Anda di halaman 1dari 37

TECHNDLDCIES FDP INTECPATINC

WIN0 FAPhS TD THE CPI0


nterIm Feport
CDNTFACT NU|8EF: K/EL/000J4J
UFN NU|8EF: 06/801













The 0T drIves our ambItIon of
'prosperIty for all' by workIng to
create the best envIronment for
busIness success In the UK.
We help people and companIes
become more productIve by
promotIng enterprIse, InnovatIon
and creatIvIty.
We champIon UK busIness at home and
abroad. We Invest heavIly In
worldclass scIence and technology.
We protect the rIghts of workIng
people and consumers. And we
stand up for faIr and open markets
In the UK, Europe and the world.

i


TECHNOLOGIES FOR INTEGRATING WIND FARMS TO THE GRID
(Interim Report)

CohIracI humber: K/EL/000343
UPN humber: 06/801


Contractor

AREVA T&D Technology Centre (with University of Manchester)


Prepared by Prepared by Prepared by Prepared by

Mr SIamaIios Chohdrogiahhis, Dr Mike Barhes (UhiversiIy oI MahchesIer)
Dr M AIeh (APEVA T&D Techhology CehIre)








The work described in this report was carried out
under contract as part of the DTI Technology
Programme: New and Renewable Energy, which is
managed by Future Energy Solutions. The views and
judgements expressed in this report are those of the
contractor and do not necessarily reflect those of the
DTI or Future Energy Solutions.




First published 2006
APEVA

ii
Acknowledgements:

The authors would like to thank Dr Mark Osborne and Mr Antony Johnson of the National
Grid and Dr Phillip Cartwright of AREVA T&D, and Dr Olimpo Anaya-Lara and Professor
Nick Jenkins of the DTI Centre for Distributed Generation and Sustainable Electrical
Energy for their comments and suggestions.


iii
EXECUTIVE SUMMARY

This report discusses possible technologies for enabling the integration of wind farms into
the power supply network. The capabilities of each technology in complying with the
requirements set by the Grid Code are analysed. Subsequent reports associated with this
project will establish the effect of a STATCOM to enable a grid connection of wind farms to
the National Grid power supply network.

Background of the work

The Government has set targets for the UK to generate 10% of electrical energy from
renewable sources by 2010, 15% by 2015, with aspirations of 20% by 2020. By 2050 there
is a further aspirational goal to reduce emissions by 60% against a baseline year of 1990.
Drivers behind this policy include the UKs obligations within the United Nations
Framework Convention on Climate Change, the Kyoto Protocol, carbon reduction
measures set by the EU and the enhancement of security of energy supply within the UK.
The UK possesses a third of the EUs potential for offshore wind generation, yet in 2000
renewables supplied only 1.3% of the UKs electricity compared with 16.7% in Denmark
and 3.2% in Germany. In order to help facilitate the Government targets from renewable
sources, the UK Renewables Obligation and the Climate Change Levy together will
provide the renewables industry with 1billion/year of support. To meet the Governments
aspiration to reduce CO
2
emissions, a penetration of 25% to 35% renewables will be
required. As the total installed capacity of wind farms increases, the reliable and secure
operation of the electrical power transmission network needs careful consideration.
Hitherto, wind farms were considered essentially as negative loads which did not
contribute to voltage or frequency support. However, for wind energy to become a
substantial quantity within the British generation portfolio, the technical capabilities of wind
farms need to be enhanced. Grid Code compliance is a necessity and further research and
development is needed to fully ensure that wind farm designs meet the requirements
specified by the GB Grid Code.

The work performed

To illustrate the technical challenges of integrating large wind farms into the network, an
analysis of the GB Grid Code was carried out with particular attention paid to:
Voltage control
Fault ride through capability
Fault current contribution
Frequency control.

The main three wind generation technologies reviewed were:
Doubly Fed Induction Generators (DFIGs)
Fixed Speed Induction Generators (FSIGs)
Direct drive synchronous generators through back-to-back voltage source
converters.


iv
The four technical architectures for the connection of offshore wind farms to the power
supply network reviewed were:
AC connection
AC connection along with employment of Dynamic Reactive Compensation devices
Voltage Source (VS) HVDC link
Line Commutated (LC) HVDC link.

The GB Grid Code requirements apply to the whole wind farm as seen at the Point of
Connection (PoC) rather than at the turbine terminals. The philosophy behind this decision
avoids discrimination between different wind turbine technologies and promotes flexibility
for wind farm developers. Since the Grid Code requirements are based on the behaviour
of the windfarm at the PoC, a discussion of potentially Grid Code compliant configurations
has to include representative systems from all major generator technologies, and possible
combinations with available approaches for connecting offshore wind farms into the power
supply network. A qualitative comparison of the technical capabilities of these
combinations was conducted and is presented in tabular form.

Finally, this report provides an overview of the current state-of-the-art modelling
techniques used for various components of a wind farm. Generic models of a DFIG, a
FSIG generator and a STATCOM device are shown through simulation in the
PSCAD/EMTDC software modelling package.

Findings

For cases of wind farms connected through AC submarine cables, without a
supplementary reactive compensation device, the performance at the PoC depends mainly
on the employed wind turbine generator technology. If a dynamic reactive compensation
device is used, then this complementary technology has the potential to enhance the
behaviour of the wind farm, especially for fault ride-through, reactive power and voltage
control capability. In the case of a DC connection to shore, the performance of the wind
farm as seen at the PoC depends mostly on the inherent technical capabilities of the
employed DC technology.

Doubly-Fed Induction Generators (DFIGs) have been extensively researched and the
capability of DFIGs to comply with the requirements for power quality is well documented.
In cases where the electrical distance between the wind farm and the PoC is large, the use
of a static and/or dynamic reactive compensation device may be necessary in order that
the required reactive power capability of the farm is satisfied, as measured at the PoC.
System frequency control has been demonstrated in the Horns Rev wind farm by
appropriate blade pitch angle control of the individual wind turbines. Fault ride through
capability is starting to be demonstrated in the public domain through simulations and site
tests. However, there is a lack of information on the impact of a wind farm comprising
DFIG wind turbines on the existing network under transient and fault conditions.

A wind farm comprised of Fixed Speed Induction Generators (FSIGs) and connected to
the grid with an AC connection without dynamic VAr compensation, cannot meet the new
requirements of the Grid Code. Particular issues are voltage control and the supply of
reactive power to the network during faults, and here the use of a dynamic reactive
compensation device can significantly enhance the wind farms technical capabilities.
Frequency control can be provided by a wind farm with FSIG wind turbines when they

v
possess blade angle controllers. For transient stability, recent research shows that a
combination of improved machine design and supplementary controls during faults can aid
fault ride through capability but this must be coordinated with Grid Code requirements for
post-fault real power output recovery. Still, further work is needed for an assessment of the
impact of a STATCOMs rating and electrical point of connection in technical and
economical terms.

Theoretically a wind farm connected to the power supply network through a VS HVDC
scheme can meet all the requirements set in the GB Grid Code. Evaluation is necessary
on a site specific basis for the performance of the scheme during faults in the power
supply network, especially to establish whether the scheme can provide active power
support to the network during voltage sags at the PoC. The principal drawback of VS
HVDC to date is the relatively large power losses, and the limited DC voltage (150kV)
compared to conventional HVDC links (500kV) based on present technology. It is credible
to consider that VS HVDC will become more attractive in the future, as the technologies
behind self-commutated switches and multilevel inverters progress, increasing device
power ratings, decreasing losses and reducing capital costs.

Even though the use of Line Commutated HVDC links is a mature technology, its
application for connecting offshore wind farms to the power supply network has been
proposed only recently. One particular technical issue for wind turbines employing power
electronic interfaces, is that they may have limited capability to provide reactive power
support and voltage control to the offshore wind farms DC link rectifier terminal. If the wind
turbine generators are FSIGs then the issue is more prevalent since the turbines are
essentially passive in terms of voltage control. Configurations where a synchronous source
is used at the wind farm to provide the necessary e.m.f for operation of the rectifier
thyristor converters have been proposed. This type of configuration henceforth referred to
as Hybrid HVDC, can have a similar behaviour to a VS HVDC link even though additional
auxiliary equipment may be needed. However, research is needed especially in the areas
of ride through capability and power system frequency response.

vi
CONTENTS PAGE

EXECUTIVE SUMMARY (iii)

CONTENTS PAGE (vi)

1. BACKGROUND 1

2. GB GRID CODE CONNECTION REQUIREMENTS 2

2.1 Reactive Power and Voltage Control 2
2.2 Fault Ride-through Capability 3
2.3 Frequency Range of Operation and Power Frequency Characteristic 4
2.4 Frequency control 4
2.5 Power Quality Harmonics 4
2.6 Power Quality Flicker 4
2.7 Further Requirements for Wind Farms 5

3. WIND TURBINE GENERATION TECHNOLOGIES 6

3.1 Fixed Speed Induction Generators (FSIG) Wind Turbines 6
3.2 Doubly Fed Induction Generators (DFIG) Wind Turbines 6
3.3 Direct Drive Synchronous Generator Wind Turbine 6

4. MAIN TYPES OF CONNECTION TO THE TRANSMISSION NETWORK 8

4.1 Synchronous connections 8
4.1.1 AC Connection 8
4.1.2 AC Connection plus Dynamic Reactive Compensation 9
4.2 Asynchronous connection 10
4.2.1 Voltage Source HVDC 10
4.2.2 Line Commutated HVDC 11

5. WINDFARM TECHNICAL CAPABILITIES IN RESPECT TO GENERATOR
TECHNOLOGY AND TYPE OF CONNECTION
13

6. MODELLING ASPECTS 15

6.1 Offshore Wind Farm 15
6.2 Mechanical Dynamics of the Wind Turbine Generators 15
6.3 Power Electronic Converters 16

7. MODEL DEVELOPMENT AND EVALUATION 17

7.1 Wind farm with DFIGs and no STATCOM 21
7.2 Wind farm with FSIGs+STATCOM 22
7.3 Comparison between average and switched models 23
8. CONCLUDING REMARKS AND FURTHER WORK 24


vii

REFERENCES 25

GLOSSARY OF TERMS 29

1
1. BACKGROUND

Under present plans, without new-builds, only one nuclear energy plant will still be
operating in the UK by 2025. The 2003 Energy White Paper [1] does not rule out the
construction of new nuclear plant but points out that its current economics make it an
unattractive option for new, carbon-free generating capacity and there are also important
issues of nuclear waste disposal to be resolved. For coal, the White Paper concludes
Coal fired generation will also have an important part to play in widening the diversity of
the energy mix provided ways can be found to materially reduce its carbon emissions.
Although it also notes that: Domestic coal production is likely to continue to decline as
existing pits reach the ends of their geological and economic lives. This leaves
renewables, particularly wind, with an important part to play in the UK energy production
mix to ensure diversity and security of supply.

The DTI and Carbon Trust have recently commissioned an impact study on renewables [2]
to assess the impact on the electricity network of Government aspirations. Broadly the
findings were that 72% of the 2010 target can be met by 2006, but significant barriers need
to be addressed for further expansion of renewables, particularly planning uncertainty and
reinforcement of the transmission and distribution networks. The level of reinforcement
required to the transmission and distribution systems in Scotland and North-West England
requires expenditure in the order of at least 1.4bn for the transmission network and
780m for the distribution network [2] to ensure a secure and stable system. Furthermore
the transmission value quoted above is based on power flows and system security but
excludes transient stability and fault levels [3]. Intermittency is not thought to be a
significant issue affecting the development of renewables for present 2010 targets, but
balancing costs may increase substantially for the 2020 target.

Hitherto, wind farms were considered essentially as negative loads which did not
contribute to voltage or frequency support. However, when wind energy becomes a
substantial percentage within the British generation portfolio, the technical capabilities of
wind farms need to be enhanced. As a consequence of this, the three Transmission
Licensees in Great Britain (England and Wales National Grid, Scottish Power
Transmission and Scottish Hydro Electric Transmission Limited) revised the Grid Code
requirements. Grid Code compliance is a necessity and further testing and verification is
needed to fully ensure that wind farm designs satisfy the requirements for fault ride
through specified by the latest GB Grid Code. A summary of the Grid Code process is
given in [4].


2
2. SUMMARY OF THE GB GRID CODE CONNECTION
REQUIREMENTS

The requirements for new and renewable forms of generation including DC converters
were introduced into the Grid Code on the 1
st
June 2005. The key issues of the
Connection Conditions of the GB Grid Code [5] specifically concerning wind farms are
reviewed. The technical requirements are defined for the whole wind farm as seen at the
Point of Connection to the power supply network and not individually for each generator.

2.1 Reactive Power and Voltage Control

According to CC.6.3.2, CC.6.3.6, and CC.6.3.8 of the Grid Code, wind farms have to be
able to provide automatic voltage control at the Point of Connection (PoC) by continuous
changes to their reactive power output, according to a slope characteristic whose exact
specifications will be specified in a site specific bilateral agreement. The general reactive
power capability requirement of a wind farm is defined in CC.6.3.2 and depicted in figure 1.
This reactive power capability applies when all wind turbines are in service. If fewer
machines are operating then the required maximum amount of reactive power is
diminished proportionally with the number of machines running.


Point A is equivalent (in MVAr) to: 0.95 leading p.f. at Rated MW output
Point B is equivalent (in MVAr) to: 0.95 lagging p.f. at Rated MW output
Point C is equivalent (in MVAr) to: -5% of Rated MW output
Point D is equivalent (in MVAr) to: +5% of Rated MW output
Point E is equivalent (in MVAr) to: -12% of Rated MW output

Figure 1: Required reactive power capability of wind farms
(Figure 1 of Connection Conditions of the Grid Code [5])


3
2.2 Fault Ride-through Capability

The required fault behaviour of a wind farm can be summarised into four requirements:
For system faults that last up to 140 ms, the wind farm has to remain connected to
the network. For supergrid voltage dips of duration greater than 140 ms, the wind
farm has to remain connected to the system for any dip-duration on or above the
heavy black line of figure 2.
During system faults and voltage sags, a wind farm has to supply maximum
reactive current to the Grid System without exceeding the transient rating of the
plant.
For system faults that last up to 140 ms, upon the restoration of voltage to 90% of
nominal, a wind farm has to supply active power to at least 90% of its pre-fault
value within 0.5 sec. For voltage dips of duration greater than 140 ms, a wind farm
has to supply active power to at least 90% of its pre-fault value within 1 sec of
restoration of voltage to 90% of nominal.
During voltage dips lasting more than 140 ms, the active power output of a wind
farm has to be retained at least in proportion to the retained balanced supergrid
voltage.
It should be noted that in cases where less than 5% of the turbines are running, or under
very high wind speed conditions where more than 50% of the turbines have been shut
down, a wind farm is permitted to trip.



Figure 2: Required ride through capability of wind farms for supergrid voltage dips of
duration greater than 140 ms (Figure 5 of Connection Conditions of the Grid Code [5])

4
2.3 Frequency Range of Operation and Power Frequency Characteristic

According to CC.6.1.3 of the Grid Code wind farms have to be able to operate
continuously for system frequencies between 47.5 and 52 Hz and for at least 20 sec for
system frequencies between 47 and 47.5 Hz.

With the generating plant operating at full power output at frequencies between 47 Hz and
50.4 Hz, the minimum active power output as a function of system frequency deviations
from 50 Hz is described in CC.6.3.3 and is depicted in figure 2 of the Grid Code
Connection Conditions. For frequencies between 50.4 Hz and 52 Hz the power output of a
generator has to decrease at a minimum of 2% of output power for every 0.1 Hz rise of
system frequency above 50.4 Hz (Grid Code Reference: Balancing Code 3.7.2).

2.4 Frequency Control

According to CC.6.3.6 (a) and CC.6.3.7, wind farms are required to have the capability to
contribute to frequency control by adjusting their power output. Accordingly, CC.6.3.7
obliges wind farms to be fitted with a fast acting proportional frequency control device. This
is equivalent to a speed governor controller of a synchronous generator in a thermal plant.

When in Frequency Sensitive Mode, generators (both synchronous and non-
synchronous) have to provide primary, secondary and high frequency response, which are
defined in CC.A.3.1 and in figures CC.A.3.2 and CC.A.3.3 of the Grid Code. The minimum
level of frequency response that generators have to achieve is defined in figure CC.A.3.1.

2.5 Power Quality Harmonics

The harmonic content from a wind farm as a whole has to be limited in accordance with
the values set in Engineering Recommendation G5/4.

2.6 Power Quality Flicker

A specific form of harmonic distortion caused especially by wind farms with fixed speed
induction generators is flicker, i.e. voltage variations of frequency 8 to 10 Hz that can
cause light variations from incandescent lamps that are especially annoying to the human
eye and may even trigger epilepsy episodes [6]. Short Term and Long Term Flicker
Severities at the PoC have to be within the limits of Engineering Recommendation P28.

5
2.7 Further Requirements for Wind Farms

During phase to earth and phase-phase to earth faults, phase-to-earth voltages can rise
up to 140% in England & Wales and 150% in Scotland. This should be taken into account
when establishing the protection settings and voltage ratings of wind farm equipment.

6
3. WIND TURBINE GENERATION TECHNOLOGIES

3.1 Fixed Speed Induction Generators (FSIG) Wind Turbines

The FSIG wind turbine is a simple squirrel cage induction generator, which can be directly
coupled to the electricity supply network. The frequency of the network determines the
rotational speed of the stators magnetic field, while the generators rotor speed changes
as its electrical output changes. However, because of the well known steep torque-slip
characteristic of the induction machine, the operating range of the generator is very
limited. The wind turbine is therefore effectively fixed speed. FSIGs do not have the
capability of independent control of active and reactive power, which is their main
disadvantage. Their great advantage is their simple and robust construction, which leads
to lower capital cost. In contrast to other generator topologies, FSIGs offer no inherent
means of torque oscillation damping which places greater burden and cost on their
gearbox.

3.2 Doubly Fed Induction Generators (DFIG) Wind Turbines

The DFIG is a wound rotor induction generator whose rotor is fed via slip rings by a
frequency converter. The stator is directly coupled to the electrical power supply network.
As a result of the use of the frequency converter, the network frequency is decoupled from
the mechanical speed of the machine and variable speed operation is possible, permitting
maximum absorption of wind power. Since power ratings are a function of slip, DFIGs
operate over a range of speeds between about 0.75 and 1.25 pu of synchronous
frequency, which requires converter power ratings of approximately 25% [7]. A great
advantage of the DFIG wind turbine is that it has the capability to independently control
active and reactive power. Moreover, the mechanical stresses on a DFIG wind turbine are
reduced in comparison to a FSIG. Due to the decoupling between mechanical speed and
electrical frequency that results from DFIG operation, the rotor can act as an energy
storage system, absorbing torque pulsations caused by wind gusts [8]. Other advantages
of the DFIG include reduced flicker and acoustic noise in comparison to FSIGs. The main
disadvantages of DFIG wind turbines in comparison to FSIGs are their increased capital
cost and the need for periodic slip ring maintenance.

3.3 Direct Drive Synchronous Generator Wind Turbine

An alternative to the much-used induction machine generator is the use of a multipole
synchronous generator, fed through a power electronic AC/DC/AC stage. The excitation of
the synchronous generator can be given either by an electrical excitation system or by
permanent magnets. The AC/DC/AC converter acts as a frequency converter and
decouples the generator from the Grid. It consists of two back-to-back voltage source
converters, usually with IGBT switches, which can independently control the active power
transfer through the dc link and the reactive power output at each converter terminal [9].
The speed range is generally similar to that of DFIGs [10]. The multipole construction of
the synchronous generator leads to a low mechanical rotational speed of the generator
rotor and can permit direct coupling to the wind turbine. The possibility of reducing the
number of stages in the gearbox or eliminating it completely is often quoted as an
advantage of direct drive synchronous generator wind turbines. However set against this is
the greater VA rating of the power electronic converter compared with DFIGs and the
larger physical generator size.

7

As a result of the increased mechanical stresses experienced by FSIG wind turbines at
present there is a practical limit to the rating of commercial models of this technology. All
present commercial models for multi-MW wind turbines in the range above 3MW are either
DFIGs, or synchronous generators coupled to the network through back-to-back
converters.

8
4. CONNECTION OPTIONS TO THE TRANSMISSION NETWORK

The following section describes the technical issues associated with various types of
connection for offshore wind farms and existing transmission and distribution technology.

Connection options of an offshore windfarm to the Grid can be divided into two categories:
AC connection options
The wind turbines are connected to the power supply network through AC
submarine cables. The electrical field of the generators rotates synchronously with
the power supply network.
DC connection options
The wind turbines are connected to the power supply network through a DC
transmission scheme. The offshore wind farm is decoupled electrically from the
power supply network. The electrical frequency of the offshore wind farm is
determined by the control of the DC scheme. As a consequence the electrical field
of the generators may run with a frequency different to the frequency of the power
supply network.

4.1 AC connection options
4.1.1 AC Connection

For a simple AC connection, the behaviour of the wind farm is defined solely by the
inherent technical capabilities of the wind turbines. In this report the term AC connection
denotes this case.

Referring to GB standardised voltage levels, the two principal options for the AC
connection of large offshore wind farms are the use of multiple 33 kV links, which is the
cheapest option for distances of a few kilometres, or voltages of 132 kV. The latter option
enables a reduction in power losses [11, 12] but will necessitate an offshore platform to
accommodate the step-up transformer(s) and circuit breaker(s). The presently favoured
strategy for relatively large offshore wind farms [13] appears to be the use of 33kV cable to
interconnect the offshore wind farm arrays, feeding to a higher voltage central
transmission link for connection to the shore. The final choice of voltage level(s) for
connection however is not specified by the GB Grid Code and belongs to the wind farm
developer.

The major problem of connection by AC submarine cables is their capacitance. Cables
present a high shunt capacitance in comparison to overhead lines. The capacitive
charging currents increase the overall current of the cable and thus reduce the power
transfer capability of the cable (it is thermally limited). In addition transient voltage stresses
are present requiring suitable switchgear. As the length of the cable increases, so does the
capacitance of the cable and the charging currents. As a consequence, a single cable can
only transmit a certain amount of power for a given distance, after which more cables in
parallel are required. One way of addressing this issue would be to compensate the
capacitive losses along the length of the cable. This is difficult for submarine cables so the
next best solution is to compensate losses from both ends of the cable. The compensation
is usually achieved by fixed shunt inductors. However, even this form of compensation will
only provide a limited improvement to the cable rating. A further limiting factor to be
considered is the voltage drop between sending and receiving end, which in most cases
has to be contained in the range of 10% - 15% [14].

9

Another issue that should be taken into consideration at the design phase of an AC
connection for an offshore wind farm is the potential danger of resonance between the
shore step-up transformer and the capacitance of the cable at low harmonic frequencies
inherent in the power system [14]. Resonance phenomena can cause unacceptably high
harmonic currents at the critical frequencies, which will exert high stresses in the
submarine cable(s). Also the amplification of the harmonic currents may in turn cause
amplified harmonic voltages at the offshore wind farm [15].

The overall costs of AC transmission cables are higher than the costs of DC transmission
cables because they can transfer less power per cable core. However, an advantage of
AC transmission is that it requires the smallest and thus the cheapest offshore platform
which represents a major capital cost component. Furthermore, one should include the
higher investment and maintenance costs and operational losses of power converters in
the lifetime cost of DC transmission. If transmission requirements are fulfilled through AC
connection, then an AC connection would appear to be preferable over a HVDC link [14].

4.1.2 AC Connection plus Dynamic Reactive Compensation

Until now Dynamic Reactive Compensation devices in conjunction with wind farms have
been used mainly for voltage control and mitigation of flicker [16, 17]. However, by
specification, these devices can enhance transient stability of the wind turbine generators
and may also contribute to fault current [5, 18, 19]. Consequently, Dynamic Reactive
Compensation devices, along with the inherent capabilities of the generators, will
determine the steady-state and dynamic behaviour of the wind farm as a whole.

There are three main types of Dynamic Reactive Compensation devices:
Static Var Compensators (SVCs)
STATic COMpensators (STATCOMs)
Synchronous Condensers
This report will concentrate on STATCOMs which provide the most capabilities for
enhancing the steady-state and especially the dynamic behaviour of an offshore wind
farm.

A STATCOM is a shunt voltage source inverter coupled to the network through a
transformer or an inductor. A STATCOM can provide or absorb reactive power in a
continuous range. Its output does not depend on the value of the dc link capacitor.
Theoretically the dc link capacitance could have a very small value, however in practice its
value should be large enough in order that dc voltage ripple is small and potential
resonances with the coupling reactance are avoided [20, 21].

There are two main classifications of STATCOM, depending on the switching technique
that is used. A Pulse Width Modulation technique (e.g. Space Vector PWM) permits
independent control of both the amplitude and the angle of the output AC voltage of the
STATCOM, while 180 degree conduction mode or fixed Selective Harmonic Elimination
Modulation techniques lead to a dependent relationship between amplitude and angle of
the output AC voltage. The first technique has higher switching losses and was avoided for
large ratings in the past, but the progress of technical innovation and the reducing cost of
IGBTs makes the use of PWM switching techniques progressively more promising. The
advantage of higher switching frequencies is that smaller harmonic filters can be used. In

10
addition, if a form of energy storage is incorporated to the STATCOM, PWM switching
allows independent control of its real and reactive power output [22-24], enhancing its
capabilities significantly. Possible applications of the combination of STATCOM and
energy storage could include power oscillation damping [24] and load-levelling [23]. If a
sufficient amount of energy storage was incorporated, the latter approach could have an
application for the planning uncertainty inherent in weather forecasting. However, technical
challenges for the incorporation of energy storage into a practical STATCOM remain, and
the current cost of energy storage devices is presently restrictive. The widespread use of
STATCOMs and energy storage for load levelling is therefore unlikely in the near future.

The response time of STATCOMs is less than one cycle [25], making it the fastest
Dynamic Reactive Compensation device available. Compared with SVCs which behave
like a variable susceptance, a STATCOM effectively appears as an emf on the AC side.
Hence its reactive current injection to the system does not depend on the bus voltage and
it can provide, in principle, a sustained fault current during disturbances depending on the
rating of the DC capacitor. Its main disadvantage is the increased capital cost and losses
due to the use of self-commutated power electronic switches. If the application permits, a
good balance between costs and operational behaviour can be achieved with the use of a
STATCOM in conjunction with an SVC [26].

4.2 DC connection options
4.2.1 Voltage Source HVDC

Progress in power electronic switches, especially in IGBTs, and the advent of multilevel
inverter configurations has made direct current transmission possible based on voltage
source converters. Under current technology, Voltage Source HVDC (VS-HVDC) links
have a maximum rated power of approximately 500 MW for DC voltages of 150kV [27].
As more sophisticated configurations of voltage source converters are developed, along
with advances in self-commutated switches, the maximum voltage rating of VS HVDC links
will increase, and consequently their maximum power ratings.

VS-HVDC is inherently bipolar, with each cable carrying half the power [28]. For the
connection of an offshore wind farm to the power supply network, two DC substations
would be required; one onshore and the other at an offshore platform to house the power
electronic converter at the wind farm along with its step-up transformer and harmonic
filters. The use of DC power transmission through submarine cables offers a number of
advantages.

Firstly, when HVDC transmission is employed there is no potential danger of resonance
between the submarine cables and the onshore step-up transformer, since the
capacitance of the cables and the inductance of the transformer are electrically decoupled.
Secondly, submarine cables do not have capacitive losses with DC current transmission.
The power handling capability of DC cables is larger than that of AC cables for the same
insulation level, however one must take into account the additional AC/DC converter
power losses which are not present in an AC system.

In contrast to a classical Line Commutated HVDC scheme a Voltage Source HVDC link
has the capability of independent control between the active power that it transfers and the
reactive power that it absorbs or injects at each terminal [28]. The reactive power
capability at each terminal will depend on the VA rating of the respective converter. VS-

11
HVDC links can also be connected to weak or even completely passive networks without
the need of additional plant, since they employ self-commutated active switches and may
contribute to short circuit power [29]. In comparison to Line Commutated HVDC, VS-HVDC
requires smaller harmonic filters and the substation has an overall smaller footprint. For
the case of offshore wind farms this is important, because of the high construction cost of
offshore platforms. It should be noted however that a VS-HVDC offshore substation
platform will generally still be larger than an AC platform.

The drawback of a VS HVDC link is the increased substation capital cost and the losses
which, when capitalised, represent the bulk of the operating cost. The higher substation
capital cost is attributable to the present price of self-commutated switches. A VS-HVDC
substation is 10 times more expensive than an AC substation [28]. Using present
technology, relatively high losses at the two voltage source converters can result in losses
between 5% and 7% [14]. In comparison, the losses of Line Commutated HVDC, including
transformer losses, have been reduced down to approximately 1.2% [30]. Also, the copper
losses are higher in the case of VS-HVDC, because such schemes must operate with
lower DC voltage than Line Commutated HVDC, resulting in a higher current. Finally, due
to its bipolar nature, VS-HVDC always needs two cables, whereas monopolar Line
Commutated HVDC may use one integrated return cable, reducing cable laying costs.

4.2.2 Line Commutated HVDC

A conventional Line Commutated HVDC (LC HVDC) connection based on thyristor
converters is an alternative to Voltage Source HVDC. Line Commutated HVDC is a well
proven technology for the transmission of bulk power over long distances both on land and
through submarine cable [31]. With present technology LC HVDC links have a maximum
rating of 1300-1500 MW at 500kV dc voltage [14].

LC HVDC can be either mono-polar or bi-polar. The mono-polar link has one high voltage
dc cable, with the return path provided physically by ground or water, or by using a metallic
return path [31]. The use of a metallic return path is necessary when earth restistivity is
high or electromagnetic interference created would be beyond acceptable limits. The
metallic return path is typically a 24kV XLPE cable [14], however cables with an integrated
return path have recently been developed [32] for which case just one cable is required
reducing the cable laying cost. The bi-polar configuration has two converters, one has
positive and the other has negative polarity with respect to ground, and they are of equal
rating connected in series on the dc side. Two high voltage dc cables are used, each
carrying half the power. If each group of converters and cables is designed to carry full
power, and stray currents are within acceptable limits (or a metallic return path has been
used), then in the case of a fault in one group, a switch-over is possible and the scheme
can operate as a fully rated mono-polar link. This will increase the availability at the
expense of increased capital cost.

Drawbacks of the technology are that it needs an active source of voltage at both ends
and LC-HVDC converters always draw reactive power into both terminals. Reactive power
consumption depends on the transmitted real power and it is generally about 50% of the
real power flow [31], but this is largely compensated for by the capacitive harmonic AC
filters. Line Commutated HVDC on its own cannot feed passive systems or be connected
to weak networks and is prone to commutation failure due to voltage drops.


12
De Oliveira [33] and Andersen [34] describe how Line Commutated HVDC may be used to
connect wind farms to the grid. It is necessary to connect a supplementary voltage source
at the farm end in order to provide the necessary commutation voltage for the thyristor
valves and to provide reactive power compensation during normal operation. However, the
supplementary voltage source may only need to be small in rating, in the case of DFIGs or
PMSGs having their own voltage source converters. The additional voltage source can be
a rotating synchronous condenser or a static compensator.

Comparing the two possible HVDC technologies, LC-HVDC is at present probably the best
solution for transmission of very large powers and/or over long distances. Contemporary
Voltage Source HVDC technology has a maximum capacity up to around 500 MW for
distances approximately up to 500 km [14, 27]. For connection of larger wind farms and/or
longer distances, multiple cables and converters have to be used, which increases the
cost prohibitively [35].

13
5. WINDFARM TECHNICAL CAPABILITIES IN RESPECT TO
GENERATOR TECHNOLOGY AND TYPE OF CONNECTION

Generator Technology
FSIG DFIG, PMSG
RP CS CS
FRT NC C
FC C C
G
r
i
d

C
o
d
e

c
o
m
p
l
i
a
n
c
e

PQ CS C
A
C

(
N
o
t
e
:

T
h
e
r
e

c
a
n

b
e

s
w
i
t
c
h
e
d

s
h
u
n
t

e
l
e
m
e
n
t
s
)

Design
considerations
Smaller offshore platform than DC schemes
Conventional connection option
Avoid possibility of harmonic resonance at design stage
Increased losses at the submarine cable
RP C C
FRT C C
FC C C
G
r
i
d

C
o
d
e

c
o
m
p
l
i
a
n
c
e

PQ C C
A
C

+
D
y
n
a
m
i
c

R
e
a
c
t
i
v
e

C
o
m
p
e
n
s
a
t
i
o
n

Design
considerations
Smaller offshore platform than DC schemes, but could be
larger than standard AC designs without dynamic reactive
compensation
Conventional connection option
Avoid possibility of harmonic resonance at design stage
Increased losses at the submarine cable
RP C C
FRT FR FR
FC FR FR
G
r
i
d

C
o
d
e

c
o
m
p
l
i
a
n
c
e

PQ C C
V
o
l
t
a
g
e

S
o
u
r
c
e


H
V
D
C

Design
considerations
Smaller offshore platform than LC HVDC but larger than AC
New technology
Large conversion losses

RP C C
FRT FR FR
FC FR FR
G
r
i
d

C
o
d
e

c
o
m
p
l
i
a
n
c
e

PQ C C
T
y
p
e

o
f

C
o
n
n
e
c
t
i
o
n

L
i
n
e

C
o
m
m
u
t
a
t
e
d

H
V
D
C

Design
considerations
Lower losses
Untested technology for connection of offshore windfarms
Larger offshore platform
Needs emf sources at both ends
Additional reactive compensation at PoC may be necessary



14
KEY

C: Can be made Compliant (depending on engineering design)
CS: Case specific (depending on grid and connecting cable)
FC: Frequency control
FR: Further research is needed
FRT: Fault ride through capability
NC: Non Compliant
PQ: Power quality
RP: Reactive power and voltage control

15
6. MODELLING ASPECTS

Computer simulation is a most powerful tool to investigate the means and capabilities of
different technologies for integrating medium and large offshore wind farms to the power
network. The following is a discussion on the key issues in the literature relating to how the
various components of a wind farm should be modelled.

6.1 Offshore Wind Farm

For stability studies the most pessimistic scenario corresponds to rated mechanical input
to every wind turbine. Representation of the whole wind farm as a coherent machine is
adequate for transient stability studies under fault conditions [10, 36]. Hence, transient
studies with one equivalent machine and rated mechanical input are sufficient for
determining the ride-through capability of the wind farm. For steady state studies a
detailed representation of the wind farm at the design stage is necessary for determining
the required reactive compensation for complying with the Grid Code. Having determined
the exact layout of the wind farm and of the reactive compensation, it is possible to
represent the wind farm as a coherent machine connected to the power supply network
through an equivalent submarine cable. The parameters of the equivalent submarine cable
is chosen in such a way that the simplified model of the wind farm with the one coherent
machine and the detailed representation of the wind farm have, at the PoC, the same
steady state behaviour [36]. Since the major aim of this project is to draw general
guidelines for Grid Code compliance of offshore wind farms, rather than the design of a
specific offshore windfarm, a coherent machine representation of the wind turbines is
adopted. This study approach is considered reasonable both for transient studies under
faults and reactive power capability.

6.2 Mechanical Dynamics of the Wind Turbine Generators

In many studies, the mechanical dynamics are not considered and the wind turbine rotor,
mechanical drive train, gearbox and generator rotor are modelled as a lumped mass with
an equivalent inertia. It is common for this modelling approach to be adopted for
synchronous machines in stability studies [31]. However, wind turbines have very different
mechanical characteristics than conventional plant. They show large inertias and low shaft
stiffness. The interaction between the wind turbine and electrical generator could give rise
to low frequency oscillations that can limit the transient stability of the farm. Representing
the mechanical system as a lumped mass may give optimistic results especially for FSIGs
for which a two-mass representation of the drive train is typically necessary [37]. For
DFIGs and Direct Drive Synchronous Generators, mechanical and electrical frequency are
decoupled and torsional oscillations can be damped by a controller. Therefore a lumped
mass representation is reasonable for DFIGs and Direct Drive Synchronous Generators.
Consequently a two-mass representation of the mechanical system is adopted for FSIGs,
while lumped mass model is adopted for DFIGs.

16
6.3 Power Electronic Converters

Typically power electronic converters have been modelled in system studies using
average time invariant models [35]. Even though time variant representation is feasible in
electromagnetic simulation software like PSCAD/EMTDC, such models increase the
simulation time. One other disadvantage is that switching noise obscures the slower time-
constant response of the machine and control dynamics which are the more important
features for the examination of the wind farm behaviour. Time invariant models of voltage
source converters are based on their representation as controllable ideal voltage sources.
The respective equations have already been analysed for DFIG [38, 39], Voltage Source
HVDC [35] and STATCOM [21]. Average time invariant models of the back-to-back VSIs of
a DFIG and of the VSI of a STATCOM are adopted.

17
7. MODEL DEVELOPMENT AND EVALUATION

Qualitative validation of a generic model of a wind farm with DFIGs (figure 3) and a model
of a wind farm with FSIGs+STATCOM (figure 4) has been undertaken in PSCAD/EMTDC.
This has been based on an aggregated wind farm model with assumed parameters, as
given in Tables 7.1, 7.2 and 7.3. For the evaluation of the model the distance to the power
supply network was 20km. In both cases the aggregated rated power of the wind turbines
is 504MW. The Short Circuit Level of the power supply network is 5000MVA and the X/R
ratio is 14.3.
o
86 5000

Figure 3: Generic model of a 504MW wind farm with DFIGs and without a STATCOM
o
86 5000

Figure 4: Generic model of a 504MW wind farm with FSIGs+250MVAr STATCOM

Table 7.1 FSIG data
Electrical generator data
Rating 3MVA
Stator Voltage (L-L, Rms) 0.69kV
Ls 0.1413pu
Lr 0.05pu
Lm 4.134pu
Rs 0.006pu
Rr 0.007pu
Mechanical system data
Equivalent turbine-blade inertia
(referred to generator side)
3sec
Generator rotor inertia 0.5sec
Shaft stiffness 0.6pu/electrical rad
Step-up transformer data
Rating 3MVA
V
secondary
(L-L, Rms) 0.69kV
V
primary
(L-L, Rms) 33kV
L 0.06pu
No load losses 0.2%
Copper losses 0%
Fixed power factor capacitors (installed at 0.69kV terminals)
3.33 pu



18
Table 7.2 STATCOM data
Voltage Source Inverter data
DC link capacitance (C
dc
): 3.5 pu
Grid side converter coupling inductance (L
grid
): 0.15 pu
Grid side converter coupling resistance (R
grid
): 0.02 pu
AC output voltage (L-L, Rms) 20 kV
Coupling transformer data
Rating
Equal to the rating of the
VSI
V
secondary
(L-L, Rms) 20kV
V
primary
(L-L, Rms) 400kV
L 0.18pu
No load losses 0.5%
Copper losses 0%

Table 7.3 DFIG data
Electrical generator data
Rating 4.5MVA
Stator Voltage (L-L, Rms) 1kV
Ls 0.09241pu
Lr 0.09955pu
Lm 3.95279pu
Rs 0.00488pu
Rr 0.00549pu
H 3.5sec
Range of optimum power absorption 0.75-1.25 pu
Equation of optimum torque
2
r opt
56 . 0 T =
Maximum capacitive reactive power capability 33% of rating
3-winding transformer data
Rating 4.5MVA
V1 (L-L, Rms) 1kV
V2 (L-L, Rms) 0.4kV
V3 (L-L, Rms) 33kV
L12 0.08pu
L23 0.08pu
L12 0.001pu
No load losses 0.35%
Copper losses 0%
Frequency converter data
C 3.5pu
Vdc 0.7kV
Lcoupling 1 pu (per 0.4kV base)
Grid side VSI
Rcoupling 0.017pu (per 0.4kV base)
Lcoupling 0.2pu (per 0.69kV base)
Rotor side VSI
Rcoupling 0pu (per 0.69kV base)

The DFIG generator is controlled according to the scheme presented by the DTI Centre for
Sustainable Electricity and Distributed Generation [7]. According to the scheme, the rotor

19
side Voltage Source Inverter (VSI) controls the rotor speed and the reactive power output
of the machine in a decoupled manner. The grid side VSI is reactive power neutral and its
main purpose is to keep the dc link voltage at its nominal value, establishing the power
balance between the two back-to-back converters. The control strategy for the rotor speed
maintains optimal power extraction from the wind. Reactive power can be controlled either
to keep a constant power factor at the terminals of the machine or to control the voltage at
a specific electrical point. In the examined study case, reactive power is controlled in order
to regulate the PoC voltage. In addition, the DC link of the DFIG is equipped with a
breaking resistor for overvoltage protection.

The control strategy of the STATCOM follows the work of Schauder and Mehta [22]. A
STATCOM with IGBT switches and a SV-PWM switching technique is considered. This
allows independent control of the amplitude and the phase of the STATCOM injected AC
voltage. The reactive power output of the STATCOM is controlled in order to regulate the
PoC voltage.

The qualitative validation procedure is based on:
Comparison of the behaviour of the models with previous research.
Comparison between the results obtained with the developed average time
invariant models (average models) and the results obtained with more detailed
models of the VSIs where the power electronic switches are represented
(switched models).

The simulation scenario for both wind farms comprises a step change in real power from
0.4pu to 0.8pu at 5sec, a step change in the PoC voltage command from 1pu to 1.02pu at
15sec, and a 3-phase bolted fault at 20sec at the PoC lasting 140ms. The results are
shown in figures 5 and 6, and are discussed in the next sections

The goal of the study cases at this stage was to evaluate the operation of the main system
components: FSIG and DFIG wind turbine models and STATCOM models. These models
will be used to assess windfarm behaviour against a test-matrix of Grid Code
requirements. Architectures comprising FSIG turbines with a STATCOM, DFIG turbines
only and DFIG turbines with a STATCOM will be considered in these subsequent studies.









20
0 5 10 15 20 25 30
0.8
0.9
1
1.1
1.2
Time (sec)
w

(
p
u
)
0 5 10 15 20 25 30
0
0.5
1
Time (sec)
P
o
C

V
o
l
t
a
g
e

(
p
u
)
0 5 10 15 20 25 30
-200
0
200
400
600
Time (sec)
P

f
a
r
m

(
M
W
)
0 5 10 15 20 25 30
-500
-250
0
250
Time (sec)
Q

f
a
r
m

(
M
V
A
r
)
0 5 10 15 20 25 30
0
50
100
150
200
250
Time (sec)
R
o
t
o
r

C
u
r
r
e
n
t

(
%
)
0 5 10 15 20 25 30
50
100
150
200
Time (sec)
D
C

l
i
n
k

V
o
l
t
a
g
e

(
%
)
average
switched

Figure 5: Wind farm with DFIGs and without a STATCOM

0 5 10 15 20 25 30
0.9
1
1.1
Time (sec)
w

(
p
u
)
0 5 10 15 20 25 30
0
0.5
1
Time (sec)
P
o
C

V
o
l
t
a
g
e

(
p
u
)
0 5 10 15 20 25 30
-200
0
200
400
600
800
Time (sec)
P

f
a
r
m

(
M
W
)
0 5 10 15 20 25 30
-800
-600
-400
-200
0
200
400
Time (sec)
Q

f
a
r
m

(
M
V
A
r
)
0 5 10 15 20 25 30
-200
0
200
400
Time (sec)
Q
s
t
a
t
c
o
m

(
M
V
A
r
)
0 5 10 15 20 25 30
50
100
150
Time (sec)
D
C

l
i
n
k

V
o
l
t
a
g
e

(
%
)
average
switched

Figure 6: Wind farm with FSIGs+STATCOM


21
7.1 Wind farm with DFIGs and without a STATCOM

In this study case with the assumed DFIG parameters, the windfarm can meet the Grid
Code reactive power capability requirement without the employment of a Dynamic
Reactive Compensation device. This, however, may change if different submarine cable
lengths and/or reactive power capability of the machine are considered.

For the equivalent DFIG wind turbine, the capability of independent control of real and
reactive power output is validated in figure 5. The change of the real power output at
t=5sec does not affect the reactive power output, and the change in the reactive power at
t=15sec does not affect the real power output. When the step-change in the mechanical
input occurs at t=5sec, it takes approximately 10sec in order that electrical output reaches
0.8pu, indicating that the dynamic performance of the speed control is dominated by the
combined turbine and machine inertia [7]. A transient DFIG rotor overcurrent occurs as a
result of the sudden rotor flux change at fault occurrence [7, 41-44]. As the stator voltage
is collapsed during the fault, the grid side converter cannot remove the power that is fed
into the dc link by the rotor side converter, resulting in a dc link overvoltage [45]. The
braking resistor keeps the DC link voltage below the threshold of 150%. A magnified view
of the behaviour of the farm during the fault is presented in figure 7.

19.5 20 20.5 21 21.5 22
1.1
1.15
1.2
Time (sec)
w

(
p
u
)
19.5 20 20.5 21 21.5 22
0
0.5
1
Time (sec)
P
o
C

V
o
l
t
a
g
e

(
p
u
)
19.5 20 20.5 21 21.5 22
-200
0
200
400
600
Time (sec)
P

f
a
r
m

(
M
W
)
19.5 20 20.5 21 21.5 22
-500
-250
0
250
Time (sec)
Q

f
a
r
m

(
M
V
A
r
)
19.5 20 20.5 21 21.5 22
0
50
100
150
200
250
Time (sec)
R
o
t
o
r

C
u
r
r
e
n
t

(
%
)
19.5 20 20.5 21 21.5 22
50
100
150
200
Time (sec)
D
C

l
i
n
k

V
o
l
t
a
g
e

(
%
)
average
switched

Figure 7: Fault behaviour of windfarm with DFIGs and without a STATCOM

22
7.2 Wind farm with FSIGs+STATCOM

In this study case a STATCOM of 250 MVA with a 130% overload capability for 1 sec had
to be used in order to ensure dynamic stability. Note that it may be possible with further
system and control development to reduce the STATCOM rating, but the main purpose
here was to develop the study case and the machine and STATCOM models.

At t=5sec when the step change in the mechanical power input occurs and at fault
clearance, the equivalent FSIG wind turbine shows oscillations in speed and real power.
The observed low frequency power oscillations are caused by the stiff coupling between
mechanical system and electrical behaviour of Fixed Speed Induction Generators and are
characteristic of FSIG wind turbines [37, 45]. The STATCOM regulates the voltage at the
PoC effectively, as can be seen in Figure 6 when at t=15sec the voltage command
changes. During the fault and after fault clearance the STATCOM provides maximum
reactive current, compensating the large amounts of reactive power that the equivalent
FSIG draws as a result of its overspeeding [20]. During the fault the PoC voltage collapses
and STATCOM cannot draw enough real power to compensate for its losses, resulting in a
DC link voltage drop. At fault clearance an overshoot is observed in the DC link voltage of
the STATCOM, which however is not dangerous. This overshoot is characteristic of PI
controllers and can be ameliorated with integrator anti-windup [47].

19.5 20 20.5 21 21.5 22
0.9
1
1.1
Time (sec)
w

(
p
u
)
19.5 20 20.5 21 21.5 22
0
0.5
1
Time (sec)
P
o
C

V
o
l
t
a
g
e

(
p
u
)
19.5 20 20.5 21 21.5 22
-200
0
200
400
600
800
Time (sec)
P

f
a
r
m

(
M
W
)
19.5 20 20.5 21 21.5 22
-800
-600
-400
-200
0
200
400
Time (sec)
Q

f
a
r
m

(
M
V
A
r
)
19.5 20 20.5 21 21.5 22
-200
0
200
400
Time (sec)
Q
s
t
a
t
c
o
m

(
M
V
A
r
)
19.5 20 20.5 21 21.5 22
50
100
150
Time (sec)
D
C

l
i
n
k

V
o
l
t
a
g
e

(
%
)
average
switched

Figure 8: Fault behaviour of windfarm with FSIGs+STATCOM


23
7.3 Comparison between average and switched models

For both wind farms, the agreement between the results with average and switched
models is very good, indicating that the average models give a sufficient representation.
Under steady state conditions these studies demonstrate that the main discrepancies are
in the real power output in the case of DFIGs. In the case of the STATCOM in the
FSIG+STATCOM studies the main discrepancy is in reactive power. These are shown
magnified for the period between 25 and 30sec in Figure 9. Studies showed that when the
switching frequency was increased, the discrepancies reduced, as expected since the
switching function then more nearly approximates the average model. Using an average
model also reduces secondary effects such as noise coupled to the PLL control block,
which otherwise cause further discrepancies. The additional noise produced when the
converters are represented as non-averaged, time-variant devices is evident in figure 9.

During faults, the agreement between the average and switched model is very good for
DFIGs. The main observed difference when the switching function is represented, is the
additional noise in the dc link and the rotor current. For the STATCOM, the main
discrepancy is the pronounced overshoot of the DC link voltage at fault clearance. Still, the
difference is small and the results produced lie on the conservative side. The discrepancy
can be attributed to the higher losses in the VSI as a result of the switching harmonic
currents, which are represented only in the switched model, and introduce additional
damping in the DC link voltage loop.

25 25.5 26 26.5 27 27.5 28 28.5 29 29.5 30
80
85
90
95
Time (sec)
Q
s
t
a
t
o
c
m

(
M
V
A
r
)
25 25.5 26 26.5 27 27.5 28 28.5 29 29.5 30
380
385
390
395
Time (sec)
P

f
a
r
m

(
M
W
)
average
switched

Figure 9: Main discrepancies between average and switched models

24
8. CONCLUDING REMARKS AND FURTHER WORK

This report has summarised the technical options for connecting an offshore windfarm to
the mainland transmission and distribution network. Each combination has benefits and
drawbacks depending on the size, technology, location and architecture of the windfarm.
Given the scale of any such project, the balance of initial project costs, lifetime, operating
and maintenance costs need to be evaluated

Preliminary investigations showed that for the case of an offshore windfarm with DFIGs
and an AC connection to shore, cable length, PoC connection voltage, short circuit level of
the system and the machine characteristics in terms of reactive power limits and steady
state voltage limits have a significant impact in determining whether additional reactive
power compensation is needed at the PoC. Such issues are important for ensuring the
wind farm can comply with the Grid Code requirements for reactive power capability. For
the case of an offshore windfarm with FSIGs+STATCOM and an AC connection to shore,
the STATCOM rating is mainly determined for meeting the Grid Code requirement on fault
ride-through, not the steady state reactive power capability obligation. However, the
required size of the STATCOM depends considerably on cable length, PoC connection
voltage level, the systems strength at the PoC and the electrical and mechanical
characteristics of the wind turbine generators.

Subsequent studies will concentrate on qualitative aspects of STATCOM performance and
the required rating to achieve Grid Code compliance, on the sensitivity associated with
varying:
System fault level
Connection voltage at the PoC
Wind farm rating
Cable length
STATCOM placement.


25
REFERENCES

[1] "Our energy future - creating a low carbon economy," Energy White Paper, DTI,
2003.
[2] The Carbon Trust and DTI Renewables Network Impact Study, Mott MacDonald,
April 2004, Carbon Trust,.
[3] "Annex 2: Transmission Network Topography Analysis," The Carbon Trust and DTI
Renewables Network Impact Study, Mott MacDonald, December 2003, Carbon
Trust.
[4] A. Johnson and N. Tleis, "The development of Grid Code requirements for new and
renewable forms of generation in Great Britain," Fifth International Workshop on
Large-Scale Integration of Wind Power and Transmission Networks for Offshore
Wind Farms, Glasgow, Scotland, 2005.
[5] "The Grid Code, Issue 3, Revision 12," National Grid Electricity Transmission plc,
available at
http://www.nationalgrid.com/uk/Electricity/Codes/gridcode/gridcodedocs/.
[6] N. Jenkins, R. Allan, P. Crossley, D. Kirchen, and G. Strbac, "Embedded
Generation," IEE Power and Energy Series, 2000.
[7] O. Anaya-Lara and N. Jenkins, "Modelling and Control of DFIGs for Wind Energy
Generation," Notes in the context of MSc in "Power Systems Engineering",
University of Manchester, 2005.
[8] S. Muller, M. Deicke, and R. W. De Doncker, "Doubly fed induction generator
systems for wind turbines," IEEE Industry Applications Magazine, pp. 26-33, 2002.
[9] S. Achilles and M. Poller, "Direct Drive Synchronous Machine Models for Stability
Assessment of Wind Farms," DIgSILENT GmbH, available at
http://www.digsilent.de/Consulting/Publications/DirectDrive_Modeling.pdf.
[10] V. Akhmatov, "Analysis of Dynamic Behaviour of Electric Power Systems with Large
Amount of Wind Power," PhD Thesis, Electric Power Engineering, Orsted-DTU,
Technical University of Denmark, 2003.
[11] Garrad Hassan & Partners, Tractebel Energy Engineering, Ris National
Laboratory, Kvaerner Oil & Gas, and Energi & Miljoe Undersoegelser (EMU), "Final
Report: Concerted Action on Offshore Wind Energy in Europe," available at
http://www.offshorewindenergy.org/, 2001.
[12] G. Takoudis, G. Ault, S. Gair, and J. McDonald, "Method for Assessing Offshore
Wind Farm Cable Reliability Incorporating Cost Effectiveness of Redundancy," Fifth
International Workshop on Large-Scale Integration of Wind Power and
Transmission Networks for Offshore Wind Farms, Galsgow, Scotland, 2005.
[13] J. R. Kristoffersen and P. Christiansen, "Horns Rev offshore windfarm: its main
controller and remote control system," Wind Engineering, pp. 351-9, 2003.
[14] G. Balog and N. Christl, "Transmission Strategies for Offshore Windmill Parks,"
Fifth International Workshop on Large-Scale Integration of Wind Power and
Transmission Networks for Offshore Wind Farms, Galsgow, Scotland, 2005.
[15] S. Santoso, H. W. Beaty, R. C. Dugan, and M. F. McGranaghan, "Electrical Power
Systems Quality," 2d ed: McGraw Hill, 2002.
[16] ELTRA, "Power Quality Improvements of Wind Farms," Fredericia, 1998.
[17] C. Rasmussen, P. Jorgensen, and J. Hasvager, "Improving voltage quality in
Eastern Denmark with a Dynamic Phase Compensator," Fifth International
Workshop on Large-Scale Integration of Wind Power and Transmission Networks
for Offshore Wind Farms, Galsgow, Scotland, 2005.

26
[18] V. Akhmatov, H. Knudsen, A. H. Nielsen, J. K. Pedersen, and N. K. Poulsen,
"Modelling and transient stability of large wind farms," International Journal of
Electrical Power & Energy Systems, pp. 123-44, 2003.
[19] S. K. Salman and A. L. J. Teo, "Improvement of fault clearing time of wind farm
using reactive power compensation," Proceedings of 2001 Power Tech, 10-13 Sept.
2001, pp. 6 pp. vol.2, 2001.
[20] X. Wu, A. Arulampalam, C. Zhan, and N. Jenkins, "Application of a Static Reactive
Power Compensator (STATCOM) and a Dynamic Braking Resistor (DBR) for the
stability enhancement of a large wind farm," Wind Engineering, pp. 93-106, 2003.
[21] S. Dong, W. Zhonghong, J. Y. Chen, and Y. H. Song, "Harmonic resonance
phenomena in STATCOM and relationship to parameters selection of passive
components," IEEE Transactions on Power Delivery, pp. 46-52, 2001.
[22] C. Schauder, "Vector analysis and control of advanced static VAr compensators,"
IEE Proceedings-Generation, Transmission and Distribution, pp. 299-306, 1993.
[23] S. Atcitty and S. Ranade, "Summary of State-of-the-Art PCS System Configurations
and Recommendations for Future Research and Development," Sandia Report
SAND98-2019, 1998.
[24] G. Ariyoshi, K. Murata, K. Harada, and K. Yamasaki, "Load leveling using EDLCs
under PLL control," IEICE Transactions on Fundamentals of Electronics,
Communications and Computer Sciences International Technical Conference on
Circuits/Systems, Computers and Communications (ITC-CSCC'99), 13-15 July
1999, pp. 1014-22, 2000.
[25] Siemens Power Transmission and Distribution Inc, "Benefit of Static Compensator
(STATCOM) plus Superconducting Magnetic Energy Storage (SMES) in the
Transmission Network," Spring 2001 Energy Storage Association Meeting, 2001.
[26] C. Schauder, E. Stacey, M. Lund, L. Gyugyi, L. Kovalsky, A. Keri, A. Mehraban, and
A. Edris, "AEP UPFC project: installation, commissioning and operation of the +or-
160 MVA STATCOM (phase I)," IEEE Transactions on Power Delivery, pp. 1530-5,
1998.
[27] C. Horwill, A. J. Totterdell, D. J. Hanson, D. R. Monkhouse, and J. J. Price,
"Commissioning of a 225 Mvar SVC incorporating A +or-75 Mvar STATCOM at
NGC's 400 kV East Claydon substation," presented at Proceedings of AC and DC
Transmission, 28-30 Nov. 2001, London, UK, 2001.
[28] D. Wensky and J. Bernauer, "Conceptual Design of Offshore Wind Power Grid
Connections under Special Consideration of Minimised Investment and Life Cycle
Costs for Losses," Fifth International Workshop on Large-Scale Integration of Wind
Power and Transmission Networks for Offshore Wind Farms, Galsgow, Scotland,
2005.
[29] T. Ackermann, "Transmission systems for offshore wind farms," IEEE Power
Engineering Review, pp. 23-7, 2002.
[30] Y. Jiang-Hafner, M. Hyttinen, and B. Paajarvi, "On the short circuit current
contribution of HVDC Light," presented at Proceedings Asia Pacific Conference and
Exhibition of the IEEE-Power Engineering Society on Transmission and Distribution,
6-10 Oct. 2002, Yokohama, Japan, 2002.
[31] M. Takasaki, N. Gibo, and T. Hayashi, "Evaluation of the relation between voltage
sourced converter performance and design parameters," presented at 7th
International Conference on AC-DC Power Transmission, Nov 28-30 2001, London,
United Kingdom, 2002.
[32] P. Kundur, "Power System Stability and Control," McGraw-Hill, Inc., 1994.

27
[33] I. Atkinson, C. Harvey, M. Smith, P. Damgaard, M. Haeusler, M. Kuhn, P. Lips, M.
Wohlmuth, G. Balog, and K. Stenseth, "The Moyle interconnector," Power
Engineering Journal, pp. 117-128, 2002.
[34] M. De Oliveira, M. Poloujadoff, A. Le Du, and P. G. Therond, "Supply of an entirely
passive AC system through an HVDC link," International Journal of Electrical Power
and Energy System, pp. 111-116, 1994.
[35] B. R. Andersen and L. Xu, "Hybrid HVDC system for power transmission to island
networks," IEEE Transactions on Power Delivery, pp. 1884-90, 2004.
[36] P. Cartwright, "Power electronics based applications for the increased penetration
of wind power into electrical networks," PhD Thesis, Electrical Engineering and
Electronics, UMIST, 2004.
[37] P. D. Hopewell, W. W. Price, N. W. Miller, and W. Liu, "Modelling and Simulation of
Wind Turbine Generators in Large Offshore Applications," Fifth International
Workshop on Large-Scale Integration of Wind Power and Transmission Networks
for Offshore Wind Farms, Galsgow, Scotland, 2005.
[38] V. Akhmatov and H. Knudsen, "An aggregate model of a grid-connected, large-
scale, offshore wind farm for power stability investigations-importance of windmill
mechanical system," International Journal of Electrical Power & Energy Systems,
pp. 709-17, 2002.
[39] P. Cartwright, L. Holdsworth, J. B. Ekanayake, and N. Jenkins, "Co-ordinated
voltage control strategy for a doubly-fed induction generator (DFIG)-based wind
farm," IEE Proceedings: Generation, Transmission and Distribution, pp. 495-502,
2004.
[40] V. Akhmatov, "Variable-speed wind turbines with doubly-fed induction generators. I.
Modelling in dynamic simulation tools," Wind Engineering, pp. 85-108, 2002.
[41] V. Akhmatov, "Variable-speed wind turbines with doubly-fed induction generators.
II. Power system stability," Wind Engineering, pp. 171-88, 2002.
[42] J. B. Ekanayake, L. Holdsworth, and N. Jenkins, "Comparison of 5th order and 3rd
order machine models for doubly fed induction generator (DFIG) wind turbines,"
Electric Power Systems Research, pp. 207-15, 2003.
[43] J. Ekanayake and N. Jenkins, "Comparison of the response of doubly fed and fixed-
speed induction generator wind turbines to changes in network frequency," IEEE
Transactions on Energy Conversion, pp. 800-2, 2004.
[44] M. Hogdahl and J. G. Nielsen, "Modelling of the Vestas V80 VCS wind turbine with
low voltage ride-through," Fifth International Workshop on Large-Scale Integration
of Wind Power and Transmission Networks for Offshore Wind Farms, Galsgow,
Scotland, 2005.
[45] T. Sun, Z. Chen, and F. Blaabjerg, "Transient stability of DFIG wind turbines at an
external short-circuit fault," Wind Energy, pp. 345-360, 2005.
[46] S. K. Salman and A. L. J. Teo, "Windmill modelling consideration and factors
influencing the stability of a grid-connected wind power based embedded
generator," presented at 2003 IEEE Power Engineering Society General Meeting,
13-17 July 2003, Toronto, Ont., Canada, 2003.
[47] G. F. Franklin, J. D. Powell, and A. Emami-Naeini, "Feedback control of dynamic
systems," World Student Series, 1994.


28
GLOSSARY OF TERMS

BC Balancing Code
CC Connection Conditions
DFIG Doubly Fed Induction Generator
DTI Department of Trade and Industry
FSIG Fixed Speed Induction Generator
HVDC High Voltage Direct Current
IGBT Insulated Gate Bipolar Transistor
LC HVDC Line Commutated HVDC
NGC National Grid Company plc.
PCC Point of Common Coupling
p.f. Power factor
PWM Pulse Width Modulation
PLL Phase Locked Loop
SHEM Selective Harmonic Elimination Modulation
SHETL Scottish Hydro-Electric Transmission Limited
SPT Scottish Power Transmission Limited
STATCOM Static Compensator
SVC Static Var Compensator
T&D Transmission & Distribution
VS HVDC Voltage Source HVDC

Anda mungkin juga menyukai