be defined as follows
i L

for - < cot < --

i =0

for - - < cot < - and < cot < -

2n

i = -Id

for - n < cot <

2n

,
and

2n

2n

< cot < n

(2.41)

Static power conversion 37


cot = 0

- JI/2

JI

(a)

(b)

Figure 2.17

Idealised phase-current waveforms on the primary side


a Star-star transformer connection
b Delta-star transformer connection

The Fourier series for such a waveform is


i = - ^ Id(cos cot - \ cos bent + % cos 7cot - \ i cos 1 lot + ...)

(2.42)

with harmonic orders determined from the expression


n = 6k 1

(2.43)

where k is an integer.
The magnitude of the nth harmonic is given by
In

*-d

(2.44)

nn
and that of the fundamental
(2.45)

For the star-delta or delta-star transformer connection, the current


waveform is given by Figure 2.176 and the Fourier series is
/rf(cos oot+% cos hot - % cos lot - \ x cos 1 lot + ...)

(2.46)

38

High voltage direct current transmission

Eqns. 2.42 and 2.46 are the same excepting that harmonics 5, 7 (k = odd
numbers in eqn. 2.43) are of opposite sequence, and therefore with two
bridges in series as above, only the harmonics corresponding to n = \2k
1 will enter the AC system.
The current waveform and harmonic spectrum of a double-bridge 12pulse configuration are illustrated in Figure 2.18 (with the overlap angle
ignored).
If the commutation angle is taken into account, the current waveform
for the star-star connection has been defined in Section 2.6.2; the characteristic fifth, seventh, 11th and 13th harmonics, as a percentage of the
fundamental (I\), are illustrated in Figures 2.19-2.22, inclusive.2 It is seen
that the harmonics decrease with increases in commutation angle (u), the
rate of decrease being greater for higher harmonics. For the same u,

1/V3

* * Time

(a)

1.0
0.8

a.

<

0.6
0.4
0.2

J L
11

13

23 25

Frequency

x Fundamental
(b)

Figure 2.18

Idealised phase-current waveform with 12-pulse operation


a Current waveform
b Frequency domain representation

Static power conversion 39


20

19

cr

\ \
IT
2
c

18

Q>
(D

17

CO

16

a = 30\Y
a = 20

15

a = 10 \
14

10

20

30

a =
40

Angle of overlap u (degrees)

Figure 2.19

Variation of fifth-harmonic current in relation to angle of delay and


overlap

changes in a make little difference outside the range between a = 0 and a =


10. A simplifying assumption for the analysis of eqn. 2.18 can thus be
made by using a = 0.
Harmonics tend to reach a minimum at approximately u = 360/n and
then increase slightly. It may also be noted that, during normal operation,
a is small (say up to 10) and u is large (say 20), whereas during disturbances, when a is nearly 90, u is very small and the harmonics approach
their maximum.
For inverter operation a and (a + u) should be replaced by the extinction
angle y and the angle of advance /?, respectively

2.11 Noncharacteristic harmonics


The ideal conditions, used in the last section to calculate the characteristic
harmonics produced by HVDC converters, are not met in practice and, as

a = 0

10

20

30

Angle of overlap u (degrees)

Figure 2.20

Variation of seventh-harmonic current in relation to angle of delay


and overlap

a result, relatively small quantities of noncharacteristic harmonics are


always present.
The interaction of current and voltage frequency components between
HVDC converters and their interconnected AC and DC systems is very
complex, encompassing a time scale of minutes for transformer coresaturation instability, to multiples of one cycle for subsynchronous generator shaft torsional interactions, to fractions of one cycle for supersynchronous harmonic and other waveform distortion problems. In the
supersynchronous region the characteristic harmonics are the most significant and filters are normally used to limit their penetration into the AC
and DC systems.
Noncharacteristic harmonics arise from a number of causes, typically a
system imbalance, presence of fundamental-frequency current on the DC
link and AC-system nonlinearities. Harmonically-unrelated frequencies
exist whenever the fundamental supply frequencies of the interconnected
AC systems are not identical. This effect is more critical in the case of

Static power conversion 41

2
c
CD
CO

(0
W
CO

10

20

30

40

Angle of overlap u (degrees)

Figure 2.21

Variation of 1 lth-harmonic current in relation to angle of delay and


overlap

0)
CO
T3
C

10

20

30

40

Angle of overlap u (degrees)

Figure 2.22

Variation ofUthharmonic
overlap

current in relation to angle of delay and

42

High voltage direct current transmission

back-to-back interconnectors because of the close coupling which exists


between the two AC systems.
To some extent, every DC scheme experiences crossmodulation effects
at noncharacteristic frequencies. A clear understanding of the crossmodulation phenomena will help to design control-system philosophies to minimise the cost of additional plant components, often after commissioning
of the link.
With reference to AC-current harmonics, the term uncharacteristic indicates frequencies other than those determined by the expression p.k 1
(for k = 0, ... , n) where p is the pulse number of the valve group, i.e. six
for a single bridge group and 12 for a double bridge group.
Strictly speaking, a harmonic is an integer multiple of the fundamental
frequency. However, here the term harmonic is used in a broader sense, to
represent also interharmonics (frequencies of noninteger multiple orders)
and subharmonics (frequencies below the fundamental). Therefore, in this
section the factor k is used to indicate frequencies as a multiple of the
fundamental; k may or may not be an integer.

2.11.1 Harmonic crossmodulation6


The harmonic transfers through an HVDC converter are best explained
using modulation theory.7'8
The function of the HVDC converter as a modulator is twofold. First, it
takes a three-phase positive-sequence AC voltage waveform and, by switching consecutively through the phases, ensures that a DC voltage is always
applied on the DC side of the converter. In a twelve-pulse converter, every
30 degrees there is a thyristor switching that connects a combination of
phases which maintains a constant average voltage on the DC side. In this
way, the fundamental frequency waveform is modulated down to DC.
Secondly, the same switching pattern takes the DC current on the DC
side, and switches it onto the AC-side phases in such a way that a fundamental-frequency positive-sequence AC current exists on the AC side.
The voltage and current relationships across the converter can be
expressed as follows
(2-47)

h = >VAC *DC

(2-48)

where F^DC and >VAC are transfer functions for the voltages and current,
respectively, and W = 0, 120, 240 for each of the three phases.
In the absence of commutation overlap (Figure 2.23 - in dotted lines),
the transfer functions for the voltage and current modulation are rectangular waveforms and can be expressed as9

Static power conversion 43


n=\

cos n

Yvc= Z An cos n (coi t + 2 V 3 )


n=l

(2.49)

where
An = 4/n . Vw sin WV2 . cos n V 6

(2.50)

This process is illustrated in Figure 2.24, which shows the modulated


output current on the AC side of the converter in response to a DC
current which contains a ripple frequency.
When the commutation angle is taken into account, the voltage transfer
functions (Figure 2.23a - continuous line) are shown as in eqn. 2.49 but
with the value of An replaced by
A,M = A n . c o s n %

(2.51)

where \i is the commutation angle.


Regarding the current transfer function, the assumption of a linearlychanging commutation current during the commutation (Figure 2.236 continuous line) leads to the following alternative expression for the coefficient An
^

(2.52)

For an accurate quantitative assessment, the analysis must include the


effects of pulse-position and pulse-duration modulation as affected by the
modulating frequencies as well as explicit representation of the converter
controller.
The switching pattern is synchronised with the AC-side fundamental
frequency, and as such contains a large fundamental component. As
switching is an on-off process, harmonics of the fundamental are present
as well. These manifest themselves as the characteristic harmonics on both
sides of the converter, i.e. on the DC side, harmonics 12n, and on the AC
side, harmonics 12n + 1 in positive sequence, and 12n - 1 in negative
sequence, where n is an integer. These components are always present,
even under ideal (undistorted AC voltage and DC current) operating
conditions.
Any noncharacteristic frequencies on the AC or DC sides are subjected
to exactly the same modulation process by the converter.
Because the converter switches consecutively between phases, the phase
sequence of the noncharacteristic frequencies is very important, and

44

High voltage direct current transmission

Aaji

Aaji

Aaji

Aa(i

-1

(a)
Y

\j/AC

-1

Figure 2.23

Transfer functions F^ D c and *VAC


a Transfer function to DC voltage
b Transfer function to AC current

/ ac /phase

+1

switching
ing function

n
; modulated output
.a
o

modulating function
DC + ripple

Figure 2.24

Idealised switching function and modulating function giving modulated AC output

Static power conversion 45

should normally be stated. However, the converter is neither sensitive to


nor generates zero-sequence components on the AC side.
The situation is further complicated by the fact that the switching
pattern of the HVDC converter bridges is affected by distortion in the ACvoltage and DC-current waveforms. Usually, DC current directly affects
the firing instants of the thyristors through the converter control, and the
commutation period is directly affected by both AC voltage and DC
current. These effects have surprising significance; although the switching
pattern varies only a little, it affects the transfer of kilo-amperes to the AC
side, and kilo-volts to the DC side. Even a very small noncharacteristic
frequency affecting the thyristor switching instants leads to significant
noncharacteristic distortion around the converter.

Transfer of noncharacteristic harmonic voltage from the AC to the DC side


The largest component of the thyristor switching function is at the fundamental frequency. This component modulates positive-sequence voltage at
harmonic multiple k+ 1, and negative-sequence voltage at harmonic multiple k - 1, to multiple k on the DC side. The voltages are switched across in
the same ratio as the fundamental is, except that the firing-delay angle
does not reduce the magnitude but merely phase shifts the voltage. Thus,
if the maximum DC voltage (zero firing-delay angle and no voltage drop
due to commutation) is 550 kV and the AC voltage 345 kV, with a commutation period |n, then harmonic voltages will be increased by 550/345 cos
OY2) across the converter. To a good approximation, the percentage
voltage distortion on the AC side is preserved in its transfer to the DC
side. This mechanism is not unique to HVDC converters; in fact any
three-phase rectifier or inverter demonstrates similar characteristics.
A secondary mechanism comes from the AC-voltage distortion affecting
the commutation period, which modifies the harmonic voltage level on the
DC side by up to ten per cent, and the phase angle of that distortion by up
to 0.3 radians. This effect is frequency dependent, as well as being closely
related to variables such as firing-delay angle and commutation-period
duration.
Further noncharacteristic harmonic voltages appear on the DC side,
associated with the other components of the thyristor switching pattern. A
positive-sequence voltage at harmonic multiple k + 1 on the AC side will
appear as a voltage at multiples k and 12n + k on the DC side; k may or
may not be an integer. The same frequencies will appear for an AC-side
negative-sequence voltage at harmonic multiple k - 1. These higher-order
noncharacteristic harmonics would be expected to be at a level reduced by
a factor of approximately l/(12n) from the terms described in the first
paragraph of this section, if the switching instants remained unaffected by
the distortion.

46

High voltage direct current transmission

However, the same secondary mechanism also applies to these harmonics. Unfortunately, the spectrum that appears as a result of commutationperiod variation does not reduce nearly as quickly with increasing order.
Thus, for the terms 12 k, commutation-period variation is as important
a mechanism as the direct transfer, and for higher orders it is substantially
more important. This variation is more difficult to describe and generalise.
However, these harmonics can be expected to be present at levels up to 20
% of the terms described in the first paragraph of this section.
In summary, the frequencies on the DC side from a positive-sequence
harmonic multiple k + 1, or a negative-sequence harmonic multiple k - 1,
can be written as
/DC = (12n k)f0

for n e (0,1,2,3, ... )

(2.53)

Transfer of noncharacteristic harmonic current from the DC to the AC side


The fundamental component of the switching function modulates current
at harmonic multiple k to multiple k + 1 in positive sequence and k - 1 in
negative sequence on the AC side. Considering just DC on the DC side, it
becomes multiple +1 in positive sequence, and multiple -1 in negative
sequence on the AC side; negative frequency in negative sequence is
equivalent to positive frequency in positive sequence, and the two components add together to give the total AC current. Harmonic currents are,
therefore, switched across at half the ratio of the direct current, i.e. if the
DC current is 2 kA and the fundamental-frequency AC current 2.6 kA,
then harmonic currents will be changed by (1/2) (2.6/2) across the converter.
A secondary mechanism comes from the DC-current distortion affecting
the commutation period, which modifies the harmonic level on the AC
side by up to 15 %, and the phase angle of that distortion by up to 0.2
radians. This effect is frequency dependent, and is closely related to variables such as firing-delay angle and commutation-period duration.
Further noncharacteristic harmonics appear, associated with the other
components in the thyristor switching pattern. For the DC-side current
harmonic multiple k, harmonics (12n + 1) + k appear on the AC side in
positive sequence, and harmonics (12n - 1) + k appear on the AC side in
negative sequence. Once again, these higher-order noncharacteristic
harmonics could be expected to be at a level reduced by a factor of
approximately l/(12n + 1), as appropriate, from the terms described in
the first paragraph of this section.
The same secondary mechanism applies to these harmonics as well, and
the spectrum which appears as a result of commutation-period variation
reduces more quickly than that for DC voltage but not as quickly as the
characteristic harmonics. Thus, for the terms where n = 1, commutationperiod variation is as important a mechanism as the direct transfer, and
for higher orders it is more important. These harmonics can be expected

Static power conversion 47

to be present at levels of up to 20 % of the first-order terms described in


the first paragraph of this section.
In summary, the AC-side frequencies resulting from a DC-side harmonic multiple k, can be written
/ AC = ((12m 1) k)f0

for m e (0,1,2,3, ...)

(2.54)

in phase sequences as described.


Effect of switching instant variation
The effect of firing-angle or commutation-period variation consequential
to AC-voltage or DC-current distortion has been described above in a
limited way. The spectrum that appears from this sort of variation encompasses all the frequencies described above. Thus, if there is a negativesequence fundamental-frequency imbalance, even if no second-harmonic
current can flow on the DC side, there will still be third-harmonic positivesequence current on the AC side as a direct result of the variation of the
commutation period, and the full DC-side DC current. Further to this,
additional frequencies appear but in rapidly diminishing magnitudes. For
a harmonic multiple k on the DC side, if m is an integer, additional DCside harmonic multiples of 12n + mk can be expected, and on the AC
side, multiples (12n + 1) mk will appear. In most cases, these will be at
very low levels. Figure 2.25 shows the DC-side voltages and AC-side
currents that could be expected for a firing-angle order sinusoidal variation of 3 degrees, at a frequency 4.5 times the fundamental. The proliferation of frequencies, at low and high orders, should be noted. This
shows how the spectrum generated by switching-instant variation alone
diminishes only slowly with increasing order.
Harmonic transfer across the DC link
When the DC link interconnects separate power systems, either of the
same nominal (but in practice slightly different) frequency or of different
nominal frequencies, there will be a wide range of harmonic and nonharmonic frequency transfers. These can be divided into two groups:
1 Frequencies at terminal 1 caused by the characteristic DC-voltage
harmonics (\2nf2) and their consequential currents from terminal 2. These
are represented in the expression
/ACI = (12m 1)/! 12n/2

(2.55)

where m, n e (0,1,2,3 ... ) which can have any frequency including


frequencies below the fundamental.
The frequency-conversion schemes, and particularly the back-to-back
schemes such as Sakuma, represent the worst condition for noninteger

48

High voltage direct current transmission


0.06-.

11

0.050.0413

5 0.030.02-

23 25

0.010

10

15

1.1.

35

20
25
30
harmonic order
(a)

35

.1 1.

47 49
40

45

50

12

1024

O
Q

36

5-

48

, 35
1

10

15

'

20
25
30
harmonic order

35

..i .1
40

45

50

(b)

Figure 2.25 a DC-voltage harmonic spectrum, modulation frequency = 4.5 a)0


b AC-current harmonic spectrum, modulation frequency = 4.5 COQ
harmonic frequencies. In this case, with small smoothing reactors the DCside coupling is likely to be strong which means that the flow of harmonically-unrelated currents on the DC side can be large. In six-pulse operation
such schemes can produce considerable subharmonic content even under
perfect AC-system conditions. However, 12-pulse converters do not
produce subharmonic content under symmetrical and undistorted ACsystem conditions. These will produce interharmonic currents as defined
by eqn. 2.55, which for m= n= 1, f% = 60 and/! = 50 results in a 70 Hz AC
current and, through the AC-system impedance, voltage. The latter will

Static power conversion 49

beat with the main frequency f\ producing some 20 Hz flicker. However,


the levels of 70 Hz expected from this second-order effect will normally
be too small to be of consequence.
When the link interconnects two isolated AC systems of the same
nominal frequency such as the New Zealand scheme, and the two ends of
the DC link are different in frequency by a smaller frequency Af0, then
the characteristic harmonics are different by !2nAf0. If a DC-side voltage
at frequency I2n(fo + A/b) is generated by one converter, then at the other
converter this will be modulated down again by a characteristic frequency
in the thyristor switching pattern as per eqn. 2.55, i.e.
(12m l)/o 12n(/o
which on the AC side, among other frequencies, includes (for m = n)
/o 12nA/0
The latter will beat with the fundamental component at a frequency
\2nAf0, which at some values of n will allow flicker-inducing currents to
flow.
2 Frequencies caused in system 1 by unbalance or distortion in the
supply voltage of system 2.
Negative-sequence voltages at frequencies (k - l)f2 produce the following
noncharacteristic frequencies on the DC side
/ DC = (12n k)f2

(n= 0,1,2, ...)

(2.56)

Crossmodulation of these current components produces the following


frequencies in system 1
/ AC i = (12m 1)/! (12n k)f2

(2.57)

Let us first consider a frequency-conversion scheme with a sinusoidal but


negative-sequence unbalanced voltage in system 2, i.e. (k - 1) = 1 (and
therefore k - 2). Substituting m = n = 0 and k = 2 into eqn. 2.55 yields
currents (and therefore voltages) at frequencies
/ACI = /i 2/2

(2.58)

One of these frequencies (/i - 2/2) will beat with the fundamentalfrequency voltage of system 1 at a frequency
/ i + ( / i - 2 / 2 ) = 2(/i-/2>

(2.59)

which for a 50/60 Hz conversion scheme becomes 20 Hz. This is a flickerproducing frequency. This same frequency will be referred to generator
rotor-shaft torque at 20 Hz, which may excite mechanical resonances.
Again, this type of crossmodulation effect is most likely to happen in
back-to-back schemes owing to the stronger coupling between the two

50

High voltage direct current transmission

converters, although it is also possible with any HVDC scheme in the


presence of a suitable resonance.
Now, consider two AC systems of the same nominal (but slightly different) frequency.
Substituting m = n = 0 and k = 2 for fundamental frequency fo into eqn.
2.57, a current and resultant voltage (through the AC-system impedance)
of frequency
/o2(/o + A/o)
which leads to fo + 2A/o is induced on the AC side. This will either beat
with the fundamental frequency fo at frequency 2A/o, or produce generator/motor shaft torques at 2A/Q. This frequency is generally too low to
produce flicker but may induce mechanical oscillations.
Substituting m = n = 1 and k = 2 in eqn. 2.57 gives, among others, a
current (and thus voltage) at the frequency
and for f0 = 50 Hz and A/ = 1 Hz the resulting AC current (and thus
voltage) in system 1 is
13 x 50- 14 x 49 = 36 Hz
This distorting voltage will, therefore, beat with the fundamental producing 14 Hz flicker. However, the subharmonic levels expected from this
second-order effect will normally be too small to be of consequence.

2.12 Harmonic transfer generalisation


The main transfer relationships discussed in Sections 2.10 and 2.11 have
been collected together in Figure 2.26 for the case of a 12-pulse HVDC
link interconnecting two AC systems of frequencies f\ and / 2 , respectively.
These expressions are equally applicable to another pulse number, p, by
replacing the number 12 by p.
The exciting harmonic sources, surrounded by a circle, are multiples,
integer or noninteger, of the frequency in system 1.
(k\) is a current harmonic source, whereas (k\-]) and (k\+l) are voltage
harmonic sources.
The resulting harmonic orders in system 2 are related to the frequency
of system 2.
The DC column refers to the DC side of the link, the ACi+ and ACi"
columns represent the positive and negative sequences of system 1 and
AC2+ and AC2~ represent the positive and negative sequences of system 2,
respectively.

AC 2 +

AC 2

,,2::r:,,

ACf

DC

:::,":;:

::;,

::;.,

rTrh

(12n/c-|) Ulf2-^
(12/77-1)

k^lf2

12n 1 ^

(12/77-1) /C-,

(12/77 1) (12/7/C-|)f,//2 (12^+1)1(12/71^)^7/2

Figure 2.26

Harmonic transfers across a 12-pulse HVDC link; the encircled elements indicate harmonic sources and m, n
(1,2,3
...)

52

High voltage direct current transmission


\

(a)

(c)

Figure 2.27

Sustained unbalanced voltages on single converter bridge


a Three-phase voltages
b Direct voltage
c Direct current
d Three-phase currents

2.13 Quantified effects of system asymmetries


An unbalanced voltage source supplies imperfect references to the converter-control system, which in the early schemes modulated the pulse width
of the AC-current waveforms. This effect was eliminated by the use of an
equidistant firing control.11 However, the AC-voltage unbalance also
produces noncharacteristic modulation of the rectified voltage (and therefore DC-current) waveform; this problem is not completely eliminated by

Static power conversion 53


Table 2.1

Asymmetry test cases

Case 1

supply voltage asymmetry


2 phases = 100%
1 phase
= 99%

Case 2

transformer reactance asymmetry


5 phases = 20%
1 phase (star-delta) = 21%

Case 3

differences in transformer reactances


star-star
= 20%
star-delta = 21%

Case 4

differences in transformer ratios


star-star
= 100%
star-delta = 100.5%

Case 5

firing-angle asymmetry
5 valves
= 15
1 valve
= 15.2

the modern controllers. Figure 2.27 illustrates this effect for an unrealistic
case of unbalance.12 As predicted by Figure 2.26, the result is a high level
of second harmonic on the DC side and of third harmonic on the AC side.
Under more realistic conditions the levels of asymmetry and distortion to
be expected are relatively small and can be determined very precisely by
steady-state three-phase AC-DC conversion analysis.
The importance of various factors of asymmetry has been considered by
CIGRE SC1413 with reference to the 12-pulse configuration. The document considered the five different cases listed in Table 2.1; in practice, the
global effect results from a combination of all the asymmetries and the
quantities may be larger or smaller than those individually calculated.
Table 2.2 contains the levels of the first 25 harmonics resulting from
each of the five test cases.
The main conclusions of the CIGRE report were:

supply system asymmetry only produces triplen harmonics;


errors in the transformer ratios and differences in the transformer
reactances produce residual harmonics associated with six (rather than
12) pulse conversion;
differences between the three phases of the transformer reactances
produce harmonics of all orders;

54

High voltage direct current transmission

Table 2.2
Harmonic
order
1

Harmonic currents generated by ifae asymmetries of Table 2.


Reference:
ideal
100.000

Case 1

Case 2

100.0000

100.0000

0.0690

0.1300

Case 3
100.000

Case 4
100.0000

2
3
4
5

0.2080

0.414

0.0666

6
7

0.1840

0.366

0.5160

8
9
10
11

0.0430

0.0904

3.090

3.1000

3.0800

2.940

3.0800

1.490

1.5000

1.4900

1.390

1.4900

0.0225

0.0472
0.0683

0.135

0.0133

0.0617

0.122

0.0172

100.0000
0.0279
0.0268
0.0253
0.0234
0.0213
0.0189
0.0163
0.0136
0.0109

12
13
14

Case 5

15
16
17

3.0900
0.0056
1.4900
0.0009
0.0011
0.0028
0.0041
0.0032

18
19
20
21

0.0136

0.0339

0.0058
0.0062
0.0062

22
23
24

0.842

0.8510

0.8480

0.801

0.8430

0.0060
0.8450

25

0.597

0.6020

0.5960

0.542

0.5960

0.0048
0.5980

errors in the firing angles also produce harmonics of all orders, odd
and even; however, for practical (permitted) tolerances the values of
the noncharacteristic odd harmonics are much lower than those caused
by the other sources of asymmetry;
the principal causes of triplen harmonics are asymmetries in the transformer reactances;
the effects owing to differences between the phase reactances and in
the firing angles depend also on the transformer connection, being
more important in the absence of delta connections.

Static power conversion 55

2.14 References
1 UHLMANN, E.: 'Power transmission by direct current' (Springer-Verlag,
Berlin-Heidelberg, 1975), Section 1
2 ADAMSON, C, and HINGORANI, N.G.: 'High voltage direct current power
transmission' (Garraway Ltd, London, 1960), Chaps. 2 and 3
3 ARRILLAGA, J., ARNOLD, C.P., and HARKER, B.J.: 'Computer modelling of
electrical power systems' (John Wiley Ltd, London, 1983), Chap. 3
4 KIMBARK, E.W.: 'Direct current transmission' (Wiley Interscience, New York,
1971)
5 CIGRE Working Group 14.07: 'Guide for planning DC links terminating at AC
systems locations having low short-circuit capacities. Part 1: AC/DC interaction
phenomena'. Report 68, June 1992
6 ARRILLAGA, J., and WOOD, A.R.: 'Harmonic cross-modulation in HVDC
transmission'. International colloquium on HVDC and FACTS, Paper 6.7, Johannesbury, September, 1997
7 SWARTZ, M., BENNETT, W.R., and STEIN, S.: 'Communication systems and
techniques' (McGraw-Hill, 1966)
8 PERSSON, E.V.: 'Calculation of transfer functions in grid controlled converter
system', IEE Proc, May 1970, 117, (5), pp. 989-97
9 HU, L., and YACAMINI, R.: 'Harmonic transfer through converters and
HVDC links', IEEE Trans. Power Electron., July 1992, 7, (4), pp. 514-25
10 WOOD, A.R.: 'An analysis of non-ideal HVDC converter behaviour in the
frequency domain and a new control proposal'. PhD thesis, University of
Canterbury, New Zealand, 1993
11 AINSWORTH, J.D.: 'The phase-locked oscillator - a new control system for
controlled static convenors', IEEE Trans., 1968, PAS-87, (3), pp. 859-65
GIESNER, D.B., and ARRILLAGA, J.: 'Behaviour of h.v.d.c. link;
lanced a.c. fault conditions', Proc. IEE, 1972, 119, (2), pp. 209-15
13 CIGRE WG 14-03: 'AC harmonic filters and reactive compensation for HVDC
with particular reference to non-characteristic harmonics'. 1989

Chapter 3

Harmonic elimination

3.1 Introduction
Since the commutation reactance is low in relation to the DC smoothing
reactance, an HVDC converter acts, from the AC point of view, as a source
of harmonic currents (high internal impedance) and from the DC point of
view, as a source of harmonic voltage (low internal impedance). The
orders and levels of such harmonics have been discussed in Chapter 2.
Excessive levels of harmonic current must be prevented as they will
cause voltage distortion, extra losses and overheating, as well as interference with external services (e.g. telephone and railway signals).
The obvious place to eliminate the harmonics is the source itself. In
theory, characteristic harmonics could be eliminated either by some
complex converter configuration (which would be uneconomical), or by
the use of a series filter preventing the harmonics from arising (which
would upset the correct operation of the converter).
Therefore, accepting that the appearance of harmonics is an inherent
property of the static-conversion process, it will be necessary to reduce
their penetration into the AC and DC systems.
Any solution which increases the pulse number reduces the harmonic
orders penetrating into both sides of the converter and should be fully
exploited. Beyond the economic range of higher pulse configurations,
harmonic elimination will normally require the use of filters.
These are now considered separately.

3.2 Pulse number increase


The relationship between pulse number and harmonic order, discussed in
Chapter 2, indicates that the higher the pulse number, the higher the

Harmonic elimination 57
Converter
busbar

Figure 3.1

Converter
transformers

6-Pulse
bridges

D.C. line

12-pulse converter configuration

frequency of the lowest order harmonic produced. T h e use of increased


pulse numbers has the following disadvantages:
(a) Increased levels of lower order harmonics when converters are
temporarily out of service during maintenance.
(b) Increased number of transformers, both in service and spares.
(c) Increased complexity of transformer connections and the consequent
problems of insulation.
Moreover, as the harmonic order increases its amplitude decreases and it is
normally cheaper to eliminate it substantially by filtering.
With HVDC schemes only simple transformer connections are used.
This is due to the problem of insulating the transformers so that they withstand the alternating voltages combined with the high direct voltages. A
pulse number of 12 is easily obtained with star-star and star-delta transformer connections in parallel. This configuration, shown in Figure 3.1, is
now generally accepted for HVDC transmission.

3.3 Design of AC filters


33.1 Design criteria
An ideal in filter design is the elimination of all the detrimental effects
caused by waveform distortion, and particularly telephone interference.
However, this ideal criterion is unrealistic because of the difficulty of estimating in advance the harmonic flow throughout the AC system. It is also
uneconomic and, in the case of telephone interference, the problem can

58

High voltage direct current transmission

normally be solved more economically by taking some of the preventive


action in the telephone system itself.
A more practical solution is the reduction of harmonic voltage to an
acceptable level at the converter terminals. The flow of harmonic current
causes no special problem provided that the system harmonic impedance is
small and therefore a criterion based on harmonic voltage rather than
current is more convenient for filter design.
Typical specified factors to be taken into account in filter design are the
voltage distortion caused by individual harmonics (FJ, the total voltage
distortion defined as

and the telephone influence factor (TIF).


The TIF gives an approximation to the effect of the distorted voltage or
current waveform of a power line on telephone noise, without considering
the geometrical aspects of coupling. The harmonic frequencies that are
sensitive to the ear are given high weighting factors, since even if the
harmonic magnitudes are small, these harmonics may result in unacceptable telephone noise. The TIF is defined as
1/2

where
Kf= 5000(/1000) = 5/
Pf= C-message weighting
Vf= r.m.s. voltage of frequency / o n the power line
and
11/2

v=
3.3.2 Design factors
Two basic concepts in filter design are filter size and quality. The size of a
filter is defined as the reactive power that the filter supplies at fundamental frequency, which is substantially equal to the fundamental reactive
power supplied by the capacitors. The total size of all the branches of a
filter is determined by the reactive-power requirements of the converter
and by how much this requirement can be more economically supplied by

Harmonic elimination

59

the AC generators, extra shunt capacitors, synchronous condensers or


static VAR systems (SVS).
The quality of a filter (Q) expresses the sharpness of tuning and is therefore defined differently for tuned and high-pass filters. The sharpness of
tuning of a resonant filter branch increases with the ratio of its resonance
inductance or capacitance to its resistance, whereas in the case of a highpass filter, the sharpness increases in inverse proportion to that ratio.
The high Q or tuned, filter is sharply tuned to one or two of the lower
harmonic frequencies such as the fifth and seventh. The low Q or
damped, filter provides a low impedance over a broad band of frequencies
and is often used to eliminate the higher-order harmonics, e.g. 17th up. It
is normally referred to as a high-pass filter.

=i=c

Figure 3.2

a Single-tuned shunt filter circuit


b Single-tuned shunt filter impedance

60

High voltage direct current transmission

T
(a)

a
E

Figure 33

a Damped shunt filter circuit


b Damped shunt filter impedance against frequency

Figures 3.2 and 3.3 show typical circuit diagrams and characteristics of
the two types and Figure 3.4 illustrates their incorporation within the
conventional six-pulse HVDC converter configuration.
The diagram in Figure 3.5 indicates that the harmonic current generated by the converter divides between the shunt filters and the AC
network. To be efficient, the filter needs to be of much lower impedance
than the AC network and ideally must not resonate with the AC-network
impedance.
Therefore, the key to good filter design is a clear understanding of the
two components of the equivalent circuit, i.e.:
(a) The harmonic source (discussed in Chapter 2).
(b) The impedance of the AC network at harmonic frequencies.

Harmonic elimination
Converter
busbars
a b c A.C. circuit
breaker (CB)

61

6-pulse
bridge

Converter
transformer

5th, 7th, 11th, 13th harmonic and


high-pass (HP) filters per phase

Figure 3.4

Typical filter arrangement for a six-pulse converter configuration

Harmonic current into network

Harmonic current from source

- 0 Harmonic
-A source

Filter

A.C. network
i

Harmonic current into filter

Figure 3.5

Simplified circuit for the harmonic source, filter and AC-network


impedance

62

High voltage direct current transmission

3.3.3 Network impedance


Kimbark1 explains that, in general, the harmonic impedance of the AC
network at the point of filtering will exhibit the following characteristics:
(a) Alternation of resonance (low resistance) and antiresonance (high
resistance) as the frequency increases.
(b) Lower maximum impedance at high loads than at low loads.
(c) Great change in network impedance due to line outages.
(d) Resonances in the AC network are the rule rather than the exception.
(e) The harmonic impedances bear no relationship to the fundamentalfrequency short-circuit level.
(f) Loads provide some damping.
(g) This damping increases with frequency.
(h) On cable systems, the impedances to higher-order harmonics (15th to
25th) are lower than on overhead-line systems.
Figures 3.6 and 3.7 show typical impedance loci in the New Zealand
(South Island) 220 kV network for different frequencies and under different operating conditions.
500

"A
\

250 -

\11OO
T1101

ISA

><

7 650^
/
640>

-250

f 630
J
635

1660

680

-500

Figure 3.6

250

500

750

Harmonic impedances for a strong 220 kVAC system

Harmonic elimination

63

500

570

575

-500 L

250

500

750

1000

R(Q)

Figure 3. 7 Harmonic impedances for a weak 220 kV AC system

The impedance loci illustrate the difficulty of estimating harmonic


impedances, even under balanced conditions. Moreover, some harmonic
effects depend on the harmonic content of the three phases simultaneously, e.g. communication interference arising from harmonics in the
AC system is usually caused by the zero-sequence components of harmonic
currents. If converter-generated harmonics are the source of the interference then the zero-sequence currents arise solely because of the AC system
unbalance (i.e. no zero sequence currents are generated by the converter).
The high-voltage transmission lines are often untransposed and this
causes the electrical parameters to be different for each phase. Under
these conditions the sequence networks are mutually coupled. That is, a
current flow of one sequence induces voltages and currents to flow in the
other sequences.2
As the system harmonic admittances vary with network configuration
and load patterns, large amounts of data are generated.
Considering the large number of studies involved in filter design, it is
prohibitive to represent the whole system with the same degree of detail
for every possible operating condition. The detail of component representation depends on their relative position with respect to the harmonic
source as well as their size as compared with that of the harmonic source.
Any local plant components such as synchronous compensators, static
capacitors and inductors etc., will need to be explicitly represented.

64

High voltage direct current transmission

As the high-voltage transmission system has relatively low losses, it is also


necessary to consider the effect of plant components with large (electric)
separation from the harmonic source. It would thus be appropriate to
model accurately at least all the primary transmission network using the
equivalent Pi model.3 Moreover, owing to the standing wave effect on
lines and cables, a very small load connected via a line or cable can have a
dramatic influence on the system response at harmonic frequencies.
It is recommended to consider the loads on the secondary transmission
network in order to decide whether these should be modelled explicitly or
as an equivalent circuit. If these loads are placed directly on the secondary
side of the transformer their damping can be overestimated when using
equivalents which are too simple.4
The required size of network representation has to be evaluated in each
case. However, for the purpose of specification, it may be better to add a
reasonable margin rather than spending too much effort in getting very
accurate results, as the knowledge about parameters to be used is limited
and there may be a great deal of uncertainty about future configurations.
Radial parts of the system or neighbouring interconnected systems that
remain invariant when performing multiple case studies can be replaced
by frequency-dependent equivalent circuits.5
The measured or calculated harmonic impedances of a given ACnetwork configuration, viewed from the location of a harmonic source, are

Figure 3.8

Traditional boundary for AC-network impedance

Harmonic elimination

100

200

300

65

400

resistance, Q.

Figure 3.9

The annular-sector concept

often displayed using polar impedance/frequency loci, such as those


shown in Figures 3.6 and 3.7.
In the past, the impedance circle,6 shown in Figure 3.8, encompassing
all evaluated harmonic impedance loci, was used for all harmonics and the
maximum voltage distortion was derived by computer search techniques.
This approach led to unduly pessimistic filter designs, particularly at loworder harmonics. Besides, such an approach requires considerable computing and engineering time, often not available at the tendering stage. Two
practical alternatives currently used by the industry are annular sectors
and discrete polygons.
The annular-sector approach, illustrated in Figure 3.9, restricts the
geometric area applicable to each harmonic by setting up upper and lower
limits to the magnitude and phase of the harmonic impedance. Taking
into account all the relevant operating conditions, a comprehensive scatter
plot is produced for each harmonic on the impedance plane; all these
points are then encompassed by two circles and a sector and the resulting
values of Z\, Z2, 6x and 62 are tabulated.
This approach was used in the design of the filters attached to the
expansion of the New Zealand HVDC link and the information obtained
is shown in Table 3.1.
In the discrete polygon concept a distinction is made between low and
high harmonic orders. At the lower harmonics discrete points are obtained
for the different operating conditions as for the annular sector. Encompassing these points by a polygon results in a set of polygons for each
harmonic of interest.

Table 3.1
Harmonic
order "

2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50

Boundaries of Benmore 220 kVAC-system harmonic-impedance sectors


Zi
(ohms)
47.60
37.20
85.70
71.20
70.10
66.60
114.90
97.30
156.90
168.90
93.60
121.80
198.38
117.70
99.70
97.70
140.90
304.40
604.20
657.90
291.20
128.30
146.50
204.80
341.90
525.10
1319.60
460.90
97.00
431.30
449.80
372.60
333.00
130.50
238.20
368.50
209.80
172.20
169.90
143.10
149.60
145.00
105.90
73.30
74.90
136.50
102.20
55.10
43.10

z2

e2

(degrees)

(ohms)

(degrees)

87
77
78
61
58
78
78
63
63
55
48
79
55
43
51
75
78
82
76
52
0
86
23
71
59
52
46
-49
0
43
-14
5
-22
25
8
-13
0
-40
-73
6
-69
0
-53
-58
29
-55
-65
-60
-2

19.7
26.6
45.5
29.2
30.8
47.6
55.7
70.9
81.1
109.6
24.1
51.8
129.7
35.5
43.1
54.7
84.6
146.0
236.4
163.0
65.9
37.4
72.6
62.0
157.8
200.3
381.7
179.0
56.0
162.8
118.2
107.6
82.4
41.6
102.8
210.3
136.8
114.4
88.7
66.2
70.5
69.2
52.0
37.5
37.1
90.6
41.8
28.4
35.6

82
30
49
17
29
40
54
24
26
-38
-10
56
-19
-10
24
53
68
68
10
-69
-39
-35
0
58
51
1
-61
-72
-34
-34
-53
-54
-69
-5
-37
-65
-33
-52
-76
-12
-79
-70
-70
-70
11
-63
-71
-69
-22

Harmonic elimination

67

At higher harmonic frequencies, e.g. 14th to 49th, the scatter of the R +


jX values and hence the boundary of the encompassing polygons would
become increasingly large. Additionally, the impedances would begin to
extend into the capacitive region of the impedance plane. From detailed
information of the particular system involved, it is possible to decide on
the use of a realistic outer boundary with a single geometrical shape,
without introducing an unacceptable degree of pessimism into the filter
design studies.
A computer technique is then used to search each polygon in turn to
evaluate the system impedance which maximises voltage distortion at, or
current injection into, the point of common coupling.
This approach was used in the design of the AC harmonic filters on the
2000 MW Cross-Channel HVDC scheme.7 Individual search areas were
defined for harmonics 1-13 as shown in Figures 3.10, 3.11 and 3.12 for 24
defined operating conditions as follows:
bus reactor in and out;
filters at nearby busbars in and out;
single-circuit outages;
double-circuit outages;
minimum generating plant;
maximum generating plant.
In the case of the Cross-Channel scheme, a circle of centre 750 + jO Q and
10090807060-

2 5040302010-

Figure 3.10

10

20 30
R, Q.

40

50

Harmonic impedances for harmonic orders two to five

68

High voltage direct current transmission

180160140120X, a 100-

806040200

Figure 3.11

20

40

60

80 100 120

Harmonic impedances for harmonic orders six to nine

radius 750 Q (as shown in Figure 3.13) was considered sufficient to encompass all possible impedance loci derived from the 24 operating conditions
considered.
These figures indicate that the first harmonic to exhibit a resonance
condition is the 13th, whereas a generalised impedance circle approach
300-.

250-

13

200150X, Q

10050-

0-50-

40

80 120 160 200 240 280 320 360


R, 1

Figure 3.12

Harmonic impedances for harmonic orders 10 to 13

Harmonic elimination

69

14-49

\ 200 400 600 800 100012001

-800J

Figure 3.13

Harmonic impedances for harmonic order 13, and envelope of


harmonic impedance loci for harmonic orders 14 to 49

would have allowed even low-order harmonics (2nd, 3rd) to exhibit resonance.
In this particular application a further refinement was introduced.
Having chosen the particular worst (resonance) condition from the
polygon search areas, the remaining system impedances for harmonic
numbers two to 25 were chosen from a number of tables of harmonic
impedance, from the column which included the impedance closest to the
resonant impedance. For harmonic numbers greater than 25, the network
impedance was chosen from the impedance circle of Figure 3.13 to maximise the voltage distortion at each harmonic.
The calculated R jX values used in the polygons are the equivalent
Thevenin impedances of the entire network reduced to the Sellindge 400
kV busbar. These include the harmonic impedances of individual plant
items such as transmission lines, generators, transformers, etc.
It must be understood that the quantitative impedance plots used in this
scheme cannot be taken as typical and used as a default option in other
schemes. For instance, in cases of AC networks with long EHV or UHV
lines, the first resonant frequency may even occur below the second
harmonic.
The discrete polygon approach provides a realistic way of representing
the AC network for the purposes of AC-filter design. It avoids the pessimism of a generalised approach using a single search area, and offers a
technique which provides acceptably quick solution times for the highly
iterative task of filter design.
The cumulative effect of the existing nonlinearities will impose an extra
burden on the converter plant (filters in particular) at the bus under investigation. Traditionally the harmonic currents injected by the converter are
increased by some percentage (typically ten per cent) to take into account

70

High voltage direct current transmission

the presence of existing harmonics. When the AC system is represented by


impedances there is no other alternative.

3.3.4 Circuit modelling


Filter design is a complex subject that requires accurate modelling of the
behaviour of the harmonic source and of the AC-system configuration and
parameters.
The harmonic source itself, discussed in Chapter 2, is very dependent
on the AC-system conditions. At present, the converter harmonic currents
are calculated using three-phase terminal voltage sources with levels of
asymmetry and distortion specified from the experience of earlier
schemes.
Normally computer programmes are used in the calculation of the
distortion caused by individual harmonics, the total distortion caused by a
given set of harmonic current injections and the telephone influence
factor (TIF).
The programmes determine the most critical combination for each
harmonic and the corresponding distortion factor. The highest harmonic
voltage (i.e. maximum distortion) occurs when the parallel impedance of
the filters and AC network is maximum (i.e. with parallel resonance). It is,
however, unrealistic to expect that parallel resonance occurs at every
frequency and only two maximum single distortions are normally considered for the calculation of the total distortion and TIF (i.e. no resonance is
assumed for the remaining harmonic currents). Any shunt capacitors
present at the converter terminals must be included in the calculations.
However, a detailed assessment of the noncharacteristic harmonics
requires iterative computer simulation of the interaction between the ACside, converter and DC-side systems.8

3.3.5 Tuned filters9


The early DC schemes relied almost entirely on shunt-harmonic tuned
filters, each of which consisted of a series RLC circuit tuned to the
frequency of a low characteristic harmonic. The impedance of a singletuned filter is
Zf= R+j[coL- 1/icoQ]

(3.1)

which at the resonant frequency is a pure resistance R. The passband of


the filter is defined as being bounded by the frequencies at which the
filter's reactance and resistance are equal (i.e. when the impedance angle is
45 and its modulus y/2R).
Figure 3.2 shows a typical impedance curve for the single-tuned filter
and the following equations define the filter characteristics

Harmonic elimination 71
(3.2)
where fn is the resonant frequency
Xo = con L= \/(cDnC) = ^J(L/O = inductive or capacitive
reactance at resonance

(3.3)

The quality factor Qcan be expressed either as


Q= Xo/R

(3.4)

Q= a)n/PB (where PB is the passband in rad/s)

(3.5)

C=l/((onRQ

(3.6)

or

L=RQ/con
(3.7)
Often, two single-tuned filters are replaced by a double-tuned filter. This
has proved more economical because it uses only one common inductor
and the power loss at fundamental frequency is lower.
In practice, a filter is not always tuned exactly to the frequency of the
harmonic that it is intended to suppress, for the following reasons:
(a) Variations of the power-system frequency, which result in proportional
changes in the harmonic frequency.
(b) Changes in the inductance and capacitance of the filter owing to
ageing and temperature variations.
(c) The accuracy of the actual tuning is restricted by the discrete nature of
tuning steps.
The total detuning is
S = Aco/con = Af/fn + \ (AL/Ln + AC/Cj

(3.8)

where fn is the nominal system frequency, Ln is the nominal inductance


and Cn is the nominal capacitance. In terms of S and Q the filter impedance can be written as
Zf = R[l +jQS(2 + S)/(l +d)]

(3.9)

We are normally interested in small frequency deviations, i.e. 5 < 1 and


therefore
Z/ R(l+j2dQ)
Very often, admittances are used instead of impedance, i.e.

zf
where

(3.10)

72

High voltage direct current transmission


Gf= l/[R(l

+ 4Q2<52)]

Bf=-2Q5/[R(1

(3.12)

+ 4(g52)]

(3.13)

The harmonic voltage Vn on the AC terminals is then given by


(3.14)

Vn=iy(Yf+Yj
or

i2 \
'

'-

90s

ZVd

1 1 2 -i/2
' '

(3.15)

where Ysn = Gsn + jBsn is the AC-network admittance at the harmonic


frequency n.
The purpose of the filter is to minimise the harmonic voltage given by
eqn. 3.15. With reference to eqn. 3.15, the filter variables that can be
altered by the designer are the filter size and its quality factor. An
optimum filter size is decided by overall cost, and the effect of the filter
capacitance on filter cost is illustrated in Figure 3.14.
The filter quality is selected to achieve optimal filter operation, i.e. the
selected value should inject minimum harmonic current into the network
for the network condition that is assumed. A larger Q reduces the filter
losses and the harmonic voltage (when the filter is correctly tuned), but it
increases the risk of parallel resonance between the filters and the network.
However, some of the variables in eqn. 3.15 are not under the control of
the designer and have to be selected pessimistically; these are:
(a) The frequency deviation, 8, which is set to the greatest value that is
expected to exist.
(b) The network admittance, Ysn. If an accurate specification of system
impedance could be made, the filter design would be technically satisfactory and cost effective. This is not normally the case, however, and
various criteria used in the Ysn selection have been described in Section
3.3.3.

3.3.6 Self-tuned filters10


The detuning effects can be compensated for by continuous adjustment of
the capacitor or inductor (normally the latter), thus permitting the use of
high values of Q without having to increase the components ratings. Moreover, the resulting low value of the filter resistance increases the filter efficiency.
Self tuning consists of an on/off servo control which measures the
harmonic frequency reactive power in the filter and automatically adjusts
the main filter tuning to near resonance, i.e. with filter-arm harmonic
VARs within the prescribed limits.

Harmonic elimination

73

Cost

Capacitance

Figure 3.14

Relative costs of filter branch versus capacitance


a Cost of losses
b Cost of inductor
c Cost of capacitor
d Total'cost

The additional cost of providing the inductor variation has to be justified by savings in the capacitor cost and by improved performance. Whenever the detuning effects are small or when the required fundamental
frequency reactive power of the filter is high, self tuning is not normally
considered economic. However, a new self-tuning concept discussed in
Chapter 11 may provide an economical alternative.

3.3.7 High-pass filters


A high-pass damped filter presents a capacitive reactance at the fundamental frequency and a low, predominantly resistive, impedance over a wide
band of higher-order harmonics.

74

High voltage direct current transmission

II
(a)

Figure 3.15

ib)

High-pass filters
a Second-order filter
b Third-order filter

Since the sharpness of the high-pass filter tuning increases with the ratio
R/XQ, the Qof this filter normally refers to that ratio (i.e. the inverse of the
expression used for the resonant filters). Typical values of Qare between
0.5 and 5.
The second and third-order high-pass filters shown in Figures 3.15a and
b are extensively used in HVDC schemes. These are designed to reduce
the injection of harmonics above the 17th order into the AC system.
When designing such damping filters the Q is chosen to obtain the best
characteristic over the required frequency band, and there is no optimal Q
as with tuned filters. Because of their bandwidth there is no sensitivity to
fundamental-frequency deviation or component-value drift.

33.8 Example of recent filter arrangement


A typical filter/shunt capacitor configuration for a modern 1000 MW
converter station is illustrated in Figure 3.16a. The station consists of two
500 MW, 12-pulse groups, one on each side of earth. The filters are
divided into two groups, each including 11th and 13th single-tuned arms
and a high-pass arm (tuned to the 24th harmonic), as shown in Figure
3.166.

33.9 Type C damped filters


With the ratings of some HVDC links being of the same order as the
system short-circuit level, there is an increased probability of low-order
harmonic resonance between the system impedance and the filter capacitance. Series or parallel resonance will result depending on whether the

Harmonic elimination

75

240 kV

Groupi

Group 2

Group 3

11/13/HP
11/13/HP C-Bank
160 MVAr 160 MVAr 128 MVAr
(a)

Group 3:

Group 1/2:

C1=s

C2:

11th arm
56 MVAr

13th arm
40 MVAr

24th high pass


64 MVAr

shunt capacitors
64 MVAr 64 MVAr

(b)

Figure 3.16 Filter configuration for a modern 1000 MW converter station


a Single-line diagram
b Detail of filter and capacitor groups
low-harmonic source is within the AC system or converter stations, respectively.
By way of example, a high probability of third-harmonic resonance had
been expected on the British side of the 2000 MW Cross-Channel link. To
overcome the problem it was decided to design half the filters for a
minimum impedance at around the third harmonic frequency. However,
the use of damped filters for low-order harmonics involves large fundamental power loss in the damping resistor.
The power loss of conventional damped filters can be reduced by a
Type C filter,11 illustrated in Figure 3.17, where the resistor is bypassed by
a fundamental-frequency tuned arm (C2 - L). This circuit is more susceptible to frequency variations because of the fundamental-frequency tuning,
but exhibits much lower losses.

76

High voltage direct current transmission

Figure 3.17

C-type bandpass filter

3.3.10 Simplified filtering for 12-pulse converters


Conventional filter design, based on the use of separate tuned filters of the
series resonant type for the 11th and 13th harmonics and a high-pass filter
for the higher-order harmonics, provides a more effective reduction of
harmonics than is normally required. In conventional design the
minimum size of the filters is usually determined by the available
economic size of capacitor units and by the minimum amount of reactivepower compensation required at the converter's terminal.
Therefore, the filter design can be simplified, either by replacing the
tuned filters for harmonics 11 and 13 by a single filter of the damped
type, or by replacing all the individual filters by a single damped filter. In
the first case, the damped filter replacing the two tuned filters should be
tuned to about the 12th harmonic with a fairly high (2(20 to 50) and the
damped filter used for the higher harmonics has a much smaller Q (two to
four). In the second case, the single damped filter is also tuned to about
the 12th harmonic but a fairly low Q has to be chosen (two to six) to
achieve a sufficiently low impedance at higher harmonics.
The advantages of the damped filter are:
(a) The performance and loading are less sensitive to temperature,
system-frequency deviations and component tolerances.
(b) Because a wide spectrum of harmonic frequencies is filtered, the
considerable cost of subdividing the filter into several separate arms is
avoided; this also leads to a reduced site area.
(c) Maintenance is reduced.
(d) Uncharacteristic harmonics are also absorbed, subject to the filter Q
and centre frequency.
(e) The need to carry out tuning on site is reduced or eliminated.
(f) It is easier and cheaper to split the filter into smaller subgroups for
reactive power control; sharing the harmonic current between these
subgroups presents no problem.

Harmonic elimination

77

On the other hand, damped filters need to be bigger in terms of fundamental MVAR to achieve the same level of filtering performance as do
tuned resonant filters. The harmonic losses in tuned resonant filters are
usually lower than those in damped filters, although the opposite is true
for the fundamental-frequency losses.

3.4 DC-side filters


On the DC side of HVDC converters the voltage harmonics generate
harmonic currents with amplitudes which depend on known elements
such as the delay and extinction angles, the overlap angle and the impedance of DC circuits (i.e. smoothing reactors, damping circuits, surge capacitors and the line itself).
In contrast to the AC filters discussed above, the DC filters:

do not carry fundamental-frequency power and therefore have substantially lower losses;
do not need to provide reactive power, their only function being
harmonic mitigation;
have, in their main capacitor, to withstand the full pole to neutral DC
voltage.

Also, the harmonic impedances do not change with the operating conditions and it is therefore possible to use tuned filters with higher Q factors
and thus smaller capacitors and reactors.
The criteria to be met by the filters relate mainly to telephone interference from the DC line. The CCITT directives (published in 1989) require
that:

the induced voltage measured at subscribers' sets, for safety reasons,


should not normally exceed 60 volts r.m.s.;
the psophometric e.m.f induced should not exceed 1 mV, if psophometric weightings are used for noise measurements, or 20 dBrnc of
noise when using the C message weighting.

The performance requirements are assessed by the equivalent disturbing


current (ieq) and test-line methods.
In the first method, all harmonic currents on the DC line are reduced to
an equivalent disturbing current (Ieq) at a single frequency, under the
assumption that this current should cause the same interference effect on
the telephone lines, i.e.
/ n=m

kq = J E (/ x Hn x Cnf

78

High voltage direct current transmission

where
= harmonic order: the upper limit usually considered ranges
between 2.5 and 5 kHz
Cn = normalised weighting factor at harmonic n, referred to 800 Hz
or 1000 Hz for the psophometric and C message weights,
respectively
Hn = frequency-dependent factor, taking into account variations (if
significant) of the mutual impedance among the HVDC line
and the telephone lines, of shielding and of telephone circuit
balance

In the second method, the performance requirements are specified as a


longitudinal induced voltage, properly weighted as indicated above, on a
hypothetical sample telephone line, which is assumed to be parallel to the
main line, 1 km long and 1 km far away; the ground resistivity to be used
in the calculation is to be specified.
Comprehensive studies must be carried out at the planning stage, in
order to decide whether to use filters or reroute parts of the transmission
lines away from telephone systems.

/12 (A)

i i

0.18-

0.14-

0.10

0.06-

100

Figure 3.18

200

300

400

500
Distance (miles)

12th-harmonic current profile along a DC transmission line

Harmonic elimination 79
0.9 H

200 kV

0.4-0.6 H

ololo

San Dalmazio
(Italian mainland)

400 KV

L6

2.5 H P
r 07
> 1000
ohms

0.7 j
mH'

280
mH

63
ohms
!
:30

Celilo
Sylmar: same as Celilo except neutral
bus surge capacitor 5 11F

(a) Sardinia

(b) Pacific intertie

0.83 H

0.5H

533 kV

0.21
HF--

0.5 J .

12

244

0.3-- 0.035-L

mH 9

8.8 I

110

Ohms <
Dorsey

0.29 i
F ^

<;

ohms <

7c

-V

16-

X 0.018

- f

^F

Songo (Mozambique) Same as Apollo


(d) Cabora Bassa

0.06 H

MF

Apollo (South Africa)

f^F X

(c) Nelson River Bipole 1

ohms> ohmsS ohmsS

3.25 4:

Radisson: Same as Dorsey except neutral


bus surge capacitor 55 jiF

12&24

0.75 H

0.4 JJ

0.9^

A]
A]

120
mH

123
mH

1000
fohms

20
mH

ohms ,

Arrowhead (Duluth)
Center: Same as Arrowhead

(e) Square Butte

Figure 3.19

Dorsey (Winnipeg)
Henday: Same as Dorsey

{f) Nelson River Bipole 2

DC filter circuits of various HVDC schemes

The assessment of interference levels requires detailed information


about the harmonic voltage and current profiles along the HVDC line;
electromagnetic induction from harmonic currents is normally the main
problem. Moreover, since both ends of the link contribute to the disturbance, it is necessary to obtain the profiles from each end and add their
effects.

80

High voltage direct current transmission

Computer programmes are used for this purpose8'12 which calculate the
self impedance, mutual impedance and capacitive matrices for any
frequency and line.
As an estimated worst case, the results are then combined by adding the
root of the sum of squares (of the peak profile) derived from each end of
the link. At each harmonic a profile along a single equivalent conductor is
obtained. This is determined from the vector addition of the harmonic
current values from each of the DC conductors and the DC overhead
ground wire. The equivalent conductor is assumed to be located along the
centre of the DC lines. A typical 12-harmonic current profile, with filters
included, is illustrated in Figure 3.18.
Typical types and locations of DC filters in several existing schemes are
shown in Figure 3.19.

3.5 Active filters


3.5.1 AC-side active cancellation
The complexity and high cost of conventional filters, together with their
inability to correctly reduce resonances at noncharacteristic harmonics, has
stimulated the development of power-electronic compensation techniques,
generally described as active filters.
The prospective application of active filtering to HVDC converters was
suggested as early as 19711S for the elimination of harmonic currents by
magnetic-flux compensation, as illustrated in Figure 3.20. A current transformer is used to detect the harmonic components coming from the
Secondary current /2

A A

'

Signal processing
circuti

ioh

r-4"
F3

.J

I
I

I
Figure 3.20

Basic configuration of harmonic-current elimination method by flux


compensation

Harmonic elimination

81

nonlinear load. These are fed, through an amplifier, into the tertiary
winding of a transformer in such a manner as to cause cancellation of the
harmonic currents concerned. The main area of concern with this system
is the coupling of the output of the amplifier to the tertiary winding in
such a way that the fundamental current flow does not damage the amplifier. A quaternary winding and filter are used, as shown in Figure 3.20, to
reduce the fundamental current in the amplifier output.
Another important difficulty is the transfer of the amplified compensating current waveform from the low to high-voltage side of the transformer.
A small-scale prototype was developed and discussed in Reference 13; its
extension to a 300 MW converter was alleged to require a 750 kW amplifier. Thus the replacement of the lower-order characteristic harmonics
does not appear to be a viable proposition at the moment.

3.5.2 DC-side active cancellation


As indicated above, the complete replacement of present passive-filter
schemes is not envisaged. However, some of the main difficulties encountered on the AC side do not apply on the DC side and an interesting
hybrid solution has already been implemented.14'15
The principle of the passive/active filter concept is that the harmonics
in the DC-line current are measured, and a controller reproduces a
current in counter phase with the disturbance. This signal is then amplified in a high-power amplifier and finally fed into the neutral bus end of
the DC filter by a transformer.
The main components of the active filter are: a harmonic current transformer (HCT), a computerised controller, a pulse-width modulated
(PWM) power amplifier, and a high-frequency transformer together with
a transient overvoltage protection. These are shown in Figure 3.21.
The harmonic current transformers consist of a Rogowski coil and electronic circuits which convert the measurement to optical information that
is fed to ground potential by an optical fibre.
The controller is realised in the same hardware and software environment as the control equipment for the HVDC pole. A digital signal processor handles the fast controller mathematics in conjunction with a singleboard computer.
The power amplifier consists of a large number of digitally-modulated
transistors (using the pulse width modulator principle) working in parallel
in switched bridges. The output ranges up to 4 kHz at a peak voltage of
about 300 V and a power of 100 kVA.
The high-frequency transformer is of a dry isolated type and has a ratio
1:20, with the low-voltage side connected to the amplifier.
The transient overvoltage protection is realised by two antiparallel thyristors of the same type as the thyristors used in the HVDC valve. The
protection is triggered by overvoltage or overcurrent.

82

High voltage direct current transmission


smoothing reactor

arrester

protection
circuit

neutral bus

Figure 3.21

Active DC filter

3.6 References
1 KIMBARK, E.W.: 'Direct current transmission' (Wiley Interscience, New York,
1971)
2 KUUSSAARI, M., and PESONEN, A.J.: 'Measured power line harmonic
currents and induced telephone noise interference with special reference to
statistical approach'. CIGRE, Paris, 1976, paper 36-05
3 ARRILLAGA, J., BRADLEY, D., and BODGER, P.S.: 'Power system harmonics' (John Wiley 8c Sons, Chichester, 1985)
4 ARRILLAGA, J., et ah 'A.c. system modelling for a.c. filter design, an overview
of impedance modelling', Electra, 1996, (164)
5 WATSON, N.R.: 'Frequency-dependent a.c. system equivalents for harmonic
studies and transient converter simulation'. PhD thesis, University of Canterbury, New Zealand, 1987
6 LAURENT, P.G., GARY, C., and CLADE, J.: 'D.c. interconnection between
France and Great Britain by submarine cables'. CIGRE, Part III, Paper 331,
1962
7 BAKER, W.P.: 'Measured impedances of power systems'. International conference on Harmonics in power systems, UMIST, Manchester, England, 1981
8 ARRILLAGA, J., SMITH, B.C., WATSON, N.R., and WOOD, A.R.: 'Power
system harmonic analysis' (J. Wiley 8c Sons, Chichester, September 1997)
9 CLARKE, CD., and JOHANSON-BROWN, M.J.: 'The application of selftuned harmonic filters to h.v.d.c. convenors'. IEE Conference on High voltage
DC transmission, Publication 22, 1966, pp.275-76

Harmonic elimination

83

10 STANLEY, C.H., PRICE, I J., and BREWER, G.L.: 'Design and performance
of a.c. filters for 12-pulse n.v.d.c. schemes', in IEE Conf. Publ. 154 on 'Power
electronics-power semiconductors and their applications', 1977
11 OUELETTE, K.R., and LEWIS, D.W.: 'Harmonic interference from d.c.
lines'. Manitoba Power Conference EHV-DC, Winnipeg, 1971, pp.543-78
12 HARRISON, R.E., and KRISHNAYYA, P.C.S.: 'System considerations in the
application of d.c. filters for h.v.d.c. transmission'. CIGRE, Paris 1978, Paper
14-09
13 SASAKI, H., and MACHIDA, T.: 'A new method to eliminate AC harmonic
currents by magnetic flux compensation - considerations on basic design',
IEEE Trans. Power Appar. Syst. 1971, PAS-90, (5), pp.2009-19
14 JANSSON, B., et at 'New features of the Skagerrak 3 HVDC intertie', International Colloquium on HVDC and FACTS, Wellington (NZ), 1993, paper 6.3
15 NYMAN, A., EKENSTIERNA, B., and WALDHAUER, H.: 'The Baltic cable
HVDC project'. CIGRE, 1996, paper 14-105

Chapter 4

HVDC system development

4,1 Basic DC system configurations


HVDC transmission systems can be configured in many ways to take
into account cost, flexibility and operational requirements. Five basic
configurations are illustrated in Figure 4.1 in order of increasing complexity.1
The simplest is the back-to-back interconnection (Figure 4.1a) in which
the two converters are on the same site, as there is no transmission line.
They can be designed more economically (a 15 % to 20 % saving in
converter plant is quoted) than those of long distance schemes, and are
designed for relatively low voltages (50 to 150 kV), although still use the
highest current rating of the single thyristors. The two units are identical
and each can be used on the rectification or inversion modes as ordered
by the system control.
In the monopolar link (Figure 4.16), the two converter stations are
joined by a single conductor line, and earth (or the sea) is used as the
return conductor, requiring two electrodes capable of carrying the full
current.
Next, and the most common configuration, is the bipolar link shown in
Figure 4.1c, which consists of two monopolar systems combined, one at
positive and one at negative polarity with respect to ground (middle). Each
monopolar side can operate on its own with ground return; but if the two
poles have equal current, they cancel each other's ground current to practically zero. In such cases, the ground path is used for limited periods in
an emergency, when one pole is temporarily out of service.
The remaining two cases are multiterminal. In the parallel-connected
three-terminal configuration (Figure 4.Id) converters 1 and 2 operate as
rectifiers, and converter 3 operates as an inverter. Alternatively with an
automatic reversal of voltage, along with firing-angle control, converters 1
and 2 operate as inverters and 3 as a rectifier. Other combinations of recti-

HVDC system development 85


converter 2

converter 1

(a)

(c)

ela-

L*

\\< i a -

A
A

.,. j o

A
A

converter 1

converter 2

y \
y

converter 3

-cm

-Q
A
Figure 4.1

- x i xi

Five basic configurations

fiers and inverters can be created by first reversing the connections of a


terminal with mechanical switches.
The series connection of Figure 4.1^, although still unused, is an attractive proposition for small taps (using some five to ten per cent of the total
line power) because of the comparatively high cost of the full voltage
parallel tapping alternative.

86

High voltage direct current transmission

4.2 Mercury-arc schemes


The fast development of the graded electrode mercury-arc valve in the late
1940s paved the way for a first commercial application of an HVDC technology in 1954. This scheme was immediately followed by others of
increased ratings.
The success of mercury-arc-based HVDC technology can best be illustrated with reference to some of the early schemes. A summary of the
main factors justifying their existence is now given based on the technical
background of the previous and subsequent chapters.
Sweden-Gotland link (1954)2
The first in commercial operation, the Gotland DC link was initially
designed as a 20 MW, 100 kV monopolar submarine link to supply power
to the island of Gotland. It was justified economically as an alternative to
the establishment of extra thermal generation on the island. The distance
(96 km) was too large for transmission by AC cable.
English Channel (1961)
The original 160 MW, 100 kV DC submarine link interconnecting
Great Britain and France was built to take advantage of peak load and
generation diversity between the two countries. It was marginally justifiable in terms of distance (only 64 km), particularly as the North Sea
cannot be used as a return path because of interference with ships'
magnetic compasses. However, the relatively small power tie would have
been difficult to control with an AC cable interconnection, as the British
power system has no automatic load-frequency control. The provision of
such control throughout the British grid would have been far more expensive than the alternative DC scheme.
Volgograd-Donbass (1962-65)
This was the first overhead transmission scheme and was built as a reinforcement of an existing AC weak transmission system. The transmission
distance is 470 km and the link rating 720 MW at 400 kV and 900A.
The scheme used an alternative design of mercury-arc valve with a single
(air-cooled) anode and an oil-cooled tank. The bipolar scheme uses four
bridges per pole and two valves in series each for 50 kV and 900A.
New Zealand link (1965)
A 600 MW, + 250 kV DC partly overhead (570 km) and partly submarine
(40 km) link was installed to transfer to the North Island the surplus of

HVDC system development 87

energy from hydro resources in the South Island. Given the relatively
short submarine interconnection, the scheme was feasible using conventional AC technology. However, the depth of the Cook Strait would have
restricted the use of AC cable to relatively low voltages and thus resulted
in many parallel circuits. Under such conditions, and given the reasonably
large overhead transmission, the DC solution was found to be more
economical.
Konti-Skan (1965)3
This is a monopolar 250 MW, 250 kV DC (using earth return), partly overhead, partly submarine interconnection between Sweden and Denmark.
The cable section, although sufficiently long (87 km), is divided into two
parts and an AC interconnection would have been feasible. In fact, the
economic justification for DC was very marginal, but it offered the advantage of future further extension (by means of a second pole) to 500 MW, a
power rating beyond the stability limit of an alternative AC interconnection between the two countries.
Sakuma interconnection (1965)
The first zero-distance frequency interconnector, this link was built to
interchange up to 300 MW (at 125 kV) in either direction between the
50 and 60 Hz power systems in Japan in case of AC-network disturbances
or lack of energy in either of the systems. In this case, there was no practical alternative to the AC-DC-AC converting station.
Sardinia-Italy (mainland) (1967)
The monopolar 200 MW at 200 kV DC link (with earth and sea return)
was mainly installed to provide frequency support to the Sardinian AC
network by power/frequency control of the DC link.
Transmission alternates between submarine cable and overhead line as it
passes through Sardinia, the sea, Corsica, the sea and the Italian mainland.
The total cable distance of 121 km justified the use of DC.
This scheme has now been replaced by a three-terminal 300 MW link
with a 50 MW tapping station in Corsica which went into operation in
1987.
Vancouver Island (1968)
In parallel with two existing 132 kV AC cables, this 74 km DC link was
built mainly to reinforce the supply of energy to the Island. It contains
overhead and submarine cable lengths. A 312 MW pole was built between

88

High voltage direct current transmission

1968 and 1969 with mercury-arc valves. A second pole, of equal power,
was added in 1979 using thyristor technology.
Pacific intertie (1970)
Unlike the other schemes discussed above, the Pacific DC intertie runs in
parallel with two other AC circuits at 500 kV (60 Hz).
This DC scheme was partly justified in terms of distance (1372 km of
overhead transmission) and partly by the availability of additional control
features which were expected to help damp out the power oscillations
experienced in the parallel AC transmission system already in existence. It
consisted of a 1440 MW at 400 kV bidirectional link to take advantage
of the seasonal diversity in load and generation between the northwest and
southwest areas of the United States. The original link was replaced in
1986 by a 1920 MW link using thyristor valves.
Kingsnorth scheme (1974)
The main purpose of the 640 MW (at + 266 kV) three-terminal Kingsnorth-Beddington-Willesden DC scheme in the UK was the reinforcement of an existing AC system in areas of high load density without
increasing the short-circuit level. It would have been difficult to justify the
use of DC purely in terms of distance, as the two underground transmission distances were marginal (59 km and 82 km, respectively). This link is
now retired from service.
Nelson River Bipole 1 (19 73- 77)
HVDC transmission was selected to transport bulk power from distant
hydro generation to the loading centre (in this case Winnipeg). The
distance involved is 895 km with an initial nominal capacity of 1620 MW
(at + 450 kV) and an ultimate capacity of about 6500 MW. The first stage
of the Nelson River scheme used the highest power mercury-arc valves
ever developed (i.e. 150 kV, 1800A); this stage also made history by being
the last case of successful competition for mercury-arc over the solid-state
technologies.

4.3 Evolution of the modern solid-state HVDC scheme


In spite of the successful operation of the mercury-arc schemes, the incidence of arc backs, considerable maintenance and voltage limitations
encouraged the development of the solid-state technology.

HVDC system development 89

The impact of the thyristor technology was not just because of valve cost
(which initially was much the same as the cost of the mercury-arc valve)
but also because of the overall economic effect of station plant and layout.
Even before the commissioning of the last two mercury-arc schemes
(Nelson River and Kingsnorth) the small experience gained with thyristor
valves was sufficient to discourage any further development of the
mercury-arc technology.
The technical reasons in favour of solid-state HVDC transmission are
basically the same as those given for the mercury-arc schemes discussed
above. However, a brief look at the solid-state schemes in existence will
help us to understand the great progress made with the new technology.
Eel River (1972)
If the Nelson River made history for being the last of the old technology,
another Canadian River (the Eel) started the new technology. To be fair,
solid-state additions to previously-existing schemes had already been operating in Sweden and the UK. However, the Eel River scheme (320 MW at
2 x 80 kV) was the first large HVDC project specifically designed for
thyristor valves. It is a back-to-back asynchronous interconnection of two
systems (New Brunswick and Hydro-Quebec) of nominally equal
frequency but drifting in relation to each other.
Each station consists of two converter bridges with a total of 4800 aircooled thyristors placed in changeable 40-unit modules. It was necessary to
parallel four thyristor units per converter arm.
Cahora-Bassa (1978)4
A decision was taken in 1969 to use solid-state valves to transmit 1920 MW
at + 533 kV between the Zambia River in Mozambique and Johannesburg,
separated by 1360 km. This was the first scheme, whether AC or DC, in
the megavoltage range (between poles). It was also the first case of international bulk power transmission. The first stage of the scheme started
commercial operation in 1978. It used over 36000 thyristors, oil cooled
and oil insulated in an outdoor-valve layout involving four bridges per
pole at each end of the link. It is interesting to consider that switching,
which used to be the curse of power-transmission circuits, had become the
basis of power controllability. In fact, the normal operation of the
complete Cahora-Bassa scheme required of the order of 4 x 106 ON and
OFF individual switchings per second.
Skagerrak (1976)5
A further step from the Inga-Shaba project led to this scheme where
the four individual valves (per phase) of a 12-pulse converter are

90

High voltage direct current transmission

combined into a single vertical quadruple valve, also of modular design


for easy maintenance. This arrangement has become standard in later
schemes.
The Skagerrak scheme provides 500 MW (1000 in the second state) at
250 kV of interconnection between Norway and Denmark. It includes 127
km of cable transmission and 113 km of overhead line. DC was justified
partly in terms of distance and partly by the different capacities of the
Norwegian and European systems.
Square Butte (1977)
The cost of a generating plant in North Dakota and electrical-energy transport by a 750 km HVDC link between North Dakota and Minnesota was
found to be a cheaper alternative to the cost of coal transportation for the
use of local generation in Minnesota. Moreover, the use of DC offered a
solution to the system-stabilisation problem of an alternative AC transmission scheme.
The thyristor valves of the Eel River scheme had been shown to possess
considerably higher availability than the rest of the circuitry needed for
six-pulse operation. This led to the selection of a 12-pulse converter per
pole with a rating of 500 MW 250 kV.
CU(1979)
Again, the DC solution was justified in this scheme for base-load bulk
power transmission from the Coal Creek lignite mine generating plant to
the Dickinson area in the USA. It is a 702 km 1000 MW 400 kV scheme
and uses the, by then standard, air-cooled and air-insulated quadruple
valves.
Hokkaido-Honshu (1979)6
Despite the short transmission distance (167 km) the use of HVDC was
justified by the 43 km submarine crossing and the need for an effective
control of frequency variations as well as mutual emergency power assistance during fault occurrences.
It is a bipolar 125 kV scheme and was developed in three stages, i.e.
150 MW (1979), 300 MW (1980) and 600 MW (1993). In the final stage,
light-triggered water-cooled thyristor valves of Japanese manufacture were
used.
Inga-Shaba (1982)
The rapid progress made in thyristor architecture became apparent in this
project, approved by the Zaire Government in 1974 to exploit the Zaire
River hydro potential. In the first stage, the DC system transmits 560 MW

HVDC system development 91

over 1700 km at 500 kV. The scheme includes double valves of a


modular design and air insulated for indoor installation.
Gotland II (1983-87)
The mercury-arc pioneering link (Gotland I), described in Section 4.2, was
replaced between 1983 and 1987 by a 320 MW thyristor scheme.
Nelson River Bipole 2 (1985)7
The final stage of the Nelson River Bipole 2 scheme is rated 2000 MW at
+ 500 kV. The 12-pulse group quadruple valves in this scheme are water
cooled. With the experience of this scheme, water became the most acceptable coolant in the later schemes.
Itaipu scheme (1986)
The culmination of solid-state technology is the Itaipu HVDC scheme
which involves some 6000 MW of DC-power transmission between Paraguay and Brazil over a distance of 796 km. The total transmission is
12600 MW, half of which uses 800 kV AC. The compromise reached was
based on the different frequencies of the two countries (50 and 60 Hz)
and the agreement for each country to have the right to use half of the
power generated at Itaipu. Consequently, half of the generators are
designed for 50 Hz and half for 60 Hz. The scheme uses two bipolar DC
lines at 600 kV, each consisting of four (two per pole) 12-pulse watercooled valve converters and the power rating of the individual valves is
788 MW.
Des Cantons-Comerford (1986)8
New England (USA) contracted to purchase from Hydro-Quebec three
million MW hours per year for 11 years, and a DC bipolar 690 MW at
450 kV link was built. However, the final plan was a 1500 km long
multiterminal scheme with further terminals at Sandy Pond (1800 MW),
Nicolet (2138 MW) and Radisson (2250 MW). The multiterminal scheme
(Phase II of the project) was commissioned in several stages owing to its
complexity, between 1990 and 1992. However, the Phase I DC terminals,
Des Cantons and Comerford were not integrated into the multiterminal
DC network because of anticipated performance problems.
Cross-Channel link (1986)
Almost 20 years after the commissioning of the first cross-channel link
between England and France, an agreement to proceed with the construe-

92

High voltage direct current transmission

tion of a 2000 MW link was issued by EDF and CEGB in 1981 to exploit
the different daily load curves and generating plant mix of the French
and British systems. The converter stations are linked by eight cables
operating at + 270 kV. The scheme consists of two 1000 MW bipoles
between Sellindge in SE England and Bonningues-les-Calais in Northern
France.
The main novel feature of the scheme is the use of different converter
equipment (with valves and control of different manufacturers) at each
end of the link. On the English side the converter plant includes three
high-speed static compensators of the saturated-reactor type to provide, in
conjunction with switched capacitor banks, the required reactive power
control capability for load and load rejection conditions.
Sileru-Barsoor (1989)
The Indian-built 196 km long 100 MW 100 kV six-pulse monopolar link
was initially developed with the purpose of gaining design and operational
experience on HVDC transmission. Two further stages were planned, first
a 12-pulse extension and then a second extension to make the link bipolar
with a final rating of 400 MW 200 kV.
Fenno-Skan (1989)9
This is a 200 km sea-cable interconnection to exchange up to 500 MW at
400 kV between Finland and Sweden. It is a monopolar link with sea
return. Its main purpose is to use the available peak-power reserve of both
countries, transfer power from North Sweden to Southern Sweden via the
Finnish network and increase the power-transmission capability on the AC
interconnections owing to higher stability limits.
Gesha (1991)10
Long-distance bulk-power transmission justified the use of DC for the
Gezhouba-Shanghai interconnection, a 1200 MW, 500 kV scheme and
the first HVDC in China.
New Zealand hybrid (1992)11
The expanded configuration of the Cook Strait link involved combining
the earlier four mercury-arc bridges into two 12-pulse groups in parallel in
one pole (at +270 kV) and adding a new 12-pulse thyristor valve group on
the other pole (at -350 kV). The total power capability became 1240 MW
and required new submarine cables and substantial reinforcement of the
overhead transmission insulation.

HVDC system development 93


Rihand-Delhi

(1992)12

This scheme was built to transfer bulk power from the Rihand-Singrauli
thermal power-generating complex to the load centre around Delhi. Over
a distance of 814 km a nominal power of 1500 MW is transferred at
500 kV.
Baltic cable (1994)13
With a 255 km cable length, this 600 MW, 450 kV link between Sweden
and Germany offered a considerable leap in cable technology. The scheme
has also provided experience in the use of deep-hole electrodes and active
DC filters.
Kontek (1995)14
This link rated at 600 MW and 400 kV utilises the highest rated cable in
commercial service. It has both sea and land-cable sections (totalling 120
km) and is used for economy energy transfers between Denmark and
Germany.

4.3.1 Frequency conversion


As explained earlier, the Sakuma scheme in Japan (1965) pioneered the
use of HVDC for the interconnection of systems of different nominal
frequencies (50 and 60 Hz); the early scheme using mercury valves was
decommissioned in 1993 and was then replaced by thyristor valves of the
same capacity.
A second frequency-conversion scheme, the Shin-Shinano link, 5 was
built in Japan in 1977 as a back-to-back interconnector. In 1992, the original solid-state valves were replaced by light-triggered, air-insulated, watercooled thyristor valves.16 A further 300 MW frequency converter, the
Higashi-Shimizu link, is also planned to be completed by 1998 in Japan.
The Japanese experience has also led to two more modest monopolar
interconnections in South America, both rated at 50 MW. The Acaray
link,1 commissioned in 1981, was built to supply energy to Paraguay
during periods of shortage of water and export to Brazil the surplus hydro
energy. It is also used to stabilise the Paraguay frequency by adding
frequency control to the DC link. The Uruguaiana link18 (between Argentina and Brazil), commissioned in 1986, was built to permit energy
exchange between the two countries for system operation optimisation.

43.2 Asynchronous back-to-back interconnections


As early as 1972 the, until then, mercury-arc based Swedish-British HVDC
transmission monopoly was surprisingly overtaken by German and North

94

High voltage direct current transmission

American suppliers with their solid-state technology. As explained in


Section 1.6, the Eel River was the first scheme developed entirely with the
new technology. Apart from the frequency-conversion links (described in
Section 4.3.1), the Eel River was also the first asynchronous back-to-back
interconnection between two systems of the same nominal frequency. A
common denominator of the back-to-back philosophy was the elimination
of the synchronous tie between two interconnected AC systems and the
provision of a flexible and more reliable interchange of energy.
North American schemes
The unusual map of North America shown in Figure 4.2 highlights four
blocks of AC transmission connected together by back-to-back DC links. It
shows the significant progress made with HVDC back-to-back, particularly
in the decade of the 1980s.
There are now 20 links in North America and half of these are back-toback connections.

Figure 4.2

. ^ . *> back-back HVDC tie


1

1 HVDC line

Unusual map of North America highlighting four blocks of AC transmission connected together by DC links

HVDC system development 95

In chronological order, the following back-to-back links have already


been commissioned in North America:
(i)

Eel River (Canada, 1972) - 100 MW 100 kV (described in section


1.6).
(ii)
David Hamil (USA, 1977) - a monopole 100 MW 50 kV link
installed to provide a stable power-transfer path between the eastern
and western parts of the 230 kV systems at Stegall, Nebraska.
(iii) Eddy Country (USA, 1983) - a monopolar 82 kV 200 MW link
between the eastern and western electric grids in Eddy Country
(New Mexico) installed to transport energy from the South Western
Public Service Co. to El Paso Electric Co. and Texas New Mexico
Power.
(iv) Chateauguay (Canada, 1984)21 - a double monopolar scheme rated
100 MW at 140 kV to transmit surplus power from the hydro plants
in northern Quebec.
(v)
Highgate (USA, 1985) - a monopolar 56 kV, 200 MW link between
Hydro Quebec and Vermont Electric Power Co., to compensate for
the loss of power caused by the shutdown of the Vermont Yankee
Nuclear plant.
(vi) Madawaska (Canada, 1985) - a monopolar 130.5 kV, 350 MW link
between Hydro Quebec and New Brunswick to supply local load in
New Brunswick.
(vii) Miles City (USA, 1985)22 - a monopolar 82 kV, 200 MW link
between East and West USA-Canadian transmission systems to
restore power-transfer capability across the federal system. Also to
transfer cheaper energy to the loads on the other side of the boundary.
(viii) Oklaunion (USA, 1985)23 - this monopolar 82 kV, 200 MW unit
connects the Southpower pool to the Electric Reliability Council
Commission of Texas.
(ix) Virginia Smith (USA, 1987)24 - a monopolar 50 kV, 200 MW unit
installed at Sidney (Nebraska) to interchange energy between the
East and West United State-Canadian networks.
(x)
McNeill (Canada, 1989) - a monopolar 42 kV, 150 MW link to trade
energy between Alberta Power and Saskatchewan Power.
European schemes

(xi)
(xii)

Vyborg (CIS - Finland, 1981)19 - three independent, parallel backto-back units totalling 1065 MW at 85 kV to provide contracted
energy supply from Russia to Finland.
Duernrohr (Austria, 1983)20 - a monopolar 550 MW at 145 kV link
installed at Duernrohr to exchange energy between East and West
Europe.

96

High voltage direct current transmission

(xiii) Elzenricht (Germany, 1993)25 - this monopolar link, rated 600 MW


and 160 kV, located at the Elzenricht substation in NE Bavaria, was
the only practical solution for interconnecting Germany and
Czechoslovakia, two power grids with very different power-system
control philosophies and management practices. In addition to the
economic benefits, this link is expected to improve air pollution by
reducing significantly the generation of power near the common
border.
(xiv) Vienna South East (Austria, 1993) - this tie between Austria and
Hungary has the same characteristics as the Elzenricht link
described immediately above.
The Indian expansion
Figure 4.3 shows a map of India divided into four electrical regions

HVDC back to back


back to back proposed


HVDC line
HVDC line under
construction or proposed

Figure 4.3

Interconnections using HVDC between India's regions

HVDC system development 97

reasonably close to the four geographical areas and indicating their


prospective asynchronous interconnection by means of back-to-back
HVDC links.
(xv)

Western-Northern - a 2 x 250 MW 70 kV link (Vindhyachal) was


commissioned in 1990 interconnecting the western and northern
regions for emergency support, for improving the dynamic performance of their respective systems and to exchange power under
normal conditions.
(xvi) Western-Southern26 - still under construction at the time of writing,
this 2 x 500 MW, 205 kV link will control power flow via an existing 200 km 400 kV double circuit AC line from Chandrapur
(western region) to Ramagundam (southern region). Perhaps the
most innovative (and controversial) aspect of this scheme is the
absence of smoothing reactors.
(xvii) Eastern-Southern - Vijayawada-Gajuwaka link, expected completion by 1999, will exchange up to 500 MW power between the
eastern and southern regions.
(xviii) Eastern-Northern - the feasibility of a 500 MW interconnection (the
Sasaram project) has already been established and its financial viability is under consideration.

4.4 Operation reliability


A recent CIGRE survey27 on the reliability of HVDC systems for the year
1993 shows a perfect record of valve behaviour in 13 different schemes
commissioned in the past decade. The corresponding record for the
number of forced outages in those schemes owing to other causes was 128,
with a total outage time of 600 hours.
The number of outages for the year 1994 was very similar, although the
resulting forced outages increased to 970 hours. However, of these, only
five outages and 13 lost hours were caused by the valves. The causes of
unavailability were almost equally divided between transmission-line and
auxiliary-equipment faults.
Typical thyristor failure rates of 0.1 per cent per operating year are
currently achieved, but even such failures do not affect thyristor valve
operation, owing to built-in thyristor redundancy; i.e. the faulty thyristors
(or auxiliary equipment) can wait for replacement until the next scheduled
maintenance period.

4.5 References
1 HINGORANI, N.G.: 'High voltage DC transmission: a power electronics workhorse', IEEE Spectr., April 1996, pp.63-72

98

High voltage direct current transmission

2 NYMAN, ADIELSON, LOFGREN, and MUDER: 'Supply of a small isolated


AC system by local generation with an HVDC interconnection'. CIGRE, 1990,
paper 14-302
3 VON GEIJER, G., SMEDSFELT, S., AHLGREN, L., and ANDERSEN, E.:
'The Konti-Skan HVDC project', Electra, 1979, (63)
4 KANNGIESSER, K.W., et ah 'Commissioning of the Cahora-Bassa HVDC
system'. CIGRE, 1978, paper 14-13
5 HAUGE, O., VIKANES, S., ANDERSEN, E., and STYRBRO, G.: 'The Skagerrak HVDC transmission scheme system design features and service experiment'. CIGRE, 1978, paper 14-04
6 TAKEUCHI, T., et ah 'Hokkaido-Honshu HVDC link'. CIGRE, 1980, paper
14-03
7 BERIGER, C, ETLER, P., HENGSBERGER, J., and THIELE, G.: 'Design of
water cooled thyristor valve groups for extension of Manitoba Hydro HVDC
system'. CIGRE, 1976, paper 14-05
8 DONAHUE, J.A., and BRADLEY, D.: 'Multiterminal commissioning of the
Sandy Pond HVDC converter terminal'. International Colloquium on HVDC
and FACTS, paper 3.3, Wellington, 1993
9 EMBAIE, T., et ah 'Fenno-Skan HVDC link as a part of interconnected AC/
DC system'. CIGRE, 1988, paper 14-02
10 HAMMEL, A., KOELSCH, H., and DACHLER, D.: 'Control features of the
HVDC Gezhouba-Shanghai transmission scheme'. APS COM-91 international

conference on Advances in power system control, operation and management, 1991,

pp.680-85
11 GLEADOW, J., O'BRIEN, M.T., and FLETCHER, J.E.: 'Principal features of
the DC hybrid link'. International colloquium on HVDC andflexibleAC power
transmission systems, Paper 1.3, Wellington, NZ, 1993
12 DUBE, S.K., et ah 'Operating experience of the Rihand-Dadri 500 kV
HVDC transmission'. IEE conference on AC and DC transmission, 1996, pp. 1348
13 NYMAN, A., EKENSTIERNA, B., and WALDHAUER, H.: 'The Baltic cable
HVDC project'. CIGRE, 1996, 14-105
14 NIELSEN, T.G., CANELHAS, A., and HANSEN, B.S. 'KONTEK - the
HVDC link with the longest land cable in the world'. International colloquium
on HVDC and FACTS, Paper 6.1, Wellington, NZ, September 1993
15 YASUDA, M., MIZUSHIMA, K., KATO, Y., and SEKI, A.: 'Shin Shinano
frequency converter station'. CIGRE, 1978, paper 14-02
16 SENDA, T., et ah 'New technology applied to the recent HVDC converter
stations in Japan'. CIGRE, 1992, paper 14-02
17 RUIZ DIAZ, O., et ah 'Acaray HVDC back-to-back station'. CIGRE, 1982,
paper 14-02
18 SAKAI, T., et ah 'The Uruguaiana frequency converter station interconnecting
the Brazilian and Argentina AC system . Toshiba Rev., 1985, (153)
19 EMELUANOV, V., HEIKKILA, H., and MAKELA, L.: 'USSR-Finland HVDC
intertie'. CIGRE, 1978, paper 14-11
20 KANNGIESSER, K.W., MORAW, G., and POVH, D.: 'Commissioning and
system tests for the HVDC back-to-back tie Duernrohr/Austria'. CIGRE, 1984,
paper 14-01
21 LIPS, P., and THIELE, G.: 'Design and testing of thyristor valves for the
HVDC back-to-back tie Chateauguay'. Proc. int. conf. DC power transmission,
Montreal, June 1984, pp.228-33
22 DOHERTY, R.D., JOHNSON, R.K., SCHWEITZER, S.F., and WEAVER,
T.L.: 'Miles City converter station - early operating experience'. CIGRE, 1986,
paper 14-03
23 KOLODZIEJ, E., and HINGORANI, N.G.: 'Layout consideration for the
Oklaunion HVDC tie'. IEEE PES, tenth conference and exposition on Overhead
and underground transmission and distribution, September 1986
24 KLENK, E., et ah 'Advanced concepts and commissioning experiences with the
Sidney converter station'. CIGRE, 1988, paper 14-10
25 SCHMITT, et ah 'System performance and basic design aspects for the Elzen-

HVDC system development 99


richt 600 MW back to back HVDC converter station', IEE.Conf. Publ on AC
and DC power transmission, (345), London, 1991, pp. 171-76
26 WHEELER, J.D., DAVIDSON, C.C., WILLIAMS, J.D.G., and ROY, A.K.:
'Design aspects of the Chandrapur 2 x 500 MW back-to-back HVDC scheme'.
CIGRE, 1996, paper 14-104
27 CHRISTOFERSEN, D.J., ELAHI, H., and BENNETT, M.G.: 'A survey of the
reliability of HVDC systems throughout the world during 1993-1994'. CIGRE,
1996, paper 14-101

Chapter 5

Control of HVDC converters


and systems

A CONVERTER CONTROL
5.1 Basic philosophy
The ideal control system for an HVDC converter should meet the following requirements: 1
(a) Symmetrical firing of the valves under steady-state conditions.
(b) Instant of firing to be decided with regard to permissible values of
commutation voltage (rectifier) and commutation margin (inverter).
(c) Minimal reactive-power consumption in the converters, subject to the
condition that it is achieved without an unacceptable risk of commutation failure.
(d) Insensitivity to normal variations in voltage and frequency of the AC
supply network.
(e) Some degree of prediction of the optimum instant of firing in the
inverters, based on actual network voltage and direct current, subject
to the condition that it is achieved without an unacceptable risk of
commutation failure.
(/) Current-control characteristics with sufficient speed and stability
margin to cope with changing reference values and disturbances.
(g) Continuous operating range from full rectification to full inversion.
The theory presented in Section 2.5 is based on perfectly symmetrical and
sinusoidal waveforms with the firing angles (a) occurring at exactly equal
intervals and in the appropriate cyclic sequence. Deviations from such
ideal conditions give rise to two basically different control methods which
are discussed in Sections 5.2 and 5.3.

Control of HVDC converters and systems 101

A.C. line
voltage
m

/A \

Critical
level
trigger
circuit

Variable
delay

-L.
Grid
pulse
(valve 1)

w
A A

/ \

1 ~
lds\

Figure 5.1

y-

Amplifier

Constants current-control system

5.2 Individual phase control


This is the method used in the early HVDC converters. The firing instants
are determined individually for each valve, so that a constant delay (or
extinction) angle is maintained for all the valves in the steady state with
respect to the earliest firing instant (i.e. the voltage crossing).
In rectification the constant delay angle is normally determined from a
negative-feedback control loop, involving the set current and the actual
monitored current, as shown in Figure 5.1.
To maintain safe inverter operation with minimum reactive power
requirements the individual firings require:

a continuous calculation of the available voltage integral for commutation;


a continuous calculation of the required voltage integral for safe
commutation.

Optimum firing is achieved when the results of these two coincide.


The relationship governing the commutation process relies on the fact
that the time integral of the commutating voltage, i.e. the voltage integral,
is equal to the overall voltage change produced by the commutating
current, ic. With reference to Figure 5.2 such a relationship can be
expressed as
$Z~7oy/2Vc sin(ajt)d((ot) = 2(coL) $ dic
or
J2VC cos a + y/2 Vc cos y0 = 2wLId

Moreover, during large or small system disturbances the actual current at

102

High voltage direct current transmission

V2VC-

Figure 5.2

Typical six-pulse inverter under predictive CEA control


a Positive and negative direct voltages with respect to the transformer neutral
b Voltage across valve 1, and commutating voltages
c Valve current i\, i3 and ib
d Valve current z'2, i4 and i6

Control of HVDC converters and systems 103

the end of the commutation period will be different from the magnitude
anticipated by the controller, and compensation is made for the rate of
change of current. Thus the equation used as a basis for a predictive
constant extinction angle (CEA) is2
y[2Vc cos a + y/2Vc cos y0 = 2OJL (Id+ u)
dcot

(5.1)

where a is the firing delay angle, y0 is the preset CEA, u is the commutation angle and Vc is the commutating voltage (r.m.s.), e.g. VC(RB) is used to
compute eqn. 5.1 for valve 1.
This type of control has the advantage of being able to achieve the
highest direct voltage possible under asymmetrical or distorted supply
waveforms. On the other hand, any deviations from the ideal voltage
waveforms will break the 120 symmetry of the current waveform and
thus cause extra waveform distortion, as explained in Section 2.11.
Operating difficulties encountered in the early schemes indicated that
the distorted converter currents cause AC-voltage distortion which can
influence the firing pulse spacing via the control system, and often reinforce the original distortion.
In theory, this situation can be improved by placing control filters
between the AC system and the control system, so as to attenuate the
harmonics. However, the use of control filters has disadvantages too, like
the filter's inherent phase error, which varies with system frequency, and
its inability to attenuate negative-sequence fundamental voltages, the effect
of which is precisely to cause irregular firing-pulse spacings. Moreover,
with filters included, the control system will ignore the presence of harmonics on the AC voltages, whereas the valves will respond to the actual
voltages reaching them, including harmonics.

5.3 Equidistant firing control


The difficulties encountered with the original scheme encouraged the
development of an alternative control philosophy which could get away
from the voltage waveform dependence. A new principle, initially referred
to as the phase-locked oscillator, appeared in the late 1960s3 and largely
achieved the target.
The basis of this control system, illustrated in Figure 5.3, is a voltagecontrolled oscillator which delivers a train of pulses at a frequency directly
proportional to a DC control voltage, Vc.
The train of pulses is fed to a six-stage ring counter in which only one
stage is on at a time; the ON stage is stepped cyclically from positions 1 to
6 by the oscillator pulses. As each ring-counter stage turns on, it produces
a short pulse at the output (once per cycle). Therefore the complete set of

104

High voltage direct current transmission

Firing pulses
to converter

Figure 5.3

Principle of the phase-locked oscillator control system

six output pulses normally occurs at successive intervals of 60. The STOP
pulses are also obtained from the ring counter but two stages later (e.g. the
START pulse for valve 1 is from stage 1 and the STOP pulse for valve 1 is
from stage 3, normally 120 later). One oscillator and one ring counter per
bridge constitute the basic control hardware.
The various control modes only differ in the type of control loop which
provides the oscillator control voltage, Vc.
The phase of each firing pulse will have some arbitrary value relative to
the AC-line voltage, i.e. an arbitrary value of converter firing angle a.
However, when the three-phase AC-line voltages are symmetrical fundamental sine waves, a is the same for each valve.
In practice the simple independent oscillator would drift in frequency
and phase relative to the AC system; hence some method of phase locking
the oscillator to the AC system is required. This is normally achieved by
connecting Vc in a conventional negative-feedback loop for constant
current or constant extinction angle, as described in the following two
sections.

5.3.1 Constant-current loop


With reference to Figure 5.3, let us first consider the constant-current
loop, i.e. only signal Vc\ being effective. This voltage is obtained from the
amplified difference (error) between the current reference and the
measured DC-line current; this forms a simple negative-feedback control

Control of HVDC converters and systems 105

loop, tending to hold current constant at a value very close to the reference.
To visualise the operation of this loop, imagine that the current is nearly
equal to the reference, such that the amplified error (Vc) happens to be
precisely that value required to give an oscillator frequency of six times
the supply frequency. The ring-counter outputs, and thus the valve-gate
pulses, will have a certain phase with respect to the AC-system voltage.
Suppose, further, that this phase, which is identical to firing angle a,
happens to be such as to give the correct converter DC output voltage,
which, with the particular back e.m.f. of the DC link, results in the correct
DC-line current. This is steady-state operation.
The loop is self correcting against disturbances of any source. For
instance, a drop of back e.m.f. in the DC system causes a temporary
current increase, which reduces Vc and hence slows down the oscillator,
thus retarding its phase and finally increasing the firing angle a. This
tends to decrease the current again, and the system settles down to the
same current, with the same Vc and oscillator frequency but a different
phase, i.e. different a.
The control system will also follow system frequency variations, in which
case the oscillator has to change its frequency; this results in different Vc
and hence current, but the current error is made small by using high-gain
amplification.
This constant-current scheme is the main control mode during rectification; it is also used during inversion whenever the inverter has to take over
current control, as explained in Section 5.5.
The control system response is fast but, in practice, its effect will be
slowed down by the relative slower response of the DC line which includes
capacitance, inductance and smoothing reactance.

5.3.2 Inverter extinction-angle control


This control mode is implemented by a negative-feedback loop very
similar to the current loop and is also shown in Figure 5.3. The difference
between the measured y and the y setting is amplified and provides Vc as
before. However, it differs in that y is a sampled quantity rather than a
continuous quantity. For each valve the extinction angle is defined as the
time difference between the instant of current zero and the instant when
the anode voltage next crosses zero, going positive. Typical waveforms of
the y-measuring technique are shown in Figure 5.4.
For each bridge there are six values of y to be measured, which under
symmetrical steady-state operation are identical. Under unbalanced conditions, however, the valve with greatest risk of commutation failure is the
one having the smallest y; this smallest measured y produces the most
negative output and thus causes its diode (signal A in Figure 5.3) to
conduct and produce the negative feedback voltage VC2. During steady-

106

High voltage direct current transmission

M. .

Figure 5.4

OH

(a) Firing pulse, valve 3

(b) Anode current, valve 1

(c) Anode-cathode voltage, valve 1

(d) y-Timing waveform


(e) Voltage Vc2

Waveforms of y-control circuit

state operation and full inversion, Vc2 controls the oscillator holding the
smallest y at a predetermined value by closed-loop control.
Under these conditions Vc\ is zero because the inverter constant current
(CC) setting is less than the DC-line current (determined by the rectifier
CC control); hence the inverter CC loop is trying to decrease oe by making
Vc\ as low as possible. The minimum Vc\ is clamped to zero volts and thus
during normal inverter operation the CC loop is ineffective.
Component B is an additional feedback voltage (V^) applied during the
transient conditions when y < ymin. A sudden impulse is then applied to
the voltage-controlled oscillator, which has an integrating characteristic
and thus can suddenly shift the phase (i.e. angle a) by an appropriate
amount.

5.3.3 Transition from extinction-angle to current control


The current setting for inverter operation is below that of the current in
the DC line, which is normally kept at a higher constant level by the
sending-end terminal (explained in Section 5.5). Under such conditions
Vc\ in Figure 5.3 is zero, the current-control amplifier is saturated and the
converter operates on constant-extinction angle y at full inversion.
In the event of a sudden AC-system voltage rise at the inverter end, or a
DC-line voltage reduction, the direct current will decrease; the current
amplifier then comes out of saturation and Vc\ becomes positive. This
additional input to the oscillator causes operation of a larger y, i.e.
advances the firings and the converter takes over current control.

5.3.4 Other equidistantfiring-controlschemes


Various versions of the original phase-locked oscillator control system have
subsequently appeared.

Control of HVDC converters and systems 107

1Aa

Figure 5.5

Firing pulses of alternative equidistant firing control

The basic principle of an early alternative4 is illustrated in Figure 5.5,


where a control function Vcf of linear slope initiates the firing signal at the
intersection with the controller output voltage Vc. At that instant Vcf
returns to its initial value and starts again. The time between consecutive
impulses is therefore determined by the magnitude of Vc and the slope of
Let us assume that the value of Vc in Figure 5.3 has been selected so that
the impulse interval is precisely 60 of the actual system frequency, and
that the position of the impulse corresponds with a particular delay angle,
a. If Vc remains unchanged, the sequence of firing instants is determined
by the dotted vertical lines. With a temporary increase in Vc the impulse
spacing increases and the firing sequence is determined by the full vertical
lines. For each impulse occurring while Vc is higher, the delay angle is
increased by the amount Aa. Following the return to the original Vc the
total increase (3Aa in the illustration) is retained.
Some schemes involve an element of prediction of the extinction angle,
similar to the conventional individual phase-control scheme; the prediction
is made using the latest values of AC voltage and DC current in the
commutation equation, as described in Section 5.2.
One of these versions, entitled 'equidistant firing predictive-type control'?

claims that the prediction is effective for the incoming firing, and uses a
feedback loop to update the predictive model for the subsequent firings.
In both the basic phase-locked oscillator and the equidistant-firing
predictive schemes, a change in the control voltage directly changes the
frequency of the oscillator, and the synchronisation of the oscillator takes
place with the help of the main current-control loop or the extinctionangle control loop.

108

High voltage direct current transmission

In an alternative equidistant-firing-control scheme,6 illustrated in Figure


5.6, a voltage-controlled current source charges the capacitor C. When the
voltage across this capacitor exceeds a given control voltage, Vc\, by an
amount A V/2, the voltage comparator produces an output pulse and the
capacitor is rapidly discharged to the voltage Vc\ - AV/2; the charging of
the capacitor then begins once again. The output pulses of the voltage
comparator are distributed to the converter valves via a shift register.
In Figure 5.6, the frequency of the oscillator is determined by Vc2
which results from the sum of two signals, VC2\ and VC22- ^c2i *s proportional to the AC-system frequency and Vc22 is the output of a slow-acting
a-control loop. The phase of the control pulses is determined by the
control voltage, Vc\, which is derived (see Figure 5.7) either from a
proportional-integral acting current controller or from a proportionalintegral acting extinction angle controller. The cos"1 circuit compensates
for the nonlinear characteristic of the converter. To ensure inverter stability it is important to keep negative deviations of the extinction angle as
small as possible, and the measured y is presented to the controller
through a shaping network. The additional disturbing magnitudes
provide the element of prediction.

5.3.5 Application to 12-pulse converter groups


When bridges are connected in series, the common direct current is determined by one controller per pole and the control pulses are directed to the
individual bridges from a common ring counter. With 12-pulse converter
groups it is thus necessary to double the number of control impulses per
cycle. In this case the ring counter provides the control pulses to the 12
different valves of the two bridges, keeping four valves activated simultaneously.

5.3.6 Comparative merits


For the normal working range, equidistant-firing control is preferable to
individual phase control because of the reduction of abnormal harmonics. As with any control system, its firing-pulse spacing may be modulated by any noncharacteristic harmonics entering via feedbacks, e.g. of
DC current. Thus its description as equidistant-firing control is somewhat
anomalous; however the phase-locked oscillator control system is very
stable when suitably designed, and this is not usually an important effect.
It is often claimed that, whenever the DC power contributes considerably to maintaining the network stability, such a system may benefit from a
changeover to individual phase control under asymmetrical-network fault
conditions. However, under asymmetries the power transmitted is limited
by the current (with individual control) or the voltage (with equidistant

Control of HVDC converters and systems 109


Pulse generator

Shift
register

Voltage
comparator

Voltage
controlled
current

source

Capacitor
discharge
circuit

ex-Control ler|

(a)

A^ci.i

vl2

1 /f A ..A
[AW2

Ar/i /i i

\A I

V V V\

tan 6

vP

(ot

* 60

5.6

nn

Pulse-phase grid-control system (PPC) ( 1970 IEEE)


a Block diagram
b Determination of the phase of the firing pulses

n_

110

High voltage direct current transmission

Disturbing
magnitudes

Y*
y-Measuring
circuit

v.

Y*
/
'

+
+

YN

V
controller

Y^,
Grid
control

ld act

Controller

Figure 5. 7 Block diagram of the closed-loop extinction-angle control with asymmetrical controller and of the current control with selecting circuit (
1970 IEEE)

control) and therefore the above conclusion must be restricted to asymmetries which require no current limiting. For large asymmetries, with the
power decreasing to very small levels (with either control), the reduction
of direct voltage during the fault is less important; in such cases the use of
equidistant-firing control provides more reliable commutations and facilitates the rapid return of power flow when the asymmetry ends.
On the subject of adding predictive control to the basic closed-loop equidistant control, simulator studies, carried out with reference to the CrossChannel link, have led to the following conclusions:
(a) In normal steady-state inverter operation, whether balanced or not, it
is essential that at least half the valves in a group are on closed-loop
control; otherwise harmonic instability occurs (except with a very
strong AC system, which is irrelevant for design purposes).
(b) The addition of a predictive y control which is not normally in operation, but standing by at a few degrees less than normal, reduces the
probability of commutation failure owing to very small disturbances
from the AC system.
(c) However, some magnification of the original disturbance is inherent
in all predictive y controls of whatever type.
(d) The net effect for moderate AC-system disturbances (remote fault) is
to cause so much magnification that the distortion is beyond the
capabilities of predictors, and commutation failure then follows in
many cases where it would not have occurred without the predictive
addition.

Control of HVDC converters and systems 111

(e) Following large disturbances, commutation failure occurs initially


regardless of the type of control.
(/) During a sustaine'd single-phase fault at medium distance, after the
initial shock an inverter may settle to some form of steady unbalance
operation, giving finite power, provided the closed-loop control is
given suitable characteristics. The addition of predictive y control
generally gives unstable operation in this mode; hence the well known
argument that predictive (individual-phase) control gives more power
in this mode is a fallacy, again except for the case of infinite busbars,
which is not of practical interest.
(g) The use of inverse-cosine linearisation gives a small advantage on infinite busbars, since a slightly higher loop gain can be used. However,
on a finite system impedance it has a detrimental effect, as for operation near to 0 or 180 the control loop causes large changes of a for a
small change of the feedback quantity (ID, y, etc.); this produces large
changes of reactive power and AC voltage, which tends to destabilise
the system. A simple linear integral loop is a good compromise.

B - DC SYSTEM CONTROL
5.4 Basic philosophy
From an operational point of view, the use of constant-current control
provides a greater safeguard against disturbances. Therefore, if there is a
choice and the economics are right, constant current should be used as the
basic control philosophy. This, however, is not possible with conventional
AC-power systems which, being normally multiended, require reasonably
constant voltages at the point of common coupling.
The use of current conversion as the basis for rectification and inversion
has already been justified in Section 2.2. Moreover, present DC schemes
consist of point-to-point system interconnections and constancy of DCvoltage levels is not a primary consideration because these are not directly
available to consumers. There are, however, other considerations influencing the control philosophy, among them featuring prominently the overvoltages resulting from open circuiting and load rejections and the high
resistive losses resulting from constant-current transmission at low power
levels.
A hybrid voltage/current philosophy is possible with a DC transmission
scheme to suit the needs of the particular operating conditions. This is
achieved by adjusting the DC-voltage levels on both sides of the link, by
means of on-load tap-changer control on the steady state, and by thyristor
control following large or small changes of operating conditions at either

112

High voltage direct current transmission

end of the link. The DC current is only limited by the small resistance of
the transmission line and is therefore very sensitive to such variations.
It will be shown in the following sections that the provision of current
controllers at both ends, combined with transformer on-load tap changing,
offers a perfectly satisfactory solution to this problem; thus the use of
current control is universally accepted in HVDC transmission.

5.5 Characteristics and direction of DC-power flow


Most DC schemes in existence are provided with bidirectional power-flow
capability. This property is inherent in the case of AC transmission, where
the direction of power flow is determined by the sign of the phase-angle
difference of the voltages at the two ends of the line; the power flow direction is in fact independent of the actual voltage magnitudes.
On the other hand, in DC transmission the power-flow direction is
dictated by the relative voltage magnitudes at the converter terminals and
the absolute or relative phase of the AC voltages plays no part in the
process. However, this condition can be altered by exercising a type of
firing-angle control which can make the power-flow direction independent
of the terminal AC-voltage magnitudes and behave, instead, like the AC
counterpart.
The basic characteristic of a converter from full rectification to full
inversion is illustrated in Figure 5.8. The converter is assumed to be
provided with constant current and constant extinction-angle controls,
which have already been discussed in Section 5.3.
The level of the natural voltage characteristic can be adjusted by the
transformer onload tap changer. This part of the characteristic takes place
when a = 0 (i.e. with diode operation), i.e. the converter has no controllability in this region and the DC voltage (from eqn. 2.12 with a = 0) is

Vd=Vc()--Irl

(5.2)

For delay angles larger than a = 0 the converter exercises constant-current


control, i.e. to maintain a current reference (Id), which has a practically
vertical characteristic. This region is governed by eqn. 2.12 and is limited
by the need to maintain a certain minimum extinction angle required for
safe commutation, as explained in Section 2.7. When this limit is reached,
the converter again loses controllability and is governed by eqn. 2.19.
Let us now consider the specific control parameter which implements
the direction of power flow. We have already explained that each converter station is normally provided with current and extinction-angle control
facilities. The assignment of current control, either to the rectifier or to

Control of HVDC converters and systems 113

C.C. control

/3V2\

/3to/.\

(a = n -

Rectifier
i

C.C. control

Inverter

/3V2\
/ 3a)L
Wccos p+

V JI

V n

C.E.A. control

(3V2)
JI

(3V2)

Vc cos

Figure 5.8

YO

<=

/3V2\

VCCOSYO-

/3coL\

Complete control of a converter, from inversion to rectification


NV = natural voltage
CC = constant current
CEA = constant extinction angle

114

High voltage direct current transmission

Ids

(a)

Figure 5.9

(b)

Steady-state characteristics and operating point

a Under rectifier-current control


b Under inverter-current control

the inverter station, is made on consideration of the investment cost for


reactive-power compensation, the availability of reactive power, the minimisation of the losses and the total running cost. Normally, the total reactive-power compensation is least, and the utilisation of the line best, if the
rectifier is assigned the current-control task while the inverter operates on
minimum extinction-angle control.
This combination is achieved in a two-terminal DC link by providing the
power-sending station with a slightly higher current setting than the
power-receiving station. The difference between the two settings is termed
the current margin (/^w). Its effect can be better understood with reference
to Figure 5.9a where the operating current is set by the constant-current
control at the rectifier end. The inverter-end current controller then
detects an operating current which is greater than its setting and tries to
reduce it by raising its own voltage, until it hits the ceiling determined by
the minimum extinction-angle control at point A. This is the normal
steady-state operating point, which presumes a higher natural voltage characteristic at the rectifier end, a condition which may require onload tapchange action at the converter transformer.
The no-load and direct voltage regulation along the transmission link
are illustrated by the continuous line in the diagram of Figure 5.10, with
power flowing from left to right and with the operating current maintained by the rectifier controller.
Starting from the rectifier end, point a represents the maximum average
direct voltage, (Vc0)R, with the rectifier unloaded. This is reduced by firingangle control to (V^))R cos a (point b). The slope of lines bcA and -dA is
determined by the operating direct current Id, and their horizontal projections relate to the commutation reactances XCR (for be), XQI (for dA) and
line resistance Rd (for cA). Therefore, point c represents the output voltage
at the rectifier station and point A the voltage at the remote end of the

Control of HVDC converters and systems 115

d\ A
>^
-^

- ,

d'
A^~-~~\

"?^

8 8
0

fid

Figure 5.10

DC voltage profile

line. This point is also reached from the inverter-end open-circuit voltage,
(Ko)i> reduced by the extinction angle, (V^i cos y, and by the commutation reactance. It should be obvious that point A represents the same operating condition as indicated by the crossing point in the characteristics of
Figure 5.9a.
Let us now assume that there has been a substantial AC-voltage reduction at the rectifier end, such that the DC-voltage ceiling (the natural
voltage) of the rectifier becomes lower than that of the inverter. In the
absence of a current controller at the inverter, the voltage across the line is
reversed and the current reduces to zero (current through the valves
cannot reverse). However, an inverter-current controller will prevent a
current reduction below its setting by advancing its firing (i.e. reducing a
and hence inverter DC voltage), thus changing from extinction angle to
constant-current control. A new operating point, A (Figure 5.96), results at
a current reduced by the current margin. In Figure 5.10 this condition is
represented by the dotted lines and clearly shows that power flow will
continue in the same direction, in spite of the lower AC voltage at the
sending end.

5.5.1 Tap-changer control


Referring to Figure 5.9a, the task of the tap changer on the rectifier side is
to place the voltage ceiling (i.e. the a = 0 characteristic) relative to the oper-

116

High voltage direct current transmission

ating point {A), so as to minimise the reactive-power consumption subject


to a minimum a limit. This limit is used to ensure that all the anodes (in
the case of the mercury-arc valve) or the series thyristors have a sufficient
positive voltage across them to avoid the risk of misfiring.
On the inverter side, the reactive power is minimised at the y = y0 limit,
which is the normal operating condition. It is therefore possible to use its
tap changer to keep the direct voltage within any desired limits.
When the current control is transferred to the inverter (as in Figure
5.9b), the rectifier tap changer tries to raise the voltage ceiling and the tap
changer action on the inverter side must stop, since the latter no longer
determines the voltage level. Otherwise, the tap changer would try and
raise the inverter-side AC voltage and thus reduce the power factor.
It is interesting to note that the DC link, within its normal power rating,
operates as a constant-voltage circuit, i.e. the voltage is at its highest value
under no-load conditions and power is raised by increasing the current
settings. However, in order to avoid overloads, when the current limit is
reached, the transmission link acts as a constant current system, i.e. a
power change can then only be achieved by means of a voltage change
and, because of the voltage ceilings, this is only possible to a limited extent
by means of tap-changer action.

5.5.2 Reversal of power flow


The previous section explained that the direction of power flow is decided
purely by control action. It can be expected, therefore, that a change in
power direction cannot happen naturally as a result of a change in the
operating conditions, but rather following a control order resulting from
the overall requirements of the power system.
The characteristics of Figure 5.9<2 can be extended below the voltagezero line so that the rectifier and inverter ends exchange their function.
The result of this action, illustrated in Figure 5.11a, is that the two characteristics do not meet again. Station I increases the delay angle well into the
inverting region and hits the limiting extinction-angle voltage (70) Station
II advances its firing into the rectifying region and finally hits the rectifier
voltage ceiling (a = 0). In the process, the line current reduces to zero and
the complete system is blocked.
Since, as indicated in the previous section, the rectifying station
requires the higher current setting, it is thus necessary to subtract the
current margin from the reference value of Station I. This results in the
characteristics of Figure 5.116, which display one operating point of
different voltage polarity and with the roles of the two stations interchanged. Thus, the direction of power flow has been reversed without
altering the direction of current flow, which of course is fixed by the
converter valves.

Control of HVDC converters and systems 117

Idm -

Ids

(a)

Figure 5.11

(b)

a Operating point with power flowing from Station I to Station II


b Operating point after power reversal

5.5.3 Modifications to the basic characteristics


During AC-system faults at the receiving end there is a large risk of
commutation failure and, if the fault is electrically close, the inverter may
not be capable of recovering by itself. In such cases it is important to
reduce the stress on the inverter valves, and this is achieved by providing a
low-voltage-dependent current limit to the rectifier control characteristic.
The modified characteristics illustrated in Figure 5.12 consist of a branch
CD' at the rectifier and another EF' at the inverter.
The break points C and E in Figure 5.12 are typically between 70 per
cent and 30 per cent of DC voltage, and in special cases even higher
depending on AC-system requirements. The higher voltage is applied
when the receiving AC network is sensitive to disturbances which could
cause large voltage fluctuations.
The terms VDCOL (voltage-dependent current-order limit) and LVCL
(low-voltage current limit) are commonly used for the function which
reduces the current setting when the voltage decreases.
In the literature, the added branches CD' and EF' are normally given
finite slopes on the basis that the finite slope is expected to provide
smooth control; in fact this argument is not valid, since the slope constitutes a strong negative resistance which produces the opposite effect. A plain
horizontal line, as shown in Figure 5.12, is better since the system will then
operate stably down to a lower DC voltage during faults in the rectifier AC

118

High voltage direct current transmission

a = 0
B'
B

'dm

F
D'

K'H K
G H
Figure 5.12

lds

Rectifier and inverter modified characteristics

system; it will also recover faster from a total collapse of an inverter


because the AC line/cable will charge more rapidly.
Also, an unwanted voltage reversal of the inverter is generally prevented
by limiting a to some minimal value larger than 90 (shown by KK of the
characteristic). Branches DH and FG represent maximum-current limits at
low voltage. A minimum-current limit (not shown in the Figure) is also
normally provided to prevent operation with discontinuous current.
Another problem with the basic characteristics occurs when the rectifiervoltage ceiling gets very close to that of the inverter, such that the characteristics intersect in the region between I(is and Ijs - I(im. In this region both
current controllers are out of action. In practice, the current will not settle
at an intermediate point; instead, the inverter will periodically enter the
current-controlling mode. A proven countermeasure against this type of
oscillation is a small shape alteration of the inverter characteristic, as
shown in Figure 5.12. The limiting y = y() characteristic cannot be changed
without sacrificing heavily the reactive power; it is, however, sufficient to
break the characteristic to a positive-resistance slope in the transfer region
from y control to current control of the inverter (AB in Figure 5.12 instead
of AB').

5.5.4 Operational nonminimum margin angle


In this case, the margin angle y is varied around an average value in order
to stabilise the AC voltage. This can be achieved either by a direct control
of the AC voltage or indirectly by controlling the DC voltage. Another way

Control of HVDC converters and systems 119

Ymin

const. Ud
1.0 p.u.

Figure 5.13

Alternative inverter characteristic with higher than minimum


margin-angle control

of stabilising the receiving-system AC voltage is to use the inverter, rather


than the rectifier, as the current-controlling station. These modes of
control are normally used for operation in the unstable region of the ACvoltage/DC-power characteristic.
An alternative solution is shown in Figure 5.13 in which the operating
point is moved down along the part of the characteristic with positive
slope. The inverter is then able to increase the voltage at increased direct
current. This method is very effective when the inverter is connected to a
weak AC network. The normal operating point A corresponds to a value
of y higher than the minimum.

5.5.5 Power-flow control


As the primary object of HVDC transmission, power control should be the
main consideration. However, for simplicity, the early schemes still used
the current-control loop as a basis. In such a case each station is provided
with a dividing circuit, consisting of a power calculator and a high-gain
operational amplifier, which receives the power-order, voltage and current
signals and produces the current order. Other components of the station
power-control unit, illustrated in Figure 5.14, include the telecommunication equipment (TC), the constant-power order-setting unit (OS), an emergency-power controller (EPC) and a current limiter. A control error (CE) is
formed from comparison of the calculated current order (minus the
current margin in the case of the inverter station) and the measured
current.
In some early schemes each pole at both stations was equipped with
power-flow controls as shown in Figure 5.14. The power order and corresponding rate of change order could then be set by the operator at the
main station, or they could be transmitted there from a dispatch centre.
The task of the order-setting unit of such a scheme is to implement power

120

High voltage direct current transmission

AP*

CE

&f*

Figure 5.14

Example of power-flow control with additional emergency control for


frequency
TC telecommunication equipment
EPC emergency-power controller
P*o manually-set power order
Z!*P additional power order by higher-level controller
AP* step in emergency power
P*E emergency-power order
OS order-setting unit
Af thresholds for frequency deviations
A/ AC-network frequency deviation
I*i calculated current order
CE control error
Inv inverter
Rec rectifier
(Asterisks denote reference values)
( 1978CIGRE)

stepping, and synchronisation of order setting at the two ends of the link,
by telecommunications.
In most cases there is only one master controller (at one of the stations)
which sends a current order to the pole controls of the two ends of the
link. The power is monitored by multiplying voltage and current (summed
from both poles) and fed back directly to the controller. As in previous
control methods, to prevent unacceptable current orders (e.g. during start
up) limits are normally built in.

5.5.6 Frequency control


The frequency of a network interconnected to a larger one by an HVDC
transmission link can be controlled by means of a frequency feedback loop
acting on the DC-link controls, such that the small network draws the
required power change from the larger one.

Control of HVDC converters and systems 121


' Pa

+ Af

Figure 5.15

Power/frequency control characteristic

A typical case of frequency control is the Gotland link, as originally


designed, which included a synchronous compensator. This network once
started could be operated with no power feed other than the DC link.
Similarly, when the power rating of the DC line is comparable with or
greater than the rating of the running generators in the AC system to
which the line is connected, the line terminal can share in the frequency
regulation or even perform it unaided.

5.5.7 Power/frequency control


Often a combination of control modes is used, such that the control signal
applied to the current controller is normally power, and remains so as
long as the frequency remains within the predetermined limits. Outside
these limits, frequency control takes over to assist the system in difficulty.
However, when the maximum-rated transmission power is reached, the
frequency controller becomes inoperative. This combination is illustrated
in Figure 5.15.
It should be remembered that the DC link is insensitive to frequency
variations, unless some sensitivity is deliberately added to the control
system. Without it, a constant-power flow can overspeed a receiving system
which has lost part of its load and a sending system can eventually collapse
if the required DC-link power is more than the connected generation can
produce. In general, therefore, the incorporation of an element of
frequency control is recommended.

5.6 Different control levels


For reasons of reliability, the controls of an HVDC link are generally
divided into four different levels,9 i.e.:

122

High voltage direct current transmission


Overall
controls

p
df/dt
Telecom.

-1

Figure 5.16

**

h
Telecom.

Block diagram of a control scheme

(a) Bridge controls - to control the firing instants of the valves within a
bridge and to define the y0 and a min limits.
(b) Pole controls - to co-ordinate the bridges in a pole to provide the
ordered current, with minimum harmonic generation.
(c) Master controls - to provide co-ordinated current orders to all the
poles.
(d) Overall controls - to provide the current orders to the master controls
in response to required functions such as power-transfer control,
system-frequency control, system damping or combinations of these.
A simplified block diagram of a multibridge control is illustrated in Figure
5.16. The bridge controls, and in particular the valve-firing circuits,
contain most of the components involved, and are therefore kept independently for each bridge unit (or converter group in modern 12-pulse
schemes) for reliability reasons.

Control of HVDC converters and systems 123

5.6.1 Overall control co-ordination


The need for an overall control policy is discussed here with reference to
the Nelson River scheme. It should be made clear that the specific needs
of this scheme, and therefore the solutions adopted, may not apply generally. However, the overall control philosophy used in the Nelson River
project, illustrated in the simplified diagram of Figure 5.17, helps us to
understand the degree of dynamic interaction which can be achieved in
modern AC-DC systems.
The following brief description of the main control functions shown in
Figure 5.17 has been extracted from the literature:10
(a) The master power controller is designed to accept a power order (Poi)
from the operator or the automatic-generation control. This power
order is divided by the measured bipole voltage (V\) to determine the
current order (/01).
(b) The frequency-based capability control (FBCC) provides protection
against the HVDC collector system overloading. DC power is reduced
if the collector-system frequency drops below 59 Hz.
(c) The excess power-order transfer (EPOT) controller is designed to fully
utilise the combined available HVDC system capability. The EPOT
controller transfers ordered power, which is in excess of the capability
of one bipole, to the other operating bipole. This controller is acti-

Henday

Radisson

Ont. U.S.230 U.S. 500


ties kVTies kVTie

Figure 5.17

A simplified schematic of controls on the Nelson River HVDC system

124

(d)

(e)

if)

ig)
ih)
ii)

High voltage direct current transmission


vated only when a major outage occurs on the bipole, such as a valve
group or pole block.
The HVDC power reduction for the tie-line trips controller is
designed to ensure system stability upon loss of interconnection lines.
DC power is reduced by an amount equal to tie-line loading prior to
tripping.
The undervoltage DC-reduction controller is designed to reduce DC
power whenever Dorsey voltage starts to collapse. A fixed amount of
DC reduction releases MVARs both from the HVDC link and the AC
system to restore the voltage. This control does not react to faults.
The allocator accepts DC reductions from the tie-line reduction and
undervoltage-reduction controllers. Total DC reduction is then
summed and allocated to each bipole according to a preset power
order.
The sending-end frequency control (SEFC) minimises oscillations in
the collector system.
The receiving-end frequency control (REFC) minimises oscillations in
the receiving system.
The receiving-end damping control (REDC) operates to prevent
changes in the angle of the Dorsey 230 kV bus voltage.

5.6.2 Hierarchical power control at the New Zealand link11


The recently upgraded New Zealand HVDC system has a hybrid configuration. In this configuration, the existing mercury-arc converters were
paralleled onto one pole (pole 1) for operation at 270 kV, and new thyristor converters together with new submarine power cables for operation at
350 kV formed the other pole (pole 2). The term hybrid was used because
the configuration adopted was partly a voltage uprating, and partly a
current uprating. The single-line diagram is shown in Figure 5.18.
The control system is microprocessor based with built-in redundancy.
For maximum reliability, a hierarchical power-control arrangement is
provided which uses three levels of controls; half-pole control (270 kV
pole only), pole controls and bipole controls.
The HVDC link is usually operated in constant-power mode with
balanced pole currents even though the voltages differ. The bipole control
order is subdivided in proportion to the pole voltages to produce separate
power orders for each pole. A fast power-transfer feature is included
which operates in the event of the loss of one pole and transfers power to
the in-service pole subject to its short-time overload capability.
Both the frequency stabilisation and the spinning-reserve sharing functions are de-activated during the short time when the Haywards voltage
stabilisation is active, as this has the highest priority.
The scheme includes an integrated supervisory control and data acquisition (SCADA) unit to permit full remote control and monitoring of alarms

Control of HVDC converters and systems 125


110kV
34.7km

110kV
system

1x350kV 1430A

bipole +270kV/-350kV
nominal rating 992MW, 1600A
continuous overload rating 1240MW 2000A

^ t 99_su|?manne_cables_
2x350kV 1430A

Haywards

Figure 5.18

Benmore

1240 MW hybrid upgrade of the New Zealand link

from either end as well as power-setting control, reactive-power control


and monitoring from the North and South Island control centres.
The control system of the new thyristor converter and the paralleled
mercury-arc group has a number of special features to enhance the operation of the connected AC systems; these are:

voltage stabilisation modulation of the DC power for high-power transfer to the North Island to maintain stability following disturbances;
frequency stabilisation modulation to reduce severe frequency excursions and to stabilise the frequency on both islands;
spinning-reserve sharing to allow a reduction in the spinning-reserve
requirement on both islands through automatically adjusting the DCpower transfer to share the spinning-reserve generators;
a constant-frequency control, for use only when Haywards (and the
associated local area) is isolated or islanded from the rest of the North
Island AC grid. This is manually initiated following a disconnection
and controls the frequency in the Wellington region to 50 Hz. This
control feature has improved the reliability of supply in the Wellington
region by using the HVDC link without assistance from any North
Island generation plants;
runbacks in power transfer following the loss of critical elements in the
DC link or the connected AC systems, and for loss of communications;
a thermal image/overload characteristic control function which
continuously determines the maximum time duration allowable at
various overload levels, based on the predisturbance thermal loading

126

High voltage direct current transmission

of the submarine cables. The power transfer onto the overloaded pole
is automatically reduced to the maximum continuous acceptable cable
current after the inherent overload capability has been utilised. This
assists the dispatch of spinning reserve plant on the AC grid and minimises AC-grid frequency excursions and load shedding following a
sudden forced-pole outage;
a fully automated reactive-power/AC-bus voltage controller at each
station, to control all the equipment which affects reactive power
supply. The control includes switching of AC filters and shunt reactors, generator and synchronous condenser outputs, DC-voltage level,
and interconnecting transformer tap changers. This ensures the
optimum co-ordination of equipment for any DC-power level and
system configuration;
sustained reduced voltage level operating provision, of 250 kV for the
350 kV pole, to cater for conditions such as serious insulation pollution
problems on the DC line. This is achieved by a combination of the tapchanger range on the converter transformers and some delay-angle
control. During reduced voltage operation the DC current is limited to
1200 A.

5.7 Telecommunication requirements12


Without special arrangements to co-ordinate the operation at both ends of
the link, a reduction of rectifier (or an increase of inverter) current order
by more than the current margin, relative to the inverter (or rectifier)
current order, can cause a complete DC-voltage shutdown. This problem is
normally avoided by the following procedure:
(a) When the master controller operates from the sending (i.e. rectifying)
end, an increase in current order is implemented immediately at the
local end, and as soon as possible (only subject to the unavoidable telecommunication delay) at the remote end. On the other hand a
decrease in current order updates the remote current order first, and
the local current-order change waits for a true check-back signal from
the remote end, via a return telecommunication channel indicating the
receipt of an error-free signal.
(b) When the master controller is housed at the inverter end, the above
procedure still applies but with the words increase and decrease interchanged.
(c) If an error is detected in either telecommunication receiver then the
local current order is not updated; the return check back is sent as a
zero.
Functions (a)-(c) are obtained by telecommunication coupling logic, which
requires two registers, respectively, for present and last current order, with

Control of HVDC converters and systems 127

a comparator and a few simple gates. This logic is inserted between the
source of master current order and the telecommunication system, and
also supplies the output to the local pole controls. Its effect is to ensure
that the sequence of current-order updating is never such as to decrease
the effective current margin, even in the presence of errors in the telecommunication systems, and to freeze both current orders for a detected error
in either telecommunication channel.
Telecommunication technologies used in earlier HVDC schemes
included microwave radio, carrier on the power conductors of private
wire, rented wire and the use of the public telephone system.
All of these are liable to interference and even occasional failures.
Carrier systems, using an HVDC line or cable, are affected by a continuous
source of interference (i.e. the converters) which is difficult to filter
because of the high impedance of the smoothing reactors; the use of the
lowest possible bandwidth is thus recommended to reduce the effective
noise.
To avoid noise the analogue signals are converted to digital form and
transmitted with an error-checking signal.
Digital signals are normally sent in regular blocks of binary bits, at a
block rate which depends on the telecommunication medium and on the
desired rate of response. The main information includes the current order
and some logical signals such as the power-flow direction and fast shutdown orders.
A number of relatively slow signals, such as the interchange of manuallyset power orders or supervisory signals, can be sent by submultiplexing.
Recent schemes make extensive use of fibre-optic cables, which are
immune from interference. The New Zealand scheme, for instance,
contains a high-capacity fully-redundant communication link plus a physically-independent back-up link. It is used for control, protection, data,
status functions and speech facilities, between Haywards and Benmore.11
The high-capacity redundant link utilises fibre-optic ground and submarine wire as far as Islington near Christchurch, and finally a microwave link
to Benmore.
The fibre-optic ground wires, containing 12 fibres, were installed in
place of the original ground wires, and two 40 km 12 fibre submarine
cables were installed across Cook Strait at the same time as the 350 kV
submarine power cables were installed. This section of the main link
has a signal capacity of 565 megabits/s and the microwave section 8 megabits/s.
The main communications link transmits all the signals which are critical to the successful operation of the DC hybrid link. When operating on
the back-up link, which has a limited capacity of 256 kbits/s, the HVDC
control is limited to the constant-current mode which is reliant on some
operator manual intervention and does not include the use of any of the
modulation features.

128

High voltage direct current transmission

5.8 References
1 TARNAWECKY, M.Z.: 'H.v.dx. transmission control schemes', Manitoba
power conference EHV-DC, Winnipeg, 1971, pp.699-741
2 HINGORANI, N.G., and CHADWICK, P.: 'A new constant extinction angle
control for a.c./d.c./a.c. static converters', Trans. IEEE, 1968, PAS-87, (3)
pp.866-72
3 AINSWORTH, J.D.: 'The phase-locked oscillator - a new control system for
controlled static convenors', Trans. IEEE, PAS-87, (3), 1968, pp.859-65
4 UHLMANN, E.: 'Power transmission by direct current' (Springer-Verlag,
Berlin/Heidelberg, 1975) p. 147
5 EKSTROM, A., and LISS, G.: 'A refined h.v.dx. control system', Trans. IEEE,
1970, PAS-89, (5/6)
6 RUMPF, E., and RANADE, S.: 'Comparison of suitable control systems for
h.v.d.c. stations connected to weak a.c. systems. Part I: New control systems.
Part II: Operational behaviour of the h.v.d.c. transmission', Trans. IEEE, 1972,
PAS-91, pp.549-64
7 JOTTEN, R., BOWLES, J.P., LISS, G., MARTIN, C.J.B., and RUMPF, E.:
'Control of h.v.d.c systems - The state of the art'. CIGRE, Paris, 1978, paper
14-10
8 CIGRE WG 14-07: 'Guide for planning DC links terminating at AC system
locations having low short-circuit capacities'. Part 1: AC-DC interaction
phenomena, 1992
9 BOWLES, J.B.: 'Control systems for h.v.d.c. transmission'. Report to CEAHVDC Subsection, Edmonton, 1975
10 CHAND, J., RASHWAN, M.M., and TISHINSKI, W.K.: 'Nelson River HVDC
system-operating experience'. IEE Conf., Publ. 205 on 'Thyristor and variable
static equipment for A.C. and D.C. transmission' (London, 1981), pp.223-26
11 O'BRIEN, M.T., FLETCHER, D.E., and GLEADOW, J.C.: 'Principal features
of the New Zealand DC hybrid link'. International colloquium on HVDC and
FACTS, Wellington, 1993, paper 1.3-1
12 AINSWORTH, J.D.: 'Telecommunication for h.v.d.c.'. IEE Conf. Publ. 205 on
'Thyristor and variable static equipment for A.C. and D.C. transmission'
(London, 1981) pp. 190-93

Chapter 6

Interaction between AC and DC systems

6.1 Introduction
A DC link can be operated according to the basic control modes described
in Chapter 5, and thus remain passive to any special needs of the interconnected AC systems. Alternatively, the link can be provided with more
adaptive dynamic controls which respond to critical deviations from the
expected operating conditions in the AC or DC systems.
The exclusive use of the basic controls often gives rise to unwanted
interaction between the AC and DC systems, which is manifested in a
variety of voltage, harmonic and power instabilities. When full advantage
is taken of the fast and adaptable converter controllability, a more useful
interaction can be achieved which manifests itself in stable AC and DC
system operation.
AC-DC system interactions are concerned with voltage stability, overvoltages, resonances and recovery from disturbances:

voltage-stability criteria are used to determine the type of voltage


control and of reactive-power supply;
the level of temporary overvoltage (TOV) influences station design,
including thyristor valve and surge arrester ratings; the lower the value
of SCR the higher the potential value of TOV;
shunt capacitors are used in converter stations as part of the AC filters
and VAR compensation; the larger the ratio of shunt capacitor MVAR
to AC-system short-circuit MVA, the lower will be the resonant
frequency;
recoveries from AC and DC faults are slower with weak systems (i.e.
high impedance sources), although modern controls are less affected
by the system impedance than those used in earlier schemes.

Thus, the degree of interaction depends on the relative strength of the


AC-DC system, a concept that must first be clarified.

130

High voltage direct current transmission

6.2 System strength definition


A CIGRE document released in 19921 provides valuable assistance to the
understanding of these interactions. Their influence on station design and
performance is assessed with reference to the AC-DC system strength, a
relative term which is generally expressed by the short-circuit ratio, i.e. the
ratio of the AC-system short-circuit capacity to DC-link power. A strong
system is defined as having an SCR above three, and the SCRs of weak
and very weak systems range between three and two and below two,
respectively.
However, the SCR concept has to be used with caution when considering
different types of interaction, because:

neither the short-circuit capacity nor the DC-link power remains


constant for a given scheme;
the system impedance calculated from the short-circuit capacity does
not include the filters and other shunt components connected at the
converter terminals;
the calculated system impedance at fundamental frequency does not
increase linearly with frequency;
the short-circuit capacity normally defines the transient impedance (for
use in fundamental-frequency overvoltage assessment) or the subtransient impedance (for use in harmonic studies);
the presence of other converters or nonlinearities in the AC system is
not taken into account in the derivation of the SCR.

Two important alternative definitions are the effective (ESCR) and critical
(CSCR) short-circuit ratios.
The ESCR allows for the presence of AC filters and shunt capacitors at
the converter terminals, as shown in Figure 6.1a, and is therefore defined
as

The CSCR corresponds to the operation at the maximum available power


(MAP), described in Section 2.9, and a typical value for the inverter is two.
For SCR values below the CSCR the operation is in the unstable region of
the AC voltage versus DC power characteristic.
Other, less utilised, expressions defined in the CIGRE document are the
operating short-circuit ratio (OSCR) and the (^effective short-circuit ratio
(QESCR).

6.2.1 AC-system Thevenin equivalent


With reference to Figure 6.1a, the various impedances connected to the
converter station can be combined into an equivalent Thevenin source

Interaction between AC and DC systems 131


-o

ZS

Id

-o
(a)

l=lp-jlt

Figure 6.1

AC system representation at the converter busbar

a Simplified diagram showing shunt capacitors (C) and synchronous


compensators (SC)
b Thevenin equivalent circuit

impedance Zsti for the purpose of defining the strength of the AC system
as shown in Figure 6.1&.
Zst is the equivalent of Zs (source), Xsc (synchronous compensators), Xc
(filters and static compensation) and Z/ (local load) in parallel. Th^ equivalent Thevenin source voltage, Est, results from the vectorial addition of
Vc/yj3 (line to neutral) and LZst, as shown in Figure 6.2.
As the DC link is integrated into the AC-power system, it is convenient
to express the converter equations in the same per-unit system as for the
AC network.
A convenient factor in the per-unit notation is the converter rating, r,
defined as the ratio of converter MVA to DC power (P^), i.e.
(6.1)
and substituting eqn. 2.14
r=(3y/2/n)Vc/Vd

(6.2)

It is standard practice to refer the parameters of a power-plant component


to its rated power and voltage. In the case of the converter bridge we will
use as reference base parameters the converter transformer MVA and

132

High voltage direct current transmission

(a)

Figure 6.2

Voltage regulation
a Rectifier end
b Inverter end

nominal voltage. The base impedance is therefore given by


(6.3)

If MVAp is the short-circuit level of the AC system at the converter terminal, the system impedance Zst can be expressed as
Zst=V2c/MVAF

(6.4)

and in per unit


zst = Zst/ZBASE = (VV MVAF)(M V A / V2C) = M VAC/MVAF

(6.5)

Using the converter rating factor, r, the per-unit system impedance can be
expressed as
zst = rPd/MVAF = r/ESCR

(6.6)

6.3 Voltage interaction


The AC-terminal voltages at the converter stations depend on the active
and reactive-power characteristics of the converter. To minimise voltage
variations it is essential to control the reactive-power supply to match the
converter reactive-power demand.

Interaction between AC and DC systems 133

It has already been explained in Chapter 2 that an HVDC transmission


link consumes reactive power at both ends, which can be typically 60 % of
the power transmitted at full load. Moreover, during transients the reactivepower demand may vary widely, the duration of such variation depending
to a large extent on the characteristics of the DC-link control system.
With the combination of rectifier constant current and inverter constant
extinction angle controls, both backed by transformer onload tap changers,
the reactive against active-power variation is nonlinear, as shown in Figure
2.13. The minimum values of firing delay, extinction angle and commutation reactance, however, are determined by other considerations such as the
need to reduce the risk of valve maloperation and to limit the valve stresses.
If the presence of local generators can be guaranteed (e.g. close to a
rectifier) it is always more economical to supply most reactive power from
these, with minimum-size filters to reduce harmonics.
For other cases it may be necessary to provide full reactive-power
compensation of the converter, sometimes with extra compensation for
AC-system loads. A large proportion of the reactive power is supplied by
the shunt harmonic filters, and additional compensation has so far been
provided in the form of shunt capacitor banks or synchronous compensators. The filters and capacitors are normally connected to the high-voltage
terminal, whereas the synchronous compensators are connected through a
transformer tertiary winding. Other permutations have also been used,
e.g. filters on a tertiary winding and synchronous compensators on separate transformers.
The size of the individual filters is the result of a compromise between
economy (which demands the larger size) and the ability of the AC system
to accept the step changes in system voltage caused by filter switchings.
Switching is often needed to control reactive power at DC loads lower than
the nominal. There is a further difficulty in discrete filter switching, which
relates to the nonlinear relationship between the increase of reactive power
and harmonic current with load.
If the extra cost can be justified, the use of dynamic compensation, in
the form of either synchronous compensators or controllable static
compensators, can reduce or eliminate the step switching of filter
branches. This is discussed in more detail in a later section.

6.3.1 Dynamic voltage regulation


During HVDC-link disturbances the voltage-control requirements depend
on the nature and location of the disturbance. The reactive-power
consumption, although possibly higher initially, is partially or totally eliminated following the disturbance, with the result of considerable dynamic
overvoltage regulation.
With reference to the voltage diagram in Figure 6.2, let us consider the
effect of total load rejection (i.e. / = 0) at the rectifier and at the inverter

134

High voltage direct current transmission

ends. The maximum voltage regulation will occur if the disturbance takes
place when the phase angle of the converter (c|)) is equal to the phase angle
of the equivalent AC-system impedance (4>.^), and is calculated as follows2
(6.7)

Est=V+LZst
Dividing throughout by the base voltage
Est/Vc=V/Vc+ZJ/Vc

(6.8)

and substituting eqns. 6.3 and 6.6


esi/y/S = (vA/3) + (// Fc)(r/ESCR)( V2C/M VAC)

(6.9)

or
est=v+r/ESCR

(6.10)

With an ESCR of 2.5 and under rated conditions


(6.11)

est=l+0Ar

In order to calculate r, let us refer to the DC-voltage equation (eqn. 2.12),


i.e.
Vd = Vc0 cos a - (2Xc/n)Id

(6.12)

which can be more conveniently written as


(6.13)

= cos a Substituting Vc0 = (3^2/n)Vc and eqn. 6.2 yields


- = cos a - XcId/(j2Vc)
r
The commutation reactance in per unit is
xc=Xc/[Vc/(j3l)]

(6.14)

(6.15)

and since / = (y^6/7c)/^, eqn. 6.15 can be changed into


XcId/(j2Vc)

= (n/6)xc

(6.16)

which substituted in eqn. 6.14 finally produces


- = cos a
xc
(6.17)
r
6
Thus, for a system having a commutation reactance of 0.2 p.u. (to the base
of transformer MVA), and a rated firing angle a (or y) of 15, the value of
r in eqn. 6.17 is approximately 1.15, which substituted into eqn. 6.11
results in a temporary overvoltage equal to 1.46.
Under these particular severe conditions a regulation of 46 % can therefore be expected.

Interaction between AC and DC systems 135

In order that (|> = fyst (at a rectifier) under rated conditions, either the
phase angle of the system impedance needs to be of the order of 30, an
unusually low value, or the rectifier should operate at an unusually high
firing angle. Typical impedance angles of the order of 70-85 are found
in practice; the more resistive the AC system appears at the rectifier end,
the higher regulation overvoltages may be anticipated.
At the inverter, however, it can be seen (Figure 6.2 b) that under normal
operating conditions <fyst is considerably less than cj). Therefore, the more
reactive the AC system appears, the higher the regulation overvoltage.
With an impedance of 0.4 p.u., a maximum value of est - 1.29 p.u. can be
anticipated, thus causing a 29 % regulation, following full-load rejection;
however, the expected impedance angle in practice is of the order of 70.
The regulation (dynamic) overvoltages are therefore more significant at
the rectifier end. At both terminals the effective impedance angle is as
important in determining the overvoltages as is the magnitude of the
impedance. For links from hydro sources, the increase of frequency
following load rejection will produce even higher dynamic overvoltages.
This is an unacceptable situation for local consumers and must be allowed
for in the insulation co-ordination of the converter station.
In practice, transformers start to saturate at typically 1.2 to 1.25 p.u. AC
voltage and the fundamental-frequency overvoltage will therefore be a
little lower, with some distortion.
Single line-to-ground faults are also a source of dynamic overvoltage on
the other phases or pole. As a result of an AC phase-to-earth fault, the
mutual coupling between phases causes a voltage increase in the other
phases; for a network effectively earthed the overvoltage is limited to a
peak value of ^/2(0.8) of the line-to-line r.m.s.
On the other hand, following a voltage drop in the AC network, the
initial effect is a fall in power. The power controller of the DC link then
increases the current reference to try and restore the ordered power; the
extra current increases the reactive demand and tends to reduce the ACsystem voltage further. With very weak AC systems this could lead to
voltage collapse; however, power controllers always have built-in limits to
avoid excessive action.
By far the most important case is that of a nearby three-phase short
circuit, assuming that the converters are blocked permanently during the
fault, with all the capacitors left on. This condition produces full magnetising inrush current on all transformers after fault removal, which results in
substantial fundamental and harmonic overvoltages. Such overvoltages
constitute in practice the determining condition for most valve, surge
arrester and insulation voltage ratings.
Dynamic-compensation equipment is used to reduce the dynamicvoltage regulation, to help in the recovery of the AC system from faults
and to reduce the disturbances resulting from DC-load variation or from
the switching of filter banks.

136

High voltage direct current transmission

Ideally, to meet such a comprehensive range of duties, the compensation


will generally need to have both reactive-power absorption and generation
capabilities. In each particular application, the system short-circuit ratio
and the stability of the DC link and compensator controls must be considered, when trying to decide the type and dynamic range to be used.
A preliminary comparison of the technical characteristics of reactivepower compensators for use with HVDC transmission, with an extensive
bibliography on the subject, has been presented at CIGRE.3
The main types of dynamic compensation already used, or under
consideration, for HVDC schemes are: synchronous compensator, AC selfsaturated reactors, thyristor-controlled reactors and thyristor-switched
capacitors.
The slow response of synchronous compensators may be a problem
particularly in the absence of local generation. However, the synchronous
compensator reduces the sensitivity to transients by increasing the SCR. It
also increases the resonance frequency of the system, since it reduces the
need for extra shunt capacitance.
It is often claimed that the last three alternatives (i.e. the static ones)
improve the voltage stability of the AC network and thus help HVDCcontrol stability and speed of response. In practice, the so-called fast static
compensators rarely improve voltage stability, except that they act as limiters if the voltage tries to rise too much. In the steady state the practicable
thyristor-controlled reactors appear to cause some destabilisation, although
this can be made acceptable; saturated reactors appear to be better in this
respect.
The other main problem of the fast static compensators appears when
the system e.m.f. falls even slightly; in such a case, both the thyristorcontrolled and the saturated reactors tend to go to a zero-current condition, and stability then depends only on the bare AC system plus all capacitors. Normally the latter account for 0.6 p.u. per converter plus an extra
0.4 p.u. for load compensation, and under these conditions the net ESCR
could be very low (e.g. 1.5), thus making stability difficult.

6.4 Dynamic stabilisation of AC systems


A power system is stable if after a disturbance it returns to a condition of
equilibrium. This is manifested, not by the constancy of absolute rotational
speed of the various machines involved, but rather by these machines
swinging together until a new common speed is reached. The power
exchanged between them is determined by their relative angular position
and, therefore, when the equilibrium is disturbed, their rotor positions
must give rise to corrective power flow leading to the new state of equilibrium.

Interaction between AC and DC system

137

If the angle between the machines increases steadily, the system is transiently unstable. If the machines fall out of step after a period of increasing
oscillations around the equilibrium point, the system is dynamically
unstable. Dynamic instability is rare in tightly connected systems, which
are usually well damped for their characteristic frequencies of electromechanical swing (between 1 and 2 Hz).
However, when large systems are connected by long relatively-weak
interties, low-frequency swing modes result. T h e response of the powersystem controls to the synchronising swings associated with these lowfrequency modes can produce sufficient negative damping to cancel the
natural positive damping of the system. When this happens, oscillations of
increasing amplitude occur.
An example of dynamic instability4 is the northern and southern parts
of the Western US power system, which are connected by the parallel
Pacific AC and DC Interties with ratings of 2500 and 1400 MW, respectively. The AC Intertie has a long history of negatively-damped 1/3 Hz
oscillations resulting from interactions between generators with automatic
voltage regulators and system loads. As a result of these oscillations, and
because the oscillatory tendency imposed a constraint on the amount of
surplus north west hydro power which could be transmitted to the south
west, a control system to modulate the Pacific DC Intertie was developed.
Damping in the Pacific Intertie is produced by small-signal modulation
of the DC power in proportion to the frequency difference across the AC
Intertie. This was accomplished by processing the AC lntertie power
measured at the northern end, to obtain a filtered signal proportional to
the derivative of AC: power at frequencies near 1/3 Hz. This signal is
applied, through a & 3 016 (+ 40 MW) limiter, to the current regulator at
the northern terminal of the DC Intertie; thus the current setting changes
are well within the current margin.
Figure 6.3 shows the results of field tests with and without modulation,
wherein series-capacitor compensation was First switched in and then
bypassed. It is also possible to use AC current, rather than power, as the
modulation (error) signal. Current is more linear with respect to large
swing angles, hence will be more effective when AC-system oscillations
approach stability limits.
Successful operation of DC modulation was a key factor permitting
uprating of the AC Pacific Intertie from 2100 MW to 2500 MW.
Detecting the effect of the modulating frequency he. the l/S Hz harmonic power in the case of the Pacific Intertie) is less effective than detecting
absolute phase change directly (a method used in the Nelson River
scheme); power measurement provides a signal which levels off near 90" of
the phase difference between the e.m.f.s 6.e. right where most response is
urgently needed to try and prevent pole slipping). In fact, in most
machine-swing problems, the net peak survival angle is about 130" and yet
above 90" the power-measurement method actually gives a reduced output.

138

High voltage direct current transmission


With modulation

1 25 MW

3 seconds
D.C. power
A.C. power

Without modulation

iA/v

T50MW

T5OMW

A.C. power

Figure 6.3

System response to AC-intertie series capacitor bypass

6.4.1 Large-signal modulation


Although the small-signal modulation described in the last section is suitable for maintaining the state of equilibrium, it is inadequate for the
damping of large disturbances.
Large-signal modulation is thus needed to regain the equilibrium state
following large disturbances. A large-signal modulation scheme4 has been
added to the Square Butte HVDC system in the form of a frequency-sensitive power control (FSPC).
A block diagram of the scheme is shown in Figure 6.4. The frequency
deviation (A/) at the rectifier station is first filtered to eliminate torsional
interaction at 11.5 Hz of the 400 MW dedicated generating plant. The
output of the stabiliser (A/) is limited to 20 % of the rated DC-link current
(1000 A); the scheme has been designed to cope with a 20 % transient overload.
In order to maintain the current margin between the two ends, the
output of the FSPC is also communicated to the inverting station by means
of a microwave link using parallel tone channels; each tone channel
communicates a 50 A (0.05 p.u.) step change in current order.

6.4.2 Controlled damping of DC-interconnected systems


With an AC-tie line, if one of the interconnected systems is in difficulty
following a disturbance, the line is normally tripped to prevent the distur-

Interaction between AC and DC systems 139


Current order (limited to 1.1 p.u.)
Limiter
Signal conditioning
network (FSPC)
transfer function

Frequency
detector

A/

0.2

Rate
limiter

Limiter
1.3

1 p.u y
-0.2

/so

_j0.2

' p.u. To current


regulator

ms

Digitizer/ limiter

Rectifier
(center terminal)
A/"

7-

0.05

16.7 ms

Inverter
(arrowhead
terminal)

Figure 6.4

Communication system
delay

1 p.uy
/50

Im
current margin
(0.1 p.u.)

ms
Rate
limiter

p.u. I

_j0.15

To current

Limiter

Square Butte FSPC block diagram

bance affecting the other system, and thus the system in difficulty loses an
essential infeed.
An HVDC link, on the other hand, even with the basic controls, shields
one system from disturbances on the other. Although the specified power
flow can continue, the option is available to vary the power setting to help
the system in difficulty to the extent which the healthy system can allow,
without putting itself in difficulty, and subject to the rating on the link.
Although the policy of providing controls which enable the HVDC link
to assist actively in the damping of disturbances should be encouraged, it
must be considered that the DC link contains negligible energy storage
and therefore any action to damp a disturbance at one end must naturally
produce some disturbance at the other. In some cases such assistance is
readily acceptable, for example when the local system has no directlyconnected consumers and it can be designed for greater than normal
frequency variations. In effect, this allows the inertia of the system to be
used to provide the energy for damping the distant system. Another
example is where one system is very small as compared with the other,
such as the case of an off-shore system where the total load is insignificant
compared with the size of the mainland network.
With appropriate control, a disturbance originating in either system can
be shared in a predetermined manner, and the resulting system oscillations
can be damped simultaneously. Unlike transient stability, where the DC
link must have the necessary overload capability to get through the first

140

High voltage direct current transmission

swing, dynamic stability can be achieved without overloading the DC link.


Moreover, if the DC link is already operating close to its full capacity,
substantial damping can be achieved exclusively by DC-power reductions
at the appropriate instants.5

6.43 Damping of subsynchronous resonances6


The torsional oscillation modes of turbine generators can interact with the
system oscillation modes of the electric power-transmission system. With a
pure AC transmission system, the interaction is dominated by a synchronising component which gives rise to the system modes of oscillation. A relatively small negative-damping component is also present, owing to the
resistance of the AC transmission line.
The torsional modes of vibration of the turbine-generator shaft are
normally stable when connected to an AC transmission system, because of
the relatively large positive damping contributed by the damping windings
and mechanical damping resulting from steam flow, friction etc. With the
addition of series-capacitor compensation in the AC network, however, the
negative-damping contribution of the AC system is dramatically increased
when the electrical and mechanical resonant frequencies are close. Under
these conditions the torsional modes of vibration can become unstable, a
phenomenon which is commonly referred to as subsynchronous resonance
(SSR).
When the turbine generator is connected to an HVDC system, a slightly
different situation exists. In the absence of control the HVDC system would
appear as a load on the turbine generator which has a positive-damping
characteristic. In practice, the presence of the current-control loop changes
the impact of the DC system to one of negative damping. This effect can be
better understood by considering the case where the DC line has no resistance, and the inverter is connected to a system having infinite capacity.
The inverter voltage will then be constant, and the use of constant-current
control will result in a constant-power load on the turbine generator. A
constant-power load can be shown to have a pure negative-damping characteristic. The presence of a parallel AC transmission system will decouple
the turbine generator from the influence of the HVDC system, and will
allow the generator damper windings to add more positive damping.
The negative-damping contributions owing to current or power control
with the HVDC system only occur within the bandwidth of the regulators.
Typically, these control functions have bandwidths in the neighbourhood
of ten to 20 Hz. At frequencies above the bandwidth, the DC system
approaches the situation of no control, which provides positive damping.
Hence, only the torsional modes of vibration having frequencies below
about 20 Hz will, in general, have the potential for adverse interaction
with the DC system. This must be treated only as a typical conclusion,
however, since individual systems may have wider bandwidths.

Interaction between AC and DC systems 141

This interaction occurred at the Square Butte project through two


different control paths; the auxiliary-damping control, discussed earlier,
and the current control.
In the case of the auxiliary-damping control, the interaction was due to
high gain and phase lag in the lower range of the subsynchronous
torsional-frequency region. Thus, an effective method of eliminating
interaction through this control path was by notch filtering the auxiliary
damping control at 11.5 Hz.
The torsional instability due to current regulator response was more
difficult to solve owing to its inherent negative electrical damping which
cannot be eliminated. However, it was possible to minimise the magnitude
of the interaction at the frequencies at which the negative damping
occurred.
The potential destabilisation of torsional oscillations due to HVDC
systems is similar to that caused by series-compensated AC transmission
lines. However, the interaction with DC systems can be solved relatively
simply by providing power-modulation control to cancel the negativedamping impact of the basic constant-power control loop. Series compensation, on the other hand, will require special equipment such as series
blocking filters and shunt-reactive control devices.7
A subsynchronous damping-control (SSDC) concept developed under
EPRI RP1425-1 consists of a wide-bandwidth controller sensitive to
generator speed if available, or to the frequency of an AC signal synthesised from voltage and current measured at the DC converter terminal.
With such a design, the SSDC can be made to provide a positive damping
contribution over the entire range of subsynchronous torsional frequency.
Another method of damping torsional oscillations is to modulate the
firing instant with a suitable phase and amplitude with regard to the
generator oscillations. The input signal for this modulation can be derived
either from the speed of the generator or from the bus frequency. When
using the generator speed as the input signal, the only required signal
processing is to pass the signal through a bandpass filter to eliminate the
low intermachine-oscillation frequencies and the high-frequency noise.
Considerable positive-damping contributions may be achieved in the
subsynchronous torsional-frequency range with such a damping method.
However, in some applications it may not be possible to use such a wide
bandwidth damping controller. In these situations, it might be possible to
focus on one torsional mode at a time by using bandpass-filtering techniques.

6.4.4 Active and reactive-power co-ordination


The degree of DC-power modulation that can be achieved is restricted by
terminal reactive-power constraints. With only current or power modulation, an increase in active-power transfer will be accompanied by a larger

142

High voltage direct current transmission


\t

A P.,

Modulation
controller

\Elt

AP,,

Band pass
f.lter

Figure 6.5

P,<
control

control

E,,
control

Band pass
filter

r-

Generic AC-DC system-controller structure showing the modulated


converters and a simple interarea AC system; bandpass filters are
indicated in each modulation channel

increase in terminal reactive-power requirements (see Figure 2.13) and this


effect is particularly noticeable during severe system disturbances. The
reactive-power variations can cause current-control mode transitions
between the rectifier and inverter ends, and hence DC-current changes
equal to the DC-system margin current.
Co-ordination between the active and reactive-power modulation can be
achieved by DC-system voltage modulation. An increase in DC voltage will
increase the DC-power transfer as well as the power factor at both terminals, and hence decrease the reactive consumption as a percentage of
active power transmitted.
A general AC-DC system configuration8 is shown in Figure 6.5. Each
terminal is provided with an independent control loop, shown as a current
controller at the sending end and as a voltage controller at the receiving
end. The settings of these control loops are fixed by the DC modulation
controller, the inputs of which can be either the bus frequencies or ACline power flows.

6 A. 5 Transient stabilisation of AC systems


Where system disturbances result in the reduction of transmission capability, the generating source will usually accelerate. Remote sources may
decelerate as load exceeds generation as a result of a fault which decreases
power into that area. When the fault is cleared, the generation and the
remaining transmission experience a transient swing which may lead to

Interaction between AC and DC systems 143

instability. In particular, long fault-clearance times can cause a loss of


synchronism.
If the loss of synchronism is irrelevant, as in the case of an HVDC link
connecting generation to load areas, it is advantageous to increase the
sending-end DC-link power in the post-fault period in response to the
increase of generator speed. This action will remove energy from the
generator, reduce its speed and thus reduce the angular displacement
between the generator and the AC receiving system. The suggested magnitude of the modulation applied for this purpose is in the range of 20 to 40
% of the DC-link rating. Some systems have been designed with temporary overload limits as high as 65 %; in other cases even higher modulation
limits have been utilised after taking into account the AC-system powertransfer need, the AC-voltage support (VAR) capability and the DC-system
design ratings. Similarly, for receiving-end phase angle or speed changes,
DC-link power can be controlled to correct this condition within the limits
imposed by the controllability of the receiving-end phase angle, the DClink capability and the energy that may be taken from the generation
source.
The thyristor valves used in HVDC transmission are rated to withstand
considerable overloads without adverse effects to avoid unnecessary
protective action. This capability provides the basis for first-peak transientstability improvement. The particular strategy, i.e. current increase or
decrease, temporary power reversal etc., will vary from scheme to scheme,
and in each case it can be assessed with the help of a multimachine transient-stability programme combined with a small-step transient-converter
simulation programme,9 such as the EMTDC. The two solutions are
carried out separately and compared periodically to ensure agreement on
interface quantities. The time steps required for the DC solution are far
shorter than those that would be used for the normal transient-stability
solution, and therefore during this calculation a Thevenin-equivalent AC
network is used.

6.5 AC-DC frequency interactions 10


The mechanism of AC-DC harmonic interaction has already been
discussed in Chapter 2 and is summarised in Figure 6.6. The presence of a
harmonic distortion at k times fundamental frequency on the DC side of a
12-pulse HVDC converter will produce on the AC side, positive-sequence
harmonics of orders A + 1, 13 + A, 25 k, 12^+1 k; n= 3,4,..., and
negative-sequence harmonics of orders k- \, 11 + k, 23 + k, 12n - 1
k; n = 3,4,... These harmonic sequences are reflected back to the DC side
as the kh harmonic and various high-order harmonics of 12 k, 24 k,
I2n k; n = 3,4,... Among these harmonics, the most significant terms
are the first-order ^th harmonic on the DC side, and the positive-sequence

144

High voltage direct current transmission


AC side

DC side

positive sequence harmonics


/c+1
13/c
+
25/c
12n+1 k, n = 3,4,...

negative sequence harmonics


/c-1
11 k
<
23/c
12/7-1 k, n = 3,4,-

Figure 6.6 AC-DC harmonic interactions

k + 1 and negative-sequence - 1 harmonics on the AC side. The higher


harmonics are an order of magnitude smaller than the lower-order harmonics. Therefore, for most analyses, particularly those with small distortion
levels, it is reasonable to ignore the contribution from high-order harmonics. Figure 6.6 gives a representation of the harmonic orders which can
be expected on each side of an HVDC converter.
With the high-order harmonics ignored, the presence of a secondharmonic distortion on the DC side will result in a positive-sequence thirdharmonic and a negative-sequence fundamental-frequency component on
the AC side.
The interaction mechanism can be extended to nonharmonic frequencies. For instance, if there is a distortion near the fundamental frequency
such as at 51 Hz on the DC side, the distortions on the AC side would be
near the second harmonic (101 Hz) for the positive-sequence component
and near DC (1 Hz) for the negative-sequence component.
As the frequency of the DC-side distortion approaches fundamental
frequency, the lower corresponding frequency component on the AC side,
which is in the negative-sequence format, will be approaching 0 Hz, i.e.
approaching DC.

6.6 Harmonic instabilities


Harmonic instability at a converter terminal can be defined as the generation and/or magnification of noncharacteristic frequencies by a DC system
containing, initially, no unbalance or asymmetry.

Interaction between AC and DC systems 145

AC-DC systems with low short-circuit ratios (SCR) often experience


problems of waveform instability. The low SCR indicates a high ACsystem impedance, with an inductance which may resonate with the reactive-compensation capacitors and the harmonic filters installed at converter terminals. These resonant frequencies can be low, possibly as low as
the second harmonic. The resonances can be excited under certain operating conditions or in the event of fault, and the small initial distortion
may develop to an instability. Instability related to the interaction of
harmonics (or any frequencies) has been customarily referred to as
harmonic instability.
The problem of harmonic instability in AC-DC systems was first identified in relation to the individual valve-firing control11 and an alternative
principle, equidistant-firing control,12 was adopted for new installations.
Since then, other forms of harmonic instability have been identified, involving complementary resonances, composite resonances, crossmodulation
and transformer core saturation.

6.6.1 Instability caused by individual firing control


This section reproduces the basic analysis and conclusions contained in a
classic paper by Ainsworth.11 Using simple analysis, based on one converter and one harmonic, and assuming linear behaviour for the small levels
of harmonic-voltage distortion involved, the paper demonstrated the existence of a problem which dramatically changed conventional control
philosophy.
In the circuit of Figure 6.7 a spurious harmonic e.m.f. of magnitude En
and harmonic order n (of positive or negative sequence) is added to the
voltage source, Est.
The AC-system impedance is assumed to be Zn at the harmonic
frequency of order n and zero at other frequencies. The DC-reactor inductance is assumed infinite, i.e. the DC-line current is constant.

Figure 6. 7 Equivalent circuit per phase for AC system of finite impedance

146

High voltage direct current transmission

The harmonic voltage Vn on the converter AC busbars then depends on


impedance Zn and its phase angle, (j)m9 harmonic order, n, firing angle, a,
and relative phase, 9, of the interfering e.m.f. The latter is random and
therefore only the maximum Vn as 9 varies is of importance. The harmonic magnification due to converter operation is
M=VjEn
Neglecting transformer saturation effects, M is given by the following
equations (for a constant-a control system)
(a)

orn=6p-2
Mmax = 1 + x cos (4>sn - na) + \x2 - x
+ xcos(</>m- na)+-\\
4 JJ

(6.18)

where
(6.19)
(b) orn=6p+2
Mmax = jl + x cos {(j)sn -m)+-\~/2

(6.20)

In both cases the voltage magnitude per phase is the same.


For n = 3jfr, both sequences (orders +n and -n) are obtained in busbar
voltage and current. Phase voltages are then unequal, but the maximum M
for the highest of the three-phase voltages is the same as in the first equation.
For high Zn, high magnifications can occur. In the worst cases, Zn
greater than some critical value gives infinite magnification, i.e. instability.
In these cases, it is obvious that the instability can be initiated from any
asymmetry, however small, whether originating in the AC system, as postulated, or elsewhere, even with a perfectly adjusted control system. In practice, Vn will not rise indefinitely but will be limited at a considerable value
by nonlinearity, which is ignored in this analysis. An example of harmonic
instability obtained in a physical model is illustrated in Figure 6.8.
In an HVDC converter, although shunt harmonic filters are normally
provided to reduce total AC-system impedance to low values at normal
harmonic orders (5, 7, 11, 13 etc.), it is not economical to provide filters
for the intermediate abnormal harmonics, and the total Zn may be high
for these. The normal harmonics may, in any case, be readily shown to
produce the same change of firing-pulse time for every valve but not to
affect their relative spacing, and hence may be ignored.

Interaction between AC and DC systems 147

ib)

Figure 6.8 Model test on six-pulse converter with realistic AC system (short-circuit
ratio = 3) and conventional harmonic filter: constant-en control
system, a = 32; control-system-filter Qfactor = 2.5
a AC-line voltages
b DC-bridge voltage
It is, however, clear from the equations above that if Zn (in per unit) can
be guaranteed to be less than
0.5 for n = 2, 4, 6, 7 etc.
1.0 for n= 3, 9, 15 etc.
there is no danger of instability, although magnifications may still occur.
This can, in principle, be achieved by the addition of extra shunt filters to
the main circuit, but this is usually very expensive.
If a shunt filter is used for the normal harmonics, i.e. orders 5, 7, 11, 13
etc., partial resonance may occur between the AC system and the filter at
the abnormal harmonics; the requirement then virtually implies that the
conductance component of the AC-system harmonic admittance must be
greater than 2 p.u. or, expressed in another way, the AC harmonic impedance on a polar diagram must be within a circle of radius 0.25 p.u.
centred on (0.25, 0).
Except for the case of a relatively small converter connected to a large
AC system, this is found to be a somewhat stringent requirement, and in
certain typical cases the calculated maximum circle radius has been about
1-2 p.u., implying stable operation only up to 0.125-0.25 of rated current,
with the worst combination of a and system-impedance angle.
The main practical effects of large harmonic magnification or instability
are:
(a) Excessive harmonic voltages and currents in the AC and DC systems.
Instability owing to even one system resonance, in general, produces
distortion containing all harmonic orders, i.e. 2, 3, 4 etc. Usually, any
local overvoltage resulting from this is small but interference elsewhere may be unacceptable.

148

High voltage direct current transmission

(b) Operation approaching full inversion may be impossible owing to


continuously repeated commutation failures.
It should be emphasised that, as shown by the equations, operation is critically dependent on a; a converter which shows harmonic instability at one
value of a owing to an AC-system resonance, may be stable with only small
distortion at a value of a different by only, say, 20, depending on the
frequency of the resonance.
In many cases it will be desirable to take steps to prevent the magnification or instability discussed. The method used in early schemes was the
addition of a control-system filter in the supply of AC-timing voltages
from the main system to the control system. However, this method by
itself did not eliminate harmonic instabilities and extra expensive filters
had to be added in some schemes.
The constant-extinction angle control systems are analogous to the
constant-a control system but they, in effect, control to a constant negative
time delay (the extinction angle y), measured back from the instant corresponding to the second zero crossing (at a = 180) on the appropriate ACline voltage waveform.
Thus, harmonic magnification or instability can occur with constant
extinction angle control systems, as for constant-a systems, if the ACsystem harmonic impedance Zn is high. The results can, in practice, be
worse, since practical predictive systems can only operate on smooth waveforms; with badly distorted waveforms, they will make many mistakes
which, in the worst cases, produce continuous-commutation failures and
collapse of operation.
It appears to be an inescapable conclusion that purely predictive systems
are not workable for constant-extinction-angle control of inverters on very
high-impedance AC systems. Some improvement is, however, obtained by
the use of control-system filters.

6.6.2 Composite resonances13


The term resonance is most often used with reference to isolated parts of
an overall system, usually being either the AC or the DC system. This sort
of electrical resonance is well defined, being the frequency at which the
capacitive and inductive reactances of the circuit impedance are equal. At
the resonant frequency, a parallel resonance has a high impedance and a
series resonance has a low impedance. This approach has led to the
concept of a complementary resonance; a high-impedance parallel resonance on the AC side coupled with a low-impedance series resonance at an
associated frequency on the DC side.
When the DC and AC systems are interconnected by a static converter,
the system impedances interact via the converter characteristics to create
entirely different resonant frequencies. The term composite resonance is

Interaction between AC and DC systems 149

used here to describe this sort of resonance, emphasising its dependence


on all the components of the system. A special case of composite resonance, involving a converter-transformer core-saturation contribution, is
discussed in Section 6.6.3.
A composite resonance may be excited by a relatively small distortion
source in the system, or by an imbalance in the converter components or
control. The amplification of a small source by the resonant characteristics
of the system can present problems, and should be taken into account if
steady-state harmonic sources are expected.
Further to this, the converter impedance comprises several contributions. First, there is the AC side and converter transformer impedance,
which usually sums to be largely inductive. Secondly, there is the end-ofcommutation period dynamics, which is such that if the DC current out of
the converter increases, the DC voltage reduces. This impedance looks
mainly resistive. Finally, the constant-current control modifies the converter DC-terminal voltage according to the DC current. This can also be
described as an impedance although, over a range of frequencies, the
resistive component of this impedance will be negative.
A true instability results when, at the composite resonant frequency, the
resistance of the overall circuit is negative. This can occur at noninteger
frequencies, and is driven by conversion from the fundamental frequency
and DC components to the composite resonance frequency via the converter control. Light damping, or ringing, during fault recovery, indicates
that the negative resistance offered by the current controller is close to the
natural resistance of the circuit.
This instability is demonstrated using the CIGRE model HVDC link.14
The model was designed to present a difficult case for control, having a
parallel resonance at the second harmonic on the AC side, and a series
resonance at fundamental frequency on the DC side. The rectifier DC
terminals are chosen as the point where the system impedances are added.
Two alternative constant-current-control gains are used at the rectifier,
selected to have slightly negative and slightly positive damping factors.
The constant-current-control gain has a strong effect on the compositeresonance damping factor through its effect on the circuit resistance.
Table 6.1 gives the selected constant-current-control gains and the calculated composite-resonance damping factors. Rather higher gains than
would normally be employed are chosen to show light positive and negative damping.
Dynamic simulations are run for the two cases involving a three-phase
fault on the inverter AC busbar for 70 ms to excite the composite resonance.
The results of dynamic simulation runs are shown in Figure 6.9.
Example 1 demonstrates the composite instability predicted by the analysis,
and example 2 shows the lightly damped resonance, also predicted by the
analysis.

150

High voltage direct current transmission


Table 6.1 Constant-current-control gains for test case examples
Ex.
no
1
2

current, pu
1.10000.8800 0.6600 0.4400 0.2200
0.0000
0.0000
current, pu
1.39801.11840.83880.55920.27960.0000 0.0000
current, pu
1.3570-j
1.08560.81420.5428 0.27140.0000
0.0000

0.2000

Proportional
gain rad/kA

Integral time
constant

Composite resonance
damping factor

0.0714
0.9341

0.0093
0.0107

-5.1
+ 2.2

0.4000

0.6000

0.8000

1.0000

1.2000

1.4000

1.6000

1.8000 2.0000
time, s

1.2000

1.4000

1.6000

1.8000

1.2000

1.4000

1.6000

1.8000

Example 1

0.2000

0.4000

0.6000

0.8000

1.0000
Example 2

0.2000

0.4000

0.6000

0.8000

1.0000
Example 3

2.0000
time, s

2.0000
time, s

Figure 6.9 Rectifier DC-current transient responses

6.6.3 Transformer-core-related harmonic instability


Transformer-core saturation is a well known source of harmonic current.
The saturation can occur as a result of DC magnetisation or overexcitation.
Transformer textbooks usually refer to odd ordered harmonics being
generated, with predominance of third harmonic. They also explain that,
with balanced AC voltages, the third harmonic can be absorbed in a delta
winding.
The case of a converter transformer is very special in this respect
because the converter, under nonideal operating conditions, can produce
nonzero sequence triplen harmonics, even harmonics and direct current.
Transformer-core saturation often has an additional amplifying effect
for harmonic instability. There have been several reported incidences of

Interaction between AC and DC systems 151

core saturation instability, at the Kingsnorth,15 Nelson River,16 Blackwater17 and Chateauguay18 schemes. Despite these incidences, there is
little information on the nature of the phenomenon and this may have led
to some incidences being misinterpreted as another type of harmonic
instability or resonance. The control solutions are very similar, typically
involving some sort of firing-angle modulation and in some cases the
installation of additional harmonic filters.
The mechanism of the converter transformer-core saturation instability
phenomenon can be demonstrated using the block diagram of Figure
6.10. If a small level of positive-sequence second-harmonic voltage distortion exists on the AC side of the converter, a fundamental-frequency
distortion will appear on the DC side. Through the DC-side impedance, a
fundamental-frequency current will flow, resulting in a positive-sequence
second-harmonic current and a negative-sequence DC flowing on the AC
side. The negative-sequence DC will begin to saturate the converter transformer, resulting in a multitude of harmonic currents being generated,
including a positive-sequence second-harmonic current. Associated with
this current will be an additional contribution to the positive-sequence
second-harmonic voltage distortion and in this way the feedback loop is
completed. The stability of the system is determined by the characteristics
of this feedback loop.
In Figure 6.10 the aforementioned instability feedback loop does not
involve the entirety of the negative-sequence DC produced by the converter. This is because, owing to the dynamics of the instability, the DC-side
distortion is never exactly at the fundamental frequency and, therefore,
the negative-sequence DC is not a true DC but is varying slowly. The level

AC side

DC side

f
AC side
^
second harmonic \

positive sequence
second harmonic
voltage distortion

\^ impedance J

ideal transformer

positive sequence
second harmonic current distortion
multitude of
distortions at
many frequencies

transformer
core
saturation

\
majority of negative
sequence DC current
distortion

a small part of negative


sequence DC current distortion

Figure 6.10

converter
switching
action

converter
switching
action

-ideal transformer

Mechanism of core saturation instability

fundamental
frequency voltage
distortion
DC side
fundamental
frequency
impedance
fundamental
frequency current
distortion

152

High voltage direct current transmission

of DC component in the transformer-valve side current will be changing


and is in fact transferring itself between phases in a cyclic manner.
However, since the variation is near DC, the phrase 'negative-sequence
DC is used to refer to this extremely slowly varying distortion which is
oriented in a negative-sequence format as explained in the preceding
section. This variation, although sufficiently slow to cause transformer
saturation, is also sufficiently fast for a percentage of it to pass through
the transformer and into the AC system. The faster the variation of this
negative-sequence DC, the more of it will pass through the transformer
and the less of it will saturate the transformer, and vice versa. The portion
that passes through the converter transformer is distributed into the AC
network and may tend to DC bias other transformers in the system, but it
is unlikely to cause significant saturation to further contribute to the build
up of the instability. For accurate prediction of an instability, the division
of current into magnetising and transferred components is important, but
if only an indication of the risk of instability is required, it may be
neglected.
The onset of core-saturation instability is closely related to the saturation
level of the converter transformer. In this respect, the instability is broadly
divided into two categories, distinguished by their starting conditions. The
first type has a spontaneous nature as it develops under normal operating
conditions without any external stimulus. System imbalances or asymmetry
in the converter firings will result in low levels of transformer saturation
which can ultimately develop into an instability. Study of this type of
instability requires the evaluation of the transformer response at low
saturation levels. The second type is referred to as kick-started instability
which may see substantial transformer saturation as the starting condition.
Some disturbances may impress a high level of saturation on the converter
transformer and consequently result in the development of core saturation
instability after the disturbance. For this latter category of instability, the
transformer response at high levels of saturation has to be determined.
The techniques used to analyse this instability can be grouped into the
three categories of direct-frequency domain,17'19 iterative-frequency
domain20 and time-domain simulation.21
To demonstrate this type of interaction the behaviour of the CIGRE
benchmark model14 was analysed by PSCAD/EMTDC simulation with a
small fundamental-frequency modulation added to the rectifier firingangle order for 0.5 seconds. The growing distortion in DC current (shown
in Figure 6.11a) and the corresponding increase in the magnitude of the
negative-sequence quasi DC magnetising current content (shown in Figure
6.11 &) are a clear indication of an instability.
The converter transformer-core saturation instability can be prevented
by operating the system away from the unstable conditions or, in other
words, providing sufficient damping at the relevant frequencies. This type
of action may involve modification to the system impedances by filter

Interaction between AC and DC systems 153


1.25n
1.00
kA 0-750.500.250.00
0.5

DC current

(a)

1.0

1.5

2.0

2.5
time, s

3.0

3.5

4.0

4.5

magnitude of -ve seq. DC harmonic current distortion on / mag

1.0

Figure 6.11

1.5

2.0

2.5

3.0

3.5

4.0

4.5

EMTDC simulation results of the modified CIGRE HVDC benchmark model

retuning, or tuning of the converter-controller parameters or the converter


steady-state operating parameters. Although the purpose of the changes to
these parameters is to modify the system response at the frequencies
related to this instability, it usually also affects the system response at other
frequencies. The design of such preventative measures has to ensure that
other system requirements or constraints are still met after the modifications. These actions can be broadly regarded as passive measures.
On the other hand, active measures can be applied to stabilise the system
when the development of the instability is detected. This type of solution
has been used to prevent core saturation instability in existing schemes,
with some sensing instruments estimating the level of transformer saturation and appropriate action taken in accordance with the extent of the
saturation. Active measures should be designed to function at a certain
limited range of frequencies without altering the system response significantly under normal operating conditions. This will allow the system to be
operated as usual, but with the added security of some stabilising action
when instability is suspected.
Owing to the great differences in the characteristics of the various
HVDC systems, it is difficult to pinpoint which is the best solution to
counter this instability. The high dependency of the system stability on the
properties of the HVDC scheme suggests that the most appropriate solution for one system may not suit the others. Moreover, each HVDC
scheme usually has its own unique requirements and restrictions which
have to be fulfilled. It is therefore necessary to undertake independent
analysis for different systems or for a similar system under different operating conditions.

154

High voltage direct current transmission


DC current

1.25-]

1.00

KA 0.750.500.250.00
0.5

(a)
1.0

1.5

2.0

2.5
time, s

3.0

3.5

4.0

4.5

magnitude of -ve seq. DC harmonic current distortion on / mag

2.0

Figure 6.12

2.5
time, s

EMTDC simulation results with the addition of an auxiliary highpass filter to the converter controller

The addition of a high-pass filter to the converter controller has been


found to be effective for this particular test case, but had to be properly
tuned to avoid exciting a composite resonance at about 70 Hz.22 With the
high pass filter included the dynamic-simulation results of Figure 6.12
show that the instability has been eliminated.
A vulnerable HVDC rectifier system is likely to have the following impedance profile:

a low and predominantly capacitive DC-side impedance at the fundamental frequency with the presence of a series resonance near to but
higher than the fundamental frequency;
a high and predominantly inductive AC-side second-harmonic impedance with the presence of a parallel resonance near to but higher
than the second-harmonic frequency;
a high AC-side resistance near 0 Hz.

On the other hand, a susceptible inverter system will possess an opposite


reactive characteristic with inductive DC-side impedance at fundamental
frequency, and capacitive AC-side second-harmonic impedance. A high 0
Hz resistance is also observed at the unstable inverter station, but the two
reactive components have the dominant role in determining the system
stability.
The common use of HVDC back-to-back interties to interconnect large
and weak AC networks has resulted in low-order resonances at the converter terminals, making them prone to core saturation instability. However,
with comparable network sizes at both the rectifier and inverter ends, this
harmonic instability is most likely to occur only at one end of the scheme.
This is because of the opposite reactive requirements of the impedances

Interaction between AC and DC systems 155

for the instability to occur at either end. Moreover, the high resistance at
the unstable end will be reflected onto the DC side as additional damping
which tends to stabilise the opposite-end system. Therefore, for the backto-back scheme, it is necessary to consider the consequential impact on the
stability of the remote-end system when undertaking any modification at
the local end.
Besides the system impedances, the stability of the AC-DC system is
strongly dependent on the response of the converter controller. Considering the stringent reactive requirements for the instability to develop, the
onset of this harmonic instability almost certainly involves a destabilising
contribution from the converter controller. This suggests the possibility of
preventing the onset of the instability through proper tuning of the
converter controller.

6.7 AC-DC interaction following disturbances


6.7.1 AC-side fault recovery
AC-busbar voltage reductions during AC-system faults may cause commutation failures in some or all of the connected valve groups. During the
period of commutation failures, the valve groups cannot deliver power
into the AC network.
The importance of commutation failures during system faults depends
on the sensitivity of the receiving AC system to the energy deficit during
the failure, and the converter behaviour during the subsequent recovery
period. If the recovery period is not smoothly controlled, the effects on
the AC system can be aggravated.
When inverters are operated with commutation margin angles (y) of 18
(on 60 Hz), AC-voltage reductions to less than about 85 % may frequently
cause commutation failures. However, with y of 20 or larger, commutation
failures are not likely to occur for such or even greater voltage reductions.
Very weak systems will have difficulty in providing reactive power at the
rate required for fast DC-system recovery. Also, such systems will produce
severe AC-voltage distortion owing to magnetising inrush currents at reenergisation of the converter transformers upon fault clearing. The
converter controls have difficulty in operating correctly with such highly
distorted AC voltages. This can result in more commutation failures and
slow recovery.
When more than one DC link or other large AC-DC conversion plant
are connected to the AC system, the added DC power of all converters is
relevant to the recovery. In some cases, particularly with very low SCR
systems, it may be better to stagger the converter recoveries.
Depending on the characteristics of the AC and DC systems and the
control strategy used, typical times for a DC system to recover to 90 % of

156

High voltage direct current transmission

its prefault power, following AC-fault clearing, range from 100 ms to


300 ms.
Modern DC control systems are capable of resynchronising and recommencing correct operation of the DC system within two cycles of clearing a
severe AC fault, such as a three-phase-to-ground short circuit. Also the
gains and time constraints of the control systems are such that they do not
limit or increase the recovery time set by the main system characteristics,
and there is no significant delay in changeover between different control
modes at the same converter, for instance from constant y to constant
current.
To obtain good DC-system recovery, without further commutation failures, control strategy alternatives can include delayed or slow ramp recovery, reduced current level, reduced power level at recovery and a switch of
DC-system control mode (master control level) from constant-power
control to constant-current control.
The voltage-dependent current order limit (VDCOL) function can have
an important role in determining the DC-system recovery from faults,
particularly in a weak receiving-end AC system.
If the VDCOL function is activated during an inverter AC-system fault,
the result will be to decrease the DC current and hence the inverter reactive-power consumption, thus helping to support the AC-system voltage.
In the case of severe single-line-to-ground faults, the VDCOL may also
help to recover normal commutation, and thus some power transfer can
resume during the fault. Following fault clearing, the removal of the
VDCOL function current limit may be delayed and ramped so as to
maximise the recovery rate, while avoiding subsequent commutation failures.
The fault developments and related protective philosophies are
discussed in detail in Chapter 8.

6.7.2 DC-side fault recovery


Unlike AC-system faults, where the AC voltage and system behaviour are
interactive with the DC behaviour, DC-line faults are mainly a matter of
total-energy loss at the receiving AC system. However, it is possible to draw
down the rectifier AC voltage during the controlled de-energisation if the
rectifier system is weak. This results from the sudden and large reactive
demand during large converter firing-angle changes. This is not likely to
be of much significance unless the voltage reduction is very severe and if
special problems exist in the rectifier system.
The most common causes of DC-line faults usually result in a singlepole fault, with the healthy pole remaining unaffected in terms of power.
In some cases, where the DC link is operating below full load, the healthy
pole can quickly increase its power to compensate for the temporary interruption of power flow on the faulted pole.

Interaction between AC and DC systems 157


Since the detection and control action are relatively fast, the most significant factors affecting the energy loss are the arc deionisation time, the
number of restart attempts that may be required to clear a particular fault
and the recovery rate.
Typically, all actions including detection, forced retard and controlled
restart to 90 % power, but excluding the deionisation time, require less
than 50 ms. Depending on many factors including flashover mechanism
and air conditions, the deionisation time required may be of the order of
100 to 500 ms to ensure a successful restart.

6.8 References
1
2
3
4
5
6

7
8
9
10
11
12
13
14
15

CIGRE WG 14.07, 'Guide for planning DC lines terminating at AC system


locations having low short-circuit capacities, Part I: AC-DC interaction
phenomena', 1992
BOWLES, J.B.: 'Alternative techniques and optimisation of voltage and reactive power control at h.v.d.c. converter stations'. IEEE conference on Overvoltages and compensation on integrated A. C.-D.C. Systems, Winnipeg, 1980
LE DU, A.: 'Use of static or synchronous compensators in h.v.d.c. systems'.
CIGRE Study Committee 14, Rio de Janeiro, 1981
GRUND, C.E., POHL, R.V., CRESAP, R.L., and BAHRMAN, M.P.: 'Increasing power transfer capabilities of a.c./d.c. transmission systems by coordinated
dynamic control'. Symposium sponsored by the Division of Electric Energy
Systems, U.S. Dept of Energy, Phoenix, Arizona, 1980, pp.37l-87
UHLMANN, E.: Tower transmission by direct current' (Springer-Verlag,
Berlin/Heidelberg, 1975), p. 169
HINGORANI, N.G., NILSSON, S., BAHRMAN, M., REEVE, J., LARSEN,
E.V., and PIWKO, R.J.: 'Subsynchronous frequency stability studies of energy
systems which include h.v.d.c. transmission'. Symposium sponsored by the
Division of Electric Energy Systems, U.S. Dept of Energy, Phoenix, Arizona,
1980, pp.389-98
IEEE SSR Working Group: 'Countermeasures to subsynchronous resonance
problems'. IEEE PES Summer Meeting, Paper F79 754-4, Vancouver, 1979
GRUND, C.E., POHL, R.V., and REEVE, J.: 'Increased performance of
h.v.d.c. power modulation by active and reactive power coordination and
modern control design'. IEE Conf. Publ. 205 on 'Thyristor and variable static
equipment for A.C. and D.C. transmission' (London, 1981), pp. 176-81
TURNER, K.S.: 'Transient stability analysis of integrated a.c. and d.c. power
systems'. PhD thesis, University of Canterbury, New Zealand, 1980
CHEN, S.: 'Analysis of HVDC converter transformer core saturation instability, and design of a data acquisition system for its assessment'. PhD thesis,
University of Canterbury, New Zealand, 1996
AINSWORTH, J.D.: 'Harmonic instability between controlled static converters
and a.c. networks'. Proc. IEE, 1967, 114 (7), pp.949-57
AINSWORTH, J.D.: 'The phase locked oscillator - a new control system for
controlled static converters'. IEEE Trans., 1968, PAS-87 (3), pp.859-65
WOOD, A.R. and ARRILLAGA, J.: 'Composite resonance, a circuit approach
to the waveform distortion dynamics of an HVDC converter'. IEEE Trans.,
1995, PD-10 (4), pp.1882-888
SZECHTMAN, M., WESS, T., and THIO, C.V.: 'First benchmark model for
HVDC control studies'. Electra, 1991, (135), pp.55-75
AINSWORTH, J.D.: 'Core saturation instability in the Kingsnorth HV-d.c.
link'. Paper presented to CIGRE study committee No. 14, Winnipeg, Canada,
1977

158

High voltage direct current transmission

16 CHAND, J., and TANG, D.: 'Experience with Resonances and Oscillations in
the Nelson River HVDC System'. HVDC system operating conference, Winnipeg,
Canada, 1987
17 STEMMLER, H.: 'HVDC Back-to-back Interties on Weak a.c. Systems, Second
Harmonic Problems, Analyses and Solutions'. CIGRE symposium 09-87,
Boston, 1987, paper no. 300-08, 1-5
18 HAMMAD, A.E.: 'Analysis of second harmonic instability for the Chateauguay
HVdc/SVC scheme', IEEE Trans. Power Deliv., 1992, 7 (1), pp.410-415
19 CHEN, S., WOOD, A.R., and ARRILLAGA, J.: 'HVDC Converter Transformer Core Saturation Instability: A Frequency Domain Analysis'. IEE Proc,
Gener. Transm. Distrib., 1996 143 (1), pp.75-81
20 YACAMINI, R., and de Oliveira, J.C.: 'Instability
'Instat
in HVDC Schemes at Low
Integer Harmonics'. IEE Proc. C, 1980, 127 (3), pp.179-188
21 BURTON, R.S.: 'Report on Harmonic Effects on HVDC Control and Performance'. CEA 337 T 750, prepared by Manitoba HVDC Research Centre, 1994

Chapter 7

Main design considerations

7.1 Introduction
A typical design sequence for an HVDC transmission scheme should
include the following steps:
(a) Identify the main operational objectives to be met, i.e. energy considerations, MW loading requirements and maintenance.
(b) Identify any technical constraints which may have to be accepted, e.g.
the maximum voltage and current ratings of submarine cables, limitations of earth return etc.
(c) Adopt voltage and current ratings.
(d) Decide the overall control requirements, e.g. constant-power control,
short-term overload, damping characteristics, constant extinction-angle
control, constant ideal (noload) direct voltage, etc.
(e) Develop converter-station arrangements.
if) Design the transmission line.
(g) Assess the capital equipment cost, the operating costs and the cost of
losses.
Steps (a) to (/) must be critically reviewed to assess the effect of any
permissible parameter variation on (g).
Although the basic principles of rectification and inversion apply
equally to the mercury-arc and thyristor technologies, the design layout of
the converter plant is greatly influenced by the switching-device technology.
As there are still several schemes using mercury valves, this Chapter
starts with a brief exposition of their components and layout. Most of the
Chapter, however, is devoted to thyristor-converter technology.

160

High voltage direct current transmission

7.2 Mercury-arc circuit components


7.2.1 Valve group
In mercury-arc converters the six-pulse bridge constitutes the valve group,
although under normal conditions two phase-shifted groups (12-pulse)
operate together on each pole. This arrangement is exemplified by the
simplified diagram of Figure 7.1, which represents the original New
Zealand HVDC scheme.
Besides the main bridge valves and bridge transformers, the valve group
includes a bypass valve, which provides a path for the DC current of the
series-connected bridges during temporary bridge disturbances. 1 For
disturbance times exceeding the bypass rating capability a high-speed
bypass switch is automatically closed across its terminals. Permanent isolation of the valve bridge for maintenance also requires two isolating
switches.
The main circuit plant components associated with mercury-arc valves
are:
(a) Current dividers in series with each of the parallel anodes to achieve
current sharing between them.
(b) Anode reactors in series with each valve to reduce the rate of change
of current.
(c) Cathode reactors to damp the high-frequency current oscillations
owing to the commutation process, which would otherwise radiate
through the wall bushings.
(d) Valve-damping circuits in parallel with each valve to control the rate of
change of the voltage high-frequency oscillations owing to the commutation process.
(e) High-voltage isolating transformers to supply auxiliary power for the
excitation system, grid-control circuits, vacuum pump, air-duct fan and
heaters.
(J) Surge diverters across the valve group, between phases and from phase
to ground.
(g) Some schemes also include transient-voltage-suppression capacitors
connected across the valve group to improve the transient performance.

7.2.2 Converter station


Again with reference to Figure 7.1, each pole includes the following associated equipment in the DC switchyard:
(a) A smoothing reactor.
(b) A line surge capacitor.
(c) Surge diverters across the smoothing reactor and between line and
ground.

Main design considerations 161

Figure 7.1 Diagram of one pole 0fa mercury-arc HVDC station

162

High voltage direct current transmission

(d) Voltage divider and DC-current transformers for monitoring


purposes.
(e) Some schemes also include sixth and 12th shunt harmonic filters.
The main plant components of the AC system illustrated in Figure 2.1a
are:
(a) Six single-phase, three-winding bridge transformers per pole.
(b) Synchronous condensers connected to the delta-tertiary windings of
the converter transformers. These are often needed to supply the
extra reactive power required by the converter and network, and also
to increase the subtransient short-circuit ratio of the AC system.
(c) A filter bank per pole, which in this terminal includes tuned shunt
branches for the fifth, seventh, 11th and 13th current harmonics and
a high-pass filter for the 17th and higher orders.

7.2.3 Mercury-arc converter layout


A typical valve house2 comprises a long, reinforced-core building with a
steel-framed protected metal-clad annex on each side.
The central three-storey core comprises a valve-transport corridor with
ventilating and air-conditioning rooms above and air-pressure rooms
below in the basement. Along the outdoor station side there are several
valve halls with reinforced-concrete dividing walls in between and each
valve group functions as a self-contained unit with its own ventilating
system. Filtered air is drawn into the valve hall via louvres opposite the
base of each valve and the air exhausts via roof vents.
The annex comprises a clean workshop sanitary room, assembly room,
degassing room and control room.
Radio disturbance can be transmitted by direct radiation from the valve
acting as a dipole. To suppress this, effective screening of the valve-hall
enclosure is normally required, particularly when the converter station is
close to a residential area. The steel cladding of the outside walls and roof
of the valve halls is bonded to provide an effective shielding. A fine-mesh
screen is embedded in the concrete floors and walls, in addition to the
normal reinforcing. Expanded metal screens, well bonded and with phosphor-bronze pressure contacts on all gates, are provided between the valve
halls and the transport corridor.

7.3 Thyristor valves


7.3.1 Electrical considerations
Because of the limited voltage rating of the individual thyristors, many of
them must be connected in series to constitute an HVDC valve. The series

Main design considerations 163

saturating
reactor

firing databack
optical fibres

Figure 7.2

Electrical circuit of the thyristor level

connection of thyristors requires additional passive components to distribute the OFF state voltage uniformly between them and to protect the
individual thyristors from overvoltage, excessive rate-of-rise of voltage (dv/
dt) and rate-of-rise of inrush current (di/di). The thyristor, together with its
local voltage-grading and thyristor-triggering circuits, known as a thyristor
level, is the building block of the valve architecture. The circuit of a typical
thyristor level, shown in Figure 7.2, contains the following components:
(i)

(ii)

(iii)

A saturating reactor, which consists of several strip-wound steel


cores mounted on a bar primary, and presents a large inductance in
series with the stray capacitances of the external circuit. This component is required to protect the thyristors from damage immediately
after firing. However, excessive inductance is also undesirable
because it increases the reactive power absorbed by the converter.
The saturating reactor avoids this problem by presenting a high
inductance only at low current. At higher currents the steel saturates
and causes the incremental inductance to decrease; at full-load
current the effect of the reactor is practically negligible.
Good voltage distribution, achieved by several components acting
over different frequency ranges. The direct voltage is distributed by
a DC grading resistor (/JQ) which is also used as a voltage divider to
measure the voltage across the thyristor level.
Voltage distribution in the range from power frequency up to a few
kilohertz, which is ensured by a complementary pair of RC grading
circuits (RD and CD). This frequency range includes the natural
frequency characteristic of the voltage oscillations which occur at

164

(iv)

High voltage direct current transmission

the end of a conduction interval. Appropriate component values are


chosen to limit the magnitude of this voltage oscillation,
Insulation failures within the converter can subject the valve to
voltage oscillations of much higher frequency, at which the above
RC circuits are not effective. To protect the thyristor levels from
severe voltage during such events, a capacitive grading circuit is also
included. Thus, a fast-grading capacitor (CFG) is used to discharge
via part of the saturating reactor to limit its contribution to thyristor
inrush current.

Normally, the thyristors are triggered into conduction at a particular pointon-wave determined by the control system. The command to fire the valve
is sent via optical fibres from a valve-base electronics cubicle at earth
potential to every thyristor in the valve. The optical signals are decoded by
a gate-electronics unit located adjacent to each thyristor, which then generates a pulse of current to trigger the thyristor. The gate-electronics unit
derives the necessary power for its operation from the RC grading circuit
during the OFF state interval.
Thyristors can be damaged by excessive forward voltage or forward rate
of change of voltage (dv/dt), especially at the higher junction temperatures
which may occur during faults. They are particularly vulnerable during
the recovery period following turn off, when even a modest forward
voltage may cause uncontrolled conduction. However, they can be
protected by arranging for the gate electronics to trigger the thyristor into
conduction independently of the central control system.
In marginal cases, some thyristors may block forward voltage but others
do not. In the limit, this could result in the whole of the valve-winding
voltage being applied to a single thyristor, so that the thyristor would be
destroyed if it were not protected. A back-up triggering system, based on a
breakover diode (BOD) is used for this purpose; it comprises a series string
of small overvoltage-triggered thyristors, connected from the anode to the
gate of the main thyristor via a current-limiting resistor. When the
forward voltage across the main thyristor threatens to exceed the
maximum safe value, the semiconductor elements of the BOD conduct,
and pass a heavy pulse of current to the gate of the main thyristor, rapidly
triggering it into conduction. If the thyristor-level components are suitably
rated, the BOD can operate repetitively in the event of a failure of the
gate electronics, thereby preventing consequential failure of the thyristor.
By connecting a suitable number of thyristor levels in series, a valve of
the necessary voltage rating can be constructed.
Modern DC schemes are designed exclusively for 12-pulse operation. A
simplified diagram of the 12-pulse converter group is shown in Figure 7.3.
Besides the extensive protective measures incorporated at the individual
thyristor level, it is also necessary to provide overall valve protection
against reverse overvoltage. Figure 7.3 shows that this is accomplished by

Main &sign conszderations 165


SOUTH

WEST
400/93 kV
single-phase
three-winding
transtorrners

r----------___

.*-

2415 A. 205 kV DC

_____

400/93 kV

4W kV AC
switchyard

Figure 7.3 Schematic arrangewilt (fconuerter equipment for m pole


connecting a gapless metal-oxide surge arrester across the valve. This surge
arrester constitutes the primary protection of the valve against overvoltages
of external origin. The arrester has to withstand continuous operation
while subjected to the valve OFF state voltage, including the periodic
switching transients occurring every cycle. The arrester provides a protective level typically around 70 % higher than the peak of the normal operating voltage of the valve.
T h e number of series-connected thyristors required to meet the operating conditions for a particular scheme is determined by the protective level
of the arrester and a test withstand margin (typically 15 %). Extra redundant levels are included to allow the equipment to remain in service after
a small number of thyristor failures.

7.3.2Mechanical considerations
T h e power thyristors used in the valves are constructed from monocrystalline silicon wafers of typically 100 mm nominal diameter. Thyristors of
this type require a very high clamping pressure to maintain adequate
thermal and electrical contact between the silicon and the electrical connections and heatsinks; a 100 mm thyristor requires a clamping force of 80
kN (eight tonnes). T o provide this clamping force in a compact and lowcost manner without losing the capability to replace a thyristor in the event
of its failure, the elastic properties of a very advanced composite material
are used. T h e banded-pair assembly, shown in Figure 7.4, consists of thick
bands of filament-wound void-free glass-reinforced polymer, which apply
the necessary clamping force to a series-connected pair of thyristors and
the associated heatsinks. Removal of a thyristor is accomplished by using a
hydraulic tool to separate the two heatsinks which are in contact with the

166 High voltage direct current transmission

Figure 7.4 Bandedjmir mmbly

Figure 7.5 DC semitier

Main d e s i p considerations 167


FRONT ELEVATION

SIDE ELEVATION

nt.:

UC.

onc
valve
A i:

AC

I)
c

DC

tier
assemb Iy

AT,

AT:

DC

Figure 7.6 A quadrivalve comprising four vertically-stacked valves

particular thyristor while retaining sufficient clamping force on the other


thyristor to ensure that its pressure contacts are not compromised. A particular advantage of this system is that it is not necessary to open either the
power circuit or the cooling circuit in order to replace a thyristor.
T h e banded pairs are built into frames to form a tier assembly. A typical
tier, shown in Figure 7.5, contains 14 series-connected thyristors. T h e
components surrounding the thyristor provide thermal, mechanical and
electrical protection. At the front of the tier are the gate electronics which
convert the optical signal to an electrical firing pulse for the thyristor.
The individual tier assemblies are stacked and separated from each
other by composite column insulators which employ the same void-free
glass-reinforced polymer used for the clamping bands. These insulators
provide a strong and compact alternative to conventional ceramic insulators and give the valve stack good seismic capability.
A sketch of the quadrivalve comprising four vertically-stacked valves is
shown in Figure 7.6 and a complete valve group (for 12-pulse operation)
in Figure 7.7.

7.3.3 Valve-coolingsystem
T h e thyristors produce considerable heat loss, typically 30 to 40 W/cm2 (or
over 1 MW for a typical quadruple valve), and an efficient cooling system
is thus essential. Each thyristor unit is normally provided with a double
heat sink and the heat is taken away from the sinks by circulating water.
High-purity water combines superb cooling with high dielectric strength.

168

High voltage direct current transmission

Figure 7. 7 Thyristor valves at Sellindge converter station

The purity of the water must be very high to remove all ionic components,
which would otherwise cause the coolant to bypass the electrical insulation
of the valve. Thus, the conductivity of the pure-water coolant is monitored
continuously and is controlled to less than 0.5 jiS/cm by the use of ionexchange resins. The heat generated by the thyristor valves is rejected to
air by evaporating coolers.
The liquid coolant is distributed to every thyristor level in the valve
through electrically-insulating polyethylene hoses. Water-cooling systems,
however, require careful design to prevent leakage (which would have
disastrous consequences) and corrosion. A water-cooling system is normally
placed in the basement under the valve hall, as illustrated in Figure 7.8.

7.3.4 Valve-control circuitry


Electrically-triggered thyristors require a local electronics unit for the
generation of trigger pulses as well as protection and monitoring. All

Main design considerations 169

Figure 7.8

500 kV valve hall with mechanical auxiliaries (Brown Boveri Rev.)


1 = quadruple valve
2 = cooling water system
3 = base electronics
4 = ventilation of the hall with air filter

signal communication within and to the valves across potential differences


is performed using light pulses transmitted by light guides (fibre optics).
This applies to both the firing signals for the individual thyristors in the
valve, and the feedback signals from each thyristor level to the valve
control equipment.
These feedback signals also make it possible to monitor the state of each
individual thyristor. Microcomputers are used in the control room to
process the information from the valve. A faulty thyristor is immediately
detected and the exact position of the defective thyristor is reported. Since
each valve contains a somewhat larger number of thyristors than is actually
needed, the converter can continue to operate, even if some thyristors are
defective; these are only replaced during the planned regular maintenance.

170

High voltage direct current transmission

Thyristor level
(Local firing
control and
protection)

125 level valve


with 1 level per
module.
(Simplified)

125 Fibre optics


(thyristor
firing
*^*
commands)

Valve
(Current
K H * measurement
-triplicated)

Control room
(Main valve
firing control
and protection
-triplicated)

3 Fibre optics
(Valve current
data)
Valve firing
command
(duplicated)

Ground level
at valve base
(Signal interface)

'Start' 'Stop'
(From central control system)

Figure 7.9

Location and basic functions of the Cross-Channel valve electronic


systems

The auxiliary power needed for the thyristor firing is obtained from the
voltage across the thyristor.
The location and basic functions of the Cross-Channel4 valves are shown
in Figure 7.9; they are divided into a number of thyristor levels or
modules acting independently. As a result, marginal differences in protective settings, or tolerances in valve components, can cause the protective
circuits at some levels to operate. This may lead to cascade turn on, with
the last level to fire experiencing a greater duty than that occurring under
normal turn-on conditions. The valve-circuit components are rated to withstand such duty.
In the absence of cascade turn on, the levels that have been protectively
fired will conduct the valve grading current. If the disturbance causing a
protective level to operate were now to reverse the valve voltage to a value
approaching the protective level of the valve-surge arrester, then those

Main design considerations 171

levels which were conducting would be driven by the valve-grading current


to a prospective negative voltage considerably higher than the reverse
voltage rating of the thyristors. However, the thyristors are normally
provided with a high reverse-avalanche capability which limits the reversevoltage excursion by conducting the valve-grading network current in an
avalanche mode, until the valve voltage is more evenly distributed.
The technology to manufacture fully self-protected light-triggered thyristors is now becoming available.
The use of light-triggered thyristors reduces the number of valve
components, the valve-tier volume and the valve cost. It also improves
thyristor-level reliability.

13.5

Valve tests5

Tests on thyristor levels


The routine tests carried out on each thyristor in the factory include:
recovery-charge check;
recovery-time check;
gate-voltage and gate-current check;
gate-controlled delay-time check;
nonrepetitive inrush-current test at overvoltage protection triggering;
nonrepetitive peak-reverse avalanche voltage test;
repetitive voltage measurements;
on-state voltage-drop measurements;
surge on-state current test;
helium leakage test.
The final assembly of the device with the heat sink is further tested for
forward-voltage drop, current balance if the design involves paralleling of
thyristors on the same heat sink, rate of change of current and gating
capability.
Test specifications also require checking of the protective devices (for
forward and reverse protection) at this stage.
Each resistor is tested for poor mechanical joints or defective connections; the test stresses the resistor thermally and mechanically simultaneously.
All valve commutating and storage capacitors are required to pass a
temperature cycling test of typically 25C to 80C.
Reactors and pulse transformers are individually tested for corona
extinction at a level at least 1.25 times the maximum repetitive voltage
stress expected.

172

High voltage direct current transmission

Tests on valve modules

Assembled thyristor modules contain all the basic components of a valve


and can therefore be subjected to the complete testing procedure, as
required by the IEC 700 standard, which includes:

connection tests: to check that the connections within the thyristor level
are correct;
voltage-grading circuit test: to monitor tolerances on grading-circuit
components against given ranges;
hydraulic-pressure test: to check leakage detection and thyristor clamping
pressure;
current cycling: to determine maximum temperature swing;
on-state voltage-drop test: measurement of on-state voltage drop at
nominal current;
partial-discharge test;
short-circuit test: to check the satisfactory operation of the thyristor and
its monitoring circuits during short-circuit conditions;
impedance test: to check for short or open circuit in the voltage divider
circuit;
firing test: to check thyristor turn on in response to firing signal;
protective-firing test: to check the protective-firing level against specified
tolerance and the co-ordination of this protection with the thyristor
forward-voltage withstand capability;
recovery protection test: to check the recovery protection level at two separate instants against specified tolerances, and the co-ordination of the
protection with the thyristor-withstand capability;
reverse-blocking test: to check the thyristor reverse-blocking voltage withstand at the test voltage.

Operational type tests


The operational tests are intended to verify the valve performance under
conditions of normal, overload and abnormal operation and also under
transient fault conditions.
A typical test circuit consists of two six-pulse bridges connected back-toback with a valve represented by three series-connected thyristor levels. It
includes filter banks and rated values of the effective stray capacitances
across the valves, the impedances in the commutation path and the saturating valve reactors to ensure that the turn-on and turn-off stresses owing to
the commutation process are representative of service conditions.
The load tests include:

a heat-run test to demonstrate that the valve operates correctly under


the worst normal operating conditions;

Main design considerations

173

a temperature-rise test to verify the thermal design of all critical heatproducing components of the valve module and its components;
a periodic-firing and extinction test to demonstrate correct behaviour
of the valve when subjected to periodic voltage and current stresses.

The remaining operational tests are:

minimum alternating voltage: to demonstrate the correct functioning of


thyristors and auxiliary circuits at a specified minimum-voltage level;
intermittent direct current: to demonstrate satisfactory operation of the
refiring circuitry at low-load levels when the current becomes discontinuous;
short-circuit current with subsequent blocking: to test the fault-current
suppression capability of the valve;
short-circuit current: to demonstrate that the valve can withstand three
loops of the highest overcurrent without thermal or electrical degradation;
forward recovery: to test thyristor withstand and protection against
forward-voltage transients during the thyristor recovery period.

The last two tests above are not included in IEC 7005 but are being recommended for inclusion in a future revision of the standard.6

Dielectric type tests


The dielectric tests are intended to verify the insulation integrity of the
valve, and the immunity of its components from interference; they also
ensure that partial discharges are within specified limits for various types
of overvoltages.
The dielectric tests normally carried out on the single valve units include
DC-withstand voltage, AC-withstand voltage, switching impulse, lightning
impulse and steep-front impulse.
The complete quadrivalve structure is also subjected to switching and
impulse tests, as well as a DC-corona voltage test.

Additional tests
Each scheme has specific nonstandard requirements and a supplementary
test programme results from detailed discussions between the purchaser
and the supplier. For instance, in the case of the New Zealand upgrade,
extra tests were carried out on a representative section of a valve to verify
the valve fire withstand capability. Snap-back tests on the quadrivalve unit
(including the associated arresters) were also performed to demonstrate its
dynamic response and verify seismic requirements.7

174

High voltage direct current transmission

Figure 7.10

Cutaway view of the valve hall showing the locations of the major
components

7.3.6 Valve-hall arrangement


Figure 7.10 is a cutaway model of the valve hall arrangement of the 500
MW Chandrapur back-to-back converters.
The six quadrivalves are arranged within the valve hall, a large building
(30 x 25 x 10 m high), with space around them for maintenance access,
electrical clearances and connections. To provide a containment for the
radio-frequency interference generated by the valve-switching transients,
the valve-hall walls are screened and the transformer connections are
entirely enclosed within the valve hall.

7.4 Station layout


A typical set of 12-pulse converter components is shown in Figure 7.118
which relates to the CU HVDC project. All the components enclosed
within the thick rectangle are located inside the valve building. The transformers are connected in star-star and star-delta, respectively, to provide
the necessary 30 phase shift for 12-pulse operation.
The AC harmonic filters consist of tuned branches for the 11th and
13th harmonics and a high-pass branch tuned to the 24th harmonic. A

Main design considerations 175

10

Figure 7.11

Main circuit diagram for one pole of a converter station (ASEA


Journal)
1 surge arrester
2 converter transformer
3 air-core reactor
4 thyristor valve
5 smoothing reactor
6 director-voltage measuring divider
7 DC filter
8 current measuring transducer
9 DC line
10 electrode line

high-pass DC filter tuned to the 12th harmonic is placed on the DC side.


Extra shunt capacitors are installed at the Dickinson station only, since the
generators at Coal Creek can provide the necessary additional reactive
power. To limit inrush currents and overvoltages during transformer energisation, the converter breakers in both stations are provided with preinsertion resistors.
A DC smoothing reactor is located on the low-voltage side and air-core
reactors on the line side of the converters; the latter to limit any steepfront surges entering the station from the DC side. Additional air-core
reactors are installed in each phase on the AC side to reduce the rate of
rise of current during thyristor turn ON.
The thyristor valves are protected by phase-to-phase surge arresters. The
three top valves, connected to the pole bus, are exposed to higher overvoltages in connection with specific but rare incidents, and are further
protected by arresters across each valve. The indoor arrester connected to

176

High voltage direct current transmission

the low-voltage side of the valve protects the reactor. Pole and electrode
arresters supplement the overvoltage protection.
The measuring equipment, i.e. a voltage divider, current measuring
transductors and current transformers, provide the necessary input signals
for the control and protection circuits.
The switching components (i.e. isolators and circuit breakers), are of
conventional design on the AC side of the converters. Several switches are
also used on the DC side. Conventional oil-minimum circuit breakers are
used to interrupt small currents for the switching of the neutral bus load
and for the changeover from single-pole metallic return to bipolar operation.
Also, an HVDC circuit breaker is used to achieve ground-to-metallic
return transfer; this breaker is designed to interrupt 1500 A and to absorb
an energy of 2 MJ.
The area of the modern thyristor station is only a fraction of that
needed for earlier mercury-arc converter stations.
Figure 7.12 shows a typical layout for a 1000 MW bipolar HVDC
station9 and gives a clear indication of the relative space taken up by the
various plant components. The major proportion of the space is taken by
external plant, particularly the capacitors used in the form of harmonic
filters and for voltage support.
The layout of the valve hall, which apart from the valves contains surge
arresters, phase reactors and the line reactor, is better explained with
reference to the sketch in Figure 7.13. The transformers (on the right of
the picture) and the smoothing reactor (on the left) are placed close to
the hall walls with their bushings passing through the wall. The location
of the equipment inside the hall is designed to combine a low probability
of internal flashovers with the best possible utilisation of the space available.
The floor area of the valve, service and control rooms is only a small
fraction of the total station area. The auxiliary power equipment used for
cooling and air conditioning is placed immediately under the valve hall.
The building normally contains a steel structure designed to act as a
Faraday cage to reduce electromagnetic radiation from the valve hall
which might cause radio interference.

7.5 Relative costs of converter components


A survey10 among converter manufacturers has provided information on
costs, losses, overload capabilities and reliability of individual converter
components. The data collection used as a basis a system consisting of two
bipoles of conventional equipment; the bipole included four 12-pulse
groups, each rated at 800, 1200, 1000 and 2000 MW at a daily mean
temperature of 30.

Next Page

Main design considerations 177


_1

11

Figure 7.12

Station layout for a bipolar HVDC station (ASEA Journal)


1 DC and electrode lines
2 DC switchyard
3 DC smoothing reactors
4 valve hall, pole 1
5 service building with control room
6 valve hall, pole 2
7 converter transformers
8 AC harmonic filters
9 high-pass filter
10 11th harmonic filter
11 13th harmonic filter
12 shunt capacitors
13 AC switchyard

Chapter 8

Fault development and protection

8.1 Introduction
DC converter stations form an integral part with the AC-power system, and
their basic protection philosophy is thus greatly influenced by AC-system
protection principles.
There are, however, two technical reasons which influence some departure from the conventional protection philosophy, i.e. the limitations of
DC circuit breakers and the speed of controllability of HVDC converters.
Furthermore, the series connection of converter equipment also presents
some special problems not normally encountered in AC substations.
As with AC protective systems, DC safety margins should be based on
statistical risk evaluations, distinguishing between independent disturbances and the possible cascading of faults. For a given disturbance, the
protective system must also be capable of disconnecting only the lowest
necessary level and for the minimum time interval.
The characteristics of internal (within the converter) and external faults
are quite different and are considered separately.

8.2 Converter disturbances


According to the origin of the malfunction, converter disturbances can be
divided into three broad groups, i.e.:
(a) Malfunction of the valves or their associated equipment. The main
types are: misfire, firethrough and backfire (only in mercury-arc
valves).
(b) Commutation failure, the most common disturbance during inverter
operation. This fault often follows other internal or external disturbances.

Fault development and protection 201

(c) Short circuits within the converter station. Although these faults are
rare, they must be taken into consideration in converter design.

8.2.1 Misfire and firethrough


Misfire is the failure to fire a valve during a scheduled conducting period
and firethrough is the failure to block a valve during a scheduled nonconducting period. These faults are caused by various malfunctions in the
control and firing equipment.
The effect of these faults is more critical when they occur at the inverter
end. With rectifier operation they do not constitute a serious disturbance
unless they are sustained, in which case they can introduce voltage and
current oscillations on the DC side.
By way of example, Figure 8.1 illustrates the development of a firethrough in valve V\ at the instant B during inverter operation. The valve
voltage Vi is indicated by the thick broken line in Figure 8.1ft; this valve
can firethrough at any time after instant A, although the scheduled firing
instant is F. If the cause of the firethrough persists, the fault will recur at
instant G, as the thick dotted lines indicate.
It must be pointed out that the idealised waveforms of Figure 8.1 are
only valid in the presence of infinite smoothing inductance and a very
large short-circuit ratio. In practice, the current will change considerably
during the disturbance and, with it, the level of distortion of the voltage
waveforms (as explained in Section 8.3).

8.2.2 Commutation failure


This fault is the result of a failure of the incoming valve to take over the
direct current before the commutating voltage reverses its polarity, taking
into account the need for sufficient extinction time.
A true commutation failure is due to varying conditions in the external
AC or DC circuits combined with inadequate predictive control of the
inverter extinction angle. Either a low alternating voltage, a high DC
current or both, can prevent completion of the commutation process in
sufficient time for safe commutation; in such cases the direct current is
shifted back from the incoming valve to the previously-conducting valve.
Figure 8.2 illustrates the idealised development of a single commutation
failure. For simplicity the fault is created by introducing some delay in the
firing of the incoming valve, i.e. valve V3 is fired at instant B, instead of
the normal instant A Since the commutating voltage (phases RY) becomes
positive after instant C, the incoming valve V3 eventually ceases conducting
at E (shown by current waveform e) and the direct current commutates
back to the preceding valve Vi (waveform c).
When valve V4 is fired at D (waveform j), a three-phase short circuit is
briefly established by the conduction of the four valves V\ to V4 until

202

High voltage direct current transmission

G HI

Figure 8.1

Single firethrough of valve V1 in typical inverter

S3

S5

Fault development and protection 203

Y
*x.

B
/

. P6
,

(a)

' .' 1

'

'(PS)

'

' . Fl

SI

.(.PIT

1Cl

. 1 C6 '
1
1

'

S4

'

C4\

..

J-C2

1 1

A''
|

If

If

(b)

Cl

06

C5

C5

C3\

"

b
RY
h

I Id

(0

M n

V f

I^

RB

1 I

C1

RY

RB

11

i
i
l
1

!
i

'

Cl

VB

VR

V
I

l'4
1

ih)
CA

Figure 8.2

(Of

~"V

if
1

. '

C p / l l C i i
C D E F G

'
'6

C 2 ,
H

C5

\ ^
, C 3 |
I
J

C4

C 5 ,

i
1
1
06
KL

Single commutation failure from valve V] to valve V3

o,f;
/

'5

>/|
A B

(Of

"V

'3

(g)

(Of

IV

ci

Cl

(Of

204

High voltage direct current transmission

instant F, when the commutation from F2 to V4 is completed. Between D


and / the bridge is bypassed by the conduction of valves V\ and V4, during
this period there are no alternating currents in the converter transformer,
but the direct current increases in practice as a result of the temporary
voltage collapse of the inverter. If a normal commutation takes place from
V4 to VQ between H and /, the bridge normal voltage is gradually re-established.
However, owing to the DC-current rise, caused by the temporary DCvoltage collapse at the inverter end, commutation from valve V2 to V4 may
also be unsuccessful, thus causing a double successive commutation
failure. In this case, illustrated in Figure 8.3, successive valves V\ and V2
are the only conducting valves after instant G, and the inverter output
voltage Vd reverses for nearly half a cycle (as shown in Figure 8.3b). Such
development would, in practice, increase the DC current rapidly and, as a
result, subsequent commutations may also fail, as indicated by the thick
dotted lines in Figure 8.3b.
As a result of the DC short circuit at the inverter end, the transformer is
either partially or totally bypassed and the DC-line current exceeds the
current in the AC lines. This effect has been used to detect the occurrence
of commutation failures.
After the occurrence of a commutation failure, the next firing instant is
advanced by constant extinction-angle control. If the failure is caused by
low alternating voltage following an AC disturbance, upon clearance of
this disturbance the normal voltage will return and prevent further
commutation failures.
In the event of recurring commutation failures the valve group should
be blocked. This action, as explained in Section 8.2.5, is often combined
with bridge or valve-group bypass in the case of a multigroup converter
station.
Voltage-waveform distortion following AC-switchgear switching operations at the inverter station can lead to commutation failures. Countermeasures used to avoid these are temporary shifts of the commutation margin
angle prior to the switching operation and the application of circuit breakers with closing resistors.
The probability of commutation failure can be reduced by increasing
the minimum extinction angle allowed in normal operation. This,
however, increases the VAR compensation required and a compromise is
reached where a reasonably low probability of commutation failure is
acceptable. As a guideline, the minimum margin angle is fixed to avoid
commutation failures during voltage reductions of up to 15 %.1 Therefore, only disturbances resulting in converter voltages below 85 % of the
nominal are considered in the assessment of the converter transient reliability (TR). TR, a concept used in CIGRE survey reports,2 is defined as the
ratio of the number of times that the HVDC system returns to its predisturbance power transfer level after a disturbance, to the number of distur-

Fault development and protection 205

CA

AB

Figure 8.3

Double-successive commutation failure from valve Vj to valve V3 and


valve V2 to valve V4

206

High voltage direct current transmission

bances (in per cent). Another parameter of interest in this respect is the
number of commutation-failure starts (CFS). A CFS is the initiation of a
distinct and separate commutation-failure event.
Table 8.1 shows reported cases of TR and CFS in 1994 in 24 different
schemes. The high recovery level (one hundred per cent in most cases) is a
clear indication of the progress made in the reliability of recent HVDC
schemes. Another interesting piece of information extracted from the
Table is the low probability of commutation failure (in the New Zealand
scheme only one CF for every 20 x 106 successful ones!).
Table 8.1

Transient reliability and number of commutationfailure starts (CFS)


1994
Recordable AC faults

System

Skagerrak 1 and 2
Skagerrak 3
Vancouver Pole 2
Square Butte
Shin-Shinano 1
Shin-Shinano 2
Nelson River BP2
Hokkaido-Honshu
CU
Durnrohr
Gotland II
Miles City
Highgate
Cross-Channel 1 and 22
IPP4
Virginia Smith
Konti-Skan 2
Vindhyachal
Fennoskan
Rihand-Delhi
SACOI 25
New Zealand Pole 2
Wein Sudost
Sakuma

Number

Transient
reliability %

Total no.
of CFS1

8
8
21
68
0
3
3
4
13
69
1
237
37
103
22
8
NR
11
NR
48
NR
55
16
0

100.0
100.0
95.2
98.5
100.0
66.7
100.0
100.0
98.6
100.0
100.0
100,0
100.03
95.5
100.0
100.0
100.0
100.0
97.9
NR
100.0
100.0

6
40
2
21
0
3
19
4
49
53
1
97
28
60
NR
NR
27
0
6
47
28
35
NR
4

1 CFS from all causes (internal and external)


2 data is total for Cross-Channel 1 and 2
3 data is for Sellindge terminal only
4 data is for one terminal only
5 data is for two terminals

Fault development and protection 207

8.2.3 Backfire
Although backfires, or conduction in the reverse direction, can occur (and
have occurred) on thyristors, both as external flashover and as failure of
all thyristors in a valve, this fault is only discussed with reference to
mercury-arc valves and is caused by the combined effect of:

high reverse voltage across the valve;


high rate of rise of initial voltage jump;
high rate of fall of current at the instant of initial voltage jump.

The reverse voltage is higher during rectification (refer to Figure 2.3) and
therefore the backfire probability is much higher on this mode of operation.
Having lost its unidirectional conduction property, the backfiring valve,
together with the forward-conducting valve on the same side of the bridge,
provides a path for uncontrolled phase-to-phase short circuit in the converter transformer. Self recovery is not normally possible with backfires and
total blocking is ordered upon detection of a single fault. However, blocking is not always possible and back-up AC breaker action is often needed
to clear the fault.
The current in the forward-conducting valve during this condition
reaches typical peak values of 10 p.u.3 The combination of a high-voltage
jump following current extinction in the forward-conducting valve, and
the large current magnitude shortly before extinction (i.e. the ionisation
level met by the recovery voltage), often produces what is called a consequential backfire in this valve. This constitutes a most serious condition in
mercury-arc rectifiers, as the converter valves and transformers have to be
rated to withstand large overcurrents prior to fault clearance by the AC
circuit breaker.

8.2.4 Internal short circuit


Although rare, short circuits can occur at various locations of the converter
station, as shown in Figure 8.4. These can be caused by maloperation of
earthing switches, deteriorating insulators or surge-arrester failures, particularly during transient overvoltages.
A flashover across a nonconducting valve (Figure 8Aa) produces a
phase-to-phase short circuit with a very large overcurrent on the conducting valve.
The largest stress is produced during rectification with a small firing
delay, and the worst instant is immediately after commutation, e.g. across
valve V\ in Figure 8.4; in this case the current in valve V3 is only limited by
the transformer leakage reactance and the system source impedance.

208

High voltage direct current transmission


Bridge 1
Convenor
transformers

.4

Possible locations of internal AC-DC short-circuit faults in typical


12-pulse thyristor converter
a Faults across a nonconducting valve
b Faults across bridge terminals
c Faults across AC phases on the valve side of converter transformer
d Ground faults at a DC terminal of a bridge
e Ground faults at an AC phase on the valve side of converter transformer
/ Ground faults at the station pole or DC busbar

8.2.5 Bypass action


Many of the valve faults are of a temporary nature and can be eliminated
by a temporary absence of conduction.
In mercury-arc schemes this is achieved by the use of a bypass valve
across the converter bridge (see Figure 7.1). This valve is kept blocked

Fault development and protection 209

when the bridge unit is conducting in the normal manner. When it is


necessary to stop the bridge from conducting, the bypass valve is fired
while the main bridge valves are blocked. Once a bypass valve has fired, it
can be blocked only by first interrupting its current so that its grid can
regain control. In the case of a rectifier, assuming that its bypass valve is
carrying current, when the bridge valves are restarted a positive voltage is
established across the bypass valve, the cathode of which becomes positive
with respect to its anode; the bridge valves then take over the current from
the bypass valve, which stops conducting since it cannot conduct in the
reverse direction.
In the case of the inverter, the bypass valve will not stop until its cathode
is made positive with respect to its anode; the necessary reversal of polarity
may be accomplished by a temporary advance of the angle /? to greater
than 60.
A combination of a bypass and two series switches (shown in Figure 7.1)
permits bridge isolation for more permanent outages.

8.2.6 Bypass action in thyristor bridges4


The absence of backfires in thyristor valves permits a simpler bypass
scheme without the need for a bypass valve. Instead, one of the main
bridge arms provides the necessary bypass. A healthy-arm pair can always
be found to relieve a temporary disturbance in one of the bridge valves, as
shown in Figure 8.5.
Blocking of a converter through bypass pairs involves the same series of
operations, in principle, as blocking through a bypass valve; i.e. blocking
of the main firing pulses and the simultaneous injection of continuous
firing pulses to a bypass pair.
The selection of blocking sequences applicable to bypass pairs is particularly important to give the faulty valves the best chance to recover, without
resorting to the operation of the isolators associated with the converter.

3!i

Figure 8.5

Bridge using a bypass pair (1-4), two series switches and a bypass
switch

210

High voltage direct current transmission

The criterion for this selection, after a repetitive failure, is that none of
the valves of the selected bypass pair should have been involved in the
fault.
Although this will be satisfied only by one bypass pair, there are various
alternatives according to the selected blocking instant. Ideally, bypass
action should be carried out immediately after the fault, using the valve
which was the last to conduct and its opposite; e.g. if at the instant of the
blocking signal, valve 4 is commutating to 6, then the bypass pair to be
used should be 3, 6.
The selection of the bypass pair and the blocking sequences is thus
simple, with the exception of double successive commutation failure; with
this fault, two of the valves involved in the fault form a bypass pair, and
each one of the remaining two valves belongs to either of the two remaining bypass pairs; the simple criterion used to select the bypass pair for a
single commutation failure cannot be applied in this case. Generally,
however, not all the four valves will be faulty or directly responsible for
the occurrence of the fault, and there will be at least one bypass pair
through which blocking will be possible.
The selection of the most suitable bypass pair for blocking depends on
the cause of the fault. If the fault detectors do not provide sufficient
discrimination of the initial cause of the fault and its subsequent development, the selection of the bypass pair and final blocking will be slower.
Resumption of normal operation simply demands the restoration of
firing pulses with suppression of the blocking pulses. Inverter deblocking
by these means is much simpler than in the case when a conventional
bypass valve is used, since momentary rectifier action by advancing the
firing angle, /?, is not required.

8.3 Simulation of practical disturbances


The waveforms illustrated in Figures 8.1-8.3 apply only under idealised
conditions in which the converter fault has no effect on the commutating
voltage. In practice, however, such an ideal system never exists and even a
single commutation failure produces considerable waveform distortion,
caused by the temporary harmonic current mismatch between the converter and the filters. The latter cannot accommodate instantaneously the
current changes caused by the disturbance, and the difference penetrates
into the AC system, causing transient-voltage harmonic distortion in
proportion to the system impedance.
Although the converter is expected to recover from commutation failures caused by normal variations in the AC-system voltage and direct
current, the situation is rather different following large disturbances.
It is thus essential to be able to predict the behaviour of the DC link at
the design stage, and for this purpose the manufacturers have in the past

Fault development and protection 211

made extensive use of scaled-down physical models with detailed representation of the controls. However, the losses in physical models cannot be
scaled down in proportion to the power and, as a result, all the oscillations
are subjected to excessive damping; the valve voltage drop is often
reduced by means of electronic compensation in the models. Moreover,
the so-called physical models are also restricted in the number of converter
groups and AC-system components.
It has been the feeling among some manufacturers in the past that
mathematical simulation and computer models could not be trusted to
represent the converter behaviour during disturbances. Such reservations,
however, were based on lack of data and the need for large computer
requirements. The tendency has been to keep the mathematics simple,
while increasing the complexity of the experimental simulators.
More recently, the computer requirements, i.e. memory and calculation
times, have become less important and the modelling of the AC-DC
system behaviour more realistic.5 Real-time digital simulators (RTDS)6 are
now becoming used widely, either by themselves or combined with scaleddown physical converter models via digital-to-analogue conversion and
signal amplification. It appears, therefore, that the main restriction is the
availability of reliable data, a restriction which affects equally the physical
simulator.
Some existing AC-DC models, whether physical or mathematical, tend to
concentrate on the representation of the DC-link components, and the
dynamic behaviour of the AC-system plant, and particularly the generators,
is oversimplified. Other mathematical models put the emphasis on the ACnetwork behaviour and use a simplified equivalent for the DC system. Yet
it should be obvious that one without the other will lead to inaccurate
prediction of the complete fault development, including recovery.
This point is illustrated with reference to the recovery transient following a three-phase short circuit at the inverter end, when the faulted AC
system is represented as a time-invariant equivalent circuit; the results, illustrated in Figure 8.6a, show recovery voltages much in excess of the steadystate values. When the dynamic behaviour of the generators in the fault
system is represented in detail, i.e. the equivalent AC system during and
subsequent to the fault is made time variant, the results, illustrated in
Figure 8.66, show that the recovery voltages are well below the nominal
voltage levels.

8.4 AC-system faults


Following an AC fault, the depressed voltage at the converter terminals of
an HVDC link will either reduce or eliminate the power transmitted by the
link. Under such conditions optimum control strategies must be applied so
that normal operation is resumed as soon as possible after fault clearance.

I
3

9.0

10.

11.0
Time (cycles)

Figure 8.6a

Inverter AC voltages following a three-phase fault using a time-invariant AC circuit model

Fault development and protection 213

214

High voltage direct current transmission

However, the DC current cannot be reset instantly to the original value


because of the time constants involved in the DC controls and DC circuit.
Moreover, the AC fault will have altered the reactive-power requirements
and, consequently, the converter voltages.
Since the DC-power transfer consists of the product of voltage (V^) and
current (Ij), any control strategy aiming at fast power recovery needs to
take into account both the voltage and current behaviour.

8.4.1 Three-phase faults


The severity of a three-phase short circuit is greatly reduced as compared
with an alternative AC interconnection because the DC link, owing to its
fast current controller, feeds virtually no additional current into the fault.
If the fault occurs on the rectifier side no special control action needs to
be taken. Provided there is some commutating voltage, the rectifier will
continue operating with the highest possible direct voltage and, when the
fault is cleared, the rectifier can again recover without special action from
the control system.
A short circuit occurring sufficiently close to the inverter end causes
commutation failures, thus producing large DC-current peaks. These are
minimised by quickly reducing the firing angle of the valves (i.e. giving
more time for valve extinction).
Typical simulation results of a fault close to the inverter, and cleared
after five cycles, are illustrated by the DC-power transfer behaviour in
Figure 8.7. In practice, a low voltage limit is applied following an AC fault
close to the receiving end, as has been explained in Section 5.5.3.
The speed of recovery is a question of optimisation having regard for
the AC-system impedance and voltage and current gradients. Figure 8.8
illustrates a TNA simulated study8 of the resumption of power transmission following a three-phase fault at the inverter end. It refers to a backto-back scheme with a short-circuit ratio of three, and the Figure shows the
behaviour with optimum gradients of current and voltage.

8.4.2 Unsymmetrical faults


In the case of an unsymmetrical fault there is sufficient commutating
voltage for continued operation of the link under reduced power conditions; modern thyristor schemes, however, derive the firing circuitry power
from the valve-winding AC voltage and any protection scheme must take
into account that this constraint is met, if operation is to continue. In practice, the gate-control power supply contains sufficient energy to continue
operating for a considerable time during a disturbance (e.g. three-quarters
of a second in the case of the new Cross-Channel UK valves).
The question of re-establishing full power transmission quickly after
fault clearance is very similar to the case of a three-phase short circuit.

10

20

30

40

50

60

70

80

90

100

110
Time (cycles)

Figure 8. 7 Converter DC power following a three-phase fault at the inverter end


Pi = inverter power waveform
Pr = rectifier power waveform
IsO

216

High voltage direct current transmission

100 mA

50 V
Ud

Figure 8.8

TNA study of a three-phase fault at the inverter end

Because of the lack of symmetry, pronounced double-frequency modulation is introduced on the DC side^ which, in the presence of a weak AC
system, will produce heavy oscillations. In extreme cases, such as a line-toline ungrounded fault, it may be advisable to interrupt operation while the
fault persists.
Figure 8.9 shows the response to a staged 60 ms single phase-to-ground
fault placed on a 220 kV line near Haywards in the New Zealand system
with the bipole transmitting 600 MW south.

8.5 DC-line fault development


The main characteristic of a DC-line short circuit is that once started,
owing to any permanent or temporary fault, it will not be extinguished by
itself until the current is brought down to zero and the arc deionised. DC
faults caused by lightning on overhead lines are often self clearing,
because they deionise at the current zeros owing to line oscillations;
however, this can hardly be guaranteed.
As the fault occurs, the line voltage collapses, the rectifier current tends
to rise and the inverter current tends to fall. The inverter will then
advance its firing angle, if necessary beyond 90 into rectification (causing
a small reversal of voltage on its side of the line) to provide sufficient
voltage to maintain the inverter set current.

Fault development and protection 217


2000-.

1000-

2.6

2.7

2.8

2.9

3.1

3.4

3.5

3.2

3.3

3.6

time, s
2000-.

1000Q
CO

2.6

2.7

2.8

2.9

3.1

3.2

3.3

3.4

3.5

3.6

time, s
2000-1

2.6
time, s
100

2.6

Figure 8.9

Staged AC fault

The rectifier, on the other hand, will increase its firing delay and maintain its own current setting.
Thus normal converter control is not adequate to reduce the fault
current to zero; however, by suitable control action, the current can be
reduced to zero very rapidly as compared with AC protection. In practice,
this is achieved by driving the two terminals temporarily into inversion

218

High voltage direct current transmission

and thus clearing the energy stored in the DC circuit faster. This action
requires a reversal of the rectifier voltage, following fault detection. The
inverter already has the correct polarity, but it must be prevented from
going into rectification by setting a limit to its firing-angle advance, ft.

8.5.1 Fault detection11


The voltage and current gradients of the travelling waves set up by the
fault provide the basis for fast fault detection and discrimination. Their
polarity with respect to the line voltage holds sufficient information to
identify bipolar and monopolar faults, as well as the poles involved. A
monopolar fault can also induce overvoltage on the healthy pole owing to
mutual coupling.
Fault location and line characteristics affect the overvoltage magnitude
at the terminals, but converter controls have practically no effect on the
first wave reflections at the terminals, which in general involve the peak
overvoltages.
In the case of a line fault, the rate of fall of voltage at the rectifier terminals is higher than that for converter or AC-system faults, since in the
latter case there is much more inductance in the circuit. However, with
high resistance ground faults close to the inverter end of a long line, information based exclusively on voltage magnitude and rate of change may
not be sufficiently reliable.
For fast DC-line discrimination it may be better to use the weighted sum
of the direct voltage (Vj) and current (Ij) gradient, i.e.

which is directly related to the travelling waves initiated by the fault and
contains information from which fault type and location can be determined. In a bipolar DC line each pole will require this type of detection.
Although the above considerations have been made in relation to the
rectifier end, the inverter end should be equipped with a similar detection
scheme (but with different settings) to ensure fast arc extinction.
The sensitivity of the settings of the wave-front detectors has to be
assessed by means of actual line tests. For instance, in the case of the
Nelson River Bipole 2,12 and for faults at the remote end of the DC line,
the steepness values obtained from an early simulation were rather different from those encountered in the actual tests, i.e. 0.3 kV//is on an HVDC
simulator, 0.7 kV/fis from digital computer simulation and 3.7 kV/fis for
the actual plant tests. The digital mode has since been improved and
provides more accurate simulation.
DC cable faults are generally permanent and fast detection is not
normally used; it is, however, important to provide a very reliable faultdetection and location scheme.

Fault development and protection 219

8.5.2 Fault clearing and recovery


As indicated earlier, on detection of a DC-line fault, the rectifier firing
angle is delayed into the inverting region (say a = 120 to 135) to speed
up the rectifier current collapse, and is kept at that value until arc extinction and deionisation are likely to be completed. Similarly, to ensure that
the inverter end maintains its correct line-voltage polarity, it is necessary to
limit the inverter firing angle advance (say, p < 80).
However, the presence of capacitance (particularly with cable transmission) and inductance can create large overswings and polarity reversal of
the DC-line voltage at the inverter end.
On completion of the deionisation period the restart procedure can
begin in order to restore normal voltage and prefault power. If re-energisation at full voltage is not acceptable (e.g. owing to wet or dirty insulators), then a lower voltage may be used bypassing one or more of the
bridges. However, this type of action is not available on the modern 12pulse schemes.
A starting order is needed to release the emergency control systems of
the converters during the fault. The restart time required will depend on
the properties of the DC line and the converter controls. Computer studies
carried out in the New Zealand DC system11 indicate that a better performance is achieved by overriding conventional current control and using an
exponential function to control the recharging of the DC line. Following
completion of the deionisation period, the rectifier firing angle is stepped
from, say, 125 to 90 over one firing instant; subsequent firing action is
then controlled by the restart function
where OCQ is the control angle which will give nominal line voltage, A9 is the
elapsed time since the beginning of restart control action (i.e. from when a
= 90) and k is a constant controlling the rate of response.
Throughout the recharge period the inverter continues operating under
extinction angle control (i.e. yt = y0)-

8.5.3 Overall dynamic response


The results of simulator studies to optimise the fault development of the
Nelson River Bipole 2 scheme12 are illustrated in Figure 8.10. The faultclearing and recovery action at the rectifier end is controlled by a fixedrate ramp as can be seen from the trace rectifier a-order. The inverter is
clamped at a fixed-delay angle during the fault and during restart.
The response of the realHVDC transmission scheme to a DC-line faultprotection operation (no real fault involved) is illustrated in Figure 8.11.
The DC-line response is more oscillatory in the actual system; otherwise
the simulated and actual responses are very similar.

220
250

kV

225

kV

High voltage direct current transmission


Rectifier D.C. line voltage

Inverter D.C Line voltage

HUB!! ll!!!!!!!!!!!!!!!!!!!!!!

1800 A
1800 A

Rectifier D.C. current l<*


Rectifier D.C. Current ordei

! liillliiiiiiiiiiiiiiiiiii ifiliiii

1800 A

IIIIIIIIIIIIIH!!!!!!!!"!!!!!!!!!

1800

Inverter D.C. current /<*>


Inverter O.C. current order
Rectifier a-order

180"
Inverter a-order

120*
180*
83 ms

Figure 8.10

DC-line fault on an HVDC simulator after the introduction of an


a-ramp in the rectifier and an oc-clamp in the inverter ( 1980
IEEE)
Inverter a-order

180

Inverter bipole power

422 MW
222 kV

1900 A
1900 A
250 kV
1900 A

1900 A

Inverter D.C. Line voltage

Inverter D.C. Current order


Inverter D.C. current /dS

Rectifier D.C. line voltage


Rectifier D.C. Current /dS

Rectifier D.C. Current order

Rectifier a-order
180

Figure 8.11

HVDC system response to DC-line fault protection operation


(redrawn from oscillograms of both substations) ( 1980 IEEE)

Fault development and protection 221

A typical DC-fault development (relating to the New Zealand system


parameters) obtained with a digital model5 and with the strategy described
in Section 8.5.2 is illustrated in Figure 8.12.
The results illustrated in Figures 8.10-8.12 clearly indicate that the
behaviour of a DC system on a line short circuit compares very favourably
with a three-phase short circuit in an AC transmission line. This is owing
to the presence of the smoothing reactor and the speed of current controllers.
The fault current does not exceed two to three times rated current, and
after about 10 to 20 ms no more than the set current remains. As explained
in Sections 8.5.1 and 8.5.2, HVDC transmission schemes involving overhead-line transmission are provided with various DC-line protection functions, i.e. to detect the line fault, to force the DC-line current to zero in
order to extinguish the earth fault and finally to establish a fast recovery.
The DC voltage and current are normally restored at controlled rates to
minimise recovery transients. The fast control permits very short interruption times in the range of 100 to 200 ms.

8.6 Overcurrent protection


With mercury-arc converters the highest overcurrents are produced as a
result of backfires, particularly when they develop into a consequential
backfire in another valve. This situation, described in Section 8.2.3, causes
overcurrents of typically ten per unit and often requires AC circuitbreaker tripping to protect the converter plant, particularly the transformer.
In the case of thyristor-based HVDC converters the overcurrent protection settings are decided with reference to the overload capability of the
thyristor valves, because of their smaller heat dissipation compared to
conventional plant. A typical valve-overload capability is shown in Figure
8.13, which also indicates suggested tripping levels for various disturbances.13
The steady-state short-circuit current of an uncontrolled rectifier is
limited by the increase in overlap angles, which leads to simultaneous
commutations on both halves of the bridge and zero DC-voltage periods.
However, this situation will not be permitted to develop in practice as the
fault current is normally reduced by control action. Of more concern,
therefore, is the transient short-circuit behaviour which, as indicated in
Section 8.3, requires detailed transient simulation of the converter and AC
and DC-systems behaviour. Such studies are necessary to provide adequate
overcurrent protection without unnecessary trippings.
Disturbances causing overcurrents in the converter valves can be divided
into three groups, i.e. internal faults, DC-line faults and inverter-end
commutation failures.

I
I

0.0
12.0

14.0

16.0

18.0

20.0
Time (cycles)

Figure 8.12a

Direct current waveform during a DC-line fault


(i) Rectifier end
(ii) Inverter end

12.0

Figure 8.12b

Direct voltage waveforms during a DC-line fault


(Hi) Rectifier-line end
(iv) Inverter-line end

14.0

16.0

18.0

20.0
Time (cycles)

224

High voltage direct current transmission


id

1\

Valve capability

100
Internal faults

External faults j Overload

Group diff. protection i Line protection I Overload protection


Overcurrent
protection
Fast tripping

j Auto reclosure | Automatic limitation


| of set values
program
I
V
|
| Line undervol- |
I tage protection |
I Set value
I reduction

Figure 8.13

Disturbances
in normal operation

Normal switch off

J
'

Co-ordination of overcurrent protection ( 1978 CIGRE)

Commutation failures may cause long duration overstress if they occur


as a consequence of another fault in the receiving-end system. Their effect
on the rectifier-end current is relatively small, since they are separated
from the rectifier valves by two smoothing reactors and the DC line.
The protection of internal and DC-line faults is considered next under
the subjects of valve-group and DC-line protection, respectively.

8.6.1 Valve-group protection


Internal faults cause severe overcurrents because the impedance between
the fault location and the AC-system source is small. A terminal-to-terminal short circuit across one valve, although a rare event, produces the
highest overcurrent stress in other valves; typically, a current peak of ten
per unit can be expected. Normally, fast suppression of the firing pulses
should block the short-circuit current, provided that the valves are capable
of withstanding the recovery voltage immediately after the fault. However,
if the valve is not able to block, the only way to avoid repetitive overcurrent peaks is by immediate tripping of the AC circuit breaker.
A typical13 overcurrent protection scheme, used for the detection of
internal faults in a modern 12-pulse group converter, is illustrated in
Figure 8.14. The group differential protection compares the rectified AC
current with the DC current. This unit-protection scheme provides speed
and selectivity.
If the comparison shows that the DC is greater than the AC current
(when the firing angle is in the inverting region) this indicates a commutation failure and the only action required is an automatic increase in the

Fault development and protection 225

Figure 8.14

Overcurrent protection for a converter station with one 12-pulse


group per pole ( 1978 CIGRE)

extinction angle, y. However, if the condition persists, a temporary block is


initiated.
The case of the AC being greater than the DC current indicates either a
backfire (only with mercury-arc converters) or a short circuit in the valve
group and requires circuit-breaker tripping.
A nonunit overcurrent protection scheme is also used as a back up, with
a higher tripping level, to avoid tripping action during faults outside the
station which can be cleared by control action.
Finally, the detection of ground faults on the DC side relies on a poledifferential protection scheme, which blocks the converter valves and trips
the AC breaker of the affected pole.

8,6.2 DC-line protection


DC faults are more frequent than internal short circuits and are mainly
caused by lightning. The current amplitude is limited by the smoothing
reactor and by control action to typically two to three per unit.
The detection scheme includes rate of change of voltage (which
responds after some 3 ms) and a slower back-up line undervoltage relay
(which responds after, say, 50 ms). With parallel lines it is also common to
include a rate of change of current comparison scheme to indicate which
parallel line is faulty.

226

High voltage direct current transmission

The rate of change is set to discriminate between ground faults and


lightning surges not resulting in ground faults; commutation failures at
the other end of the line produce less steep voltages at the rectifier end
(since there are two smoothing reactors in between) and are also ignored.
A back-up undervoltage unit is also used to detect high-resistance DCline faults, where the rate of decrease of DC-line voltage is slow. Although
normal current control can limit the current to a small value, this action is
not sufficient to extinguish the fault arc. The fastest way of bringing the
line current to zero is to delay the rectifier firing into the inverting region;
as a result both stations are temporarily inverting to transfer the energy
stored in the DC-circuit electric and magnetic fields into the two AC
systems.
DC-line protection is not needed in back-to-back interconnections as the
two converters are located in the same building.

8.6.3 Filter protection


The capacitor banks are made up of series-connected racks of capacitor
units in parallel, each unit having an external fuse. However, to prevent
frequent prolonged shut downs to replace fuses, the filter arms are also
equipped with overcurrent relays. These have an extremely inverse characteristic and respond equally to fundamental or harmonic-frequency overcurrents.
A restricted earth-fault relay is also provided which operates during an
earth fault within any filter arm.
The filter banks are switched by circuit breakers and the filter arms
within the banks by disconnect switches.

8,7 References
1 CIGRE WG 14.02.: "Commutation failure in HVDC transmission systems due to
AC system faults'. Electro, December 1996, (169), pp.59-85
2 CHRISTOFERSEN, D.J., ELAHI, H., and BENNETT, M.G.: 'A survey of the
reliability of HVDC systems throughout the world during 1993-1994'. Paris
1996, CIGRE, paper 14-101
3 ARRILLAGA, J., and GIESNER, D.B.: 'Recovery of mercury-arc h.v.d.c. interconnectors from backfire faults'. Proc. IEE, 1972, 119 (11), pp.1611-15
4 MORALES, M.: 'Sequential arrangements for the elimination of bypass valves
in high voltage direct current converters'. PhD thesis, Manchester University,
UK, 1965
5 ARRILLAGA, J., ARNOLD, C.P., and HARKER, B J.: 'Computer modelling of
electrical power systems' (John Wiley, UK, 1983)
6 RTDS Technologies Inc.: 'Real-time digital simulator users manuals'. Rev.
February 1996
7 HEFFERNAN, M.D.: 'Analysis of a.c.-d.c. system disturbances'. PhD thesis,
University of Canterbury, New Zealand, 1980
8 KAUFHOLD, W., and POVH, D.: 'Recovery of the h.v.d.c. transmission after

Fault development and protection 227


faults in the a.c. system'. IEE Conf. Publ. 205 on 'Thyristor and variable static
equipment for A.C., and D.C. transmission', London, 1981, pp. 171-75
9 GIESNER, D.B., and ARRILLAGA, J.: 'Behaviour of h.v.d.c. links under unbalanced a.c. fault conditions', Proc. IEE, 1972, 119 (2), pp.209-15
10 DE SILVA, J.R., et al: 'Commissioning of the New Zealand HVDC hybrid link'.
International Colloquium on HVDC and FACTS, Wellington 1993, paper 1.6
11 HEFFERNAN, M.D., ARRILLAGA, J., TURNER, K.S., and ARNOLD, C.P.:
'Recovery from temporary h.v.d.c. line faults', Trans. IEEE, 1980, PAS-100, (4),
pp. 1864-70
12 MAZUR, G., CARRYER, R., RANADE, S.T., and WEB, T.: 'Converter control
and protection of the Nelson River h.v.d.c. Bipole 2 - Commissioning and first
year of commercial operation'. Trans. IEEE, 805M6T4-2, Minneapolis, 1980
13 KAUFERLE, J., and POVH, D.: 'Concepts of overvoltage and overcurrent
protection of h.v.d.c. converters'. CIGRE, Paris, 1978, paper 14-08

Chapter 9

Transient overvoltages and insulation


co-ordination

9.1 Introduction
There are some fundamental differences between the types of overvoltage
experienced in AC and DC transmission schemes, the main differences
resulting from:
(a) The commutation phenomena between the converter devices which,
even during normal operating conditions, result in complex voltage
waveforms as shown in Figure 9.1.
(b) The combination of direct and alternating (or transient) voltage.
The bridge units of multigroup converters are normally connected in
series on the DC side and in general they are symmetrically placed with
respect to earth, as shown in Figure 9.2. The number of energy-storing
elements increases in proportion to the number of bridges on the same
side of earth and this results in growing voltage-to-earth stresses on the
bridge valves as their separation with respect to earth increases; the stress
across each valve is the same, however.
According to their source of origin converter overvoltages can be
divided into:

overvoltages excited by disturbances on the DC side;


overvoltages excited by disturbances on the AC side;
fast transients produced at the converter itself;
fast transients of external origin (i.e. lightning or switching-type
surges).

These are first considered separately and their effect is taken into account
in the insulation co-ordination of the converter station.

Transient overvoltages and insulation co-ordination 229


Peak transient at load disconnection
(commutation overshoot included)

Highest continuous peak


(commutation overshoot included)

Figure 9.1

Voltage waveforms during normal operation


a Voltage across a rectifier bridge
b Phase-to-phase voltage on the valve side of the converter transformer

P
Figure 9.2

Multigroup converter with bridges in series

9.2 Overvoltages excited by disturbances on the DC side


In order of increasing relevance, three types of DC-system disturbance
need to be discussed.
(i)

(ii)

When a short circuit occurs in one pole of an HVDC bipolar transmission system, a transient overvoltage is induced on the healthy
pole. However, the level of overvoltage will be limited to under two
per unit by control action and does not constitute a decisive factor
for the insulation coordination.
The re-energisation or deblocking of converter bridges can be a
cause of large transient overvoltage. Figure 9.3 shows the overvoltage which can develop on the DC line when the rectifier is
deblocked with full rectifier voltage, against an open inverter end.1

230

High voltage direct current transmission

Deblocking with full rectifier voltage against an open inverter end

Figure 9.3

(iii)

However, the start control normally raises the rectifier voltage


gradually and the condition shown in Figure 9.3 can occur only as a
result of a fault in the scheme.
The most severe voltage oscillations across a valve will occur in the
event, unlikely but possible, of an earth fault on the valve side of
the DC-reactor's bushing; however, this overvoltage does not affect
the line insulation. Figure 9.4 shows the maximum rate of rise of
calculated overvoltage across the thyristor valves of the Cahora
Bassa scheme2 with an earth fault at point a or b.

U(kV)

>

600-

500

s U'i(a)

400-

A7'A /

300200100-

0
-100-

.2 0 4 0.6 0.8 1

t/julS

-200-

Figure 9.4

Calculated maximum rate of rise of earth-fault overvoltages across the


valves with faults inside the converter station ( 1974 IEEE)

Transient overvoltages and insulation co-ordination 231

9.3 Harmonic overvoltages excited by AC disturbances


The subject of steady-state harmonic distortion has been discussed in
Chapters 3 and 6; this section considers the problem of transient harmonic
overvoltages excited by AC disturbances.
AC harmonic filters present a capacitive impedance for frequencies
below their lowest tuned frequency. If the AC system is inductive, as is the
usual case, there is the possibility of a parallel resonance between the filter
and the system at one of these low harmonics of the supply frequency (e.g.
at the third or fourth harmonic in the presence of fifth-harmonic filters),
and the lower resonant frequencies will usually result in increased overvoltages.
Following disturbances such as transformer switching, load rejection,
AC-system faults etc., the resonant circuit can be excited, generating
harmonic voltages superimposed on the fundamental frequency, an
example of which can be seen in Figure 9.5.
When transformers are driven into saturation they draw harmonic
currents and, if the AC system exhibits high impedance as a result of resonance at one of these harmonic frequencies, then harmonic voltages of
considerable amplitude can be superimposed on the fundamentalfrequency voltage.
As explained in Section 6.3.1, the most important overvoltages occur
during three-phase-faults clearance.
The level of harmonic overvoltage is very dependent on the effective
system resistance, and therefore on the impedance angle of the system at
the harmonic frequency.
Figure 9.6 indicates the statistical overvoltages as a function of impedance angle and system impedance, derived from many hundreds of tests
carried out on the IREQ HVDC transmission simulator.3 The Figure illustrates that the lower the impedance angle the smaller the harmonic overvoltages. The presence of local load at a rectifier station therefore tends to

1.0 |

0.5 I
|
0.0

-0.5
-1.0
Figure 9.5 Harmonic voltage distortion following an AC fault

232

High voltage direct current transmission

3rd harmonic resonance


2.0
System impedance
constant at 0.33 p.u.

System impedance angle


constant at 85

4th harmonic resonance


1-0 I , r
2.5 3.0

4.0

5.0

6.0

7.0

System short circuit ratio = ,


Z(pu)

Figure 9.6

1.0

75

80

85

System impedance angle

Statistical overvoltages on transformer energisation as a function of


system impedance and impedance angle ( 1980 IEEE)

reduce the harmonic (transient) overvoltages, in contrast to the increase in


regulation (dynamic) overvoltages described in Section 6.3.
Unbalanced AC faults cause the injection of second-harmonic voltage
and current on the DC line,4 which may excite oscillations if the natural
frequency of the DC line coincides with this frequency.

9,4 Overvoltages owing to converter disturbances


Internal converter disturbances such as a repetitive misfire or commutation
failure in one valve can inject fundamental-frequency voltage on the DC
line. An example of fundamental-frequency resonance which occurred
during early operation of the Cahora-Bassa scheme5 is shown in Figure
9.7.
Another important source of overvoltage on the DC line occurs when
the inverter firings are blocked without bypass action. In this case, two
valves (per bridge) remain conducting and a large AC voltage is injected
on the DC line. Voltage and current oscillations on the DC line may then
cause current extinction at the inverter followed by a considerable overvoltage, owing to the rectifier continuing to provide energy into the DC
system.
A properly designed converter current-control system is very effective in
reducing DC overvoltages and oscillations. Even when the line is resonant
to the frequency of the injected AC voltage, a good controller will limit
the overvoltages to less than fifty per cent of the line voltage.

Transient overvoltages and insulation co-ordination 233

VWW'wx,

Figure 9.7

Line current and voltage recorded at the inverter during missing


puke condition in a rectifier bridge ( 1980 CIGRE)

9.5 Fast transients generated on the DC system


This category includes the following:
(a) Fast front surges of external origin and of shapes similar to the lightning surge. The front and tail times of the lightning surge tend to fall
within reasonably closely-defined limits; for most purposes it is taken
that a voltage surge associated with lightning may be represented by a
unidirectional double-exponential wave, rising to crest from zero in
about 1 fisy and falling to half the amplitude in 50 us. This is the socalled standard lightning impulse, or 1.2/50 us wave.
(b) Slow front surges with shapes similar to the switching surges, e.g. with
250 jus front and 2500 us tail.
The characteristics of lightning and switching surges are well documented
and are only considered here in as much as they affect the insulation coordination of the converter plant.

9.5.1 Lightning surges


The steep wavefront of a lightning surge travelling along the DC overhead
line is first slowed down by the line surge impedance. The propagation of
the incident surge into the converter station depends on the characteristics
of the smoothing reactor and DC-line surge capacitor.
When calculating the effect of lightning surges the DC-line surge capacitor is normally replaced by its capacitance and an associated inductance
(representing the connections inside the capacitor and to the line); the DC

234

High voltage direct current transmission


kV
2000-

1500

1000

500
Microseconds

Figure 9.8 Lightning surge IQ- 5 kA propagating into the converter station
a Main-circuit scheme
h Surge-capacitor voltage
parameter T:risetime of the lightning surge on the DC line before
the surge is influenced by the surge capacitor
surge capacitor C() = 0.1 fiF
surge capacitor connections LQ = 100 fxH
stray capacitance across the smoothing reactor C\\ = 1000 pF
DC-line wave impedance ZQ

reactor coil behaves as an inductance in parallel with an effective interturn


capacitance as shown in Figure 9.8a.
By way of example the effect of a 5 kA incident surge reaching the
converter termination is illustrated in Figure 9.86, using the rise time T as
a variable parameter.6 These sharp voltage oscillations can be reduced
substantially by providing ground wires in the vicinity of the converter
station combined with low tower-footing resistance. For a typical increase
of surge-front time of 1 fis/km, Figure 9.8 shows that the provision of a 2
km long ground wire can reduce the rise time of the surge across the
surge capacitor to 2 /is.

Transient overvoltages and insulation co-ordination 235

9.5.2 Switching-type surges


Owing to capacitive and inductive coupling between the conductors, earth
faults on one pole give rise to surges on the healthy pole similar to those
caused by switching (i.e. long front and long tail).
For the analysis of surge transfers to the converter side the stray inductance of the surge capacitor and the shunt capacitance of the smoothing
reactor can be neglected because of their small time constants (relative to
the wavefront time).
Transient network analyser tests indicate that the peak can reach up to
1.8, depending on line distance and termination.
For a converter in its normal conducting mode, the distribution of the
switching surge is governed by the amount of inductance existing between
the points concerned and ground.
The voltages across both the upper and lower bridges are mainly determined by the voltage drops across the conducting phases of the transformer belonging to the bridge in question.
Switching surges of long duration and high energy content can cause
current levels of sufficient magnitude to produce current extinction;8 this
is illustrated in Figure 9.9b. Since the inductance of the smoothing reactor
is much larger than that of the converter transformer leakage, the incoming surge has little effect on the bridge voltages while the valves are
conducting. However, when the incoming surge extinguishes the valve
current, the station side of the smoothing reactor is connected to earth
through the valve grading, damping circuits and stray capacitances only;
this causes an overswing of higher value than the incoming surge.
Figure 9.9 shows the overvoltage distribution for three bridges and the top
transformer connections.
It is worth pointing out that overvoltages are often the result of certain
sequences of events occurring in practice; caution must therefore be used
in using purely speculative theoretical studies with possibly incorrect
assumptions.

9.6 Surges generated on the AC system


Switching surges generated in the AC side are inductively transferred to
the valve side via the transformer. Since the valve-group voltage only
consists of phase-to-phase components, the zero sequence is not included
in the converter voltage.
A single-phase superimposed surge is transferred to the valve side in
different ways in the star-star and star-delta connected transformers of a
two series-connected valve group, the total peak voltage across the group
being given by6

236

High voltage direct current transmission

Group 3

Group 2

Group 1

Figure 9.9

DC-line switching surge causing valve current extinction


a Circuit scheme after valve-current extinction
b Current and voltage
A inductive voltage distribution between bridges
B capacitive voltage distribution between bridges
Id line current
Ufi direct voltage/bridge
^B kfc> % voltage to ground from DC busbars
L7D1 voltage to ground from AC bus group 3
UL voltage to ground on the DC line

Transient overvoltages and insulation co-ordination 237

where n is the transformer turn ratio (star-star), Vm is the maximum


permissible steady-state operating phase voltage and Vo is the peak value
of the superimposed switching surge.
The converter stations and transmission lines in their vicinity are
normally shielded to prevent direct lightning strokes. Strokes distant from
the converter station are attenuated by the time they reach the converter
station. Moreover, the steep-front overvoltages generated during lightning
are subjected to a considerable front prolongation due to the capacitive
termination. The tuned AC filters and capacitor banks have considerable
damping effect on the incoming wave; most of the filter damping is due to
the high-pass branch. At low load, however, a number of capacitor banks

2 p.u.

p*W

1.7

1-5 "

50
Microseconds

Figure 9.10

Capacitive and inductive voltage transformation through the converter, transformer of a lightning surge with the amplitude 2 p.u.
superimposed on the operating voltage on one phase
parameter Ct: capacitive coupling between the valve and line windings of the transformer
transformer ratio 1:1

238

High voltage direct current transmission

may be disconnected but the maximum probable overvoltage factor is not


likely to exceed 2.2 p.u.
Lightning surges in the AC system can be transferred to the DC side via
the transformer both electrostatically and electromagnetically. Generally,
the electrostatic transfer can be neglected since only waves with a steep
wave front are transferred via the stray capacitance.
An example illustrating the capacitive and inductive transformation of a
lightning surge is shown in Figure 9.10. The voltage increase after 60 /is is
seen to depend upon electromagnetic transfer.
However, when the operating voltage in the AC-system side of the
converter transformer is much higher than the corresponding voltage on
the DC side, the capacitive transformation may be of importance; for
transformer ratios larger than four the capacitively transferred component
is larger than the electromagnetic.

9.7 Fast transient phenomena associated with the converter


plant
Comprehensive computer programmes are required to assess the internal
distribution of transient-voltage stresses within the converter upon the
arrival of surges of either internal or external origin.
The physical behaviour and circuit configuration of thyristor and
mercury-arc converters are very different in this respect and are thus
considered separately.

9. 7.1 Mercury-arc converters


With reference to the general model of a 12-pulse mercury-arc converter
suitable for fast transient studies, shown in Figure 9.11, the following
criteria apply to the energy-storing elements during operating conditions:
(a) The anode-to-cathode self capacitance is only effective when the valve
is not conducting, the self capacitance of transformer windings is only
effective in the idle phase and the cathode-to-neutral capacitance is
effective at the start and end of every commutation.
(b) The capacitor, damping circuit and anode reactor associated with the
valve exist or disappear according to the three different states of the
valve, i.e.
conducting - all of them are short circuited,
starting conduction - all of them exist,
stopping conduction - only the anode reactor disappears.
Moreover, the transient response of the passive circuit depends on the
initial values of the instantaneous nodal voltages and injected nodal
currents at the instant when the discontinuity takes place.

Transient overvoltages and insulation co-ordination 239

240

High voltage direct current transmission

40

600

Figure 9.12

Converter-voltage distribution for a 1000/1/23


(nonconducting converter)

a
b
c
d
e

incident surge

On valve side of DC reactor


Across upper valve of upper bridge
Across upper bridge
Across lower bridge
Across upper valve of lower bridge

To represent the different operating conditions encountered in practice,


the general circuit must be modified by eliminating or short circuiting
some of the branches.
By way of example, the propagation of a 1000/1/23 incident surge
across the 0.5 H smoothing reactor, and the internal overvoltage distribution in a typical double-bridge mercury-arc converter,9 are illustrated in
Figure 9.12.
The equivalent circuit of Figure 9.11 can also be used to optimise the
design of the damping circuits used to limit commutation oscillations. This
is carried out by performing a large number of studies, representative of
the different types of normal and abnormal commutations,10 with delay
angles producing the highest voltage jump in each case.
For the analysis of arc-quenching phenomena,11 the faulty valve is simulated by an open switch shunted by the valve stray capacitance in series
with the anode reactor represented by a parallel combination of inductance, resistance and capacitance.

9. 7.2 Thyristor converters


The modelling of transient phenomena in thyristor converters must take
into account the thyristor nonlinearities such as recovered charge, leakage
and displacement-current phenomena associated with the various thyristor

Transient overvoltages and insulation co-ordination 241

operating states and the diversity which exists in those properties between
the individual thyristors of the valve structure. These characteristics interact in a complex manner with other converter components such as the
saturating inductors, transformers, damping and grading circuits, stray
capacitances and inductances of the valves, busbars, transformer windings
etc.
Computer simulation12 is used to assess the internal distribution of transient voltage stresses and other related information such as cascade turn
on, overvoltage limitation, protection and co-ordination, valve recovery at
turn off, voltage unbalance along a series-connected string of thyristors
and transient overvoltage disturbances.
If, in a valve equipped with independent overvoltage firing at each
voltage level, a component failure causes one level in the valve to rely on
this protection for triggering, abnormal voltage excursion and inrush
current in excess of the normal will be imposed on the afflicted level.
Since this operating regime can persist repetitively until the next scheduled
maintenance, it is of crucial importance to valve-component ratings.
Thyristor valve

Simplified
external circuit

Figure 9,13

Simplified equivalent circuit for cascade turn-on investigations

242

High voltage direct current transmission

When the levels in a valve fire noncoherently, the voltage across the last
level to turn on will rise at a relatively fast rate. The overvoltage protection
gates the thyristor when its switching-threshold level is exceeded. However,
because of the turn-on delay of the thyristor, a finite time passes before a
thyristor impedance falls sufficiently to establish a safe conduction path.
The simulation programme is then used to ensure that during this interval
the thyristor and other components are not overstressed owing to excessive
current, voltage or rate of rise of voltage, and that all components are
adequately rated.
A simplified equivalent circuit used to study the cascade firing is shown
in Figure 9.13 for a valve employing n levels with saturating inductance

Cascade firing
O)

Time (\xs)
(a) Thyristor voltage waveforms.

Cascade firing

CD

Time (us)
{b) Thyristor current waveforms.

Figure 9.14

Typical thyristor voltage and current waveforms for normal and


cascade firing

Transient overvoltages and insulation co-ordination 243


distributed equally between the levels. The circuit external to the valve
represents the bridge in respect of valve-inrush current over the period of
interest. The valve is split up into two parts, one part representing the late
firing level and the other simulating the rest of the (n - 1) levels; the impedances of the (n - 1) levels are taken with minimum tolerance and those of
the late firing level are at maximum tolerance (this maximises the voltage
on the last level to turn on).
The equivalent circuit of Figure 9.13 is initially charged to the appropriate voltage level. At the beginning of the calculation the (n - 1) levels turn
on simultaneously, with their thyristors represented by time-dependent
resistors/During this time, the thyristor in the last level to turn on is simulated by a voltage-dependent capacitor. When the voltage across the thyristor reaches the protection threshold, the thyristor is gated to turn on after
a specified delay. The impedance of the thyristor begins to fall in a
manner determined by its time-dependent switching characteristics, the
sequence being initiated when the gate current attains a preset level.
Typical waveforms of thyristor voltage and inrush current are shown in
Figure 9.14 for the last level to fire. Corresponding waveforms for normal
coherent turn on are also shown for comparison.

9.8 Insulation co-ordination


The generic purpose of insulation co-ordination is the selection of the
most economical combination of plant insulation and overvoltage protection to ensure satisfactory performance of all the insulation around the
converter plant. More specifically, the subject is concerned with the need
to protect the converter plant components against large transient overvoltages for which they cannot be economically designed or compensated.
The limitation of overvoltages, essential to economic power-plant
design, is carried out in two ways:
(a) By suitable system design.
(b) By suitable co-ordination between insulation and surge-arrester protection.

9.8.1 System design


On the DC side, overvoltages can be reduced by various means, i.e.:

shielding the converter station and transmission lines;


suitable design of the converter control equipment;
use of damping circuits;
selecting system parameters to try and avoid resonance under fault
conditions.

244

High voltage direct current transmission

On the AC side, voltage support equipment is essential to provide voltage


control during transient and dynamic system disturbances. The type and
amount of compensating equipment requires detailed studies to determine
the relationship between DC-system recovery (i.e. changes in convertercontrol mode, rate of ramping of the DC current on restart and stabilising
signals) and voltage-support equipment and response times.
The different characteristics of each type of voltage-support equipment
have marked effects on the performance of the AC-DC system which must
be studied and evaluated. For example, an AC-system fault may result in
sufficient voltage reduction and distortion to cause commutation failures
at the inverter. With synchronous condensers, as they contribute to the
short-circuit level, such a problem will be less likely than with static VAR
systems (SVS).
The occurrence of commutation failures during conditions of low or
distorted AC busbar voltage, either during a fault or on recovery, requires
realistic representation of the converter valves with their controls and of
the AC system (in particular, system damping and effect of generator
control).
It appears that a combination of static and synchronous compensators
can be the best solution for many applications. The short-circuit capacity
of the synchronous condenser, coupled with the speed of response of the
SVS, should provide better overall voltage control and dynamic performance.

9.8.2 Surge arresters


There is an important difference in the behaviour of AC and DC surge
diverters. Owing to the parallel connections, the AC system has a low
impedance and the follow current in an AC surge diverter, even with
strong arc-suppression characteristics, has little effect on the AC-system
voltage against which the diverter has to reseal. On the other hand, DC
schemes have large smoothing reactors and other inductive components
i.e. the diverter has to suppress the direct current in a highly inductive
circuit. As a result, unless the suppression is done in a controlled way, the
surge diverter might cause further overvoltages. Some early HVDC
schemes used protective gaps but the development of self-sealing surge
arresters changed the situation. Arresters with series gaps and siliconcarbide valve elements with nonlinear resistance have been extensively
used in HVDC schemes.
However, metal-oxide arresters13 have now taken over the overvoltage
protection of HVDC converters. They consist mainly of zinc-oxide (ZnO)
but contain additives of other metal oxides. The main advantages of the
new type are their high discharge capability and the lack of gap spark-over
transient; the latter property is particularly important in the protection of
thyristor valves.

Transient overvoltages and insulation co-ordination 245

Typically, a ZnO disc can carry thousands of amps at twice the nominal
voltage and thus permit the elimination of series-connected spark gaps.
The zinc-oxide arrester voltage characteristic has a very definite knee
and is extremely flat; hence, the arrester will not permit the voltage to rise
without shunting a substantial current to ground. Furthermore, in the case
of a switching surge, all arresters connected to a busbar share the discharge
duty; they draw an increased current out of the overvoltage source and
hence contribute to the damping effect.
With gapless metal-oxide arresters, the arrester elements are continuously subjected to the normal operating voltage of the AC-DC system. The
number of series-connected elements are selected so that only a very low
current flows under normal applied voltage.
An important advantage of the metal-oxide arresters is their ability to
parallel units to achieve the needed energy capability. However, metaloxide arresters have a given physical ratio between the maximum DC
operating voltage and the protective level at internal overvoltages. This
ratio is about 1.9 p.u. at the present state of the art. For arresters with
this protective level only moderate energy capability is required even at
high internal overvoltages. A parallel gap can, if necessary, be added
across part of the metal-oxide arrester in order to limit overvoltages to
lower values.14 The idea is to reduce the arrester current at the operating
voltage by a larger number of elements. However, when an overvoltage
occurs, part of the series-connected elements is bridged over by the parallel gap to reduce the protective level. The use of this arrester needs
careful co-ordination to consider at which conditions the gap should
operate and ensure that it reseals after the operation. It is also possible to
trigger the gap using a signal from the voltage-measuring devices or from
DC control.
An overvoltage limiter can also be designed using antiparallel-connected
thyristor valves in series with the metal-oxide arrester, or using other material such as a metallic resistance. The limiter can be triggered depending
on overvoltage conditions or by DC control. With such equipment even
lower protective levels can be achieved.

9,8.3 Application of surge arresters


As a general principle, a surge arrester must be set to discharge, or divert,
overvoltages higher than the highest normal operating voltage and lower
than the breakdown voltage of the insulation under protection.
Most overvoltages within a converter station are of the switching surge
type. Although lightning surges caused by thunder storms do not enter the
converter bridges, the valves are exposed to lightning-surge-type stresses
during ground faults within the station. It is thus necessary to test the
HVDC equipment with some standard voltage waveforms similarly to
HVAC equipment.

246

High voltage direct current transmission


(825)

825
575

(825)
O

rr^

D.C. arrester
A.C. arrester
(BID BIL to ground
[BIL] BIL across device

To remote
ground electrode

Figure 9.15

Square Butte insulation co-ordination ( 1980 IEEE)

Four examples of surge arrester protection are now discussed:


(i) Square Butte: Figure 9.15 illustrates the insulation levels and protective
margins for the Square Butte converters.15 In this scheme the DC arresters
are zinc oxide of early design and thus include a series gap.
DC arresters are applied on the line, across the bridges, on the neutral
and on the 125 kV bus. Conventional arresters are used across the smoothing reactor, on the transformer primary and across each valve. In addition
to valve arresters, each thyristor level is equipped with forward overvoltage
protection. The valve-arrester characteristics are selected to co-ordinate
with the valve inverse voltage rating and the operating voltage of the valve
forward-protection circuits. In this manner, the valve is transiently
protected against overvoltages, whether owing to actual system disturbances or to potential control problems.
It is interesting to note that phase-to-phase arresters have been omitted
in this scheme because the phase-to-phase insulation is sufficiently high to
be protected by the arresters across the valves (two of which are connected
between the transformer phases).
(ii) Cross-Channel 2000 MW scheme:16 In this example, the emphasis is on
the protection and insulation levels adopted for the AC and DC-systems
equipment.
Figure 9.16 shows the insulation co-ordination of the AC-system equipment. Surge arresters are connected on the 400 kV system from phase to
earth, their main function being the limitation of the maximum energy

Transient overvoltages and insulation co-ordination 247


D.C. Line

20 kA
910 kV
400 kV
Switchyard

60 kA
430 kV

All voltages in kV
Withstands given at 1.2/50 us and 250/2500 \xs

Figure 9.16

AC-equipment insulation co-ordination-arrester protection levels at


8/20 us and current specified

stored in the filter capacitors; this in turn reduces the energy absorption
duty imposed on other surge arresters.
Regarding the filter components, during transient conditions the
prospective voltage across these components may be even higher than the
phase-to-earth voltage. The insulation level of the resistors and reactors
can thus be substantially reduced by using a surge arrester in parallel with
these components.
The surge arresters of the DC-converter equipment are illustrated in
Figure 9.17. The arrester across the thyristor valve is determined mainly
by considerations of maximum continuous operating voltage. For an
economic valve design the protection level of the valve arresters should be
kept as low as possible. In addition to the discharge energy present in the
DC system, valve arresters may be exposed to severe discharge duty during
fault recovery in the AC system; as a result, the arresters incorporate
several parallel columns of zinc-oxide blocks.
The DC-cable arrester is mainly determined by considerations of
required protection level. In the absence of overhead lines, fast transient

248

High voltage direct current transmission

Figure 9.17

DC-equipment insulation co-ordination-withstand voltages in kV and


at 1.2/50 us and 250/2500 jus

overvoltages of significant amplitude will only be caused by flashover to


earth.
(iii) British Columbia Hydro (Stage IV) 140 kV valves:17 Figure 9.18 graphi-

cally illustrates the valve insulation co-ordination and clearly indicates that
the thyristor characteristics are far in excess of the arrester characteristics.
The central column quantities are in per unit (referred to the rate DC
voltage).
The commutation transient peak at 90 firing delay must be less than
the minimum sparkover of the arrester; moreover, sufficient margin must
be allowed for the commutation transient during normally expected overvoltages. Excessive reduction of the commutation transient by damping
resistor-capacitor circuits is avoided (considering the increased losses) by
introducing an inverse-time overvoltage protection scheme which inhibits
operation at near 90 firing delay during excessive AC-system overvoltage.

Transient overvoltages and insulation co-ordination 249


4.5

4.0

Valve arrester characteristics

3.5

5 microsecond test
(negative only)

(475))
Switching surge test

3.0

Maximum impulse sparkover


Maximum 60 Hz and switching
surge sparkover
Minimum 60 Hz sparkover
Minimum switching surge
Minimum impulse sparkover

(356)
(347^

Rated reseal voltage

(237)

2.5

lir

' Maximum permissible dynamic


a.c. overvoltage
(arrester limited)

J240)

Crest commutation transient


at a = 90

1.5
168)
Scale: per unit of 140/kV
(crest volts)

1.0
B.C. Hydro

Figure 9.18

(140)

A.C. line to line crest


Rated valve direct voltage

BC Hydro valve-insulation co-ordination ( 1978 CIGRE)

(iv) A more recent example of arrester co-ordination is shown in


Figure 9.19. It relates to the 600 MW back-to-back Etzenricht station18
between Germany and the Czech Republic that went into operation in
1993. The Figure shows the arrangement of ZnO arresters for one side of
the station.
Arresters of type A are installed close to the converter transformer-line
side bushing to limit overvoltages on the primary and secondary side.
These arresters are designed for the worst energy discharge duty after a
solid ground fault in the substation, followed by the recovery fundamental
frequency and saturation overvoltages assuming that the converter valves
are blocked.

250

High voltage direct current transmission

420 kV

Figure 9.19

Arrester protective scheme

As an option, the back-to-back station is designed to operate with an


isolated neutral connection. In this case an arrester type iV protects the
primary-side transformer neutral; it is designed for fundamentalfrequency overvoltages at the neutral terminal of the converter transformers.
The AC-filter arresters, types Fl and F2, protect the low-voltage filter
components against transient overvoltages during faults limited by the ACbus arresters type A. Lightning-surge type and switching-surge type stresses
are considered to protect the components properly against overvoltages.
The maximum continuous operating voltage and the protective level of
the shunt-reactor arresters (type L) are significantly higher than those of
type A arresters.
B-type arresters protect the individual thyristor valves against overvoltages when the valve is blocked. The main consideration in this case is the
transferred switching-surge overvoltage from the AC side, which is limited
to the switching-surge protective level of the A AC-bus arresters. Therefore, the arrester duties and the thyristor-valve protection level are calculated considering the transfer impedance of the converter transformers
between AC side and valve-arrester side.

9.9 Considerations on cable overvoltage protection


Most of the internal overvoltages occurring in the DC cables have amplitudes lower than 1.5 p.u. and durations in the range of 100 ms. They do
not influence the insulation co-ordination of the cable and there is no
need for their limitation by overvoltage protection, as the DC-cable insulation capability for internal overvoltages at the given operating voltage is
higher.
However some faults, combined with control malfunctioning, can lead
to overvoltages higher than 1.5 p.u. Depending on the assumptions made
regarding possible control failures and on the transmission data used -

Transient overvoltages and insulation co-ordination 251


Table 9.1 Energy dissipation in kWs/kV of metal-oxide arresters when the DC
circuit is resonant at fundamental frequency under normal current control
Arrester protective
level
1.9 p.u.
1.7 p.u.
1.5 p.u.

DC cable for 250 kV

DC cable for 400 kV*

2.8
6.8
10.0

7.2
17.4
25.6

*extrapolated values

especially if a resonance condition at the fundamental frequency is present


in the DC circuit - overvoltages exceeding 2 p.u. can occur. These overvoltages should be limited by surge arresters to reduce the cost of DC-cable
insulation. Examples of these occurrences are loss of firing pulses in inverter or inverter blocking, rectifier start against an inverter open end and a
fault on a DC line in mixed overhead/cable transmission schemes.
Considering that the present ratio between maximum DC operating
voltage and internal overvoltage protecting levels of metal-oxide arresters
is about 1.9, the DC-cable overvoltages must be limited to this value if no
additional measures are taken. For arresters with this protective level only
moderate energy capability is required, as shown in Table 9.1. 19
In mixed DC overhead line/DC-cable transmission, lightning overvoltages can stress the DC-cable insulation. During direct strokes on the DC
line and back flashovers, surges propagate towards the cable terminals and
part of their discharge energy is transmitted into the cable. For long
cables, since the surge impedance of the cable is much lower than that of
the overhead line, the voltage on the cable is reduced and damped before
reflections in the cable can lead to high overvoltages. Therefore, the
reflections are changed into overvoltages of the switching-surge type and
of lower amplitude.
However, if a short cable section is inserted into the DC overhead-line
transmission (e.g. only a few kilometres) multiple-wave reflections in the
cable can result in high overvoltages. To limit these overvoltages it is
important to provide measures which prevent direct strokes and back
flashovers close to the line/cable junction, e.g. the use of low tower
footing resistance and good DC-line shielding by earth wires.
The DC cables must also be protected against lightning by arresters at
both ends.

9.10 References
1 UHLMANN, E., and FLISBERG, G.: 'H.v.d.c. insulation co-ordination, Part 1:
Generation of overvoltages', Direct Curr. Power Electron., 1971, 2, (1), pp.8-14

252

High voltage direct current transmission

2 HEISE, W., BURGER, U., KAUFERLE, J., and POVH, D.: 'The Cahora-Bassa
D.C. transmission system: overvoltage protection and insulation co-ordination'.
IEEE PES winter meeting, paper T74 050-1, New York, 1974
3 BOWLES, J.P.: 'Overvoltages in h.v.d.c. transmission systems caused by transformer magnetising inrush currents', IEEE Trans., 1974, PAS-93, pp.487-93
4 GIESNER, D.B., and ARRILLAGA, J.: 'Behaviour of HVDC links under unbalanced AC fault conditions', Proc. IEE, 1972, 119, (2), pp.209-15
5 RAYNHAM, E.F., and GOOSEN, P.V.: 'Anollo inverter station h.v.d.c. operating experience'. Presented to CIGRE committee 14, Rio de Janeiro, 1981
6 UHLMANN, E., and FLISBERG, G.: 'H.v.d.c. insulation co-ordination - Part 2:
Distribution of overvoltages', Direct Curr. Power Electron., 1971, 2, (3), pp. 104-11
7 CLERICI, A.: 'Transient overvoltages caused by earth fault on bipolar dx.
lines'. IEE Conf. Publ. 107 on 'High voltage D.C. and/or A.C. power transmission',

London, 1973, pp. 196-200


8 BREUER, G.D., CSUROS, L., HUGUM, R.W., KAUFERLE, J., POVH, D., and
SCHEI, A.: 'H.v.d.c. surge diverters and their application for overvoltage
protection on h.v.d.c. schemes'. CIGRE conference 1972, paper 33-14
9 ARRILLAGA, J., and EL-BATAL, S.: 'Lightning-surge distribution in h.v.d.c.
convertors', Proc. IEE, 1973, 120, (5), pp.595-600
10 ARRILLAGA, J., and EL-BATAL, S.: 'Internal oscillations in multibridge
h.v.d.c. convenors', Proc. IEE, 1972, 119, (9), pp.1351-59
11 ARRILLAGA, J., and EL-BATAL, S.: 'Arc-quenching transients in HVDC
convenors', Proc. IEE, 1973, 120, (11), pp.1397-1402
12 DISEKO, NX., WOODHOUSE, M.L., THANAWALA, H.L., ANDERSEN,
B.R., CRAWSHAW, A.M., and ROWE, J.E.: 'Application of a digital computer
program to transient analysis and design of h.v.d.c. and ax. thyristor valves'.
IEE Conf. Publ. 205 on 'Thyristor and variable static equipment for A.C. and
D.C. transmission', London, 1981, pp. 167-70
13 KREGGE, J.S., and SAKSHANG, E.G.: 'Zinc oxide arrester experience and
application at h.v.d.c. stations'. IEEE conference on Overvoltages and compensation on integrated A.C.-D.C. systems, Winnipeg, 1980, pp.65-69
14 BUI-VAN, Q., BEAULIEU, G., and ROSENQUIST, R.: 'Overvoltage studies
for the St Lawrence River 500 kV DC cable crossing'. IEEE/PES winter
meeting, paper 91WM 121-4, PWRD, 1991
15 BAHRMAN, M.P.: 'Overvoltage and VAR compensation on the Square Butte
h.v.d.c. system'. IEEE conference on Overvoltages and compensation on integrated
A.C.-D.C. systems, Winnipeg, 1980
16 ANDERSEN, B.R., DISEKO, N.L., and ROBINSON, A.A.: 'Insulation co-ordination for the U.K. terminal of the 2000 MW h.v.d.c. cross-channel scheme'.
IEE Conf. Publ. 205 on Thyristor and variable static equipment for A.C. and
D.C. transmission, London, 1981, pp. 199-203
17 DEMAREST, D.M., and STAIRS, CM.: 'Solid state valve test procedures and
field experience correlation'. Paris, 1978 CIGRE paper 14-12
18 GAMPENRIEDER, R., et at: 'Design goals, specification, studies and commissioning of the 600 MW HVDC back-to-back station Elzenricht'. Paris, 1994
CIGRE paper 14-105
19 POVH, D., and LUONI, G.: 'Impact of overvoltages on design of HVDC
cables'. Paris, 1992 CIGRE paper 14-104

Chapter 10

DC versus AC transmission

10.1 General considerations


High-voltage transmission serves a dual purpose, i.e. system interconnection and bulk-energy transfer.
With reference to system interconnection, the need to operate the whole
system in perfect synchronism often prevents the transfer of power by
alternating current:

the jeconomic power ratings of such interconnections are often small in


relation to the installed capacity of the systems to be interconnected; in
such cases an AC tie line may not be able to cope with the power flow
and stability control problems. An alternative DC interconnection
provides a fast and flexible power flow control, regardless of the conditions in the AC systems, and can provide stability improvement for the
two interconnected systems;
AC interconnections always result in a reduction of the overall system
impedance and hence in increases of the short-circuit levels; these may
exceed the capability of the existing circuit breakers or cause unacceptable electrical and mechanical stresses on the system equipment;
if the systems to be interconnected have different frequencies
(normally 50 and 60 Hz), an AC tie line is not possible;
even with network systems of the same nominal frequency but
controlled according to different principles, an AC interconnection is
often impractical.

As far as bulk-energy transfer is concerned, there are various alternatives,


not all of them involving electric-power transmission, and an economic
assessment is essential in each case. Whenever the transmission distance is
sufficiently large, and restricting the choice to the electrical alternatives,
the case of HVDC transmission (shown schematically in Figure 10.1) is

254

High voltage direct current transmission


0 . . . 2000 km

(a)

Figure 10.1

(a)

(b)

(c)

(b)

Typical DC transmission system (ASEA Journal)


a = AC system
b = converter station
c = DC line
F = filter

well established in spite of the relatively high cost of the dual conversion
required.
In less obvious decisions, the accounting procedures used in the
economic comparison must include the cost of lines, terminals, any special
apparatus needed for voltage support (see Figure 10.2), short-circuit
limitation, etc. The energy lost and the plant needed to supply it must also
be capitalised.
200 . . . 250 km

2 0 0 . . . 250 km

200 . . . 250 km
SC

-#-

rr

-X-

(c)

Figure 10.2

(b)

(c)

(b)

svs
(a)

Typical long-distance AC transmission system (ASEA Journal)


a = AC system
b = AC transmission line
c = substation
SVS = static VAR source
SC = series capacitors

However, the basic costs alone are not decisive and allowance must be
made for other considerations such as:

DC versus AC transmission 255

operational reliability, flexibility and performance during disturbances;


the consequences of shutdowns owing to maintenance and forced
outages;
maximum loading capability as well as the continuous and short-time
overload;
transmission-system development and the possibility of a staged installation programme.

Moreover, to try and achieve a meaningful and generally applicable


comparison in marginal cases is a very difficult task. Among the factors
responsible for the complexity of a generalised theory are:

the wide range of practical situations involving differing conditions


among countries, e.g. overhead line costs vary from country to country
by a factor as high as 2.5, and the cost of the converter terminals varies
very little;
the lack of technical comparability, given the rather different degrees
of freedom of AC and AC-DC power systems;
the need to consider the long-term effects on overall system design and
cost when choosing among alternative plans for system development;
the rapid strides being made in the technology of both AC and DC
transmission; in this respect, transmission-line costs have experienced a
large increase in recent years. As the line cost is relatively lower in the
case of HVDC transmission this effect has affected the AC alternative
more; however, the appearance of FACTS devices is now having a
similar beneficial effect on AC-transmission costs.

Thus, a precise economic determination is only possible in terms of a


specific situation, taking into account the long-range development of the
system, its probable future pattern of load growth and generation
resources, and the many other factors affecting the planning of power
systems.
This Chapter provides a brief comparison of the AC and DC technologies with reference to bulk-energy transfer.
The escalating costs of bulk-energy transfer in the first half of the
century kept alive the memory of the initial supremacy of direct current as
a transmission channel and encouraged its revival. However, the fact that
transmission by DC requires a less expensive line or cable for the same
power capacity has to be weighed against the high cost of the AC to DC
and DC to AC power-conversion terminals.

10.2 Power-carrying capability of AC and DC lines 1


a If, for a given insulation length, the ratio of continuous-working withstand voltages is

256

High voltage direct current transmission


DC withstand voltage
ri

\ 1U. 1 /

"

(r.m.s.) AC withstand voltage


various experiments on outdoor DC overhead-line insulators have demonstrated that, owing to unfavourable effects, there is some precipitation of
pollution on one end of the insulators and a safe factor under such conditions is k = 1. However, if an overhead line is passing through a reasonably
clean area, k may be as high as y/2, corresponding to the peak value of
r.m.s. alternating voltage. For cables, however, k equals at least two and
here the prospect for DC is obviously very encouraging.
b As discussed in Chapter 9, a transmission line has to be insulated for
overvoltages expected during faults, switching operations, etc. AC transmission lines are normally insulated against overvoltages of more than
four times the normal r.m.s. voltage; this insulation requirement can be
met by insulation corresponding to an AC voltage of 2.5 to three times the
normal rated voltage
AC insulation level _ 9 ~
rated AC voltage (Ep)

(\c\9\

say, for AC.


On the other hand, with suitable converter control the corresponding
HVDC transmission ratio, i.e.
h

DC insulation level

(iQ 3)

rated DC voltage (Vd)


need only be 1.7.
Thus, for a DC pole-to-earth voltage Vd and AC phase-to-earth voltage
Ep, the following relations exist
..
.
insulation length required for each ACAphase
insulation ratio = :
:
:

insulation length required tor each DC pole


AC insulation level 1 / | DC insulation level
AC withstand level J' L DC withstand level
and substituting eqns. 10.1, 10.2 and 10.3
insulation ratio =

. v

(10.4)

c Consider a new DC transmission system to compare with a three-phase


AC system transmitting the same power and having the same percentage
losses and the same size of conductor. The DC system is considered to
have two conductors at plus and minus Vd to earth
power in the AC system:

SEpI^ (assuming that cos <f> = 1)

DC versus AC transmission 257


power in the DC system:
AC losses:
DC losses:

2IdVd
3/ 2 L #
2I2dR

Equating line losses


3/2Li?=2/V?

(10.5)

Id=(yfi/y/2)lL

(10.6)

or

Equating powers
3EpI^ = 2IdVd

(10.7)

Vd=(j3/j2)Ep

(10.8)

or

and substituting eqn. 10.8 in eqn. 10.4


insulation ratio = (kkx/h(yj2/ft)

(10.9)

For the values of k, k\ and k recommended above, the above ratio is equal
to 1.2 for overhead lines and 2.4 for cables.

DC transmission capacity of an existing three-phase double-circuit AC line


The AC line can be converted into three DC circuits, each having two
conductors at + Vd to earth, respectively; thus
power transmitted by AC:

Pa = 6EpI^

(10.10)

power transmitted by DC:

Pd = 6 VdId

(10.11)

On the basis of equal current and insulation


/L = /<*

(10.12)

Vd-{kkx/k^)Ep

(10.13)

derived from eqn. 10.4 equated to 1. The power ratio is therefore


= =(**i)/*2
Pa

(10.14)

Ep

and since the actual losses are the same, the percentage power-loss ratio
will be the inverse of eqn. 10.14. Thus, for the same values of k, k\ and k^
as in a above, the power transmitted by overhead lines can be increased to
147 %, with the percentage line losses reduced to 68 %; corresponding
figures for cables are 294 % and 34 %, respectively.

258

High voltage direct current transmission

10.3 A comparison of AC and DC transmission


characteristics
Overhead transmission: With reference to Figure 10.3, the power transfer in
a transmission line is determined approximately by the expression
^SR

where V represents r.m.s. fundamental-frequency voltage, X the series


reactance, 9 the relative voltage phase shift (or load angle) and suffixes S,
R indicate the sending and receiving ends, respectively.
Power-flow disturbances are quickly reflected in load-angle oscillations
and, therefore, for reasons of stability, the load angle is kept at relatively
low values under normal operating conditions (about 30). For a line
loaded with the natural (or surge) impedance this angle puts a limit to the
maximum series reactance (0.5 per unit) and therefore to the transmission
distance. Beyond this point, to avoid instability owing to faults on the line,
it is normal practice to provide parallel lines. With very high voltage longdistance transmission (Figure 10.2), it is also usual to include switching
stations with series capacitors and other means of voltage support such as
synchronous condensers or static compensators.
Moreover, the AC-line conductor has to be designed to transmit the
charging current (or capacitive reactive power) under light-load conditions.
This effect causes overvoltages and requires shunt-reactor compensation.
Since under steady-state conditions, inductance and capacitance have no
effect on a DC line, the above difficulties do not arise with HVDC transmission and there is no need for intermediate switching stations.
On the other hand, the operation of a converter, whether as a rectifier
or as an inverter, involves a consumption of reactive power. Since the reactive power need is substantial, it is usually supplied as near as possible to

Load

Figure 103

Power flow in an AC system

Load

DC versus AC transmission 259

the converter station, partly by the capacitance of the AC filters and often
by additional shunt capacitors. With weak AC systems, the AC-voltage
regulation with varying load conditions may demand the use of synchronous condensers or static compensators.
With DC no stability problems occur because the AC systems are
decoupled and the power flow can be freely and rapidly adjusted by
converter control. For a given conductor size the power-carrying capability
of the DC line in principle is only limited by thermal considerations and,
therefore, should the extra current be justified (in terms of ohmic loss),
the DC conductors can carry substantially higher power levels.
On the other hand, overloading is more restricted in DC transmission.
The silicon-controlled rectifier has a very small thermal capacity and thus
the modern converter valve is only designed to handle temporary overcurrents under fault conditions or to damp AC-system oscillations. If longterm overload capability is desired, this can be achieved by appropriate
overrating of the thyristors and permitting higher temperatures at the
converter plant.
Cable transmission: High-voltage transmission by cable is rarely used because
of the higher cost and longer repair times; it is normally restricted to
underwater crossings and infeed to urban centres.
The high-voltage cables have a low series inductance and large shunt
capacitance. Moreover, their loading, owing to the lower surge impedance
and thermal limitations, is usually below 0.3 times the surge-impedance
level. Therefore, high charging reactive powers are required, which
considerably limit the length of AC-cable transmission. For instance, at
50 Hz the charging current varies typically from 5.5 A/km for a 132 kV
cable to about 15 A/km for a 380 kV cable. With a 4.52 cm2, 380 kV cable
of 600 A thermal limit, the charging current for a 40 km length equals the
thermal limit and no useful load can therefore be carried. Similarly, a
2.58 cm2, 450 A, 132 kV cable has a critical length of about 80 km.
These critical lengths may be extended by inserting shunt reactors. Even
with a one hundred per cent compensation by means of two reactors, one
at each end, the power transmission capacity is only 86.6 % at critical
length and reduces to zero at twice the critical length. With two intermediate reactors, each providing 100 % compensation, dividing the line in
three equal parts, the critical length will increase to three times. Moreover,
intermediate compensation is impractical in the case of underwater links.

10.4 Other considerations


Equivalent reliability criterion
Before any final cost comparison can be made it is essential to perform
reliability studies of the AC and DC transmission alternatives to ensure
that they are reasonably equivalent in this respect.

260

High voltage direct current transmission

The reliability of a two-pole DC transmission line is often compared


with that of a double three-phase AC line. If this criterion is accepted, the
DC alternative offers a very economical transmission corridor, particularly
in terms of the right-of-way requirement. An illustration of this effect is
shown in Figure 10.12, which compares the tower structure and land to
transport 2000 MW. T h e AC alternative requires a 100 m corridor for
maintenance and repairs, whereas the DC solution only needs half of that.
However, the validity of such an assumption is questionable if a tower
fault, rather than an insulator fault, needs to be considered. A realistic
comparison should include both the probability of occurrence of the two
types of fault and their associated power transmission loss.
A reliability study carried out to assess the transfer of some 4760 MW in
British Columbia 2 over a distance of about 1000 km used a minimum of
two parallel DC transmission lines (as in the AC case) as a basis for cost
comparisons. T h e compared alternatives were:

a two-line 765 kV series compensated scheme;


a two bipole 600 kV DC scheme.

Reliability evaluations were performed in two stages:


(a) A Monte Carlo simulation of the radial-transmission alternatives,
which provided figures of availability of different power-transfer
levels and frequency of departures from these levels. T h e studies used
information obtained from manufacturers of HVDC equipment,
which included their experience with plant availability, overload
capabilities and forced outage rates for each HVDC component.
(b) A loss of load probability evaluation for the whole British Columbia
Hydro system using the information from (a) as input data. An initial
comparison without major spares produced very unfavourable relative
availability for the DC solution. T h e addition of a spare smoothing
reactor and a spare converter transformer at each terminal provided
similar availability for the two alternatives which were used for the
final economic comparison.
Effect of losses and discount rates
The AC and DC transmission alternatives are costed using a cash-flow and
present-worth analysis; the present worth of the estimated losses should be
included.
In the British Columbia Hydro study discussed in the previous section2
the component costs received from the HVDC manufacturers were carefully analysed, to remove inconsistencies, and often averaged.
It is interesting to consider the size of the effect that varying the value of
losses has on the economic comparison. This is illustrated in Figure 10.4
for various alternatives in the British Columbia Hydro scheme. The DC

DC versus AC transmission 261

2000

CD

1500

1a
s

x
1000

500

1.0

2.0

3.0

4.0

5.0

Cost of losses ($ x 1O~ )

Figure 10A

Cost comparison with varying losses

alternative produced the lowest capital cost for low values of the losses.
However, the two-line 765 kV scheme became cheaper when losses were
evaluated at about $6.4 x 1(T3 per kWh. It should be pointed out that this
situation occurs because of the lower DC-voltage level used (i.e. 600 kV).
With a 765 kV DC line the relative cost of the DC alternative would
reduce as the value of losses increased.
Some allowance is made in the economic analysis for the difference
between cost of money and inflation. In the above example, such a difference was considered equivalent to a discount rate of three per cent. The
effect of varying the discount rate is illustrated in Figure 10.5.
Earth return capability
If the circumstances are right, earth may be used as one of the conductors,
making it possible to transmit power by one conductor only, with substantial saving in capital cost. For example the Gotland scheme in Sweden
consists of only one submarine cable at negative polarity, sea being used as
the other conductor. Use of earth, which is a substantially low-resistance
path, also results in a considerable reduction of transmission losses. No use
can ever be made of earth with AC systems, because of the associated
inductive effects, excepting when it is absolutely necessary to use the rails

262

High voltage direct current transmission


2000
\
\

i
"
CO

1500

X
\V

- 2 x 1159 kV

\v

1000

2 x 765k\T

\A

2 x yutr

-2 >' ( 600 i kV dc-

0)

500

3 x 765

10

Discount rate 1%)

Figure 10.5

Cost comparison with varying discount rate

in electrified railway and even in that case many precautions have to be


taken.
The use of earth as a DC conductor has difficulties as well, owing to
possible interference with communication and railway signalling circuits,
corrosion of pipes and cable sheaths etc., but in many cases these difficulties do not arise or can be overcome without too much expense.
When operating on earth return, especially via a land electrode, a DCvoltage gradient will be established. If large DC currents are involved, the
land electrode potential rise can be of the order of several hundred volts
and the associated voltage gradient may extend over 100 km of countryside. AC systems within this zone form parallel paths for the DC groundreturn current, between widely separated grounded transformer neutrals.
Neutral grounding resistors or capacitors may be necessary to limit the DC
current in the AC systems to low values (of the order of 5A or less) in
much the same way as is necessary to protect AC systems against interference from geomagnetically-induced currents.
In many cases only a temporary use of earth may be permissible so that
a system with two conductors at Vd is adopted. In such a case, if a fault
results in a loss of one conductor, fifty per cent of the rated power can still
be supplied through the other conductor and earth. Compared to this a

DC versus AC transmission 263

fault on any one conductor of a three-phase AC line results in a complete


shut down of transmission.
Favourable routes and right-of-way

Since the relative transmission cost of DC is lower than AC, particularly by


cable, and as there are no technical limitations with DC regarding the
length of transmission, it is possible to consider alternative and more
favourable routes, from the point of view of way leaves or security from
faults etc.
The scope of DC for transmission of power by cable into or near big
cities is enhanced owing to its much higher circuit-loading capacity,
compared to AC cables. Many city centres are now so congested that it is
physically impossible to find room for another cable duct or trench, and
the possibility exists of converting present AC-cable routes into DC-cable
routes (perhaps by using the existing cables) thereby greatly increasing the
circuit capacity; this possibility is discussed further in Section 10.8.
Phased system development

A DC terminal normally consists of several converters connected in series


on the DC side. The DC-line voltage can, if desired, be raised gradually by
the installation of an increased number of converters and the power capacity of the DC link will be increased accordingly. Both the line and the
terminal can in the beginning have a single-pole arrangement using earth
return. The capacity can be doubled later by adding another pole.

10.5 Infeeds at lower voltage levels


When planning bulk-power infeed into load centres the use of underground cable and DC transmission will often be favoured. In such cases
the choice of the AC voltage level at the point of DC-AC power inversion
is an important economic parameter.
Large urban areas are normally supplied from a primary high-voltage
AC system (say 500 kV) and one or more levels of secondary-distribution
systems (of typically 132 kV). As load increases, each of the distribution
systems will need reinforcing and eventually as they reach their specified
maximum fault level, new distribution supply points will be required (not
normally interconnected). To provide the expanding load, new points of
generation have to be developed connected to the primary system; when
the fully interconnected primary system reaches its maximum specified
fault level it has to be sectionalised. At this point in time a new primary
system, at a higher voltage, is normally introduced and to this the new
generation is connected.

264

High voltage direct current transmission

r-1 -i

network j
100 j

Dist

J I
i

L-

(a)

1
I D1
!
!

Jr"
|
i
|

Distribution
network
200

]
i
!
I

(b)

Figure 10.6 a General pattern of development with synchronous sets


b Pattern of development with asynchronous and synchronous sets
An alternative system development, which could postpone or eliminate
the introduction of a new primary system, consists of HVDC power injections from the new generation points to appropriate AC-system locations.
The contribution of HVDC converter stations to the AC network shortcircuit capacity is negligible as long as no rotating synchronous compensation is required; filter-circuit capacitors and capacitor banks have very
short time constants during discharge and do not really contribute to the
short-circuit power rating.
The resultant effect of such asynchronous-power injection on the overall
system is illustrated in Figure 10.6. If asynchronous-connected generation
were connected at point A, the corresponding amount of reinforcement
would be avoided in the primary-transmission network. If it were
connected at D\ and D2, reinforcement of both primary and secondary
networks would be retarded, resulting in greater gain.
Some studies3 have shown the economic advantage of HVDC power
injection at the lowest practicable network-system level of the metropolitan
load area.

10.6 Examples of the application of the break-even distance


A recent contribution4 on this topic indicates that giving comparative cost
numbers can become a believer's contest. The contribution includes three
cases of long-distance transmission.

DC versus AC transmission 265


400

$218/kW+
$157/kW+

100

1000

500

1500

2000

HVDC converter capacity, MW


Figure 10. 7a

Approximate costs of two converter stations against scheme capacity

_ 2000 1000S 500-

r' iooi 508

105100

200

300

400

500

600

700

breakeven distance, miles

Figure 10. 7b Break-even distance for HVDC compared to AC transmission

The first case is a 500 kV AC line for 2000 MW at an approximate cost


of $500 000 per mile (or $250 per MW-mile). The converter costs of the
alternative DC solution are about $100 per kW (see Figure 10.7a). Allowing $40 per kW for the AC substations, reactive compensation gives an
added HVDC substation cost of $60 per kW, or a total of $120 million to
be saved from the line cost to break even. If the HVDC line costs $150 per
MW-mile, thus saving $100 per MW-mile, it would take 600 miles of 2000
MW HVDC transmission to break even with AC.
Two further examples are given to illustrate that, as the power transmission is decreased, so is the break-even distance. One is a 138 kV 200 MW

266

High voltage direct current transmission

link that results in a break-even distance of 360 miles. The other case is
particularly interesting because of the very low power involved, i.e. only
10 MW. With 35 kV AC, plus reactive compensation, the scheme may cost
some $40 000 per mile (i.e. $4000 per MW-mile). The cost of the two DC
converters will be of the order of $4 million ($400 per kW for two converters). Assuming that the cost of AC substations and reactive compensation
is about $1 million, the additional DC-converter cost for break even is $3
million. If the 30 to 50 kV bipolar DC line costs $25 000 per mile, then
this line saves $15,000 per mile which represents a break-ev^n distance of
200 miles.
A representative curve of break-even distances as a function of power
can be expected to be within the shaded region of Figure \0.1b.
The subject of light-power HVDC transmission is considered further in
Chapter 11 with reference to self-commutated schemes.

10.7 Environmental effects5


The environmental effects of overhead transmission lines have been
causing more concern in recent years, the principal effects being: audible
noise, electric field, radio and television interference and visual impact.

10.7.1 Electric field6


The two main causes of electric fields are the potential difference between
the overhead conductor and earth and the space-charge clouds produced
by conductor corona. The highest electric-field strength occurs directly

E
>

bipolar

15

10

-60

-30

0
distance, m

Figure 10.8

Electric field ofmonopolar and bipolar 450 kV overhead lines

DC versus AC transmission 267


- potential

earth electrode

Figure 10.9

Step voltage in the vicinity of an earth electrode

under the overhead conductor, both with monopolar and bipolar transmissions, and is approximately 21 kV/m for a 450 kV overhead-power transmission line (see Figure 10.8). Moreover, the electric field may be
strengthened further by external factors such as the weather, seasonal
variations and relative humidity.
The electric-field problem is less severe in DC because of the lack of
steady-state displacement current; thus HVDC lines require much less
right-of-way (ROW) than the horizontal AC configuration and less height
than the AC delta configuration of HVAC lines of comparable rating.
Land electrodes create potential differences on the earth's surface,
termed step voltages, which can cause shock currents (Figure 10.9). Assuming a typical human-body resistance of 1000 Q, a limit value of 5 mA is
recommended as the maximum safe current that can flow through the
body. If a higher electric-field strength is anticipated, then there will be a
proportionately higher shock current in the vicinity of the electrode. In
that case the area is deemed to be hazardous and may have to be fenced
off.

10.7,2 Radiated interference


The commutation processes between the thyristor valves and the operation-related harmonics cause disturbances in the kilohertz and megahertz
regions of the radio-frequency spectrum, as shown in Figure 10.10.
High-frequency oscillations following valve switchings propagate to the
overhead lines through the converter transformers. The most effective

268

High voltage direct current transmission


1

10

100

1k

10k

100k

1M

AC-harm
i
50

100M

1G

PLC
i
2.5

telephone
interference

Figure 10.10

10M

100

2.5

400

30

Rl

400
corona
400

400

Various interference bands

method of controlling radio-interference radiation is the use of electromagnetic shielding of the valve hall. The converter station is built with
reinforced concrete thus providing an inherent shielding, and the shielding effectiveness can be improved by additional wire meshes or liner
panels on the walls and ceiling of the hall.
Figure 10.11 shows the shielding effectiveness against frequency for a
typical valve hall with an additional wire mesh on the inner walls of the
hall.8 These levels are of little concern to radio or television interference.
Corona on the surface of the high-voltage overhead-power transmission
line is the principal source of radiated noise. The conductor corona
process depends on the magnitude of the electric-field strength, the
diameter of the line, its surface characteristics and weather conditions.

60

1/

40

^ 1

1.00

10.00

100.00

Frequency (MHz)

Figure 10.11

Typical shielding effectiveness of a valve-hall shielding

DC versus AC transmission 269

The radio-interference level of an HVDC overhead-power transmission


line is lower than that of a corresponding AC line. At a distance of
approximately 30 m from a 450 kV HVDC line, the interference level of a
0.5 MHz frequency is approximately 40 dB(//V/m). The corresponding
figure for a 380 kV AC line is around 50 dB(/iV/m).
The 30-400 kHz frequency band resulting from valve commutations
and corona discharges also produces power-line carrier (PLC) frequency
interference along high-voltage overhead-power transmission lines.

10.7.3 Acoustic noise


Acoustic noise from industrial plant, although relatively minor when
compared with road and rail traffic, often provokes the main complaints.
Subjective perceptions of acoustic-noise nuisance are dependent on the
amplitude, frequency and duration of the noise. The permitted acousticnoise emission limiting values for industrial plant depend on local conditions (for example, whether the area concerned is industrial or domestic
and local authorities' interpretation of statutory regulations) but are generally between 35 and 40 dB(A).
HVDC transmission systems contain a large number of subassemblies
and components which cause noise, the most powerful sources being the
transformers. Transformer acoustic noise is primarily dependent on the
flux density within the iron core, rather than the load. In standard transformer types, the core noises which occur during no-load operation are
ten to 20 dB(A) higher than those at rated load current. With converter
transformers, on the other hand, the sum of all the load noises is approximately ten dB(A) higher than the no-load noises9 and the frequency
content of their emitted noise is spread almost evenly across a band from
300 to 3000 Hz. Other sources of acoustic noise are cooling towers, chillers, filters, smoothing reactors and filter capacitors.
Reactors radiate noise as a function of AC harmonics and their mechanical stiffness and impedance. The frequency content peaks in the range 500
to 1300 Hz.
Filter capacitors have a complex sound-emission pattern owing to the
interaction from multiple individual point sources.
Measurements taken at the Haywards terminal of the New Zealand
link10 have indicated that the principal contribution to the measured noise
at the boundary is coming from the HP 24 filter capacitor in the 600 1300 Hz frequency range, with some contribution from the smoothing
reactor at 600 Hz.
There are three main ways to reduce or control noise emission from
equipment to achieve a specified level at a boundary. These are:

specification and purchase of quiet equipment;


enclosure of equipment to attenuate noise emission;

270

High voltage direct current transmission

separation of noisy equipment either by distance or shielding from


other objects in the direct or reflected-noise path.

Generally, the most effective strategy is to reduce the noise at source by


using inherently low-noise components. However, this is not always the
most cost effective approach as the features needed to reduce noise can
compromise the operation of the plant.
Noise-damping covers and walls, or even complete housings, can be used
to reduce the sound level of a typical HVDC station by ten dB(A) at a
distance of some 350 m. However, further acoustic noise reductions, especially at shorter distances, will have a significant influence on the overall
plant cost.

10.7.4 Visual impact and space requirements


In contrast with three-phase transmission systems, it is possible to reduce
the space requirements and visual effect of DC lines considerably.
Figure 10.12 shows typical mast designs and route widths (rights-of-way)
for alternative AC and DC transmission systems with identical transmission
power (2000 MW).6

10.8 Existing 11AC transmission facilities converted for use


with DC
Normally AC transmission lines are not loaded to their maximum thermal
rating and the firm-power capability is lower in any multicircuit AC
arrangement than it could be if the same facilities were used to transmit
power by DC. If any AC link is used for DC, the conductors can form
poles of a DC system which may be operated independently, if necessary,
up to the thermal rating.
The insulation, chosen to satisfy AC-system requirements, could generally sustain a DC voltage to earth equalling the peak value of the AC
voltage to earth, or exceeding it if heavy atmospheric pollution is not a
limiting factor. The possibility of using earth return under outage conditions offers an added attraction.
Similar arrangements could be made for both land and submarine-type
cables. This principle has already been applied in the case of DC transmission to Vancouver Island.
The conversion capability is not restricted to overhead lines and cables.
Capacitor units for series or shunt compensation, if suitably specified initially, could be reused at converter stations to constitute harmonic filters.
Existing AC switchgear could also be re-applied at converter stations to
provide isolation facilities and, in the event of schemes with earth return,
AC switchgear could be used for high-speed changeover duty.

DC versus AC transmission 271

500 kV DC
route width: 50m

800 kV AC
85m

2 x 500 kV AC
100m

Figure 10.12

Typical tower structures and rights-of-way for alternative transmission systems (2000 MW)

Local generation
and compensation
800 MW, high
security load

Local generation
and compensation
1100 MW high
security load

ttt

(*

ttt

(0

(b)

(a)

Local generation
and compensation
2200 MW high
security load

t
T-

HI
M l

4^

220 kV ac

M MM

I 220 kV
330 kV ac

330 kV ac

f- 330 kV

220 kV ac

210 km
T

tif

98 km

2 x 350 m m 2
ACSR conds

123 km
330 kV ac

r>-330 kVac

- r 330 kV
Hydro generation
Hydro generation

Figure 10.13

Power-transmission reinforcements
a Existing 330 kV AC system
b One AC circuit converted for DC
c Two AC circuits converted for DC

Hydro generation

DC versus AC transmission 273

An additional benefit will be a reduction in the fault level.


The major problems expected in the implementation of this method of
increasing power transmission capability are:

withdrawal from service during the period of changeover;


the need, in many schemes, to provide loads at intermediate points.

Some practical examples illustrating conversion possibilities are shown in


Figure 10.13.
This method is a practical alternative to the addition of further parallel
AC circuits or the introduction of a higher AC voltage. It thereby provides
an economical means of overcoming potential amenity problems which, as
explained in the previous section, may eventually emerge as the limiting
factor of public acceptance of very high-voltage overhead transmission.

10.9 Very long-distance transmission 12 ' 13


The use of ultra high-voltage direct current (Tjjjyrjc) j e voitages above
the present highest in use, 600 kV, is an attractive proposition for bulk
energy transfer in the range 3000-5000 MW and distances above 1000 km.
The use of 800 kV is considered achievable in the near future,14 but at the
top of the range considered (5000 MW and 4000 km) the optimum transmission voltage is around 1000-1100 kV.13
In recent years various proposals have been made for very long transmission systems involving large amounts of power; among them are the
possible transmission 30000-60000 MW scheme from Inga Hydro Plant
on the Zaire River up to Europe (5000-6000 km),15 the so-called world
interconnection from the USA to Europe through Alaska-Siberia16 and
the use in south-east Brazil of the large Amazon hydro power.l There are
also several possible schemes in China ranging from 6000 to 15 000 MW.
An empirical assessment of the DC-terminal cost against voltage is
shown in Figure 10.14, based on ABB's experience. Up to + 600 kV, the
actual costs have been taken into account; for voltages above + 600 kV a
cost increase above the natural trend has been considered, to take care of
the development cost necessary in a very limited market as foreseen now.
Moreover, a possible optimistic assumption of a large market for UHV DC
projects would permit the use of the natural trend of terminal costs for
voltages above + 600 kV and would enhance the adoption of the highest
voltages.
For each of the line lengths considered (1000-2000-3000-4000 km) and
for each one of the five values of power (1000-2000-3000-4000-5000
MW), Figure 10.15 defines the optimum voltage which results in
minimum cost.
However, there are many issues to be resolved before very long-distance
bulk-power interconnections become a practical reality. These have

274

High voltage direct current transmission


190-j

180170-

realistic trend in a limited


market for UHV

160g 150| 140130120-

"natural" trend in a
conveniently large market

110100
400

600

800

1000

kV

Figure 10.14 DC terminal cost against voltage

recently been highlighted by CIGRE SC-14 working group 14.70,18 with


reference to the concepts of dependence and interdependence.
Dependence is often perceived as a major obstacle to gaining the full
benefits of interconnection, because of a reluctance to give up some autonomy. This applies particularly to energy import, as the energy-deficient
country usually has self-sufficiency concerns. Interconnection of power
systems involves some technical and supply dependencies, and a degree of
financial dependence. Lack of close co-ordination among the participating
systems will precipitate the risk of large-system failure.

1200

1200
kV

|LOSSE:S2000$/kW

1000

5000 MV
4000 MV f

/
_ _

_1

800

400

,_'

1000

2000

3000

'

1000 MV

4000

km

I
(LOSSES * 1000 $/kW

1000

5000 MV^

800

2000 MV

_'
600

3000 MW

kV

___^~*

600

400

4000 MVf
3000MVf
2000 MVr
1000MVf

1000

2000

3000

4000

km

Figure 10.15 Optimum voltage against line length with different capitalised cost
of losses

DC versus AC transmission 275


transmission of electricity
length

HVDC
OHTL

AC
power
RE

transmission of gas
length
GAS
field

gas duct

SE

Figure 10.16

RE

Alternative energy-transmission systems

The supply options to countries over which a transmission-line interconnection passes but which are not directly connected must also be
addressed. This becomes a technical issue, especially if the load required
to be fed is very small compared to the main interconnection. Some possible solutions to this problem are discussed in Chapter 11.
The main competition to HVDC for the power ratings and distances
considered appears to be from bulk gas transmission, with the electricity
generation needed for consumption in locations far from the gas field.
The two alternatives are shown in Figure 10.16.
Considering that the costs of DC terminals and of combined-cycle plants
are not affected by the transmission length, the transmission of electricity
is more attractive than gas transmission for:

longer distances;
higher cost areas for transmission lines and gas ducts;
lower prices for gas at the gas field;
smaller powers.

For a distance of 5000 km and power in the range from 1000 to 5000 MW,
the cost of kWh with gas transmission is from 1.9 times to 1.2 times higher
than for electricity transmission. The range covers differences in the gas
price (i.e. two or five c$/m3) and high or low transmission costs (as applying in Europe and Africa, respectively).

276

High voltage direct current transmission


Table 10.1 Costs (in millions of US$)

Transmission
Power plant
Total
Gas Gm3/year

975
1819
2793

2292
1500
3792

5.7

4.8

For 3000 km and 3000 MW and low price for both overhead transmission and gas ducts, the cost of the two alternatives (G-gas and ^-electricity)
are shown in Table 10.1.
In any case, up to powers around 5000-8000 MW, electricity transmission is the preferred option.

10.10 References
1 ADAMSON, C, and HINGORANI, N.G.: 'High voltage direct current power
transmission' (Garraway Ltd, London, 1960), Chap. 1
2 HARDY, J.E., TURNER, F.P.P., and ZIMMERMAN, L.A.: 'A.c. or d.c, one
utility's approach'. IEE Conf. Publ. 205 on Thyristor and variable equipment
for A.C. and D.C. transmission' London 1981, pp.241-46
3 EHMKE, B., and HARDERS, C.F.: 'Planning aspects of h.v.d.c. power transmission into metropolitan load centres'. Symposium sponsored by the Division of
Electric Energy Systems USDOE, Phoenix, Arizona, 1980, pp.63-75
4 HINGORANI, N.G.: 'Dc technology for rural transmission'. CIGRE international colloquium on HVDC and FACTS, Johannesburg 1997, paper 3.1
5 LAFOREST, J.J., LINDH, C.B., and STAMBACH, M.R.: Techniques for
determining overhead line cost data for comparison of a.c. and d.c. transmission alternatives'. Symposium sponsored by the Division of Electric Energy
Systems USDOE, Phoenix, Arizona, 1980, pp. 143-61
6 SCHMIDT, G., FIEGL, B., and KOLBECK, S.: 'HVDC transmission and the
environment'. Power Eng. /., October 1996, pp.204-10
7 EPRI Research Project 1467-1: 'HVDC ground electrode design'. Electric Power
Research Institute, August 1981
8 JACKEL, B.W.: 'Investigations on radio interference and power line carrier
interference of a back-to-back converter'. IEE Conf. Publ. 423 on 'A.C. and D.C.
power transmission', London 1996, pp.58-63
9 REIPLINGER, E.: 'Lasta bhangige transformatorengerausche', Sonderdruck aus
Elektrotechnischer Zeitschrift (etz), 3, 1989
10 COAD, J.N.O.: 'Audible noise design and testing for the DC hybrid link
project'. International Colloquium on HVDC and FACTS, Wellington, 1993,
paper 2.2-1
11 JONES, K.M., and KENNEDY, M.W.: 'Existing AC transmission facilities
converted for use with dc'. IEE Conf Publ. 107 on 'High voltage DC and/or AC
power transmission', London 1973, pp.253-60
12 ASPLUND, G., et al: 'A novel approach on UHVDC 800 kV station and equipment design'. International Colloquium on HVDC and FACTS, Wellington,
1993, paper 7-3
13 CLERICI, A., LONGHI, A., and TELLINI, B.: 'Long distance transmission: the

DC versus AC transmission 277


DC challenge'. IEE Conf. Publ. 423 on AC and DC power transmission, London,
April 1996, pp.86-92
14 KRISHNAYYA, P.C.S., et at. 'An evaluation of the R. & D. requirements for
developing HVDC converter stations for voltages above + 600 kV\ CIGRE,
Paris, 1988, paper 14-07
15 PARIS, L.: 'Remote renewable energy resources made possible by high voltage
interconnection: the Grand Inga case'. IEEE PES WM New York - IEEE Power
Eng. Rev., June 1992
16 MEISEN, P.: 'Worldwide interconnections may be an idea whose time has
come', Trans. Distrib. Int., December 1992
17 PRACA, J.C.G., et at. 'Amazon transmission challenge - comparison of technologies'. CIGRE, Paris, 1992, paper 14/37/38-01
18 BAKER, M.H., HEPBURN, A., and LEWIS, W.P: 'HVDC economic assessment
in a Southern African context'. International colloquium on HVDC and FACTS,
Johannesburg, 1997, paper 6.1

Chapter 11

New concepts in HVDC converters


and systems

11.1 Introduction
HVDC technology took a big step forward around 20 years ago when thyristor valves succeeded the mercury-arc valves previously used. The converter-station concept introduced at that time, however, has remained
practically unchanged since then, even though great improvements in
equipment and subsystems have taken place. At the same time there have
been substantial advances in conventional AC technology and, particularly,
in the application of power electronics to make power transmission more
flexible and economical. Such competition is now exciting a continuous
stream of new HVDC concepts and techniques with the aim of improving
performance and reducing costs and delivery times. The main advances
are discussed in this Chapter.

11.2 Advanced devices


11.2A Thyristor development1
Improvements in thyristor ratings (voltage and current) and characteristics
have a high influence on the costs of the valves and other equipment in a
converter station. Work is under way to raise the voltage capability of the
present thyristor wafer from 8 kV to 10 - 12 kV. Larger wafer diameters
are also under development, although the timing of their introduction will
depend on the demand for them in HVDC and other markets such as flexible AC transmission systems and industrial applications.
The light-triggered thyristor has also been developed, demonstrated and
even used commercially in one HVDC project in the United States and
three in Japan. The objective here is to eliminate the electronic circuits

New concepts in HVDC converters and systems 279

presently used at each thyristor level for converting the turn-on light
signals into electrical pulses. Direct light triggering of the devices using
optical fibres requires powerful light sources at ground level; Japanese
manufacturers have developed light-emitting diodes for that purpose.
However, the elimination of the protecting circuit at each thyristor level
requires higher overvoltage margins for the light-triggered thyristors; a
thyristor self turn on, when voltage or rate of rise of voltage is too high, is
now under development in Europe.

7 7.2.2 Gate turn-off semiconductors (GTO)


GTOs are making a significant impact in power-electronics design; the
turn-off feature leads to such new circuit concepts as self-commutated,
pulse-width modulated (PWM), soft-switching, voltage-driven and multistep converters, and to circuits which operate at higher internal switching
frequencies (several hundreds of hertz). These, in turn, reduce the harmonic content and allow operation at unity and even leading power factors.
Attention is now turning towards their application to HVDC transmission. In this respect, GTOs are attractive for DC-power conversion into AC
systems which have little or no voltage support. In such cases, the synchronous condensers or static VAR systems used in some present schemes will
not be needed. Before GTOs can be regularly used for DC transmission,
series connection of the semiconductors must be confidently developed
and losses should be reduced. If we accept the inevitability of GTO
converters for DC transmission, the system designer is then faced with
determining a control strategy which must control DC power, AC voltage
and AC-system frequency. At present, however, GTO ratings are much
lower than those of thyristors, and their cost and losses are almost twice
those of thyristors; these three aspects (ratings, cost and losses) have a large
influence on the cost of the other equipment in a complete converter
station.
In Japan several power companies and the Central Research Institute of
the Electric Power Industry, are funding a research and development
programme for the application of GTOs to the reinforcement of powersystem interconnections (see Section 11.5.1).

11.3 New concepts for thyristor converters2


In the face of competition from the new semiconductor devices, the thyristor-based converter technology is not remaining at a standstill. Under the
banner HVDC 2000, ABB has recently proposed a new generation of
HVDC converter stations, incorporating the latest developments; these
affect a number of technical areas and are aimed at improved performance, design simplicity and reduced construction time.

280

High voltage direct current transmission

control and service budding


valve-cooling

Figure 11.1

HVDC2000

The key features of HVDC 2000, shown in Figure 11.1, include:

a capacitor-commutated converter (CCC);


actively-tuned AC filters;
air-insulated outdoor thyristor valves;
active DC filters.

A concise description of these new components is given next (except for


active filters, which have been described in Chapter 3).

11.3.1 Capacitor-commutated converter3'4


A capacitor-commutated converter (CCC) is a conventional HVDC converter provided with commutation capacitors between the converter transformer and the valves, as shown in Figure 11.2. The basic function of this
concept is that the capacitors contribute to the valve commutation voltage.
This contribution makes it possible to operate the CCC with much lower
reactive-power consumption compared to a conventional converter. The
rating of the filter bank is, therefore, considerably reduced and the filters

Figure 11.2

Basic CCC single-line diagram

New concepts in HVDC converters and systems 281


Table 11.1 Cost comparison of voltage-control devices
Economic evaluation
(% of station capital cost)
Voltage-control device

Synchronous condenser (1 unit)


Synchronous condenser (2 units)
Thyristor-controlled reactor
Thyristor-switched capacitor
Metal-oxide varistor
Series capacitor

capital
cost

capitalised
operational
cost

total
lifetime
cost

21
40
17
18
14
13

11
15
6
5

32
55
23
23
14
14

negligible
1

can be connected throughout the operating range, from minimum to


maximum load, while still fulfilling the reactive-power requirements.
Consequently, the switching of filter banks, used in a conventional converter, can be eliminated.
A recent study carried out by the Manitoba HVDC Research Centre has
indicated that series-capacitor compensation is one of the most economical
methods of AC-voltage control for HVDC converter stations.5 Table 11.1
summarises their findings with reference to an 810 MW inverter station.
Further, CCC gives a more robust and stable dynamic performance of
the inverter station, especially when inverters are connected to weak AC
systems and/or long DC cables. With weak AC systems the AC load-rejection voltages are also reduced owing to the low reactive-power consumption.
Although, in principle, the capacitors could be connected on the AC side
of the converter transformers, it is not possible to completely avoid ferroresonance problems with such a configuration. The location of the capacitors between the converter transformers and the valve bridge results in full
control of the capacitor currents and complete elimination of the risk of
ferroresonance.
A long DC cable has large capacitive energy storage. In the event of a
temporary AC-voltage reduction in the inverter AC system, for example
caused by a remote single-phase-to-ground fault, the DC cable will partially
discharge into the AC system. Thus the transient current increase at the
inverter end will not be immediately detected at the rectifier end, resulting
in a delay before the rectifier reduces the direct current. Therefore, ACvoltage collapse may occur unless the inverter is able to counteract the
current increase by raising the terminal voltage. However, this is not possible as the conventional inverter operating on minimum commutation
margin has a negative-impedance characteristic. On the other hand, a CCC

282

High voltage direct current transmission

inverter operated at minimum commutation margin has an almost


constant or slightly positive impedance, which improves the response to a
transient-current increase compared to a conventional converter.
If required, increased commutation margins can be achieved, without
increasing the reactive-power consumption of the converter station, by
reducing the capacitance of the commutating capacitors in order to
increase their contribution to the commutation voltage.

113.2 Continuously-tuned AC filters6


Detuning of conventional filters is caused by network frequency excursions and component variations, e.g. capacitance changes due to temperature differences.
A filter in which tuning can be adjusted to follow frequency variations
and component variations can be designed with a high Q factor to provide
a low impedance for the harmonics. Automatic tuning will also ensure that
all risks of resonances and current-amplification phenomena are eliminated, implying that the ratings of the AC-filter components can be
reduced.
In the proposed filter, tuning is achieved by variation of the inductance
by means of a transverse DC magnetic field. This field is perpendicular to
the main flux direction and does not affect the linearity of the magnetising process. A cross section of the reactor is shown in Figure 11.3. The
insulating tube

coil winding for


/ filter current

VJ

tube-shaped core
Figure 11.3

direct current control


winding

Cross section of the self-tunedfilterreactor

New concepts in HVDC converters and systems 283

control winding, placed at ground potential, is insulated from the main


winding by an epoxy-filament cylinder.
With the continuously-tuned filter, separation of filtering and reactivepower compensation is possible, since filters with low reactive-power
generation can be built.
In combination with the capacitor commutated-converter concept
(described in Section 11.3.1), where only limited reactive-power compensation by means of shunt filters is needed, the continuously-tuned filter is an
attractive solution. With automatically-tuned filters for the 11th and 13th
harmonics and conventional high-pass filters, only between 0.12 and 0.15
p.u. total reactive power is needed.
With such a small AC filter, the harmonic resonance, formed by the ACsystem inductance in parallel with the shunt-filter capacitance, will be
sufficiently high, i.e. above the third harmonic, and as a consequence the
use of low-order filters will not be required.

11.33 Outdoor valves


Two of the early HVDC projects with thyristor valves used oil-insulated
outdoor valves. However, this technology was soon abandoned and all
other commercial HVDC transmissions have thyristor valves of indoor
design. Nevertheless, there are a number of disadvantages with the present
scheme. Among them are the large, costly valve buildings with complex
interfaces to the electrical equipment, risk of a valve-building fire, and risk
of flashovers across large wall bushings.
The following advantages have motivated revisiting the outdoor valve
design concept:

the civil content is greatly reduced;


the number of interfaces between different activities in a project is also
reduced;
wall bushings subject to high DC stresses are eliminated;
the thyristor valves can be delivered as completely-assembled and tested
units, like the other components.

A prototype of an outdoor valve has been long-term tested since mid-1992


at the Stenkullen converter station of the Konti-Skan link. The new valve
unit, shown in Figure 11.4, is an air-insulated live tank valve, using similar
thyristor and reactor modules to those in the indoor valve. Normally, each
unit represents one valve function, and therefore 12 units are needed for
the 12-pulse converter. The present ratings for the 12-pulse group are 275
kVand 1200 A.
For the active part of the thyristor valve standard thyristor and reactor
modules are used. The DC-pole side of the valve is connected directly to
the valve tank, whereas the AC-phase connection is made through a
bushing.

284

High voltage direct current transmission

thyristorand
reactor
modules

rs v

, , ,-V;-,r-r-^
,.>--....*.<

bushing

valve
housing

communication
channel

valve cooling and


cooling control

valve base
electronics

valve control and


monitoring

air-cooled
liquid coolers

valve control signals j ~ *

Figure 11.4 Basic elements of the outdoor valve

The valve tank comprises a supporting structure covered with steel


sheets and the tank is thermally insulated to any ambient climate conditions.
Outdoor valves eliminate the need for large diameter wall bushings. The
insulation between valve-tank potential and ground is achieved by verticalsupport insulators, and small wall bushings provide the insulation between
the valve structure and its own tank.
Since one end of the valve is electrically connected to the tank, the clearance between the tank and the live parts inside is determined solely by the
voltage across the valve.
The cooling and valve-control equipment are located at ground potential. The cooling water, fibre optics and ventilation air are all supplied to
the valve via the main vertical insulator. The enclosure of this insulator is
a glass-fibre pipe with silicone-rubber sheds on the outside to provide the
necessary creepage distance^
The 12 outdoor valve units will need a ground area somewhat larger
than that occupied by a valve building, but this will have little impact on
the total station-area requirement. More important may be the lower

New concepts in HVDC converters and systems 285

height of the outdoor valves compared to that of a traditional valve building-

11.4 Compact converter stations


Many prospective HVDC applications involve at least one terminal in the
vicinity of metropolitan areas and often the potential converter sites are
subject to air pollution. Such cases can benefit from compacting techniques.
An early report by EPRI7'8 described the development of gas-insulated
bus systems and compact valves of a so-called dead-tank design. An SF6
gas-insulated bus is used to interconnect the DC potheads, smoothing
reactors, valves and converter transformers. The use of a gas bus in combination with air-insulated valves and conventional valve buildings considerably reduces valve-building size and cost. It also avoids pollution deposits
on exposed portions of a DC yard, thus eliminating the regular silicongrease applications in current use.
The removal of heat, achieved by the use of liquid freon, is probably the

rv~

COOLERS

GAS INSULATED
3 4 5 kV AC BUS

GAS INSULATED
3 4 5 kV AC BUS

SMOOTHING
REACTOR
BASEMENT PARKING
AND STORAGE

Figure 11.5 Section of converter building

286

High voltage direct current transmission

only realistic approach for dead-tank designs; it may also be an attractive


alternative to water for valve-cooling systems, since liquid freon is a much
superior dielectric and leaks do not cause flashovers.
A conceptual design of a 500 kV 2400 MW HVDC compact station
has been recently carried out as part of Consolidated Edison's ongoing
programme on HVDC equipment for urban environments.9 A cross
section of the converter building with the main components involved is
shown in Figure 11.5. The project uses conventional suspended air-insulated, water-cooled thyristor valves. The single-phase transformers, with a
fourth winding for filter connection, use oil to water-glycol heat exchangers of a low audible-noise design; the same cooling system is used for the
smoothing reactors, located next to the building in a special sound-proof
enclosure. The AC filters' capacitors are air-insulated and the reactors use
nonflammable silicon-based oil. All these components, as well as switching
equipment for the associated gas-insulated 345 kV cable circuits, can be
installed in a 37 m high building on a 61 x 122 m site.

11.5 GTO-based voltage-source converters


The replacement of thyristors by GTOs permits self commutation of the
converter with the following advantages:

the commutation does not fail when system voltage is decreased or


distorted;
the low-order harmonics are greatly reduced and therefore the harmonic filters can be small;
no local reactive-power supply is required, even when it is used with
low short-circuit level power systems;
the active power supplied through the DC line and the reactive-power
output from each terminal are independently controlled;
the response is faster owing to the increased switching frequency pulsewidth modulation (PWM) process. ,

series filter

rectifier

voltage source GTO inverter

Figure 11.6 Voltage-source inverter (Manitoba HVDC Centre Bulletin)

New concepts in HVDC converters and systems 287

Therefore, instead of using the present thyristor configuration (a currentsource converter), the preferred option seems to be the voltage-source
converter. A GTO voltage-sourced inverter (VSI) can deliver power to a
weak AC system and even to a system totally devoid of power generation,
as shown in Figure 11.6.

11.5.1 A GTO back-to-back HVDC link10


The power utilities and Central Research Institute of Electric Power Industry (CRIEPI) in Japan are jointly developing a 300 MW back-to-back
(BTB) self-commutated interconnector.
The BTB system-circuit configuration is shown in Figure 11.7 and its
main specifications are listed in Table 11.2. Each terminal consists of four
self-commutated converters. The windings on the AC-system side of the
converter transformers are connected in series. On the DC side, the four
converters are connected in parallel and a shunt capacitor is placed
between the two terminals of the link. The converter uses a pulse width
modulation (PWM) scheme at a switching frequency of 450 Hz for the 50
Hz system and 540 Hz for the 60 Hz system. The increased pulse number
(nine) reduces the harmonic content and improves the steady-state and
dynamic performance of the converter. However, the losses also increase
with the switching frequency.
Active and reactive powers are controlled independently of each other.
A voltage-margin method similar to the current-margin control of present
schemes is used as the basis of link DC-power transfer control. The simplified BTB diagram of Figure 11.8 shows the components involved in the

system A

system B *

Figure 11.7 System configuration of a 300 MW back-to-back interconnector

288

High voltage direct current transmission

Table 11.2 Fundamental specifications of 300 MW self-commutated converter

system

converter

transformer

Item

Specification

rating capacity

300MW-100MVar
(316 MVA)
AC 275 kV
total < 1 %
each order < 0.5 %

rating voltage
voltage distortion
converter type
power device
insulation
cooling
control
configuration

voltage-source type
GTO
air insulated
pure water cooling
nine-pulse.PWM control
four-stage (or 8 stage)

configuration

direct step up

voltage-margin method (the reactive-power control system is omitted in


this diagram).
Figure 11.9 shows the operating point for the case when the DC voltage
is controlled by terminal A and the direction of active power flow is determined by the lower-limit value of the DC voltage (i.e. that of terminal B).
The power-flow level is then controlled by adjusting the lower limit of the
DC-voltage control at terminal B.
Changing the terminal having the voltage margin will cause the power
flow to be reversed.
The reactive-power control system can be switched between a constant
reactive-power control mode and a system-voltage control mode to maintain the AC-system voltage at a specified level.

power flow
terminal A

terminal B

system A

system B

AEdp
PWMh

Figure 11.8 BTB system

|DC-AVR|(J-Edp

New concepts in HVDC converters and systems 289


DC voltage

operating
' point

. DC volt,
ref. B
voltage
margin

I
j
terminal B
lower limit

terminal A
lower limit
power flow

Figure 11.9

Active power against DC-voltage characteristics (terminal B has the


voltage margin)

11.5.2 HVDC light


Voltage-source converter technology can extend the economical power
range of HVDC transmission down to a few megawatts and even feed
totally passive industrial or distribution systems with no other sources of
supply. The DC feeder then controls the frequency and voltage levels.
A pilot 3 MW, + 10 kV, ten km long scheme has recently been
completed by ABB (from Hellsjon to Grangesberg) in Sweden. This
project uses series-connected insulated-gate bipolar-transistor (IGBT)
converters without transformers at either end.
A compact ten MW converter station is likely to occupy less than 200 m2
and be fully automatic under remote supervision. Possible applications of
the HVDC light concept include the supply of power to islands, infeed to
cities, remote small-scale generation, off-shore generation and small
tappings from large point-to-point HVDC schemes.11

11.6 DC cable developments


Current-dependent voltage control (CDVC) has been applied in a few
limited cases, to exploit the increased dielectric strength of mass impregnated paper insulation which becomes available when load heating reduces
the viscosity of the impregnating oil and the associated internal pressure
rise makes conditions within the cable insulation similar to those in an oilfilled cable. The applied voltage may then be increased and decreased
over the higher voltage range while the current is maintained at rated
value to gain additional load capacity.
Further opportunities may arise in new schemes to exploit this technique
and to develop it further.

290

High voltage direct current transmission

XLPE insulation has so far been developed for use in commercial installations up to 150 kV DC (Gotland Island DC link) and research and development work for use at higher voltages is continuing.
PPLP (polypropylene laminated-paper insulation) has been under investigation for some years, primarily for use in very high-voltage AC oil-filled
cables where it has particular advantages arising from the ability of the
insulating material to withstand higher stress levels. Reduced insulation
thickness and lower dielectric losses result. AC cables are being produced
using PPLP insulation for voltages up to 800 kV. The relatively hightemperature coefficient of expansion, and the swelling characteristics of
PPLP when impregnated with oil compounds, currently pose difficulties
for use in DC cables. Research is continuing, but so far no commercial
project has been installed.

11.7 Direct connection of generators to HVDC converters


The arrangement traditionally used for power generation connected to
HVDC transmission is shown in Figure 11.10. In that configuration the
generating units feed, via the generator transformer, a common AC
busbar to which the AC harmonic filters are connected.
A new concept termed unit connection (or direct connection)12 has been
investigated by a joint working group of CIGRE SC11 and SC1413 study
committees. The modified configuration, illustrated in Figure 11.11, elimTo AC network
Pole 1
12 Pulse Convenor
Electrode

Pole 2

i
Figure 11.10

Conventional arrangement in which the rectifier station is built


independently of the power station; AC filters and two transformation levels are required

New concepts in HVDC converters and systems 291


Polel
12 Pulse Convenor

Electrode

Pole 2

Figure 11.11

Unit-connected HVDC power station, in which generators and


converters are integrated into single units; series-parallel and/or
parallel combination of units at the DC side is also allowed

inates the generator transformer and the AC busbar. The generators are
directly connected to the converter transformers and the harmonic
currents produced by the 12-pulse unit-connected scheme are absorbed by
the generator so that the need for AC filters is eliminated. Moreover,
voltage control can be exercised entirely by the generator excitation and
transformer on-load tap changers are no longer needed.
The direct-connected scheme is considered an attractive proposition for
electrical generation from remote sources of power, such as hydro and
low-grade coal fields, when a new development supplies little or no local
load. Its potential for variable-speed operation can be used to optimise the
efficiency of hydro sets under different load conditions and under varying
water heads. This property can also be useful in pump-storage and windpower applications. Two interesting variations of the basic direct connection of Figure 11.11 are the group connection and the uncontrolled rectifier.1 In the group connection, illustrated in Figure 11.12, the matching
of generator and converter ratings is no longer required, the flexibility of
operation and of maintenance is enhanced, and although variable-speed
operation is still allowed, the group of synchronised generators will
require some form of joint speed control.
Direct connection with uncontrolled (diode) rectifiers leads to the
highest degree of simplification of the sending end, and brings in additional savings on firing-control gear and on the valves themselves, as
diodes are less expensive and have reduced losses. However, the acceptability of diode rectification is subject to its ability to clear and recover
from DC faults and receiving-end commutation failures without greatly
increased delays.

292

High voltage direct current transmission


Polel
12 Pulse Convenor

Electrode

Pole 2

Figure 11,12 Group-connected HVDC power station, in which machines and


converters can be combined in groups via a transfer bus

Possible limitations of the unit connection scheme are:

converter harmonic currents flow through the generator stator


winding, causing additional losses and ripple torques;
outage criteria may require some additional equipment, if generatorconverter ratings are individually matched; note, however, that this
does not apply to the group connection;
the practice of shutting down some generator sets at partial load in
large hydro stations may be limited to a reduced number of steps.

Some approximate cost benefits suggested by the working group are given
in Table 11.3 and recent tests carried out at the Benmore end of the New
Zealand link15 have shown the capability of standard generators to cope
with the harmonic content in the absence of filters.

11.8 Small HVDC tappings


The need for and possibility of building multiterminal HVDC transmission have been discussed at least since the early 1960s. Studies have been
made regarding the feasibilities of series16 and parallel connections17 of
the converter stations on the DC side. The general consensus is that a
parallel connection, or constant DC-voltage multiterminal HVDC scheme,
is to be preferred when the power ratings of the different terminals are of
the same order of magnitude. Two such schemes have already been built,
i.e. Sardinia-Corsica-Italy and Quebec-New England.
There is also a need to tap small amounts of power at some locations
along the route of some long-distance overhead-transmission systems.

Table 11.3

Capital cost breakdown of sending-end arrangements assuming conventionalfixed-speed50 or 60 Hz generation; figures


from average typical cost, in per cent
Conventional
station,
Fig. 11.10

Generator
Generator transformers
Generator breakers
Generator's transfer bus
AC switchgear, capacitors
and AC filter banks
Converter transformers and
smoothing reactors
Converter valves
Valve control and auxiliaries
Valve cooling
DC filters, electrodes and electrode lines
DC-line breakers
Subtotal
HVDC station site development, land,
civil work, labour, engineering and
supervision
Total

8.0 %
2.0 %

Unit,
Fig. 11.11
thyristors

Group,
Fig. 11.12
thyristors

Unit,
Fig. 11.11
diodes

1.0 %

1.0 %

1.0 %

3.0 %
1.5 %

2.0 %

20.0 %
15.0 %

14.0 %

14.0

14.0 %

15.0 %
6.0 %
2.0 %
4.0 %
72.0 %

15.0 %
5.0 %
2.0 %
4.0 %

15.0 %
5.0 %
2.0 %
4.0 %

41.0 %

45.5 %

12.0 %
1.0 %
2.0%
4.0 %
1.5 %
37.5 %

28.0 %

20.0 %

22.0 %

18.0 %

100.0%

61.0 %

67.5 %

55.5 %

Comments

4 % gen. overprice assumed


main single item saved
higher stresses in group
major overall saving in
unit-connected schemes
OLTC dispensed with in
unit-connected schemes
20 % savings with diodes
no valve controls with diodes
independent of arrangement
independent of arrangement
for line faults and CFs
reduced requirements of land,
civil works and possible valve
hall-powerhouse integration

294

High voltage direct current transmission


station B

I"

tapping station
converter 1

Figure 11.13

converter_2__

/<#

'd3

converter 3

DC transmission line with a series tapping

However, the conventional technique, with line-commutated currentsource converters in parallel on the DC side, is too expensive for this application, as the converter has to be designed for the full-line voltage. Moreover, the local AC network is usually very weak in such cases, thus
requiring the installation of expensive synchronous compensators.
The connection of small tapping stations in series on the DC side offers
certain advantages in such cases, with regard to both costs and performance. Also, the availability of new semiconductor devices with turn-off
capability will make it possible to use choppers or forced-commutating
converters.
A series-connected tapping station with forced-commutated converters is
illustrated in Figure 11.13. 8 The power is supplied to the local AC
network through a voltage-source forced-commutated converter (shown as
a simple six-pulse converter). In a real installation it is foreseen that a
higher pulse number will have to be used to reduce the harmonic content.
Electromagnetic transient simulations have shown that disturbances in
the series tap have little effect on the main converters. However, the cost
for such stations is still expected to be fairly high.

11,9 References
1 HINGORANI, N.G.: 'High voltage DC transmission: a power electronics workhorse', IEEE Spectr., April 1996, pp.63-72

New concepts in HVDC converters and systems 295


2 CARLSSON, L., et al: 'New concepts in HVDC converter station design'.
CIGRE, Paris, 1996, paper 14-102
3 REEVE, J., BARON, J.A., and HANLEY, G.A.: 'A technical assessment of artificial commutation of hvdc converters with series capacitors', IEEE Trans.,
1968, PAS-87, (10), pp. 1830-40
4 JONSSON, T., and BJORKLUND, P.E.: 'Capacitor commutated converters for
HVDC. IEEE/KTH Stockholm conference on Power tech., paper SPT PE 023-0366, Sweden, 1995
5 WOODFORD, D.A., and ZHENG, F.: 'Series compensation of DC links'.
CIGRE symposium on Power electronics in electric power systems paper 420-04,

6
7
8
9
10
11
12
13

Tokyo, 1995
HOLMGREN, T., et al: 'A test installation of a self-tuned AC filter in the
Konti-Skan 2 HVDC link'. IEEE/KTH Stockholm conference on Power tech.,
paper SPT PE 02-06-0367, Sweden, 1995
ASPLUND, G., et al: 'Outdoor thyristor valve for HVDC IEEE/KTH Stockholm conference on Power Tech., paper SPT PE 02-05-0365, Sweden 1995
FLAIRTY, C, HINGORANI, N.G., LEBOW, M., and WILSSON, S.: 'EPRI's
compact, gas-insulated DC converter stations'. USDOE symposium on Incorporating HVDC power transmission into system planning, Phoenix, 1980, pp.507-18
LEBOW, M., et al: 'A study of a + 500 kV, 2400 MW compact converter
station'. IEE Conf. Publ. 345, London, 1991, pp. 165-70
HORIUCHI, S., et al: 'Control system for high performance self-commutated
power converter'. CIGRE, Paris, 1996, paper 14-304
ASPLUND, G., ERIKSSON, K., and SVENSSON, K.: 'DC transmission based
on voltage source converters'. CIGRE SC-14 colloquium, South Africa, 1997,
paper 5.7
KRISHNAYYA, P.C.S.: 'Block and double-block connections of HVDC power
stations infeed'. IEEE EHV/UHV conference, Vancouver, Canada, 1973
CIGRE JWG 11/14-09: 'Direct connection of generators to hvdc converters:
main characteristics and comparative advantages', Electra, 1993, (149), pp.18-

14 BOWLES, J.P.: 'Multiterminal HVDC transmission systems incorporating


diode rectifier stations'. IEEE Trans., 1981, PAS-100, (4)
15 MCDONALD, S.J., ENRIGHT, W., ARRILLAGA, J., and O'BRIEN, M.T.:
'Harmonic measurements from a group connected generator for HVDC
converter scheme'. IEEE Trans. Power Deliv., 1995, 10, (4), pp. 1937-43
16 REEVE, J., and ARRILLAGA, J.: 'Series connection of converter stations in an
h.v.d.c. transmission'. Direct Current, 1965, 10, pp.72-8
17 LAMM, U., UHLMANN, E., and DANFORS, P.: 'Some aspects of tapping of
h.v.d.c. transmission systems'. Direct Currents, 1963, 8, pp. 124-29
18 EKSTROM, A., and LAMELL, P.: 'HVDC tapping station: power tapping
from a DC transmission line to a local AC network'. IEE Conf. Publ. 345,
London, 1991, pp. 126-30

INDEX

Index Terms

Links

A
AC-DC system interaction

70

123

140

143

177

193

254

195

269

129

AC system faults, see Short circuit faults


AC switchyard
AC transmission
Acoustic noise
Active filters
Anode

80
3

13

160

161

Arcback (see also Backfire)

11

Asynchronous interconnection

89

97

264

Anode reactor

Average direct voltage

255

93

26

B
Backfire
consequential
Bipolar link
Blocking of valves

200

207

221
84

182

209

224
Break-even distance

221

264

Bridge converter

13

Bridge inverter

27

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

210

Index Terms

Links

Bridge rectifier

13

Bridges in parallel

18

24

Bridges in series

174

228

Bypass switch

161

208

Bypass valve

160

208

Bypassing of bridges

208

C
Cables

184

198

259

13

160

289
Capacitor commutated converter
Cathode

280
3

CEA control, see Constant extinction


angle control
Charging current

259

Commutating voltage

18

101

Commutation (see also Overlap angle)

13

18

101

110

117

148

155

201

224

18

20

34

115

134

Commutation transients

229

248

Compact converter station

285

Complementary resonance

148

Composite resonance

148

154

Computer modelling

79

210

267
Commutation angle, see Overlap angle
Commutation failure
Commutation reactance

241

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

238

Index Terms

Links

Constant current control

104

111

113

112

113

149
Constant current transmission

111

Constant extinction angle control

103
148

Control
characteristics

112

117

equidistant

103

106

filters

103

148

87

120

individual phase-control

101

145

148

instability of

145
87

119

120

103

107

39

42

47

84

154

181

192

214

249

86

88

160

238

21

22

constant current, see Constant current


control
constant extinction angle, see Constant
extinction angle control

frequency

power
predictive

145

tap changer, see On-load tap-changing


Converter
back to back

287
bridge circuit
compact
mercury-arc
multibridge

13
285

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

186

Index Terms

Links

Converter (Cont.)
phase-shift

19

35

57

42

50

160
pulse number

33
56

rating factor

132

134

six-pulse

25

34

twelve-pulse

21

38

42

50

108

143

Cross-modulation

39

42

145

Current extinction

232

235

Current margin

114

127

Current setting

112

D
DC line, fault

216

DC link
bipolar
construction in stages

85

122

194

11

26

79

176

181

197

235

286

161

197

257

265

28

45

263

monopolar, see Monopolar link


DC reactor

DC switches
DC transmission, advantages

270
Deblocking

229

Deionisation of fault arc

219

Delay angle (see also Gate control)

14
103

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

Index Terms

Links

Direct voltage, see Average direct


voltage
Displacement factor

30

Distortion of alternating voltages, see


Voltage distortion
Dynamic compensation

133

135

E
Earth return

261

Economics
AC versus DC

254

Electric field

266

Electrode, earth

190

EMTDC

143

Environmental effects

266

270

Equivalent disturbing current

77

Extinction angle

28

39

155

204

118

F
FACTS

Filter
admittance of

71

cost of

73

damped

60

73

237
second-order

74

third-order

74

type-C

74

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

76

Index Terms

Links

Filter (Cont.)
DC

77

double-tuned

71

196

high-pass, see Filter, damped


impedance of

64

protection of

226

quality of

58

71

74
133

77
reactive power of

58

73

self-tuned

72

282

sharpness of tuning

58

single-tuned

70

size

58

tuned

70

74

Filter design

57

69

Firethrough

201

Firing angle, see Delay angle


Forced commutation
Fourier analysis
Frequency control

294
27

33

120

Frequency conversion

87

93

Frequency deviation

71

138

G
Gate control

14

Gate turn-off, see GTO


Generators feeding DC link
unit connection

290
290

Grading electrode

Grid control

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

76

Index Terms
GTO

Links
7

279

H
Harmonic distortion

210

Harmonic filters, see Filter


Harmonic impedance

58

59

148

Harmonic instability

129

144

150

Harmonics
AC characteristic

33
33

43

at no overlap

34

37

at overlap

39

direct-voltage

34

even

144

149

magnification of

110

144

noncharacteristic

39

43

46

80

144

62

70

145

resonance

232

147
sequence components

45

63

triplen

75

144

I
IGBT

Instability, see Harmonic instability and


Power instability
Insulated gate bipolar transistor, see
IGBT
Insulation co-ordination

135

164

243
This page has been reformatted by Knovel to provide easier navigation.

229

Index Terms

Links

Interaction, see AC-DC system


interaction
Interharmonics

42

Inversion

27

48

L
Light guides

90

169

225

233

135

231

171
Lightning surges

216
238

Load rejection

133
281

M
Margin angle, see Extinction angle
Master power controller

122

Maximum available power (MAP)

32

Maximum power curve (MPC)

32

Mercury-arc valve

123

160

Microcomputers

124

169

Misfire

201

Monopolar link

84

182

263
MOSFET
Multigroup converters
Multiterminal DC
series connection

7
229
84

91

85

294

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

238

190

Index Terms

Links

N
Network harmonic impedance
effect on filter design

62

loci in complex plane

62

Noise

72

179

184

30

111

112

141

163
42

O
On-load tap-changing

180
Oscillations

50

Overhead lines

182

Overlap angle

28

34

Overvoltages

147

229

internal

189

232

P
Per unit quantities

134

Phase-locked oscillator

103

Physical models

211

Power capability of overhead lines

255

Power control

119

Power instability

129

107

Power factor (see also Displacement


factor)

28

142

Power/frequency control

87

121

Power losses

75

257

Power modulation

137

141

Power reversal

116

143

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

Index Terms

Links

Protection
differential

224

filter

226

overcurrent

221

valve group

224

Pulse duration modulation

43

Pulse number

33

Pulse position modulation

43

224

42

Q
Q, see Filter, quality of and Sharpness
of tuning

R
Radio interference
Reactive power
Rectifier (see also Converter)

176

268

11

28

114

133

141

281

24

Re-energisation of DC line

219

Reliability

121

255

259

Resonance

50

62

68

136

147

231

Reversal of power, see Power reversal


Ripple

12

RTDS

211

S
Sea return, see Earth return
Self-saturated reactor

136

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

Index Terms

Links

Short-circuit faults

200

207

211

214
unsymmetrical

111

135

214

Short-circuit ratio

129

136

145

124

259

Static VAR system

244

259

Subsynchronous resonance

140

Surge arrester

165

175

Surge capacitor

161

233

Surge diverter

160

Switching surges

198

197
Silicon controlled rectifier, see Thyristoi
Smoothing reactor, see DC reactor
Stability, of AC transmission

244

235

237

20

63

133

197

244

259

Tap changer control

30

115

Telecommunications

126

245
Synchronous compensator

Telephone influence factor

58

70

Telephone interference

57

77

190
Tests

171

Thyristor

180

converter

162

light triggered

278

module

167

station layout

174

240
172

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

184

Index Terms

Links

Thyristor (Cont.)
valve

143

162

176

192

283
cooling

167
286

quadruple
tests

89
171

Thyristor-controlled reactor

136

Thyristor-switched capacitors

136

Transformer connections
for increased pulse number
Transformer, converter
leakage
saturation effects

167

36

57

36
174

178

196

150

182

18
135
231

tertiary winding

21

Transient overvoltages

190

231

Transient stability

139

142

143

228

233

238

49

54

63

117

156

Transients fast (see also Lightning and


Switching surges)

247
Triple harmonics

150

U
Unbalanced AC voltage

V
Valve, see Mercury-arc and Thyristor
VDCOL

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

Index Terms

Links

Voltage distortion

49

Voltage regulation

134

Voltage source converter

286

Voltage stability

135

58

W
Weak AC systems

135

216

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

69

You might also like